nlin0105006/sm1.tex
1: \documentstyle[aps,psfig]{revtex} 
2: \tightenlines 
3: \newlength{\figwidth} 
4: \setcounter{equation}{0} 
5: \setlength{\figwidth}{3.4in} 
6:  
7:  
8: \title{Exact and semiclassical Husimi distributions of Quantum Map 
9: Eigenstates} 
10: \author{M. Saraceno$^1$ and A. G. Monastra$^2$} 
11: \address{$^1$Departamento de F\'{\i}sica, Comisi\'on Nacional de Energ\'{\i}a
12: At\'omica (CNEA), \\ Av. del Libertador 8250, 1429 Buenos Aires, Argentina
13: \\ $^2$Laboratoire de Physique Th\'eorique et Mod\`eles Statistiques,
14: B\^at. 100, \\ 91405 Orsay Cedex, France} 
15:  
16: \begin{document} 
17: \maketitle {\begin{abstract} The projector onto single quantum map eigenstates
18: is written only in terms of powers of the evolution operator, up to half the
19: Heisenberg time, and its traces. These powers are semiclassically
20: approximated, by a complex generating function, giving the Husimi distribution
21: of the eigenstates. The results are tested on the Cat and Baker maps.  
22: \end{abstract}} 
23:  
24:  
25: \section{Introduction} 
26: The relation between the classical invariant structures of a dynamical 
27: system and its 
28: quantum counterparts, the eigenenergies and eigenstates, is not 
29: completely understood and 
30: is an intense subject of investigation. When the system is integrable 
31: the eigenenergies 
32: are determined by the well-known Bohr-Sommerfeld or EBK quantization 
33: rules and the 
34: eigenfunctions have a very intuitive phase space structure: they are 
35: essentially 
36: localized on the quantized tori and they present characteristic quantum 
37: features 
38: --interference fringes for the Wigner function \cite{berry1} or zeros 
39: for the Husimi 
40: function \cite{voros}-- away from them. No such clear picture is 
41: available for the 
42: eigenstates of chaotic systems and much work
43: \cite{heller,bogo1,vergini,nonnenmacher} has gone 
44: into trying to disentangle the purely statistical aspects of the 
45: eigenfunctions from the 
46: specific dynamical features such as scars and localization. 
47:  
48: Although a lot of recent work deals with the study of the statistical 
49: aspects of chaotic 
50: wavefunctions \cite{mirlin} we concentrate here on the 
51: purely dynamical 
52: aspects and explore how well the semiclassical formulae can reproduce 
53: the details of 
54: individual eigenfunctions. 
55:  
56: With regards to spectral properties the theoretical tool for this 
57: analysis is the 
58: Gutzwiller trace formula \cite{gutzw} relating the spectral density to 
59: the periodic 
60: orbits. As is well known, this relationship is fraught with 
61: uncertainties both due to 
62: the approximations involved, as to the fact that it diverges where it is 
63: supposed to 
64: locate the single eigenenergies and to the sheer complexity of adding 
65: the contributions 
66: of an exponentially increasing number of periodic orbits. Some of these 
67: difficulties can 
68: be overcome by resummation techniques \cite{bogo-keat} that limit the 
69: exponential 
70: increase by partially restoring the unitarity of the propagator, lost in 
71: the 
72: semiclassical approximation. 
73:  
74: In the present work we extend these resummation methods to the 
75: calculation of 
76: eigenfunctions of chaotic maps. We utilize Fredholm theory to construct 
77: a projector on 
78: single eigenstates whose expression is directly amenable to 
79: semiclassical 
80: approximations. These approximations share the same advantages - and the
81: drawbacks- of the resummation techniques available for spectral properties.
82:  
83: In section 2 we construct explicitly the projector on single 
84: eigenstates. This is 
85: achieved exactly in terms of a finite sum on powers of the quantum map 
86: propagator and 
87: its traces. The cut-off time is half the Heisenberg time due to the 
88: explicit imposition 
89: of unitarity. In section 3 these formulae are implemented 
90: semiclassically by a) 
91: approximating the traces in terms of actions and instabilities of 
92: periodic orbits in the 
93: usual way, and b) by computing the propagator in the coherent state 
94: representation in 
95: terms of specially devised complex generating functions. 
96:  
97: The final result is a semiclassical expression for the Husimi function 
98: of an eigenstate 
99: in terms of properties of a finite set of periodic orbits. 
100:  
101: In section 4 we apply the method to the cat map --where the 
102: semiclassical approximation 
103: is exact -- and to the baker map where we can test the approximation. 
104:  
105: The formulae are here applied as a test to simple maps, but they can be 
106: easily adapted 
107: to more realistic Hamiltonian systems by selecting an appropriate 
108: Poincar\'e section 
109: \cite{sara-simo}. 
110:  
111:  
112:  
113: \section {Projector on single eigenstates} 
114:  
115: For a map, the density of Floquet states can be expressed in the usual 
116: way in terms of 
117: traces by the formula 
118:  
119: \begin{equation} 
120: d(\epsilon) = \frac{2\pi}{N}+\sum_{n=1}^{\infty} [tr \hat U^n  e^{- i 
121: \epsilon n} ] + 
122: C.C. 
123: \end{equation} 
124: $1/N$ plays the role of Plank's constant $h$, $n$ is the discrete time, 
125: and $\hat U$ is 
126: the unitary quantization of the map. The density of states results in a 
127: $2\pi$ periodic 
128: function with unit $\delta$-spikes at the eigenangles $\epsilon_k$, 
129: defined by  
130:  
131: \begin{equation} 
132: \hat U | \psi_k \rangle = e^{i \epsilon_k} | \psi_k \rangle \ . 
133: \end{equation} 
134:  
135: Semiclassical approximations using this formula require the calculation 
136: of traces of 
137: powers of the map for large values of $n$ which are in turn related to 
138: classical orbits 
139: of the map by the Gutzwiller-Tabor \cite{gutzw} trace formula 
140:  
141: \begin{equation} 
142: b_n = tr \hat U^n \approx \sum_{\gamma} A_{\gamma} \exp \left[ 
143: i(S_{\gamma}/\hbar - 
144: \mu_{\gamma} \pi /2) \right] \label{gutz} \ , 
145: \end{equation} 
146: where $\gamma$ label the $n$-periodic orbits of the map, $S_{\gamma}$ is 
147: the action, 
148: $\mu_{\gamma}$ is the Maslov index, and $A_{\gamma}$ is a coefficient 
149: related to the 
150: monodromy matrix. The approximation is valid for fixed $n$ in the limit 
151: as $N 
152: \rightarrow \infty$ (but not vice-versa). 
153:  
154: A more efficient scheme starts from the spectral determinant 
155:  
156: \begin{equation} 
157: P(s) = \det (\hat I - s \hat U) 
158: \end{equation} 
159: which vanishes at $s = s_k=e^{ - i \epsilon_k}$ and which can be 
160: expanded as 
161:  
162: \begin{equation} 
163: P(s) = \sum_{n=0}^N \beta_n s^n \ . 
164: \end{equation} 
165: The coefficients $\beta_n$ can be computed recursively from the traces 
166: as 
167:  
168: \begin{equation} 
169: \beta_n = -\frac{1}{n} \sum_{j=1}^n \beta_{n-j} b_j  \label{betarec} 
170: \end{equation} 
171: and therefore when computed semiclassically as in (3) they are expressed in 
172: terms of linear 
173: combinations of periodic orbits (pseudo orbits \cite{pseudo}) of period
174: up to $n$. 
175:  
176: The advantage over the calculation in term of traces is that only $N$ 
177: coefficients need 
178: to be calculated. Moreover as a consequence of unitarity the symmetry 
179:  
180: \begin{equation} 
181: \beta_{N-j} = (-1)^N \det \hat U ~ \bar{\beta}_j \label{rever} 
182: \end{equation} 
183: can be imposed, so that only $N/2$ coefficients are needed and therefore 
184: only periodic 
185: trajectories for times up to $N/2$ (i.e. half the Heisenberg time) are 
186: involved for a 
187: semiclassical calculation. This is now an optimal encoding of the 
188: eigenvalues in terms 
189: of traces, with the $N$ real phases $\epsilon_k$ expressed exactly in 
190: terms of the $N/2$ 
191: complex coefficients $\beta_n$ 
192:  
193: It is convenient to define another function \cite{bogo2}, proportional 
194: to the spectral 
195: determinant, which has the same roots. Utilizing the symmetry 
196: (\ref{rever}), for an 
197: even $N$, this function can be written 
198:  
199: \begin{equation} 
200: Z(s) = \frac{P(s)}{\beta_{N/2} ~ s^{N/2}} = 1 + \frac {1}{\beta_{N/2}} 
201: \sum_{n=0}^{N/2-1} \beta_n ~ s^{n-N/2} + \frac{1}{\bar{\beta}_{N/2}} 
202: \sum_{n=0}^{N/2-1} 
203: \bar{\beta}_n ~ s^{N/2-n} \ . 
204: \end{equation} 
205:  
206: When $s$ is on the unit circle $1/s=\bar s$, and then the second sum is 
207: the complex 
208: conjugate of the first. Then the function $Z(s)$ becomes real on 
209: $|s|=1$. 
210:  
211:  
212: Unitarity also implies that the spectral determinant has all its zeros 
213: on the unit 
214: circle. However when the coefficients are computed semiclassically this 
215: is no longer 
216: guaranteed. If the symmetry (\ref{rever}) is imposed the resulting 
217: spectral determinant 
218: becomes self-inversive, which is a necessary -but by no means 
219: sufficient- condition for 
220: unitarity \cite{bbl}. Self-inversive polynomials have the following properties 
221: \cite{marsden}: 
222:  
223: 1) All zeros are either on the unit circle or symmetric with respect to 
224: it. Thus as a 
225: function of a continuous parameter the zeros of a self-inversive 
226: polynomial can only 
227: leave the unit circle in pairs by degenerating on it. 
228:  
229: 2) If $P(s)$ has only single roots then $dP/ds$ has no zeros on the unit 
230: circle. 
231:  
232:  
233:  
234:  
235:  
236:  
237:  
238: \subsection {Green Operator} 
239:  
240:  
241: Very similar techniques can be implemented for the calculation of matrix 
242: elements and 
243: therefore of eigenfunctions. To obtain the eigenfunctions of the unitary 
244: matrix $\hat 
245: U$, we define a Green operator  
246:  
247: \begin{equation} 
248: \hat G (s) = \frac {f(\hat U)}{\hat I - s \hat U} \ , 
249: \end{equation} 
250: which, as an analytic function of $s$, has poles at $s = s_k = e^{ - i 
251: \epsilon_k}$, and 
252: the residues are proportional to the projectors $\hat P_k = |\psi_k 
253: \rangle \langle 
254: \psi_k |$. To obtain normalized projectors, we instead define a {\it 
255: normalized} Green 
256: operator 
257:  
258: \begin{equation} 
259: \hat g (s) = \frac {\hat G (s)}{tr[\hat G (s)]} \ . 
260: \end{equation} 
261:  
262: At the poles, the singularities cancel and we obtain $\hat g (s_k) = 
263: \hat P_k$. $f(\hat 
264: U)$ is left arbitrary for the moment, and specific choices -made below- 
265: will be used to 
266: obtain formulae with different properties. 
267:  
268: We can expand the preceding expression utilizing the transpose cofactor 
269: matrix to write 
270: the Green operator 
271:  
272: \begin{equation} 
273: \hat g (s) = \frac {f(\hat U) ~ C^t (\hat I - s \hat U)}{ tr[f(\hat U) ~ 
274: C^t (\hat I - s 
275: \hat U)]} \label{gs} \ . 
276: \end{equation} 
277: The transpose cofactor matrix can be written as a finite power series in 
278: $s$ 
279:  
280: \begin{eqnarray}  
281: C^t (\hat I - s \hat U) &=& \sum_{n=0}^{N-1} s^n \hat X _n \ , \\ 
282: \hat X _n &=& \sum_{i=0}^n \beta_i ~ \hat U ^{n-i} \label{xn} \ . 
283: \end{eqnarray} 
284:  
285: By the unitarity of $\hat U$, the $\hat X _n$ matrices satisfy the 
286: following symmetry 
287:  
288: \begin{equation} 
289: \hat X _{N-j} = - \hat U^{\dagger} (-1)^N \det \hat U ~ \hat X 
290: _{j-1}^{\dagger} \ . 
291: \label{reverx}
292: \end{equation} 
293:  
294:  
295: Utilizing this symmetry two choices of $f(\hat U)$ lead to expansions 
296: with useful 
297: properties: 
298:  
299: \noindent a) For $f(\hat U)= \hat U$, for even $N$, the corresponding 
300: Green function is 
301:  
302: \begin{equation} 
303: \hat g_a (s) = \frac{ \sum_{n=0}^{N/2-1} s^n ~ \hat U \hat X_n - \det 
304: \hat U ~ s^{N-1} 
305: \sum_{n=0}^{N/2-1} s^{-n} ~ \hat X^{\dagger}_n } { \sum_{n=0}^{N/2-1} 
306: s^n ~ tr 
307: \left(\hat U \hat X_n \right) - \det \hat U ~ s^{N-1} \sum_{n=0}^{N/2-1} 
308: s^{-n} ~ tr 
309: \left(\hat X^{\dagger}_n \right)}  \ . 
310: \end{equation} 
311:  
312: Replacing the expansion of the $\hat X_n$ operators in powers of $\hat 
313: U$, we obtain 
314:  
315: \begin{equation} 
316: \hat g_a(s) = \frac{\sum_{i=0}^{N/2-1} c_i (s) ~ \hat U ^{i+1} - \det 
317: \hat U ~ s^{N-1} 
318: \sum_{i=0}^{N/2-1} \bar{c_i} (1/\bar s) ~ \hat U^{\dagger ~ i} }  
319: {\sum_{i=0}^{N/2-1} 
320: c_i (s) ~ tr \left(\hat U ^{i+1} \right) - \det \hat U ~ s^{N-1} 
321: \sum_{i=0}^{N/2-1} 
322: \bar{c_i} (1/\bar s) ~ tr \left(\hat U^{\dagger ~ i} \right) }  
323: \label{greenfinal} \ , 
324: \end{equation} 
325: in which the projector is expressed as a finite combination of the first 
326: $N/2$ powers of 
327: the propagator. 
328:  
329: The coefficients $c_i$ are obtained from the $\beta_n$ 
330:  
331: \begin{equation} 
332: c_i (s) = \sum_{n=i}^{N/2-1} \beta_{n-i} ~ s^n \label{cis} \ . 
333: \end{equation} 
334:  
335: This normalized Green operator has no singularities as $s$ goes around 
336: on the unit 
337: circle. This can be demonstrated because the denominator is equal to  
338: $-dP/ds$ (see 
339: Appendix A). Then by the second property of the self-inversive 
340: polynomials it is 
341: guaranteed that this denominator never takes the zero value for $s$ over 
342: the unit 
343: circle. This property, guaranteed by unitarity, will remain true --even 
344: when the traces 
345: and propagator are calculated semiclassically-- as long as self 
346: inversiveness is 
347: maintained. 
348:  
349:  
350: \noindent b) For $f(\hat U)= \hat I + s \hat U$ , $\hat G(s)$ in (9) is 
351: the   
352: Cayley transform of $\hat U$ and is therefore hermitian on the unit 
353: circle. In 
354: this case 
355: \begin{equation} 
356: \hat g_b (s) = \frac{\sum_{n=0}^N s^n ~ \hat Y_n}{\sum_{n=0}^N s^n ~ tr 
357: \left(\hat Y_n 
358: \right) } 
359: \end{equation} 
360: with 
361: \begin{equation} 
362: \hat Y_n = \hat X_n + \hat U \hat X_{n-1} = 2 \hat X_n - \beta_n \hat I 
363: \label{yn} \ . 
364: \end{equation} 
365: These new operators satisfy a simpler symmetry relation than the 
366: preceding $\hat X_n$, 
367: analogous to the symmetry of the $\beta_n$ coefficients 
368:  
369: \begin{equation} 
370: \hat Y_{N-j} = - (-1)^N \det \hat U ~ \hat Y^{\dagger}_j \ . 
371: \label{revery}
372: \end{equation} 
373: In this case, for even $N$, the normalized Green operator can be 
374: written as 
375:  
376: \begin{equation} 
377: \hat g_b(s) = \frac{\hat{\Sigma} (s) - \hat{\Sigma}^{\dagger} (1/\bar 
378: s)} {tr\left[ 
379: \hat{\Sigma} (s) \right] - tr \left[ \hat{\Sigma}^{\dagger} (1/\bar s) 
380: \right]} 
381: \label{gb} \ , 
382: \end{equation} 
383: where 
384:  
385: \begin{equation} 
386: \hat{\Sigma} (s) = \frac{1}{\beta_{N/2}} \left[ \frac{1}{2} \hat Y_{N/2} 
387: + 
388: \sum_{n=0}^{N/2-1} s^{n-N/2} ~ \hat Y_n  \right] \ . 
389: \end{equation} 
390:  
391: In this expression, more symmetrical than (\ref{greenfinal}), the 
392: numerator is 
393: anti-hermitian on the unit circle, and the denominator is purely 
394: imaginary. The basic 
395: operator $\hat{\Sigma} (s)$ is a Fourier transform of the $\hat Y_n$ 
396: operators and their 
397: traces. When utilizing this finite representation, they then become the 
398: basic objects of 
399: study in the time domain, in place of the propagator and its traces. 
400:  
401: When $\hat U$ is a finite matrix, $\hat g_a(s)$ and $\hat g_b(s)$ are 
402: both rational 
403: analytic functions of $s$. They are identical at the eigenvalues $s=s_k$ 
404: (on the unit 
405: circle). However their zero and singularity structure is quite 
406: different. 
407:  
408: In particular $\hat g_a(s)$, as noted, has a denominator which does not 
409: vanish on the 
410: unit circle, even when the coefficients are computed approximately, as 
411: long as 
412: self-inversiveness is imposed on them. The operator $\hat g_a(s)$ then, 
413: has no 
414: singularities in a small strip enclosing the unit circle and can be 
415: followed continuously 
416: from eigenvalue to eigenvalue. A disadvantage is that $\hat g_a(s)$ is 
417: not hermitian 
418: (except at the values $s=s_k$). On the contrary $\hat g_b(s)$ is 
419: hermitian on the unit 
420: circle, but its singularities (the zeroes of the denominator) are 
421: precisely on the unit 
422: circle, in between two successive eigenvalues. 
423:  
424: It should be noticed that, although for simplicity all the above formulae
425: are written for even $N$, they are easily adapted to the odd case, with
426: minimal changes.  
427: 
428:  The preceding operator expressions for $\hat g_a(s)$ and $\hat g_b(s)$ 
429: provide a common structure that underlies all further semiclassical
430: approximations. Depending on the representation used to evaluate them, 
431: they can yield results for transition probabilities, phase space 
432: distributions, response functions or wave function correlations. This is 
433: achieved by linking the spectral properties of the map to a {\it finite 
434: number } of traces and  matrix elements of the propagator. To apply them
435: one needs the semiclassical evaluation of the propagator traces as in
436: (\ref{gutz})- which are representation independent- and also of the 
437: semiclassical propagator in the chosen representation.
438: For example if they are computed in the coordinate representation they 
439: provide  expressions for $|\langle x| \psi_k \rangle |^2$. If the propagator 
440: is computed in the Weyl representation \cite{ozorio} they yield the Wigner 
441: distribution of eigenfunctions \cite{fishman}. 
442: In all cases the unitarity of the propagator - which is generally only 
443: approximate in semiclassical treatments - is partially restored by the 
444: imposition of the symmetries (\ref{rever},\ref{reverx},\ref{revery}).
445: In the next section the above formalism is developed in the coherent state 
446: representation to obtain expressions for semiclassical Husimi distributions. 
447:  
448:  
449: \section{Semiclassical approximation in the coherent state representation} 
450:  
451: The formulae derived in the previous section are exact and for maps can, 
452: at first sight, 
453: seem just a complicated way of looking at a simple algebraic eigenvalue 
454: problem. 
455: However, they are prepared in such a way as to make the semiclassical 
456: transition as 
457: clear as possible and to unveil the necessary approximations involved in 
458: such a 
459: transition,  
460:  
461: The coherent state mean value of $\hat g(s)$ is 
462:  
463: \begin{equation} 
464: {\cal H} (s, z, \bar z) \equiv \frac{\langle z | \hat g(s) | z 
465: \rangle}{\langle z | z \rangle} \ . 
466: \end{equation} 
467: This is a phase space distribution, analytic in $s$, which becomes the 
468: Husimi 
469: distribution of the eigenstates $|\psi_k \rangle$ at the values $s_k = 
470: e^{- i 
471: \epsilon_k}$.  
472:  
473: Its properties, reflecting the zeroes and singularities of $\hat g(s)$ 
474: can be followed 
475: continuously in the $s$ complex plane. Its calculation involves the $N/2$ 
476: coherent state return amplitudes 
477:  
478: \begin{equation} 
479: C_n (z, \bar z) \equiv \frac{\langle z | \hat U^n | z \rangle}{\langle z 
480: | z \rangle} 
481: \end{equation} 
482: in addition of the $N/2$ traces 
483:  
484: \begin{equation} 
485: b_n=tr \hat U^n = \int dz ~ d\bar z ~ C_n (z,\bar z) \ . 
486: \end{equation}  
487: 
488: The first admits immediately a well known semiclassical approximation in 
489: the time 
490: domain, namely the Gutzwiller-Tabor trace formula (\ref{gutz}). The 
491: stationary phase 
492: approximations involved in the derivation are of course valid for fixed 
493: time $n$ and for 
494: $\hbar \rightarrow 0$, or equivalently, $N \rightarrow \infty$, and 
495: their use in the 
496: regime $n \approx N/2$ needed for the exact calculation is questionable. 
497:  
498: The semiclassical calculation of the return amplitudes $C_n (z, \bar z)$ can 
499: be done by stationary phase calculation of the integral 
500:  
501: \begin{equation} 
502: \langle z_f| \hat U^n |z_i \rangle = \int \int \langle z_f | q_f \rangle 
503: \langle q_f | 
504: \hat U^n | q_i \rangle \langle q_i | z_i \rangle ~ dq_i ~ dq_f 
505: \label{intzq} \ , 
506: \end{equation} 
507: where 
508: \begin{equation} 
509: \langle z| q \rangle = \frac{\exp \left\{ -\frac{1}{2 \hbar} \left[ z^2 
510: + q^2 - 2 \sqrt 
511: 2 z q \right] \right\} }{\sqrt [4]{\pi \hbar}} 
512: \end{equation} 
513: is the coherent state in coordinate representation, and the 
514: propagator has the usual Van Vleck form 
515: \begin{equation} 
516: \langle q_f| \hat U^n |q_i \rangle = \frac{1}{\sqrt{2\pi i \hbar}} 
517: \sum_{\gamma} 
518: \sqrt{\frac{\partial ^2 S^{(n)}_{\gamma} (q_i,q_f) }{\partial q_i 
519: \partial q_f}} ~ \exp \left[ 
520: ~ \frac{i}{\hbar} S^{(n)}_{\gamma} (q_i,q_f) \right] \label{vleck} \ . 
521: \end{equation} 
522: The sum must be made over all the paths in the phase space of the 
523: corresponding 
524: classical map that connect $q_i$ with $q_f$ in $n$ steps. Additional phases 
525: connecting the different branches, the Maslov indices, must be added to 
526: $S^{(n)}_{\gamma}$. 
527:  
528: The calculation of the integral (\ref{intzq}) with the Van Vleck 
529: approximation 
530: (\ref{vleck}) can be done as usual by stationary phase and the result 
531: has the structure 
532: of another Van Vleck formula with complex coordinates for a generating 
533: function. 
534:  
535: \begin{equation} 
536: \langle z_f | \hat U^n |z_i \rangle = \sum  \sqrt{ -i \frac{\partial ^2 
537: G_3}{\partial 
538: \bar z_i \partial z_f} } ~ \exp \left[ -\frac{i}{\hbar} G_3(\bar 
539: z_i,z_f) \right] 
540: \label{zuzsem} \ . 
541: \end{equation} 
542:  
543: This new classical generating function is a complex Legendre transform 
544: of the action $S$ 
545:  
546: \begin{equation} 
547: G_3 = -S + \frac{1}{2} (p_f q_f - p_i q_i) + \frac{i}{4} (q_f^2 + p_f^2 
548: + q_i^2 + p_i^2) 
549: \label{gzz} \ , 
550: \end{equation} 
551: or equivalently, in the complex coordinates 
552:  
553: \begin{eqnarray} 
554: z &=& \frac{1}{\sqrt{2}} \left(q - i p \right)  \\ 
555: \bar z &=& \frac{1}{\sqrt{2}} \left(q + i p \right) \ ,   
556: \end{eqnarray} 
557: it can be written as 
558:  
559: \begin{eqnarray} 
560: G_3 =  -S + \frac{i}{4} \left(z_f^2 + 2 z_f \bar z_f - \bar z_f^2 
561: \right) - \frac{i}{4} 
562: \left( z_i^2 - 2 z_i \bar z_i - \bar z_i^2 \right) \label{g1z} \ . 
563: \end{eqnarray} 
564:  
565: This function depends explicitly on $\bar z_i,z_f$ variables, and yields 
566: the equations 
567: of motion 
568:  
569: \begin{equation} 
570: z_i = -i \frac{\partial G_3}{\partial \bar z_i} \ , ~~~~~~~~~  \bar z_f = -i 
571: \frac{\partial 
572: G_3}{\partial z_f} \label{motion} \ . 
573: \end{equation} 
574:  
575: These are algebraic equations that define the map implicitly provided 
576: $\frac{\partial^2 
577: G_3}{\partial \bar z_i \partial z_f} \neq 0$ where the sum must be made 
578: over complex 
579: paths that connect the point $\bar z_i$ to $z_f$. These formulae show 
580: that the return amplitude $C_n$ has as exponent the function 
581: $G_3(\bar z,z) -i \bar z z$ which using 
582: (\ref{motion}) can be shown to be stationary at the periodic points. 
583: Quadratic approximation of the generating function at these points yields 
584: the coherent state return amplitudes as a sum of Gaussian peaks centered 
585: on them 
586:  
587: \begin{equation} 
588: C_n(z,\bar z) = \sum_{\gamma} \frac{1}{\sqrt{u_{\gamma}}} \exp \left[ - 
589: \frac{1}{\hbar} {\cal G}^{\gamma} (z,\bar z) + \frac{i}{\hbar} S_{\gamma} 
590: \right] \label{cn} \ . 
591: \end{equation} 
592:  
593: \begin{equation} 
594: {\cal G}^{\gamma} (z,\bar z) = \frac{1}{u_{\gamma}} \left[ \frac{\bar 
595: v_{\gamma}}{2} 
596: (\bar z - \bar z_{\gamma})  ^2 + (u_{\gamma}-1) (\bar z - \bar 
597: z_{\gamma}) (z - 
598: z_{\gamma}) - \frac{v_{\gamma}}{2} (z - z_{\gamma})^2 \right] 
599: \label{gexp} \ ,  
600: \end{equation} 
601: where $z_{\gamma},\bar z_{\gamma}$ are the position of the periodic 
602: points, $S_{\gamma}$ 
603: are the corresponding actions, and $u_{\gamma},v_{\gamma}$ and $\bar 
604: v_{\gamma}$ are the 
605: elements of the complex monodromy matrix defined by 
606:  
607: \begin{equation} 
608: \left(  
609: \begin{array}{cc} 
610: \partial z_f / \partial z_i & \partial z _f / \partial \bar z_i  \\ 
611: \partial \bar z_f / \partial z_i & \partial \bar z_f / \partial \bar z_i  
612: \end{array} 
613: \right) = \left(  
614: \begin{array}{cc} 
615: u & \bar v   \\ 
616: v & \bar u  
617: \end{array} 
618: \right) 
619: \end{equation} 
620: calculated on the corresponding periodic point. 
621:  
622: Using (\ref{cn}) for the return amplitudes and (\ref{gutz}) for the 
623: traces, the 
624: Husimi distributions of eigenstates are explicitly written in terms of 
625: properties of 
626: periodic points of all periods up to half the Heisenberg time $N/2$. 
627:  
628: Although the semiclassical approximations are available in the time 
629: domain for the 
630: traces $b_n$ and the return amplitudes $C_n(z,\bar z)$ it is clear from 
631: our formulae 
632: that the essential ingredients for the calculations are the coefficients 
633: $\beta_n$ and 
634: the operators $\hat X_n$ (or $\hat Y_n$). These are related by 
635: (\ref{betarec}) and 
636: (\ref{xn}) to $b_n$ and $C_n$ and therefore can be recursively computed 
637: from them. This 
638: procedure yields the $\beta_n$ as linear combinations of {\it products} 
639: --the so called 
640: pseudo-orbits-- of terms involving many periodic orbits of different 
641: periods and with 
642: signs implying many possible cancellations between long and short 
643: orbits. The same 
644: situation prevails for the $\hat X_n$ (or $\hat Y_n$). The expressions 
645: yield linear 
646: combinations of return amplitudes at many different times. In fact $\hat 
647: X_n(z,\bar z)$ 
648: give the phase space distribution of pseudo-orbits and it is easy to 
649: show that 
650: $\beta_n=(N-n) ~ tr \hat X_n$. In view of the cancellations involved it 
651: would be very 
652: desirable to have a direct semiclassical derivation --both for $\beta_n$ 
653: and $\hat 
654: X_n$-- that would not go through the indirect procedure of approximating 
655: $b_n$ and $\hat 
656: U_n$, and then recursively computing $\beta_n$ and $\hat X_n$. However 
657: to our knowledge 
658: such a derivation is not available. 
659:  
660:  
661:  
662:  
663:  
664:  
665: \section{Numerical calculations} 
666:  
667:  
668: We will test the methods proposed on two simple well known maps: the cat 
669: map, and the 
670: baker map. 
671:  
672: For the cat map 
673:  
674: \begin{eqnarray} 
675: \left(  
676: \begin{array}{c} 
677: q_f  \\ 
678: p_f  
679: \end{array} 
680: \right) = \left(  
681: \begin{array}{cc} 
682: a & b   \\ 
683: c & d 
684: \end{array} 
685: \right)  \left(  
686: \begin{array}{c} 
687: q_i  \\ 
688: p_i 
689: \end{array} 
690: \right)  (mod ~ 1) 
691: \end{eqnarray} 
692: with $a,b,c$ and $d$ integer numbers, and $ad-bc=1$, the Van Vleck 
693: formula in the 
694: $q_i,q_f$ representation is exact \cite{hannay} and the corresponding 
695: one in the 
696: coherent state representation can be computed explicitly. 
697:  
698: The transformation to complex coordinates leads to the map 
699:  
700: \begin{eqnarray} 
701: \left(  
702: \begin{array}{c} 
703: z_f  \\ 
704: \bar z_f  
705: \end{array} 
706: \right) = \left(  
707: \begin{array}{cc} 
708: u & \bar v   \\ 
709: v & \bar u 
710: \end{array} 
711: \right)  \left(  
712: \begin{array}{c} 
713: z_i  \\ 
714: \bar z_i 
715: \end{array} 
716: \right)  
717: \end{eqnarray} 
718: with 
719:  
720: \begin{eqnarray} 
721: u &=& \frac{1}{2} \left [ (a+d) + i(b-c) \right ] \ , \\ 
722: v &=& \frac{1}{2} \left [ (a-d) + i(b+c) \right ] \ . 
723: \end{eqnarray} 
724: and its quantization with (\ref{vleck}) is obtained exactly by Gaussian 
725: integration. The coherent state return amplitudes are computed with the 
726: formulae (\ref{cn}) and (\ref{gexp}), with the 
727: advantage that the coefficients of the monodromy matrix are the same for 
728: all the 
729: periodic points of a given period. Since the cat map is defined on the 
730: torus we use 
731: periodized coherent states \cite{voros}. With this result we can test 
732: our formulae 
733: keeping in mind that for this special case the semiclassical 
734: approximation is exact. 
735:  
736: In Fig.1 we first look at quantities in the time domain, i.e. as a 
737: function of $n$. The 
738: correlations $|C_n(z, \bar z)|^2$ display for short times Gaussian peaks 
739: at the 
740: n-periodic points. 
741:  
742: However we have seen that the irreducible information about eigenstates 
743: is carried by 
744: more complicated operators such as the $\hat X_n$ in (\ref{xn}) or the 
745: $\hat Y_n$ in 
746: (\ref{yn}). These operators carry the phase space information that 
747: corresponds to the 
748: pseudo-orbits in the resummation of trace formula. In fact many 
749: different powers of 
750: $\hat U$ contribute to them and give rise to complex interference 
751: structures in phase 
752: space. The pseudo-orbit contributions --contained in the coefficient 
753: $\beta_n$-- are 
754: obtained from the fact that $\beta_n = (N-n) ~ tr \hat X_n$. 
755:  
756: In Fig.2 we have displayed the first few operators $\hat U \hat X_n$ 
757: appearing in the 
758: expansion of $g_a(s)$. For very short times we notice the simple 
759: superposition of the 
760: patterns of $C_n(z,\bar z)$. In particular $\hat U \hat X_2$ shows peaks 
761: at the period 
762: 1, period 2 and period 3 periodic orbits. For longer times ($\hat U \hat 
763: X_3$) the 
764: interferences between orbits of different periods are very strong and 
765: one can barely 
766: notice the period 4 orbits, while the dominant structure is of period 1. 
767: This is in 
768: striking contrast with the regular appearance of periodic points in the 
769: return amplitudes. Notice also the totally unexpected --and for us 
770: unexplained-- appearance of a strong  
771: periodicity (of period three) in the phase space patterns, which is 
772: totally invisible in the $C_n$ 
773: coefficients. We cannot give an explanation to this behavior except 
774: noting that a direct 
775: semiclassical calculation of these coefficients would be very desirable 
776: and might be 
777: more effective than one which goes through the calculation with periodic 
778: points. 
779:  
780: The smooth dependence of the operator $\hat g(s)$ on the quasi energy 
781: variable allow us 
782: to consider the exact response function as a distribution in phase space 
783: depending 
784: continuously on $s$, thus allowing an {\it exact} interpolation for all 
785: values of $s$. 
786: The Husimi eigen-distributions are particular values of this function 
787: which 
788: should be positive and display a pattern of $N$ zero values, in 
789: accordance 
790: with the general properties of Husimi distributions of pure states on 
791: the 
792: torus. \cite{voros} 
793:  
794: We display in Fig.3 the real part of ${\cal H}_a (e^{-i \epsilon}, z, 
795: \bar z)$ for 
796: values of $\epsilon$ near an eigenvalue. The distribution shows a very 
797: stable pattern of 
798: minima that gradually develops into $N=14$ zeros at 
799: $\nu=\epsilon/2\pi=0.4167$. At this 
800: point $Z(e^{-i \epsilon})=0$, the distribution is positive and we 
801: verify that the 
802: imaginary part vanishes for every phase space point. This is the signal 
803: \cite{voros} 
804: that a true projector has been reached and that the Husimi distribution 
805: is the modulus 
806: of an analytic function with $N$ zeros. Past the eigenvalue the 
807: distribution can become
808: negative but the pattern of minima remains very stable. At the values 
809: where $Z(e^{-i 
810: \epsilon})$ has an extremum, in between zeros, we show in the appendix 
811: that the 
812: distribution becomes flat with value $1/N$. At this point the pattern of 
813: minima changes 
814: and picks up the structure of zeros of the next eigenfunction. 
815:  
816: For cat maps the semiclassical formulae are exact and therefore the above 
817: results merely confirm that the whole scheme is consistent and accurate but
818: they do not really test the semiclassical approximation. 
819: Towards this purpose we consider another simple linear map --the baker's-- 
820: where the semiclassics is not exact. The classical $T$-iteration of the 
821: baker map from the $(q_0, p_0)$ to $(q_T, p_T)$ real coordinates 
822: \cite{baker}, takes the complex form
823:  
824: \begin{eqnarray} 
825: z_T &=& u_T ~ z_0 + v_T ~ \bar {z_0} - w_T \\ 
826: \bar z_T &=& v_T ~ z_0 + u_T ~ \bar {z_0} - \bar w_T 
827: \end{eqnarray} 
828:  
829: \begin{eqnarray} 
830: u_T &=& \frac{2^T + 2^{-T}}{2} = \cosh (T \ln 2) \\ 
831: v_T &=& \frac{2^T - 2^{-T}}{2} = \sinh (T \ln 2) \\ 
832: w_T &=& \frac{1}{\sqrt{2}} \left( \gamma + i ~ 2^{-T} \gamma ^{\dagger} 
833: \right) 
834: \end{eqnarray} 
835: $\gamma$ is the integer part of $2^T q_0$ and $\gamma^{\dagger}$ is the number 
836: obtained after 
837: inversion of their binary digits. Each possible $\gamma$ defines a periodic 
838: point $z_{\gamma}=(\gamma-i\gamma^{\dagger})/\sqrt{2}(2^T-1)$. The return 
839: amplitudes $C_n (z, \bar z)$ are expressed as a sum over them 
840:  
841: \begin{equation} 
842: C_n(z,\bar z) = \frac{1}{\sqrt{u_T}}  \sum_{\gamma=1}^{2^T-1} \exp \left[ - 
843: \frac{1}{\hbar} {\cal 
844: G}^{\gamma} (z,\bar z) +  \frac{i}{\hbar} S_{\gamma} \right] \ , 
845: \end{equation} 
846: where the complex generating function takes the standard form 
847: \begin{equation} 
848: {\cal G}^{\gamma} (z,\bar z) = \frac{1}{u_T} \left[ \frac{v_T}{2} (\bar z - 
849: \bar z_{\gamma}) 
850: ^2 + (u_T-1) (\bar z - \bar z_{\gamma}) (z - z_{\gamma}) - \frac{v_T}{2} (z - 
851: z_{\gamma})^2 
852: \right]  
853: \end{equation} 
854: and  
855: \begin{equation} 
856: S_{\gamma} = \frac{\gamma ~ \gamma ^{\dagger}}{2^T-1} 
857: \end{equation} 
858: is the action of the corresponding periodic orbit. 
859:  
860: In Fig.4 we show the exact passage of the distribution through an 
861: eigenvalue ($N=16$) where - as expected - the same general pattern as the 
862: previous example  can be observed. 
863: 
864: In Figs.5 and 6 we test the semiclassical
865: approximation by comparing both the spectrum and the eigendistributions with
866: the exact results. Fig.5 shows the $Z$ function for a small stretch of the 
867: unit circle comprising three zeroes ($ N=20 $) and showing that the 
868: semiclassical zeroes are approximated well within a fraction of the mean 
869: eigenvalue separation. 
870:  In Fig.6 we show the three eigendistributions: the top three are the
871:  exact ones obtained at the exact zeroes, while the bottom three are
872:  the corresponding semiclassical ones (computed at the semiclassical zeroes).
873:  Considering that the distributions are obtained by combining the classsical
874:  information of $ 2^{N/2}\approx 1000 $ periodic points, it is remarkable 
875: how well the approximation captures the essential structure
876:  of the eigenfunctions. One important point should be noticed: in the 
877: semiclassical case there is no guarantee that the distribution calculated at
878:  a zero of $Z$ will correspond exactly to a one dimensional projector 
879: (i.e. to a pure state) and therefore its Husimi distribution may not have 
880: exactly $N$  zeroes and may be even negative. This is clearly the case in 
881: the first semiclassical state at $\nu=0.4734$. The large white area has 
882: negative values - and is not contoured - but corresponds very well with 
883: the central area where the zeroes of the exact distribution are located. 
884: The positive top of the distribution is roughly well located but it looks 
885: like the whole distribution
886:  had been shifted towards negative values, while mantaining a very accurate
887:  correspondence of the maxima and minima. The second state at $\nu=0.5445$
888:  reproduces very well the top part of the distribution, while now the minima
889:  have become positive, but again maintaining a very accurate correspondence
890:  with the exact pattern of zeroes.  The third state at $\nu=0.6088$ 
891: reproduces in very fine detail the patterns of the exact state. Overall 
892: we have observed that this state of affairs is quite typical: the position 
893: of the maxima and minima of the semiclassical states are very well 
894: reproduced but the relative heights of peaks and valleys are not always 
895: accurate.
896: 
897:  
898:  
899: \section{Conclusions} 
900: We have presented a method that can, in principle, provide a reliable 
901: way to compute 
902: single Husimi eigendistributions and which has the same advantages --and 
903: drawbacks-- as 
904: the resummation methods based on the spectral determinant for the 
905: calculation of single 
906: eigenvalues. The final results (\ref{greenfinal}) and (\ref{gb}) for the 
907: Green operators 
908: when approximated semiclassically give approximations to the projectors 
909: on single eigenstates as sums of 
910: Gaussian centered on periodic points with periods up to half the 
911: Heisenberg time. Similar 
912: expressions, with contributions not so well localized, can be obtained 
913: for Wigner 
914: functions or probability distributions. The usual exponential 
915: proliferation of periodic 
916: orbits in a chaotic system is the main obstacle to applications in the 
917: deep semiclassical 
918: regime. These are however not limited to simple maps as the method has 
919: been applied also 
920: to the Bunimovich stadium in \cite{sara-simo}. 
921:  
922: We acknowledge many important discussions with F. Simonotti and P. Leboeuf, 
923: and partial support from ANPCyT program PICT97-01015 and ECOS-Secyt A98E03. 
924:  
925:  
926: \appendix 
927: \section{Normalized Green operator at $Z$-function extremes} 
928:  
929:  
930: To calculate the denominator of $\hat g_a$ we use a well known property 
931: of the 
932: parametrized matrices  
933:  
934: \begin{equation} 
935: \frac{d}{ds} \det \left[ \hat M(s) \right] = \det \left[ \hat M(s) 
936: \right] tr \left[ 
937: \frac{d}{ds} \left[ \hat M(s) \right] \hat M^{-1} (s) \right] \ . 
938: \end{equation} 
939:  
940: If we use this property for $\hat M(s)=\hat I - s \hat U$, then its 
941: determinant is the 
942: characteristic polynomial, and we have 
943:  
944: \begin{equation} 
945: \frac{d}{ds} P(s) = P(s) ~ tr \left[ - \hat U (\hat I - s \hat U)^{-1} 
946: \right] \ . 
947: \end{equation} 
948: Then if $P(s)$ is introduced into the trace, we obtain the transpose 
949: cofactor matrix, 
950: and we demonstrate 
951:  
952: \begin{equation} 
953: tr \left[ \hat U ~ C^t \left( \hat I - s \hat U \right) \right] = - 
954: \frac{d}{ds} P(s) \ , 
955: \end{equation} 
956: that it is exactly the denominator of $\hat g_a$. Recalling the definition 
957: of $Z(s)$ function 
958:  
959: \begin{equation} 
960: Z(s) = \frac{P(s)}{\beta_{N/2} s^{N/2}} \ ,  
961: \end{equation} 
962: we calculate its derivative respect to the real parameter $\epsilon$ 
963: that rounds the 
964: unit circle $(s=e^{-i \epsilon})$ 
965:  
966: \begin{equation} 
967: \frac{dZ}{d \epsilon} =  \frac{dZ}{ds} \frac{ds}{d \epsilon} = 
968: \frac{i~s^{-N/2}}{ 
969: \beta_{N/2} } \left[ \frac{N}{2} P(s) - s \frac{d}{ds} P(s)  \right] \ . 
970: \end{equation} 
971: Then at $s_{\alpha}=e^{-i \epsilon_{\alpha}}$ where $Z$, as a function 
972: of $\epsilon$, 
973: reach a maximum or a minimum, we have the following identity 
974:  
975: \begin{equation} 
976: \frac{dP}{ds} (s_{\alpha}) = \frac{N}{2 s_{\alpha}} P(s_{\alpha}) \ . 
977: \end{equation} 
978:  
979: The denominator of the $\hat g_a$ normalized Green operator is equal to 
980: the derivative 
981: of the spectral determinant. Then, in a diagonal basis of $\hat U$, 
982: $\hat g_a$ 
983: specialized at a extreme of the $Z$ function is 
984:  
985: \begin{eqnarray} 
986: \hat g_a(s_{\alpha}) &=&  \frac{ \sum_{n=1}^N (1/s_n)  \prod_{k \neq n} 
987: (1-s_{\alpha} 
988: /s_k) |\psi_n \rangle \langle \psi_n| } {-\frac{dP}{ds} (s_{\alpha})} 
989: \nonumber \\ 
990: &=& - \frac{ \sum_{n=1}^N (1/s_n)  \prod_{k \neq n} (1-s_{\alpha}/s_k) 
991: |\psi_n \rangle 
992: \langle \psi_n| } { \frac{N}{2s_{\alpha}} P(s_{\alpha})} \nonumber \\ 
993: &=& - \frac{2}{N} \frac{s_{\alpha}}{\prod_{k} (1-s_{\alpha} / s_k)} 
994: \sum_{n=1}^N (1/s_n) 
995: \prod_{k \neq n} (1-s_{\alpha}/s_k) |\psi_n \rangle \langle \psi_n| \ . 
996: \end{eqnarray} 
997: Introducing the spectral determinant of the denominator into the sum we 
998: obtain 
999:  
1000: \begin{equation} 
1001: \hat g_a(s_{\alpha}) = - \frac{2}{N} \sum_{n=1}^N \frac{(s_{\alpha} 
1002: /s_n)}{1-(s_{\alpha}/ s_n) } |\psi_n \rangle \langle \psi_n| \ . 
1003: \end{equation} 
1004:  
1005: We can separate the real and the imaginary part of the prefactor, 
1006: knowing that both 
1007: $s_{\alpha}$ and $s_n$ are on the unit circle. 
1008:  
1009: \begin{equation} 
1010: \frac{(s_{\alpha}/s_n)}{1-s_{\alpha} /s_n} = -\frac{1}{2} + \frac{i}{2} 
1011: \frac{ \Im 
1012: (s_{\alpha}/s_n) }{\left[ 1- \Re (s_{\alpha}/s_n) \right]} \ . 
1013: \end{equation} 
1014:  
1015: Contrary to the imaginary part, we see that the real part is independent 
1016: of the index $n$ or $\alpha$. Then the 
1017: normalized Green operator at the extremes of the $Z$ function can be 
1018: written as 
1019:  
1020: \begin{eqnarray} 
1021: \hat g(s_{\alpha}) &=& - \frac{2}{N} \sum_{n=1}^N \left( -\frac{1}{2} + 
1022: \frac{i}{2} 
1023: B_{{\alpha} n} \right)  |\psi_n \rangle \langle \psi_n| \\ 
1024: &=& \frac{1}{N} \sum_{n=1}^N  |\psi_n \rangle \langle \psi_n|  - 
1025: \frac{i}{N} 
1026: \sum_{n=1}^N B_{{\alpha} n} |\psi_n \rangle \langle \psi_n| 
1027: \end{eqnarray} 
1028: and splits into an hermitian and antihermitian part. We have demonstrated 
1029: here that the 
1030: hermitian part is exactly the identity operator over $N$. Then, the real 
1031: part of the 
1032: expectation value of $\hat g(s_{\alpha})$ will be always $1/N$, 
1033: independent of the 
1034: quantum state. 
1035:  
1036:  
1037: \begin{thebibliography}{99} 
1038:  
1039: \bibitem{berry1} M. V. Berry, Les Houches Lecture Notes, Summer School 
1040: on Chaos and 
1041: quantum Physics, M.-J. Giannoni, A. Voros, and J. Zinn-Austin, eds., 
1042: Elsevier Science 
1043: Publishers B. V. (1991); {\sl Proc. Roy. Soc.} {\bf A 243}, 219 (1989) 
1044: \bibitem{voros} P. Leboeuf and A. Voros, {\sl J. Phys.} {\bf A 23}, 1765 
1045: (1990) 
1046: \bibitem{heller} E. J. Heller, {\sl Phys. Rev. Lett.} {\bf 53}, 1515 
1047: (1984); Wavepacket 
1048: Dynamics and Quantum Chaology in {\sl Chaos and Quantum Physics}, M.-J. 
1049: Giannoni, A. 
1050: Voros, and J. Zinn-Austin, eds., Elsevier Science Publishers, Amsterdam 
1051: (1990) 
1052: \bibitem{bogo1} E. B. Bogomolny {\sl Physica} {\bf D 31}, 169 (1988) 
1053: \bibitem{vergini} D. A. Wisniacki and E. Vergini, {\sl Phys. Rev. E} 
1054: {\bf 62}, 4513 
1055: (2000) 
1056: \bibitem{nonnenmacher}S. Nonnenmacher, A. Voros {\sl J. Stat. Phys.}
1057: {\bf 92}, 431 (1998)
1058: \bibitem{mirlin} A. D. Mirlin, Lecture course given at the International 
1059: School {\sl Enrico Fermi} on New Directions in Quantum Chaos, Varenna, 
1060: July 1999 (cond-mat/0006421) 
1061: \bibitem{gutzw} M. C. Gutzwiller, {\sl J. Math. Phys.} {\bf 8}, 1979 
1062: (1967) 
1063: \bibitem{bogo-keat} E. B. Bogomolny and J. P. Keating, {\sl Phys. Rev. 
1064: Lett.} {\bf 77}, 
1065: 6091 (1994) 
1066: \bibitem{agamfish} O. Agam and S. Fishman, {\sl Phys. Rev. Lett.} {\bf 
1067: 73}, 806 (1994) 
1068: \bibitem{sara-simo} M. Saraceno and F. Simonotti, {\sl Phys. Rev. E} 
1069: {\bf 61}, 6527 
1070: (2000) 
1071: \bibitem{pseudo} M. V. Berry and J. P. Keating, {\sl J. Phys.} {\bf A 
1072: 23}, 4839 (1990) 
1073: \bibitem{bogo2} E. B. Bogomolny, {\sl Nonlinearity} {\bf 5}, 805 (1992) 
1074: \bibitem{bbl} E. B. Bogomolny, O. Bohigas and P. Leboeuf, {\sl Phys. Rev. 
1075: Lett.} {\bf 68}, 2726 (1992) 
1076: \bibitem{marsden} M. Marsden, {\sl Geometry of Polynomials}, American 
1077: Mathematical 
1078: Society, Providence, Rhode Island (1966). 
1079: \bibitem{berry-keat} M. V. Berry and J. P. Keating, {\sl Proc. Roy. 
1080: Soc.} {\bf A 437}, 
1081: 151 (1992) 
1082: \bibitem{ozorio}A. M. Ozorio de Almeida, {\sl Phys. Rep. }{\bf 295},265 (1998)
1083: \bibitem{fishman} S. Fishman, B. Georgeot and R. E. Prange, {\sl J. 
1084: Phys.} {\bf A 29}, 
1085: 919 (1996); S. Fishman in {\sl Supersymmetry and Trace Formulae: Chaos 
1086: and Disorder}, 
1087: Lerner {\sf et al.} Eds., NATO ASI {\bf B} Vol. 370, Kluwer Academic, 
1088: New York (1999) 
1089: \bibitem{hannay} J. H. Hannay and M. V. Berry, {\sl Physica} {\bf 1D}, 
1090: 267 (1980) 
1091: \bibitem{baker} M. Saraceno and A. Voros, {\sl Physica D} {\bf 79}, 206 
1092: (1994) 
1093:  
1094: \end{thebibliography} 
1095:  
1096: \newpage 
1097:  
1098: %************************************************************* 
1099: \begin{figure} 
1100: \centerline{\psfig{figure=fig1.epsi}} 
1101: \vspace{0.2cm} 
1102: \caption{\label{ucat} Modulo of the coherent state return amplitude for the 
1103: first six powers of the Arnold cat map propagator. The torus is quantized 
1104: with periodic boundary conditions, and $N=128$. Linear grey scale between 
1105: 0 (white), and the maximum (black) of each picture.} 
1106: \end{figure} 
1107: %************************************************************* 
1108: \begin{figure} 
1109: \centerline{\psfig{figure=fig2.epsi}} 
1110: \vspace{0.2cm} 
1111: \caption{\label{uxcat} Modulo of the coherent state mean value for the 
1112: $\hat U \hat X_n$ operators. Same parameters and scales as Fig. 1} 
1113: \end{figure} 
1114: %************************************************************* 
1115: \begin{figure} 
1116: \centerline{\psfig{figure=fig3.epsi}} 
1117: \vspace{0.2cm} 
1118: \caption{\label{gcat} Real part of ${\cal H}_a (s,z,\bar z)$, for $s = \exp 
1119: (-i 2 \pi \nu)$. 
1120: Linear grey scale between 0 (white), and $1/\sqrt{N}$ (black). 
1121: Logarithmic level curves 
1122: at $1/N, 1/N^2,1/N^3,...$. Arnold cat map with periodic boundary 
1123: conditions, and $N=14$. 
1124: There is an eigenvalue of the propagator near $\nu=0.4167$ (3rd. 
1125: picture). The exact and 
1126: semiclassical computation gives the same values.} 
1127: \end{figure} 
1128: %************************************************************* 
1129: \begin{figure} 
1130: \centerline{\psfig{figure=fig4.epsi}} 
1131: \vspace{0.2cm} 
1132: \caption{\label{gbak} The same as Fig. 3, but for the baker map, with 
1133: anti-periodic boundary 
1134: conditions, and $N=16$. There is an eigenvalue of the propagator near 
1135: $\nu=0.4606$ (4th. 
1136: picture).} 
1137: \end{figure} 
1138: %************************************************************* 
1139: \begin{figure} 
1140: \centerline{\psfig{figure=newfig5.eps}} 
1141: \vspace{0.2cm} 
1142: \caption{\label{zexsem} Exact $Z$ function (bold line), and its semiclassical 
1143: approximation (dashed line), for the baker map, with anti-periodic boundary 
1144: conditions, and $N=20$. The zeroes correspond to the eigenvalues}
1145: \end{figure} 
1146: %************************************************************* 
1147: \begin{figure} 
1148: \centerline{\psfig{figure=newfig6.eps}} 
1149: \vspace{0.2cm} 
1150: \caption{\label{gbaksem} Exact and semiclassical eigenstates in correspondence
1151: with the zeroes in Fig.5. The gray scale and the contours are as in Fig.3. }
1152: \end{figure} 
1153: %************************************************************* 
1154:  
1155:  
1156:  
1157:  
1158: \end{document} 
1159:  
1160:  
1161:  
1162:  
1163:  
1164:  
1165:  
1166:  
1167:  
1168:  
1169:  
1170: