1: \documentclass[12pt]{article}
2: \usepackage{psfig}
3: \usepackage{amssymb,amsmath,amstext,amsthm,amsfonts}
4: \usepackage[dvips]{graphicx}
5: \usepackage[ansinew]{inputenc}
6:
7: \title{Random perturbations of non-uniformly\\ expanding
8: maps\thanks{ Work partially supported by FCT through
9: {Centro de Matemática da Universidade do Porto.}}}
10:
11:
12: \author{José F. Alves, Vítor Araújo}
13:
14: \date{\today}
15:
16: \textwidth=15.5cm \oddsidemargin=.5cm
17: \begin{document}
18:
19: \newcommand{\mcup}{\mbox{$\bigcup$}}
20: \newcommand{\mcap}{\mbox{$\bigcap$}}
21:
22: \def \RR {{\mathbb R}}
23: \def \ZZ {{\mathbb Z}}
24: \def \NN {{\mathbb N}}
25: \def \PP {{\mathbb P}}
26: \def \TT {{\mathbb T}}
27:
28: \def \ra {\rightarrow }
29: \def \wh {\widehat }
30: \def \un{\underline }
31: \def \dist {\mbox{dist}}
32: \def \ov {\overline}
33: \def \supp {\mbox{supp}\, }
34: \def \wlim {\mbox{$w^*$-}\lim_{n\ra\infty}\, }
35: \def \distp{\mbox{d}_\PP }
36:
37: \def \al {\alpha } \def \be {\beta } \def \de {\delta }
38: \def \ga {\gamma } \def \ep {\epsilon } \def \vfi {\varphi
39: } \def \th {\theta } \def \si {\sigma }
40:
41: \def \cf {\mathcal{F}}
42: \def \cm {\mathcal{M}}
43: \def \cn {\mathcal{N}}
44: \def \cq {\mathcal{Q}}
45: \def \cp {\mathcal{P}}
46: \def \cc {\mathcal{C}}
47: \def \ch {\mathcal{H}}
48:
49:
50: \newcommand{\dem}{\begin{proof}}
51: \newcommand{\cqd}{\end{proof}}
52:
53:
54: \newcommand{\qand}{\quad\text{and}\quad}
55:
56:
57: \newtheorem{maintheorem}{Theorem}
58: \renewcommand{\themaintheorem}{\Alph{maintheorem}}
59: \newcommand{\cmt}{\begin{maintheorem}}
60: \newcommand{\fmt}{\end{maintheorem}}
61:
62:
63: \newtheorem{T}{Theorem}[section]
64: \newcommand{\ct}{\begin{T}}
65: \newcommand{\ft}{\end{T}}
66:
67: \newtheorem{Corollary}[T]{Corollary}
68: \newcommand{\cco}{\begin{Corollary}}
69: \newcommand{\fco}{\end{Corollary}}
70:
71: \newtheorem{Proposition}[T]{Proposition}
72: \newcommand{\cpr}{\begin{Proposition}}
73: \newcommand{\fpr}{\end{Proposition}}
74:
75: \newtheorem{Lemma}[T]{Lemma}
76: \newcommand{\cle}{\begin{Lemma}}
77: \newcommand{\fle}{\end{Lemma}}
78:
79: \newtheorem{Remark}[T]{Remark}
80: \newcommand{\cre}{\begin{Remark}}
81: \newcommand{\fre}{\end{Remark}}
82:
83: \newtheorem{Definition}[T]{Definition}
84: \newcommand{\cd}{\begin{Definition}}
85: \newcommand{\fd}{\end{Definition}}
86:
87: %\newcommand{\dem}{\par\medbreak\noindent{\em Proof}.\enspace}
88: %\newcommand{\cqd}{\,\,\,$\sqcup\!\!\!\!\sqcap\bigskip$}
89:
90:
91: \maketitle
92:
93: \begin{abstract} We give both sufficient conditions and
94: necessary conditions for the stochastic stability of
95: non-uniformly expanding maps either with or without
96: critical sets. We also show that the number of probability
97: measures describing the statistical asymptotic behaviour of
98: random orbits is bounded by the number of SRB measures if
99: the noise level is small enough.
100:
101: As an application
102: of these results we prove the stochastic stability of certain
103: classes of non-uniformly expanding maps introduced in
104: \cite{V} and \cite{ABV}. \end{abstract}
105:
106:
107: \section{Introduction}
108:
109: In broad terms, Dynamical Systems theory is mostly
110: interested in describing the typical behaviour of orbits as
111: time goes to infinity, and understanding how this behaviour
112: is modified under small perturbations of the system. This
113: work concentrates in the study of the latter problem from a
114: probabilistic point of view.
115:
116: Given a map $f$ from a manifold $M$ into itself, let
117: $(x_n)_{n\ge1}$ be the orbit of a given point $x_0\in M$,
118: that is $x_{n+1}=f(x_n)$ for every $n\ge 1$. Consider the
119: sequence of time averages of Dirac measures $\delta_{x_j}$
120: along the orbit of $x_0$ from time $0$ to $n$. A special
121: interest lies on the study of the convergence of such time
122: averages for a ``large" set of points $x_0\in M$ and the
123: properties of their limit measures. In this direction, we
124: refer the work of Sinai \cite{Si} for Anosov
125: diffeomorphisms, later extended by Ruelle and Bowen
126: \cite{BR,R} for Axiom A diffeomorphisms and flows. In the
127: context of systems with no uniform hyperbolic structure
128: Jakobson \cite{J} proved the existence of such measures for
129: certain quadratic transformations of the interval
130: exhibiting chaotic behaviour. Another important
131: contribution on this subject was given by Benedicks and
132: Young \cite{BY}, based on the previous work of Benedicks
133: and Carleson \cite{BC1,BC2}, where this kind of measures
134: were constructed for Hénon two dimensional maps exhibiting
135: strange attractors. The recent work of Alves, Bonatti and
136: Viana \cite{ABV} shows that such measures exist in great
137: generality for systems exhibiting some non-uniformly
138: expanding behaviour.
139:
140: The notion of stability that most concerns us can be
141: formulated in the following way. Assume that, instead of
142: time averages of Dirac measures supported on the iterates
143: of $x_0\in M$, we consider time averages of Dirac measures
144: $\delta_{x_j}$, where at each iteration we take $x_{j+1}$
145: close to $f(x_j)$ with a controlled error. One is
146: interested in studying the existence of limit measures for
147: these time averages and their relation to the analogous
148: ones for unperturbed orbits, that is, the stochastic
149: stability of the initial system.
150:
151: Systems with some uniformly hyperbolic structure are quite
152: well understood and stability results have been established
153: in general, see \cite{Ki1,Ki2} and \cite{Yo}. The
154: knowledge of the stochastic behaviour of systems that do
155: not exhibit such uniform expansion/contraction is still
156: very incomplete. Important results on this subject were
157: obtained by Benedicks, Young \cite{BY}, Baladi and Viana
158: \cite{BaV} for certain quadratic maps of the interval.
159: Another important contribution is the announced work of
160: Benedicks and Viana for Hénon-like strange attractors. As
161: far as we know these are the only results of this type for
162: systems with no uniform expanding behaviour.
163:
164: In this work we present both sufficient conditions and
165: necessary conditions for the stochastic stability of
166: non-uniformly expanding dynamical systems.
167: As an application of these results we prove
168: that the classes of non-uniformly expanding maps introduced
169: in \cite{V} and \cite{ABV} are stochastically stable.
170:
171:
172:
173:
174:
175: \subsection{Statement of results}
176: Let $f:M\ra M$ be a smooth map defined on a compact riemannian
177: manifold $M$.
178: We fix some normalized
179: riemannian volume form $m$ on $M$ that we call {\em Lebesgue measure}.
180:
181: Given $\mu$ an $f$-invariant Borel probability measure on
182: $M$, we say that $\mu$ is an \emph{SRB measure}
183: if, for a positive
184: Lebesgue measure set of points $x\in M$, the averaged
185: sequence of Dirac measures along the orbit $(f^n(x))_{n\geq
186: 0}$ converges in the weak$^*$ topology to $\mu$, that is,
187: \begin{equation}\label{average}
188: \lim_{n\ra +\infty}
189: \frac{1}{n}\sum_{j=0}^{n-1}\vfi\big(f^n(x)\big)
190: =\int \vfi\, d\mu
191: \end{equation} for every continuous map $\vfi:M\to \RR$. We
192: define the {\em basin} of $\mu$ as the set of those points
193: $x$ in $M$ for which (\ref{average}) holds for all
194: continuous $\vfi$. The maps to be considered in this work
195: will only have a finite number of SRB measures whose basins
196: cover the whole manifold $M$, up to a set of zero Lebesgue
197: measure.
198:
199: We are interested in studying random perturbations of the
200: map $f$. For that, we take a continuous map
201: \[
202: \begin{array}{rccl}
203: \Phi:& T &\longrightarrow& C^2(M,M)\\
204: & t &\longmapsto & f_t
205: \end{array}
206: \]
207: from a metric space $T$ into the space of $C^2$ maps from
208: $M$ to $M$, with $f=f_{t^*}$ for some fixed $t^*\in T$.
209: Given $x\in M$ we call the sequence $\big(f_{\un
210: t}^n(x)\big)_{n\ge1}$ a \emph{random orbit} of $x$, where
211: $\un t$ denotes an element $(t_1,t_2,t_3,\ldots)$ in the
212: product space $T^{\NN}$ and
213: $$
214: f^n_{\un t}= f_{t_n}\circ \cdots \circ
215: f_{t_1}\quad \mbox{for}\quad n \ge1.
216: $$
217: We also take a family $(\th_\ep)_{\ep>0}$ of probability
218: measures on $T$ such that $(\supp\th_\ep)_{\ep>0}$ is a
219: nested family of connected compact sets and
220: $\supp\th_\ep\rightarrow \{t^*\}$ when $ \ep\to 0$.
221: We will also assume some quite general nondegeneracy
222: conditions on $\Phi$ and $(\th_\ep)_{\ep>0}$ (see the beginning of Section
223: \ref{s.stationary}) and refer to $\{\Phi,(\th_\ep)_{\ep>0}\}$
224: as a {\em random perturbation} of $f$.
225:
226: In the context of random perturbations of a map we say
227: that a Borel probability measure $\mu^\ep$ on $M$ is
228: \emph{physical } if for a positive Lebesgue measure set of
229: points $x\in M$, the averaged sequence of Dirac
230: probability measures $\delta_{f_{\un t}^n(x)}$ along random
231: orbits $\big(f_{\un t}^n(x) \big)_{n\geq 0}$ converges in
232: the weak$^*$ topology to $\mu^\ep$ for $\th_\ep^\NN$ almost
233: every $\un t\in T^\NN$. That is,
234: \begin{equation}\label{pertaverage}
235: \lim_{n\ra +\infty}
236: \frac{1}{n}\sum_{j=0}^{n-1}\vfi\big(f_{\un t}^n(x)\big)
237: =\int \vfi \,d\mu^\ep \quad\mbox{for all continuous $\vfi\colon
238: M\rightarrow\RR$}
239: \end{equation}
240: and $\th^\NN_\ep$ almost every $\un t\in
241: T^\NN.$
242: We denote the set of points $x\in M$ for which
243: (\ref{pertaverage}) holds by $B(\mu^\ep)$ and call it the
244: \emph{basin of $\mu^\ep$}.
245: The map $f\colon M\ra M$ is said to be {\em stochastically
246: stable} if the weak$^*$ accumulation points (when $\ep>0$
247: goes to zero) of the physical probability measures of $f$
248: are convex linear combinations of the (finitely many) SRB
249: measures of~$f$.
250:
251: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
252:
253: \subsubsection{Local diffeomorphisms}
254:
255: Let $f:M\ra M$ be a $C^2$
256: local diffeomorphism of the manifold $M$.
257: We say that $f$ is {\em non-uniformly expanding} if there
258: is some constant $c>0$ for which
259: \begin{equation} \label{liminf1}
260: \limsup_{n\ra +\infty}\frac{1}{n}
261: \sum_{j=0}^{n-1}\log\|Df(f^j(x))^{-1}\|\leq -c<0
262: \end{equation}
263: for
264: Lebesgue almost every $x\in M$. It was proved in \cite{ABV}
265: that for a non-uniformly expanding local diffeomorphism $f$
266: the following holds:
267: \begin{itemize}
268: \item[(P)]
269: {\em There is a finite number of
270: ergodic absolutely continuous (\emph{SRB}) $f$-invariant
271: probability measures $\mu_1,\dots ,\mu_p$ whose basins
272: cover a full Lebesgue measure subset of $M$. Moreover,
273: every absolutely continuous $f$-invariant probability
274: measure $\mu$ may be written as a convex linear
275: combination of $\mu_1,\dots ,\mu_p$: there are real numbers
276: $w_1,\dots,w_p\geq 0$ with $w_1+\cdots+w_p=1$ for which
277: $\mu=w_1\mu_1+\cdots+w_p\mu_p$}.
278: \end{itemize}
279: The proof of the previous result was based on the
280: existence of $\al$-hyperbolic times
281: for the points in $M$:
282: given $0<\al<1$, we say that $n\in\ZZ^+$ is a {\em
283: $\al$-hyperbolic time} for the point $x\in M$ if
284: \begin{equation}\label{htimesc}
285: \prod_{j=n-k}^{n-1}
286: \|Df(f^j(x))^{-1}\|\leq \al^k \quad\mbox{for every}\quad
287: 1\leq k\leq n.
288: \end{equation}
289: The existence of (a positive frequency of) $\al$-hyperbolic
290: times for points $x\in M$ is a consequence of the
291: hypothesis of non-uniform expansion of the map $f$ and
292: permits us to define a map $h:M\to\ZZ^+$ giving the first
293: hyperbolic time for $m$ almost every $x\in M$.
294:
295:
296: In the context of random perturbations of a non-uniformly
297: expanding
298: map we are also able to prove a result on the finitness
299: of physical measures.
300:
301: \cmt \label{t.finite1}
302: Let $f\colon M\rightarrow M$ be a $C^2$ non-uniformly expanding
303: local diffeomorphism. If $\ep>0$ is sufficiently small, then there
304: are physical measures $\mu^\ep_1,\dots, \mu^\ep_l$
305: (with $l$ not depending on
306: $\ep$)
307: such that:
308: \begin{enumerate} \item for each $x\in M$ and
309: $\th_\ep^\NN$ almost every $\un t\in T^\NN$, the average of
310: Dirac measures $\delta_{f_{\un t}^n(x)}$ converges in the
311: weak$^*$ topology to some $\mu^\ep_i$ with $1\le i\le l$;
312: \item for each $1\le i\le l$ we have
313: $$
314: \mu^\ep_i=\wlim\frac1n\sum_{j=0}^{n-1}\int\big(f_{\un t}^j\big)_*
315: \big(m\mid B(\mu^\ep_i)\big)\,d\th_\ep^\NN(\un t),
316: $$
317: where $m\mid B(\mu^\ep_i)$ is the normalization of the
318: Lebesgue measure restricted to $B(\mu^\ep_i)$;
319: \item if $f$ is
320: transitive, then $l=1$.
321: \end{enumerate}
322: \fmt
323:
324:
325: We say that the map $f$ is {\em non-uniformly expanding for
326: random orbits} if there is some constant $c>0$ such that
327: for $\ep>0$ small enough
328: \begin{equation} \label{liminf2}
329: \limsup_{n\ra +\infty}\frac{1}{n}
330: \sum_{j=0}^{n-1}\log\|Df(f_{\un t}^j(x))^{-1}\|\leq -c<0,
331: \end{equation}
332: for $\th_\ep^\NN\times m$ almost every $(\un t,x)\in
333: T^\NN\times M$. Similarly to the deterministic situation,
334: condition (\ref{liminf2}) permits us to introduce a notion
335: of $\al$-hyperbolic times for points in $T^\NN\times M$
336: and define a map
337: $$
338: h_\ep\colon T^\NN\times M\ra \ZZ^+
339: $$
340: by taking $h_\ep(\un t,x)$ the first $\al$-hyperbolic time
341: for the point $(\un t,x)\in T^\NN\times M$ (see
342: Section~\ref{s.distortion}). Assuming that
343: %$h_\ep\in L^1(\th^\NN_\ep\times m)$
344: $h_\ep$ is integrable with respect to $\th^\NN_\ep\times
345: m$,
346: then
347: \begin{equation}\label{c.unif}
348: \|h_\ep\|_1=\sum_{k=0}^{\infty}k\,
349: (\th_\ep^\NN\times m)
350: \big(\{(\un t,x)\colon h_\ep(\un t,x)=k\}\,\big)
351: <\infty.
352: \end{equation}
353: %for small $\ep>0$.
354: We say that the family $(h_\ep)_{\ep>0}$ has {\em uniform
355: $L^1$-tail}, if the series in (\ref{c.unif}) converges uniformly
356: to $\|h_\ep\|_1$ (as a series of functions on the variable $\ep$).
357:
358:
359:
360:
361: \cmt \label{t.stc}
362: Let $f\colon M\rightarrow M$ be a non-uniformly expanding $C^2$
363: local diffeomorphism.
364: \begin{enumerate}
365: \item If $f$ is stochastically stable,
366: then $f$ is non-uniformly expanding for random orbits.
367: \item If $f$ is non-uniformly expanding for random orbits
368: and $(h_\ep)_\ep$ has uniform $L^1$-tail,
369: then $f$ is stochastically stable.
370: \end{enumerate}
371: \fmt
372:
373: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
374:
375: \subsubsection{Maps with critical sets}
376:
377: Similar results to those presented for random perturbations
378: of local diffeomorphisms will also be obtained for maps
379: with critical sets in the sense of~\cite{ABV}. We start by
380: describing the class of maps that we are going to consider. Let
381: $f\colon M\ra M$ be a continuous map of the compact manifold $M$
382: that fails to be a $C^2$ local diffeomorphism on a critical set $\cc\subset M$ with zero Lebesgue measure. We assume that $f$ {\em behaves like a power of the distance} close to
383: the critical set $\cc$:
384: there are constants $B>1$ and $\be>0$ for which
385:
386: \begin{itemize}
387: \item[(S1)]
388: \hspace{.1cm}$\displaystyle{\frac{1}{B}\dist(x,\cc)^{\be}\leq
389: \frac{\|Df(x)v\|}{\|v\|}\leq B\dist(x,\cc)^{-\be}}$;
390: \item[(S2)]
391: \hspace{.1cm}$\displaystyle{\left|\log\|Df(x)^{-1}\|-
392: \log\|Df(y)^{-1}\|\:\right|\leq
393: B\frac{\dist(x,y)}{\dist(x,\cc)^{\be}}}$;
394: \item[(S3)]
395: \hspace{.1cm}$\displaystyle{\left|\log|\det Df(x)^{-1}|-
396: \log|\det Df(y)^{-1}|\:\right|\leq
397: B\frac{\dist(x,y)}{\dist(x,\cc)^{\be}}}$;
398: \end{itemize}
399: for every $x,y\in M\setminus \cc$ with
400: $\dist(x,y)<\dist(x,\cc)/2$ and $v\in T_x M$. Given
401: $\delta>0$ we define the $\delta$-{\em truncated distance}
402: from $x\in M$ to $\cc$
403: $$ \dist_\delta(x,\cc)= \left\{
404: \begin{array}{ll} 1 & \mbox{if }\dist(x,\cc)\geq \delta,\\
405: \dist (x,\cc) & \mbox{otherwise.} \end{array} \right. $$
406:
407: Assume that $f$ is a non-uniformly expanding map, in the
408: sense that there is $c>0$ such that the limit in
409: (\ref{liminf1}) holds for Lebesgue almost every $x\in M$
410: (recall that we are taking $\cc$ with zero Lebesgue
411: measure) and, moreover, suppose that the orbits of $f$ have
412: {\em slow approximation to the critical set}: given small
413: $\gamma >0$ there is $\delta >0$ such that
414: \begin{equation}
415: \label{limsup1}
416: \limsup_{n\ra +\infty}\frac{1}{n}
417: \sum_{j=0}^{n-1}-\log\dist_\delta(f^j(x), \cc)\leq \gamma
418: \end{equation}
419: for Lebesgue almost every $x\in M$. The results in
420: \cite{ABV} show that in this situation we obtain the
421: same conclusion on the finiteness of SRB measures for such an $f$, also holding property (P).
422:
423: In order to prove the stochastic stability of maps with
424: critical sets we need to restrict the class of
425: perturbations we are going to consider: we take maps $f_t$
426: with the same critical set $\cc$ and impose that
427: \begin{equation}\label{e.perturbation}
428: Df_t(x)=Df(x) \quad\mbox{for every $x\in M\setminus\cc$ and $t\in T$}.
429: \end{equation}
430: This may be implemented, for instance, in parallelizable
431: manifolds (with an additive group structure, e.g. tori
432: $\TT^d$ or cylinders $\TT^{d-k}\times\RR^k$) by considering
433: $$T=\{t\in \RR^d\colon \|t\|\leq \ep_0\}$$
434: for some
435: $\ep_0>0$, $\th_\ep$ the normalized Lebesgue measure on the
436: ball of radius $\ep\leq \ep_0$, and taking $f_t=f+t$; that
437: is, adding at each step a random noise to the unperturbed
438: dynamics.
439:
440:
441:
442: For the case of maps with critical sets we also need to
443: impose an analog of condition~(\ref{limsup1}) for random
444: orbits; we assume {\em slow approximation of random orbits
445: to the critical set}: given any small $\gamma >0$ there is
446: $\delta >0$ such that \begin{equation}
447: \label{limsup2}
448: \limsup_{n\ra +\infty}\frac{1}{n}
449: \sum_{j=0}^{n-1}-\log\dist_\delta(f^j_{\un t}(x), \cc)\leq
450: \gamma
451: \end{equation}
452: for $\th_\ep^\NN\times m$ almost every $(\un t,x)\in
453: T^\NN\times M$ and small $\ep>0$. Results similar to
454: those presented for local diffeomorphisms on the finiteness
455: of physical measures can also be obtained in this case.
456:
457:
458: \cmt \label{t.finite2}
459: Let $f\colon M\rightarrow M$ be
460: a $C^2$ non-uniformly expanding map behaving
461: like a power of the distance close to the critical set
462: $\cc$, and whose orbits have slow
463: approximation to $\cc$. If $f$ is non-uniformly expanding for
464: random orbits and random orbits have slow
465: approximation to $\cc$,
466: then we arrive at the
467: same conclusions of Theorem~\ref{t.finite1}.
468: \fmt
469:
470:
471: The property of non-uniform expansion
472: for random orbits, together with the slow approximation of random
473: orbits to the critical set permit us to introduce a notion
474: of $(\al,\delta)$-hyperbolic times for points in $(\un
475: t,x)\in T^\NN\times M$ and define a map
476: $$
477: h_\ep\colon T^\NN\times M\ra \ZZ^+,
478: $$
479: by taking $h_\ep(\un t, x)$ the first
480: $(\al,\delta)$-hyperbolic time for the point $(\un t,x)\in
481: T^\NN\times M$, see Section \ref{s.distortion}. Assuming
482: that $h_\ep$ is integrable with respect to $\th_\ep\times
483: m$, then we obtain an analog to (\ref{c.unif}), which
484: enables us to define a notion of {\em uniform $L^1$-tail}
485: exactly in the same way as before.
486:
487: Due to the fact that $\log\|Df^{-1}\|$ is not a continuous map
488: (it is not even everywhere defined) we are not able to
489: present in this context a similar to Theorem~\ref{t.stc} in
490: all its strength. However, we obtain the same kind of
491: conclusion of the second item of Theorem~\ref{t.stc}.
492:
493: \cmt \label{t.sts}
494: Let $f\colon M\rightarrow M$ be
495: non-uniformly expanding $C^2$ map behaving like a power of the
496: distance close to its critical set $\cc$ and whose orbits have slow
497: approximation to $\cc$. Assume that $f$ is non-uniformly
498: expanding for random orbits and random orbits have slow
499: approximation to $\cc$. If $(h_\ep)_\ep$ has uniform
500: $L^1$-tail, then $f$ is stochastically stable. \fmt
501:
502: As a major application of the previous theorem we are
503: thinking of a class of maps on the cylinder $S^1\times \RR$
504: introduced in~\cite{V}. Subsequent
505: works~\cite{Al} and~\cite{AV} showed that such systems are
506: topologically mixing (thus transitive) and have a unique
507: SRB measure. In Section \ref{s.applications} we prove that
508: Viana maps satisfy the hypotheses of Theorem~\ref{t.sts},
509: hence being stochastically stable. An application of
510: Theorem~\ref{t.stc} will also be given in
511: Section~\ref{s.applications} for an open class of local
512: diffeomorphisms introduced in~\cite[Appendix A]{ABV}.
513:
514:
515:
516: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
517:
518:
519:
520:
521: \section{Distortion bounds} \label{s.distortion}
522:
523: In this section we generalize some of the results in
524: \cite{Al} and \cite{ABV} for the setting of stochastic
525: perturbations of a non-uniformly expanding map. These
526: results will be proved in the setting of maps with critical sets.
527: Then everything follows in the same way for local
528: diffeomorphisms if we think of $\cc$ as being equal to
529: the empty set, with the only exception of a particular point
530: that we clarify in Remark \ref{r.hyp} below (due to
531: the fact that we are not assuming condition
532: (\ref{e.perturbation}) for maps with no critical sets).
533: For the next definition we take
534: $0<b<\min\{1/2,1/(2\beta)\}$.
535:
536: \cd Given $0<\al<1$ and $\delta>0$, we say that $n\in\ZZ^+$ is a
537: $(\al,\delta)$-hyperbolic time for $(\un t,x)\in T^\NN\times M$
538: if
539: $$
540: \prod_{j=n-k}^{n-1}
541: \|Df_{t_{j+1}}(f^j_{\un t}(x))^{-1}\|\leq\al^k
542: \quad\mbox{and}\quad
543: \dist_\delta(f_{\un t}^{n-k}(x),\cc)\geq \al^{bk}
544: $$
545: for every $1\leq k\leq n$.
546: \fd
547:
548: The following lemma, due to Pliss \cite{Pli}, provides the
549: main tool in the proof of the existence of hyperbolic times
550: for points with non-uniform expansion on random orbits.
551:
552: \cle \label{l.pliss} Let $H\ge c_2 > c_1 >0$ and
553: $\zeta={(c_2-c_1)}/{(H-c_1)}$. Given real numbers
554: $a_1,\ldots,a_N$ satisfying
555: $$
556: \sum_{j=1}^N a_j \ge c_2 N
557: \qand a_j\le H \;\;\mbox{for all}\;\; 1\le j\le N,
558: $$
559: there
560: are $l>\zeta N$ and $1<n_1<\ldots<n_l\le N$ such that $$
561: \sum_{j=n+1}^{n_i} a_j \ge c_1\cdot(n_i-n) \;\;\mbox{for
562: each}\;\; 0\le n < n_i, \; i=1,\ldots,l. $$
563: \fle
564: \dem See \cite[Lemma 3.1]{ABV}.
565: \cqd
566:
567:
568:
569: \cpr \label{p.hyp} There are $\al>0$ and $\delta>0$ for
570: which $\th_\ep^\NN\times m$ almost every $(\un t,x)\in
571: T^\NN\times M$ has some $(\al,\delta)$-hyperbolic time.
572: \fpr
573: \dem
574: Let $(\un t,x)\in T^{\NN}\times M$ be a point
575: satisfying~(\ref{liminf2}). For large $N$ we have $$
576: -\sum_{j=0}^{N-1} \log \left\| Df ( f^j_{\un t} (x) )^{-1}
577: \right\| \ge \frac{c}2 N >0, $$ by definition of
578: non-uniform expansion on random orbits. Fixing $\rho>\be$
579: we see that condition (S1) implies
580: \begin{equation}\label{ineq1} \left| \log \left\|
581: Df(x)^{-1} \right\| \right| \le \rho \left| \log {\rm
582: dist\,}(x,\cc) \right| \end{equation} for every $x$ in a
583: neighborhood $V$ of $\cc$. Now we take $\ga_1>0$ so that
584: $\rho\ga_1\le c/10$ and let $\de_1>0$ be small enough to
585: get
586: \begin{equation}\label{ineq2}
587: -\sum_{j=0}^{N-1} \log {\rm dist\,}_{\de_1} (f^j_{\un
588: t}(x), S) \le \ga_1 N\quad\mbox{for large $N$},
589: \end{equation} which is possible after
590: property~(\ref{limsup1}) of slow approximation to $\cc$.
591: Moreover, fixing $H\ge \rho|\log \de|$
592: sufficiently large in order that it be also an upper bound
593: for for the set $\{ -\log \| Df^{-1}_{\un t}\| : t\in T, \;
594: x\in M\setminus V\}$, then the set
595: $$
596: E=\{ 1\le j\le N: -\log
597: \|Df(f^{j-1}_{\un t}(x))^{-1}\|>H \}
598: $$
599: is such that
600: $f^{j-1}_{\un t}(x)\in V$ for all $j\in E$ and
601: $$
602: \rho\left|\log {\rm dist\,}(f^{j-1}_{\un t}(x),\cc)\right|
603: > -\log \left\| Df(f^{j-1}_{\un t}(x))^{-1} \right\| > H
604: \ge \rho|\log \de|
605: $$
606: i.e., ${\rm dist\,}(f^{j-1}_{\un
607: t}(x),\cc)<\de_1$, in particular
608: $\dist_{\de_1}(f^{j-1}_{\un t}(x),\cc)= \dist(f^{j-1}_{\un
609: t}(x),\cc)<\de_1$ for all $j\in E$. Hence, defining
610: $$
611: a_j=\left\{ \begin{array}{lcr} -\log \left\|
612: Df(f^{j-1}_{\un t}(x))^{-1} \right\| & \mbox{if} & j\not\in
613: E \\ 0 & \mbox{if} & j\in E \end{array} \right. $$ it holds
614: $a_j\le H$ for $1\le j \le N$, and~(\ref{ineq1})
615: and~(\ref{ineq2}) imply $$ -\sum_{j\in E} \log \left\|
616: Df(f^{j-1}_{\un t}(x))^{-1}\right\| \le \rho\sum_{j\in E}
617: \left| \log \dist (f^{j-1}_{\un t}(x),\cc) \right| \le
618: \rho\ga_1 N. $$
619: Since $\rho\ga_1\le c/10$ we deduce
620: $$
621: \sum_{j=1}^N a_j = \sum_{j=1}^N \left( -\log \left\|
622: Df(f^{j-1}_{\un t}(x))^{-1} \right\| \right) - \sum_{j\in
623: E} \left( -\log \left\| Df(f^{j-1}_{\un t}(x))^{-1}
624: \right\| \right) \ge \frac25 c N.
625: $$
626: By the previous arguments we may apply Lemma \ref{l.pliss}
627: to the sequence $a_j$ with $c_1=c/5$ and $c_2=2c/5$ (we may
628: suppose $H>c_1$ too by increasing $H$ if needed). Thus
629: there are $\zeta_1>0$ and $l_1>\zeta_1 N$ times $1\le q_1 <
630: \ldots < q_{l_1} \le N$ such that
631: \begin{equation}\label{ineq3} \sum_{j=n+1}^{q_i} -\log
632: \left\| Df(f^{j-1}_{\un t}(x))^{-1} \right\| \ge
633: \sum_{j=n+1}^{q_i} a_j \ge \frac{c}2(q_i-n) \end{equation}
634: for every $0\le n <q_i, \; i=1,\ldots,l_1$. We observe
635: that~(\ref{ineq3}) is just the first part of the
636: requirements on $(\al,\de)$-hyperbolic times for $(\un
637: t,x)$ if $\al=\exp(c/5)$.
638:
639: Now we apply again Lemma~\ref{l.pliss}, this time to the
640: sequence $a_j=\log \dist_{\de_2} (f^{j-1}_{\un t}(x),\cc)$,
641: where $\de_2>0$ is small enough so that for $\ga_2>0$ with
642: $2\ga_2 (bc)^{-1}<\zeta_1$ we have by
643: assumption~(\ref{limsup1}) $$ \sum_{j=0}^{N-1} \log
644: \dist_{\de_2} (f^j_{\un t}(x) , \cc) \ge -\ga_2
645: N\quad\mbox{for large $N$}. $$ Defining $c_1=bc/2$, $c_2=
646: -\ga_2$, $H=0$ and $$ \zeta_2=\frac{c_2-c_1}{H-c_1}=
647: 1-\frac{2\ga_2}{bc}, $$ Lemma~\ref{l.pliss} ensures that
648: there are $l_2\ge\zeta_2 N$ times $1\le r_1 < \ldots <
649: r_{l_2} \le N$ satisfying \begin{equation}\label{ineq4}
650: \sum_{j=n+1}^{r_i} \log \dist_{\de_2} (f^{j+1}_{\un t}(x),
651: \cc) \ge \frac{bc}2 (r_i -n ) \end{equation} for every
652: $0\le n < r_i$, $i=1,\ldots,l_2$. Let us note that the
653: condition on $\ga_2$ assures $\zeta_1+\zeta_2>1$. So if
654: $\zeta=\zeta_1+\zeta_2-1$, then there must be
655: $l=(l_1+l_2-N)\ge \zeta N$ and $1\le n_1 <\ldots <n_l \le
656: N$ for which~(\ref{ineq3}) and~(\ref{ineq4}) both hold.
657: This means that for $1\le i \le l$ and $1\le k \le n_i$ we
658: have
659: $$
660: \prod_{j=n_i-k}^{n_i} \left\| Df (f^j_{\un
661: t}(x))^{-1} \right\| \le \al^k \qand
662: \dist_{\de_2}(f^{n_i-k}_{\un t}(x),\cc)\ge \al^{bk},
663: $$
664: and
665: hence these $n_i$ are $(\al,\de)$-hyperbolic times for
666: $(\un t, x)$, with $\de=\de_2$ and $\al=\exp(c/5)$. It
667: follows that for $\th_\ep^\NN\times m$ almost every $(\un
668: t,x)\in T^\NN\times M$ there are (positive frequency of)
669: times $n\in\ZZ^+$ for which
670: \begin{equation}\label{e.simple}
671: \prod_{j=n-k}^{n-1} \|Df(f^j_{\un t}(x))^{-1}\|\leq\al^k
672: \quad\mbox{and}\quad
673: \dist_\delta(f_{\un t}^{n-k}(x),\cc)\geq \al^{bk}
674: \end{equation}
675: for every $1\leq k\leq n$. Now the conclusion of the lemma
676: is a direct consequence of assumption (\ref{e.perturbation}).
677: \cqd
678:
679: \cre\label{r.hyp}
680: In the setting of
681: random perturbations of a local diffeomorphism $f$
682: we may also derive
683: from the first part of (\ref{e.simple}) the existence
684: of hyperbolic times for $\th_\ep^\NN\times m$
685: almost every $(\un t,x)\in T^\NN\times M$ without assuming
686: condition (\ref{e.perturbation}).
687: Actually, let $(\un t,x)$ be a point in $ T^\NN\times M$ for which the
688: first part of (\ref{e.simple}) holds. Taking the perturbations
689: $f_t$ in a sufficiently small $C^1$-neighborhood of $f$, then
690: $$
691: \|Df_t(y)^{-1}\|\leq \frac{1}{\sqrt\al}\|Df(y)^{-1}\|
692: $$
693: for every $y\in M$, which together with (\ref{e.simple}) gives
694: $$
695: \prod_{j=n-k}^{n-1} \|Df_t(f^j_{\un t}(x))^{-1}\|\leq
696: \prod_{j=n-k}^{n-1} \frac{1}{\sqrt\al}\|Df(f^j_{\un t}(x))^{-1}\|
697: \leq\al^{k/2}.
698: $$
699: In the context of maps with no critical sets this $n$ may
700: be defined as a $\sqrt\al$-hyperbolic time for $(\un t,x)$
701: and all the results that we present below hold with $\sqrt\al$-hyperbolic times replacing $(\al,\delta)$-hyperbolic times for maps with critical sets.
702: \fre
703:
704: Proposition \ref{p.hyp} allows us to introduce a map
705: $$
706: h_\ep \colon T^\NN\times M\ra \ZZ^+,
707: $$
708: by taking $h_\ep(\un t,x)$ as the first
709: $(\al,\delta)$-hyperbolic time for $(\un t,x)\in
710: T^\NN\times M$. We assume henceforth that the family
711: $(h_\ep)_{\ep>0}$ has uniform $L^1$-tail. For the next lemma we
712: fix $\de_1>0$ in such a way that
713: $4\delta_1<\min\{\de,\de^{\be}|\log \al|\}.$
714:
715: \cle \label{l.contr}
716: Given any $1\le j \le n$, we have
717: \begin{equation*}\label{l.contr1}
718: \|
719: Df(y)^{-1} \| \le \al^{-1/2} \| Df( f^{n-j}_{\un t}(x)
720: )^{-1} \|
721: \end{equation*}
722: for every $y$ in the ball of radius $2\de_1\al^{j/2}$ around
723: $f^{n-j}_{\un t}(x)$.
724: \fle
725: \dem We are assuming
726: $\dist_{\de}(f^{n-j}_{\un t}(x),\cc)\ge \al^j$ since $n$ is
727: a $(\al,\de)$-hyperbolic time for $(\un t,x)$. This means
728: that
729: $$
730: \dist (f^{n-j}_{\un t}(x),\cc)= \dist_{\de}(f^{n-j}_{\un
731: t}(x),\cc)\ge \al^{bj} \;\;\mbox{or else}\;\; \dist
732: (f^{n-j}_{\un t}(x),\cc)\ge \de.
733: $$
734: Either way it holds
735: $\dist(y,f^{n-j}_{\un t}(x)) \ge \dist (f^{n-j}_{\un
736: t}(x),\cc)/2$ because $b<1/2$ and $\de_1<\de/4<1/4$ for all
737: $y$ in the ball of radius $2\de_1\al^{j/2}$ around $
738: f^{n-j}_{\un t}(x)$. Therefore condition (S2) implies
739: $$ \log\frac{\| Df(y)^{-1}
740: \|}{\|Df(f^{n-j}_{\un t}(x))^{-1}\|} \le
741: B\frac{\dist(f^{n-j}_{\un t}(x),y)} {\dist(f^{n-j}_{\un
742: t}(x),\cc)^\be} \le B\frac{2\de_1\al^{j/2}}{\min\{\al^{b\be
743: j}, \de^\be\}}.
744: $$
745: But $\al,\de <1$ and $b\be<1/2$ so
746: $\al^{j/2}<\al^{b\be j}$ and thus the right hand side of
747: the last expression is bounded from above by
748: $2B\de_1\de^{-\be}$. The assumptions on $\de_1$ assure this
749: last bound to be smaller than $\log \al^{-1/2}$, which
750: implies the statement. \cqd
751:
752:
753:
754:
755: \cpr\label{p.contr} There is $\de_1>0$ such that if $n$ is
756: $(\al,\delta)$-hyperbolic time for $(\un t,x)\in
757: T^\NN\times M$, then there is a neighborhood $V_n(\un t,x)$
758: of $x$ in $M$ such that \begin{enumerate} \item $f_{\un
759: t}^n$ maps $V_n(\un t,x)$ diffeomorphically onto the ball
760: of radius $\de_1$ around $f_{\un t}^n(x)$; \item for every
761: $1\leq k\leq n$ and $y,z\in V_k(\un t,x)$
762: $$ \dist(f_{\un
763: t}^{n-k}(y),f_{\un t}^{n-k}(z)) \leq \al^{k/2}\dist(f_{\un
764: t}^{n}(y),f_{\un t}^{n}(z)).
765: $$
766: \end{enumerate}
767: \fpr
768: \dem
769: The proof will be by induction on $j\ge1$. First we show
770: that there is a well defined branch of $f^{-j}$ on a ball
771: of small enough radius around $f^j_{\un t}(x)$. Now we
772: observe that Lemma \ref{l.contr} gives for $j=1$
773: $$
774: \| Df(y)^{-1} \| \le \al^{-1/2}\|Df(f^{n-1}_{\un
775: t}(x))^{-1}\| \le \al^{1/2},
776: $$
777: because $n$ is a
778: $(\al,\de)$-hyperbolic time for $(\un t,x)$. This means
779: that $f$ is a $\al^{-1/2}$-dilation in the ball of radius
780: $2\de_1\al^{1/2}$ around $ f^{n-1}_{\un t}(x)$.
781: Consequently there is some neighborhood $V_1(\un t, x)$ of
782: $f^{n-1}_{\un t}(x)$ inside the ball of radius
783: $2\de_1\al^{1/2}$ that is diffeomorphic to the ball of
784: radius $\de_1$ around $ f^n_{\un t}(x) $ through $f_{t_n}$,
785: when $f$ is a map with critical set
786: satisfying~(\ref{e.perturbation}).
787:
788: For $j\ge1$ let us suppose that we have obtained a
789: neighborhood $V_j(\un t,x)$ of $f^{n-j}_{\un t}(x)$ such
790: that $f_{t_n}\circ\cdots\circ f_{t_{n-j+1}}\mid V_j(\un
791: t,x)$ is a diffeomorphism onto the ball of radius $\de_1$
792: around $ f^n_{\un t}(x) $ with
793: \begin{equation}\label{p.contr2}
794: \|Df(f_{t_{n-j+i+1}}\circ\cdots\circ
795: f_{t_{n-j+1}}(z))^{-1}\| \le
796: \al^{-1/2}\|Df(f^{n-j+i+1}_{\un t}(x))^{-1}\|
797: \end{equation} for all $z\in V_{j}(\un t,x)$ and $0\le
798: i<j$. Then, by Lemma \ref{l.contr} and under the assumption
799: that $n$ is a $(\al,\de)$-hyperbolic time for $x$,
800: \begin{eqnarray*}
801: \big\| D\big(f_{t_n}\circ\cdots\circ
802: f_{t_{n-j}}(y)\big)^{-1} \big\| &\le& \prod_{i=0}^j \big\|
803: Df_{t_{n-j+i}}\big (f_{t_{n-j+i-1}}\circ\cdots\circ
804: f_{t_{n-j}}(y)\big)^{-1} \big\| \\ &\le& \prod_{i=0}^j
805: \al^{-1/2} \big\| Df_{t_{n-j+i}}\big(f^{n-j+i-1}_{\un
806: t}(x)\big)^{-1} \big\| \\ &\le&
807: (\al^{-1/2})^{j+1}\cdot\al^{j+1}=\al^{(j+1)/2}
808: \end{eqnarray*}
809: for every $y$ on the ball of radius
810: $2\de_1\al^{(j+1)/2}$ around $f^{n-j-1}_{\un t}(x)$ whose
811: image $f_{t_{n-j}}(y)$ is in $V_{j}(\un t,x)$ (above we
812: convention $f_{t_{n-j+i-1}}\circ\cdots\circ
813: f_{t_{n-j}}(y)=y$ for $i=0$).
814:
815:
816: This shows that the derivative of $f_{t_n}\circ\cdots\circ
817: f_{t_{n-j}}$ is a $\al^{-(j+1)/2}$-dilation on the
818: intersection of
819: $
820: f^{-1}_{t_{n-j}}\big(V_{j}(\un t,x)\big)
821: $
822: with the ball of radius $2\de_1\al^{(j+1)/2}$ around
823: $ f^{n-j-1}_{\un t}(x)$,
824: and hence there is an
825: inverse branch of $f_{t_n}\circ\cdots\circ f_{t_{n-j}}$
826: defined on the ball of radius $\de_1$ around $f^n_{\un
827: t}(x) $. Thus we may define $V_{j+1}(\un t,x)$ as the image
828: of the ball of radius $\de_1$ around $ f^n_{\un t}(x) $
829: under this inverse branch, and recover the induction
830: hypothesis for $j+1$. In this manner we get neighborhoods
831: $V_{j}(\un t, x)$ of $f^{n-j}_{\un t}(x)$ as above for all
832: $1\le j\le n$. \cqd
833:
834: \cco\label{c.dist}
835: There is a constant $C_1>0$ such that if
836: $\un t\in T^\NN$, $n$ is a $(\al,\de)$-hyperbolic time for
837: $x\in M$ and $y,z\in V_n(\un
838: t,x)$, then
839: $$
840: \frac{1}{C_1}\leq \frac{|\det Df_{\un
841: t}^n(y)|}{|\det Df_{\un t}^n(z)|} \leq C_1.
842: $$
843: \fco
844: \dem
845: For $1\le k\le n$ the distance between $f^k_{\un t}(x)$ and
846: either $f^k_{\un t}(y)$ or $f^k_{\un t}(z)$ is smaller than
847: $\al^{(n-k)/2}$ which is smaller than $\al^{b(n-k)}\le
848: \dist(f^k_{\un t}(x),\cc)$. So, by (S3) we have
849: \begin{eqnarray*} \log\frac{|\det Df_{\un t}^n(y)|} {|\det
850: Df^n_{\un t}(z)|} &=& \sum_{k=0}^{n-1} \log \frac{|\det
851: Df_{t_{k+1}}(f^k_{\un t}(y))|} {|\det Df_{t_{k+1}}(f^k_{\un
852: t}(z))|}\\ & \le &\sum_{k=1}^{n-1}\log\frac{|\det
853: Df(f^k_{\un t}(y))|} {|\det Df(f^k_{\un t}(z))|} \\ &\le&
854: \sum_{k=0}^{n-1} 2B\frac{\al^{(n-k)/2}}{\al^{b\be(n-k)}},
855: \end{eqnarray*} and it is enough to take $C_1\le \exp
856: \left( \sum_{i=1}^\infty 2B \al^{(1/2-b\be)i} \right)$,
857: recalling that $b\be<1/2$ and also~(\ref{e.perturbation}).
858: \cqd
859:
860: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
861:
862:
863: \section{Stationary measures} \label{s.stationary}
864:
865: As mentioned before, we will assume the random
866: perturbations of the non-uniformly expanding map $f$
867: satisfy some {\em nondegeneracy conditions}: there
868: exists $0<\ep_0<1$ such that for every $0<\ep<\ep_0$ we may
869: take $n_0=n_0(\ep)\in\NN$ for which the following holds:
870:
871: \begin{enumerate}
872: \item there is $\xi=\xi(\ep)>0$ such
873: that $ \left\{ f^n_{\un t}(x) \colon \un{t}\in
874: (\supp\th_\ep)^{\NN} \right\} $ contains the ball of
875: radius~$\xi$ around $f^n(x)$ for all $x\in M$ and
876: $n\ge n_0$;
877: \item
878: $(f^n_x)_*{\th}^{\NN}_\ep\ll m$ for all $x\in M$ and
879: $n\ge n_0$.
880: \end{enumerate}
881: Here $(f^n_x)_*{\th}^{\NN}_\ep$
882: is the push-forward of ${\th}^{\NN}_{\ep}$ to $M$ via
883: $f^n_x:T^{\NN}\ra M$, defined as $f_x^n(\un t)=f_{\un t}^n(x)$.
884: %for $\un t\in T^\NN$.
885: Condition~1 means that
886: perturbed iterates cover a full neighborhood of the
887: unperturbed ones after a threshold for all sufficiently
888: small noise levels. Condition~2 means that sets of
889: perturbation vectors of positive ${\th}_{\ep}^{\NN}$
890: measure must send any point $x\in M$ onto subsets of $M$
891: with positive Lebesgue measure after a finite number of
892: iterates.
893:
894: In~\cite[Examples 1 \& 2]{Ar1} it was shown that given any
895: smooth map $f:M\ra M$ of a compact manifold we can always
896: construct a random perturbation satisfying the
897: nondegeneracy conditions 1 and 2, if we take $T=\RR^p$,
898: $t^*=0$ and $\th_\ep$ is equal to the normalized
899: restriction of the Lebesgue measure to the ball of radius
900: $\ep$ around $0$, for a sufficiently big number $p\in\NN$
901: of parameters. For parallelizable manifolds the random
902: perturbations which consist in adding at each step a
903: random noise to the unperturbed dynamics, as described in
904: the Introduction, clearly satisfy nondegeneracy conditions
905: 1 and 2 for $n_0=1$.
906:
907: \medskip
908:
909: In the context of random perturbations of a map, we say that a set $A\subset M$ is {\em invariant} if $f_t(A)\subset A$, at least
910: for $t\in\supp(\th_\ep)$ with $\ep>0$ small.
911: The usual invariance of a measure with respect to a
912: transformation is replaced by the following one: a
913: probability measure $\mu$ is said to be \emph{stationary},
914: if for every continuous $\vfi:M\to\RR$ it holds
915: \begin{equation}
916: \label{eq.stationary}
917: \int\vfi\,d\mu=
918: \int\int\vfi\big(f_t(x)\big)\,d\mu(x)\,d\th_\ep(t).
919: \end{equation}
920:
921: \cre\label{re.accinvariant}
922: If $(\mu^\ep)_{\ep>0}$ is a
923: family of stationary measures having $\mu_0$ as a weak$^*$
924: accumulation point when $\ep$ goes to $0$, then it follows
925: from~(\ref{eq.stationary}) and the convergence of
926: $\supp(\th_\ep)$ to $\{t^*\}$ that $\mu_0$ must be
927: invariant by $f=f_{t^*}$.
928: \fre
929: It is not difficult to see (cf.~\cite{Ar1}) that a
930: stationary measure $\mu$ satisfies
931: $$
932: x\in\supp(\mu)
933: \quad\Rightarrow\quad f_t(x)\in\supp(\mu) \quad\mbox{for
934: all} \quad t\in\supp(\th_\ep)
935: $$
936: just by continuity of
937: $\Phi$. This means that if $\mu$ is a
938: stationary measure, then $\supp(\mu)$ is an invariant set.
939: Nondegeneracy condition 1 ensures that the interior of
940: $\supp(\mu)$ is nonempty.
941:
942: Let us write $\supp(\mu)$ as a disjoint union $\bigcup_i C_i$
943: of connected components and consider only those $C_i$ for
944: which $m(C_i)>0$ -- this collection is nonempty
945: %\marginpar{??}
946: since
947: $\supp(\mu)$ contains open sets. Moreover each $f_t$ must
948: permute these components for $t\in\supp(\th_\ep)$, because
949: $f_t(C_i)$ is connected by continuity,
950: $f_t(C_i)\subset\supp(\mu)$ by invariance, and
951: $m(f_t(C_i))>0$ since we have $(f_t)_*m\ll m$.
952:
953: The connectedness of $C_i$ and continuity of $\Phi$ guarantee that
954: the abovemention perturbation of the components $C_i$
955: induced by $f_t$ does not depend on $t\in\supp(\th_\ep)$.
956: Indeed, supposing that $t,t^\prime\in\supp(\th_\ep)$ are
957: such that $$ f_t(C_i)\subset C_j \qand
958: f_{t^\prime}(C_i)\subset C_{j^\prime}, $$ then fixing some
959: $z\in C_i$ we have that $\{ f_t(z)\colon t\in\supp(\th_\ep)\}$
960: is a connected set intersecting both $C_j$ and
961: $C_{j^\prime}$ inside $\supp(\mu)$, and so $C_j =
962: C_{j^\prime}$.
963:
964:
965: We will show that these connected components are periodic
966: under the action induced by $f_t$ with $
967: t\in\supp(\th_\ep)$. After this, we may use nondegeneracy
968: condition 1 to conclude that each component contains a ball
969: of uniform radius and thus that each component satisfies
970: $m(C_i)> \mbox{const} >0$. Hence there existing only a
971: finite number of such components.
972:
973: At this point it is useful to introduce the skew-product map
974: \[
975: \begin{array}{rccc} F: & T^\NN\times M &\longrightarrow &
976: T^\NN\times M\\
977: & (\un t, z) &\longmapsto & \big(\sigma(\un t),f_{t_1}(z)\big)
978: \end{array}
979: \]
980: where $\sigma$ is the left shift on
981: sequences $\un t=(t_1,t_2,\dots)\in T^\NN$.
982: It is easy to check that the product measure $\th_\ep^\NN\times\mu$
983: is $F$-invariant, as so is the set
984: $\supp(\th_\ep^\NN\times\mu)=\supp(\th_\ep)^\NN\times\supp(\mu)$.
985:
986:
987: \cle \label{pr.cycles}
988: The support of a stationary measure
989: $\mu$ contains a finite number of connected components
990: arranged in cycles permuted by the action of $f_t$ for
991: $t\in\supp(\th_\ep)$.
992: \fle
993: %We remark that this \emph{does not say} that the
994: %support has \emph{only} a finite number of
995: %connected components.
996: \dem
997: Is is enough to obtain that each connected component $C_i$
998: is periodic under the action of $f_t$ for $t\in\supp(\th_\ep)$, in
999: the sense that $f^p_{\un t}(C_i)\subset C_i$ for some
1000: $p\in\NN$ and all $\un t\in\supp(\th_\ep^\NN)$.
1001: There are components $C_i$ with nonempty interior,
1002: since the interior of $\supp(\mu)$ is nonempty.
1003: So we may take a component
1004: $C_i$ that contains some ball $B$.
1005: Then
1006: we have $m(B)>0$ and so
1007: $(\th_\ep^\NN\times\mu)(\supp(\th_\ep^\NN)\times B)>0$.
1008: Poincar\'e Recurrence Theorem now guarantees there is
1009: $(\un t,x)\in \supp(\th_\ep^\NN)\times B$
1010: such that the $F$-orbit
1011: of $(\un t,x)$ has the same $(\un t,x)$ as an accumulation
1012: point. We see that there must exist some $p\in\NN$ such
1013: that $f^p_{\un t}(x)\in B\subset C_i$.
1014: In view of the independence of the permutation on the
1015: choice of $\un t$, we conclude that $C_i$ is sent inside
1016: itself by $f_{\un t}^p$ for all $\un
1017: t\in\supp(\th_\ep^\NN)$. \cqd
1018:
1019: It is clear that the cycles obtained above are invariant
1020: sets.
1021: We are now ready to decompose $\mu$ into some simpler
1022: measures. For that we need the following result.
1023:
1024: \cle \label{l.restrict} The normalized restriction of a
1025: stationary measure to an invariant set is a
1026: stationary measure.
1027: \fle
1028: \dem See \cite[Lemma~8.2]{Ar1}. \cqd
1029:
1030: We define an {\em invariant domain} in $M$ as a finite
1031: collection $(U_0,\dots,U_{p-1})$ of pairwise separated open
1032: sets, that is, $\ov{U}_i\cap\ov{U}_j = \emptyset$ if $i\neq
1033: j$, such that
1034: $f^k_{\un t}(U_i)\subset U_{(k+i)\bmod p}$
1035: for all $k\ge1$,
1036: $i=0,\dots,p-1$ and $\un t\in\supp(\th_\ep^\NN)$.
1037:
1038: In order to get the separation of the connected
1039: components in a cycle, we may unite those components $C_i$
1040: and $C_j$ such that $\ov{C}_i\cap\ov{C}_j\neq\emptyset$ and
1041: observe that the permutation now induced in the new sets by
1042: $f_t$ also does not depend on the choice of
1043: $t\in\supp(\th_\ep)$. In this manner we construct invariant
1044: domains inside the support of any stationary probability
1045: measure.
1046:
1047: The next step is to look for \emph{minimal invariant
1048: domains} with respect to the natural order relation of
1049: inclusion of sets.
1050: Let $D=(U_0,\dots,U_{p-1})$ and $D'=(W_0,\dots,W_{q-1})$ be
1051: invariant domains.
1052: On the one hand, $D=D'$ if there are $i,j\in\NN$ such that
1053: $U_{(i+k) \bmod p} = W_{(j+k) \bmod q}$ for all $k\ge1$,
1054: which implies $p=q$ because the open sets that form each
1055: invariant domain are pairwise disjoint. On the other hand,
1056: we say $D\prec D'$ if there are $i,j\in\NN$ such that
1057: $U_{i\bmod p} \subsetneq W_{j\bmod q}$
1058: and $U_{(i+k) \bmod p} \subset W_{(j+k)
1059: \bmod q}$ for all $k\ge1$.
1060:
1061: \cle \label{l.minimal}
1062: In the partially ordered family of
1063: all invariant domains in $M$, with respect to the relation
1064: $\prec$, the number of $\prec$-minimal domains is finite.
1065: Moreover, every invariant domain contains at least one
1066: minimal domain.
1067: \fle
1068:
1069: \dem The proof relies in showing that Zorn's Lemma can be
1070: applied to this partially ordered set and that minimal
1071: domains are pairwise separated. See~\cite[Section~3]{Ar1}.
1072: \cqd
1073:
1074:
1075:
1076:
1077: Let us now fix $x\in M$ and consider
1078: \begin{equation}
1079: \label{pushforward1}
1080: \mu_n(x)=\frac1n\sum_{j=0}^{n-1} (f^j_x)_*\th_\ep^\NN.
1081: \end{equation}
1082: Since this is a sequence of probability measures on the
1083: compact manifold $M$, then it has weak$^*$ accumulation
1084: points.
1085:
1086:
1087: \cle \label{p.muinvariant}
1088: Every weak$^*$ accumulation point
1089: of $\big(\mu_n(x)\big)_n$ is stationary and absolutely
1090: continuous with respect to the Lebesgue measure.
1091: \fle
1092: \dem Let $\mu$ be a weak$^*$ accumulation point of
1093: $\big(\mu_n(x)\big)_n$. We may write
1094: $$
1095: \int \int \vfi \big( f_t(x) \big) \,
1096: d\mu(x) \, d\th_\ep(t) = \int \lim_{k\to+\infty}
1097: \frac1{n_k} \sum_{j=0}^{n_k-1} \int \vfi \left(
1098: f_t\big(f_{\un t}^j(x)\big) \right) \, d\th_\ep^\NN(\un
1099: t)\,d\th_\ep(t)
1100: $$
1101: for each continuous $\vfi:M\to\RR$.
1102: Moreover dominated convergence ensures that we may exchange
1103: the limit and the outer integral sign and, by definition of
1104: $f^j_{\un t}(x)$, we get
1105: $$
1106: \lim_{k\to\infty} \frac1{n_k}
1107: \sum_{j=0}^{n_k-1} \int \vfi \big( f_{\un t}^{j+1}(x)\big)
1108: \, d\th_\ep^\NN(\un t) = \int \vfi \, d\mu,
1109: $$
1110: according to the definition of $\mu$. Thus
1111: (\ref{eq.stationary}) must hold and $\mu$ is stationary.
1112:
1113: Noting that $C^0(M,\RR)$ is dense in $L^1(M,\mu)$ with the
1114: $L^1$ norm, we see that (\ref{eq.stationary}) holds for all
1115: $\mu$-integrable functions $\vfi:M\to\RR$. In particular,
1116: if $E\subset M$ is such that $m(E)=0$, then
1117: \begin{eqnarray*}
1118: \int 1_E \,d\mu
1119: &=&\int\int 1_E\big(f_{t}(x)\big) \,d\mu(x)\,d\th_\ep(t) \\
1120: &=&\int\int 1_E\big(f_{t}(x)\big) \,d\th_\ep(t)\,d\mu(x) \\
1121: &=&\int\int\int1_E\big(f_{t}(f_{s}(x))\big)\,
1122: d\th_\ep(t)\,d\mu(x)\,d\th_\ep(s)\\
1123: &=&\int\int 1_E\big(f^2_{\un t}(x)\big)\,d\th_\ep^\NN(\un t)\,d\mu(x)\\
1124: &=&\int (f^2_x)_*\th_\ep^\NN(E) \,d\mu(x).
1125: \end{eqnarray*}
1126: This process may be iterated to yield
1127: $$
1128: \mu(E)=\int (f^{n_0}_x)_*\th_\ep(E)\,d\mu(x)
1129: $$
1130: and, since $(f^{n_0}_x)_*\th_\ep\ll m$ by nondegeneracy
1131: condition 2, we must have $\mu(E)=0$.
1132: \cqd
1133:
1134:
1135: Clearly if $x\in M$ belongs to some set of an
1136: invariant domain $(U_0,\dots,U_{p-1})$,
1137: then $\mu_n(x)$ have supports contained in
1138: $\ov{U}_0\cup\dots\cup\ov{U}_{p-1}$ for all $n\ge1$
1139: and any weak$^*$ accumlation point $\mu$ of
1140: $(\mu_n(x))_n$ is a stationary measure with
1141: $\supp(\mu)\subset\ov{U}_0\cup\dots\cup\ov{U}_{p-1}$.
1142: We will now see these measures are physical.
1143:
1144:
1145:
1146: \cle
1147: \label{l.minimal=physical}
1148: If $(U_0,\dots,U_{p-1})$ is a minimal invariant
1149: domain, then there is a unique absolutely
1150: continuous stationary measure $\nu$ such that
1151: $\supp(\nu)\subset\ov{U}_0\cup\dots\cup\ov{U}_{p-1}$.
1152: Moreover, this $\nu$ is a physical measure
1153: and $\supp(\nu)=\ov{U}_0\cup\dots\cup\ov{U}_{p-1}$.
1154: \fle
1155:
1156: \dem Let us assume $n_0=1$ for simplicity
1157: (see~\cite[Section~7]{Ar1} for the general case) and let us
1158: consider a stationary absolutely continuous probability
1159: measure $\nu$ with
1160: $\supp(\nu)\subset\ov{U}_0\cup\dots\cup\ov{U}_{p-1}$. We
1161: first show the ergodicity of $\nu$, in the sense that
1162: $\th_\ep^\NN\times\nu$ is $F$-ergodic. It turns out that to
1163: be $F$-ergodic it suffices that either $\nu(G)=0$ or
1164: $\nu(G)=1$ for every Borel set $G\subset M$
1165: satisfying
1166: \begin{equation}
1167: \label{eq.ergodic}
1168: 1_G(x)=\int 1_G \left(f_t(x)\right)\,d\th_\ep(t)
1169: \end{equation} for $\nu$ almost every $x$ (cf.~\cite{Ar1}
1170: and~\cite{V2}). So let us take $G$ such that $\nu(G)>0$ and
1171: $G$ satisfies the left hand side of~(\ref{eq.ergodic}).
1172: Then it must be $m(G)>0$ because $\nu\ll m$ and there is a
1173: closed set $J\subset G$ such that $m(G\setminus J)=0$ and
1174: also $\nu(G\setminus J)=0$. Hence $J$ also satisfies the
1175: left hand side of~(\ref{eq.ergodic}) because of
1176: nondegeneracy condition 2 (with $n_0=1$), since
1177: $$
1178: \int 1_E(f_t(x))\,d\th_\ep(t)=(f_x)_*\th_\ep^\NN(E).
1179: $$
1180: This means that when $x\in J$ we have $f_t(x)\in J$ for
1181: $\th_\ep$ almost all $t\in\supp(\th_\ep)$. Since a set of
1182: $\th_\ep$ measure 1 is dense in $\supp(\th_\ep)$ (we are
1183: supposing $\th_\ep$ to be positive on open sets) and
1184: $f_t(x)$ varies continuously with $t$,
1185: we see that $f_t(x)\in J$ for all $t\in\supp(\th_\ep)$
1186: because $J$ is closed.
1187: We then have that the interior of $J$ is nonempty by
1188: condition 1 on random perturbations and we may apply the
1189: methods of decomposition into connected components as
1190: before (Lemma~\ref{pr.cycles}). In this manner we
1191: construct an invariant domain inside $J$ which, in turn, is
1192: inside a minimal invariant domain. This contradicts
1193: minimality and so we conclude that $J$ must contain
1194: $\ov{U}_0\cup\dots\cup\ov{U}_{p-1}$. Thus we have
1195: $\nu(G)=\nu(J)=1$ proving $\th_\ep^\NN\times\nu$ to be
1196: $F$-ergodic.
1197:
1198: Now, given $\vfi:M\to\RR$ continuous we consider the map
1199: $\psi=\vfi\circ\pi$ from $T^\NN\times M$ to $\RR$,
1200: where $\pi:T^\NN\times M\to M$ is
1201: the natural projection. The Ergodic Theorem then ensures
1202: $$
1203: \lim_{n\to+\infty}\frac1n\sum_{j=0}^{n-1}
1204: \psi(F^j(\un t,x))=\int \psi \, d(\th_\ep^\NN\times\nu)
1205: $$
1206: for $\theta_\epsilon^\NN\times\nu$ almost all $ (\un t,x)$,
1207: which is just the same as
1208: \begin{equation}
1209: \label{eq.ergodic2}
1210: \lim_{n\to+\infty}\frac1n\sum_{j=0}^{n-1}
1211: \vfi(f^j_{\un t}(x))=\int\vfi\,d\nu
1212: \end{equation}
1213: for $\theta_\epsilon^\NN\times\nu$ almost all $ (\un
1214: t,x)$.
1215: Finally considering the ergodic basin
1216: $B(\nu)$, defined as the set of points $ x\in M$ for which
1217: $$
1218: \lim_{n\to+\infty}
1219: \frac1n\sum_{j=0}^{n-1}\vfi(f^j_{\un t}(x))=
1220: \int\vfi\,d\nu
1221: $$
1222: for all $\vfi\in C^0(M,\RR)$ and
1223: $\th_\ep^\NN$ almost every $\un t\in T^\NN$,
1224: it is easy to
1225: see that $B(\nu)$ satisfies
1226: (\ref{eq.ergodic}) in the place of $G$ and we must have
1227: as before $B(\nu)\supset \ov{U}_0\cup\dots\cup\ov{U}_{p-1}$.
1228:
1229: This shows that if another stationary absolutely continuous
1230: probability measure $\tilde \nu$ is such that $\supp(\tilde
1231: \nu)\subset \ov{U}_0\cup\dots\cup\ov{U}_{p-1}$, then the
1232: basins of $\nu$ and $\tilde \nu$ must have nonempty intersection.
1233: Thus these
1234: measures must be equal. Moreover $\nu\big(B(\nu)\big)=1$ and so, by
1235: absolute continuity, $m\big(B(\nu)\big)>0$ and thus $\nu$ is a physical
1236: probability. \cqd
1237:
1238:
1239: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1240:
1241: \section{The number of physical measures}
1242: \label{s.physical}
1243:
1244: In this section we will prove that the number $l$ of physical
1245: measures is bounded by the number $p$ of SRB measures. Moreover we will present examples of dynamical systems for which $l=p$ and $l<p$.
1246:
1247: Let $\mu_1,\ldots,\mu_l$ be the physical measures supported
1248: on the minimal invariant domains in $M$, which exist by
1249: Lemmas~\ref{pr.cycles} and \ref{l.minimal}
1250: through~\ref{l.minimal=physical}. If $\mu$ is an absolutely
1251: continuous stationary measure, its restrictions to the
1252: minimal invariant domains of $M$, normalized when not equal
1253: to the constant zero measure, are absolutely continuous
1254: stationary measures by Lemma~\ref{l.restrict}. After
1255: Lemma~\ref{l.minimal=physical} these restrictions must be
1256: the physical measures $\mu_1,\ldots,\mu_l$ of the minimal
1257: domains. Hence $\mu$ must decompose into a linear
1258: combination of physical measures.
1259: Moreover, the union of
1260: $\supp(\mu_1),\dots,\supp(\mu_l)$ must contain
1261: $\supp(\mu)$, except possibly for a $\mu$ null set. In
1262: fact, if the following set function
1263: $$
1264: \mu-\mu\big(\supp(\mu_1)\big)\mu_1-\cdots
1265: -\mu\big(\supp(\mu_l)\big)\mu_l
1266: $$
1267: were nonzero, then its normalization $\mu^\prime$ would be
1268: an absolutely continuous stationary measure, and the above
1269: decomposition could be applied to $\mu^\prime$,
1270: thus giving another minimal domain inside $\supp(\mu)$.
1271: Clearly this cannot happen. We then have a convex linear
1272: decomposition
1273: \begin{equation}
1274: \label{eq.decomposition}
1275: \mu=\al_1\mu_1+\dots+\al_l\mu_l
1276: \end{equation}
1277: where $\al_i=\mu(\supp(\mu_i))\ge0$ and
1278: $\al_1+\dots+\al_l=1$. We will see that this decomposition
1279: is uniquely defined.
1280:
1281: \medskip
1282:
1283:
1284: We remark that so far we did not use more than the continuity
1285: of the map $f$.
1286: %\marginpar{verdade?}
1287: For the next result we assume that $f:M \to M$ is a $C^2$
1288: non-uniformly expanding map whose orbits have slow
1289: approximation to the critical $\cc$ (possibly the emptyset)
1290: with $m(\cc)=0$. This result contains the assertions of the
1291: first two items of Theorem~A (if we think of
1292: $\cc=\emptyset$) and Theorem~C.
1293:
1294: \cpr
1295: \label{p.finitemeasures}
1296: If $\ep>0$ is small enough, then there exist physical measures
1297: $\mu^\ep_1,\dots,\mu^\ep_l$
1298: (with $l$ not depending on $\ep$) such that
1299: \begin{enumerate}
1300: \item for $x\in M$ there is a $\th_\ep^\NN$ mod $0$
1301: partition $T_1(x),\dots,T_l(x)$ of $T^\NN$ such that
1302: $$
1303: \mu_i^\ep=\wlim \frac1n \sum_{j=1}^{n-1} \de_{f_{\un t}^j(x)}
1304: \quad \mbox{ if and only if } \quad \un t \in T_i(x);
1305: $$
1306: \item for each $i=1,\dots,l$ we have
1307: $$
1308: \mu_i^\ep=\wlim\frac1n\sum_{j=0}^{n-1}\int (f^j_{\un t})_*
1309: \big(m\mid B(\mu_i^{\ep})\big) \, d\th_\ep^\NN(\un t),
1310: $$
1311: where $m\mid B(\mu_i^{\ep})$ is the normalized
1312: restriction of Lebesgue measure to $B(\mu_i^{\ep})$.
1313: \end{enumerate}
1314: \fpr
1315: \dem
1316: Take $x\in M$ and let $\mu$ be a weak$^*$ accumulation
1317: point of the sequence $(\mu_n(x))_n$ defined
1318: in~(\ref{pushforward1}). We will prove that this is the
1319: only accumulation point of~(\ref{pushforward1}) by showing
1320: that the values of the $\al_1,\dots,\al_l$ in
1321: decomposition (\ref{eq.decomposition}) depend only on $x$
1322: and not on the subsequence that converges to $\mu$. The
1323: definition of the average in~(\ref{pushforward1}) implies
1324: that there is a subset
1325: of parameter vectors $\un t\in\supp(\th_\ep^\NN)$
1326: with positive $\th_\ep^\NN$ measure
1327: for which there is $j\ge1$ such that $f^j_{\un
1328: t}(x)\in\supp(\mu_i)$.
1329: We define for $i=1,\dots,l$
1330: $$
1331: T_i(x)=
1332: \left\{ \un t\in\supp(\th_\ep^\NN): f^j_{\un t}(x) \in
1333: \supp(\mu_i) \quad\mbox{for some}\quad j\ge1 \right\}.
1334: $$
1335: We clearly have
1336: $$
1337: T_i(x)=\mbox{$\bigcup$}_{j\ge1} T_i^j(x) \quad\mbox{where}\quad T_i^j(x)=
1338: \{ \un t\in\supp(\th_\ep^\NN): f^j_{\un t}(x)\in \supp(\mu_i) \}
1339: $$
1340: and $T_i^j(x)\subset T_i^{j+1}(x)$ for all $i,j\ge1$, since the supports of
1341: stationary measures are themselves invariant. In addition,
1342: since $\mu$ is a regular (Borel) probability measure, we
1343: may find for each $\eta>0$ an open set $U$ and a closed set $K$
1344: such that $K\subset\supp(\mu_i)\subset U$ with
1345: $\mu(U\setminus K)<\eta$ and $\mu(\partial U)=\mu(\partial
1346: K)=0$. In fact, there is an at most countable number of
1347: $\delta$-neighborhoods of $\supp(\mu_i)$ whose boundaries have
1348: positive $\mu$ measure, and likewise for the compacts coinciding with the complement of the $\delta$-neighborhood of $M\setminus\supp(\mu_i)$. Then,
1349: taking $\alpha_i=\mu(\supp(\mu_i))$ we have
1350: \begin{eqnarray*}
1351: \al_i+\eta\ge \mu(U) &=&
1352: \lim_{k\to+\infty}\frac1{n_k}\sum_{j=0}^{n_k-1}
1353: \th_\ep^\NN\{ \un t\in T^\NN: f^j_{\un t}(x)\in U \} \\ &
1354: \ge & \limsup_{k\to+\infty} \frac1{n_k}\sum_{j=0}^{n_k-1}
1355: \th_\ep^\NN\big(T_i^j(x)\big) \end{eqnarray*} for some sequence of
1356: integers $n_1<n_2<n_3<\cdots$, and likewise for
1357: \begin{eqnarray*} \al_i-\eta\le \mu(K)
1358: &=&
1359: \lim_{k\to+\infty}\frac1{n_k}\sum_{j=0}^{n_k-1}
1360: \th_\ep^\NN\{ \un t\in T^\NN:
1361: f^j_{\un t}(x)\in K \}
1362: \\
1363: & \le &
1364: \liminf_{k\to+\infty}
1365: \frac1{n_k}\sum_{j=0}^{n_k-1}
1366: \th_\ep^\NN\big(T_i^j(x)\big),
1367: \end{eqnarray*}
1368: where $\eta>0$ is arbitrary.
1369: This shows
1370: $$
1371: \al_i=\mu(\supp(\mu_i))= \lim_{k\to\infty}
1372: \frac1{n_k}\sum_{j=0}^{n_k-1} \th_\ep^\NN\big(T_i^j(x)\big).
1373: $$
1374: We also have
1375: $$
1376: \th_\ep^\NN\big(T_i(x)\big)=\lim_{j\to\infty}\th_\ep^\NN\big(T_i^j(x)\big)
1377: =\lim_{n\to\infty}\frac1n\sum_{j=0}^{n-1}
1378: \th_\ep^\NN\big(T_i^j(x)\big) =\al_i
1379: $$
1380: which shows that the $\al_i$ depend only on the random
1381: orbits of $x$ and not on the particular sequence $(n_k)_k$.
1382: Thus we see that the sequence of measures in
1383: (\ref{pushforward1}) converges in the weak$^*$ topology.
1384: Moreover the sets $T_1(x),\dots,T_l(x)$ are pairwise disjoint by
1385: definition and their total $\th_\ep^\NN$ measure equals
1386: $\al_1+\cdots+\al_l=1$, thus forming a $\th_\ep^\NN$ modulo
1387: zero partition of $T^\NN$. We observe that if $\un t\in
1388: T_i(x)$, then $f^n_{\un t}(x)\in \supp(\mu_i)\subset
1389: B(\mu_i)$ for some $n\ge1$ and $i=1,\dots,l$. This means
1390: this $\th_\ep^\NN$ modulo zero partition of $T^\NN$
1391: satisfies the first item of the proposition.
1392:
1393: Now fixing $i=1,\dots,l$, for all $x\in B(\mu_i)$ (the
1394: ergodic basin of $\mu_i$) it holds that $$
1395: \lim_{n\to+\infty}\frac1n\sum_{i=0}^{n-1} \vfi(f_{\un
1396: t}^j(x)) = \int \vfi\, d\mu_i $$ for $\th_{\ep}^\NN$ almost
1397: every $\un t\in T^\NN$. Recall that $m(B(\mu_i))>0$ by the
1398: definition of physical measure. Using dominated convergence
1399: and integrating both sides of the above equality twice,
1400: first with respect to the Lebesgue measure $m$, and then
1401: with respect to $\th_{\ep}^\NN$, we arrive at the statement
1402: of item 2.
1403:
1404: Recall that up until now the noise level $\ep>0$ was kept
1405: fixed. For small enough $\ep>0$ the measures
1406: $\mu_i=\mu_i^\ep$ depend on the noise level, but we will
1407: see that the number of physical measures is constant.
1408:
1409: Fixing $i\in\{1,\dots,l\}$ we let $x$ in the interior of
1410: $\supp(\mu_i^\ep)$ be such that the orbit $(f^j(x))_j$ has
1411: infinitely many hyperbolic times. Recall that $f\equiv
1412: f_{t^*}$ is non-uniformly expanding (possibly with
1413: criticalities). Then there is a big enough hyperbolic time
1414: $n$ so that $V_n(\un t^*, x)\subset\supp(\mu_i^\ep)$, by
1415: Proposition~\ref{p.contr}, where we take $\un
1416: t^*=(t^*,t^*,t^*,\dots)$. Since $t^*\in\supp(\th_\ep)$ and
1417: $\supp(\mu_i^\ep)$ is invariant under $f_t$ for all
1418: $t\in\supp(\th_\ep)$, we must have
1419: $$
1420: f^n_{\un t^*}\big(V_n(\un
1421: t^*,x)\big)= B\big(f^n_{t^*}(x),\de_1\big)
1422: \subset\supp(\mu_i^\ep),
1423: $$
1424: where $\de_1>0$ is the constant given by
1425: Proposition~\ref{p.contr} and
1426: $B\big(f^n_{t^*}(x),\de_1\big)$ is the ball of radius
1427: $\de_1$ around $f^n_{t^*}(x)$.
1428:
1429: On the
1430: one hand, we deduce that the number $l=l(\ep)$ is bounded
1431: from above by some uniform constant $N$ since $M$ is
1432: compact. On the other hand, since each invariant set must
1433: contain some physical measure (by Lemma~\ref{l.minimal}),
1434: we see that for $0<\ep^\prime<\ep$ there must be some
1435: physical measure $\mu^{\ep^\prime}$ with
1436: $\supp(\mu^{\ep^\prime})\subset\supp(\mu^\ep)$. In fact
1437: $\supp(\mu^\ep)$ is invariant under $f_t$ for every
1438: $t\in\supp(\th_{\ep^\prime})\subset\supp(\th_\ep)$. This
1439: means the number $l(\ep)$ of physical measures is a
1440: nonincreasing function of $\ep>0$. Thus we conclude that
1441: there must be $\ep_0>0$ such that $l=l(\ep)$ is constant
1442: for $0<\ep<\ep_0$, ending the proof of the proposition.
1443: \cqd
1444:
1445: \cre \label{r.transitivity} Observe that if the map $f:M\to
1446: M$ is transitive, then every stationary measure must be
1447: supported on the whole of $M$, since the support is
1448: invariant and has nonempty interior. According to the
1449: discussion above, there must be only one such stationary
1450: measure, which must be physical. \fre
1451:
1452: We note that the number $l$ of physical measures for small
1453: $\ep>0$ and the number $p$ of SRB measures for $f$ are
1454: obtained by different existential arguments. It is natural
1455: to ask if there is any relation between $l$ and $p$.
1456:
1457: \cpr If $p\ge1$ is the number of SRB measures of $f$ and
1458: $l\ge1$ is the number of physical measures of the random
1459: perturbation of $f$, then for $\ep>0$ small enough we have $l\le
1460: p$.
1461: \fpr
1462: \dem
1463: We start by observing that if $p=1$ then every weak$^*$
1464: accumulation point of a family $(\mu_i^\ep)_{\ep>0}$ of
1465: physical measures when $\ep\to0$ must equal the unique SRB
1466: measure $\mu_1$ for $f$. Hence the weak$^*$ limit of
1467: $(\mu_i^\ep)_{\ep>0}$ when $\ep\to0$ exists and equals
1468: $\mu_1$ for all $i=1,\ldots,l$. Then there must be a single
1469: physical measure $\mu_1^\ep$ for all small enough $\ep>0$.
1470: In fact, let us assume there are distinct families
1471: $(\mu_1^\ep)_{\ep>0}$ and $(\mu_2^\ep)_{\ep>0}$ of physical
1472: measures as given by Proposition~\ref{p.finitemeasures}.
1473: Take $x\in\supp(\mu_1)$ and a sequence of continuous maps
1474: $\vfi_n:M\to\RR$ such that
1475: $$ \vfi_n\ge0\quad,\quad
1476: \vfi_n\mid B(x, 1/n)\equiv 1 \qand \vfi_n \mid \big(
1477: M\setminus B(x,2/n)\big) \equiv 0,
1478: $$ for all big $n\in\NN$, where $B(x,r)$ denotes the ball of radius
1479: $r$ around $x$ for each $r\ge0$. If we fix $n$, then $\mu_1(\vfi_n)>0$. Thus for all
1480: small enough $\ep>0$ we must have $\mu_1^\ep(\vfi_n)>0$ and
1481: $\mu_2^\ep(\vfi_n)>0$. Hence the distance between
1482: $\supp(\mu_1^\ep)$ and $\supp(\mu_2^\ep)$ is smaller than
1483: $2/n$.\ Since $n$ may be arbitrarily large we see that
1484: $\supp(\mu_1^\ep)$ and $\supp(\mu_2^\ep)$ get arbitrarily
1485: close when $\ep\to0$. Thus $\mu_1^\ep$ and $\mu_2^\ep$ must
1486: coincide for small $\ep>0$ because
1487: Proposition~\ref{p.finitemeasures} and its proof show that
1488: these families, when distinct, are at a uniform distance
1489: apart (since $\{\supp(\mu_1^\ep)\}_{\ep>0}$ and
1490: $\{\supp(\mu_2^\ep)\}_{\ep>0}$ are nested families of
1491: pairwise disjoint compact sets).
1492:
1493: In general, if $p>1$ then the weak$^*$ accumulation points
1494: of a family $(\mu_i^\ep)_{\ep>0}$ for $i=1,\ldots, l$
1495: when $\ep\to0$ are convex linear combinations
1496: $\al_1\mu_1+\cdots+\al_p\mu_p$ of the $p$ SRB measures of
1497: $f$. The preceding argument implies that two distinct
1498: families $(\mu_i^\ep)_{\ep>0}$ and $(\mu_j^\ep)_{\ep>0}$ of
1499: physical measures cannot have weak$^*$ accumulation points
1500: expressed as linear convex combinations $$
1501: \al_1\mu_1+\cdots+\al_p\mu_p\qand
1502: \al_1^\prime\mu_1+\cdots+\al_p^\prime\mu_p $$ with both
1503: $\al_k$ and $\al_k^\prime$ nonzero for some $k=1,\ldots,p$
1504: (take $x\in\supp(\mu_k)$ and repeat the arguments in the
1505: above paragraph). Hence $l\le p$. \cqd
1506:
1507: The reverse inequality does not hold in general, as the
1508: following examples show: it is possible for two distinct
1509: SRB measures to have intersecting supports and, in this
1510: circumstance, the random perturbations will mix their
1511: basins and there will be some physical measure whose
1512: support overlaps the supports of both SRB measures.
1513:
1514: \begin{figure}[htb]
1515: \begin{center}
1516: \includegraphics[width=8cm]{parabolas.ps}
1517: \caption{\label{fig.1}
1518: map for which $1=l<p=2$}
1519: \end{center}
1520: \end{figure}
1521:
1522: \medskip
1523:
1524: The first example is the map $f:[-3,1]\to[-3,1]$ whose
1525: graph is figure~\ref{fig.1}: $$ f(x)= \left\{
1526: \begin{array}{lcl} 1-2x^2 & \mbox{if} & -1\le x \le 1 \\
1527: 2(x+2)^2 -3 & \mbox{if} & -3 \le x \le -1 \end{array}
1528: \right. . $$ The dynamics of $f$ on $[-1,1]$ and $[-3,-1]$
1529: is conjugated to the tent map $T(x)=
1530: 1-2|x|$ on $[-1,1]$. Thus understanding $f$ as a circle map
1531: through the identification $S^1=[-3,1]/\{-3,1\}$, this is a
1532: non-uniformly expanding map with a critical set satisfying
1533: conditions (S1)-(S3) and
1534: there are two ergodic absolutely continuous (thus SRB)
1535: invariant measures $\mu_1, \mu_2$ whose supports are
1536: $[-3,-1]$ and $[-1,1]$ respectively. Moreover defining
1537: $\Phi(t)= R_t\circ f,
1538: $
1539: where $R_t:S^1\to S^1$ is the
1540: rotation of angle $t$ and $\th_\ep=(2\ep)^{-1}(m\mid
1541: [-\ep,\ep])$ for small $\ep>0$, we have that $\{ \Phi,
1542: (\th_\ep)_{\ep>0} \}$ is a random perturbation satisfying
1543: nondegeneracy conditions 1 and 2. Since
1544: $\supp(\mu_1)\cap\supp(\mu_2)=\{-1\}$ we have that for
1545: $\ep>0$ small enough there must be a single physical
1546: measure $\mu^\ep$. Indeed, by property (P) any weak$^*$
1547: accumulation point of a family of physical measures must
1548: have $-1$ in its support.
1549: \begin{figure}[htb]
1550: \begin{center} \includegraphics[width=8cm]{parabolas2.ps}
1551: \caption{\label{fig.2} map for which $l=p=2$} \end{center}
1552: \end{figure}
1553:
1554: \medskip
1555:
1556:
1557: The second example is defined on the interval $I=[-7,2]$.
1558: We take the map $q_a(x)=a-x^2$ on $[-2,2]$ for some
1559: parameter $a\in(1,2)$ satisfying Benedicks-Carleson
1560: conditions (see~\cite{BC1} and~\cite{BC2}), and the
1561: ``same" map on $[-7,-3]$ conveniently conjugated:
1562: $p_a(x)=(x+5)^2-5-a$. Then the two pieces of graph are
1563: glued together in such a way that we obtain a smooth map
1564: $f:I\to I$ sending $I$ into its interior, as
1565: figure~\ref{fig.2} shows. The intervals
1566: $I_q=[q_a^2(0),q_a(0)]$ and $I_p=[p_a(-5),p_a^2(-5)]$ are
1567: forward invariant for $f$, and then we can find slightly
1568: larger intervals $I_1\supset I_p$ and $I_2\supset I_q$
1569: that become trapping regions for $f$. So, taking
1570: $ \Phi(t)= f+t, $
1571: and $\theta_\ep$ as in the previous
1572: example with $0<\ep<\ep_0$ for some $\ep_0>0$ small enough,
1573: then $\{ \Phi, (\th_\ep)_{\ep}\}$ is a random perturbation
1574: of $f$ leaving the intervals $I_1$ and $I_2$ invariant by
1575: each $\Phi(t)$. Moreover, Lebesgue almost every $x\in I$
1576: eventually arrives at one of these intervals. Then
1577: by~\cite{BC1} and \cite{BY} the map $f$ is non-uniformly
1578: expanding and has two SRB measures with supports contained
1579: in each trapping region. Finally $f$ admits two distinct
1580: physical measures whose supports are contained in $I_1$
1581: and $I_2$ respectively, for $\ep_0>0$ small enough, see
1582: \cite{BaV}.
1583:
1584:
1585:
1586: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1587:
1588:
1589: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1590:
1591:
1592: \section{Stochastic stability} \label{s.stochastic}
1593:
1594: In this section we will prove the first item of
1595: Theorem~\ref{t.stc} and Theorem~\ref{t.sts}. The second
1596: item of Theorem~\ref{t.stc} may be obtained in the same way
1597: as Theorem \ref{t.sts}, if we think of $\cc$ as being equal
1598: to the empty set and take into account Remark~\ref{r.hyp}.
1599:
1600: We start by proving the first item of Theorem \ref{t.stc}.
1601: Assume that $f$ is a stochastically stable non-uniformly
1602: expanding local diffeomorphism. We know from
1603: Proposition~\ref{p.finitemeasures} that there is a finite
1604: number of physical measures $\mu_1^\ep,\dots \mu_l^\ep$ and
1605: for each $x\in M$ there is a $\th_\ep^\NN$ mod $0$
1606: partition $T_1(x),\dots,T_l(x)$ of $T^\NN$ for which
1607: $$
1608: \mu_i^\ep=\wlim \frac1n \sum_{j=1}^{n-1}\de_{f_{\un t}^j(x)}
1609: \quad \mbox{for each}\quad \un t \in T_i(x).
1610: $$
1611: Furthermore, since we are taking $f$ a local
1612: diffeomorphism, then $\log\|(Df)^{-1}\|$ is a continuous
1613: map. Thus, we have for each $x\in M$ and $\th^\NN_\ep$
1614: almost every $\un t\in T^\NN$
1615: $$
1616: \lim_{n\ra \infty}\frac{1}{n}\sum_{j=0}^{n-1}
1617: \log\|Df\big(f_{\un t}^j(x)\big)^{-1}\|=
1618: \int\log\|(Df)^{-1}\|d\mu_i^\ep
1619: $$
1620: for some physical measure $\mu_i^\ep$ with $1\le i\le l$.
1621: Hence, for proving the non-uniform expansion of $f$ on
1622: random orbits it suffices to show that there is $c_0>0$
1623: such that if $\mu^\ep=\mu_i^\ep$ for some $1\le i\le l$
1624: then
1625: $$
1626: \int\log\|(Df)^{-1}\|d\mu^\ep<c_0\quad\mbox{for small $\ep>0$}.
1627: $$
1628:
1629: \cle\label{l.aprox}
1630: Let $\vfi\colon M\ra \RR$ be a continuous map.
1631: Given $\de>0$ there is $\ep_0>0$ such that if $\ep\leq\ep_0$, then
1632: $$
1633: \left|\int \vfi d\mu^\ep -
1634: \int \vfi d\mu_\ep\right|<\de,
1635: $$
1636: for some absolutely continuous $f$-invariant probability
1637: measure $\mu_\ep$.
1638: \fle
1639: \dem
1640: We will use the following auxiliary result: \emph{Let $X$
1641: be a compact metric space, $K\subset X$ a closed (compact)
1642: subset and $(x_t)_{t>0}$ a curve in $X$ (not necessarily
1643: continuous) such that all its accumulation points (as $t\to
1644: 0^{+}$) lie in $K$. Then for every open neighborhood $U$ of
1645: $K$ there is $t_0>0$ such that $x_t\in U$ for every
1646: $0<t<t_0$. } Indeed, supposing not, there is a sequence
1647: $(t_n)_{n}$ with $t_n\to 0^{+}$ when $n\to\infty$ such that
1648: $x_{t_n}\not\in U$. Since $X$ is compact this means that
1649: $(x_t)_{t>0}$ has some accumulation point in $X\setminus
1650: U$, thus outside $K$, contrary to the assumption.
1651:
1652: Now, the space $X=\PP(M)$ of all probability measures in
1653: $M$ is a compact metric space with the weak$^{*}$ topology,
1654: and the convex hull $K$ of the (finitely many) SRB measures
1655: of $f$ is closed. Hence, considering the curve
1656: $(\mu_\ep)_\ep$ in $\PP(M)$, we are in the context of the
1657: above result, since we are supposing $f$ to be
1658: stochastically stable. A metric on $X$ topologically
1659: equivalent to the weak$^{*}$ topology may be given by
1660: $$
1661: \distp(\mu,\nu)=\sum_{k=1}^\infty \frac1{2^n} \left| \int
1662: \varphi_n \, d\mu - \int \varphi_n \, d\nu \right|
1663: $$
1664: where
1665: $\mu,\nu\in\PP(M)$ and $(\varphi_n)_{n\ge 1}$ is a
1666: dense sequence of functions in $C^0(M,\RR)$,
1667: see~\cite{Man}.
1668:
1669: Let $\varphi:M\to\RR$ continuous be given and let us fix
1670: some $\de>0$. There must be $n\in\NN$ such that $\| \varphi
1671: - \varphi_n \|_0 < \de/3$ and, by the auxiliary result in
1672: the beginning of the proof, there exists, for some $\ep_0>0$ and
1673: every $0<\ep<\ep_0$, a probability measure
1674: $\mu_\ep\in\PP(M)$ for which $\distp(\mu^\ep,\mu_\ep)<\de
1675: (3\cdot 2^n)^{-1}$. This in particular means that
1676: $$
1677: \frac1{2^n} \left| \int \varphi_n \, d\mu^\ep - \int
1678: \varphi_n \, d\mu_\ep \right| < \frac{\de}{3\cdot 2^n},
1679: $$
1680: by the
1681: definition of the distance $\distp$, which implies
1682: $$
1683: \left| \int \varphi_n \, d\mu^\ep - \int
1684: \varphi_n \, d\mu_\ep \right| < \frac{\de}3.
1685: $$
1686: Hence we get
1687: \begin{eqnarray*}
1688: \lefteqn{\left| \int \varphi \, d\mu^\ep - \int \varphi \,
1689: d\mu_\ep \right|\leq}\hspace{.5cm}\\
1690: & \le &
1691: \left| \int \varphi \, d\mu^\ep - \int
1692: \varphi_n \, d\mu^\ep \right| + \left| \int \varphi_n \,
1693: d\mu^\ep - \int \varphi_n \, d\mu_\ep \right| + \left| \int
1694: \varphi_n \, d\mu_\ep - \int \varphi \, d\mu_\ep \right| \\
1695: & < & \frac{\de}3 +\frac{\de}3 +\frac{\de}3 =
1696: \de, \end{eqnarray*}
1697: which completes the proof of the lemma.\cqd
1698:
1699: Now we take $\vfi=\log\|(Df)^{-1}\|$ and $\de=c/2$ in the
1700: previous lemma, where $c>0$ is the constant given by the
1701: non-uniform expansion of $f$ (recall (\ref{liminf1})). For
1702: each $\ep\leq\ep_0$ let $\mu_\ep$ be the measure given by
1703: Lemma~\ref{l.aprox}. Since property (P) holds, there are
1704: real numbers $w_1(\ep),\dots ,w_p(\ep)\geq 0$ with
1705: $w_1(\ep)+\cdots+w_p(\ep)=1$ for which
1706: $\mu_\ep=w_1(\ep)\mu_1+\cdots+w_p(\ep)\mu_p$. Since each
1707: $\mu_i$ is an SRB measure for $1\leq i\leq p$, we have for
1708: Lebesgue almost every $x\in B(\mu_i)$
1709: $$
1710: \int\log\|(Df)^{-1}\|d\mu_i=\lim_{n\ra +\infty}\frac{1}{n}
1711: \sum_{j=0}^{n-1}\log\|Df(f^j(x))^{-1}\|\leq -c<0.
1712: $$
1713: This implies
1714: $$
1715: \int\log\|(Df)^{-1}\|d\mu_\ep\leq -c,
1716: $$
1717: and so, by Lemma \ref{l.aprox} and the choice of $\delta$,
1718: $$
1719: \int\log\|(Df)^{-1}\|d\mu^\ep\leq -c/2.
1720: $$
1721: This completes the proof of the first item of Theorem B.
1722:
1723: \medskip
1724:
1725: Now we go into the proof of Theorem \ref{t.sts}.
1726: In order to prove that $f$ is stochastically stable, and
1727: taking into account property (P), it suffices to prove
1728: that the weak$^*$ accumulation points of any family
1729: $(\mu^\ep)_{\ep>0}$, where each $\mu^\ep$ is a physical
1730: measure of level $\ep$, are absolutely continuous with respect
1731: to the Lebesgue measure. Let $\mu^\ep$ be a physical measure
1732: of level $\ep$ for some small $\ep>0$
1733: and define
1734: for each $n\geq 1$
1735: $$ \mu_n^\ep = \frac1n\sum_{j=0}^{n-1}\int (f^j_{\un t})_{*}
1736: \big(m\mid{B(\mu^\ep)\big)}\, d\th_\ep^\NN(\un t).
1737: $$
1738: We know from Proposition~\ref{p.finitemeasures} that each
1739: $\mu^\ep$ is the weak$^*$ limit of the sequence
1740: $(\mu_n^\ep)_n$. We will prove Theorem~\ref{t.sts} by
1741: providing some useful estimates on the densities of the
1742: measures $\mu_n^\ep$. Define for each $\un t\in T^\NN$ and
1743: $n\geq 1$
1744: $$
1745: H_n(\un t)=\{ x\in B(\mu^\ep)\colon \mbox{ $n$ is a
1746: $(\al,\delta)$-hyperbolic time for $(\un t,x)$ }\},
1747: $$
1748: and
1749: $$
1750: H^*_n(\un t) =\{ x\in B(\mu^\ep)\colon \mbox{ $n$ is the first
1751: $(\al,\delta)$-hyperbolic time for $(\un t,x)$ }\}.
1752: $$
1753: $ H^*_n(\un t)$ is precisely the set of those points
1754: $x\in B(\mu^\ep)$ for which $h_\ep(\un t,x)=n$
1755: (recall the definition of
1756: the map $h_\ep$).
1757: For
1758: $n,k\geq 1$ we also define $R_{n,k}(\un t)$ as the set of
1759: those points $x\in M$ for which $n$ is a
1760: $(\al,\delta)$-hyperbolic time and $n+k$ is the first
1761: $(\al,\delta)$-hyperbolic time after $n$, i.e. $$
1762: R_{n,k}(\un t)= \left\{x\in H_n(\un t)\colon \:f^n_{\un
1763: t}(x)\in H^*_k(\si^n\un t)\: \right\}, $$ where $\si\colon T^\NN \rightarrow T^\NN$ is the shift map
1764: $\si(t_1,t_2,\dots)=(t_2,t_3,\dots)$.
1765: Considering the
1766: measures
1767: $$
1768: \nu^\ep_n=\int ( f_{\un
1769: t}^n)_*\big(m\mid H_n(\un t)\big)d\th_\ep^\NN(\un t)
1770: $$
1771: and
1772: $$
1773: \eta_n^\ep=\sum_{k=2}^\infty\sum_{j=1}^{k-1}\int (f_{\un
1774: t}^{n+j})_*\big(m\mid R_{n,k}(\un t)\big) d\th_\ep^\NN(\un
1775: t),
1776: $$
1777: we may write
1778: $$
1779: \mu_n^\ep\leq
1780: \frac{1}{n}\sum_{j=0}^{n-1}(\nu_j^\ep+\eta_j^\ep).
1781: $$
1782:
1783:
1784: \cpr\label{p.dens}
1785: There is a constant $C_2>0$ such that for every $n\geq 0$
1786: and $\un t\in T^\NN$
1787: $$
1788: \frac{d}{dm}(f_{\un t}^n)_*\big(m\mid H_n(\un t)\big)\leq C_2.
1789: $$
1790: \fpr
1791: \dem Take $\de_1>0$ given by Proposition \ref{p.contr}. It is
1792: sufficient to prove that there is some uniform constant
1793: $C>0$ such that if $A$ is a Borel set in $M$ with diameter
1794: smaller than $\delta_1/2$ then
1795: $$
1796: m\big(f_{\un t}^{-n}(A)\cap H_n(\un t)\big)\leq C m(A).
1797: $$
1798: Let $A$ be a Borel set in $M$ with diameter smaller than
1799: $\delta_1/2$ and $B$ an open ball of radius $\delta_1/2$
1800: containing $A$. We may write
1801: $$
1802: f_{\un t}^{-n}(B)=\bigcup_{k\geq 1}B_k,
1803: $$
1804: where $(B_k)_{k\geq 1}$ is a (possibly finite) family of
1805: two-by-two disjoint open sets in $M$. Discarding those
1806: $B_k$ that do not intersect $H_n(\un t)$, we choose for
1807: each $k\geq 1$ a point $x_k\in H_n(\un t)\cap B_k$. For
1808: $k\geq 1$ let $V_n(\un t,x_k)$ be the neighborhood of $x_k$
1809: in $M$ given by Proposition \ref{p.contr}. Since $B$ is
1810: contained in $B\big(f_{\un t}^n(x_k), {\de_1}\big)$, the
1811: ball of radius $\de_1$ around $f_{\un t}^n(x_k)$, and
1812: $f_{\un t}^n$ is a diffeomorphism from $V_n(\un t,x_k)$
1813: onto $B\big(f_{\un t}^n(x_k), {\de_1}\big)$, we must have
1814: $B_k\subset V_n(\un t,x_k)$ (recall that by our choice of
1815: $B_k$ we have $f_{\un t}^n(B_k)\subset B$). As a
1816: consequence of this and Corollary \ref{c.dist}, we have for
1817: every $k$ that the map $f_{\un t}^n\mid B_k\colon
1818: B_k\rightarrow B$ is a diffeomorphism with bounded
1819: distortion:
1820: $$ \frac{1}{C_1}\leq \frac{|\det Df_{\un
1821: t}^n(y)|}{|\det Df_{\un t}^n(z)|} \leq C_1 $$ for all
1822: $y,z\in B_k$.
1823: This finally
1824: gives \begin{eqnarray*} m\big(f_{\un t}^{-n}(A)\cap H_n(\un
1825: t)\big) &\leq & \sum_{k}m\big(f_{\un t}^{-n}(A\cap B)\cap
1826: B_k\big)\\ &\leq & \sum_{k}C_1\frac{m(A\cap
1827: B)}{m(B)}m(B_k)\\ &\leq & C_2 m(A), \end{eqnarray*} where
1828: $C_2>0$ is a constant only depending on $C_1$, on the
1829: volume of the ball $B$ of radius $\delta_1/2$, and on the
1830: volume of $M$.\cqd
1831:
1832:
1833:
1834:
1835: It follows from Proposition \ref{p.dens} that
1836: \begin{equation}\label{dens}
1837: \frac{d\nu_n^\ep}{dm}\leq C_2
1838: \end{equation}
1839: for every $n\geq 0$ and small $\epsilon>0$. Our goal now is to
1840: control the density of the measures $\eta_n^\ep$
1841: in such a way that we may assure the absolute
1842: continuity of the weak$^*$ accumulation points
1843: of the measures $\mu^\ep$ when $\ep$ goes to zero.
1844:
1845:
1846:
1847:
1848: \cpr\label{p.dens2} Given $\zeta>0$, there is
1849: $C_3(\zeta)>0$ such that for every $n\geq 0$ and $\ep>0$
1850: we may bound $\eta_n^{\ep}$ by the sum of two non-negative
1851: measures, $\eta_n^{\ep} \leq \omega^{\:\ep}+\rho^{\:\ep}$,
1852: with
1853: $$
1854: \frac{d\omega^{\:\ep}}{dm}\leq C_3(\zeta)\qand
1855: \rho^{\:\ep}(M)<\zeta.
1856: $$
1857: \fpr
1858: \dem
1859: Let $A$ be some Borel set in $M$.
1860: We have for each $n\geq 0$
1861: \begin{eqnarray*}
1862: \eta_n^\ep(A)
1863: &=&
1864: \sum_{k=2}^\infty\sum_{j=1}^{k-1} \int
1865: m\big( f_{\un t}^{-n-j}(A)\cap R_{n,k}(\un t)\big)
1866: d\th_\ep^\NN(\un t)\\
1867: &\leq &
1868: \sum_{k=2}^\infty\sum_{j=1}^{k-1} \int m\big( f_{\un
1869: t}^{-n} \big(f_{\si^n\un t}^{-j}(A)\cap H^*_k(\si^n\un t)\big)\cap
1870: H_n(\un t)\big)d\th_\ep^\NN(\un t)\\
1871: &=&
1872: \sum_{k=2}^\infty\sum_{j=1}^{k-1} \int\nu_n^\ep\big(
1873: f_{\si^n\un t}^{-j}(A)\cap H^*_k(\si^n\un t)\big) d\th_\ep^\NN(\un
1874: t) \\
1875: &\leq &
1876: \sum_{k=2}^\infty\sum_{j=1}^{k-1} C_2\int
1877: m\big( f_{\un t}^{-j}(A)\cap
1878: H^*_k(\un t)\big)d\th_\ep^\NN(\un t).
1879: \end{eqnarray*}
1880: (in this last inequality we used that $\th^\NN_\ep$
1881: is $\si$-invariant and estimate (\ref{dens}) above). Let
1882: now $\zeta>0$ be some fixed small number. Since we are
1883: assuming $(h_\ep)_\ep$ with uniform $L^1$-tail this means
1884: that there is some integer $N=N(\zeta)$ for which
1885: $$
1886: \sum_{j=N}^{\infty}k\int
1887: m\big(H^*_k(\un t)\big)d\th^\NN_\ep(\un t)<\frac{\zeta}{C_2}.
1888: $$
1889: We take
1890: $$
1891: \omega^{\:\ep}=C_2\sum_{k=2}^{N-1}\sum_{j=1}^{k-1}
1892: \int(f_{\un t}^j)_*\big(m\mid H^*_k(\un t)\big)d\th_\ep^\NN(\un t)
1893: $$
1894: and
1895: $$
1896: \rho^{\:\ep}=C_2\sum_{k=N}^\infty\sum_{j=1}^{k-1}
1897: \int(f_{\un t}^j)_*\big(m\mid H^*_k(\un t)\big)d\th_\ep^\NN(\un t).
1898: $$
1899: For this last measure we have
1900: $$
1901: \rho^{\:\ep}(M)=C_2\sum_{k=N}^\infty\sum_{j=1}^{k-1} \int
1902: m\big( H^*_k(\un t)\big)d\th_\ep^\NN(\un t) \leq C_2
1903: \sum_{k=N}^\infty k\int m\big( H^*_k(\un t)\big)d\th_\ep^\NN(\un t)
1904: <\zeta.
1905: $$
1906: On the other hand, it follows from the definition of
1907: $(\al,\de)$-hyperbolic times that there is some constant
1908: $a=a(N)>0$ such that
1909: $\dist\big(H_k(\un t), \cc\big)\geq a
1910: $ for $1\leq k\leq N$. Defining $\Delta\subset M$ as the
1911: set of those points in $M$ whose distance to $\cc$ is
1912: greater than $a$, we have
1913: $$
1914: \omega^{\:\ep}\leq C_2\sum_{k=2}^{N-1}\sum_{j=1}^{k-1}
1915: \int(f_{\un t}^j)_*(m\mid \Delta)\:d\th_\ep^\NN(\un t),
1916: $$
1917: and this last measure has density bounded by some uniform
1918: constant, as long as we take the maps $f_t$ in a
1919: sufficiently small neighborhood of $f$ in the $C^1$
1920: topology.
1921: \cqd
1922:
1923: It follows from this last proposition and (\ref{dens}) that the weak$^*$
1924: accumulation points of $\mu^\ep$ when $\ep\to0$ cannot
1925: have singular part, thus being absolutely continuous with
1926: respect to the Lebesgue measure. Moreover, the
1927: weak$^*$ accumulation points of a family of
1928: stationary measures are always $f$-invariant
1929: measures, cf. Remark~\ref{re.accinvariant}. This together with (P)
1930: gives the stochastic stability of $f$.
1931:
1932: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1933:
1934: \section{Applications}\label{s.applications}
1935:
1936: In this section we will apply Theorems \ref{t.stc} and
1937: \ref{t.sts} to certain examples of
1938: non-uniformly expanding maps.
1939: Before we explicit the examples we have in mind let us give
1940: a practical criterion for proving that the family of
1941: hyperbolic time maps $(h_\ep)_\ep$ has uniform $L^1$-tail.
1942:
1943: If we look at the proof of Proposition \ref{p.hyp} we
1944: realize that what we did was fixing some positive number
1945: $c_0$ smaller than $c$, and then, for $\th_\ep^\NN\times m$
1946: almost every $(\un t,x)\in T^\NN\times M$, we took a
1947: positive integer $N_\ep=N_\ep(\un t,x)$ for which $$
1948: %\frac{1}{N_\ep}
1949: \sum_{j=0}^{N_\ep-1}\log\|Df(f_{\un t}^j(x))^{-1}\|\leq
1950: -c_0N_\ep \quad \mbox{and} \quad
1951: % \frac{1}{N_\ep}
1952: \sum_{j=0}^{N_\ep-1}-\log\dist_\delta(f_{\un t}^j(x),
1953: \cc)\leq \gamma N_\ep, $$ for suitable choices of
1954: $\delta>0$ and $\gamma>0$. This permits us to introduce a
1955: map
1956: $$N_\ep\colon T^\NN\times M\ra \ZZ^+$$
1957: whose existence provides a first hyperbolic time map
1958: $$
1959: h_\ep\colon T^\NN\times M\ra \ZZ^+\quad\mbox{with}\quad h_\ep \leq N_\ep
1960: $$
1961: (recall the proof of Proposition \ref{p.hyp}).
1962: Thus, the integrability of the map $h_\ep$ is implied by the
1963: integrability of the map $N_\ep$, which is in practice
1964: easier to handle.
1965:
1966: In the examples we are going to study below we will show
1967: that there is a sequence of positive real numbers
1968: $(a_k^\ep)_k$ for which
1969: $$ (\th_\ep^\NN\times
1970: m)\left(\big\{(\un t,x)\in T^\NN\times M\colon N_\ep(\un
1971: t,x)>k\big\}\right)\leq a_k^\ep \quad \mbox{and} \quad
1972: \sum_{k=1}^{\infty} ka_k^\ep <\infty,
1973: $$
1974: This gives the integrability of $h_\ep$ with respect to the
1975: measure $\th_\ep^\NN\times m$. The fact the family
1976: $(h_\ep)_\ep$ has uniform $L^1$-tail can be proved by
1977: showing that the sequence $(a_k^\ep)_k$ may be chosen not
1978: depending on $\ep>0$.
1979:
1980: Now we are ready for the applications of Theorems
1981: \ref{t.stc} and \ref{t.sts}. We will describe firstly a
1982: class of local diffeomorphisms introduced in \cite[Appendix
1983: A]{ABV} that satisfies the hypotheses of Theorem
1984: \ref{t.stc}, and then a class of maps (with critical sets)
1985: introduced in \cite{V} satisfying the hypotheses of Theorem
1986: \ref{t.sts}.
1987:
1988:
1989: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1990:
1991: \subsection{Local diffeomorphisms} \label{ss.localdiffeos}
1992: Now we follow~\cite[Appendix A]{ABV} and describe robust
1993: classes of maps (open in the $C^2$ topology) that are
1994: non-uniformly expanding local diffeomorphisms and
1995: stochastically stable. Let $M$ be a compact Riemannian
1996: manifold and consider
1997: \[
1998: \begin{array}{rccl}
1999: \Phi:& T &\longrightarrow& C^2(M,M)\\
2000: & t &\longmapsto & f_t
2001: \end{array}
2002: \]
2003: a continuous family of $C^2$ maps, where $T$
2004: is a metric space.
2005: We begin with an essentially
2006: combinatorial lemma.
2007:
2008:
2009: \cle \label{l.frequency} Let $p,q\ge1$ be integers and
2010: $\si>q$ a real number. Assume $M$ admits a measurable cover
2011: $\{ B_1,\ldots,B_p,B_{p+1},\ldots,B_{p+q}\}$ such that for
2012: all $t\in T$ it holds \begin{enumerate} \item $|\det
2013: Df_t(x)|\ge \si$ for all $x\in B_{p+1}\cup\ldots\cup
2014: B_{p+q}$; \item $\left(f_t\mid B_i\right)$ is injective for
2015: all $i=1,\ldots,p$. \end{enumerate} Then there is $\zeta>0$
2016: such that for every Borel probability $\th$ on $T$ we have
2017: \begin{equation}\label{eq.frequency} \#\{0\le j < n:
2018: f^j_{\un t}(x) \in B_1\cup\ldots\cup B_p \} \ge \zeta n
2019: \end{equation} for $\th^\NN\times m$ almost all $(\un
2020: t,x)\in T^\NN\times M$ and large enough $n\ge1$. Moreover
2021: the set $I_n$ of points $(\un t,x)\in T^\NN\times M$ whose
2022: orbits do not spend a fraction $\zeta$ of the time in
2023: $B_1\cup\ldots\cup B_p$ up to iterate $n$ is such that
2024: $(\th^\NN\times m)(I_n)\le \tau^n$ for some $0<\tau<1$ and
2025: for large $n\ge1$. \fle
2026:
2027: \dem
2028: Let us fix $n\ge1$ and $\un t\in T^\NN$.
2029: For a sequence
2030: $\un i=(i_0,\ldots,i_{n-1})\in\{1,\ldots,p+q\}^n$
2031: we write
2032: $$
2033: [\un i]=
2034: B_{i_0}\cap (f^1_{\un t})^{-1}(B_{i_1})\cap\cdots
2035: \cap (f^{n-1}_{\un t})^{-1}(B_{i_{n-1}})
2036: $$
2037: and define
2038: $g(\un i)=\#\{ 0\le j < n : i_j \le p \}$.
2039:
2040: We start by observing that for $\zeta>0$ the number of
2041: sequences $\un i$ such that $g(\un i)<\zeta n$ is bounded
2042: by $$ \sum_{k<\zeta n} {n \choose k} p^k q^{n-k} \le
2043: \sum_{k\le \zeta n} {n \choose k} p^{\zeta n} q^n. $$ Using
2044: Stirling's formula (cf.~\cite[Section 6.3]{BV}) the
2045: expression on the right hand side is bounded by $(e^\gamma
2046: p^\zeta q)^n$, where $\gamma>0$ depends only on $\zeta$ and
2047: $\gamma(\zeta)\to 0$ when $\zeta\to0$.
2048:
2049: Assumptions 1 and 2 ensure $m([\un i])\le
2050: \si^{-(1-\zeta)n}$ (recall that $m(M)=1$). Hence the
2051: measure of the union $I_n(\un t)$ of all the sets $[\un i]$
2052: with $g(\un i)<\zeta n$ is bounded by $$ \si^{-(1-\zeta)n}
2053: (e^\gamma p^\zeta q)^n. $$ Since $\si>q$ we may choose
2054: $\zeta$ so small that $e^\gamma p^\zeta q <
2055: \si^{(1-\zeta)}$. Then $m(I_n(\un t))\le \tau^n$ with
2056: $\tau=e^{\gamma+\zeta-1}\cdot p^\zeta \cdot q <1$ for big
2057: enough $n\ge N$. Note that $\tau$ and $N$ do not depend on
2058: $\un t$. Setting $$ I_n=\mcup_{\un t\in T^\NN} \big(\{\un
2059: t\} \times I_n(\un t)\big) $$ we also have $(\th^\NN\times
2060: m)(I_n)\le \tau^n$ for all big $n\ge N$ and for every Borel
2061: probability $\th$ on $T$, by Fubini's Theorem. Since
2062: $\sum_n (\th^\NN\times m)(I_n) < \infty$ then
2063: Borel-Cantelli's Lemma implies $$ (\th^\NN\times m) \left(
2064: \mcap_{n\ge1} \mcup_{k\ge n} I_k \right) =0 $$ and this
2065: means that $\th^\NN\times m$ almost every $(\un t,x)\in
2066: T^\NN\times M$ satisfies~(\ref{eq.frequency}). \cqd
2067:
2068: \cle \label{pr.nonunifexp.localdiffeo} Let $\{
2069: B_1,\ldots,B_p,B_{p+1},\ldots,B_{p+q}\}$ be a measurable
2070: cover of $M$ satisfying conditions 1 and 2 of
2071: Lemma~\ref{l.frequency}. For $0<\lambda<1$ there are
2072: $\eta>0$ and $c_0>0$ such that, if $f_t$ also satisfies for
2073: all $t\in T$ \begin{enumerate} \item[3.] $\| Df_t(x)^{-1}
2074: \| \le \lambda <1$ for $x\in B_1,\ldots,B_p$; \item[4.]
2075: $\|Df_t(x)^{-1}\| \le 1+\eta$ for $x\in
2076: B_{p+1},\ldots,B_{p+q}$; \end{enumerate} then we have
2077: for $f\equiv
2078: f_{t^*}$, where $t^*$ is some given point in $T$,
2079: \begin{equation}\label{eq.nonunifexp.sum}
2080: \limsup_{n\to+\infty}\frac1n\sum_{j=0}^{n-1} \log\|
2081: Df(f^j_{\un t}(x))^{-1}\| \le -c_0 \end{equation} for
2082: $\th^\NN\times m$ almost all $(\un t,x)\in T^\NN\times M$,
2083: where $\th$ is any Borel probability measure on $T$.
2084: Moreover the first hyperbolic time map $h:T^\NN\times M\to
2085: \ZZ^{+}$ satisfies $$ (\th^\NN\times m) \{ (\un t,x)\in
2086: T^\NN\times M: h(\un t,x) > k \} \le a_k \qand
2087: \sum_{k=1}^\infty ka_k < \infty $$ with $(a_k)_k$
2088: independent of the choice of $\th$. \fle
2089:
2090: \dem Let $\zeta>0$ be the constant provided by
2091: Lemma~\ref{l.frequency}. We fix $\eta>0$ sufficiently small
2092: so that $\lambda^\zeta(1+\eta)\le e^{-c_0}$ holds for some
2093: $c_0>0$ and take $(\un t,x)$
2094: satisfying~(\ref{eq.frequency}). Conditions 3 and 4 now
2095: imply \begin{equation}\label{eq.nonunifexp.prod}
2096: \prod_{j=0}^{n-1}\|Df(f^j_{\un t}(x))^{-1}\| \le
2097: \lambda^{\zeta n} (1+\eta)^{(1-\zeta)n} \le e^{-c_0 n}.
2098: \end{equation} for large enough $n$. This
2099: means~(\ref{eq.nonunifexp.prod}) holds for $\th^\NN\times
2100: m$ almost every $(\un t,x)\in T^\NN\times M$.
2101:
2102: We observe that if $h(\un t,x)=k$, then $1\le n < k$
2103: cannot be hyperbolic times for $(\un t,x)$. Hence $(\un
2104: t,x)\in I_n$ for all $n=1,\ldots,k-1$. In particular $$
2105: (\th^\NN\times m)\{ (\un t,x)\in T^\NN\times M: h(\un
2106: t,x)=k \} \le (\th^\NN\times m)(I_{k-1}) \equiv a_k $$ and
2107: $\sum_k k a_k\le\sum_k k\tau^{k-1} < \infty$. \cqd
2108:
2109:
2110:
2111: Now we will show that families of $C^2$ maps satisfying
2112: conditions 1 through 4 of Lemmas~\ref{l.frequency} and
2113: \ref{pr.nonunifexp.localdiffeo} contain open sets of
2114: families in the $C^2$ topology. Let $M$ be a
2115: $n$-dimensional torus $\TT^n$ and $f_0:M\to M$ a uniformly
2116: expanding map: there exists $0<\lambda <1$ such that
2117: $\|Df_0(x)v\|\ge\lambda^{-1}\|v\|$ for all $x\in M$ and
2118: $v\in T_x M$. Let also $W$ be some small compact domain in
2119: $M$ where $f_0\mid W$ is injective. Observe that $f_0$ is a
2120: volume expanding local diffeomorphism due to the uniform
2121: expansion.
2122:
2123: Modifying $f_0$ by an isotopy inside $W$ we may obtain a
2124: map $f_1$ which coincides with $f_0$ outside $W$, is volume
2125: expanding in $M$, i.e., $| \det Df_1(x) | >1$ for all $x\in
2126: M$, and has bounded contraction on $W$ near 1: $\| Df_1(x)
2127: ^{-1} \| \le 1+\eta$ for every $x\in W$ and some $\eta>0$
2128: small. This new map $f_1$ may be taken $C^1$ close to $f_0$
2129: and we may consider a $C^2$ map $f_2$ arbitrarily $C^1$
2130: close to $f_1$.
2131:
2132: Now any map $f$ in a small enough $C^2$ neighborhood of
2133: $f_2$ admits $\si>1$ such that $|\det Df(x)|\ge\si$ for all
2134: $x\in M$ and, for $x$ outside $W$, we have $\|
2135: Df(x)^{-1}\|\le\lambda$. If the $C^2$ neighborhood is taken
2136: sufficiently small then we maintain $\|Df(x)^{-1}\|\le
2137: 1+\eta$ for $x\in W$ and for some small $\eta>0$. Let us
2138: take $B_1,\ldots,B_p, B_{p+1}=W$ a partition of $M$ into
2139: measurable sets where the restriction $f\mid B_i$ is
2140: injective for $i=1,\ldots,p+1$. Then any continuous family
2141: of $C^2$ maps $\Phi:T\to C^2(M,M)$ together with a family
2142: $(\th_\ep)_{\ep>0}$ of Borel probability measures in the
2143: metric space $T$, satisfying $\supp(\th_\ep)\to\{t^*\}$
2144: when $\ep\to0$ and $f_{t^*}\equiv f$, for some $t^*\in T$,
2145: is such that $f$ is non-uniformly expanding for random
2146: orbits and $(h_\ep)_{\ep>0}$ has uniform $L^1$-tail -- by
2147: Lemma~\ref{pr.nonunifexp.localdiffeo} with $q=1$ and
2148: $T=\supp(\th_\ep)$ for small enough $\ep>0$. Theorem B then
2149: shows
2150:
2151: \cco \label{co.nonunifexp} There are open sets ${\cal U}
2152: \subset C^2(M,M)$
2153: of maps in the $C^2$ topology
2154: such that every map $f\in {\cal U}$ is a stochastically
2155: stable non-uniformly expanding local diffeomorphism. \fco
2156:
2157:
2158: \subsection{Viana maps}
2159: \label{ss.vianamaps}
2160:
2161: In what follows we describe the class of non-uniformly
2162: expanding maps (with critical sets) introduced by M. Viana,
2163: referring the reader to \cite{V}, \cite{Al} and \cite{AV}
2164: for details. Then we show that those maps satisfy the
2165: hypotheses of Theorem \ref{t.finite2} and Theorem
2166: \ref{t.sts}.
2167:
2168: Let $a_0\in(1,2)$ be such that
2169: the critical point $x=0$ is pre-periodic for the quadratic
2170: map $Q(x)=a_0-x^2$. Let $S^1=\RR/\ZZ$ and $b:S^1\rightarrow
2171: \RR$ be a Morse function, for instance, $b(s)=\sin(2\pi
2172: s)$. For fixed small $\alpha>0$, consider the map
2173: \[ \begin{array}{rccc} \hat f: & S^1\times\RR
2174: &\longrightarrow & S^1\times \RR\\
2175: & (s, x) &\longmapsto & \big(\hat g(s),\hat q(s,x)\big)
2176: \end{array}
2177: \]
2178: where $\hat g$ is the uniformly expanding
2179: map of the circle defined by $\hat{g}(s)=ds$ (mod $\ZZ$)
2180: for some $d\ge16$, and $\hat q(s,x)=a(s)-x^2$ with
2181: $a(s)=a_0+\al b(s)$. It is easy to check that for $\al>0$
2182: small enough there is an interval $I\subset (-2,2)$ for
2183: which $\hat f(S^1\times I)$ is contained in the interior of
2184: $S^1\times I$. Thus, any map $f$ sufficiently close to
2185: $\hat f$ in the $C^0$ topology has $S^1\times I$ as a
2186: forward invariant region. We consider from here on these
2187: maps $f$ close to $\hat f$ restricted to $S^1\times I$.
2188:
2189: Taking into account the expression of $\hat f$ it is not
2190: difficult to check that $\hat f$ (and any map $f$ close to
2191: $\hat f$ in the $C^2$ topology) behaves like a power of the
2192: distance close to the critical set.
2193:
2194:
2195: \subsubsection{Non-uniform expansion}
2196: \label{ss.non-uniform}
2197:
2198: The results in
2199: \cite{V} show that if the map $f$ is sufficiently close to
2200: $\hat f$ in the $C^3$ topology then $f$ has two positive
2201: Lyapunov exponents almost everywhere:
2202: there is a constant $\lambda>0$ for which
2203: $$\liminf_{n\rightarrow+\infty}\frac{1}{n}\log
2204: \|Df^n(s,x)v\|\geq \lambda$$ for Lebesgue almost every
2205: $(s,x)\in S^1\times I$ and every non-zero $v\in
2206: T_{(s,x)}(S^1\times I)$. In fact, the method used for
2207: showing this result also gives that $f$ is a non-uniformly
2208: expanding map and its orbits have slow approximation to the
2209: critical set, as we now explain. For the sake of clearness,
2210: we start by assuming that $f $ has the special form
2211: \begin{equation}
2212: f(s,x)=(g(s),q(s,x)),\quad\mbox{with}\quad\partial_xq(s,x)=0
2213: \quad\mbox{if and only if}\quad x=0, \label{assumption}
2214: \end{equation}
2215: and describe how the conclusions in \cite{V}
2216: are obtained for each $C^2$ map $f$ satisfying
2217: \begin{equation}
2218: \|f-\hat f\|_{C^2}\leq\al\quad\mbox{on}\quad S^1\times
2219: I.\label{alfa}
2220: \end{equation}
2221: Then
2222: we explain how these conclusions extend to the general
2223: case, using the existence of a central invariant foliation,
2224: and we show how the results in \cite{V} give the
2225: non-uniform expansion and slow approximation of orbits to the critical
2226: set for each map $f$ as in (\ref{alfa}).
2227:
2228:
2229: The estimates on the derivative rely on a
2230: statistical analysis of the returns of orbits to the
2231: neighborhood $S^1\times (-\sqrt{\al},\sqrt{\al}\,)$ of the
2232: critical set $\cc=\{(s,x):x=0\}$.
2233: We set
2234: $$J(0)=I\setminus(-\sqrt\al,\sqrt\al)\qand
2235: J(r)=\{x\in I:|x|<e^{-r}\}\quad\mbox{for
2236: $r\ge0$}.$$
2237: From here on we only consider points $(s,x)\in S^1\times I$
2238: whose orbit does not hit the critical set $\cc$. This constitues
2239: no restriction in our results, since the set of those
2240: points has full Lebesgue measure.
2241:
2242: For each integer $j\geq 0$ we
2243: define $(s_j,x_j)=f^j(s,x)$ and
2244: $$r_j(s,x)=\min\left\{r\ge0:x_j\in
2245: J(r)\right\}. $$
2246: Consider, for some small constant $0<\eta<1/4$,
2247: $$
2248: G=\bigg\{0\le j<n:
2249: r_j(s,x)\ge\bigg(\frac12-2\eta\bigg)\log\frac1\al
2250: \bigg\}.
2251: $$
2252: Fix some integer $n\ge1$ sufficiently large (only depending
2253: on $\al>0$).
2254: The results in \cite{V} show that if we take
2255: $$
2256: B_2(n)=\big\{(s,x):\:\mbox{there is $1\le j<n$ with $x_j\in J([\sqrt n])$\:}\},
2257: $$
2258: where $[\sqrt n]$ is the integer part of $\sqrt n$, then we
2259: have
2260: \begin{equation}
2261: \label{e.b2}
2262: m(B_2(n))\le\mbox{const}\,e^{-\sqrt n/4}
2263: \end{equation}
2264: and, for
2265: every small $c>0$ (only depending on the quadratic map
2266: $Q$),
2267: \begin{equation}
2268: \label{e.principal}
2269: \log\prod_{j=0}^{n-1}|\partial_xq(s_j,x_j)|\ge
2270: 2cn-\sum_{j\in G}r_j(s,x)\quad\mbox{for}\quad (s,x)\notin B_2(n),
2271: \end{equation}
2272: see \cite[pp. 75 \& 76]{V}.
2273: Moreover, if we define for $\gamma>0$
2274: $$
2275: B_1(n)=\bigg\{(s,x)\notin B_2(n):\sum_{j\in G}r_j(s,x)\ge
2276: \gamma n\bigg\},
2277: $$
2278: then, for small $\gamma>0$, there is a constant $\xi>0$ for which
2279: \begin{equation}
2280: \label{e.b1}
2281: m\big(B_1(n)\big)\le e^{-\xi n},
2282: \end{equation}
2283: see~\cite[p. 77]{V}.
2284: Taking into account the definitions of $J(r)$ and $r_j$, this shows that if we take
2285: $\delta=(1/2-2\eta)\log(1/\al)$, then
2286: \begin{equation}
2287: \nonumber
2288: \sum_{j=0}^{n-1}-\log\dist_\delta(f^j(x),\cc)\leq \gamma n
2289: \quad\mbox{for}\quad(s,x)\notin B_1(n)\cup B_2(n).
2290: \end{equation}
2291: This in particular gives that almost all orbits have slow
2292: approximation to $\cc$.
2293:
2294: On the other hand, we have for $(s,x)\in
2295: S^1\times I$
2296: \begin{equation}\label{e.forma}
2297: \big( Df(s,x)\big)^{-1}=
2298: \frac{1}{\partial_x q(s,x)\partial_s g(s)}
2299: \left( \begin{array}{cc}
2300: \partial_x q(s,x) & 0 \\
2301: -\partial_{s}q(s,x) & \partial_s g(s)
2302: \end{array} \right).
2303: \end{equation}
2304: Since all the norms are equivalent in finite dimensional
2305: Banach spaces, it no restriction for our purposes
2306: if we take the norm of $\big(Df(s,x)\big)^{-1}$ as the
2307: maximum of its entries. From (\ref{assumption}) and (\ref{alfa})
2308: we deduce that for small $\alpha$
2309: $$
2310: |\partial_s g|\geq d-\alpha,\quad |\partial_s q|\leq
2311: \alpha|b^\prime|+\alpha\leq 8\alpha\quad
2312: \mbox{and}\quad|\partial_xq|
2313: \leq |2x|+\alpha\leq 4,
2314: $$
2315: which together with (\ref{e.forma}) gives
2316: $$
2317: \big\|\big(Df(s,x)\big)^{-1}\big\|
2318: =
2319: |\partial_xq(s,x)|^{-1},
2320: $$
2321: as long as $\al>0$ is taken sufficiently small. This
2322: implies
2323: \begin{equation}
2324: \label{e.soma}
2325: \sum_{j=0}^{n-1}\log\|Df(s_j,x_j))^{-1}\|=-
2326: \sum_{j=0}^{n-1}\log|\partial_xq(s_j,x_j)|
2327: \end{equation}
2328: for every $(s,x)\in S^1\times I$.
2329: If we choose $\gamma<c$, then we have
2330: \begin{equation}
2331: \label{e.soma2}
2332: \sum_{j=0}^{n-1}\log|\partial_xq(s_j,x_j)|=
2333: \log\prod_{j=0}^{n-1}|\partial_xq(s_j,x_j)|\ge cn
2334: \end{equation}
2335: for every $(s,x)\notin B_1(n)\cup B_2(n)$ (recall (\ref{e.principal}) and
2336: the definition of $B_1(n)$).
2337: We conclude from (\ref{e.soma}) and (\ref{e.soma2}) that
2338: $$
2339: \sum_{j=0}^{n-1}\log\|Df(s_j,x_j))^{-1}\|\le
2340: -cn
2341: \quad
2342: \mbox{for}\quad(s,x)\notin B_1(n)\cup B_2(n),
2343: $$
2344: which, in view of the estimates on the Lebesgue measure of
2345: $B_1(n)$ and $B_2(n)$, proves that $f$ is a non-uniformly
2346: expanding map.
2347:
2348:
2349:
2350: \medskip
2351:
2352: Now we describe how in \cite{V} the
2353: same conclusions are obtained without assuming
2354: (\ref{assumption}).
2355: Since $\hat f$ is strongly expanding in the horizontal
2356: direction, it follows from the methods of \cite{HPS} that
2357: any map $f$ sufficiently close to $\hat f$ admits a unique
2358: invariant central foliation $\cf^c$ of $S^1\times I$ by
2359: smooth curves uniformly close to vertical segments, see
2360: \cite[Section 2.5]{V}. Actually, $\cf^c$ is obtained as
2361: the set of integral curves of a vector field $(\xi^c,1)$
2362: in $S^1\times I$
2363: with $\xi^c$ uniformly close to zero.
2364: The previous analysis can then be
2365: carried out in terms of the expansion of $f$ along this
2366: central foliation $\cf^c$. More precisely, $|\partial_x
2367: q(s,x)|$ is replaced by
2368: $$
2369: |\partial_c q(s,x)| \equiv |Df(s,x)v_c(s,x)|,
2370: $$ where
2371: $v_c(s,x)$ is a unit vector tangent to the foliation at
2372: $(s,x)$. The previous observations imply that $v_c$ is
2373: uniformly close to $(0,1)$ if $f$ is close to $\hat f$.
2374: Moreover, cf. \cite[Section 2.5]{V}, it is no restriction
2375: to suppose $|\partial_c q(s,0)|\equiv 0$, so that
2376: $\partial_c q(s,x)\approx |x|$, as in the unperturbed case.
2377: Indeed, if we define the {\em critical set} of $f$ by
2378: $$
2379: \cc=\{(s,x)\in S^1\times I:\partial_c q(s,x)=0\}.
2380: $$
2381: by an easy
2382: implicit function argument it is shown in \cite[Section
2383: 2.5]{V} that $\cc$ is the graph of some $C^2$ map $\eta :
2384: S^1\rightarrow I$ arbitrarily $C^2$-close to zero if
2385: $\alpha$ is small. This means that up to a change of
2386: coordinates $C^2$-close to the identity we may suppose that
2387: $\eta\equiv 0$ and, hence, write for $\al>0$ small
2388: $$
2389: \partial_c q(s,x)=x\psi(s,x)\quad\mbox{with $|\psi+2|$ close
2390: to zero}.
2391: $$
2392: This provides an analog to the second part of assumption
2393: (\ref{assumption}).
2394: At this point, the
2395: arguments apply with $\partial_xq(s,x)$ replaced by
2396: $\partial_c q(s,x)$, to show that orbits have slow
2397: approximation to the critical set $\cc$ and
2398: $
2399: \prod_{i=0}^{n-1} |\partial_c q(s_i,x_i)|
2400: $ grows
2401: exponentially fast for Lebesgue almost every $(s,x)\in
2402: S^1\times I$. A matrix formula for
2403: $\big(Df^n(s,x)\big)^{-1}$ similar to that in (\ref{e.forma}) can
2404: be obtained if we replace the vector $(0,1)$ in the
2405: canonical basis of the space tangent to $S^1\times I$ at $(s,x)$ by
2406: $v_c(s,x)$, and consider the matrix of $\big(Df^n(s,x)\big)^{-1}$
2407: with respect to the new
2408: basis.
2409:
2410: \medskip
2411:
2412: For future reference, let us make some considerations on
2413: the way the sets $B_1(n)$ and $B_2(n)$ are obtained. Let
2414: $X:S^1\ra I$ be a smooth map whose graph in $ S^1\times I$
2415: is nearly horizontal (see the notion of admissible curve in
2416: \cite[Section~2]{V} for a precise definition). Denote $\wh
2417: X_n(s)=f^n\big(s,X(s)\big)$ for $n\ge0$ and $s\in S^1$.
2418: Take
2419: some leaf $L_0$ of the foliation $\cf^c$.
2420: Letting $L_n=f^n(L_0)$ for $n\ge 1$, we define a sequence
2421: of Markov partitions $(\cp_n)_n$ of $S^1$ in the following
2422: way:
2423: $$
2424: \cp_n=\left\{[s^\prime,s^{\prime\prime})\colon (s^\prime,s^{\prime\prime})
2425: \mbox{ is a connected component of }
2426: \widehat{X}_n^{-1}\big((S^1\times I)\setminus L_n\big)\right\}.
2427: $$
2428: It is easy to check that $\cp_{n+1}$ refines $\cp_n$ for
2429: each $n\ge 1$ and
2430: $$
2431: (d+\mbox{const}\,\al)^{-n}\leq |\omega|\leq (d-\mbox{const}\,\al)^{-n}
2432: $$
2433: for each $\omega\in\cp_n$. Due to the large expansion of
2434: $f$ in the horizontal direction, we have that if $J\subset
2435: I$ is an interval with $|J|\le\al$, then for each
2436: $\omega\in\cp_n$
2437: \begin{equation}
2438: \label{e.cor2.3}
2439: m\big(\{s\in \omega\colon \wh X_j(s)\in S^1\times
2440: J\}\,\big)\le \mbox{const}\,\sqrt{|J|}\,m(\omega)
2441: \end{equation}
2442: see \cite[Corollary~2.3]{V}. The estimate (\ref{e.b2}) on
2443: the Lebesgue measure of $B_2(n)$ is now an easy consequence
2444: of (\ref{e.cor2.3}). For that we only have to compute the
2445: Lebesgue measure of $B_2(n)$ on each horizontal line of
2446: $S^1\times I$ and integrate. The estimate (\ref{e.b2}) on
2447: the Lebesgue measure of $B_1(n)$ is is obtained by mean of
2448: a large deviations argument applied to the horizontal
2449: curves in $S^1\times I$.
2450:
2451:
2452: \cre \label{r.uniform} The choice of the
2453: constants $c, \xi,\gamma$ and $ \delta $ only depends on
2454: the quadratic map $Q$ and $\al>0$. In particular the decay
2455: estimates on the Lebesgue measure of $B_1(n)$ and $B_2(n)$
2456: depend only on the quadratic map $Q$ and $\al>0$.
2457: \fre
2458:
2459:
2460:
2461:
2462: \subsubsection{Random perturbations}
2463:
2464: Let $f$ be
2465: close to $\hat f$ in the $C^3$ topology. As we have
2466: seen before, it is no restriction if we assume that
2467: $\cc=\{(s,x)\in S^1\times I\colon x=0\}$ is the critical set of $f$.
2468: Fix $\{\Phi,(\theta_\ep)_\ep\}$ a random perturbation of
2469: $f$ for which (\ref{e.perturbation}) holds.
2470: Our goal now is to prove that any
2471: such $f$ satisfies the hypotheses of
2472: Theorems~\ref{t.finite2} and \ref{t.sts} for $\ep>0$
2473: sufficiently small, and thus conclude
2474: that $f$ is stochastically stable. So, we want to show
2475: that if $\ep>0$ is small enough then
2476: \begin{itemize}
2477: \item $f$ is non-uniformly expanding for random orbits;
2478: \item random orbits have slow approximation to the critical set
2479: $\cc$;
2480: \item the family of hyperbolic time maps $(h_\ep)_\ep$ has uniform $L^1$-tail.
2481: \end{itemize}
2482:
2483:
2484:
2485: We remark that in the estimates we have obtained for
2486: $\log\|(Df(s_j,x_j))^{-1}\|$ and
2487: $\log\dist_\delta(x_j,\cc)$ over the orbit of a given point
2488: $(s,x)\in S^1\times I$, we can easily replace the iterates
2489: $(s_j,x_j)$ by random iterates $(s^j_{\un t},x^j_{\un t})=
2490: f_{\un t}^j(s,x)$. Actually, the methods used for obtaining
2491: estimate (\ref{e.principal}) rely on a delicate
2492: decomposition of the orbit of a given point $(s,x)$ from
2493: time 0 until time $n$ into finite pieces according to its
2494: returns to the neighborhood $S^1\times
2495: (-\sqrt\al,\sqrt\al)$ of the critical set. The main tools
2496: are \cite[Lemma~2.4]{V} and \cite[Lemma~2.5]{V} whose
2497: proofs may easily be mimicked for random orbits. Indeed,
2498: the important fact in the proof of the referred lemmas is
2499: that orbits of points in the central direction stay close
2500: to orbits of the quadratic map $Q$ for long periods, as long
2501: as $\al>0$
2502: is taken sufficiently small. Hence, such results can easily
2503: be obtained for random orbits as long as we take $\ep>0$
2504: with $\ep\ll\al$ and perturbation vectors $\un
2505: t\in\supp(\theta_\ep)$.
2506:
2507: Thus, the
2508: procedure of \cite{V} described in Subsection
2509: \ref{ss.non-uniform} applies to this situation, and we are
2510: able to prove that there is $c>0$, and for $\gamma>0$ there
2511: is $\de>0$, such that
2512: $$
2513: \sum_{j=0}^{n-1}\log\|Df(s^j_{\un t},x^j_{\un t}))^{-1}\|\le
2514: -cn
2515: \qand
2516: \sum_{j=0}^{n-1}-\log\dist_\delta(x^j_{\un t},\cc)\leq \gamma n
2517: $$
2518: for $(s,x)\notin B_1(n)\cup B_2(n)$, where
2519: $B_1(n)$ and $ B_2(n)$ are subsets $S^1\times I$ with
2520: $$
2521: m\big(B_1(n)\big)\le e^{-\xi n}\qand
2522: m(B_2(n))\le\mbox{const}\,e^{-\sqrt n/4}
2523: $$
2524: for some constant $\xi>0$ depending only on $\gamma$.
2525: This gives the
2526: non-uniform expansion and slow approximation to the critical set
2527: for random orbits. Moreover, the arguments show that we may
2528: take the map $N_\ep$ with
2529: $$ (\th_\ep^\NN\times
2530: m)\left(\big\{(\un t,x)\in T^\NN\times M\colon N_\ep(\un
2531: t,x)>n\big\}\right)\leq \mbox{const}\,e^{-\sqrt n/4},
2532: $$
2533: thus giving that the family of first hyperbolic time maps
2534: has uniform $L^1$-tail (see the considerations at the beginning
2535: of Section \ref{s.applications}).
2536:
2537:
2538: \medskip
2539:
2540:
2541: For the sake of completeness, an explanation is required on
2542: the way the Markov partitions $\cp_n$ of $S^1$ can be
2543: defined in this case, in order to obtain the estimates on
2544: the Lebesgue measure of $B_1(n)$ and $B_2(n)$. We consider
2545: $M=S^1\times I$ and define the skew-product map
2546: \[
2547: \begin{array}{rccc} F: & T^\NN\times M &\longrightarrow &
2548: T^\NN\times M,\\
2549: & (\un t, z) &\longmapsto & \big(\sigma(\un t),f_{t_1}(z)\big)
2550: \end{array}
2551: \]
2552: where $\si$ is the left shift map. Writing
2553: $f_t(z)=\big(g_t(z),q_t(z)\big)$ for $z=(s,x)\in S^1\times
2554: I$, we have that $q_t(s,\cdot)$ is a unimodal map close to
2555: $\hat q$ for all $s\in S^1$ and $t\in \supp(\theta_\ep)$
2556: with $\ep>0$ small.
2557:
2558:
2559: \cpr\label{p.foliation}
2560: Given $\un t\in T^\NN$ there is a
2561: $C^1$ foliation $\cf^c_{\un t}$ of $M$ such that if $L_{\un t}(z)$
2562: is the leaf of $\cf^c_{\un t}$ through a point $z\in M$,
2563: then
2564: \begin{enumerate}
2565: \item $L_{\un t}(z)$ is a $C^1$ submanifold of
2566: $M$ close to a vertical line in the $C^1$ topology;
2567: \item $f_{t_1}\big(L_{\un t}(z)\big)$ is contained in $L_{\si\un t}\big(f_{t_1}(z)\big)$.
2568: \end{enumerate}
2569: \fpr
2570: \dem
2571: This will be obtained as a consequence of the fact that
2572: the set of vertical lines constitutes a normally expanding
2573: invariant foliation for $\hat f$. Let $\ch$ be the space of
2574: continuous maps $\xi:T^\NN\times M\rightarrow [-1,1]$
2575: endowed with the sup norm, and define the map
2576: $A:\ch\rightarrow\ch$ by
2577: $$
2578: A\xi(\un t,z)=\frac{\partial_xq_{t_1}(z)\xi( F(\un t,
2579: z))-\partial_xg_{t_1}(z)} {-\partial_s q_{t_1}(z)\xi(F(\un
2580: t,z) )+\partial_s g_{ t_1}(z)}, \quad \un
2581: t=(t_1,t_2,\dots)\in T^\NN \qand z\in M.
2582: $$
2583: Note that $A$ is well-defined,
2584: since
2585: $$
2586: |A\xi(\un t,z)|\leq \frac{(4+\al+\epsilon)+\al+\epsilon}
2587: {-(\mbox{const }\alpha+\epsilon )+(d-\al-\epsilon)}
2588: <1\label{campo}
2589: $$
2590: for small $\alpha>0$ and $\epsilon>0$. Moreover,
2591: $A$ is a contraction on $\ch$: given $\xi,\zeta\in\ch$ and
2592: $(\un t,z)\in T^\NN\times M$ then
2593: \begin{eqnarray*}
2594: \lefteqn{|A\xi(\un t,z)-A\zeta(\un t,z)|}\hspace{2cm}\\
2595: &\leq &
2596: \frac{|\mbox{det}Df_{t_1}(z)|\cdot|\xi(\un t, z)-\zeta(\un t,z)|}
2597: {\big|\big(\!-\partial_s q_{t_1}(z)\xi(F(\un t,z))+\partial_s
2598: g_{t_1}(z)\big)\cdot\big (\!-\partial_s
2599: q_{t_1}(z)\zeta(F(\un t,z))+\partial_s
2600: g_{t_1}(z)\big)\big|}\\ &\leq &
2601: \frac{\big((d+\alpha+\epsilon)(4+\alpha+\epsilon)+\alpha+
2602: \epsilon\big)\cdot\big|\xi(\un t, z)-\zeta(\un t,z)\big|}
2603: {(d-\mbox{const}\alpha-\epsilon)^2}.
2604: \end{eqnarray*}
2605: This last
2606: quantity can be made smaller than $|\xi(\un t,z)-\eta(\un
2607: t, z)|/2$, as long as $\alpha$ and $\epsilon$ are chosen
2608: sufficiently small. This shows that $A$ is a contraction on
2609: the Banach space $\ch$, and so it has a unique fixed point
2610: $\xi^c\in\ch$.
2611:
2612: It is no restriction for our purposes if we
2613: think of $T$ as being equal to $\supp(\theta_{\ep})$ for
2614: some small $\ep$.
2615: Note that the map $A$ depends
2616: continuously on $F$ and for $\ep>0$ small enough the fixed
2617: point of $A$ is close to the zero constant map. This holds
2618: because we are choosing $\supp(\theta_{\ep})$ close to
2619: $\{t^*\}$, $f_{t^*}=f$ and $f$ close to $\hat f$. Then, for
2620: $\ep>0$ small enough, we have $\xi^c(\un t, \cdot)$ uniformly
2621: close to $\xi^c(\un t^*, \cdot)$ and it is not hard to
2622: check that $\xi^c_0=\xi^c(\un t^*, \cdot)$ is precisely the map
2623: whose integral leaves of the vector field $(\xi_0^c,1)$
2624: give the invariant foliation $\cf^c$ associated to $f_{t^*}=f$.
2625: Since this foliation depends continuously on the dynamics
2626: and for $f=\hat f$ we have $\xi_0^c\equiv 0$
2627: (see \cite[Section~2.5]{V}), we finally deduce
2628: that $\xi^c(\un t, \cdot)$ is uniformly close to zero for
2629: small $\ep>0$.
2630:
2631:
2632: We have defined $A$ in such a way
2633: that if we take $E^c(\un t, z)=\mbox{span}\{(\xi^c(\un
2634: t,z),1)\}$, then for every $\un t\in T^\NN$ and $z\in
2635: S^1\times I$
2636: \begin{equation} Df_{t_1}(z)E^c(\un t,z)\subset
2637: E^c(F(\un t,z)).\label{ec} \end{equation}
2638: Now, for fixed $\un t\in T^\NN$, we take $\cf^c_{\un t}$ to
2639: be the set of integral curves of the vector field
2640: $z\rightarrow (\xi^c(\un t ,z),1)$ defined on $S^1\times
2641: I$. Since the vector field is taken of class $C^0$, it does
2642: not follow immediately that through each point in
2643: $S^1\times I$ passes only one integral curve. We will prove
2644: uniqueness of solutions by using the fact that the map $f$
2645: has a big expansion in the horizontal direction.
2646:
2647: Assume, by
2648: contradiction, that there are two distinct integral curves
2649: $Y, Z\in\cf^c_{\un t}$ with a common point. So we may take
2650: three distinct nearby points $z_0,z_1,z_2\in S^1\times I$
2651: such that $z_0\in Y\cap Z$, $z_1\in Y$, $z_2\in Z$ and
2652: $z_1,z_2$ have the same $x$-coordinate. Let $X$ be the
2653: horizontal curve joining $z_1$ to $z_2$. If we consider
2654: $X_n=\pi_2\circ F^n(\un t,X)$ for $n\ge 1$, where $\pi_2$
2655: is the projection from $T^\NN\times S^1\times I$ onto
2656: $S^1\times I$, we have that the curves $X_n$ are nearly
2657: horizontal and grow in the horizontal direction (when $n$
2658: increases) by a factor close to $d$ for small $\al$ and
2659: $\ep$, see \cite[Section 2.1]{V}. Hence, for large $n$,
2660: $X_n$ wraps many times around the cylinder $S^1\times I$.
2661: On the other hand, since $Y_n=\pi_2\circ F^n(\un t,Y)$ and
2662: $Z_n=\pi_2\circ F^n(\un t,Z)$ are always tangent to the
2663: vector field $z\rightarrow \big(\xi^c(\sigma^n\un
2664: t,z),1\big)$ on $S^1\times I$, it follows that all the
2665: iterates of $Y_n$ and $Z_n$ have small amplitude in the
2666: $s$-direction. This gives a contradiction, since the closed
2667: curve made by $Y$, $Z$ and $X$ is homotopic to zero in
2668: $S^1\times I$ and the closed curve made by $Y_n$, $Z_n$ and
2669: $X_n$ cannot be homotopic to zero for large $n$. Thus, for
2670: fixed $\un t\in T^\NN$ we have uniqueness of solutions of
2671: the vector field $z\rightarrow (\xi^c(\un t,z),1)$, and
2672: from (\ref{ec}) it follows that $\cf^c_{\un t}$ is an
2673: $F$-invariant foliation of $M$ by nearly vertical leaves.
2674: \cqd
2675:
2676: Now, using the foliations given by the previous proposition
2677: we are also able to define the Markov partitions of $S^1$
2678: in this setting. Given any smooth map $X:S^1\ra I$ whose
2679: graph is nearly horizontal, denote
2680: $\wh X_{\un t}^n
2681: (s)=f_{\un t}^n\big(s,X(s)\big)$ for $n\ge0$ and $s\in
2682: S^1$. Take
2683: some leaf
2684: $L_{\un t}^0$ of the foliation $\cf^c_{\un t}$.
2685: Letting $L^n_{\un t}=f^n_{\un t}(L_{\un t})$ for $n\ge 1$,
2686: we define the sequence of Markov partitions $(\cp^n_{\un t
2687: })_n$ of $S^1$ as
2688: $$
2689: \cp^n_{\un t}=\left\{[s^\prime,s^{\prime\prime})\colon
2690: (s^\prime,s^{\prime\prime})
2691: \mbox{ is a connected component of }
2692: (\widehat{X}^n_{\un t})^{-1}\big((S^1\times I)\setminus L^n_{\un t
2693: }\big)\right\}.
2694: $$
2695: It is easy to check that $\cp^{n+1}_{\un t }$ refines
2696: $\cp^n_{\un t }$ for each $n\ge 1$ and, taking $\ep\ll\al$,
2697: $$
2698: (d+\mbox{const}\,\al)^{-n}\leq |\omega|\leq (d-\mbox{const}\,\al)^{-n}
2699: $$
2700: for each $\omega\in\cp^n_{\un t }$. This permits to obtain
2701: estimates for the Lebesgue measure of the sets $B_1(n)$ and
2702: $B_2(n)$ exactly in the same way as before also with the
2703: constants only depending on the quadratic map $Q$ (cf.
2704: Remark \ref{r.uniform}).
2705:
2706:
2707:
2708: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2709:
2710:
2711: \begin{thebibliography}{ABV}
2712:
2713: \bibitem[Al]{Al} J. F. Alves, {\em SRB measures for
2714: non-hyperbolic systems with multidimensional expansion},
2715: Ann. Scient. Éc. Norm. Sup., $4^e$ série, 33 (2000), 1-32.
2716:
2717: \bibitem[ABV]{ABV} J. F. Alves, C. Bonatti, M. Viana, {\em
2718: SRB measures for partially hyperbolic systems with mostly
2719: expanding central direction}, Invent. Math. (to appear).
2720:
2721: \bibitem[AV]{AV} J. F. Alves, M. Viana, {\em Statistical
2722: stability for robust classes of maps with non-uniform
2723: expansion}, preprint CMUP 1999.
2724:
2725: \bibitem[Ar]{Ar1} V. Ara\'ujo, {\em Attractors and time
2726: averages for random maps}, Ann. de l'Inst. H. Poincar\'e -
2727: Anal. non-Lin. (to appear).
2728:
2729: %\bibitem[Ar2]{Ar2} V. Ara\'ujo, {\em Infinitely many
2730: %stochastically stable attractors}, preprint CMUP 2000.
2731:
2732: \bibitem[BV]{BaV} V. Baladi, M. Viana, {\em Strong
2733: stochastic stability and rate of mixing for unimodal maps},
2734: Ann. Scient. Éc. Norm. Sup., $4^e$ série, 29 (1996),
2735: 483-517.
2736:
2737: \bibitem[BC1]{BC1} M. Benedicks and L. Carleson, {\em On
2738: iterations of $1-ax^2$ on $(-1,1)$}, Ann. Math. { 122}
2739: (1985), 1-25.
2740:
2741: \bibitem[BC2]{BC2} M. Benedicks and L. Carleson, {\em The
2742: dynamics of the H\'enon map}, Ann. Math. { 133} (1991),
2743: 73-169.
2744:
2745: \bibitem[BY1]{BY} M. Benedicks, L.-S. Young, {\em
2746: Absolutely continuous invariant measures and random
2747: perturbations for certain one-dimensional maps}, Erg. Th.
2748: \& Dyn. Sys. { 12} (1992), 13-37.
2749:
2750: \bibitem[BY2]{BY2} M. Benedicks and L.-S. Young, {\em
2751: SRB-measures for certain H\'enon maps}, Invent. Math. {
2752: 112} (1993), 541-576.
2753:
2754: \bibitem[BV]{BV} C. Bonatti, M. Viana, {\em SRB measures
2755: for partially hyperbolic systems with mostly contracting
2756: central direction}, Israel J. Math. (to appear).
2757:
2758: \bibitem[BR]{BR} R. Bowen and D. Ruelle, {\em The ergodic
2759: theory of Axiom A flows}, Invent. Math. { 29} (1975),
2760: 181-202.
2761:
2762: \bibitem[HPS]{HPS} M. Hirsch, C. Pugh, M. Shub, {\em
2763: Invariant manifolds}, Lect. Notes in Math. 583, Springer
2764: Verlag, 1977.
2765:
2766: \bibitem[Ja]{J} M. Jakobson, {\em Absolutely continuous
2767: invariant measures for one-parameter families of
2768: one-dimensional maps}, Comm. Math. Phys. { 81}
2769: (1981), 39-88.
2770:
2771: \bibitem[Ki1]{Ki1} Yu. Kifer, {\em Ergodic theory of
2772: random perturbations}, Birkh\"auser, Boston Basel, 1986.
2773:
2774: \bibitem[Ki2]{Ki2} Yu. Kifer, {\em Random perturbations
2775: of dynamical systems}, Birkh\"auser, Boston Basel, 1988.
2776:
2777: \bibitem[Ma]{Man} R. Ma\~n\'e, \emph{Ergodic theory and
2778: differentiable dynamics}, Springer-Verlag, 1987.
2779:
2780: \bibitem[Pl]{Pli} V. Pliss, {\em On a conjecture due to
2781: Smale}, Diff. Uravnenija, { 8} (1972), 262-268.
2782:
2783: \bibitem[Ru]{R} D. Ruelle, {\em A measure associated with
2784: Axiom A attractors}, Amer. Jour. Math. { 98} (1976),
2785: 619-654.
2786:
2787: \bibitem[Si]{Si} Ya. Sinai, {\em Gibbs measures in ergodic
2788: theory}, Russ. Math. Surv. { 27}, n. 4, (1972), 21-69.
2789:
2790: \bibitem[Vi1]{V} M. Viana, {\em Multidimensional
2791: non-hyperbolic attractors}, Publ. Math. IHES { 85} (1997),
2792: 63-96.
2793:
2794: \bibitem[Vi2]{V2} M. Viana, {\em Stochastic dynamics of
2795: deterministic systems}, Lect. Notes XXI Braz. Math Colloq.,
2796: IMPA, 1997.
2797:
2798: \bibitem[Yo]{Yo} L.-S. Young, {\em Stochastic stability of
2799: hyperbolic attractors}, Erg. Th. \& Dyn. Sys. {6} (1986),
2800: 311-319.
2801:
2802: \end{thebibliography}
2803:
2804: \bigskip
2805:
2806: \noindent Jos\'e Ferreira Alves ({\tt jfalves@fc.up.pt}) \\
2807: V\'\i tor Ara\'ujo ({\tt vdaraujo@fc.up.pt})\\
2808: Centro de Matem\'atica da Universidade do Porto\\
2809: Pra\c ca Gomes Teixeira,
2810: 4099-002 Porto, Portugal\\
2811:
2812:
2813:
2814: \end{document}
2815: