nlin0207049/sgr9.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %       Kottos-Smilansky
3: %       Long paper on chaotic scattering on Graphs
4: %       June 2002
5: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6: %-----------------------LaTex------------------------------
7: \documentclass[12pt]{iopart}
8: \usepackage{iopams}
9: \usepackage{epsf}
10: \usepackage{euscript}
11: %\usepackage{graphicx}
12: %%% Various definitions of bold Z, Q, R, N. %%%
13: \def\bbbone{{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l}
14: {\rm 1\mskip-4.5mu l} {\rm 1\mskip-5mu l}}}
15: \def\bbbn{{\rm I\!N}}
16: \def\bbbr{{\rm I\!R}}
17: \def\bbbq{{\mathchoice {\setbox0=\hbox{$\displaystyle\rm Q$}\hbox{\raise
18: 0.15\ht0\hbox to0pt{\kern0.4\wd0\vrule height0.8\ht0\hss}\box0}}
19: {\setbox0=\hbox{$\textstyle\rm Q$}\hbox{\raise
20: 0.15\ht0\hbox to0pt{\kern0.4\wd0\vrule height0.8\ht0\hss}\box0}}
21: {\setbox0=\hbox{$\scriptstyle\rm Q$}\hbox{\raise
22: 0.15\ht0\hbox to0pt{\kern0.4\wd0\vrule height0.7\ht0\hss}\box0}}
23: {\setbox0=\hbox{$\scriptscriptstyle\rm Q$}\hbox{\raise
24: 0.15\ht0\hbox to0pt{\kern0.4\wd0\vrule height0.7\ht0\hss}\box0}}}}
25: \def\bbbz{{\mathchoice {\hbox{$\sf\textstyle Z\kern-0.4em Z$}}
26: {\hbox{$\sf\textstyle Z\kern-0.4em Z$}}
27: {\hbox{$\sf\scriptstyle Z\kern-0.3em Z$}}
28: {\hbox{$\sf\scriptscriptstyle Z\kern-0.2em Z$}}}}
29: 
30: \begin{document}
31: \renewcommand{\baselinestretch}{1.5}
32: \title{ Quantum Graphs: A simple model for Chaotic Scattering}
33: \author{Tsampikos Kottos$^1${\footnote {corresponding author:
34: tsamp@chaos.gwdg.de}}
35: and Uzy Smilansky$^2$ \\
36: $^1$ Max-Planck-Institut f\"ur Str\"omungsforschung, 37073 G\"ottingen,
37: Germany,\\
38: $^2$ Department of  Physics of Complex Systems,
39: The Weizmann Institute of Science, 76100 Rehovot, Israel}
40: \date{\today }
41: 
42: \begin{abstract}
43: We connect quantum graphs with infinite leads, and turn them to scattering systems.
44: We show that they display all the features which characterize quantum scattering
45: systems with an underlying classical chaotic dynamics: typical poles, delay time and
46: conductance distributions, Ericson fluctuations, and when considered statistically,
47: the ensemble of scattering matrices reproduce quite well the predictions of 
48: appropriately defined Random Matrix ensembles. The underlying classical dynamics
49: can be defined, and it provides important parameters which are needed for the quantum
50: theory. In particular, we derive exact expressions for the scattering matrix, and an
51: exact trace formula for the density of resonances, in terms of classical orbits,
52: analogous to the semiclassical theory of chaotic scattering. We use this in order to
53: investigate the origin of the connection between Random Matrix Theory and the
54: underlying classical chaotic dynamics. Being an exact theory, and due to its relative
55: simplicity, it offers new insights into this problem which is at the fore-front of
56: the research in chaotic scattering and related fields.\\
57: 
58: \hspace {-0.5cm} submitted to J. Phys. A Special Issue -- Random Matrix Theory
59: \end{abstract}
60: 
61: 
62: %-------------------------------------------------------------------------
63: 
64: \section{\bf Introduction}
65: \label{sec:introduction}
66: 
67: Quantum graphs of one-dimensional wires connected at nodes were introduced already
68: more than half a century ago to model physical systems. Depending on the envisaged
69: application the precise formulation of the models can be quite diverse and ranges
70: from solid-state applications to mathematical physics 
71: \cite{A85,FJK87,ML77,A81,CC88,NYO94,Ibook,MT01,Z98,V98,E95,R83}. Lately, quantum 
72: graphs attracted also the interest of the quantum chaos community because they can 
73: be viewed as typical and yet relatively simple examples for the large class of 
74: systems in which classically chaotic dynamics implies universal correlations in the 
75: semiclassical limit \cite{KS97,KS99,A99,BK99,SS00,T00,BSW02,BG00,K01}. Up to now we 
76: have only a limited understanding of the reasons for this universality, and quantum 
77: graph models provide a valuable opportunity for mathematically rigorous investigations 
78: of the phenomenon. In particular, for quantum graphs an exact trace formula exists 
79: \cite{R83,KS97,KS99} which is based on the periodic orbits of a mixing classical 
80: dynamical system. Moreover, it is possible to express two-point spectral correlation 
81: functions in terms of purely combinatorial problems \cite{BK99,SS00,T00,BSW02}.
82: 
83: By attaching infinite leads at the vertices, we get non-compact graphs, for which a 
84: scattering theory can be constructed \cite{KS00,BG01,TM01}. They display many of the 
85: feature that characterize scattering systems with an underlying chaotic classical 
86: dynamics \cite{M73,S89,GR89,RBUS90,EVP99}, and they are the subject of the present 
87: paper. The {\it quantum} scattering matrix for such problems can be written explicitly, 
88: together with a trace formula for the density of its resonances. These expressions 
89: are the analogues of the corresponding semiclassical approximations available in the 
90: theory of chaotic scattering \cite{M73,S89,GR89}, albeit here they are exact. With 
91: these tools we analyze the distribution of resonances, partial delay times, and the 
92: statistics of the fluctuating scattering amplitudes and cross sections. Moreover, 
93: we address issues like the statistical properties of the ensemble of the scattering 
94: matrices, and conductance distribution. Finally we analyze the effect of non-uniform  
95: connectivity on the statistical properties of the scattering matrix. A main part of 
96: our analysis will be focused on the comparison of the statistical properties of the 
97: above quantities with the Random Matrix Theory (RMT) predictions 
98: \cite{M75,MPS85,JSA92,HILSS92,BB94,FS96b,FS96,FS97,BFB97,GM98,MB99}.
99: 
100: The paper is structured in the following way. In section (\ref {sec:definitions}), 
101: the mathematical model is  introduced and the main definitions are given. Section 
102: (\ref{sec:s-matrix}) is devoted to the derivation of the  scattering matrix for
103: graphs. The trace formula for the  density of resonances of quantum graphs is
104: presented in section (\ref {sec:classical}) which includes also the analysis of 
105: the underlying classical  system. The next section, is dedicated to the analysis 
106: of various statistical  properties of the $S$-matrix. Our numerical data are
107: compared both with the predictions of RMT and with the semiclassical expectations. 
108: Finally, in section (\ref {sec:star}), we analyze the class of star-graphs for which 
109: several results can be analytically derived. Our conclusions are summarized in 
110: section (\ref {sec:conclusions}).
111: 
112: %-------------------------------------------------------------------------
113: 
114: \section{\bf Quantum Graphs: Definitions}
115: \label{sec:definitions}
116: 
117: We start by considering a {\it compact} graph ${\cal G}$. It consists of $V$ {\it 
118: vertices} connected by $B$ {\it bonds}. The number of bonds which emanate from each 
119: vertex $i$ defines the valency $v_i$ of the corresponding vertex (for simplicity we 
120: will allow only a single bond between any two vertices). The graph is called $v-regular$ 
121: if all the vertices have the same valency $v$. The total number of bonds is $B={1\over 
122: 2} \sum_{i=1} ^V v_{i}$. Associated to every graph is its connectivity matrix $C$. It 
123: is a square matrix of size $V$ whose matrix elements $C_{i,j}$ take the values $1$ if 
124: the vertices $i,j$ are connected with a bond, or $0$ otherwise. The bond connecting 
125: the vertices $i$ and $j$ is denoted by $b \equiv (i,j)$, and we use the convention 
126: that $i<j$. It will be sometimes convenient to use the ``time reversed" notation, where 
127: the first index is the larger, and $\hat b \equiv (j,i)$ with $j>i$. We shall also use 
128: the directed bonds representation, in which $b$ and $\hat b$  are distinguished as two 
129: directed bonds conjugated by time-reversal. We associate the  natural metric to the 
130: bonds, so that $x_{i,j}\  (x_{j,i})$  measures the distance from the vertex $i\  (j)$ 
131: along the bond.  The length of the bonds are denoted by $L_{b}$ and we shall henceforth 
132: assume that they are {\it rationally independent}. The mean length is defined by 
133: $\left<L \right>\equiv (1/B) \sum_{b=1}^B L_b$ and in  all numerical calculations 
134: bellow it will be taken to be $1$. In the {\it directed- bond} notation $L_{b} = 
135: L_{\hat b}$.
136: 
137: The {\it scattering} graph ${\tilde {\cal G}}$ is obtained by adding leads which
138: extend from  $M (\leq V)$ vertices to infinity.  For simplicity we connect at most
139: one lead to any vertex. The valency of these  vertices increases to ${\tilde v}_i
140: =v_i+1$. The $M$ leads are denoted by the index $i$ of the vertex to which they are
141: attached while $x_i$ now measures the distance from the vertex along the lead $i$.
142: 
143: The Schr\"odinger operator (with $\hbar=2m=1$) is defined on the graph ${\tilde {\cal
144: G}}$ in the following way: On the bonds $b$, the components $\Psi_b$ of the total wave
145: function $\Psi$ are solutions of the one - dimensional equation
146: \begin{equation}
147: \label{schrodinger}
148: \left( -i{\frac{{\rm d\ \ }}{{\rm d}x}}-A_b\right) ^2\Psi _b(x)=k^2\Psi
149: _b(x), \,\,\,\,\,\,\,\,\,\,\,\, \bigskip\ b=(i,j)
150: \end{equation}
151: where $A_b$ (with $\Re e(A_{b})\ne 0$ and $A_{b}= -A_{\hat b}$) is a ``magnetic
152: vector potential" which breaks the time reversal symmetry. In most applications
153: we shall assume that all the $A_{b}$'s are equal and the bond index will be dropped.
154: The components of the wave functions on the leads, $\Psi_i(x)$, are solutions of
155: \begin{equation}
156: \label{schrodinger1}
157: -{\frac{{\rm d^2\ \ }}{{\rm d}x^2}}\Psi _i(x)=k^2\Psi_i(x),
158: \,\,\,\,\,\,\,\,\,\,\,\, \bigskip\ i=1,...,M.
159: \end{equation}
160: At the vertices, the wavefunction satisfies boundary conditions which ensure current 
161: conservation. To implement the boundary conditions, the components of the wave function 
162: on each of the bonds $b$ and the leads $i$ are expressed in terms of counter propagating  
163: waves with a wave-vector $k$:
164: \begin {eqnarray}
165: &&{\rm On\ the \  bonds:\ } \Psi_{b} = a_{b} {\rm e}^{i(k+A_{b})x_{b}}
166: +c_{ b} {\rm e}^{i(-k+A_{b})x_{b}} \nonumber  \\
167: &&{\rm On\  the \ leads\ :\ } \Psi_i =  I_{i} {\rm e}^{-ikx_{i}} + O _{i}
168: {\rm e}^{ ikx_{i}} \ .
169: \label{wfun1}
170: \end{eqnarray}
171: The amplitudes $a_b,c_b$ on the bonds and $I_i,O_i$ on the lead are related  by
172: \begin{eqnarray}
173: \label{vertexcond}
174: {\hspace {-20mm}}
175: {\left (
176: \begin {array} {c}
177:    O_i  \\  a_{i,j_1}\\ \cdot  \\   a_{i,j_{v_i}}
178:   \end {array}
179: \right )} =
180: \Sigma^{(i)}
181: {\left ( \begin{array} {c}
182:    I_i  \\ c_{j_1,i} \\ \cdot  \\ c_{j_{v_i},i}
183:   \end {array} \right ) }, \quad\quad \quad\quad
184: \Sigma^{(i)}=
185: {\left (
186: \begin{array}{cccc}
187: \rho^{(i)}  & \tau^{(i)}_{j_1} & \cdot   & \tau^{(i)}_{j_{v_i}} \\
188: \tau^{(i)}_{j_1} & \tilde \sigma^{(i)}_{j_1,j_1} & \cdot  &
189: \tilde \sigma^{(i)}_{j_1,j_{v_i}} \\
190: \cdot  & \cdot &\cdot &\cdot \\
191: \tau^{(i)}_{j_{v_i}} & \tilde \sigma^{(i)}_{{j_{v_i}},j_1} &
192: \cdot & \tilde \sigma^{(i)}_{j_{v_i},j_{v_i}} \\
193: \end{array}
194: \right)}.
195: \end{eqnarray}
196: These equalities impose the  boundary conditions at the vertices. The vertex scattering 
197: matrices $\Sigma^{(i)}_{j,j'}$, are $\tilde v_i \times \tilde v_i$ unitary symmetric 
198: matrices, and $j,j'$ go over all the $v_i$ bonds and the lead which emanate from $i$. 
199: The unitarity of $\Sigma^{(i)}$   guarantees current conservation at each vertex.
200: 
201: On the right hand side of (\ref{vertexcond}), the vertex scattering matrix $\Sigma^{(i)}$
202: was written explicitly in terms of the vertex reflection amplitude $\rho^{(i)}$, the
203: lead-bond transmission amplitudes $\{ \tau^{(i)}_{j}\}$, and the $v_i\times v_i$ bond-bond
204: transition matrix $\tilde \sigma^{(i)}_{j,j'}$, which is {\it sub unitary} ($|\det \tilde
205: \sigma^{(i)}|<1 $), due to the coupling to the leads.
206: 
207: Graphs  for which there are no further requirements on the $\Sigma^{(i)}$ shall be 
208: referred  to as {\it  generic}. It is often convenient to compute the vertex scattering
209: matrix  from a requirement that the wave function is continuous and satisfies $Neumann$
210: boundary conditions at that vertex. These graphs shall be referred to as Neumann graphs, 
211: and the resulting $\tilde \Sigma^{(i)}$ matrices read:
212: \begin {equation}
213: \label{Neumann}
214: \tilde \sigma^{(i)} _{j,j'}= {2 \over {\tilde v}}  -\delta _{j,j'};\quad
215: \tau^{(i)}_j =   {2 \over {\tilde v}};\quad
216: \rho^{(i)} ={2 \over {\tilde v}}-1\ .
217: \end{equation}
218: 
219: Vertices which are not coupled to leads have $\rho ^{i}=1 ,\ \tau^{(i)}_j =0$, while
220: the bond-bond transition matrix $\tilde \sigma^{(i)}_{j.j'}={2 \over v}-\delta _{j,j'}$ 
221: is unitary.
222: 
223: %-------------------------------------------------------------------------
224: 
225: \section{\bf The S-matrix for Quantum Graphs}
226: \label{sec:s-matrix}
227: 
228: It is  convenient to discuss first graphs with leads connected to all the vertices
229: $M=V$. The generalization to an arbitrarily number $M\leq V$ of leads (channels) is
230: straightforward and will be presented at the end of this section.
231: 
232: To derive the scattering matrix, we first write the  bond wave functions using the 
233: two  representations which are conjugated by   ``time reversal":
234: \begin{eqnarray}
235: \label{wfun2}
236: \Psi_{b}(x_b) &=& a_{b} {\rm e}^{i(k+A_{b})x_{b}}
237: +c_{ b} {\rm e}^{i(-k+A_{b})x_{b}}\  = \nonumber \\
238: \Psi _{\hat b}(x_{\hat b}) &=&a_{\hat b}{\rm e}^{i(k+A_{\hat b})x_{\hat b}}+
239: c_{\hat b}{\rm e}^  {i(-k+A_{\hat b})x_{\hat b}}\  = \nonumber  \\
240: &=&a_{\hat b}{\rm e}^  {i(-k-A_{\hat b})x_b}{\rm e}^  {i(k+A_{\hat b})L_b} +
241: c_{\hat b}{\rm e}^  {i(-k+A_{\hat b})L_b}{\rm e}^  {i(k-A_{\hat b})x_b} \ .
242: \end{eqnarray}
243: Hence,
244: \begin{equation}
245: c_b=a_{\hat b}{\rm e}^  {i(k+A_{\hat b})L_b},\,\,\,\,
246: a_b=c_{\hat b}{\rm e}^  {i (-k+A_{\hat b})L_b }.\label{wfun3}
247: \end{equation}
248: In other words, but for a phase factor, the outgoing wave from the vertex $i$ in the 
249: direction $j$ is identical to the incoming wave at $j$ coming from $i$.
250: 
251: Substituting $a_b$ from Eq.~(\ref{wfun3}) in Eq.~(\ref{vertexcond}), and solving for 
252: $c_{i,j}$ we get
253: \begin{eqnarray}
254: \label{wfun4}
255: c_{i,j'}&=&\sum_{r,s} \left({\bf 1}-\tilde S_B(k;A) \right
256: )^{-1}_{(i,r),(s,j)} D_{(s,j)}
257: \tau _s^{(j)} I_j \nonumber \\
258: O_i&=&\rho^{(i)} I_i + \sum_{j'}\tau_{j'}^{(i)} c_{ij'}
259: \end{eqnarray}
260: where ${\bf 1}$ is the $2B\times 2B$ unit matrix. Here, the ``bond scattering matrix'' 
261: ${\tilde S}_b$ is a sub-unitary matrix in the $2B$ dimensional space of directed bonds 
262: which propagates the wavefunctions. It is defined as $\tilde S_B(k,A)=D(k;A)\tilde R $, 
263: with
264: \begin{eqnarray}
265: \label{DandT}
266: D_{ij,i^{\prime }j^{\prime }}(k,A) &=&\delta _{i,i^{\prime }}\delta
267: _{j,j^{\prime }}{\rm e}^{ikL_{ij}+iA_{i,j}L_{ij}} \\
268: \ \tilde R_{ji,nm} &=&\delta _{n,i}C_{j,i}C_{i,m}{\tilde \sigma}
269: _{ji,im}^{(i)}.
270: \nonumber
271: \end{eqnarray}
272: $D(k,A)$ is a  diagonal unitary matrix which depends only on the metric properties of 
273: the graph, and provides a phase which is due to free propagation  on the bonds. The 
274: sub-unitary matrix $\tilde R$ depends on the connectivity and on the bond-bond transition 
275: matrices ${\tilde \sigma}$. It  assigns a scattering amplitude for transitions between 
276: connected directed bonds. $\tilde R$ is sub-unitary, since
277: \begin{equation}
278: \label{subR}
279: |\det \tilde R| = \prod_{i=1}^V |\det \tilde \sigma ^{(i)}|<1.
280: \end{equation}
281: 
282: Replacing $c_{i,j'}$ in the second of Eqs.~(\ref{wfun4}) we get the following relation 
283: between the outgoing and incoming amplitudes $O_i$ and $I_j$ on the leads:
284: \begin{equation}
285: \label{inout}
286: O_i = \rho^{(i)} I_i + \sum_{j'j r s}\tau_{j'}^{(i)}
287: \left({\bf 1}-\tilde S_B(k;A) \right )^{-1}_{(i,r),(s,j)} D_{(s,j)} \tau
288: _s^{(j)} I_j
289: \end{equation}
290: Combining (\ref{inout}) for all leads $i=1,\dots ,V$, we obtain the unitary $V\times V$ 
291: scattering matrix $S^{(V)}$,
292: \begin{equation}
293: \label{scatmat}
294: S^{(V)}_{i,j}  = \delta_{i,j} \rho^{(i)}
295:   + \sum_{r,s} \tau^{(i)}_r
296: \left ({\bf 1}-\tilde S_B(k;A) \right )^{-1}_{(i,r),(s,j)} D_{(s,j)}
297: \tau _s^{(j)}.
298: \end{equation}
299: From Eq.~(\ref{scatmat}), we see that the scattering matrix may be decomposed into 
300: two parts $S^{(V)}(k)=S^{D}+S^{fl}(k)$ which are associated with two well separated 
301: time scales of the scattering process. $S^{D}(=\delta_{i,j} \rho^{i})$ is the prompt 
302: reflection at the entrance vertex and induces a ``direct" component. In general, it
303: varies very slowly with energy, and is system dependent. On the other hand the
304: ``chaotic" component of the $S$ matrix, $S^{fl}(k)$, starts by a transmission from
305: the incoming lead $i$ to the bonds $(i,r)$ with transmission amplitudes 
306: $\tau^{(i)}_{r}$. The wave gains a phase ${\rm e}^{i(k+A_{b})L_{b}}$ for each bond it 
307: traverses and a scattering amplitude $\tilde \sigma^{(i)}_{r,s}$ at each vertex. This 
308: multiple scattering inside the interaction region becomes apparent when the expansion
309: \begin{equation}
310: \label{multi}
311: ({\bf 1}-\tilde S_B(k;A))^{-1}= \sum_{n=0}^{\infty} \tilde S_B ^n(k;A)
312: \,\,\,\,\,\, 
313: \end{equation}
314: is substituted in    (\ref{scatmat}). Eventually the wave is transmitted
315: from the bond $(s,j)$
316: to the lead $j$ with an  amplitude $\tau _s^{(j)}$. Explicitly,
317: \begin{equation}
318: S^{(V)}_{i,j}  = \delta_{i,j} \rho^{(i)} +  \sum_{t \in {\cal
319: T}_{i\rightarrow j}}
320: {\cal B}_{t} {\rm e}^{i (k l_t +  \Theta_t)}
321: \label {sexplicit}
322: \end{equation}
323: where ${\cal T}_{i\rightarrow j}$ is the set of the trajectories on $\tilde
324: {\cal G}$ which lead from $i$ to $j$.  ${\cal B}_{t}$ is the amplitude
325: corresponding to a path $t$  whose length and directed length are $l_t=
326: \sum_{b\in t}L_b$ and $\Theta_t=\sum_{b\in t} L_bA_b$  respectively. 
327: Thus the
328: scattering amplitude $S^{(V)}_{i,j}$ is  a sum of a large number of partial
329: amplitudes, whose complex  interference brings about the typical irregular
330: fluctuations of $|S^{(V)}_{i,j}|^2$ as a function of $k$.
331: 
332: One of the basic concepts in the quantum theory of scattering are the 
333: resonances.
334: They represent long-lived intermediate states to which bound states of a 
335: closed
336: system are converted due to coupling to continua. On a formal level, 
337: resonances
338: show up as poles of the scattering matrix $S^{(M)}$ occurring at complex 
339: wave-numbers
340: $\kappa_n = k_n - \frac i2 \Gamma_n$, where $k_n$ and $\Gamma_n$ are the 
341: position
342: and the width of the resonances, respectively. From (\ref{scatmat}) it 
343: follows that
344: the resonances are the complex zeros of
345: \begin{equation}
346: \label {resonancecond}
347: \zeta_{\tilde {\cal G}}(\kappa) = \det \left({\bf 1} -\tilde S_B(\kappa
348: ;A)\right)=0 \ .
349: \end{equation}
350: The eigenvalues of $\tilde S_B$ are in the unit circle, and therefore the
351: resonances appear in the lower half of the complex $\kappa$ plane. Moreover
352: from
353: Eq.~(\ref{resonancecond}) it is clear that their formation is closely
354: related to the
355: internal dynamics inside the scattering region which is governed by $\tilde
356: S_B$.
357: 
358: There exists an intimate link  between the scattering matrix and the
359: spectrum of the
360: corresponding closed graph. It manifests the exterior -interior duality
361: \cite{DS92a} for  graphs. The spectrum of the closed graph is the set of
362: wave-numbers for which
363: $S^{(V)}$ has $+1$ as an eigenvalue. This corresponds to a  solution where
364: no currents flow in
365: the leads so that the conservation of current is  satisfied on the internal
366: bonds.
367: $1$ is in the spectrum of
368: $S^{(V)}$ if
369: \begin{equation}
370: \label{phsca4}
371: \zeta_{\cal G}(k)=\det\left[{\bf 1} -S^{(V)}(k)\right] = 0 \quad .
372: \end{equation}
373:   Eq.~(\ref{phsca4}) can be transformed in an alternative  form
374: \begin{equation}
375: \label{phscaa} {\hspace {-15mm}}
376: \zeta_{\cal G}(k)= \det [ {\bf 1} -\rho ]
377: {\det[{\bf 1}-D(k)R] \over \det[{\bf 1}-D(k) {\tilde R}]} = 0 \quad ; \quad
378: R_{i,r;s,j} = \tilde R_{i,r;s,j} +\delta_{r,s}
379: {\tau^{(r)}_{i}\tau^{(r)}_{j}\over 1-\rho^{(r)} }.
380: \end{equation}
381: which is satisfied once
382: \begin{equation}
383: \label{phscaa1}
384: \det[{\bf 1}-D(k)R] = 0 .
385: \end{equation}
386: In contrast to ${\tilde R}$, $R$ is a unitary matrix in the space of
387: directed bonds, and
388: therefore the spectrum is real. (\ref {phscaa1}) is the secular equation
389: for the
390: spectrum of the compact part of the graph, and it was derived in a
391: different way in \cite
392: {KS97}.
393: 
394: The difference $\delta R = R - \tilde R$ gets smaller as larger graphs are considered 
395: (for graphs with Neumann  boundary conditions it is easy to see that the difference 
396: is of order $\frac{1}{v}$). That  is, the leads are weekly coupled to the compact part 
397: of the graph, and one can use  perturbation theory for the computation of the resonance 
398: parameters. To lowest order, $(\delta R =0)$, the resonances coincide with the spectrum 
399: of the compact graph. Let $k_n$ be in the spectrum.  Hence, there exists a vector 
400: $|n \rangle$ which satisfies the equation
401: \begin{equation}
402: D(k_n) R\  |n \rangle = 1\ |n \rangle \ .
403: \end{equation}
404: To first order in $\delta R$, the resonances acquire a width
405: \begin{equation}
406: \label{pertu3}
407: \delta \kappa_n = -i  {\langle n|D(k_n)\delta R|n\rangle \over \langle n|L|n\rangle} \,.
408: \end{equation}
409: To check the usefulness of this result, we searched numerically for the 
410: true poles
411: for a few scattering graphs and compared them with the approximation 
412: ~(\ref{pertu3}).
413: In figure~1 we show the comparison for fully connected Neumann graphs
414: with $V=5,15$ and $A=0$.  As expected the agreement between the exact 
415: poles and the
416: perturbative results improves as $v$ increases.
417: 
418: \begin{figure}
419: \begin{center}
420: \epsfxsize.5\textwidth%
421: \epsfbox{sfig1.eps}
422: \caption { Poles of the $S^{(V)}$-matrix for regular Neumann graphs. The exact 
423: evaluated poles are indicated with $(\circ )$ while $(\star )$ are the results 
424: of the perturbation theory (\ref{pertu3}): (a) $V=5$ and $v=4$ and (b) $V=15$ 
425: and $v=14$.}
426: \end{center}
427: \label{fig:fig2}
428: \end{figure}
429: 
430: We finally comment that the formalism above can be easily modified for graphs where
431: not all the vertices are attached to leads. If the vertex $l$ is not attached, one
432: has to set $\rho^{(l)}=1, \tau^{(l)}_j =0$  in the definition of $\Sigma^{(l)}$.
433: The dimension of the scattering matrix is then changed accordingly.
434: 
435: For the sake of completeness we quote here an alternative expression for the $S$ 
436: matrix which applies for Neumann graphs, exclusively  \cite{KS99}. For this purpose 
437: we define the $V\times V$ matrix
438: \begin{equation}
439: \label{secu1}
440: h_{i,j}(k,A) =\left\{
441: \begin{array}{cc}
442: -\sum_{m\neq i}C_{i,m}\cot(kL_{i,m})  , &
443: i=j \\ \\
444: C_{i,j}{\rm e}^  {-iA_{i,j}L_{i,j}}(\sin (kL_{i,j}))^{-1}, & i\neq j \
445: \end{array}
446: \right.   \nonumber
447: \end{equation}
448: in terms of which,
449: \begin{equation}
450: S^{(V)} = (i {\bf 1} + h(k))^{-1} (i {\bf 1}-h(k))
451: \label{phsca3}
452: \end{equation}
453: where ${\bf 1}$ is the $V\times V$ unit matrix. $S^{(V)}$ is unitary 
454: since $h(k)$ is
455: hermitian.
456: 
457:   In the case of graphs connected to leads at an arbitrary  set of $M<V$
458: vertices with indices
459: $\left \{i_l \right\}, \ \ l=1,\cdots,M$, the $M \times M$ scattering
460: matrix $S^{(M)}$
461: has to be modified in the following way
462: \begin{equation}
463: S^{(M)} = 2i W \left ( h(k) + i W^{T} W \right ) ^{-1} W^{T} -{\bf 1} \ .
464: \label{phsca5}
465: \end{equation}
466: Here $W_{i_l,j}=\delta_{i_l,j}$ is the $M\times V$ leads - vertices coupling
467: matrix, ( $W={\bf 1}$ when $M=V$). This form of the $S$ matrix is
468: reminiscent of the
469: expression  which was introduced by Weidenm\"uller to generalize the
470: Breit-Wigner theory
471: for  many channels and internal states. However, (\ref {phsca5}) is an
472: exact expression
473: which involves no truncations.
474: 
475:   It follows from (\ref {phsca5}) that in the present case, one can identify
476: the poles of the
477: $S$ matrix with the zeros of
478: \begin{equation}
479: \label{resonN}
480: \zeta_{\tilde {\cal G}}(\kappa) = \det \left(h(\kappa) + i W^{T} W
481: \right)=0 \ .
482: \end{equation}
483: Its main advantage over (\ref{resonancecond}) is that it involves a determinant of a 
484: matrix of much lower dimension.
485: 
486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487: \section{\bf The trace formula for resonances and  classical scattering on
488: graphs}
489: \label{sec:classical}
490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
491: 
492: In the spectral theory of bounded hamiltonian systems, the most fundamental object of 
493: study is the spectral density which consists of a sum of $\delta$ functions at the 
494: spectral points. For open systems, it is replaced by the resonance density, which
495: is defined on the real $k$ line and consists of an infinite sum of Lorentzians which
496: are centered at $k_n=\Re e \kappa_n$ and have width $\Gamma _n =-2\Im m \kappa_n$,
497: where $\kappa_n$ are the complex poles of the scattering matrix. In the present section 
498: we shall express the resonance density for scattering graphs in terms of periodic 
499: orbits of their compact part. This is the analogue of the trace formula for bounded 
500: graphs.
501: 
502: 
503: %-----------------------------------------------------------------------------------
504: \subsection{\bf The Trace Formula}
505: 
506: The resonance density $d_R(k)$ can be deduced from the total phase $\Phi(k) = \frac{1}
507: {i} \ln \det S^{(M)}(k) $ of the scattering matrix \cite {Krein}:
508: \begin{equation}
509: \label{resden}
510: d_R(k)\equiv{1\over 2\pi} {d\Phi\over dk} = -i{1\over 2\pi}{\partial\over
511: \partial k}\ln \det S^{(M)}(k).
512: \end{equation}
513: It is a smooth function for real $k$, and can be interpreted as the mean length of
514: the {\it delay} associated with the scattering at wave-number $k$ 
515: \cite{S89}.
516: 
517: Using Eq.~(\ref{scatmat}) and performing standard manipulations \cite{MW69}, we obtain 
518: the following expression for the phase $\Phi(k)$
519: \begin{equation}
520: \Phi(k)- \Phi(0) = -2 \Im m \ln \det (I- \tilde S_B(k;A)) + {\cal L}k \, .
521: \label{tra}
522: \end{equation}
523: where ${\cal L} =2\sum_{b=1}^B L_b $ is twice the total length of the 
524: bonds of
525: $\tilde {\cal G}$.
526: 
527: Using the expansion
528: \begin{equation}
529: \label{expan}
530: \ln \det (I- \tilde S_B(k;A)) = -\sum_{n=1}^{\infty} {1\over n} {\rm tr}
531: {\tilde
532: S_B}^n(k;A)
533: \end{equation}
534: we rewrite Eq.~(\ref{tra}) as
535: \begin{equation}
536: \label{tra1}
537: \Phi(k)= \Phi(0) + {\cal L}k +
538: 2 \Im m \sum_{n=1}^{\infty} {1\over n} {\rm tr} {\tilde S_B}^n (k;A).
539: \end{equation}
540: 
541: On the other hand, using Eq.~(\ref{DandT}) we can write ${\rm tr} {\tilde
542: S_B}^n(k;A)$
543: as sums over $n-$periodic orbits on the graph
544: \begin{equation}
545: \label{ppo}
546: {\rm tr} ({\tilde S_B}^n(k;A)) = \sum_{p\in {\cal P}_{n}} n_p {\tilde {\cal
547: A}_p}^r
548: {\rm e}^  {i(l_p k+\Theta_p)r} \quad ,
549: \end{equation}
550: where the sum is over the set ${\cal P}_n$ of primitive periodic orbits 
551: whose
552: period $n_p$ is a divisor of $n$, with $r=n/n_p$ (primitive periodic orbits
553: are those which
554: cannot be written as a repetition of a shorter periodic orbit). The 
555: amplitudes
556: ${\tilde {\cal A}}_p$ are the products of the bond-bond scattering
557: amplitudes $\tilde
558: \sigma^{(i)}_{b,b'}$ along the primitive loops i.e.
559: \begin{equation}
560: \label{ampli}
561: {\tilde {\cal A}_p} = \prod_{i=1}^{n_p} {\tilde \sigma^{(i)}_{b,b'}} \ .
562: \end{equation}
563: Substituting Eqs.~(\ref{tra1},\ref{ppo}) in Eq.~(\ref{resden}) one gets the resonance
564: density
565: \begin{equation}
566: \label{resdens}
567: d_R(k)  = {1\over 2 \pi}{\cal L}+{1\over \pi}\Re e\sum_{n=1}^{\infty}\sum_{p\in 
568: {\cal P}_{n}} n_p l_p r  {\tilde {\cal A}_p}^r {\rm e}^  {i(l_p k+\Theta_p)r}
569: \end{equation}
570: Eq.~(\ref{resdens}) is an {\it exact} trace formula for the resonance density. The
571: first term on the right hand side of Eq.~(\ref{resdens}) corresponds to the smooth
572: resonance density, while the second provides the fluctuating part. We notice that
573: the mean resonance spacing is given by
574: \begin{equation}
575: \label{mrd}
576: \Delta= {2\pi\over {\cal L}} \simeq {2 \pi\over 2 B \langle L\rangle}
577: \end{equation}
578: and it is the same as the mean level spacing obtained for the corresponding bounded 
579: graph ${\mathcal G }$ \cite{KS97}.
580: 
581: %------------------------------
582: \subsection{\bf Classical Dynamics}
583: 
584: We conclude this section with the discussion of the classical dynamics on the graph
585: ${\tilde {\cal G}}$. A classical particle moves freely as long as it is on a bond.
586: The vertices are singular points, and it is not possible to write down the analogue
587: of the Newton equations there.  In \cite {KS97,KS99} it was shown that it is possible
588: to define a classical evolution on the graph: A {\it Poincar\'{e} section} on the
589: graph consists of the discrete set of directed bonds. The phase-space density at a
590: (topological) time $n$ is the set of occupation probabilities $\rho_b(n)$ of the
591: directed bonds, and the classical evolution is governed by a Markovian master equation.
592: Applied to the compact part of a scattering graph it reads,
593: \begin{equation}
594: \label{master}
595: \tilde \rho _b(n+1)=\sum_{b^{\prime }} \tilde  U_{b,b^{\prime }}
596: \tilde \rho _{b^{\prime }}(n)
597: \end{equation}
598: where the transition matrix $ \tilde U_{b,b^{\prime }}$ is given by the 
599: corresponding
600: {\it quantum} transition probability
601: \begin{equation}
602: \tilde U_{ij,nm}= \left|\tilde R_{ij,nm}\right|^2=\delta _{j,n} |
603: \sigma^{(j)}_{ij,jm}|^2.
604:   \label{cl3}
605: \end{equation}
606:   Notice that $\tilde U$ does not involve any metric information on the 
607: graph.
608: 
609:   Due to loss of flux to the leads $\sum_{b'}{\tilde U}_{bb'}<1$, and the
610: phase-space measure is not preserved, but rather, decays in time. The
611: probability to remain on $\tilde {\cal G}$ is
612: \begin{equation}
613: {\tilde P}(n) \equiv \sum_{b=1}^{2B}{\rho}_b(n) = \sum_{b,b'}{\tilde
614: U}_{bb'}{\rho}_{b'}(n-1)
615: \simeq  {\rm e}^  {-\Gamma_{cl}n}{\tilde P(0)}
616: \end{equation}
617: where $\exp (-\Gamma_{cl})$ is the largest eigenvalue of the ``leaky" 
618: evolution
619: operator ${\tilde U}_{bb'}$.
620: 
621: For the $v$-regular graph with $\tilde \sigma^{(i)}$ given by (\ref 
622: {Neumann})
623: the spectrum of $\tilde U$ is restricted to the interior of a circle with
624: radius given by the maximum eigenvalue $\nu_1=(v-1)\tau^2+\rho^2$ with
625: corresponding eigenvector $|1>=(1/2B)(1,1,\cdots,1)^T$. Hence the decay rate
626: $\Gamma_{cl}=-\ln \nu_1$ for regular Neumann graphs take the simple form
627: \begin{equation}
628: \label{Ndrate}
629: \Gamma_{cl}=-\ln \left(1-\tau^2\right) \approx (2/(1+v))^2.
630: \end{equation}
631: We notice that removing the leads from the vertices and turning ${\tilde
632: {\cal G}}$ into a compact graph ${\cal G}$ we get $\Gamma_{cl}= 0$ since in
633: this case $\tau^{(i)}=0$ (and $\rho^{(i)}=1$) and the phase-space measure
634: is preserved as expected.
635: 
636: The inverse decay rate $T_{cl} = \Gamma_{cl}^{-1}$, gives the average classical delay 
637: time that the particle spends within the interaction region. Injecting a particle from 
638: the leads to the scattering domain, its probability to be on any bond randomizes, 
639: because at each vertex a Markovian choice of one out of $v$ directions is made. The 
640: longer a particle remains within the interaction regime, the more scattering events it 
641: experiences. The set of trapped trajectories whose occupancy decays exponentially in 
642: time is the analogue of the strange repeller in generic Hamiltonian systems displaying 
643: ``chaotic scattering".
644: 
645: 
646: %-------------------------------------------------------------------------
647: 
648: \section{\bf Statistical Analysis of the $S-$matrix}
649: \label{sec:statistics}
650: 
651: %---------------------------------------------------------------------------
652: 
653: So far we developed the scattering theory of graphs, pointing out their similarity
654: with scattering systems which display chaotic scattering in the classical limit. 
655: Due to the interference of a large number of amplitudes, the $S-$matrix fluctuates 
656: as a function of $k$, and its further analysis calls for a statistical approach 
657: which will be the subject of the present section. We shall show that quantum graphs 
658: possess typical poles, delay time and conductance distributions, Ericson fluctuations 
659: of the scattering amplitudes, and when considered statistically, the ensemble of 
660: scattering matrices are very well reproduced by the predictions of RMT. At the same 
661: time deviations from the universal RMT results, which are related to the system-
662: specific properties of some graphs, will be  pointed out. The study  of these 
663: deviations is especially convenient for graphs because of the transparent and simple 
664: scattering theory developed in terms of scattering trajectories.
665: 
666: An important parameter which is associated with the statistical properties of the 
667: $S$-matrix is the Ericson parameter defined through the scaled mean resonance width 
668: as:
669: \begin{equation}
670: \label {eripar}
671: \left<\gamma\right>_k \equiv  {\left<\Gamma_n\right>_k  \over \Delta}
672: \end{equation}
673: where $\langle \cdot \rangle _k$ denotes spectral averaging and $\Delta$ is the
674: mean spacing between resonances. The Ericson parameter determines whether the
675: resonances overlap $(\langle \gamma\rangle_k >1)$ or are isolated $(\langle
676: \gamma\rangle_k <1)$. Typical example for the two extreme situations are shown
677: in figure~2.
678: 
679: \begin{figure}
680: \begin{center}
681: \epsfxsize.5\textwidth%
682: \epsfbox{sfig2.eps}
683: \end{center}
684: \caption{ Representative examples of scattering cross sections 
685: $|S_{i,j}^{(V)}|^2$
686: for Neumann graphs in a regime of  (a) isolated resonances ($V=15$, 
687: $v=14$ with
688: corresponding $\left<\gamma\right>_k=0.59$) and (b) overlapping 
689: resonances ($V=49$,
690: $v=14$ with corresponding $\left<\gamma\right>_k\simeq 2$). The real 
691: parts of the
692: resonances in this energy interval are indicated by the vertical lines.}
693: \end{figure}
694: 
695: The degree of resonance overlap determines the statistical properties of the
696: $S$-matrix.  We shall show in the sequel that the mean width  can be
697: approximated  by the classical decay rate $\langle \gamma\rangle_k=
698: \gamma_{cl}$. For the $v$
699: regular graphs discussed above, we have
700: \begin{equation}
701: \label{clmr}
702: \gamma_{cl}\equiv {\Gamma_{cl}\over \Delta} \approx {4\over 2\pi} 
703: {v\over 1+v}
704: {V\over 1+v}.
705: \end{equation}
706: where we made use of Eqs.~(\ref{mrd},\ref{Ndrate}).
707: Thus changing $v$ and $V$ we can control the degree of overlap allowing to
708: test various phenomena.
709: 
710: In what follows, unless explicitly specified, we shall consider regular
711: graphs with
712: one lead attached to each vertex, i.e. $M=V$. Finally, the widths are
713: always scaled
714: by the mean spacing $\Delta$ i.e. $\gamma_n\equiv {\Gamma_n \over \Delta}$.
715: 
716: %---------------------------------------------------------------------------
717: \subsection{\bf The resonance width distribution }
718: 
719: The resonance width distribution can be computed for a given graph in the following 
720: way. Consider the complex $\kappa=x+iy$ plane, where the zeros of the secular function
721: \begin{equation}
722: \label {resonancecond1}
723: \zeta_{\tilde {\cal G}}(\kappa) = \det \left({\bf 1} -\tilde S_B(\kappa ;A)\right) = f(x,y) +i
724: g(x,y)
725: \ .
726: \end{equation}
727: are the poles of the $S$ matrix (resonances). The variables $(x,y)$
728: are expressed in units of $\Delta$. On average, the number of resonances
729: with real part, $x \in
730: [0,X]$ is $X$. The density of resonance widths reads
731: \begin{equation}
732: \hspace{-10mm}
733: {\mathcal P}(\gamma) =\lim_{ X \rightarrow \infty} \frac{1}{ X } \int_0^X
734: \delta\left (f(x,y= - \gamma)\right ) \delta\left (g(x,y= -  \gamma)\right )
735: \left |\frac{{\rm d}\zeta_{\tilde {\cal G}}(\kappa)}{ {\rm d} \kappa}\right |^2_{y= -
736: \gamma}   {\rm d}x
737: \ .
738: \label{eq:disty}
739: \end{equation}
740: We used the Cauchy-Riemann theorem for the evaluation of the Jacobian $(f_xg_y - f_y 
741: g_x)$ which multiplies the two $\delta$ functions that locate the complex zeros of 
742: $\zeta_{\tilde {\cal G}}$. We now recall that the $\kappa$ dependence of $\tilde S_B
743: (\kappa ;A)$ comes from the factors ${\rm e}^{i\kappa L_b}$ in (\ref {DandT}). This
744: implies that for a given $y$, $\zeta_{\tilde {\cal G}}$ is a quasi periodic function 
745: of $x$. Moreover, expanding the determinant (\ref {resonancecond1}) it is not difficult 
746: to show that the frequencies involved can be written as linear combinations of the bond 
747: lengths $\lambda =\sum  q_b L_b $ with integer coefficients $q_b =0, 1, {\rm or}\ 2$. 
748: Since the bond lengths are rationally independent, we find that $\zeta_{\tilde {\cal G}}$
749: depends on a {\it finite}  number of incommensurate frequencies. The expression (\ref
750: {eq:disty}) can be regarded as a ``time"  integral over a trajectory on a multidimensional
751: incommensurate torus, which covers the torus ergodically. Hence, the integral can be
752: replaced by a phase space average,
753: 
754: \begin{equation}
755: \hspace{-10mm}
756: {\mathcal P}(\gamma) =
757: \int_0^{2\pi}\frac {{\rm d} \psi_1}{2\pi} \cdots \int_0^{2\pi}  \frac 
758: {{\rm d}
759: \psi_J}{2\pi} \
760: \delta\left (f({\vec \psi} , - \gamma)\right )\ \delta\left (g({\vec \psi} ,
761: -\gamma)\right ) \
762: \left |\frac{{\rm d}\zeta_{\tilde {\cal G}}({\vec \psi,-\gamma})}{ {\rm d}
763: \kappa}\right |^2_ .
764: \label{eq:distyergod}
765: \end{equation}
766: Here ${\vec \psi}$ denotes  the vector of independent angles on the $J$ dimensional
767: torus. Although the above formula provides a general framework, its application to
768: actual graphs is a formidable task.
769: 
770: An important feature of the distribution of the resonances in the complex plane can
771: be deduced by studying the secular function $\zeta_{\tilde {\cal G}}(\kappa)$. Consider
772: $\zeta_{\tilde {\cal G}}(\kappa=0)$. If one of the eigenvalues of the matrix $\tilde R$
773: (\ref {DandT}) takes the value $1$, $\zeta_{\tilde {\cal G}}(k=0) =0$ and because of
774: the quasi-periodicity of $\zeta_{\tilde {\cal G}}$, its zeros reach any vicinity of
775: the real axis infinitely many times. The largest eigenvalue of the $\tilde R$ matrix
776: for $v$-regular Neumann graphs is $1$, and therefore the distribution of resonance 
777: widths is  finite in the vicinity of $\gamma=0$. For generic graphs, the spectrum of 
778: $\tilde R$ is inside  a circle of radius $\lambda_{max} <1$. This implies that the 
779: poles are excluded from a strip just under the real axis, whose width can be estimated 
780: by
781: \begin{equation}
782: \label{gapKS}
783: \Gamma_{gap}=-2 \ln( |\lambda_{max}|)/L_{max}.
784: \end{equation}
785: where $L_{max}$ is the  maximum bond length. The existence of a gap is 
786: an important
787: feature of the resonance width distribution ${\cal P}(\gamma)$ for 
788: chaotic scattering
789: systems.
790: 
791: A similar argument was used recently in \cite{BG01} in order to obtain an {\it upper}
792: bound for the resonance widths.  It is
793: \begin{equation}
794: \label{upl}
795: \Gamma_{max}\sim -2 \ln( |\lambda_{min}|)/L_{min}\ ,
796: \end{equation}
797: where $\lambda_{min}$ and $L_{min}$ are the minimum eigenvalue and bond length,
798: respectively.
799: 
800: The distribution of the complex poles  for a generic fully connected graph with $V=5$
801: is shown in figure~3a. The vertical line which marks the region from which resonances
802: are excluded was computed using (\ref {gapKS}).
803: \begin{figure}
804: \begin{center}
805: \epsfxsize.5\textwidth%
806: \epsfbox{sfig3.eps}
807: \caption{ (a) The 5000 resonances of a single realization of a complete $V=5$ graph 
808: with $A\ne 0$ and $M=V$. The solid line marks the position of the gap $\gamma_{\it gap}$. 
809: (b) The distribution of resonance widths ${\cal P}(\gamma)$. The solid line is the RMT 
810: prediction (\ref{FSpole}). The difference ${\cal P}(\gamma)-{\cal P}_{CUE} (\gamma)$ 
811: is shown in the inset.}
812: \end{center}
813: \end{figure}
814: 
815: Random matrix theory can provide an expression for the distribution of resonances. In the
816: case of non-overlapping resonances,  perturbation theory  shows that the resonance widths 
817: are distributed according to the so-called $\chi^2$ distribution
818: \begin{equation}
819: \label{chi2}
820: {\cal P}(\gamma) = \frac {(\beta M/2)^{\beta M/2}}{\langle \gamma\rangle_k{\it \Gamma}
821: (\beta M/2)} \left({\gamma\over\langle \gamma\rangle_k}\right)^{\beta M/2 -1} 
822: \exp(-\gamma \beta M/2\langle \gamma\rangle_k ),
823: \end{equation}
824: where $\beta=1(2)$ for systems which respect (break) time-reversal symmetry, and 
825: ${\it \Gamma}(x)$ is the gamma-function. Once $\langle \gamma\rangle_k$ becomes large 
826: enough the resonances start to  overlap, and (\ref{chi2}) does not hold. In the general 
827: case, Fyodorov and Sommers \cite{FS96,FS97} proved that the distribution of scaled 
828: resonance widths for the unitary random matrix ensemble, is given by
829: \begin{equation}
830: \label{FSpole}
831: {\cal P}(\gamma) = \frac {(-1)^M}{\Gamma(M)} \gamma^{M-1} {d^M\over 
832: d\gamma^M}
833: \left({\rm e}^  {-\gamma \pi g} {\sinh(\gamma\pi)\over(\gamma\pi)}\right)
834: \end{equation}
835: where the parameter $g={2\over (1-\langle S^D\rangle_k^2)} -1$ controls the degree of
836: coupling with the channels (and is assumed that $g_i=g \,\,\forall i=1,...M$). For
837: $g\gg 1$ (i.e. weak coupling regime) Eq.~(\ref{FSpole}) reduces to
838: (\ref{chi2}).
839: 
840: In the limit of $M\gg 1 $, Eq.~(\ref{FSpole}) reduces to the following expression
841: \cite{FS97}
842: \begin{equation}
843: {\cal P}(\gamma)=
844: \left\{
845: \begin{array}{lll}
846: {M\over 2\pi \gamma^2}&\ {\rm for}\ &{M\over \pi(g+1)}<
847: \gamma<{M\over\pi(g-1)} \\\nonumber
848: \ 0&\ \ \ \ & \ \ \ \ \ \  {\rm otherwise} \nonumber
849: \end{array}
850: \right.  \ .
851: \label{largeM}
852: \end{equation}
853: It shows that in the limit of large number of channels there exist a  strip in the 
854: complex $\kappa-$ plane which is free of resonances. This is in agreement with 
855: previous findings \cite{GR89,M75,SZ88}. In the case of maximal coupling i.e. $g=1$, 
856: the power law (\ref{largeM}) extends to  infinity, leading to divergencies of the 
857: various moments of $\gamma$'s. Using  (\ref{FSpole}) we recover the well known 
858: Moldauer-Simonius relation \cite{M75} for the mean resonance width \cite{FS97}
859: \begin{equation}
860: \label{SMrw}
861: \langle \gamma\rangle _k = -\frac {\sum_{i=1}^{V} \ln (|\langle S^{D}\rangle _k|^2)}{2\pi}.
862: \end{equation}
863: 
864: The resonance width distribution for a $V=5$ regular and generic graph is shown in
865: figure~3b together with the RMT prediction, which reproduces the numerical distribution
866: quite well. Figure~4 shows a similar comparison for a Neumann graph. The relatively
867: high abundance of resonances in the vicinity of the real axis conforms with the
868: expectations.
869: 
870: \begin{figure}
871: \begin{center}
872: \epsfxsize.5\textwidth%
873: \epsfbox{sfig4.eps}
874: \end{center}
875: \label{figure:n-resdist}
876: \caption {  Resonance width distribution ${\cal P}(\gamma)$ for a complete
877: Neumann graph with $M=V=5$ and $A\ne 0$. The solid line is the RMT prediction
878: (\ref{FSpole}). }
879: \end{figure}
880: 
881: 
882: %--------------------------------------------------------------------
883: 
884: \subsection{\bf The Form Factor}
885: 
886: To investigate further the dynamical origin of the resonance fluctuations, we study
887: the resonance two-point form factor $K(t)$. The main advantage of $K(t)$ is that it
888: allows us to study the resonance fluctuations in terms of classical orbits. It is
889: defined as
890: \begin{equation}
891: \label{rescor}
892: K_R(t)\equiv  \int {\rm d}\chi \ {\rm e}^{i 2\pi \chi {\cal L} t}
893: R_2(\chi).
894: \end{equation}
895: where $t$ measures lengths in units of the Heisenberg length $l_H={\cal L}$
896: and $
897: R_2(\chi)$ is the excess probability density of finding two resonances at a
898: distance
899: $\chi$,
900: \begin{equation}
901: \label{corr1}
902: \hspace {-15mm}
903: R_2(\chi;k_0) =\langle \tilde d_R  (k+ {\chi \over 2})\ \tilde d_R(k-
904: {\chi\over 2})\rangle_k = \langle {\Delta\over 2k_0}\int_{-k_0} ^{k_0}
905: {\tilde d}_R  (k+ {\chi \over 2})\ {\tilde d}_R(k- {\chi\over 2})
906: dk\rangle_{k_0}
907: \ .
908: \end{equation}
909: Above $\langle \ldots \rangle _{k_0}$ indicate averaging oven a number of spectral 
910: intervals of size $\Delta_k=2k_0$, centered around $k_0$. Here $\tilde d_R(k)$ is 
911: the oscillatory part of $d_R(k)$ (see Eq.~(\ref{resdens})). Substituting the latter 
912: in Eqs.~(\ref {rescor},\ref{corr1}) we obtain $K_R(t)$ in terms of periodic orbits 
913: and their repetitions
914: \begin{equation}
915: \label{ff1}
916: K_R(t)= {2{\cal N}\over{\cal L}^2}\left|\sum_p\sum_r l_p {\tilde {\cal 
917: A}}_p^r
918: {\rm e}^{i (kl_a+ \Theta_a)} \delta_{\cal N} (t-rl_p/{\cal L})\right|^2 .
919: \end{equation}
920: where $\delta_{\cal N}(x) = (\sin({\cal N}x/2))/({\cal N}x/2)$ and ${\cal N}={\Delta_k
921: \over\Delta}$. A similar sum contributes to the spectral form factor of the compact
922: graph \cite{KS97,KS99}. However, the corresponding amplitudes are different
923: due to the fact that ${\tilde {\cal A}}_p$ contains also the information about the
924: escape of flux to the leads.
925: 
926: \begin{figure}
927: \begin{center}
928: \epsfxsize.5\textwidth%
929: \epsfbox{sfig5.eps}
930: \caption{The form factors $K_R(t)$ for a complete $V=5$ graph, with either generic or
931: Neumann  boundary conditions, $A\neq 0$ and $M=V$. The data were averaged over $5000$ 
932: spectral intervals and smoothed on small $t$ intervals. Upper inset: $K_R(t)$ for small 
933: times.  Solid line correspond to the numerical data for the pentagon with generic boundary 
934: conditions while dashed line is the approximant $K_R(t)\approx K(t) exp(-\gamma_{cl} 
935: t)$. Lower inset: $K_R(t)$ calculated with high resolution. The lengths of  periodic 
936: orbits of the close graph are indicated by arrows.}
937: \end{center}
938: \end{figure}
939: 
940: Assuming that all periodic orbits decay at the same rate, one would 
941: substitute
942: ${\tilde {\cal A}}_p$ with ${\cal A}_p \exp(-n_p \gamma_{cl}/2)$ where 
943: ${\cal
944: A}_p$ is the weight assign to the periodic orbit of the corresponding 
945: close system.
946: Then Eq.~(\ref{ff1}) takes the following simple form
947: \begin{equation}
948: \label{corr6}
949: K_R(t) \approx K(t) {\rm e}^  {-\gamma_{cl}t}\,\,\,,
950: \end{equation}
951: where $K(t)$ is the form factor of the compact system \cite{EVP99}. 
952: Notice that
953: for $t\ll \gamma_{cl}^{-1}$ the resonance form factor $K_R(t)$ is equal 
954: to $K(t)$.
955: This is  reasonable since an open system cannot be distinguished from a 
956: closed
957: one during short times. This simple approximation is checked in the 
958: inset of
959: figure~5 (see dashed line) and it is shown to reproduce the numerical 
960: data rather
961: well in the domain $t\le 5$. The asymptotic decay is dominated by the 
962: resonances
963: which are nearest to the gap, and it cannot be captured by the crude 
964: argument
965: presented above. For generic graph, $K_R$ decays exponentially but with 
966: a rate
967: given by $\gamma_{as} =\gamma_{gap}$ (the best fit, indicated in 
968: figure~5 by the
969: dashed line, give $\gamma_{as}$ which agrees with  $\gamma_{gap}$ within 
970: $30\%$).
971: For the graph with Neumann boundary conditions, $\gamma_{gap}=0$ and one 
972: expects
973: an asymptotic power-law decay. The corresponding best fit (see dashed 
974: line in
975: figure~5) shows $t^{-2}$.
976: 
977: In Hamiltonian systems in more than one dimension, the size of the 
978: spectral interval
979: $\Delta_k$ where the spectral average is performed is limited by the 
980: requirement
981: that the smooth part of the spectral density is approximately constant. 
982: Here instead
983: we can take arbitrarily large spectral intervals since the smooth 
984: spectral density
985: is constant \cite{KS97,KS99}. This way, one can reach the domain where 
986: the function
987: $K_R(t)$ is composed of arbitrarily sharp spikes which resolve 
988: completely the length
989: spectrum for lengths which are both smaller and larger than $l_H$. In 
990: the inset of
991: figure~5 we show the numerical $K(t)$ calculated with high enough 
992: resolution. In the
993: same inset we mark with arrows the location of the lengths of short 
994: periodic orbits.
995: We notice that as long as $t{\cal L}$ is shorter than the length of the 
996: shortest
997: periodic orbit, $K(t)=0$. With increasing $t$, the periodic orbits 
998: become exponentially
999: dense and therefore the peaks start to overlap, giving rise to a quasi 
1000: continuum
1001: described approximately by Eq.~(\ref{corr6}).
1002: 
1003: 
1004: %----------------------------------------------------------------------------
1005: \subsection{\bf Ericson fluctuations}
1006: 
1007: As was mentioned above, the scattering cross sections are dominated by 
1008: either isolated
1009: resonances, or by overlapping resonances whose fluctuations follow a 
1010: typical pattern.
1011: These patterns were first discussed by Ericson \cite {E60} in the frame 
1012: of nuclear
1013: physics and were shown to be one of the main attributes of chaotic 
1014: scattering \cite
1015: {S89}. The transition between the two regimes is controlled by the 
1016: Ericson parameter
1017: (\ref {eripar}). Typical fluctuating cross sections are shown in figure~2.
1018: 
1019: \begin{figure}
1020: \begin{center}
1021: \epsfxsize.5\textwidth%
1022: \epsfbox{sfig6.eps}
1023: \end{center}
1024: \label{figu:autocorr}
1025: \caption{ The autocorrelation function $C(\chi,\nu=1)$ for regular 
1026: graphs with
1027: Neumann boundary conditions. ($\circ$) correspond to a graph with
1028: $\gamma=0.59$
1029: (isolated resonance regime), ($\star$) to a graph with $\gamma=1.36$
1030: (intermediate
1031: regime) while ($\diamond$) to a graph with $\gamma \simeq 2$ (overlapping
1032: resonances
1033: regime). The solid lines correspond to the theoretical expression
1034: (\ref{ericscl}):
1035: (a) The real part of $C(\chi,\nu=1)$ ; (b) The imaginary part of
1036: $C(\chi,\nu=1)$.}
1037: \end{figure}
1038: 
1039: 
1040: A convenient measure for Ericson fluctuations is the autocorrelation 
1041: function
1042: \begin{equation}
1043: \label{ericcor}
1044: C(\chi; \nu) =  {1\over\Delta j} \sum _{ j=j_{min}}^{j_{max}}\left 
1045: \langle  \
1046: S^{(M)}_{j,j+\nu}(k+{\chi\over 2}) \ S^{(M)\star}_{j,j+\nu}
1047: (k-{\chi \over 2}) \ \right  \rangle_{k}
1048: \end{equation}
1049: where $\Delta j = j_{max}-j_{min}+1$. To evaluate Eq.~(\ref{ericcor}) we
1050: substitute
1051: the expression of the $S$-matrix from Eq.~(\ref{sexplicit}) and  split the
1052: sum over
1053: trajectories into two distinct parts: the contributions of short
1054: trajectories are
1055: computed explicitly by following the multiple scattering expansion up to
1056: trajectories
1057: of length $l_{max}$. The contribution of longer orbits are approximated by
1058: using the
1059: diagonal approximation, which results in a Lorentzian with a width
1060: $\gamma_{Er}$.
1061: Including explicitly up to $n=3$ scattering events we get,
1062: \begin{eqnarray}
1063: \label{ericscl}
1064: C(\chi;\nu ) &\approx & G {\rm e}^  {il_{max}\chi} \frac {
1065: \gamma_{Er}}{\gamma_{Er}-i\chi}
1066: +\frac1{\Delta j}\sum_{j=j_{min}}^{j_{max}} [ \tau^4 {\rm e}^  {i\chi
1067: L_{j,j+\nu}}
1068: +\tau^4 \rho^4 {\rm e}^  {3i\chi L_{j,j+\nu} } \nonumber \\
1069: &+& \tau^6 \sum_{m\neq j,j+\nu} {\rm e}^  {i\chi(L_{j,m}+L_{m,j+\nu})} ]
1070: \end{eqnarray}
1071: where $G$ is determined by the condition $C(\chi=0;\nu)=1$. The 
1072: interplay between the
1073: contributions of long and short periodic orbits is shown in figure~6. 
1074: For overlapping
1075: resonances, the autocorrelation function is well reproduced by a 
1076: Lorentzian while for
1077: the case of isolated resonances one can clearly see the contributions of 
1078: short paths.
1079: 
1080: \begin{figure}
1081: \begin{center}
1082: \epsfxsize.5\textwidth%
1083: \epsfbox{sfig7.eps}
1084: \caption{ The mean resonance width $\left<\gamma\right>_k$, autocorrelation
1085: width $\gamma_{\it Er}$, the classical expectation $\gamma_{cl}$, and 
1086: the RMT
1087: prediction (\ref{SMrw}) vs. $V$ for various graphs with Neumann boundary
1088: conditions and constant valency $v=14$.}
1089: \end{center}
1090: \end{figure}
1091: 
1092: In figure~7 we report the mean resonance width $\langle \gamma\rangle_k$ calculated
1093: numerically for  various graphs, the parameter $\gamma_{\rm Er}$ extracted from the
1094: best fit of the  numerical $C(\chi)$ with Eq.~(\ref{ericscl}), together with the RMT
1095: prediction Eq.~(\ref{SMrw}), and the classical expectation given by Eq.~(\ref{clmr}).
1096: The results justify the use  of the classical estimate especially in the limit $V
1097: \rightarrow \infty$ for fixed $v/V$ (which is the analogue of the semiclassical limit).
1098: In this limit, the RMT and the classical estimate coincide.
1099: 
1100: %----------------------------------------------------------------------------
1101: 
1102: \subsection{\bf $S-$matrix statistics}
1103: 
1104: One can check further the applicability of RMT by studying the entire $S^{(M)}-$matrix 
1105: distribution function. The probability density of $M\times M$ unitary $S^{(M)}$-matrices 
1106: is defined in a $M+\beta M(M-1)/2$ parameter space and was found first in \cite{MPS85} 
1107: to be given by the Poisson Kernel
1108: \begin{equation}
1109: \label{sRMT}
1110: d{\cal P}_{\bar S}(S^{(M)})=p_{\bar S}(S)
1111: d\mu(S)=C_{\beta}\frac{[\det(1-{\bar S}
1112: {\bar S^{\dagger}})]^{(\beta M +2-\beta)/2}}{|\det(1-{\bar S}{\bar
1113: S^{\dagger}})|^{(\beta M +2-\beta)}} d\mu_{\beta}(S)
1114: \end{equation}
1115: where $C_{\beta}$ is a normalization constant which depend on the symmetry class. All
1116: system-specific relevant informations are included in the ensemble average $S^{(M)}-$
1117: matrix, defined as ${\bar S}_{i,j}= \langle S^D\rangle _k$. For regular graphs, ${\bar 
1118: S}$ is proportional to the unit matrix i.e. ${\bar S}_{i,j}= \frac{1-v}{1+v}\delta_{i,j}$,
1119: while the eigenvectors are distributed uniformly and independent from the eigenphases.
1120: The invariant measure $d\mu_{\beta}(S)$ is given as
1121: \begin{equation}
1122: \label{sRMT1}
1123: d\mu_{\beta}(S^{(M)}) = \prod_{i<j} |{\rm e}^  {i\phi_i}-{\rm e}^ 
1124: {i\phi_j}|^{\beta}
1125: \prod_{i=1}^Md\phi_i d\Omega
1126: \end{equation}
1127: where $d\Omega$ is the solid angle on the $M$ dimensional unit hyper-sphere.
1128: 
1129: \begin{figure}
1130: \begin{center}
1131: \epsfxsize.5\textwidth%
1132: \epsfbox{sfig8.eps}
1133: \caption{ The ${\cal P}(\psi)$ and ${\cal P}(\omega)$ distributions for a $2
1134: \times 2$ scattering matrix $S$. The solid lines correspond to the
1135: predictions of
1136: RMT (\ref{s2}). The upper panel correspond to $A=0$ while the lower panel to
1137: $A\neq 0$.}
1138: \end{center}
1139: \end{figure}
1140: For large $M$ values, a numerical check of this probability distribution is
1141: prohibitive. However, for $M=2$ one can find easily the exact form of
1142: (\ref{sRMT})
1143: and compare with the numerical results. The resulting distribution of the
1144: eigenphases $\phi_1,\phi_2$ of the $2\times 2$ scattering matrices of 
1145: regular
1146: graphs is given by
1147: %\widetext
1148: \begin{eqnarray}
1149: \label{s2}
1150: \hspace{-20mm}
1151: & &d{\cal P}_{\bar S} (S^{(2)}) = \\
1152: \hspace{-20mm}
1153: &=& \frac {C_{\beta} \left(4v\over (v+1)^2\right)^{\beta+2}
1154: \sin^{\beta}\left(\omega\over 2\right)}{\left[1-4{\bar 
1155: S}\cos(\frac{\omega}2)
1156: \cos\psi + 2{\bar S}^2 \left( 2\cos^2(\frac{\omega} 2)+\cos(2\psi)\right)-
1157: 4{\bar S}^3 \cos\psi \cos(\frac{\omega} 2)+{\bar 
1158: S}^4\right]^{\frac{\beta+2}2}}
1159: d\omega d\psi \nonumber
1160: \end{eqnarray}
1161: where we used the notation $\psi = \frac {\phi_1 +\phi_2}2$ and $\omega 
1162: = \phi_1
1163: - \phi_2$. Integrating Eq.~(\ref{s2}) over $\omega$, $\psi$ we get the 
1164: corresponding
1165: distribution functions ${\cal P}_{\bar S}(\omega)$ and ${\cal P}_{\bar 
1166: S}(\psi)$.
1167: Our numerical results for an ensample of $S$-matrices calculated for 
1168: different
1169: realizations of the lengths of the graphs are reported in figure~8 
1170: together with
1171: the RMT  predictions (\ref{s2}) for a regular graph with two channels. 
1172: An overall
1173: good agreement is seen both for $A=0$ and $A\neq 0$.
1174: 
1175: %-------------------------------------------------------------------------------
1176: \subsection{\bf Partial delay times statistics}
1177: 
1178: The Wigner delay time  captures the time-dependent aspects of quantum 
1179: scattering. It
1180: can be interpreted  as the typical time an almost monochromatic wave 
1181: packet remains
1182: in the interaction  region. It is defined as
1183: \begin{equation}
1184: \label{wignersmith1}
1185: T_W \ =\ {1\over 2 i M}{\rm tr} \left [  S^{(M)\,\,\dagger}
1186: \frac{dS^{(M)}(k)}{dk} \right ]\ =
1187: \ {1\over 2 M}
1188: \sum_{i=1}^M \frac{\partial\phi_i(k)}{\partial k}\ ,
1189: \end{equation}
1190: where $\phi_i$ are the eigen-phases of the $S^{(M)}$-matrix.
1191: The partial derivatives ${\partial \phi_i(k)\over \partial k }$ are the
1192: {\it partial delay times} and their statistical properties were studied
1193: extensively within the RMT \cite{FS96b,FS97,GM98}. For the one-channel 
1194: case it
1195: was found \cite{FS97,GM98} that the probability distribution of the scaled
1196: (with the mean level spacing $\Delta$) partial delay times $T_i={\Delta 
1197: \over
1198: 2\pi}{\partial \phi_i(k)\over \partial k }$ is
1199: \begin{equation}
1200: \label{wignersmith3}
1201: {\cal P}_{\bar S}^{(\beta)}(T) =\frac{({\beta\over 2})^{\beta/2}}
1202: {{\it \Gamma}({\beta\over 2})T^{2+\beta/2}}
1203: \int_0 ^{2\pi} {\cal P}(\phi)^{1+\beta/2} {\rm e}^  {-{\beta\over 2T}{\cal
1204: P}(\phi)}d\phi
1205: \end{equation}
1206: where ${\cal P}(\phi)$ is the Poisson Kernel Eq.~(\ref{sRMT}). The general
1207: case of
1208: $M$-equivalent open channels was studied in \cite{FS96b,FS97} where it was
1209: found that
1210: for broken time reversal symmetry the probability distribution of partial
1211: delay
1212: times is
1213: \begin{equation}
1214: \label{wignersmith4}
1215: {\cal P}(T)= \frac{(-1)^M}{M! T^{M+2}}\frac {\partial^M}{\partial(T^{-1})^M}
1216: \left[ {\rm e}^  {-g/T}I_0(T^{-1}{\sqrt(g^2-1)}) \right]
1217: \end{equation}
1218: where $I_0(x)$ stands for the modified Bessel function.
1219: 
1220: \begin{figure}
1221: \begin{center}
1222: \epsfxsize.5\textwidth%
1223: \epsfbox{sfig9.eps}
1224: \caption{
1225: The distribution of the scaled partial delay times $T$ for various graphs with
1226: Neumann boundary conditions. The dashed lines correspond to the RMT expectation
1227: (\ref{wignersmith3},\ref{wignersmith4}): (a) One channel and $A=0$; (b) $M=V$
1228: channels and $A\ne 0$.}
1229: \end{center}
1230: \end{figure}
1231: 
1232: To investigate the statistical properties of the partial delay times for our system we 
1233: had calculated ${\cal P}(T)$ for various graphs. The resulting distributions are shown 
1234: in figure~9  together with the RMT predictions (\ref{wignersmith3},\ref{wignersmith4}).
1235: An overall agreement is evident.  Deviations appear at the short time regime (i.e. 
1236: short  orbits), during which the ``chaotic" component due to multiple scattering is
1237: not yet fully developed \cite {RBUS90}.
1238: 
1239: %----------------------------------------------------------------------------
1240: 
1241: \subsection{\bf Conductance Distribution }
1242: 
1243: Due to the recent experimental investigation of electronic transport through mesoscopic
1244: devices  \cite{MRWHG92}, the study  of the statistical properties of the conductance 
1245: gained some importance. For a device connected to reservoirs by leads, the Landauer-
1246: B\"uttiker formula relates its conductance $G$ to the transmission coefficient $T_G$ 
1247: by the expression $G=(2{\rm e}^  2/\hbar) T_G$. When each lead supports one channel, 
1248: the transmission coefficient can be written in terms of the $S^{(M)}-$matrix as
1249: \begin{equation}
1250: \label{landauer}
1251: T_G= \sum_{i\neq j}^M |S_{i,j}^{(M)}|^2 = 1-|S_{j,j}^{(M)}|^2
1252: \end{equation}
1253: where $j$ is the input channel.
1254: \begin{figure}
1255: \begin{center}
1256: \epsfxsize.5\textwidth%
1257: \epsfbox{sfig10.eps}
1258: \caption{ Conductance distribution ${\cal P}(T_G)$ for a graph with $M=2$
1259: channels: (a) Time reversal symmetry is preserved ($A=0)$. The solid line is
1260: the RMT results Eq.~(\ref{ptrans}) where we had used Eq.~(\ref{s2}) for the
1261: Poisson Kernel while the dashed line is the approximate expression (\ref{Slarge1});
1262: (b) Broken time reversal symmetry ($A\ne 0$). The solid line is the RMT result
1263: Eq.~(\ref{Slarge2}).}
1264: \end{center}
1265: \end{figure}
1266: 
1267: In the absence of direct processes, the distribution of conductance ${\cal P}(T_G)$ 
1268: for arbitrarily number of channels was worked out within RMT, and the results 
1269: describe in a satisfactory way both the numerical calculations and the experimental 
1270: data (for a review see \cite{MB99} and references therein). However in cases where 
1271: direct processes appear significantly, one must use the Poisson's kernel (\ref{sRMT}) 
1272: in its full  generality. This is exactly the case with Neumann graphs since ${\bar S} 
1273: \neq 0$. The  probability distribution of the transmission $T_G$ is then
1274: \begin{equation}
1275: \label{ptrans}
1276: {\cal P} (T_G) = \int \delta(T_G-\sum_{i\neq j} |S_{i,j}|^2) {\cal P}_{\bar
1277: S}(S)
1278: d\mu(S).
1279: \end{equation}
1280: 
1281: For the case $M=2$ and for a diagonal ${\bar S}-$matrix (only direct
1282: reflection) with
1283: equivalent channels ${\bar S}_{11}={\bar S}_{22}={\bar S}$,
1284: Eq.~(\ref{ptrans}) can be
1285: worked out analytically \cite{BB94} in the limit of strong reflection and
1286: $T_G\ll 1$.
1287: The resulting expression is:
1288: \begin{equation}
1289: \label{Slarge1}
1290: {\cal P}_{\beta=1}(T_G)=
1291: {8\over \pi^2(1-{\bar S}^2)} T_G^{-1/2}\quad , \quad  T_G\ll (1-{\bar 
1292: S}^2)^2
1293: \end{equation}
1294: For $\beta=2$ one can  compute in a close form the whole distribution
1295: \cite{MB99}
1296: \begin{equation}
1297: \label{Slarge2}
1298: {\cal P}_{\beta=2}(T_G)= (1-{\bar S}^2) \frac{(1-{\bar S}^4)^2+2{\bar 
1299: S}^2 (1+
1300: {\bar S}^4)T_G+ 4{\bar S}^4T_G^2}{((1-{\bar S}^2)^2 + 4{\bar S}^2 
1301: T_G)^{5/2}}.
1302: \end{equation}
1303: 
1304: Our numerical results for a regular graph with two leads are plotted in 
1305: figure~10
1306: together with the RMT results 
1307: (\ref{ptrans},\ref{Slarge1},\ref{Slarge2}). Notice
1308: that,  the small conductances are emphasized, because of the presence of 
1309: direct
1310: reflection  and no direct transmission. At the same time, the most 
1311: pronounced
1312: differences between  the two symmetry classes are for $T_G\ll 1$, where for
1313: $\beta=1$ the conductance distribution diverges while it take a finite 
1314: value for
1315: $\beta = 2$.
1316: 
1317: %----------------------------------------------------------------------------
1318: 
1319: \section{\bf Scattering from star-graphs.}
1320: \label{sec:star}
1321: 
1322: So far, we have studied scattering from well connected graphs, and we have shown that
1323: many statistical properties of the scattering matrix are described by RMT. We shall
1324: dedicate this section to demonstrate the statistical properties of the $S-$matrix for
1325: graphs which have non-uniform connectivity.
1326: 
1327: A representative example of this category are the ``star" (or ``Hydra") graphs 
1328: \cite{KS99,BK99}. They consist of $v_0$ bonds, all of which emanate from a single 
1329: common vertex labeled with the index $i=0$. The vertex at $i=0$, will be referred
1330: to in the sequel as the  {\it head}. The total number of vertices for such a graph 
1331: is $V=v_0+1$, and the vertices at the end of the bonds will be labeled by $i=1
1332: \cdots v_0$. The star is a bipartite graph, a property which implies e.g., that there
1333: exists no periodic orbits of odd period \cite{KS99}. To turn a star graph  into a 
1334: scattering system we add a lead to its head.
1335: 
1336: It is a simple matter to derive the scattering matrix $S=S^{(M=1)}$ for a Neumann star. 
1337: It reduces to a phase factor, 
1338: \begin{equation}
1339: \label{contH}
1340: S(k)\equiv {\rm e}^  {i\phi(k)} = {-\sum_{i=1}^{v_0} \tan(kL_i) +
1341: i \over \sum_{i=1}^{v_0} \tan(kL_i) + i}
1342: \end{equation}
1343: 
1344: The spectrum $\{k_n\}$ of the close system can be identified as the set of wave-vectors
1345: for which $S(k)$ equals  1, which implies that no current flows in the lead. The 
1346: resulting quantization condition is
1347: \begin{equation}
1348: \label{eq:starb}
1349: 1-S(k)=0\longleftrightarrow {2\sum_{i=1}^{v_0} \tan(k_nL_i) \over
1350: \sum_{i=1}^{v_0}
1351: \tan(kL_i) + i}=0\quad,
1352: \end{equation}
1353: which is satisfied once $\sum_{i=1}^{v_0} \tan(k_nL_i)=0$. This is 
1354: identical with
1355: the condition derived in \cite{KS99}.
1356: 
1357: The poles $\{\kappa_n\}$ are the complex zeros of
1358: \begin{equation}
1359: \label{openH}
1360: \sum_{i=1}^{v_0}\tan(\kappa_nL_i) + i =0.
1361: \end{equation}
1362: To first order in $\frac {1}{v_0}$, we get
1363: \begin{equation}
1364: \label{pertH}
1365: \Gamma_n^{(1)} = {1\over\sum_{i=1}^{v_0}{L_i\over \cos(2k_nL_i) +1}}\
1366: \end{equation}
1367: which can be used as a starting point for the exact evaluation of the 
1368: poles. For the
1369: latter one has to perform a self consistent search for the complex zeros 
1370: of the
1371: secular equation (\ref{openH}). This is a time consuming process and the 
1372: correct
1373: choice of the initial conditions is very important.
1374: 
1375: In figure~11  we present our numerical results for the distribution of 
1376: rescaled
1377: resonance widths ${\cal P}(\gamma)$ for a star with $v_0=20$. The data 
1378: are in
1379: excellent agreement with the RMT expectation given in Eq.~(\ref{chi2}). 
1380: We point
1381: that in this case the coupling  to the continuum is weak since $g\simeq 
1382: 10 \gg 1$
1383: and therefore the $\chi^2$-distribution  with $M=1$ is applicable.
1384: 
1385: \begin{figure}
1386: \begin{center}
1387: \epsfxsize.5\textwidth%
1388: \epsfbox{sfig11.eps}
1389: \caption{ The rescaled resonance width distribution ${\cal P}(\gamma)$ for
1390: a star graph
1391: with $v_0=20$. The solid line is the RMT prediction Eq.~(\ref{chi2}).}
1392: \end{center}
1393: \end{figure}
1394: 
1395: Using Eq.~(\ref{contH}) we get the following relation for the scaled 
1396: delay times
1397: \begin{equation}
1398: \label{delayH}
1399: T(k) ={\Delta\over 2\pi}{2\sum_{i=1}^{v_0} {L_i\over \cos^2(kL_i)}\over 1+
1400: \left(\sum_{i=1}^{v_0}\tan k L_i\right)^2}
1401: \end{equation}
1402: which can be used to generate ${\cal P}(T)$. The latter is reported in figure~12
1403: together with the RMT prediction ~(\ref{wignersmith4}). We notice that although
1404: the tail of the distribution agrees reasonably well with the RMT prediction, there
1405: are considerable deviations at the origin.
1406: 
1407: A peculiarity of the star graph is that the mean time delay is twice larger than the
1408: expected one from semiclassical considerations. To be more specific, for a generic
1409: chaotic system coupled to $M$ channels, one has \cite{DS92a}
1410: \begin{equation}
1411: \label{taumH}
1412: \langle T\rangle _k \simeq {M+1\over M} {\Delta\over 2\pi} \langle l_p\rangle _k
1413: \end{equation}
1414: where $\langle l_p\rangle _k\simeq \Gamma_{cl}^{-1}$ is the average length of the 
1415: classical paths inside the interaction area. Thus for the star with $M=1$, we would 
1416: expect based on Eq.~(\ref{taumH}) that $\langle T\rangle _k \simeq 2 \Gamma_{cl}$. 
1417: However, due to the fact that all the periodic orbits on the  star graph are self 
1418: tracing we get an additional factor of $2$ and thus
1419: \begin{equation}
1420: \label{mtHc}
1421: \langle T\rangle _k= {\Delta\over 2\pi} 4 \Gamma_{cl}^{-1}.
1422: \end{equation}
1423: 
1424: \begin{figure}
1425: \begin{center}
1426: \epsfxsize.5\textwidth%
1427: \epsfbox{sfig12.eps}
1428: \caption {The distribution of the scaled partial delay times ${\cal P}(T)$ for a star 
1429: graph with $v_0=20$. The dashed line is the RMT prediction Eq.~(\ref{wignersmith3}).
1430: In the upper inset we show the same data in a double logarithmic plot. In the lower 
1431: inset we plot the numerical results for $<T>_k$ ($\circ$). The solid line is the 
1432: asymptotic value $1$ expected from semiclassical considerations.}
1433: \end{center}
1434: \end{figure}
1435: 
1436: The corresponding classical decay rate $\Gamma_{cl}$ can be found exactly by a direct
1437: evaluation of the eigenvalues of the classical evolution operator ${\tilde U}$. One 
1438: can easily show that
1439: \begin{equation}
1440: \label{UH1}
1441: {\tilde U} =\left( \begin{array}{ll}
1442:   0 & {\bf 1}  \\
1443: ({\tilde \sigma^{(0)}}) ^2 & 0
1444: \end{array}  \right)
1445: \end{equation}
1446: where ${\tilde \sigma}^{(0)}$ is the $v_0\times v_0$ vertex scattering matrix at
1447: the star head as defined in Eq.~(\ref{Neumann}), while ${\bf 1}$ denotes the
1448: $v_0\times v_0$ identity matrix. The square of the classical evolution operator
1449: ${\tilde U}^2$ has a block diagonal form
1450: \begin{equation}
1451: \label{UH2}
1452: {\tilde U}^2 =\left( \begin{array}{ll}
1453: ({\tilde \sigma^{(0)}}) ^2 & 0 \\
1454: 0 & ({\tilde \sigma^{(0)}}) ^2
1455: \end{array}  \right).
1456: \end{equation}
1457: and its spectrum consists of the values $1-{4\over (1+v_0)^2}$ with multiplicity
1458: 2 and $1-{4\over (1+v_0)}$ with multiplicity $2v_0-2$. Therefore the spectrum of
1459: ${\tilde U}$ is
1460: \begin{eqnarray}
1461: \label{UH3}
1462: \lambda _u &=& \pm \sqrt {1-{4\over(1+v_0)^2}}\\
1463:             &=& \pm \sqrt {1-{4\over 1+v_0}} \quad \quad {\rm with}\,\, {\rm
1464: multiplicity}
1465: \,\, v_0-1\nonumber.
1466: \end{eqnarray}
1467: For short times where the classical evolution is applicable, the dominant eigenvalue
1468: is $\lambda_u=\sqrt{1-{4\over 1+v_0}}$ leading to a classical decay rate 
1469: \begin{equation}
1470: \label{Hgammac}
1471: \Gamma_{cl} \simeq {2\over 1+v_0}.
1472: \end{equation}
1473: Substituting Eq.~(\ref{Hgammac}) we get eventually that $\langle T\rangle _k = {1+
1474: v_0\over v_0} \ \overrightarrow{\scriptstyle v_0{\rightarrow \infty }}\ 1$ which 
1475: indicate that the mean time a particle spend inside the interaction regime is
1476: proportional to the Heisenberg time. Our numerical calculations reported in the lower 
1477: inset of Fig.~12 agrees nicely with the semiclassical prediction ~(\ref{mtHc}).
1478: 
1479: Finally, we analyze the distribution of $S$ when generated over different realizations 
1480: of the lengths of the bond. For the one channel case this is equivalent with the 
1481: distribution ${\cal P} (\phi)$ of the phase of the $S$-matrix. To derive the latter, 
1482: it is convenient to rewrite the $S$-matrix in the following form
1483: \begin{equation}
1484: \label{hydraPs}
1485: S\equiv \exp(i\phi) =(-K + i)/(K + i),
1486: \end{equation}
1487: with $ K = \sum_{i=1}^v \tan(kL_i) $. The probability distribution of
1488: $K$ is
1489: \begin{eqnarray}
1490: \label{hydra2}
1491: {\cal P}(K)& = &\left \langle \delta\left(K-\sum_{i=1}^{v}
1492: \tan (kL_i) \right)\right \rangle_{L}\nonumber \\
1493: & = & \frac 1{2\pi} \int {\rm e}^  {iKx} dx \left( \frac1 {\Delta L}
1494: \int_{L_{min}}^{L_{max}} dL {\rm e}^  {-ix\tan(kL)}\right)^v \nonumber \\
1495: &=& \frac 1{\pi}
1496: \frac {v}{v^2+K^2} \ .
1497: \end{eqnarray}
1498: Thus, with $\bar {S} \equiv \langle S\rangle =\frac{1-v}{1+v}$, we get
1499: \begin{equation}
1500: \label{hydra3}
1501: {\cal P}(\phi) = {\cal P}(K) \left|\frac{dK}{d\phi}\right| =
1502: \frac 1{2\pi} \frac{1-{\bar S}^2}{1+{\bar S}^2-2{\bar S} \cos\phi} \ .
1503: \end{equation}
1504: Equation ~(\ref{hydra3}) reproduces the Poisson's kernel for a one-channel
1505: scattering
1506: matrix, derived in the framework of RMT \cite{FS97}. The conditions under
1507: which this
1508: result is derived in \cite{FS97} are fulfilled {\it exactly} in the present
1509: case. Our numerical results are reported in figure~13 and are in
1510: excellent agreement  with the theoretical prediction Eq.~(\ref{hydra3}).
1511: 
1512: \begin{figure}
1513: \begin{center}
1514: \epsfxsize.5\textwidth%
1515: \epsfbox{sfig13.eps}
1516: \end{center}
1517: \label{fig:starphase}
1518: \caption{The distribution of phases ${\cal P}(\phi)$ of the $S_H$-matrix
1519: for two stars with $v_0=15$ and $v_0=30$. The solid lines are the 
1520: corresponding
1521: theoretical predictions (\ref{hydra3}).}
1522: \end{figure}
1523: 
1524: %-------------------------------------------------------------------------
1525: 
1526: \section{\bf Conclusions}
1527: \label{sec:conclusions}
1528: 
1529: In this paper, we turned quantum graphs into scattering systems. We show that they
1530: combine the desirable features of both behaving ``typically" and being mathematically
1531: simple. Thus, we propose them as a convenient tool to study generic behavior of
1532: chaotic scattering systems.
1533: 
1534: The classical dynamics on an open graph was defined, and the classical staying
1535: probability was shown to decay in an exponential way. The resulting classical
1536: escape rate was calculated and used to describe the properties of the corresponding
1537: quantum system. The scattering matrix was written in terms of classical orbits and
1538: an exact trace formula for the resonance density was found. A gap for the resonance
1539: widths has been obtained for ``generic" graphs and its absence was explained for
1540: Neumann graphs. An analysis of the cross section autocorrelation function was 
1541: performed and its non-universal characteristics were explained in terms of the short
1542: classical scattering trajectories. Finally, due to the relative ease by which a large
1543: number of numerical data can be computed for the graph models, we had performed a
1544: detail statistical analysis of delay times, resonance widths and distribution of the
1545: $S$-matrix. Our results compares well with the predictions of RMT indicating that our
1546: model can be used in order to understand the origin of the connection between RMT and
1547: the underlying classical chaotic dynamics.
1548: 
1549: The results reported here, complete our previous investigations on graphs. We conclude
1550: that quantum graphs may serve as a convenient paradigm in the area of quantum chaos,
1551: both for spectral and scattering studies.
1552: 
1553: 
1554: \section{\bf Acknowledgments}
1555: 
1556: We acknowledge many useful discussions with Y. Fyodorov and H. Schanz. This work was
1557: supported by the Minerva Center for Nonlinear Physics, and an Israel Science Foundation
1558: Grant. (T.~K.) acknowledges a postdoctoral fellowship from the Feinberg School, The 
1559: Weizmann Institute of Science.
1560: 
1561: %-----------------------------------------------------------------------------
1562: 
1563: \begin{thebibliography}{99}
1564: 
1565: \bibitem{A85}  S. Alexander, Phys. Rev. B {\bf 27}, 1541 (1985).
1566: 
1567: \bibitem{FJK87}  C.~Flesia, R.~Johnston and H.~Kunz, Europhys. Lett. {\bf 3}
1568: , 497 (1987).
1569: 
1570: \bibitem{ML77}  R. Mitra and S. W. Lee, {\it Analytical techniques in the
1571: theory of guided waves,} Macmillan, New York.
1572: 
1573: \bibitem{A81}  P. W. Anderson, Phys. Rev. B {\bf 23}, 4828 (1981); B. 
1574: Shapiro,
1575: Phys. Rev. Lett. {\bf 48} 823 (1982).
1576: 
1577: \bibitem{CC88} J. T. Chalker and P. D. Coddington, J. Phys. C {\bf 21}, 2665
1578: (1988); Rochus Klesse and Marcus Metzler, Phys. Rev. Lett. {\bf 79}, 721
1579: (1997); R. Klesse, Ph. D. Thesis, Universitat zu K\"oln, AWOS-Verlag, 
1580: Erfurt,
1581:   1996.
1582: 
1583: \bibitem{NYO94}  Y.~Avishai, J. M. Luck, Phys. Rev. B {\bf 45}, 1074 (1992);
1584: T.~Nakayama, K.~Yakubo, R.~L.~Orbach, Rev. Mod. Phys. {\bf 66}, 381 (1994).
1585: 
1586: \bibitem{Ibook}  Y. Imry, {\it Introduction to Mesoscopic Systems}, Oxford,
1587: New York (1996); D. Kowal, U. Sivan, O. Entin-Wohlman, Y. Imry, Phys. Rev. B
1588: {\bf 42}, 9009 (1990).
1589: 
1590: \bibitem{MT01} J. Vidal, G. Montambaux, D. Doucot, Phys. Rev. B {\bf 62},
1591: R16294 (2000).
1592: 
1593: \bibitem{Z98} Zhang et al. Phys.~Rev.~Lett. {\bf 81}, 5540 (1998).
1594: 
1595: \bibitem{V98} Colin de Verdi\`ere. {\em Spectres de Graphes}. Soci\'et\'e
1596: Math\'ematique \phantom{$[$19$]$} de France, Marseille (1998).
1597: 
1598: \bibitem{E95} P.~Exner, Phys. Rev. Lett. {\bf 74}, 3503 (1995); P.~Exner,
1599: P.~Seba, Rep. Math. Phys. {\bf 28}, 7 (1989); J.~E.~Avron, P.~Exner, Y. 
1600: Last,
1601: Phys. Rev. Lett. {\bf 72}, 896 (1994); R. Carlson, Electronic Journal of
1602: Differential Equations {\bf 6}, 1 (1998).
1603: 
1604: \bibitem{R83} Jean-Pierre Roth, in: {\it Lectures Notes in Mathematics: 
1605: Theorie
1606: du Potentiel}, A. Dold and B. Eckmann, eds. (Springer--Verlag) 521-539.
1607: 
1608: 
1609: \bibitem{KS97} Tsampikos Kottos and Uzy Smilansky. Phys.~Rev.~Lett. {\bf 
1610: 79},
1611: 4794 (1997).
1612: 
1613: \bibitem{KS99}Tsampikos Kottos and Uzy Smilansky, Annals of Physics {\bf 
1614: 274},
1615: 76 (1999).
1616: 
1617: \bibitem{A99}Akkermans E, Comtet A, Desbois J, Montambaux G, Texier C,
1618: Annals of Physics {\bf 284}, 10 (2000).
1619: 
1620: \bibitem{BK99}Berkolaiko G. and Keating J.~Phys.~A {\bf 32} 7827 (1999);
1621: Berkolaiko G., Bogomolny E.B., and Keating J.~Phys.~A {\bf 34} 335 (2001).
1622: 
1623: \bibitem{SS00}H. Schanz and U. Smilansky. Phys.~Rev.~Lett. {\bf 84}, 
1624: 1427 (2000);
1625: Proceedings of the Australian Summer School on Quantum Chaos and 
1626: Mesoscopics,
1627: Canberra, Australia, 1999. [chao-dyn/9904007]; Tsampikos Kottos and Holger
1628: Schanz, Physica E {\bf 9}, 523 (2001).
1629: 
1630: \bibitem{T00} G. Tanner, J.~Phys.~A {\bf 33}, 3567 (2001).
1631: 
1632: \bibitem{BSW02} G. Berkolaiko, H. Schanz, R. S. Whitney, Phys. Rev. Lett.
1633: {\bf 88}, 104101 (2002).
1634: 
1635: \bibitem{BG00} F. Barra, P. Gaspard, J. Stat. Phys. {\bf 101}, 283 (2000);
1636: F. Barra, P. Gaspard, Phys. Rev. E {\bf 63}, 066215 (2001).
1637: 
1638: \bibitem{K01}L. Kaplan, Phys. Rev. E {\bf 64}, 036225 (2001); Tsampikos 
1639: Kottos
1640: in preparation (2002).
1641: 
1642: \bibitem{KS00} Tsampikos Kottos and Uzy Smilansky, Phys. Rev. Lett. {\bf85},
1643: 968 (2000).
1644: 
1645: \bibitem{BG01} F. Barra and P. Gaspard, Phys. Rev. E, {\bf 65}, 016205 
1646: (2002).
1647: 
1648: \bibitem{TM01} C. Texier and G. Montambaux, J.~Phys.~A {\bf 34} 10307 
1649: (2001);
1650: C. Texier, J.~Phys.~A {\bf 35} 3389 (2002).
1651: 
1652: \bibitem{M73} W. H. Miller, Adv. Chem. Phys. {\bf 25} 69 (1974).
1653: 
1654: \bibitem{S89}  U. Smilansky, in {\it Les Houches Summer School on Chaos
1655: and Quantum Physics}, M.-J. Giannoni et.al., eds. (North-Holland) 371-441
1656: (1989) (see also, {\it Les Houches Summer School on mesoscopic quantum 
1657: physics},
1658: E. Akkermans et. al., eds. (North-Holland) 373-433 (1994)).
1659: 
1660: \bibitem{GR89} P. Gaspard, in {\it ``Quantum Chaos", Proceedings of
1661: E. Fermi Summer School 1991}, G. Casati et. al., eds. (North-Holland)
1662: 307; P. Gaspard and S. A. Rice, J. Chem Phys {\bf 90}, 2225,2242,2255(1989);
1663: {\bf 91} E3279, (1989).
1664: 
1665: \bibitem{RBUS90}  R. Bl\"umel and U. Smilansky , Phys. Rev. Lett. {\bf 64},
1666: 241 (1990).
1667: 
1668: \bibitem {EVP99} B. Eckhardt, I. Varga, P. Pollner, Physica E {\bf 9}, 535
1669: (2001).
1670: 
1671: \bibitem{M75} P. A. Moldauer, Phys. Rev. C {\bf 11}, 426 (1975); P. A.
1672: Moldauer,
1673: Phys. Rep. {\bf 157}, 907 (1967); M. Simonius, Phys. Lett. {\bf 52B}, 279
1674: (1974).
1675: 
1676: \bibitem{MPS85} P. Mello, P. Pereyra, and T. H. Seligman, Ann. Phys. (NY)
1677: {\bf 161}, 254 (1985).
1678: 
1679: \bibitem{JSA92} R. A. Jalabert, A. D. Stone and Y. Alhassid, Phys. Rev. 
1680: Lett.
1681: {\bf 68}, 3468 (1992).
1682: 
1683: \bibitem{HILSS92} F. Haake, F. Izrailev, N. Lehmann, D. Saher, H-J Sommers,
1684: Z. Phys. B {\bf 88}, 359 (1992); N. Lehmann, D. Saher, V. Sokolov,  H-J
1685: Sommers,
1686: Nucl. Phys. A {\bf 582}, 223 (1995).
1687: 
1688: \bibitem{BB94} P. W. Brouwer and C. W. J. Beenakker, Phys. Rev. B {\bf 50},
1689: 11263 (1994).
1690: 
1691: \bibitem{FS96b}Y. V. Fyodorov and H.-J. Sommers, Phys. Rev. Lett. {\bf 76},
1692: 4709 (1996); Y. V. Fyodorov, D. V. Savin, and H.-J. Sommers, Phys. Rev. E
1693: {\bf 55} R4857 (1997).
1694: 
1695: \bibitem{FS96} H-J Sommers, Y. V. Fyodorov, and M. Titov, J. Phys. A:
1696: Math. Gen. {\bf 32} L77 (1999); Y. Fyodorov, H-J Sommers, Pis'ma Zh. Eksp.
1697: Teor.
1698: Fiz. {\bf 63} 970 (1996) [JETP Lett. {\bf 63},1026 (1996)]
1699: 
1700: \bibitem{FS97} Y. Fyodorov, H-J Sommers, J. Math. Phys. {\bf 38}
1701: 1918 (1997); P. Mello in {\it Les Houches Summer School on Chaos and
1702: Quantum Physics}, E. Akkermans et.al., eds (North-Holland) 437-491 (1994).
1703: 
1704: \bibitem{BFB97} P. W. Brouwer, K. M. Frahm, C. W. Beenakker, Phys. Rev. 
1705: Lett.
1706: {\bf 78}, 4737 (1997).
1707: 
1708: \bibitem{GM98} V. A. Gopar and P. Mello, Europhys. Lett. {\bf 42}, 131 
1709: (1998).
1710: 
1711: \bibitem{MB99} P. A. Mello and H. U. Baranger, Waves in Random Media 
1712: {\bf 9},
1713: 105 (1999).
1714: 
1715: \bibitem{DS92a} E. Doron, and U. Smilansky, Nonlinearity {\bf 5}, 1055 
1716: (1992);
1717: E. Doron, and U. Smilansky, Phys. Rev. Lett. {\bf 68}, 1255 (1992).
1718: 
1719: \bibitem {MW69}C. Mahaux and H. A. Weidenmuller, {\it ``Shell Model Approach
1720: in Nuclear Reactions''}, (North-Holland, Amsterdam), (1969).
1721: 
1722: \bibitem{Krein} M.~Sh. Birman and D.~R. Yafaev.
1723: \newblock The spectral shift function. {T}he work of {M}. {G}. {K}rein and
1724:    its further development.
1725: \newblock {\em St. Petersburg Math. J.}, 4:833--870, 1993.
1726: 
1727: \bibitem{SZ88} V. Sokolov and G. Zelevinsky, Phys. Lett. B {\bf202} 10 (1988); 
1728: E. Sobeslavsky, F. M. Dittes, and I. Rotter, J. Phys. A: Math. Gen. {\bf 28},
1729: 2963 (1995).
1730: 
1731: \bibitem{E60}T. Ericson, Phys. Rev. Lett. {\bf 5}, 430 (1960); T. Ericson,
1732: Ann. Phys. (NY) {\bf 23} 390 (1963).
1733: 
1734: \bibitem{MRWHG92}C. M. Marcus, A. J. Rimberg, R. M. Westervelt, P. F. 
1735: Hopkins,
1736: and A. C. Gossard, Phys. Rev. Lett. {\bf 69}, 506 (1992).
1737: 
1738: \end{thebibliography}
1739: 
1740: 
1741: \end{document}
1742: