0809.2726/nuy.tex
1: \documentclass[11pt]{article} %\input epsf.tex
2: \usepackage{amsmath,amsfonts,graphicx,mathrsfs} %bbm}
3: %\include{graphics}
4: 
5: 
6: %***********************************************************************
7: \setlength{\textheight}{22.0cm} \setlength{\topmargin}{-1cm}
8: \setlength{\textwidth}{16.0cm} \setlength{\parskip}{0.12cm}
9: \setlength{\rightmargin}{0.7cm} \hoffset=-1.00 true cm
10: \setlength{\parindent}{0cm} \addtolength{\abovedisplayskip}{2.0mm}
11: \addtolength{\belowdisplayskip}{2.0mm}
12: \addtolength{\abovedisplayshortskip}{2.0mm}
13: \addtolength{\belowdisplayshortskip}{2.0mm}
14: \addtolength{\abovedisplayskip}{2.0mm}
15: \addtolength{\belowdisplayskip}{2.0mm}
16: \addtolength{\abovedisplayshortskip}{2.0mm}
17: \addtolength{\belowdisplayshortskip}{2.0mm}
18: \addtolength{\footnotesep}{2.0mm}
19: 
20: 
21: %*********************************************************************
22: \renewcommand{\theequation}{\thesection.\arabic{equation}}
23: \newcommand{\be}{\begin{align}}
24: \newcommand{\ee}{\end{align}}
25: \newcommand{\bear}{\begin{eqnarray}}
26: \newcommand{\eear}{\end{eqnarray}}
27: 
28: 
29: %*****Some macros for making vectors and matrices:********************
30: \newcommand{\mvec}[2]{\left(\begin{array}{c}#1\\ #2\end{array}\right)}
31: \newcommand{\bigvec}[3]{\left(%
32: \begin{array}{c}#1\\ #2\\ #3\end{array}\right)}
33: \newcommand{\mat}[4]{\left(%
34: \begin{array}{cc}#1&#2 \\#3&#4\end{array}\right)}
35: \newcommand{\bigmat}[9]{\left(%
36: \begin{array}{ccc}#1&#2&#3\\#4&#5&#6\\#7&#8&#9\end{array}\right)}
37: \newcommand{\str}{\mathrm {Str}}
38: \newcommand{\tr}{\mathrm{Tr}}
39: \newcommand{\sdet}{\mathrm {Sdet}}
40: %*****Some macros for Dirac's bra and ket notation:*******************
41: \newcommand{\cba}{\bar c}
42: \newcommand{\dba}{\bar d}
43: \newcommand{\1}{{\sf 1 \!\! 1}}
44: \newcommand{\ba}{\begin{array}}
45: \newcommand{\ea}{\end{array}}
46: \newcommand{\lae}{\begin{array}{c}\,\sim\vspace{-21pt}\\<\end{array}}
47: \newcommand{\gae}{\begin{array}{c}\,\sim\vspace{-21pt}\\>\end{array}}
48: \newcommand{\nn}{\nonumber}
49: \newcommand{\diag}{\textrm{diag}}
50: \newcommand{\s}{\ \! }
51: \newcommand{\zbar}{\bar{z}}
52: \newcommand{\qbbar}{\bar{q}_b}
53: \newcommand{\qfbar}{\bar{q}_f}
54: \def\vbr{$\vphantom{\sqrt{F_e^i}}$}% vertical brace for tables
55: 
56: \parindent0pt
57: \parskip10mm
58: 
59: 
60: \begin{document}
61: \title{\bf On the mean density of complex eigenvalues for an ensemble of
62: random matrices with prescribed singular values}
63: \author{Yi Wei $^{1,2}$ and Yan V. Fyodorov $^{2,1}$\\\\
64:  $^1$ Isaac Newton Institute for Mathematical Sciences, Cambridge,
65: UK
\\$^2$ School of Mathematical Sciences, University of Nottingham,
66: UK}
67: \date{}
68: 
69: \maketitle \vspace*{-0mm} \abstract{Given any fixed $N \times N$
70: positive semi-definite diagonal matrix $G\ge 0$ we derive the
71: explicit formula for the density of complex eigenvalues for random
72: matrices $A$ of the form $A=U\sqrt{G}$} where the random unitary
73: matrices $U$ are distributed on the group $\mathrm{U(N)}$
74: according to the Haar measure.
 
75: 
76: 
77: 
78: \section{Introduction}
79: The question of characterizing the locus of complex eigenvalues
80: for an $N\times N$ matrix $A$ with prescribed singular values, that is
81:  eigenvalues $g_i\ge 0,\,\, i=1,\ldots, N$ of $ A^{*} A$,
, with $A^*$ being Hermitian conjugate of $A$, 
82:   was considered in classical papers by Horn
83: and Weyl \cite{sing}. In the present paper we provide a kind of
84: statistical answer to that question. Define
85: $G=\diag(g_1,\ldots,g_N)\ge 0$, multiply the matrix $\sqrt{G}$ by a
86: general unitary transformation $U$ from the left and average over
87: the unitary group $\mathrm{U}(N)$ with the invariant (Haar) measure. This
88: construction induces a natural measure on the set of matrices
89: $ A=U\sqrt {G}$ with given singular values $g_i$ and in this way
90: provides us with the corresponding random matrix ensemble. The
91: complex eigenvalues $z_i$ of such matrices $A$ will cover
92: generically an annular domain $R_{min}<|z|<R_{max}$ 
93: in the complex plane with some density $\rho(z)$. In such 
94: an approach the statistical
95: characterization of the locus of complex eigenvalues amounts to
96: knowing the profile of the ensemble averaged value of that density  
97: for a given set of singular values.  It is easy to understand that such mean density
98:  can depend only on $|z|$. The corresponding explicit formula is
99: provided in {\bf Theorem 2.1} which is the main result of
100: the paper.
101: 
102: Apart from being a rather nontrivial mathematical problem,
103: understanding the statistical properties of complex eigenvalues of
104: the above-mentioned ensemble is motivated by its applications in the
105: domain of Quantum Chaotic Scattering. In this capacity the problem
106: attracted attention for some time, and a few partial results were
107: obtained previously in several limiting
108: cases\cite{trunc,prel,FSJETP,rev03,yb05}. Below we give a brief
109: description of the physical context related to the problem.
110: 
111: 
112:  One of  very useful instruments in the
113: analysis of classical Hamiltonian systems with chaotic dynamics
114: are the so-called area-preserving chaotic maps, see e.g.
115: \cite{Haakebook} and references therein. They appear naturally either as a mapping of
116: the Poincar\'{e} section onto itself, or as a result of a
117: "stroboscopic" description of Hamiltonians which are periodic
118: functions of time. Their quantum mechanical analogues are unitary
119: operators which act on Hilbert spaces of finite large dimension
120: $N$, and are often referred to as evolution, scattering or Floquet
121: operators, depending on the given physical context. Their
122: eigenvalues consist of $N$ points on the unit circle
123: (eigenphases). Numerical studies of various classically chaotic
124: systems suggest that the eigenphases conform statistically quite
125: accurately the results obtained for unitary random matrices of a
126: particular symmetry (Dyson circular ensembles).
127: 
128: Let us now imagine that a system represented by
129:  a chaotic map ("inner world") is embedded in a
130: larger physical system ("outer world") in such a way that it
131: describes particles which can come inside the region of chaotic
132: motion and leave it after some time via $M$ open channels. Models
133: of such type appeared in various disguises for example, in
134:  \cite{KolRev,JSB,OKG,Prange}
135:  and most
136: recently discussed in much detail in relation to properties of dielectric 
137: microresonators in \cite{KNS}. A natural
138: mathematical framework allowing to deal efficiently with such a
139: situation was suggested in \cite{FSJETP}, see also \cite{rev03}
140: and we mention here only its gross features. For a closed quantum
141: system characterized by a wavefunction $\Psi$
142:  the "stroboscopic" (discrete-time) dynamics amounts to a
143: linear unitary map $\Psi(n)\to\Psi(n+1)$, such that
144: $\Psi(n+1)=\hat{u}\Psi(n)$. The unitary evolution operator
145: $\hat{u}$ describes the "closed" inner state domain decoupled both
146: from input and output spaces. Then a coupling that makes the
147: system open must convert the evolution operator $u$ to a
148: contractive operator $\hat{A}$ such that ${\bf
149: 1}-\hat{A}^{*}\hat{A}\ge 0$. It is easy to show that one can
150: always choose $\hat{A}=\hat{u} \sqrt{{\bf
151: 1}-\hat{\tau}\hat{\tau}^{*}}$ where the matrix $\hat{\tau}$
152: is a rectangular $N\times M,  M \le N $ diagonal with the entries
153: $\tau_{ij}=\delta_{ij}\tau_j\, , \, 1\le i\le N\, ,\, {1\le j\le
154: M} ,\quad 0\le \ \tau_j\le 1$. With $\hat{u}$ replaced by $\hat{A}$,  the equation
155: $\Psi(n+1)=\hat{A}\Psi(n)$ then describes an irreversible decay of
156: any initially prepared state $\Psi(0)\ne 0$, assuming that 
157:  external input is absent during the subsequent evolution.
158: The complex eigenvalues $z_k$ of the operator $\hat{A}$ all belong
159: to the interior of the unit circle $|z|<1$ and play the role of
160: resonances for the discrete time systems. Let us mention that
161:  various aspects of
162:  resonances associated with quantum chaotic open
163: maps recently attracted considerable attention\cite{KNS,Nonnem}.
164: 
165: 
166: The relation with the random matrix construction
167:  employed in this paper is now obvious.
168: It amounts to replacing the true
169: evolution operator  
170: $\hat{u}$ with its random matrix analogue $u$ taken from the Circular Unitary Ensemble (CUE)
171:  in accordance with the standard
172: ideas of Quantum Chaos,  and to denoting $G=
173: \sqrt{{\bf 1}-\hat{\tau}\hat{\tau}^{*}}$.
174:  In this way the task of studying resonances is reduced
175: to investigating eigenvalues of  random matrices $ A$ of the specified type.
176: \footnote{
177: Let us however note that the case most frequently encountered in
178:  direct physical applications actually
179: corresponds to choosing $u$ 
180: to be {\it unitary symmetric} \cite{KNS} taken from COE.
181:  Such a choice reflects the inherent 
182: time-reversal invariance typical for the
183: the closed quantum chaotic system. A somewhat simpler choice of unrestricted unitary random matrices from CUE
184: corresponds to system with broken time reversal invariance.}
185: 
186: 
187: 
188: 
189: The first significant random matrix result on such matrices $A$ 
190: seems to have appeared in
191: \cite{trunc} where the authors considered so-called "truncations" of  random unitary
192: matrices. In our notations that case is equivalent to
193: taking $g_1=g_2=\ldots g_M=0$, with all the rest
194: $N-M$ of $g_i$ being equal to unity. 
195: In \cite{FSJETP} the mean density of complex eigenvalues was
196: derived in the limit $N\to \infty$, with
197: $M<\infty$ being fixed and all $g_i\le 1$.   The
198: work \cite{rev03} provided some 
199: general results
200: on the joint probability
201: density of all $N$ complex eigenvalues $z_i$, as well as a few
202: formulae for the few-point correlation functions (the so-called marginal
203: distributions) of the eigenvalue densities. Those formulae however only
204: lead to
205:  explicit managable expressions again 
206:  in the same limiting case as in \cite{FSJETP}.
207: As to the results
208: valid for arbitrary finite $N$, only the simplest particular case
209: $g_1=g_2=\ldots=g_{N-1}=g,\, g_N<g$ was so far addressed by a
210: variety of methods, see \cite{yb05} for the most recent account
211: and \cite{prel} for an early consideration.
212: 
213: \section{Statement of the Main Results}
214: 
215: Our results show that the ensemble-averaged eigenvalue density function $\rho(z)=\Psi(|z|^2)$, i.e. 
216: indeed depends only on $|z|$. To write the function $\Psi$ explicitly we need to introduce
217: a few notations.
218: 
219: 
220:  Let $\mathrm{s}^l$ be the $l$-th
221: order elementary symmetric polynomials of $g_i,\ i=1,\dots,N$, e.g. $\mathrm{s}^0=1,
222: \ \mathrm{s}^1=\sum_{i=1}^Ng_i,\ \mathrm{s}^2=\sum_{i<j}^Ng_ig_j,\ \dots$, etc.
223: Let us denote $\mathrm{s}^l_{[i_1,i_2,\dots]}=\mathrm{s}^l|_{g_{i_1}=g_{i_2}=\dots=0}$.
224: Define the following functions of $\{g_i\}$ and the complex variable $z$:
225: \begin{align}
226: \mathrm{F}_-(g_i)&=-\frac{1}{{{\prod'}_{j=1}^{N}} (g_i-g_j)}N g_i^{N-1}
227: \sum\limits_{l=0}^{N-1}\mathrm{s}^l_{[i]}|z|^{2(-l-1)}\,\frac{l}{\left(\!\!\!\begin{array}{c}
228: N\!-\!1 \\ l \end{array}\!\!\!\right)}\\
229: \mathrm{F}_+(g_i)&=\mathrm{F}_-(g_i)+\mathrm{F}_\Delta(g_i)\\
230: \label{eq:fdelta}\mathrm{F}_\Delta(g_i)&
231: =\frac{1}{{\prod\limits_{j=1}^{N}}^\prime (g_i-g_j)}(g_i-|z|^2)^{N-2}
232: \sum\limits_{l=0}^{N-1}\mathrm{s}^l_{[i]}|z|^{-2(l+1)}\frac{1}{\left(\!\!\!\begin{array}{c}
233:  N \!-\!1 \\ l \end{array}\!\!\!\right)}
234: \bigg[lg_i+(N-1-l)|z|^2\bigg]\nn\\
235: &=\frac{(g_i-|z|^2)^{N-2}}{{\prod\limits_{j=1}^{N}}^\prime (g_i-g_j)}\int_0^\infty
236: \frac{Ndt}{(1+t)^{N+2}}\det\bigg(1+\frac{t}{|z|^2}G_{[i]}\bigg)
237: \bigg[N-t+\frac{g_i}{|z|^2}(Nt-1)\bigg]\;,
238: \end{align}
239: where we used the binomial function \small{$\left(\!\!\begin{array}{c}
240: N\\l\end{array}\!\!\right)=\frac{N!}{l!(N-l)!}$}.
241: In Eq.\eqref{eq:fdelta}, we defined a matrix $G_{[i]}=\diag(g_1,\dots,g_{i-1},
242: g_{i+1},\dots,g_N)$ and used the dash to denote that the index $j$ can not be equal to $i$
243: in the product. 
244: 
245: The main statement of the paper is that the density of complex eigenvalues
246: can be written in terms of the functions defined above. More precisely, we state the following\\
247: 
248: \noindent{\bf{Theorem} 2.1} Let $U\in \mathrm{U}(N)$ be an element of the unitary
249: group and $G=\diag(g_1,\dots,g_N)$ be a fixed positive diagonal matrix, such that
250: $0<g_1<\dots<g_N<\infty$. Let $U$ be distributed on $\mathrm{U}(N)$ according to the Haar measure.
251: Then the mean density $\rho(z)$ of complex eigenvalues of the matrix  $A=U\sqrt{G}$
252:  is given by
253: \begin{align}\label{eq:sum}
254: \rho(z)=\Psi(|z|^2)=\frac{1}{N}\sum_{i=1}^N \mathrm{F}_\sigma(g_i)\;,
255: \end{align}
256: where $\sigma=+$ for $|z|^2>g_i$, $\sigma=-$ for $|z|^2<g_i$.\\
257: 
258: \noindent{\bf{Remark}}:
259: Exploiting  that the function $\mathrm{F}_-$ is totally antisymmetric with respect to all $g_i$'s, we
260: can rewrite the above expression in the form:\\
261: 
262: \begin{align}\label{eq:Psi}
263: \Psi(|z|^2)=\left\{ \begin{array}{ll} 0 & |z|^2<g_1<g_2\dots<g_N\\
264: \!\! \frac{1}{ N}\sum\limits_{i=k+1}^{N}\mathrm{F}_\Delta(g_i)& g_1<\dots<g_k<|z|^2<g_{k+1}<\dots<g_N
265: \\ 0 & g_1<g_2\dots<g_N<|z|^2 \end{array}\right.
266: \end{align}
267: 
268: \noindent{\bf{Remark}}: For $N>3$, it is not difficult to show that the eigenvalue density function is 'smooth' at
269: each $|z|=g_i, 1< i< N$, that is it has a continuous derivative. 
270: When $N=3$, the density function is only 'continuous' at $g_2$ but not 'smooth'.\\
271: 
272: \noindent In the case of degenerate eigenvalues of the matrix $G$ the density function can be derived
273: from Theorem 2.1 by taking the corresponding limits as shown below. \\
274: 
275: \noindent{\bf{Corollary} 2.2}: Suppose the diagonal matrix $G$ has the following
276: degeneracies,
277: \begin{align}\label{eq:degenerate}
278: g_{k_1}=\dots=g_{k_1+i_1},\ \ g_{k_2}=\dots=g_{k_2+i_2}, \ \ \dots, \ \ g_{k_s}=\dots=g_{k_s+i_s}\;,
279: \end{align}
280: which is denoted by the short-hand notation
281: \begin{align}
282: {G}=\diag(\dots,[g_{k_1},\dots,g_{k_1+i_1}],\dots,[g_{k_s},\dots,g_{k_s+i_s}],\dots)\;.
283: \end{align}
284: Define the following two functions
285: 
286: \begin{align}
287: \mathrm{f}^{[k,i]}_n(g)\overset{\mbox{\tiny def}}{=}\frac{(g-|z|^2)^{N-2}}
288: {\prod\limits_{j=1}^{k-1}\prod\limits_{j=k+i+1}^{N}(g-g_j)}\sum\limits_{l=n}^{N-1}
289: \mathrm{s}^{l-n}_{[k,\dots,k+n]}|z|^{-2(l+1)}\frac{1}{\left(\begin{array}{c}\!\!\!
290: N\!-\!1\\ l \end{array}\!\!\!\right)}\bigg[lg+(N-1-l)|z|^2\bigg]\;,
291: \end{align}
292: 
293: \begin{align}
294: \mathrm{F}^{[k,i]}_\Delta\stackrel{\mbox{\tiny def}}{=}\sum\limits_{n=0}^{i}\frac{(-)^n}{(i-n)!}
295: \frac{d^{i-n}}{dg^{i-n}_{k+n}}\mathrm{f}^{[k,i]}_n(g_{k+n})\;.
296: \end{align}
297: The density function is then given by replacing each $\sum_{n=0}^i\mathrm{F}_\Delta(g_{k+n})$ in
298: the Theorem by $\mathrm{F}_\Delta^{[k,i]}$ and making substitutions Eq.{\eqref{eq:degenerate}}.\\
299: 
300: \noindent{\bf{Proof}:}
301: Consider the following sum: $\psi=\sum_{n=0}^i\mathrm{F}_\Delta(g_{k+n})$ when $g_{k}=\dots=g_{k+i}$.
302: Taking the limit $g_{k+i-1}\to g_{k+i}=g$ yields
303: \begin{align}
304: &\lim_{g_{k+i-1}\to g_{k+i}=g}\psi\nn\\
305: =&\left(\sum\limits_{n=0}^{i-2}
306: \mathrm{F}_\Delta(g_{k+n})+\frac{d}{dg_{k+i-1}}(g_{k+i-1}-g_{k+i})
307: (\mathrm{F}_\Delta(g_{k+i-1})+\mathrm{F}_\Delta(g_{k+i}))\right)_{g_{k+i-1}=g_{k+i}=g}\nn\\
308: =&\left(\sum\limits_{n=0}^{i-2}
309: \mathrm{F}_\Delta(g_{k+n})+\frac{d}{dg_{k+i-1}}
310: \left[\mathrm{f}_0^{[k+i-1,1]}(g_{k+i-1})-\mathrm{f}_1^{[K+i-1,1]}
311: (g_{k+i})\right]\right)_{g_{k+i-1}=g_{k+i}=g}\nn\\
312: =&\ \mathrm{F}_\Delta^{[k+i-1,1]}|_{g_{k+i-1}=g_{k+i}=g}\;.
313: \end{align}
314: In the second step we have exploited the linearity of $\mathrm{s}^l$ as a function of $g$'s.
315: Next, we take the limit $g_{k+i-2}\to g$. Repeating the procedure $i$ times, we arrive at
316: \begin{align}\label{eq:single-deg}
317: \psi|_{g_{k}=\dots=g_{k+i}=g}=\mathrm{F}_\Delta^{[k,i]}|_{g_{k}=\dots=g_{k+i}=g}\;.
318: \end{align}
319: Applying the results Eq.{\eqref{eq:single-deg}} to other degeneracies, we obtain the eigenvalue density
320: function for $U\sqrt{G}$ under the condition Eq.{\eqref{eq:degenerate}}.\hspace*{10mm}$\Box$\\
321: 
322: 
323: \noindent{{Example 1}\bf (rank-one deviation from unitary matrix):}\\ In the special case of
324: $G=\diag(g_1,[g_2,\dots,g_N])$, with $g_2=\dots=g_N=g$, the above procedure leads to
325: an especially simple formula for the mean eigenvalue density:
326: \begin{align}\label{eq:rank1}
327: \Psi=\frac{(|z|^2-g_1)^{N-2}}{(g-g_1)^{N-1}|z|^{2N}}\!\left((N-1)(|z|^{2N}+g^{N-1}g_1)+\!\!
328: \sum\limits_{k=0}^{N-2}\left[(N-2-k)g+kg_1\right]g^k|z|^{2(N-1-k)}\right)\;.
329: \end{align}
330: which coincides with the known result \cite{prel,yb05}.\\
331: 
332: \noindent{\bf{Remark}}: In Eq.{\eqref{eq:rank1}}, as $g\to g_1$, the density function $\Psi\to\infty$
333: on $[g_1,g]$ and is zero otherwise. On the other hand, the integration of $\Psi$ over the
334: region $[g_1,g]$ yields one.  We conclude that in this case the density
335: function is simply $\delta(g-|z|^2)$, as it must be for a random matrix $A=U\sqrt{g}$ which is 
336: simply proportional to CUE matrix.\\
337: 
338: \noindent {\bf{Remark}}: In fact, in our derivation of the main theorem, we can extend the
339: domain of $g_i$'s to include the origin, i.e. our formula holds for $g_i\ge0$. We illustrate
340: this observation in the following example.\\
341: 
342: \noindent{{Example 2}\bf (truncated unitary matrix):}\\
343: Consider the case  $G=\diag([g_1,\dots,g_M],[g_{M+1},\dots,g_N])$,
344: where $g_1=\dots=g_M=0$ and $g_{M+1}=\dots=g_N=1$. By Corollary, we can write
345: the density function of eigenvalues of $A=U\sqrt{G}$ as
346: \begin{align}\label{eq:truncate}
347: \Psi=&\ \mathrm{F}^{[M+1,N-M-1]}_\Delta|_{g_1=\dots=g_M=0,g_{M+1}=\dots=g_N=1}\nn\\
348: \propto&\ (1-|z|^2)^{M-1}\left(\frac{d}{d|z|^2}\right)^M\frac{1-|z|^{2N}}{1-|z|^2}\;.
349: \end{align}
350: In the first step, we have extended domains of $g$ to $[0,\infty)$ in Theorem 2.1, and
351: correspondingly, Corollary 2.2. In fact, when $G=\diag(0\cdot I_M,I_{N-M})$, eigenvalues
352: of the matrix $A$ in Example 2. coincide with those of $(N-M)\times(N-M) $ lower right sub-block
353: of a random unitary matrix, also known as the 'truncated'
354: unitary matrix. Same results as
355: Eq.{\eqref{eq:truncate}} has been obtained with a  completely different method in \cite{trunc}. \\
356: 
357: \noindent {\bf{Remark}}: As we can obviously always absorb the $\mathrm{U}^N(1)$ phase of
358: $N\times N$ complex diagonal matrix into $U$, the domain for the matrix $ G$ can
359: be defined on $\mathbb{C}_1^N$. The eigenvalue density function of $U\sqrt{G}$,
360: averaged over CUE is then obtained by substituting $g_i\to |g_i|$ into Theorem 2.1.\\
361: 
362: \noindent Finally, we compare our formula Eq.\eqref{eq:Psi} with numerical simulations. To this
363: end, we choose a fixed diagonal matrix $ G$ and generate unitary matrices according to the Haar measure.
364: We draw a histogram of the radial part of eigenvalues, $|z|$, of the matrix $U\sqrt{G}$,
365: see Fig.\ref{histo} To compare with the histogram, we define the appropriately modified density function
366: $\Psi_1(|z|)=2|z|\Psi(|z|^2)$, which is shown by the solid line. From Fig. \ref{histo}, we observe a
 very
367: good match between our formula Eq.\eqref{eq:Psi} and the results of numerical simulations.
368: 
369: 
370: \hspace{75mm}
371: 
372: \begin{figure}
373: \centering
374: \includegraphics[width=7.5cm, height=6cm]{histo-vs-curve.eps}
375: \caption{Histogram of the radial part of the eigenvalue distribution of $5\times5$
376: matrices $A=U\sqrt{G}$. Here $G=\diag(1,4,9,16,25)$ and  $U$ is the  Haar-distributed unitary
377: matrix, with sample size 100000 and bin 0.10. Solid
378: line represents the function $\Psi_1(|z|)$ as derived from Eq.\eqref{eq:Psi}.}
379: \label{histo}
380: \end{figure}
381: 
382: 
383: \section{Main steps of the proof}
384: \subsection{Colour-flavour transformation}
385: Our starting expression is the following formula \cite{FKS,rev96}
386: for the averaged density of complex eigenvalues for a general
387: finite-size non-Hermitian random matrix $A$
388: \begin{align}\label{eq:formula}
389: \rho(z)=-\frac{1}{\pi}
390: \lim_{\kappa\to0}\frac{\partial}{\partial{\bar z}}\lim_{z_b\to z}
391: \frac{\partial}{\partial{z_b}}\Bigg<\frac{\det\left(
392: \begin{array}{cc}\kappa&i(z-A)\\ i(\bar{z}-A^\dagger)&\kappa\end{array}\right)}{\det\left(
393: \begin{array}{cc}\kappa&i(z_b-A)\\ i(\bar{z}_b-A^\dagger)&\kappa\end{array}\right)}\Bigg>_U\;.
394: \end{align}
395: In our case $A=U\sqrt{G}$, where $U\in \mathrm{U}(N)$, $G=\diag(g_1,\dots,g_N)>0$.
396: Averaging over the Haar measure on the unitary group $\mathrm{U}(N)$ is denoted by $<\cdots>_U$.
397: 
398: Introduce vectors $S_a=(s_a^i)$ with complex components and their counterparts $\chi_a=(\chi_a^i)$ 
399: with anti-commuting components (Grassmann variables), for all $i=1,\dots,N$ and $a=1,2$. 
400: This defines two sets of (graded) vectors
401: $\psi^i_{a}=\left(\begin{array}{c} s^i_{a}\\ \chi^i_{a}\end{array}\right)$ with $a=1,2$.
402: The determinants can be represented as integrals over complex variables $s$ and Grassmann variables $\chi$
403: in the standard way:
404: \begin{align}\label{eq:int-before-cft}
405: <\cdots>_U\propto&\int dU \int dS_1 dS_2 \exp\left[-\kappa(S_1^\dagger S_1+S_2^\dagger S_2)
406: -i(z_bS^\dagger_1S_2+\bar{z}_b S^\dagger_2S_1)\right]\nn\\
407: &\int d\chi_{1}d\bar{\chi}_1d\chi_{2}d\bar{\chi}_2\
408: \exp\left[-\kappa(\chi_1^\dagger \chi_1+\chi_2^\dagger \chi_2)
409: -i(z\chi^\dagger_1\chi_2+\bar{z}\chi^\dagger_2\chi_1)\right]\nn\\
410: &\ \exp i\left[\bar\psi_{1}^iA_{ij}\psi_2^j+\bar\psi_{2}^iA_{ij}^\dagger\psi_1^j\right]\;,
411: \end{align}
412: where we defined $dS_1\,dS_2=\prod_{a=1}^2\prod_{i=1}^N d\bar{s}_a^ids_a^i$.
413: Note that by Eq.{\eqref{eq:formula}}, averaging over the unitary group
414: in the above expression should be carried out after performing the integral over the graded vectors $\{\psi,\bar\psi\}$.\\
415: 
416: Next we change the order of integration over $\{\psi,\bar\psi\}$ and $\mathrm{U}(N)$, which is possible due
417: to the fact that $\mathrm{U}(N)$ is compact and the integral is bounded. The integration over the unitary
418: group can be  performed explicitly by exploiting the colour-flavour transformation discovered 
419: by Zirnbauer \cite{cft}:
420: \begin{align}\label{eq:cft}
421: \int dU \exp \mathrm{i}\left[\bar\psi_{1}^iA_{ij}\psi_2^j+\bar\psi_{2}^iA_{ij}^\dagger\psi_1^j\right]
422: =\int D(Q,\tilde{Q})\ \exp \mathrm{i}
423: \left[ \bar{\psi}_{1}^iQ\psi_{1}^i+g_i\bar{\psi}_{2}^i\tilde{Q}\psi_{2}^i\right]\;.
424: \end{align}
425: Such a transformation trades the integration over $\mathrm{U}(N)$, where N can be an arbitrary
426: large integer for the integration over a considerably simpler $2\times2$ graded matrices $Q$ defined as
427: \begin{align}\label{eq:cfta}
428: Q=\left(\begin{array}{cc} q_b&\eta_1\\ \eta_2&q_f\end{array}\right),
429: \hspace*{5mm} \tilde{Q}=\left(\begin{array}{cc} \bar{q}_b&\sigma_1\\ \sigma_2&-\bar{q}_f\end{array}\right)\;.
430: \end{align}
431: Such $Q$ belongs to a Riemannian symmetric superspace \cite{sym-space} of the type $\mathrm{AIII|AIII}$.
432: Here, $\eta$'s and $\sigma$'s are anti-commuting Grassmann variables. The so-called boson-boson
433: and fermion-fermion blocks of $Q$ are given by
434: \begin{align}\label{eq:cftb}
435: q_b\in\mathrm{U(1,1)/U(1)\times U(1)}=\mathrm{H}^2\ {\mathrm{and}}\ q_f\in\mathrm{U(2)/U(1)\times U(1)=S^2}\;.
436: \end{align}
437: The invariant measure on this domain is defined as
438: \begin{align}
439: D(Q,\tilde Q)=\sdet^N(1-\tilde QQ)dQd\tilde{Q}\;.
440: \end{align}
441: where $dQd\tilde{Q}=d^2q_bd^2q_fd\sigma_1d\sigma_2d\rho_1d\rho_2$. 
442: 
443: 
444: After the colour-flavour transformation Eq.{\eqref{eq:cft}}, we get
445: \begin{align}\label{eq:qcft}
446: <\cdots>_U\ \propto F(\kappa)=\int d\bar{\psi}d\psi\int D(Q,\tilde Q)\exp-(\bar{\psi}_1^i,\bar{\psi}_2^i)
447: \left(\begin{array}{cc}\kappa-iQ&\mathrm{i}Z \\ & \\ \mathrm{i}\bar{Z} &
448: \kappa-ig_i\tilde{Q}\end{array} \right)\left(\begin{array}{l}\psi_1^i\\ \psi_2^i\end{array}\right)\;,
449: \end{align}
450: where $d\bar{\psi}d\psi=dS_1^2dS_2^2d\chi_{1}d\bar{\chi}_1d\chi_{2}d\bar{\chi}_2$ and we defined
451: \begin{align}
452: Z=\left(\begin{array}{cc}z_b&0\\0&z\end{array}\right)\ \ \mathrm{and}\ \
453: \bar{Z}=\left(\begin{array}{cc}\zbar_b&0\\0&\zbar\end{array}\right)\;.
454: \end{align}
455: For a fixed complex number $z$ and a given diagonal matrix $ G$, the integral in Eq.{\eqref{eq:qcft}},
456:  defines a function $F(\kappa)$ of the variable $\kappa$. Using
457: the integral representation for Bessel functions we can show that $F(\kappa)$ is analytic in
458: the half plane $\mathrm{Re}\ \kappa\!>\!0$.
459: 
460: \subsection{Integration over $ Q$ and analytic continuation}
461: 
462: Direct evaluation of the integral over
463: the graded matrix $Q$ followed by the integration over $\psi$ in Eq.{\eqref{eq:qcft}} is very
464: difficult. In fact, it is already a highly involved task in a much simpler case where $Q$ is a complex
465: number and $\psi$ is a complex vector, see \cite{yb05} for the 
466: corresponding calculation in such a case. A natural way out could be changing the
467: order of integration in Eq.{\eqref{eq:qcft}} in order to integrate first over $\{\psi,\bar\psi\}$
468: by using the standard Gaussian integral formula for graded vectors. However, extra care must be
469: taken in performing such a change. To understand this consider the integral involving the boson-boson 
470: part of the supermatrix
471: $Q$ and the complex vectors $S_1$ and $S_2$,
472: \begin{align}\label{eq:ibosonic}
473: I_{\mathrm{bosonic}}=\int dS_1dS_2\int_{|q_b|\le1} dq_b^2\
474: \mathrm{e}^{-\kappa(S_1^\dagger S_1+S_2^\dagger S_2)-(iz_bS_1^\dagger S_2+i\bar{z}_bS_2^\dagger S_1)
475: +iq_bS_1^\dagger S_1+i\bar{q}_bg_i\bar{S}_2^iS_2^i}\;,
476: \end{align}
477: where we have omitted the trivial Grassmann integrals. Changing the order of integration in
478: Eq.\eqref{eq:qcft} we arrive at
479: \begin{align}\label{eq:itilde}
480: \tilde{I}_{\mathrm{bosonic}}=\int_{|q_b|\le1} dq_b^2\int dS_1dS_2\
481: \mathrm{e}^{-\kappa(S_1^\dagger S_1+S_2^\dagger S_2)-(iz_bS_1^\dagger S_2+i\bar{z}_bS_2^\dagger S_1)
482: +iq_bS_1^\dagger S_1+i\bar{q}_bg_i\bar{S}_2^iS_2^i}\;.
483: \end{align}
484: It is clear that $\tilde{I}_{\mathrm{bosonic}}$ is only well-defined in $\mathrm{Re}\,\kappa\in(1,\infty)$.
485: For $\kappa\to 0$, which is the limit we have to perform in the very end
486: of the calculation, the integration over the boson-boson
487: domain forbids changing integration order in Eq.{\eqref{eq:qcft}}.
488: Actually, such a problem was first noticed in \cite{yb05}, and solved by modifying
489: in a non-trivial way the domain of integration Eq.\eqref{eq:cftb} over bosonic variables
490:  in the colour-flavour transformation. After such a modification one can 
491: actually carry out the required change of integration order for any $\kappa>0$.
492: We shall however see that one can work in the standard parametrisation Eq.\eqref{eq:cftb} in the 
493: allowed region $\mathrm{Re} \kappa\!\!>\!\!1$, and then continue to  $ 0<\mathrm{Re}\,\kappa<1$
494: exploiting analytic properties of the function $F(\kappa)$.
495: 
496: Let us from now on work in the domain $\mathrm{Re}\ \kappa\!\!>\!\!1$.
497: Substituting the transformation Eq.{\eqref{eq:cft}} into Eq.{\eqref{eq:int-before-cft}},
498: changing the order of integrations over graded vectors $\{\psi,\bar\psi\}$ and the graded matrix
499: $Q$ and integrating out $\{\psi,\bar\psi\}$, we get
500: \begin{align}\label{eq:int-after-cft}
501: G(\kappa)=&\int dQd\tilde{Q}\ \sdet^N(1-\tilde{Q}Q)\prod^N_{i=1}\sdet^{-1}
502: \left[\begin{array}{cc}\kappa-iQ&i \left(\begin{array}{cc}z_b&0\\0&z\end{array}\right)\\
503: & \\ i\left(\begin{array}{cc}\zbar_b&0\\0&\zbar\end{array}\right) &
504: \kappa-ig_i\tilde{Q}\end{array} \right]\;.
505: \end{align}
506: Performing the integration over the supermatrix $Q$ is still a rather involved
507: technical problem. We provide below
508: a few comments related to it.
509: 
510: 
 The Grassmann variables can be integrated out at any stage, and
511: and we find it convenient to carry out that integration at the very 
512: beginning. On the other hand, it turns out to be important that the integration over the boson-boson
513: part of $Q$ (i.e. $q_b$) should be performed before the fermion-fermion part $q_f$. In fact, a quick
514: inspection of the $q_f$ integrals in Eq.\eqref{eq:int-after-cft} shows that they diverge logarithmically.
515: However, those logarithmic divergences are actually a
516: spurious feature of the colour-flavour transformation. The correct
517: way of treating such divergencies when
518:  performing any supersymmetric colour-flavour transformation
519:  is to integrate first over the boson-boson sub-manifold. 
520: Then, combining all terms which are logarithmically divergent, one can show that
521: the divergent parts cancel each other and the result is actually finite.
522: 
523: To perform the integration over the boson-boson part of $Q$, we introduce polar coordinates
524: $q_b=\sqrt{r}e^{\mathrm{i}\theta}$, where $r\in[0,1], \theta\in[0,2\pi]$ so that
525: $dq_bd\bar{q}_b=drd\theta$. In Eq.\eqref{eq:int-after-cft}, $G(\kappa)$ is defined for
526: $\mathrm{Re}\ \kappa\!\!>\!\!1$. For simplicity, we focus on the real 
527: $\kappa>1$. And further more, we assume all $g_i<1$, for $i=1,\dots,N$. 
528: This is done for convenience only and does not reduce generality 
529: as we can always scale the $ G$-matrix by the magnitude of the largest eigenvalue,
530:  and at the end of the calculation to scale it
531: back. Under these assumptions we can integrate 
532: over the angular variable $\theta$ by residue theorem. In this way 
533: the result of the original integration
534: naturally splits into a sum of contributions from different
535: residues. 
536: 
537: It is crucial that after integrating over $\theta$, we can show $G(\kappa)$ is analytic
538: in the half plane $\mathrm{Re}\ \kappa\!\!>\!\!0$. Therefore we are allowed to
539: make the required
540: analytic continuation
541: on $G(\kappa)$ to $\mathrm{Re}\ \kappa\!\!>\!\!0$. Since both $F(\kappa)$ and $G(\kappa)$
542: are now analytic functions on $\mathrm{Re}\ \kappa\!\!>\!\!0$ and $F(\kappa)=G(\kappa)$ on
543: $\kappa\!\!>\!\!1$, we conclude that $F(\kappa)=G(\kappa)$. We emphasis that this
544: continuation is possible to
545: carry out only after angular integration is performed
546: under the assumption $\kappa\!>\!1$.
547: 
548: 
 Knowing that we should take the limit $\kappa\to0$ in the end of the calculation, we can use the condition 
549: $\kappa<<1$ to simplify significantly the integration over the radial
550: part of $q_b$. It turns out that one has to distinguish
551: two essentially different cases: $g_i>|z|^2$ and
552: $g_i<|z|^2$. Each of these two cases yields different result when integrating
553: over $q_b$ which
554: explains why we have to
555: distinguish $F_+(g_i)$ from $F_-(g_i)$ in the final expression.
556: Integration over the fermion-fermion part of $Q$ uses certain properties of
557: elementary symmetric functions but is otherwise straightforward. Finally, taking
558: derivatives with
559: respect to
560: $z_b$ and $z$ and letting
561: $\kappa\to 0$, we arrive after straightforward but still cumbersome calculations to
562:  the formula Eq.\eqref{eq:sum}.\\
563: 
564: \section{Open problems}
565: In conclusion, we would like to mention a few open  problems and possible extensions along 
566: the lines of the present work. An interesting problem would be to investigate 
567: the density of complex eigenvalues in the limit 
568: $N\to \infty$ assuming that the matrix $g$ has a finite limiting density $\nu(g)=\frac{1}{N}\sum_i\delta(g-g_i)$
569:  of eigenvalues $g_i$ in an interval of the  $g-$axis. A special variant of the problem is to assume that 
570: $g_i$ are eigenvalues of some random Hermitian matrix with rotationally invariant measure. This case is in fact
571: equivalent to the so-called Feinberg-Zee problem, see \cite{FZ}, which attracted a considerable interest recently.
572: We hope to address it in our future publications. 
573: 
574: Another important extension would be to replace matrices $U$ by unitary symmetric random matrices, or 
575: to take them from some other groups (e.g. orthogonal). The corresponding colour-flavour transformations are known,
576: but the calculations seem to be extremely challenging technically.  
577: 
578: 
579: 
580: 
581: 
582: 
583: \section*{Acknowledgements}
584: This research was completed during the 2008 programme "Anderson
585: Localisation: 50 years after" at the Isaac Newton Institute for
586: Mathematical Sciences where both authors were supported by
587: visiting Fellowships. We gratefully acknowledge Prof. B.
588: Khoruzhenko for many clarifying discussions.
589: We also are grateful to S. Nonnenmacher and J. Keating for their stimulating 
590: interest in this work. 
591: The research in Nottingham was performed in the framework of EPSRC
592: grant EP/C515056/1"Random Matrices and Polynomials: a tool to
593: understand complexity".
594: 
595: 
596: 
597: 
598: 
599: \begin{thebibliography}{99}
600: \bibitem{sing} Horn A,  On the eigenvalues of a matrix with prescribed singular values, 
601:   {\it Proc. Am. Math. Soc} {\bf 5} (1954) 4-7; \\
602: Weyl H,  Inequalities between the two kinds of eigenvalues of a linear transformation, 
603:   {\it Proc.Nat.Acad.Sci. USA} {\bf 35} (1949)  408-411.
604: 
605: \bibitem{trunc} {\. Z}yczkowski K and  Sommers H-J,
606: Truncations of random unitary matrices, {\it J. Phys. A: Math.
607: Gen.} {\bf 33} (2000) 2045-2057.
608: 
609: \bibitem{prel} Fyodorov YV, 2001 in {\it Disordered and
610: Complex Systems} edited by P.Sollich et al. {\it AIP Conference
611: Proceedings } {\bf 553} 191 (Melville NY) [arXiv:nlin.CD/0002034].
612: 
613: \bibitem{FSJETP} Fyodorov YV and  Sommers H-J, 
614:   Spectra of random contractions and scattering theory for discrete-time systems,  {\it JETP Letters } {\bf 72} (2000)
615: 422-427.
616: 
617: \bibitem{rev03} Fyodorov YV  and  Sommers H-J, Random matrices close to Hermitian or unitary: 
618: overview of methods and results, {\it J. Phys. A:Math.Gen.} {\bf 36} (2003) 3303-3347.
619: 
620: \bibitem{yb05} Fyodorov YV, Khoruzhenko BA, On absolute moments of characteristic polynomials of a 
621: certain class of complex random matrices., {\it Commun. Math. Phys.} {\bf 273} (2007) 561-599.
622:  \bibitem{Haakebook} Haake F. 1999 {\it Quantum Signatures of Chaos} (Springer, 2nd ed.).
623: 
624: \bibitem{KolRev}  Gl\"{u}ck M, Kolovski AR and Korsch HJ, 
625:  Wannier-Stark resonances in optical and semiconductor superlattices, 
626: {\it Phys.Rep} {\bf 366} (2002) 103-182.
627: 
628: \bibitem{JSB} Jacquod P, Schomerus H, and Beenakker CWJ, 
629:  Quantum andreev map: A paradigm of quantum chaos in superconductivity, {\it Phys.Rev.Lett.} 
630: {\bf 90} (2003) 207004 (4p); \\
631: Tworzydlo J, et al., Dynamical model for the quantum-to-classical crossover of shot noise,  
632: {\it Phys.Rev. B} {\bf 68} (2003)  115313 (6pp);\\ 
633: Schomerus H and Jacquod P,  Quantum-to-classical correspondence in open chaotic systems, 
634: {\it J. Phys. A: Math.Gen.} {\bf 38} (2005) 10663-10682.
635: \bibitem{OKG} Ossipov A, Kottos T, and Geisel T,
636:  Fingerprints of classical diffusion in open 2D mesoscopic systems in the metallic regime, 
637:  {\it Europh. Lett.} {\bf 62} (2003) 719-725.
638: 
639: \bibitem{Prange} Prange RE , Resurgence in Quasiclassical Scattering, {\it Phys.Rev.Lett.} (2003)
640: {\bf 90} 070401 (4pp).
641: \bibitem{KNS} Keating JP, Novaes M, and Schomerus H,  Model for chaotic dielectric microresonators, 
642: {\it Phys.Rev. A} {\bf 77} (2008)  013834 (9pp).
643: 
644: \bibitem{Nonnem}
645: Nonnenmacher S, Zworski M, Fractal Weyl laws in discrete models of chaotic scattering, 
646: {\it  J. Phys. A: Math.Gen.} {\bf  38} 10683-10702 ;
647: 
648: Nonnenmacher S, Rubin M, Resonant eigenstates for a quantized chaotic system, 
649: {\it Nonlinearity} {\bf  20} (2007) 1387-1420 ; 
650: 
651: Nonnenmacher S, Schenck E , Resonance distribution in open quantum chaotic systems, 
652:  Preprint : arXiv:0803.1075.
653: 
654: 
655: \bibitem{FKS} Fyodorov YV, Khoruzhenko BA and Sommers H-J, 
656: Almost-Hermitian random matrices: eigenvalue density in the complex plane,  
657: {\it Physics Letters} {\bf 226} (1997)  46-52. 
658: 
659: \bibitem{rev96} Fyodorov YV and Sommers H-J, Statistics of resonance poles,
660:  phase shifts and time delays in quantum chaotic scattering: 
661: random matrix approach for systems with broken time-reversal invariance {\it J. Math.
662: Phys.} {\bf 38} (1997) 1918-1981.
663: 
664: \bibitem{cft} Zirnbauer MR, Supersymmetry for systems with unitary disorder: circular ensembles, 
665:   {\it J. Phys. A: Math.Gen.} {\bf 29} (1996) 7113-7136.
666: 
667: \bibitem{sym-space} Zirnbauer MR,  Riemannian symmetric superspaces and their origin in random-matrix theory, 
668:  {\it J. Math. Phys.} {\bf 37} (1996) 4986-5018.
669: 
670: 
671: \bibitem{FZ} Feinberg J and Zee A,
672:  Non-Gaussian Non-Hermitean Random Matrix Theory:
673:  phase transitions and addition formalism, {\it  Nucl. Phys. B} {\bf  501} (1997) 643-669;
674: 
675: Feinberg J, Scalettar R. and Zee A,  
676: Single Ring Theorem and the Disk-Annulus Phase Transition,
677: {\it  J Math Phys} {\bf 42} (2001) 5718-5740.
678: 
679: \end{thebibliography}
680: 
681: \end{document}
682: