1: \documentclass{amsart}
2:
3: %VT
4: \usepackage{fullpage}
5:
6: \numberwithin{equation}{section}
7:
8:
9: %%%%%%%packages
10: \usepackage{amssymb}
11: \usepackage{enumerate}
12: %\usepackage{showkeys}
13: %\usepackage[bottom,first]{draftcopy}
14: \usepackage{changebar}
15: %\usepackage{xspace}
16:
17: %%%%%%%%%Pagination
18: \hfuzz=15pt
19:
20: %%%%%%%%%Theorems
21: \newtheorem{thm}{Theorem}[section]
22: \newtheorem{lem}[thm]{Lemma}
23: \newtheorem{cor}[thm]{Corollary}
24: \newtheorem{sublem}[thm]{Sub-lemma}
25: \newtheorem{prop}[thm]{Proposition}
26: \newtheorem{conj}{Conjecture}\renewcommand{\theconj}{}
27: \newtheorem{defin}{Definition}
28: \newtheorem{cond}{Condition}
29: \newtheorem{exmp}{Example}
30: \newtheorem{rem}[thm]{Remark}
31:
32: %%%%%%%%%%mathcal
33: \newcommand\B{{\mathcal B}}
34: \newcommand\Co{{\mathcal C}}
35: \newcommand\D{{\mathcal D}}
36: \newcommand\F{{\mathcal F}}
37: \newcommand\Ka{{\mathcal K}}
38: \newcommand\Lp{{\mathcal L}}
39: \newcommand\Or{{\mathcal O}}
40: \newcommand\M{{\mathcal M}}
41: \newcommand\T{{\mathcal T}}
42:
43: %%%%%%%%%%%%mathbb
44: \newcommand\A{{\mathbb A}}
45: \newcommand\C{{\mathbb C}}
46: \newcommand\E{{\mathbb E}}
47: \newcommand\N{{\mathbb N}}
48: \newcommand\R{{\mathbb R}}
49: \newcommand\To{{\mathbb T}}
50: \newcommand\Z{{\mathbb Z}}
51:
52: %%%%%%%%%%%greek
53: \newcommand\ve{\varepsilon}
54: \newcommand\eps{\epsilon}
55: \newcommand\vf{\varphi}
56:
57: %%%%%%%%%%%%%%%vaious
58: \newcommand\Id{\text{\bf Id}}
59: \newcommand{\marginnote}[1]
60: {\mbox{}\marginpar{\raggedright\hspace{0pt}{$\blacktriangleright$ #1}}}
61: \newcommand{\cs}{\operatorname{C}}
62: \newcommand\nn{\nonumber}
63:
64:
65: %%%%%%%%%%%%%%%%%%Figures
66: \setlength{\unitlength}{1mm}
67:
68: \begin{document}
69:
70: \title[Convergence to equilibrium for intermittent symplectic
71: maps]{Convergence to equilibrium for intermittent symplectic maps}
72:
73: \author{Carlangelo Liverani, Marco Martens}
74: \address{Carlangelo Liverani\\
75: Dipartimento di Matematica\\
76: II Universit{\`a} di Roma (Tor Vergata)\\
77: Via della Ricerca Scientifica, 00133 Roma, Italy.}
78: \email{{\tt liverani@mat.uniroma2.it}}
79: \address{Marco Martens\\
80: University of Groningen, Department of Mathematics
81: P.O. Box 800, 9700 AV Groningen, The Netherlands}
82: \email{{\tt m.martens@math.rug.nl}}
83: \date{\today}
84: \thanks{One of us (C.L.) would like to thank S.Vaienti and M.Benedicks
85: for helpful discussions. In addition, we acknowledge the support from
86: the ESF Program PRODYN, I.B.M., M.I.U.R. and NSF}
87: \begin{abstract}
88: We investigate a class of area preserving non-uniformly hyperbolic
89: maps of the two torus. First we establish some results on the
90: regularity of the invariant foliations, then we use this knowledge
91: to estimate the rate of mixing.
92: \end{abstract}
93: \maketitle
94:
95: %VT
96: \thispagestyle{empty}
97: \input{imsmark}
98: \SBIMSMark{2005/01}{January 2005}{}
99:
100: \section{Introduction}
101: \label{sec:intro}
102: In recent years the physics community has devoted an increasing
103: attention to anomalous properties of physical systems (e.g., anomalous
104: transport, anomalous diffusion, anomalous conductivity, etc.). Such
105: properties have proven
106: relevant in many fields such as thermal conductivity, kinetic
107: equations, plasma physics, etc. and they are widely believed to be
108: dynamical in nature. In fact, such phenomena seem to depend on the
109: weak chaotic properties of the underling dynamics, see \cite{Za} and
110: references therein for a detailed discussion. The
111: basic idea is that, while uniformly hyperbolic dynamics gives rise to
112: normal transport properties (consider for example the diffusive
113: behavior in a finite horizon Lorentz gas \cite{BuSi})
114: non-uniform hyperbolicity gives rise to different behavior
115: (e.g. the anomalous diffusion believed to occur in infinite horizon
116: Lorentz gas \cite{Bl}) due to weaker
117: mixing properties (e.g. polynomial decay of correlations) and regions
118: in which the motions is rather regular and where the systems spend a
119: substantial fraction of time ({\em sticky} regions).
120:
121: Unfortunately, the theoretical understanding of dynamical models with
122: polynomial decay of correlations is extremely limited, hence the
123: necessity to rigorously investigate relevant toy models. The only well
124: understood cases are expanding one dimensional maps with a neutral
125: fixed point. Such maps were proposed as a model of intermittent
126: behavior in fluids (\cite{PM}) and have been widely studied. It has been
127: proven that such maps enjoy polynomial decay of correlations with the
128: rate depending on the behavior of the fixed point \cite{LSV, Yo1,
129: Hu1, Hu2, Sa, Go1}. In addition, when the
130: decay of correlation is sufficiently slow, the observables do not
131: satisfy the Central Limit Theorem or the Invariance Principle but
132: rather, when properly rescaled, some stable law (\cite{Zw, Go2, PoSh}).
133:
134: Some, more partial, results exist for multidimensional expanding
135: maps \cite{PoYu} as well. Yet, the usual physically relevant models are
136: connected to Hamiltonian dynamics and, to our knowledge, no rigorous
137: results are available in such a situation. The simplest case which
138: retain some Hamiltonian flavor is clearly a two dimensional area
139: preserving map. In fact, mixing area preserving maps of the two
140: dimensional torus with a neutral fixed point (the simplest type of
141: sticky set) have been investigated
142: numerically \cite{ArPr} predicting the possibility of a polynomial decay
143: of correlations.
144: In this paper we consider a class of non-uniformly hyperbolic symplectic maps \eqref{eq:map}
145: of the two torus where the hyperbolicity breaks down because of a non-hyperbolic fixed point. In fact, the linearized
146: dynamics at the fixed point is a shear \eqref{eq:der}. We prove that
147: the decay of correlations is polynomial and, more precisely, decays at
148: least as $n^{-2}$ and, in some sense, one cannot expect much
149: more. Note that the example treated numerically in \cite{ArPr} is a
150: special case of the present setting. In \cite{ArPr} the predicted
151: decay was $n^{-2.5}$. This emphasizes the difficulty to investigate
152: such issues and the strong need for more theoretical work on the subject.
153:
154:
155: The result in the present paper is based on a precise quantitative
156: analysis of the angle between the stable and the unstable
157: direction. This angle turns out to degenerate approaching the origin
158: (where the non hyperbolic fixed point is located). Once such a control
159: is achieved it is possible to obtain a bound on the expansion and
160: contraction in the system. Such expansion turns out to be only
161: polynomial, in contrast with the uniformly hyperbolic case where it is
162: exponential. In turn, the bound on the expansion allows to study the
163: regularity of the stable and unstable foliation. It turns out that
164: they are $\Co^1$ away from the origin. This suffice to apply a simple
165: random approximation technique that allows estimating the speed of the
166: correlations.
167:
168: As the rate of convergence to equilibrium is of order $n^{-2}$, see Theorem
169: \ref{thm:main}, the
170: Central Limit Theorem holds for zero average observable, see Corollary
171: \ref{cor:clt},
172: so the model does not exhibit anomalous statistical behavior in this
173: respect. Yet, it clearly exhibits an intermittent behavior and it shows the
174: mechanism whereby slow decay of correlations may arise. The present
175: work emphasizes the need to carry out similar studies in cases where the
176: set producing intermittency has a more complex structure than a simple
177: isolated point.
178:
179: The paper is organized as follows: section \ref{sec:results} details
180: the model and makes precise the results. Section \ref{sec:fix} studies
181: the local dynamics at the fixed point and, in particular the
182: properties of its stable and unstable manifolds. This can be achieved
183: in many way, here we find most efficient to apply a variational
184: technique. Section \ref{sec:narrow} establishes a precise bound for
185: the angle between the stable and the unstable direction at each
186: point. As anticipated, such a bound yields an a priory bound on the
187: expansion and contractions rates in the systems, these are obtained in
188: section \ref{sec:expansion}. The latter result suffices to apply
189: standard distortion estimates that, in turn, allow to prove precise
190: results on the regularity of the invariant foliation and the
191: holonomies, see section \ref{sec:regularity} and section
192: \ref{sec:holo} respectively. Next, in section
193: \ref{sec:random} we introduce a random perturbation of the above map
194: and investigate its statistical properties that, thanks to the added
195: randomness, can be addressed fairly easily. The relevance of the above
196: random perturbation is that the limit of zero noise allows to easily obtain a
197: bound on the rate of mixing in the original map, we do this in section
198: \ref{sec:decay}. Finally, in section \ref{sec:lower}, we show that the
199: obtained bound is close to being optimal. The paper ends with Remark
200: \ref{rem:problems} pointing to the unsatisfactory nature of some of
201: the present results and the need to investigate the related open
202: problems.
203:
204:
205:
206: \section{The model and the results}
207: \label{sec:results}
208: For each $h\in\Co^{\infty}(\To^1,\To^1)$ we define the map
209: $T:\To^2\to\To^2$ by\footnote{\label{foot:uno} Note that the following formula is
210: equivalent, by the symplectic change of variable $q=x-y$, $p=y$, to
211: the map
212: \[
213: \widetilde T(q,p)=\begin{cases}
214: q+p&\quad\mod 1\\
215: p+h(q+p)&\quad \mod 1
216: \end{cases}
217: \]
218: which belongs to the standard map family. Yet, the functions $h$
219: considered here differ substantially from the sine function which would
220: correspond to the classical Chirikov-Taylor well known example.}
221: \begin{equation}\label{eq:map}
222: T(x,y)=\begin{cases}
223: x+h(x)+y&\quad\mod 1\\
224: h(x)+y&\quad\mod 1
225: \end{cases}
226: \end{equation}
227: We moreover require the following properties
228: \begin{enumerate}
229: \item $h(0)=0$ (zero is a fixed point);
230: \item $h'(0)=0$ (zero is a neutral fixed point)
231: \item $h'(x)>0$ for each $x\neq 0$ (hyperbolicity)
232: \end{enumerate}
233: Note that conditions (2--3) imply that zero is a minimum for $h'$, which forces
234: \[
235: h''(0)=0;\quad h'''(0)\geq 0.
236: \]
237: We will restrict to the generic case
238: \begin{enumerate}
239: \item[(4)] $h'''(0)>0$.
240: \end{enumerate}
241: In order to simplify the discussion we will also assume the following
242: symmetry
243: \begin{enumerate}
244: \item[(5)] $h(-x)=-h(x)$.
245: \end{enumerate}
246:
247: This means that we can write
248: \begin{equation}
249: \label{eq:h}
250: h(x)=bx^3+\Or(x^5).
251: \end{equation}
252: \begin{rem}
253: Note that two facts implied by the above assumptions are not necessary and could be
254: done away with at the price of more extra work: the hypothesis that
255: there is only one neutral fixed point (finitely many neutral periodic
256: orbits would make little difference) and the symmetry (5). We assume
257: such facts only to simplify the presentation of the arguments.
258: \end{rem}
259: Since the derivative of the map is given by
260: \begin{equation}\label{eq:der}
261: DT=\begin{pmatrix}
262: 1+h'(x)&1\\
263: h'(x)&1
264: \end{pmatrix}
265: \end{equation}
266: $\det(DT)=1$, thus the Lebesgue measure $m$ is an invariant measure (the
267: maps are symplectic). From now on we will consider the dynamical
268: systems $T: (\To^2,m)\rightarrow (\To^2,m)$.
269:
270: Formula \eqref{eq:der} and property (3) imply that
271: the cone $\Co_+=\{v\in\R^2\;|\;Q(v):=\langle v_1,\,v_2\rangle\geq 0\}$
272: is invariant for $DT$. In additions, it is easy to check that $D_\xi
273: T^2 \Co_+\subset \hbox{int }\Co_+\cup\{0\}$ for all
274: $\xi\in\To^2\backslash \{0\}$. From
275: this and the general theory, see \cite{LW}, follows immediately
276:
277: \begin{thm}
278: The above described dynamical systems are non-uniformly hyperbolic and mixing.
279: \end{thm}
280: \begin{exmp}
281: An interesting concrete example for the above setting is given by the function
282: $h(x):=x-\sin x$.
283: \end{exmp}
284: The question remains about the rate of mixing, this is the present
285: topic.
286: \begin{rem}
287: In the following by $\cs$ we designate a generic
288: constant depending only on $T$. Accordingly, its value may vary from
289: an occurrence to the next. In the instances when we will need a constant
290: of the above type but with a fixed value we will use sub-superscripts.
291: \end{rem}
292: \begin{thm}\label{thm:main}
293: For each $f,g\in\Co^{1}(\To^2,\R)$, $\int f=0$, holds
294: true\footnote{In fact, a slightly sharper bound holds, see \eqref{eq:sharp}.}
295: \[
296: \left|\int f g\circ T^n\right|\leq \cs
297: \|f\|_{\Co^{1}}\|g\|_{\Co^{1}}n^{-2}(\ln n)^4.
298: \]
299: \end{thm}
300: \begin{rem}
301: As in other similar cases \cite{LSV, L, Po} the logarithmic
302: correction is almost certainly
303: an artifact of the technique of the proof. It could probably be
304: removed by using a more sophisticated (and thus more technically
305: involved) approach. See also section \ref{sec:lower}.
306: \end{rem}
307: Form Theorem \ref{thm:main} many facts follow, just to give an
308: example let us mention the following result that can be obtained from
309: Theorem 1.2 in \cite{L3}.
310: \begin{cor}[CLT]\label{cor:clt}
311: Given $f\in\Co^1$, $\int f=0$, the random variable
312: \[
313: \frac 1{\sqrt n}\sum_{i=0}^{n-1}f\circ T^i
314: \]
315: converges in distribution to a Gaussian variable with zero mean and
316: finite variance $\sigma$. In addition, $\sigma=0$ iff there exists
317: $\vf\in L^1$ such that $f=\vf-\vf\circ T$.\footnote{In particular this means
318: that the average of $f$ on each periodic orbit must be zero.}
319: \end{cor}
320: The rest of the paper is devoted to the proof of
321: Theorem \ref{thm:main} that will find its conclusion in section
322: \ref{sec:decay}. The basic
323: fact needed in the proof, a fact of independent
324: interest and made quantitatively precise in Lemma \ref{lem:distreg1}, is
325: the following.
326: \begin{thm}\label{thm:mainlem}
327: The stable and unstable distributions are $\Co^1$ in $\To^2\backslash\{0\}$.
328: \end{thm}
329:
330:
331:
332: \section{The fixed point manifolds}\label{sec:fix}
333:
334:
335: As usual we start by studying the local dynamics near the fixed
336: point. The first basic fact is the existence of stable and unstable
337: manifolds. This is rather standard, yet since we need some
338: quantitative information we will construct them explicitly.
339:
340: Instead of constructing them via usual fixed point arguments it turns
341: out to be faster to use a variational method.
342:
343: \subsection{A variational argument}
344:
345: Let us consider, in a neighborhood of zero, the function
346: \begin{equation}\label{eq:generating}
347: L(x,x_1):=\frac 12 (x-x_1)^2+G(x)\;;\quad G(x):=\int_0^xh(z) dz.
348: \end{equation}
349: By setting
350: \[
351: \begin{split}
352: &y:=-\frac{\partial L}{\partial x}=x_1-x-h(x)\\
353: &y_1:=\frac{\partial L}{\partial x_1}=x_1-x
354: \end{split}
355: \]
356: we have $(x_1,y_1)=T(x,y)$, that is {\sl $L$ is a generating function
357: for the map \eqref{eq:map}}.
358:
359: Then, for each $a\in\R$, we define the Lagrangian $\Lp_a:\ell^2(\N)\to\R$ by
360: \begin{equation}\label{eq:lagrange}
361: \Lp_a(x):=\sum_{n=1}^\infty L(x_n,x_{n+1})+L(a,x_1).
362: \end{equation}
363: The justification of the above definition rests in the following
364: Lemma.
365: \begin{lem}\label{lem:critical}
366: For each $a\in(-1,1)$, holds true $\Lp_a\in\Co^1(\ell^2(\N))$. In addition, if
367: $x\in \ell^2(\N)$ is such that $D_x\Lp_a=0$, then setting $x_0=a$ and
368: $y_n:=x_{n+1}-x_n-h(x_n)$, we have $T^n(x_0,y_0)=(x_n,y_n)$.
369: \end{lem}
370: \begin{proof}
371: First of all \eqref{eq:h} implies that there exists $\cs>0$ such that
372: $|G(x)|\leq \cs x^4$. It is then easy to see that $\Lp_a$ is well defined
373: for each sequence in $\ell^2(\N)$.
374:
375: Next, for each $n\in\N$ let us define
376: $(\nabla\Lp_a)_n:=\partial_{x_n}\Lp_a$.
377: Clearly,
378: \[
379: \begin{split}
380: (\nabla\Lp_a)_1(x)&=2x_1-x_2+h(x_1)-a\\
381: (\nabla\Lp_a)_n(x)&=2x_n-x_{n+1}-x_{n-1}+h(x_n).
382: \end{split}
383: \]
384: Of course, for $x\in\ell^2(\N)$, $ \nabla\Lp_a(x)\in\ell^2(\N)$, it is
385: then trivial to check
386: that $D_x\Lp_a(v)=\langle \nabla\Lp_a(x), v\rangle$.
387: The last statement follows by a direct computation.
388: \end{proof}
389:
390: By the above Lemma it is clear that one can obtain the stable
391: manifolds of the fixed point from the critical points of $\Lp_a$, it
392: remains to prove that such critical points do exist. We will start by
393: considering the case $a\geq 0$.
394:
395: Define
396: \begin{equation}\label{eq:fixedset}
397: Q_B:=\{x\in\ell^2(\N)\;|\;\,|x_n-\frac A{n+c}|\leq B(n+c)^{-\frac
398: 32}\}
399: \end{equation}
400: where $A:=\sqrt{\frac 2b}$; $c:=\frac Aa$.
401:
402: It is immediate to check that $Q_B$ is compact and convex. In
403: addition, if $a$ is sufficiently small, then $G$ is strictly convex on
404: $[-2a,2a]$ which implies that $\Lp_a|_{Q_B}$ is strictly
405: convex. Accordingly, $\Lp_a$ has minimum in $Q_B$, moreover the strict
406: convexity implies that such a minimum is unique, for $a$ fixed.
407:
408: Let us call $x(a)$ the point in $Q_B$ where $\Lp_a$ attains its
409: minimum.
410: \begin{lem}\label{lem:stab}
411: For $a$ small enough, $D_{x(a)}\Lp_a=0$.
412: \end{lem}
413: \begin{proof}
414: Suppose that $\partial_{x_n}\Lp_a(x(a))\neq 0$ for some $n\in\N$, for
415: example suppose it is negative. Then $x(a)$ is on the border of $Q_B$,
416: say $x(a)_n=A(n+c)^{-1}+B(n+c)^{-\frac 32}$, otherwise we could increase
417: $x(a)_n$ and decrease $\Lp_a$ still remaining in $Q_B$, contrary to
418: the assumption. But then
419: \[
420: \begin{split}
421: \partial_{x_n}\Lp_a(x(a))&=2x(a)_n-x(a)_{n-1}-x(a)_{n+1}+h(x(a)_{n})\\
422: &=\frac {2A}{n+c}+\frac{2B}{(n+c)^{\frac
423: 32}}-x(a)_{n-1}-x(a)_{n+1}+h(\frac {A}{n+c}+\frac{B}{(n+c)^{\frac
424: 32}}) \\
425: &\geq -\frac{2A}{[(n+c)^2-1](n+c)}+\frac
426: {2A}{(n+c)^3}+\frac{3bA^2B}{(n+c)^\frac 72}+\Or((n+c)^{-4})\\
427: &+\frac{B(n+c)^3\{2(1-(n+c)^{-2})^{\frac 32}-(1-(n+c)^{-1})^{\frac
428: 32} -(1+(n+c)^{-1})^{\frac 32}}{(n+c)^{\frac 32}[(n+c)^2-1]^{\frac
429: 32}}\\
430: &=\frac{(6-\frac {15}4)B}{(n+c)^{\frac 72}}+\Or((n+c)^{-4})\geq 0
431: \end{split}
432: \]
433: provided $a$ is sufficiently small. We have thus a contradiction. The
434: other possibilities are analyzed similarly.
435: \end{proof}
436:
437: To conclude we need some information on the regularity of $x(a)$ as a
438: function of $a$. Unfortunately, the
439: implicit function theorem does not applies since $D^2\Lp_a$ does not
440: have a spectral gap, yet for our purposes a simple estimate
441: suffices.
442: \begin{lem}\label{lem:lipman}
443: $x_1(a)$ is a Lipschitz function of $a$. Moreover, when derivable,
444: \[
445: |y_0(a)'|\leq \cs |x(a)_0|.
446: \]
447: \end{lem}
448: \begin{proof}
449: By Lemma \ref{lem:stab} it follows, for each $a,a'$ sufficiently small
450: \[
451: \partial_{x_n}\Lp_{a'}(x(a'))=\partial_{x_n}\Lp_{a}(x(a))=0
452: \]
453: that is
454: \[
455: \partial_{x_n}\Lp_{a'}(x(a'))-\partial_{x_n}\Lp_{a'}(x(a))=\partial_{x_n}\Lp_{a}(x(a))
456: -\partial_{x_n}\Lp_{a'}(x(a))
457: \]
458: which yields
459: \begin{equation}\label{eq:strange}
460: \begin{split}
461: &(2+h'(\xi_1))\zeta_1-\zeta_2=a'-a\\
462: &(2+h'(\xi_n))\zeta_n-\zeta_{n+1}-\zeta_{n-1}=0,
463: \end{split}
464: \end{equation}
465: where $\zeta_n:=x(a')_n-x(a)_n$ and $\xi_n\in[x(a)_n,x(a')_n]$.
466:
467: Notice that, if $|\zeta_n|\geq|\zeta_{n-1}|$, then
468: \[
469: |\zeta_{n+1}|=|(2+h'(\xi_n))\zeta_n-\zeta_{n-1}|\geq
470: 2|\zeta_n|-|\zeta_{n-1}|\geq |\zeta_n|.
471: \]
472:
473: Thus, by induction, if $|\zeta_n|\geq |\zeta_{n-1}|$, then $|\zeta_m|\geq
474: |\zeta_{n-1}|$ for each $m\geq n$, which would imply $\zeta_{n-1}=0$ since
475: $\zeta\in\ell^2(\N)$. But then
476: $(2+h'(\xi_n))\zeta_n=\zeta_{n+1}$, that is
477: $|\zeta_{n+1}|\geq|\zeta_n|$. Accordingly, again by induction,
478: $\zeta_m=0$ for each $m\geq n-1$. This means that we can restrict
479: ourselves to the case $\zeta_n\neq 0$, $|\zeta_n|\geq |\zeta_{n+1}|$.
480: Hence,
481: \[
482: |a'-a|=|(2+h'(\xi_1))\zeta_1-\zeta_2|\geq 2|\zeta_1|-|\zeta_2|\geq
483: |\zeta_1|.
484: \]
485: That is $|x(a')_n-x(a)_n|\leq |x_1(a')-x_1(a)|\leq |a'-a|$.
486:
487: Finally, summing \eqref{eq:strange} over $n\in\N$,
488: \[
489: \sum_{n=1}^\infty h'(\xi_n)\zeta_n=-\zeta_1+a'-a.
490: \]
491: Thus, where all the $x(a)_n$ are differentiable (a full measure set),
492: $|x(a)_n'|\leq|x(a)_1'|\leq 1$ and
493: \begin{equation}\label{eq:derivxy}
494: \begin{split}
495: x(a)_1'&=1-\sum_{n=1}^\infty h'(x(a)_n)x(a)_n'\\
496: y(a)_0'&=-\sum_{n=0}^\infty h'(x_n)x(a)_n'.
497: \end{split}
498: \end{equation}
499: Accordingly,
500: \[
501: |y_0(a)'|\leq 6b\sum_{n=0}^\infty x_n^2\leq \cs a.
502: \]
503: \end{proof}
504:
505: Clearly, the above Lemma implies that, calling $(x,\gamma_s(x))$ the
506: graph of the stable manifold, $\gamma_s\in Lip(-1,1)$. The case $a\leq
507: 0$ and the unstable manifolds can be treated similarly, yet there
508: exists a faster--and more
509: instructive--way.
510:
511: \subsection{Reversibility}
512:
513: Notice that the map $T$ is {\sl reversible}
514: with respect to the transformations\footnote{While the reversibility for
515: $\Pi$ is a general fact, the one for $\Pi_1$ depends on the simplifying
516: symmetry hypothesis (5).}
517: \begin{equation}\label{eq:invol}
518: \Pi(x,y):=(x, -y-h(x)); \quad \Pi_1(x,y):=(-x, y+h(x))
519: \end{equation}
520: Indeed, $\Pi^2=\Pi_1^2=\Id$ and $\Pi T\Pi=\Pi_1 T\Pi_1=T^{-1}$.
521:
522: \begin{rem}\label{rem:rev}
523: The reversibility implies that, for $x\geq 0$,
524: $(x,\gamma_u(x))=\Pi(x,\gamma_s(x))$, and, for $x\leq 0$,
525: $(x,\gamma_u(x))=\Pi_1(-x,\gamma_s(-x))$ is the unstable manifold of
526: zero.
527: \end{rem}
528:
529: \subsection{A quasi-Hamiltonian}
530:
531: To study the motion near the fixed point it is helpful to find a local
532: ``Hamiltonian'' function. By Hamiltonian function we mean a function
533: that is locally invariant for the dynamics. Such a function can be
534: computed as a formal power series starting by the relation $H\circ
535: T=H$. In fact, we are interested only in a suitable approximation. A
536: direct computation yields that, by defining $G(x):=\int_0^x h(z)dz$ and
537: \begin{equation}\label{eq:hamiltonian}
538: H(x,y):=\frac 12 y^2-G(x)+\frac 12 h(x)y-\frac 1{12}h'(x)y^2+\frac 1 {12}h(x)^2,
539: \end{equation}
540: holds true\footnote{In fact, setting $(x_1,y_1):=T(x,y)$, holds
541: \[
542: \begin{split}
543: &H(x_1,y_1)-H(x,y)=yh(x)+\frac 12 h(x)^2-h(x)y-h(x)^2-\frac
544: 12h'(x)y_1^2-\frac 16 h''(x)y_1^3+\frac 12h'(x)y_1^2\\
545: &+\frac 12 h(x)^2
546: +\frac 14 h''(x)y_1^3-\frac 1{12}h''(x)y_1^3-\frac
547: 16h'(x)h(x)y_1+\frac 16 h'(x)h(x)y_1+\Or(x^8+x^4y^2+y^4).
548: \end{split}
549: \]
550: }
551: \begin{equation}\label{eq:appham}
552: H(T(x,y))-H(x,y)=\Or (x^8+y^4).
553: \end{equation}
554: This approximate conservation law suffices to obtain rather precise
555: information on the near fixed point dynamics.\footnote{The reader
556: should be aware that it is possible to do much better, that is to
557: obtain an exponentially precise conservation law, see \cite{L2},
558: \cite{BG}.}
559: The first application
560: is given by the following information on the stable manifold.
561: \begin{lem}\label{lem:manifzero}
562: For $x\geq 0$ sufficiently small holds
563: \[
564: \gamma_s(x)= -A^{-1}x^2+\Or(x^3).
565: \]
566: \end{lem}
567: \begin{proof}
568: Using the notation of Lemma \ref{lem:stab}, for fixed $a$ we get
569: \begin{equation}\label{eq:unoy}
570: y_n=x_{n+1}-x_n-h(x_n)=\Or((n+c)^{-\frac 32}).
571: \end{equation}
572: Hence
573: \[
574: H(x_n,y_n)=\Or((n+c)^{-3}).
575: \]
576: Using equation \eqref{eq:appham} we have
577: \[
578: H(x_0,y_0)=H(x_n,y_n)+\Or(\sum_{i=0}^{n-1}x_i^8+y_i^4)
579: \]
580: that is
581: \[
582: |H(x_0,y_0)|\leq \cs \left\{(n+c)^{-3}+
583: \sum_{i=0}^{\infty}(n+c)^{-6}\right\}\leq \cs
584: \left\{(n+c)^{-3}+c^{-5}\right\}.
585: \]
586: Taking the limit for $n$ to infinity in the above expression and
587: remembering the definition of $c$ follows
588: \[
589: |H(x_0,y_0)|\leq\cs x_0^{5}.
590: \]
591: Since equation \eqref{eq:unoy} implies $|y_0|=\Or(x_0^{\frac 32})$,
592: from \eqref{eq:hamiltonian} we have
593: \[
594: y_0^2-\frac b2 x_0^4+bx_0^3y_0=\Or(x_0^5)
595: \]
596: from which the lemma follows.
597: \end{proof}
598: According to Lemma \ref{lem:manifzero}, the local picture of the
599: manifolds is given by Figure \ref{fig:man}.
600: \begin{figure}[ht]\
601: \centering
602: \put(-40,0){\line(1,0){80}}
603: \put(0,-20){\line(0,1){40}}
604: \put(-6.5,-15){{\tt Fat sector}}
605: \put(30,2){{\tt Thin sector}}
606: \thicklines
607: \qbezier(-40, 10)(0,-10)(40,10)
608: \qbezier(-40, -10)(0,10)(40,-10)
609: \put(-39,10){$W^s$}
610: \put(-39,-14){$W^u$}
611: \put(40,11){$W^u$}
612: \put(40,-9){$W^s$}
613: \put(0,-24){\ }
614: \caption{The manifolds of the fixed point}\label{fig:man}
615: \end{figure}
616:
617: \subsection{Manifold regularity}
618: Since in the previous section we have seen that the manifold are
619: Lipschitz curves, we can define the dynamics restricted to the unstable
620: manifold:
621: \begin{equation}\label{eq:fu}
622: f_u(x):=x+h(x)+\gamma_u(x).
623: \end{equation}
624:
625: Our next task is to obtain sharper information on the manifolds
626: regularity.\footnote{Of course, the manifold should be as smooth as $h$, but
627: this results is not needed in the following while we do need an
628: explicit bound on the curvature.}
629: \begin{lem}\label{lem:manifoldreg}
630: The unstable manifold of the fixed point is $\Co^2$, apart from zero.
631: \end{lem}
632: \begin{proof}
633: It is clearly enough to show that $\gamma_u\in \Co^2$ apart from
634: zero. To do so call $u(x)=\gamma_u'(x)$ (the derivative exists almost
635: everywhere since $\gamma_u$ is Lipschitz). The tangent vector to the
636: unstable manifold has the form $(1,u)$. On the other hand
637: \[
638: \begin{split}
639: \lambda_u(f_u^{-1}(x))
640: \begin{pmatrix}1\\u(x)\end{pmatrix}&:=\begin{pmatrix}
641: 1+h'(f_u^{-1}(x))&1\\
642: h'(f_u^{-1}(x))&1
643: \end{pmatrix}
644: \begin{pmatrix}1\\u(f_u^{-1}(x))\end{pmatrix}\\
645: &=:\lambda_u(f_u^{-1}(x))\begin{pmatrix}1\\F(f_u^{-1}(x),u(f_u^{-1}(x)))
646: \end{pmatrix}
647: \end{split}
648: \]
649: where
650: \begin{equation}\label{eq:unsdef}
651: \begin{split}
652: &\lambda_u(x):=1+h'(x)+u(x)=f'_u(x)\\
653: &F(x,u):=1-\frac{1}{1+h'(x)+u}.
654: \end{split}
655: \end{equation}
656: Accordingly, setting $x_i:=f_u^{-i}(x)$, holds
657: $u(x_i)=F(x_{i+1},u(x_{i+1}))$. Next, let $v_i:=2A^{-1}x_i$, then a
658: direct computation yields $F(x_i,v_i)-v_{i-1}=\Or(x_i^3)$, thus
659: \[
660: |u(x_i)-v_i|\leq |F(x_{i+1},u(x_{i+1}))-F(x_{i+1}, v_{i+1})|+\cs x_i^3
661: \leq |u(x_{i+1})- v_{i+1}|+\cs x_i^3.
662: \]
663: By induction, and equation \eqref{eq:fixedset}, it follows
664: $|u(x)-2A^{-1}x|\leq \cs\sum_{i=0}^\infty x_i^3\leq \cs x^2$.
665:
666: On the other hand, given a different
667: point $z$, it holds
668: \[
669: \begin{split}
670: u(x)-u(z)&=F(x_1,u(x_1))-F(z_1,u(z_1))\\
671: &=\lambda_u(x_1)^{-1}\lambda_u(z_1)^{-1}
672: \left\{(u(x_1)-u(z_1))+h'(x_1)-h'(z_1))\right\}.
673: \end{split}
674: \]
675: Iterating the above equation yields
676: \begin{equation}\label{eq:unstreg}
677: \begin{split}
678: u(x)-u(z)=&\lambda_{u,n}(x)^{-1}\lambda_{u,n}(z)^{-1}(u(x_n)-u(z_n))\\
679: &\ \ +\sum_{k=1}^{n}\lambda_{u,k}(x)^{-1}\lambda_{u,k}(z)^{-1}(h'(x_k)-h'(z_k))
680: \end{split}
681: \end{equation}
682: where $\lambda_{u,n}(x):=\prod_{k=1}^{n}\lambda_u(f_u^{-k}(x))$.
683: Next, let $x=a$, then, accordingly to Lemma
684: \ref{lem:stab}, equation \eqref{eq:fixedset} and Remark \ref{rem:rev}, we have
685: $x_i= A(i+c)^{-1}+\Or((i+c)^{3/2})$. This means that, in a
686: sufficiently small neighborhood of zero, and for $z$ sufficiently
687: close to $x$ holds
688: \[
689: \begin{split}
690: \lambda_{u,n}(x)&\geq
691: \prod_{k=1}^{n}\left(1+v_k-\cs x_i^2\right)
692: \geq \prod_{k=1}^{n}\left(1+\frac 2{k+c}-\cs (k+c)^{-3/2}\right)\\
693: &\geq e^{\sum_{k=1}^{n}\frac{2}{k+c}-\cs (k+c)^{-3/2}}
694: \geq \cs (A^{-1}an+1)^{2}
695: \end{split}
696: \]
697: provided $x$ is close
698: enough to zero. The same estimate holds for $\lambda_{u,n}(z)$.
699:
700: This implies that $u$ is continuous. Indeed, for each $\ve>0$ choose
701: $n_0(x)\in\N$ such that $\lambda_{u,n_0(x)}(x_n)^{-1}\leq \ve$, then
702: \[
703: |u(x)-u(z)|\leq \sum_{k=1}^{n_0(x)}|h'(x_k)-h'(z_k)|+\ve
704: \]
705: and we can thus choose $z$ close enough to $x$ such that
706: $|u(x)-u(z)|\leq 2\ve$. Note that this implies the continuity of the
707: $\lambda_{u,n}$ as well.
708:
709: To conclude, we choose $n(z)$ such that $\cs(A^{-1}an(z)+1)^{-4}\geq
710: |x-z|^{1+\alpha}$, for some $\alpha>0$,
711: accordingly
712: \[
713: u(x)-u(z)=\sum_{k=1}^{n(z)}\lambda_{u,k}(x)^{-1}
714: \lambda_{u,k}(z)^{-1}(h'(x_k)-h'(z_k))+{\mathcal
715: O}(|x-z|^{1+\alpha})
716: \]
717: Since the series is uniformly convergent we have
718: \begin{equation}\label{eq:unstder}
719: u'(x)=\sum_{k=1}^{\infty}\lambda_{u,k}(x)^{-3}h''(x_k)
720: \end{equation}
721: from which the lemma follows.\footnote{Remark that to obtain the
722: result on a larger neighborhood it suffices to iterate the unstable
723: manifold forward.}
724: \end{proof}
725:
726:
727: \begin{rem}
728: All the above results for the unstable manifold $\gamma_u$ have the
729: obvious counterpart for the stable manifold $\gamma_s$ that can be
730: readily obtained via reversibility, see Remark \ref{rem:rev}.
731: \end{rem}
732:
733:
734: \section{A narrower cone field}\label{sec:narrow}
735:
736: Here our goal is to estimate the angle between stable and unstable manifolds.
737:
738:
739: More precisely, we wish to prove that there exists two constants $K_+,K_-
740: \in\R^+$ such that the cone field
741: $\Co_*(\xi):=\{(1,u)\in\R^2\;|\;K_-(|x|+\sqrt{|y|})\leq u\leq
742: K_+(|x|+\sqrt{|y|})\}$ contains the unstable direction (by
743: reversibility we can also define
744: the stable cone field $\Co^-_*$).
745:
746: \begin{prop}\label{prop:dista}
747: For each $\xi\in\To^2$ holds true $E^u(\xi)\in\Co_*(\xi)$,
748: $E^s(\xi)\in\Co_*^-(\xi)$.
749: \end{prop}
750:
751: The rest of the section is devoted to the proof of Proposition
752: \ref{prop:dista}.
753:
754: Clearly a problem arises only in a neighborhood of zero. Accordingly
755: the first step is to gain a better understanding of the dynamics near
756: zero.
757:
758:
759: \subsection{Near fixed point dynamics}\label{sub:nearfix}
760: For each $\delta>0$ let $Q_\delta:=[-\delta,\delta]^2$ be a square
761: neighborhood of zero. The manifolds of the fixed point divide
762: $Q_\delta$ into four sectors: two thin and two fat (see Figure
763: \ref{fig:man}). We will discuss explicitly the dynamics in the two
764: sectors below the unstable manifold (the other two being identical by
765: symmetry).
766:
767: \begin{lem}\label{lem:zdyn1}
768: For each $(x,y)\in\To^2$, $y\leq \gamma_u(x)$, let
769: $(x_n,y_n):=T^n(x,y)$, then it holds true
770: \[
771: x_n\leq f_u^n(x)\quad \forall n\geq 0.
772: \]
773: \end{lem}
774: \begin{proof}
775: First note that the trajectory will always remain below the unstable
776: manifold. Hence, by induction,
777: \[
778: \begin{split}
779: x_{n+1}&=x_n+h(x_n)+y_n\leq x_n+h(x_n)+\gamma_u(x_n)\\
780: &\leq f_u^n(x)+
781: h(f_u^n(x))+\gamma_u( f_u^n(x))=f_u^{n+1}(x).
782: \end{split}
783: \]
784: \end{proof}
785: The above lemma will suffice to control the dynamics in the thin
786: sector, more work is needed for the fat one. In fact, when the
787: trajectories are close to the stable or the unstable manifolds the
788: above result can still be used (possibly remembering reversibility). On the
789: other hand when the trajectory is close enough to zero its behavior
790: is drastically different from the one on the invariant manifolds.
791:
792: To define more precisely the meaning of ``close to zero'' let us
793: introduce the parabolic sector $P_M:=\{(x,y)\in Q_1\;|\;|y|\geq Mx^2\}$.
794: We consider a backward trajectory starting from $x\leq 0$, $y\leq
795: \gamma_u(x)$, the other possibilities follow by reversibility. Let, as usual, $(x_{n},
796: y_n):=T^n(x,y)$, $n\in\Z$. Let $m_+$ be the smallest integer for which
797: $(x_{-n},y_{-n})\in P_M$, $m$ the largest integer such that
798: $x_{-m}\leq 0$, and $m_-$ the largest integer for which
799: $(x_{-n},y_{-n})\in P_M$. Define, see \eqref{eq:hamiltonian},
800: \[
801: E:=H(x_{-m},y_{-m}).
802: \]
803: In addition, define the function $\Upsilon_E: [-1,1]\to\R^-$,
804: by\footnote{Computing for $y\leq0$ yields
805: \[
806: \begin{split}
807: \Upsilon_E(x)&=-\frac {h(x)}2-\sqrt{\frac{h(x)^2}4+2[G(x)+E-\frac
808: 1{12}h(x)^2](1-\frac 16 h'(x))}\\
809: &=-\sqrt{2(E+G(x))}(1+\Or(x^2))+\Or(x^3).
810: \end{split}
811: \]
812: }
813: \begin{equation}\label{eq:upsi}
814: H(x,\Upsilon_E(x))=E.
815: \end{equation}
816: Then, by \eqref{eq:appham}, and since $|y_{-n}|\geq \cs x_{-n}^2$,
817: \[
818: \begin{split}
819: H(x_{-n},y_{-n})-&H(x_{-n},\Upsilon_E(x_{-n}))=H(x_{-n},y_{-n})-H(x_{-m},y_{-m})\leq
820: \cs \sum_{k=n}^m y_{-k}^4\\
821: &\leq\cs \sum_{k=n}^m |y_{-k}^3|(x_{-k-1}-x_{-k})
822: \leq \cs|y_{-n}^3| |x_{-n}|.
823: \end{split}
824: \]
825: Accordingly, for $n\leq m$ it holds true
826: \begin{equation}\label{eq:dyny}
827: |y_{-n}-\Upsilon_E(x_{-n})|\leq \cs|y_{-n}|^2|x_{-n}|.
828: \end{equation}
829: \begin{lem}\label{lem:zdyn2}
830: In the above described situation, setting $M=\sqrt b$, the following holds true
831: \begin{enumerate}
832: \item $f_u^{-n}(x)\leq x_{-n}\quad \forall\; n\leq m$;
833: \item $f_s^{k}(x_{-n})\geq x_{-n+k}\quad \forall\; n\geq m_-,\ k\leq
834: n-m_-$;
835: \item $\sqrt{E}\leq |y_{-n}|\leq 3\sqrt{E}$ for all
836: $n\in\{m_+,\dots, m_-\}$;
837: \item $m_+\leq 2A(Eb^{-1})^{-\frac 14}$;
838: \item $ 2(12Eb)^{-\frac 14}\leq m_--m_+\leq 4(Eb)^{-\frac 14}$.
839: \end{enumerate}
840: \end{lem}
841: \begin{proof}
842: The first fact is proven as in Lemma \ref{lem:zdyn1}, the second
843: follows by reversibility. Hence, by
844: the results of section \ref{sec:fix}, for $n\leq m$, it follows
845: \begin{equation}\label{eq:nest}
846: |x_{-n}|\leq |f_u^{-n}(x)|\leq \frac{2A|x|}{|x|n+A}.
847: \end{equation}
848: Next we want to determine the points $x_{m_+}$ and $x_{m_-}$. The idea
849: is to use $\eqref{eq:dyny}$ that determines with good precision the
850: geometry of the trajectories. Let $\bar x$ be defined by
851: $\Upsilon_E(\bar x)=-M\bar x^2$. Then
852: \[
853: \bar x=-\left[\frac{2E}{M^2-\frac b2}\right]^{\frac 14}+\Or(E^{\frac
854: 54}).
855: \]
856: On the other hand, since by definition $|y_{-m_+}|\geq Mx_{-m_+}^2$
857: and $|y_{-m_++1}|\leq Mx_{-m_++1}^2$, holds
858: $|y_{-m_+}-Mx_{-m_+}^2|\leq \cs|x_{-m_+}^3|$.
859: Hence, by \eqref{eq:dyny} it follows
860: \[
861: |x_{-m_+}-\bar x|\leq \cs\frac{x_{-m_+}^3}{\min\{\bar
862: x,x_{-m_+}\}}\leq \cs \frac{x_{-m_+}^3}{x_{-m_+}-|x_{-m_+}-\bar x|}.
863: \]
864: Solving the above inequality yields
865: \[
866: |x_{-m_+}-\bar x|\leq\cs x_{-m_+}^2\leq \cs\bar x^2\leq\cs
867: \sqrt E.
868: \]
869: Analogously $|x_{-m_-}+\bar x|\leq \cs \sqrt E$.
870: \relax From this (3) and (4) easily follows. Finally,
871: \[
872: 2|\bar x|\geq |x_{m_-}-x_{m_+}|=|\sum_{n=m_+}^{m_-}y_n|\geq
873: \cs(m_+-m_-)\sqrt E.
874: \]
875: Which implies (5).
876: \end{proof}
877:
878: We are now ready to refine our knowledge of the stable and unstable
879: direction. Let us fix $\varrho\in(0,1/2)$.
880:
881: \subsection{The cone field--Outside $Q_\varrho$}
882:
883: The general idea is to take the positive cone field $\Co_+$ (which is
884: invariant and contains the unstable direction) and to push it forward
885: in order to obtain a narrower cone
886: field. First of all outside $Q_{\sqrt\varrho}$ we have (see \eqref{eq:unsdef})
887: \[
888: 1\geq F(x,u)\geq F(x,0)\geq 2b \varrho,
889: \]
890: where we have chosen $\varrho$ small enough.
891: Hence the cone field $\Co_0=\{(1,u)\;|\;1\geq u\geq 2b\varrho\}$ is invariant
892: outside $Q_{\sqrt\varrho}$. It remains to understand what
893: happens in a neighborhood of the origin of order $\sqrt\varrho$.
894:
895: Let us define $\bar u(\xi)$ by the equation $u=F(\xi,u)$. An easy
896: computation shows
897: \[
898: \bar u(\xi)= \frac{-h'(x)+\sqrt{h'(x)^2+4h'(x)}}2= \sqrt{3b}|x|+\Or(x^2).
899: \]
900: By reversibility we can restrict ourselves to the case $x\geq 0$, in this
901: case the only possibility to enter the region $Q_{\sqrt\varrho}$ is
902: via the fourth quadrant.
903:
904: Note that $F(\xi,u)\geq u$ provided $0\leq u\leq \bar u(\xi)$.
905: This means that if $\xi\not\in Q_{\sqrt\varrho}$ but $T\xi\in
906: Q_{\sqrt\varrho}$, then the lower bound of the cone $D_\xi T^n\Co_0$
907: does not decreases until $\sqrt{3b}x_i\leq 2b\varrho$, where
908: $(x_i,y_i):=T^i\xi$.
909:
910: Accordingly, the cone field $\Co_0$ is invariant
911: also in the fourth quadrant, outside the set $Q_{2\sqrt{\frac b3}\varrho}$.
912: Now consider the cone field $\Co_1(\xi):=\{(1,u)\;|\;2b\varrho\leq
913: u\leq L \bar u(\xi)\}$, $L=(3b\varrho)^{-\frac 12}$, for $\xi\in
914: Q_{\sqrt\varrho}\backslash Q_{2\sqrt{\frac b3}\varrho}$. Clearly, if
915: $|x|\geq \sqrt \varrho$, then $L\bar u(\xi)\geq 1$. Hence, as the
916: point enters $Q_{\sqrt\varrho}$, the image of $\Co_0$ is contained in
917: $\Co_1$, moreover we have already seen that the lower bound is
918: invariant provided $\xi_i\not\in Q_{2\sqrt{\frac b3}\varrho}$. Let us follow the
919: upper edge, if $u\leq L\bar u(\xi_i)$, then\footnote{Note that this
920: computation holds for all $\xi_i\in Q_{\sqrt{\varrho}}\backslash P_M$.}
921: \[
922: \begin{split}
923: F&(\xi_i,u)=F(\xi_i,u)-F(\xi_i,\bar u(\xi_i))+\bar u(\xi_i)\\
924: &\leq \frac{(L-1)\bar u(\xi_i)}{(1+h'(x_i)+L\bar u(\xi_i))(1+h'(x_i)+\bar
925: u(\xi_i))}+\bar u(\xi_i)\\
926: &\leq L\bar u(\xi_i)-(L^2-1)\bar u(\xi_i)^2+\Or(x^3_i)\\
927: &\leq L\bar u(\xi_{i+1})+\sqrt{3b}|y_{i+1}|-(L^2-1)\bar u(\xi_i)^2+\Or(x^3_i)\\
928: &\leq L\bar u(\xi_{i+1}),
929: \end{split}
930: \]
931: provided $\xi_{i+1}\not\in P_M$, with $M\leq \sqrt{3b}(L^2-1)$, which is
932: fine provided $\varrho$ is chosen small enough. The above discussion
933: can be summarized as follows.
934: \begin{lem}
935: \label{lem:conefield-one}
936: There exists $\varrho>0$: For $\xi\not\in Q_{2\sqrt{\frac b3}\varrho}$
937: the unstable distribution is contained in $\Co_0$. In addition, in the
938: set $\{\xi=(x,y)\in Q_{\sqrt{\varrho}}\setminus (P_M\cup
939: Q_{2\sqrt{\frac b3}\varrho})\;:\; xy\leq 0\}$ the unstable direction
940: is contained in $\Co_1$.
941: \end{lem}
942:
943: To conclude we need to study what happens in a neighborhood of the
944: origin of order $\varrho$. It is necessary to
945: distinguish two possibilities: one can enter below the stable
946: manifold, and hence be confined in the fat sector, or one can enter
947: above the stable manifold, thereby being bound to the thin sector.
948: We will start with the easy case: the second.
949:
950: \subsection{The cone field--Thin sector}
951:
952: If $x>0$, as soon as the trajectory, at some time $n$, enters
953: $Q_{2\sqrt{\frac b3}\varrho}$ we have that
954: $(1,u)\in\Co_0$ implies $u\geq 2b\varrho\geq \gamma_u'(x_n)$.
955: Let us consider in $Q_{2\sqrt{\frac b3}\varrho}$ the cone field
956: $\Co_2:=\{(1,u)\;|\; L\bar u(\xi)\geq u(x)\geq \gamma_u'(x)\}$. Note
957: that, upon entering in $Q_{2\sqrt{\frac b3}\varrho}$ such a cone contains
958: $\Co_1$.\footnote{Note that, in such a case, the trajectory cannot enter in $P_M$.}
959: Now
960: \[
961: F(x,u)\geq F(x,\gamma_u'(x))=\gamma_u'(x+h(x)+\gamma_u(x))\geq
962: \gamma_u'(x+h(x)+y),
963: \]
964: where we have use that $\gamma_u''(x)\geq 0$ for $x\in[0,\varrho]$,
965: provided $\varrho$ has been chosen small enough.
966: \begin{lem}
967: \label{lem:conefield-two}
968: In the region $Q_{2\sqrt{\frac b3}\varrho}\setminus(
969: P_M\cup\{\xi=(x,y)\in\To^2\;:\; y\geq \gamma^u(x) \text{ for }x>0;
970: y\leq \gamma^u(x) \text{ for }x<0\})$ the
971: unstable direction is contained in the cone field $\Co_2$.
972: \end{lem}
973: Note that the above lemma suffices for trajectories in the thin sector.
974: The situation it is not so simple in the fat sector since the lower
975: bound would deteriorate to zero. A more detailed analysis is needed.
976:
977: For each $\xi=(x,y)\in\To^2$, for which the unstable direction is defined,
978: let $(1,u(\xi))$ be the vector in the unstable direction. Define then
979: $\lambda_{u,n}(\xi, u)$ and $F_n(\xi,u)$ as in formulae \eqref{eq:unsdef}
980: and \eqref{eq:unstreg} and similarly define the stable
981: quantities. That is
982: \begin{equation}\label{eq:constex}
983: \begin{split}
984: &D_{T^{-n}\xi} T^{n}(1,u)=:\lambda_{u,n}(\xi,u)(1,F_n(\xi,u))\\
985: &D_{\xi} T^{-n}(1,-v)=:\mu_{s,n}(\xi,v)(1,-F^-_n(\xi,v))
986: \end{split}
987: \end{equation}
988:
989: \subsection{The cone field--Fat sector}
990: First of all notice that the trajectory can enter
991: $Q_{2\sqrt {\frac b3} \varrho}$ either in $P_M$ or outside.
992: Since the cone field $\Co_2$ for $x\geq
993: 0$, $\xi\not\in P_M$ contains the unstable vector (Lemma
994: \ref{lem:conefield-two}), we have a good control on the
995: unstable vector in both cases until we enter in $P_M$.
996: Upon entering $P_M$, we will obtain a very sharp control on the
997: evolution of the edges of the cone. Let $\xi\not\in P_M$, $T\xi\in
998: P_M$, and let $\ell_+-1>0$ be the smallest integer such that
999: $\xi_n\not\in P_M$. By equation \eqref{eq:unsdef}, we have
1000: \[
1001: u_n:=F_n(\xi_n,u)=\sum_{i=1}^{n}
1002: \lambda_{u,i}(\xi_{n-i},u_{n-i})^{-1}h'(x_{n-i})+\lambda_{u,n}(\xi_n,u)^{-1}
1003: u.
1004: \]
1005: Then, for each $n< \ell_+$, holds true
1006: \[
1007: u_n\leq \sum_{i=1}^n h'(x_{n-i})+u\leq \frac {\cs }{M}
1008: \left|\sum_{i=1}^n y_i\right|+u.
1009: \]
1010: By Lemma \ref{lem:zdyn2}-(3),(5), it follows that
1011: we have, for $u\in\Co_2(\xi)$,
1012: \[
1013: u_n\leq \cs_+\sqrt{|y_n|}.
1014: \]
1015: Moreover, remembering \eqref{eq:unsdef} and that $u\in\Co_2(\xi)$, yields
1016: \[
1017: u_n\geq e^{-2n\cs_+\sqrt{|y_n|}}u\geq \cs u\geq \cs_-\sqrt{|y_n|}.
1018: \]
1019: Consequently, if for $\xi=(x,y)$ we define the cone
1020: $\Co_3(\xi)=\{\cs_-\sqrt{|y|}\leq u\leq \cs_+\sqrt{|y|}\}$., then the
1021: above results can be written as follows.
1022: \begin{lem}
1023: \label{lem:conefield-three}In $P_M$ the unstable direction
1024: is contained in the cone field $\Co_3$.
1025: \end{lem}
1026: Finally we have to follow the trajectory outside $P_M$ until it exits
1027: from $Q_{2\sqrt{\frac b3}\varrho}$. The upper bound can be treated as
1028: before. Not so for the lower bound.
1029:
1030: Let $\xi=(x,y)$ be a point in the
1031: fat sector, $x\leq 0$, $x_{-1}\geq 0$. Then, remembering subsection
1032: \ref{sub:nearfix}, let $E:=H(x,y)$, $u_0=0$ and $u_{n+1}:=F(x_n,
1033: u_n)$. Clearly, $D_{(x,y)}T^n\Co_+\subset\{(1,u)\in\R^2\;|\; u\geq
1034: u_n\}\cup\{0\}$.
1035: \begin{lem}
1036: \label{lem:flowdir}
1037: In the situation described above, for each $n\in\N$, holds true
1038: \[
1039: F(x_n,\Upsilon_E'(x_n))-\Upsilon_E'(x_{n+1})=\Or(|y_n|^{3/2}).
1040: \]
1041: \end{lem}
1042: \begin{proof}
1043: Notice that, since the trajectory lies below the unstable manifold,
1044: $|y|\geq \cs x^2$. It is then convenient to keep track of the orders
1045: of magnitude only in terms of powers of $y$.
1046: \[
1047: F(x_n,\Upsilon_E'(x_n))=\Upsilon_E'(x_n)-\Upsilon_E'(x_n)^2
1048: +h'(x_n)+\Or(|y_n|^{3/2}).
1049: \]
1050: On the other hand, differentiating \eqref{eq:upsi}, one gets
1051: \[
1052: \Upsilon_E'(x)=\frac{h(x)}{\Upsilon_E(x)}
1053: -\frac{h(x)^2}{2\Upsilon_E(x)^2} -\frac{h'(x)}2+\Or(|\Upsilon_E|^{3/2}).
1054: \]
1055: Accordingly, by \eqref{eq:dyny},
1056: \[
1057: \begin{split}
1058: \Upsilon_E'(x_{n+1})&=\frac{h(x_n)+h'(x_n)\Upsilon_E(x_n)}{\Upsilon_E(x_n)+h(x_n)}
1059: -\frac{h(x_n)^2}{2\Upsilon_E(x_n)^2}-\frac{h'(x_n)}2
1060: +\Or(|y_n|^{3/2})\\
1061: &=\Upsilon_E'(x_{n})-\Upsilon_E'(x_{n})^2 +h'(x_n)+\Or(|y_n|^{3/2}),
1062: \end{split}
1063: \]
1064: from which the Lemma easily follows.
1065: \end{proof}
1066: Since $F$ is a contraction in $u$,
1067: we can estimate
1068: \[
1069: \begin{split}
1070: |u_n-\Upsilon_E'(x_n)|&=|F(x_{n-1}, u_{n-1})-F(x_{n-1},\Upsilon_E'(x_{n-1}))|
1071: +\Or(|y_{n-1}|^{3/2})\\
1072: &\leq |u_{n-1}-\Upsilon_E'(x_{n-1})|+\Or(|y_{n-1}|^{3/2})\\
1073: &\leq|\Upsilon_E'(x_0)|+\Or\left(\sum_{k=0}^{n-1}|y_k|^{3/2}\right)\\
1074: &=\Or\left(|y_0|+\sum_{k=0}^{n-1}\sqrt{|y_k|}(x_{k}-x_{k+1})\right)
1075: =\Or(|y_n|).
1076: \end{split}
1077: \]
1078: We have thus proved that there exists a constant $\cs_0>0$ such that
1079: \begin{equation}\label{eq:fatcone}
1080: u_n\geq \Upsilon_E'(x_n)-\cs_0\Upsilon_E(x_n).
1081: \end{equation}
1082: Hence outside $P_M$ the image of the cone will belong to the
1083: cone field $\Co_3:=\{(1,u)\in\R^2\;|\; u(\xi)\geq
1084: \Upsilon'_{E(\xi)}(x)-\cs_0 \Upsilon_{E(\xi)}(x)\}$. Note that, upon exiting $P_M$,
1085: $\Upsilon_E'(x_n)-\cs_0\Upsilon_E(x_n)\geq \cs_-'\sqrt{|y_n|}$,
1086: provided $\varrho$ is chosen small enough.
1087: The Proposition follows by choosing $\varrho$ small enough and
1088: remembering Lemmata \ref{lem:conefield-one}, \ref{lem:conefield-two}
1089: and \ref{lem:conefield-three}.
1090:
1091:
1092: \section{An a priori expansion bound}
1093: \label{sec:expansion}
1094:
1095: The results of the previous section allow to obtain the following nice
1096: estimate on the expansion in the system.
1097: \begin{lem}\label{lem:expansion}
1098: There exists $K>0$ such that, for each
1099: $\xi=(x,y)\in\To^2\backslash \{0\}$, $n\in\N$ and
1100: $(1,u)\in\Co_*(T^{-n}\xi)=\Co_*(\xi_{-n})$, holds true
1101: \[
1102: \lambda_{u,n}(\xi,u)\geq e^{-K|x|} \left(K^{-1}
1103: |x|n+1\right)^2 .
1104: \]
1105: \end{lem}
1106: \begin{proof}
1107: Let us fix $\delta>0$. On the one hand, if the trajectory lies outside
1108: of $Q_\delta$, then
1109: we have an exponential expansion, on the other hand, if the backward
1110: trajectory enjoys $|x_{-n}|\geq |x_0|$, then equation \eqref{eq:unsdef}
1111: and Proposition \ref{prop:dista} imply
1112: \begin{equation}
1113: \label{eq:trivialexp}
1114: \lambda_{u,n}(\xi,u)\geq (1+K_-|x_0|)^n\geq e^{-K|x_0|} \left(K^{-1}
1115: |x_0|n+1\right)^2 .
1116: \end{equation}
1117: We say that the backward orbit of $\xi$ (up to time $n$) passes $p$ times
1118: thru
1119: $Q_\delta$ if $\{0\leq k\leq n\;:\;\xi_{-k}\in Q_\delta\}$ consists of
1120: $p$ intervals. The Lemma holds for orbits that pass zero-times thru
1121: $Q_\delta$. Suppose it holds for orbits that pass $p$ times. Let
1122: $\xi_{-n}\in Q_\delta$ and let $m<n$ be the last time $\xi_{-m}\not\in
1123: Q_\delta$ but it passed already $p$ times in $Q_\delta$. Moreover,
1124: suppose that the Lemma holds in $Q_{2\delta}$. Accordingly, for each
1125: $n\geq l$ such that $\xi_{-n}\in Q_\delta$ holds
1126: \[
1127: \begin{split}
1128: \lambda_{u,n}(\xi,u)&\geq e^{-K|x_0|} \left(K^{-1}
1129: |x_0|m+1\right)^2e^{-2K\delta} \left(2K^{-1}
1130: \delta(n-m)+1\right)^2\\
1131: &\geq e^{-K|x_0|} \left(K^{-1}
1132: |x_0|m+1\right)^2\left(K^{-1}\delta(n-m)+1\right)^2\\
1133: &\geq e^{-K|x_0|} \left(K^{-1}|x_0|n+1\right)^2,
1134: \end{split}
1135: \]
1136: provided $\delta$ has been chosen small enough and since it must be $n-m\geq \cs
1137: \delta^{-1}$.
1138: Thus to prove the Lemma it suffices to prove it for the pieces of
1139: trajectories in $Q_\delta$. There are two cases: a trajectory enters in the thin
1140: sector or in the fat one. Let us consider the thin sector first.
1141: Set $u_{-j}:=F_{n-j}(\xi_{-j},u)$.
1142: By the usual distortion estimates follows
1143: \[
1144: \begin{split}
1145: \lambda_{u,n}(\xi,u)&=\prod_{j=1}^n(1+h'(x_{-j})+u_{-j})\geq
1146: \prod_{j=1}^n(1+h'(x_{-j})+\gamma_u'(x_{-j}))\\
1147: &\geq
1148: \prod_{j=1}^n(1+h'(f_u^{-j}(x))+\gamma_u'(f_u^{-j}(x)))
1149: =\prod_{j=1}^n(f_u)'(f_u^{-j}(x))\\
1150: &\geq\prod_{j=1}^n\frac
1151: {|f_u^{-j+1}(x)-f_u^{-j}(x)|}{|f_u^{-j}(x)-f_u^{-j-1}(x)|}
1152: e^{-\cs|f_u^{-j+1}(x)-f_u^{-j}(x)|}\\
1153: &\geq e^{-\cs|x_0|}\frac{|x_0|^2}{|f_u^{-n}(x_0)|^2}.
1154: \end{split}
1155: \]
1156: Now, notice that $f_u^{-1}(x)\leq \frac{x}{1+\cs x}$, hence
1157: $f_u^{-n}(x_0)\leq \frac{|x_0|}{1+\cs n|x_0|}$. Thus,
1158: \begin{equation}
1159: \label{eq:bah}
1160: \lambda_{u,n}(\xi,u)\geq e^{-\cs|x_0|}(1+n\cs|x_0|)^2.
1161: \end{equation}
1162: \relax For the fat sector we need only to consider the cases in which $x_0\not\in
1163: Q_\delta$ and $x_0\in Q_\delta$, $x_0\leq 0$ since if $x_0>0$ the
1164: backward trajectory increases the $x$ coordinate. In such cases we
1165: have\footnote{Again, $E$ is chosen to be the energy associated
1166: to the point of the orbit closer to the origin.}
1167: \[
1168: \begin{split}
1169: \lambda_{u,n}(\xi,u)&=\prod_{j=1}^n(1+h'(x_{-j})+u_{-j})\geq
1170: \prod_{j=1}^n(1+\Upsilon_E'(x_{-j})-\cs_0\Upsilon_E(x_{-j}))\\
1171: &\geq e^{\sum_{j=1}^n\Upsilon_E'(x_{-j})-2\cs_0\Upsilon_E(x_{-j})}\\
1172: &\geq e^{-\sum_{j=1}^n\frac{\Upsilon_E'(x_{-j})}{\Upsilon_E(x_{-j})}
1173: (x_{-j-1}-x_{-j}) -3\cs_0\sum_{j=1}^n(x_{-j-1}-x_{-j})} \\
1174: &\geq e^{-\cs|x_0|}e^{-\int_{x_0}^{x_{-n}}\frac{\Upsilon_E'(z)}{\Upsilon_E(z)}dz}
1175: = e^{-\cs|x_0|}\frac{\Upsilon_E(x_0)}{\Upsilon_E(x_{-n})}.
1176: \end{split}
1177: \]
1178:
1179: Let $n_*\in\N$ be the last integer for which $|G(x_{-n})|\geq
1180: E$, then for $n\leq n_*$ we have
1181: \[
1182: \lambda_{u,n}(\xi)\geq e^{-\cs|x_0|}\sqrt{\frac
1183: {E+G(x_0)}{E+G(x_{-n})}}
1184: \geq e^{-\cs|x_0|}\sqrt{\frac
1185: {G(x_{-n})+G(x_0)}{2G(x_{-n})}}.
1186: \]
1187: On the other hand comparing the backward motion with the
1188: backward motion on the stable manifold, as we did before with the
1189: unstable,\footnote{Here we use the inequality
1190: \[
1191: \sqrt{\frac{1+(1+a)^4}2}\geq (1+\frac a2)^2.
1192: \]
1193: }
1194: \[
1195: \lambda_{u,n}(\xi,u)\geq
1196: e^{-\cs|x_0|}\sqrt{\frac{1+(n|x_0|\cs+1)^4}2}
1197: \geq e^{-\cs|x_0|}(1+n|x_0|\cs)^2.
1198: \]
1199: Next, let us consider $n\in\{n_*,\dots, m\}$, where $m$ is the larger
1200: integer such that $x_{-m}\leq 0$, we have $2\sqrt E\geq
1201: \Upsilon_E(x_{-n})\geq \sqrt {2E}$.
1202: \[
1203: \begin{split}
1204: \lambda_{u,n-n_*}(\xi,u)
1205: &\geq e^{-\cs_2|x_{n_*}|}
1206: e^{-\int_{x_{-n_*}}^{x_{-n}}\frac{\Upsilon_E'(z)}{\Upsilon_E(z)}dz}
1207: \geq e^{-\cs|x_0|}\frac{\Upsilon_E(x_{-n_*})}
1208: {\Upsilon_E(x_{-n})}\\
1209: &\geq e^{-\cs|x_0|}\sqrt{\frac 32}\geq e^{-\cs|x_0|}(1+\cs
1210: |x_{-n_*}|(n-n_*))^2,
1211: \end{split}
1212: \]
1213: where, in the last line, we used Lemma \ref{lem:zdyn2}-(5).
1214: By symmetry it will be enough to wait another time $m$ to have
1215: $|x_{-2m}|\geq \frac 12|x_0|$, after which the expansion is assured by
1216: the estimate \eqref{eq:trivialexp}.
1217: \end{proof}
1218:
1219: Next we need to have similar estimates for the stable contraction. By
1220: \eqref{eq:constex}
1221: \begin{equation}\label{eq:stabdef}
1222: \begin{split}
1223: \mu_s(v)&:=1+v=\mu_{s,1}(\xi,v)\\
1224: F^-_1(\xi,v)&=h'(x-y)+\frac v{1+v}.
1225: \end{split}
1226: \end{equation}
1227: It is immediate to check that
1228: $D_{(x,y)}T^{-1}(1,-v)=\mu_s(v)(1,-F^-_1((x,y),v))$ and
1229: $\mu_{s,n}(\xi,v_{0}):=\prod_{i=0}^{n-1}\mu_s(v_{-i})$, where
1230: $v_0=v$ and $v_{-i-1}:=F^-_1(T^{-i}\xi,v_{-i})$.
1231:
1232: An interesting way to transform information on expansion into information on
1233: contraction is to use area preserving.
1234: \begin{lem}\label{lem:areap}
1235: Let $\xi\in\To^2$, then for each $n\in\N$, $u,v\geq 0$ let $u_{-n}=u$,
1236: $v_0=v$, $\xi_{-n}=T^{-n}\xi$ and $u_{-k+1}=F(\xi_{-k+1}, u_{-k})$,
1237: $v_{-k-1}=F^-(\xi_{-k},v_{-k})$. Then
1238: \[
1239: \mu_{s,n}(\xi, v_0)(v_{-n}+u_{-n})=\lambda_{u,n}(\xi, u_{-n})(u_0+v_0).
1240: \]
1241: \end{lem}
1242: \begin{proof}
1243: Calling $\omega$ the standard symplectic form we have
1244: \[
1245: \begin{split}
1246: \mu_{s,n}(\xi,v_0)(v_{-n}+u_{-n})&=\omega(D_{\xi}T^{-n}(1,-v_0),(1,u_{-n}))\\
1247: &= \omega((1,-v_0), D_{\xi_{-n}} T^n(1,u_{-n}))=\lambda_{u,n}(\xi,
1248: u_{-n})(u_0+v_0).
1249: \end{split}
1250: \]
1251: \end{proof}
1252: %\begin{changebar}
1253: The following is an immediate corollary of Lemmata \ref{lem:areap} and
1254: \ref{lem:expansion}.
1255: \begin{cor}\label{lem:contraction}
1256: For each $\xi=(x,y)\in \To^2$ and $n\in\N$ holds
1257: \[
1258: \mu_{s,n}(\xi, v_0)\geq e^{-\cs|x_{0}|} (\cs^{-1}|x_{0}|n+1)^2\frac
1259: {u_0+v_0}{u_{-n}+v_{-n}} \quad\forall n\in\N.
1260: \]
1261: \end{cor}
1262: %\end{changebar}
1263:
1264: All the other expansion estimates can be obtained by reversibility.
1265:
1266: \section{Distributions--regularity}
1267: \label{sec:regularity}
1268:
1269: Let $(1,u(\xi)),\,(1,-v(\xi))$ be the unstable and stable directions,
1270: respectively. We will then use the short hand
1271: $\lambda_{u,n}(\xi):=\lambda_{u,n}(\xi,u(\xi_{-n}))$ and
1272: $\mu_{s,n}(\xi):=\mu_{s,n}(\xi,v(\xi_{n}))$.
1273: \begin{lem}\label{lem:distreg0}
1274: The unstable distribution is continuous in $\To^2$.
1275: \end{lem}
1276: \begin{proof}
1277: Notice that, for $\xi=(x,y)$, $\xi_n:=T^n\xi$, iterating formula
1278: \eqref{eq:unsdef}, in analogy with \eqref{eq:unstreg}, holds true
1279: \begin{equation}\label{eq:unstreg-bis}
1280: \begin{split}
1281: u(x)-u(z)=&\lambda_{u,n}(x)^{-1}\lambda_{u,n}(z)^{-1}(u(x_{-n})-u(z_{-n}))\\
1282: &\ \ +\sum_{k=1}^{n}\lambda_{u,k}(x)^{-1}\lambda_{u,k}(z)^{-1}
1283: (h'(x_{-k})-h'(z_{-k}))
1284: \end{split}
1285: \end{equation}
1286: By Lemma \ref{lem:expansion}, we can take the limit $n\to\infty$ in
1287: the above formula provided $x\neq 0$, and obtain a
1288: uniformly convergent series from which the continuity follows.
1289: If $\xi\neq 0$ then $x_{-1}\neq 0$ and \eqref{eq:unsdef} implies
1290: \begin{equation}
1291: \label{eq:uma}
1292: u(\xi)=\lambda_{u}(\xi_{-1})^{-1}h'(x_{-1})
1293: +\lambda_{u}(\xi_{-1})^{-1}u(\xi_{-1}) ,
1294: \end{equation}
1295: hence the continuity at $\xi\neq 0$ follows. We are left with the
1296: continuity at the origin, but this is already implied by Proposition
1297: \ref{prop:dista}.
1298: \end{proof}
1299:
1300: This means that we can extend the invariant unstable distribution
1301: (that, up to now, where defined--by Pesin theory--only almost everywhere) to a
1302: continuous everywhere defined vector field. The same statement holds
1303: for the stable vectors by reversibility.
1304:
1305: Given a continuous vector field there exists integral curves. Since
1306: we do not know yet if the vector fields are Lipschitz, it does not
1307: follows automatically that from a given point there exits only one
1308: integral curve, yet this follows by standard dynamical
1309: arguments. Clearly such integral curves are nothing else than the
1310: stable and unstable manifolds that are therefore everywhere
1311: defined. In addition, remember that, by general hyperbolic theory, the
1312: foliations are absolutely continuous, it follows that the above
1313: everywhere defined foliations are continuous. Unfortunately, for the
1314: following much sharper regularity information is needed, this is
1315: obtained in the rest of the section.
1316:
1317: Let us call $\partial^u, \partial^s$ the derivative along the unstable
1318: and the stable vector fields, respectively.
1319:
1320: \begin{lem}\label{lem:uderbound}
1321: The vector field $u$ is $\Co^1$ along the unstable manifolds, apart from the
1322: origin, moreover
1323: \[
1324: |\partial^u
1325: u(\xi)|\leq C\quad
1326: \forall \xi\neq 0.
1327: \]
1328: \end{lem}
1329: \begin{proof}
1330: If $\xi$ is outside of a
1331: neighborhood of the origin of size $\delta$, then by Lemma
1332: \ref{lem:expansion}, \eqref{eq:unstreg-bis} we have, in analogy with
1333: the arguments leading to \eqref{eq:unstder},
1334: \begin{equation}
1335: \label{eq:uder-u-bond}
1336: |\partial^u
1337: u(\xi)|=\left|\sum_{k=1}^{\infty}\lambda_{u,k}(x)^{-3}h''(x_k)\right| \leq \cs
1338: \sum_{n=0}^{\infty}(\delta n+1)^{-6}\leq \cs.
1339: \end{equation}
1340: Since the series converges uniformly the $\Co^1$ property follows. To
1341: obtain a uniform bound more work is needed.
1342: If $|\xi|<\delta$, formula \eqref{eq:uma} implies
1343: \[
1344: |\partial^u u(\xi)|\leq\lambda_u(\xi_{-1})^{-3}|h''(x_{-1})|+
1345: \lambda_u(\xi_{-1})^{-3} |\partial^u u(\xi_{-1})|=:\Psi(\xi_{-1},
1346: |\partial^u u(\xi_{-1})|).
1347: \]
1348: A simple computation, remembering Proposition \ref{prop:dista}, shows
1349: that
1350: \[
1351: \Psi(\xi,\varrho)\leq7b|x|+\frac{\varrho}{1+3|u(\xi)|}\leq
1352: 7b|x|+\frac{\varrho}{1+3K_-|x|} \leq \varrho ,
1353: \]
1354: provided $\varrho\geq \frac{7b(1+3K_-)}{3K_-}$. Accordingly, for $\rho$
1355: large enough, we have
1356: $|\partial^u u(\xi)|\leq \rho$, for all $\xi$.
1357: \end{proof}
1358:
1359: It remains to investigate the regularity of the unstable distribution
1360: along the stable direction.
1361:
1362: \begin{lem}\label{lem:distreg1}
1363: The unstable distributions are $\Co^1$ along stable manifolds, apart
1364: from the origin. Moreover
1365: \[
1366: |\partial^s u(\xi)|\leq \cs \quad \forall \xi\neq 0.
1367: \]
1368: \end{lem}
1369: \begin{proof}
1370: Let us fix some arbitrary neighborhood of the origin.
1371: Let $x,z\in W^s$ outside such a neighborhood. Let $W^s_0$ be the
1372: piece of stable manifold between such two points. Clearly
1373: $W^s_{n}:=T^nW^s_0$ grows for negative $n$. Let $n(x,z)$ be the largest
1374: integer for which $|W^s_{-n}|\leq |W^s_0|^{\frac 14}$. Our first
1375: result is a distortion bound.
1376: \begin{sublem}\label{slem:dist}
1377: For each $n\leq n(x,z)$ and $\xi\in
1378: W^s_{0}$, holds
1379: \[
1380: \cs^{-1}\frac{|W^s_{-n}|}{|W^s_{0}|}\leq \mu_{s,n}(\xi)\leq
1381: \cs\frac{|W^s_{-n}|}{|W^s_{0}|} .
1382: \]
1383: \end{sublem}
1384: \begin{proof}
1385: If the backward orbit spends at least half of the time outside the
1386: neighborhood, then $W^s_{-n}$
1387: grows exponentially fast, hence $n(x,z)\leq \cs\ln|W^s_0|^{-1}$ and
1388: $\sum_{i=0}^{n(x,z)}|W^s_{-i}|\leq \cs$. If this is not the case,
1389: the worst possible situation is when $W^s_{-m}$ is the closest
1390: to the origin and all the trajectory lies in the neighborhood. In such a
1391: case, letting $m:=n(x,z)$,
1392: \[
1393: |W^s_{m}|=\int_{W^s_0}\mu_{s,m}(z)dz\geq
1394: \int_{W^s_0}\frac{\theta(z)(\cs^{-1}|z|m+1)^2}{2\theta(T^{-m}z)} dz,
1395: \]
1396: where $\theta(\zeta)=u(\zeta)+v(\zeta)$ is the separation between the
1397: stable and the unstable directions at the point $\zeta$ and we have
1398: used Lemma \ref{lem:contraction}. Now Proposition \ref{prop:dista} and
1399: Lemma \ref{lem:zdyn2}-(1) imply $\theta(T^{-m}z)\leq \cs m^{-1}$
1400: outside the parabolic sector, while Lemma \ref{lem:zdyn2}-(3,4,5) show
1401: that the same estimates remain in $P_M$ as well. Accordingly,
1402: \[
1403: |W^s_{-m}|\geq \cs m^3|W^s_0|.
1404: \]
1405: That is $m\leq \cs |W^s_0|^{-\frac 14}$, and
1406: \[
1407: \sum_{i=0}^{n(x,z)}|W^s_{-i}|\leq n(x,z)|W^s_0|^{\frac 14}\leq \cs.
1408: \]
1409: The above estimate readily implies that, for each $\xi,\eta\in W^0_s$,
1410: \[
1411: e^{-\cs |W^s_{-i}|}
1412: \leq\frac{\mu_s(\xi_{-i},v(\xi_{-i}))}{\mu_s(\eta_{-i},v(\eta_{-i}))}
1413: \leq e^{\cs |W^s_{-i}|},
1414: \]
1415: where we have used Lemma \ref{lem:uderbound} for the stable manifold.
1416: Accordingly,
1417: \[
1418: e^{-\cs\sum_{i=0}^m |W^s_{-i}|}
1419: \leq\frac{\mu_{s,n}(\xi)}{\mu_{s,n}(\eta)}
1420: \leq e^{\cs \sum_{i=0}^m|W^s_{-i}|},
1421: \]
1422: from which the Lemma readily follows.
1423: \end{proof}
1424: By Lemma \ref{lem:areap} it follows, letting again $m:=n(x,z)$,
1425: \[
1426: \begin{split}
1427: \lambda_{u,m}(x)^{-1}\lambda_{u,m}(z)^{-1}
1428: &=\lambda_{u,m}(x)^{-1}\sqrt{\lambda_{u,m}(z)^{-2}}\\
1429: &\leq
1430: \cs\left(\theta(x_{-m})\mu_{s,m}(x)
1431: \sqrt{\lambda_{u,m}(z)\theta(z_{-m})\mu_{s,m}(z)}\right)^{-1}
1432: \end{split}
1433: \]
1434: As before the worst case is clearly when $W^s_{-m}$ is the closest to
1435: the origin.
1436: In such a case, consider that at least one of the two end points of
1437: $W^s_{-n(x,z)}$ must be at a distance $\cs |W^s_{-n(x,z)}|$ from the
1438: fixed point, let us say $T^{-n(x,z)}z$, hence
1439: $\theta(T^{-n(x,z)}z)\geq \cs|W^s_{-n(x,z)}|$. In addition,
1440: $\theta(x_{-m})\geq \cs m^{-1}$. Indeed, this follows from Lemma
1441: \ref{lem:zdyn2}-(3,4,5) if the trajectory ends in $P_M$. If the
1442: trajectory lies outside $P_M$ then it approaches the origin slower than
1443: the dynamics $x-\cs x^2$, which implies $x_{-m}\geq \cs
1444: m^{-1}$. Furthermore by using the above facts, Lemma \ref{lem:expansion}, Sub-lemma
1445: \ref{slem:dist} and the definition of the stopping time $m$ yields
1446: \[
1447: \lambda_{u,m}(x)^{-1}\lambda_{u,m}(z)^{-1}\leq \cs|x-z|^{\frac
1448: 34}m\frac 1{m} \sqrt{\frac{|x-z|^{\frac
1449: 34}}{|x-z|^{\frac 14}}} \leq \cs|x-z|.
1450: \]
1451: Since we know that $u$ is a uniformly continuous function it follows
1452: \[
1453: \lim_{z\to
1454: x}\lambda_{u,m}(x)^{-1}\lambda_{u,m}(z)^{-1}
1455: \frac{|u(T^{-m}x)-u(T^{-m}z)|}
1456: {|x-z|}=0.
1457: \]
1458: Accordingly, by formula \eqref{eq:unstreg},
1459: \[
1460: \begin{split}
1461: u'(x)=&\lim_{z\to x}
1462: \sum_{n=0}^{m}\lambda_{u,n}(x)^{-1}\lambda_{u,n}(z)^{-1}
1463: \frac{h'(T^{-n}x)-h'(T^{-n}z)}{x-z}\\
1464: =& \lim_{z\to x}
1465: \sum_{n=0}^{m}\lambda_{u,n}(x)^{-1}\lambda_{u,n}(z)^{-1}
1466: h''(T^{-n}\zeta_{n}) \mu_{s,n+1}(\zeta_n),
1467: \end{split}
1468: \]
1469: for some $\zeta_n\in W_0$.
1470: But $|h''(T^{-n}\zeta_n)|\leq \cs\theta(T^{-n}z)$, hence
1471: \[
1472: \begin{split}
1473: &\lambda_{u,n}(x)^{-1}\lambda_{u,n}(z)^{-1}\theta(T^{-n}x)
1474: \mu_{s,n+1}(x)\leq \cs\lambda_{u,n}(z)^{-1}\\
1475: &\lambda_{u,n}(x)^{-1}\lambda_{u,n}(z)^{-1}\theta(T^{-n}z)
1476: \mu_{s,n+1}(z)\leq \cs\lambda_{u,n}(x)^{-1}.
1477: \end{split}
1478: \]
1479: Remembering Sub-Lemma \ref{slem:dist} the uniform convergence of the
1480: series follows and yields the formula
1481: \begin{equation}
1482: \label{eq:uder-s-bound}
1483: \partial^s u(x)=\sum_{n=0}^{\infty}\lambda_{u,n}(x)^{-2}\mu_{s,n+1}(x)
1484: h''(T^{-n}x).
1485: \end{equation}
1486: Given the arbitrariness of the neighborhood of zero, the above formula
1487: holds for each $x\neq 0$ and, since the series converges uniformly,
1488: the $\Co^1$ property follows. We can now conclude the Lemma.
1489: By Lemma \ref{lem:areap} follows
1490: \[
1491: \begin{split}
1492: |\partial^s u(x)|&\leq\cs \sum_{n=0}^\infty
1493: \lambda_{u,n}(x)^{-2}\mu_{s,n}(x)\theta(T^{-n}x) \\
1494: &\leq \cs \sum_{n=0}^\infty\lambda_{u,n}(x)^{-1}\theta (x)
1495: \leq \cs \sum_{n=0}^\infty\frac{|x|}{(|x|n+1)^2}\leq \cs .
1496: \end{split}
1497: \]
1498: \end{proof}
1499: \begin{rem}
1500: Notice that the symmetrical statements follow by reversibility.
1501: \end{rem}
1502: The final result on the regularity of the foliations can be stated as
1503: follows.
1504: \begin{lem}
1505: \label{lem:c1-reg}
1506: The stable and unstable vector fields are $\Co^1(\To^2\setminus\{0\})$ and, more
1507: precisely, for each $\xi\in\To^2\setminus\{0\}$,
1508: \[
1509: |Du(\xi)|\leq \cs \theta(\xi)^{-1}.
1510: \]
1511: \end{lem}
1512: \begin{proof}
1513: The $\Co^1$ property follows from Lemma 19.1.10 of \cite{KH}.
1514: Then the size of the derivative can be easily estimated by the size of
1515: the partial derivatives in the stable and unstable directions divided by
1516: the angle between them.
1517: \end{proof}
1518: \begin{rem}
1519: In fact, it is likely that with a little more work one can show that
1520: the foliations are $\Co^{\frac 32-\ve}$, but we do not investigate this
1521: possibility since it is not needed in the following.
1522: \end{rem}
1523: \section{Holonomy}
1524: \label{sec:holo}
1525: There exists $\cs_1>0$ such that, given two close by stable manifolds
1526: $W^s_1$, $W^s_2$ we can define the {\em unstable holonomy}
1527: $\Psi^u:W^s_1\to W^s_2$ by $\{\Psi^u(\xi)\}:=W^u(\xi)\cap W^s_2$.
1528: Let $D_r:=\{\zeta=(z_1,z_2)\in\R^2\;:\;|z_1|\leq r;|z_2|\leq r^2\}$.
1529: \begin{lem}\label{lem:holo}
1530: For each $W^s_1,W^s_2$ disjoint from $D_r$ and $\xi\in W^s_1$, holds
1531: \[
1532: |J\Psi^u(\xi)-1|\leq \cs r^{-1}\|\Psi^u(\xi)-\xi\|.
1533: \]
1534: Provided $\|\Psi^u(\xi)-\xi\|\leq \cs_1 r$.
1535: \end{lem}
1536: \begin{proof}
1537: Let $\gamma_s,\tilde\gamma_s:[-\delta,\delta]\to\R^2$ be $W^s_1,
1538: W^s_2$, respectively, parametrized by arc-length. Also, let
1539: $\Gamma:[-\delta,\delta]^2\to\R^2$, be such that $\Gamma(0,0)=\xi$,
1540: $\Gamma(s,0)=\gamma_s(s)$
1541: and $\Gamma(s,t)$ be the unstable manifold, parametrized by
1542: arc-length, of $\Gamma(s,0)$ and, finally,
1543: $\Gamma(0,\rho):=\Psi^u(\xi)$. Note that $\Gamma(s,t)$ can be obtained
1544: integrating the unstable vector field starting from $\Gamma(s,0)$,
1545: hence Lemma \ref{lem:c1-reg} and the standard results on the continuity
1546: with respect to the initial data imply $\Gamma\in\Co^1$. By the
1547: transversality of the stable and unstable manifolds there exist
1548: $\tau,\sigma:[-\delta,\delta]\to\R$ such that
1549: $\Gamma(s,\tau(s))=\Psi^u(\gamma_s(s))=\tilde\gamma(\sigma(s))\in
1550: W^s_2$. Calling $\eta(s)$ the unit vector perpendicular to $\tilde
1551: \gamma'(s)$, by the implicit function theorem, it follows
1552: \begin{equation}
1553: \label{eq:implicit}
1554: \begin{split}
1555: \tau'(s)&=-\frac{\langle \eta(s),\partial^s\Gamma(s,\tau(s))\rangle}
1556: {\langle\eta(s), \partial_t\Gamma(s,\tau(s))\rangle}\\
1557: \sigma'(s)&=\langle\tilde\gamma_s'(s),\partial^s\Gamma(s,\tau(s))\rangle
1558: -\frac{\langle\tilde
1559: \gamma'_s(s),\partial_t\Gamma(s,\tau(s))\rangle\;\langle
1560: \eta(s),\partial^s\Gamma(s,\tau(s))\rangle}
1561: {\langle\eta(s), \partial_t\Gamma(s,\tau(s))\rangle}
1562: \end{split}
1563: \end{equation}
1564: where, clearly, $\sigma'(s)=J\Psi^u(\gamma_s(s))$. Calling
1565: $v^u(\eta)$, $\eta\in\To^2$,
1566: the unit vector in the unstable direction at $\eta$ and $v^s(\eta)$ the
1567: stable one, one has
1568: $\partial_t\Gamma(s,\tau(s))=v^u(\Gamma(s,\tau(s)))$. On the other
1569: hand, setting $V(s,t):=\partial^s\Gamma(s,t)-v^s(\Gamma(s,t))$, holds
1570: $V(s,t)=0$ for $t=0$, but for $t\neq 0$, in general, it will be
1571: $V(s,t)\neq 0$. Yet, it is possible to estimate it by differentiating
1572: $
1573: \Gamma(s,t)=\Gamma(s,0)+\int_0^tv^u(\Gamma(s,t'))dt'
1574: $
1575: which yields
1576: \[
1577: \partial^s\Gamma(s,t)=v^s(\Gamma(s,0))+\int_0^t Dv^u(\Gamma(s,t'))
1578: \partial^s\Gamma(s,t')dt'.
1579: \]
1580: Lemmata \ref{lem:c1-reg} and \ref{lem:distreg1}
1581: imply that $\|Dv^u\|\leq \cs r^{-1}$ and $\|Dv^u v^s\|=|\partial^s
1582: v^u|\leq \cs$, hence
1583: \[
1584: \|V(s,t)\|\leq \cs r^{-1}\int_0^t\|V(s,t')\|dt'+\cs t.
1585: \]
1586: By Gronwal, it follows, provided $t\leq \cs \rho$ and $\rho\leq \cs_1
1587: r$, for $\cs_1$ small enough,
1588: \begin{equation}
1589: \label{eq:Vest}
1590: \|V(s,t)\|\leq \cs t.
1591: \end{equation}
1592: Accordingly, by the second of \eqref{eq:implicit} and \eqref{eq:Vest},
1593: it follows
1594: \[
1595: |\sigma'(0)-1|\leq \cs r^{-1}\rho.
1596: \]
1597: \vskip-.5cm
1598: \end{proof}
1599:
1600: \section{Random perturbations}
1601: \label{sec:random}
1602:
1603: The density of a measure with respect to Lebesgue evolves as
1604: \[
1605: \Lp f:=f\circ T^{-1}.
1606: \]
1607: We will then construct a random perturbation by introducing the
1608: convolution operator
1609: \begin{equation}\label{eq:conv}
1610: Q_\ve f(x):=\int_{\To^2}q_\ve(x-y)f(y)dy.
1611: \end{equation}
1612: Where we assume
1613: \begin{enumerate}
1614: \item $q_\ve(\xi):=\ve^{-2}\bar q(\ve^{-1}\xi)$; $\bar q\in\Co^{\infty}(\R^2,\R^+)$;
1615: \item $\int_{\R^2}\bar q(\xi)d\xi=1$;
1616: \item $\bar q(\xi)=0$ for each $\|\xi\|\geq 1$;
1617: \item $\bar q(\xi)=1$ for each $\|\xi\|\leq \frac 12$.
1618: \end{enumerate}
1619: We define then
1620: \begin{equation}\label{eq:random}
1621: \Lp_\ve:=Q_\ve\Lp^{n_\ve},
1622: \end{equation}
1623: where $n_\ve$ will be chosen later.
1624:
1625: Notice that
1626: \begin{equation}\label{eq:ker1}
1627: \Lp_\ve^2f(x)=\int_{\To^4} q_\ve(x-y)q_\ve(T^{-n_\ve}y-T^{n_\ve}z)f(z)
1628: dz dy :=\int_{\To^2}\Ka_\ve(x,z)f(z)dz.
1629: \end{equation}
1630:
1631: We have thus a kernel operator that can be investigated with rather
1632: coarse techniques. It turns out to be convenient to define the
1633: associated kernel
1634: \begin{equation}\label{eq:kernel2}
1635: \bar\Ka_\ve(x,z):=\Ka_\ve(x,T^{-n_\ve}z)=\int_{\To^2}q_{\ve}(x-y)
1636: q_\ve(T^{-n_\ve}y-z)m(dy).
1637: \end{equation}
1638: For further use let us define
1639: \begin{equation}\label{eq:levelset}
1640: \begin{split}
1641: &D_r:=\{z=(z_1,z_2)\in\To^2\;|\;|z_1|\leq r;\; |z_2|\leq r^2\}\\
1642: &B_r(\xi):=\{\eta\in\To^2\;;\;\|\xi-\eta\|<r\}.
1643: \end{split}
1644: \end{equation}
1645: The following is a relevant fact used extensively in the sequel.
1646: \begin{lem}
1647: \label{lem:levelset-escape}
1648: There exists $\cs_3,R>0$ such that, for each $\delta<R^2$, if
1649: $B_\delta(\xi)\subset D_{R/2}$, then there exists $\eta\in
1650: B_{\frac 34\delta}(\xi)$ such that $T^n B_{\frac \delta 4} (\eta)\cap
1651: D_{R}=\emptyset$ for some $n\leq \cs_3 \delta^{-\frac 12}$.
1652: \end{lem}
1653: \begin{proof}
1654: %\begin{changebar}
1655: If $\xi$ belongs to the first or third
1656: quadrant, then $T^n\xi$ is escaping from the origin.
1657: In such a case, if $B_{\frac 58\delta}(\xi)$ belongs to the thin sector we choose
1658: $\eta\in B_{\frac{\delta}2} (\xi)$, $|y|\geq \frac 38 \delta $.
1659: Clearly, if $(x,y)\in B_{\frac \delta 4}(\eta)$, $\cs x^2\geq |y|\geq
1660: \frac \delta 4$. On the other hand
1661: $|y_n|\geq n|x|^3$ while $|x_n|\leq |x|+n\cs|x_n|^2$. So, if $n\leq
1662: \cs |x|^{-1}\leq \cs\delta^{-\frac 12}$, we have $|y_n|\geq
1663: \cs|x_n|^2$. After that we can compare the dynamics with one of the
1664: type $x\mapsto x+\cs x^2$, hence after a time at most $\cs
1665: \delta^{-\frac 12}$ the $T^n B_{\frac \delta 4}(\eta)$ will exit $D_R$.
1666: If the above does not apply, then
1667: one can take a ball of radius $\delta/4$ belonging completely to the
1668: fat sector and centered at a point in $B_\delta(\xi)$. Then the results
1669: of subsection \ref{sub:nearfix} easily implies the lemma.
1670: %\end{changebar}
1671: If, on the contrary, $\xi$ belong to the second or fourth quadrant, then its
1672: trajectory may approach the origin in an arbitrary manner (even
1673: asymptotically, if the point belongs to the stable manifold). In such
1674: a case we can take a point $\eta\in B_{\delta}(\xi)$ at, at least, a vertical
1675: distance $\frac 34\delta$ from the stable manifold and such that
1676: $B_{\frac 34\delta}(\eta) \subset B_\delta(\xi)$. Again from the
1677: results of subsection \ref{sub:nearfix} it follows that such a ball
1678: will exit $D_R$ in a time at most $\cs_3\delta^{-\frac12}$.
1679: \end{proof}
1680:
1681:
1682: \begin{lem}\label{lem:estim}
1683: There exit constants $\sigma,\cs_3>0$ such that if $n_\ve=\cs_3\ve^{-\frac 12}$ holds
1684: \[
1685: \bar\Ka_\ve(x,z)\geq \sigma\quad \forall x,z\in\To^2.
1686: \]
1687: \end{lem}
1688: \begin{proof}
1689: It is trivial to see that
1690: \[
1691: \bar\Ka_\ve(x,z)\geq \ve^{-4}m(B_{\ve/2}(x)\cap T^{n_\ve}B_{\ve/2}(z)).
1692: \]
1693: Accordingly, by Lemma \ref{lem:levelset-escape} there exists two
1694: balls, of radius $\frac \ve 8$, $\bar B_1\subset B_{\ve/2}(z)$ and
1695: $\bar B_2\subset B_{\ve/2}(x)$ that are outside of $D_{r}$, $r\geq
1696: \sqrt{\frac \ve 2}$, and whose images will be outside
1697: of a neighborhood of the origin or order one in a time less than $\cs
1698: \ve^{-\frac 12}$, forward and backward in time, respectively. Given two
1699: unstable manifolds in $B_1$ at a distance larger than $\bar c r\ve$, for
1700: some appropriate $\bar c$, then no stable
1701: manifold will intersect both manifolds inside the ball $B_1$. We
1702: can thus consider $\cs r^{-1}$ unstable manifolds such that no stable
1703: manifolds intersect two of them in $B_1$. Around each such
1704: manifold we can construct a strip by moving along the stable manifold
1705: by $\cs \ve$. We obtain in this way $\cs r^{-1}$ disjoint strips
1706: each of area $\cs r\ve^2$, whose union covers a fixed fraction of the area of
1707: $B_1$. After a time less that $\cs \ve^{-\frac 12}$ such strips will be outside a
1708: neighborhood of zero, their length may have increase considerable, if so
1709: we will subdivide them into strips of length $\ve$. Since now the
1710: stable and unstable manifold are at a fixed angle and by the usual
1711: distortion arguments, such strips are essentially rectangular. At this point,
1712: by Lemma \ref{lem:expansion}, it will suffice to wait a time
1713: $\ve^{-\frac 12}$ to insure that each such strip will acquire length
1714: at least $\frac 12$ in the unstable direction. We thus iterate for such
1715: a time and, if one strip becomes longer than one, we subdivide it into
1716: pieces of length between $\frac 12$ and one. Finally, fix some box
1717: $\Lambda$ of some fixed size $C$ away from the origin with sides
1718: approximately parallel either to the stable or to the unstable
1719: directions.
1720: By mixing it suffices to wait a fixed time to be sure that a fixed
1721: percentage of each one of the above mentioned strips will intersect
1722: the box. In addition, it is possible to insure that such strips cut
1723: the box from one stable side to the other.
1724:
1725: We can then write $m(B_{\ve/2}(x)\cap
1726: T^{n_\ve}B_{\ve/2}(z))=m(T^{-n_\ve/2}B_{\ve/2}(x)\cap
1727: T^{n_\ve/2}B_{\ve/2}(z))$ since the same considerations done above for
1728: the unstable manifold can be done, iterating
1729: backward, for the stable manifold it follows that a fixed percentage of
1730: $T^{-n_\ve/2}B_{\ve/2}(x)$ and a fixed percentage of
1731: $T^{n_\ve/2}B_{\ve/2}(z))$ will intersect $\Lambda$ and hence each
1732: other. In fact each one of the above constructed strips in the unstable
1733: direction will intersects each one of the strips in the stable
1734: direction. By the usual distortion estimates, this implies that the
1735: intersection among any two such strip has a measure proportional to the
1736: product of the measure of the two strips, hence
1737: \[
1738: m(B_{\ve/2}(x)\cap T^{n_\ve}B_{\ve/2}(z))\geq \cs
1739: m(B_{\ve/2}(x))m(B_{\ve/2}(z)),
1740: \]
1741: and the lemma.
1742: \end{proof}
1743:
1744: \begin{lem}\label{lem:l1}
1745: For each $f\in L^1$, $\int f=0$ holds
1746: \[
1747: \|\Lp_\ve^n f\|_1\leq (1-\sigma)^{n/2}\|f\|_1.
1748: \]
1749: \end{lem}
1750: \begin{proof}
1751: Note that $\Lp_\ve 1=
1752: \Lp_\ve^*1=1$ and let $\M^+_\ve=\{x\in\To^2\;|\;\Lp_\ve^2 f\geq 0\}$;
1753: $\M_+=\{x\in\To^2\;|\; f\geq 0\}$,
1754: then, since $\int \Lp_\ve f=\int f=0$,
1755: \[
1756: \begin{split}
1757: \|\Lp_\ve^2 f\|_1&=2\int_{\M^+_\ve}dx\,\Lp_\ve^2 f
1758: =2\int_{\M^+_\ve}dx\int_{\To^2}dy\, \Ka_\ve(x,\,y)
1759: f(y)\\
1760: &=2\int_{\To^2}dy\, f(y)\int_{\M^+_\ve}dx[ \Ka_\ve(x,\,y)-\sigma]\\
1761: &\leq 2\int_{\M_+}dy\, f(y)\int_{\To^2}dx [\Ka_\ve(x,\,y)-\sigma]
1762: =2(1-\sigma)\int_{\M_+}dy\, f(y)\\
1763: &=(1-\sigma)\|f\|_1.
1764: \end{split}
1765: \]
1766: \end{proof}
1767: Let $v^{u,s}=(v^{u,s}_1,v^{u,s}_2)$ be the unit tangent vector fields in the
1768: unstable and stable direction, respectively. Clearly
1769: $|\partial^u(\Lp^i f)|\leq |\partial^u f|$, while
1770: $|\partial^s(g\circ T^i)\leq |\partial^s g|$.
1771: \begin{lem}\label{lem:appr}
1772: For each $f,g\in\Co^{1}(\To^2,\R)$ holds
1773: \[
1774: \left|\int Q_\ve f g-\int f g\right|\leq \cs
1775: \ve\{\|f\|_\infty+\|\partial^u f\|_{L^1(\nu)}\}
1776: \{\|g\|_\infty+\|\partial^s g\|_{L^1(\nu)}\},
1777: \]
1778: where $\nu$ is the measure defined by $\nu(h):=\int
1779: d\rho \int_{\partial D_{\rho}}h$.
1780: \end{lem}
1781: \begin{proof}
1782: It is convenient to introduce the following change of
1783: variables. For each $x,y\in\To^2$ close enough, let us call
1784: $[x,y]=W^u_\delta(x)\cap W^s_\delta(y)$, note that by Lemma
1785: \ref{prop:dista} such a point is
1786: always well defined provided $d(x,y)\leq \cs \delta^2$. We consider then
1787: the change of variable $\Phi:\To^2\times \To^2\to\R^2\times\To^2$,
1788: \begin{equation}\label{eq:changv}
1789: \begin{split}
1790: \xi:&=x-y\\
1791: \eta:&=[x,y].
1792: \end{split}
1793: \end{equation}
1794: Due to the absolute continuity of the holonomies the above change of
1795: variable is absolutely continuous. Clearly, $\partial^u_x \eta=0$ since
1796: moving along $W^u(x)$
1797: does not change the intersection point with $W^s(y)$. On the other
1798: hand $\partial^s_x \eta=J\Psi^u_x v^s(\eta)$, since moving $x$ along
1799: $W^s(x)$ moves $\eta$ on $W^s(y)$ by an amount determined exactly by the unstable
1800: holonomy $\Psi^u$ between $W^s(x)$ and $W^s(y)$. By similar arguments
1801: and a straightforward computations
1802: \[
1803: \begin{split}
1804: \partial_{x_1}\eta&=v_2^u(x)\det\begin{pmatrix} v^s(x)
1805: &v^u(x)\end{pmatrix}^{-1}J\Psi^u_x v^s(\eta)\\
1806: \partial_{x_2}\eta&=v_1^u(x)\det\begin{pmatrix} v^u(x)
1807: &v^s(x)\end{pmatrix}^{-1}J\Psi^u_x v^s(\eta)\\
1808: \partial_{y_1}\eta&=v_2^s(y)\det\begin{pmatrix} v^u(y)
1809: &v^s(y)\end{pmatrix}^{-1}J\Psi^s_y v^u(\eta)\\
1810: \partial_{y_2}\eta&=v_1^s(y)\det\begin{pmatrix} v^s(y)
1811: &v^u(y)\end{pmatrix}^{-1}J\Psi^s_y v^u(\eta).
1812: \end{split}
1813: \]
1814: In fact, calling $\theta(x)$ the sine of the angle between stable and
1815: unstable directions at the point $x$ and $v_\perp$ the orthogonal
1816: unit vector to $v$, holds
1817: \[
1818: \begin{split}
1819: \partial_x\eta&=J\Psi^u_x\theta(x)^{-1}\, |v^s(y)\rangle\langle v^u_\perp(x)|\\
1820: \partial_y\eta&=J\Psi^s_y\theta(y)^{-1}\, |v^u(x)\rangle\langle v^s_\perp(y)|.
1821: \end{split}
1822: \]
1823: Accordingly,
1824: \begin{equation}\label{eq:jactot}
1825: \begin{split}
1826: J\Phi&:=\det\begin{pmatrix}
1827: \Id &-\Id\\
1828: \partial_{x}\eta&
1829: \partial_{y}\eta
1830: \end{pmatrix}=\det\begin{pmatrix}
1831: \partial_{x}\eta+
1832: \partial_{y}\eta
1833: \end{pmatrix}\\
1834: &=J\Psi^u_x\, J\Psi^s_y\,\theta(x)^{-1}\theta(y)^{-1}\langle v^s_\perp(y),\,
1835: v^u(x)\rangle^2\\
1836: &=J\Psi^u_x\, J\Psi^s_y\,\theta(x)^{-1}\theta(y)^{-1}\det\begin{pmatrix}
1837: v^u(x)&v^s(y)
1838: \end{pmatrix}^2.
1839: \end{split}
1840: \end{equation}
1841: Before starting computing we need to collect some facts.
1842: \begin{sublem}
1843: \label{slem:domains}
1844: If $x\in\partial D_{2\rho}$, $\rho\geq r\geq \cs_4\sqrt\ve$, then, for
1845: $C_4$ large enough,
1846: \begin{enumerate}[i)]
1847: \item if $\|x-y\|\leq \ve$, then $y\not\in D_\rho$.
1848: \item $\|x-\eta\|\leq \frac\ve\rho$.
1849: \item if $\zeta\in W^u(x)$ and $\|\zeta-x\|\leq\frac\ve\rho$, then
1850: $\zeta\not\in D_\rho$.
1851: \end{enumerate}
1852: \end{sublem}
1853: \begin{proof}
1854: The first inequality follows since $D_{2\rho}$ has a vertical size
1855: $4\rho^2$. Thus $2\rho^2-\ve\geq 2\rho^2-\frac
1856: 1{\cs_4^2}\rho^2\geq\rho^2$. Such an estimate and Proposition
1857: \ref{prop:dista} imply that the angle between $W^u(x)$ and $W^u(y)$ is
1858: at least $4K_-\rho^{-1}$, thus (ii). Finally, Proposition
1859: \ref{prop:dista} implies that, if $\zeta=(z_1,z_2)$,
1860: $z_2>2\rho^2-\frac{\ve^2}{\rho^2}\geq\rho^2$.
1861: \end{proof}
1862: We can now start computing the integral.
1863: \[
1864: \begin{split}
1865: \int_{\To^4}dxdy f(x)g(y)q_\ve(x-y)&=\int_{D_{\cs_4 r}^c}dx\int_{\To^2}dy
1866: f(x)g(y)q_\ve(x-y) +\Or(\|f\|_\infty\|g\|_\infty r^3)\\
1867: &\geq\int_{\Phi(D_{\cs_4 r}^c\times \To^2)}\!\!\!\!\!\! d\eta d\xi
1868: f(x)g(y)q_\ve(\xi)J\Phi^{-1}+
1869: \|f\|_\infty\|g\|_\infty \Or(r^3)
1870: \end{split}
1871: \]
1872: Next, from formula \eqref{eq:jactot} and Sub-lemma \ref{slem:domains}-$(i)$
1873: follows
1874: \begin{equation}\label{eq:jac-bound}
1875: |J\Phi(x,y)-1|\leq \cs\frac\ve{\rho^2}.
1876: \end{equation}
1877: Hence,
1878: \[
1879: \begin{split}
1880: \int_{\To^4}dxdy f(x)g(y)q_\ve(x-y)
1881: &=\int_{\To^2} d\eta
1882: f(\eta)g(\eta)+\|f\|_\infty\|g\|_\infty\Or\left(\int_{D_r^c}|1-J\Phi^{-1}|\right)\\
1883: &\quad+\Or\left(\int_{\Phi(D_{\cs_4 r}^c\times \To^2)} d\eta d\xi
1884: [f(x)-f(\eta)]g(y)q_\ve(\xi)\right) \\
1885: &\quad+\Or\left(\int_{\Phi(D_{\cs_4 r}^c\times \To^2)} d\eta d\xi
1886: f(\eta)[g(y)-g(\eta)]q_\ve(\xi)\right)\\
1887: &\quad +\|f\|_\infty\|g\|_\infty \Or(r^3).
1888: \end{split}
1889: \]
1890: To conclude we must compute the various error terms.
1891: For each $f\in L^\infty$, holds
1892: \[
1893: \int_{D_R}dx\,f(x)=\frac 23\int_0^Rd\rho\, \frac \rho{1+\rho}\int_{\partial
1894: D_\rho}ds\, f(s).
1895: \]
1896: Remembering
1897: \eqref{eq:jac-bound} and applying Fubini
1898: \[
1899: \int_{D_r^c}|1-J\Phi^{-1}|\leq \cs\|f\|_\infty\int_r^1 d\rho\,
1900: \rho^2\frac\ve{\rho^2}\leq \cs\|f\|_\infty\ve.
1901: \]
1902: Next, let $\gamma_u^\eta:[-\delta,\delta]\to\To^2$ be the unstable manifold
1903: of $\eta$, parametrized by arc-length, and let $s(\eta,\xi)$ be such
1904: that $\gamma_u^\eta(s(\eta,\xi))=x$. Recalling Sub-lemma \ref{slem:domains},
1905: \[
1906: \begin{split}
1907: \left|\int_{\Phi(D_{\cs_4 r}^c\times \To^2)}\!\!\!\!\!\!\!\!\!\!\!\! d\eta d\xi\,
1908: [f(x)-f(\eta)]g(y)q_\ve(\xi)\right|&\leq
1909: \|g\|_\infty\int_{\Phi(D_{\cs_4 r}^c\times \To^2)} \!\!\!\!\!\!\!\!\!\!\!\!d\eta d\xi
1910: \int_0^{s(\eta,\xi)}dt\,|\partial^u f(\gamma_u^\eta(t))|q_\ve(\xi)\\
1911: &\leq\cs \|g\|_\infty\int_{D_{r}^c\times
1912: \To^2} d\eta' d\xi\, |\partial^u f(\eta')|\frac{\|\xi\|}{|\theta(\eta')|}q_\ve(\xi)\\
1913: &\leq \cs \|g\|_\infty\int_0^1 d\rho \;\ve\int_{\partial
1914: D_\rho}|\partial^uf|\\
1915: &\leq \cs \|g\|_\infty\|\partial^uf\|_{L^1(\nu)}\ve.
1916: \end{split}
1917: \]
1918: Analogously,
1919: \[
1920: \left|\int_{D_{r}^c\times \To^2} d\eta d\xi
1921: f(\eta)[g(y)-g(\eta)]q_\ve(\xi)\right|\leq \cs
1922: \|f\|_\infty\|\partial^sg\|_{L^1(\nu)}\ve.
1923: \]
1924: We can finally collect all the above estimates and obtain
1925: \[
1926: \begin{split}
1927: \int_{\To^4}dxdy &f(x)g(y)q_\ve(x-y)=\int_{\To^2} d\eta
1928: f(\eta)g(\eta)\\
1929: &+(\|f\|_\infty+\|\partial^uf\|_{L^1(\nu)})(\|g\|_\infty
1930: +\|\partial^sg\|_{L^1(\nu)}) \Or(r^3+\ve)
1931: \end{split}
1932: \]
1933: from which the lemma follows by choosing $r=\ve^{\frac 13}$.
1934: \end{proof}
1935:
1936:
1937: \section{Decay of correlations}
1938: \label{sec:decay}
1939: Here we put together the results of the previous section to prove
1940: Theorem \ref{thm:main}.
1941:
1942: Let $f,g\in\Co^{1}(\To^2,\R)$, $\int f=0$, then
1943: \[
1944: \begin{split}
1945: \int\Lp^{kn_\ve}f g&=\sum_{i=0}^{k-1}\int\Lp^{n_\ve
1946: i}(\Lp^{n_\ve}-\Lp_\ve)\Lp_\ve^{k-i-1}f g+\int\Lp^k_\ve fg\\
1947: &=\sum_{i=0}^{k-1}\int(\Lp^{n_\ve}-\Lp_\ve)\Lp_\ve^{k-i-1}f g\circ T^{n_\ve
1948: i}+\Or(e^{-\cs k}\|f\|_1\,\|g\|_\infty)\\
1949: &=\sum_{i=0}^{k-1}\int(\Id-Q_\ve)\Lp^{n_\ve}\Lp_\ve^{k-i-1}f g\circ T^{n_\ve
1950: i}+\Or(e^{-\cs k}\|f\|_1\,\|g\|_\infty) ,
1951: \end{split}
1952: \]
1953: where we have used Lemma \ref{lem:l1}. To conclude, by using Lemma
1954: \ref{lem:appr}, we need to estimate the $L^1$ norm of
1955: $\partial^u(\Lp^{n_\ve}\Lp_\ve^{j}f)$. Since $|D\Lp_\ve^{j}f|\leq \cs
1956: \ve^{-1}|f|_\infty$, for $j>0$,
1957: \[
1958: \begin{split}
1959: \|\partial^u(\Lp^{n_\ve}\Lp_\ve^{j}f)\|_{L^1(\nu)}&\leq
1960: \int_0^{\cs_4\sqrt \ve}d\rho\int_{\partial
1961: D_\rho}|\partial^u(\Lp^{n_\ve}\Lp_\ve^{j}f)|
1962: +\int_{\cs_4\sqrt \ve}^1d\rho\int_{\partial
1963: D_\rho}|\partial^u(\Lp^{n_\ve}\Lp_\ve^{j}f)|\\
1964: &\leq \cs\|f\|_\infty \int_0^{\cs_4\sqrt \ve}d\rho\,\ve^{-1}\rho+
1965: \cs|f|_\infty\int_{\cs_4\sqrt \ve}^1d\rho\int_{\partial
1966: D_\rho}\frac{\ve}{\rho^2}\ve^{-1}\\
1967: &\leq \cs \|f\|_\infty\ln\ve^{-1},
1968: \end{split}
1969: \]
1970: where we have used Lemma \ref{lem:expansion} and Sub-lemma
1971: \ref{slem:domains}.\footnote{The above estimate is not sharp. With
1972: some extra work one could avoid the $\ln\ve^{-1}$, yet this would not
1973: change in any substantial way the result, so we chose to keep the
1974: presentation as short as possible.}
1975:
1976:
1977: Thus
1978: \begin{equation}
1979: \label{eq:sharp}
1980: \begin{split}
1981: \left|\int\Lp^{n}f
1982: g\right|=(\|f\|_{\infty}+\|\partial^u f\|_{L^1(\nu)})\|g\|_{\Co^1}\Or(n\ve^{\frac
1983: 32}\ln\ve^{-1}+
1984: e^{-\cs n\sqrt \ve}).
1985: \end{split}
1986: \end{equation}
1987: Clearly the best choice is $\ve=\cs (n^{-1}\ln n)^2$ which implies the
1988: Theorem.
1989:
1990:
1991:
1992: \section{Lower bound}
1993: \label{sec:lower}
1994: In this section we prove a lower bound.
1995: \begin{lem}\label{lem:lower}
1996: If there exists a sequence $\gamma_n$ such that, for each
1997: $f,g\in\Co^1$ holds
1998: \[
1999: \left|\int_{\To^2}f\circ T^{-n} g-\int_{\To^2}f\int_{\To^2}g\right| \leq
2000: (\|\partial^u f\|_{L^1(\nu)}+\|f\|_{L^1(\nu)})|g|_{\Co^1}\gamma_n,
2001: \]
2002: then there exists $\cs>0$ such that
2003: \[
2004: \gamma_n\geq n^{-2}\cs.
2005: \]
2006: \end{lem}
2007: \begin{proof}
2008: Let $g\geq 0$ be a smooth function supported away from zero (let us say that
2009: the support of $g$ does not intersect $D_{\frac 12}$). Next, let
2010: $\xi_0=(x_0,y_0)=(\frac 12,y_0)\in W^u(0)$, $\xi_n=T^{-n}\xi_0$. For
2011: each point $\eta$ in a neighborhood of $\xi_n$ let $z_1$ be the
2012: distance, along $W^u(0)$, between $\xi_n$ and $W^u(0)\cap W^s(\eta)$,
2013: and $z_2$ the distance, along $W^s(\xi_n)$, between $\xi_n$ and
2014: $W^u(\xi_n)\cap W^s(\eta)$. By construction
2015: $\Xi_n(\eta):=(z_1,z_2)$ is a map from a neighborhood of $\xi_n$ to a
2016: neighborhood of the origin with the property that the map transforms
2017: the stable and unstable foliation into the standard foliation given by
2018: the Cartesian coordinates. Clearly, $\Xi_n^{-1}(s,0)=\gamma^u(s)$ (the
2019: unstable manifold of the origin parametrized by arc length and such that
2020: $\gamma^u(0)=\xi_n$), while $\Xi_n^{-1}(0,s)=\gamma_n^s(s)$ (the unstable
2021: manifold of $\xi_n$ parametrized by arc length). Finally we define
2022: $f_n:=(\alpha_n\beta_n)\circ \Xi_n$ with
2023: $\alpha_n(z_1):=\varsigma(\cs_5 n z_1)$, $\beta_n(z_2):=\varsigma(\cs_5n^{-1}
2024: z_2)$, for $C_5$ small enough, and
2025: \[
2026: \varsigma(x):=\begin{cases}
2027: 1-|x+1|\quad &|x+1|\leq 1\\
2028: 0&|x+1|> 1.
2029: \end{cases}
2030: \]
2031: In other words, $f_n$ is a function essentially supported on a
2032: neighborhood left of $\xi_n$ of order $n^{-1}$ in the unstable direction
2033: and the stable. Accordingly, the supports of $f_n\circ T^{-k}$ and $g$
2034: are disjoint for all $k\leq \cs n$. Lemma \ref{lem:holo} implies that
2035: the the support is essentially a rhombus of size $n^{-1}$ and angle
2036: $n^{-1}$.
2037:
2038: Thus we have
2039: \[
2040: |g|_{\Co^1}(\|\partial^u f_n\|_{L^1(\nu)}+\|f_n\|_{L^1(\nu)})\gamma_n\geq
2041: \int_{\To^2}f_n\int_{\To^2}g\geq \cs n^{-3}.
2042: \]
2043: On the other hand, using again Lemma \ref{lem:holo},
2044: \[
2045: \|\partial^u f_n\|_{L^1(\nu)}\leq \cs n\|(\alpha'\beta)\circ
2046: \Xi_n\|_{L^1(\nu)}\leq \cs n^{-1}
2047: \]
2048: which yields the Lemma.
2049: \end{proof}
2050:
2051: \begin{rem}
2052: \label{rem:problems} Note that the norms in Lemma \ref{lem:lower} and
2053: in Theorem \ref{thm:main} (even in the stronger version given by
2054: \eqref{eq:sharp}) are different. It is not obvious that, putting the
2055: $L^\infty$ norm instead of the $L^1(\nu)$ one keeps the same rate of
2056: mixing. More generally, it is well known that in the uniformly
2057: hyperbolic setting the smoothness of the function can have an influence
2058: on the mixing rate. An analogous effect may arise in the present
2059: setting but it remains to be investigate. A related problem that needs to be
2060: addressed is the higher dimensional analogous of the present model
2061: where the fixed point has different possibility of losing full
2062: hyperbolicty. It is clear that the present
2063: result is only the starting point and not the end of the story.
2064: \end{rem}
2065:
2066:
2067: \begin{thebibliography}{99}
2068: \footnotesize
2069: \bibitem{ArPr} Artuso, R., Prampolini, A., {\em Correlation decay for
2070: an intermittent area-preserving map}, Physics Letters A, {\bf 246}
2071: (1998) 407--411.
2072: \bibitem {BG} G.Benettin, A.Giorgilli, {\em On the Hamiltonian
2073: interpolation of near-to-identity symplectic mappings with application
2074: to symplectic integration algorithms}, Journal of Statistical Physics,
2075: {\bf 74}, n. 5/6, (1994), 1117--1143.
2076: \bibitem {Bl} Bleher, P. M., {\em Statistical properties of
2077: two-dimensional periodic Lorentz gas with infinite horizon},
2078: J. Statist. Phys. {\bf 66} (1992), no. 1-2, 315--373.
2079: \bibitem {BuSi}Bunimovich, L. A.; Sinai, Ya. G., {\em Statistical
2080: properties of Lorentz gas with periodic configuration of scatterers},
2081: Comm. Math. Phys. {\bf 78} (1980/81), no. 4, 479--497.
2082: \bibitem{Go1} Gou\"ezel, S., {\em Sharp polynomial estimates for the
2083: decay of correlations}, Israel J. Math. {\bf 139} (2004), 29--65.
2084: \bibitem{Go2} Gou\"ezel, S., {\em Central limit theorem and stable laws
2085: for intermittent maps}, Probab. Theory Related Fields {\bf 128}
2086: (2004), no. 1, 82--122.
2087: \bibitem{KH} A.Katok, B.Hasselblatt, {\em Introduction to the Modern
2088: Theorey of Dynamical Systems}, Encyclpedia of Mathematical and its
2089: Applicatios, {\bf 54}, Cambridge University press, Cambridge (1995).
2090: \bibitem{Hu1} Hu, H., {\em Decay of correlations for piecewise smooth
2091: maps with indifferent fixed points}, Ergodic Theory Dynam. Systems
2092: {\bf 24} (2004), no. 2, 495--524.
2093: \bibitem{Hu2}Hu, H., {\em Statistical properties of some almost
2094: hyperbolic systems} Smooth ergodic theory and its applications
2095: (Seattle, WA, 1999), 367--384, Proc. Sympos. Pure Math., {\bf 69},
2096: Amer. Math. Soc., Providence, RI, 2001.
2097: \bibitem{L} Liverani C., {\em Flows, Random Perturbations and Rate of
2098: Mixing}, Ergodic Theory and Dynamical Systems, 18, 6,
2099: pp. 1421--1446 (1998).
2100: \bibitem{L2} Liverani C.,{\em Birth of an elliptic island in a chaotic
2101: sea}, Mathematical Physics Electronic Journal, {\bf 10}, 1
2102: (2004).e
2103: \bibitem{L3} Liverani C.,{\em Central Limit Theorem for Deterministic
2104: Systems}, International Conference on Dynamical Systems, Montevideo
2105: 1995, a tribute to Ricardo Ma\~ne, Pitman Research Notes in Mathemaics
2106: Series, {\bf 362}, editors F.Ledrappier, J.Levovicz, S.Newhouse, (1996).
2107: \bibitem{LSV} Liverani C.,, Saussol B., Vaienti S., {\em A Probabilistic
2108: Approach to Intermittency}, Ergodic Theory and Dynamical Systems, {\bf
2109: 19}, pp. 671--685 (1999).
2110: \bibitem{LW} Liverani C., Wojtkowski M., {\em Ergodicity in Hamiltonian
2111: Systems}, in Dynamics Reported {\bf 4}, Springer, Berlin, Heidelbeg,
2112: (1995).
2113: \bibitem{Ma} Maume, V. {\em Projective metrics and mixing properties
2114: on towers}, Trans. Am. Math. Soc. {\bf 353}, 3371--3389 (2001).
2115: \bibitem{Po} Pollicott, M., {\em Rates of mixing for potentials of
2116: summable variation} Trans. Am. Math. Soc. {\bf 352}, 843--853 (2000).
2117: \bibitem{PoSh} Pollicott, M., Sharp, R. {\em Invariance principles
2118: for interval maps with an indifferent fixed point},
2119: Comm. Math. Phys. {\bf 229} (2002), no. 2, 337--346.
2120: \bibitem{PoYu} Pollicott, M., Yuri, M. {\em Statistical properties of
2121: maps with indifferent periodic points}, Commun. Math. Phys. {\bf
2122: 217}, 503--520 (2001).
2123: \bibitem{PM} Y. Pomeau, P. Manneville, {\em Intermittent transition to
2124: turbulence in dissipative dynamical systems}, Communication in
2125: Mathematical Physics, {\bf 74}, 189--197 (1980).
2126: \bibitem{Sa} Sarig, O.M., {\em Thermodynamic Formalism for Countable
2127: Markov Shifts}, Ergodic Theory Dyn. Syst. {\bf 19}, 1565--1593 (1999)
2128: \bibitem{Wa} Wang X.J., {\em Statistical physics of temporal
2129: intermittency}, Phys. Rev. A {\bf 40}, 6647 (1989).
2130: \bibitem{Yo1} Young L.-S., {\em Recurrence times and rates of
2131: mixing}. Israel J. Math. {\bf 110} (1999), 153--188.
2132: \bibitem{Yu} Yuri M.,
2133: {\em Decay of correlations for certain multi-dimensional maps},
2134: Nonlinearity, {\bf 9}, n. 6 (1996) 1439--1461.
2135: \bibitem{Za} Zaslavsky, G. M.,
2136: {\em Chaos, fractional kinetics, and anomalous transport},
2137: Phys. Rep. {\bf 371} (2002), no. 6, 461--580.
2138: \bibitem{Zw} Zweim\"uller, R., {\em Stable limits for probability
2139: preserving maps with indifferent fixed points}, Stoch. Dyn. {\bf 3}
2140: (2003), no. 1, 83--99.
2141: \end{thebibliography}
2142: \end{document}
2143: