cond-mat0504513/qds.tex
1: %\documentstyle[prl,aps,amssymb,epsfig,graphicx,tabularx,multicol]{revtex}
2: \documentclass[11pt]{article}
3: 
4: \usepackage{graphicx}
5: \usepackage{amsmath}
6: \usepackage{amssymb}
7: \usepackage[english,francais]{babel}
8: 
9: %\usepackage{showkeys}
10: 
11: %Projet d'article : Alain Comtet, Jean Desbois et Christophe Texier
12: %24 Juin 2004
13: %update: 05/04/2005
14: %Commande pour le bulletin de liaison du laboratoire de probabilites
15: %                                             (envoye le 07/04/2005)
16: %
17: %Apres rapports des referees...
18: %update: 27/07/2005
19: %
20: %Dernier nettoyage pour nouvelle version sur cond-mat: 06/09/2005
21: %
22: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
23: %%%%%%%%%%%%%  Journal of Physics A: Mathematical and General  %%%%%%%%%%%%%%%%
24: %
25: % Reference number: 198380 Surname: comtet
26: %
27: % Title: Functionals of the Brownian motion, localization and metric graphs
28: %
29: % Date      Status
30: %01 Aug 2005    Accepted for publication
31: %22 Jul 2005    Resubmission
32: % 6 Jul 2005    We ask to extend the deadline to the 27/07 
33: %       (I was in Russia from 24/06 to 05/07 and Alain in vacations afterwards)
34: %22 Jun 2005    Reports send to author (deadline: 20/07)
35: %03 Jun 2005    Manuscript for adjudication 
36: %03 Jun 2005    Referee report received
37: %03 Jun 2005    Article sent to Adjudicator
38: %27 May 2005    Referee report received
39: %25 Apr 2005    Referee did not report
40: %25 Apr 2005    Article sent to referee
41: %21 Apr 2005    Referee did not report
42: %21 Apr 2005    Article sent to referee
43: %19 Apr 2005    Article sent to referee
44: %19 Apr 2005    Article sent to referee
45: %18 Apr 2005    New submission received and acknowledged
46: %
47: % Reference: J. Phys. A: Math. Gen. {\bf38}, R341-R383 (2005)
48: %
49: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: \textwidth= 16cm
52: \oddsidemargin= 0.5cm
53: \evensidemargin=-0.5cm
54: \topmargin=-2cm
55: \textheight= 24cm
56: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% AMS Fonts %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
57: \newfont{\ensmathquatorze}{msbm10 scaled 1400}
58: \newfont{\ensmathonze}{msbm10 scaled 1100}
59: \newfont{\ensmathdix}{msbm10}
60: \newfont{\ensmathneuf}{msbm10 scaled 833}
61: \newfont{\ensmathhuit}{msbm10 scaled 694}
62: \newfam\ensmathfam                        
63: \textfont\ensmathfam=\ensmathonze        
64: \scriptfont\ensmathfam=\ensmathdix       
65: \scriptscriptfont\ensmathfam=\ensmathhuit
66: \def\ensmf{\fam\ensmathfam\ensmathonze}         % Usage: {\ensmf texte...}
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68: \def\be{\begin{equation}}
69: \def\ee{\end{equation}}
70: \def\benn{$$}
71: \def\eenn{$$}
72: \def\bea{\begin{eqnarray}}
73: \def\eea{\end{eqnarray}}
74: \def\beann{\begin{eqnarray*}}
75: \def\eeann{\end{eqnarray*}}
76: 
77: \def\typ{\mbox{\tiny typ}}
78: 
79: \renewcommand{\leq}{\leqslant}
80: \renewcommand{\geq}{\geqslant}
81: 
82: \def\eqdef{\stackrel{\mbox{\tiny\rm def}}{=}}     % egal a ... par definition
83: \def\eqloi{\stackrel{\mbox{\tiny\rm (loi)}}{=}}     % egal en loi
84: \def\eqlaw{\stackrel{\mbox{\tiny\rm (law)}}{=}}     % egal en loi
85: \newcommand{\ket}[1]{|\kern.3ex#1\kern.3ex\rangle}
86: \newcommand{\bra}[1]{\langle\kern.3ex #1 \kern.3ex|}
87: \newcommand{\APPROX}[1]{                % approximativement egal a .. si ..
88:    {{\raisebox{-.3cm}{$\textstyle\simeq$}} \atop {\scriptstyle{#1}}}}
89: \newcommand{\APPROXs}[1]{               % similaire a ... si ...
90:    {{\raisebox{-.3cm}{$\textstyle\sim$}} \atop {\scriptstyle{#1}}}}
91: \newcommand{\leadto}[1]{                
92:    {{\raisebox{-.3cm}{$\textstyle\longrightarrow$}} \atop {\scriptstyle{#1}}}}
93: \newcommand{\mean}[1]{\left\langle #1 \right\rangle} % moyenne statistique
94: \newcommand{\smean}[1]{\langle #1 \rangle} % moyenne statistique
95: \newcommand{\cum}[1]{\langle\langle #1 \rangle\rangle} % cumulant
96: 
97: \newcommand{\EXP}[1]{{\mbox{\large e}}^{#1}}         % exponentielle
98: %\renewcommand{\cosh}[1]{\mathop{\mathrm{ch}}\nolimits #1} % cosinus hyperbolique
99: %\renewcommand{\sinh}[1]{\mathop{\mathrm{sh}}\nolimits #1} % sinus hyperbolique
100: %\renewcommand{\tanh}[1]{\mathop{\mathrm{th}}\nolimits #1} % tangente hyperbolique
101: \newcommand{\argcosh}{\mathop{\mathrm{argch}}\nolimits}
102: \newcommand{\argsinh}{\mathop{\mathrm{argsh}}\nolimits}
103: 
104: \newcommand{\re}{\mathop{\mathrm{Re}}\nolimits}      % partie reelle
105: \newcommand{\im}{\mathop{\mathrm{Im}}\nolimits}      % partie imaginaire
106: \newcommand{\tr}[1]{\mathop{\mathrm{Tr}}\nolimits\left\{ #1 \right\}}  % Trace
107: \newcommand{\cotg}{\mathop{\mathrm{cotg}}\nolimits}  % cotangente
108: \newcommand{\sign}{\mathop{\mathrm{sign}}\nolimits}  % signe
109: \renewcommand{\min}[2]{\mathop{\mathrm{min}}\nolimits\left( #1 , #2\right)}  
110: \renewcommand{\max}[2]{\mathop{\mathrm{max}}\nolimits\left( #1 , #2\right)}  
111: 
112: \newcommand{\proba}{\mathop{\mathrm{Prob}}\nolimits}  
113: \newcommand{\heav}{\mathop{\mathrm{\rm Y}}\nolimits}  
114: 
115: \newcommand\Dslash{D\hspace{-0.25cm}/}               % D slash de Dirac
116: 
117: \def\NN{{\ensmf N}}                 % entiers 
118: \def\ZZ{{\ensmf Z}}                 % entiers relatifs
119: \def\QQ{{\ensmf Q}}                 % rationnels 
120: \def\RR{{\ensmf R}}                 % ensemble des reels
121: \def\CC{{\ensmf C}}                 % ensemble des complexes
122: 
123: \def\I{{\rm i}}                  % le i mathematique 
124: \def\D{{\rm d}}                  % la differenciation
125: \def\Dc{{\rm D}}                 % dérivée covariante
126: 
127: \newcommand{\dz}{\partial_{z}}
128: \newcommand{\dzb}{\partial_{\bar z}}
129: \newcommand{\drond}[2]{\frac{\partial #1}{\partial #2}} % derivee partielle
130: 
131: \newcommand\dphi{{\cal D}\phi{\cal D}\bar\phi\:} % mesure pour l'integrale de 
132:                                                  % chemin
133: \newcommand\ab{{\alpha\beta}}
134: \newcommand\ba{{\beta\alpha}}
135: \newcommand\mb{{\mu  \beta}}
136: \newcommand\bm{{\beta\mu}}
137: \newcommand\lab{l_{\alpha\beta}}
138: 
139: \newcommand\ide{\mbox{\bf 1}}%\hspace{-0.1cm}{\rm I}} % matrice identite.
140: 
141: % Pour ajuster la hauteur des diagrammes:
142: %=======================================
143: \newcommand{\diagram}[3]{\raisebox{#3}{\includegraphics[scale=#2]{#1}}}
144: 
145: \def\sw{S_{\rm w}}
146: 
147: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
148: %%%%%%%%%%%%%%%%%%%%%%%%  MACRO JEAN %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
149: %\newcommand{\lab}{l_{\alpha\beta}}
150: 
151: \newcommand{\xab}{x_{\alpha\beta}}
152: \newcommand{\xba}{x_{\beta\alpha}}
153: 
154: \newcommand{\xaa}{x_{ab}}
155: \newcommand{\xbb}{x_{ba}}
156: \newcommand{\Gab}{G_{(\alpha\beta )}}
157: \newcommand{\Gba}{G_{(\beta\alpha )}}
158: \newcommand{\Gaa}{G_{(ab )}}
159: \newcommand{\Gbb}{G_{(ba )}}
160: \newcommand{\Gpaa}{G'_{(ab )}}
161: \newcommand{\Gpbb}{G'_{(ba )}}
162: 
163: \newcommand{\Gpab}{G'_{(\alpha\beta )}}
164: \newcommand{\Gpba}{G'_{(\beta\alpha )}}
165: 
166: \newcommand{\dg}{\partial_{\gamma}}
167: \newcommand{\pal}{\psi_{\alpha\beta}}
168: \newcommand{\pbe}{\psi_{\beta\alpha}}
169: \newcommand{\ppal}{\psi'_{\alpha\beta}}
170: \newcommand{\ppbe}{\psi'_{\beta\alpha}}
171: \newcommand{\fal}{\phi_{(\alpha\beta )}}
172: \newcommand{\fbe}{\phi_{(\beta\alpha )}}
173: \newcommand{\fpal}{\phi'_{(\alpha\beta )}}
174: \newcommand{\fpbe}{\phi'_{(\beta\alpha )}}
175: 
176: \newcommand{\cab}{\chi_{\alpha\beta}}
177: \newcommand{\cba}{\chi_{\beta\alpha}}
178: \newcommand{\cpab}{\chi'_{\alpha\beta}}
179: \newcommand{\cpba}{\chi'_{\beta\alpha}}
180: 
181: \newcommand{\caa}{\chi_{ab}}
182: \newcommand{\cbb}{\chi_{ba}}
183: 
184: \newcommand{\cabi}{\chi_{\alpha\beta_i}}
185: \newcommand{\cbia}{\chi_{\beta_i\alpha}}
186: \newcommand{\cpabi}{\chi'_{\alpha\beta_i}}
187: \newcommand{\cpbia}{\chi'_{\beta_i\alpha}}
188: \newcommand{\sg}{\sqrt {\gamma }}
189: \newcommand{\pab}{\psi_{ab}}
190: \newcommand{\pba}{\psi_{ba}}
191: \newcommand{\ppab}{\psi'_{ab}}
192: \newcommand{\ppba}{\psi'_{ba}}
193: 
194: \newcommand{\dpl}{ \frac{ \D \psi_{\alpha\beta}}{  \D x_{\alpha\beta}} }
195: \newcommand{\dpt}{ \frac{ \D \psi_{\beta\alpha}}{ \D x_{\alpha\beta}} }
196: \newcommand{\dpa}{ \frac{ \D \psi_{ab}} {  \D x_{ab}} }
197: \newcommand{\dpb}{ \frac{ \D \psi_{ba}} { \D x_{ab}} }
198: \newcommand{\dpi}{ \frac{ \D \psi_{\alpha\beta_i}}{ \D x_{\alpha\beta_i}} }
199: \newcommand{\dpj}{ \frac{ \D \psi_{\beta_i\alpha}}{ \D x_{\alpha\beta_i}} }
200: \newcommand{\wal}{W_{\alpha\beta}}
201: \newcommand{\wla}{W_{\beta\alpha}}
202: \newcommand{\wab}{W_{ab}}
203: \newcommand{\wba}{W_{ba}}
204: \newcommand{\Wal}{{\cal W}_{\alpha\beta}}
205: \newcommand{\Wab}{{\cal W}_{ab}}
206: \newcommand{\Wla}{{\cal W}_{\beta\alpha}}
207: \newcommand{\Wba}{{\cal W}_{ba}}
208: 
209: %\def\fp{\displaystyle}
210: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
211: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
212: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
213: 
214: \begin{document}
215: \selectlanguage{english}
216: 
217: %Alain: Brownian functionals in quantum physics~: localization and graphs
218: 
219: \title{Functionals of the Brownian motion,\\ localization and metric graphs}
220: 
221: \author{Alain Comtet$^{(a,b)}$, Jean Desbois$^{(a)}$ and 
222:         Christophe Texier$^{(a,c)}$}
223: 
224: %\date{\today}
225: %\date{April 14, 2005}
226: \date{July 22, 2005}
227: 
228: \maketitle
229: 
230: \noindent
231: \hspace{1cm}
232: \begin{minipage}[t]{13cm}
233: {\small
234: 
235:   \noindent
236:   $^{(a)}$
237:   \begin{minipage}[t]{12.5cm}
238:     Laboratoire de Physique Th\'eorique et Mod\`eles Statistiques,
239:     UMR 8626 du CNRS.\\
240:     Universit\'e Paris-Sud,
241:     B\^at. 100, F-91405 Orsay Cedex, France.
242:   \end{minipage}
243: 
244:   \vspace{0.15cm}
245: 
246:   \noindent
247:   $^{(b)}$
248:   \begin{minipage}[t]{12.5cm}
249:     Institut Henri Poincar\'e, 11 rue Pierre et Marie Curie, F-75005 Paris,
250:     France.
251:   \end{minipage}
252: 
253:   \vspace{0.15cm}
254: 
255:   \noindent
256:   $^{(c)}$
257:   \begin{minipage}[t]{12.5cm}
258:     Laboratoire de Physique des Solides,
259:     UMR 8502 du CNRS.\\
260:     Universit\'e Paris-Sud,
261:     B\^at. 510, F-91405 Orsay Cedex, France.
262:   \end{minipage}
263:   }
264: \end{minipage}
265: 
266: \begin{abstract}
267: We review several results related to the problem of a quantum
268: particle in a random environment.
269: 
270: In an introductory part, we recall how several functionals of the Brownian
271: motion arise in the study of electronic transport in weakly disordered
272: metals (weak localization).
273: 
274: Two aspects of the physics of the one-dimensional strong localization are
275: reviewed~:  some properties of the scattering by a random potential (time
276: delay distribution) and a study of the spectrum of a random potential on a
277: bounded domain (the extreme value statistics of the eigenvalues).
278: 
279: Then we mention several results concerning the diffusion on graphs, and more
280: generally the spectral properties of the Schr\"odinger operator on graphs. The
281: interest of spectral determinants as generating functions characterizing the
282: diffusion on graphs is illustrated.
283: 
284: Finally, we consider a two-dimensional model of a charged particle coupled to
285: the random magnetic field due to magnetic vortices. We recall the connection
286: between spectral properties of this model and winding functionals of the
287: planar Brownian motion.
288: \end{abstract}
289: 
290: \noindent
291: PACS numbers~: 72.15.Rn~; 73.20.Fz~; 02.50.-r~; 05.40.Jc.
292: 
293: %72.   Electronic transport in condensed matter
294: %72.10.-d   Theory of electronic transport; scattering mechanisms
295: %72.10.Bg   General formulation of transport theory 
296: %72.15.Rn Localization effects (Anderson or weak localization)
297: 
298: %73.   Electronic structure and electrical properties of surfaces, interfaces,
299: %      thin films, and low-dimensional structures 
300: %73.23.-b     Electronic transport in mesoscopic systems 
301: %73.20.Fz Weak or Anderson localization
302: 
303: %02.50.-r Probability theory, stochastic processes, and statistics
304: %05.40.Jc Brownian motion
305: 
306: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
307: 
308: \tableofcontents
309: 
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
312: 
313: \section{Introduction}
314: 
315: \subsection{Weak and strong localization}
316: 
317: At low temperature, the electric conductivity $\sigma$ of metals and weakly
318: disordered semiconductors is determined by the scattering of electrons on
319: impurities. It is given by the Drude formula
320: \be
321: \sigma_0=\frac{n_e e^2 \tau_e}{m},
322: \ee
323: where $e$ and $m$ are the charge and mass of the electron respectively. $n_e$
324: is the electronic density and $\tau_e$ the elastic scattering time\footnote{
325:   The elastic scattering time $\tau_e$ is the time characterizing the 
326:   relaxation of the direction of the momentum of the electron. The 
327:   conductivity is proportional to $\tau_e$ for isotropic scattering by 
328:   impurities. For anisotropic scattering the conductivity involves a 
329:   different time  $\tau_{\rm tr}$ called the ``transport'' time
330:   \cite{AshMer76}
331:   (see ref. \cite{AkkMon04} for a discussion within the perturbative approach).
332: }. 
333: This purely classical formula is only valid in a regime where quantum
334: mechanical effects can be neglected. This is the case if the elastic mean free
335: path of electrons $\ell_e=v_F\tau_e$ is large compared with the De Broglie
336: wave length $\lambda_F= 2\pi\hbar /mv_F$ corresponding to  the
337: Fermi energy ($v_F$ is the Fermi velocity).
338: 
339: \vspace{0.25cm}
340: 
341: \noindent%{\bf(1)} 
342: {\it Strong localization.--} When these two length scales
343: are of the same order $\ell_e\approx\lambda_F$ (strong disorder) the fact that
344: the electrons are quantum objects must be taken into account and the wave like
345: character of these particles is of primary importance. It is indeed this wave
346: character which is responsible for the localization phenomenon. The multiple
347: scattering on impurities distributed randomly in space creates random phases
348: between these different waves which can interfere destructively. These
349: interference effects reduce the electronic conductivity. In the extreme case
350: of very strong disorder the waves do not propagate anymore and the system
351: becomes insulating. This is the {\it strong localization} phenomenon which was
352: conjectured by Anderson in 1958.
353: 
354: \vspace{0.25cm}
355: 
356: \noindent%{\bf(2)} 
357: {\it Weak localization.--} In the 80's it was realized that
358: even far from the strong localization regime the quantum transport is
359: affected by the disorder. Diagrammatic techniques used in the weak 
360: disorder limit  $\lambda_F\ll\ell_e$, were initiated by
361: the works of Al'tshuler, Aronov, Gor'kov, Khmel'nitzki{\u\i}, Larkin and Lee
362: \cite{GorLarKhm79,AltKhmLarLee80} (see ref.~\cite{AltLee88} for an 
363: introduction and ref.~\cite{AkkMon04} for a recent presentation). 
364: In this regime, called the {\it weak localization} regime, the
365: Drude conductivity gets a small sample dependent correction whose
366: average, denoted $\smean{\Delta\sigma}$, is called the ``weak localization
367: correction''. $\smean{\cdots}$ denotes averaging with respect to the 
368: random potential.
369: From the experimental side, this phenomenon is well 
370: established and has been the subject of many studies (see 
371: refs.~\cite{Ber84,AroSha87} for review articles).
372: 
373: 
374: \vspace{0.25cm}
375: 
376: \noindent%{\bf(3)} 
377: {\it Phase coherence and dimensional reduction.--} The localization phenomenon
378: comes from the interplay between the wave nature of electronic transport and
379: the disorder. This manifestation of quantum interferences requires that the
380: phase of the electronic wave is well defined, however several mechanisms limit
381: the phase coherence of electrons in metals, among which are the effect of the
382: vibrations of the crystal (electron-phonon interaction) or the
383: electron-electron interaction. We introduce a length scale $L_\varphi$, the
384: phase coherence length, that characterizes the length over which phase
385: breaking phenomenon becomes effective. The lack of phase coherence in real
386: systems is the reason why the strong localization regime has not been
387: observed in experiments on metals. In dimension $d=3$ the strong localization
388: regime is only expected to occur for sufficiently strong
389: disorder\footnote{
390:   Following the scaling ideas initiated by Thouless \cite{LicTho75} and
391:   Wegner \cite{Weg76,Weg79} it was shown in ref.~\cite{AbrAndLicRam79} that 
392:   the localization-delocalization transition exists only for dimension 
393:   $d\geq3$.
394:   In $d=1$ and $d=2$, the fully coherent system is always strongly localized,
395:   whatever the strength of the 
396:   disorder is.\label{Wegner}
397: }. 
398: However, strong localization can also be observed in the weak disorder
399: limit ($\lambda_F\ll\ell_e$) by reducing the dimensionnality\footnote{
400:   The effective dimension of the system is obtained by comparing the sample 
401:   size with the phase coherence length $L_\varphi$. For example, a long 
402:   wire of length $L$ and of transverse dimension $W$ is effectively in 
403:   a 1d regime if $W\ll L_\varphi\ll L$.
404: }. It has been shown in the framework of random matrix
405: theory that the localization length of a weakly disordered quasi-1d wire
406: behaves as $\lambda\sim N_c\ell_e$, where $N_c$ is the number of
407: conducting channels\footnote{
408:   The localization length predicted by the random matrix theory (RMT) can 
409:   be found in ref.~\cite{Bee97}~: 
410:   $\lambda_{\rm RMT}=[\beta(N_c-1)+2]\ell_e$, 
411:   where $\beta=1,2,4$ is the Dyson index describing orthogonal, unitary
412:   and symplectic ensembles, respectively. 
413:   Note that one must be careful with the coefficient involved in 
414:   this relation since the definitions of $\lambda_{\rm RMT}$ and $\ell_e$
415:   differ slightly in RMT and in perturbation theory.
416: } \cite{Dor82,Dor88,MelPerKum88,Bee97}.
417: Therefore the strong localization regime is expected to occur when coherence is
418: kept at least over a scale $\lambda$. At low temperature, in the absence of
419: magnetic impurity, phase breaking mechanisms are dominated by
420: electron-electron interaction, which leads to a divergence of the phase
421: coherence length $L_\varphi(T)\propto(N_c/T)^{1/3}$ predicted in
422: ref.~\cite{AltAroKhm82} and verified in several experiments like
423: \cite{WinRooChaPro86,ThoPepAhmAndDav86,EchGerBozBogNil93b}
424: (see ref.~\cite{PieGouAntPotEstBir03} for recent measurements down to 
425: $40\:$mK). 
426: Therefore the
427: temperature below which strong localization might be observable is given by
428: $L_\varphi(T_*)\sim\lambda$, which leads to $T_*\sim1/(N_c^2d\,\tau_e)$.
429: %(this equality also gives
430: %the threshold beyond which perturbation theory breaks down \cite{AltAro85}).
431: The crossover temperature in metals is out of the experimental range, however
432: it becomes reachable in wires etched at the interface of two semiconducting
433: materials, when the number of conducting channels is highly reduced, and the
434: manifestation of the strong localization has been observed in 
435: ref.~\cite{GerKhaMikBozBog97}.
436: %,KhaGerBog98b}.
437: 
438: \vspace{0.25cm}
439: 
440: \noindent%{\bf(4)} 
441: {\it Strictly one-dimensional case.--} In a weakly disordered and
442: coherent quasi-1d wire, the weak localization regime only occurs for length
443: scales intermediate between the elastic mean free path and the localization
444: length $\ell_e\ll L\ll\lambda$. In the strictly one dimensional case,
445: since $\lambda\simeq4\ell_e$ for weak disorder\footnote{
446:   The elastic mean free path $\ell_e=v_F\tau_e$ is given by the self 
447:   energy $1/(2\tau_e)=-\im\Sigma^{\rm R}(E)$. For example, for a Gaussian
448:   disorder with local correlations, $\smean{V(x)V(x')}={w}\delta(x-x')$, 
449:   we obtain $1/\tau_e\simeq2\pi\rho_0{w}$ for a weak disorder, where 
450:   $\rho_0$ is the free density of states.
451:   In one dimension $\rho_0=1/(\pi v_F)$, therefore 
452:   $\ell_e\simeq v_F^2/(2{w})$, 
453:   which coincides with the {\it high energy} (weak disorder) expansion of the 
454:   localization length 
455:   $\lambda\simeq2v_F^2/{w}$~\cite{AntPasSly81,LifGrePas88}.
456: }, 
457: such a regime does not exist and the system is
458: either ballistic ($\ell_e\gg L$) or strongly localized ($\ell_e\ll L$).
459: 
460: \vspace{0.25cm}
461: 
462: \noindent%{\bf(5)} 
463: Anderson localization in one dimension has been studied by
464: mathematicians and mathematical physicists from the view point of spectral
465: analysis and in connection with limit theorems for products of random
466: matrices \cite{BouLac85}. A breakthrough was the proof of wave localization 
467: in one dimension
468: by Gol'dshtein, Molchanov \& Pastur \cite{GolMolPas77}. Although some
469: progress has been made, the multidimensional case is still out of reach 
470: and the subject of weak localization has almost not been touched in the
471: mathematical literature. One of the main fields of interest in the last twenty
472: years is the investigation of random Schr\"odinger operators in the presence
473: of magnetic fields (see for example the 
474: refs.~\cite{Weg83,BreGroItz84,DesFurOuv95,DesFurOuv96,Fur97,Fur00}~; 
475: for recent results on Lifshitz tails with magnetic field see 
476: refs.~\cite{Fur00,LesWar04}).
477: From the physics side significant advances have been realized. 
478: Field theoretical methods based on supersymmetry provide a general framework
479: for disordered systems and also allow establishing some links with quantum
480: chaos~\cite{Efe97}.
481: 
482: \vspace{0.25cm}
483: 
484: While writing this review, we have tried to collect a large list of references
485: which is however far from being exhaustive.  Ref.~\cite{LifGrePas88}
486: provides a reference book on strong localization, mostly focused on spectrum
487: and localization properties (see also ref.~\cite{Luc92} for a review on 1d
488: discrete models and Lifshitz tails). A recent text about disorder and random
489: matrix theory is ref.~\cite{Efe97}.  Many excellent reviews have been written
490: on weak localization among which \cite{Ber84,ChaSch86} (for the role of
491: disorder and electron-electron interaction, see the book \cite{EfrPol85} 
492: or \cite{LeeRam85}). A recent reference is~\cite{AkkMon04}.
493: 
494: 
495: 
496: \subsection{Overview of the paper}
497: 
498: Section \ref{sec:wl} shows that several functionals of the Brownian motion
499: arise in the study of electronic transport in weakly disordered metals or
500: semiconductors (weakly localized). 
501: The brief presentation of weak localization given in
502: sections~\ref{sec:brownian1} and~\ref{sec:brownian2} follows the heuristic
503: discussion of refs.~\cite{AltLee88,ChaSch86}, which is based on the picture
504: proposed by Khmel'nitzki{\u\i} \& Larkin. In spite of its heuristic
505: character, it allows drawing suggestive connections with well known
506: functionals of the Brownian motion.
507: 
508: The sections \ref{sec:expfun} and \ref{sec:evss} deal with problems of strong
509: localization in one dimension. 
510: A powerful approach to handle such problems is the phase
511: formalism\footnote{
512:   The phase formalism is a continuous version of the Dyson-Schmidt method
513:   \cite{Dys53,Sch57}. A nice presentation can be found
514:   in ref.~\cite{Luc92}.
515: } (presented in refs.~\cite{AntPasSly81,LifGrePas88} for instance). This
516: formalism leads to a broad variety of stochastic processes.
517: Section~\ref{sec:expfun} discusses scattering properties of a random potential
518: and section~\ref{sec:evss} studies spectral properties of a Schr\"odinger
519: operator defined on a finite interval.
520: 
521: In section \ref{sec:graphs}, we review some results obtained for networks of
522: wires (graphs), which can be viewed as systems of intermediate dimension
523: between one and two. We will put the emphasis on spectral determinants, which
524: appear to be an efficient tool to construct several generating functions
525: characterizing the diffusion on graphs (or its discrete version, the random
526: walk).
527: 
528: Finally, in section \ref{sec:magimp}, we show that the physics of a two
529: dimensional quantum particle submitted to the magnetic field of an assembly of
530: randomly distributed magnetic vortices involves fine properties of the
531: planar Brownian motion.
532: 
533: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
534: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
535: 
536: \section{Functionals of the Brownian motion in the context of 
537:          weak localization\label{sec:wl}}
538: 
539: \subsection{Feynman paths, Brownian motion and weak localization
540:             \label{sec:brownian1}}
541: 
542: In the path integral formulation of quantum mechanics, each trajectory is
543: weighted with a phase factor $\EXP{\I S/\hbar}$ where $S$ is the classical
544: action evaluated along this trajectory. The superposition principle states
545: that the amplitude of propagation of a particle between two different points
546: is given by the sum of amplitudes over all paths connecting these two points.
547: This formulation of quantum mechanics is most useful when there is a small
548: parameter with respect to which one can make a quasiclassical expansion. One
549: encounters a similar situation in the derivation of geometrical optics starting
550: from wave optics. In this case the small parameter is the ratio of the wave
551: length to the typical distances which are involved in the problem and the
552: classical paths are light rays. In the context of weak localization the small
553: parameter is $\lambda_F/\ell_e$ and the quasiclassical approximation amounts
554: to sum over a certain subset of Brownian paths. The fact that Brownian paths
555: come into the problem is not so surprising if one goes back to our previous
556: physical picture of electrons performing random walk due to the scattering by
557: the impurities. In the continuum limit, describing the physics at length
558: scales much larger than $\ell_e$, this random walk may be described as a
559: Brownian motion. It can be shown that $\smean{\Delta\sigma}$ is related to the
560: time integrated probability for an electron to come back to its initial
561: position (see ref.~\cite{ChaSch86} for a heuristic derivation or 
562: ref.~\cite{AkkMon04} for a recent derivation in space representation)~:
563: %
564: %A probability is the square modulus of a quantum amplitude. This latter 
565: %is given by a sum of contributions associated to all possible paths.
566: %In the limit $\lambda_F\ll\ell_e$, the phases of these amplitudes are
567: %large random variables, therefore the time reversed paths play a special 
568: %role~:
569: %
570: %Since a probability is the square modulus of a sum of quantum amplitudes,
571: %whose phases are large random variables in the regime $\lambda_F\ll\ell_e$,
572: %the time reversed paths play a special role~: because the weak localization is
573: %an average quantity, reversed paths survive disorder averaging thanks to phase
574: %cancellation and give a contribution that can be written
575: %
576: \be\label{WL}
577: \smean{\Delta\sigma}= -\frac{2e^2 D}{\pi\hbar} \int_{\tau_e}^\infty \D t\, 
578: {\cal P}(\vec r,t|\vec r,0)\,\EXP{-t/\tau_\varphi}
579: \:.\ee
580: where $D$ is the diffusion constant and the factor $2$ accounts for spin 
581: degeneracy. 
582: ${\cal P}(\vec r,t|\vec r\,',0)$ is the Green's function of the diffusion 
583: equation
584: \be
585: \left(\drond{}{t}-D\Delta\right)\,{\cal P}(\vec r,t|\vec r\,',0)
586: = \delta(\vec r-\vec r\,')\,\delta(t)
587: \:.\ee
588: In eq.~(\ref{WL}) the exponential damping at large time describes the lack of
589: phase coherence due to inelastic processes. The phase coherence time
590: $\tau_\varphi$ can be related to the phase coherence length by
591: $L_\varphi^2=D\tau_\varphi$. We have also introduced a cut-off at short time
592: in eq.~(\ref{WL}), that takes into account the fact that the diffusion
593: approximation is only valid for times larger than~$\tau_e$.
594: 
595: 
596: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
597: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
598: 
599: \subsection{Planar Brownian motion~: stochastic area, winding and
600:             magnetoconductance\label{sec:brownian2}}
601: 
602: Weak localization corrections are  directly related to the
603: behaviour of the probability of return to the origin and thus to recurrence
604: properties of the diffusion process. It therefore follows that dimension $d=2$
605: plays a very special role. Consider for instance a thin film whose thickness
606: $a$ is much less than $L_\varphi$. The sample is effectively two dimensional 
607: and the correction to the conductivity is given by
608: \be
609: \smean{\Delta\sigma}= - \frac{e^2}{\pi^2\hbar} \ln(L_\varphi/\ell_e).
610: \ee
611: On probabilistic grounds the logarithmic scaling is of course not unexpected
612: here, it is the same logarithm which occurs in asymptotic laws of the planar
613: Brownian motion \cite{PitYor86,LeG92}. 
614: The neighbourhood recurrence of the planar
615: Brownian motion favors the quantum interference effects and leads to a
616: reduction of the electrical conductivity. A more striking effect is predicted
617: if one applies a constant and homogeneous magnetic field ${\cal B}$ over the
618: sample. In this case the classical action contains a coupling to the magnetic
619: field $S=e{\cal B}A$ where $A$ is a functional of the path given by the line
620: integral
621: \be
622: A= \frac12 \int (x \D y -y \D x) 
623: \:.\ee
624: Properly interpreted, this line integral is nothing but the stochastic area
625: of planar Brownian motion, whose distribution was
626: first computed by P.~L\'evy before the discovery of
627: the Feynman path integral. The weak localization correction reads
628: \be
629: \smean{\Delta\sigma({\cal B})}-\smean{\Delta\sigma(0)}= 
630: \frac{e^2}{2\pi^2\hbar}\int_0^\infty \frac{\D t}{t}\,
631: \EXP{-t/\tau_\varphi}\,
632: \left(1-{\rm E}\!\left[\EXP{2\I e{\cal B}A/ \hbar}\right]\right)
633: \:.\ee
634: The coupling to the magnetic field now appears with an additional factor $2$
635: coming from the fact that the weak localization describes quantum
636: interferences of reversed paths. The expectation ${\rm E}[\cdots]$ is taken
637: over Brownian loops for a time $t$. The magnetoconductivity is given by
638: \cite{AltKhmLarLee80,Ber84,ChaSch86}
639: \be\label{B}
640: \smean{\Delta\sigma({\cal B})}-\smean{\Delta\sigma(0)}= 
641: \frac{e^2}{2\pi^2\hbar}\left[
642:   \psi\left(\frac12+\frac{\phi_0}{8\pi {\cal B} L_\varphi^2} \right)
643:   - \ln\left(\frac{\phi_0}{8\pi {\cal B} L_\varphi^2} \right)
644: \right]
645: \simeq\frac43 \frac{e^2}{\hbar} 
646: \left(\frac{{\cal B}L_\varphi^2}{\phi_0}\right)^2
647: \:,\ee
648: where $\psi(z)$ is the digamma fonction and $\phi_0=h/e$ is the quantum flux.
649: The {\it r.h.s} corresponds to the weak field limit ${\cal
650: B}\ll\phi_0/L_\varphi^2$. This increase of the conductivity with magnetic
651: field is opposite to the behaviour expected classically. This phenomenon is
652: called ``positive (or anomalous) magnetoconductance''. Experimentally this
653: effect is of primary importance~: the magnetic field dependence allows one to
654: distinguish the weak localization correction from other contributions and
655: permits to extract the phase coherent length $L_\varphi$. This expression fits
656: the experimental results remarkably well\footnote{
657:   Note that in materials with strong spin orbit scattering, like gold, the
658:   effect is reversed and a negative magnetoconductance is observed at small
659:   magnetic fields~\cite{HikLarNag80,Ber84}. 
660: } \cite{Ber84}.
661: 
662: Strictly speaking eq.~(\ref{B}) only holds for an infinite sample. In the case
663: of bounded domains the variance depends on the geometry of the system and can
664: be computed explictly for rectangles and strips \cite{DesCom92}. The case of
665: inhomogeneous magnetic fields can be treated along the same line. Consider for
666: instance a magnetic vortex carrying a flux $\phi$ and threading the sample at
667: point $0$. In this case the functional which is involved is not the stochastic
668: area but the index of the Brownian loop with respect to $0$ (winding number).
669: The probability distribution of the index has been computed independently by
670: Edwards \cite{Edw67} in the context of polymer physics and Yor \cite{Yor80} in
671: relation with the Hartman-Watson distribution. For a planar Brownian motion
672: started at $r$ and conditioned to hit its starting point at time 1, the
673: distribution of the index when $n$ goes to infinity is
674: \be
675: {\rm Proba}\left[{\rm Ind}=n\right] \simeq
676: \frac{1}{2\pi^2 n^2}\, {\rm K}_0 (r^2)\,\EXP{-r^2}
677: \:,\ee
678: where ${\rm K}_0(x)$ is the modified Bessel function of second kind (MacDonald 
679: function).
680: The fact that the even moments of this law are infinite is reflected in the
681: non analytic behaviour of the conductivity
682: \be\label{eq9}
683: \smean{\Delta\sigma(\phi)}-\smean{\Delta\sigma(0)}\propto \left|\phi \right|
684: \:.\ee
685: It is interesting to
686: compare with eq.~(\ref{B}) where the quadratic behaviour is given by the
687: second moment of the stochastic area \cite{Mon95}. The behaviour given
688: by eq.~(\ref{eq9}) has been observed experimentally \cite{Gei89,BenKliPlo90}
689: and a theoretical interpretation is provided in ref.~\cite{RamShe87}.
690: 
691: 
692: 
693: \subsection{Dephasing due to electron-electron interaction and 
694:                 functionals of Brownian bridges}
695: 
696: \noindent
697: {\it Dephasing in a wire~: relation with the area below a Brownian bridge.--}
698: Another interesting example of a non trivial connection between a physical
699: quantity and a functional of the Brownian motion occurs in the study of
700: electron-electron interaction and weak localization
701: correction in a quasi-1d metallic wire. 
702: In ref.~\cite{AltAroKhm82}, Al'tshuler, Aronov \& Khmel'nitzki{\u\i} (AAK)
703: proposed to model the effect of the interaction between an electron and its
704: surrounding environment by the interaction with a fluctuating classical field.
705: The starting point is a path integral representation of the probability
706: ${\cal P}(\vec r,t|\vec r\,',0)$ 
707: in which is included the effect of the fluctuating 
708: field that brings a random phase.
709: After averaging over Gaussian fluctuations of the field, given by the
710: fluctuation-dissipation theorem, AAK obtained
711: \be\label{AAK}
712: \smean{\Delta\sigma}=-2\frac{e^2D}{\pi\sw}
713: \int_0^\infty\D t\:\EXP{-\gamma t}
714:   \int_{x(0)=x}^{x(t)=x}{\cal D}x(\tau)\,
715:   \EXP{-\frac1{4D}\int_0^t\D\tau\,\dot x(\tau)^2
716:        -\frac{e^2T}{\sigma_0\sw}\int_0^t\D\tau\,|x(\tau)-x(t-\tau)|}
717: \:,\ee
718: where $T$ is the temperature (the Planck and Boltzmann constants are 
719: set equal to unity $\hbar=k_B=1$), and $\sw$ is the area of the section of the 
720: quasi-1d wire. 
721: In contrast with eq.~(\ref{WL}) where the loss of phase coherence was
722: described phenomenologically by an exponential damping\footnote{
723:   Note that the introduction of an exponential damping describes rigorously 
724:   several effects~: the loss of phase coherence due to spin-orbit 
725:   scattering or spin-flip \cite{HikLarNag80}, the penetration of a
726:   weak perpendicular magnetic field in a quasi 1d wire \cite{AltAro81}.
727:   In this latter case the effect of the magnetic field is taken into 
728:   account through $\gamma=\frac13(e{\cal B}W)^2$ where $W$ is the width of the 
729:   wire of rectangular section.
730: } with the parameter
731: $\tau_\varphi$, here the electron-electron interaction affects the weak
732: localization correction through the introduction in the action of the 
733: functional of the Brownian bridge\footnote{
734:   A Brownian bridge, $(x(\tau),\,0\leq\tau\leq1\,|\,x(0)=x(1))$, is a
735:   Brownian path conditioned to return to its starting point.
736: }
737: \be\label{Atilde}
738: \widetilde{\cal A}_t=\int_0^t\D\tau\,|x(\tau)-x(t-\tau)|
739: \:.\ee
740: The additional damping $\exp{-\gamma t}$ in eq.~(\ref{AAK}) describes the 
741: loss of phase coherence due to other phase breaking mechanisms.
742: By using a trick which makes the path integral local in time, AAK have 
743: computed explicitely the path integral and obtained~:
744: \be\label{AAKres}
745: \smean{\Delta\sigma}=\frac{e^2}{\pi\sw}\,L_N\,
746: \frac{{\rm Ai}(\gamma\tau_N)}{{\rm Ai}'(\gamma\tau_N)}
747: \:,\ee
748: where ${\rm Ai}(z)$ is the Airy function.
749: The Nyquist time $\tau_N=(\frac{\sigma_0\sw}{e^2 T\sqrt{D}})^{2/3}$, gives the
750: time scale over which electron-electron interaction is effective and therefore
751: plays the role of a phase coherence time. We have also introduced the
752: corresponding length $L_N=\sqrt{D\tau_N}$. 
753: Note that the $T$ dependence of $\tau_N\propto T^{-2/3}$ directly reflects the
754: scaling of the area with time~:  
755: $\widetilde{\cal A}_t\eqlaw t^{3/2}\widetilde{\cal A}_1$. We stress that the 
756: AAK theory makes a quantitative prediction for the dependence of the phase
757: coherence length $L_N$ as a function of the temperature, which has been
758: verified experimentally for a wide range of parameters~: on metallic (Gold)
759: wires \cite{EchGerBozBogNil93b,PieGouAntPotEstBir03} 
760: (the behaviour of $L_N\propto(\sw/T)^{1/3}$ was
761: observed before in Aluminium and Silver wires \cite{WinRooChaPro86}), and on
762: wires etched at the interface of two semiconductors \cite{ThoPepAhmAndDav86}.
763: 
764: 
765: 
766: \begin{figure}[!ht]
767: \begin{center}
768: \includegraphics[scale=0.9]{bridge.eps}
769: \hspace{1cm}
770: \includegraphics[scale=0.9]{bridgesym.eps}
771: \end{center}
772: \caption{
773:          Left~: {\it The functional ${\cal A}_t$ gives the absolute area 
774:          below a Brownian bridge $x(\tau)$.}
775:          Right~: {\it  The functional $\widetilde{\cal A}_t$ measures
776:          the area between the Brownian bridge $x(\tau)$ and its 
777:          time reversed counterpart $x(t-\tau)$.
778:          The two functionals are equal in law [eq.~(\ref{AAtilde})].}
779:          \label{fig:bridge}}
780: \end{figure}
781: 
782: It is interesting to point out that the result of AAK can be interpreted as 
783: the Laplace transform of the distribution of the functional (\ref{Atilde}), 
784: ${\rm E}[\exp{-p \widetilde{\cal A}_t}]$. This functional represents the area
785: between a Brownian bridge and its time reversed  counterpart ({\it cf.}
786: figure~\ref{fig:bridge}b), where ${\rm E}[\cdots]$ describes averaging over
787: Brownian bridges. The conjugate parameter $p$ is played in (\ref{AAK}) by the
788: temperature~:  $p=\frac{e^2T}{\sigma_0\sw}$.
789: The result (\ref{AAKres}) has also been derived in the probability 
790: literature. 
791: Let us consider the functional
792: \be\label{A}
793: {\cal A}_t=\int_0^t\D\tau\,|x(\tau)|
794: \:,\ee
795: giving the absolute area below the Brownian bridge starting from the origin,
796: $x(0)=x(t)=0$ ({\it cf.} figure~\ref{fig:bridge}a). The double Laplace
797: transform of the distribution of ${\cal A}_t$ has been computed first by
798: Cifarelli and Regazzini \cite{CifReg75} in the context of economy and
799: independently by Shepp \cite{She82}, and led to
800: $
801: \int_0^\infty\frac{\D t}{\sqrt{t}}\EXP{-\gamma t}
802: {\rm E}[\EXP{-\sqrt{2}{\cal A}_t}]
803: =-\sqrt\pi\,\frac{{\rm Ai}(\gamma)}{{\rm Ai}'(\gamma)}
804: $, 
805: which is equivalent\footnote{
806:   Note that the $1/\sqrt{t}$ in the integral computed by Shepp corrects the 
807:   fact that the averaging over Brownian curves in (\ref{AAK}) is not 
808:   normalized to unity, whereas ${\rm E}[\cdots]$ is.
809: } to the result (\ref{AAKres}).
810: The connection between the two results is clear from the equality in 
811: law\footnote{
812:   For a given process $x(\tau)$ the two functionals ${\cal A}_t$ and 
813:   $\widetilde{\cal A}_t$ are obviously different, however they are 
814:   distributed according to the same  probability distribution. 
815:   They are said ``equal in law''.
816:   \label{footnoteEL}
817: }~:
818: \be\label{AAtilde}
819: {\cal A}_t \eqlaw\widetilde{\cal A}_t
820: \ee
821: which follows from\footnote{
822:   The relation (\ref{anicerel}) is easily proved by using that a Brownian
823:   bridge $(x(\tau),\,0\leq\tau\leq1\,|\,x(0)=x(1)=0)$ can be written in terms 
824:   of a free Brownian motion  $(B(\tau),\,\tau\geq0\,|\,B(0)=0)$ as~:
825:   $x(\tau)=B(\tau)-\tau\,B(1)$.
826: }
827: \begin{equation}
828:  \label{anicerel}
829:  x(\tau)-x(t-\tau) \eqlaw x(2\tau) 
830:  \hspace{0.5cm} \mbox{ for } \hspace{0.5cm}
831:  \tau\in[0,t/2]
832:  \:.
833: \end{equation}
834: The relation (\ref{AAtilde}) allows us to understand more deeply the trick 
835: used by AAK to make the path integral (\ref{AAK}) local in time.
836: The distribution of the area ${\cal A}_t$ has been also studied by Rice in 
837: ref.~\cite{Ric82}. 
838: Interestingly, the inverse Laplace transform of eq.~(\ref{AAKres}) obtained by
839: Rice has been rederived recently independently in ref.~\cite{MonAkk05} in 
840: order to analyze
841: the loss of phase coherence due to electron-electron interaction in a time
842: representation.
843: 
844: The study of statistical properties of the absolute area below a Brownian
845: motion was first addressed by Kac \cite{Kac46} for a free Brownian motion,
846: long before the case of the Brownian bridge studied by Cifarelli \&
847: Regazzini and by Shepp. Later on, it has been extended to other
848: functionals of excursions and meanders\footnote{
849:   An excursion is a part of a Brownian path between two consecutive zeros
850:   and a meander is the part of the path after the last zero.
851: %  A Brownian excursion~: 
852: %  $(e(\tau),\,0\leq\tau\leq1\,|\,e(\tau)>0,\,e(0)=e(1)=\varepsilon)$.
853: %  A Brownian meander~:
854: %  $(m(\tau),\,0\leq\tau\leq1\,|\,m(\tau)>0,\,m(0)=\varepsilon)$.
855: %  Where $\varepsilon\to0$ is a regulator.
856: } \cite{JeaPitYor97,PerWel96}, 
857: which arise in a number of seemingly unrelated problems such as computer 
858: science, graph theory \cite{FlaLou01}  and statistical physics
859: \cite{MajCom04,MajCom05}.
860: 
861: \vspace{0.25cm}
862: 
863: 
864: \noindent{\it Dephasing in a ring.--}
865: Very recently, the question of dephasing due to electron-electron interaction
866: in a weakly disordered metal has been raised again in ref.~\cite{LudMir04}. In 
867: particular it has been shown that the effect of the geometry
868: of the ring and the effect of electron-electron interaction combine 
869: in a nontrivial way leading to a behaviour of the 
870: harmonics of the magnetoconductance that differs from the one predicted by
871: eq.~(\ref{WL}) in ref.~\cite{AltAroSpi81}.
872: For the ring, the functional describing the effect of electron-electron 
873: interaction on weak localization is now given by
874: \begin{equation}
875:   \label{functring}
876:   {\cal R}_t = \int_0^t\D\tau\,
877:   |x(\tau)|\, \left( 1 - \frac{|x(\tau)|}{L} \right) 
878: \end{equation}
879: instead of the area defined by eq.~(\ref{A}). 
880: $x(\tau)$ is a Brownian path\footnote{
881:   As for the case of the wire, the functional involved in the study of
882:   dephasing in the ring ~\cite{TexMon05b} involves the difference
883:   $x(\tau)-x(t-\tau)$ instead of the bridge $x(\tau)$. We have used
884:   the equality in law (\ref{anicerel}) which implies the following property~:
885:   given a Brownian bridge $(x(\tau),\,0\leq\tau\leq{t}\,|\,x(0)=x(t)=0)$,
886:   for any even function $f(x)$ we have
887:   $\int_0^t\D\tau\,f(x(\tau)-x(t-\tau))\eqlaw\int_0^t\D\tau\,f(x(\tau))$.
888: } on a ring of perimeter $L$ such that $x(0)=x(t)=0$.
889: The question has been re-examined in more detail in ref.~\cite{TexMon05b},
890: where the double Laplace transform of the distribution
891: $
892: \int_0^\infty\D t\,\EXP{-\gamma t}\,
893: \frac{1}{\sqrt{t}}\EXP{-\frac{(nL)^2}{2t}}\,{\rm E}_n[\EXP{-{\cal R}_t}]
894: $
895: has been derived. ${\rm E}_n[\cdots]$ denotes averaging over 
896: Brownian bridges defined on a circle, with winding number $n$. Note that the 
897: Laplace transform of the distribution ${\rm E}_n[\EXP{-{\cal R}_t}]$ 
898: has been also studied in ref.~\cite{TexMon05b}.
899: 
900: 
901: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
902: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
903: 
904: \section{Exponential functionals of Brownian motion and Wigner time delay
905:          \label{sec:expfun}}
906: 
907: \subsection{Historical perspective}
908: 
909: Exponential functionals of the form,
910: \begin{equation}
911: \label{deffe}
912: A_t^{(\mu)}= \int_0^t \D s\, \exp{-2(B(s) +\mu s)}
913: \end{equation}
914: where $(B(s),\,s\geq0,\,B(0)=0)$ is an ordinary Brownian motion have been the 
915: object of many studies in mathematics \cite{Yor00}, mathematical finance and 
916: physics \cite{ComMonYor98}. 
917: In the physics of classical disordered systems, the starting point was the
918: analysis of the series (Kesten variable)
919: \be
920: Z= z_1 +z_1z_2+z_1z_2z_3+\cdots
921: \ee
922: where the $z_i$ are independent and identical random variables.
923: Several papers have been devoted to the study of this random variable in the
924: physics \cite{CalLucNiePet85,DerHil83} and mathematics \cite{Kes73,Ver79}
925: literature.
926: It was  realized that $Z$ is the discretized version of $ A_\infty^{(\mu)}$
927: defined in eq.~(\ref{deffe}),
928: which may be interpreted as a trapping time in the context of classical
929: diffusion in a random medium. The tail of the
930: probability distribution $P(Z)$ controls the anomalous diffusive behaviour of a
931: particle moving in a one-dimensional random force field \cite{BouComGeoLeD90}. 
932: The functional
933: $A_t^{(\mu)}$ also arises in the study of the transport properties of
934: disordered samples of finite length \cite{MonCom94,OshMogMor93}.
935: In the context of one dimensional localization the fact that the norm of the
936: wave function is distributed\footnote{
937:   When $\mu>0$, $A_t^{(\mu)}$ possesses a limit distribution for $t\to\infty$.
938:   Moreover, we have the equality in law (see footnote \ref{footnoteEL})~:
939:   $A_\infty^{(\mu)}\eqlaw 1/\gamma^{(\mu)}$, where 
940:   $\gamma^{(\mu)}$ is distributed according to a $\Gamma$-law~:
941:   $P(\gamma)=\frac1{\Gamma(\mu)}\gamma^{\mu-1}\EXP{-\gamma}$.
942:   \label{foot:gammalaw}
943: } as $ A_\infty^{(1)}$ is mentioned in the book of Lifshitz {\it et al}
944: \cite{LifGrePas88}.
945: 
946: 
947: 
948: \subsection{Wigner time delay\label{sec:wtd}}
949: 
950: As discussed in the introduction, the localization of quantum states in one
951: dimension is well understood. However since real systems are not infinite,
952: asking about the nature of the states which are not in the bulk is a perfectly
953: legitimate question. It was pointed out by Azbel in ref.~\cite{Azb83} 
954: that, when a disordered region is connected to a region free of disorder,
955: the localized states acquire a finite lifetime which
956: shows up in transport through sharp resonances,
957: refered as ``Azbel resonances''. This picture was used in
958: ref.~\cite{JayVijKum89} where it was argued that these resonances could be
959: probed in scattering experiments and lead to an energy dependent random time
960: delay of the incident electronic wave. Motivated by this result and by
961: developments in random matrix theory we consider the scattering problem for
962: the one dimensional Schr\"odinger equation (in units $\hbar=2m=1$)
963: \be
964: -\frac{\D^2}{\D x^2}\psi(x)+V(x)\psi(x)=E\,\psi(x)
965: \ee
966: defined on the half line $x\geq 0$  with the Dirichlet boundary conditions
967: $\psi(0)=0$.
968: The potential $V(x)$ has its support on the interval $[0,L]$.
969: Outside of this interval, the scattering state of energy $E=k^2$ is given by
970: \be\label{scattstate}
971:   \psi_E(x)=\frac1{\sqrt{hv_E}}
972:   \left( \EXP{-\I k(x-L)} +\EXP{\I k(x-L)+\I\delta} \right)
973: \:,\ee 
974: where $h=2\pi$ is the Planck constant (in unit $\hbar=1$) and 
975: $v_E=\D E/\D k=2k$ the group velocity.
976: Eq.~(\ref{scattstate})
977: represents the superposition of an incoming plane wave incident from
978: the right and a reflected
979: plane wave caracterized by its phase shift $\delta(E)$. 
980: The Wigner time delay, defined by the relation $\tau=\D\delta/\D{E}$,
981: can be understood as the time spent by the wave packet of energy $E$ 
982: in the disordered region (this interpretation is only valid 
983: at high energy~; see 
984: refs.~\cite{HauSto89,But90,LanMar94,CarNus02} for review articles on time 
985: delay and traversal times).
986: This representation of the time delay has been used in many papers both
987: in the mathematics and physics literature. Starting from this representation
988: one can derive a system of stochastic differential equations which can be
989: studied in certain limiting cases. However, to understand the universality 
990: of the statistical properties of the Wigner time in the high energy limit,
991: we follow a different approach below.
992: 
993: 
994: 
995: 
996: \subsubsection*{Universality of time delay distribution at high energy}
997: 
998: Using the expression of the scattering state (\ref{scattstate}) and the
999: definition $\tau=\D\delta/\D{E}$, we can obtain the so-called Smith formula
1000: \cite{Fri58,Smi60} relating the time delay to the wave function in the bulk~:
1001: \be\label{Smith}
1002:   \tau(E)= 2\pi\int_0^L \D x\, |\psi_E(x)|^2 
1003:   -\frac{1}{2E}\sin\delta(E)
1004: \:.\ee
1005: Following
1006: ref.~\cite{AntPasSly81} one can parametrize the wave function in the bulk in
1007: terms of its phase and modulus. 
1008: For this purpose we rewrite the Schr\"odinger equation as a set of two 
1009: coupled first order differential equations, $\D\psi_E/\D x=\psi'_E$
1010: and $\D\psi'_E/\D x=(V(x)-E)\psi_E$, and perform the change of variables~:
1011: $\psi_E(x)=\EXP{\xi(x)}\sin\theta(x)$ and 
1012: $\psi'_E(x)=k\EXP{\xi(x)}\cos\theta(x)$. One obtains a new set of 
1013: first order differential equations $\D\theta/\D x=k-\frac1kV(x)\sin^2\theta$
1014: and $\D\xi/\D x=\frac1{2k}V(x)\sin2\theta$. Up till now everything is exact. 
1015: If we now consider the high energy limit\footnote{
1016:   To define precisely the high energy limit, let us introduce the integral
1017:   of the correlation function of the disorder
1018:   ${w}=\int\D x\,\smean{V(x)V(0)}$. The high energy limit corresponds to
1019:   $k\gg{w}^{1/3}$.
1020: }
1021: we can neglect the second term on the {\it r.h.s} of eq.~(\ref{Smith})
1022: and integrate out the phase which is a fast variable~; we obtain
1023: \be\label{timedelay0}
1024: \tau = \frac{1}{k}\int_0^L \D x \:\EXP{2(\xi(x)-\xi(0))}
1025: \:.
1026: \ee
1027: This representation of the time delay holds for any realization of the
1028: disordered potential. Moreover, one can prove under rather mild conditions
1029: on the correlations of the random potential
1030: that $\xi(x)$ is a Brownian motion with drift 
1031: $\xi(x)=x/\lambda+\sqrt{1/\lambda}\,B(x)$ where
1032: $B(x)$ is an ordinary Brownian motion and $\lambda$ the localization length. 
1033: Using the scaling properties of the Brownian motion gives the following
1034: identity in law
1035: \be\label{timedelay}
1036: \tau \eqlaw \frac{\lambda}{k}\int_0^{L/\lambda} \D x\: \EXP{-2(B(x)+x)}
1037: = \frac{\lambda}{k}\, A_{L/\lambda}^{(1)}
1038: \:,
1039: \ee
1040: where $A_L^{(1)}$ has been defined in eq.~(\ref{deffe}).
1041: This representation of the time delay as an exponential functional of the
1042: Brownian motion, first established in ref.~\cite{FarTsa94} by a different
1043: method, allows us to obtain a number of interesting results
1044: \cite{ComTex97,TexCom99}~:
1045: 
1046: \vspace{0.25cm}
1047: 
1048: \noindent
1049: ({\it i}) Existence of a limit distribution for fixed\footnote{
1050:   From the remark of the footnote \ref{foot:gammalaw}, we notice that 
1051:   $1/\tau$ is distributed according to an exponential law.
1052: } $\tau$ and $ L\rightarrow\infty$
1053: \be
1054: P(\tau)=\frac{\lambda}{2k\tau^2}\, \EXP{- \frac{\lambda}{2k\tau}} 
1055: \:.\ee
1056: This result is reminiscent of the random matrix theory prediction in
1057: spite of the fact that this theory does not apply to systems that are
1058: strictly one-dimensional\footnote{
1059:   The random matrix theory describes the regime of weak localization or 
1060:   systems whose classical dynamics is chaotic. The distribution of 
1061:   time delay in this framework has the same functional form 
1062:   $P(\tau)\propto\frac1{\tau^{2+\mu}}\EXP{-\tau_0/\tau}$ 
1063:   \cite{FyoSom96a,GopMelBut96,BroFraBee97,OssFyo05} with two differences~: 
1064:   ({\it i}) the exponent $2+\mu>2$,
1065:   ({\it ii}) the time scale $\tau_0$.
1066: }. 
1067: 
1068: \vspace{0.25cm}
1069: 
1070: \noindent
1071: ({\it ii}) Linear divergence of the first moment, $\smean{\tau}=L/k$, and 
1072: exponential divergence of the higher moments 
1073: $\smean{\tau^n}\propto\EXP{2n(n-1)L/\lambda}$.
1074: This divergence reflects a log-normal tail of the distribution for a finite
1075: length $L$.
1076: 
1077: \subsubsection*{Time delay and density of states}
1078: 
1079: The relation (\ref{Smith}) also provides another interpretation to these 
1080: results. In the situation under consideration, the scattering state is 
1081: directly related to the local density of states (LDoS) by 
1082: $\rho(x;E)=\bra{x}\delta(E-H)\ket{x}=|\psi_E(x)|^2$. Therefore the time delay
1083: can be interpreted as the DoS of the disordered region\footnote{
1084:   The relation between time delay and DoS is sometimes refered as Krein-Friedel
1085:   relation \cite{Fri52,Kre53}. This relation has been originally introduced in
1086:   ref.~\cite{BetUhl37} in the context of statistical physics 
1087:   (see also ref.~\cite{DasMaBer69} and \S77 of ref.~\cite{LanLif66e}). 
1088:   Note that it has been recently
1089:   discussed in the context of graphs in refs.~\cite{Tex02,TexBut03,TexDeg03}.
1090: }. 
1091: This establishes a relation between the results given above and the work of
1092: Al'tshuler \& Prigodin~\cite{AltPri89} where the distribution of the LDoS
1093: was studied for a white noise potential using the method of
1094: Berezinski{\u\i}\footnote{
1095:   The method of Berezinski{\u\i} blocks has been introduced to study
1096:   specifically the case of white noise disordered potential in 
1097:   one dimension \cite{Ber74}.
1098:   This powerful approach has been widely used and has allowed to derive
1099:   numbers of important results like in refs.~\cite{GogMelRas76,GorDorPri83} 
1100:   for example.
1101: }.
1102: 
1103: 
1104: \subsubsection*{Time delay for Dirac Hamiltonian at the threshold energy}
1105: 
1106: The representation (\ref{timedelay}) of the time delay as an exponential
1107: functional of the Brownian motion only holds in the weak disorder ({\it i.e.}
1108: high energy) limit. A similar representation can be derived for the random
1109: mass Dirac model at the middle of the spectrum. 
1110: The Dirac Hamiltonian~is 
1111: \be\label{HDirac}
1112: H_{\rm D}=\sigma_2\,\I\frac{\D}{\D x} + \sigma_1\phi(x)
1113: \:,\ee
1114: where $\sigma_i$ are the Pauli matrices. $\phi(x)$ can be interpreted as a
1115: mass\footnote{
1116:   The Hamiltonian (\ref{HDirac}) with $\phi(x)$ a white noise was introduced
1117:   by Ovchinnikov \& Erikmann \cite{OvcEri77} as a model of 
1118:   one-dimensional semiconductor with a narrow fluctuating gap.
1119:   It is interesting to point out that the Dirac equation $H_{\rm D}\psi=k\psi$ 
1120:   also has an interpretation in the context of superconductivity, as 
1121:   linearized Bogoliubov-de~Gennes equations for a real random superconducting
1122:   gap $\phi(x)$.
1123:   Finally it is worth mentioning that more general 1d Dirac Hamiltonians 
1124:   with several kinds of disorder (mass term, potential term and magnetic 
1125:   field) have been studied in refs.~\cite{Boc99,Boc00}. 
1126: }. 
1127: The Dirac equation $H_{\rm D}\psi=k\,\psi$ possesses a
1128: particle-hole symmetry reflected in the symmetry of the spectrum with respect
1129: to $k=0$ (the dispersion relation of the Dirac equation is linear 
1130: in absence of the mass term and energy is equal to momentum).
1131: The middle of the spectrum is an interesting point where the divergence in the
1132: localization length signals the existence of a delocalized state
1133: (see footnote \ref{footnoteSUSY}). These
1134: properties should therefore show up in the probability distribution of the 
1135: Wigner time delay. 
1136: Since the dispersion relation of the Dirac equation is linear, the time 
1137: delay is now defined as $\tau=\D\delta/\D k$.
1138: The following representation has been derived 
1139: in ref.~\cite{SteCheFabGog99}~:
1140: \be \label{tdsusy}
1141: \tau= 2\int_0^L \D x\, \EXP{2\int_0^x \phi(y)\D y}
1142: \:.\ee
1143: In contrast with
1144: eqs.~(\ref{timedelay0},\ref{timedelay}), which have been obtained after
1145: averaging over the fast phase variable, the representation (\ref{tdsusy}) is
1146: exact (for $k=0$). If $\phi(x)$ is a white noise, then
1147: \begin{equation}
1148: \label{ChenTD}
1149: \tau = \frac{2}{g}\,A_{g L}^{(0)}
1150: \:.\end{equation}
1151: In this case since the drift vanishes, it is known \cite{MonCom94,Yor00} that
1152: there is no limiting distribution when $L\rightarrow\infty$. The physical
1153: meaning of the lognormal tail
1154: \be
1155: P(\tau)\sim \frac{1}{2\tau\sqrt{2\pi gL}}\, \EXP{-\frac{1}{8gL} \ln^2(g\tau)}
1156: \ee
1157: in the limit $\tau\rightarrow\infty$ is discussed in detail in
1158: ref.~\cite{SteCheFabGog99}.
1159: 
1160: 
1161: 
1162: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1163: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1164: 
1165: \section{Extreme value spectral statistics\label{sec:evss}}
1166: 
1167: The density of states (DoS) and the localization length of a one-dimensional
1168: Hamiltonian with a random potential can be studied by the phase formalism
1169: \cite{AntPasSly81,LifGrePas88} (see also ref.~\cite{Luc92} for discrete
1170: models).
1171: In the previous section we pointed out
1172: that the physics of the strong localization in one-dimension can also be
1173: probed from a different angle by considering the scattering of a plane wave by
1174: the random potential. This has led to the study of the Wigner time delay. In
1175: this section we show that the phase formalism also allows describing finer
1176: properties of the spectrum such as the probability distribution of the $n$-th
1177: eigenvalue. This is an instance of an {\it extreme value statistics} ~: given
1178: a ranked sequence of ${\cal N}$ random variables
1179: $x_1 \leq x_2 \leq \ldots\leq x_{\cal N}$, 
1180: the problem is to find the distribution of the $n$-th of these variables in a
1181: given interval. In the particular case of uncorrelated random variables, the
1182: extreme value distributions have been studied by E.~Gumbel
1183: \cite{Gum35,Gum54,Gum58}. This problem becomes much more complicated when the
1184: random variables are correlated, a case which has recently attracted a lot of
1185: attention. Such extreme value problems have appeared recently in a variety of
1186: problems ranging from disordered systems \cite{BouMez97,CarLeD01,DeaMaj01} to
1187: certain computer science problems such as growing search trees \cite{MajKra03}.
1188: In the case of the spectrum of a random Hamiltonian, the eigenvalues are in
1189: general correlated variables, apart from the case of strongly localized
1190: eigenstates \cite{Mol81}. Below we show that, in the one dimensional case,
1191: this problem is related to studying the first exit time distribution of a 
1192: one-dimensional diffusion process.
1193: 
1194: \mathversion{bold}
1195: \subsection{Distribution of the $n$-th eigenvalue~: relation with a first 
1196:             exit time distribution}
1197: \mathversion{normal}
1198: 
1199: We consider a Schr\"odinger equation $H\varphi(x)=E\varphi(x)$ on a finite
1200: interval $[0,L]$. The spectral (Sturm-Liouville) problem is further defined by
1201: imposing suitable boundary conditions. We choose the Dirichlet boundary
1202: conditions $\varphi(0)=\varphi(L)=0$. The spectrum of $H$ is denoted by 
1203: ${\rm Spec}(H)=\{E_0<E_1<E_2<\cdots\}$. Our purpose is to compute the
1204: probability 
1205: \be
1206: W_n(E) = \smean{\delta(E-E_n)}
1207: \ee
1208: for the eigenvalue $E_n$ to be at energy $E$ (the bracket $\smean{\cdots}$
1209: means averaging over the random potential).
1210: Note that the sum of these distributions 
1211: $\frac1L\sum_{n}W_n(E)=\rho(E)$ is the average DoS per unit length.
1212: We now show how the calculation of $W_n(E)$ can be cast into  a first exit time
1213: problem.
1214: 
1215: \subsubsection*{Random Schr\"odinger Hamiltonian}
1216: 
1217: We consider the Hamiltonian 
1218: \be\label{hamil}
1219: H=-\frac{\D^2}{\D x^2} + V(x)
1220: \:,\ee
1221: where $V(x)$ is a Gaussian white noise random potential~: $\smean{V(x)}=0$ and
1222: $\smean{V(x)V(x')}={w}\,\delta(x-x')$. 
1223: 
1224: We replace the Sturm-Liouville problem by a Cauchy problem~:
1225: let  $\psi(x;E)$ be the solution of the Schr\"odinger equation
1226: $H\psi(x;E) = E\psi(x;E)$ with the boundary conditions $\psi(0;E)=0$
1227: and $\frac{\D}{\D x}\psi(0;E)=1$. 
1228: %It follows that $\psi(x;E)$ is real.
1229: The boundary condition $\psi(L;E)=0$ is fulfilled whenever the energy $E$ 
1230: coincides with an eigenvalue $E_n$ of the Hamiltonian. In this case,
1231: the wave function 
1232: $\varphi_n(x)=\psi(x;E_n)/\left[\int_0^L\D{x'}\,\psi(x';E_n)^2\right]^{1/2}$ 
1233: has $n$ nodes in the interval $]0,L[$, and two nodes at the boundaries. 
1234: 
1235: \begin{figure}[!ht]
1236: \hspace{3cm}
1237: \begin{tabular}{ll}
1238: \underline{$E<E_0$}           &  \diagram{psi1b.eps}{1}{-0.75cm}  
1239: \\
1240: \underline{$E=E_0$}           &  \diagram{psi2b.eps}{1}{-0.75cm}  
1241: \\
1242: \underline{$E_0<E<E_1$}       &  \diagram{psi3b.eps}{1}{-1.40cm}  
1243: \\[0.5cm]
1244: \hspace{1.5cm}{\huge$\vdots$} & \hspace{1cm}{\huge$\vdots$\hspace{2cm}$\vdots$}
1245: \end{tabular}
1246: \caption{\it
1247:          The probability for the $n$-th level $E_n$ of the spectrum to be 
1248:          at $E$ is also the probability for the $n+1$-th node of $\psi(x;E)$ 
1249:          to be at $L$.\label{fig:psideE}}
1250: \end{figure}
1251: 
1252: Let us denote by $\ell_m$ ($m\geq1$), the length between two consecutive nodes.
1253: We consider the Ricatti variable
1254: \be
1255: z(x;E) = \frac{\D}{\D x} \ln|\psi(x;E)|
1256: \:,\ee
1257: which obeys the following equation:
1258: \be\label{Riceq}
1259: \frac{\D}{\D x} z = -E - z^2 + V(x)
1260: \:,\ee
1261: with initial condition $z(0;E)=+\infty$.
1262: This equation may be viewed as a Langevin equation for a particle located at
1263: $z$ submitted to a force $-\partial U(z)/\partial z$ deriving from the 
1264: unbounded potential 
1265: \be\label{potU}
1266: U(z)=Ez+\frac{z^3}{3}
1267: \ee 
1268: and to a random white noise $V(x)$. 
1269: 
1270: Each node of the wave function corresponds to $|z(x)|=\infty$. At ``time''
1271: $x=0$ the ``particle'' starts from $z(0)=+\infty$ and eventually ends at
1272: $z(\ell_1-0^+)=-\infty$ after a ``time'' $\ell_1$. Just after the first node
1273: it then starts again from  $z(\ell_1+0^+)=+\infty$, due to the continuity of
1274: the wave function. It follows from this picture that the distance
1275: $\ell_m$ between two consecutive nodes may be viewed as the ``time'' needed
1276: by the particle to go through the interval $]-\infty,+\infty[$ (the
1277: ``particle'' is emitted from $z=+\infty$ at initial ``time'' and absorbed
1278: when it reaches $z=-\infty$).
1279: The distances $\ell_m$ are random variables, interpreted as times needed
1280: by the process $z$ to go from $+\infty$ to $-\infty$.
1281: These random variables are statistically independent because each time the
1282: variable $z$ reaches $-\infty$, it loses the memory of
1283: its earlier history since it is brought back to the same initial condition
1284: and $V(x)$ is $\delta$-correlated. This remark is a crucial point for the
1285: derivation of $W_n(E)$.
1286: Interest in these random ``times'' $\ell_m$ lies in their relation
1287: with the distribution of the eigenvalues. Indeed the probability that
1288: the energy $E_n$ of the $n$-th excited state is at $E$ is also
1289: the probability that the sum of the $n+1$ distances between the nodes is
1290: equal to the length of the system: $L=\sum_{m=1}^{n+1}\ell_m$
1291: (this is illustrated on figure~\ref{fig:psideE}).
1292: Since the $\ell_m$ are independent and identically distributed
1293: random variables   
1294: $\proba\left[L=\sum_{m=1}^{n+1}\ell_m\right]$ is readily obtained from the
1295: distribution $P(\ell)$ of one of these variables. 
1296: 
1297: We introduce the intermediate variable ${\cal L}(z)$ giving the time needed by
1298: the process starting at $z$ to reach $-\infty$~: therefore we have 
1299: $\ell={\cal L}(+\infty)$. The Laplace transform of the distribution
1300: $
1301: h(\alpha,z) = 
1302: \smean{ \EXP{-\alpha{\cal L}} \:|\: z(0)=z;\:z({\cal L})=-\infty }
1303: $ 
1304: obeys \cite{Gar89}~:
1305: \be
1306: G_z \, h(\alpha,z) = \alpha \, h(\alpha,z)
1307: \:,\ee
1308: where the backward Fokker-Planck generator is
1309: \be\label{bfpegen}
1310: G_z = -U'(z) \partial_z + \frac{{w}}{2} \partial_z^2
1311: \:.\ee
1312: The boundary conditions are
1313: $\partial_z h(\alpha,z)|_{z=+\infty}=0$ and $h(\alpha,-\infty)=1$.
1314: By Laplace inversion we can derive $P(\ell)$ since
1315: $h(\alpha,+\infty)=\int_0^\infty\D\ell\,P(\ell)\EXP{-\alpha\ell}$.
1316: 
1317: The solution of this problem is given in ref.~\cite{Tex00}.
1318: Here we only consider the limit $E\to-\infty$, which corresponds to 
1319: the bottom of the spectrum. 
1320: In this regime the dynamics of the Ricatti variable $z$ can be read off 
1321: from the shape of the potential (see figure~\ref{fig:potUneg}).
1322: 
1323: \begin{figure}[!ht]
1324: \begin{center}
1325: \includegraphics{potUneg.eps}
1326: \end{center}
1327: \caption{\it The potential (\ref{potU}) related to the deterministic force 
1328:          to which the Ricatti variable is submitted in eq.~(\ref{Riceq}).
1329:          \label{fig:potUneg}}
1330: \end{figure}
1331: 
1332: The particle falls rapidly in the well from which it can only escape by a
1333: fluctuation of the random force. Therefore the time needed to go from
1334: $+\infty$ to $-\infty$ is dominated by the time spent in the well, which is
1335: distributed according to the Arrhenius formula. It follows that the average
1336: ``time'', which is related to the inverse of the integrated density of states
1337: (IDoS) per unit length $N(E)=\int_0^E\D E'\,\rho(E')$, behaves as
1338: $
1339: \mean{\ell} = N(E)^{-1}
1340: \simeq \frac{\pi}{\sqrt{-E}} \exp\frac{8}{3{w}}(-E)^{3/2}
1341: $.
1342: This simple picture, first provided by Jona-Lasinio \cite{Jon83} allows
1343: recovering the exponential tail obtained by several methods in 
1344: refs.~\cite{FriLlo60,Hal65,AntPasSly81,LifGrePas88,ItzDro89}\footnote{
1345:   It is interesting to quote ref.~\cite{KucSad98} where it was shown that 
1346:   this non perturbative exponential tail can be obtained by a simple counting
1347:   method of ``skeleton'' diagrams assuming that they all have the same value
1348:   (this approach was also applied to the 3d case).
1349: }.
1350: The distribution of the length $\ell$ is a Poisson law (see \cite{Gar89}
1351: or appendix of ref.~\cite{Tex00}):
1352: \be\label{dol-}
1353: P(\ell) = N(E)\,\exp{-\ell\,N(E)}
1354: \:.\ee
1355: Using this result we can show that \cite{Tex00} 
1356: \be\label{TheResult}
1357: W_n(E) = L\rho(E) \, \frac{(L\,N(E))^n}{n!} \, \EXP{-L\,N(E)} 
1358: \:.\ee
1359: This result has a clear meaning~: $L\rho(E)$ gives the probability 
1360: to find any level at $E$ and the factor $\frac{x^n}{n!}\EXP{-x}$ ``compels''
1361: the number of states below $E$, $x=LN(E)$, to be close to $n$.
1362: We may go further and write
1363: \be
1364: W_n(E) = \frac{1}{\delta E_n} \ 
1365:          \omega_n\left(\frac{E - E_n^{\typ}}{\delta E_n}\right)
1366: \:,\ee
1367: where the typical value of the energy is 
1368: $E_n^{\typ}(L)=-(\frac{3{w}}{8}\ln\tilde L)^{2/3}$,
1369: while the scale of the fluctuations reads 
1370: $
1371: \delta E_n =\frac{ {w}^{2/3} }{2\sqrt{n+1}}(3\ln\tilde L)^{-1/3}
1372: $, 
1373: where $\tilde L=\frac{L\,{w}^{1/3}}{2\pi(n+1)}$.
1374: The function
1375: \be\label{TheResultbis}
1376:   \omega_n(X)=\frac{(n+1)^{n+\frac12}}{n!}
1377:   \exp{\left(\sqrt{n+1}\:X -(n+1)\,\EXP{{X}/{\sqrt{n+1}}}\right)}
1378: \:\ee
1379: has the form of a Gumbel law for {\it uncorrelated} random variables.
1380: The fact that the eigenvalues are uncorrelated is a consequence of the 
1381: strong localization of the eigenfunctions \cite{Mol81}.
1382: 
1383: The work summarized in the above paragraph, which appeared in
1384: ref.~\cite{Tex00}, generalizes the result of McKean \cite{McK94} for the
1385: ground state.
1386: 
1387: A similar result was obtained by Grenkova {\it et al} \cite{GreMolSud83} for
1388: the model of $\delta$-impurities with random positions, in the limit of low
1389: impurity density which has no counterpart in the model we considered 
1390: here\footnote{
1391:   See footnote \ref{FrischLloyd}.
1392: }.
1393: However both models describe the same physics of strongly localized
1394: eigenstates.
1395: 
1396: 
1397: 
1398: \subsubsection*{Supersymmetric random Hamiltonian}
1399: 
1400: Random Schr\"odinger operators have been investigated through a wide range of
1401: models. Depending on the physical context, there are indeed many ways to model 
1402: the
1403: disorder. In the previous paragraph we have assumed that the potential $V(x)$
1404: is a white noise. We have also mentioned that if the potential is a 
1405: superposition of $\delta$-potentials randomly distributed along the line,
1406: although the spectra of the two models are quite different\footnote{
1407:   The model of $\delta$-impurities with random positions was introduced 
1408:   and studied by Schmidt \cite{Sch57} but often refered as the 
1409:   Frisch \& Lloyd model \cite{FriLlo60}. Compared to the model for 
1410:   white noise potential, which is characterized by one parameter, the Frisch \&
1411:   Lloyd model is characterized by two parameters~: the 
1412:   strength of the $\delta$-potential, and their density.
1413:   In the limit of high density of impurities this model is equivalent to 
1414:   the white noise potential model. The limit of low density presents 
1415:   different spectral singularities (Lifshitz singularity and, for 
1416:   negative weight of $\delta$-potentials, an additional Halperin singularity
1417:   in the negative part of the spectrum).\label{FrischLloyd}
1418: }, 
1419: the extreme spectral statistics are the same. The localization properties and
1420: the statistics of the time delay at high energy are also similar for both
1421: models \cite{Tex99,TexCom99}. These two models belong to the same class of
1422: random Hamiltonians with a random scalar potential with short range
1423: correlations. They are both continuous versions of discrete
1424: tight binding models with on-site random potential. This is the case of
1425: so-called {\it diagonal disorder}, since the random potential appears on the
1426: diagonal matrix elements of the tight binding Hamiltonian in the basis of
1427: localized orbitals\footnote{
1428:   Note that the continuum limit of tight binding Hamiltonian with diagonal 
1429:   disorder can lead to different continous models. 
1430:   Let us consider the discrete model
1431:   $H_{i,j} = -\delta_{i,j+1}-\delta_{i,j-1}+\delta_{i,j}V_i$. 
1432:   For $V_i=0$ the spectrum is $E_{k}=-2\cos(k)$, with $k\in]-\pi,\pi]$.
1433:   (A) If the continuous limit is taken by considering the band 
1434:   edge ($k\sim0$), one is led to the continuous model (\ref{hamil}).
1435:   (B) If the band center is considered instead ($k\sim\pi/2$), the 
1436:   spectrum can be linearized and one is led to a Dirac Hamiltonian,
1437:   like in refs.~\cite{Kel64,OvcEri77}.
1438:   This point has been recently rediscussed in ref.~\cite{SchTit03}.
1439:   \label{footnoteDD}
1440: }.
1441: 
1442: Other interesting models can be constructed by introducing disorder in the
1443: hoppings, instead. We refer to such models as {\it off-diagonal disorder}. The
1444: Dirac Hamiltonian (\ref{HDirac}) introduced above provides a continuum limit
1445: of such a model\footnote{
1446:   The continuum limit of a tight binding Hamiltonian with random hoppings
1447:   has been discussed in ref.~\cite{TakLinMak80} (see also the review in 
1448:   ref.~\cite{Mon95}).
1449:   As pointed in footnote \ref{footnoteDD}, random Dirac Hamiltonians 
1450:   can also appear as continuum limit of the band center of a 
1451:   discrete models with diagonal disorder. 
1452: }.
1453: Dirac Hamiltonians appear naturally in several contexts of condensed matter
1454: physics. The existence of symmetries in the Dirac Hamiltonian\footnote{
1455:   A complete classification of symmetries of disordered Hamiltonians 
1456:   extending the famous Wigner-Dyson ensembles of random matrix theory has 
1457:   been provided in refs.~\cite{Ver94,Zir96,AltZir97}
1458:   (see also the recent review article \cite{HeiHucZir05}). 
1459: }  
1460: (particle-hole, chiral,...) can lead to interesting features in the presence
1461: of disorder, which has attracted some attention
1462: (see for example
1463: refs.~\cite{NerTsvWen94,AltZir97,BroMudFur00,BroMudFur03,BocSerZir00,Boc00,EvaKat03},
1464: the last of these references gives a brief overview).
1465: The square of the Dirac Hamiltonian, 
1466: $H_{\rm D}^2=-{\D_x^2} + \phi(x)^2 +\sigma_3 \phi'(x)$,
1467: is related to the pair of supersymmetric Schr\"odinger isospectral 
1468: Hamiltonians $H_\pm=-{\D_x^2} + \phi(x)^2 \pm \phi'(x)$. When 
1469: $\mean{\phi(x)}=0$ we may forget the sign and simply consider~: 
1470: \be\label{Hsusy}
1471: H_S=-\frac{\D^2}{\D x^2} + \phi(x)^2 + \phi'(x)
1472: \:.\ee
1473: Such Hamiltonians appear in a variety of problems ranging from
1474: the 1d classical diffusion in a random force \cite{BouComGeoLeD90,IglMon05},
1475: electronic structure of polyacetylene \cite{TakLinMak80}, 
1476: spin Peierls chains \cite{FabMel97,FabMel97b,SteFabGog98} and also in a 
1477: continuum limit of the random field XY model \cite{GurCha03}
1478: (see ref.~\cite{ComTex98} for a short review on supersymmetric disordered
1479: quantum mechanics). Spectral and localization properties 
1480: have been studied in detail when $\phi(x)$ is a white noise 
1481: \cite{OvcEri77,LifGrePas88,BouComGeoLeD87,BouComGeoLeD90} and also
1482: when $\phi(x)$ has a finite correlation length as in a random telegraph process
1483: \cite{ComDesMon95} (this last case has found some application in the context
1484: of spin Peierls chains).
1485: 
1486: In the high energy limit, the localization properties and the statistics of
1487: the time delay do not show any difference with respect to the case of diagonal
1488: disorder. However the low energy properties are quite different. This is
1489: easily understood by noticing that, due to its relation to the Dirac 
1490: Hamiltonian, the supersymmetric Hamiltonian can be
1491: factorized as $H_S=Q^\dagger Q$ where $Q=-\D_x+\phi(x)$ and
1492: $Q^\dagger=\D_x+\phi(x)$. In particular, such a structure enforces a positive
1493: spectrum~:  ${\rm Spec}(H_S)\subset\RR^+$. 
1494: When $\phi(x)$ is white noise of zero mean, $\smean{\phi(x)}=0$ and
1495: $\smean{\phi(x)\phi(x')}=g\,\delta(x-x')$, the case considered below, the
1496: DoS presents a logarithmic Dyson singularity~: the integrated density 
1497: of states (IDoS) reads
1498: $N(E)\sim1/\ln^2E$ \cite{OvcEri77,BouComGeoLeD87,BouComGeoLeD90}, 
1499: while the localization length diverges logarithmically\footnote{
1500:   A more complete picture of the properties of this model at $E=0$ is given in
1501:   several works~:
1502:   (A) moments and correlations of the zero mode are studied in
1503:   refs.~\cite{SheTsv98,ComTex98}. 
1504:   (B) Additionally to the statistical properties of the time delay 
1505:   , eq.~(\ref{ChenTD}),
1506:   (C) the distribution of the transmission probability is derived in
1507:   ref.~\cite{SteCheFabGog99}. In particular it was shown that the 
1508:   average transmission through an interval of length $L$ decreases like 
1509:   $\smean{T}\propto1/\sqrt{L}$, which is slower than in the diffusive regime.
1510:   (D) The existence of a finite conductivity was
1511:   demonstrated in ref.~\cite{GogMel77}.
1512:   \label{footnoteSUSY}
1513: } $\lambda(E)\sim\ln(1/E)$.
1514: 
1515: It is interesting to investigate the extreme value statistics of the spectrum
1516: in the low energy regime, where we expect properties quite different from the
1517: one obtained for the diagonal disorder. The derivation follows closely 
1518: that in the previous case, however the relevant 
1519: random processes and the approximations are 
1520: different. The first step is to decouple the Schr\"odinger equation
1521: $H_S\varphi(x)=k^2\varphi(x)$ into two first order differential equations 
1522: ({\it i.e.} go back to the Dirac equation)~:
1523: \bea
1524: Q^\dagger \chi(x) &=& k \varphi(x) \\
1525: Q \varphi(x)      &=& k \chi(x)
1526: \:.\eea
1527: Then we may use the phase formalism by introducing a phase variable
1528: and an envelope variable~: 
1529: $\varphi(x)=\EXP{\xi(x)} \sin\vartheta(x)$ and 
1530: $\chi(x)=-\EXP{\xi(x)} \cos\vartheta(x)$.
1531: The phase variable obeys a stochastic differential equation with 
1532: a noise multiplying a trigonometric function of the phase.
1533: For convenience we introduce an additive process $\zeta(x)$ defined
1534: as $\zeta(x)=\pm\frac12\ln|\tan\vartheta(x)|$.
1535: The sign $+$ is chosen for 
1536: $(\vartheta\ {\rm mod}\ \pi)\in[0,\pi/2]$ and the sign $-$ for 
1537: $(\vartheta\ {\rm mod}\ \pi)\in[\pi/2,\pi]$. This new process obeys 
1538: the stochastic differential equation
1539: \be \label{eqzeta}
1540: \frac{\D}{\D x}\zeta = k\cosh2\zeta \pm \phi(x)
1541: \:.\ee
1542: Between two nodes of the wave function, the variable $\zeta$ crosses twice the
1543: interval $]-\infty,+\infty[$. Note that when $\phi(x)$ is a white noise of
1544: zero mean, the sign $\pm$ can be disregarded. The study of the time required
1545: by the process to cross the interval can be performed by the same method as
1546: above~:  we introduce the ``time'' $\tilde\Lambda(\zeta)$ needed 
1547: to go from $\zeta$ to $+\infty$.
1548: The Laplace transform of the distribution
1549: $h(\alpha,\zeta) = 
1550: \smean{ \EXP{-\alpha\tilde\Lambda} \:|\:\zeta(0)=\zeta;\ \zeta(\tilde\Lambda)
1551: =+\infty }$
1552: obeys a diffusion equation 
1553: $
1554: (k\cosh2\zeta\:\partial_\zeta + \frac{g}{2}\partial_\zeta^2)
1555: h(\alpha,\zeta) = \alpha \: h(\alpha,\zeta)
1556: $
1557: that involves the backward Fokker-Planck operator related to the 
1558: Langevin equation (\ref{eqzeta}).
1559: 
1560: In the low energy limit $k\ll g$ we expect that most of the ``time''
1561: $\Lambda\equiv\tilde\Lambda(-\infty)$ is spent in the region where the
1562: potential is almost flat (see figure~\ref{fig:potsusy}), therefore we replace
1563: the diffusion equation for $h(\alpha,\zeta)$ by the free diffusion equation
1564: $
1565: \frac{g}{2}\partial_\zeta^2h(\alpha,\zeta) = \alpha \: h(\alpha,\zeta)
1566: $ 
1567: on the finite interval $\zeta\in[\zeta_-,\zeta_+]$, with a reflecting boundary
1568: condition at one side $\partial_\zeta h(\alpha,\zeta_-)=0$ and 
1569: $h(\alpha,\zeta_+)=1$ at the other side (which corresponds actually to 
1570: the absorption at $\zeta=\zeta_+$). The coordinates $\zeta_\pm$ are
1571: the points where the deterministic force and the white noise have equal 
1572: strengths.
1573: 
1574: Now, we can obtain $h(\alpha,\zeta)$ straightforwardly.
1575: \be
1576: \smean{\EXP{-\alpha\Lambda}} \simeq h(\alpha,\zeta_-) 
1577: = \frac{1}{\cosh\sqrt{\alpha/N(E)}}
1578: \:,\ee
1579: where $N(E)={g}/{2\ln^2(g/k)}$ is the IDoS per unit length.
1580: 
1581: 
1582: \begin{figure}[!ht]
1583: \begin{center}
1584: \includegraphics{potUsusyb.eps}
1585: \end{center}
1586: \caption{\it The potential related to the deterministic force in 
1587:          eq.~(\ref{eqzeta}) felt by the  
1588:          variable $\zeta$.\label{fig:potsusy}}
1589: \end{figure}
1590: 
1591: %Strictly speaking there should be a $\pm$ sign in front of the $\phi(x)$, but
1592: %since $\phi(x)$ is a white noise of zero mean the sign plays no role and is
1593: %disregarded. eq.~(\ref{Riceq}) is a Langevin equation for a ``particle'' of
1594: %position $\zeta$ traveling from $-\infty$ to $+\infty$ in a potential
1595: %$U(\zeta)=-\frac{k}{2}\sinh2\zeta$.
1596: %As in the case of diagonal disorder, we study the distribution of the 
1597: %``time'' required by the variable $\zeta$ to go from $-\infty$ to $+\infty$.
1598: %We introduce the characteristic function for the  ``time'' 
1599: %$\tilde\Lambda(\zeta)$ 
1600: %needed to go from $\zeta$ to $+\infty$~:
1601: %$h(\alpha,\zeta) = 
1602: %\smean{ \EXP{-\alpha\tilde\Lambda} \:|\:\zeta(0)=\zeta;\ \zeta(\tilde\Lambda)
1603: %=+\infty }$
1604: %with boundary conditions 
1605: %$\partial_\zeta h(\alpha,\zeta)\big|_{-\infty}=0$ and 
1606: %$h(\alpha,+\infty)=1$.
1607: %It may be obtained by solving
1608: %$
1609: %(k\cosh2\zeta\:\partial_\zeta + \frac{g}{2}\partial_\zeta^2)
1610: %h(\alpha,\zeta) = \alpha \: h(\alpha,\zeta)
1611: %$.
1612: %We study this equation in the low energy limit $k\ll g$ by noticing that t
1613: %he effect of the random ``force'' $\phi(x)$ and  the deterministic ``force'' 
1614: %$-\partial_\zeta U(\zeta)$ decouple. Denoting by
1615: %$\zeta_\pm$  the two points where the deterministic force $k\cosh2\zeta$
1616: %is of the same order of magnitude as the Langevin force we then have the
1617: %following picture. The force
1618: %$-\partial_\zeta U(\zeta)$ brings the particle initially at $-\infty$ to 
1619: %$\zeta_-=-\frac12\ln(g/k)$ in a short time (of order $1/g$).
1620: %Then the particle diffuses normally to $\zeta_+=\frac12\ln(g/k)$
1621: %due to the Langevin ``force'' $\phi(x)$. At $\zeta_+$ the particle is 
1622: %brought back
1623: %to $+\infty$ mostly by the deterministic force again.
1624: %On $[\zeta_-,\zeta_+]$ the dynamics of the particle can be replaced 
1625: %by a normal diffusion with a reflecting 
1626: %boundary at $\zeta_-$ and an absorbing boundary at $\zeta_+$.
1627: 
1628: If $\Lambda\equiv\tilde\Lambda(-\infty)$ is the ``time'' needed to cross
1629: the interval, 
1630: the distance between two nodes of the wave function is a sum of two such
1631: (independent) random variables $\ell=\Lambda_1+\Lambda_2$. Therefore its 
1632: distribution is given by inverse Laplace transformation of 
1633: $\smean{\EXP{-\alpha\ell}}=1/\cosh^2\sqrt{\alpha/N(E)}$.
1634: We obtain~:
1635: \be\label{dol0susy2}
1636: P(\ell) = N(E) \, \varpi_0(N(E) \ell)
1637: \:,\ee
1638: where $\varpi_0(x)$ is the inverse Laplace transform of $\cosh^{-2}\sqrt{s}$~:
1639: \bea\label{pi0susy}
1640: \varpi_0(x) &=&  \heav(x) \sum_{m=0}^\infty
1641: \left[\pi^2(2m+1)^2x - 2\right] \EXP{ -\frac{\pi^2}{4}(2m+1)^2 x } 
1642: \APPROX{x\to\infty} \pi^2 x \EXP{ -\frac{\pi^2}{4} x }\\
1643: &=& \frac{4}{\sqrt{\pi}}\:\frac{\heav(x)}{x^{3/2}} 
1644: \sum_{m=1}^\infty(-1)^{m+1} m^2\EXP{-m^2/x}
1645: \APPROX{x\to0} \frac{4}{\sqrt{\pi}}\:\frac{\heav(x)}{x^{3/2}}\:\EXP{-1/x}
1646: \:,\eea
1647: where $\heav(x)$ is the Heaviside function.
1648: 
1649: We find the distribution of the ground state energy~:
1650: \be
1651: W_0(E) = L\rho(E)\,\varpi_0(LN(E))
1652: \:,\ee
1653: with the following limiting behaviour~:
1654: \bea \label{W0lim1}
1655: W_0(E) &\simeq& \frac{8}{\sqrt{2\pi gL}} \frac1E 
1656: \exp-\frac{\ln^2(g^2/E)}{2gL} 
1657: \hspace{1.2cm}{\rm for} \hspace{0.5cm} E\ll g^2\EXP{-\sqrt{2gL}}\\
1658: \label{W0lim2}
1659: &\simeq& \frac{8\pi^2g^2L^2}{E\ln^5(g^2/E)}
1660: \exp-\frac{\pi^2gL}{2\ln^2(g^2/E)} 
1661: \hspace{0.5cm} {\rm for} \hspace{0.5cm} g^2\EXP{-\sqrt{2gL}}\ll E\ll g^2
1662: \:.\eea
1663: In ref.~\cite{Tex00}, an integral representation of $W_n(E)$ is also given
1664: together with an explicit expression for $W_1(E)$.
1665: 
1666: It is also interesting to point out that the ground state distribution is 
1667: characterized by a typical value $E_0^{\typ} \simeq g^2 \EXP{-g L}$, 
1668: a median value $E_0^{\rm med} \sim g^2 \EXP{-\sqrt{gL}}$,
1669: and a mean value
1670: $\smean{E_0} \sim g^{2}\, (gL)^{1/2}\, \EXP{-C(gL)^{1/3}}$
1671: where $C$ is a numerical constant (see ref.~\cite{Tex00}).
1672: This latter expression has also been obtained in ref.~\cite{MonOshComBur96} 
1673: where upper and lower bounds were found using a perturbative 
1674: expression for the ground state energy as a functional of $\phi(x)$.
1675: 
1676: The distribution (\ref{pi0susy}) was obtained in ref.~\cite{LeDMonFis99} in
1677: the context of the classical diffusion by using a real space renormalization
1678: group method. In this case the distribution is interpreted as the distribution
1679: of the smallest relaxation time.
1680: 
1681: In summary, these two examples illustrate the fact that extreme value
1682: spectral statistics provides an information on correlations of eigenvalues~:
1683: the extreme value statistics of independent and identically distributed 
1684: random variables have been
1685: classified by Gumbel \cite{Gum54}. Therefore extreme value distribution for
1686: eigenvalues that differ from one of the three Gumbel's laws indicates level
1687: correlations.
1688: 
1689: 
1690: 
1691: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1692: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1693: 
1694: \section{Trace formulae, spectral determinant and diffusion on graphs\label{sec:graphs}}
1695: 
1696: \subsection{Introduction}
1697: 
1698: Up to now we have discussed several questions related to the physics of 
1699: weak and strong localization. In the previous section we have considered
1700: spectral properties while in the sections \ref{sec:wl} and \ref{sec:expfun}
1701: we mostly discussed transport properties.
1702: 
1703: In the present section we are going to review several results related to the
1704: study of the Laplace operator on metric graphs.  An object at the core of our
1705: discussion is the spectral determinant of the Laplace operator, formally
1706: defined as $S(\gamma)=\det(\gamma-\Delta)$, where $\gamma$ is a spectral
1707: parameter. This quantity encodes the information on the spectrum of the
1708: Laplace operator on the graph. The subsection \ref{sec:description} recalls
1709: the basic conventions required to describe metric graphs. The subsections
1710: \ref{sec:tf} and \ref{sec:sd} review general results on trace formulae and
1711: spectral determinants. These subsections will appear at first sight quite
1712: technical and unrelated to the previous sections. However we will see that
1713: there exists a close relation with the question of quantum transport~: the
1714: Laplace operator is the generator of the diffusion on the graph and its
1715: spectral properties play a central role in the study of transport. This
1716: connection, already evoked in the introduction and section~\ref{sec:wl}, will
1717: be emphasized again in subsection~\ref{sec:wlrg} where the explicit relation
1718: between quantum transport and spectral determinant is recalled.
1719: 
1720: One interest for spectral determinants is that they can be used as generating
1721: functions for various quantities characterizing the diffusion on the graph.
1722: This will be illustrated by using the connection with trace formulae
1723: (section~\ref{sec:tf}) and further exploited in subsection~\ref{sec:wlrg} in
1724: the context of quantum transport.  The efficiency of the method stems from the
1725: fact that, even though $S(\gamma)$ seems to be a complicated object at first
1726: sight, involving an infinite number of eigenvalues, it can be expressed as the
1727: determinant of a finite size matrix. This relation, established by Pascaud \&
1728: Montambaux \cite{Pas98,PasMon99}, allows computing easily and systematically
1729: the spectral determinant for arbitrary graphs.
1730: 
1731: The study of the Laplace operator on metric graphs (or {\it quantum graphs})
1732: is not restricted to transport and
1733: appears in many physical contexts ranging from 
1734: organic molecules \cite{RudSch53}, superconducting networks \cite{Ale83},
1735: phase coherent transport in networks of weakly disordered wires
1736: \cite{DouRam85,PasMon99,AkkComDesMonTex00,TexMon04}, 
1737: transport in mesoscopic networks 
1738: \cite{Sha82,Sha83,ButImrAzb84,GefImrAzb84,AvrSad91,VidMonDou00,TexMon01} 
1739: or metallic agregates \cite{CarAku04}.
1740: We are not going to review this history and refer the interested reader
1741: to the references~\cite{Ale83,AvrRavZur88,Col98,KotSmi99,AkkComDesMonTex00,KotSmi03} 
1742: (for a recent review see the appendix of Pavel Exner in 
1743: ref.~\cite{AlbGesHoeHol04}).  
1744: Mathematical aspects of quantum graphs are discussed in the recent issues of 
1745: Waves Random Media {\bf14} (2004) and Journal of Physics A~{\bf38} (n$^o$22) 
1746: (June 2005). See in particular the refs.~\cite{Kuc04,Kuc05}.
1747: 
1748: 
1749: \subsection{Description of metric graphs\label{sec:description}}
1750: 
1751: We collect some background material from the theory of graphs and establish
1752: our notations and conventions (illustrated on figure~\ref{fig:exampleqds}).
1753: 
1754: \noindent{\bf Vertices, bonds, arcs.--}
1755: Let us consider a network of $B$ wires (bonds) connected at $V$ vertices. The
1756: latter are labelled with Greek indices $\alpha$, $\beta$,... Therefore, the
1757: bonds of the graph can be denoted by a couple of Greek indices $(\ab)$. The
1758: oriented bonds, denoted as {\it arcs}, will also play an important role. The
1759: two arcs related to the bond $(\ab)$ will be labelled by $\ab$ and $\ba$ or
1760: more simply with Roman letters $i$, $j$,...
1761: 
1762: \noindent{\bf Adjacency matrix.--}
1763: The basic object is the adjacency
1764: $V\times V$~matrix $a_{\alpha\beta}$ characterizing the topology of the
1765: network~:  $a_{\alpha\beta}=1$ if $\alpha$ and $\beta$ are connected by a
1766: bond~; $a_{\alpha\beta}=0$ otherwise\footnote{
1767:   Note that this definition assumes that two vertices are connected by 
1768:   at most one bond and that a bond never forms a loop.
1769:   We emphasize that {\it this does not imply any particular restriction on 
1770:   the topology of the graph}~: one can always introduce a vertex on 
1771:   a bond, which separates the bond into two bonds, without modifying 
1772:   the properties of the graph. The numbers of vertices and bonds can 
1773:   always be made arbitrary large and this is partly a matter of choice. 
1774:   The case of closed bonds (forming a loop) or vertices connected by 
1775:   multiple bonds requires a simple generalization of the formalism
1776:   presented here (see for example ref.~\cite{Pas98} and appendix C of 
1777:   ref.~\cite{AkkComDesMonTex00}). This generalization allows in 
1778:   particular to minimize $B$ and $V$, which makes the computation 
1779:   sometimes easier.
1780: }.
1781: The connectivity of the vertex $\alpha$, denoted by $m_\alpha$, is 
1782: related to the adjacency matrix by $m_\alpha=\sum_\beta a_{\alpha\beta}$.
1783: Summed over the remaining index, the adjacency matrix gives  the 
1784: number of arcs~: $\sum_{\alpha,\beta} a_{\alpha\beta}=2B$.
1785: 
1786: \noindent{\bf Orbits.--}
1787: A path is an ordered set of arcs such that the end of an arc coincides with
1788: the begining of the following arc. The equivalence class of all closed paths
1789: equivalent by cyclic permutations is called an {\it orbit}.
1790: An orbit is said to be {\it primitive} when it can not be decomposed as 
1791: a repetition of a shorter orbit.
1792: 
1793: \begin{figure}[htbp]
1794:   \centering
1795:   \includegraphics[scale=1.25]{exampleqds.eps}
1796:   \caption{{\it An example of graph with 11 vertices and 13 bonds.
1797:            The arrows  show the orientations of the arc $i$ 
1798:            and the reversed arc $\bar i$.
1799:            The blue curve is an example of (primitive) orbit with one 
1800:            backtracking.}}
1801:   \label{fig:exampleqds}
1802: \end{figure}
1803: 
1804: \noindent{\bf Scalar functions.--}
1805: The graphs we consider here are not simply topological objects but have also
1806: some metric properties~: the bond $(\alpha\beta)$ is characterized by its 
1807: length $l_{\alpha\beta}$ and identified with the interval 
1808: $[0,l_{\alpha\beta}]$ of $\RR$. A scalar function $\psi(x)$ is defined
1809: by  its  components $\varphi_{\ab}(x_{\ab})$ on each bond,
1810: where $x_\ab$ is the coordinate measuring the distance along the bond 
1811: from the vertex $\alpha$ (note that $x_\ab+x_\ba=l_\ab$).
1812: 
1813: \noindent{\bf Continuous boundary conditions.--}
1814: When studying the Laplace operator $\Delta$ acting on scalar functions, 
1815: boundary conditions at the vertices must be specified in order to 
1816: ensure self adjointness of the operator.
1817: Let us introduce the notations~:
1818: \bea
1819: \label{phiab}
1820: \varphi_\ab &\equiv& \varphi_\ab(x_\ab=0) \\
1821: \label{phiprimeab}
1822: \varphi'_\ab &\equiv& \frac{\D\varphi_\ab}{\D x_\ab}(x_\ab=0)
1823: \:,\eea
1824: for the value of the function and its derivative at the vertex $\alpha$, along
1825: the arc $\ab$. 
1826: The continuous boundary conditions assume~:\\
1827: ({\it i}) continuity of the function at each vertex~: all the 
1828: components $\varphi_{\alpha\beta}$ for all $\beta$ neighbours of $\alpha$
1829: are equal. The value of the function at the vertex is denoted 
1830: $\varphi(\alpha)$.\\
1831: ({\it ii})
1832: $
1833: \sum_\beta a_{\alpha\beta} \varphi'_{\alpha\beta}=
1834: \lambda_\alpha\varphi(\alpha)
1835: $
1836: where the adjacency matrix in the sum constrains it to run over the
1837: neighbouring vertices of $\alpha$. Therefore the sum runs over all wires
1838: issuing from the vertex. The real parameter $\lambda_\alpha$ allows 
1839: describing several boundary conditions~: $\lambda_\alpha=\infty$ enforces the
1840: function to vanish, $\varphi(\alpha)=0$, and corresponds to the 
1841: Dirichlet boundary
1842: condition. $\lambda_\alpha=0$ corresponds to the Neumann boundary condition.
1843: We refer to the case of a finite $\lambda_\alpha$  as ``mixed boundary 
1844: conditions''.
1845: In the problem of classical diffusion, the Dirichlet condition describes the 
1846: connection to a reservoir that absorbs particles, while the Neumann condition
1847: ensures conservation of the probability current and describes an internal 
1848: vertex.
1849: 
1850: \noindent{\bf General boundary conditions.--}
1851: Thanks to the continuity hypothesis, the simple boundary conditions we have 
1852: just described allow us to introduce vertex variables $\varphi(\alpha)$ 
1853: (this is convenient since the number of vertices $V$ is  usually smaller than 
1854: the number of arcs $2B$).
1855: However these are not the most general boundary conditions, and in general the
1856: different components associated to the arcs issuing from a given vertex do not
1857: have {\it a priori} the same limit at the vertex. The most general boundary
1858: conditions can be written as~:
1859: \be
1860: C\,\varphi + D\,\varphi' =0
1861: \ee
1862: where $\varphi$ and $\varphi'$ are the column vectors of dimension $2B$
1863: collecting the variables (\ref{phiab}) and (\ref{phiprimeab}), respectively.
1864: $C$ and $D$ are two square $2B\times2B$ matrices. 
1865: In the arc formulation, the information on the topology of the graph is 
1866: encoded in these two matrices.
1867: The self adjointness of the Laplace operator is ensured if these two 
1868: matrices satisfy the following conditions \cite{KosSch99}~:
1869: ({\it i}) $CD^\dagger$ is self-adjoint.
1870: ({\it ii}) The  $2B\times4B$ matrix $(C,D)$ must have maximal rank.
1871: Note that the choice of $C$ and $D$ is not unique.
1872: The introduction of boundary conditions without continuity at the vertices has 
1873: been motivated physically in 
1874: refs.~\cite{Sha83,GefImrAzb84,ButImrAzb84,AvrExnLas94,Exn95,Exn96a}.
1875: 
1876: \noindent{\bf Scattering theory interpretation.--}
1877: We can give a more clear physical meaning to these conditions in a
1878: scattering setting. Let us consider the equation 
1879: $-\Delta\varphi(x)=E\varphi(x)$
1880: for a positive energy $E=k^2$. We decompose the component on the bond as the
1881: superposition of an incoming and an outgoing plane wave~:
1882: $\varphi_\ab(x_\ab)=I_\ab\,\EXP{-\I kx_\ab}+O_\ab\,\EXP{\I kx_\ab}$. It is
1883: convenient to collect the amplitudes in column vectors $I$ and $O$. The vector
1884: $I$ contains the incoming plane wave amplitudes and the vector $O$ 
1885: the outgoing plane wave amplitudes. Both are related by a vertex scattering
1886: matrix~:  $O=Q\,I$. The self-adjointness of the Schr\"odinger operator is now
1887: ensured by imposing the unitarity of the scattering matrix~: $Q^\dagger Q=1$.
1888: The relation between the two formulations yields~:
1889: \be\label{relQDC}
1890: Q=(\sqrt{\gamma}\,D-C)^{-1}(\sqrt{\gamma}\,D+C)
1891: \:.\ee 
1892: For the discussion below it is convenient to introduce the parameter
1893: $\gamma$ related to the energy by $E=-\gamma=k^2+\I0^+$ (the matrix $Q$ is
1894: unitary for $E>0$ only). 
1895: An example of a matrix $Q$ for one vertex without continuity of the wave
1896: function was given in refs.~\cite{Sha83,GefImrAzb84,ButImrAzb84}. This
1897: particular choice has become popular in mesoscopic physics (see also 
1898: ref.~\cite{TexMon01} where a more convenient parametrization was provided). 
1899: 
1900: \noindent
1901: Since we are considering here compact graphs, we are dealing only with discrete
1902: spectrum. The study of noncompact graphs (with some wires of infinite length),
1903: which have continuous spectra, requires a scattering theory approach,
1904: initiated by the work of Shapiro \cite{Sha82,Sha83} and discussed in 
1905: refs.~\cite{GerPav88,AvrSad91,Ada92,KosSch99,KotSmi00,BarGas01,KotSmi03,TexMon01,Tex02,TexBut03,TexDeg03}.
1906: 
1907: 
1908: \noindent{\bf Magnetic fluxes.--}
1909: The Laplace operator arises in the context of diffusion equation but also 
1910: in quantum mechanics. In this last case a natural generalization is to 
1911: introduce a magnetic field. This is achieved by introducing a 1-form 
1912: $A(x)\D x$ along the wires (the derivative must then be replaced by
1913: a covariant derivative~: $\D_x\to\Dc_x=\D_x-\I A(x)$). We denote by 
1914: $\theta_\ab=\int_\alpha^\beta\D x\,A(x)$ the corresponding line integral
1915: along the arc $\ab$.
1916: In the context of classical diffusion, magnetic fluxes and winding numbers are
1917: conjugated variables.
1918: 
1919: 
1920: \subsection{Trace formulae and zeta functions\label{sec:tf}}
1921: 
1922: Trace formulae play an important role in spectral theory. A famous example is
1923: the Selberg trace formula which may be viewed as an extension of the Poisson
1924: summation formula to non commutative groups \cite{Sel56}. An analogous formula
1925: in physics is the Gutzwiller trace formula \cite{Gut90} that has been
1926: extensively used in the context of quantum chaos and mesoscopic physics.
1927: Although one is exact and the other only a semiclassical approximation, both
1928: of them express the partition function (or the density of states, whose
1929: Laplace transform gives the partition function) as a sum over closed
1930: geodesics. They provide a connection between quantum properties (spectrum) and
1931: classical properties (classical trajectories). Below we discuss two examples
1932: of exact trace formulae that have been derived for graphs and their relation
1933: to spectral determinants.
1934: 
1935: 
1936: \subsubsection*{Roth's trace formula}
1937: 
1938: It expresses the trace of the heat kernel (partition function
1939: $Z(t)=\tr{\EXP{t\Delta}}$) as an infinite series of contributions of periodic
1940: orbits on the graph. This remarkable formula, due to Roth \cite{Rot83,Rot83a},
1941: applies to graphs with continuous boundary conditions with $\lambda_\alpha=0$.
1942: It it easy to include magnetic fluxes additionally~:
1943: \be\label{Roth}
1944: Z(t)=
1945: \frac{\cal L}{2\sqrt{\pi t}} + \frac{V-B}{2} + \frac{1}{2\sqrt{\pi t}}
1946: \sum_{{\cal C}} l(\widetilde{\cal C}) \alpha({\cal C})
1947: \EXP{-\frac{l({\cal C})^2}{4t}+\I\theta({\cal C})}
1948: \:.\ee
1949: ${\cal L}$ is the ``volume'' of the graph, {\it i.e.} the total length
1950: ${\cal L}=\sum_{(\ab)}l_\ab$.
1951: The sum runs over all orbits ${\cal C}=(i_1,i_2,\cdots,i_n)$ constructed in
1952: the graph. 
1953: $l({\cal C})=l_{i_1}+\cdots+l_{i_n}$ is the total length of the orbit,
1954: and $\theta({\cal C})$ the magnetic flux enclosed by it.
1955: $\widetilde{{\cal C}}$ designates the primitive orbit associated with a 
1956: given orbit ${\cal C}$.
1957: The weight $\alpha({\cal C})$ depends on the connectivity of the vertices 
1958: visited by the orbit~:
1959: $\alpha({\cal C})=\epsilon_{i_1i_2}\epsilon_{i_2i_3}\cdots\epsilon_{i_ni_1}$.
1960: The matrix $\epsilon$ couples the arcs of the graph\footnote{
1961:   Note that the parameter $\epsilon_{ij}$ has a simple interpretation in 
1962:   the scattering formulation \cite{TexMon01}~: it is the 
1963:   probability amplitude to be transmitted from arc $i$ to arc $j$.
1964:   The arc matrices $\epsilon$ and $Q$ are related by 
1965:   $\epsilon_{ij}=Q_{i\bar j}$ where $\bar j$ denotes the reversed arc.
1966: }~:\\
1967: $\bullet$ if $i$ ends at vertex $\alpha$ and $j$ starts from it, we have 
1968: $\epsilon_{ij}=2/m_\alpha$, where 
1969: $m_\alpha$ is the connectivities of the vertex.\\
1970: $\bullet$ If moreover $i$ and $j=\bar i$ are the reversed arcs
1971: $\epsilon_{i\bar i}=2/m_\alpha-1$.\\
1972: $\bullet$ Otherwise $\epsilon_{ij}=0$.
1973: 
1974: It is worth mentioning that the Roth trace formula has found recently some
1975: pratical applications to analyze magnetoconductance measurements on large
1976: square networks\footnote{   
1977:   More precisely, the expansion given below by eq.~(\ref{traceformulaS}) for
1978:   the Laplace transform of the partition function is relevant in this case.
1979: }~\cite{FerAngRowGueBouTexMonMai04}.
1980: 
1981: 
1982: \subsubsection*{Ihara-Bass trace formula}
1983: 
1984: Instead of considering metric graphs we now turn to graphs viewed as purely 
1985: combinatorial structures consisting of vertices connected by bonds 
1986: of equal lengths ($l_\ab=1$). 
1987: All the information is therefore encoded in the adjacency matrix $A$.
1988: In this setting, the Ihara $\zeta$-function is defined as
1989: \be
1990: \zeta(u)^{-1}  
1991: = \prod_{\widetilde{\cal C}_B} \left( 1 - u^{l(\widetilde{\cal C}_B)} \right)
1992: \:,\ee
1993: where the infinite product extends over all primitive backtrackless orbits
1994: $\widetilde{\cal C}_B$.
1995: The Ihara-Bass trace formula relates this infinite product to the determinant
1996: of a finite size matrix \cite{Bas92,StaTer96}~:
1997: \be\label{IharaZfct}
1998: \zeta(u)^{-1}  = (1-u^2)^{B-V}\,\det((1-u^2)\ide - u\,A + u^2\,Y)
1999: \ee
2000: where $\ide$ is the identity matrix, 
2001: $A$ is the adjacency matrix ($A_\ab\equiv a_\ab$) and $Y$ is
2002: the diagonal matrix encoding all connectivities~: $Y_\ab=\delta_\ab m_\alpha$.
2003: This relation was derived by Ihara \cite{Iha66} for regular graphs (all
2004: vertices with same connectivity) and later generalized to arbitrary graphs by
2005: Bass~\cite{Bas92}.
2006: The formalism that we have developed for metric graphs is flexible enough
2007: to describe these combinatorial structures in the same setting.
2008: We will see below that the $\zeta$-function is in fact directly related to the 
2009: spectral determinant. Moreover the Ihara-Bass trace formula can be further 
2010: generalized to include backtrackings.
2011: 
2012: 
2013: \subsection{Spectral determinant\label{sec:sd}}
2014: 
2015: In the physics literature spectral determinants arise in evaluating path
2016: integrals which are quadratic in the fluctuation around a given background
2017: field. A well known technique for regularizing such quadratic path integrals
2018: is the $\zeta$-function regularization. Given a certain operator ${\cal O}$
2019: whose eigenvalues $E_n$ are known, one defines the following $\zeta$-function~:
2020: $\zeta(s)=\sum_n E_n^{-s}$. This expression, which converges for $s$
2021: sufficiently large, can be analytically extended to a meromorphic function
2022: regular at the origin. The corresponding regularized determinant is then
2023: $\det{\cal O}=\exp(-\zeta'(0))$ ($\equiv\prod_nE_n$ formally). 
2024: One can find a general
2025: discussion on functional determinants in ref.~\cite{For87}.
2026: 
2027: %The success of functional methods and path integration has led to generalize
2028: %the concept of determinant to space of infinite dimensions. In this case
2029: %it is not possible to perform the infinite product over the eigenvalues
2030: %$\{E_n\}$. A possible way to define a determinant is to introduce the 
2031: %$\zeta$-function $\zeta(s)=\sum_n E_n^{-s}$, the determinant being defined 
2032: %as $\exp(-\zeta'(0))$ (see ref.~\cite{For87} for a general discussion on 
2033: %functional determinants).
2034: 
2035: In the context of graphs another regularization of the determinant of the
2036: Laplace operator has been used and has proved to be directly related to several
2037: physical quantities. If we introduce the trace of the resolvent
2038: $g(\gamma)=\sum_n(\gamma+E_n)^{-1}$, the spectral determinant is defined as
2039: $S(\gamma)=\exp(\int^\gamma\D\gamma'\,g(\gamma'))$ (it can be formally written
2040: as\footnote{
2041:   The limit $\gamma\to0$ of $S(\gamma)$ has been
2042:   discussed in ref.~\cite{AkkComDesMonTex00}. 
2043: }
2044: $S(\gamma)=\prod_n(\gamma+E_n)$). Moreover, the spectral parameter $\gamma$
2045: has in some cases a physical meaning (see section~\ref{sec:wlrg}).
2046: 
2047: \mathversion{bold}
2048: \subsubsection*{Laplace operator $\Delta$ with continuous boundary conditions}
2049: \mathversion{normal}
2050: 
2051: Pascaud \& Montambaux have shown in refs.~\cite{Pas98,PasMon99} that 
2052: $S(\gamma)=\det(\gamma-\Delta)$ can be 
2053: related to the determinant of a $V\times V$-matrix~:
2054: \be\label{spedet}
2055: S(\gamma) =
2056: \gamma^{\frac{V-B}2}
2057: \prod_{(\alpha\beta)}\sinh(\sqrt\gamma l_{\alpha\beta})\:
2058: \det M
2059: \:,\ee
2060: where the product runs over all bonds of the network.
2061: The $V\times V$-matrix $M$ is defined as~:
2062: \be\label{defM}
2063: M_\ab=\delta_\ab
2064: \left(\frac{\lambda_\alpha}{\sqrt\gamma} + \sum_\mu a_{\alpha\mu}
2065:       \coth(\sqrt{\gamma}l_{\alpha\mu})\right)
2066: -a_\ab\frac{\EXP{-\I\theta_\ab}}{\sinh(\sqrt{\gamma}l_\ab)}
2067: \:,\ee
2068: where the adjacency matrix constrains the sum to run over all vertices
2069: $\mu$ connected to $\alpha$. The matrix $M$ encodes all information about the
2070: network~: topology (matrix $a_{\alpha\beta}$), lengths of the wires ($l_\ab$),
2071: magnetic fluxes ($\theta_\ab$), boundary conditions ($\lambda_\alpha$). Below,
2072: we  give several examples which show how $S(\gamma)$ is related to the
2073: characteristic function of various interesting functionals of Brownian
2074: curves on a graph.
2075: 
2076: The expression (\ref{spedet}) was originally derived in ref.~\cite{PasMon99}
2077: by constructing the Green's function in the graph and eventually integrating
2078: it~:
2079: $\int\D x\,\bra{x}\frac1{\gamma-\Delta}\ket{x}=\drond{}{\gamma}\ln S(\gamma)$.
2080: A more direct derivation using path integral was later obtained  in 
2081: ref.~\cite{AkkComDesMonTex00}.
2082: 
2083: The relation between the Roth trace formula and the result (\ref{spedet})
2084: was addressed in ref.~\cite{AkkComDesMonTex00}. The main difficulty to
2085: establish this connection is to go from vertex variables to the arc language
2086: of eq.~(\ref{Roth}). A first step is to relate the determinant of the
2087: vertex-matrix, eq.~(\ref{spedet}), to the determinant of an arc-matrix~:
2088: \be\label{LfctS}
2089: S(\gamma)=\gamma^{\frac{V-B}{2}}\EXP{\sqrt{\gamma}\,{\cal L}}
2090: \,\det(\ide - QR)
2091: \:.\ee
2092: $R$ is the $2B\times2B$~matrix 
2093: $R_{ij}=\delta_{i\bar j}\,\EXP{-\sqrt{\gamma} l_i + \I\theta_i}$, where 
2094: $\bar j$ denotes the reversed arc.
2095: The matrix $Q$ was introduced above (eq.~(\ref{relQDC})) and is related to 
2096: $\epsilon$ by $Q_{ij}=\epsilon_{i\bar j}$. 
2097: For the continuous boundary conditions with $\lambda_\alpha=0$ it is given by
2098: $Q_{ii}=2/m_\alpha-1$, 
2099: $Q_{ij}=2/m_\alpha$ if $i$ and $j$ both issue from the vertex $\alpha$.
2100: $Q_{ij}=0$ in other cases.
2101: Eq.~(\ref{LfctS}) holds for the most simple boundary conditions~: 
2102: continuous with $\lambda_\alpha=0$. The general case is discussed below.
2103: Expanding the determinant by using 
2104: $\ln\det(\ide - QR)=- \sum_{n=1}^\infty \frac{1}{n}\tr{(QR)^n}$, 
2105: we eventually express the spectral determinant as an infinite product over the 
2106: primitive orbits~:
2107: \be\label{zetafctS}
2108: S(\gamma)=\gamma^{\frac{V-B}{2}}\EXP{\sqrt{\gamma}\,{\cal L}}
2109: \prod_{\widetilde{\cal C}}
2110: \left(  
2111:   1 - \alpha(\widetilde {\cal C})
2112:   \EXP{-\sqrt{\gamma}\,l(\widetilde {\cal C})+\I\theta(\widetilde {\cal C})}
2113: \right)
2114: \:.\ee
2115: This shows that the spectral determinant is a zeta function 
2116: (references on zeta functions on graphs are refs.~\cite{StaTer96,Che99}).
2117: The last step to connect this formula to Roth's trace formula is to 
2118: notice that 
2119: \be\label{traceformulaS}
2120: \drond{}{\gamma}\ln S(\gamma)=\frac{{\cal L}}{2\sqrt\gamma}+\frac{V-B}{2\gamma}
2121: +\frac{1}{2\sqrt\gamma}\sum_{{\cal C}} l(\widetilde {\cal C})
2122: \alpha({\cal C})\EXP{-\sqrt{\gamma}\,l({\cal C})+\I\theta({\cal C})}
2123: \:,\ee
2124: where the sum now runs over all orbits (if ${\cal C}$ is not primitive, 
2125: $\widetilde{\cal C}$ designates the related primitive orbit).
2126: Finally we perform an inverse Laplace transform of this expression,
2127: $\int_0^\infty\D t\,Z(t)\EXP{-\gamma t}=\drond{}{\gamma}\ln S(\gamma)$
2128: and eventually recover eq.~(\ref{Roth}).
2129: Examples of applications are studied in ref.~\cite{AkkComDesMonTex00}. 
2130: 
2131: 
2132: \mathversion{bold}
2133: \subsubsection*{Schr\"odinger operator $-\Delta+V(x)$ with general boundary conditions}
2134: \mathversion{normal}
2135: 
2136: The result (\ref{spedet}) of Pascaud \& Montambaux has been generalized by one
2137: of us. In refs.~\cite{Des00,Des00a} a similar formula was obtained for the
2138: spectral determinant of the Schr\"odinger operator $-\Delta+V(x)$ (the Hill
2139: operator), where $V(x)$ is a scalar potential defined on the graph. In
2140: ref.~\cite{Des01} the formula was further extended to describe general
2141: boundary conditions as well. As an illustration we construct the generating
2142: function of the number of closed orbits with a given number of backtrackings
2143: \cite{Iha66,WuKun99}.
2144: 
2145: \vspace{0.25cm}
2146: 
2147: The starting point is to introduce 
2148: two linearly independent solutions of the differential equation
2149: $(-\D_x^2+V_\ab(x)+\gamma)f(x)=0$ on $[0,l_\ab]$. We associate each 
2150: solution with an arc. Let us denote $f_\ab(x_\ab)$ the function
2151: satisfying 
2152: \be
2153: f_\ab(0)=1 
2154: \hspace{0.5cm} \mbox{ and } \hspace{0.5cm}
2155: f_\ab(l_\ab)=0
2156: \:.\ee 
2157: Therefore a second solution of the differential equation is naturally denoted 
2158: $f_\ba(x_\ba)=f_\ba(l_\ab-x_\ab)$.
2159: The Wronskian of these two solutions, defined as
2160: $W_\ab=f_\ab(x_\ab)\frac{\D f_\ba(x_\ba)}{\D x_\ab}
2161:       -\frac{\D f_\ab(x_\ab)}{\D x_\ab}f_\ba(x_\ba)$, 
2162: is constant along the bond~:
2163: $W_\ab=W_\ba=-f'_\ab(l_\ab)=-f'_\ba(l_\ab)$.
2164: If we consider the case $V(x)=0$, the solution is simply
2165: $
2166: f_\ab(x_\ab)
2167: =\frac{\sinh\sqrt\gamma(l_\ab-x_\ab)}{\sinh\sqrt\gamma l_\ab}
2168: \equiv\frac{\sinh\sqrt\gamma x_\ba}{\sinh\sqrt\gamma l_\ab}
2169: $.
2170: 
2171: All the required information about the potential is contained in the 
2172: $2B\times 2B$ arc-matrix $N$, defined as
2173: \be\label{n161}
2174: N_{\alpha\beta,\mu\eta}  = 
2175: \delta_{\alpha\mu}  \delta_{\beta\eta} 
2176: f'_\ab(0) 
2177: - \delta_{\alpha\eta}\delta_{\beta\mu} f'_\ab(l_\ab)
2178: \:.\ee
2179: This matrix couples a given arc to itself and to its reversed arc, only.
2180: If we assume that {\it the matrices $C$ and $D$ are independent on the 
2181: spectral parameter $\gamma$}, it was shown in ref.~\cite{Des01} that~:
2182: \be\label{n40} 
2183:  S(\gamma ) = \det(\gamma-\Delta+V(x)) = 
2184:  \prod_{(\alpha\beta)}  \frac{1}{\wal}\ \det(C+DN)
2185: \ee
2186: where the product runs over all bonds. 
2187: Functional determinants on a segment of $\RR$ with general boundary conditions
2188: at the boundaries have been studied by McKane \& Tarlie \cite{McKTar95}
2189: using the formalism developed by Forman~\cite{For87}.
2190: 
2191: It is also interesting to encode the information on the potential
2192: $V(x)$ in the matrix $R$ defined as\footnote{
2193:   Let us remark that, for the free case ($V(x)\equiv 0$),
2194:   we recover the expression of the matrix $R$ given above, coupling the arc 
2195:   $\ab$ to the reversed arc $\ba$ only~:
2196:   $
2197:   R_{\alpha\beta,\mu\eta} 
2198:   =\delta_{\alpha\eta}\delta_{\beta\mu} \EXP{-\sqrt{\gamma} \lab }
2199:   $. 
2200: }~:
2201: \be\label{n41} 
2202:   R \equiv  ( \sqrt{\gamma} \ide + N )( \sqrt{\gamma} \ide - N  )^{-1}
2203: \:.\ee
2204: Then
2205: \bea\label{n42} 
2206:  S(\gamma) =  \prod_{(\alpha\beta)} \frac{1}{ \wal } \:
2207:  \frac{1}{\det( \ide + R )} \: \det (C - \sqrt{\gamma} D ) \:
2208:  \det( \ide - Q R ) 
2209: \eea
2210: where the matrix $Q=(\sqrt{\gamma} D - C )^{-1}(\sqrt{\gamma} D + C )$
2211: was defined above by eq.~(\ref{relQDC}).
2212: Let us explain the structure of eq.~(\ref{n42})~: the term 
2213: $\det( \ide + R )\prod_{(\alpha\beta)}W_\ab$ contains informations on the 
2214: potential only\footnote{
2215:   The matrix $N$ encodes the information about the potential on the bonds 
2216:   through $f'_\ab(0)$ and $f'_\ab(l_\ab)$.
2217:   This information can also be introduced through transmission $t_\ab$ and 
2218:   reflection $r_\ab$ amplitudes by the potential $V_\ab(x)$. This relation
2219:   is developed in ref.~\cite{TexMon01}.
2220:   It is interesting to point out that 
2221:   $\det( \ide + R )\prod_{(\alpha\beta)}W_\ab
2222:   =2^B\gamma^{B/2}\prod_{(\alpha\beta)}t_\ab
2223:   =2^B\gamma^{B/2}\prod_{(\alpha\beta)}R_{\ab,\ba}$.
2224: }.
2225: The factor $\det (C -\sqrt{\gamma}D)$ contains only information on the 
2226: topology of the graph.
2227: The most interesting part is the last term $\det(\ide-QR)$ combining both
2228: informations. In particular this last part generates the infinite
2229: contributions of primitive orbits in (\ref{zetafctS}).
2230: 
2231: 
2232: 
2233: 
2234: \subsubsection*{Permutation-invariant boundary conditions.} 
2235: 
2236: In the previous section the matrices describing the graph are $2B\times2B$ arc
2237: matrices. A simplifiaction can be brought by passing to vertex variables. In
2238: the arc formulation, $C$ and $D$ define the topology of the graph~: two arcs
2239: coupled by $C$ and/or $D$ issue from the same vertex. Therefore it is possible
2240: to organize the basis of arcs in such a way that the matrices $C$ and $D$ have
2241: similar block diagonal structures. The matrices are made of $V$ square blocks,
2242: each corresponding to a vertex. A given block, of dimension
2243: $m_\alpha\times{m_\alpha}$ and denoted $C_{\alpha}$ (and $D_{\alpha}$),
2244: corresponds to the $m_\alpha$ arcs issuing from the vertex $\alpha$. If we
2245: assume that the boundary conditions are invariant under any permutation of the
2246: nearest neighbours of $\alpha$, then it is possible to introduce vertex
2247: variables. In this case we can write:
2248: \begin{eqnarray}
2249:    C_{\alpha } &=&  c_{\alpha } \ide  + t_{\alpha }  F_{\alpha }
2250:    \label{bci} \\ 
2251:    D_{\alpha } &=&  d_{\alpha } \ide  + w_{\alpha }  F_{\alpha }
2252:    \label{bci1} 
2253: \end{eqnarray}
2254: where 
2255: %$\ide$ is the unit matrix and 
2256: $F_{\alpha }$ is a matrix with all
2257: its elements equal to $1$. The boundary conditions at the vertex $\alpha$ are
2258: characterized by the four parameters $c_{\alpha }$, $d_{\alpha }$,
2259: $t_{\alpha }$ and $w_{\alpha }$ (note however that this choice
2260: is not unique).
2261: 
2262: \vspace{0.25cm}
2263: 
2264: Now, let us show  that, for boundary conditions
2265: given by eqs.~(\ref{bci},\ref{bci1})
2266: and $V(x)\ne 0$, the spectral determinant can be expressed in terms
2267: of the vertex $V \times V$-matrix.
2268: 
2269: We proceed as before but, this time, we consider, for each bond, two other
2270: independent solutions, $\cab(x_\ab)$ and $\cba(l_\ab-x_\ab)=\cba(x_\ba)$, of
2271: the equation $(-\D_{x_\ab}^2+V_\ab(x_\ab)+\gamma)\,\chi(x_\ab)=0$ that 
2272: satisfy the following conditions:
2273: \bea
2274: c_{\alpha}\, \chi_\ab(0)     + d_{\alpha}\,\chi'_\ab(0) &=& 1 \\
2275: c_{\beta} \, \chi_\ab(l_\ab) - d_{\beta} \,\chi'_\ab(l_\ab) &=& 0
2276: % c_{\alpha }\cab (\alpha ) + 
2277: % d_{\alpha }\frac{ \D \cab }{ \D \xab } (\alpha ) = 1
2278: %   \qquad &;& \qquad
2279: % c_{\beta }\cab (\beta ) + 
2280: % d_{\beta }\frac{ \D \cab }{ \D \xba } (\beta ) = 0     
2281: % \label{p1}  \\
2282: % c_{\alpha }\cba (\alpha ) + 
2283: % d_{\alpha }\frac{ \D \cba }{ \D \xab } (\alpha ) = 0
2284: %  \qquad &;& \qquad
2285: % c_{\beta }\cba (\beta ) + 
2286: % d_{\beta }\frac{ \D \cba }{ \D \xba } (\beta ) = 1        
2287: \label{p4}  
2288: \eea
2289: 
2290: 
2291: \noindent
2292: We denote by $\Xi_\ab$ the Wronskian of $\cab$ and $\cba$.
2293: Following the same steps as before, we  get 
2294: the spectral determinant (up to a multiplicative constant)~:
2295: \be\label{p19} 
2296:  S(\gamma) =  \prod_{(\alpha\beta)}\frac{1}{ \Xi_\ab }\ \det M
2297: \ee
2298: where $M$ is the $V \times V$-matrix~:
2299: %\bea 
2300: % M_{\alpha\alpha } &=& 1+ t_{\alpha }
2301: %  \sum_\mu a_{\alpha\mu} \chi_{\alpha\mu} (0)   +
2302: %  w_{\alpha } \label{p12} \sum_\mu a_{\alpha\mu} \chi_{\alpha\mu}'(0)  \\
2303: % M_{\alpha\beta } &=& ( c_{\alpha } w_{\alpha } -  t_{\alpha } d_{\alpha } )
2304: % \Xi_\ab   \qquad  \mbox{ \ if $(\alpha\beta )$ is a bond}   \label{p13}    \\
2305: %                &=& 0 \qquad \ \mbox{ otherwise} \nonumber
2306: %\eea
2307: \begin{eqnarray} \label{p12}
2308:   M_{\ab} = 
2309:   \delta_{\ab}
2310:   \left(
2311:     1 + \sum_\mu a_{\alpha\mu} 
2312:         \left[
2313:           t_{\alpha }\,\chi_{\alpha\mu}(0)+w_{\alpha }\,\chi_{\alpha\mu}'(0)
2314:         \right]
2315:   \right)
2316:   + a_\ab\, 
2317:   \left[ c_{\alpha } w_{\alpha } -  t_{\alpha } d_{\alpha } \right]\,\Xi_\ab 
2318: \:.
2319: \end{eqnarray}
2320: The eq.~(\ref{n40}), expressing the determinant in terms of arc matrices, and
2321: the eq.~(\ref{p19}), expressing it in terms of vertex matrix, have been
2322: derived up to multiplicative constants independent on $\gamma$. We can
2323: establish a precise relation by comparing their behaviour for
2324: $\gamma\to\infty$. We end up with~:
2325: \begin{equation}\label{p20} 
2326:  \prod_{(\alpha\beta)}\frac{1}{ \wal }\ \det(C+DN)
2327:  =  \prod_{(\alpha\beta)}\frac{1}{ \Xi_\ab }\ \det M
2328: \:.
2329: \end{equation}
2330: 
2331: 
2332: \mathversion{bold}
2333: \subsubsection*{Free case ($V(x)=0$). Applications~: counting backtrackings}
2334: \mathversion{normal}
2335: 
2336: In this subsection we show an application of generalized boundary conditions
2337: to count backtrackings (the figure~\ref{fig:exampleqds} shows an 
2338: orbit with one backtracking).
2339: 
2340: We now consider  the case $V(x)\equiv0$ still with 
2341:  permutation-invariant boundary conditions.
2342: 
2343: \noindent
2344: With the notations
2345: $$
2346:     \eta_{\alpha } = \frac{c_{\alpha }   +  \sg d_{\alpha } }
2347:                             { c_{\alpha }  -  \sg d_{\alpha } }
2348:          \qquad , \qquad                      
2349:      \rho_{\alpha } = \frac{\mu_{\alpha }^- - \mu_{\alpha }^+ }
2350:                             {1+ m_{\alpha } \mu_{\alpha }^- }
2351:     \qquad \mbox{and} \qquad
2352: \mu_{\alpha }^{\pm } = \frac{t_{\alpha }   \pm  \sg w_{\alpha } }
2353:       { c_{\alpha }  \pm  \sg d_{\alpha } }
2354: $$
2355: eqs.~(\ref{p20}, \ref{n41}, \ref{relQDC}) lead to~:
2356: \be\label{p804} 
2357: \det (\ide - Q R) = 2^{-V} \prod_{\alpha } \left( \rho_{\alpha }
2358: \eta_{\alpha } \right) \prod_{(\alpha\beta)}\left( 
2359:  1- \eta_{\alpha } \eta_{\beta } \EXP{- 2 \sg \lab}   \right)  \; 
2360:  \det{M} 
2361: \ee
2362: and the $V\times V$-matrix $M$ takes the form~:
2363: \bea
2364: \label{p805}
2365: {M}_{\alpha\beta} = 
2366: \delta_{\alpha\beta} \left(
2367:   \frac{2}{ \rho_{\alpha }\eta_{\alpha } } - 
2368:   \frac{ m_{\alpha }}{\eta_{\alpha } } +  \frac{1}{ \eta_{\alpha }} 
2369:   \sum_\mu  a_{\alpha\mu}
2370:   \frac{1+\eta_{\alpha }\eta_{\mu} \EXP{-2\sg l_{\alpha\mu}}  }
2371:        {1-\eta_{\alpha }\eta_{\mu } \EXP{-2\sg l_{\alpha\mu}} }
2372: \right)
2373: -a_{\alpha\beta}
2374:    \frac{2\,  \EXP{- \sg l_{\alpha\beta }}                         }
2375:         {1-\eta_{\alpha }\eta_{\beta }\EXP{-2\sg l_{\alpha\beta }} }
2376: \:.\eea
2377: Note that the expression (\ref{defM}) is recovered 
2378: for $\eta_\alpha=1$ and $\rho_\alpha=2/(m_\alpha+\lambda_\alpha/\sqrt\gamma)$.
2379: 
2380: For permutation-invariant boundary conditions, the matrices $C$, $D$ and
2381:  $Q$  (eq.~(\ref{relQDC})) are block-diagonal. The block $Q_{\alpha }$ 
2382: takes the simple form:
2383: \be\label{p900}
2384:      Q_{\alpha } = \eta_{\alpha } 
2385:      \left( -\ide + \rho_{\alpha }  F_{\alpha } \right)
2386: \ee
2387: The only non-vanishing elements of the matrix $QR$ are~:
2388: \be\label{p901}
2389:     (QR)_{\alpha\beta,\mu\alpha}= \left(
2390:     \rho_{\alpha }\eta_{\alpha } - \eta_{\alpha } \; \delta_{\beta\mu } 
2391:  \right) \; \EXP{- \sg l_{\alpha\mu }} 
2392: \ee
2393: 
2394: \vskip.3cm
2395: 
2396: \noindent
2397: We call 
2398: $\rho_{\alpha }\eta_{\alpha }$  the transmission factor
2399: at vertex $\alpha $ and $\rho_{\alpha}\eta_{\alpha}-\eta_{\alpha}$
2400: the reflection factor.
2401: By using the same expansion as above we write
2402: \be\label{p902}
2403:  \det(\ide - QR ) 
2404: = \prod_{\widetilde{\cal C}} \left( 1- \mu(\widetilde{\cal C} )
2405:  \EXP{-\sqrt\gamma l(\widetilde{\cal C})} \right)
2406: \:,\ee
2407: where the product is taken over all primitive orbits $\widetilde{C}$ whose
2408: lengths are denoted by $l(\widetilde{C})$.
2409: An orbit being a succession of arcs
2410: $\ldots  ,\tau\alpha,\alpha\beta,\ldots $  with, in $\alpha$,
2411: a reflection (if $\tau =\beta $) or a transmission  (if $\tau \ne \beta $),
2412: the weight $\mu(\widetilde{C})$, in eq.~(\ref{p902}),
2413: will be the product of
2414: all the reflection -- or transmission -- factors along~$\widetilde{C}$.  
2415: 
2416: 
2417: \vspace{0.25cm}
2418: 
2419: \noindent 
2420: {\it Graphs with wires of equal lengths.--}
2421: We consider the case of equal lengths $\lab=l$~; then we can choose
2422: $\gamma=1$ without loss of generality and introduce the notation
2423: $u\equiv\EXP{-l}$.
2424: It is clear from eq.~(\ref{p900}) that a backtracking at vertex $\alpha$ 
2425: brings a factor $\eta_{\alpha}$ in the weight $\mu(\widetilde{C})$. 
2426: In order to count backtrackings one has to choose the boundary conditions
2427: $\rho_{\alpha }\eta_{\alpha }=1$ and $\eta_{\alpha}=\eta$. Eq.~(\ref{p804})
2428: takes the simple form~:
2429: \be\label{p903}
2430: \prod_{\widetilde{C}_m} \left( 1 - (1-\eta )^{n_R(\widetilde{C}_m )} u^m
2431:  \right) = (1- \eta^2 u^2 )^{B-V} \det 
2432:  \left( ( 1- \eta^2 u^2  ) \ide +\eta u^2\, Y - u\, A   \right)
2433:  \equiv  Z^{-1}
2434: \ee
2435: $m$ is the number of arcs of the primitive orbit $\widetilde{C}_m$ and
2436: $n_R(\widetilde{C}_m)$ is the number of reflections (backtrackings) 
2437: occuring along  $\widetilde{C}_m$. 
2438: $Y$ is the $V\times V$-matrix 
2439: $Y_{\alpha\beta }=\delta_{\alpha\beta } \; m_{\alpha }$ and $A$ is the
2440: adjacency matrix.
2441: 
2442: Setting $\eta=1$ implies $n_R(\widetilde{C}_m)=0$ in the left-hand side of
2443: eq.~(\ref{p903})~: we recover Ihara-Bass formula \cite{Iha66,Bas92,StaTer96} 
2444: where
2445: only primitive orbits without tails and backtrackings are kept. (Ihara
2446: \cite{Iha66} established this formula for a regular graph~; the proof for a
2447: general graph is given in refs.~\cite{Bas92,StaTer96} using a direct counting
2448: technique).
2449: 
2450: %Note that some functional relation between generating functions of lengths and
2451: %number of backtrackings of open paths has been derived in ref.~\cite{Bar99}
2452: %for regular graphs. When applied to closed orbits the functional relation is a
2453: %simple consequence of the relation (\ref{p903}).
2454:  
2455: \vspace{0.25cm}
2456: 
2457: Now, let us consider closed random walks with a given number
2458: of backtrackings.
2459: 
2460: Eq.~(\ref{p903}) provides a non-trivial generalization of the Ihara-Bass
2461: formula (an independent derivation is also given in ref.~\cite{Bar99}).
2462: As an application let us consider the problem of enumerating $m$-steps
2463: random walks with $p$-backtracking steps~\cite{WuKun99}.
2464: Taking $Z$ in (\ref{p903}), we get: 
2465: \be\label{p1949}
2466: u \; \frac{ \D \ln Z}{ \D u} = \sum_{m=2}^{\infty } \; \sum_{p=0}^m \; 
2467: \sum_{\alpha =1}^V \; N_m^p(\alpha ) \; (1-\eta )^p \; u^m 
2468: \ee
2469: \noindent 
2470: where $ N_m^p(\alpha ) $ is the number of $m$-steps closed 
2471: random walks on the graph starting at $\alpha$, with $p$ backtrackings.
2472: 
2473: \vspace{0.25cm}
2474: 
2475: \noindent 
2476: {\it Example~: The complete graph.--}
2477: For the complete graph\footnote{
2478:   The complete graph $K_V$ with $V$ vertices is the $V-1$-simplex~: 
2479:   each vertex is connected to all other vertices (see figure~\ref{fig:k5}).
2480: } $K_V$, we get the results: 
2481: \bea
2482:   N_2^0(\alpha) &=& 0                              \nonumber \\
2483:   N_3^0(\alpha) &=&    (V-1)(V-2)                  \nonumber \\
2484:   N_4^0(\alpha) &=&    (V-1)(V-2)(V-3)             \nonumber \\
2485:   N_5^0(\alpha) &=&    (V-1)(V-2)(V-3)(V-4)        \nonumber \\
2486:   N_6^0(\alpha) &=&    (V-1)(V-2)(V^3-9V^2+29V-32) \label{p1950} 
2487: \eea
2488: \noindent 
2489: and also:
2490: \bea
2491:  N_2^1(\alpha) &=&  N_3^1(\alpha ) \ = \ N_4^1(\alpha ) \ = \ 0   \nonumber \\
2492:  N_5^1(\alpha) &=&  5(V-1)(V-2)(V-3)     \nonumber \\
2493:  N_6^1(\alpha) &=&  6(V-1)(V-2)(V-3)^2  \label{p1952} 
2494: \eea
2495: Note that these expressions have been obtained when vertices are all 
2496: characterized by the same parameter $\eta$. Introducing different parameters
2497: $\eta_\alpha$ allows counting the backtrackings at a given vertex.
2498: 
2499: \begin{figure}[!ht]
2500: \begin{center}
2501: \includegraphics[scale=0.6]{k5.eps}
2502: \end{center}
2503: \caption{\it The complete graph $K_5$ (the $4$-simplex)~: 
2504:          all vertices are connected to each 
2505:          other by bonds of equal lengths.\label{fig:k5}}
2506: \end{figure}
2507: 
2508: 
2509: 
2510: \subsection{Quantum chaos on metric graphs}
2511: 
2512: It has been recently realized that metric graphs are interesting 
2513: models for  quantum chaos. The paper of Kottos \& Smilansky
2514: \cite{KotSmi97} has stimulated several works on spectral statistics and
2515: level correlations \cite{KotSmi99,BarGas00,Tan01b}, and  
2516: renewed the interest in the Roth trace formula \cite{Rot83}.
2517: Progress in the understanding of universality of spectral statistics for
2518: generic quantum graphs has been achieved by Gnutzmann \& Altland
2519: \cite{GnuAlt04}. One of the input of this work is the observation that the
2520: spectral average for a given graph with incommensurate bond lengths is
2521: equivalent to an average over a certain ensemble of unitarity matrices. Star
2522: graphs \cite{BerKea99,KeaMarWin03} have provided simple examples of systems
2523: with intermediate level statistics similar to the one observed in {\v S}eba
2524: billiards and a precise connection has been established in
2525: ref.~\cite{BerBogKea01}. There is however an interesting class of graphs which
2526: do not enter in this category \cite{Tan05}. Chaotic scattering and transport
2527: properties in open graphs (with some infinitely long wires) have also been
2528: studied in refs.~\cite{KotSmi00,BarGas01,KotSmi03}. Since 
2529: quantum chaos is not the
2530: central subject of our review,  this list of papers is not exhaustive~: we
2531: refer the reader interested in this topic to the review
2532: papers~\cite{KotSmi99,KotSmi03,Tan05} and the PhD thesis~\cite{Win03}.
2533: 
2534: 
2535: \subsection{Weak localization and Brownian motion on graphs\label{sec:wlrg}}
2536: 
2537: Our initial physical motivation for the study of the spectral determinant on
2538: graphs was based on the observation that the weak localization correction to
2539: the conductivity is directly expressed in terms of the spectral determinant.
2540: We see from eq.~(\ref{WL}) that
2541: \be\label{WLsd} 
2542: \smean{\Delta\sigma}
2543: = - \frac{2 e^2}{\pi\,{\rm Volume}}
2544: \drond{}{\gamma} \ln S(\gamma)   
2545: \:,\ee
2546: where the spectral parameter is related to the phase coherence
2547: length $\gamma=1/L_\varphi^2$.
2548: The expressions (\ref{WLsd},\ref{spedet},\ref{defM}), due to Pascaud \&
2549: Montambaux \cite{Pas98,PasMon99,AkkComDesMonTex00}, improve the
2550: approach initiated by Dou{\c c}ot \& Rammal \cite{DouRam85,DouRam86}.
2551: 
2552: 
2553: 
2554: \subsubsection*{Nonlocality of the quantum transport in arbitrary networks}
2555: 
2556: We must stress that the formula (\ref{WLsd}) corresponds to a uniform
2557: integration of the cooperon, defined as 
2558: $P_c(x,x)=\bra{x}\frac1{\gamma-\Delta}\ket{x}$, on
2559: the graph $\smean{\Delta\sigma}\propto-\int\D x\,P_c(x,x)$. This approach is
2560: limited to the case of regular graphs, where all wires play the same role. In
2561: other terms, for a nonregular network, the quantity (\ref{WLsd}) does not
2562: correspond to a quantity measured in a transport experiment. For arbitrary
2563: networks, the cooperon must be integrated over the network with appropriate
2564: nontrivial weights that depend on the topology of the whole network and the
2565: way it is connected to external contacts. Such a generalization was provided in
2566: ref.~\cite{TexMon04}.
2567: 
2568: 
2569: \subsubsection*{Windings in a loop connected to a network} 
2570: 
2571: As we have mentioned in the section~\ref{sec:brownian2},  magnetoconductance 
2572: oscillations due to a magnetic flux and winding properties are
2573: closely related.
2574: For example, if we consider an isolated ring of perimeter $L$ pierced by a 
2575: magnetic flux $\phi$, the well-known behaviour of the harmonics
2576: \begin{equation}
2577:   \label{eq:AAS}
2578:   \smean{\Delta\sigma_n}=\int_0^{2\pi}\frac{\D\theta}{2\pi}
2579:   \smean{\Delta\sigma(\theta)}\,\EXP{-\I n\theta}
2580:   \propto\EXP{-|n|L/L_\varphi}
2581: \end{equation} 
2582: is a direct consequence of the fact that the winding around the ring scales 
2583: with time as $n_t\propto t^{1/2}$ (normal diffusion).
2584: $\theta=4\pi\phi/\phi_0$ is the reduced flux  and $\gamma=1/L_\varphi^2$.
2585: This effect was predicted by Al'tshuler, Aronov \& Spivak (AAS) in 
2586: ref.~\cite{AltAroSpi81} and observed in
2587: experiments on cylinder films~\cite{ShaSha82,AltAroSpiShaSha82}.
2588: 
2589: Recently it has been noticed that the fact that the ring is connected to arms,
2590: which is necessary to perform a transport experiment, can strongly affect the
2591: harmonics. If we consider a ring connected to $N_a$ long arms, in the limit
2592: $L_\varphi\ll L$, the harmonics are still given by the AAS behaviour
2593: (\ref{eq:AAS}), however when the perimeter is smaller than $L_\varphi$, the
2594: harmonics behave as
2595: $\smean{\Delta\sigma_n}\propto\EXP{-|n|\sqrt{N_a L/L_\varphi}}$.
2596: To clarify the origin of this behaviour, the winding of Brownian trajectories
2597: around a ring connected to another network has been recently examined in
2598: ref.~\cite{TexMon05}. This analysis is based on the fact that the winding
2599: number distribution can be expressed in terms of the spectral determinant. Let
2600: us briefly describe this approach.
2601: 
2602: \begin{figure}[!ht]
2603: \begin{center}
2604: \includegraphics[scale=1]{ringwitharm.eps}
2605: \end{center}
2606: \caption{\it A diffusive ring attached to a long arm. The parameter $\lambda_1$
2607:          describes the boundary condition at vertex 1~: $\lambda_1=\infty$
2608:          for the Dirichlet boundary and $\lambda_1=0$ for the Neumann boundary
2609:          condition.
2610:          \label{fig:ringwitharm}}
2611: \end{figure}
2612: 
2613: \begin{figure}[!ht]
2614: \begin{center}
2615: \includegraphics[scale=0.5]{ringmesh2.eps}
2616: \hspace{1cm}
2617: \includegraphics[scale=0.5]{ringbethe.eps}
2618: \end{center}
2619: \caption{\it A diffusive ring attached to {\rm (a)~:} an infinite square 
2620:          lattice,
2621:          {\rm (b)~:} an infinite Bethe lattice of connectivity $z=3$. 
2622:          To ensure that the Bethe lattice is regular, it must be embedded
2623:          in a constant negative curvature surface.
2624:          \label{fig:ringmesh}}
2625: \end{figure}
2626: 
2627: We consider a ring to which is attached an arbitrary network at vertex $0$
2628: (the figures \ref{fig:ringwitharm} and \ref{fig:ringmesh} give examples
2629: of such a situation). Our aim is to understand how the winding around the ring
2630: is affected by the presence of the network. To answer this question we
2631: introduce the probability to start from a point $x$ on the graph and come back
2632: to it after a time $t$, conditioned to wind $n$ times around the loop~:
2633: \be 
2634: {\cal P}_n(x,t|x,0) 
2635: = \int_{x(0)=x}^{x(t)=x}{\cal D}x(\tau)
2636: \:\EXP{-\frac14\int_0^t\D\tau\,\dot x(\tau)^2} 
2637: \delta_{n,{\cal N}[x(\tau)]} 
2638: \ee
2639: where ${\cal N}[x(\tau)]$ is the winding number of the path
2640: $x(\tau)$ around the loop.  For simplicity, we consider the case where the
2641: initial point is the vertex $0$.
2642: The computation of ${\cal P}_n(0,t|0,0)$ requires some local information
2643: (the eigenfunctions of the Laplace operator at point $x=0$). 
2644: On the other hand the spectral determinant encodes a global information
2645: since it results from a spatial integration of the Green's function
2646: of the Laplace operator over the network. 
2647: However we have shown in ref.~\cite{TexMon05} that if we consider mixed 
2648: boundary conditions at vertex $0$ described by a parameter $\lambda_0$
2649: (see the definition of the boundary conditions given above), 
2650: we can extract the probability of interest\footnote{
2651:   The fact that some ``local'' information, such as ${\cal P}_n(0,t|0,0)$,
2652:   can be extracted from a more ``global'' object, like the spectral 
2653:   determinant, by introducing the mixed boundary conditions with parameter
2654:   $\lambda_0$ has been also used in the context of scattering theory in
2655:   graphs in refs.~\cite{TexBut03,TexDeg03}. 
2656:   The derivative in eq.~(\ref{P0S}) can be understood as a functional
2657:   derivative since $\lambda_0$ plays a role similar to the weight of 
2658:   a $\delta$-potential at $x_0$. 
2659:   The use of functional derivatives in scattering theory for 
2660:   mesoscopic systems has been fruitfully used by
2661:   B\"uttiker and co-workers (see ref.~\cite{But02} and references therein).
2662: } as follows~:
2663: \be\label{P0S}
2664: \int_0^\infty\D t\,{\cal P}_n(0,t|0,0)\,\EXP{-\gamma t}
2665: =\int_{0}^{2\pi}\frac{\D\theta}{2\pi}\:\EXP{-\I n\theta}\:
2666: \frac{\D}{\D\lambda_0}\ln S^{(\lambda_0)}(\gamma)\Big|_{\lambda_0=0}
2667: \:,\ee
2668: where $S^{(\lambda_0)}(\gamma)$ is the spectral determinant for 
2669: mixed boundary condition at vertex $0$ and with a magnetic flux 
2670: $\theta$ piercing the ring.
2671: This formula allows us to express the probability as
2672: \be\label{Wn00}
2673:   \int_0^\infty\D t\:{\cal P}_n(0,t|0,0)\:\EXP{-\gamma t} = 
2674:   \frac{1}{2\sqrt\gamma}
2675:   \frac{\sinh\sqrt\gamma L}{\sinh\sqrt{\gamma} L_{\rm eff}(\gamma)}
2676:   \:\EXP{-|n|\sqrt{\gamma} L_{\rm eff}(\gamma)} 
2677: \:,\ee
2678: where all the information about the nature of the network attached to the 
2679: ring is contained in the effective perimeter $L_{\rm eff}(\gamma)$, defined 
2680: as~:
2681: \be\label{leffnet}
2682: \cosh\sqrt{\gamma} L_{\rm eff}(\gamma)
2683: =\cosh\sqrt{\gamma}L
2684: +\frac{\sinh\sqrt{\gamma}L}{2\,({M}_{\rm net}^{-1})_{00}}
2685: \:.\ee
2686: The matrix $M_{\rm net}$ describes the network in the absence of the loop
2687: (it is given by eq.~(\ref{defM})~; for the network of 
2688: figure~\ref{fig:ringmesh}a, ${M}_{\rm net}$ describes the infinite 
2689: square network without the ring).
2690: The matrix element $({M}_{\rm net}^{-1})_{00}$ has a clear
2691: meaning~: it is the Green function of the Laplace operator in the 
2692: network (without the loop) computed  at the position where 
2693: the ring is attached~:
2694: $\sqrt{\gamma}(M_{\rm net}^{-1})_{00}=\bra{0}\frac1{\gamma-\Delta}\ket{0}$. 
2695: 
2696: The effective perimeter $L_{\rm eff}(\gamma)$ probes the winding at time
2697: scale $t\sim1/\gamma$~: precisely, the winding number scales with time as
2698: $n_t\sim\sqrt{t}/L_{\rm eff}(1/t)$.
2699: 
2700: \vspace{0.25cm}
2701: 
2702: Two interesting examples are
2703: 
2704: \vspace{0.25cm}
2705: 
2706: \noindent$\bullet$ {\it Ring attached to an infinite wire 
2707: (figure~\ref{fig:ringwitharm}).--}
2708: When a wire of length $b$ with the Dirichlet boundary at one end is 
2709: attached to the ring, it is easy to see that eq.~(\ref{defM}) gives
2710: $({M}_{\rm net}^{-1})_{00}=\tanh\sqrt\gamma b\leadto{b\to\infty}1$.
2711: At large time $t\gg L^2$, the effective perimeter behaves 
2712: as $L_{\rm eff}\simeq\sqrt{L}\gamma^{-1/4}$. This behaviour is related 
2713: to a scaling of the winding number with  time
2714: \be\label{scalingwire}
2715: n_t \propto t^{1/4}
2716: \:.\ee
2717: The full distribution for the winding number $n$ is given in
2718: ref.~\cite{TexMon05}. The exponent $1/4$  that characterizes anomalously slow
2719: winding around the loop originates from the fact that the diffusive trajectory
2720: spends a long time in the infinite wire, which increases the effective
2721: perimeter at such time scales. This problem is also related to the anomalous
2722: diffusion along the skeleton of a comb, studied in
2723: refs.~\cite{WeiHav86,BalHavWei87} by different methods. It is interesting to
2724: use this picture. Let us consider a random walk along the sites of a line where
2725: the diffusive particle is trapped during a time $\tau$ on each site. The
2726: trapping time is distributed according to a broad distribution
2727: $P_1(\tau)\propto\tau^{-1-\mu}$ with $0<\mu<1$. It follows that the distance
2728: scales with time as $n_t\sim t^{\mu/2}$ \cite{BouGeo90}. If we go back to the
2729: problem of diffusion along the skeleton of a comb (or the winding in the ring
2730: connected to the long arm), the arm plays the role of the trap. The
2731: distribution of the trapping time is given by the first return probability of
2732: the one-dimensional diffusion~:  $P_1(\tau)\propto\tau^{-3/2}$ and we recover
2733: eq.~(\ref{scalingwire}).
2734: 
2735: \vspace{0.25cm}
2736: 
2737: \noindent$\bullet$ {\it Ring attached to a square network
2738: (figure~\ref{fig:ringmesh}a).--}
2739: When studying the winding around the loop, it is important to know
2740: whether the Brownian motion inside the network attached to the ring is
2741: recurrent or not. Let us consider the case where the network attached is a
2742: $d$-dimensional hypercubic network. For $d>2$ the Brownian motion is
2743: known to be transient where as, for $d=2$ it is neighbourhood recurrent
2744: (in $d=1$ the Brownian motion is pointwise recurrent). Therefore we
2745: expect the dimension 2 to play a special role. In the large time limit, when 
2746: $t\gg{L^2}$ and $t\gg{a^2}$ ($a$ is the lattice spacing), we find an 
2747: effective length
2748: $\sqrt\gamma\, L_{\rm eff}\simeq\sqrt{\frac{2\pi L}{a\ln(4/\sqrt{\gamma}a)}}$,
2749: that corresponds to a scaling of the winding around the loop
2750: \be\label{scalingsn}
2751: n_t \propto (\ln t)^{1/2}
2752: \:.\ee
2753: This result can be obtained in the same way as for the ring connected to 
2754: the arm. This time the plane acts as a trap. 
2755: The distribution of the trapping time is given by the first return
2756: probability on a square lattice, which is known to behave at large times 
2757: like $P_1(\tau)\propto1/(\tau\ln^2\tau)$ \cite{BarTac93}, from which
2758: we can recover eq.~(\ref{scalingsn}).
2759: 
2760: 
2761: \subsubsection*{Occupation time and local time distribution on a graph} 
2762: 
2763: Another set of problems concern occupation times~: {\it i.e.} time spent by a
2764: Brownian particle in a given region. An example of such a problem is provided
2765: by the famous arc-sine law for the 1d Brownian motion that gives the
2766: distribution of the time spent by a Brownian motion
2767: $(x(\tau),0\leq\tau\leq{t}\,|\,x(0)=0)$ on the half line $\RR^+$. This result
2768: was derived long ago by P.~L\'evy \cite{Lev48}. It has been extended by Barlow,
2769: Pitman \& Yor \cite{BarPitYor89} for a particular graph (star graph with arms
2770: of infinite lengths)~: instead of an infinite line, these authors consider $n$
2771: semi-infinite lines originating from the same point and study the joint
2772: distribution of the times spent on each branch. More recently, this problem
2773: has been reconsidered in the case of arbitrary graphs \cite{Des02}. It may be
2774: stated as follows~: consider a Brownian motion $x(\tau)$ on a graph, starting
2775: from a point $x_0$ at time $0$ and arriving at a point $x_1$ at time $t$. Let
2776: $T_{\ab}$ denotes the time spent on the wire $(\ab)$. This functional is
2777: defined as $T_{\ab}[x(\tau)]=\int_0^t\D\tau\,{\rm Y}_{\ab}(x(\tau))$ where the
2778: function ${\rm Y}_{\ab}(x)$ is $1$ for $x\in(\ab)$ and $0$ otherwise. Our aim
2779: is to compute the Laplace transforms of the joint distribution
2780: \begin{equation}
2781:   \label{eq:octimedef}
2782:   \mean{ \EXP{-\sum_{(\ab)} \xi_\ab\,T_{\ab}} }
2783:   = \int\D x_1\, {\cal F}(x_1,x_0;t;\{\xi_\ab\}) 
2784: \end{equation}
2785: with
2786: \begin{equation}
2787:   {\cal F}(x_1,x_0;t;\{\xi_\ab\})
2788:   =\int_{x(0)=x_0}^{x(t)=x_1} {\cal D}x(\tau)\,
2789:   \EXP{  -\frac14\int_0^t\D\tau\,\dot x^2
2790:          -\sum_{(\ab)} \xi_\ab\,T_{\ab}[x(\tau)]
2791:       }
2792: \:.\end{equation}
2793: The definition (\ref{eq:octimedef}) takes into account averaging over the
2794: final point $x_1$. A conjugate parameter $\xi_\ab$ is introduced for each
2795: variable $T_{\ab}$, {\it i.e.} each bond. A closed expression of the Laplace
2796: transform of the joint distribution (\ref{eq:octimedef}) has been derived in
2797: ref.~\cite{Des02}. The result is given as a ratio of two determinants,
2798: an expression reminiscent of the one of Leuridan \cite{Leu00}, although the 
2799: connection is not completely clear.
2800: 
2801: Similar methods have been also applied in ref.~\cite{ComDesMaj02} to study the
2802: distribution of the local time
2803: $T_{x_0}[x(\tau)]=\int_0^t\D\tau\,\delta(x(\tau)-x_0)$.
2804: 
2805: \vspace{0.25cm}
2806: 
2807: A simplification occurs when the final point coincides with the initial point.
2808: Then it is possible to relate the characteristic function to a single spectral
2809: determinant. We do not develop the general theory here but instead consider an
2810: example close to the one studied in the previous subsection.  Let us consider
2811: the graph in figure~\ref{fig:ringwitharm} and ask the following question~: for
2812: a Brownian motion starting from $x_0$ at time $0$ and coming back to it at 
2813: time $t$,
2814: what is the distribution of the time $T_{\rm arm}[x(\tau)]$ spent in the arm
2815: if in addition the Brownian motion is constrained to turn $n$ times around the
2816: ring~? It is natural to introduce the following function~:
2817: \begin{equation}
2818:   \label{eq:winoc}
2819:   {\cal F}_n(x_0,x_0;t,\xi) =
2820:   \int_{x(0)=x_0}^{x(t)=x_0} {\cal D}x(\tau)\,
2821:   \EXP{  -\int_0^t\D\tau\,
2822:         \left( 
2823:           \frac14\dot x^2 + \xi\,{\rm Y}_{\rm arm}(x)
2824:         \right)
2825:       }\:
2826:   \delta_{n,{\cal N}[x(\tau)]} 
2827: \end{equation}
2828: where ${\cal N}[x(\tau)]$ is the winding number around the ring.
2829: This function is related to the Laplace transform of the distribution
2830: of the functional $T_{\rm arm}[x(\tau)]$ 
2831: \begin{equation}
2832:   \label{disTbras}
2833:   \mean{\EXP{-\xi\, T_{\rm arm}[x]}}_{{\cal C}_n} 
2834:   = \frac{{\cal F}_n(x_0,x_0;t,\xi)}{{\cal F}_n(x_0,x_0;t,0)}
2835: \:,\end{equation}
2836: where $\mean{\cdots}_{{\cal C}_n}$ denotes averaging over curves of 
2837: winding~$n$. The denominator ensures normalization.
2838: The Laplace transform of (\ref{eq:winoc}) is given by a relation similar
2839: to eq.~(\ref{P0S})
2840: \begin{equation}
2841:   \int_0^\infty\D t\,{\cal F}_n(x_0,x_0;t,\xi)\,\EXP{-\gamma t}
2842:   =\int_{0}^{2\pi}\frac{\D\theta}{2\pi}\:\EXP{-\I n\theta}\:
2843:   \frac{\D}{\D\lambda_0}\ln S^{(\lambda_0)}(\gamma)\Big|_{\lambda_0=0}  
2844: \end{equation}
2845: where the appropriate spectral determinant is built as follows~:
2846: ({\it i}) since the starting point is fixed at $0$, we introduce mixed 
2847: boundary conditions at this point, with a parameter $\lambda_0$ that will be
2848: used to extract the ``local information''.
2849: ({\it ii}) A magnetic flux $\theta$ is introduced (conjugate to the winding
2850: number).
2851: ({\it iii}) The spectral parameter is shifted in the arm as 
2852: $\gamma\to\gamma+\xi$ to introduce the variable $\xi$ conjugate to the time 
2853: $T_{\rm arm}$.
2854: 
2855: We choose the vertex $0$ as initial condition and impose the Dirichlet
2856: boundary condition (which is achieved by setting $\lambda_1=\infty$) at the
2857: end of the arm of length $b$. The spectral determinant is found
2858: straightforwardly (for an efficient calculation of $S(\gamma)$ for a graph
2859: with loops, see ref.~\cite{Pas98} or appendix~C of
2860: ref.~\cite{AkkComDesMonTex00})~:
2861: \begin{equation}
2862:   S^{(\lambda_0)}(\gamma) =
2863:   \frac{\sinh\sqrt{\gamma} L\,\sinh\sqrt{\gamma+\xi}\,b}
2864:        {\sqrt{\gamma}\sqrt{\gamma+\xi}}
2865:   \left[
2866:     2\sqrt{\gamma}\frac{\cosh\sqrt{\gamma} L-\cos\theta}{\sinh\sqrt{\gamma} L}
2867:     +\lambda_0
2868:     +\sqrt{\gamma+\xi}\coth\sqrt{\gamma+\xi}\,b
2869:   \right]
2870: \end{equation}
2871: The terms in the brackets correspond to the matrix\footnote{
2872:   The introduction of the conjugate parameters $\{\xi_\ab\}$ corresponds to
2873:   shift the spectral parameter $\gamma$ on each wire as
2874:   $\gamma\to\gamma+\xi_\ab$.  The matrix to be generalized is not $M$, given
2875:   by eq.~(\ref{defM}), but ${\cal M}=\sqrt{\gamma}M$.
2876: } ${\cal M}$ (since $\lambda_1=\infty$ we 
2877: can consider only the element ${\cal M}_{00}$). The first term is the 
2878: contribution of the loop, the second comes from the boundary condition and 
2879: the last one comes from the arm. It immediately follows that 
2880: \begin{equation}
2881:   \label{relation1}
2882:   \int_0^\infty\D t\,{\cal F}_n(0,0;t,\xi)\,\EXP{-\gamma t}
2883:   =  
2884:   \frac{1}{2\sqrt\gamma}
2885:   \frac{\sinh\sqrt\gamma L}{\sinh\sqrt{\gamma} L_{\rm eff}}
2886:   \:\EXP{-|n|\sqrt{\gamma} L_{\rm eff}} 
2887: \end{equation}
2888: with
2889: \begin{equation}
2890:   \label{Leffwinoc}
2891:   \cosh\sqrt\gamma L_{\rm eff} = \cosh\sqrt\gamma L
2892:   +\frac12\sqrt{1+{\xi}/{\gamma}}\,\sinh\sqrt{\gamma} L\,
2893:    \coth\sqrt{\gamma+\xi}\,b
2894:   \:.
2895: \end{equation}
2896: Let us consider an infinitely long arm $b\to\infty$. If we are interested on
2897: time scales $t\gg\tau_L$, where $\tau_L=L^2$ is the Thouless time over which
2898: the ring is explored, eq.~(\ref{Leffwinoc}) gives 
2899: $\sqrt{\gamma}L_{\rm eff}\simeq(\gamma+\xi)^{1/4}\sqrt{L}$, therefore from
2900: eq.~(\ref{relation1}) we see that 
2901: ${\cal F}_n(0,0;t,\xi)\simeq\EXP{-\xi t}\,{\cal F}_n(0,0;t,0)$.
2902: The inverse Laplace transform of eq.~(\ref{disTbras}) leads to
2903: $\smean{\delta(T-T_{\rm arm}[x])}_{{\cal C}_n}\simeq\delta(T-t)$, which means
2904: that the Brownian motion spends almost all the time in the arm (this simple
2905: result confirms the picture presented to explain the scaling of the winding
2906: around the ring of the form $n_t\propto t^{1/4}$).
2907: 
2908: 
2909: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2910: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2911: 
2912: \section{Planar Brownian motion and charged particle in random magnetic field\label{sec:magimp}}
2913: 
2914: A model of random magnetic field describing a charged particle moving in a
2915: plane and subjected to the random magnetic field 
2916: ${\cal B}(\vec r)=\vec\nabla\times \vec A(\vec r)$ due to an ensemble of
2917: magnetic Aharonov-Bohm vortices is described by the following
2918: Hamiltonian~\cite{DesFurOuv95,Fur97}~:
2919: \bea\label{Him}
2920: H   = \frac12\left(\vec p - \vec A(\vec r) \right)^2 
2921:     + \frac12 \, {\cal B}(\vec r) 
2922:     = \frac12
2923: \left(
2924:   \vec p - \alpha\sum_i 
2925:   \frac{\vec u_z\times(\vec r-\vec r_i)}{(\vec r-\vec r_i)^2}
2926: \right)^2 
2927: +\pi\alpha\sum_i \delta(\vec r-\vec r_i)
2928: \eea
2929: where $0\leq\alpha<1$. We have set $\hbar=e=m=1$.
2930: $\alpha$ is the magnetic flux per vortex, in unit of the quantum flux 
2931: $\phi_0=h/e$~:  $\alpha=\phi/\phi_0=\phi/(2\pi)$.
2932: The model is periodic in $\alpha$ with a period $1$. 
2933: $\vec u_z$ is the unit vector perpendicular to the plane. $\vec r_i$ are the 
2934: positions of the vortices in the plane~: they are uncorrelated random  
2935: variables (Poisson distribution).
2936: The $\delta$ interactions (coupling to the magnetic field) are necessary 
2937: in order to define properly the model 
2938: \cite{JacPi90,McCOuv91,Ouv94,BerLoz94,ComMasOuv95}.
2939: 
2940: By comparison with scalar impurities discussed in the introduction, the
2941: magnetic nature of the scatterers gives rise to rather different
2942: properties\footnote{
2943:   For scalar impurities, the presence of a magnetic field also strongly
2944:   affects the spectral properties. The case of a Gaussian disorder projected
2945:   in the lowest Landau level (LLL) of a strong magnetic field has been studied
2946:   in ref.~\cite{Weg83}. Other disordered potentials have been considered
2947:   later.  In particular it was shown in ref.~\cite{BreGroItz84} 
2948:   that for a weak density of impurities, the spectrum can display power law
2949:   singularities. A physical interpretation of such power law singularities has
2950:   been discussed by Furtlehner in ref.~\cite{Fur97,Fur00}.  The relation
2951:   between the model of scalar impurities projected in the LLL of a strong
2952:   magnetic field and the model of magnetic vortices was discussed in
2953:   refs.~\cite{DesFurOuv96,Fur97}.  Finally we mention a recent work on
2954:   Lifshitz tails in the presence of magnetic field \cite{LesWar04}.
2955: }~: it was shown in ref.~\cite{DesFurOuv95} that the spectrum is
2956: reminiscent of a Landau spectrum with Landau levels broadened by disorder, in
2957: the limit of vanishing magnetic flux $\alpha\to0$ (see below). This analysis
2958: was performed using perturbative arguments supported by numerical
2959: simulations.
2960: The latter use a relation between the average DoS of
2961: this particular model and some winding properties of the Brownian motion.
2962: 
2963: 
2964: Let $Z(t)$ denotes the partition function for a given distribution of vortices
2965: and $Z_0(t)$ the partition function without fluxes. The ratio of partition
2966: functions can be written as a ratio of two path integrals
2967: \be\label{pi1}
2968: \frac{Z(t)}{Z_0(t)}=
2969: \frac{
2970:   \int \D \vec a 
2971:   \int_{\vec r(0)=\vec a}^{\vec r(t)=\vec a}{\cal D}\vec r(\tau) 
2972:   \:\EXP{\int_0^t (-\frac{1}{2} \dot{\vec r}^2 + \I\vec A\cdot\dot{\vec r}) 
2973:     \D\tau }
2974:      }
2975:      {
2976:   \int \D\vec a
2977:   \int_{\vec r(0)=\vec a}^{\vec r(t)=\vec a}{\cal D}\vec r(\tau) 
2978:   \:\EXP{-\int_0^t \frac{1}{2}\dot{\vec r}^2  \D\tau } 
2979:      }
2980: =
2981: \left\langle 
2982:   \EXP{ \I\oint_{\cal C}  \vec A \cdot \D\vec r}
2983: \right\rangle_{\cal C}
2984: \:,\ee
2985: where $Z_0(t)={V}/({2\pi t})$ is the free partition function ($V$ is the
2986: (infinite) area of the plane) and $\smean{\cdots}_{\cal C}$ stands for 
2987: averaging over all closed Brownian curves of the plane.
2988: 
2989: In order to average $Z(t)$ over the Poissonian distribution of vortices, let
2990: us consider, for a while, our problem on a square lattice with lattice spacing
2991: $a$. Let $N_i$ be the number of vortices in square $i$. 
2992: the magnetic flux through any closed random walk ${\cal C}$ on this lattice 
2993: can be written as~:
2994: \be\label{cl}
2995: \oint_{\cal C}  \vec A \cdot \D \vec r = \sum_i 2 \pi \alpha N_i n_i 
2996: \ee
2997: where $n_i$ is the number of times the square $i$ has been wound around by 
2998: ${\cal C}$.
2999: 
3000: Averaging $Z(t)$ with the Poisson distribution ($\mu$ being the mean density of
3001: vortices)
3002: \be\label{pois}
3003:  P(N_i) \; = \; \frac{ ( \mu a^2 )^{N_i} }{N_i!}\,\EXP{- \mu a^2 }  
3004: \ee
3005: gives
3006: \be\label{zbar}
3007: \mean{ Z(t) }_{ \{\vec r_i\} } 
3008: =Z_0(t) 
3009: \left\langle 
3010:   \EXP{ \I\oint_{\cal C}  \vec A \cdot \D\vec r}
3011: \right\rangle_{ {\cal C},\{\vec r_i\} }
3012: =
3013: Z_0(t) 
3014: \left\langle   
3015: \exp\left(     
3016:  \mu \sum_n S_n(\EXP{2\I\pi\alpha n} - 1)
3017:  \right) 
3018: \right\rangle_{\cal C}
3019: \ee
3020: where $\smean{\cdots}_{\{\vec r_i\}}$ denotes averaging over the positions
3021: $\{\vec r_i\}$ of the vortices.
3022: 
3023: The quantity $S_n$ denotes the area of the locus of points around which the
3024: curve ${\cal C}$ has wound $n$ times. This result was derived for a random
3025: walk on a lattice but is also obviously valid off lattice as well, for 
3026: Brownian curves. The winding sectors of a closed
3027: curve are displayed in figure~\ref{fig:jean1}.
3028: 
3029: \begin{figure}[!ht]
3030: \begin{center}
3031: \includegraphics[scale=.4,angle=0]{wind_sect.eps}
3032: %\hspace{1cm}
3033: %\includegraphics{winding_sn.eps}
3034: \caption{
3035:   \it A closed curve with its $n$-winding sectors. The label $n$
3036:   of a sector is  the winding number of any point inside 
3037:   this sector. \label{fig:jean1}}
3038: \end{center}
3039: \end{figure}
3040: 
3041: \noindent
3042: The random variable $S_n$ scales like $t$.  In particular, we know
3043: \cite{ComDesOuv90,GroFri03} its expectation $\smean{S_n}=\frac{t}{2\pi n^2}$.
3044: Moreover, it has been shown in  \cite{Wer94} that the variable
3045: $n^2S_n$ becomes more and more peaked when $n$ grows:
3046: $
3047: P(n^2 S_n = X) \leadto{n\to\infty} \delta(X-\frac{t}{2\pi })
3048: $ (the average area of the $n=0$-sector has been recently studied in 
3049: ref.~\cite{GarTru05}).
3050: 
3051: Thus, extracting   $t$, eq.~(\ref{zbar}) rewrites as
3052: \be\label{zbar1}
3053: \mean{ Z(t) }_{ \{\vec r_i\} }  = Z_0(t) 
3054: \int\D S\D A\:P(S,A) \:\EXP{- \mu t ( S + \I A )}
3055: \equiv 
3056: Z_0(t) 
3057: \left\langle \EXP{- \mu t ( S + \I A )}\right\rangle_{\cal C}
3058: \:,\ee
3059: where  $P(S,A)$ is the joint distribution of the 
3060: rescaled ($t$ independent) variables $S$ and $A$ defined as
3061: \bea
3062: \label{s} 
3063: S&=&\frac{2}{t} \sum_n S_n \sin^2 (\pi \alpha n)\\
3064: \label{a}   
3065: A&=&\frac{1}{t} \sum_n S_n \sin (2\pi \alpha n) 
3066: \:.\eea
3067: The averages of these two variables are given by~:
3068: \bea
3069:    \mean{S} &=& \pi\alpha (1 - \alpha) 
3070: \\
3071:    \mean{A} &=& 0
3072: \:.\eea
3073: With eq.~(\ref{zbar1}), we observe that $\frac1V\mean{Z(t)}_{\{\vec r_i\}}$
3074: has the scaling form ${F(\mu t)}/{t}$. Thus, its inverse Laplace transform,
3075: the average density of states per unit area, is a function of only $E/\mu$ and
3076: $\alpha$. Moreover, it is easy to realize that $\mean{Z(t)}_{\{\vec r_i\}}$ is
3077: even in $\alpha$ (each Brownian curve in $\{{\cal C}\}$ comes with its time
3078: reversed) and periodic in $\alpha$ with period 1. Thus, one can restrict to
3079: $0\leq\alpha\leq\frac{1}{2}$. Let us focus on the two limiting cases 
3080: $\alpha\to0$ and $\alpha={1}/{2}$.
3081: 
3082: \vspace{0.25cm}
3083: 
3084: \mathversion{bold}
3085: \noindent$\bullet$ {\bf Limit $\alpha \to 0$~: Landau spectrum.}\\
3086: \mathversion{normal}
3087: When  $\alpha \to 0$, a careful analysis shows that
3088: \be\label{atou0}
3089: \mean{Z(t)}_{ \{\vec r_i\} } 
3090: \simeq 
3091: Z_{\smean{{\cal B}}} \: \EXP{-\frac12\smean{{\cal B}}t }
3092: \ee
3093: where 
3094: $
3095: Z_{\smean{{\cal B}}}=Z_0(t)
3096: \frac{\smean{{\cal B}}t/2}{\sinh(\smean{{\cal B}}t/2)}
3097: $
3098: is the partition function of a charged particle in a uniform magnetic field
3099: $\mean{{\cal B}}=2\pi\mu\alpha$ (we recall that the impurity $i$ carries a
3100: magnetic field $2\pi\alpha\delta(\vec r-\vec r_i)$). The system of random
3101: vortices is, thus, equivalent to the uniform average magnetic field, albeit 
3102: with  an additional positive shift in the Landau spectrum~: the 
3103: inverse Laplace transform of the partition function (\ref{atou0})
3104: gives the Landau spectrum made of equally spaced infinitely degenerated
3105: levels. This corresponds to the oscillating behaviour shown on
3106: figure~\ref{fig:dosimpmag}.  The origin of the shift can be traced back to
3107: the presence of the repulsive $\delta$ interactions that have been added to
3108: the Hamiltonian to define properly the model.
3109:    
3110: \vspace{0.25cm}
3111: 
3112: \mathversion{bold}
3113: \noindent$\bullet$ 
3114: {\bf Half quantum flux vortices ($\alpha=1/2$) -- The spectral singularity at $E=0$.}\\
3115: \mathversion{normal}
3116: In this case the variable (\ref{a}) vanishes, $A\equiv0$, implying that
3117: $
3118: \mean{Z(t)}_{\{\vec r_i\}} = Z_0(t) 
3119: \smean{ \exp \left( - \mu t  S  \right) }_{\cal C}
3120: $ 
3121: where now
3122: \be\label{snodd}
3123:  S= \frac{2}{t} \sum_{n\ {\rm odd}} S_n  
3124: \:.\ee
3125: Performing  
3126: the inverse Laplace transform, we get the average density of states
3127:   per unit area:
3128: \be\label{dos1/2} 
3129: \left\langle \rho (E)
3130:  \right\rangle = \rho_0 (E)\ \int_0^{E/\mu} \D S\, P(S)
3131: \ee
3132: where $\rho_0(E)=\frac{1}{2\pi}$  is the free density of states per unit area
3133: and  $P(S)$ is the probability distribution of $S$.
3134: 
3135: \begin{figure}[!ht]
3136: \begin{center}
3137: \includegraphics[scale=.5,angle=0]{jean_im05fit.eps}
3138: \caption{\it 
3139:  The  average density of states at $\alpha =0.5$ (the free density of states
3140:  is constant $\rho_0=1/(2\pi)$). Circles are  simulation
3141:  results. The full line is the fit discussed in the  text.
3142:  \label{fig:jean2}}
3143: \end{center}
3144: \end{figure}
3145: 
3146: We may use this result to determine the nature of the singularity in the
3147: average DoS. As displayed in figure~\ref{fig:jean2} (where we have taken
3148: $\mu=1$), $\smean{\rho(E)}$ increases monotonically from $0$ to $\rho_0(E)$
3149: with a depletion of states at the bottom of the spectrum. Circles are the 
3150: result
3151: of numerical simulations (on a 2D square lattice, we have generated 10000
3152: closed random walks of 100000 steps each). The full line is a low-energy
3153: fit of the quantity $\smean{\rho(E)}/\rho_0$ by the function
3154: \begin{equation}
3155:   \label{specsing}
3156:   \frac{\rho_{\rm fit}(E)}{\rho_0} = a\, \EXP{-b(\mu/E)^2}
3157: \end{equation}
3158: with $a=2.8$ and $b=1.4$ (see figure~\ref{fig:jean2}).
3159: It is worth stressing that this behaviour is quite different from the one 
3160: expected for a disorder due to a scalar potential. The famous Lifshitz 
3161: argument, applicable to the low energy DoS for a low concentration of 
3162: {\it scalar} impurity, leads instead to\footnote{
3163:   The Lifshitz argument applies to the DoS for the random Hamiltonian
3164:   $H=-\Delta+\sum_i u(\vec r-\vec r_i)$ in dimension~$d$, 
3165:   where $u(\vec r)$ is a sharply peaked scalar potential.
3166:   For a low density of impurities $\mu$, the DoS behaves as 
3167:   $\rho_{\rm Lif.}(E) \sim \exp({-c_d\,\mu/E^{d/2}})$ at low energy.
3168:   \label{footnoteLif}
3169: }
3170: a behaviour 
3171: \begin{equation}
3172: \rho_{\rm Lif.}(E) \sim \EXP{-c_2\,\mu/E}\:,
3173: \end{equation}
3174: where $c_2$ is a constant. Our choice for this fit is motivated by a recent
3175: numerical work \cite{Ric03} concerning the area $\cal{A}$ of the outer
3176: boundary of planar random loops. In this work, the author suggests that the
3177: limit distribution of $\cal{A}$ is the Airy distribution implying, for small
3178: $\cal{A}$ values, a behaviour of the type 
3179: $\exp(-\mbox{const.}/{\cal A}^2)/{\cal A}^2$. Remarking that 
3180: ${\cal A}=\sum_{n\ {\rm even}}S_n+\sum_{n\ {\rm odd}}S_n$, it is natural to
3181: expect for the distribution of the random variable $S$, eq.~(\ref{snodd}), a
3182: behaviour at small $S$ that is roughly given by $\exp(-\mbox{const.}/S^2)$.
3183: Thus, we deduce the form $\rho_{\rm fit}(E)$ for the low-energy fit of
3184: $\smean{\rho(E)}$.
3185: 
3186: \vspace{0.25cm}
3187: 
3188: \mathversion{bold}
3189: \noindent$\bullet$ {\bf Transition between $\alpha=0.5$ and $\alpha\to0$.}\\
3190: \mathversion{normal}
3191: Finally, when $\alpha$ grows from $0$ to $0.5$, the oscillations in the
3192: spectrum must disappear at some critical value $\alpha_c$ (see
3193: figure~\ref{fig:dosimpmag}). Numerical simulations, specific heat
3194: considerations and, also, diagrammatic expansions \cite{DesFurOuv95} give
3195: $\alpha_c\sim0.3$.
3196: 
3197: 
3198: \begin{figure}[!h]
3199: \begin{center}
3200: 
3201: \includegraphics[scale=0.4]{dosimpmag1.eps}
3202: \hspace{0.5cm}
3203: \includegraphics[scale=0.4]{dosimpmag2.eps}
3204: 
3205: \vspace{0.25cm}
3206: 
3207: \includegraphics[scale=0.4]{dosimpmag3.eps}
3208: \hspace{0.5cm}
3209: \includegraphics[scale=0.4]{dosimpmag4.eps}
3210: 
3211: \caption{\it Average density of states of the Hamiltonian (\ref{Him}) 
3212:          for different values
3213:          of the flux per tube $\phi=\alpha\phi_0$. 
3214:          The average magnetic field reads $\mean{{\cal B}}=2\pi\mu\alpha$.
3215:          We recall that the corresponding Landau spectrum is given by 
3216:          $E_n=\mean{{\cal B}}(n+1)$, for $n\geq0$.
3217:          From ref.~\protect\cite{DesFurOuv95}.\label{fig:dosimpmag}}
3218: \end{center}
3219: \end{figure}
3220: 
3221: 
3222: 
3223: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3224: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3225: 
3226: %\newpage%\ \newpage
3227: 
3228: \section{Conclusion}
3229: 
3230: A non experienced reader may have the feeling that the two topics which have
3231: been covered in this review, namely one-dimensional disordered systems and
3232: quantum graphs, are essentially disjoint. In fact there are many interesting
3233: links between these two topics, both at a methodological and a conceptual
3234: level. The use of metric graphs for modelling quantum phenomena observed in
3235: disordered metals goes back to the pioneering work of Shapiro \cite{Sha82}
3236: and Chalker \& Coddington \cite{ChaCod88}. These systems have been used to
3237: study quantum localization and more recently as a model system for spectral
3238: statistics (for a recent review, see ref.~\cite{Tan05}).  Another set of
3239: similar questions is provided by the study of scattering properties of chaotic
3240: graphs \cite{KotSmi03} and disordered systems \cite{FyoSom97} (see also
3241: section~\ref{sec:wtd} and ref.~\cite{OssFyo05} for recent developments). We
3242: have also seen that apart from its interest to study the spectral properties
3243: of metric graphs, the spectral determinant also allows studying several
3244: properties of networks of quasi-one-dimensional weakly disordered 
3245: wires~\cite{Pas98,PasMon99,AkkComDesMonTex00}.
3246: 
3247: As a conclusion we would like to mention several open problems~:
3248: \\
3249: $\bullet$ A study of spectral statistics in the case of graphs with a random
3250: Schr\"odinger operator [there is still a factorised structure but the matrix
3251: $R$ is now given by eq.~(\ref{n41})].
3252: \\
3253: $\bullet$ A probabilistic understanding of the star graphs using the tools of
3254: excursion theory developed by Barlow {\it et al}. \cite{BarPitYor89} in the
3255: context of the Brownian spider. A first step would be to recover those
3256: probabilistic results (the joint law of the occupation time inside the
3257: branches) by a spectral approach. It is however not excluded that the
3258: probabilistic approach could provide a key to a deeper understanding of those
3259: quantum systems.  In the context of classical systems such probabilistic
3260: approaches have been very useful, {\it e.g.} recent studies of the Stochastic
3261: Loewner Equation have made enormous progress in understanding the statistical
3262: physics of a class of two-dimensional systems. This calls for new
3263: probabilistic techniques for quantum systems as well.
3264: \\
3265: $\bullet$ Several functionals of the Brownian motion (\ref{A},\ref{functring})
3266: appear when studying the important question of dephasing due to
3267: electron-electron interaction in networks of quasi-one-dimensional weakly
3268: disordered wires. The fact that such simple functionals appear relies on the
3269: translation invariance of the two particular problems studied (an infinite
3270: wire \cite{AltAroKhm82} and an isolated ring \cite{TexMon05}). For a network
3271: with arbitrary topology, the relevant functional of the Brownian bridge
3272: $x(\tau)$ is given by $\int_0^t\D\tau\,W(x(\tau),x(t-\tau))$ where
3273: $W(x,x')=\frac12[P_d(x,x)+P_d(x',x')]-P_d(x,x')$ with
3274: $-\Delta{}P_d(x,x')=\delta(x-x')$.  It now involves a nonlocal functional in
3275: time which is difficult to handle. Progress in this direction would allow 
3276: clarifying the interplay between the electron-electron interaction and the
3277: geometrical effect and help in analyzing recent experimental results
3278: \cite{FerAngRowGueBouTexMonMai04,BauMalMonSamSchTex05}.
3279: \\
3280: $\bullet$ The question of extreme value spectral statistics was addressed in
3281: the framework of random matrix theory \cite{TraWid93} and these studies have
3282: found several applications in the context of out-of-equilibrium statistical
3283: physics (see the review \cite{TraWid02}). However this question was first
3284: addressed in the context of one-dimensional disordered systems
3285: \cite{GreMolSud83}. The study of supersymmetric random Hamiltonian, for which
3286: the bottom of the spectrum plays a special role, has emphasized the interest
3287: in extreme value spectral statistics \cite{Tex00} (in particular it indicates
3288: level correlations). It would be interesting to extend such studies to other
3289: models and find some physical situations where these results would be
3290: applicable.
3291: \\
3292: $\bullet$ A beautiful heuristic argument was provided by Lifshitz to explain
3293: the spectral singularity for Hamiltonians with a weak concentration of
3294: localized scalar impurities (see for example the book \cite{LifGrePas88}).
3295: Other mechanisms should be invoked to explain the nature of the quantum states
3296: responsible for the low energy power law behaviour in the case of a uniform
3297: strong magnetic field with $\delta$-impurities \cite{Fur97,Fur00}.  For the
3298: model of randomly distributed magnetic fluxes, the numerical simulations for
3299: $\alpha=1/2$ have suggested the new type of singular behaviour
3300: (\ref{specsing}). It would be interesting to provide heuristic arguments to
3301: understand more deeply the origin of this singularity.
3302: 
3303: 
3304: 
3305: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3306: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3307: 
3308: \section*{Acknowledgments}
3309: 
3310: The article gives an overview of several joint works involving~: \'Eric
3311: Akkermans, Cyril Furtlehner, Satya Majumdar, Gilles Montambaux, C\'ecile
3312: Monthus, St\'ephane Ouvry and Marc Yor. We thank them for fruitful
3313: collaborations. Meanwhile we have also benefited from stimulating discussions
3314: with Marc Bocquet, Eug\`ene Bogomolny, Oriol Bohigas, H\'el\`ene Bouchiat,
3315: Markus B\"uttiker, David Dean, Richard Deblock, Meydi Ferrier, Sophie
3316: Gu\'eron, Jean-Marc Luck, Gleb Oshanin, Leonid Pastur and Denis Ullmo.
3317: We are pleased to acknowledge Satya Majumdar for his careful reading and 
3318: valuable remarks.
3319: 
3320: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3321: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3322: 
3323: \newpage
3324: 
3325: %\bibliographystyle{chris4}
3326: %\bibliography{biblio}
3327: 
3328: \addcontentsline{toc}{section}{\protect\bibname}
3329: \begin{thebibliography}{200}
3330: 
3331: \bibitem{AbrAndLicRam79}
3332: E.~Abrahams, P.~W. Anderson, D.~C. Licciardello and T.~V. Ramakrishnan, Scaling
3333:   theory of localization: absence of quantum diffusion in two dimensions, Phys.
3334:   Rev. Lett. {\bf 42}(10), 673 (1979).
3335: 
3336: \bibitem{Ada92}
3337: V.~Adamyan, Scattering matrices for microschemes, Oper. Theory: Adv. \& Appl.
3338:   {\bf 59}, 1 (1992).
3339: 
3340: \bibitem{AkkComDesMonTex00}
3341: E.~Akkermans, A.~Comtet, J.~Desbois, G.~Montambaux and C.~Texier, On the
3342:   spectral determinant of quantum graphs, Ann. Phys. (N.Y.) {\bf 284}, 10--51
3343:   (2000).
3344: 
3345: \bibitem{AkkMon04}
3346: {\'E}.~Akkermans and G.~Montambaux, {\em Phy\-si\-que m\'esos\-co\-pi\-que des
3347:   \'elec\-trons et des pho\-tons\/}, EDP Sciences, CNRS \'editions (2004).
3348: 
3349: \bibitem{AlbGesHoeHol04}
3350: S.~Albeverio, S.~A. Gesztesy, R.~Hoegh-Krohn and H.~Holden, {\em Solvable
3351:   models in quantum mechanics\/}, A. M. S. Chelsea Publishing, second edition
3352:   (2004).
3353: 
3354: \bibitem{Ale83}
3355: S.~Alexander, Superconductivity of networks. A percolation approach to the
3356:   effects of disorder, Phys. Rev.~B {\bf 27}(3), 1541 (1983).
3357: 
3358: \bibitem{AltZir97}
3359: A.~Altland and M.~R. Zirnbauer, Nonstandard symmetry classes in mesoscopic
3360:   normal-superconducting hybrid structures, Phys. Rev.~B {\bf 55}(2), 1142
3361:   (1997).
3362: 
3363: \bibitem{AltAro81}
3364: B.~L. Al'tshuler and A.~G. Aronov, Magnetoresistance of thin films and of wires
3365:   in a longitudinal magnetic field, JETP Lett. {\bf 33}(10), 499 (1981).
3366: 
3367: \bibitem{AltAroKhm82}
3368: B.~L. Altshuler, A.~G. Aronov and D.~E. Khmelnitsky, Effects of
3369:   electron-electron collisions with small energy transfers on quantum
3370:   localisation, J.~Phys.~C: Solid St. Phys. {\bf 15}, 7367 (1982).
3371: 
3372: \bibitem{AltAroSpi81}
3373: B.~L. Al'tshuler, A.~G. Aronov and B.~Z. Spivak, The Aaronov-Bohm Effect in
3374:   disordered conductors, JETP Lett. {\bf 33}(2), 94 (1981).
3375: 
3376: \bibitem{AltAroSpiShaSha82}
3377: B.~L. Al'tshuler, A.~G. Aronov, B.~Z. Spivak, D.~{\relax{Yu}}. Sharvin and
3378:   {\relax{Yu}}.~V. Sharvin, Observation of the Aaronov-Bohm Effect in hollow
3379:   metal cylinders, JETP Lett. {\bf 35}(11), 588 (1982).
3380: 
3381: \bibitem{AltKhmLarLee80}
3382: B.~L. Al'tshuler, D.~E. Khmel'nitzki{\u\i}, A.~I. Larkin and P.~A. Lee,
3383:   Magnetoresistance and Hall effect in a disordered two-dimensional electron
3384:   gas, Phys. Rev.~B {\bf 22}(11), 5142 (1980).
3385: 
3386: \bibitem{AltLee88}
3387: B.~L. Al'tshuler and P.~A. Lee, Disordered Electronic Systems, Physics Today
3388:   {\bf 41}, 36 (december 1988).
3389: 
3390: \bibitem{AltPri89}
3391: B.~L. Altshuler and V.~N. Prigodin, Distribution of local density of states and
3392:   NMR line shape in a one-dimensional disordered conductor, Sov. Phys. JETP
3393:   {\bf 68}(1), 198 (1989).
3394: 
3395: \bibitem{AntPasSly81}
3396: T.~N. Antsygina, L.~A. Pastur and V.~A. Slyusarev, Localization of states and
3397:   kinetic properties of one-dimensional disordered systems, Sov. J. Low Temp.
3398:   Phys. {\bf 7}(1), 1--21 (1981).
3399: 
3400: \bibitem{AroSha87}
3401: A.~G. Aronov and {\relax{Yu}}.~V. Sharvin, Magnetic flux effects in disordered
3402:   conductors, Rev. Mod. Phys. {\bf 59}(3), 755 (1987).
3403: 
3404: \bibitem{AshMer76}
3405: N.~W. Ashcroft and N.~D. Mermin, {\em Solid State Physics\/}, Saunders College
3406:   (1976).
3407: 
3408: \bibitem{AvrExnLas94}
3409: J.~E. Avron, P.~Exner and Y.~Last, Periodic Schr{\"o}dinger operators with
3410:   large gaps and Wannier-Stark ladders, Phys. Rev. Lett. {\bf 72}(6), 896
3411:   (1994).
3412: 
3413: \bibitem{AvrRavZur88}
3414: J.~E. Avron, A.~Raveh and B.~Zur, Adiabatic quantum transport in multiply
3415:   connected systems, Rev. Mod. Phys. {\bf 60}, 873 (1988).
3416: 
3417: \bibitem{AvrSad91}
3418: J.~E. Avron and L.~Sadun, Adiabatic quantum transport in networks with
3419:   macroscopic components, Ann. Phys. (N.Y.) {\bf 206}, 440 (1991).
3420: 
3421: \bibitem{Azb83}
3422: M.~{\relax{Ya}}. Azbel, Resonance tunneling and localization spectroscopy,
3423:   Solid State Commun. {\bf 45}(7), 527 (1983).
3424: 
3425: \bibitem{BalHavWei87}
3426: R.~C. Ball, S.~Havlin and G.~H. Weiss, Non-Gaussian random walks, J.~Phys.~A:
3427:   Math. Gen. {\bf 20}, 4055 (1987).
3428: 
3429: \bibitem{BarPitYor89}
3430: M.~Barlow, J.~Pitman and M.~Yor, {\em Une extension multidimensionnelle de la
3431:   loi de l'arc sinus\/}, volume 1372 of {\em Lecture Notes in Maths\/}, p. 294,
3432:   Springer (1989).
3433: 
3434: \bibitem{BarGas00}
3435: F.~Barra and P.~Gaspard, On the level spacing distribution in quantum graphs,
3436:   J.~Stat. Phys. {\bf 101}, 283 (2000).
3437: 
3438: \bibitem{BarGas01}
3439: F.~Barra and P.~Gaspard, Transport and dynamics on open quantum graphs, Phys.
3440:   Rev.~E {\bf 65}, 016205 (2001).
3441: 
3442: \bibitem{Bar99}
3443: L.~Bartholdi, Counting paths in graphs, L'Enseignement Math\'ematique {\bf 45},
3444:   83 (1999).
3445: 
3446: \bibitem{BarTac93}
3447: A.~V. Barzykin and M.~Tachiya, Diffusion-influenced reaction kinetics on
3448:   fractal structures, J. Chem. Phys. {\bf 99}, 9591 (1993).
3449: 
3450: \bibitem{Bas92}
3451: H.~Bass, The Ihara-Selberg zeta function of a tree lattice, Internat. J. Math.
3452:   {\bf 3}, 717 (1992).
3453: 
3454: \bibitem{BauMalMonSamSchTex05}
3455: C.~Bau{\"e}rle, F.~Mallet, G.~Montambaux, L.~Saminadayar, F.~Schopfer and
3456:   C.~Texier, in preparation  (2005).
3457: 
3458: \bibitem{Bee97}
3459: C.~W.~J. Beenakker, Random-Matrix theory of quantum transport, Rev. Mod. Phys.
3460:   {\bf 69}(3), 731--808 (1997).
3461: 
3462: \bibitem{BenKliPlo90}
3463: S.~J. Bending, K.~{von Klitzing} and K.~Ploog, Weak localization in a
3464:   distribution of magnetic flux tubes, Phys. Rev. Lett. {\bf 65}(8), 1060
3465:   (1990).
3466: 
3467: \bibitem{Ber74}
3468: V.~L. Berezinski\u{i}, Kinetics of a quantum particle in a one-dimensional
3469:   random potential, Sov. Phys. JETP {\bf 38}(3), 620 (1974).
3470: 
3471: \bibitem{BerLoz94}
3472: O.~Bergman and G.~Lozano, Aharonov-Bohm scattering, contact interactions and
3473:   scale invariance, Ann. Phys. (N.Y.) {\bf 229}, 416 (1994).
3474: 
3475: \bibitem{Ber84}
3476: G.~Bergmann, Weak localization in thin films, Phys. Rep. {\bf 107}, 1 (1984).
3477: 
3478: \bibitem{BerBogKea01}
3479: G.~Berkolaiko, E.~B. Bogomolny and J.~P. Keating, Star graphs and {\v S}eba
3480:   billiards, J.~Phys.~A: Math. Gen. {\bf 34}, 335 (2001).
3481: 
3482: \bibitem{BerKea99}
3483: G.~Berkolaiko and J.~P. Keating, Two-point spectral correlations for star
3484:   graphs, J.~Phys.~A: Math. Gen. {\bf 32}, 7827 (1999).
3485: 
3486: \bibitem{BetUhl37}
3487: E.~Beth and G.~E. Uhlenbeck, Physica {\bf 4}, 915 (1937).
3488: 
3489: \bibitem{Boc99}
3490: M.~Bocquet, Some spectral properties of the one-dimensional disordered Dirac
3491:   equation, Nucl. Phys.~B [FS] {\bf 546}, 621 (1999).
3492: 
3493: \bibitem{Boc00}
3494: M.~Bocquet, {\em Cha\^{\i}nes de spins, fermions de Dirac et syst\`emes
3495:   d\'esordonn\'es\/}, Ph.D. thesis, \'Ecole Polytechnique (2000), available at
3496:   {\sf http://tel.ccsd.cnrs.fr/documents/archives0/00/00/15/60/}.
3497: 
3498: \bibitem{BocSerZir00}
3499: M.~Bocquet, D.~Serban and M.~R. Zirnbauer, Disordered 2d quasiparticles in
3500:   class D: Dirac fermions with random mass, and dirty superconductors, Nucl.
3501:   Phys.~B [FS] {\bf 578}, 628 (2000).
3502: 
3503: \bibitem{BouComGeoLeD87}
3504: J.-P. Bouchaud, A.~Comtet, A.~Georges and P.~{Le~Doussal}, The relaxation-time
3505:   spectrum of diffusion in a one-dimensional random medium: an exactly solvable
3506:   case, Europhys. Lett. {\bf 3}, 653 (1987).
3507: 
3508: \bibitem{BouComGeoLeD90}
3509: J.-P. Bouchaud, A.~Comtet, A.~Georges and P.~{Le~Doussal}, Classical diffusion
3510:   of a particle in a one-dimensional random force field, Ann. Phys. (N.Y.) {\bf
3511:   201}, 285--341 (1990).
3512: 
3513: \bibitem{BouGeo90}
3514: J.-P. Bouchaud and A.~Georges, Anomalous diffusion in disordered media:
3515:   Statistical mechanisms, models and physical applications, Phys. Rep. {\bf
3516:   195}, 267 (1990).
3517: 
3518: \bibitem{BouMez97}
3519: J.-P. Bouchaud and M.~M\'ezard, Universality classes for extreme value
3520:   statistic, J.~Phys.~A: Math. Gen. {\bf 30}, 7997 (1997).
3521: 
3522: \bibitem{BouLac85}
3523: P.~Bougerol and J.~Lacroix, {\em Products of random matrices with applications
3524:   to Schr\"odinger operators\/}, Birkh\"auser (1985).
3525: 
3526: \bibitem{BreGroItz84}
3527: {\'E}.~Br{\'e}zin, D.~J. Gross and C.~Itzykson, Density of states in the
3528:   presence of a strong magnetic field and random impurities, Nucl. Phys.~B [FS]
3529:   {\bf 235}, 24 (1984).
3530: 
3531: \bibitem{BroFraBee97}
3532: P.~W. Brouwer, K.~M. Frahm and C.~W. Beenakker, Quantum mechanical time-delay
3533:   matrix in chaotic scattering, Phys. Rev. Lett. {\bf 78}(25), 4737 (1997).
3534: 
3535: \bibitem{BroMudFur00}
3536: P.~W. Brouwer, C.~Mudry and A.~Furusaki, Density of States in Coupled Chains
3537:   with Off-Diagonal Disorder, Phys. Rev. Lett. {\bf 84}(13), 2913 (2000).
3538: 
3539: \bibitem{BroMudFur03}
3540: P.~W. Brouwer, C.~Mudry and A.~Furusaki, Universality of delocalization in
3541:   unconventional dirty superconducting wires with broken spin-rotation
3542:   symmetry, Phys. Rev.~B {\bf 67}, 014530 (2003).
3543: 
3544: \bibitem{But90}
3545: M.~B{\"u}ttiker, Traversal, reflection and dwell time for quantum tunneling, in
3546:   {\em Electronic properties of multilayers and low-dimensional semiconductors
3547:   structures\/}, edited by J.~M. {Chamberlain et al.}, p. 297, Plenum Press,
3548:   New York (1990).
3549: 
3550: \bibitem{But02}
3551: M.~B{\"u}ttiker, Charge densities and charge noise in mesoscopic conductors,
3552:   Pramana J. Phys. {\bf 58}, 241 (2002), (cond-mat/0112330).
3553: 
3554: \bibitem{ButImrAzb84}
3555: M.~B{\"u}ttiker, Y.~Imry and M.~{\relax{Ya}}. Azbel, Quantum oscillations in
3556:   one-dimensional normal-metal rings, Phys. Rev.~A {\bf 30}(4), 1982 (1984).
3557: 
3558: \bibitem{CarAku04}
3559: F.~Carlier and V.~M. Akulin, Quantum interference in nanofractals and its
3560:   optical manifestation, Phys. Rev.~B {\bf 69}, 115433 (2004).
3561: 
3562: \bibitem{CarLeD01}
3563: D.~Carpentier and P.~{Le~Doussal}, Glass transition of a particle in a random
3564:   potential, front selection and entropic phenomena in Liouville and
3565:   Sinh-Gordon models, Phys. Rev.~E {\bf 63}, 026110 (2001).
3566: 
3567: \bibitem{ChaSch86}
3568: S.~Chakravarty and A.~Schmid, Weak localization: the quasiclassical theory of
3569:   electrons in a random potential, Phys. Rep. {\bf 140}(4), 193 (1986).
3570: 
3571: \bibitem{ChaCod88}
3572: J.~T. Chalker and P.~D. Coddington, Percolation, quantum tunnelling and the
3573:   integer Hall-effect, J.~Phys.~C: Solid St. Phys. {\bf 21}(14), 2665 (1988).
3574: 
3575: \bibitem{Che99}
3576: L.~O. Chekhov, A spectral problem on graphs and $L$-functions, Russ. Math.
3577:   Surv. {\bf 54}(6), 1197 (1999).
3578: 
3579: \bibitem{CifReg75}
3580: D.~M. Cifarelli and E.~Regazzini, Contributi intorno ad un test per
3581:   l'homogeneita tra du campioni, Giornale Degli Economiste {\bf 34}, 233--249
3582:   (1975).
3583: 
3584: \bibitem{Col98}
3585: Y.~{Colin de Verdi\`ere}, {\em Spectres de Graphes\/}, Soci\'et\'e
3586:   Math\'ematique de France (1998).
3587: 
3588: \bibitem{ComDesMaj02}
3589: A.~Comtet, J.~Desbois and S.~N. Majumdar, The local time distribution of a
3590:   particle diffusing on a graph, J.~Phys.~A: Math. Gen. {\bf 35}, L687 (2002).
3591: 
3592: \bibitem{ComDesMon95}
3593: A.~Comtet, J.~Desbois and C.~Monthus, Localization properties in
3594:   one-dimensional disordered supersymmetric quantum mechanics, Ann. Phys.
3595:   (N.Y.) {\bf 239}, 312--350 (1995).
3596: 
3597: \bibitem{ComDesOuv90}
3598: A.~Comtet, J.~Desbois and S.~Ouvry, Winding of planar Brownian curves,
3599:   J.~Phys.~A: Math. Gen. {\bf 23}, 3563 (1990).
3600: 
3601: \bibitem{ComMasOuv95}
3602: A.~Comtet, S.~Mashkevich and S.~Ouvry, Magnetic moment and perturbation theory
3603:   with singular magnetic fields, Phys. Rev.~D {\bf 52}(4), 2594 (1995).
3604: 
3605: \bibitem{ComMonYor98}
3606: A.~Comtet, C.~Monthus and M.~Yor, Exponential functionals of Brownian motion
3607:   and disordered systems, J. Appl. Probab. {\bf 35}, 255 (1998).
3608: 
3609: \bibitem{ComTex97}
3610: A.~Comtet and C.~Texier, On the distribution of the Wigner time delay in
3611:   one-dimensional disordered systems, J.~Phys.~A: Math. Gen. {\bf 30},
3612:   8017--8025 (1997).
3613: 
3614: \bibitem{ComTex98}
3615: A.~Comtet and C.~Texier, One-dimensional disordered supersymmetric quantum
3616:   mechanics: a brief survey, in {\em Supersymmetry and Integrable Models\/},
3617:   edited by H.~Aratyn, T.~D. Imbo, W.-Y. Keung and U.~Sukhatme, Lecture Notes
3618:   in Physics, Vol. 502, pp. 313--328. Springer (1998), Proceedings of a
3619:   workshop held at Chicago, IL, USA, 12-14 June 1997 (also available as
3620:   cond-mat/9707313).
3621: 
3622: \bibitem{DasMaBer69}
3623: R.~Dashen, S.-K. Ma and H.~J. Bernstein, $S$-matrix formulation of statistical
3624:   mechanics, Phys. Rev. {\bf 187}(1), 345 (1969).
3625: 
3626: \bibitem{CalLucNiePet85}
3627: C.~de~Calan, J.-M. Luck, T.~M. Nieuwenhuizen and D.~P{\'e}tritis, On the
3628:   distribution of a random variable occuring in 1D disordered systems,
3629:   J.~Phys.~A: Math. Gen. {\bf 18}, 501 (1985).
3630: 
3631: \bibitem{CarNus02}
3632: C.~A.~A. de~Carvalho and H.~M. Nussenzveig, Time delay, Phys. Rep. {\bf 364},
3633:   83 (2002).
3634: 
3635: \bibitem{DeaMaj01}
3636: D.~S. Dean and S.~N. Majumdar, Extreme Value Statistics of Hierarchically
3637:   Correlated Variables: Deviation from Gumbel Statistics and Anomalous
3638:   Persistence, Phys. Rev.~E {\bf 64}, 046121 (2001).
3639: 
3640: \bibitem{DerHil83}
3641: B.~Derrida, Singular behaviour of certain infinite products of random
3642:   $2\times2$ matrices, J.~Phys.~A: Math. Gen. {\bf 16}, 2641 (1983).
3643: 
3644: \bibitem{Des00}
3645: J.~Desbois, Spectral determinant of Schr\"odinger operators on graphs,
3646:   J.~Phys.~A: Math. Gen. {\bf 33}, L63 (2000).
3647: 
3648: \bibitem{Des00a}
3649: J.~Desbois, Time-dependent harmonic oscillator and spectral determinant on
3650:   graphs, Eur. Phys. J.~B {\bf 15}, 201 (2000).
3651: 
3652: \bibitem{Des01}
3653: J.~Desbois, Spectral determinant on graphs with generalized boundary
3654:   conditions, Eur. Phys. J.~B {\bf 24}, 261 (2001).
3655: 
3656: \bibitem{Des02}
3657: J.~Desbois, Occupation times distribution for Brownian motion on graphs,
3658:   J.~Phys.~A: Math. Gen. {\bf 35}, L673 (2002).
3659: 
3660: \bibitem{DesCom92}
3661: J.~Desbois and A.~Comtet, Algebraic areas enclosed by Brownian curves on
3662:   bounded domains, J.~Phys.~A: Math. Gen. {\bf 25}, 3095 (1992).
3663: 
3664: \bibitem{DesFurOuv95}
3665: J.~Desbois, C.~Furtlehner and S.~Ouvry, Random magnetic impurities and the
3666:   Landau problem, Nucl. Phys.~B [FS] {\bf 453}, 759--776 (1995).
3667: 
3668: \bibitem{DesFurOuv96}
3669: J.~Desbois, C.~Furtlehner and S.~Ouvry, Random magnetic impurities and the
3670:   $\delta$ impurity problem, J.~Phys.~I (France) {\bf 6}, 641--648 (1996).
3671: 
3672: \bibitem{Dor82}
3673: O.~N. Dorokhov, Transmission coefficient and the localization length of an
3674:   electron in $N$ bound disordered chains, JETP Lett. {\bf 36}, 318 (1982).
3675: 
3676: \bibitem{Dor88}
3677: O.~N. Dorokhov, Solvable model of multichannel localization, Phys. Rev.~B {\bf
3678:   37}, 10526 (1988).
3679: 
3680: \bibitem{DouRam85}
3681: B.~Dou{\c c}ot and R.~Rammal, Quantum oscillations in normal-metal networks,
3682:   Phys. Rev. Lett. {\bf 55}(10), 1148 (1985).
3683: 
3684: \bibitem{DouRam86}
3685: B.~Dou{\c c}ot and R.~Rammal, Interference effects and magnetoresistance
3686:   oscillations in normal-metal networks: 1. weak localization approach, J.
3687:   Physique {\bf 47}, 973--999 (1986).
3688: 
3689: \bibitem{Dys53}
3690: F.~J. Dyson, The dynamics of a disordered linear chain, Phys. Rev. {\bf 92},
3691:   1331 (1953).
3692: 
3693: \bibitem{EchGerBozBogNil93b}
3694: P.~M. Echternach, M.~E. Gershenson, H.~M. Bozler, A.~L. Bogdanov and
3695:   B.~Nilsson, Nyquist phase relaxation in one-dimensional metal films, Phys.
3696:   Rev.~B {\bf 48}(15), 11516 (1993).
3697: 
3698: \bibitem{Edw67}
3699: S.~F. Edwards, Statistical mechanics with topological constraints, Proc. Phys.
3700:   Soc. {\bf 91}, 513 (1967).
3701: 
3702: \bibitem{Efe97}
3703: K.~Efetov, {\em Supersymmetry in disorder and chaos\/}, Cambridge University
3704:   Press (1997).
3705: 
3706: \bibitem{EfrPol85}
3707: A.~L. Efros and M.~Pollak (editors), {\em Electron-electron interactions in
3708:   disordered systems\/}, North-Holland (1985).
3709: 
3710: \bibitem{EvaKat03}
3711: S.~N. Evangelou and D.~E. Katsanos, Spectral statistics in chiral-orthogonal
3712:   disordered systems, J.~Phys.~A: Math. Gen. {\bf 36}, 3237 (2003).
3713: 
3714: \bibitem{Exn95}
3715: P.~Exner, Lattice Kronig-Penney Models, Phys. Rev. Lett. {\bf 74}(18), 3503
3716:   (1995).
3717: 
3718: \bibitem{Exn96a}
3719: P.~Exner, Contact interactions on graph superlattices, J.~Phys.~A: Math. Gen.
3720:   {\bf 29}(1), 87 (1996).
3721: 
3722: \bibitem{FabMel97}
3723: M.~Fabrizio and R.~M\'elin, Coexistence of antiferromagnetism and dimerization
3724:   in a disordered spin-Peierls model: exact results, Phys. Rev. Lett. {\bf
3725:   78}(17), 3382 (1997).
3726: 
3727: \bibitem{FabMel97b}
3728: M.~Fabrizio and R.~M\'elin, Enhanced magnetic fluctuations in doped
3729:   spin-Peierls systems: a single-chain-model analysis, Phys. Rev.~B {\bf
3730:   56}(10), 5996 (1997).
3731: 
3732: \bibitem{FarTsa94}
3733: W.~G. Faris and W.~J. Tsay, Time delay in random scattering, SIAM J. Appl.
3734:   Math. {\bf 54}, 443 (1994).
3735: 
3736: \bibitem{FerAngRowGueBouTexMonMai04}
3737: M.~Ferrier, L.~Angers, A.~C.~H. Rowe, S.~Gu{\'e}ron, H.~Bouchiat, C.~Texier,
3738:   G.~Montambaux and D.~Mailly, Direct measurement of the phase coherence length
3739:   in a GaAs/GaAlAs square network, Phys. Rev. Lett. {\bf 93}, 246804 (2004).
3740: 
3741: \bibitem{FlaLou01}
3742: P.~Flajeolet and G.~Louchard, Analytic variations on the Airy distribution,
3743:   Algorithmica {\bf 31}, 361 (2001).
3744: 
3745: \bibitem{For87}
3746: R.~Forman, Functional determinants and geometry, Invent. math. {\bf 88}, 447
3747:   (1987).
3748: 
3749: \bibitem{Fri52}
3750: J.~Friedel, The distribution of electrons round impurities in monovalent
3751:   metals, Phil. Mag. {\bf 43}, 153 (1952).
3752: 
3753: \bibitem{Fri58}
3754: J.~Friedel, Metallic alloys, Nuovo Cimento Suppl. {\bf 7}, 287 (1958).
3755: 
3756: \bibitem{FriLlo60}
3757: H.~L. Frisch and S.~P. Lloyd, Electron levels in a one-dimensional random
3758:   lattice, Phys. Rev. {\bf 120}(4), 1175 (1960).
3759: 
3760: \bibitem{Fur97}
3761: C.~Furtlehner, {\em \'Etude du spectre de Landau pour un champ magn\'etique
3762:   al\'eatoire en dimension 2\/}, Ph.D. thesis, Universit\'e Paris 6 (1997).
3763: 
3764: \bibitem{Fur00}
3765: C.~Furtlehner, Lifshitz-like argument for low-lying states in a strong magnetic
3766:   field, Eur. Phys. J.~B {\bf 18}, 297 (2000).
3767: 
3768: \bibitem{FyoSom96a}
3769: Y.~V. Fyodorov and H.-J. Sommers, Parametric correlations of scattering phase
3770:   shifts and fluctuations of delay times in few-channel chaotic scattering,
3771:   Phys. Rev. Lett. {\bf 76}, 4709 (1996).
3772: 
3773: \bibitem{FyoSom97}
3774: Y.~V. Fyodorov and H.-J. Sommers, Statistics of resonance poles, phase shift
3775:   and time delays in quantum chaotic scattering: Random matrix approach for
3776:   systems with broken time-reversal invariance, J. Math. Phys. {\bf 38}(4),
3777:   1918 (1997).
3778: 
3779: \bibitem{GarTru05}
3780: C.~Garban and J.~A. {Trujillo Ferreras}, The expected area of the filled planar
3781:   Brownian loop is $\pi/5$, unpublished  (2005), preprint math.PR/0504496.
3782: 
3783: \bibitem{Gar89}
3784: C.~W. Gardiner, {\em Handbook of stochastic methods for physics, chemistry and
3785:   the natural sciences\/}, Springer (1989).
3786: 
3787: \bibitem{GefImrAzb84}
3788: Y.~Gefen, Y.~Imry and M.~{\relax{Ya}}. Azbel, Quantum oscillations and the
3789:   Aharonov-Bohm effect for parallel resistors, Phys. Rev. Lett. {\bf 52}(2),
3790:   129 (1984).
3791: 
3792: \bibitem{Gei89}
3793: A.~K. Geim, Nonlocal magnetoresistance of bismuth films in nonuniform field of
3794:   Abrikosov vortices, JETP Lett. {\bf 50}(8), 389 (1989).
3795: 
3796: \bibitem{GerPav88}
3797: N.~I. Gerasimenko and B.~S. Pavlov, Scattering problems on noncompact graphs,
3798:   Theor. Math. Phys. {\bf 74}, 230 (1988).
3799: 
3800: \bibitem{GerKhaMikBozBog97}
3801: M.~E. Gershenson, {\relax{Yu}}.~B. Khavin, A.~Mikhalchuk, H.~M. Bolzer and
3802:   A.~L. Bogdanov, Crossover from Weak to Strong Localization in
3803:   Quasi-One-Dimensional Conductors, Phys. Rev. Lett. {\bf 79}(4), 725 (1997).
3804: 
3805: \bibitem{GnuAlt04}
3806: S.~Gnutzmann and A.~Altland, Universal Spectral Statistics in Quantum Graphs,
3807:   Phys. Rev. Lett. {\bf 93}, 194101 (2004).
3808: 
3809: \bibitem{GogMel77}
3810: A.~A. Gogolin and V.~I. Mel'nikov, Conductivity of one-dimensional metal with
3811:   half-filled band, Sov. Phys. JETP {\bf 46}, 369 (1977).
3812: 
3813: \bibitem{GogMelRas76}
3814: A.~A. Gogolin, V.~I. Mel'nikov and E.~I. Rashba, Conductivity in a disordered
3815:   one-dimensional system induced by electron-phonon interaction, Sov. Phys.
3816:   JETP {\bf 42}(1), 168 (1976).
3817: 
3818: \bibitem{GolMolPas77}
3819: I.~{\relax{Ya}}. Gol'dshtein, S.~A. Molchanov and L.~A. Pastur, A pure point
3820:   spectrum of the stochastic one-dimensional Schr\"odinger operator, Funct.
3821:   Anal. and App. {\bf 11}, 1 (1977).
3822: 
3823: \bibitem{GopMelBut96}
3824: V.~A. Gopar, P.~A. Mello and M.~B{\"u}ttiker, Mesoscopic capacitors: a
3825:   statistical analysis, Phys. Rev. Lett. {\bf 77}(14), 3005 (1996).
3826: 
3827: \bibitem{GorDorPri83}
3828: L.~P. Gor'kov, O.~N. Dorokhov and F.~V. Prigara, Structure of wave functions an
3829:   AC conductivity id disordered one-dimensional conductors, Sov. Phys. JETP
3830:   {\bf 58}(4), 852 (1983).
3831: 
3832: \bibitem{GorLarKhm79}
3833: L.~P. Gor'kov, A.~I. Larkin and D.~E. Khmel'nitzki{\u\i}, Particle conductivity
3834:   in a two-dimensional random potential, JETP Lett. {\bf 30}(4), 228 (1979).
3835: 
3836: \bibitem{GreMolSud83}
3837: L.~N. Grenkova, S.~A. Mol\v{c}anov and J.~N. Sudarev, On the Basic States of
3838:   One-Dimensional Disordered Structures, Commun. Math. Phys. {\bf 90}, 101
3839:   (1983).
3840: 
3841: \bibitem{GroFri03}
3842: A.~Grosberg and H.~Frisch, Winding angle distribution for planar random walk,
3843:   polymer ring entangled with an obstacle, and all that:
3844:   Spitzer-Edwards-Prager-Frisch model revisited, J.~Phys.~A: Math. Gen. {\bf
3845:   36}, 8955 (2003).
3846: 
3847: \bibitem{Gum35}
3848: E.~J. Gumbel, Les valeurs extr\^emes des distributions statistiques, Ann. de
3849:   l'Institut Henri Poincar\'e {\bf V}, 115 (1935).
3850: 
3851: \bibitem{Gum54}
3852: E.~J. Gumbel, Statistical Theory of Extreme Values and Some Practical
3853:   Applications, National Bureau of Standards Applied Mathematics Series {\bf
3854:   33} (1954), Issued February 12.
3855: 
3856: \bibitem{Gum58}
3857: E.~J. Gumbel, {\em Statistics of Extremes\/}, Columbia University Press, New
3858:   York (1958).
3859: 
3860: \bibitem{GurCha03}
3861: V.~Gurarie and J.~T. Chalker, Bosonic excitations in random media, Phys. Rev.~B
3862:   {\bf 68}, 134207 (2003).
3863: 
3864: \bibitem{Gut90}
3865: M.~G. Gutzwiller, {\em Chaos in Classical and Quantum Mechanics\/}, volume~1 of
3866:   {\em Interdisciplinary Applied Mathematics\/}, Springer Verlag, New York
3867:   (1990).
3868: 
3869: \bibitem{Hal65}
3870: B.~I. Halperin, Green's functions for a particle in a one-dimensional random
3871:   potential, Phys. Rev. {\bf 139}, A104 (1965).
3872: 
3873: \bibitem{HauSto89}
3874: E.~H. Hauge and J.~A. St{\o}vneng, Tunneling times: a critical review, Rev.
3875:   Mod. Phys. {\bf 61}(4), 917 (1989).
3876: 
3877: \bibitem{HeiHucZir05}
3878: P.~Heinzner, A.~Huckleberry and M.~R. Zirnbauer, Symmetry classes of disordered
3879:   fermions, Commun. Math. Phys. {\bf 257}, 725 (2005).
3880: 
3881: \bibitem{HikLarNag80}
3882: S.~Hikami, A.~I. Larkin and Y.~Nagaoka, Spin-Orbit Interaction and
3883:   Magnetoresistance in the Two Dimensional Random System, Prog. Theor. Phys.
3884:   {\bf 63}(2), 707 (1980).
3885: 
3886: \bibitem{IglMon05}
3887: F.~Igloi and C.~Monthus, Strong disorder RG approach of random systems, Phys.
3888:   Rep. {\bf 412}, 277 (2005).
3889: 
3890: \bibitem{Iha66}
3891: Y.~Ihara, On discrete subgroup of the two by two projective linear group over
3892:   $p$-adic field, J. Math. Soc. Japan {\bf 18}(3), 219 (1966).
3893: 
3894: \bibitem{ItzDro89}
3895: C.~Itzykson and J.-M. Drouffe, {\em Th\'eorie statistique des champs\/},
3896:   Inter\'editions--{\sc Cnrs}, Paris (1989), Tomes 1 et 2.
3897: 
3898: \bibitem{JacPi90}
3899: R.~Jackiw and S.~Y. Pi, Classical and quantal nonrelativistic Chern-Simons
3900:   theory, Phys. Rev.~D {\bf 42}, 3500 (1990).
3901: 
3902: \bibitem{JayVijKum89}
3903: A.~M. Jayannavar, G.~V. Vijayagovindan and N.~Kumar, Energy dispersive
3904:   backscattering of electrons from surface resonances of a disordered medium
3905:   and $1/f$ noise, Z. Phys. B - Condens. Matter {\bf 75}, 77 (1989).
3906: 
3907: \bibitem{JeaPitYor97}
3908: M.~Jeanblanc, J.~Pitman and M.~Yor, The Feynman-Kac formula and decomposition
3909:   of Brownian paths, Comput. Appl. Math. {\bf 16}(1), 27 (1997).
3910: 
3911: \bibitem{Jon83}
3912: G.~Jona-Lasinio, Qualitative theory of stochastic differential equations and
3913:   quantum mechanics of disordered systems, Helv. Phys. Act. {\bf 56}, 61
3914:   (1983).
3915: 
3916: \bibitem{Kac46}
3917: M.~Kac, On the average of a certain Wiener functional and a related limit
3918:   theorem in calculus of probability, Trans. Am. Math. Soc. {\bf 59}, 401
3919:   (1946).
3920: 
3921: \bibitem{KeaMarWin03}
3922: J.~P. Keating, J.~Marklof and B.~Winn, Value distribution of the eigenfunctions
3923:   and spectral determinants of quantum star graphs, Commun. Math. Phys. {\bf
3924:   241}, 421 (2003).
3925: 
3926: \bibitem{Kel64}
3927: L.~V. Keldysh, Deep levels in semiconductors, Sov. Phys. JETP {\bf 18}(1), 253
3928:   (1964).
3929: 
3930: \bibitem{Kes73}
3931: H.~Kesten, Acta Math. {\bf 131}, 208 (1973).
3932: 
3933: \bibitem{KosSch99}
3934: V.~Kostrykin and R.~Schrader, Kirchhoff's rule for quantum wires, J.~Phys.~A:
3935:   Math. Gen. {\bf 32}, 595 (1999).
3936: 
3937: \bibitem{KotSmi97}
3938: T.~Kottos and U.~Smilansky, Quantum Chaos on Graphs, Phys. Rev. Lett. {\bf
3939:   79}(24), 4794 (1997).
3940: 
3941: \bibitem{KotSmi99}
3942: T.~Kottos and U.~Smilansky, Periodic Orbit Theory and Spectral Statistics for
3943:   Quantum Graphs, Ann. Phys. (N.Y.) {\bf 274}(1), 76 (1999).
3944: 
3945: \bibitem{KotSmi00}
3946: T.~Kottos and U.~Smilansky, Chaotic Scattering on Graphs, Phys. Rev. Lett. {\bf
3947:   85}(5), 968 (2000).
3948: 
3949: \bibitem{KotSmi03}
3950: T.~Kottos and U.~Smilansky, Quantum graphs~: A simple model for chaotic
3951:   scattering, J.~Phys.~A: Math. Gen. {\bf 36}, 3501 (2003).
3952: 
3953: \bibitem{Kre53}
3954: M.~G. Krein, Trace formulas in perturbation theory, Matem. Sbornik {\bf 33},
3955:   597 (1953).
3956: 
3957: \bibitem{KucSad98}
3958: E.~Z. Kuchinskii and M.~V. Sadovskii, Combinatorics of Feynman diagrams for the
3959:   problems with Gaussian random field, Sov. Phys. JETP {\bf 86}, 367 (1998).
3960: 
3961: \bibitem{Kuc04}
3962: P.~Kuchment, Quantum Graphs: I. Some basic structures, Waves Random Media {\bf
3963:   14}, S107 (2004).
3964: 
3965: \bibitem{Kuc05}
3966: P.~Kuchment, Quantum Graphs: II. Some spectral properties of quantum and
3967:   combinatorial graphs, J.~Phys.~A: Math. Gen. {\bf 38}, 4887 (2005).
3968: 
3969: \bibitem{LanLif66e}
3970: L.~D. Landau and E.~Lifchitz, {\em Physique statistique\/}, Mir (1966), tome 5.
3971: 
3972: \bibitem{LanMar94}
3973: R.~Landauer and T.~Martin, Barrier interaction time in tunneling, Rev. Mod.
3974:   Phys. {\bf 66}(1), 217 (1994).
3975: 
3976: \bibitem{LeDMonFis99}
3977: P.~{Le~Doussal}, C.~Monthus and D.~S. Fisher, Random walkers in one-dimensional
3978:   random environments: Exact renormalization group analysis, Phys. Rev.~E {\bf
3979:   59}(5), 4795 (1999).
3980: 
3981: \bibitem{LeG92}
3982: J.-F. {Le~Gall}, {\em Some properties of planar Brownian motion\/}, Lectures
3983:   Notes in Maths, Springer, Berlin (1992), p.~1527.
3984: 
3985: \bibitem{LeeRam85}
3986: P.~A. Lee and T.~V. Ramakrishnan, Disordered electronic systems, Rev. Mod.
3987:   Phys. {\bf 57}, 287 (1985).
3988: 
3989: \bibitem{LesWar04}
3990: H.~Leschke and S.~Warzel, Quantum-Classical Transitions in Lifshitz Tails with
3991:   Magnetic Fields, Phys. Rev. Lett. {\bf 92}, 086402 (2004).
3992: 
3993: \bibitem{Leu00}
3994: C.~Leuridan, Th\'eor\`eme de Ray-Knight dans un arbre~: une approche
3995:   alg\'ebrique, Pr\'epublication de l'institut Fourier n$^{\rm o}509$  (2000),
3996:   {\sf http://www-fourier.ujf-grenoble.fr/prepublications.html}.
3997: 
3998: \bibitem{Lev48}
3999: P.~{L\'evy}, {\em Processus stochastiques et mouvement brownien\/}, \'Editions
4000:   Jacques Gabay, Paris (1948).
4001: 
4002: \bibitem{LicTho75}
4003: D.~C. Licciardello and D.~J. Thouless, Constancy of Minimum Metallic
4004:   Conductivity in Two Dimensions, Phys. Rev. Lett. {\bf 35}(21), 1475 (1975).
4005: 
4006: \bibitem{LifGrePas88}
4007: I.~M. Lifshits, S.~A. Gredeskul and L.~A. Pastur, {\em Introduction to the
4008:   theory of disordered systems\/}, John Wiley \& Sons (1988).
4009: 
4010: \bibitem{Luc92}
4011: J.-M. Luck, {\em Syst\`emes d\'esordonn\'es unidimensionnels\/}, CEA,
4012:   collection Al\'ea Saclay, Saclay (1992).
4013: 
4014: \bibitem{LudMir04}
4015: T.~Ludwig and A.~D. Mirlin, Interaction-induced dephasing of Aharonov-Bohm
4016:   oscillations, Phys. Rev.~B {\bf 69}, 193306 (2004).
4017: 
4018: \bibitem{MajCom04}
4019: S.~N. Majumdar and A.~Comtet, Exact maximal height distribution of fluctuating
4020:   interfaces, Phys. Rev. Lett. {\bf 92}(22), 225501 (2004).
4021: 
4022: \bibitem{MajCom05}
4023: S.~N. Majumdar and A.~Comtet, Airy Distribution Function: From the Area Under a
4024:   Brownian Excursion to the Maximal Height of Fluctuating Interfaces, J. Stat.
4025:   Phys. {\bf 119}, 777 (2005).
4026: 
4027: \bibitem{MajKra03}
4028: S.~N. Majumdar and P.~L. Krapivsky, Extreme Value Statistics and Traveling
4029:   Fronts: Various Applications, Physica A {\bf 318}, 161 (2003).
4030: 
4031: \bibitem{McCOuv91}
4032: J.~McCabe and S.~Ouvry, Perturbative three-body spectrum and the third virial
4033:   coefficient in the anyon model, Phys. Lett.~B {\bf 260}, 113 (1991).
4034: 
4035: \bibitem{McKTar95}
4036: A.~J. McKane and M.~B. Tarlie, Regularization of functional determinants using
4037:   boundary perturbations, J.~Phys.~A: Math. Gen. {\bf 28}, 6931 (1995).
4038: 
4039: \bibitem{McK94}
4040: H.~P. McKean, A Limit Law for the Ground State of Hill's Equation, J.~Stat.
4041:   Phys. {\bf 74}(5/6), 1227 (1994).
4042: 
4043: \bibitem{MelPerKum88}
4044: P.~A. Mello, P.~Pereyra and N.~Kumar, Macroscopic approach to multichannel
4045:   disordered conductors, Ann. Phys. (N.Y.) {\bf 181}, 290 (1988).
4046: 
4047: \bibitem{Mol81}
4048: S.~A. Mol\v{c}anov, The Local Structure of the Spectrum of the One-Dimensional
4049:   Schr\"odinger Operator, Commun. Math. Phys. {\bf 78}, 429 (1981).
4050: 
4051: \bibitem{MonAkk05}
4052: G.~Montambaux and E.~Akkermans, Non exponential quasiparticle decay and phase
4053:   relaxation in low dimensional conductors, Phys. Rev. Lett. {\bf 95}, 016403
4054:   (2005).
4055: 
4056: \bibitem{Mon95}
4057: C.~Monthus, {\em \'Etude de quelques fonctionnelles du mouvement brownien et de
4058:   certaines propri\'et\'es de la diffusion unidimensionnelle en milieu
4059:   al\'eatoire\/}, Ph.D. thesis, Universit\'e Paris 6 (1995), Ann. Phys.
4060:   (France) {\bf20}, 341 (1995).
4061: 
4062: \bibitem{MonCom94}
4063: C.~Monthus and A.~Comtet, On the flux distribution in a one-dimensional
4064:   disordered system, J.~Phys.~I (France) {\bf 4}, 635--653 (1994).
4065: 
4066: \bibitem{MonOshComBur96}
4067: C.~Monthus, G.~Oshanin, A.~Comtet and S.~Burlatsky, Sample-size dependence of
4068:   the ground-state energy in a one-dimensional localization problem, Phys.
4069:   Rev.~E {\bf 54}(1), 231 (1996).
4070: 
4071: \bibitem{NerTsvWen94}
4072: A.~A. Nersesyan, A.~M. Tsvelik and F.~Wenger, Disorder effects in
4073:   two-dimensional $d$-wave superconductors, Phys. Rev. Lett. {\bf 72}(16), 2628
4074:   (1994).
4075: 
4076: \bibitem{OshMogMor93}
4077: G.~Oshanin, A.~Mogutov and M.~Moreau, Steady flux in a continuous-space Sinai
4078:   chain, J.~Stat. Phys. {\bf 73}, 379 (1993).
4079: 
4080: \bibitem{OssFyo05}
4081: A.~Ossipov and Y.~V. Fyodorov, Statistics of delay times in mesoscopic systems
4082:   as a manifestation of eigenfunction fluctuations, Phys. Rev.~B {\bf 71},
4083:   125133 (2005).
4084: 
4085: \bibitem{Ouv94}
4086: S.~Ouvry, $\delta$ perturbative interactions in Aharonov-Bohm and anyons
4087:   models, Phys. Rev.~D {\bf 50}, 5296 (1994).
4088: 
4089: \bibitem{OvcEri77}
4090: A.~A. Ovchinnikov and N.~S. Erikmann, Density of states in a one-dimensional
4091:   random potential, Sov. Phys. JETP {\bf 46}, 340 (1977).
4092: 
4093: \bibitem{Pas98}
4094: M.~Pascaud, {\em Magn\'etisme orbital de conducteurs m\'esoscopiques
4095:   d\'esordonn\'es et propri\'et\'es spectrales de fermions en interaction\/},
4096:   Ph.D. thesis, Universit\'e Paris 11 (1998).
4097: 
4098: \bibitem{PasMon99}
4099: M.~Pascaud and G.~Montambaux, Persistent currents on networks, Phys. Rev. Lett.
4100:   {\bf 82}, 4512 (1999).
4101: 
4102: \bibitem{PerWel96}
4103: M.~Perman and J.~A. Wellner, On the distribution of Brownian areas, Ann. Appl.
4104:   Probab. {\bf 6}(4), 1091 (1996).
4105: 
4106: \bibitem{PieGouAntPotEstBir03}
4107: F.~Pierre, A.~B. Gougam, A.~Anthore, H.~Pothier, D.~Esteve and N.~O. Birge,
4108:   Dephasing of electrons in mesoscopic metal wires, Phys. Rev.~B {\bf 68},
4109:   085413 (2003).
4110: 
4111: \bibitem{PitYor86}
4112: J.~Pitman and M.~Yor, Asymptotic laws of planar Brownian motion, Ann. Probab.
4113:   {\bf 14}(3), 733 (1986).
4114: 
4115: \bibitem{RamShe87}
4116: J.~Rammer and A.~L. Shelankov, Weak localization in inhomogeneous magnetic
4117:   fields, Phys. Rev.~B {\bf 36}(6), 3135 (1987).
4118: 
4119: \bibitem{Ric82}
4120: S.~O. Rice, The integral of the absolute value of the pinned Wiener process --
4121:   Calculation of its probability density by numerical integration, Ann. Probab.
4122:   {\bf 10}(1), 240 (1982).
4123: 
4124: \bibitem{Ric03}
4125: C.~Richard, Area distribution of the planar random loop boundary, J.~Phys.~A:
4126:   Math. Gen. {\bf 37}, 4493 (2003).
4127: 
4128: \bibitem{Rot83a}
4129: J.-P. Roth, Le spectre du Laplacien sur un graphe, in {\em Colloque de
4130:   Th\'eorie du potentiel - Jacques Deny\/}, p. 521, Orsay (1983).
4131: 
4132: \bibitem{Rot83}
4133: J.-P. Roth, Spectre du Laplacien sur un graphe, C. R. Acad. Sc. Paris {\bf
4134:   296}, 793 (1983).
4135: 
4136: \bibitem{RudSch53}
4137: K.~Rudenberg and C.~Scherr, J. Chem. Phys. {\bf 21}, 1565 (1953).
4138: 
4139: \bibitem{Sch57}
4140: H.~Schmidt, Disordered one-dimensional crystals, Phys. Rev. {\bf 105}(2), 425
4141:   (1957).
4142: 
4143: \bibitem{SchTit03}
4144: H.~Schomerus and M.~Titov, Band-center anomaly of the conductance distribution
4145:   in one-dimensional Anderson localization, Phys. Rev.~B {\bf 67}, 100201
4146:   (2003).
4147: 
4148: \bibitem{Sel56}
4149: A.~Selberg, Harmonic analysis and discontinuous groups in weakly symmetric
4150:   Riemannian spaces, with applications to Dirichlet series, J. Ind. Math. Soc.
4151:   {\bf 20}, 47 (1956).
4152: 
4153: \bibitem{Sha82}
4154: B.~Shapiro, Renormalization-Group transformation for the Anderson transition,
4155:   Phys. Rev. Lett. {\bf 48}(12), 823 (1982).
4156: 
4157: \bibitem{Sha83}
4158: B.~Shapiro, Quantum conduction on a Cayley tree, Phys. Rev. Lett. {\bf 50}(10),
4159:   747 (1983).
4160: 
4161: \bibitem{ShaSha82}
4162: D.~{\relax{Yu}}. Sharvin and {\relax{Yu}}.~V. Sharvin, Magnetic-flux
4163:   quantization in a cylindrical film of a normal metal, JETP Lett. {\bf 34}(5),
4164:   272 (1982).
4165: 
4166: \bibitem{SheTsv98}
4167: D.~G. Shelton and A.~M. Tsvelik, Effective theory for midgap states in doped
4168:   spin-ladder and spin Peierls systems: Liouville quantum mechanics, Phys.
4169:   Rev.~B {\bf 57}(22), 14242 (1998).
4170: 
4171: \bibitem{She82}
4172: L.~A. Shepp, On the integral of the absolute value of the pinned Wiener
4173:   process, Ann. Probab. {\bf 10}(1), 234 (1982), [Acknowledgement of priority:
4174:   Ann. Probab. {\bf 19}, 1397 (1991)].
4175: 
4176: \bibitem{Smi60}
4177: F.~T. Smith, Lifetime matrix in collision theory, Phys. Rev. {\bf 118}, 349
4178:   (1960).
4179: 
4180: \bibitem{StaTer96}
4181: H.~M. Stark and A.~A. Terras, Zeta functions of finite graphs and coverings,
4182:   Adv. Math. {\bf 121}, 124 (1996).
4183: 
4184: \bibitem{SteCheFabGog99}
4185: M.~Steiner, Y.~Chen, M.~Fabrizio and A.~O. Gogolin, Statistical properties of
4186:   localization-delocalization transition in one dimension, Phys. Rev.~B {\bf
4187:   59}(23), 14848 (1999).
4188: 
4189: \bibitem{SteFabGog98}
4190: M.~Steiner, M.~Fabrizio and A.~O. Gogolin, Random mass Dirac fermions in doped
4191:   spin-Peierls and spin-ladder systems: one-particle properties and boundary
4192:   effects, Phys. Rev.~B {\bf 57}(14), 8290 (1998).
4193: 
4194: \bibitem{TakLinMak80}
4195: H.~Takayama, Y.~R. {Lin-Liu} and K.~Maki, Continuum model for solitons in
4196:   polyacetylene, Phys. Rev.~B {\bf 21}(6), 2388 (1980).
4197: 
4198: \bibitem{Tan01b}
4199: G.~Tanner, Unitary-stochastic matrix ensembles and spectral statistics,
4200:   J.~Phys.~A: Math. Gen. {\bf 34}, 8485 (2001).
4201: 
4202: \bibitem{Tan05}
4203: G.~Tanner, From quantum graphs to quantum random walks, unpublished  (2005),
4204:   preprint quant-ph/0504224.
4205: 
4206: \bibitem{Tex99}
4207: C.~Texier, {\em Quelques aspects du transport quantique dans les syst\`emes
4208:   d\'esordonn\'es de basse dimension\/}, Ph.D. thesis, Universit\'e Paris 6
4209:   (1999), available at {\sf
4210:   http://ipnweb.in2p3.fr/$\sim$lptms/membres/texier/research.html}.
4211: 
4212: \bibitem{Tex00}
4213: C.~Texier, Individual energy level distributions for one-dimensional diagonal
4214:   and off-diagonal disorder, J.~Phys.~A: Math. Gen. {\bf 33}, 6095--6128
4215:   (2000).
4216: 
4217: \bibitem{Tex02}
4218: C.~Texier, Scattering theory on graphs (2): the Friedel sum rule, J.~Phys.~A:
4219:   Math. Gen. {\bf 35}, 3389--3407 (2002).
4220: 
4221: \bibitem{TexBut03}
4222: C.~Texier and M.~B{\"u}ttiker, Local Friedel sum rule in graphs, Phys. Rev.~B
4223:   {\bf 67}, 245410 (2003).
4224: 
4225: \bibitem{TexCom99}
4226: C.~Texier and A.~Comtet, Universality of the Wigner time delay distribution for
4227:   one-dimensional random potentials, Phys. Rev. Lett. {\bf 82}(21), 4220--4223
4228:   (1999).
4229: 
4230: \bibitem{TexDeg03}
4231: C.~Texier and P.~Degiovanni, Charge and current distribution in graphs,
4232:   J.~Phys.~A: Math. Gen. {\bf 36}, 12425--12452 (2003).
4233: 
4234: \bibitem{TexMon01}
4235: C.~Texier and G.~Montambaux, Scattering theory on graphs, J.~Phys.~A: Math.
4236:   Gen. {\bf 34}, 10307--10326 (2001).
4237: 
4238: \bibitem{TexMon04}
4239: C.~Texier and G.~Montambaux, Weak localization in multiterminal networks of
4240:   diffusive wires, Phys. Rev. Lett. {\bf 92}, 186801 (2004).
4241: 
4242: \bibitem{TexMon05b}
4243: C.~Texier and G.~Montambaux, Dephasing due to electron-electron interaction in
4244:   a diffusive ring, unpublished  (2005), preprint cond-mat/0505199, to appear
4245:   in Phys. Rev. B {\bf72}.
4246: 
4247: \bibitem{TexMon05}
4248: C.~Texier and G.~Montambaux, Quantum oscillations in mesoscopic rings and
4249:   anomalous diffusion, J.~Phys.~A: Math. Gen. {\bf 38}, 3455--3471 (2005).
4250: 
4251: \bibitem{ThoPepAhmAndDav86}
4252: T.~J. Thornton, M.~Pepper, H.~Ahmed, D.~Andrews and G.~J. Davies,
4253:   One-Dimensional Conduction in the 2D Electron Gas of a GaAs-AlGaAs
4254:   Heterojunction, Phys. Rev. Lett. {\bf 56}(11), 1198 (1986).
4255: 
4256: \bibitem{TraWid93}
4257: C.~A. Tracy and H.~Widom, Level-spacing distribution and the Airy kernel, Phys.
4258:   Lett. B {\bf 305}, 115 (1993).
4259: 
4260: \bibitem{TraWid02}
4261: C.~A. Tracy and H.~Widom, Distribution functions for largest eigenvalues and
4262:   their applications, in {\em Proceedings of the ICM\/}, volume~1, p. 587,
4263:   Beijing (2002).
4264: 
4265: \bibitem{Ver94}
4266: J.~Verbaarschot, The spectrum of the Dirac operator near zero virtuality for
4267:   $N_c=2$ and chiral random matrix theory, Nucl. Phys.~B [FS] {\bf 426}, 559
4268:   (1994).
4269: 
4270: \bibitem{Ver79}
4271: W.~Verwaat, On a stochastic difference equation and a representation of non
4272:   negative infinitely divisible random variables, Adv. Appl. Probab. {\bf 111},
4273:   750 (1979).
4274: 
4275: \bibitem{VidMonDou00}
4276: J.~Vidal, G.~Montambaux and B.~Dou\c{c}ot, Transmission through quantum
4277:   networks, Phys. Rev.~B {\bf 62}, R16294 (2000).
4278: 
4279: \bibitem{Weg79}
4280: F.~Wegner, The mobility edge problem: continuous symmetry and a conjecture, Z.
4281:   Phys. B {\bf 35}(3), 207 (1979).
4282: 
4283: \bibitem{Weg83}
4284: F.~Wegner, Exact density of states for Lowest Landau Level in white noise
4285:   potential. Superfield representation for interacting systems, Z. Phys. B -
4286:   Condens. Matter {\bf 51}, 279 (1983).
4287: 
4288: \bibitem{Weg76}
4289: F.~J. Wegner, Electrons in disordered systems. Scaling near the mobility edge,
4290:   Z. Phys. B {\bf 25}(4), 327 (1976).
4291: 
4292: \bibitem{WeiHav86}
4293: G.~H. Weiss and S.~Havlin, Some properties of a random walk on a comb
4294:   structure, Physica A {\bf 134}, 474 (1986).
4295: 
4296: \bibitem{Wer94}
4297: W.~Werner, Sur les points autour desquels le mouvement brownien plan tourne
4298:   beaucoup, Probab. Theory Rel. {\bf 99}, 111 (1994).
4299: 
4300: \bibitem{WinRooChaPro86}
4301: S.~Wind, M.~J. Rooks, V.~Chandrasekhar and D.~E. Prober, One-Dimensional
4302:   Electron-Electron Scattering with Small Energy Transfers, Phys. Rev. Lett.
4303:   {\bf 57}(5), 633 (1986).
4304: 
4305: \bibitem{Win03}
4306: B.~Winn, {\em The Laplacian on a graph and quantum chaology\/}, Ph.D. thesis,
4307:   University of Bristol (2003), available at {\sf
4308:   http://www.math.tamu.edu/$\sim$bwinn/thesis.htm}.
4309: 
4310: \bibitem{WuKun99}
4311: F.~Y. Wu and H.~Kunz, Restricted random walks on a graph, Ann. Combin. {\bf 3},
4312:   475 (1999).
4313: 
4314: \bibitem{Yor80}
4315: M.~Yor, Loi de l'indice du lacet brownien et distribution de Hartman-Watson,
4316:   Z.~Wahrscheinlichkeit. {\bf 53}, 71 (1980).
4317: 
4318: \bibitem{Yor00}
4319: M.~Yor, {\em Exponential functionals of Brownian motion and related
4320:   processes\/}, Springer (2000).
4321: 
4322: \bibitem{Zir96}
4323: M.~R. Zirnbauer, Riemannian symmetric superspaces and their origin in
4324:   random-matrix theory, J. Math. Phys. {\bf 37}, 4986 (1996).
4325: 
4326: \end{thebibliography}
4327: 
4328: 
4329: \end{document}
4330: