nlin0608069/math7.tex
1: \documentclass[12pt]{amsart}
2: \usepackage[T1]{fontenc}
3: \usepackage{amssymb}
4: \usepackage{a4wide}
5: \usepackage{graphicx}
6: \usepackage{verbatim}
7: \usepackage{amsmath}
8: \usepackage{version}
9: \usepackage{yhmath}
10: 
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% Textclass specific LaTeX commands.
12:  \theoremstyle{plain}    
13:  \newtheorem{thm}{Theorem}%%[section]
14:  \numberwithin{equation}{section} %% Comment out for sequentially-numbered
15: %\numberwithin{figure}{section} %% Comment out for sequentially-numbered
16:  \theoremstyle{plain}
17:  \newtheorem{prop}{Proposition} %%Delete [thm] to re-start numbering
18:  \theoremstyle{plain}    
19:  \newtheorem{lem}{Lemma} %%Delete [thm] to re-start numbering
20:  \theoremstyle{plain}    
21:  \newtheorem{cor}{Corollary} %%Delete [thm] to re-start numbering
22:  \theoremstyle{plain}    
23:  \newtheorem{conj}{Conjecture} %%Delete [thm] to re-start numbering
24:  \theoremstyle{plain}    
25:  \newtheorem{question}{Question} %%Delete [thm] to re-start numbering
26: \theoremstyle{definition}
27: \newtheorem{rem}{Remark} %%Delete [thm] to re-start numbering
28: \theoremstyle{definition}
29: \newtheorem{Def}{Axioms} %%Delete [thm] to re-start numbering
30: 
31: 
32: 
33: \begin{document}
34: \input{def.tex}
35: 
36: \title%[Resonant eigenstates in quantum chaotic scattering ]
37: {Resonant eigenstates for a quantized chaotic system}
38: 
39: \author[S. Nonnenmacher]{St\'ephane Nonnenmacher}
40: \author[Mathieu Rubin]{Mathieu Rubin}
41: \address{Service de Physique Th\'eorique, 
42: CEA/DSM/PhT, Unit\'e de recherche associ\'e CNRS,
43: CEA/Saclay,
44: 91191 Gif-sur-Yvette, France}
45: \email{snonnenmacher@cea.fr}
46: 
47: \begin{abstract}
48: We study the spectrum of quantized open maps, 
49: as a model for the resonance spectrum of quantum
50: scattering systems. We are particularly interested in open
51: maps admitting a fractal repeller. Using the ``open baker's map'' 
52: as an example, we numerically
53: investigate the exponent appearing in the Fractal Weyl law
54: for the density of resonances; we show that this exponent is not 
55: related with the ``information dimension'', but rather the Hausdorff
56: dimension of the repeller. We then 
57: consider the {\em semiclassical measures} associated with 
58: the eigenstates: we prove that these measures are
59: conditionally invariant with respect to the classical dynamics.
60: We then address the problem
61: of classifying semiclassical measures among conditionally invariant ones. 
62: For a solvable model, the ``Walsh-quantized'' open baker's map, we manage to
63: exhibit a family of semiclassical measures with simple self-similar properties.
64: \end{abstract}
65: 
66: 
67: \maketitle
68: 
69: %%%%%%%%%%%%%%%%%%%%%%
70: %%%%%%%%%%%%%%%%%%%%%%
71: \section{Introduction}
72: %%%%%%%%%%%%%%%%%%%%%%
73: %%%%%%%%%%%%%%%%%%%%%%
74: 
75: %%%%%%%%%%%%%%%%%%%%%%
76: \subsection{Quantum scattering on $\IR^D$ and resonances} 
77: %%%%%%%%%%%%%%%%%%%%%%
78: 
79: In a typical scattering system, particles of positive energy
80: come from infinity, interact with a localized potential $V(q)$, and then leave
81: to infinity. The corresponding  quantum Hamiltonian 
82: $H_\hbar=-\hbar^2\Delta +V(q)$ has an
83: absolutely continuous spectrum on the positive axis.
84: Yet, the Green's function $G(z;q',q)=\la q'|(H_\hbar-z)^{-1}|q\ra$ admits
85: a meromorphic continuation from the upper half plane $\set{\Im z>0}$ to (some
86: part of) the lower
87: half-plane $\set{\Im z<0}$. This continuation generally has 
88: {\it poles} $z_j=E_j-\i\Gamma_j/2$, $\Gamma_j>0$,
89: which are called {\it resonances} of the scattering system.
90: 
91: The probability density of the corresponding ``eigenfunction'' $\varphi_j(q)$
92: decays in time like $\e^{-t \Gamma_j/\hbar}$, so physically $\varphi_j$ represents
93: a metastable state with decay rate $\Gamma_j/\hbar$, or lifetime $\tau_j=\hbar/\Gamma_j$.
94: In the semiclassical limit $\hbar\to 0$, we will call ``long-living''
95: the resonances $z_j$ such that $\Gamma_j=\cO(\hbar)$, equivalently with lifetimes 
96: bounded away from zero.
97: 
98: The eigenfunction $\varphi_j(q)$ is meaningful only
99: near the interaction region, while its behaviour outside that region
100: (exponentially increasing outgoing waves) is clearly unphysical.
101: As a result, one practical method to compute resonances (at least approximately)
102: consists in adding a {\it smooth absorbing potential} $-\i W(q)$ to the Hamiltonian
103: $\hat H$,
104: thereby obtaining a nonselfadjoint operator $H_{W,\hbar}=H_\hbar-\i W(q)$.
105: The potential $W(q)$ is supposed to vanish
106: in the interaction region, but is positive outside: its effect is to {\em absorb}
107: outgoing waves, as opposed to a real positive potential which would reflect
108: the waves back into the interaction region. Equivalently, the (nonunitary)
109: propagator $\e^{-\i H_{W,\hbar}/\hbar}$ kills wavepackets localized oustide the
110: interaction region. 
111: 
112: The spectrum of $H_{W,\hbar}$ in some neighbourhood of the
113: positive axis is then made of discrete eigenvalues $\tilde{z}_j$ associated with 
114: square-integrable eigenfunctions $\tilde\varphi_j$. 
115: Absorbing Hamiltonians of the type of $H_{W,\hbar}$ have been widely used in
116: quantum chemistry to study reaction or dissociation dynamics
117: \cite{LeforWyatt83,SeidMill92}; in those works it is implicitly assumed that eigenvalues
118: $\tilde{z}_j$ close to the real axis are small perturbations of the resonances
119: $z_j$, and that the corresponding eigenfunctions $\varphi_j(q)$, $\tilde\varphi_j(q)$ 
120: are close to one another in the interaction region.
121: Very close to the real axis (namely, for $|\Im\tilde{z}_j|=\cO(\hbar^n)$
122: with $n$ sufficiently large), one can prove that this is indeed the case \cite{Stef05}.
123: Such very long-living resonances are possible when the classical dynamics admits a
124: trapped region of positive Liouville volume.
125: In that case, resonances and the associated eigenfunctions can be approximated
126: by quasimodes of an associated closed system \cite{TangZw98}. 
127: 
128: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
129: \subsection{Resonances in chaotic scattering} 
130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
131: 
132: We will be interested in a different situation, where the set of 
133: trapped trajectories has volume zero, and is a fractal hyperbolic repeller. 
134: This case encompasses the
135: famous 3-disk scatterer in 2 dimensions \cite{GaspRice89}, or its smoothing,
136: namely the 3-bump
137: potential introduced in \cite{Sjo90} and numerically studied in \cite{Lin02}.
138: Resonances
139: then lie deeper below the real line (typically, $\Gamma_j\gtrsim \hbar$),
140: and are not perturbations of an associated real spectrum. 
141: Previous studies have focussed in counting the number of resonances in small
142: disks around the energy $E$, in the semiclassical r\'egime. Based on the
143: seminal work of Sj\"ostrand \cite{Sjo90}, several authors have conjectured
144: the following Weyl-type law:
145: \bequ\label{e:Weyl-law}
146: \#\set{z_j\in \Res(H_\hbar)\,:\,|E-z_j|\leq \gamma\hbar}\sim C(\gamma)\hbar^{-d}\,.
147: \end{equation}
148: Here the exponent $d$ is related to the trapped set at energy $E$: the latter has
149: (Minkowski) dimension $2d+1$. 
150: This asymptotics was numerically checked by Lin and Zworski for
151: the 3-bump potential
152: \cite{Lin02,LSZ03}, and by 
153: Guillop\'e, Lin and Zworski for scattering on a hyperbolic surface \cite{GLZ04}. 
154: However, only the upper bounds for the number
155: of resonances could be rigorously proven \cite{Sjo90,Zw99,GLZ04,SjoZw05}. 
156: 
157: To avoid the
158: complexity of ``realistic'' scattering systems,
159: one can study simpler models, namely quantized open maps on a compact phase space, for instance
160: the quantized open baker's map studied in \cite{NonZw05,NonZw06,KNPS06} (see \S\ref{e:open-map}). 
161: Such a model is meant to mimick
162: the propagator of the nonselfadjoint Hamiltonian $H_{W,\hbar}$, in the case where the
163: classical flow at energy $E$ is chaotic in the interacting region. The above
164: fractal Weyl law has a direct counterpart in this setting; such a fractal scaling
165: was checked for the open kicked rotator in \cite{SchoTwo04}, 
166: and for the ``symmetric'' baker's map in \cite{NonZw05}.
167: In \S\ref{s:infodim?} we numerically check this fractal law for an 
168: asymmetric version of the open baker's map; apart from
169: extending the results of \cite{NonZw05}, this model allows to
170: specify more precisely the dimension $d$ appearing in the scaling law. 
171: To our knowledge, so far the only
172: system for which the asymptotics corresponding to \eqref{e:Weyl-law} 
173: could be rigorously proven is the ``Walsh quantization'' of 
174: the symmetric open baker's map \cite{NonZw06}, which will be described 
175: in \S\ref{s:walsh}.
176: 
177: After counting resonances, the next step consists in studying the 
178: long-living resonant eigenstates $\varphi_j$ or $\tilde\varphi_j$.
179: Some results on this matter 
180: have been announced by M.Zworski and the first author in \cite{NonZw-gap}.
181: Interesting numerics were performed by M.~Lebental and coworkers for a 
182: model of open stadium billiard, relevant to describe an experimental micro-laser cavity
183: \cite{Leben06}. In the framework of quantum open maps, eigenstates of the open Chirikov
184: map have been numerically studied by Casati {\it et al.} \cite{CasMasShep99}; the authors
185: showed that, in the semiclassical
186: limit, the long-living eigenstates concentrate on the classical hyperbolic repeller. They
187: also found that the phase space structure of the eigenstates are very correlated with their 
188: decay rate. In this paper we will consider general ``quantizable'' open maps,
189: %mainly focus on a different map, namely the open baker, 
190: and formalize the above observations into rigorous statements. Our main result is
191: Theorem~\ref{thm:main} (see \S\ref{s:semiclass}), which 
192: shows that the {\it semiclassical measure} associated
193: to a sequence of quantum eigenstates is (up to subtleties due to discontinuities) 
194: necessary an {\em eigenmeasure} of the classical open map.
195: Such eigenmeasures are necessarily supported on the
196: the {\em backward trapped set}, which, in the case of a chaotic dynamics, 
197: is a fractal subset of the phase space: this motivated
198: the denomination of ``quantum fractal eigenstates'' used in \cite{CasMasShep99}.
199: Let us mention that eigenstates of quantized open maps have been studied 
200: in parallel by J.~Keating and coworkers \cite{KNPS06}.
201: Our theorem~\ref{thm:main} provides a rigorous version of
202: statements contained in their work.
203: 
204: Inspired by our experience with closed
205: chaotic systems, in \S\ref{s:abundance} we attempt to
206: classify
207: semiclassical measures among all possible eigenmeasures, in particular
208: for the open baker's map.
209: In the case of
210: the ``standard'' quantized open baker, the classification remains open.
211: In \S\ref{s:walsh} we consider a solvable model,
212: the Walsh quantization of the open baker's map, introduced in \cite{NonZw05,NonZw06}. 
213: For that model, one can explicitly construct some semiclassical measures and partially
214: answer the above questions. A further study of semiclassical measures for
215: the Walsh model will appear in a joint publication with J.~Keating, M.~Novaes and M.~Sieber.
216: 
217: \subsection*{Acknowledgments}
218: M.~Rubin thanks the Service de Physique Th\'eorique for hospitality 
219: in the spring 2005, during which this work was initiated. S.~Nonnenmacher
220: has been partially supported from the grant ANR-05-JCJC-0107-01
221: of the Agence Nationale de la Recherche.
222: Both authors are grateful to J.~Keating, M.~Novaes and M.~Sieber for
223: communicating their results concerning the Walsh-quantized baker before 
224: publication, and for interesting discussions. We also thank M.~Lebental
225: for sharing with us her preliminary results on the open stadium billiard,
226: and the anonymous referees for their stimulating comments.
227: 
228: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
229: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
230: \section{The open baker's map}
231: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
232: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
233: \subsection{Closed and open symplectic maps}\label{e:open-map} 
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: 
236: Although many of the results we will present deal with a particular family of maps on the
237: 2-torus, namely the family of baker's maps, we start by some general considerations on
238: open maps defined on a compact metric space $\hM$, equipped with a probability measure $\mu_L$. 
239: We borrow some ideas and notations from the 
240: recent review of Demers and Young \cite{DemYou06}. We start with a invertible map $\hT:\hM\to\hM$,
241: which we assume to be piecewise smooth, and to preserve the measure $\mu_L$ (when
242: $\hM$ is a symplectic manifold, $\mu_L$ is the Liouville measure).
243: We then {\em dig a hole} in $\hM$, that is a certain subset $H\subset \hM$, 
244: and decide that points falling in the hole are no more iterated, but rather
245: ``disappear'' or ``go to infinity''. The hole is assumed to be a Borel subset of $\hM$.
246: 
247: Taking $M\defeq \hM\setminus H$, we are thus lead to consider 
248: the {\em open map} $T=\hT_{|M}:M\to \hM$, or equivalently, say that
249: $T$ sends points in the hole to infinity. By 
250: iterating $T$, we see that any point $x\in M$ has a certain time of escape $n(x)$, which is the
251: smallest integer such that $\hT^n(x)\in H$ (this time can be 
252: infinite). For each $n\in\IN^*$ we call 
253: \bequ
254: \label{e:M^n}
255: M^n=\set{x\in M,\ n(x)\geq n}=\bigcap_{j=0}^{n-1} \hT^{-j}(M)\,.
256: \eequ
257: This is the domain of definition of
258: the iterated map $T^n$. The {\em forward trapped set} for the open map $T$ is
259: made of the points which will never escape in the future:
260: \bequ\label{e:forward-TS}
261: \Gamma_- = \bigcap_{n\geq 1}M^n = \bigcap_{j= 0}^\infty \hT^{-j}(M)\,.
262: \eequ
263: These definitions allow us to split
264: the full phase space into a disjoint union
265: \bequ\label{e:split}
266: \hM=\big(\bigsqcup_{n=1}^\infty M_n\big)\sqcup \Gamma_-,
267: \eequ
268: where $M_1\defeq H$, and for each $n\geq 2$, $M_n\defeq M^{n-1}\setminus M^{n}$ 
269: is the set of points 
270: escaping exactly at time $n$.
271: 
272: We also consider the backward evolution given by the inverse map $\hT^{-1}$. 
273: The hole for this backwards 
274: map is the set $H^{-1}=\hT(H)$, and we call
275: $T^{-1}$ the restriction of $\hT^{-1}$ to $M^{-1}=\hT(M)$ 
276: (the ``backwards open map''). We also define $M^{-n}=\bigcap_{j=1}^n \hT^{j}(M)$, 
277: $M_{-1}=H^{-1}$, $M_{-n}=M^{-n+1}\setminus M^{-n}$ ($n\geq 2$). This leads to the
278: the {\em backward trapped set} 
279: \bequ\label{e:backward-TS}
280: \Gamma_+ = \bigcap_{j=1}^\infty \hT^{j}(M)\,,\quad\text{and the {\em trapped set} }
281: K=\Gamma_-\cap \Gamma_+\,.
282: \eequ
283: We also have a ``backward partition'' of the phase space:
284: \bequ\label{e:split2}
285: \hM=\big(\bigsqcup_{n=1}^\infty M_{-n}\big)\sqcup \Gamma_+.
286: \eequ
287: The dynamics of the open map $T$ is interesting if 
288: the trapped sets are not empty. This is be the case for the 
289: open baker's map we will study more explicitly (see \S\ref{s:baker-trapped}
290: and Fig.~\ref{f:Cantor}).
291: 
292: \medskip
293: 
294: In order to quantize the maps $\hT$ and $T$, one needs further assumptions.
295: In general, $\hM$ is
296: a {\em symplectic manifold}, $\mu_L$ its Liouville measure, and $\hT$ a canonical
297: transformation on $\hM$. Also, we will assume that $T$ is sufficiently regular (see
298: \S\ref{s:quantum}).
299: 
300: Yet, one may also consider ``quantizations'' on more general phase spaces, 
301: like in the axiomatic framework of Marklof and O'Keefe \cite{MarOK05}. 
302: As we will explain in \S\ref{s:Lipschitz}, 
303: the ``Walsh quantization'' of the baker's map is easier to analyze if we consider it
304: as the quantization of an open map on a certain symbolic space, which
305: is not a symplectic manifold. By extension, we also call ``Liouville'' the 
306: measure $\mu_L$ for this case.
307: 
308: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
309: \subsection{The open baker's map and its symbolic dynamics}\label{s:open-baker} 
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
311: We present the closed and open maps which will be our central examples: the baker's
312: maps and their associated symbolic shifts.
313: 
314: \subsubsection{The closed baker}
315: The phase space of the baker's map is the 2-dimensional torus
316: $\hM=\t2\simeq [0,1)\times [0,1)$. A point on $\t2$ is described with
317: the coordinates $x=(q,p)$, which we call respectively position (horizontal) 
318: and momentum (vertical), to insist on the symplectic structure $dq\wedge dp$.
319: %%%%%%%%%%%%%%%
320: \begin{figure}[htbp]
321: \begin{center}
322: \includegraphics[width=13cm]{bak32bis.eps}
323: \caption{\label{f:3-baker} Sketch of the closed baker's map $\hB_{\vr}$, and its
324: open counterpart $B_{\vr}$, for the case $\vr=\vr_{sym}=(1/3,1/3,1/3)$. 
325: The three rectangles form a Markov partition.}
326: \end{center}
327: \end{figure}
328: %%%%%%%%%%%%%%%
329: We split $\t2$ into three vertical rectangles $R_i$ with widths $(r_0,r_1,r_2)\defeq \vr$
330: (such that $r_0+r_1+r_2=1$, with all $r_\ep>0$), and first define a
331: closed baker's map on $\t2$ (see Fig.~\ref{f:3-baker}):
332: \bequ\label{e:closed-baker}
333: (q,p)\mapsto \hB_{\vr}(q,p)\defeq (q',p')= \left\{  \begin{array}{ll}
334: \big(\frac{q}{r_0},\,  p\,r_0\big) & \text{if}\  0 \leq q < r_0 \\
335: \big(\frac{q-r_0}{r_1}, \,  p\,r_1+r_0 \big)& \text{if}\  r_0 \leq q < r_0+r_1 \\
336: \big(\frac{q-r_0-r_1}{r_2}, \,  p\,r_2+r_1+r_0\big)& \text{if}\  1-r_2 \leq q < 1  \,.
337: \end{array}\right. 
338: \eequ
339: This map is invertible and symplectic on $\t2$. 
340: It is discontinuous on the boundaries of the rectangles $R_i$ but smooth (actually, affine)
341: inside them.
342: We now recall how $\hB_{\vr}$ can be conjugated with a symbolic dynamics. 
343: We introduce the right shift $\hat\sigma$ on the {\em symbolic space} 
344: $\Sigma=\set{0,1,2}^{\IZ}$:
345: $$
346: \bep=\ldots\ep_{-2}\ep_{-1}\cdot\ep_0\ep_1\ldots\in \Sigma\longmapsto 
347: \hat\sigma(\bep)=\ldots\ep_{-2}\ep_{-1}\ep_0\cdot\ep_1\ldots\,.
348: $$
349: The symbolic space $\Sigma$ can be mapped to the 2-torus as follows: every
350: bi-infinite sequence $\bep\in\Sigma$ is mapped to the point $x=J_{\vr}(\bep)$ with coordinates
351: \bequ\label{e:conjugation}
352: q(\bep)=\sum_{k=0}^\infty r_{\ep_0}r_{\ep_1}\cdots r_{\ep_{k-1}}\,\alpha_{\ep_k}\,,\qquad
353: p(\bep)=\sum_{k=1}^\infty r_{\ep_{-1}}r_{\ep_{-2}}\cdots r_{\ep_{-k+1}}\,\alpha_{\ep_{-k}}\,,
354: \eequ
355: where we have set $\alpha_0=0$, $\alpha_1=r_0$, $\alpha_2=r_0+r_1$. 
356: The position coordinate (unstable direction) depends on symbols 
357: on the right of the comma, while the momentum coordinate (stable direction)
358: depends on symbols on the left.
359: The map $J_{\vr}:\Sigma\to\t2$ is
360: surjective but not injective, for instance the sequences $\ldots \ep_{n-1}\ep_n 000\ldots$ 
361: (with $\ep_n\neq 0$) and $\ldots \ep_{n-1}(\ep_n-1) 222\ldots$ have the same image on $\t2$. 
362: For this reason, it is convenient to restrict $J_{\vr}$ to the subset $\Sigma'\subset\Sigma$
363: obtained by removing from $\Sigma$ the sequences ending by $\ldots 2222 \ldots$ on the
364: left or the right.
365: $\Sigma'$ is invariant through the shift $\hat\sigma$. 
366: The map $J_{\vr|\Sigma'}:\Sigma'\to\t2$ is now bijective, and it conjugates
367: $\hat\sigma_{|\Sigma'}$ with the baker's map $\hB_{\vr}$ on $\t2$:
368: \bequ\label{e:shift-A}
369: \hB_{\vr} = J_{\vr|\Sigma'}\circ \hat\sigma\circ (J_{\vr|\Sigma'})^{-1}\,.
370: \eequ
371: Any finite sequence 
372: $\bep=\ep_{-m}\ldots\ep_{-1}\cdot\ep_{0}\ldots\ep_{n-1}$ represents a {\em cylinder} 
373: $[\bep]\subset\Sigma$, which consists in the sequences sharing the same symbols 
374: between indices $-m$ and $n-1$. 
375: We call $J_{\vr}([\bep])$
376: a {\em rectangle} on the torus (sometimes we will note it $[\bep]$).
377: This rectangle has sides
378: parallel to the two axes; it has width $r_{\ep_0}r_{\ep_1}\cdots r_{\ep_{n-1}}$
379: and height $r_{\ep_{-1}}\cdots r_{\ep_{-m}}$.
380: 
381: For any triple, the Liouville measure on $\t2$ is the push-forward through $J_{\vr}$ 
382: of a certain Bernoulli measure on
383: $\Sigma$, namely $\mu^\Sigma_L= \nu^{\Sigma_-}_{\vr}\times \nu^{\Sigma_+}_{\vr}$
384: (see \S\ref{s:Bernoulli}).
385: 
386: %=======================================
387: \subsubsection{The open baker}\label{s:baker-trapped}
388: %=======================================
389: 
390: We choose to take for the hole the middle rectangle $H=R_1=\set{r_0\leq q<1-r_2}$.
391: We thus obtain an ``open baker's map'' $B_{\vr}$ defined on $M=\t2\setminus H$:
392: $$
393: (q,p)\mapsto B_{\vr}(q,p)\defeq (q',p')= \left\{  \begin{array}{ll}
394: \big(\frac{q}{r_0},\,  p\,r_0 \big) & \text{if}\  0 \leq q < r_0 \\
395: \quad\infty& \text{if}\  r_0 \leq q < 1-r_2 \\
396: \big(\frac{q-1+r_2}{r_2}, \,  p\,r_2+1-r_2\big)& \text{if}\  1-r_2 \leq q < 1 \,.
397: \end{array}\right. 
398: $$ 
399: The ``inverse map'' $B_{\vr}^{-1}$
400: is defined on the set $M^{-1}=B_{\vr}(M)$, that is outside the backwards hole 
401: $H^{-1}=B_{\vr}(H)=\set{r_0\leq p<1-r_2}$ (see Fig.~\ref{f:3-baker}, right).
402: 
403: Due to the choice of the hole, this open map is still easy to analyze 
404: through symbolic dynamics:
405: the hole $R_1$ is the image through $J_{\vr}$ of the set $\set{\ep_0=1}\cap \Sigma'$.
406: Let us define as follows the {\it open shift}
407: $\sigma$ on $\Sigma$:
408: \bequ\label{e:shift-B}
409: \bep\in \Sigma%=\ldots\ep_{-2}\ep_{-1}\cdot\ep_0\ep_1\ldots%\stackrel{B_{\vr}}{\longmapsto} 
410: \longmapsto \sigma(\bep)\defeq 
411: \begin{cases}\infty\quad\mbox{if}\quad \ep_0=1\\
412: \hat\sigma(\bep)%\ldots\ep_{-2}\ep_{-1}\ep_0\cdot\ep_1\ldots=\hat\sigma(\bep)
413: \quad \mbox{if}\quad\ep_0\in\set{0,2}\,.
414: \end{cases}
415: \eequ
416: $J_{\vr|\Sigma'}$ conjugates the open shift $\sigma_{|\Sigma'}$
417: with $B_{\vr}$ as in \eqref{e:shift-A}. Similarly, the backwards open shift 
418: $\sigma^{-1}$, which kills the sequences
419: s.t. $\ep_{-1}=1$ and otherwise moves the comma to the left, is conjugated with
420: $B_{\vr}^{-1}$.
421: 
422: \medskip
423: 
424: These conjugations allow to easily characterize the various trapped sets of $B_{\vr}$.
425: On the symbolic space $\Sigma$, the forward (resp. backward) trapped set of the
426: open shift $\sigma$ is given by the sequences $\bep$ such that 
427: $\ep_{n}\in\set{0,2}$ for all $n\geq 0$ (resp. for all $n<0$). To obtain the
428: trapped sets for the baker's map, we restrict ourselves on $\Sigma'$ and conjugate
429: by $J_{\vr|\Sigma'}$. The image sets are given by the direct products 
430: $\Gamma_-=\cC_{\vr}\times [0,1)$, $\Gamma_+=[0,1)\times \cC_{\vr}$ 
431: and $K=\cC_{\vr}\times\cC_{\vr}$,
432: where $\cC_{\vr}$ is (up to a countable set) 
433: the Cantor set on $[0,1)$ adapted to the partition $\vr$ 
434: (see Fig.~\ref{f:Cantor}).
435: %%%%%%%%%%%%%%%%%%%%
436: \begin{figure}[htbp]
437: \begin{center}
438: \includegraphics[width=16cm]{cantor3bis.eps}
439: \caption{\label{f:Cantor} In black we approximate the forward (left), 
440: backward (center)
441: and joint (right) trapped sets
442: for the symmetric open baker $B_{\vr_{sym}}$. 
443: On the left (resp. centre), red/gray scales (from white to
444: dark) correspond to points escaping 
445: at successive times in the future (resp. past), that is to sets $M_{n}$
446: (resp. $M_{-n}$), for $n=1,2,3,4$.}
447: \end{center}
448: \end{figure}
449: %%%%%%%%%%%%%%%%%
450: 
451: For future use, we define the subset $\Sigma''\subset\Sigma$ by
452: \bequ\label{e:Sigma''}
453: \Sigma''\defeq \Sigma\setminus
454: \big(\set{\ldots 222\cdot 0\ep_1\ep_2\ldots}\cup\set{\ldots 222\cdot 2\ep_1\ep_2\ldots}\cup 
455: \set{\ldots \ep_{-2}\ep_{-1}\cdot \ep_0 222\ldots}\big)\,.
456: \eequ
457: We notice that $\Sigma''\varsupsetneq\Sigma'$, and that $J_{\vr}(\Sigma\setminus \Sigma'')$ is
458: a subset of the discontinuity set of the map $B_{\vr}$ (see \S\ref{s:quantum-baker2}
459: and Fig.~\ref{f:gamma+}). 
460: The map $J_{\vr}$ realizes a kind of {\it semiconjugacy} between $\sigma_{|\Sigma''}$ and $B$:
461: \bequ\label{e:semiconj}
462: \forall \bep\in \Sigma'',\qquad  J_{\vr}\circ\sigma(\bep)=B_{\vr}\circ J_{\vr}(\bep).
463: \eequ
464: Above it is understood that both sides are ``sent to infinity'' if $\ep_{0}=1$. 
465: 
466: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
467: \subsection{Eigenmeasures of open maps}\label{s:eigenmeasures0} 
468: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
469: 
470: Before defining eigenmeasures of
471: open maps, we briefly recall how invariant measures emerge in the study of the
472: quantized {\it closed} maps.
473: 
474: \subsubsection{Quantum ergodicity for closed chaotic systems}
475: \label{s:eigenstates-IM} 
476: 
477: The quantum-classical correspondence between a closed symplectic map
478: $\hT$ and its quantization $\hT_{N}$ (see \S\ref{s:quantum}) 
479: has one important consequence:
480: in the semiclassical limit $N\to\infty$, stationary states of the quantum system (that is,
481: eigenstates of $\hT_{N}$) should reflect the stationary properties of the
482: classical map, namely its invariant measures. 
483: To be more precise, 
484: to any sequence of eigenstates $(\psi_N)_{N\to \infty}$ of the quantum map, 
485: one can associate at least one {\bf semiclassical measure} (see \S\ref{s:semicl-measure}).
486: Omitting problems due to the discontinuities of $\hT$, the quantum-classical
487: correspondence implies that a semiclassical measure $\mu$ must be 
488: invariant w.r.to $\hT$:
489: \bequ\label{e:invariant}
490: \hT^*\,\mu=\mu\quad\Longleftrightarrow 
491: \quad\text{for any Borel set } S\subset \t2,\quad \mu(\hT^{-1}(S))=\mu(S)\,.
492: \eequ
493: $\mu$ is then also invariant w.r.to the inverse map $\hT^{-1}$.
494: 
495: If the map $\hT$ is ergodic w.r.to the Liouville measure $\mu_{L}$ on $\hM$, 
496: the quantum ergodicity theorem (or Schnirelman's theorem \cite{Schni74})
497: states that, for ``almost any'' sequence $(\psi_N)_{N\to \infty}$,
498: there is a unique associated semiclassical measure, which is $\mu_{L}$ itself.
499: 
500: Such a theorem was first proven for eigenstates of the Laplacian on compact
501: Riemannian manifolds with ergodic geodesic flow \cite{Zel87,CdV85}, then for more general 
502: Hamiltonians \cite{HelMarRob87}, billiards \cite{GerLei93,ZelZwo96}
503: and maps \cite{BouzDB96,Zel96}. Quantum ergodicity for piecewise smooth
504: maps was proven in a general setting in \cite{MarOK05}, and the particular case of the
505: baker's map was treated in \cite{DENW06}. Finally, in \cite{AN06} it was shown that
506: the (closed) Walsh-baker's map, seen as the quantization of the right shift $\hat\sigma$
507: on $(\Sigma,\mu_L)$, also satisfies quantum ergodicity with respect to $\mu_L$.
508: 
509: It is generally unknown whether there exist ``exceptional sequences'' of eigenstates, 
510: converging to a different invariant measure. 
511: The absence of such sequences is expressed by the
512: quantum unique ergodicity conjecture \cite{RudSar94}, which has been proven only for
513: systems with arithmetic properties \cite{Linden06,KurRud00}. This conjecture has 
514: been disproved for some specific systems enjoying
515: large spectral degeneracies at the quantum level, allowing for
516: sufficient freedom to build up partially localized eigenstates \cite{FNdB03,AN06,Kelmer05}. 
517: Some special eigenstates of the standard quantum baker with interesting multifractal
518: properties have been numerically identified \cite{MeenakLaksh04}, 
519: but their persistence in the semiclassical limit remains unclear.
520: 
521: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
522: \subsubsection{Eigenmeasures of open maps}\label{s:eigenmeasures} 
523: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
524: 
525: We now dig the hole $H=\hM\setminus M$, and consider the open map $T=\hT_{|M}$. 
526: Eigenmeasures (also called conditionally invariant measures) of maps ``with holes'' have been less
527: studied than their invariant counterparts. The recent article of Demers and Young \cite{DemYou06}
528: summarizes most of the properties of these measures, for maps enjoying various dynamical
529: properties. Below we describe some of these properties for general maps, before 
530: being more specific in the
531: case of the baker's map.
532: 
533: A probability measure $\mu$ on $\hM$ which is invariant
534: through $T$ {\em up to a multiplicative factor} will be called
535: an eigenmeasure of $T$:
536: \bequ\label{e:eigenmeasure}
537: T^*\,\mu=\Lambda_\mu\,\mu
538: \quad\Longleftrightarrow\quad 
539: \text{for any Borel set } S\subset\hM,\quad \mu(T^{-1}(S))=\Lambda_\mu\,\mu(S)
540: \,. 
541: \eequ
542: Here $\Lambda_\mu\in [0,1]$ (or rather $\gamma_\mu=-\log\Lambda_\mu$) 
543: is called the ``escape rate'' or ``decay rate'' of the eigenmeasure $\mu$, and is
544: given by $\Lambda_{\mu}=\mu(\cS)$. It corresponds
545: to the fact that a fraction of the particles in the support of $\mu$ 
546: escape at each step. Our definition slightly differs from the one in \cite{DemYou06}:
547: their measures are supported on $M$ and normalized there, while we choose the normalization
548: $\mu(\hM)=1$.
549: 
550: Here are some simple properties:
551: %%%%%%%%%%%%%%%
552: \begin{prop}
553: Let $\mu$ be an eigenmeasure of $T$ with decay rate $\Lambda_\mu$. \\
554: If $\Lambda_\mu=0$, $\mu$ is supported in the hole $H$. \\
555: If $\Lambda_\mu=1$, $\mu$ is supported in the trapped set $K$ (invariant measure).\\
556: If $0<\Lambda_\mu<1$, $\mu$ is supported on the set $\Gamma_+\setminus K$.
557: \end{prop}
558: %%%%%%%%%%%%%%%
559: 
560: %%%%%%%%%%%%%%%
561: \begin{figure}[htbp]
562: \begin{center}
563: \includegraphics[width=8cm]{cantor6bis.eps}
564: \caption{\label{f:Cantor6} Various components $\Gamma_+^{(n)}$
565: of the forward trapped set for the open baker $B_{\vr_{sym}}$. 
566: Various red/grey scales (from light to dark) correspond to $n=1,2,3,4$.}
567: \end{center}
568: \end{figure}
569: %%%%%%%%%%%%%%%
570: Following the decomposition \eqref{e:split}, the set $\Gamma_+\setminus K$ can be split into a 
571: disjoint union (see Fig.~\ref{f:Cantor6}):
572: $$
573: \Gamma_+\setminus K=\bigsqcup_{n\geq 1} \Gamma_+^{(n)},\quad\text{where}\qquad
574: \Gamma_+^{(n)}\defeq \Gamma_+\cap M_{n}=
575: T^{-n+1}\Gamma_+^{(1)},
576: \quad n\geq 1\,.
577: $$
578: Following \cite[Thm. 3.1]{DemYou06}, any
579: eigenmeasure can be constructed as follows.
580: %&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&
581: \begin{prop}\label{p:eigen-decompo}
582: Take some $\Lambda\in [0,1)$ and $\nu$ an arbitrary Borel probability
583: measure on $\Gamma_+^{(1)}=\Gamma_+\cap H$.
584: Define the probability measure $\mu$ on $\t2$ as follows:
585: \bequ\label{e:eigen-decompo}
586: \mu= %(1-\Lambda) \sum_{n\geq 0}\Lambda^n\,(\hT^*)^n\,\nu=
587: (1-\Lambda) \sum_{n\geq 0}\Lambda^n\,(T^*)^n\,\nu\,.
588: \eequ
589: Then $T^*\,\mu=\Lambda\,\mu$.
590: All $\Lambda$-eigenmeasures of $T$ can be written this way.
591: \end{prop}
592: %&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&&
593: This construction shows that, as long as $\Gamma_+^{(1)}$ is not empty 
594: (that is, there exists at least a point $x_0$ trapped in the past but escaping
595: in the future), there are plenty of eigenmeasures. 
596: 
597: The case where the map $\hT$ is uniformly hyperbolic has been studied in detail
598: by Chernov and Markarian \cite{Mark96,CherMar97}. The set $\Gamma_+^{(1)}$ is then 
599: uncoutable, and it is foliated by the unstable foliation. These authors 
600: focussed on eigenmeasures of the open map $T$ which are
601: absolutely continuous along the unstable direction. 
602: Even with this condition, one has plenty of eigenmeasures 
603: (we remind that
604: for a closed hyperbolic map, unstable absolute continuity is satisfied for a 
605: unique invariant measure, namely the SRB measure).
606: 
607: %=======================================
608: \subsubsection{Pure point eigenmeasures}\label{s:pp} 
609: %=======================================
610: 
611: Applying the recipe of Proposition~\ref{p:eigen-decompo} to the Dirac measure $\delta_{x_0}$ on
612: an arbitrary point $x_0\in\Gamma_+^{(1)}$, we obtain a simple, pure point $\Lambda$-eigenmeasure
613: supported on the backwards trajectory $\big(x_{-n}=T^{-n}(x_0)\big)_{n\geq 0}$. 
614: For any $\Lambda\in (0,1)$, we call this eigenmeasure
615: \bequ\label{e:mu-traject}
616: \mu_{x_0,\Lambda}\defeq (1-\Lambda)\sum_{n\geq 0} \Lambda^n\,\delta_{x_{-n}}\,.
617: \eequ
618: We recall that the pure point {\em invariant} 
619: measures of $T$ are localized on periodic orbits of $T$
620: on $K$ (which, for the case of a horseshoe map $T_{|K}$, form a countable family). 
621: On the opposite, 
622: for each $\Lambda\in [0,1)$ the family of pure point eigenmeasures 
623: $\mu_{x_0,\Lambda}$ is labelled by all $x_0\in \Gamma_+^{(1)}$ (an uncountable set).
624: 
625: %=======================================
626: \subsubsection{Natural eigenmeasure}\label{s:natural} 
627: %=======================================
628: 
629: As discussed in \cite{DemYou06}, one may search for the definition of a {\em natural} eigenmeasure of
630: $T$. The simplest (``ideal'') definition reads as follows: for any 
631: initial measure $\rho$ absolutely continuous w.r.to $\mu_L$, and
632: such that $\rho(\Gamma_-)> 0$,
633: \bequ\label{e:SRB-open}
634: \mu_{nat}=\lim_{n\to\infty} \frac{T^{*n}\,\rho}{\|T^{*n}\,\rho\|}\,,\qquad\text{where}\ 
635: \|\mu\|\defi \mu(\hM)\,.
636: \eequ
637: It was shown in \cite{Mark96,CherMar97}
638: that, for a hyperbolic open map, the limit exists and is independent of $\rho$. Besides,
639: the natural measure is then absolutely continuous along the unstable direction.
640: Yet, for a general open map the above limit does not necessarily exist, 
641: or it may depend on the initial distribution $\rho$ \cite{DemYou06}.
642: 
643: In case \eqref{e:SRB-open} holds, 
644: the eigenvalue $\Lambda_{nat}=\mu_{nat}(M)$ associated with this measure is generally called
645: ``the decay rate of the system'' by physicists.
646: For a {\em closed} chaotic map $\hT$, 
647: the natural measure is the Liouville 
648: measure $\mu_{L}$ (indeed, $\hT^{*n}\rho\to \mu_{L}$ is equivalent with
649: the fact that $\hT$ is {\em mixing} with respect to $\mu_{L}$). 
650: As explained in \S\ref{s:eigenstates-IM}, the quantum ergodicity theorem 
651: shows that this particular invariant measure is ``favored'' by quantum mechanics.
652: One interesting question we will address is the relevance of
653: $\mu_{nat}$ with respect to the quantized open map $T_N$.
654: 
655: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
656: \subsubsection{Bernoulli eigenmeasures of the open baker}\label{s:Bernoulli} 
657: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
658: 
659: We now focus on the open baker $B_{\vr}$ and the open shift $\sigma$ it is conjugated with,
660: and construct a family of eigenmeasures, called Bernoulli eigenmeasures. Some of
661: these measures will appear in \S\ref{s:walsh} as
662: semiclassical measures for the Walsh-quantized open baker.
663: 
664: 
665: The first equality in \eqref{e:conjugation} maps
666: the set $\Sigma_+$ of one-sided sequences $\cdot \ep_0\ep_1\ldots$ to
667: the position interval $[0,1]$. By a slight abuse, we also call this
668: mapping $J_{\vr}$. We will first construct Bernoulli measures on the
669: symbol spaces $\Sigma_+$ and $\Sigma$, and then push them on 
670: the interval or the torus using $J_{\vr}$. 
671: 
672: Let us recall the definition of Bernoulli measures on $\Sigma_+$. Choose
673: a {\em weight distribution} $\vP=(P_0,P_1,P_2)$, 
674: where $P_\ep\in[0,1]$ and $P_0+P_1+P_2=1$. We define the measure $\nu^{\Sigma_+}_{\vP}$
675: on $\Sigma_+$ as follows: 
676: for any $n$-sequence $\bep=\cdot\ep_0\ep_1\cdots\ep_{n-1}$, the weight of the cylinder $[\bep]$
677: is given by 
678: $$
679: \nu^{\Sigma_+}_{\vP}([\bep])=P_{\ep_{0}}\cdots P_{\ep_{n-1}}\,.
680: $$
681: The push-forward on $[0,1]$ of this measure will also be called a Bernoulli measure,
682: and called $\nu_{\vr,\vP}=J_{\vr}^*\,\nu^{\Sigma_+}_{\vP}$.
683: If we take $P_\ep=r_\ep$ for $\ep=0,1,2$, we recover $\nu_{\vr,\vr}=\nu_{Leb}$ the Lebesgue 
684: measure on $[0,1]$. 
685: If for some $\ep\in \set{0,1,2}$ we take $P_\ep=1$, we get for $\nu_{\vr,\vP}$ 
686: the Dirac measure at the point $q(\cdot \ep\ep\ep\ldots)$, which takes the respective
687: values $0$, $\frac{r_0}{1-r_1}$ and $1$. 
688: For any other distribution $\vP$,
689: the Bernoulli measure $\nu_{\vr,\vP}$ is purely singular
690: continuous w.r.to the Lebesgue measure. Fractal properties of the measures 
691: $\nu_{\vr,\vP}$ were studied in \cite{HJPPS86}. 
692:  If the weight $P_\ep$ vanishes for a single $\ep$ (e.g. $P_1=0$),
693: $\nu_{\vr,\vP}$ is supported on a Cantor set (e.g. $\cC_{\vr}$).
694: 
695: A Bernoulli measure on $\nu^{\Sigma_-}_{\vP}$ can be defined similarly.
696: By taking products of two Bernoulli measures, one easily constructs eigenmeasures of 
697: the open shift $\sigma$ or the open baker $B_{\vr}$.
698: %%%%%%%%%%%%%%%%%%%
699: \begin{prop}
700: Take any weight distribution $\vP$ such that $P_1<1$. Then the following hold.
701: 
702: $i)$ there is a unique auxiliary distribution,
703: namely $\vP^*=\set{\frac{P_0}{P_0+P_2},0,\frac{P_2}{P_0+P_2}}$,
704: such that the product measure
705: $$
706: \mu^\Sigma_{\vP}\defeq \nu^{\Sigma_+}_{\vP}\times\nu^{\Sigma_-}_{\vP^*}
707: $$
708: is an eigenmeasure of the open shift $\sigma$. The corresponding decay rate is
709: \bequ
710: \Lambda_{\vP}=1-P_1=P_0+P_2\,.
711: \eequ
712: $ii)$ for any $\vr$, the push-forward 
713: $\mu_{\vr,\vP}=J_{\vr}^*\,\mu^\Sigma_{\vP}$
714: is an eigenmeasure of the open baker $B_{\vr}$.
715: \end{prop}
716: %%%%%%%%%%%%%%%%%%%
717: By definition, the product measure has the following weight on a cylinder 
718: $[\bep]=[\ep_{-m}\ldots\ep_{n-1}]$:
719: $$
720: \mu^\Sigma_{\vP}([\bep])=P^*_{\ep_{-m}}\ldots P^*_{\ep_{-1}}\,P_{\ep_{0}}\cdots P_{\ep_{n-1}}\,.
721: $$
722: The proof of the first statement is straightforward.
723: There is a slight subtlety concerning push-forwards of $\sigma$-eigenmeasures, 
724: which is resolved in the following 
725: %============
726: \begin{prop}\label{l:eigenmeasure-push}
727: Let $\mu^\Sigma$ be an eigenmeasure of the open shift $\sigma:\Sigma\to\Sigma$. 
728: If $\mu^\Sigma$ does not charge the subset $\Sigma\setminus\Sigma''$ described in \eqref{e:Sigma''}, 
729: that is if $\mu^\Sigma(\Sigma\setminus\Sigma'')=0$, then its push-forward
730: $\mu=J_{\vr}^*\mu^{\Sigma}$ on $\t2$ is an eigenmeasure of $B_{\vr}$.
731: \end{prop}
732: %============
733: \begin{proof}
734: Take any Borel set $S\in\t2$.
735: Then the following identities hold:
736: $$
737: \mu(B_{\vr}^{-1}(S))\defeq \mu^\Sigma(J_{\vr}^{-1}\circ B_{\vr}^{-1}(S))
738: =\mu^\Sigma(\sigma^{-1}\circ J_{\vr}^{-1}(S))%\\
739: =\Lambda \, \mu^\Sigma(J_{\vr}^{-1}(S))
740: \defeq \Lambda \, \mu(S)\,.
741: $$
742: The second equality is a consequence of the semiconjugacy \eqref{e:semiconj}, which implies
743: the fact that the symmetric difference of the sets $J_{\vr}^{-1}\circ B_{\vr}^{-1}(S)$ and 
744: $\sigma^{-1}\circ J_{\vr}^{-1}(S)$
745: is necessarily a subset of $\Sigma\setminus\Sigma''$.
746: \end{proof}
747: The condition $\mu^\Sigma(\Sigma\setminus\Sigma'')=0$ in the proposition cannot be removed. 
748: Indeed, by applying the 
749: construction of proposition~\ref{p:eigen-decompo} to an initial ``seed'' 
750: supported on $\set{\ldots 222\cdot 1\ep_1\ldots}$, one obtains an eigenmeasure of $\sigma$ 
751: charging $\Sigma''$, and such that its push-forward is not an eigenmeasure of $B_{\vr}$.
752: 
753: To prove the second point of the Proposition, we remark that 
754: the only Bernoulli eigenmeasure $\mu^\Sigma_{\vP}$ charging $\Sigma\setminus\Sigma''$
755: is the Dirac measure on the sequence $\ldots 2222\ldots$, which is
756: pushed-forward to the delta measure at the origin of $\t2$. 
757: $\hfill\square$
758: 
759: If we take $\vP=\vr$, the push-forward is
760: the natural eigenmeasure $\mu_{\vr,\vr}=\mu_{nat}$ of the open baker $B_{\vr}$.
761: If $\vP\neq\vP'$, the Bernoulli eigenmeasures $\mu^\Sigma_{\vP}$ and
762: $\mu^\Sigma_{\vP'}$
763: are {\em mutually singular} (there exists disjoint Borel subsets $A$, $A'$ of $\Sigma$
764: such that $\nu_{\vr,\vP}(A)=\nu_{\vr,\vP'}(A')=1$),
765: eventhough they may share the same decay rate.
766: Except in the case $\vP=(1,0,0)$, $\vP'=(0,0,1)$, the push-forwards 
767: $\mu_{\vr,\vP}$ and $\mu_{\vr,\vP'}$ are also mutually singular.
768: 
769: 
770: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
771: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
772: \section{Quantized open maps}
773: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
775: 
776: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
777: \subsection{``Axioms'' of quantization}\label{s:quantum}
778: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
779: For appropriate $2D$-dimensional compact symplectic manifolds $\hM$, one may
780: define a sequence of ``quantum'' Hilbert spaces $(\hn)_{N\to\infty}$ of 
781: finite dimensions $N$, which are related with Planck's constant by
782: $N\sim \hbar^{-D}$. 
783: Quantum states are normalized vectors in $\hn$. One
784: also wants to quantize observables, that is functions $a\in C^\infty(\hM)$,
785: into operators $\Op_N(a)$ on $\hn$. An invertible (resp. open) map $\hT$ (resp. $T$)
786: is quantized into a family of
787: a {\em unitary} (resp. contracting) operators $\hT_N$ (resp. $T_N$), 
788: which satisfies certain properties when $N\to\infty$ (see below). 
789: 
790: For the example we will treat explicitly (the open baker's map), 
791: the propagators of the closed and open maps are related by
792: $T_N=\hT_N\circ\Pi_{M,N}$,
793: where $\Pi_{M,N}$ is a projector associated with the subset $M$: it kills 
794: the quantum states microlocalized in the hole, while keeping unchanged
795: the states microlocalized inside $M$ \cite{SaVa96}. 
796: Yet, in general the propagator
797: $T_N$ can be defined without having to construct $\hT_N$ beforehand.
798: 
799: The proof of our main result, Theorem~\ref{thm:main}, only uses some
800: ``minimal'' properties of the quantized observables and
801: maps. These properties were presented as
802: ``quantization axioms'' by Marklof and O'Keefe in the case of closed maps \cite{MarOK05}.
803: We adopt the same approach, namely define quantization through these ``minimal axioms'',
804: and state our result in this general framework.
805: Afterwards, we will check that these
806: axioms are satisfied for the quantized open baker.
807: 
808: %================================
809: \subsubsection{Axioms on observables}\label{s:axioms-obs}
810: %================================
811: 
812: All axioms will describe properties of the quantum operators in the semiclassical limit 
813: $N\to\infty$. With a slight abuse of notation, we 
814: write $A_N\sim B_N$ when two families of operators $(A_N)$, $(B_N)$ on $\hn$ satisfy
815: $$
816: \norm{A_N-B_N}_{\cL(\hn)}\Nto8 0\,.
817: $$
818: The axioms concerning the quantization of observables
819: read as follows \cite[Axiom~2.1]{MarOK05}. 
820: \begin{Def}\label{d:axiom-obs}
821: For any $a\in C^\infty(\hM)$, the operators
822: $\Op_N(a)$ on $\hn$ must satisfy, in the limit $N\to\infty$:
823: \bequ
824: \begin{split}\label{e:axiom-obs}
825: \Op_N(\bar{a})\sim \Op_N(a)^{\dagger}\,&\qquad \text{(asymptotic hermiticity)}\\
826: \Op_N(a)\,\Op_N(b)\sim \Op_N(ab)\,&\qquad \text{(0-th order symbolic calculus)}\\
827: \lim_{N\to\infty} N^{-1}\,\Tr \Op_N(a)=\int_{\hM}a\,d\mu_{L}&\qquad \text{(normalization)}\,.
828: \end{split}
829: \eequ
830: \end{Def}
831: These axioms are satisfied by all standard quantization recipes (e.g. geometric or
832: Toeplitz quantization on K\"ahler manifolds \cite{Zel96}). Notice that they do not
833: involve the symplectic structure on $\hM$, and can thus be extended to more general phase
834: spaces.
835: 
836: %================================
837: \subsubsection{Semiclassical mesures}\label{s:semicl-measure}
838: %================================
839: From there, we may define semiclassical measures associated with sequences of (normalized)
840: quantum states $(\psi_N\in\hn)_{N\to\infty}$. Such a sequence is said to converge to the
841: distribution $\mu$ on $\hM$ iff, for any observable $a\in C^\infty(\hM)$, 
842: \bequ\label{e:semiclass-measure}
843: \la \psi_N,\,\Op_N(a)\psi_N\ra\Nto8 \int_{\hM}a\,d\mu\,.
844: \eequ
845: From the above axioms, one can show that $\mu$ is necessarily 
846: a probability measure on $\hM$,
847: which is called the semiclassical measure associated with $(\psi_N\in\hn)$.
848: By weak compactness, from any sequence $(\psi_N\in\hn)$ one can always extract a
849: subsequence $(\psi_{N_k})$ converging in the above sense to a certain measure. The latter
850: is called {\it a} semiclassical measure of the sequence $(\psi_N)$.
851: 
852: %==============================
853: \subsubsection{Axioms on maps}\label{s:axiom-maps}
854: %==============================
855: 
856: In the axiomatic framework of \cite{MarOK05}, the conditions satisfied by 
857: the unitary propagators $(\hT_N)$ quantizing a closed map $\hT$ 
858: consist in some form of quantum-classical correspondence (Egorov's theorem), 
859: when evolving quantum observables $\Op_N(a)$ through these propagators. 
860: However, if $\hT$ is discontinuous (or nonsmooth) at some
861: points, its propagator $\hT$ will exhibit diffraction phenomena 
862: around these ``singular'' points,
863: which alter the propagation properties. At the classical level,
864: a smooth observable $a\in C^\infty(\hM)$ is transformed into 
865: a smooth observable $a\circ \hT\in C^\infty(\hM)$ only if $a$ vanishes
866: near the singular points of $\hT^{-1}$.
867: As a result, the quantum-classical correspondence
868: is a reasonable axiom only if the observable $a$ is
869: supported ``far away'' from the singular set of $\hT^{-1}$..
870: 
871: We now specify our assumptions on the open map $T:M\to\hM$. Firstly, the
872: hole $H$ will be a ``nice'' set, that is a set with nonempty interior and such 
873: that $\partial H$ has Minkowski content zero 
874: ($i.e.$ the volume of its $\eps$-neighbourhood vanishes when $\eps\to 0$).
875: We also assume that $M$, the domain of definition of $T$, can be decomposed using 
876: finitely or countably many open connected sets $O_i$:
877: $\overline{M}=\overline{\cup_{i\geq 1}O_i}$, with $O_i\cap O_j=\emptyset$, and for each $i$, 
878: $T_{|O_i}:O_i\to T(O_i)$ is a smooth canonical diffeomorphism, 
879: with all derivatives uniformly bounded. 
880: Let us split the hole into
881: $H=\ring{H}\sqcup DH$.
882: The {\it continuity set} (resp. {\it discontinuity set}) of the map $T$ is defined as
883: $$
884: C(T)=\sqcup_{i\geq 1}O_i\subset \ring{M}\,,
885: \quad\text{resp.}\quad D(T)\defi (M\setminus C(T))\sqcup DH\,.
886: $$
887: We assume that $D(T)$ has Minkowski content zero. We have the decomposition
888: $\hM=C(T)\sqcup \ring{H}\sqcup D(T)$. A similar decomposition holds for the inverse map:
889: \bequ\label{e:inverse-decompo}
890: \hM=C(T^{-1})\sqcup \ring{H}^{-1}\sqcup D(T^{-1}),\quad\text{with}\quad 
891: C(T^{-1})=T\big(C(T)\big)\,.
892: \eequ
893: Adapting the axioms of \cite{MarOK05} to the case of open maps, we set as follows
894: the characteristic
895: property of the operators $(T_N)_{N\to\infty}$ quantizing $T$.
896: \begin{Def}\label{d:axiom}
897: We say that the operators $\big(T_N\in\cL(\hn)\big)_{N\to\infty}$ quantize
898: the open map $T$ iff
899: \begin{itemize}
900: \item for $N$ large enough, $\norm{T_N}_{\cL(\hn)}\leq 1$ 
901: \item for any observable $a\in C^\infty_c\big(C(T^{-1})\sqcup \ring{H}^{-1}\big)$, we
902: have in the limit $N\to\infty$
903: \bequ\label{e:axiom2-map}
904: T_N^{\dagger}\,\Op_N(a)\,T_N \sim \Op_N\big(\bbbone_{M}\times(a\circ T)\big)\,.
905: \eequ
906: \end{itemize}
907: Here $C^\infty_c(S)$ indicates the smooth functions compactly supported inside $S$,
908: and $\bbbone_{M}$ is the characteristic function on $M$.
909: \end{Def}
910: Notice that, if $a$ is supported inside $C(T^{-1})\sqcup \ring{H}^{-1}$, 
911: the function $\bbbone_{M}\times(a\circ T)$ is well-defined,
912: smooth and supported inside $C(T)$. 
913: The factor $\bbbone_{M}$ ensures that
914: an observable supported inside $H^{-1}$ is ``semiclassically killed'' 
915: by the evolution through $T_N$.
916: The condition \eqref{e:axiom2-map} reminds of the definition of a
917: ``quantized weighted relation'' introduced in \cite{NonZw06}, but it is less precise
918: (it only describes the lowest order in $\hbar$).
919: 
920: \begin{rem}
921: Going back to the problem of potential scattering mentioned in the introduction,
922: we expect the operators $T_N$ to share some spectral
923: properties with the propagator $\exp\big(-\i H_{W,\hbar}/\hbar\big)$ of the
924: ``absorbing Hamiltonian'', in the semiclassical limit.
925: The eigenvalues $\set{\lambda_{j}}$ of $T_{N}$ 
926: should be compared with $\set{\e^{-\i \tilde{z}_j/\hbar}}$, or with 
927: $\set{\e^{-\i z_j/\hbar}}$, where $\set{\tilde{z}_j}$ (resp. $\set{z_j}$) are the
928: eigenvalues of $H_{W,\hbar}$ (resp. the resonances of $H_\hbar$). Similarly, 
929: the eigenstates of $T_N$ should share some microlocal properties with the eigenfunctions
930: $\tilde\varphi_j$ of $H_{W,\hbar}$ (resp. the resonant states $\varphi_j$ of $H_\hbar$)
931: inside the interaction region. Accordingly, we will sometimes 
932: call ``resonances'' and ``resonant
933: eigenstates'' the eigenvalues/states of quantized open maps.
934: \end{rem}
935: 
936: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
937: \subsection{The quantum open baker's map}\label{s:quantum-baker}
938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
939: 
940: For any dimension $N=(2\pi\hbar)^{-1}$, the quantum Hilbert space $\hn$ 
941: adapted to the torus phase space is spanned by the
942: orthonormal {\em position basis} $\set{\bq_j,\ j=0,\ldots,N-1}$, localized at the 
943: discrete positions $q_j=\frac{j}N$. 
944: 
945: There are several standard ways to quantize observables on $\t2$: Weyl quantization,
946: Toeplitz (or anti-Wick) quantizations \cite{BouzDB96}, or Walsh quantization \cite{AN06}. 
947: Weyl and anti-Wick quantizations are
948: equivalent with each other in the semiclassical limit, in the sense that 
949: $\Op_N^{W}(a)\sim  \Op_N^{AW}(a)$ for any smooth observable. 
950: On the opposite, the
951: Walsh quantization (see \S\ref{s:walsh0}) is not equivalent 
952: to the previous ones.
953: 
954: After recalling the definition of the anti-Wick quantization on $\t2$ and
955: the ``standard'' quantization of the baker's map,
956: following the original approach of Balazs-Voros, Saraceno, 
957: Saraceno-Vallejos \cite{BV89,Sar90,SaVa96}, we check that the latter
958: satisfies Axioms~\ref{d:axiom}
959: with respect to the anti-Wick quantization.
960: 
961: %=====================================================
962: \subsubsection{Anti-Wick quantization of observables}\label{s:Husimi} 
963: %=====================================================
964: We recall the definition and properties of coherent states on $\t2$,
965: which we use to construct the anti-Wick quantization of observables,
966: and by duality the Husimi representation
967: of quantum states \cite{BouzDB96}.
968: The Gaussian coherent state in $L^2(\IR)$, localized at the phase space
969: point $\bx=(q_0,p_0)\in\IR^2$ and with squeezing parameter $\ssigma>0$, 
970: is defined by the normalized wavefunction
971: \begin{equation}
972:   \label{e:planeCS}
973:   \Psi_{\bx,\ssigma}(q)\defeq  \left( \frac{\ssigma}{\pi\hbar}\right)^{1/4}
974:    \,\e^{-\i \frac{p_0 q_0}{2\hbar}}\,
975: \e^{\i \frac{p_0 q}{\hbar}}\,
976: \e^{-\ssigma \frac{(q-q_0)^2}{2\hbar}}\,.
977: \end{equation}
978: When $\hbar=(2\pi N)^{-1}$, 
979: that state can be periodized on the torus, to yield the torus coherent state
980: $\psi_{\bx,\ssigma}\in\hn$ with following components in the basis $\set{\bq_j}$:
981: \bequ\label{e:torusCS}
982: \la\bq_j, \psi_{\bx,\ssigma}\ra=\frac{1}{\sqrt{N}}\,\sum_{\nu\in\IZ}
983: \Psi_{\bx,\ssigma}(j/N+\nu),\qquad
984: j=0,\ldots,N-1\,.
985: \eequ
986: For $s>0$ fixed, these states are asymptotically normalized when $N\to\infty$.
987: 
988: To any squeezing $\ssigma>0$ and inverse 
989: Planck's constant $N\in\IN$ we
990: associate the anti-Wick (or Toeplitz) quantization
991: \bequ\label{e:anti-Wick}
992: f\in C^\infty(\t2)\longmapsto
993: \OpAW(f)\defi \int_{\t2}  |\psi_{\bx,\ssigma}\ra\la\psi_{\bx,\ssigma}|\;f(\bx)\;N\,d\bx\,,
994: \eequ
995: which satisfies the Axioms~\eqref{d:axiom-obs} \cite{BouzDB96}.
996: By duality, this quantization defines, for any state $\psi\in\hn$, 
997: a Husimi distribution $H^\ssigma_\psi$:
998: $$
999: \forall f\in C^\infty(\t2),\qquad H^\ssigma_\psi(f)\defeq \la\psi,\OpAW(f)\,\psi\ra\,.
1000: $$
1001: For $\norm{\psi}=1$, this distribution is a probability measure, with density given by 
1002: the (smooth, nonnegative) Husimi function 
1003: \bequ\label{e:Husimi}
1004: H^\ssigma_{\psi}(\bx)= N\,|\la \psi_{\bx,\ssigma},\psi\ra|^2\,,\quad
1005: \bx\in\t2\,.
1006: \eequ
1007: Applying the definition \eqref{e:semiclass-measure} to the present framework, 
1008: a sequence of states $(\psi_N\in\hn)_{N\to\infty}$ converges
1009: to the measure $\mu$ on $\t2$ iff, for any given $\ssigma>0$, the Husimi measures
1010: $(H^{\ssigma}_{\psi_N})_{N\to\infty}$ weak-$*$ converge to the measure $\mu$.
1011: 
1012: Following Schubert's work \cite{Roman-PhD}, 
1013: one can also consider anti-Wick 
1014: quantizations  (and dual Husimi measures)
1015: in which the squeezing parameter $\ssigma$ in the integral \eqref{e:anti-Wick} 
1016: depends on the phase space point.
1017: Adapting the proofs of \cite{Roman-PhD} to the torus setting, one
1018: shows that all these quantizations are equivalent to one another:
1019: %+++++++++++++++++++++++++++++
1020: \begin{prop}\label{p:quant}
1021: Choose two functions  $\ssigma_1,\ssigma_2\in C^\infty(\t2,(0,\infty))$.
1022: Then the two associated anti-Wick quantizations become close
1023: to one another when $N\to\infty$:
1024: $$
1025: \forall f\in C^\infty(\t2),\qquad \norm{{\rm Op}^{{\rm AW}\!,\ssigma_1}_N(f)-
1026: {\rm Op}^{{\rm AW}\!,\ssigma_2}_N(f)}=\cO(N^{-1})\,.
1027: $$
1028: \end{prop}
1029: %+++++++++++++++++++++++++++++
1030: %As a result, the convergence of $(H_{\psi_N}^{\ssigma})$ to the semiclassical 
1031: %measure $\mu$ is independent of the choice of $\ssigma$.
1032: 
1033: %=====================================================
1034: \subsubsection{Standard quantization of the baker's map}\label{s:quantum-baker2}
1035: %=====================================================
1036: Strictly speaking, the quantization 
1037: of the closed baker's map $\hB_{\vr}$ is well-defined only 
1038: if the coefficients $\vr$ are such that
1039: \bequ\label{e:condition}
1040: N\,r_i=N_i\in\IN,\quad i=0,1,2\,.
1041: \eequ
1042: Yet, in the
1043: semiclassical limit $N\to\infty$ one can, if necessary,
1044: slightly modify the $r_i$ by amounts $\leq 1/N$ in order to satisfy this condition: such
1045: a modification is irrelevant for the classical dynamics. Assuming 
1046: \eqref{e:condition}, the quantization of $\hB_{\vr}$ on $\hn$ is given by the
1047: following unitary matrix in the position basis \cite{BV89}:
1048: \bequ\label{e:A_N}
1049: \hB_{\vr,N}=F_N^{-1}\begin{pmatrix}F_{N_0}&&\\&F_{N_1}&\\&&F_{N_2}\end{pmatrix}\,.
1050: \eequ
1051: Here $F_{N_i}$ denotes the $N_i$-dimensional discrete Fourier transform, 
1052: \bequ\label{e:DFT}
1053: (F_{N_i})_{jk}={N_i}^{-1/2}\,\e^{-2\i\pi jk/N_i},\quad j,k=0,\ldots N_i-1.
1054: \eequ
1055: Since we already have a quantization for the $\hB_{\vr}$ and the hole is the 
1056: rectangle $H=R_1=\set{r_0\leq q<1-r_2}$, a natural choice to quantize $B_{\vr}$ is to
1057: project on the positions $q\in [0,1)\setminus [r_0,1-r_2)$ and then apply $\hB_{\vr,N}$.
1058: One gets the following open propagator in the position basis \cite{SaVa96}:
1059: \bequ\label{e:B_N}
1060: B_{\vr,N}=F_N^{-1}\begin{pmatrix}F_{N_0}&&\\&0&\\&&F_{N_2}\end{pmatrix}\,.
1061: \eequ
1062: This is the ``standard'' quantization of the open baker's map $B_{\vr}$. These
1063: matrices are obviously contracting on $\hn$.
1064: The semiclassical
1065: connection between the matrices $\hB_{\vr,N}$ (resp. $B_{\vr,N}$) and the classical map 
1066: $\hB_{\vr}$ (resp. $B_{\vr}$) has
1067: been analyzed in detail in \cite{NonZw06}; this analysis implies the property
1068: \eqref{e:axiom2-map} with respect to the Weyl quantization.
1069: %%%%%%%%%%%%%%%%%%%%
1070: \begin{figure}[htbp]
1071: \begin{center}
1072: \includegraphics[width=.6\textwidth]{gamma+3.eps}
1073: \caption{\label{f:gamma+} On the left, we show the backward trapped set $\Gamma_+$ for the
1074: symmetric open baker $B_{\vr_{sym}}$ (black), and the discontinuity set $D(B_{\vr_{sym}}^{-1})$ 
1075: (pink/light gray). Their union gives $\Gamma_+\sqcup D_{-\infty}$, which contains
1076: the support of semiclassical measures (see Thm.~\ref{thm:main}, $i)$).
1077: The dotted lines indicate identical points on $\t2$. 
1078: On the right, we show the backwards image 
1079: $B_{\vr_{sym}}^{-1}(D(B_{\vr_{sym}}^{-1}))$ involved in 
1080: Thm~\ref{thm:main}, $iii)$ (pink/light gray), and the projection on $\t2$ of the set 
1081: $\Sigma\setminus\Sigma''$ defined in \eqref{e:Sigma''} (red/dark gray).} 
1082: \end{center}
1083: \end{figure}
1084: %%%%%%%%%%%%%%%%%%%%
1085: Below we give an alternative proof of the property \eqref{e:axiom2-map} for the open
1086: propagator $B_{\vr,N}$, with respect to the
1087: anti-Wick quantization. The proof uses the semiclassical propagation of 
1088: coherent states analyzed in \cite{DENW06}. 
1089: 
1090: To start with, we precisely give the continuity set
1091: of $B_{\vr}^{-1}$:
1092: $$
1093: C(B_{\vr}^{-1})=\tR_0\sqcup \tR_2\,,
1094: $$
1095: where $\tR_0=\set{q\in(0,1),\ p\in (0,r_0)}\subset B_{\vr}(R_0)$ 
1096: and similarly for $\tR_2$: in the case of the symmetric baker,
1097: these are the open grey rectangles on 
1098: the right plot of Fig.~\ref{f:3-baker}.
1099: On the other hand, the hole $H^{-1}=B_{\vr}(R_1)=\set{q\in [0,1),\ p\in [r_0,1-r_2)}$ (this is 
1100: the white strip, including the vertical side). 
1101: The discontinuity set $D(B_{\vr_{sym}}^{-1})$
1102: is shown in Fig.~\ref{f:gamma+} (pink/light gray lines in the left plot).
1103: 
1104: Any observable 
1105: $a\in C^\infty_c\big(C(B_{\vr}^{-1})\sqcup \ring{H}^{-1}\big)$ is supported 
1106: at a distance $\geq\delta>0$
1107: from the discontinuity set $D(B_{\vr}^{-1})$. 
1108: Let select some smooth $\ssigma\in C^\infty(\t2,(0,\infty))$, and consider
1109: the corresponding anti-Wick quantization (see \eqref{e:anti-Wick}).
1110: From the definition \eqref{e:anti-Wick}, the operator 
1111: ${\rm Op}^{{\rm AW}\!,\ssigma}(a)$ only involves coherent states
1112: $\psi_{\bx,\ssigma(\bx)}$ located at distance $\geq\delta>0$ from $D(B_{\vr}^{-1})$.
1113: Adapting the proof of \cite[Prop.5]{DENW06} to the general open baker $B_{\vr}$,
1114: one can show the following propagation for such states:
1115: \bequ\label{e:CS-evol}
1116: B_{\vr,N}^\dagger\; \psi_{\bx,\ssigma(\bx)}=\bbbone_{M^{-1}}(\bx)\,\e^{i \theta(\bx,N)} \,
1117: \psi_{\bx',\ssigma'(\bx')}
1118: +\cO_{\hn}(\e^{-N\,C(\ssigma,\del)})\,,\qquad N\to\infty\,.
1119: \eequ
1120: Here $\bx'=B_{\vr}^{-1}(\bx)$, $\ssigma'(\bx')=\ssigma(\bx)/r_\eps^2$ if $\bx'$ lies inside the
1121: rectangle $R_{\eps}$, and $\theta(\bx,N)$ is a phase which can be explicitly computed. 
1122: 
1123: Through the symplectic 
1124: change of variable $\by=B_{\vr}^{-1}(\bx)$ for $\by\in M$, one gets
1125: $$
1126: B_{\vr,N}^\dagger\,{\rm Op}^{{\rm AW}\!,\ssigma}(a)\,B_{\vr,N}= 
1127: {\rm Op}^{{\rm AW}\!,\ssigma'}(\bbbone_{M}\times (a\circ B_{\vr})) +
1128: \cO_{\cB(\hn)}(\e^{-N\,C'(\ssigma,\del)})\,.
1129: $$
1130: The function $\ssigma'$ is obtained by taking $\ssigma'(\by)=\ssigma(B_{\vr}(\by))/r_\eps^2$ for 
1131: $\by\in B_{\vr}^{-1}(\supp a\cap \tR_{\eps})$, and smoothly extending the function to 
1132: $\ssigma'\in C^\infty(\t2,(0,\infty))$.
1133: Since the quantizations with parameters $\ssigma$ and $\ssigma'$ are equivalent 
1134: (Prop.~\ref{p:quant}), we have proven
1135: the property \eqref{e:axiom2-map} for the family $(B_{\vr,N})$. $\hfill\square$
1136: 
1137: 
1138: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1139: \section{Fractal Weyl law for the quantized open baker} 
1140: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1141: In this section, which mainly presents numerical results, 
1142: we exclusively consider the open baker's maps $B_{\vr}$ and their 
1143: quantizations \eqref{e:B_N}. Our aim is to investigate the precise notion of dimension
1144: entering the fractal Weyl law conjectured below, through a numerical study which
1145: complements the one performed in \cite{NonZw05}. Still, we believe that most
1146: statements should hold as well if we replace $B_{\vr}$ by a map $T$ obtained by
1147: restricting an Anosov diffeomorphism $\hT$ outside some 
1148: ``small hole'', as described in \cite{CherMar97}.
1149: 
1150: From the explicit formula \eqref{e:B_N}, the subspace of $\hn$ spanned by 
1151: $\set{\bq_j,\ N_0\leq j<N-N_2}$ is in the kernel of $B_{\vr,N}$. We call
1152: ``nontrivial'' the spectrum of $B_{\vr,N}$ on the complementary subspace.
1153: That spectrum is situated inside the unit disk. In the semiclassical 
1154: limit $N\to\infty$, most of it accumulates 
1155: near the origin, which corresponds to ``short-living'' eigenvalues \cite{SchoTwo04}.
1156: We rather focus on ``long-living'' eigenvalues, situated in some annulus away from the origin. 
1157: By analogy with the case of potential scattering \eqref{e:Weyl-law}, a fractal Weyl law
1158: for the semiclassical density of ``long-living'' eigenvalues was conjectured in \cite{NonZw06}:
1159: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1160: \begin{conj}[Fractal Weyl Law]\label{conj:FWL}
1161: Let $B_{\vr,N}$ be the quantized open baker's map described in \S\ref{s:quantum-baker}.  
1162: Then, for any radius $0<r<1$, there exists $C_r\geq 0$ such that
1163: \bequ\label{e:FWL}
1164: n(N,r)\defeq
1165: \#\set{ \lambda \in \Spec ( B_{\vr,N} ),\ :\  |\lambda|\geq r}=C_r\,N^{d}\,+o(N^{d }),
1166: \qquad N\to\infty\,.
1167: \eequ
1168: The eigenvalues
1169: are counted with multiplicities, and $2d$ is an appropriate 
1170: fractal dimension of the trapped set $K$. 
1171: \end{conj}
1172: %%%%%%%%%%%%%%%%%%%%%%%%%%%
1173: %So far, this conjecture could be proven only for the Walsh quantization $\tB_N$
1174: %of the symmetric baker $B_{\vr_{sym}}$, see \S\ref{s:Walsh-baker}.
1175: %This Weyl law was numerically checked for the open kicked rotator \cite{SchoTwo04}, and
1176: %for the ``standard'' quantum open baker $B_{\vr_{sym},N}$, especially 
1177: %for $N$ along geometric sequences $N=N_o\,3^k$, $k\geq 1$ \cite{NonZw05}. 
1178: 
1179: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1180: \subsection{Which dimension plays a role?}\label{s:infodim?} 
1181: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1182: %%%%%%%%%%%%%%%%%%%%
1183: \begin{figure}[htbp]
1184: \begin{center}
1185: \includegraphics[width=8cm]{ecart-r0=0.1-1-32.2-3.eps}
1186: \caption{\label{f:fit} Standard deviations when fitting the Weyl law \eqref{e:FWL} 
1187: to various dimensions, integrated on $0.1\leq r\leq 1$. The two marks on the horizontal
1188: axis indicate the theoretical values $d_I$ and $d_0$.}
1189: \end{center}
1190: \end{figure}
1191: %%%%%%%%%%%%%%%%%%%%
1192: %%%%%%%%%%%%%%%%%%%%
1193: \begin{figure}[htbp]
1194: \begin{center}
1195: \includegraphics[width=10cm]{spect-asymB-unscaled-BW.eps}
1196: \includegraphics[width=10cm]{spect-asymB-rescaled-BW2.eps}
1197: \caption{\label{f:asymB}Top: spectral counting function for the asymmetric baker $B_{\vr_{asym}}$, 
1198: for various values of Planck's constant $N$. Bottom: same curves vertically
1199: rescaled by $N^{-d_0}$. The thick tick mark indicates the radius $\sqrt{\Lambda_{nat}}$ 
1200: corresponding to the natural measure.}
1201: \end{center}
1202: \end{figure}
1203: %%%%%%%%%%%%%%%%%%%%
1204: In the proofs for upper bounds of the Weyl law \eqref{e:Weyl-law}, 
1205: the exponent $d$ is defined in terms of the 
1206: {\em  upper Minkowski dimension} of the trapped set $K$ \cite{Sjo90,Zw99,GLZ04,SjoZw05}. 
1207: In the case of the open baker $B_{\vr}$, we therefore expect that
1208: the exponent $d$ appearing in the conjecture (resp. $d+1$) is given by the
1209: Minkowski dimension of the Cantor set $\cC_{\vr}$ (resp. $\Gamma_+$), which is equal to
1210: its box, Hausdorff and packing dimensions. We call this theoretical
1211: value $d_0$.
1212: 
1213: For the symmetric baker $B_{\vr_{sym}}$, the Hausdorff dimension $d_H(\Gamma_+)=d_0+1$ 
1214: happens to be equal to the Hausdorff dimension of the
1215: natural measure $\mu_{nat}$, defined by
1216: $$
1217: d_H(\mu_{nat})=\inf_{A\subset\t2,\ \mu_{nat}(A)=1} d_H(A)\,.
1218: $$
1219: For a nonsymmetric baker's map $B_{\vr}$, those two Hausdorff dimensions 
1220: only satisfy the inequality
1221: $d_H(\mu_{nat})\leq d_H(\Gamma_+)$ (see the explicit expressions below).
1222: We want to investigate a possible ``role'' of the natural eigenmeasure $\mu_{nat}$
1223: regarding the structure of the quantum spectrum. It is therefore legitimate to
1224: ask the following 
1225: \begin{question}\label{qu0}
1226: Is the correct exponent in the Weyl law \eqref{e:FWL}
1227: given by $d_I=d_H(\mu_{nat})-1$ instead of $d_0=d_H(\Gamma_+)-1$?
1228: \end{question}
1229: Here the suffix $I$ indicates that $d_I$ is sometimes called
1230: the {\em information dimension} \cite{HJPPS86}.
1231: As mentioned above, both dimensions are equal for the symmetric baker $B_{\vr_{sym}}$,
1232: for which the Weyl law \eqref{conj:FWL} has been numerically tested in \cite{NonZw05}.
1233: They are also equal in the case of a closed map on $\t2$:
1234: in that case, the Weyl law has exponent $1$ (the whole spectrum
1235: lies on the unit circle), and we have $d_H(\t2)=d_H(\mu_{L})=2$. 
1236: 
1237: For a nonsymmetric baker $B_{\vr}$, the two dimensions
1238: take different values:
1239: $$
1240: d_0\quad \text{is the solution of}\quad r_0^d+r_2^d=1,\qquad\text{while}\quad
1241: d_I=\frac{r_0\log p_0+r_2\log p_2}{r_0\log r_0+r_2\log r_2}\,.
1242: $$
1243: To answer the above question, we
1244: considered a very asymmetric baker, taking $\vr_{asym}$ with $r_0=1/32$, $r_2=2/3$.
1245: The two dimensions then
1246: take the values $d_0\approx 0.493$, $d_I\approx 0.337$. We computed the counting function
1247: $n(N,r)$ for several radii $0.1\leq r\leq 1$ and several values of $N$. We then tried
1248: to fit the Weyl law \eqref{e:FWL} with an exponent $d$ varying in a certain range,
1249: and computed the standard deviations (see Fig.~\ref{f:fit}). 
1250: The numerical result is unambiguous: the best fit clearly occurs away from $d_I$, 
1251: but it is close to $d_0$. This numerical test rules out the possibility 
1252: that $d_I$ provides the correct exponent of the Weyl law, and suggests to 
1253: indeed take $d=d_0$.
1254: 
1255: To further illustrate the Weyl law for the asymmetric baker $B_{\vr_{asym}}$,  
1256: we plot in Fig.~\ref{f:asymB} (top) the counting functions
1257: $n(N,r)$ as a function of $r\in(0,1)$, for several values of $N$.
1258: On the bottom plot, we rescale $n(N,r)$ by the power $N^{-{d_0}}$: 
1259: the rescaled curves almost perfectly
1260: overlap, indicating that the scaling \eqref{e:FWL} is correct.
1261: \begin{rem}\label{r:fill}
1262: On figure~\ref{f:asymB} (right) the rescaled 
1263: counting function seems to converge to a function which is strictly decreasing
1264: on an interval $[\lambda_{\min},\lambda_{\max}]$, where $\lambda_{\min}\approx 0.1$, 
1265: $\lambda_{\max}\approx 0.9$. This implies that the spectrum of $B_{\vr_{asym},N}$ becomes
1266: dense in the whole annulus $\set{\lambda_{\min}\leq|\lambda|\leq\lambda_{\max}}$, 
1267: when $N\to\infty$. 
1268: Therefore, at this heuristic level, 
1269: for any $\lambda\in [\lambda_{\min},\lambda_{\max}]$
1270: one may consider sequences of 
1271: eigenvalues $(\lambda_N)_{N\geq 1}$ with the property $|\lambda_N|\Nto8 \lambda$.
1272: In particular, we may consider sequences converging to $\sqrt{\Lambda_{nat}}$.
1273: \end{rem}
1274: 
1275: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1276: \section{Localization of resonant eigenstates}\label{s:localization}
1277: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1278: 
1279: We now return to the general framework of an open map $T$
1280: on a subset $M$ of a compact phase space $\hM$, which is
1281: quantized into a sequence of operators $(T_N)_{N\to\infty}$ according to Axioms~\ref{d:axiom}.
1282: We want to study the semiclassical measures associated with the long-living eigenstates of $T_N$.
1283: 
1284: To start with, we fix some $\lambda_m\in(0,1)$, such that for $N$ large enough, 
1285: $\Spec T_N\cap \set{|\lambda|\geq \lambda_m}\neq \emptyset$. We
1286: can then consider sequences of eigenstates $(\psi_N)_{N\to\infty}$ such that, for $N$
1287: large enough, 
1288: \bequ\label{e:eigenstates}
1289: T_N\,\psi_N=\lambda_N\,\psi_N\,,\qquad \norm{\psi_N}=1\,,\qquad 
1290: |\lambda_N|\geq \lambda_{m}\,.
1291: \eequ
1292: The role of the (quite arbitrary) lower bound $\lambda_{m}>0$ is to ensure 
1293: that the eigenstates we consider are ``long-living''.
1294: Up to extracting a subsequence, we can assume that
1295: $(\psi_N)$ converges to a certain semiclassical measure $\mu$ 
1296: (see \S\ref{s:semicl-measure}). 
1297: 
1298: \medskip
1299: 
1300: To state our result, we first
1301: we need to analyze the continuity sets of the backward iterates of $T$. 
1302: For any $n\geq 1$, the map $T^{-n}$ is defined on the set $M^{-n}$. Its 
1303: continuity set $C(T^{-n})$ can be obtained iteratively through 
1304: $$
1305: C(T^{-n})=T\big(C(T^{-n+1})\cap C(T)\big)\,,\quad n\geq 2\,.
1306: $$
1307: The set $M_{-n}$ of points escaping exactly at 
1308: time $(-n)$ can also be split between its continuity subset
1309: its continuity subset
1310: $$
1311: CM_{-n}\defeq C(T^{-n+1})\cap \ring{M}_{-n}=T^{n-1}(C(T^{n-1})\cap \ring{H}^{-1})\,,
1312: $$
1313: which is a union of open connected sets, and its discontinuity subset
1314: $DM_{-n}=M_{-n}\setminus CM_{-n}$, which has Minkowski content zero (in the case $n=1$,
1315: we take $CM_{-1}=\ring{H}^{-1}$).
1316: Using this splitting of the $M_{-n}$, the decomposition \eqref{e:split2} can be recast into:
1317: \bequ\label{e:decompo-final}
1318: \hM=C_{-\infty}\sqcup D_{-\infty} \sqcup\Gamma_+,\quad\text{where}\quad
1319: C_{-\infty}=\big(\bigsqcup_{n=1}^\infty CM_{-n}\big),\quad
1320: D_{-\infty}=\big(\bigsqcup_{n=1}^\infty DM_{-n}\big)\,.
1321: \eequ
1322: The set $C_{-\infty}$ consists of points
1323: which eventually fall in the hole when evolved through $T^{-1}$, 
1324: and remain at finite distance from $D(T^{-1})$ all along their transient trajectory.
1325: 
1326: We can now state our main result concerning semiclassical measures.
1327: %%%%%%%%%%%%%
1328: \begin{thm}\label{thm:main}
1329: Assume that a sequence of eigenstates $(\psi_N)_{N\to\infty}$ of the 
1330: open quantum map $T_N$, with eigenvalues $|\lambda_N|\geq\lambda_m>0$,
1331: converges to the semiclassical measure $\mu$ on $\hM$. Then the following hold.
1332: 
1333: i) the support of $\mu$ is a subset of $\Gamma_+\sqcup D_{-\infty}$.
1334: 
1335: ii) If $\mu\big(C(T^{-1})\big)>0$, there exists $ \Lambda\in [\lambda_{m}^2,1]$ such
1336: that the eigenvalues $(\lambda_N)_{N\to\infty}$ satisfy
1337: $$
1338: |\lambda_N|^2\xrightarrow{N\to\infty} \Lambda\,.
1339: $$ 
1340: For any Borel subset $S$ not intersecting 
1341: $D(T^{-1})$, one has
1342: $\mu(T^{-1}(S))=\Lambda\,\mu(S)$. 
1343: 
1344: iii) If $\mu(D(T^{-1}))=\mu\big(T^{-1}(D(T^{-1}))\big)=0$, 
1345: then $\mu$ is an eigenmeasure of $T$, with decay rate $\Lambda$. 
1346: \end{thm}
1347: %%%%%%%%%%%%%
1348: \begin{proof}
1349: To prove the first statement, let us choose some $n\geq 1$, and take an observable
1350: $a\in C^\infty_c(CM_{-n})$. Every point $x\in\supp a$ has the property that
1351: for any $0\leq j <n-1$, $T^{-j}(x)\in C(T^{-1})$, while $T^{-n+1}(x)\in \ring{H}^{-1}$.
1352: Applying iteratively the property \eqref{e:axiom2-map}, one finds that in the 
1353: semiclassical limit,
1354: $$
1355: (T_N^{\dagger})^{n}\,\Op_N(a)\,(T_N)^{n}\sim 0\,.
1356: $$
1357: We take into account the fact that $\psi_N$ is a right eigenstate of $T_N$:
1358: $$
1359: \la \psi_N,\,\Op_N(a)\,\psi_N\ra=
1360: |\lambda_N|^{-2n}\,\la \psi_N,(T_N^{\dagger})^{n}\,\Op_N(a)\,(T_N)^{n}\,\psi_N\ra\,.
1361: $$
1362: Using $|\lambda_N|\geq \lambda_m$, these two expressions imply
1363: $\mu(a)\defeq \int a\,d\mu=0$.
1364: Since $n$ and $a$ were arbitrary, this shows $\mu(C_{-\infty})=0$, which is the first statement.
1365: 
1366: From the assumption in $ii)$, we may select $a\in C^\infty_c(C(T^{-1}))$ 
1367: such that $\int a\,d\mu>0$. The first iterate $a\circ T$ is supported in $C(T)\subset M$.
1368: Applying \eqref{e:axiom2-map}, we get
1369: \bequ\label{e:eigenstate}
1370: |\lambda_N|^{2}\,\la\psi_N,\,\Op_N(a)\,\psi_N\ra \sim 
1371: \la \psi_N,\,\Op_N(a\circ T)\,\psi_N\ra\,.
1372: \eequ
1373: For $N$ large enough, the matrix element
1374: $\la\psi_N,\,\Op_N(a)\,\psi_N\ra > \mu(a)/2$, so we may divide the
1375: above equation by this element, and obtain
1376: $$
1377: |\lambda_N|^2\Nto8 \frac{\mu(a\circ T)}{\mu(a)}\,.
1378: $$
1379: We call $\Lambda\geq |\lambda_m|^2$ this limit, which is obviously 
1380: independent of the choice of $a$. For any Borel set
1381: $S\subset C(T^{-1})$, one can approximate $\bbbone_{S}$ by smooth functions supported
1382: in $C(T^{-1})$, to prove that 
1383: $$
1384: \mu(T^{-1}(S))=\Lambda\,\mu(S)\,.
1385: $$
1386: If $S\subset \ring{H}^{-1}$, we know that $\mu(S)=0$ and $T^{-1}(S)=\emptyset$ ,
1387: so the above equality still makes sense. This proves $ii)$.
1388: 
1389: To obtain $iii)$, we
1390: split any Borel set $S$ into $S=(S\cap D(T^{-1}))\sqcup CS$.
1391: From $ii)$, we have $\mu(T^{-1}(CS))=\Lambda\,\mu(CS)$.
1392: By assumption, $\mu(S\cap D(T^{-1}))=\mu(T^{-1}(S\cap D(T^{-1})))=0$, so
1393: we get $\mu(T^{-1}(S))=\Lambda\,\mu(S)$.
1394: \end{proof}
1395: 
1396: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1397: \subsection{Semiclassical measures of the open baker's map} \label{s:semiclass}
1398: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1399: 
1400: In this section we apply the above theorem to the case of some open baker's map $B_{\vr}$,
1401: quantized as in \S\ref{s:quantum-baker2}. We also numerically compute the Husimi
1402: measures $H^1_{\psi_N}$ associated with some quantum eigenstates (we choose the isotropic
1403: squeezing $\ssigma=1$ by convenience): although we cannot really go to the semiclassical 
1404: limit, we hope that for $N\sim 1000$ these Husimi measures already give some idea of the
1405: semiclassical measures.
1406: 
1407: \subsubsection{Applying Theorem~\ref{thm:main}}
1408: 
1409: To simplify the presentation we will restrict the discussion to the symmetric baker 
1410: $B_{\vr_{sym}}$, which 
1411: will be denoted by $B$ in short. The discontinuity set $D(B^{-1})$, its backwards image 
1412: and the trapped sets were described in \S\ref{s:baker-trapped} and \S\ref{s:quantum-baker2}, and
1413: plotted in Fig.~\ref{f:gamma+}.
1414: The sets $M_{-n}$ and their continuity subsets are also simple to describe using symbolic dynamics.
1415: For any $n=1$, we know that $CM_{-1}=\ring{H}^{-1}$ is the open rectangle 
1416: $\set{q\in [0,1),\ p\in (1/3,2/3)}$. 
1417: For $n\geq 2$, 
1418: $CM_{-n}$ is the union of $2^{n-1}$ horizontal rectangles, indexed by the $n$-sequences 
1419: $\bep=\ep_{-1}\ldots\ep_{-n}$ such that $\ep_{-i}\in\set{0,2}$, $i=1,\ldots,n-1$, while
1420: $\ep_{-n}=1$. Each such rectangle is
1421: of the form $\set{q\in (0,1),\ p\in (p(\bep), p(\bep)+3^{-n})}$, where 
1422: $p(\bep)\equiv \cdot\,\ep_{-1}\ldots\ep_{n}$ in ternary decomposition. 
1423: Some of those  rectangles are shown in Fig.~\ref{f:Cantor} (center).
1424: The union of all these rectangles (for $n\geq 1$) makes up
1425: $C_{-\infty}$. Its complement $\Gamma_+\sqcup D_{-\infty}$ is given by
1426: $$
1427: \Gamma_+\sqcup D_{-\infty}=\big([0,1)\times \cC_{\vr_{sym}}\big)\cup D(B^{-1})\,.
1428: $$
1429: This set is shown in Fig.~\ref{f:gamma+} (left).
1430: %%%%%%%%%%%%%%%%%%%%
1431: \begin{figure}[htbp]
1432: \begin{center}
1433: \rotatebox{-90}{\includegraphics[width=.33\textwidth]{3states-nat.ps}}
1434: %\rotatebox{-90}{\includegraphics[width=6cm]{3states-sym.ps}}
1435: \rotatebox{-90}{\includegraphics[width=.33\textwidth]{3states-2maps.ps}}
1436: %\rotatebox{-90}{\includegraphics[width=6cm]{2states9-3.ps}}
1437: \caption{\label{f:eigenstates} Husimi densities of (right) 
1438: eigenstates of $B_{\vr,N}$ (black=large values, white=0).
1439: Top: 3 eigenstates of $B_{\vr_{sym},N}$ with $|\lambda_N|\approx \sqrt{\Lambda_{nat}}$.
1440: Bottom left, center: two eigenstates of $B_{\vr_{sym},N}$ with different $|\lambda_N|$. 
1441: Bottom right: one eigenstate of $B_{\vr_2,N}$.}
1442: \end{center}
1443: \end{figure}
1444: %%%%%%%%%%%%%%%%%%%%
1445: 
1446: In Figure~\ref{f:eigenstates}, we plot the Husimi densities $H^1_{\psi_N}$ of some
1447: (right) eigenstates of $B_{\vr,N}$, for the symmetric baker and 
1448: an asymmetric one, $\vr_{2}=(1/9,5/9,1/3)$.
1449: 
1450: \begin{rem}
1451: We notice that all Husimi functions are indeed
1452: very small in the horizontal rectangles $M_{-n}$ for $n=1,2$ (in case $N\leq 1500$),
1453: and also $n=3$ for $N=4200$. Using \eqref{e:CS-evol}, 
1454: one can refine the proof of Thm.~\ref{thm:main} to show that $H^1_{\psi_N}(x)=\cO(\e^{-cN})$ 
1455: for $x\in C_{-\infty}$, $N\to\infty$. 
1456: \end{rem}
1457: 
1458: \begin{rem}\label{r:rem2}
1459: Although this is not proven in our theorem, all the Husimi functions we have computed
1460: are very small on the set $D(B^{-1})\setminus \Gamma_+\subset \set{q=0}$
1461: (see the left plot in Fig.~\ref{f:gamma+}). 
1462: On the other hand, some of these Husimi functions are large on
1463: $D(B^{-1})\cap \Gamma_+$,
1464: so we cannot rule out the possibility that 
1465: some semiclassical measures $\mu$ charge this set. 
1466: For instance, in some of the Husimi plots ($e.g.$
1467: the bottom left in Fig.~\ref{f:eigenstates}), we clearly see
1468: a strong peak at the origin, and
1469: lower peaks on other points of $D(B^{-1})\cap \Gamma_+$.
1470: We call ``diffractive'' the components of an eigenstate localized on $D_{-\infty}$. 
1471: \end{rem}
1472: 
1473: From these observations, we state the following
1474: \begin{conj} \label{c:eigenm}
1475: Let $B_{\vr}$ be an open baker's map, quantized by \S\ref{s:quantum-baker2}. Then,
1476: \begin{itemize} 
1477: \item all long-living semiclassical measures are supported on 
1478: $\overline{\Gamma_+}$
1479: \item ``almost all'' long-living eigenstates are non-diffractive, 
1480: that is, their weight
1481: on $D(B^{-1})$ and $B^{-1}(D(B^{-1}))$ are negligible
1482: in the semiclassical limit. 
1483: The corresponding semiclassical measures are then eigenmeasures of $B_{\vr}$.
1484: \end{itemize}
1485: \end{conj}
1486: In the next section we further comment on the above Husimi plots.
1487: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1488: \subsection{Abundance of semiclassical measures}\label{s:abundance} 
1489: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1490: 
1491: Let us draw some consequences from Theorem~\ref{thm:main} and the following remarks.
1492: Statement $ii)$ of the theorem strongly constrains the 
1493: converging sequences of eigenstates: a sequence $(\psi_N)$ can
1494: converge to some measure $\mu$ (with $\mu(C(B^{-1})) > 0$) only if the
1495: corresponding eigenvalues $(\lambda_N)$ asymptotically approach the circle of radius 
1496: $\sqrt{\Lambda}$. 
1497: 
1498: We take for granted the density argument in Remark~\ref{r:fill} and assume that
1499: a ``dense interval'' $[\lambda_{\min},\lambda_{\max}]\subset (0,1]$
1500: exists for any open baker's map $B_{\vr}$. Hence,
1501: for any $\Lambda\in [\lambda_{\min}^2,\lambda_{\max}^2]$ there exist
1502: many sequences of eigenstates of $B_{\vr,N}$, 
1503: such that $|\lambda_N|^2\Nto8 \Lambda$. From our Conjecture~\ref{c:eigenm}, almost any
1504: semiclassical measure associated with such a sequence will be an eigenmeasure with
1505: decay rate $\Lambda$. Therefore, two converging sequences $(\psi_N)$, $(\psi_N')$
1506: associated with limiting decays $\Lambda\neq\Lambda'$ will necessarily converge
1507: to different eigenmeasures. This already shows that the semiclassical eigenmeasures
1508: generated by all possible sequences in the annulus 
1509: $\set{\lambda_{\min}\leq|\lambda|\leq\lambda_{\max}}$ form an uncountable family.
1510: 
1511: According to \S\ref{s:eigenmeasures}, for each decay rate $\Lambda\in (0,1)$,
1512: there exist uncountably many eigenmeasures. A natural question thus concerns the
1513: variety of semiclassical measures associated with a given 
1514: $\Lambda\in [\lambda_{\min}^2,\lambda_{\max}^2]$:
1515: 
1516: \begin{question}\label{qu1}$\ $
1517: 
1518: For a given $\Lambda\in [\lambda_{\min}^2,\lambda_{\max}^2]$, what are the
1519: semiclassical measures of $B_{\vr}$ of decay rate $\Lambda$?
1520: \begin{itemize}
1521: \item is there a unique such measure? 
1522: \item otherwise, is some limit measure  ``favored'', in the sense that ``almost all'' sequences
1523: $(\psi_N)$ with $|\lambda_N|^2\to \Lambda$ converge to $\mu$?
1524: \item can the {\em natural measure} $\mu_{nat}$ be obtained as a semiclassical measure?
1525: \end{itemize}
1526: \end{question}
1527: The same type of questions were asked by Keating and coworkers in ref.~\cite{KNPS06}. 
1528: At present we are unable to answer them rigorously for the quantum open baker $B_{\vr,N}$. 
1529: 
1530: From a heuristic point of view we notice the following features 
1531: in the plots of Fig.~\ref{f:eigenstates}. The three
1532: top plots correspond to eigenvalues $|\lambda_N|$ close to the value 
1533: $\sqrt{\Lambda_{nat}}\approx 0.8165$. However, the Husimi measures on the left and center
1534: seem very different from the natural measure (the latter is approximated by the black rectangles
1535: in Fig.~\ref{f:gamma+}). These two Husimi functions seem to charge different 
1536: parts of $\Gamma_+$. Only the rightmost state seems compatible with a convergence 
1537: to $\mu_{nat}$.
1538: 
1539: As commented in Remark~\ref{r:rem2}, the bottom-left state has strong concentrations on
1540: $D(B^{-1})\cap\Gamma_+$. For the various values of $N$ we have investigated, this 
1541: concentration seems characteristic of the eigenstate with the largest 
1542: $|\lambda_N|$. The discontinuities of $B_{\vr}$ manage to ``trap'' those quantum states 
1543: better than periodic orbits on $C(B^{-1})\cap\Gamma_+$. Such states seem ``very diffractive''.
1544: 
1545: Finally, the bottom-center state, with a smaller eigenvalue, clearly
1546: shows a selfsimilar structure along the horizontal direction, with a
1547: probability inside the hole higher than for the top eigenstates. 
1548: This feature had already been noticed in \cite{KNPS06} for averages over Husimi 
1549: functions with comparable $|\lambda_N|$.
1550: 
1551: In section~\ref{s:walsh} we will address Questions~\ref{qu1} for
1552: a different quantization of the symmetric open baker, namely 
1553: the Walsh-quantized open baker, where one can compute some semiclassical
1554: measures explicitly.
1555: 
1556: Before that, we explain how to construct approximate eigenstates
1557: ({\em pseudomodes}) for the quantum baker $B_{\vr_{sym},N}$.
1558: 
1559: %%%%%%%%%%%%%%%%%%%%%%%%%%
1560: \subsection{Pseusomodes and pseudospectrum} 
1561: %%%%%%%%%%%%%%%%%%%%%%%%%%
1562: 
1563: From the explicit representation of eigenmeasures given in Proposition~\ref{p:eigen-decompo},
1564: it is possible to construct approximate eigenstates of $T_N$ by 
1565: backward propagating wavepackets localized on the set $\Gamma_+^{(1)}=\Gamma_+\cap H$.
1566: 
1567: We call these approximate eigenstates 
1568: {\it pseudomodes}, by analogy with the recent literature on nonselfadjoint semiclassical
1569: operators (see e.g. \cite{DSZ04,BU03}). Those papers deal with pseudodifferential operators
1570: obtained by quantizing complex-valued observables, and they construct
1571: pseudomodes of error $\cO(\hbar^\infty)$, microlocalized at a single phase space point.
1572: 
1573: Our pseudomodes will be less precise and will be microlocalized on a countable set. 
1574: We will restrict ourselves to the case of the
1575: open symmetric baker $B=B_{\vr_{sym}}$, but our construction works for 
1576: any $B_{\vr}$, and can probably be extended to other maps or phase spaces.
1577: 
1578: Inspired by the pointwise eigenmeasures \eqref{e:mu-traject}, we construct 
1579: approximate eigenstates by backwards evolving coherent states localized in $\Gamma_+^{(1)}$.
1580: Precisely, for any $\bx_0\in \Gamma_+^{(1)}$, $\ssigma>0$ and $\lambda\in \IC$, $|\lambda|<1$, 
1581: $N\in\IN$, we define the following quantum state:
1582: \begin{equation}\label{e:quasim}
1583: \Psi^\ssigma_{\lambda,x_0}\defeq \sqrt{1-|\lambda|^2}\;
1584: \sum_{n\geq 0}\lambda^n\,B_{N}^{\dagger n}\,\psi_{\bx_0,\ssigma}\,,
1585: \end{equation}
1586: where $\psi_{\bx_0,\ssigma}$ is the coherent state defined in \S\ref{s:Husimi}, and $B_N$ is
1587: the standard quantization of $B$.
1588: %++++++++++++++++
1589: \begin{prop}
1590: Consider the symmetric open baker $B=B_{\vr_{sym}}$ and its quantization $(B_N)$.
1591: Fix $\lambda\in\IC$ with $|\lambda|<1$. 
1592: 
1593: $i)$ Choose $\vareps>0$ small and call $\alpha=(1-\vareps)\frac{\log 1/|\lambda|}{\log 3}$. 
1594: For any $N\in\IN^*$, 
1595: one can choose a point $x_0(N)\in \Gamma_+^{(1)}$ and a squeezing 
1596: parameter $\ssigma=\ssigma(N)>0$ such that the 
1597: states\\ $\big(\Psi_N\defi \Psi^{\ssigma(N)}_{\lambda,x_0(N)}\big)_{N\to\infty}$ defined
1598: in \eqref{e:quasim} satisfy
1599: $$
1600: \norm{\Psi_N}=1+\cO(N^{-\alpha}),\qquad 
1601: \norm{(B_N-\lambda)\Psi_N}=\cO(N^{-\alpha})\,,\qquad N\to\infty.
1602: $$
1603: $ii)$ for any $x_0\in \Gamma_+^{(1)}$, one can 
1604: select the points $\big(x_0(N)\big)$ such that
1605: $x_0(N)\Nto8 x_0$. In that case, the sequence $(\Psi_N)_{N\to\infty}$
1606: converges to the eigenmeasure $\mu_{x_0,\Lambda}$ described in \eqref{e:mu-traject},
1607: with $\Lambda=|\lambda|^2$.
1608: \end{prop}
1609: %++++++++++++++++
1610: This proposition shows that, for
1611: any $\alpha>0$ and $N$ large enough,
1612: the $N^{-\alpha}$-pseudospectrum of $B_N$ contains the disk 
1613: $\set{|\lambda|\leq 3^{-\alpha(1+\vareps)}}$.
1614: We notice that the errors are not very small, and increase when $|\lambda|\to 1$. We should
1615: emphasize that, in our numerical trials, we never found eigenstates of $B_N$ with
1616: Husimi measures looking like $\mu_{x_0,\Lambda}$.
1617: 
1618: \subsubsection*{Proof of the proposition}
1619: To control the series \eqref{e:quasim}, 
1620: we would like to ensure that the evolved
1621: coherent state $B_{N}^{\dagger j}\,\psi_{x_0(N),\ssigma(N)}$
1622: remains close to an approximate coherent state 
1623: $\psi_{x_{-j},\ssigma_{-j}}$ (as in \eqref{e:CS-evol})
1624: up to large times $j$. 
1625: For this we need the points $x_{-j}=B^{-j}(x_0(N))$ to stay ``far'' 
1626: from the discontinuity set
1627: $D(B^{-1})$, and we also need all $\psi_{x_{-j},\ssigma_{-j}}$ to be microlocalized
1628: in a small neighbourhood of $x_{-j}$. 
1629: Due to the hyperbolicity of $B^{-1}$, the second condition constrains
1630: the times $j$ to be smaller than the {\it Ehrenfest time} \cite{DENW06}
1631: \bequ\label{e:Ehrenf}
1632: n_E=(1-\vareps)\frac{\log N}{\log 3}\,.
1633: \eequ
1634: Here $\vareps>0$ is the small parameter in the statement of the proposition.
1635: 
1636: To identify a good ``starting point'' $x_0(N)$, we set $n=[n_E]$, 
1637: and consider the following subset of $\t2$, for $\delta,\gamma>0$:
1638: $$
1639: \cD_{n,\delta,\gamma}\defi\set{(q,p)\in\t2,\ q\in (1/3+\delta,2/3-\delta),\ 
1640: \forall k\in\IZ,\ |p-\frac{k}{3^n}|>\gamma}\cap\Gamma^{(1)}_+\,.
1641: $$
1642: We first show that this set is nonempty if $\delta<1/9$ and 
1643: $\gamma=3^{-n}\gamma'$, $\gamma'<1/9$.
1644: Take $q\in (1/3+\delta,2/3-\delta)$ arbitary, and select $p=p(\bep)$ with 
1645: the following properties: take all indices $\ep_{-j}\in\{0,2\}$, $j\geq 1$, 
1646: so that $p\in \cC_{\vr_{sym}}$,
1647: and require furthermore that the word $\ep_{-n-1}\ep_{-n-2}\in\set{02,20}$. 
1648: The point $(q,p)$ is then in $\cD_{n,\delta,\gamma}$. 
1649: 
1650: Let us take any point $x_0(N)$ in that set. It automatically lies in 
1651: $C(B^{-n})$, so its backwards iterates stay away from 
1652: $D(B^{-1})$ at least until the time $n$.
1653: Let us select the squeezing parameter
1654: $\ssigma(N)=s_0=N^{-1+\vareps}$. A simple adaptation
1655: of \cite[Prop.5]{DENW06} implies the following estimate:
1656: \bequ\label{e:semiclass1}
1657: \exists c,C>0,\ \ \forall j,\  0\leq j\leq n_E,\qquad
1658: \norm{B_{N}^{\dagger j}\,\psi_{\bx_0,\ssigma_0}-\e^{\i\theta_j}\,\psi_{x_{-j},\ssigma_{-j}}}\leq
1659: C\,\e^{-c\,N^\vareps}\,.
1660: \eequ
1661: Here we can take $c=\min(\delta^2,\gamma^{\prime 2})$, and $C$
1662: is uniform w.r.to the initial point $x_0$.
1663: From the values of $x_0$, $\ssigma_0$, one checks that 
1664: the components 
1665: $\la \bq_k,\psi_{x_0,\ssigma_0}\ra$ for $k/N\not\in [1/3,2/3]$ 
1666: are of order $\cO(\e^{-cN^\vareps})$: 
1667: that state is (very) localized in the hole $H=R_1$, so
1668: $$
1669: B_{N}\,\psi_{\bx_{0},\ssigma_{0}}= \cO(\e^{-c\,N^\vareps})\,.
1670: $$
1671: Similarly,
1672: for any $j\leq n_E$, the state $\psi_{x_{-j},\ssigma_{-j}}$ is localized inside a certain
1673: connected component of $M_{j+1}$ (one of the the pink/grey rectangles in Fig.~\ref{f:Cantor}, left). 
1674: In particular, the components of 
1675: $\la \bq_k,\psi_{x_{-j+1},\ssigma_{-j+1}}\ra$ are exponentially small in $H$. Therefore,
1676: $$
1677: \forall j,\  0\leq j\leq n_E,\qquad 
1678: B_{N}\,B^\dagger_{N}\psi_{\bx_{-j},\ssigma_{-j}}= 
1679: \psi_{\bx_{-j},\ssigma_{-j}}+\cO(\e^{-c\,N^\vareps})\,.
1680: $$
1681: Summing all terms $j\leq n_E$ in \eqref{e:quasim} 
1682: and estimating the remaining series using $\norm{B_N}\leq 1$, we obtain
1683: $$
1684: \norm{(B_{N}-\lambda)\,\Psi^\ssigma_{\lambda,x_0}}= \cO(|\lambda|^{n_E})\,.
1685: $$
1686: From the definition of $n_E$, we have $|\lambda|^{n_E}=N^{-\alpha}$.
1687: 
1688: The asymptotic normalization of $\Psi^\ssigma_{\lambda,x_0}$ is proven by estimating the
1689: overlaps between coherent states $\psi_{x_{-j},\ssigma_{-j}},\psi_{x_{-j'},\ssigma_{-j'}}$
1690: for $j,j'\leq n_E$. Because the sets $M_{j+1}$
1691: and $M_{j'+1}$ are disjoint for $j\neq j'$, the above mentioned localization
1692: properties imply that
1693: $$
1694: \forall j,j'\leq n_E,\qquad
1695: \la \psi_{x_{-j},\ssigma_{-j}},\psi_{x_{-j'},\ssigma_{-j'}}\ra=\delta_{j',j}+\cO(\e^{-c\,N^\vareps}),
1696: $$
1697: and the normalization estimate follows.
1698: This achieves the proof of $i)$.
1699: 
1700: To prove $ii)$, let us consider an arbitrary $x_0=(q_0,p_0)\in \Gamma^{(1)}_+$. 
1701: If $q_0\not\in\set{1/3,2/3}$,
1702: we take $\delta$ small enough such that $q_0\in (1/3+\delta,2/3-\delta)$ and set $q_0(N)=q_0$.
1703: Otherwise, we may let $\delta$ slowly decrease with $N$, and find 
1704: a sequence $\big(q_0(N)\big)$ such
1705: that $q_0(N)\in (1/3+\delta(N),2/3-\delta(N))$ and $q_0(N)\Nto8 q_0$.
1706: On the other hand, 
1707: $p_0\in\cC_{\vr_{sym}}=p(\bep)$ for a certain sequence $\bep$, with all 
1708: $\ep_{-j}\in\set{0,2}$. For each $N$, we
1709: inspect the word $\ep_{-n-1}\ep_{-n-2}$ of that sequence, where $n=[n_E]$ (see \eqref{e:Ehrenf}). 
1710: If this word is in the set $\set{02,20}$ we keep $p_0(N)=p_0$, otherwise we 
1711: replace this word by $02$ (keeping the other symbols unchanged) 
1712: to define $p_0(N)$. The point $x_0(N)=(q_0(N),p_0(N))$ is then in 
1713: $\cD_{n,\delta,\gamma}$, and 
1714: $x_0(N)\Nto8 x_0$.
1715: 
1716: The convergence to the measure $\mu_{x_0,\Lambda}$ is due to
1717: the localization of the coherent states $\psi_{x_{-j}, s_{-j}}$ for $j\leq n_E$.
1718: 
1719: $\hfill\square$
1720: 
1721: 
1722: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1723: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1724: \section{A solvable toy model: Walsh quantization of the open baker}\label{s:walsh}
1725: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1726: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1727: In this section we study an alternative quantization of the open symmetric
1728: baker $B_{\vr_{sym}}$, introduced in \cite{NonZw05,NonZw06}. 
1729: A similar quantization
1730: of the closed baker $\hB_{\vr_{sym}}$ was proposed and 
1731: studied in \cite{SchaCav00,TraSco02,ErmSara06},
1732: mainly motivated by research in quantum computation.
1733: From now on, we will drop the
1734: index $\vr_{sym}$ from our notations, and call $B=B_{\vr_{sym}}$.
1735: 
1736: A ``simplified'' quantization of $B$ was introduced in \cite{NonZw05,NonZw06}, 
1737: as a $N\times N$ matrix $\tB_N$, obtained by only keeping the ``skeleton''  
1738: of $B_N$ (see \eqref{e:B_N}). Although $\tB_N$ can be defined for any 
1739: $N$, its spectrum can be explicitly computed only 
1740: if $N=3^k$ for some $k\in\IN$.
1741: As explained in those references, $\tB_N$ can be interpreted as
1742: the ``Weyl'' quantization of a  
1743: multivalued map $\tB$ built upon $B$. In the present work we will stick to a different
1744: interpretation of $\tB_N$, valid in the case $N=3^k$: one can then introduce
1745: a (Walsh) quantization for observables on $\t2$, which is not equivalent with
1746: the anti-Wick quantization of \S\ref{s:Husimi}. We will check below that
1747: the matrices $\tB_N$ satisfies property \eqref{e:axiom2-map} with respect to that
1748: quantization and the open baker $B$. 
1749: Finally, this Walsh
1750: quantization is also suited to quantize observables on
1751: the symbol space $\Sigma$: the matrices $\tB_N$ are then quantizing the
1752: open shift $\sigma$. We will see that this interpretation is more ``convenient'', because
1753: it avoids problems due to discontinuities. 
1754: 
1755: We first recall the definition of the
1756: Walsh quantization of observables on $\t2$ (or $\Sigma$), and the associated
1757: Walsh-Husimi measures.
1758: 
1759: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1760: \subsection{Walsh transform and coherent states} \label{s:walsh0}
1761: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1762: 
1763: \subsubsection{Walsh coherent states}
1764: The Walsh quantization of observables on $\t2$ uses the decomposition of the quantum
1765: Hilbert space $\hn$ into a tensor product of ``qubits'', namely
1766: $\hn=(\IC^3)^{\otimes k}$ (clearly, this makes sense only for $N=3^k$). 
1767: Each discrete position $q_j=j/N$ can be represented by its ternary sequence
1768: $q_j= 0\cdot\, \ep_0\ep_2\cdots\ep_{k-1}$, with symbols $\ep_i\in\set{0,1,2}$. 
1769: Accordingly, each position eigenstate $\bq_j$ can be represented as a tensor product:
1770: $$
1771: \bq_j=e_{\ep_0}\otimes e_{\ep_2}\otimes\ldots\otimes e_{\ep_{k-1}}
1772: =|[\ep_0\ep_2\ldots\ep_{k-1}]_k\ra\,,
1773: $$
1774: where $\set{e_0,\,e_1,\,e_2}$ is the canonical (orthonormal) basis of $\IC^3$. The notation
1775: on the right hand side emphasizes the fact that this state is 
1776: associated with the cylinder $[\bep]=[\bep]_k$
1777: with $k$ symbols on the right of the comma, no symbol on the left. Its
1778: image on $\t2$ is a rectangle $[\bep]_k$ of height unity and width $3^{-k}=1/N$.
1779: 
1780: Walsh quantization consists in replacing the discrete Fourier transform
1781: \eqref{e:DFT} on $\hn$ by the Walsh(-Fourier) transform $W_{N}$, which is a unitary operator
1782: preserving the tensor product structure of $\hn$. 
1783: We define it through its inverse $W_{N}^*$, which maps the position basis to the orthonormal basis 
1784: of ``Walsh momentum states'': for any $j\equiv \ep_0\ldots\ep_{k-1}$, 
1785: $$
1786: \bp_j= W_{N}^*\,\bq_j\defeq F^*_3 e_{\ep_{k-1}}\otimes F^*_3 e_{\ep_{k-2}}\otimes\ldots
1787: \otimes F^*_3 e_{\ep_1}\otimes F^*_3 e_{\ep_0}
1788: $$
1789: (here $F_3^*$ is the inverse Fourier transform on $\IC^3$).
1790: To agree with our notations of \S\ref{s:open-baker}, we will index the symbols relative to the
1791: momentum coordinate by {\em negative} integers, so the Walsh momentum states will be denoted by
1792: $$
1793: \bp_j=|[\ep_{-k}\ldots\ep_{-1}]_0\ra\,,\quad\text{where}\quad 
1794: p_j=j/N=0 \cdot\,\ep_{-1}\ldots \ep_{-k}\,.
1795: $$
1796: This momentum state is associated with a rectangle of
1797: height $3^{-k}$ and width unity. 
1798: 
1799: More generally, for any $\ell\in\set{0,\ldots,k}$ and any sequence
1800: $\bep=\ep_{-k+\ell}\ldots\ep_{-1}\cdot\ep_0\ldots\ep_{\ell-1}$, one can construct
1801: a (Walsh-)coherent state
1802: \bequ\label{e:Walsh-CS}
1803: |[\bep]_{\ell}\ra  \defeq 
1804: e_{\ep_0}\otimes\ldots e_{\ep_{\ell-1}}\otimes 
1805: F_3^* e_{\ep_{-k+\ell}}\otimes\ldots F_3^* e_{\ep_{-1}} \,.
1806: \eequ
1807: This state is localized in the rectangle $[\bep]_\ell$ with height $3^{-k+\ell}$ and
1808: width $3^{-\ell}$, so it still has area $1/N$ 
1809: (all such rectangles are minimal-uncertainty, or ``quantum'' rectangles).
1810: 
1811: Like the squeezing parameter $\ssigma$ of Gaussian wavepackets (see \S\ref{s:Husimi}),
1812: the index $\ell$ describes the aspect ratio of the coherent state. When going
1813: to the semiclassical limit $k\to\infty$, we will
1814: always select $\ell(k)\sim k/2$, which corresponds to an ``isotropic'' squeezing. 
1815: One important
1816: difference between Gaussian and Walsh coherent states lies in the fact that the latter are
1817: {\em strictly} localized both in momentum and position. Another difference is that, for
1818: each $\ell$,  
1819: the $\ell$-coherent states make up a finite orthonormal basis of $\hn$, 
1820: instead of a continuous overcomplete family.
1821: 
1822: %==========================================
1823: \subsubsection{Walsh quantization for observables on $\Sigma$}\label{s:Lipschitz}
1824: %==========================================
1825: 
1826: As explained in \cite{AN06}, it is more natural to Walsh-quantize observables
1827: on the symbol space $\Sigma$ than on $\t2$.
1828: Indeed, if one equips $\Sigma$ with the metric structure
1829: $$
1830: d_{\Sigma}(\bep,\bep')=\max(3^{-n_+},3^{-n_-}),\quad n_+=\min\set{n\geq 0,\,\ep_n\neq \ep'_n},\ 
1831: n_-=\min\set{n\geq 0,\,\ep_{-n-1}\neq \ep'_{-n-1}}\,,
1832: $$
1833: the closed shift $\hat\sigma$ then acts as a {\it Lipschitz} map on $\Sigma$, and
1834: the hole $\set{\ep_0=1}$ is at finite distance from its complement in $\Sigma$.
1835: Indeed, the ``lift'' from $\t2$ to $\Sigma$ has the effect to
1836: ``blow up'' the lines of discontinuity of $B$.
1837: 
1838: The conjugacy $J:\Sigma\to\t2$ is also Lipschitz, so
1839: any Lipschitz function $F\in Lip(\t2)$ is pushed to 
1840: a function $f=F\circ J \in Lip(\Sigma)$. However, the converse is not true:
1841: if we use the inverse map $(J_{|\Sigma'})^{-1}:\t2\to\Sigma'$ and take an arbitrary
1842: $f\in Lip(\Sigma)$, the function $F= f\circ (J_{|\Sigma'})^{-1}$
1843: is generally discontinuous on $\t2$.
1844: 
1845: Let us select some $\ell\sim k/2$. The Walsh-quantization of a function $f\in Lip(\Sigma)$ is 
1846: defined as the following operator on $\hn$:
1847: \bequ\label{e:Walsh-quantiz}
1848: \Op_N^\ell(f)\defeq 3^k \sum_{[\bep]_{\ell}} 
1849: |[\bep]_{\ell}\ra \la[\bep]_{\ell}|\;\int_{[\bep]_{\ell}}f\,d\mu_L
1850: \eequ
1851: Here the sum goes over all ``quantum'' cylinders $[\bep]_{\ell}$, that is
1852: over all $3^k$ sequences $\bep=\ep_{-k+\ell}\ldots\ep_{-1}\cdot\ep_0\ldots\ep_{\ell-1}$.
1853: The integral over $f$ is performed using the uniform Bernoulli measure $\mu_L$, which
1854: is equivalent to the Liouville measure on $\t2$ (see the end of \S\ref{s:open-baker}).
1855: Notice the formal similarity of this quantization with the anti-Wick 
1856: quantization \eqref{e:anti-Wick}.
1857: 
1858: It is shown in \cite[Prop.3.1]{AN06} that this quantization satisfies the 
1859: Axioms~\ref{d:axiom-obs} (with Lipschitz observables), and that 
1860: two quantizations $\Op_N^{\ell_1}$, $\Op_N^{\ell_2}$ are semiclassically 
1861: equivalent if both $\ell_1,\,\ell_2\sim k/2$.
1862: 
1863: By duality, we define the Walsh-Husimi measure of a quantum state $\psi_N$. The corresponding 
1864: density is constant in each $\ell$-cylinder $[\bep]_{\ell}$:
1865: \bequ\label{e:WH}
1866: \forall \bal\in[\bep]_{\ell}\,,\qquad WH^\ell_{\psi_N}(\bal)=3^k\,|\la [\bep]_\ell , \psi_N\ra|^2\,.
1867: \eequ
1868: This density is originally defined on $\Sigma$, but can be 
1869: pushed-forward to $\t2$.
1870: Semiclassical measures
1871: are defined as the weak-$*$ limits of sequences $(WH^\ell_{\psi_N})$,
1872: where $N=3^k\to\infty$, and $\ell=\ell(k)\sim k/2$. One first obtains a measure $\mu^\Sigma$ 
1873: on $\Sigma$, which can be pushed on $\t2$ into $\mu=J^*\mu^\Sigma$ (equivalently,
1874: $\mu$ is the weak-$*$ limit of the Walsh-Husimi measures on $\t2$).
1875: 
1876: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1877: \subsection{Walsh quantization of the open baker}\label{s:Walsh-baker} 
1878: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1879: 
1880: We now recall the Walsh quantization of the open baker 
1881: $B=B_{\vr_{sym}}$, as defined in \cite{NonZw05,NonZw06}.
1882: Mimicking the standard quantization \eqref{e:B_N},
1883: we replace the Fourier transforms $F_{N}^*$, $F_{N/3}$ by their Walsh analogues
1884: $W_N$, $W_{N/3}$ (with $N=3^k$, $k\geq 0$), so that the Walsh-quantized open baker
1885: is given by the following matrix in the position basis:
1886: \bequ
1887: \widetilde B_{N}\defeq W_N^*\begin{pmatrix}W_{N/3}&&\\&0&\\&&W_{N/3}\end{pmatrix}\,.
1888: \eequ 
1889: For any set of vectors $v_0,\,\ldots\, v_{k-1}\in \IC^3$, this operator acts as follows
1890: on any tensor product state $v_0\otimes\ldots\otimes v_{k-1}$:
1891: \bequ\label{e:B_k}
1892: \widetilde B_{N}\big(v_0\otimes v_1\ldots\otimes v_{k-1}\big) =
1893: v_1\otimes\ldots\otimes v_{k-1}\otimes \widetilde F_3^* v_0\,.
1894: \eequ
1895: Here $\widetilde F_3^*=F_3^*\,\pi_{02}$, where $\pi_{02}$ is the orthogonal projector on 
1896: $\IC e_0\oplus \IC e_2$ in $\IC^3$.
1897: 
1898: One can generalize the quantum-correspondence
1899: of \cite[Prop.3.2]{AN06} to the open shift $\sigma$, and prove the Walsh version of
1900: Axioms~\ref{d:axiom}:
1901: %================
1902: \begin{prop}\label{p:Walsh-Egorov}
1903: 
1904: $i)$ Take any $f\in Lip(\Sigma)$. Then, in the limit $N=3^k\to\infty$, $\ell\sim k/2$, 
1905: $$
1906: \tB_N^\dagger\,\Op^\ell_N(f)\,\tB_N \sim  
1907: \Op^{\ell}_N\big(\bbbone_{\set{\ep_0\neq 1}}\times f\circ\sigma\big)\,.
1908: $$
1909: Notice that the function 
1910: $\bbbone_{\set{\ep_0\neq 1}}\times f\circ\sigma\in Lip(\Sigma)$.
1911: 
1912: $ii)$ If we take $F\in Lip_c(C(B^{-1})\sqcup \ring{H}^{-1})$ and $f=F\circ J$, we have
1913: $$
1914: \bbbone_{\set{\ep_0\neq 1}}\times f\circ\sigma=\big(\bbbone_{M}\times F\circ B\big)\circ J\,.
1915: $$
1916: \end{prop}
1917: %================
1918: The points $i)-ii)$ show that the family $(\tB_N)$
1919: satisfies the Axioms~\ref{d:axiom} with respect to the map $B$ on $\t2$ and
1920: the quantization \eqref{e:Walsh-quantiz} of observables in $Lip(\t2)$. 
1921: The point $i)$ alone shows that $(\tB_N)$ is a quantization 
1922: of the open shift $\sigma$ on $\Sigma$. As noticed above,
1923: the latter interpretation allows to get rid of problems of discontinuities.
1924: \begin{proof}
1925: Applying \eqref{e:B_k} to a coherent state $|[\bep]_\ell\ra$, we get the exact 
1926: evolution
1927: $$
1928: \tB_N^\dagger\,|[\bep]_\ell\ra=(1-\delta_{\ep_{-1},1})\,|[\sigma^{-1}(\bep)]_{\ell+1}\ra\,.
1929: $$
1930: That is, the coherent state $|[\bep]_\ell\ra$ is either killed if $\sigma^{-1}([\bep]_\ell)=\infty$, 
1931: or transformed into a coherent state associated with the cylinder 
1932: $[\sigma^{-1}(\bep)]_{\ell+1}=\sigma^{-1}([\bep]_{\ell})$.
1933: This exact expression, which is the quantum counterpart of the classical shift \eqref{e:shift-B},
1934: should be compared with the approximate expression \eqref{e:CS-evol}.
1935: From there, a straighforward computation shows that, for any $f\in Lip(\Sigma)$:
1936: $$
1937: \tB_N^\dagger\,\Op^\ell_N(f)\,\tB_N = 
1938: \Op^{\ell+1}_N\big((1-\delta_{\ep_{0},1})\times f\circ\sigma\big)\,.
1939: $$
1940: The semiclassical equivalence $\Op^{\ell+1}_N\sim \Op^{\ell}_N$
1941: finishes the proof of $i)$.
1942: 
1943: To prove $ii)$, we remark that, 
1944: if $F\in Lip_c(C(B^{-1})\sqcup\ring{H}^{-1})$ and $f=F\circ J$,
1945: the function $F\circ B$ is supported away from $D(B)$, and both functions
1946: $\bbbone_{\set{\ep_0\neq 1}}\times f\circ\sigma$ and 
1947: $\bbbone_{\set{\ep_0\neq 1}}\times F\circ B\circ J$ are supported inside $\Sigma''$. 
1948: The semiconjugacy \eqref{e:semiconj} shows that these two functions are equal.
1949: Finally, we notice that, for $\bep\in\Sigma''$, one has
1950: $\bbbone_{\set{\ep_0\neq 1}}(\bep)=\bbbone_{M}\circ J(\bep)$.
1951: \end{proof}
1952: 
1953: %==========================================
1954: \subsubsection{Spectrum of the Walsh open baker}
1955: %==========================================
1956: 
1957: The simple expression \eqref{e:B_k} allows to explicitly compute
1958: the spectrum of $\tB_{N}$ (see \cite[Prop. 5.5]{NonZw06}). 
1959: That spectrum is determined by the two nontrivial
1960: eigenvalues $\lambda_-,\ \lambda_+$ of the matrix 
1961: $\widetilde F_3^*$. These eigenvalues have moduli $|\lambda_+|\approx 0.8443$, 
1962: $|\lambda_-|\approx 0.6838$.
1963: The spectrum of $\widetilde B_{N}$ has a gap: the long-living eigenvalues are 
1964: contained in the annulus $\set{|\lambda_-|\leq |\lambda|\leq |\lambda_+|}$,
1965: while the rest of the spectrum lies at the origin.
1966: Most of the eigenvalues are degenerate. If we count multiplicities, the long-living
1967: ($\equiv$ nontrivial) spectrum satisfies the following asymptotics when $k\to\infty$:
1968: \bequ\begin{split}
1969: \forall r>0,\qquad
1970: &\#\set{\lambda_j\in\Spec(\tB_{N})\,,\ |\lambda_j|\geq r}=C_r\,2^k+o(2^k)\,,\\
1971: C_r&=\begin{cases} 1\,, & r < r_0\\
1972: 0\,, & r > r_0
1973: \end{cases}\,,\qquad r_0\defeq|\lambda_-\lambda_+|^{1/2}=3^{-1/4}\,.
1974: \end{split}\eequ
1975: The nontrivial
1976: spectrum is spanned by a subspace $\cH_{N,long}$ of dimension $2^k$.
1977: Since the trapped set for $B=B_{\vr_{sym}}$ has dimension $2d=2\,\frac{\log 2}{\log 3}$,
1978: the above asymptotics agrees with the Fractal Weyl law \eqref{e:FWL}.
1979: Although the density of resonances (counted with multiplicities) is 
1980: peaked near the circle $\set{|\lambda|=r_0}$, the spectrum (as a set) 
1981: densely fills the
1982: annulus $\set{|\lambda_-|\leq |\lambda|\leq |\lambda_+|}$ when $k\to\infty$ 
1983: (see Fig.~\ref{f:toy-log}). 
1984: %%%%%%%%%%%%%%%%%%%
1985: \begin{figure}[ht]
1986: \rotatebox{-90}{\includegraphics[width=.5\textwidth]{toy-log+.eps}}
1987: \caption{\label{f:toy-log} Nontrivial spectrum of the Walsh open baker 
1988: $\tB_N$ for $N=3^{10}$ (circles) and
1989: $3^{15}$ (crosses), using a logarithmic representation 
1990: (horizontal=$\arg\lambda_j$, vertical=$\log|\lambda_j|$). We plot horizontal lines
1991: at the extremal radii $|\lambda_{\pm}|$ of the spectrum (dashed),
1992: at the radius $r_0=|\lambda_-\lambda_+|^{1/2}$ of highest degeneracies (dotted)
1993: and at the radius corresponding to the natural measure (full).}
1994: \end{figure}
1995: %%%%%%%%%%%%%%%%%%%
1996: In the next section we construct some long-living eigenstates of $\tB_N$ and analyze
1997: their Walsh-Husimi measures. Due to the spectral degeneracies, there is
1998: generally a large freedom to select eigenstates $(\psi_N)$ associated with
1999: a sequence of eigenvalues $(\lambda_N)$. Intuitively, that freedom should 
2000: provide more possibilities for semiclassical measures.
2001: 
2002: We mention that 
2003: J.~Keating and coworkers have recently studied the eigenstates 
2004: of a slightly different version of the Walsh-baker, namely 
2005: a matrix $\tB_N'$ obtained by replacing $F_3$ by the ``half-integer Fourier transform'' 
2006: $(G_3)_{jj'}=3^{-1/2}\e^{-2\i\pi(j+1/2)(j'+1/2)/3}$ (see~\cite{KNNS07}). 
2007: 
2008: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2009: \subsection{Long-living eigenstates of the Walsh open baker} 
2010: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2011: 
2012: We first provide the analogue of 
2013: Theorem~\ref{thm:main} for the Walsh-baker. We remind that
2014: a semiclassical measure $\mu^\Sigma$ (or its push-forward $\mu$) 
2015: is now a weak-$*$ limit of some sequence of Walsh-Husimi measures
2016: $(WH^{\ell}_{\psi_N})_{N=3^k\to\infty}$,
2017: where $\psi_N$ are eigenstates of $\tB_N$, and the index $\ell\approx k/2$.
2018: 
2019: From the quantum-classical correspondence of Prop.~\ref{p:Walsh-Egorov} and
2020: using Prop.~\ref{l:eigenmeasure-push}, we deduce the following
2021: %%%%%%%%%%%%%%%%%%%%%%%%%
2022: \begin{cor}\label{thm:walsh}
2023: Let $\mu^\Sigma$ be a semiclassical measure for a sequence of long-living eigenstates 
2024: $(\psi_N,\lambda_N)$ of 
2025: the Walsh-baker $\tB_N$. Then:
2026: 
2027: $i)$ $\mu^\Sigma$ is an eigenmeasure for the open shift $\sigma$
2028: on $\Sigma$, and 
2029: the corresponding decay rate $\Lambda$ satisfies $\Lambda=\lim_{N\to\infty}|\lambda_N|^2$.
2030: 
2031: $ii)$ If $\mu^\Sigma(\Sigma\setminus\Sigma'')=0$ (where $\Sigma''$ is defined in \eqref{e:Sigma''}), 
2032: then $\mu=J^*\mu^{\Sigma}$ is an eigenmeasure
2033: of the open baker $B$, with decay rate $\Lambda$ .
2034: \end{cor}
2035: %%%%%%%%%%%%%%%%%%%%%%%%%
2036: From the structure of $\Spec(\tB_N)$ explained above, 
2037: there exist sequences of eigenvalues $(\lambda_N)_{N\to\infty}$ converging to any 
2038: circle of radius $\lambda\in [|\lambda_-|,|\lambda_+|]$. We also
2039: know that any semiclassical measure is an eigenmeasure of $\sigma$, so it is 
2040: meaningful to ask Questions~\ref{qu1} in the present framework 
2041: (setting $\lambda_{\max/\min}=|\lambda_{\pm}|$). We add the following question:
2042: are there semiclassical measures $\mu^\Sigma$ such that $\mu^\Sigma(\Sigma'')<1$?
2043: In the next section we give partial answers to these questions.
2044: 
2045: Concerning the last point in Question~\ref{qu1}, we notice that
2046: the ``physical'' decay rate for $B=B_{\vr_{sym}}$ is $\Lambda_{nat}=2/3$. The
2047: circle $\set{|\lambda|=\sqrt{\Lambda_{nat}}}$ is contained inside the annulus
2048: $\set{|\lambda_-|\leq |\lambda|\leq |\lambda_+|}$ where the nontrivial spectrum
2049: of $\tB_N$ is semiclassically dense, although it differs from
2050: the circle $\set{|\lambda|=r_0}$ where the spectral density is peaked, 
2051: see Fig.~\ref{f:toy-log}. 
2052: Still, there exist semiclassical measures of $\tB_N$ with
2053: eigenvalue $\Lambda_{nat}$, and it is relevant to ask whether $\mu_{nat}$ 
2054: can be one of these. 
2055: At present we are not able to answer that question. The next section
2056: shows that there are plenty of semiclassical measures with eigenvalue $\Lambda_{nat}$,
2057: so even if $\mu_{nat}$ is a semiclassical measure with $\Lambda=\Lambda_{nat}$, 
2058: it is certainly not the only one.
2059: 
2060: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2061: \subsection{Constructing the eigenstates of $\tB_N$}
2062: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2063: 
2064: In this section we construct one particular (right)
2065: eigenbasis of $\tB_N$ restricted to the subspace $\cH_{N,long}$ of long-living
2066: eigenstates.
2067: The construction starts from the (right) eigenvectors
2068: $v_\pm\in\IC^3$
2069: of $\widetilde F_3^*$ associated with $\lambda_{\pm}$. 
2070: Notice that these two vectors (which we take normalized)
2071: are not orthogonal to each other.
2072: For any sequence 
2073: $\bet=\eta_0\ldots\eta_{k-1}$, $\eta_i\in\set{\pm}$, we form the tensor product state 
2074: $$
2075: |\bet\ra\defeq v_{\eta_0}\otimes v_{\eta_1}\ldots\otimes v_{\eta_{k-1}}\,.
2076: $$
2077: The action \eqref{e:B_k} of $\tB_N$ implies that
2078: $$
2079: \tB_N |\bet\ra = \lambda_{\eta_0}\,|\tau(\bet)\ra\,,
2080: $$
2081: where $\tau$ acts as a cyclic shift on the sequence: 
2082: $\tau(\eta_0\ldots\eta_{k-1})=\eta_1\ldots\eta_{k-1}\eta_0$.
2083: The {\em orbit} $\set{\tau^j(\bet),\ j\in\IZ}$ contains 
2084: $\ell_{\bet}$ elements, where the {\em period} $\ell_{\bet}$ of the 
2085: sequence $\bet$ necessarily divides $k$.
2086: The states $\set{|\tau^j(\bet)\ra,\ j=0,\ldots,\ell_{\bet}-1}$ are not orthogonal 
2087: to each other, but form a linearly 
2088: independent family, which generates the $\tB_N$-invariant subspace 
2089: $\cH_{\bet}\subset \cH_{N,long}$. The eigenvalues
2090: of $\tB_N$ restricted to $\cH_{\bet}$ are of the form 
2091: $\lambda_{\bet,r}=\e^{2\i\pi r/\ell_{\bet}}\,
2092: \big(\prod_{j=0}^{\ell_{\bet}-1}\lambda_{\eta_j}\big)^{1/\ell_{\bet}}$ 
2093: (with indices $r=1,\ldots,\ell_{\bet}$), 
2094: and the corresponding eigenstates read
2095: \bequ\label{e:formula}
2096: |\psi_{\bet,r}\ra=\frac1{\sqrt{\cN_{\bet,r}}}
2097: \sum_{j=0}^{\ell_{\bet}-1} c_{\bet,r,j}\; |\tau^j(\bet)\ra\,,\qquad 
2098: c_{\bet,r,j}=\prod_{m=0}^{j-1}\frac{\lambda_{\eta_m}}
2099: {\lambda_{\bet,r}}\,,
2100: \eequ
2101: where $\cN_{\bet,r}>0$ is the factor which normalizes $|\psi_{\bet,r}\ra$.
2102: Up to a phase, this state is unchanged if $\bet$ is replaced by $\tau(\bet)$.
2103: In the next subsections 
2104: we explicitly compute the Walsh-Husimi measures of some of these eigenstates.
2105: %%%%%%%%%%%%%%%%%%%
2106: \begin{figure}[htbp]
2107: \begin{center}
2108: \rotatebox{-90}{\includegraphics[width=8.5cm]{max6.ps}}
2109: \rotatebox{-90}{\includegraphics[width=8.5cm]{min6.ps}}
2110: \caption{\label{f:maximal} Walsh-Husimi densities for the extremal eigenstates $\psi_{+,N}$,
2111: $\psi_{-,N}$ of $\tB_{N}$, with $N=3^6$, $\ell=3$. These are coarse-grained 
2112: versions of the Bernoulli measures $\mu_{\vP_{+}}$ and $\mu_{\vP_{-}}$.}
2113: \end{center}
2114: \end{figure}
2115: %%%%%%%%%%%%%%%%%%%
2116: 
2117: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2118: \subsubsection{Extremal eigenstates} \label{s:extremal}
2119: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2120: 
2121: The simplest case is provided by the sequence $\bet=++\cdots+$, which has period $1$, so that
2122: $|\psi_{+,N}\ra=|\bet\ra=v_+^{\otimes k}$ is the (unique) eigenstate 
2123: associated with the
2124: largest eigenvalue $\lambda_+$ (this is the longest-living eigenstate).
2125: For any choice of index 
2126: $0\leq \ell\leq k$, the Walsh-Husimi measure of $|\psi_{+,N}\ra$ factorizes:
2127: $$
2128: \text{for all $\ell$-rectangle }[\bep]_\ell,\qquad
2129: WH^\ell_{\psi_{+,N}}([\bep]_\ell)
2130: =\prod_{j=0}^{\ell-1} |\la v_{+},\,e_{\ep_j}\ra|^2\ 
2131: \prod_{j=-1}^{-k+\ell}|\la v_{+},\,F_3^*\, e_{\ep_j}\ra|^2\,.
2132: $$
2133: The second product involves the vector $w_+\defeq F_3 v_+$, with components
2134: $w_{+,\ep}=(1-\delta_{\ep,1})v_{+,\ep}/\lambda_+$.
2135: Following the notations of \S\ref{s:Bernoulli}, let $\mu^\Sigma_{\vP_+}$ 
2136: be the Bernoulli eigenmeasure of $\sigma$ with weights 
2137: $P_{+,\ep}=|v_{+,\ep}|^2$, $P^*_{+,\ep}=|w_{+,\ep}|^2$.
2138: The above expression shows that the Husimi measure $WH^\ell_{\psi_{+,N}}$ is equal
2139: to the measure  $\mu^\Sigma_{\vP_+}$, 
2140: {\em conditioned on the grid formed by the $\ell$-cylinders}. 
2141: Since the diameters of the cylinders decrease to
2142: zero as $k\to\infty$, $\ell(k)\sim k/2$, the Husimi measures 
2143: $(H^\ell_{\psi_{+,N}})$ converge to $\mu^\Sigma_{\vP_+}$.
2144: 
2145: One can similarly show that the Husimi functions of the eigenstates
2146: $\psi_-=v_-^{\otimes k}$, associated with the smallest nontrivial eigenvalue $\lambda_-$, 
2147: converge to the Bernoulli
2148: eigenmeasure $\mu^\Sigma_{\vP_-}$, with weights $P_{-,\ep}=|v_{-,\ep}|^2$, 
2149: $P^*_{-,\ep}=|w_{-,\ep}|^2$, where $w_{-}=F_3 v_{-}$.
2150: 
2151: In Fig.~\ref{f:maximal} we plot the Walsh-Husimi densities (pushed-forward on $\t2$)
2152: for $\psi_{+,N}$ and $\psi_{-,N}$,
2153: using the ``isotropic'' $\ell=k/2$. These give a clear idea of the
2154: selfsimilar structure of the respective semiclassical measures 
2155: $\mu_{\vP_+}$ and $\mu_{\vP_-}$. The weights have the approximate values
2156: $\vP_+\approx (0.579, 0.287, 0.134)$,
2157: $\vP_-\approx (0.088, 0.532, 0.380)$.
2158: 
2159: Considering the fact that the eigenvalues $\lambda_N$ close to the circles 
2160: of radii $|\lambda_+|$
2161: and $|\lambda_-|$ have small degeneracies, we propose the following
2162: %&&&&&&&&&&&&&
2163: \begin{conj}
2164: Any sequence of eigenstates $(\psi_N)_{N\to\infty}$ with eigenvalues
2165: $|\lambda_N|\to |\lambda_+|$ (resp.
2166: $|\lambda_N|\to |\lambda_-|$) converges
2167: to the semiclassical measure $\mu^\Sigma_{\vP_+}$ (resp. $\mu^\Sigma_{\vP_-}$).
2168: \end{conj}
2169: %&&&&&&&&&&&&&
2170: This conjecture can be proven for the version of the Walsh baker 
2171: $\tB_N'$ studied in \cite{KNNS07}: in 
2172: that case the two eigenvectors of $\tilde G^*_3$ replacing $v_{\pm}$
2173: are orthogonal to each other, which greatly simplifies the analysis.
2174: The limit measure $\mu^\Sigma_{\vP'_+}$ is then the ``uniform'' measure 
2175: on the trapped set $\set{\ep_n\neq 1,\ n\in\IZ}$, 
2176: with $\vP'_+={\vP'_+}^*=(1/2,0,1/2)$.
2177: 
2178: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2179: \subsubsection{Semiclassical measures in the ``bulk''} 
2180: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2181: 
2182: %%%%%%%%%%%%%%%%%%%
2183: \begin{figure}[htbp]
2184: \begin{center}
2185: \rotatebox{-90}{\includegraphics[width=8.5cm]{mix+-+-4-0.ps}}
2186: \rotatebox{-90}{\includegraphics[width=8.5cm]{mix++--4-0.ps}}
2187: \caption{\label{f:mix} Walsh-Husimi densities of two eigenstates $\psi_{\bet,0}$
2188: constructed from the sequences $\bet_{4}=+-+-$ (left) and $\bet'_{4}=++--$ (right).}
2189: \end{center}
2190: \end{figure}
2191: %%%%%%%%%%%%%%%%%%%
2192: 
2193: In this section we investigate some eigenstates of the form \eqref{e:formula},
2194: with eigenvalues $\lambda_N$ situated in the ``bulk'' of the nontrivial spectrum, 
2195: that is $|\lambda_N|\in (|\lambda_-|,|\lambda_+|)$. 
2196: 
2197: In the next proposition we show that there can be two different semiclassical
2198: measures with the same decay rate $\Lambda$. 
2199: This answers by the {\em negative} the first point in Question~\ref{qu1}.
2200: %%%%%%%%%%%%%
2201: \begin{prop}
2202: Choose a rational number $t=\frac{m}{n}$, with $n\geq 1$, $1\leq m\leq n-1$.
2203: 
2204: i) Select a sequence $\bet_n$ with $m$ $(+)$ and $n-m$ $(-)$. 
2205: For all $k'\geq 1$, form the repeated sequence $(\bet_n)^{k'}$, and
2206: choose $r=r(k')\in\IZ$
2207: arbitrarily. Then the sequence of eigenstates 
2208: $\big(\psi_{(\bet_n)^{k'},r(k')}\big)_{k=nk'\to\infty}$ 
2209: converges to the semiclassical measure $\mu^\Sigma_{\bet_n}$, which is
2210: a linear combination of Bernoulli measures for the iterated shift $\sigma^n$ (see \eqref{e:rho_bet_n}). 
2211: This measure is independent of the choice of $\big(r(k')\big)$, and has decay rate 
2212: $\Lambda_t\defeq\big|\lambda_-^{1-t}\,\lambda_+^{t}\big|^2$. Its push-forward is an
2213: eigenmeasure of $B$.
2214: 
2215: ii) If $\bet_n$
2216: and $\bet'_n$ are two sequences with $m$ $(+)$ and $n-m$ $(-)$, 
2217: which are not related by a cyclic permutation, then 
2218: the semiclassical measures $\mu^\Sigma_{\bet_n}$,
2219: $\mu^\Sigma_{\bet'_{n}}$ are mutually singular, and so are their
2220: push-forwards on $\t2$.
2221: \end{prop}
2222: %%%%%%%%%%%%%
2223: Note that the radius $\sqrt{\Lambda_t}$ lies in the ``bulk'' of the 
2224: nontrivial spectrum, but can be different from the radius $r_0$ where the spectral 
2225: density is peaked.
2226: In Fig.~\ref{f:mix} we plot the Husimi functions of two states $\psi_{\bet_4,0}$,
2227: $\psi_{\bet'_4,0}$
2228: constructed from two $4$-sequences $\bet_{4}$, $\bet'_{4}$ not cyclically related. 
2229: The two functions, which give a rough idea of the limit measures $\mu_{\bet_{4}}$, 
2230: $\mu_{\bet'_{4}}$, seemingly concentrate on different parts of the phase space.
2231: \begin{proof}
2232: For short we call $\bet=(\bet_n)^{k'}$ which has period $\ell_{\bet}=\ell_{\bet_n}$, and $r=r(k')$. 
2233: From \eqref{e:formula}, each state $|\psi_{\bet,r}\ra$ is a combination of 
2234: $\ell_{\bet}$ states $|\tau^j(\bet)\ra$, and its eigenvalue has exact modulus $\sqrt{\Lambda_t}$.
2235: When $k'\to\infty$, those states are asymptotically orthogonal to
2236: each other. Indeed, their overlaps can be decomposed as
2237: $$
2238: \la \bet|\tau^j(\bet)\ra=\big(\la \bet_n|\tau^j(\bet_n)\ra \big)^{k'}\,
2239: \quad j=0,\ldots,\ell_{\bet}-1\,,
2240: $$
2241: and for any $j\neq 0\bmod \ell_{\bet}$ we have $\bet_n\neq \tau^j(\bet_n)$, which implies
2242: $|\la \bet_n|\tau^j(\bet_n)\ra|\leq c$, with $c \defeq |\la v_+,v_-\ra|^2<1$.
2243: As a result, the normalization factor of $\psi_{\bet,r}$ satisfies
2244: $$
2245: \cN_{\bet,r}= \sum_{j=0}^{\ell_{\bet}-1} |c_{\bet,r,j}|^2+\cO(c^{k'})\,,\qquad k'\to\infty\,.
2246: $$
2247: To study the semiclassical measures of the sequence $(\psi_{\bet,r})_{k'\to\infty}$, 
2248: we fix some cylinder
2249: $[\bal]=[\alpha_{-l'}\ldots\alpha_{-1}\cdot \alpha_0\ldots\alpha_{l-1}]$
2250: and compute the weight of the measures $H^\ell_{\psi_{\bet,r}}$. If $k$ is large enough and
2251: $\ell\sim k/2$, the conditions $\ell>l$ and $k-\ell>l'$ are fulfilled, and this weight
2252: can be written
2253: \bequ\label{e:overlap}
2254: H^\ell_{\psi_{\bet,r}}([\bal])=
2255: \la\psi_{\bet,r}|\Pi_{[\bal]}|\psi_{\bet,r}\ra=
2256: \sum_{j,j'=0}^{\ell_{\bet}-1} c_{\bet,r,j}\,\bar{c}_{\bet,r,j'}\,
2257: \la\tau^{j'}(\bet)|\Pi_{[\bal]}|\tau^j(\bet)\ra\,,
2258: \eequ
2259: where the projector on $[\bal]$ is a tensor product operator:
2260: $$
2261: \Pi_{[\bal]}=\pi_{\alpha_0}\otimes\pi_{\alpha_1}\ldots\pi_{\alpha_{l-1}}\otimes 
2262: (I)^{\otimes k-l-l'}\otimes F_3^*\pi_{\alpha_{-l'}}F_3\otimes\ldots \otimes F_3^*\pi_{\alpha_{-1}}F_3\,.
2263: $$
2264: The tensor factor $(I)^{k-l-l'-1}$ implies that each matrix element 
2265: $\la\tau^{j'}(\bet)|\Pi_{[\bal]}|\tau^j(\bet)\ra$ contains
2266: a factor $\big(\la\tau^{j'}(\bet_n)|\tau^j(\bet_n)\ra\big)^{k'-\cO(1)}$; 
2267: for the same reasons as above, this element
2268: is $\cO(c^{k'})$ if $j\neq j'$. We are then lead to consider only
2269: the diagonal elements $j=j'$:
2270: \bequ\label{e:weight}
2271: \forall j=0,\ldots,\ell_{\bet}-1,\qquad
2272: \la\tau^{j}(\bet)|\Pi_{[\bal]}|\tau^j(\bet)\ra=
2273: \prod_{i=0}^{l-1} P_{\eta_{j+i},\alpha_i}\,
2274: \prod_{i'=1}^{l'} P^*_{\eta_{j-i'},\alpha_{-i'}}\,.
2275: \eequ
2276: As above the weights
2277: $P_{\pm,\ep}=|v_{\pm,\ep}|^2$, $P^*_{\pm,\ep}=|w_{\pm,\ep}|^2$,
2278: and the definition of $\eta_i$ was extended to $i\in\IZ$ by periodicity. We claim that
2279: the right hand side exactly corresponds to $\tilde\mu^\Sigma_{\tau^j(\bet_n)}([\bal])$,
2280: where $\tilde\mu^\Sigma_{\tau^j(\bet_n)}$ is a certain Bernoulli eigenmeasure for the iterated
2281: shift $\sigma^n$.
2282: The latter can be seen as a simple shift on the symbol space $\tilde\Sigma$ constructed from
2283: $3^n$ symbols $\teps\in\set{0,\ldots,3^n-1}$: 
2284: each $\teps$ is in one-to-one correspondence
2285: with a certain $n$-sequence $\ep_0\ldots\ep_{n-1}$.
2286: Adapting the formalism of \S\ref{s:Bernoulli}
2287: to this new symbol space,
2288: the Bernoulli measure $\tilde\mu^\Sigma_{\tau^j(\bet_n)}$ corresponds to the
2289: following weight distributions $\tilde\vP$, $\tilde\vP^*$:
2290: \bequ\label{e:weights}
2291: \text{for}\ \tilde\eps\equiv \ep_0\ldots\ep_{n-1},\qquad
2292: \tilde P_{\tilde\ep}=\prod_{i=0}^{n-1} P_{\eta_{n,j+i},\ep_i}\quad\text{and}\quad
2293: \tilde P^*_{\tilde\ep}=\prod_{i=1}^{n} P^*_{\eta_{n,j-i},\ep_{n-i}}\,.
2294: \eequ
2295: The measures $\tilde\mu^\Sigma_{\tau^j(\bet_n)}$, $j=0,\ldots,\ell_{\bet}-1$,
2296: are related to one another through $\sigma$:
2297: \bequ\label{e:class-cycle}
2298: \sigma^*\,\tilde\mu^\Sigma_{\tau^j(\bet_n)}= |\lambda_{\eta_{n,j}}|^2\,
2299: \tilde\mu^\Sigma_{\tau^{j+1}(\bet_n)}\,.
2300: \eequ
2301: Finally, the semiclassical measure associated with the sequence
2302: $\big(\psi_{\bet,r}\big)_{k=nk'\to\infty}$ is
2303: \bequ\label{e:rho_bet_n}
2304: \mu^\Sigma_{\bet_n}=
2305: \frac1{\cN_{\bet_n}}\sum_{j=0}^{\ell_{\bet}-1} C_{\bet_n,j}\,\tilde\mu^\Sigma_{\tau^j(\bet_n)}\,,\quad
2306: \text{where}\quad C_{\bet_n,j}=\prod_{m=0}^{j-1}\frac{|\lambda_{\eta_{n,m}}|^2}{\Lambda_t}\,,
2307: \quad \cN_{\bet_n}=\sum_{j=0}^{n-1}C_{\bet_n,j}
2308: \eequ
2309: This is
2310: a probability eigenmeasure of $\sigma$, with decay rate $\Lambda_t$. It only 
2311: depends on the orbit $\set{\tau^j\bet_n,\ j=0,\ldots,\ell_{\bet_n}-1}$, and not on the
2312: choice of $(r(k'))$. This measure does not charge $\Sigma\setminus\Sigma''$, so its push-forward is an
2313: eigenmeasure of $B$.
2314: 
2315: The proof of statement $ii)$ goes as follows: since $\bet_n$ and $\bet'_n$ are
2316: not cyclically related, for
2317: any $j,j'\in\IZ$, the weight distributions $\tilde\vP$ and $\tilde\vP'$ 
2318: defining respectively the 
2319: Bernoulli measures $\tilde\mu^\Sigma_{\tau^j(\bet_n)}$ and
2320: $\tilde\mu^\Sigma_{\tau^{j'}(\bet'_n)}$ (see \eqref{e:weights}) are different. 
2321: As a result, these two measures are mutually singular
2322: (see the end of \S\ref{s:Bernoulli}), and so are the 
2323: two linear combinations
2324: $\mu^\Sigma_{\bet_n}$, $\mu^\Sigma_{\bet'_n}$. Because the weights $\tilde\vP$, $\tilde\vP'$
2325: are different from $(1,0,\ldots,0)$ and $(0,\ldots,1)$, the push-forwards $\mu_{\bet_n}$,
2326: $\mu_{\bet'_n}$ are also mutually singular.
2327: \end{proof}
2328: By a standard density argument, we can exhibit semiclassical measures for arbitrary decay
2329: rates in $[|\lambda_-|^2,|\lambda_+|^2]$.
2330: \begin{cor}\label{c:eigenmeasures}
2331: Consider a real number $t\in [0,1]$, and a 
2332: sequence of rationals $(t_p = \frac{m_p}{n_p}\in [0,1]\big)_{p\to\infty}$ converging
2333: to $t$ when $p\to\infty$. 
2334: For each $p$, let $\bet_p$ be a sequence with $m_p$
2335: $(+)$ and $n_p-m_p$ $(-)$, and $\mu^\Sigma_{\bet_p}$ the 
2336: $\sigma$-eigenmeasure constructed above.
2337: Then any weak-$*$ limit of the sequence
2338: $\big(\mu^\Sigma_{\bet_p}\big)_{p\to\infty}$ is a semiclassical measure of $(\tB_N)$;
2339: it is a $\sigma$-eigenmeasure  of decay rate 
2340: $\Lambda_{t}=\big|\lambda_-^{1-t}\,\lambda_+^{t}\big|^2$, and
2341: its push-forward is an eigenmeasure of $B$. 
2342: \end{cor}
2343: Let us call $\mathfrak{M}^\Sigma(\Lambda_t)$ the family of semiclassical measures obtained
2344: this way, and 
2345: $\mathfrak{M}^\Sigma=\cup_{t\in[0,1]}\mathfrak{M}^\Sigma(\Lambda_t)$.
2346: We don't know whether the family $\mathfrak{M}^\Sigma$
2347: exhausts the full set of semiclassical measures for $(\tB_N)$. Still, 
2348: we can address the third point in Question~\ref{qu1} with respect to this family.
2349: \begin{prop}
2350: The family of eigenmeasures $\mathfrak{M}^\Sigma$ does not contain
2351: the natural measure $\mu^\Sigma_{nat}=\mu^\Sigma_{\vr_{sym}}$. The same statement
2352: holds after push-forward on $\t2$.
2353: \end{prop}
2354: \begin{proof}
2355: For any $\ep_0\in\set{0,1,2}$, let us compute the weight of a
2356: measure $\mu^\Sigma_{\bet_p}\in\mathfrak{M}^\Sigma$ on the cylinder
2357: $[\eps_0]$ (corresponding to the vertical rectangle $\overline{R_{\ep_0}}$). For the 
2358: natural measure we have
2359: $\mu^\Sigma_{nat}([\ep_0])=1/3$ for $\ep_0=0,1,2$.
2360: 
2361: From \eqref{e:weights}, for any $j\in\IZ$ one has
2362: $\tilde\mu^\Sigma_{\tau^j(\bet_p)}([\eps_0])= P_{\eta_{p,j},\ep_0}$. Combining this
2363: with \eqref{e:rho_bet_n}, we get
2364: $$
2365: \mu^\Sigma_{\bet_p}([\ep_0])=C_{+}\,P_{+,\ep_0}+ (1-C_+)\,P_{-,\ep_0},\quad\text{where}\quad
2366: C_{+}=\frac1{\cN_{\bet_p}}\sum_{j\,:\,\bet_{p,j}=(+)} C_{\bet_p,j}\,.
2367: $$
2368: For any $\mu^\Sigma\in\mathfrak{M}$, the weights $\mu^\Sigma([\ep_0])$  will take 
2369: the same form, for some $C_{+}\in [0,1]$. Using the approximate expressions for
2370: the weights given in \S\ref{s:extremal},
2371: one checks that the condition 
2372: $\mu^\Sigma([\ep_0])=1/3$ cannot be satisfied simultaneously for $\ep_0=0,1,2$.
2373: \end{proof}
2374: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2375: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2376: \section{Concluding remarks}
2377: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2378: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2379: 
2380: The main result of this paper is the semiclassical connection between, on the one hand,
2381: eigenfunctions of a quantum open map (which mimick ``resonance eigenfunctions''), on the
2382: other hand, eigenmeasures of the classical open map. 
2383: 
2384: We proved that, modulo some problems at the discontinuities of the classical map, 
2385: semiclassical measures 
2386: associated with long-living resonant:
2387: are eigenmeasures of
2388: the classical dynamics, and their decay rate is directly related with 
2389: those of the corresponding resonant eigenstates (see Thm~\ref{thm:main}).
2390: This result, which basically derives from Egorov's theorem, has been expressed
2391: in a quite general framework, and applied to the specific example of the open baker's map. 
2392: An analogue has been proven
2393: for the more realistic setting of Hamiltonian scattering \cite[Theorem~3]{NonZw-gap}.
2394: 
2395: Although the construction and classification of eigenmeasures with decay rates $\Lambda<1$
2396: is quite easy, the classification of semiclassical measures
2397: among all possible eigenmeasures remains largely open (see Question~\ref{qu1}).
2398: The solvable model provided by the
2399: Walsh-quantized baker provides some hints to these classification, in the form
2400: of an explicit family of semiclassical measures,
2401: but we have no idea whether these results
2402: apply to more general systems, not even the ``standard'' quantum open baker. Indeed,
2403: the high degeneracies of the Walsh-baker may be responsible for a nongeneric profusion 
2404: of semiclassical measures for that model.
2405: Interestingly, the natural eigenmeasure does not seem to play
2406: a particular role at the quantum level. 
2407: 
2408: A tempting way of constraining the set of semiclassical
2409: measures would be to adapt the ``entropic'' methods of \cite{AN06} to open chaotic maps. 
2410: A desirable output of these methods would be, for instance, to forbid
2411: semiclassical measures from being of the pure point type described in \S\ref{s:pp}.
2412: 
2413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2414: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2415: \begin{thebibliography}{999999}
2416: 
2417: \bibitem{AN06}
2418: N.~Anantharaman and S.~Nonnenmacher, 
2419: {\em Entropy of semiclassical measures of the Walsh-quantized baker's map}, 
2420: Ann. Henri Poincar\'e {\bf 8} (2007) 37--74
2421: 
2422: \bibitem{BV89}
2423: N.~L.~Balazs and A.~Voros, {\em The quantized baker's transformation},
2424: Ann. Phys. (NY) {\bf 190} (1989) 1--31
2425: 
2426: \bibitem{BU03}
2427: D.~Borthwick and A.~Uribe, {\em On the pseudospectra of Berezin-Toeplitz operators}, 
2428: Meth. Appl. Anal. {\bf 10} (2003) 31--65
2429: 
2430: \bibitem{BouzDB96}
2431: A.~Bouzouina and S.~De~Bi\`evre, {\it Equipartition
2432: of the eigenfunctions of quantized ergodic maps on the torus}, 
2433: Commun. Math. Phys. {\bf 178} (1996)  83--105
2434: 
2435: \bibitem{CasMasShep99}
2436: G.~Casati, G.~Maspero and D.~Shepelyansky, {Quantum fractal eigenstates}, 
2437: Physica {\bf D 131} (1999) 311--316
2438: 
2439: \bibitem{CherMar97} N.~Chernov and R.~Markarian, {\it Ergodic properties of Anosov maps
2440: with rectangular holes}, Boletim Sociedade Brasileira Matematica {\bf 28} (1997) 271--314;
2441: N.~Chernov, R.~Markarian and S.~Troubetzkoy,
2442: {\it Invariant measures for Anosov maps with small holes},
2443: Ergod. Th. Dyn. Sys. {\bf 20} (2000) 1007--1044
2444: 
2445: \bibitem{CdV85} 
2446: Y.~Colin~de~Verdi\`ere, 
2447: {\it Ergodicit\'e et fonctions propres du laplacien}, 
2448: Commun. Math. Phys. {\bf 102} (1985) 497--502
2449: 
2450: \bibitem{DENW06}
2451: M.~Degli~Esposti, S.~Nonnenmacher and B.~Winn, 
2452: {\em Quantum Variance and Ergodicity for the baker's map},
2453: Commun. Math. Phys. {\bf 263} (2006) 325--352
2454: 
2455: \bibitem{DemYou06}
2456: M.~F.~Demers and L.-S.~Young, {\em Escape rates and conditionally invariant
2457: measures}, Nonlinearity {\bf 19} (2006) 377--397
2458: 
2459: \bibitem{DSZ04}
2460: N.~Dencker, J.~Sj\"ostrand and M.~Zworski, {\em Pseudospectra of semiclassical
2461: (pseudo-)differential operators},  Comm. Pure Appl. Math. {\bf 57} (2004) 384--415
2462: 
2463: \bibitem{ErmSara06}
2464: L.~Ermann and  M.~Saraceno, {\it Generalized Quantum Baker Maps as 
2465: perturbations of a simple kernel},
2466: Phys. Rev. {\bf E 74} (2006) 046205
2467: 
2468: \bibitem{FNdB03}
2469: F.~Faure, S.~Nonnenmacher and S.~De~Bi\`evre,
2470: {\it Scarred eigenstates for quantum cat maps of minimal periods},
2471: Commun. Math. Phys. {\bf 239} (2003) 449--492
2472: 
2473: \bibitem{GaspRice89}
2474: P. Gaspard and S.A. Rice, {\em Scattering from a classically chaotic repellor},
2475:  J. Chem. Phys. {\bf 90} (1989) 2225--2241; 
2476: ibid, {\em  Semiclassical quantization of the
2477: scattering from a classical chaotic repellor},
2478: J. Chem. Phys. {\bf 90} (1989) 2242--2254;
2479: A.~Wirzba, 
2480: {\it Quantum Mechanics and Semiclassics of Hyperbolic n-Disk Scattering Systems},
2481: Physics Reports {\bf 309} (1999) 1--116
2482: 
2483: \bibitem{GerLei93} 
2484: P.~G\'erard and E.~Leichtnam, 
2485: {\it Ergodic properties of eigenfunctions for the Dirichlet problem}, 
2486: Duke Math. J. {\bf 71} (1993) 559--607
2487: 
2488: \bibitem{GLZ04} L. Guillop\'e, K. Lin, and M. Zworski, 
2489: {\em The Selberg zeta function for convex co-compact Schottky groups,}
2490: Comm. Math. Phys, {\bf 245} (2004) 149--176
2491: 
2492: \bibitem{HJPPS86}
2493: T.C.~Halsey, M.H.~Jensen, L.P.~Kadanoff, I.~Procaccia and B.I.~Shraiman, 
2494: {\it Fractal measures and their singularities: The characterization of
2495: strange sets}, Phys. Rev. {\bf A 33} (1986) 1141--1151
2496: 
2497: \bibitem{HelMarRob87} 
2498: B.~Helffer, A.~Martinez and D.~Robert, {\it Ergodicit\'e et limite semi-classique}, 
2499: Commun. Math. Phys. {\bf 109} (1987) 313--326
2500: 
2501: \bibitem{Kelmer05}
2502: D.~Kelmer, {\it Arithmetic quantum unique ergodicity for symplectic linear
2503: maps of the multidimensional torus}, to appear in Ann. of Math., \texttt{math-ph/0510079};
2504: {\it ibid}, {\it Scarring on invariant manifolds for perturbed quantized hyperbolic toral automorphisms},
2505: preprint, \texttt{math-ph/0607033}
2506: 
2507: \bibitem{KNNS07}
2508: J.P.~Keating, S.~Nonnenmacher, M.~Novaes, and M.~Sieber,
2509: {\em On the resonance eigenstates of an open quantum baker map}, 
2510: in preparation.
2511: 
2512: \bibitem{KNPS06}
2513: J.P.~Keating, M.~Novaes, S.D.~Prado and M.~Sieber,
2514: {\it Semiclassical structure of chaotic resonance eigenfunctions}, 
2515: Phys. Rev. Lett. {\bf 97} (2006) 150406 
2516: 
2517: \bibitem{KurRud00} 
2518: P.~Kurlberg and Z.~Rudnick, {\it Hecke theory and
2519: equidistribution for the quantization of linear maps of the torus}, 
2520: Duke Math. J. {\bf 103} (2000) 47--77
2521: 
2522: \bibitem{Leben06}
2523: M.~Lebental, J.-S.~Lauret, J.~Zyss, C.~Schmit and E.~Bogomolny, 
2524: {\it Directional emission of stadium-shaped microlasers}, Phys. Rev. {\bf A 75}
2525: (2007) 033806
2526: 
2527: \bibitem{LeforWyatt83}
2528: C.~Leforestier and R.E.~Wyatt, {\it Optical potential for laser induced
2529: dissociation}, J. Chem. Phys. {\bf 78} (1983) 2334--2344
2530: 
2531: \bibitem{Lin02}
2532: K.~Lin, {\it Numerical study of quantum resonances in          
2533: chaotic scattering}, J. Comp. Phys. {\bf 176} (2002) 295--329;
2534: K.~Lin and M.~Zworski, {\it Quantum resonances in          
2535: chaotic scattering}, Chem. Phys. Lett. {\bf 355} (2002) 201--205
2536: 
2537: \bibitem{Linden06}
2538: E.~Lindenstrauss, {\it Invariant measures and arithmetic quantum unique ergodicity}, 
2539: Annals of Math. {\bf 163} (2006) 165-219
2540: 
2541: \bibitem{LSZ03}
2542: W.~Lu, S.~Sridhar, and M.~Zworski,
2543: {\em Fractal Weyl laws for chaotic open systems}, 
2544: Phys. Rev. Lett. {\bf 91} (2003) 154101
2545: 
2546: \bibitem{Mark96}
2547: R.~Markarian, A.~Lopes,
2548: {\it Open billiards: invariant and conditionally invariant probabilities on Cantor sets},
2549: SIAM J. Appl. Math. {\bf 56} (1996) 651--680
2550: 
2551: \bibitem{MarOK05}
2552: J.~Marklof and S.~O'Keefe, {\em Weyl's law and quantum ergodicity for maps with 
2553: divided phase space}, Nonlinearity {\bf 18} (2005) 277--304
2554: 
2555: \bibitem{MeenakLaksh04} N.~Meenakshisundaram and A.~Lakshminarayan,
2556: {\it Multifractal eigenstates of quantum chaos and the Thue-Morse sequence},
2557: Phys. Rev. {\bf E 71} (2005) 065303;
2558: ibid, {\it Using the Hadamard and related transforms for simplifying 
2559: the spectrum of the quantum baker's map}, J. Phys. {\bf A 39} (2006) 11205-11216
2560: 
2561: \bibitem{NonZw05}
2562: S.~Nonnenmacher and M.~Zworski, {\em Fractal Weyl laws in discrete models
2563: of chaotic scattering},
2564: J. Phys. {\bf A 38} (2005) 10683--10702 (special issue on ``Trends in quantum
2565: chaotic scattering''); S.~Nonnenmacher, {\it Fractal Weyl law for open chaotic maps}, in 
2566: {\it Mathematical physics of quantum mechanics}, J.~Asch and A.~Joye Eds.,
2567: Lect. Notes in Physics {\bf 690}, Springer, Berlin, 2006.
2568: 
2569: \bibitem{NonZw06}
2570: S.~Nonnenmacher and M.~Zworski, 
2571: {\em Distribution of resonances for open quantum maps}, 
2572: Commun. Math. Phys. {\bf 269} (2007) 311--365
2573: 
2574: \bibitem{NonZw-gap}
2575: S.~Nonnenmacher and M.~Zworski, 
2576: {\em Quantum decay rates in chaotic scattering}, talk given at 
2577: \'Ecole Polytechnique, Palaiseau, France (May 2006), 
2578: \texttt{http://math.berkeley.edu/$\sim$zworski/nzX.ps.gz}
2579: 
2580: \bibitem{RudSar94}
2581: Z.~Rudnick and P.~Sarnak, {\em The behaviour of eigenstates of arithmetic
2582: hyperbolic manifolds}, Commun. Math. Phys. {\bf 161} (1994) 195--213
2583: 
2584: \bibitem{Sar90}
2585: M.~Saraceno, {\em Classical structures in the quantized baker transformation}
2586: Ann. Phys. (NY) {\bf 199} (1990) 37--60
2587: 
2588: \bibitem{SaVa96} 
2589: M.~Saraceno and R.O.~Vallejos, {\it The quantized D-transformation}, 
2590: Chaos {\bf 6} (1996) 193--199
2591: 
2592: \bibitem{SchaCav00}
2593: R.~Schack and C.M.~Caves, {\it Shifts on a finite qubit string: 
2594: a class of quantum baker's maps},
2595: Appl. Algebra Engrg. Comm. Comput. {\bf 10}, 305--310 (2000)
2596: 
2597: \bibitem{Schni74} 
2598: A.~Schnirelman, {\it Ergodic properties of eigenfunctions}, 
2599: Usp. Math. Nauk. {\bf 29} (1974) 181--182
2600: 
2601: \bibitem{SchoTwo04}
2602: H.~Schomerus and J.~Tworzyd{\l}o, {\it Quantum-to-classical crossover of
2603: quasi-bound states in open quantum systems}, 
2604: Phys. Rev. Lett. {\bf 93} (2004) 154102
2605: 
2606: \bibitem{Roman-PhD}
2607: R.~Schubert, {\it Semiclassical localization in phase space},
2608: PhD thesis, University of Ulm, 2001.
2609: 
2610: \bibitem{SeidMill92}
2611: T.~Seideman and W.H.~Miller, {\it Calculation of the cumulative reaction
2612: probability via a discrete variable representation with absorbing boundary
2613: conditions}, J.~Chem.~Phys. {\bf 96} (1992) 4412--4422
2614: 
2615: \bibitem{Sjo90}
2616: J. Sj\"ostrand,  {\em
2617: Geometric bounds on the density of resonances for semiclassical problems},
2618: { Duke Math. J.}, {\bf 60} (1990) 1--57
2619: 
2620: \bibitem{SjoZw05} J. Sj\"ostrand and M. Zworski,
2621: {\em Fractal upper
2622: bounds on the density of semiclassical resonances},
2623: to appear in Duke Math. J., \texttt{math.SP/0506307}
2624: 
2625: \bibitem{Stef05}
2626: P.~Stefanov, {\it Approximating resonances with the Complex Absorbing 
2627: Potential Method},  Commun. PDE {\bf 30} (2005) 1843--1862
2628: 
2629: \bibitem{TangZw98}
2630: S.-H.~Tang and M.~Zworski, {\it From quasimodes to resonances},
2631: Math. Res. Lett. {\bf 5} (1998) 261--272;
2632: P.~Stefanov, {\it Quasimodes and resonances: sharp lower bounds},
2633: Duke Math. J. {\bf 99} (1999) 75--92
2634: 
2635: \bibitem{TraSco02}
2636: M.~M.~Tracy and A.~J.~Scott, {\it The classical limit for a class of 
2637: quantum baker's maps}, J. Phys. {\bf A 35} (2002) 8341--8360
2638: 
2639: \bibitem{Zel87}
2640: S.~Zelditch, 
2641: {\it Uniform distribution of the eigenfunctions on compact hyperbolic surfaces},
2642: Duke Math. J. {\bf 55} (1987) 919--941
2643: 
2644: \bibitem{Zel96}
2645: S.~Zelditch, {\em Quantum ergodicity of $C^*$ dynamical systems},
2646: Commun. Math. Phys {\bf 177} (1996) 507--528
2647: 
2648: \bibitem{ZelZwo96} S. Zelditch and M. Zworski, 
2649: {\it Ergodicity of eigenfunctions for Ergodic Billiards}, 
2650: Commun. Math. Phys {\bf 175} (1996) 673--682
2651: 
2652: \bibitem{Zw99} M. Zworski, {\em Dimension of the limit set and the density of
2653: resonances for convex co-compact Riemann surfaces}, Inv. Math. {\bf 136} (1999)
2654: 353--409
2655: \end{thebibliography}
2656: 
2657: \end{document}
2658: 
2659: 
2660: