math0411590/KR.TEX
1: \documentclass[twoside]{article}
2: 
3: \usepackage{graphicx}
4: \usepackage{amsmath}
5: \usepackage{amssymb}
6: \usepackage{hhline}
7: \usepackage{epsfig}
8: 
9: \pagestyle{plain}
10: % \setlenght{\textwidth}{27pc}
11: % \setlenght{\textheight}{43pc}
12: 
13: \newtheorem{theorem}{Theorem}
14: \newtheorem{theorema}{Theorem.}
15: \newtheorem{theoremb}{Theorem}
16: \newtheorem{theoremc}{Theorem}
17: \newtheorem{theoremd}{Theorem}
18: \newtheorem{cor}[theoremd]{Corollary}
19: \newtheorem{dfn}[theoremb]{Definition}
20: \newtheorem{examp}[theorem]{Example}
21: \newtheorem{lem}[theorem]{Lemma}
22: \newtheorem{prop}[theorem]{Proposition}
23: \newtheorem{rk}[theoremc]{Remark}
24: 
25: \newenvironment{proof}[1][Proof]{\textbf{#1. }}{\qed}
26: \newenvironment{Proof}[1]{\textbf{#1. }}
27: 
28: \newcommand\bib[1]{\bibitem[#1]{#1}}
29: \renewcommand{\thetheorema}{}
30: \newcommand\abz{\hspace{12pt}}
31: \newcommand\qed{\phantom{\underline{y}}\hfill\hfill$\square$}
32: 
33: \newcommand\1{{\bf 1}}
34: \newcommand\CC{{\mathbb C}}
35: \newcommand\h{h_{\text{\rm top}}}
36: \newcommand\hm{H_{\text{\rm mult}}}
37: \newcommand\hs{H_{\text{\rm sing}}}
38: \newcommand\e{\epsilon}
39: \renewcommand\l{\lambda}
40: \newcommand\z{\sigma}
41: \newcommand\op[1]{\mathop{\rm #1}\nolimits}
42: \newcommand\po{$\!\!\!{\text{\bf.}}$ }
43: \newcommand\ls{\mathop{\overline\lim}\limits_{n \rightarrow \infty}}
44: \newcommand\N{{\mathbb N}}
45: \newcommand\Z{{\mathbb Z}}
46: \newcommand\R{{\mathbb R}}
47: \newcommand\D{{\mathcal D}}
48: \newcommand\X{{\mathcal X}}
49: \newcommand\Y{{\mathcal Y}}
50: \newcommand\T{{\mathcal T}}
51: \newcommand\vp{{\varphi}}
52: \newcommand\hps{\hskip-16pt . \hskip2pt}
53: \newcommand\hpss{\hskip-13.5pt . \hskip2pt}
54: \newcommand{\weg}[1]{}
55: 
56: \makeatletter
57: \renewcommand{\@oddhead}{\hfil Dynamics and entropy of SOC\hfil}
58: \renewcommand{\@evenhead}{\hfil Boris Kruglikov, Martin Rypdal \hfil}
59: \makeatother
60: 
61: \begin{document}
62: 
63: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65: 
66: \title{Dynamics and entropy in the Zhang model of Self-Organized Criticality}
67: \author{B. Kruglikov  \& M. Rypdal \\ ~ \\
68: {\small Institute of Mathematics and Statistics}\\
69: {\small University of Troms\o, N-9037 Troms\o, Norway}\\
70: {\small Boris.Kruglikov@matnat.uit.no;
71: Martin.Rypdal@matnat.uit.no} }
72: \date{}
73: \maketitle
74: 
75: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
76: \begin{abstract}
77: We give a detailed study of dynamical properties of the Zhang
78: model, including evaluation of topological entropy and estimates
79: for the Lyapunov exponents and the dimension of the attractor. In
80: the thermodynamic limit the entropy goes to zero and the Lyapunov
81: spectrum collapses.\!\!\footnote{Keywords: sand-pile models,
82: avalanche dynamics, skew-product systems,
83:  Lyapunov exponents, entropy, Hausdorff dimension, thermodynamic limit.}
84: \end{abstract}
85: 
86: 
87: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
88: %%%%%%%%%%%%%%%%%%%%%%%%
89: 
90: \section*{Introduction}
91: 
92: In 1987 the concept of Self-Organized Criticality (SOC) was introduced by
93: Bak, Tang and Wiesenfeld \cite{BTW}. The attempt was to give an
94: explanation of the omnipresence of fractal structures and power-law
95: statistics in nature, and the claim was that certain physical
96: systems can self-organize into stationary states, reminiscent of equilibrium
97: system at the critical point, in the sense that one has scale invariance
98: and long range correlations in space and time.
99: 
100: SOC is proposed as an explanation for variety of phenomena in
101: nature, such as earthquakes, forest fires, stock markets and
102: biological evolution \cite{J}. However, most work has been devoted
103: to the study of idealized ''sandpile-like'' computer models, such
104: as the sandpile model \cite{BTW}, the abelian sandpile \cite{DR}
105: and the Zhang model \cite{Z} that, one believes, exhibit SOC in
106: the thermodynamic limit. Despite this effort, a satisfactory
107: understanding of the model is not yet achieved. Through numerical
108: investigation it was observed that in the thermodynamic limit,
109: observables have power-law distributions. More precisely, the
110: probability distribution of an observable $s$ has the form $P(s)
111: \sim 1/s^{\tau_s}$ in the thermodynamic limit. There is no widely
112: agreed upon method for computing the SOC-exponents $\tau_s$
113: numerically and, due to the incomplete understanding of the
114: dynamics of the models and lack of a formal treatment of the
115: thermodynamic limit, it is difficult to properly explain the
116: observed behavior. Hence it is not clear what the SOC-exponents
117: really tell us about the dynamics of the SOC models.
118: 
119:  \subsection{\hpss Discussion of the Zhang model}
120: In a series of papers by  Cessac, Blanchard and Kr\"uger
121: \cite{BCK} it was proposed that deeper understanding of SOC models
122: can be achieved by studying the models in the framework of
123: dynamical system theory. They showed how a particular model, the
124: Zhang model, could be formulated as a dynamical system of
125: skew-product type with singularities, where the randomness of the
126: external driving is described by a Bernoulli shift, and the
127: threshold relaxation dynamics is given by piecewise affine maps.
128: 
129: In this paper we present a detailed study of the dynamical system
130: defined in \cite{BCK}. We prove several basic properties, some of
131: which are already stated in \cite{BCK}, before discussing
132: fundamental dynamical properties. Depending on the parameters of
133: the model, we can observe fundamentally different types of
134: behavior.
135: 
136: For low values of the threshold energy (critical energy), the
137: dynamics can be relatively simple, since the singularities only
138: effect the dynamics in a finite number of time-steps. In such
139: situations we say that singularities are removable, and we show
140: that the system permits symbolic coding. We give examples of how
141: symbolic coding provides a complete description of the
142: dynamics as a topological Markov chain. Hence the dynamics is
143: chaotic, but the essential dynamical invariants are all inherited
144: from the Bernoulli shift factor. Moreover we can identify the
145: physical invariant measure, and hence understanding of the
146: statistical properties is reduced to the theory of Markov chains.
147: 
148: As we increase the critical energy the role of the singularities
149: becomes essential. Techniques based on codings are no longer
150: applicable and a very interesting dynamics emerges. The
151: dimensional characteristics of the attractor are also sensitive to
152: the parameters of the system, as we show by generalizing the Moran
153: formula for the iterated function system (IFS). In addition, we
154: observe the situation, when the dimension of the IFS-attractor
155: increases to the maximum, while the support of the SRB-measure
156: remains fractal.
157: 
158: To measure the complexity of the dynamics we study entropy and
159: Lyapunov exponents. We show that the system is hyperbolic, with
160: one positive exponent originating in the Bernoulli shift. However,
161: due to the presence of singularities the Ruelle inequality and the
162: Pesin formula are not directly applicable. We show that the metric
163: entropy of any SRB-measure equals the topological entropy almost
164: surely, and we evaluate the latter generalizing the technique
165: developed by Buzzi \cite{B1,B2}. The result is that the Pesin
166: formula and the variational principle hold a posteriori.
167: 
168: To give a satisfactory physical interpretation of the dynamics we
169: rescale time to prevent infinitely slow driving of the system. We
170: prove that for this physical system, the Lyapunov spectrum
171: collapses completely and that the entropy goes to zero in the
172: thermodynamic limit. This implies that the expanding (chaotic)
173: properties are lost, so that we may expect power-laws statistics
174: and long range correlation effects.
175: 
176: The statistical properties we obtain hold for any SRB-measure,
177: because the most input comes from the Bernoulli shifts. The
178: existence of SRB-measures is in fact still an open problem. From
179: the general theory of dynamical systems with singularities
180: \cite{KS,P1,ST}, we can give conditions that are sufficient for
181: the existence of SRB-measures, but it is not known if these
182: conditions hold for the majority of parameters. We expect this to
183: be true (it was also conjectured in \cite{BCK}) and derive some
184: statistical corollaries.
185: 
186: Apart from the physical importance of the Zhang model, it is
187: interesting from a mathematical point of view. It can be described
188: as a piecewise affine hyperbolic map of the form
189:  $$
190: F:\Sigma_N^+\times M\to\Sigma_N^+\times M,\quad
191: ((t_0t_1t_2\dots),x)\mapsto((t_1t_2t_3\dots),f_{t_0}(x)),
192:  $$
193: where $\Sigma_N^+$ is the set of right infinite sequences from a
194: finite alphabet and $\{f_i\}$ a collection of piece-wise affine
195: non-expanding maps of $M$ to itself. Previously piecewise affine
196: expanding maps and piecewise isometries have been studied, but the
197: contracting property of the relaxation dynamics gives rise to some
198: difficulties. Therefore several methods are developed in this
199: paper, which hold far beyond the framework of the Zhang model.
200: 
201:  \subsection{\hpss Structure of the paper}
202: In Section \ref{sec_1} we describe the model and derive bounds on
203: the size and duration of avalanches. This enables us to
204: use the Poincar\'e return to reformulate the systems in a
205: skew-product form. Then we study the contraction property to
206: conclude hyperbolicity of the model (Theorem \ref{5}) and
207: describe, when degenerations occur (Theorem \ref{th_invert}; the
208: original Zhang setting $\e=0$ is not the only possibility). In
209: Section \ref{sec_2} we introduce the concept of removability of
210: singularities, which appears in the coding approach for the study
211: of the model.
212: 
213: Section \ref{sec_3} is devoted to the study of measure entropy and
214: Lyapunov spectrum. We prove in Theorem \ref{17} that the entropy
215: of an SRB-measure is always maximal. Section \ref{sec_4} concerns
216: the topological entropy. We evaluate it for the most parameter
217: values (Theorems \ref{140} and \ref{a.s.0}). We also discuss
218: nearly-Zhang models and show that the dynamical quantities do not
219: change. This is natural from the physical perspective, because SOC
220: should not be obtained through a fine tuning of parameters.
221: 
222: Section \ref{sec_5} briefly describes the dimension issues of the
223: model (Theorem \ref{theo-5} gives the asymptotic values), which
224: enters into all inter-relations involving entropy and
225: characteristic exponents. We demonstrate how the fractality occurs
226: in IFS-context, noting the difference due to singularities and
227: overlaps. In Section \ref{examples} we illustrate the most
228: important effects in the model by examples.
229: 
230: In Section \ref{sec_7} we discuss the thermodynamic limit and
231: attempt to explain appearance of the power-law statistics via a
232: reparametrization. Conclusion contains the physical implications
233: of the current investigation.
234: 
235: In Appendices \ref{app_A} and \ref{app_B} we provide bounds for
236: the entropy and the dimension, which are new in the presence of
237: singularities, overlaps and degenerations (this was designed for
238: an application to the Zhang model). The results are of interest in
239: its own and can be read independently.
240: 
241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
242: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
243: 
244: \section{\hps Basic properties of the Zhang model}\label{sec_1}
245: 
246: In the Zhang model each site on the lattice is associated with a
247: non-negative real number, which we call the energy of the site.
248: The collection of energies is called an energy configuration, and
249: can be represented as a point in $N$-dimensional space, where $N$
250: is the number of sites in the lattice. If a configuration is
251: unstable, the overcritical sites will lose some of their energy to
252: their nearest neighbors, resulting in a new energy configuration.
253: This transformation on $\R^N$ is denoted by $f$. If a
254: configuration is stable, a site is chosen at random and an energy
255: quantum $\delta=1$ is added to this site. In \cite{BCK} it was
256: shown how the relaxation and random excitation can be formulated
257: as a map of skew-product type on an extended phase-space. This
258: extended phase-space has the configuration space as one factor,
259: and the set of all possible sequences of excitations as the other
260: factor. In \cite{BCK} it was also shown how one can reformulate
261: the dynamical system by considering the return maps to the set of
262: stable configurations. This gives a simplification, in the sense
263: that each avalanche is associated with an affine transformation.
264: The set of stable configurations is partitioned into domains,
265: where each domain corresponds to an avalanche.
266: 
267: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
268: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
269: 
270: \subsection{\hpss Relaxation}
271: 
272: Take $d,L \in {\mathbb N}$ and let $\Lambda \subset {\mathbb Z}^d$
273: be the cube $[1,L]^d$ of cardinality $N:=L^d=|\Lambda|$. Let
274: $\phi:\Lambda \rightarrow \Lambda':= \{1,\dots,N\}$ be a
275: bijection. We define a metric $d_{\Lambda}$ on $\Lambda$ by
276: $$
277: d_\Lambda ({\bf k},{\bf l})=\sum_{1 \leq n \leq d} |k_n-l_n|\,,
278: $$
279: and let $d_{\Lambda'}:=\phi_{*} d_\Lambda$. In the following we
280: omit primes when it is clear from the context that we are
281: considering the metric space $(\Lambda', d_{\Lambda'})$. Elements
282: of $\Lambda$ will be called sites. We say that sites $i$ and $j$
283: are nearest neighbors if $d_{\Lambda}(i,j)=1$. The boundary
284: $\partial \Lambda$ is defined as those sites $i \in \Lambda$ that
285: have less than $2d$ nearest neighbors.
286: 
287: Fix parameters $E_c>0$ and $\epsilon \in  [0,1)$ and define
288: $f:{\mathbb R}^N_{\geq 0} \rightarrow {\mathbb R}^N_{\geq 0}$ by
289: $$
290: f(x)_i=x_i-\theta(x_i-E_c)(1-\epsilon)x_i+\frac{1-\epsilon}{2d}
291: \sum_{d_\Lambda (i,j)=1} \theta(x_{j}-E_c)x_{j}\,,
292: $$
293: where
294: $$
295: \theta(a)= \begin{cases} 1 \text{ if } a>0 \\ 0 \text{ if } a \leq
296: 0\end{cases}\,.
297: $$
298: Let $\|x\|_1=\sum_{i=1}^N |x_i|$ be the 1-norm on $\R^N$.
299: \begin{prop}\po \label{1}
300: For all $x \in \R_{\geq 0}^N$ we have
301: $$
302: \frac{1+\epsilon}{2} \|x\|_1\leq \|f(x)\|_1 \leq \|x\|_1\,,
303: $$
304: and $\|f(x)\|_1 = \|x\|_1$ if and only if $x_i \leq E_c$ for all
305: $i \in \partial \Lambda$. If there is $i \in \partial \Lambda$
306: such that $x_i>E_c$ then
307: $$
308: \|f(x)\|_1 \leq \|x\|_1 - \frac{1-\epsilon}{2d}E_c\,.
309: $$
310: \end{prop}
311: \begin{proof}
312: Let $\{x_{i_k} \}_{k=1}^m$ be the entries of the vector $x$ that
313: are greater than $E_c$. Let $n_{i_k}$ be the number of nearest
314: neighbors of $x_{i_k}$. Then
315: \begin{eqnarray*}
316: \|f(x)\|_1 &=& \sum_{i \in \Lambda} f(x)_i =\sum_{i \in \Lambda}
317: x_i-(1-\epsilon)\sum_{k=1}^m x_{i_k}+\frac{1-\epsilon}{2d}
318: \sum_{k=1}^m n_{i_k} x_{i_k} \\
319: &=&\sum_{i \in \Lambda} x_i -(1-\epsilon) \sum_{k=1}^m
320: (1-\frac{n_{i_k}}{2d})x_{i_k}\,.
321: \end{eqnarray*}
322: The statement follows from the fact that we always have $d \leq
323: n_{i_k} \leq 2d$, and $n_{i_k} = 2d$ if and only if $x_{i_k} \not
324: \in \partial \Lambda$.
325: \end{proof}
326: \\
327: 
328: We say that a site $i \in \Lambda$ of the configuration $x$ is
329: relaxed if $x_i \leq E_c$, and excited if $x_i>E_c$. A
330: configuration $x$ is called stable if all sites are relaxed. The
331: set of stable configurations is $M:=[0,E_c]^N$. For each
332: configuration $x$ we define $m(x)=\min \{n \geq 0\,|\,f^n(x) \in
333: M\}$.
334: \begin{prop}\po \label{2}
335: For all $x \in \R_{\geq 0}^N$ we have:
336: $$m(x) \leq
337: \frac{2dN}{1-\epsilon}\frac{\|x\|_1}{E_c}
338: \Big{(}\frac{2d}{1-\epsilon}+1\Big{)}^{\text{\em
339: diam}(\Lambda)/2}\,.
340: $$
341: \end{prop}
342: 
343: We need the following lemma ($[ \cdot ]$ denotes integer part):
344: 
345: \begin{lem}\po \label{3}
346: For $x \in \mathbb{R}^N_{\geq 0}$ and $n \in \mathbb{N}$ let
347: $\alpha_i(n,x)$ be the cardinality of the set $\{ l \leq
348: n\,|\,(f^l x)_i>E_c\}$.  Let $\gamma=[2d/(1-\epsilon)]+1$. If
349: $d_\Lambda(i,j)=1$, then $\alpha_j(n,x) \geq
350: [\alpha_i(n,x)/\gamma]$.
351: \end{lem}
352: 
353: \noindent
354: \begin{proof}
355: There is a finite increasing sequence $\{m_k\}$, such that
356: $(f^{m_k}(x))_i>E_c$. We claim that on each interval $(m_k,
357: m_{k+\gamma}]$ there is a number $m$ such that $(f^m(x))_j>
358: E_c$. In fact, in the opposite case
359: $$
360: \big{(}f^{1+m_{k+\gamma-1}}(x) \big{)}_j \geq \gamma
361: \frac{1-\epsilon}{2d} E_c>E_c\,.
362: $$
363: Since $[0,\alpha_i(n,x)]$ contains $\beta=[\alpha_i(n,x)/\gamma]$
364: disjoined such intervals, we get $\alpha_j(n,x) \geq \beta$. Thus
365: $\alpha_j(n,x) \geq [ \alpha_i(n,x)/\gamma ]$.
366: \end{proof}
367: \\
368: 
369: \noindent {\bf Proof of Proposition \ref{2}.} By applying
370: inductively Lemma \ref{3} we get:
371: $$
372: \alpha_j(n,x) \geq
373: \Big{[}\frac{\alpha_i(n,x)}{\gamma^{d_\Lambda(i,j)}}\Big{]}
374: $$
375: In fact, if $j=j_0,j_1,\dots j_k=i$ is a path with
376: $d_\Lambda(j_s,j_{s+1})=1$ and
377: $$
378: \Big{[} \frac{\alpha_i(n,x)}{\gamma^k} \Big{]}=t\,,
379: $$
380: then $\alpha_{j_k}(n,x) \geq t\gamma^k,\alpha_{j_{k-1}}(n,x) \geq
381: t\gamma^{k-1},\dots, \alpha_{j_0}(n,x) \geq t$. By Proposition
382: \ref{1}
383: $$
384: \alpha_j(n,x) \leq \frac{2d}{1-\epsilon} \frac{\|x\|_1}{E_c}
385: $$
386: for $j \in \partial \Lambda$, so
387: $$
388: \alpha_i(n,x) \leq \alpha(x):= \frac{2d}{1-\epsilon}
389: \frac{\|x\|_1}{E_c} \gamma^{\text{diam}(\Lambda)/2}\,.
390: $$
391: for all $i \in \Lambda$ and all $n \in \mathbb{N}$. Suppose
392: $f^m(x) \not \in M$ for all $m \leq T$. If $T>N\alpha(x)$, then
393: there must be a site $i \in \Lambda$ that is greater than $E_c$
394: for more than $\alpha(x)$ different times. This is impossible so
395: $m(x) \leq N\alpha(x)$. \qed
396: 
397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
398: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
399: 
400: \subsection{\hpss Random excitations}\label{randex}
401: 
402: Define $\Sigma_N^+ = \Lambda^{\N}$ to be the set of
403: right-infinite $\Lambda$-sequences and let $\sigma_N^+:\Sigma_N^+
404: \rightarrow \Sigma_N^+$ be the left shift. We define a map
405: $\hat{f}:\Sigma_N^+ \times {\R}_{\geq 0}^N \rightarrow \Sigma_N^+
406: \times {\R}_{\geq 0}^N$  by
407: $$
408: \hat{f}({\bf t},x)=\begin{cases}  (\sigma_N^+ {\bf t},x+e_{t_0}) &
409: \,\mbox{if} \,\, x \in M \\ ({\bf t},f(x)) & \,\mbox{if} \,\, x
410: \not \in M
411: \end{cases}\,,
412: $$
413: where $e_{1},\dots,e_N$ is the standard basis in $ \mathbb{R}^N$.
414: We denote points in $\Sigma_N^+ \times \mathbb{R}_{\geq 0}^N$ by
415: $\hat{x}=({\bf t},x)$, and we define $\pi_u$ and $\pi_s$ to be the
416: projections to $\Sigma_N^+$ and $\R_{\geq 0}^N$ respectively.
417: 
418: \begin{prop}\po \label{4}
419: For all $\hat{x} \in \Sigma_N^+ \times \R_{\geq 0}^N$ it holds:
420: $$
421: \min \{ m \geq 0 \,|\,\forall i \in \Lambda \, \exists  m' \leq
422: m:\, (\pi_s \circ \hat{f}^{m'}(\hat{x}))_i>E_c\} \leq n(E_c,
423: \epsilon, \Lambda)\,,
424: $$
425: where
426: $$
427: n(E_c,\e, \Lambda)= N(NE_c+2) \Big{(}\Big{[}\frac{2d}{1-\e}
428: \Big{]}+1 \Big{)}^{\text{{\em diam}}(\Lambda)} \,,
429: $$
430: \end{prop}
431: \begin{proof}
432: In $N[E_c]+1$ time-steps, there must be an overcritical site.
433: Since in the relaxation process there is always an overcritical
434: site, then during arbitrary subsequent $N[E_c]+2$ time-steps an
435: exited site can be found. Hence after $N\xi(N[E_c]+2)$ time-steps
436: either all sites have been overcritical or there is a site that
437: has been overcritical at least $\xi$ times. However it follows
438: from the proof of Proposition \ref{2} that if one site is
439: overcritical
440: $$\xi=\Big{(}\Big{[}\frac{2d}{1-\e}
441: \Big{]}+1 \Big{)}^{\text{diam}(\Lambda)}$$
442: times, then all sites have been overcritical at least once.
443: \end{proof}
444: \\
445: 
446: For $x \in \mathbb{R}_{\geq 0}^N$ and $i \in \Lambda$ we define
447: $\tau(i,x):= \min \{n \in \N \,|\,f^n(x+e_i) \in M\}$. Proposition
448: \ref{2} assures us that this number is finite and
449:  $$
450: \max_{x\in M}\max_{i\in \Lambda}\tau(i,x)\le
451: \tau_m(E_c,\e,\Lambda)=N^2\Bigl(1+\frac1{NE_c}\Bigr) \Big{(}
452: \frac{2d}{1-\e}+1 \Big{)}^{\text{diam}(\Lambda)/2+1}.
453: $$
454: Thus we observe that neither $n(E_c,\e, \Lambda)$ nor
455: $\tau_m(E_c,\e,\Lambda)$ are uniformly bounded in $E_c$, but there
456: is the following alternative:
457:  %("certainty relation")
458: 
459: {\it There exists a constant $C_0$, not depending on the energy
460: $E_c$, such that either $n(E_c,\e, \Lambda)\le C_0$ or
461: $\tau_m(E_c,\e,\Lambda)\le C_0$.}
462: 
463: In fact, we can set $C_0=3N^2\bigl(\frac{2d}{1-\e}+1
464: \bigr)^{\text{diam}(\Lambda)+1}$. Thus we get that either
465: relaxation happen sufficiently fast or all the sites keep being
466: excited sufficiently often (uniformly in $E_c$).
467: 
468: But there does not exist such a bound uniform in $\e$ or $N$.
469: 
470: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
471: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
472: 
473: \subsection{\hpss Return maps} \label{return maps}
474: 
475: Let $\hat{x}=({\bf t},x) \in \Sigma_N^+ \times M$. For
476: $n=1,\dots,\tau(t_0,x)$ define $C_n(\hat{x})=\{ i \in \Lambda\,|\,
477: (\pi^s \circ \hat{f}^{n} \hat{x})_i>E_c \}$, and
478: $A(\hat{x})=\big{(}C_1(\hat{x}),\dots,C_{\tau(t_0,x)}
479: (\hat{x})\big{)}$.  We call $A(\hat{x})$ the {\em avalanche} of
480: the point $\hat{x}$. Let $\hat{M}:=\Sigma_N^+ \times M$ and define
481: an equivalence relation $\sim$ on $\hat{M}$ by
482:  $$
483: \hat{x} \sim \hat{y} \Leftrightarrow A(\hat{x})=A(\hat{y})\,.
484:  $$
485: This gives a partition of $\hat{M}$. From the definition it is clear
486: that $A(\hat{x})$ depends on $t_0$ and $x$ only. Hence partition
487: elements are of the form $[i] \times M_{ij}$, where
488:  $$
489: \forall i \in \Lambda:\, \bigcup_j M_{ij}=M
490:  $$
491: and $[i]=\{{\bf t} \in \Sigma_N^+\,|\,t_0=i\}$ is the cylinder of
492: the symbol $i$. We see that for each $i \in \Lambda$, the domains
493: $M_{i1},M_{i2},\dots$ are separated by segments of at most
494: $N!^{\tau_m}=\exp{(\tau_m(E_c,\e,\Lambda)\log N!)}$ hyperplanes.
495: Hence we have a finite number of domains $M_{i1},\dots,M_{iq_i}$
496: for each $i \in \Lambda$. By definition there is a unique
497: avalanche for each partition element $[i] \times M_{ij}$. We
498: denote this avalanche by $A_{ij}$. Its duration is
499: $\tau_{ij}:=\tau(i,x)$, for $x \in M_{ij}$, and define its size to
500: be $s_{ij}=\sum_{n=1}^{\tau_{ij}} |C_n|$.
501: 
502: We define the piecewise continuous map $F:\hat{M} \rightarrow
503: \hat{M}$  by
504: $$({\bf t},x) \mapsto (\sigma_N^+ {\bf t}, F_{t_0}x)$$ where
505: $F_i(x):=f^{\tau(i,x)}(x+e_i)$. We define $F_{ij}:=F_i
506: |_{M_{ij}}$.
507: 
508:  \begin{rk}\po\label{ps.vs.mt}
509: From a mathematical point of view the formulation $(\hat{M},F)$ is
510: a simplification compared to $(\Sigma_N^+ \times \R_{\geq
511: 0}^N,\hat{f})$. However, the duration of avalanches are suppressed
512: so that all avalanches have the same duration. This is not
513: satisfactory from a physical point of view, and hence we call
514: $(\hat{M},F)$ the mathematical model and $(\Sigma_N^+ \times
515: \R_{\geq 0}^N,\hat{f})$  the physical model. We will later make a
516: rescaling of time in the physical model, so that the driving does
517: not become infinitely slow in the thermodynamic limit.
518:  \end{rk}
519: 
520: For each $x \in \mathbb{R}^N_{\geq 0}$ we define a matrix $Q(x)$
521: by
522:  $$
523: Q_{kl}(x)=\begin{cases}
524: \frac{1}{2d} \theta(x_l-E_c)   & \,\mbox{if} \,\, d_\Lambda(k,l)=1\,, \\
525:  0                             & \,\mbox{otherwise}\,,
526: \end{cases}
527:  $$
528: and a diagonal matrix $J(x)$ by
529: $J_{kl}(x)=(1-(1-\epsilon)\theta(x_l-E_c))\delta_{kl}$. Set
530:  \begin{equation}\label{defS}
531: S(x)=J(x)+(1-\epsilon)Q(x)
532:  \end{equation}
533: and observe that $f(x)=S(x)x$.  Let $x(1)=x+e_i$ and
534: $x(n)=f(x(n-1))$ for $n \in \{2,\dots,\tau(t,x)\}$. Then
535: $F_i(x)=L_i(x+e_i)$, where
536: $$
537: L_i(x)=S(x(\tau(x,i))) \dots S(x(1))\,.
538: $$
539: If $x,y \in M_{ij}$, then $\tau(i,x)=\tau(i,y)$ and the same
540: components of $x(n)$ and $y(n)$ are grater than $E_c$ for each
541: $n=1,\dots,\tau(t,x)$, so $L_i(x)=L_i(y)$. We define the linear
542: map $L_{ij}:=L_i(x)$ for $x \in M_{ij}$. We get
543: $F_i|_{M_{ij}}(x)=L_{ij}(x+e_i)$.
544: 
545:  \begin{dfn} \po \label{d1}
546: A sequence $\{(i_n,j_n)\,|\,1\leq n \leq \theta\}$ is said to be
547: admissible if
548: $$
549: \bigcap_{n=1}^\theta   (F_{i_{n-1} j_{n-1}}\circ\dots\circ F_{i_1
550: j_1})^{-1}(M_{i_{n}j_{n}})  \neq \emptyset\,,
551: $$
552:  \end{dfn}
553: 
554:  \begin{theorem}\po \label{5}
555: For all $i,j$ $\|L_{ij}\|_1 \leq 1$. Moreover for every constant
556: $c\in (0,1)$ there is a number $T \in \mathbb{N}$ such that for
557: every $\theta> T$ and admissible sequence $\{(i_n,j_n)\,|\,1\leq n
558: \leq \theta\}$ it holds:
559: $$
560: \|L_{i_\theta j_\theta} \dots L_{i_{1}j_{1}}\|_1<c\,.
561: $$
562:  \end{theorem}
563: 
564: \noindent
565: \begin{proof}
566: If $A$ is an $N \times N$ matrix we let $C_k(A)$ be its $k$-th
567: column. Observe that for any matrices $A$ and $B$ we have the
568: following formula:
569:  \begin{equation}\label{column}
570: \|C_k(AB)\|_1=\sum_l \|C_l(A)\|_1 B_{lk}\,.
571:  \end{equation}
572: By the construction: $\|C_k(S(x))\|_1 \leq 1$ for all $k \in
573: \Lambda$. Hence $\|L_{ij}\|_1 \leq 1$.
574: 
575: To prove the second statement we note that for $\e>0$ the diagonal
576: elements of the matrices $S(x)$ are non-zero and $\ge\e$.
577: Therefore
578:  $$
579: (S(x(m))S(x(m-1)))_{kl}\geq
580: \e\cdot\max\{S_{kl}(x(m)),S_{kl}(x(m-1))\}.
581:  $$
582: Moreover, $S_{kl}(x(m))>0$ if $x(m)_l>0$ and $d_\Lambda(k,l)=1$.
583: It follows that any admissible product $L_{i_\theta j_\theta}
584: \dots L_{i_1 j_1}$ of length $\theta \geq n(E_c, \epsilon,
585: \Lambda)$ is positive. By Proposition \ref{1} there must be at
586: least one column such that the sum over this column is less than
587: $1$, for some factor $L_{i_tj_t}$ and hence for the whole product.
588: Therefore the sum over each column of any admissible product of
589: length $2n(E_c,\epsilon, \Lambda)$ must be less than $1$. Let
590: $c_0<1$ be the maximal norm of all admissible products of length
591: $2 n(E_c,\epsilon,\Lambda)$.  For $k>k_0:=[\log c/\log c_0]+1$ we
592: have $c_0^k<c$ and hence $T=2k_0n(E_c,\epsilon, \Lambda)$ is the
593: required number.
594: 
595: The above argument does not apply to the case $\e=0$, and a different
596: proof must be given for this case (which actually works in general as well).
597: Take $\hat{x}\in \hat{M}$
598: and let $x(t) \in M$ be the projection of its orbit to $\hat{M}$.
599: Denote $S(x(t))$ by $S_t(\hat{x})$, and let
600: $$\tilde{S}_t(\hat{x})=S_t(\hat{x}) \cdots S_0(\hat{x})\,.$$
601: We make the following claims:
602: \begin{enumerate}
603: \item There exists $\bar{n} \in \N$ such that for all $l,m \in \Lambda$
604: and all $\hat{x} \in \hat{M}$ there is $t\le\bar{n}$ such that
605: $(\tilde{S}_t(\hat{x}))_{lm} \neq 0$.
606: \item For all $i \geq 0$ there exists $n_i \in \N$ such that
607: $\|C_m(\tilde{S}_t(\hat x))\|_1<1$ for all $t \geq n_i$, $\hat
608: x\in\hat M$ and all sites $m \in \Lambda$ with
609: $d_\Lambda(m,\partial \Lambda)\leq i$.
610: \end{enumerate}
611: 
612: The second claim for $i=\frac12\op{diam}(\Lambda)$
613: implies the statement of the theorem.
614: 
615: To see the first claim we fix $\hat x$ and let $U\subset\Lambda^2$
616: be the subset of the pairs $(l,m)$ with $(\tilde S_t)_{lm}=0$ for
617: all sufficiently large $t$. By inductively applying (\ref{column})
618: we see that the columns for $\tilde{S}_t(\hat{x})$ are non-zero
619: for all $t\geq 0$. So for all $\beta \in \Lambda$ there is $\alpha
620: \in \Lambda$ such that $(\alpha, \beta) \in \Lambda^2 \setminus
621: U$. Given sites $\alpha$ and $\beta$ we choose $t$ such that
622: $\tilde{S}_t(\hat{x})_{\alpha \beta} \neq 0$. Consider now column
623: $\alpha$ of the matrix
624: $\tilde{S}_{t+1}(\hat{x})=S_{t+1}(\hat{x})\tilde{S}_{t}(\hat{x})$.
625: If $\alpha$ is stable, i.e. $x(t+1)_\alpha\leq E_c$, then
626: $(S_{t+1}(\hat{x}))_{\alpha \alpha}>0$ and
627: $(\tilde{S}_{t+1}(\hat{x}))_{\alpha \beta} \neq 0$, so we just
628: repeat the argument. But the site $\alpha$ can not be stable for
629: more than $n(E_c,0,\Lambda)$ iterations. Hence we can with no loss
630: of generality choose $t$ such that $x(t+1)_\alpha > E_c$. Then the
631: column $\alpha$ of $S_{t+1}(\hat{x})$ has non-zero elements in all
632: position that correspond to neighbors of $\alpha$. Hence we obtain
633: that $(\alpha',\beta) \in \Lambda^2 \setminus U$ for all $\alpha'$
634: with $d_\Lambda(\alpha',\alpha)=1$. Any two points can be
635: connected by a path of neighbors, so $U=\emptyset$, and the first
636: claim follows. In fact, one can see that the bound $\bar n$ does
637: not depend on a choice of $\hat x$ and satisfies: $\bar n\le
638: \op{diam}(\Lambda)\cdot n(E_c,0,\Lambda)$.
639: 
640: To prove the second claim let us note that if
641: $\|C_k(\tilde{S}_{t}(\hat{x}))\|_1<1$, then
642: $\|C_k(\tilde{S}_{t+1}(\hat{x}))\|_1<1$ because by (\ref{column}):
643: $\|C_k(AB)\|\le\op{max}_l\|C_l(A)\|\cdot\|C_k(B)\|$.
644: 
645: We will use induction on $i$ starting from $i=0$. Take $k \in
646: \partial \Lambda$ and $t \leq \bar{n}$ such that
647: $(\tilde{S}_t(\hat{x}))_{kk}\neq 0$. If $x(t+1)_k > E_c$, then
648: $\|C_k(S_{t+1}(\hat{x}))\|_1<1$ and
649: $$
650: \|C_k(\tilde{S}_{t+1}(\hat{x}))\|_1=\sum_l
651: \|C_l(S_{t+1}(\hat{x}))\|_1 (\tilde{S}_{t}(\hat{x}))_{lk}<1\,,
652: $$
653: and so we have the desired inequality. If $x(t+1)_k \leq E_c$,
654: then $(S_{t+1}(\hat{x}))_{kk}=1$ and hence
655: $(\tilde{S}_{t+1}(\hat{x}))_{kk}\neq 0$. Then we repeat the
656: argument. Since no site can be stable for more than
657: $n(E_c,0,\Lambda)$ successive time-steps we obtain the claim for
658: $i=0$ with $n_0=n(E_c,0,\Lambda)+\bar{n}$.
659: 
660: Consider now the case $i>0$. For a site $m \in \Lambda$ with
661: $d(m,\partial \Lambda)=i$, we take $l\in \Lambda$ with
662: $d_\Lambda(l,m)=1$ and $d_\Lambda(l,\partial \Lambda)=i-1$. By the
663: first claim we find some $t\le\bar n$ such that
664: $(\tilde{S}_t(\hat{x}))_{lm} \neq \emptyset$, and by the induction
665: hypothesis for $t'\ge n_{i-1}$ we have:
666:  $$
667: \|C_l\big{(}S_{t+t'}(\hat{x})\cdots
668: S_{t+1}(\hat{x})\big{)}\|_1<1\,.
669:  $$
670: Using (\ref{column}) we obtain
671: $\|C_m\big{(}\tilde{S}_{t+t'}(\hat{x})\big{)}\|_1<1$. We can
672: choose $n_i=n_{i-1}+\bar{n}$.
673: \end{proof}
674: 
675:  \begin{lem}\po\label{6}
676: Let $E_c \geq \epsilon/(1-\epsilon)$. Then for any $\hat{x} \in
677: \hat{M}$, $n \in \N$ and $i,j \in C_n(\hat{x})$ we have
678: $d_\Lambda(i,j) \neq 1$.
679:  \end{lem}
680:  \begin{proof}
681: Take $\hat{x} \in M$ and let $E_n$ be the maximal energy of a site
682: in $C_n(\hat{x})$. Clearly $E_1 \leq E_c+1$ and
683:  $$
684: E_{n+1} \leq \max \big{\{} \max\{\epsilon E_n, E_c
685: \}+(1-\epsilon)E_n, E_c+1 \big{\}}\,.
686:  $$
687: From this we see by induction that
688:  $$
689: E_n \leq \max \big{\{} \frac{E_c}{\epsilon}, E_c+1
690: \big{\}}=\frac{E_c}{\epsilon}\,,
691:  $$
692: so $\epsilon E_n \leq E_c$ for all $n \in \N$, and this means that
693: a site cannot be overcritical in two successive time-steps (for
694: $\e=0$ the above argument does not work, but the statement holds
695: obviously).
696: 
697: All avalanches start with a single site. Let $C_1(\hat{x})=\{i\}$.
698: Then $d(i,j)=1$ for all $j \in C_2(\hat{x})$. This implies that
699: any two elements of $C_2(\hat{x})$ can be connected with a path of
700: length 2, so no two sites of $C_2(\hat{x})$ are nearest neighbors.
701: If there exists a path of even length between two points in
702: $\Lambda$, then all paths connecting these points are of even
703: length. Therefore we can repeat the argument proving by induction
704: that $d_\Lambda(i,j) \in 2 {\mathbb Z}$ for all $i,j \in
705: C_n(\hat{x})$.
706: \end{proof}
707: 
708: 
709:  \begin{prop}\po \label{7}
710: The linear maps $L_{ij}$ are all invertible whenever
711: $\epsilon\ge 1/2$ or $\e>0$ and $E_c \geq \epsilon/(1-\epsilon)$.
712: If we have $E_c \geq \epsilon/(1-\epsilon)$, then
713: $$
714: \det L_{ij}=\epsilon^{s_{ij}}\,.
715: $$
716:  \end{prop}
717: \begin{proof}
718: Take arbitrary $x \in \R^N_{\geq 0}$. First we observe that since
719: the sum over each column of $Q(x)$ is less than or equal to $1$,
720: we have $\|Q(x)v\|_1 \leq \|v\|_1$ for each $v \in \R^N$. This
721: implies that
722: \begin{eqnarray*}
723: \|S(x)v\|_1 &=&\|(J(x)+(1-\epsilon)Q(x))v\|_1 \\ &\geq&
724: \|J(x)v\|_1-(1-\epsilon)\|Q(x)v\|_1 \\
725: &\geq&(2\epsilon-1)\|v\|_1\,.
726: \end{eqnarray*} If $\epsilon>1/2$, then $S(x)v \neq 0$ for all $v \neq 0$, so we have
727: invertibility.
728: 
729: For $\e=1/2$ the claim follows since in the above chain of
730: inequalities at least one is strict if $v\ne0$. In fact, if
731: $J(x)v=\e v$, then $v_i=0$ for all relaxed sites $i$. We claim
732: that the equality $Q(x)v=v$ is impossible. To see this denote by
733: $\tilde Q$ the minor-matrix formed by the rows and columns of
734: $Q(x)$, corresponding to exited sites, and denote by $\tilde v$ be
735: the respective reduced vector. Then $\tilde Q\tilde v=\tilde v$.
736: 
737: Let $U$ be the set of overcritical sites $k$ with $v_k=\max v_l$
738: (we suppose it is positive, multiplying by $-1$ in the opposite
739: case). Choose a boundary site $k\in U$, i.e. the number of
740: neighbors $l$ to $k$ with $v_l=v_k$ is less than $2d$. Then:
741:  $$
742: v_k=\sum_l\tilde Q_{kl}v_l<v_k\sum_l\tilde Q_{kl}\le v_k.
743:  $$
744: This contradiction yields the result.
745: 
746: Finally consider the last statement about the case $E_c \geq
747: \epsilon/(1-\epsilon)$. It is proved by reducing the matrix
748: $S(x)$. If $x_i \leq E_c$, then column $C_i(S(x))$ equals
749: $(0,\dots,0,1,0,\dots0)^T$, where the 1 is in the $i^\text{th}$
750: position. We can start the decomposition of $\det S(x)$ with
751: column $i$, and hence we see that row $i$ and column $i$ can be
752: removed from $S(x)$ without changing the determinant. We remove
753: all rows and columns that correspond to relaxed sites. If
754: $\rho(x)$ is the number of overcritical sites of $x$, we get a
755: $\rho(x) \times \rho(x)$ matrix $S_{red}(x)$. If site $k$ is
756: overcritical then $J_{kk}(x)=\epsilon$. If $E_c \geq
757: \epsilon/(1-\epsilon)$, then it follows from Lemma \ref{6} that
758: all nearest neighbors of $k$ are relaxed. Hence column $k$ of
759: $Q(x)$ has only zero entries. This shows that
760: $S_{red}(x)=\mbox{diag}(\epsilon,\dots,\epsilon)$. Then
761: \begin{equation*}
762: \det S(x)=\det S_{red}(x)=\epsilon^{\rho(x)}\,,
763: \end{equation*} so $\det L_{ij}=\epsilon^{s_{ij}}$.
764: \end{proof}
765: 
766: 
767: \begin{rk}\po
768: In the original model of Zhang one has $\epsilon=0$ in which case
769: $\det L_{ij}=0$ if $L_{ij}\ne\1$. But it is not true that
770: non-trivial kernels can occur for $\epsilon=0$ only, contrary to
771: what was stated in \cite{BCK}. A simple counter-example is the
772: case $N=2$, $E_c=1/3$ and $\epsilon=1/3$. For $x_1>0$ and
773: $2x_1+3x_2 < 1$ we have:
774: $$
775: F_1 \Big{(} \begin{bmatrix} x_1 \\ x_2 \end{bmatrix} \Big{)}
776: =\frac{1}{9}
777: \begin{bmatrix} 2 & 3 \\ 2 & 3  \end{bmatrix} \begin{bmatrix} x_1+1 \\ x_2
778: \end{bmatrix}\,.
779: $$
780: and so $\det L_{12}=0$.
781: \end{rk}
782: 
783: Having non-degenerate maps in the model is more convenient from
784: the point of view of mathematical tools (though from a physical
785: viewpoint it can make no big difference between degenerate and
786: close-to-degenerate systems). Fortunately, degenerations occur
787: only for a negligible set of parameters.
788: 
789:  \begin{theorem}\po\label{th_invert}
790: The maps $L_{ij}$ are invertible for almost all $(\e,E_c)$. In
791: fact, they are invertible for the parameters complimentary to the
792: set $\Xi\subset[0,1)\times(0,\infty)$, which consists of a finite
793: set of vertical intervals for fixed $d$ and $N$.
794:  \end{theorem}
795: 
796: \begin{proof}
797: Fix an avalanche $A_{ij}$ and let $L_{ij}^\e$ be the corresponding
798: linear maps (we stress dependence on $\e$). These maps are the
799: compositions of elementary matrices $S^\e(x(\tau(x,i))) \dots
800: S^\e(x(1))$, with the factors from (\ref{defS})
801:  $$
802: S^\e(x(t))=\1+(\e-1)(\tfrac{dJ}{d\e}-Q)(x(t))
803:  $$
804: being polynomial in $\e$ and independent of the choice of $x=x(1)
805: \in M_{ij}$. The condition $\det L_{ij}^\e=0$ is equivalent to
806: $\det S^\e(x(t))=0$ for some $t$. Denoting by $\op{Sp}_-(T)$ the
807: negative part of the spectrum of $T$, we get:
808: $\e\in1+\op{Sp}_-(Q-\frac{dJ}{d\e})^{-1}$.
809: 
810: There are only finite number of possibilities for the matrix
811: $S^\e(x)$ (though a countable number for their compositions
812: $L_{ij}^\e$, the length of which grow as $E_c\to0$). Thus we
813: obtain $k=k(d,N)$ different values of $\e$ for which $\det
814: S^\e(x)=0$: $\{\e_a\}_{a=1}^k$. For each $\e_a$ there is the
815: maximal value $E_c^a$ of $E_c$ (finite if $\e_a\ne0$), where the
816: corresponding matrix $S^\e(x)$ can appear in the avalanche. Thus
817: the set of degenerate systems is
818: $\{(\e,E_c)\,|\,\e=\e_a,0<E_c<E_c^a\}$.
819:  \end{proof}
820:  \vskip4pt
821: 
822: By proposition \ref{7} $\Xi$ does not intersect the set
823: $\{\e\ge1/2\}\cup\{E_c\ge\e/(1-\e)\}$.
824: 
825: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
826: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
827: 
828: \section{\hps Removability of singularities and
829: coding}\label{sec_2}
830: 
831: The map $F$ may be considered as a piecewise affine map
832: $F:I\times M \rightarrow I \times M$, where $I=[0,1]$ and $F(t,
833: x)=(Nt \mod 1, F_{[Nt]}(x))$. The map $t \mapsto  Nt \mod 1$ is
834: not conjugated to $\sigma_N^+$ since the points $m/N^k \in I$ do
835: not have unique representations in $\Sigma_N^+$. However the sets
836: $\{m/N^k\}\times M \subset I \times M$ are singularities, and
837: following the standard approach for piecewise affine maps, should
838: be removed.
839: 
840: In some physical systems, like the Belykh family, the
841: singularities propagate, intersecting themselves transversally.
842: The Zhang model is not a general position system in this respect,
843: because singularities $\{m/N^k\}\times M \subset I \times
844: M=\hat{M}$ map into themselves, forming zero angle.
845: 
846: 
847: %%%%%%%%%%%%%%%%%%%%%%%%
848: %%%%%%%%%%%%%%%%%%%%%%%%
849: 
850: \subsection{\hpss Construction of attractors} \label{constofattr}
851: 
852: Define the (spatial) singularity set $S(F)=\cup_{ij} \partial
853: M_{ij}$. Then $U=M \setminus S(F)$  consists of a collection of
854: open connected sets $\mathcal Z=\{Z\}$.  Let $U_0:=U$ and
855: $$
856: U_n:=\bigcup_{i \in \Lambda} F_i ( U_{n-1})\cap U\,.
857: $$
858: We say that $x \in S(F)$ is a non-essential singularity of order
859: $m$ if there exists $\epsilon>0$ and $m>0$ such that
860: $$
861: \text{card}\{Z \in \mathcal Z\,|\, U_n\cap B_\epsilon(x)\cap Z
862: \neq \emptyset\} \leq 1
863: $$
864: for all $n>m$. Denote the set of non-essential singularities of
865: order $m$ by $NES(F;m)$ and let $NES(F):=\cup_{m \geq 0} NES(F;m)$
866: be the set of all non-essential singularities. Define $ES(F)=S(F)
867: \setminus NES(F)$ to be the collection of essential singularities.
868: Observe that there is a natural extension of $F$ to $V_0=U \cup
869: NES(F)$. In the following we let $F$ denote the extended map. As
870: above we define
871: $$
872: V_n:=\bigcup_{i \in \Lambda} F_i ( V_{n-1}) \cap V_0\,.
873: $$
874: Let ${\mathcal X}=\cap_{n \geq 0} V_n$ and $\D=\Sigma_N^+ \times
875: {\mathcal X}$. Clearly $F(\D)=\D$. The set $\Y=\overline{\mathcal
876: X}$ is called the {\em physical} (or spatial) attractor of $F$,
877: and $\mathcal A=\Sigma_N^+ \times \Y=\overline{\D}$ is the
878: extended attractor of $F$.
879: 
880: 
881: \begin{prop}\po \label{8}
882: $F|_\D$ is continuous.
883: \end{prop}
884: 
885: \begin{proof}
886: The set $\D$ intersects non-essential singularities only. Hence we
887: must show that if $x$ is a non-essential singularity in $\D$, then
888: the extension of each $F_i$ to $NES(F)$ is continuous at the point
889: $x$. Choose $m \in \N$ and $\varepsilon>0$ such that
890: $B_\varepsilon(x) \cap U_n$ intersects only one partition element
891: $Z\in\mathcal Z$ for $n> m$ and let $y\in
892: B_{\varepsilon/2}(x)\cap\D$. Then $B_{\varepsilon/2}(y)\subset
893: B_\varepsilon(x)$ and so $B_{\varepsilon/2}(y)\cap U_n$ intersects
894: the same partition element $Z$. So $x$ and $y$ are mapped by the
895: same affine map $F_i|_Z$ for each $i \in \Lambda$. The claim
896: follows.
897: \end{proof}
898: \vspace{5pt}
899: 
900: In general, the map $F$ does not have a continuous extension to
901: $\mathcal A$, but only to $\mathcal A\setminus(\Sigma_N^+\times
902: ES(F))$. Actually, if $x\in ES(F)\cap\Y$ lies on the boundary of
903: several continuity partitions for $F_i$, then there are several
904: extensions of $F$ to $(i,x)$. Thus we can continuously extend $F$
905: to $\mathcal A$ only when the essential singularities do not
906: intersect the attractor (are removable).
907: 
908: 
909: 
910: %%%%%%%%%%%%%%%%%%%%%%%%
911: %%%%%%%%%%%%%%%%%%%%%%%%
912: 
913: \subsection{\hpss Symbolic Coding}
914: 
915: If the singularities can affect the dynamics only for a finite
916: number of iterations, then the dynamics can be well
917: approximated by a topological Markov chain.
918: 
919: \begin{dfn}\po\label{d3}
920: We say that singularities are removable if there exists $m\in \N$
921: such that $S(F)=NES(F,m)$.
922: \end{dfn}
923: 
924: The physically most relevant observables $\phi:\hat{M} \rightarrow \R$ are those
925: that are determined by avalanches. We say that
926: $\phi$ is an {\em avalanche observable} if it is constant on
927: continuity domains $[i]\times M_{ij}$.
928: 
929:  \begin{theorem}\po \label{10}
930: If singularities are removable, then the map $F$ is well-defined
931: and continuous on $\mathcal A$ and there is a topological
932: Markov-chain $(\Sigma_A^+,\sigma_A^+)$ and a continuous
933: semi-conjugancy $g:\mathcal A \rightarrow \Sigma_A^+$ such that
934: for all $\hat{x}, \hat{y} \in {\mathcal A}$ and for all avalanche
935: observables $\phi$ we have:
936:  $$
937: g(\hat{x})=g(\hat{y}) \Rightarrow \phi(F^n(\hat x))=\phi(F^n(\hat
938: y))\,\, \forall n\geq 0\,.
939:  $$
940: The Markov-chain is determined by a
941: matrix $A$ which has a maximal eigenvalue equal to $N$.
942:  \end{theorem}
943: 
944:  \begin{rk} \po
945: It is clear that all properties related to distribution of
946: avalanche size, duration, area, etc. are invariant under a
947: semi-conjugancy such as this. Observe that for each avalanche
948: observable $\phi$ on $\mathcal A$, there is a unique observable
949: $\phi':\Sigma_A^+ \rightarrow \R$ such that $\phi=\phi' \circ g$.
950: Suppose we have a measure $\mu$ on $\mathcal A$, and let
951: $\nu=g_*\mu$. If $\phi$ is an avalanche observable on $\mathcal
952: A$, then the statistical properties of $\phi$ with respect to
953: $\mu$ are equivalent to the statistical properties of $\phi'$ with
954: respect to $\nu$. In this $(\sigma_A^+, \Sigma_A^+)$ is a good
955: approximation to $F|_{\mathcal A}$. The coding gives estimates on
956: entropy and growth of periodic points, but these estimates are
957: asymptotically no better than what we get from the trivial
958: semi-conjugancy $\hat{M} \rightarrow \Sigma_N^+$.
959:  \end{rk}
960: 
961: \begin{proof}
962: Singularities are removable so there exists an integer $m \in \N$
963: such that $\pi_s \circ F^m(\hat{M})$ only intersects trivial
964: singularities. Let $X_1,\dots, X_s$ be the closure of the
965: connected components of  $\pi_s \circ F^m(\hat{M})$. $F$ is well
966: defined and continuous on these components. Let $Y_1,\dots,Y_s$ be
967: the intersections of the components  $X_1,\dots,X_s$ with $\Y$. We
968: construct the partition ${\mathcal R}=\{[i] \times Y_k\}$ and
969: enumerate it so that ${\mathcal R}=\{R_1,\dots,R_r\}$, where
970: $r=Ns$.
971: 
972: Let $A=\|a_{ij}\|$ be the $r\times r$ matrix defined by the rule:
973: $a_{ij}=1$ if $F(R_i) \cap R_j \neq \emptyset$, and $a_{ij}=0$
974: otherwise. A sequence $R_{\omega_0}R_{\omega_1}\dots$ is legal if
975: $a_{\omega_{t-1} \omega_{t}}=1$ for all $t \in \N$. Define
976: $g:{\mathcal A} \rightarrow \Sigma_A^+$ by $g(\hat{x})= (\omega_0
977: \omega_1\dots \omega_t \dots)$, where $F^t(\hat{x}) \in
978: R_{\omega_t}$. To prove that $g$ is surjective it suffices to show
979: that for each legal sequence $R_{\omega_0}R_{\omega_1}\dots$,
980: there is a point $\hat{x} \in \pi_s \circ F^m(\hat{M})$ such that
981: $F^i(\hat{x}) \in R_{\omega_t}$ for all $i\in \N$. Note that each
982: $\omega$ can be written as a pair $(t,k)$, where $t\in
983: \{1,\dots,N\}$ and $k \in \{1,\dots,s\}$. Hence we can write
984:  $$
985: \bigcap_{n=0}^\infty F^{-n}( R_{\omega_n})=\bigcap_{n=0}^\infty
986: F^{-n}([t_n]\times Y_{k_n} )=\{{\bf t}\} \times
987: \bigcap_{n=0}^\infty F_{t_0}^{-1}\circ \dots \circ
988: F_{t_{n-1}}^{-1}(Y_{k_n})\,.
989:  $$
990: The continuous image of a connected set is connected, so for each
991: $i=1,\dots,N$ and each $k=1,\dots,s$ there is a unique $l \in
992: \{1,\dots,s\}$ such that $F_i(X_k) \subset X_l$.
993: This implies that we have a nested sequence
994: $$
995: Y_0 \subset F_{t_0}^{-1}(Y_{k_1}) \subset F_{t_0}^{-1}\circ
996: F_{t_1}^{-1}(Y_{k_1}) \subset \dots
997: $$
998: and hence the intersection is non-empty.
999: 
1000: It is clear that $g_{\mathcal R}$ is continuous (see \cite{R} for
1001: details). Since the partition ${\mathcal R}$ is a refinement of
1002: the continuity partition the conjugancy will be injective up to
1003: the classes of points that follow the same continuity domains.
1004: Hence if $\phi(F^n \hat{x}) \neq \phi(F^n \hat{y})$ for some
1005: avalanche observable $\phi$ and some $n \geq 0$, then $g(\hat{x})
1006: \neq g(\hat{y})$.
1007: \end{proof}
1008: 
1009: \noindent
1010: \begin{rk}\po\label{rkk4}
1011: Suppose we modify the Zhang model by using a full shift
1012: $(\Sigma_N, \sigma_N)$ as the excitation factor. It is then
1013: possible that the modified map $F$ is injective on $\Sigma_N
1014: \times \Y$. Since we have strict attraction in the spatial factor
1015: after a fixed number of iterations it is clear that we can then
1016: obtain an injective coding, and hence a topological conjugancy.
1017: However, if we make this modification it is not clear that
1018: $\Sigma_N \times \Y$ equals the set
1019: $$
1020: \Omega=\overline{\bigcap_{n=-\infty}^\infty F^{n}(\Sigma_N \times
1021: M)}\,.
1022: $$
1023: In fact if the maps $F_i|_\Y$ are all injective, then $F|_\Omega$
1024: is invertible, but $F|_{\Sigma_N \times \Y}$ is typically
1025: non-invertible. The reason for this is that, due to contraction, a
1026: point $x \in \Y$ does not have preimages for all the maps $F_i$
1027: and $F_i$ are invertible only on $F_i(\Y)\subset\Y$. So to obtain
1028: invertibility we must turn to the attractor $\Omega$. From a
1029: physical point of view the spatial attractor is of the great
1030: interest, so it is desirable to have an attractor which is a
1031: Cartesian product of the Bernoulli shift and the spatial attractor
1032: $\Y$.
1033: \end{rk}
1034: 
1035: We can always construct a coding of $F|_{\D}$ (even in
1036: non-removable case) by choosing a partition ${\mathcal
1037: R}=\{R_1,\dots R_r\}$, and taking $g_{\mathcal R}:\D \rightarrow
1038: \{1,\dots,r\}^\N$ to be the map sending a point $\hat{x} \in \D$
1039: to the unique sequence $\omega \in \{1,\dots r\}^\N$ such that
1040: $F^t(\hat{x}) \in R_{\omega_t}$ for all $t \geq 0$. But there is
1041: no reason, however, to expect $g_{\mathcal R}(\D)$ to be a
1042: topological Markov chain, cf. \cite{BCK}.
1043: 
1044: %%%%%%%%%%%%%%%%%%%%%%%%
1045: %%%%%%%%%%%%%%%%%%%%%%%%
1046: 
1047: \section{\hps Metric properties}\label{sec_3}
1048: 
1049: The natural volume on $\hat{M}$ is given by the product measure of
1050: the uniform Bernoulli measure on $\Sigma_N^+$ and the Lebesgue
1051: measure on $M$. By iterating this measure (and averaging) we can
1052: construct SRB-measures. However it can happen that the measures
1053: constructed are supported on essential singularities, where it is
1054: not possible to define the dynamics in such a way that the measure
1055: is invariant. Hence we must give some conditions to ensure the
1056: existence of SRB-measures. If there is an SRB-measure it is
1057: characterized by the fact that its projection to $\Sigma_N^+$
1058: coincides with the uniform Bernoulli measure. From this it follows
1059: that any SRB-measure is a measure of maximal entropy. In
1060: situations where the system allows symbolic coding the SRB-measure
1061: corresponds to the Perry measure on the topological Markov chain
1062: $\Sigma_A^+$.
1063: 
1064: %%%%%%%%%%%%%%%%%%%%%%%%
1065: %%%%%%%%%%%%%%%%%%%%%%%%
1066: 
1067: \subsection{\hpss Existence and characterization of SRB-measures}
1068: 
1069: Let $m=m^u \times m^s$, where $m^u=\mu_{\text{{\tiny Ber}}}$ is
1070: the uniform Bernoulli measure on $\Sigma_N^+$, and
1071: $m^s=\mu_{\text{{\tiny Leb}}}$ is the Lebesgue measure on $M$. We
1072: say that an invariant Borel probability measure $\mu$ on $\hat{M}$
1073: has the SRB-property if there exists a measurable invariant set $G
1074: \subset \hat{M}$ such that
1075:  \begin{enumerate}
1076: \item $m(G)>0$
1077: \item $m^u(\pi_u(G))=1$
1078: \item All points $\hat{x} \in G$ are future generic with respect to $\mu$, i.e.
1079: $$\frac{1}{n}
1080: \sum_{t=0}^{n-1} \phi(F^t \hat x) \rightarrow \int \phi
1081: \,d\mu\,,$$ for all $\hat{x} \in G$ and all continuous functions
1082: $\phi:\hat{M}\to\R$.
1083:  \end{enumerate}
1084: 
1085: For Axiom A attractors one can ensure the existence of measures
1086: for which the set of generic points has full Lebesgue measure and
1087: this is equivalent to saying that the canonical family of
1088: conditional measures on the unstable manifolds are absolutely
1089: continuous with respect to the Lebesgue measure. For
1090: non-invertible maps one can in general only expect the set of
1091: generic points to have positive measure and hence it is
1092: unreasonable to require that $m(G)=1$. Condition 2 is (for
1093: physical reasons) important in the Zhang model. It means that the
1094: statistical properties do not depend on the choice of a generic
1095: sequence ${\bf t}$ of excitations. (This is always implicitly
1096: assumed in the numerical investigations of the Zhang model that
1097: can be found in the physical literature.) Moreover, condition 2
1098: will be satisfied for the SRB-measures that can be constructed by
1099: iterating the measure $m$.
1100: 
1101: By a standard approach we can give conditions for existence of
1102: SRB-measures that hold if singularities are removable, but it is
1103: not known if these conditions hold in all non-removable
1104: situations.
1105: 
1106:  \begin{prop}\po \label{11}
1107: Let $(\e,E_c)$ does not belong to the negligible set $\Xi$ of
1108: Theorem \ref{th_invert}. If there exists $n\geq 0$, $C>0$ and
1109: $q>0$ such that
1110:  $$
1111: \forall \delta>0, \forall t \geq 0:\, m \Big{(} F^{-t}\big{(}
1112: \Sigma_N^+ \times U_\delta( ES(F;n) \big{)} \Big{)} \leq C
1113: \delta^q\,,
1114:  $$
1115: then there exists a set $\D \subset \hat{M}$ (constructed in \S
1116: \ref{constofattr}), which may intersect singularities, and a
1117: natural extension of $F$ to $\D$ such that $F(\D)=\D$. Moreover
1118: the set $\D$ carries an $F$-invariant Borel probability measure
1119: with the SRB-property.
1120:  \end{prop}
1121: 
1122: 
1123:  \begin{rk}\po
1124: Proposition \ref{11} is a simple modification of the result of
1125: Schmeling and Troubetzkoy \cite{ST}. In their paper the conditions
1126: for existence are in general too restrictive for the Zhang model.
1127: In fact, in Example A of \S \ref{examples} we show a situation
1128: where the SRB-measure constructed in \cite{ST} does not exist, but
1129: we clearly have existence of a physically relevant measure. The
1130: reason for this paradox is that one in general remove all singular
1131: points on the construction of the attractor, even if there is a
1132: natural extension of $F$ to the points of singularity.
1133: (Proposition \ref{11} obviously applies to this example since
1134: $ES(F;5)=\emptyset$.)
1135:  \end{rk}
1136: 
1137: \noindent
1138: \begin{proof}
1139: In \cite{ST} it is shown that a piecewise smooth map $f$ with
1140: singularity set $S$ has a measure, not supported on singularities,
1141: such that the set of generic points has positive Lebesgue measure.
1142: They require that the following conditions are satisfied:
1143: \begin{enumerate}
1144: \item The restrictions of $f$ to each of its continuity domains
1145: are diffeomorphisms onto their image.
1146: 
1147: \item The second differentials $D^2f_x$ does not grow too fast close
1148: to singularities. (See \cite{ST} for a more precise formulation.)
1149: 
1150: \item $f$ is hyperbolic. In this context this means that there are
1151: constants $C>0$ and $\lambda \in (0,1)$ such that for all  $x \not
1152: \in S$ there is a splitting of the tangent space at $x$ into
1153: subspaces $E^+(x)$ and $E^-(x)$. There are cones $C^+(x)$ and
1154: $C^-(x)$ around $E^+(x)$ and $E^-(x)$ that are invariant under
1155: $Df_x$ and $Df_x^{-1}$ respectively. The angles between the
1156: $C^+(x)$ and $C^-(x)$ are bounded away from zero, and for all
1157: points $x$ that do not intersect singularities in the first $n$
1158: iterations it holds:
1159: $$\|D_xf^n(v)\| \geq C^{-1}\lambda^{-n}\|v\|\,\,\text{for}\,\,\,
1160: v \in C^+(x)\,,$$ and
1161: $$\|D_xf^n(v)\| \leq C\lambda^{n}\|v\|\,\,\text{for}\,\,\,
1162: v \in C^-(x)\,.$$
1163: \item There exists $C>0$ and $q>0$ such that
1164: $m(f^{-t}(U_\varepsilon(S))) \leq C \varepsilon^q$ for all
1165: $\varepsilon>0$ and all $t\in \N$.
1166: \end{enumerate}
1167: We apply this result to the map $F|_{\Sigma_N^+ \times (U_n
1168: \setminus ES(F;n))}$. The singularity set for this map is
1169: contained in $\Sigma_N^+ \times ES(F)$, so by assumption condition
1170: 4 is satisfied. Condition 1 follows from Proposition \ref{7},
1171: condition two is obviously satisfied since $F$ is piecewise affine
1172: and condition 3 follows from Theorem \ref{5} with
1173: $E^+=\R^1\oplus0$, $E^-=0\oplus\R^N$ and $C^\pm$ being the regular
1174: cones around them (actually Theorem \ref{5} ensures hyperbolicity
1175: for some iterate $F^T$, which implies the claim).
1176: 
1177: In \cite{ST} the measures are constructed by iterating $m$,
1178: averaging and taking a weak limit. It is clear that, in the Zhang
1179: model, any measure obtained in this way will satisfy condition 2
1180: in our definition of an SRB-measure.
1181: \end{proof}
1182: \\
1183: 
1184: If an SRB-measure exists it can be characterized by a number of
1185: different properties. From a physical perspective it is reasonable
1186: to require that a relevant invariant measure should preserve the
1187: uniform  Bernoulli structure on $\Sigma_N^+$. This corresponds to
1188: the Lebesgue measure on $[0,1]$ in the alternative formulation of
1189: the map $F$, and hence to absolutely continuous measure
1190: conditional measures on the unstable space $[0,1]$.
1191: 
1192:  \begin{prop}\po \label{12}
1193: If $\mu$ is an SRB-measure on $\D$, then $\mu^u:=(\pi_u)_* \mu$ is
1194: the uniform Bernoulli measure on $\Sigma_N^+$.
1195:  \end{prop}
1196: 
1197: \noindent
1198: \begin{proof}
1199: There is a set $A=\pi_u(G)$ of full $m^u$-measure, such that all
1200: ${\bf t} \in  A$ are generic with respect to $\mu^u$. Take a
1201: continuous function $\phi:\Sigma_N^+ \rightarrow \mathbb{R}$. Then
1202:  $$
1203: \int \phi\,d\mu^u= \lim_{n \rightarrow \infty} \frac{1}{n}
1204: \sum_{t=0}^{n-1} \phi((\sigma_\Lambda^+)^t {\bf t})=\int
1205: \phi\,dm^u \,,
1206:  $$
1207: where the left equality holds for ${\bf t}\in A$ and the right one
1208: for ${\bf t}\in B$ with $B\subset\Sigma_N^+$ a subset of full
1209: $m^u$-measure (from Birkhoff ergodic theorem). Since $A\cap
1210: B\ne\emptyset$, we get: $\int \phi \,d\mu^u=\int \phi\,dm^u$ for
1211: all continuous functions $\phi$.
1212: \end{proof}
1213: 
1214: 
1215: %%%%%%%%%%%%%%%%%%%%%%%%
1216: %%%%%%%%%%%%%%%%%%%%%%%%
1217: 
1218: 
1219: \subsection{\hpss Measures of maximal entropy} \label{meaofmaxentr}
1220: 
1221: Suppose that there exists an invariant Borel probability measure
1222: $\mu$ on $\hat{M}$. Let $\mu^u:=(\pi_u)_* \mu$ and let $\{
1223: \nu_{\bf t}\}$ to be the canonical family of conditional measures
1224: on the fibers $\pi_u^{-1}(\{\bf t \})$. By the Abramov-Rokhlin
1225: formula
1226: $$
1227: h_\mu (F)=h_{\mu^u}(\sigma_N^+)+h_\mu(F|\sigma_N^+) \,,
1228: $$
1229: where
1230: $$
1231: h_\mu(F|\sigma_N^+;\mathcal Q) =  \lim_{n \rightarrow \infty}
1232: \frac{1}{n} \int H_{\nu_{{\bf t}}} \Big{(} \bigvee_{k=0}^{n-1}
1233: (F_{t_{k-1}} \circ\dots\circ F_{t_0})^{-1}({\mathcal Q}) \Big{)}
1234: \, d\mu^u({\bf t})\,,
1235: $$
1236: for a partition $\mathcal Q$ and
1237: $$
1238: h_\mu(F|\sigma_N^+)=\sup_{\mathcal Q} h_\mu(F|\sigma_N^+;\mathcal
1239: Q)\,,
1240: $$
1241: The supremum is taken over all finite measurable partitions
1242: ${\mathcal Q}$ of $\hat{M}$. This formula was originally proved
1243: for product measures by Abramov and Rokhlin \cite{AR} and extended
1244: to arbitrary skew products by Bogenschultz and Crauel \cite{BC}.
1245: 
1246: Below we use the notation $F_{\bf t}$ for the dynamics over a
1247: pre-fixed sequence ${\bf t}=(t_0t_1\dots)\in\Sigma_N^+$. By
1248: $n$-the iteration we mean the map $F^n_{\bf
1249: t}=F_{t_{n-1}}\circ\dots\circ F_{t_0}$.
1250: 
1251:  \begin{theorem}\po \label{17}
1252: If $\mu$ is an invariant Borel probability measure on $\hat{M}$,
1253: then $h_\mu(F) = h_{\mu^u}(\sigma_N^+)$.
1254:  \end{theorem}
1255: 
1256: \noindent
1257: \begin{proof}
1258: We prove the proposition by estimating $h_\mu(F|\sigma_N^+)$ from
1259: above. Let ${\mathcal Q}$ be a partition of $M$ and
1260:  $$
1261: \mathcal Q_{t_0\dots t_{n-1}}:=\bigvee_{k=0}^{n-1} (F_{t_{k-1}}
1262: \circ\dots\circ F_{t_0})^{-1}({\mathcal Q})=\bigvee_{k=0}^{n-1}
1263: F_{\bf t}^{-k}({\mathcal Q})\,.
1264:  $$
1265: Fix $\varepsilon>0$ and choose the partition $\mathcal Q$ such
1266: that $h_\mu(F|\sigma_N^+;\mathcal Q) +\varepsilon\geq
1267: h_\mu(F|\sigma_N^+)$. The maps $F_i$ are non-expanding so it
1268: follows from the Ruelle-Margulis inequality \cite{KH} that
1269:  $$
1270: \frac1n H_{\nu_{\bf t}} \Big{(}\bigvee_{k=0}^{n-1} F_{\bf
1271: t}^{-k}({\mathcal Q})\Big{)}\longrightarrow h_{\nu_{\bf t}}(F_{\bf
1272: t})=0.
1273:  $$
1274: Moreover the convergence is $\mu^u$-uniform and so the same holds
1275: for the integrals. Another way to see it is via the multiplicity
1276: notion of \S\ref{evalofentr} (then ${\mathcal Q}$ should be
1277: subordinate to each continuity partition
1278: $\{M_{ij}\,|\,j=1,\dots,q_i\}$):
1279:  $$
1280: h_\mu(F|\sigma_N^+;\mathcal Q)\le
1281: \lim_{n\to\infty}\frac1n\log\max_{|{\bf t}|\leq n}
1282: \op{mult}(\mathcal Q_{t_0\dots t_{n-1}}\cap\op{supp}(\nu_{\bf
1283: t})).
1284:  $$
1285: Therefore $h_\mu(F|\sigma_N^+)\le \e$. Let $\e\to0$.
1286: \end{proof} \\
1287: 
1288:  \begin{rk}\po
1289: Theorem \ref{17} is a partial case of Theorem 4 from \cite{KR2}.
1290:  \end{rk}
1291: 
1292: We say that an invariant measure $\mu$ is maximal if
1293: $h_\mu(F)=\sup_\nu h_\nu(F)$, where the supremum is taken over all
1294: invariant Borel probability measures on $M$. It follows from
1295: Theorem \ref{17} that $h_\mu(F)\le\log N$. So if $\h(F)>\log N$,
1296: the variational principle fails (this can happen for piece-wise
1297: affine systems, see \cite{KR1,KR2}). But we show in
1298: \S\ref{evalofentr} that the abnormal growth of $\h(F)$ does not
1299: occur in the Zhang model, at least for generic values of
1300: parameters $E_c,\e$.
1301: 
1302:  \begin{cor}\po\label{corr1}
1303: Any SRB-measure on $\D$ has entropy $h_\mu(F)=\log N$ and is hence
1304: a maximal measure.
1305:  \end{cor}
1306: 
1307:  \begin{cor}\po
1308: Suppose singularities are removable and that $\mu$ is an
1309: SRB-measure on ${\mathcal A}$. Let $g:{\mathcal A} \rightarrow
1310: \Sigma_A^+$ be the semi-conjugacy constructed in the proof of
1311: Theorem \ref{10}. If $(\sigma_A^+,\Sigma_A^+)$ is topologically
1312: transitive, then $g_*\mu$ is the Perry measure on $\Sigma_A^+$.
1313:  \end{cor}
1314: \begin{proof}
1315: A transitive topological Markov chain has a unique measure of
1316: maximal entropy. This measure is called the Perry measure
1317: \cite{KH}.
1318: \end{proof}
1319: 
1320: 
1321: 
1322: %%%%%%%%%%%%%%%%%%%%%%%%
1323: %%%%%%%%%%%%%%%%%%%%%%%%
1324: 
1325: \subsection{\hpss Hyperbolic structure} \label{hypstr}
1326: 
1327: There are several ways to define Lyapunov exponents for the Zhang
1328: model. The Zhang model can be represented as a piecewise affine
1329: map, where Bernoulli shift is represented as the expanding map $t
1330: \mapsto Nt \mod 1$ of the interval (see \S \ref{evalofentr}).
1331: Hence it is clear that there is one positive Lyapunov exponent
1332: $\chi^+_0=\log N$. We define the other exponents by introducing
1333: the co-cycle $\T : \hat{M} \rightarrow GL(N,\R)$, defined by
1334: $\T(\hat{x})=L_{ij}$, where $\hat{x} \in [i] \times M_{ij}$. For
1335: $\hat {x} \in \hat{M}$ and $v \in \R^N \setminus \{0\}$ we define
1336: $$
1337: \chi(\hat x,v)= \mathop{\overline\lim}\limits_{n \rightarrow
1338: \infty} \frac1n\log \frac{\| \T( F^{n-1}(\hat{x}) ) \dots \T(
1339: F(\hat{x}) )  \T( \hat{x} ) v\|} { \|v\| }\,.
1340: $$
1341: It is a general fact that the function $\chi(\hat{x},\cdot)$ takes
1342: at most $N$ different values $\chi^-_1(\hat{x})\ge\dots \geq
1343: \chi^-_N(\hat{x})$.
1344: 
1345:  \begin{prop}\po \label{hyp}
1346: For all $\hat{x} \in \hat{M}$ the Lyapunov spectrum is:
1347:  $$
1348: \chi^+_0=\log N>0>\chi^-_1(\hat{x})\ge \dots \geq \chi^-_N(\hat{x})
1349:  $$
1350: and for $(\e,E_c)$ outside the negligible set $\Xi$ from Theorem
1351: \ref{th_invert}:  $\chi^-_N(\hat{x})>-\infty$.
1352:  \end{prop}
1353: 
1354:  \begin{proof}
1355: From Theorem \ref{5} we know that there exists $T \in \N$ and $c
1356: \in (0,1)$ such that
1357: $$
1358: \|\T(F^{T-1}(\hat{x})) \dots \T(F(\hat{x}))\T(\hat{x})\|\leq c
1359: $$
1360: for all $\hat{x} \in \hat{M}$. It immediately follows
1361: that $\chi(\hat{x},v) \leq T^{-1} \log c <0$.
1362: 
1363: For $\epsilon\ge1/2$ and arbitrary $E_c$ or for $\e>0$ and
1364: $E_c\ge(1+\e)/(1-\e)$ all linear maps are invertible, and so for
1365: all $\hat{x} \in \hat{M}$ and all $v \in \R^N \setminus \{0\}$ we
1366: have $\chi(\hat{x},v)\geq \log k$, where $k=\min_{ij} \min
1367: \text{Sp}(L_{ij})>0$.
1368:  \end{proof} \\
1369: 
1370: If there exists a unique SRB-measure, then it follows from the
1371: Osceledec theorem that there are numbers $\chi^-_1, \dots
1372: ,\chi^-_N$ such that $\chi_i(\hat{x})=\chi_i^-$ for Lebesgue
1373: almost every $\hat{x} \in \hat{M}$. The numbers
1374: $\chi_0^+,\chi_1^-,\dots, \chi^-_N$ are the Lyapunov exponents of
1375: the Zhang model. From Proposition \ref{hyp} it follows that the
1376: Zhang model is hyperbolic in the sense that the Lyapunov spectrum
1377: consists of:
1378: $$
1379: \chi^+_0 =\log N > 0 > \chi^-_1 \ge \dots \geq
1380: \chi^-_N>-\infty\,.
1381: $$
1382: We see from Corollary \ref{corr1} that the Pesin formula
1383: $h_\mu(F)=\chi^+$ holds for any SRB-measure. If there is no
1384: SRB-measure then the Lyapunov spectrum should be defined as
1385: functions on $M$:
1386: $$
1387: \chi_i(x)=\int_{\Sigma_N^+}\chi_i({\bf t},x) \,d \mu_{\text{{\tiny Ber}}}\,,
1388: $$
1389: where $\mu_{\text{{\tiny Ber}}}$ is the uniform Bernoulli measure on $\Sigma_N^+$.
1390: 
1391: 
1392: %%%%%%%%%%%%%%%%%%%%%%%%
1393: %%%%%%%%%%%%%%%%%%%%%%%%
1394: 
1395: \subsection{\hpss Entropy of physical vs. mathematical
1396: models}\label{305}
1397: 
1398: We can reformulate the Zhang system as the map $\hat f:\hat
1399: B\to\hat B$, where $\hat B=\cup_{i\ge0}\hat f^i(\hat M)$ is a
1400: compact $\hat f$-invariant subset of $\Sigma_N^+\times\R^N$. We
1401: wish to compare this to the induced transformation $F:\hat
1402: M\to\hat M$ (cf. Remark \ref{ps.vs.mt}).
1403: 
1404: If $\mu$ is an $F$-invariant Borel probability measure on $\hat
1405: M$, then there is an associated $\hat f$-invariant Borel
1406: probability measure $\hat\mu$ on $\hat B$ (and vice versa).
1407: Abramov's theorem (\cite{Br}) relates the entropies of both
1408: systems:
1409:  \begin{equation}\label{umum}
1410: h_{\hat\mu}(\hat f)=h_\mu(F)\cdot\hat\mu(\hat M).
1411:  \end{equation}
1412: One does not need to assume ergodicity and can allow degenerations
1413: \cite{DGS}, as happens for the case of Zhang model. In ergodic
1414: situation by the recurrence theorem of Kac \cite{Br} for a
1415: $\mu$-generic point $\hat x\in\hat M$:
1416:  \begin{equation}\label{mumu}
1417: \frac1{\hat\mu(\hat M)}=\lim_{n\to\infty}\frac1n\sum_{k=0}^{n-1}
1418: \tau(F^k\hat x),
1419:  \end{equation}
1420: where $\tau(\hat x)=\tau(i,x)$ is the avalanche time initiated by
1421: addition of $e_i$ to $x\in M$ (see \S\ref{randex}). If
1422: $\mu=\mu_\text{SRB}$ is a unique SRB-measure, the above point
1423: $\hat x$ can be chosen Lebesgue generic. The resulting limit is
1424: the average avalanche time $\bar\tau$ (we discuss it in more
1425: details in \S\ref{Ssec}-\ref{S-sec}) and we obtain:
1426:  $$
1427: h_{\hat\mu}(\hat f)=h_{\mu}(F)/\bar\tau\quad\text{ resp. }\quad
1428: h_{\hat\mu_\text{SRB}}(\hat f)=h_{\mu^u}(\sigma_N^+)/\bar\tau.
1429:  $$
1430: 
1431: In general non-ergodic situation to get equality (\ref{mumu}) we
1432: should integrate the terms in right-hand side and then we again
1433: obtain the average avalanche size $\langle\tau\rangle$, but now it
1434: is the space-average. Substituting this into (\ref{umum}) we get:
1435:  \begin{equation}\label{mu-tau}
1436: h_{\hat\mu}(\hat f)=h_\mu(F)/\langle\tau\rangle.
1437:  \end{equation}
1438: 
1439: For SRB-measures this formula is indicated by the Ledrappier-Young
1440: theorem \cite{LY}, because we have only one positive Lyapunov
1441: exponent.
1442: 
1443: 
1444: %%%%%%%%%%%%%%%%%%%%%%%%
1445: %%%%%%%%%%%%%%%%%%%%%%%%
1446: 
1447: \section{\hps Topological entropy}\label{sec_4}
1448: 
1449: To calculate the topological entropy of $F$ we established in
1450: \cite{KR2} a set of inequalities, using the technique developed by
1451: J. Buzzi \cite{B1}, \cite{B2} for piecewise expanding maps and
1452: piecewise isometries, see Appendix \ref{app_A}. The contraction in
1453: the maps $F_{ij}$ provides difficulties, so several results were
1454: generalized to fit the framework of the Zhang model. It is not
1455: however true (as was widely believed, see  \cite{B2}) that the
1456: contraction does not contribute to topological (contrary to
1457: metric) entropy, the corresponding counter-example can be found in
1458: \cite{KR1}. The Zhang model has a feature common to all such
1459: examples \cite{KR2}, namely angular expansion, but still for most
1460: values of the parameters this abnormal increase of the entropy
1461: does not occur.
1462: 
1463: %%%%%%%%%%%%%%%%%%%%%%%%
1464: %%%%%%%%%%%%%%%%%%%%%%%%
1465: 
1466: \subsection{\hpss Growth of the number of continuity domains}
1467: 
1468: Let $\mathcal P=\{[i] \times M_{ij}\}$ be the partition of
1469: continuity for $F$, and enumerate the elements so that $\mathcal
1470: P=\{P_1,\dots P_r\}$. Let
1471:  $$
1472: [P_{a_0}\dots P_{a_{n-1}}]:=\bigcap_{m=0}^{n-1} F^{-m}(P_{a_m})\,,
1473:  $$
1474: and
1475:  $$
1476: \mathcal P^n=\{[P_{a_0}\dots P_{a_{n-1}}] \neq \emptyset\,|\,
1477: a_m=1,\dots,r\}\,.
1478:  $$
1479: We define the singularity entropy of $F$ by
1480:  $$
1481: \hs(F)=\lim_{n \rightarrow \infty} \frac{1}{n} \log
1482: \text{card}(\mathcal P^n)\,.
1483:  $$
1484:  \begin{rk}\po\label{rk9}
1485: Define a map $g:{\mathcal A} \rightarrow \Sigma_r^+$ by letting
1486: $g(\hat{x})$ be the unique sequence $a_0a_1\dots$ such that
1487: $F^n(\hat{x}) \in P_{a_n}$. Then
1488:  $$
1489: \hs(F)=\h(\sigma_r^+;g({\mathcal A})) \,.
1490:  $$
1491: For piecewise affine expanding maps and piecewise isometries it is
1492: clear that this also equals the topological entropy, but due to
1493: the contraction, this is not obvious in the Zhang model. In
1494: addition, if singularities are not removable, the map $g$ has
1495: discontinuities.
1496:  \end{rk}
1497: 
1498:  \begin{dfn}\po \label{d4}
1499: Call a point $x\in S(F)$ an unstable singularity if for all $i$
1500: and $k\ne l$ we have: $\lim_{y\to x}F_{ik}(y)\ne \lim_{y\to
1501: x}F_{il}(y)$.
1502:  \end{dfn}
1503: 
1504:  \begin{theorem}\po \label{ssthm}
1505: If all singularities $S(F)\cap {\mathcal Y}$ are unstable, then
1506: $\h(F)=\hs(F)$.
1507:  \end{theorem}
1508: 
1509: We need the following technical lemma:
1510:  \begin{lem} \po \label{sslemma}
1511: If the singularities in $\Y$ are unstable, then there exists a
1512: constant $\gamma>0$ such that for all $\delta>0$, $x\in\Y\cap
1513: M_{ik}$ and $y\in\Y\cap M_{il}$, $k \neq l$, we have:
1514:  $$
1515: d(x,y) < \delta \Rightarrow d(F_{i}(x),F_{i}(y))>\delta\,.
1516:  $$
1517:  \end{lem}
1518: 
1519: \noindent
1520: \begin{proof}
1521: Suppose that for all $\delta>0$ there exists $x\in M_{ik} \cap \Y$
1522: and $y \in M_{il} \cap \Y$ such that $d(x,y)<\delta$ and
1523: $d(F_{ik}(x),F_{il}(y))\leq \delta$. There exist sequences
1524: $\{x_m\} \subset M_{ik}$ and $\{y_m\} \subset M_{il}$ such that
1525: $d(x_m,y_m) \rightarrow 0$ and $d(F_{ik}(x_m),F_{il}(y_m))
1526: \rightarrow 0$. The sequence $\{x_m\}$ has a convergent
1527: subsequence $x_{m_n} \rightarrow z$. The point $z$ lies in $S(F)$
1528: and $y_{m_n} \rightarrow z$. By the continuity of the maps
1529: $F_{ik}$ and $F_{il}$ we have $F_{ik}(x_{m_n}) \rightarrow
1530: F_{ik}(z)$ and $F_{il}(y_{m_n}) \rightarrow F_{il}(z)$. Since the
1531: metric $d$ is continuous on $M \times M$ we have
1532:  $$
1533: d(F_{ik}(z), F_{il}(z))=
1534: \lim_{n \rightarrow \infty}
1535: d(F_{ik}(x_{m_n}),F_{il}(y_{m_n}))=0\,.
1536:  $$
1537: Hence $F_{ik}(z)=F_{il}(z)$. Contradiction.
1538: \end{proof} \\
1539: 
1540:  \noindent {\bf Proof of Theorem \ref{ssthm}.}
1541: Let
1542:  $$
1543: [P_{a_0}\dots P_{a_{n-1}}]=[i_0\dots i_{n-1}]\times K\,,
1544:  $$
1545: where $K \subset M$ is a convex polygon. Fix $\delta>0$ and set
1546: $k=[\log1/\delta]$. Let $z_1,\dots,z_{m(\delta)}$ be a
1547: $\delta$-spanning set for $K$. Chose ${\bf t} \in \Sigma_N^+$ and
1548: define $N^k$ sequences
1549:  $$
1550: {\bf s}_{r_0 \dots r_{k-1}}= (i_0 \dots i_{n-1} r_0 \dots r_{k-1}
1551: t_{n+k} t_{n+k-1} \dots ) \in \Sigma_N^+\,.
1552:  $$
1553: Since the maps $F_{ij}$ are contracting it is clear that the set
1554: $\{\hat{x}=({\bf s}_{r_0 \dots r_{k-1}},z_l)\}$ is a
1555: $(n,\delta)$-spanning set for $[P_{a_0}\dots P_{a_{n-1}}]$, and
1556: since the minimum number of balls needed to cover a convex polygon
1557: $K\subset M$ is bounded by $m(\delta) \leq C_N\delta^{-N}$ we see
1558: that the number of $(n,\delta)$-balls to cover $[P_{a_0}\dots
1559: P_{a_{n-1}}]$ is bounded by $m(\delta)N^k\cdot\#\{[P_{a_0}\dots
1560: P_{a_{n-1}}]\}$. Therefore we get an estimate for the number of
1561: $(n,\delta)$-balls to cover $\hat M$ and so
1562:  $$
1563: \h(F)\leq\lim_{n\to\infty}\frac{\log\bigl(
1564: C_\delta\delta^{-N}\op{card}(\mathcal P^n)\bigr)}{n}=\hs(F)\,.
1565:  $$
1566: 
1567: To see the opposite inequality let $A,B \in {\mathcal P}^n$ and
1568: take $\hat{x}_1=({\bf t},x) \in A$ and $\hat{x}_2=({\bf s},y) \in
1569: B$. Suppose that $A \neq B$ and that
1570: $t_0=s_0,\dots,t_{n-1}=s_{n-1}$. Then there is $m<n$ such that
1571: $\pi_s \circ F^m(\hat{x}) \in M_{t_mk}$ and $\pi_s \circ
1572: F^m(\hat{y}) \in M_{t_ml}$, $k \neq l$. By Lemma \ref{sslemma}
1573: there is $\gamma>0$ such that for all $\xi < \gamma$:
1574:  $$
1575: \max\{d(\pi_s \circ F^m(\hat{x}),\pi_s \circ F^m(\hat{y})),
1576: d(\pi_s \circ F^{m+1}(\hat{x}),\pi_s \circ F^{m+1}(\hat{y})) \}
1577: \geq \xi\,.
1578:  $$
1579: Therefore for $\delta$ sufficiently small, no $(n+1,\delta)$-ball
1580: can contain points of both $A$ and $B$ and the minimal
1581: $(n+1,\delta)$-spanning set has at least $\op{card}(\mathcal P^n)$
1582: elements. Then $\h(F) \geq \hs(F)$. \qed
1583: 
1584:  \weg{
1585: \begin{rk}\po
1586: The claim $\h(F)=\hs(F)$ holds in a more general situation, for
1587: instance, when the system is a skew-product $\Sigma_N^+\times
1588: M$, $F({\bf t},x)=(\sigma_N^+{\bf t},f_{t_0}x)$ as in the Zhang
1589: model. The essential feature of this extension is that the
1590: singularities of the unstable factor do not intersect under
1591: backward-iterations.
1592: \end{rk}
1593:  }
1594: 
1595: 
1596:  \begin{theorem}\po \label{nostable}
1597: For the Zhang models: $\h(\hat f)=\hs(\hat f)$, $\h(F)=\hs(F)$.
1598:  \end{theorem}
1599: 
1600: \noindent
1601:  \begin{proof}
1602: By Remark \ref{rk9} all quantities are topological entropies. The
1603: corresponding systems in the second equality are the Poincar\'e
1604: return maps for the transformations of the first equality. The map
1605: $F:\hat M\to\hat M$ is already the return map for $\hat f:\hat
1606: B\to \hat B$ by the very construction.
1607: 
1608: To achieve the same claim for the symbolic system we extend the
1609: partition $\mathcal{P}$ of $\hat M$ to a partition
1610: $\mathcal{\tilde P}$ of $\hat B$, which on
1611: $\Sigma_N^+\times(\R^N_{\ge0}\setminus M)\cap\hat B$ equals the
1612: product of the standard partition $\Sigma_N^+=\cup_i[i]$ and the
1613: partition of the spacial part by the hyperplanes $\{x_i=E_c\}$.
1614: Denote by $s\le r+N(2^N-1)$ the number of elements of the new
1615: partition $\mathcal{\tilde P}$.
1616: 
1617: Let $\mathcal{B}\subset\hat B$ be the $\hat f$-invariant closure
1618: of $\mathcal{A}$ in $\Sigma_N^+\times\R^N_{\ge0}$. Define a map
1619: $\hat g:\mathcal{B}\to\Sigma_s^+$ by letting $\hat g(\hat x)$ be
1620: the unique sequence $b_0b_1\dots$ such that $\hat f^n(\hat
1621: x)\in\mathcal{\tilde P}_{b_n}$. We wish to prove that
1622:  \begin{equation}\label{tten}
1623: \h(\hat f)=\hs(\hat f)=\h(\sigma_s^+|\hat g({\mathcal B})) \,.
1624:  \end{equation}
1625: For this it is sufficient to check that the singularities $S(\hat
1626: f)\cap\pi_s(\mathcal{B})$ are unstable for $\hat f$ (the second
1627: equality follows from the definition).
1628: 
1629: Consider a singular point $x\in\R^N_{\ge0}$. Let $y,z$ tend to $x$
1630: by two different domains of the projected partition
1631: $\mathcal{\tilde P}$ in $\R^N_{\ge0}$. Then we can subdivide
1632: $\{1,\dots,N\}=A\cup B\cup C$, where $y_i=E_c+0$, $z_i=E_c-0$
1633: (with the obvious notations instead of limits) for $i\in A$,
1634: $y_i=E_c-0$, $z_i=E_c+0$ for $i\in B$ and $y_i,z_i$ belong to the
1635: same side of $E_c$ for $i\in C$. Then for $\tilde S=S(y)-S(z)$ we
1636: have:
1637:  $$
1638: \tilde S_{ij}=\left\{
1639:  \begin{array}{ll}
1640: \e-1, & \text{ if }i=j\in A,\\
1641: 1-\e, & \text{ if }i=j\in B,\\
1642: \frac{1-\e}{2d}, & \text{ if }d_\Lambda(i,j)=1, j\in A,\\
1643: \frac{\e-1}{2d}, & \text{ if }d_\Lambda(i,j)=1, j\in B,\\
1644: 0 & \text { otherwise.}
1645:  \end{array} \right.
1646:  $$
1647: We should check that $\tilde S\cdot v\ne0$ for any vector
1648: $v\in\R^N_{\ge0}$ with components $v_i=E_c$ for $i\in A\cup B$
1649: (this implies that $x$ is unstable). Note that other components
1650: $v_i$, $i\in C$ do not contribute to the product.
1651: 
1652: Let $i$ be a site from $A$ with not all neighbors from $A$ (if the
1653: set $A$ is empty consider $B$). Denote the number of $A$-neighbors
1654: of $i$ by $k_A<2d$ and the number of $B$-neighbors by $k_B\ge0$.
1655: Then $(\tilde S\cdot v)_i=E_c(2d-k_A+k_B)\frac{\e-1}{2d}\ne0$.
1656: 
1657: Thus Theorem \ref{ssthm} implies (\ref{tten}). Moreover, the same
1658: reasons yield a more general statement. Namely, since $\hat g$ is
1659: a semi-conjugancy, we get:
1660:  \begin{equation}\label{sten}
1661: \h(\hat f|\mathcal{K})=\h(\sigma_s^+|\hat g({\mathcal{K}}))
1662:  \end{equation}
1663: for any subset $\mathcal{K}\subset \mathcal{B}$ (recall that $\hat
1664: g$ may have discontinuities, but our arguments are not injured by
1665: this fact). This subset needs not to be invariant, and in this
1666: case we should use Bowen's definition of entropy \cite{Bo}. Thus
1667: Katok's entropy formula \cite{K} implies that $h_{\hat\mu}(\hat
1668: f)=h_{\hat g_*\hat\mu}(\sigma_s^+)$ for all $\hat f$-invariant
1669: measures $\hat\mu$, which are not supported on singularities.
1670: 
1671: If the system $(\hat M,F)$ possesses a measure $\mu$ of maximal
1672: entropy, then by the obtained result the second claim of the
1673: theorem follows from (\ref{mu-tau}) and Kac's theorem \cite{PY}.
1674: In general, we can apply the above arguments to the partition of
1675: $\hat M$ by the subsets of equal return times and using the fact,
1676: that both returns of $\hat f$ and $\hat g$ have the same
1677: combinatorics, we get: $\h(F)=\h(\sigma_r^+|g({\mathcal A}))$
1678: (see, for instance, the loop equation approach \cite{Pt}).
1679:  \end{proof}
1680: 
1681: 
1682: 
1683: %%%%%%%%%%%%%%%%%%
1684: %%%%%%%%%%%%%%%%%%
1685: 
1686: \subsection{\hpss Evaluation of topological entropy} \label{evalofentr}
1687: 
1688: It was predicted on the base of variational principle in
1689: \cite{BCK} that topological entropy of the Zhang model is
1690: $\h(F)=\log N$. However this principle does not apply because the
1691: map is not well-defined (continuously) on the whole space (or
1692: thanks to non-compactness if we remove the singularities). In
1693: fact, there can be no invariant measures on the non-singular part
1694: at all.
1695: 
1696: While we support the claim that $\h(F)=\log N$, it will not be
1697: proved in full generality. We start with the asymptotic
1698: statement.
1699: 
1700: Consider the bifurcation diagram on the $(E,\e)$ strip
1701: $(0,+\infty)\times[0,1)$, where a point is critical if in its
1702: neighborhood dynamics of the Zhang model can experience avalanches
1703: of different types. Thus the strip is partitioned into different
1704: avalanche type domains. The partition depends on $L,d$. For $N=2$
1705: the diagram is shown on Figure \ref{bifdiag}.
1706: 
1707: %%
1708: %%
1709: \noindent
1710: \begin{figure}
1711: \begin{center}
1712: \includegraphics[width=5.5cm]{f16.eps}
1713: \caption{{\small Shows the avalanche type domains on
1714: $(\e,E_c)$-bifurcation diagram for $N=2$. The top domain is
1715: $E_c>\frac{1+\e}{1-\e}$. The line from infinity to the origin is
1716: $E_c=\frac{\e}{1-\e}$. We see infinitely many domains with
1717: different avalanches, accumulating to two of the axes.}}
1718: \label{bifdiag}
1719: \end{center}
1720: \end{figure}
1721: %%
1722: 
1723: Note that for all $L,d$ there is the top avalanche type domain
1724: representing the shortest avalanche time. For $N=2$ it is given by
1725: the relation $E>\frac{1+\e}{1-\e}$. Also note that some domains
1726: have $\e$-projection strictly smaller than the interval $[0,1)$.
1727: The next statement concerns only the top domain and the avalanche
1728: type domains that are adjacent to the line $\e=1$.
1729: 
1730:  \begin{theorem}\po \label{140}
1731: For the Zhang model: $0\le\h(F)-\log N\le\theta(E,\epsilon)$,
1732: where $\lim_{E_c\to\infty}\theta(E_c,\epsilon)=0$ ($\e$ fixed) and
1733: $\lim_{\e\to1}\theta(E_c,\epsilon)=0$. In the latter case $E_c$
1734: changes accordingly with $\e$ so that $(E_c,\e)$ belongs to the
1735: same avalanche type domain (then $E_c\to\infty$, though this case
1736: differs from the former).
1737:  \end{theorem}
1738: 
1739: We can remove the $N$-rational points from $\Sigma_N^+$ and then
1740: represent it as the subset of $I=[0,1]$ with points $n/N^k$ being
1741: deleted (this process does not change the topological entropy). In
1742: fact, we need to remove only points $n/N$ for the others will be
1743: deleted by inverse iterations of the map of $I\times X$ with the
1744: formula $F(t,x)=(Nt,f_{[Nt]}(x))$. Thus we represent our system as
1745: a piece-wise affine partially hyperbolic map of the subset of
1746: $\R^{1+N}$. This is important for an application of the results
1747: from Appendix \ref{app_A}.
1748: 
1749: The proof uses the notions of multiplicity and multiplicity
1750: entropy, due to G. Keller and J. Buzzi. Given a finite partition
1751: ${\mathcal Z}=\{Z_k\}$ of $\hat{M}$ we define
1752:  $$
1753: {\mathcal Z}^n=\{[Z_0\dots Z_{n-1}] \neq \emptyset\}\,,
1754:  $$
1755: where
1756:  $$
1757: [Z_0 \dots Z_{n-1}]=Z_0 \cap F^{-1}(Z_1) \cap \dots \cap
1758: F^{1-n}(Z_{n-1})\,.
1759:  $$
1760: Define multiplicity of a partition by
1761:  $$
1762: \text{mult}({\mathcal Z})=\max_{\hat{x} \in \hat{M}}
1763: \text{card}\{Z \in {\mathcal Z}\,|\,\overline{Z} \ni x\}\,,
1764:  $$
1765: If ${\mathcal Z}$ is the continuity partition of the map $F$ we
1766: often denote the multiplicity of the ${\mathcal Z}$ by
1767: $\text{mult}(F)$. Then it is clear that
1768: $\text{mult}(F^n)=\text{mult}({\mathcal Z}^n)$. The multiplicity
1769: entropy of $F$ is (the limit exists by subadditivity, cf.
1770: \cite{KH})
1771: $$
1772: H_{\text{mult}}(F) = \sup_{{\mathcal Z}} \lim_{n \rightarrow
1773: \infty} \frac{1}{n} \log \text{mult}({\mathcal Z}^n)\,,
1774: $$
1775: and we see that if ${\mathcal Z}$ is the continuity partition,
1776: then
1777: $$
1778: H_{\text{mult}}(F) = \lim_{n \rightarrow \infty} \frac{1}{n} \log
1779: \text{mult}(F^n)\,.
1780: $$
1781: 
1782: It is clear that $\hm(F)=0$ if the singularities are removable, so
1783: $\h(F)=\log N$ by the result in Appendix \ref{app_A}. For big
1784: $E_c\gg1$ the singularities are generally non-removable, but still
1785: we have the same effect asymptotically:\\
1786: 
1787: \noindent
1788:  \begin{Proof}{Proof of Theorem \ref{140}}
1789: We take $\theta(E_c,\e)=\hm(F)$. Since the singularities $t=n/N\in
1790: I$ of the map $\sigma_N^+:t\mapsto Nt$ do not intersect in inverse
1791: iterations, the multiplicity growth is only due to the spacial
1792: maps $F_i:M\to M$. Thus using the notation $F_{\bf t}$ from
1793: Section \ref{meaofmaxentr} for the dynamics over a prefixed
1794: sequence of excitations we obtain $\hm(F)=\sup_{{\bf
1795: t}\in\Sigma_N^+}\hm(F_{\bf t})$.
1796: 
1797: To show the first claim let us notice that when $\e=\op{const}$,
1798: but $E_c\to\infty$, then the avalanches map the critical part of
1799: the boundary $\{\exists i:x_i=E_c\}\subset\partial M$ far from
1800: $\partial M$, namely to the distance $\sim\gamma(\e,L,d)E_c$, see
1801: Figures \ref{pic1} for $N=2$ and \ref{pic2} for $N=3$. To reach
1802: again the boundary and experience avalanche we need many shifts
1803: $F_i$ by the basic vectors $e_i$. Thus the singularity can meet
1804: only after big number of iterations. Since the initial picture of
1805: singularities has bounded multiplicity (for $(\e,E_c)$ from the
1806: top avalanche type domain), the multiplicity decreases at least as
1807: $k/E_c$ so that it vanishes in the limit.
1808: 
1809: To prove the second claim we use the inequality
1810: $\hs(F)\le\sum\rho_i(F)$ from Appendix \ref{app_A.3}. If the
1811: avalanche type domain is fixed, the number of compositions of
1812: matrices $S(x)$ in one avalanche (see \S\ref{return maps}) is
1813: bounded. Every such a matrix tends to identity when $\e\to1$. Thus
1814: all the linear parts $L_{ij}$ of avalanche maps $F_{ij}$ tend to
1815: identity and so angular expansions $\rho_i(F)$ tend to zero.
1816:  \qed
1817:  \end{Proof}
1818:  \vskip6pt
1819: 
1820: Notice that if $E_c$ is fixed, but $\e\to1$, then the number of
1821: avalanches has unlimited grow and the previous argument do not
1822: work. However, due to estimates of \S\ref{randex} on the maximal
1823: avalanche length $\tau_m$ we conclude that $\theta(E_c,\e)\to0$ as
1824: either $E_c\to\infty$ or $\e\to1$, both quantities being related
1825: by the constraint $E_c\ge
1826: N(1-\e)^{-\frac12\op{diam}(\Lambda)-\sigma}$ for some $\sigma>0$
1827: (we don't require, but allow $N\to\infty$ as well). This statement
1828: is stronger than in the theorem.
1829: 
1830: Now we are going to prove vanishing of $\theta(E_c,\e)$ a.e. in
1831: the finite part.
1832: 
1833:  \begin{theorem}\po \label{a.s.0}
1834: For generic $(E_c,\e)$ we have: $\h(F)=\log N$.
1835:  \end{theorem}
1836: 
1837: We will prove the statement not only for the Zhang model, but also
1838: for nearly Zhang models. By this we mean the following. The map
1839: $F$ is a bundle over the Bernoulli shifts $\sigma_N^+$ with
1840: factors $F_i$ being piece-wise affine partially contracting maps.
1841: That is $M=\cup_j M_{ij}$ and for $F_{ij}=F_i|_{M_{ij}}$ we have
1842: $F_{ij}(x)=L_{ij}(x)+b_{ij}$, $b_{ij}=L_{ij}e_i$. We are going to
1843: make arbitrary small generic perturbations of the matrices
1844: $L_{ij}$ (still with spectrum within unit ball) and vectors
1845: $b_{ij}$ (one should care that $M_{ij}$ are mapped into $M$) and
1846: prove the statement for this modified system.
1847: 
1848: Due to round-off errors there is no much difference between the
1849: original and the perturbed systems in computer simulations. And
1850: from the point of view of the experiment such perturbations
1851: (instrument instability) are indispensable -- look at \cite{Ru}
1852: for the discussion of physical relevance of variation of
1853: parameters as the noise.
1854: 
1855:  \vskip4pt
1856: \noindent
1857:  \begin{Proof}{Proof of Theorem \ref{a.s.0}}
1858: We claim that for generic $(E_c,\e)$ the singularities do not
1859: multiply. Actually, some intersections of singularities are
1860: deformable as we vary the parameters, just by the transversality
1861: reasons, but the other disappear with small perturbations.
1862: 
1863: Namely, for a multi-index $\sigma=(\alpha_1,\dots,\alpha_k)$,
1864: $\alpha_s=(i_s,j_s)$ coding an orbit, denote
1865: $F_\sigma=F_{\alpha_k}\circ\dots\circ F_{\alpha_1}$ the
1866: corresponding map along the orbit. The singularities of this map
1867: are: $\op{Sing}(F_\z)=\cup_{r=1}^k
1868: F_{\z_{[r]}}^{-1}\op{Sing}(F_{\alpha_r})$, where
1869: $\z_{[r]}=(\alpha_1,\dots,\alpha_r)$.
1870: 
1871: If $z_0\in\op{Sing}(F_\z)$, then its orbit (with
1872: multi-possibilities due to singularities: mapping a singular point
1873: we extend the components of the map $F_\z$ in various ways) meets
1874: several singularity planes, i.e. for some cuts $\z_s=\z_{[r_s]}$
1875: of $\z$, $s=1,\dots,m$, we have the following system:
1876:  \begin{equation}\label{masha}
1877: F_{\sigma_1}(z_0)=z_1,\dots,\,F_{\sigma_m}(z_0)=z_1,\
1878: l_{q_1}(z_1)=0,\dots,\,l_{q_m}(z_m)=0,
1879:  \end{equation}
1880: where $l_{q_s}(z)$ are equations for the singularity hyperplanes
1881: of the corresponding map $F_i$ (there are also inequalities, which
1882: we don't mention). When $m\le N$ there are occasions, when
1883: (\ref{masha}) has a solution continuously depending on $(E_c,\e)$.
1884: However for $m>N$ this is no longer the case. In fact, considering
1885: nearly-Zhang models we see that for generic data the above system
1886: (\ref{masha}) is characterized by a collection of non-trivial
1887: polynomial in $\e$ equations (for each $E_c$).
1888: 
1889: More precisely, the set of $(L_{ij},b_{ij})$ giving trivial
1890: polynomials has positive codimension and hence zero Lebesgue
1891: measure. Uniting these sets over all choices of multi-indices
1892: $(\z_1,\dots,\z_m)$ we see that the complement has full measure
1893: and dimension and so a generic perturbation yields the data
1894: $(L_{ij},b_{ij})$ from it. Since non-trivial polynomials have only
1895: finite number of zeros, then for a generic nearly Zhang model and
1896: every $E_c$ there is a countable subset of $\{0\le\e<1\}$, so that
1897: the corresponding systems (\ref{masha}) have no solutions. This
1898: means that multiplicity of $F_{\bf t}^n$ does not grow with $n$
1899: and so $\hm(F_{\bf t})=0$.
1900: 
1901: Now let us look to the Zhang model. We restrict for simplicity of
1902: exposition to the case of two sites $N=2$. In this case the linear
1903: parts of the affine maps are compositions of $L_1=\1+(\e-1)A_1$
1904: and $L_2=\1+(\e-1)A_2$ with
1905:  $$
1906: A_1=\begin{bmatrix}1 & 0 \\ -1/2 & 0\end{bmatrix},\quad
1907: A_2=J^{-1}A_1J=\begin{bmatrix}0 & -1/2 \\ 0 & 1\end{bmatrix},
1908: \text{ where }J=\begin{bmatrix}0 & 1 \\ -1 & 0\end{bmatrix}
1909:  $$
1910: (we exclude the obvious matrix $\1$, see Example A for details).
1911: Notice that $\det L_1=\det L_2=\e$.
1912: 
1913: Suppose that (\ref{masha}) has a continuous solution in some
1914: domain of $(E_c,\e)$. Then it is algebraic in $\e$ and linear in
1915: $E_c$ (the latter is because the singularity lines within one
1916: avalanche type domain shift with velocity 1 in the direction of
1917: either $e_1=\begin{pmatrix} 1\\0 \end{pmatrix}$ or
1918: $e_2=\begin{pmatrix} 0\\1 \end{pmatrix}$. Differentiating this by
1919: $E_c$ we obtain:
1920:  \begin{equation}\label{medved}
1921: L_{\rho_1}(z_0')=z_1',\dots,L_{\rho_m}(z_0')=z_m',
1922:  \end{equation}
1923: where $L_{\rho_i}=L_{\rho_{i,t_i}}\circ\dots\circ L_{\rho_{i,1}}$
1924: for $\rho_i=(\rho_{i,1},\dots,\rho_{i,t_i})$ is the linear part of
1925: $F_{\z_i}$ ($\rho_i$ is different from $\z_i$ because $dF_{\z_i}$
1926: is a composition of several maps $L_s$). The points $z_k$, $1\le
1927: k\le m$, are constrained to the singularity lines and we can
1928: suppose these are the lines $\{x_1=E_c\}$ or $\{x_2=E_c\}$ (all
1929: other singularities are mapped to them within one avalanche). Thus
1930: $z_k'=v_k+\psi_k w_k$, where $v_k=e_1\text{ or }e_2$ and
1931: $w_i=Jv_i$ is the other basic vector, $\psi_k$ being an unknown
1932: scalar.
1933: 
1934: Let $\xi_k=L_{\rho_k}^{-1}(v_k)$, $\zeta_k=L_{\rho_k}^{-1}(w_k)$;
1935: these vectors depend meromorphically on $\e$. System
1936: (\ref{medved}) is solvable iff the affine lines
1937: $\xi_k+\psi_k\zeta_k$, $1\le k\le m$, in $\R^2$ have a common
1938: point. We can suppose $m=3$.
1939: 
1940: Denote by $\Omega(\xi,\eta)=\langle J\xi,\eta\rangle$ the standard
1941: symplectic form on $\R^2$. The above 3 lines intersect jointly iff
1942:  $$
1943: \Omega(\xi_1,\zeta_1)\Omega(\zeta_2,\zeta_3)+
1944: \Omega(\xi_2,\zeta_2)\Omega(\zeta_3,\zeta_1)+
1945: \Omega(\xi_3,\zeta_3)\Omega(\zeta_1,\zeta_2)=0.
1946:  $$
1947: Dividing by $\prod_{k=1}^3\Omega(\xi_k,\zeta_k)$ and using the
1948: fact that
1949: $\Omega(\xi_k,\zeta_k)=\det(L_{\rho_k}^{-1})=\e^{-|\rho_k|}$ we
1950: get the equivalent equation:
1951:  \begin{equation}\label{yest}
1952: \Omega(\eta_1,\eta_2)+\Omega(\eta_2,\eta_3)+\Omega(\eta_3,\eta_1)=0.
1953:  \end{equation}
1954: Here $\eta_k=\e^{-|\rho_k|}L_{\rho_k}^{-1}w_k=\tilde
1955: L_{\rho_k}w_k$, where $\tilde L_{\tau}=\tilde
1956: L_{\tau_1}\circ\dots\circ\tilde L_{\tau_t}$ for
1957: $\tau=(\tau_1,\dots,\tau_t)$ and $\tilde L_k=\e^{-1}L_k^{-1}=
1958: \1+(\e-1)\tilde A_k$ is the adjunct matrix for $L_k$, which gives
1959:  $$
1960: \tilde A_1=\begin{bmatrix}0 & 0 \\ 1/2 & 1\end{bmatrix},\quad
1961: \tilde A_2=\begin{bmatrix}1 & 1/2 \\ 0 & 0\end{bmatrix}.
1962:  $$
1963: 
1964: Now equation (\ref{yest}) holds iff there exist 3 not
1965: simultaneously zero numbers $\beta_1,\beta_2,\beta_3$ such that
1966: $\beta_1+\beta_2+\beta_3=0$ and
1967: $\beta_1\eta_1+\beta_2\eta_2+\beta_3\eta_3=0$. Since $\eta_k$ is a
1968: polynomial matrix of degree $|\rho_k|$ and the products of $A_t$
1969: are always proportional to $e_1$ or $e_2$ (depending on the
1970: left-most factor) this last equation is never satisfied if the
1971: multi-indices $\rho_k$ are different.
1972:  \qed
1973:  \end{Proof}
1974: 
1975:  \begin{rk}\po
1976: To support the usage of nearly-Zhang models note that the whole
1977: paradigm of SOC should allow generic perturbations of the data,
1978: for if there is a fine tuning of parameters, the model is
1979: unappropriate for physical explanation (of course, we should pass
1980: to the thermodynamic limit, but in practice this only means some
1981: large finite parameters).
1982:  \end{rk}
1983: 
1984: Our computer experiments did not expose any exponential growth of
1985: multiplicity in the Zhang model (though we see growth in
1986: complications of singularities), so we suggest that $\hm(F)=0$ and
1987: hence $\h(F)=\log N$ always. In addition, by the above discussion
1988: we can disregard these exceptional values of $(E_c,\e)$ even if
1989: there are any. This finishes discussion of topological entropy.
1990: 
1991: %%%%%%%%%%%%%%%%%%%%%%%%
1992: %%%%%%%%%%%%%%%%%%%%%%%%
1993: 
1994: \section{\hps Geometry of the attractor}\label{sec_5}
1995: 
1996: The construction of the spacial attractor $\Y$ can be interpreted
1997: as an iterated function system (IFS), where the maps $F_i$ are not
1998: affine as usually considered, but piecewise affine. Hence one
1999: might expect that attractors $\Y$ are fractal, but with various
2000: size characteristics, like dimension and measure, depending on
2001: parameters $E_c$ and $\epsilon$.
2002: 
2003: %%%%%%%%%%%%%%%%%%%%%%%%
2004: %%%%%%%%%%%%%%%%%%%%%%%%
2005: 
2006: \subsection{\hpss Fractal structure}
2007: 
2008: Computer experiments show that in certain cases the spacial
2009: attractor $\mathcal{Y}$ has fractal structure, see e.g. Figures
2010: \ref{frac1} and \ref{frac2}. We clearly see that $\Y$ consists of
2011: self-similar pieces. However the pieces overlap, making evaluation
2012: of the fractal dimension difficult. So we can provide only
2013: estimates of the attractor's size.
2014: 
2015: Nevertheless we observe from our experiments that Hausdorff
2016: dimension $\dim_\text{H}(\Y)$ and the Lebesgue measure
2017: $\mu_\text{Leb}(\Y)$ of the attractor grow piece-wise
2018: monotonically with $E_c$ and $\e$. Thus the following effects
2019: occur in steps:
2020: 
2021:  \begin{itemize}
2022: \item The dimension and the measure of $\Y$ vanish.
2023: \item $\dim_\text{H}(\Y)$ is positive, while the measure is zero.
2024: \item Both $\dim_\text{H}(\Y)$ and $\mu_\text{Leb}(\Y)$ are
2025: positive.
2026: \item The attractor $\Y$ contains an interior point.
2027:  \end{itemize}
2028: 
2029: We will demonstrate the dimensional part in the next section,
2030: while we disregard the observation about the measure. The reason
2031: for this is that $\mu_\text{Leb}$ is not physically motivated and
2032: we should look for an SRB-measure.
2033: 
2034: The experiments show that such a measure exists and has support
2035: lying strictly inside $\Y$. (See Figure \ref{nonfrac} and Figure
2036: \ref{orbit}. In Figure \ref{nonfrac} the attractor $\Y$ is shown
2037: for the case $N=2$, $E_c=20$ and $\e=2/3$, and Figure \ref{orbit}
2038: shows the orbit of a random initial condition. The latter
2039: corresponds to the support of the SRB-measure.) This is possible
2040: because the contraction rate of $f_{\bf t}$ is smaller for the
2041: exceptional sequences ${\bf t}\in\Sigma_N^+$, than for a generic
2042: one. Thus study of the IFS-attractor does not lead to conclusions
2043: about ergodicity or uniqueness of the SRB-measure. Still it
2044: provides an information about spacial distribution of the orbits
2045: in the Zhang dynamics.
2046: 
2047: %%
2048: %%
2049: \noindent
2050: \begin{figure}
2051: \begin{center}
2052: \includegraphics[width=5.5cm]{f7.eps}
2053: \caption{{\small  Shows the set $U_{10}$ for $N=2$, $E_c=5$ and
2054: $\epsilon=1/2$.}} \label{frac1}
2055: \end{center}
2056: \end{figure}
2057: %%
2058: %%
2059: \noindent
2060: \begin{figure}
2061: \begin{center}
2062: \includegraphics[width=5.5cm]{f8.eps}
2063: \caption{{\small  Shows the set $U_{10}$ for $N=2$, $E_c=3$ and
2064: $\epsilon=1/5$.}} \label{frac2}
2065: \end{center}
2066: \end{figure}
2067: %%
2068: %%
2069: 
2070: 
2071: %%%%%%%%%%%%%%%%%%
2072: %%%%%%%%%%%%%%%%%%
2073: 
2074: \subsection{\hpss Dimensional study of the attractor}
2075: 
2076: The fractal properties of $\Y$ do not hold for all values of
2077: parameters $(E_c,\e)$. An example where $\Y$ has integer dimension
2078: is shown in Figure \ref{nonfrac}.
2079: 
2080: It was noted in \cite{BCK} that Hausdorff (fractal) dimension of
2081: the attractor is about to increase as $E_c$ grows. The arguments
2082: were the following: For bigger $E_c$ the contraction rate
2083: decreases, so the theory of iterated function system (IFS) implies
2084: increasing of the Hausdorff dimension
2085:  $$
2086: {\frak D}_\mathcal{Y}(\e,E_c)=\dim_\text{H}(\mathcal{Y})
2087:  $$
2088: as a function of $E_c$. While this seems to be true, the statement
2089: does not hold in precise sense. For instance, for $N=2$, $E_c\in
2090: [\frac{1+\e}{1-\e},\frac{2}{1-\e}]$ the attractor is the set of 3
2091: points, while it seems to have non-zero dimension for other
2092: parameters (computer simulations clearly show this).
2093: 
2094: %%
2095: \noindent
2096: \begin{figure}
2097: \begin{center}
2098: \includegraphics[width=5.5cm]{f17.eps}
2099: \caption{{\small  Shows the set $\Y$ in the top right corner of
2100: $M$ for $N=2$, $E_c=20$ and $\epsilon=2/3$. The dimension of $\Y$
2101: is $2$ in this example.}} \label{nonfrac}
2102: \end{center}
2103: \end{figure}
2104: %%
2105: %%
2106: \noindent
2107: \begin{figure}
2108: \begin{center}
2109: \includegraphics[width=5.5cm]{f18.eps}
2110: \caption{{\small  Shows an orbit of a randomly chosen point (i.e.
2111: the support of the SRB-measure) for $N=2$, $E_c=13$ and
2112: $\epsilon=2/3$.}} \label{orbit}
2113: \end{center}
2114: \end{figure}
2115: %%
2116: 
2117: The problem is with the framework of IFS, where usually only
2118: conformal maps are considered and certain regularity of their
2119: mapping graph and overlaps is assumed. However we will show
2120: validity of the claim in the asymptotic sense:
2121: 
2122:  \begin{theorem}\po\label{theo-5}
2123: With fixed $d,L$ and generic $\e$ we have:
2124: $\lim_{E_c\to\infty}{\frak D}_\mathcal{Y}(\epsilon,E_c)=N$.
2125: Moreover, ${\frak D}_\mathcal{Y}(\epsilon,E_c)=N$ for big values
2126: $E_c\gg1/\e$.
2127: 
2128: On the other hand for all $\e\in[0,1)$ it holds:
2129: $\lim_{E_c\to0}{\frak D}_\mathcal{Y}(\epsilon,E_c)=0$.
2130:  \end{theorem}
2131: 
2132: In the above statement "generic" means both full Lebesgue measure
2133: and second Baire category. In fact, the equality holds for all
2134: $\e$ outside a countable set. It seems though that the limit
2135: statement is valid for all $\e$. Thus we see that ${\frak
2136: D}_\mathcal{Y}(\e,E_c)$ is not strictly monotone in $E_c$ for its
2137: large values as one might expect from the arguments cited before
2138: the theorem. \\
2139: 
2140:  \noindent
2141: \begin{proof}
2142: Consider first the statement about big energies $E_c\gg1/\e$. Let
2143: us start by demonstrating the idea of the proof on the example
2144: $N=2$, see Figure \ref{pic1}.
2145: 
2146: The image of the vertical continuity domain $M_{13}$ of height 1
2147: adjusted to the right-top corner is a trapezium with the slope
2148: depending on $\e$ (see (\ref{3lines}) of Section \ref{examples}
2149: for the numeration of domains). It is thin -- of constant length
2150: of horizontal section equal $\e$ near its bottom side, but long --
2151: with diameter approximately equal $E_c\frac{1-\e^2}4$. Thus if we
2152: shift $\sim C/\e$ times this domain up and then all of the shifted
2153: images horizontally to the right, so that its first coordinate
2154: satisfies the inequality $E_c-1<x_1\le E_c$, then these shifts
2155: cover an open domain, including a unit square, in the vertical
2156: strip $K_1=\{x_1\in(E_c-1,E_c]\}$ (note that we can leave a copy
2157: of the domain since this corresponds to the shift -- dropping of
2158: energy $(x_1,x_2)\mapsto(x_1,x_2+1)$ on $M_{13}\subset K_1$ before
2159: the avalanche). An easy calculation shows that such shifts cover
2160: the whole upper part of $K_1$, strictly including the continuity
2161: domain $M_{13}\subset K_1$ adjusted to $(E_c,E_c)$ and covering a
2162: vertical part of $M_{12}$. A similar scenario happen to the second
2163: coordinate. Thus in iterating the dynamics we will always have two
2164: continuity domains $M_{13}$ and $M_{23}$ adjusted to $(E_c,E_c)$
2165: and the adjacent parts of $M_{12},M_{22}$ lying in the attractor.
2166: 
2167: For general $L$ and $d$ we observe the same picture: With generic
2168: $\e$ a continuity domain $M_{ij}$ adjusted to the upper-most
2169: corner is mapped under avalanche-map $F_{ij}$ to the
2170: trapezoid-like polyhedron with irrational slopes. Its shifts cover
2171: then an open domain in each of the strips
2172: $K_i=\{x_i\in(E_c-1,E_c]\}$ and so after more shifts -- the upper
2173: part of this strip, whence the statement.
2174: 
2175: We illustrate this process on Figure \ref{pic2}. The 3 domains
2176: adjusted to the corner $(E_c,E_c,E_c)$ are mapped into interior of
2177: the cube and they have different irrational slopes (we picture
2178: them of zero thickness that corresponds to large values of
2179: $E_c\gg1$), so that their shifts cover a big open domain near the
2180: faces adjusted to the above corner.
2181: 
2182: Consider now the second statement, $E_c\ll1$. To estimate the
2183: fractal dimension from above we use the generalization of the
2184: Moran's formula from Appendix \ref{app_B}. It implies that if the
2185: IFS $f_1,\dots,f_N$ satisfies $\|f_i\|\le\delta$, then the
2186: Hausdorff dimension of the attractor admits the following
2187: estimate: ${\frak D}_\mathcal{Y}\le\log
2188: {N\vartheta}/\log\frac1\delta$, where $\vartheta$ is the maximal
2189: multiplicity of the continuity partitions for $f_i$.
2190: 
2191: Now we claim that as $E_c\to0$ we have: $\delta=\max\|f_i\|\le
2192: (E_c)^\sigma$ for some $\sigma>0$. To see this let us estimate the
2193: maximal duration of the avalanche
2194: $\bar\tau_m=\max_{(i,x)\in\Lambda\times M}\tau(i,x)$. This
2195: quantity tends to $\infty$ as $E_c\to0$, but not as fast as
2196: $\tau_m(E_c,\e,\Lambda)\sim C_1/E_c$ (see \S\ref{randex}). Namely
2197: we state that $\bar\tau_m\sim C_2\log1/E_c$. Actually, if we drop
2198: energy 1 to an arbitrary site from a configuration in $M$, then in
2199: a finite $E_c$-independent time all the sites become overcritical.
2200: They remain overcritical, while the system does not loose a
2201: substantial amount of energy. During this process the total energy
2202: is dissipating in geometric progression with an average
2203: contraction rate $1-\frac{1-\e}{N}<1$. So the duration of this
2204: stage has asymptotic $C_3\log(1/E_c)$. The remaining time to
2205: finish the avalanche has a smaller asymptotic.
2206: 
2207: By \S\ref{randex} $n(E_c,\e,\Lambda)<C_0$ for small $E_c$. Thus
2208: the proof of Theorem \ref{5} implies that for certain
2209: $E_c$-independent constant $C_n=kC_0$ for each sequence of $C_n$
2210: steps in the avalanche process the product of the corresponding
2211: $S$-matrices will have norm $\le c(\e)<1$, which is a uniform
2212: estimate in $E_c\ll1$. The number of steps in one avalanche grows
2213: as $\log1/E_c$. Therefore $\delta\le C_4
2214: c(\epsilon)^{C_n^{-1}\log1/E_c}\le\exp(-\sigma\log1/E_c)$, where
2215: $\sigma=\frac12C_n^{-1}\log\frac1{c(\e)}$ as was claimed in the
2216: estimate.
2217: 
2218: Next we claim that $\vartheta\le\vp(E_c)$ with
2219: $\vp=o((1/E_c)^\upsilon)$ $\forall\upsilon>0$. In fact, we
2220: described above the avalanche process for small energies. The
2221: first stage is finite and contributes only a bounded number of
2222: singularity hyperplanes. In its seconds stage all of the sites are
2223: excited, so there the corresponding number of singularity
2224: hyperplanes equals the duration. The last stage is shorter of time
2225: $\psi=o(\log1/E_c)$, but the number of singularity hyperplanes
2226: grows faster, but still is bounded by $e^{\psi\log N!}
2227: =o((1/E_c)^\upsilon)$ (see \S\ref{return maps}) for any
2228: $\upsilon>0$.
2229: 
2230: Finally ${\frak D}_\mathcal{Y}(\e,E_c)\le\log
2231: {N\vartheta}/\log\frac1\delta\le\frac{C'+\log\vp}{\sigma\log1/E_c}\to0$
2232: as $E_c\to0$.
2233: \end{proof}
2234: 
2235: 
2236: \begin{cor}\po
2237: For big $E_c\gg1/\e$ the system $(\mathcal{A},F)$ is not
2238: topologically transitive.
2239: \end{cor}
2240: \noindent
2241: \begin{proof}
2242: Suppose $({\bf t},x)$ is a point with the dense orbit in
2243: $\mathcal{A}$. We know that for big $E_c$ the Lebesgue measure of
2244: the spatial part $\Y$ of the attractor $\mathcal{A}$ is a positive
2245: number $\omega>0$. It follows from Theorem \ref{5} that under
2246: iterations with fixed excitation sequence $\bf t$ the volume of
2247: the spatial part $M$ decreases in geometric progression with the
2248: number of avalanches. Thus after a finite number of steps it
2249: becomes less than $\omega$. This iteration will be still a finite
2250: number of polyhedra, so that its closure does not coincide with
2251: $\Y$. Since it contains all the points $\pi_s(F^n({\bf t},x))$, we
2252: obtain contradiction.
2253: \end{proof}
2254: 
2255: \vspace{10pt}
2256: 
2257: In the case $N=2$ and $\e>1/2$, the value of $E_c$ starting from
2258: which ${\frak D}_\mathcal{Y}(\e,E_c)=2$ can be calculated
2259: precisely because even one shift of the sloped strip mentioned in
2260: the above proof overlaps with itself and is sufficient for
2261: obtaining an open domain in the attractor. This condition $\e>1/2$
2262: together with $E_c\gg1$ from the theorem ideologically coincide
2263: with the sufficient conditions for invertibility of the
2264: differentials of avalanche maps (Proposition \ref{7}). This makes
2265: an indication of a relation between this invertibility and
2266: fractality of the attractor in the spirit of Ledrappier-Young
2267: formula \cite{LY}. This
2268: latter is however unappropriate in our situation. \\
2269: 
2270: \noindent {\bf Note on the usage of the Ledrappier-Young formula.}
2271: This formula, essentially used in \cite{BCK} in the study of the
2272: Zhang model, cannot be used for the map $F:\D \to \D$ since this
2273: map is never invertible (in loc. cit. it was applied to $F^{-1}$).
2274: In addition to invertibility the Ledrappier-Young theorem is based
2275: on the SRB-property. For the map $F^{-1}$ this property is
2276: equivalent to absolute continuity of the stable foliation for $F$
2277: w.r.t.\ the measure $\mu$. If the measure has fractal support this
2278: cannot happen. Therefore all formulas based on this property may
2279: turn to be wrong. We demonstrate this in Example A of \S
2280: \ref{examples}.
2281: 
2282: On the other hand, as we have just shown in the theorem, in
2283: thermodynamic limit $E_c\to\infty$ the fractality is lost and so
2284: the absolute continuity property is restored (but only for the
2285: geometric attractor, the support of SRB-measure is smaller!). This
2286: however does not help with non-invertibility of the factor
2287: $(\Sigma_N^+,\sigma^+_N)$. Even if we change this factor to
2288: invertible two-sided sequences $(\Sigma_N,\sigma_N)$, the system
2289: remains non-invertible since not all points of the attractor
2290: (which can be quite fat) admit negative iterations (Remark
2291: \ref{rkk4}). In addition, a new negative Lyapunov exponent $-\log
2292: N$ in the first factor appears and the formulas exploited in
2293: \cite{BCK} become completely inadequate.
2294: 
2295: 
2296: %%%%%%%%%%%%%%%%%%%%%%%%
2297: %%%%%%%%%%%%%%%%%%%%%%%%
2298: 
2299: \section{\hps Examples} \label{examples}
2300: 
2301: In the examples below we consider the one-dimensional Zhang
2302: model with two sites, $N=2$. \\
2303: 
2304: \noindent
2305: {\bf Example A:} A computation shows that for $E_c \geq
2306: (1+\epsilon)/(1-\epsilon)$ we have six domains of continuity $[i]
2307: \times M_{ij}$, $i=1,2$. The domains $M_{1j}$ are given by
2308: 
2309:  \begin{equation}\label{3lines}
2310:  \begin{array}{l}
2311: M_{11} =\{x\in [0,E_c]^2\,|\,x_1+1\leq E_c \}       \\
2312: M_{12}=\{x\in [0,E_c]^2\,|\,x_1+1> E_c \text{ and }
2313: (1-\epsilon)(x_1+1)/2+x_2 \leq E_c \}\\
2314: M_{13}=\{x\in [0,E_c]^2\,|\,x_1+1>E_c \text{ and }
2315: (1-\epsilon)(x_1+1)/2+x_2 > E_c \}
2316:  \end{array}
2317:  \end{equation}
2318: and the domains $M_{2j}$ are symmetric to these. The maps $F_{ij}$
2319: are of the form $F_{ij}(x)=L_{ij}(x+e_i)$, where
2320:  \begin{center}
2321: \begin{tabular}{lll}
2322: $L_{11}= \begin{bmatrix} 1 &  0  \\ 0 & 1 \end{bmatrix}$ &
2323: $L_{12}= \begin{bmatrix} \epsilon & 0 \\ \frac{1-\epsilon}{2} & 1
2324: \end{bmatrix}$ &
2325: $L_{13}= \begin{bmatrix}  (\frac{1+\epsilon}{2})^2 &
2326: \frac{1-\epsilon}{2}
2327: \\ \epsilon \frac{\vphantom{2^{2^2}} 1-\epsilon}{2} & \epsilon \end{bmatrix}$
2328: \end{tabular}
2329: \end{center}
2330: 
2331: and
2332: \begin{center}
2333: \begin{tabular}{lll}
2334: $L_{11}= \begin{bmatrix} 1 &  0  \\ 0 & 1 \end{bmatrix}$ &
2335: $L_{12}= \begin{bmatrix} 1 & \frac{1-\epsilon}{2} \\ 0 & \epsilon
2336: \end{bmatrix}$ &
2337: $L_{13}= \begin{bmatrix} \epsilon & \epsilon \frac{1-\epsilon}{2}
2338: \\  \frac{\vphantom{2^{2^2}} 1-\epsilon}{2} & (\frac{1+\epsilon}{2})^2 \end{bmatrix}$
2339: \end{tabular}
2340: \end{center}
2341: 
2342: The maps $F_{11}$ and $F_{21}$ correspond to avalanches of size
2343: $0$, the maps $F_{12}$ and $F_{22}$ correspond to avalanches of
2344: size $1$, and the maps $F_{13}$ and $F_{23}$ correspond to
2345: avalanches of size $2$.
2346: 
2347: It was discovered in \cite{BCK} that the physical attractor $\Y$
2348: has the following simple structure:
2349: $$
2350: \Y=\Big{\{} \big{(}\frac{1+\epsilon}{1-\epsilon},
2351: \frac{\epsilon}{2-\epsilon} \big{)},
2352: \big{(}\frac{\epsilon}{2-\epsilon},\frac{1+\epsilon}{1-\epsilon}
2353: \big{)}, \big{(}\frac{1+\epsilon}{1-\epsilon},
2354: \frac{1+\epsilon}{1-\epsilon} \big{)} \Big{\}}\,.
2355: $$
2356: for $E_c \in [\frac{1+\epsilon}{1-\epsilon},\frac{2}{1-\epsilon}]$
2357: We denote these points by $a,b,c$ so that $\Y=\{a,b,c\}$. The maps
2358: $F_1|_\Y$ and $F_2|_\Y$ are permutations of $\Y$:
2359: $$
2360: F_1|_\Y=\Big{(}\begin{matrix} a & b & c \\ b & c & a \end{matrix}
2361: \Big{)}\,,\, F_2|_\Y=\Big{(}\begin{matrix} a & b & c \\ c & a & b
2362: \end{matrix} \Big{)}\,.
2363: $$
2364: %%
2365: %%
2366: %% FIGURE 1
2367: %%
2368: %%
2369: \noindent
2370: \begin{figure}
2371: \begin{center}
2372: \includegraphics[width=10.5cm]{f1.eps}
2373: \caption{{\small The figure shows the physical attractor
2374: $\Y=\{a,b,c\}$ and the maps $F_1|_\Y, F_2|_\Y$ for $E_c=7/2$ and
2375: $\epsilon=1/2$. The arrows on the picture to the left shows how
2376: the points of $\Y$ are mapped under $F_1$, and the picture on the
2377: right shows how the points are mapped under $F_2$.}} \label{figure 1}
2378: \end{center}
2379: \end{figure}
2380: %%
2381: %%
2382: %%
2383: %%
2384: %%
2385: 
2386: Figure \ref{figure 1} shows the physical attractor $\Y$ and the
2387: maps $F_1|_Y$ and $F_2|_Y$ for $E_c=7/2$ and $\epsilon=1/2$. We
2388: construct a partition ${\mathcal R}=\{R_1,\dots, R_6\}$ of
2389: $\Sigma_N^+\times Y$ by
2390: \begin{center}
2391: \begin{tabular}{lll}
2392: $R_1=[1]\times \{a\}$ & $R_2=[1]\times \{b\}$& $R_3=[1]\times \{c\}$ \\
2393: $R_4=[2]\times \{a\}$ & $R_5=[2]\times \{b\}$& $R_6=[2]\times
2394: \{c\}$
2395: \end{tabular}\,.
2396: \end{center}
2397: We let $A=\|a_{ij}\|$ be the $6 \times 6$ matrix where $a_{ij}=1$
2398: if $F(R_i) \cap R_j \neq \emptyset$ and $a_{ij}=0$ otherwise. It
2399: is easy to verify that
2400: $$
2401: A=\begin{bmatrix}
2402: 0 & 1 & 0 & 0 & 1 & 0 \\
2403: 0 & 0 & 1 & 0 & 0 & 1 \\
2404: 1 & 0 & 0 & 1 & 0 & 0 \\
2405: 0 & 0 & 1 & 0 & 0 & 1 \\
2406: 1 & 0 & 0 & 1 & 0 & 0 \\
2407: 0 & 1 & 0 & 0 & 1 & 0 \\
2408: \end{bmatrix}
2409: $$
2410: It is clear that $g_{\mathcal R}: \mathcal A \rightarrow
2411: \Sigma_A^+$ is a topological conjugancy of the maps $F|_{\mathcal
2412: A}$ and $\sigma_A^+$. The matrix $A$ is transitive and
2413: $\text{Sp}(A)=\{-1,-1,0,0,0,2\}$. Hence $\h(F|_{\mathcal A})=\log
2414: 2$.
2415: 
2416: If $\mu$ is the SRB-measure on ${\mathcal A}$, with $(\pi_u)_*\mu$
2417: being the uniform Bernoulli measure on $\Sigma_N^+$, then
2418: $(g_{\mathcal R})_* \mu$ is the Perry measure on $\Sigma_A^+$.
2419: With respect to this measure it is easy to see that the average
2420: avalanche size is $\overline{s}_0=1$. It then follows from that
2421: the sum of the negative Lyapunov-exponents is $\chi_1+\chi_2=\log
2422: \epsilon$. For instance we see that for $E_c=11/2$ and
2423: $\epsilon=2/3$ we have $h_{\mu}(F|_{\mathcal A}) >
2424: |\chi_1|+|\chi_2|$. So even though $F_1|_\Y$ and $F_2|_\Y$ are
2425: invertible, the Ruelle inequality (and therefore the Pesin
2426: formula) cannot be reversed (though this was argued in
2427: \cite{BCK}).
2428: 
2429: We also remark that for $E_c=(1+\epsilon)/(1-\epsilon)$ we have
2430: $\Y \subset S(F)$, so the standard construction of SRB-measure
2431: will fail in this case. Still there is clearly a natural invariant
2432: measure.
2433: 
2434: The example with a trivial physical attractor can be generalized
2435: to all $N$ for $d=1$. Then $\Y$ consists of $N+1$ points
2436: $z_0,z_1,\dots,z_N$ given by
2437: $$
2438: z_n=\frac{1+\epsilon}{1-\epsilon}(1,1,\dots,1)-e_n
2439: $$
2440: where $e_0=0$ and $e_1,\dots,e_N$ is the standard basis in  $\R^N$.  \\
2441: 
2442: \noindent {\bf Example B:}
2443: %%
2444: %%
2445: \noindent
2446: \begin{figure}
2447: \begin{center}
2448: \includegraphics[width=10.5cm]{f15.EPS}
2449: \caption{{\small The region in the $(\e,E_c)$-plane where the
2450: avalanches are the same as for $E_c=\e=1/3$. The point $(1/3,1/3)$
2451: is shown in the interior of the region.}} \label{avalachedomain}
2452: \end{center}
2453: \end{figure}
2454: %%
2455: %%
2456: For $E_c=\epsilon=1/3$ the dynamics is very simple. The map $F_1$
2457: has two domains of continuity: $ M_{11}=\{(x,y)\in M\,|\,y \leq
2458: -2x/3+1/3\}$ and $M_{12}=M \setminus M_{11}$. The domains of
2459: continuity for $F_2$ are given by symmetry. The maps are given by
2460: $F_{11}(x)=L_{11}(x+e_1)$ and $F_{12}(x)=L_{12}(x+e_1)$, where
2461: $$
2462: L_{11}=\frac19 \begin{bmatrix} 2 & 3 \\ 2 & 3
2463: \end{bmatrix}\,\,\text{ and }\,\,
2464: L_{12}=\frac{2}{27} \begin{bmatrix} 2 & 3 \\ 2 & 3
2465: \end{bmatrix}\,.
2466: $$
2467: We see that all four $F_{ij}$ are mappings to the diagonal line in
2468: $M$. In fact, $F_{11}(M_{11})=F_{21}(M_{21})=[p_1,p_2]$, where
2469: $p_1=(2/9,2/9)$ and $p_2=(1/3,1/3)$. The interval is contained in
2470: $M_{12} \cap M_{22}$ since the two lines of singularity intersect
2471: the diagonal in the point $(1/5,1/5)$. The images of $M_{12}$ and
2472: $M_{22}$ also coincide and is an interval $[p_1,p_3]\subset
2473: [p_1,p_2]$, where $p_3=(22/81,22/81)$. This shows that
2474: singularities are removable after one iteration. It is easy to see
2475: that for all ${\bf t} \in \Sigma_N^+$ the dynamics will contract
2476: to the fixed point $P=(4/17,4/17)$. Hence the attractor of the
2477: system is ${\mathcal A}=\Sigma_N^+ \times \{P\}$. The dynamics on
2478: the attractor is of course conjugated to $\sigma_N^+$.
2479: 
2480: The example can easily be extended to a neighborhood of
2481: $(1/3,1/3)$ in the $(\e,E_c)$-plane. In the region
2482: $$
2483: \max \Big{\{}\frac{1-2\e+13\e^2}{5+12\e-13 \e^2},
2484: \frac{\e(7\e^2-6\e+3)}{7\e^3-10 \e^2+7 \e-4} \Big{\}} \leq E_c
2485: \leq \min \Big{\{}\frac{\e}{1-\e}, \frac{1-2\e+5\e^2}{1+4\e-5\e^2}
2486: \Big{\}}
2487: $$
2488: we have the same avalanches as for $E_c=\e=1/3$. This region is
2489: shown in Figure \ref{avalachedomain}. For all points in the region
2490: the maps depend continuously on $\e$ and the lines of singularity
2491: depend continuously on $\e$ and $E_c$. The condition for an atomic
2492: spacial attractor is that the images of the domains do not
2493: intersect the singularities. For $E_c=\e=1/3$ the images are
2494: bounded away from the lines of singularity, and hence there exists
2495: an open neighborhood of $(1/3,1/3)$ in the $(\e,E_c)$-plane where
2496: the same holds, i.e. the attractor is of the form $\Sigma_N^+
2497: \times \{P\}$ (where $P$ is a point in $M$) and the dynamics is
2498: conjugated to $\sigma_N^+$. \\
2499: 
2500: \noindent {\bf Example C:} Let us consider $E_c=1/3$ and
2501: $\epsilon=1/2$. In this case there are 28 domains of continuity
2502: and 28 corresponding maps. A computer program is written to
2503: compute the sets $U_n$. The program uses exact calculations of the
2504: edges of the polygons that make up $U_n$, and hence it can be used
2505: to give rigorous ''proof by computer'' of removability of
2506: singularities. By using the program we obtain that singularities
2507: are removable. In fact $U_5 \cap {\mathcal S}=\emptyset$. The set
2508: $U_5$ consists of 13 connected components. Figure \ref{figure 2}
2509: shows the set $U_5$ and the lines of singularity, and Figure
2510: \ref{figure 3} is a schematic illustration of how these connected
2511: components are situated with respect to the lines of singularity.
2512: %%
2513: %%
2514: %% FIGURE 2
2515: %%
2516: %%
2517: \noindent
2518: \begin{figure}
2519: \begin{center}
2520: \includegraphics[width=5.5cm]{f2.eps}
2521: \caption{{\small Shows the set $U_5$ for $N=2$, $E_c=1/3$ and
2522: $\epsilon=1/2$. The points on the attractor are magnified in order
2523: to make them visible in the figure, and hence it looks as if they
2524: intersect singularities, but in fact they do not.}} \label{figure
2525: 2}
2526: \end{center}
2527: \end{figure}
2528: \noindent
2529: %%
2530: %%
2531: %% FIGURE 3
2532: %%
2533: %%
2534: \begin{figure}
2535: \begin{center}
2536: \includegraphics[width=5.5cm]{f3.eps}
2537: \caption{{\small Shows how the 13 spatial partition elements
2538: are situated with respect to the lines of singularity.}} \label{figure 3}
2539: \end{center}
2540: \end{figure}
2541: %%
2542: %%
2543: %% FIGURE 4
2544: %%
2545: %%
2546: \begin{figure}
2547: \begin{center}
2548: \includegraphics[width=5.5cm]{f4.eps}
2549: \caption{{\small Shows the matrix $A$ for the coding of Example C.
2550: Black squares are 1-s and white squares are 0-s.}} \label{figure
2551: 4}
2552: \end{center}
2553: \end{figure}
2554: %%
2555: %%
2556: %%
2557: %%
2558: %%
2559: \noindent The intersection of the connected components of $U_5$
2560: with $\Y$ are denoted by $Y_1,\dots ,Y_{13}$. Then we construct the
2561: partition $\mathcal R=\{[i] \times Y_j\}$, and enumerate the
2562: elements so that $\mathcal R=\{R_1,\dots,R_{26}\}$, where
2563: \begin{center}
2564: \begin{tabular}{llll}
2565: $R_1=[1]\times Y_1$ & $R_3=[1]\times Y_2$  & \dots &
2566: $R_{25}=[1]\times Y_{13}$
2567: \\
2568: $R_{2}=[2]\times Y_1$ & $R_4=[2]\times Y_2$  & \dots &
2569: $R_{26}=[2]\times Y_{13}$
2570: \end{tabular}\,.
2571: \end{center}
2572: We construct the $26 \times 26$ matrix $A=\|a_{ij}\|$ by letting
2573: $a_{ij}=1$ if $F(R_i) \cap R_j \neq \emptyset$, and $a_{ij}=0$
2574: otherwise. After making the computations, the matrix $A$ becomes
2575: as shown in Figure \ref{figure 4}. The black squares represent
2576: ones and white squares represent zeros. Direct computation shows
2577: that the matrix $A$ is transitive.
2578: 
2579: Since singularities are removable the map $g_{\mathcal
2580: R}:{\mathcal A} \rightarrow \Sigma_A^+$ is an avalanche conjugancy
2581: between $F_{\mathcal A}$ and $\sigma_A^+$. The SRB-measure
2582: projects to the Perry measure on $\Sigma_A^+$, so it is possible
2583: to calculate properties such as average avalanche size. In this
2584: example a computation gives $\overline{s}_0=123/17$.  The spectral
2585: radius of the matrix $A$ is $2$, and hence
2586: $\h(\sigma_A^+)=\log 2$.  \\
2587: 
2588: \noindent {\bf Example D:} In the previous examples singularities
2589: are removable. This is however not always the case. Figure
2590: \ref{figure 5} shows the set $U_{20}$ for $N=2$, $E_c=7$ and
2591: $\epsilon=1/2$. This is an example where singularities are
2592: non-removable.
2593: %%
2594: %%
2595: %% FIGURE 5
2596: %%
2597: %%
2598: \begin{figure}
2599: \begin{center}
2600: \includegraphics[width=5.5cm]{f5.eps}
2601: \caption{{\small Shows the set $U_{20}$ for $E_c=7$ and $\e=1/2$. In this example
2602: singularities are non-removable.}} \label{figure 5}
2603: \end{center}
2604: \end{figure}
2605: %%
2606: %%
2607: %% FIGURE 6
2608: %%
2609: %%
2610: \begin{figure}
2611: \begin{center}
2612: \includegraphics[width=5.5cm]{f6.eps}
2613: \caption{{\small Illustration of the fact that singularities
2614: are non-removable for $E_c=7$ and $\e=1/2$.}} \label{figure 6}
2615: \end{center}
2616: \end{figure}
2617: %%
2618: %%
2619: %%
2620: %%
2621: %%
2622: \noindent We will in the following show that singularities are
2623: non-removable for $E_c=(3+\epsilon)/(1-\epsilon)$.
2624: 
2625: Since  $(3+\epsilon)/(1-\epsilon)>(1+\epsilon)/(1-\epsilon)$, the
2626: domains of continuity are given by the formulas presented in
2627: Example A. Take the point $p=(E_c, E_c-1) \in M_{13}$. See Figure
2628: \ref{figure 6}. Observe that $p$ lies on the horizontal line
2629: $x_1=E_c-1$ and hence $p \in S(F)$. Clearly $p$ is in the interior
2630: of $M_{13}$, so
2631: $$
2632: F_1(p)=F_{13}\Big{(}\frac{3+\epsilon}{1-\epsilon},
2633: \frac{3+\epsilon}{1-\epsilon}-1\Big{)}=
2634: \Big{(}\frac{3+\epsilon}{1-\epsilon}-1,
2635: \frac{3+\epsilon}{1-\epsilon}-2 \Big{)}=p-(1,2)\,.
2636: $$
2637: Denote $q_1:=F_1(p)$ and observe the it lies on the singularity
2638: line $x_2=E_c-1$. On the other hand $q_1$ is in the interior of
2639: $M_{21}$, so $F_2(q_1)=p-(1,1)$. Denote $q_2:=F_2(q_1)$. This
2640: point also lies in the interior of $M_{21}$. Let
2641: $q_3:=F_{2}(q_2)=p-(1,0)$. It is clear that $F_{11}(q_3)=p$, and
2642: in this sense $p$ is a periodic point. However, $F_{1}(q_3)$ is
2643: not well defined since $q_3 \in \partial M_{12} \cap  \partial
2644: M_{13}$.
2645: 
2646: In the following we let $\langle z_1,\dots,z_n \rangle$ denote the
2647: open convex polygon with edges $z_1,\dots,z_n$. Define
2648: $B_\delta(p)=\langle p, p+(0,\delta),p+(-\delta,
2649: \delta),p+(-\delta,0)\rangle$. For small $\delta>0$ we have
2650: $B_\delta(p) \in M_{13}$, and hence
2651: $$
2652: F_1(B_{\delta}(p))=F_{13}(B_{\delta}(p))=\langle q_1,q_1+\delta
2653: a,q_1+\delta b,q_1+\delta c \rangle\,,
2654: $$
2655: where
2656: $$
2657: a=\Big{(} \frac{1-\epsilon}{2}, \epsilon \Big{)}\,,\,\, b=\Big{(}
2658: \frac{1-4\epsilon-\epsilon^2}{4}, \frac{1+\epsilon}{2} \epsilon
2659: \Big{)}\,,\,\, c=\Big{(}-\big{(} \frac{1+\epsilon}{2}\big{)}^2,
2660: -\frac{1-\epsilon}{2} \epsilon \Big{)} \,.
2661: $$
2662: The polygon $F_1(B_{\delta}(p))$ intersects the singularity line
2663: $x_1=E_c-1$. See Figure \ref{figure 6}. A simple computation shows
2664: that
2665: $$
2666: F_1(B_{\delta}(p)) \cap M_{11}=\langle q_1,q_1+\delta
2667: a',q_1+\delta b,q_1+\delta c \rangle\,,
2668: $$
2669: where
2670: $$a'=\Big{(}0, \big{(} \frac{2\e}{1+\e} \big{)}^2 \Big{)}\,.$$
2671: It then follows from the above discussion that
2672: $$
2673: F_1\circ F_2^2(F_1(B_{\delta}(p)) \cap M_{11})= \langle p,p+\delta
2674: a',p+\delta b,p+\delta c \rangle\,.
2675: $$
2676: It is then easy to verify that for all $\delta>0$ there is
2677: $\gamma>0$ such that $B_\gamma(p) \subset \pi_s\circ
2678: F^4(\Sigma_N^+ \times B_\delta(p))$. So for each $n \in \N$ there
2679: is $\delta_n>0$ with $B_{\delta_n}(p) \subset U_n$. The image of $
2680: B_{\delta_n}(p)$ under $F_1$ intersects singularities, and its
2681: closure contains the point $q_1$, which thus is an essential
2682: singularity. Clearly the points $p$, $q_2$ and $q_3$ are also
2683: essential singularities.
2684: 
2685: 
2686: 
2687: 
2688: %%%%%%%%%%%%%%%%%%
2689: %%%%%%%%%%%%%%%%%%
2690: 
2691: \section{\hps Statistical properties}\label{sec_7}
2692: 
2693: In order to evaluate the entropy and Lyapunov spectrum
2694: of the physical model
2695: in the thermodynamic limit we need to derive several
2696: estimates for the asymptotic behavior of observables like
2697: avalanche size, avalanche duration and "waiting-time" between
2698: avalanches. The results are derived using only the uniform Bernoulli
2699: measure on $\Sigma_N^+$, and hence hold for any SRB-measure and for
2700: time-averages.
2701: 
2702: In \S \ref{sect13} we define the thermodynamic limit as the double
2703: limit $E_c \rightarrow \infty$, $L=\sqrt[d]{N} \rightarrow
2704: \infty$, contrary to \cite{BCK}, where the thermodynamic limit is
2705: defined as the limit $L \rightarrow \infty$ only. This is
2706: important as the quasi-classical limit since equivalently means a
2707: fixed energy, but the energy quantum of Section \ref{sec_1}
2708: $\delta\to0$. As $E_c \rightarrow \infty$ we must make a scaling
2709: of time in the physical model. Otherwise the influx of energy to
2710: the system will go to zero. With this new scaling we show that the
2711: entropy goes to zero and the Lyapunov spectrum is collapsing.
2712: 
2713: We do not provide strict mathematical proofs, but still think
2714: important to include the discussion of our results from the
2715: physical point of view.
2716: 
2717: 
2718: %%%%%%%%%%%%%%%%%%
2719: %%%%%%%%%%%%%%%%%%
2720: 
2721: \subsection{\hpss Statistics of observables}\label{Ssec}
2722: 
2723: Let $\tau$ be the coordinate measuring the duration of avalanche
2724: and let $\omega$ correspond to the interval between avalanches
2725: (minimal value 1). We will also study the observable $s$ -- the
2726: avalanche size (defined in \S \ref{return maps}).
2727: While in the first case we
2728: consider only actual avalanches, so that $\tau>0$, in the second
2729: we make distinction between $s_0$ -- all avalanches including the
2730: trivial case of under-critical state ($s=0$) and $s_+$ -- the
2731: actual avalanches, so that $s_+>0$.
2732: 
2733: The reason for introducing two different avalanche size
2734: observables is the following: $s_0$ plays a crucial role in
2735: mathematical investigation of the model (see \S\ref{L.S.}), while
2736: $s_+$ is important from physical perspective. In \S\ref{sect13} we
2737: will see that physical observables should allow a thermodynamic
2738: limit.
2739: 
2740: Denote by $\bar\tau$, $\bar\omega$, $\bar s_0$, $\bar s_+$ the
2741: corresponding mean time-average quantities, each of which is a
2742: function on the space-factor $M$ and is defined as follows:
2743:  $$
2744: \bar\sigma(x)=\int_{\Sigma_N^+}\lim_{k\to\infty}\frac1k
2745: \sum_{i=0}^{k-1}\sigma\bigl(\hat f^i({\bf t},x)\bigr)\,d\mu_{\text{{\tiny Ber}}}
2746:  $$
2747: with $\mu_{\text{{\tiny Ber}}}$ being the Bernoulli measure on the
2748: one-sided shifts (by Birkhoff ergodic theorem the time-average
2749: limit exists almost everywhere and is measurable). This function
2750: is invariant in the sense:
2751: $\bar\sigma(x)=\frac1N\sum_{i=1}^N\bar\sigma(\hat f(i,x))$.
2752: 
2753: Whenever the system is ergodic with respect to an invariant measure $\mu$ the
2754: function $\bar\sigma$ is constant $\mu$-a.e.
2755: and equals the space (ensemble) average
2756: $$
2757: \langle\sigma\rangle_\mu=\int\limits_{\Sigma_N^+\times M}
2758: \hspace{-8pt}\sigma\,d\mu.
2759: $$
2760: If the system has a unique invariant SRB-measure the function
2761: $\bar\sigma$ is constant $\mu_{\text{\tiny Ber}} \times
2762: \mu_{\text{\tiny Leb}}$-a.e. and equals the space (ensemble)
2763: average
2764: $$
2765: \langle\sigma\rangle=\int\limits_{\Sigma_N^+\times M}
2766: \hspace{-8pt}\sigma\,d\mu_{\text{\tiny SRB}}.
2767: $$
2768: 
2769: We will need the maximal values of these observables in a sequel,
2770: which we denote by $\tau_{\text{max}}$, $\omega_{\text{max}}$ and
2771: $s_{\text{max}}=\max s_0=\max s_+$ respectively. We shall
2772: calculate their asymptotics in $L$ and $E_c$.
2773: 
2774: Denote by $\varphi_1\sim\varphi_2$ the asymptotic equivalence
2775: relation meaning that the ratio $\varphi_1/\varphi_2$ has
2776: subexponential grow/decay. We denote the equivalence by $\approx$,
2777: when the limit of the ratio is 1.
2778: 
2779:  \begin{lem}\po\label{lmm}
2780: For $E_c\gg1$ the maximal avalanche time and size have the
2781: asymptotics: $\tau_{\text{max}}\sim L^{\gamma_\tau}$,
2782: $\omega_{\text{max}}=L^d E_c$, $s_{\text{max}}\sim
2783: L^{d+\gamma_s}$, where $\gamma_\tau=\gamma_s=1$ for $d=1$ and
2784: $1<\gamma_\tau,\gamma_s<d$ for $d>1$.
2785:  \end{lem}
2786:  \begin{proof}
2787: Consider at first the simple case $d=1$.
2788: The maximum avalanche duration and size are achieved when all
2789: sites contribute to the avalanche, i.e.\ their energies are
2790: sufficiently big and there is one site with energy greater than
2791: $E_c-1$ to initiate the avalanche. Actually, if some site has
2792: small energy, it will serve as a boundary and the avalanche wave
2793: reflects from it (to be explained below).
2794: 
2795: Assuming $E_c>\frac\e{1-\e}$ we know from Lemma \ref{6} that in
2796: the avalanche process each site, whenever overcritical, relaxes
2797: until in the next step it receives a sufficient portion of energy
2798: to become overcritical and relax etc. In other words, the sites
2799: blink, being under- and over-critical in turn. But in this process
2800: they make overcritical their neighbors and the process propagates
2801: as a wave, with only difference that its front excites new sites,
2802: while in the traversed region there remain blinking overcritical
2803: sites.
2804: 
2805: This wave spreads along the interval $B_L^1=[1,L]\subset\Z$
2806: towards its boundary and then it reflects from it, bearing now
2807: relaxation. In fact, as the front wave reaches the boundary it
2808: losses a substantial part of the energy on the boundary sites,
2809: which thus cannot be recovered and remain undercritical for the
2810: rest of this avalanche. They influence their neighbors to stop
2811: being critical and so forth. Thus we obtain the reflected wave
2812: that, in contrast with the first one, turns overcritical sites
2813: into relaxed. When the wave hits itself, the avalanche process
2814: stops.
2815: 
2816: It is clear that the duration of this avalanche is
2817: $\tau_{\text{max}}\approx L$. The number of involved sites
2818: corresponds to the area of the triangle with a side $B_L^1=[1,L]$
2819: and height $L/2$ (recall that each site is overcritical only half
2820: time of its blinking period), whence $s_{\text{max}}\approx
2821: L^2/4$.
2822: 
2823: 
2824: Consider now the case of dimension $d>1$. Here the scenario of
2825: maximal avalanche is more complicated and consists of three stages
2826: (the proof is similar to the case $d=1$, but quite lengthy and will be suppressed).
2827: Again the maximum avalanche duration and size are achieved when
2828: all sites contribute to the avalanche, though now if some isolated
2829: site has smaller energy it serves as a boundary only once but then
2830: on the next several waves it receives the required portion of energy
2831: and follows the general scheme of motion.
2832: 
2833: At the first stage a site is excited and it initiates the
2834: rhombus-shape wave (a cube in the Manhattan metric, see Figure
2835: \ref{xxx}) that spreads to the boundary of the cube
2836: $B_L^d=[1,L]^d\subset\Z^d$ (in time $\approx L/2$ if the center of
2837: the rhombus is placed near the center of the cube).
2838: %%
2839: \noindent
2840: \begin{figure}
2841: \begin{center}
2842: \includegraphics[width=5.5cm]{f9.eps}
2843: \caption{{\small The picture is a "snapshot" of the lattice $\Lambda$ for
2844: $d=2$, $L=10$, $E_c=7$ and $\epsilon=1/2$. A single site in the marginally
2845: stable configuration has been exited, and a great avalanche is unfolding in a
2846: rhombus shape. This is the first stage of this avalanche. The white squares are
2847: overcritical sites, the gray squares are sites with energy just bellow $E_c$ and
2848: the black squares have energy approximately equal $\epsilon E_c$.}} \label{xxx}
2849: \end{center}
2850: \end{figure}
2851: %%
2852: %%
2853: The second stage begins as the wave reaches the boundary face and
2854: reflects from it (See Figure \ref{zzz}). The reflected wave is
2855: almost momentary overthrown by the coming overcritical wave, which
2856: again reflects from the boundary, come now deeper into the
2857: interior of the cube, but is overthrown too etc. If one looks
2858: along the boundary face, the reflected wave travel along towards a
2859: vertex with preserved form (like a soliton) and then disappears
2860: into this vertex (there occur strong interactions with other waves
2861: in this corner). But if one looks into the perpendicular
2862: direction, the collection of reflected waves oscillates (each
2863: reflected wave enters deeper and deeper into the cube)
2864: contributing to the avalanche duration the sum $1+2+3+\dots$,
2865: which stops with the end of the second stage (we do not specify
2866: the sum precisely because after some oscillations the wave front
2867: becomes more and more eroded by the interactions between
2868: overcritical and relaxing waves; this impairs the sum and
2869: decreases the exponent, but not too drastically).
2870: %%
2871: \noindent
2872: \begin{figure}
2873: \begin{center}
2874: \includegraphics[width=5.5cm]{f12.eps}
2875: \caption{{\small The picture shows a "snapshot" of the second
2876: stage of the avalanche shown in Figure \ref{xxx}. We can see the
2877: well-shaped (soliton-like) waves of energy near the boundary and
2878: observe how their form begins being eroded near the vertices.}}
2879: \label{zzz}
2880: \end{center}
2881: \end{figure}
2882: %%
2883: %%
2884: 
2885: The third stage begins as the main body of the overcritical sites
2886: becomes disconnected and the avalanches behaves like worms
2887: crawling along the high energy fractal-like collection of states.
2888: We illustrate this in Figure \ref{xyz}.
2889: %%
2890: \noindent
2891: \begin{figure}
2892: \begin{center}
2893: \includegraphics[width=5.5cm]{f10.eps}
2894: \caption{{\small  The picture shows a "snapshot" of the third
2895: stage of the avalanche shown in Figure \ref{xxx}. The energy
2896: configuration has a fractal structure where the avalanche can
2897: sustain in regions of high energy.}} \label{xyz}
2898: \end{center}
2899: \end{figure}
2900: %%
2901: From the description of the maximal avalanche process it is clear
2902: that the asymptotic exponents $\gamma_\tau,\gamma_s$ do not depend
2903: on the energy $E_c\gg1$. Let us denote (the sequences are
2904: increasing, so the limits exist):
2905: $$
2906: \gamma_\tau=\lim_{L\to\infty}\frac{\log\tau_{\text{max}}}{\log L}
2907: \quad\text{ and }\quad d+\gamma_s=\lim_{L\to\infty}\frac{\log
2908: s_{\text{max}}}{\log L}.
2909: $$
2910: 
2911: Duration of the first stage of avalanche is $\sim L$. The above
2912: arguments show that the exponent $\gamma_\tau'$ of the second
2913: stage is $>1$. The last stage can only increase it:
2914: $\gamma_\tau\ge\gamma_\tau'$. To see that $\gamma_\tau<d$ we note
2915: that in average the number of critical sites on the boundary is
2916: about $\kappa_L\sim L^{\gamma_\kappa}$ with $0<\gamma_\kappa<d$.
2917: Thus we have a constant flow of energy out of the system with the
2918: average speed $>\frac{1-\e}{2d}\kappa_L E_c$, and the inequality
2919: $E_{\text{tot}}\le L^dE_c$ proves the claim.
2920: 
2921: To estimate the maximal avalanche size exponent $\gamma_s$
2922: consider again the second stage of the above scenario. The number
2923: of involved sites corresponds to the volume of the prism $\Pi_L$
2924: over the cube $B_L^d$ with height
2925: $\ell\approx\frac12\tau_{\text{max}}\sim L^{\gamma_\tau'}$, i.e.
2926:  $$
2927: L^{d+\gamma_s'}\sim\op{Vol}_{d+1}(\Pi_L)=\dfrac
2928: \ell{d+1}\cdot\op{Vol}_d(B^d_L) \sim L^{d+\gamma_\tau'}.
2929:  $$
2930: Thus $\gamma_s\ge\gamma_s'=\gamma_\tau'>1$. Inequality
2931: $\gamma_s<d$ follows from the inequality  from above for the
2932: duration exponent $\gamma_\tau$.
2933: 
2934: The maximal value of the waiting time is obvious.
2935:  \end{proof}
2936: 
2937: 
2938: 
2939: %%%%%%%%%%%%%%%%%%
2940: %%%%%%%%%%%%%%%%%%
2941: 
2942: \subsection{\hpss Asymptotic of the statistical data}\label{S-sec}
2943: 
2944: Now we can study statistics of the avalanche data asymptotically
2945: (as for the thermodynamic limit).
2946: 
2947: We will need the following technical statement (informal only for
2948: $E_c>1$):
2949: \begin{lem}\po\label{lml}
2950: Almost every (w.r.t. a random excitation sequence) spatial
2951: trajectory returns to the cube $B_0=(E_c-1,E_c]^N$.
2952: \end{lem}
2953: \begin{proof}
2954: It follows from Proposition \ref{4} that $K=\{x\in M\,|\,\exists
2955: i:x_i\in(E_c-1;E_c]\}$ is a return set, i.e. every trajectory
2956: $F^n({\bf t},x)$ meets $\Sigma_N^+ \times K$. Partition the
2957: spatial part $\Y$ of the attractor according to the hyperplanes
2958: collection $\mathcal{H}=\cup_{i,m\in\N}\{x_i=E_c-m\}$.
2959: 
2960: Each such a part can be shifted by an excitation sequence to the
2961: cube $B_0$. The probability of all such sequences (where the
2962: avalanche starts from the set $B_0$ of maximal energy in all
2963: sites) is positive. Let $\rho>0$ denote the minimum of these
2964: probabilities over the finite set of all partition elements of $Y$
2965: by $\mathcal{H}$. Then the probability of not entering $B_0$ in
2966: $k$ successive avalanches is less than $(1-\rho)^k$.
2967: 
2968: Therefore since the number of avalanches tend to infinity as we
2969: iterate the dynamics, the measure of trajectories staying away from
2970: $B_0$ is zero.
2971: \end{proof}
2972: 
2973: 
2974: \begin{theorem}\po\label{tyh}
2975: We have the following asymptotic estimates valid as $E_c\to\infty$
2976: (and $N\gg1$ fixed) or $L=\sqrt[d]{N}\to\infty$ (and $E_c\gg1$
2977: fixed):
2978: \begin{enumerate}
2979: \item[\rm(i)] $\bar\tau\sim L^{\gamma_\tau}$;
2980: \item[\rm(ii)] $\bar\omega \sim E_c$;
2981: \item[\rm(iii)] $\bar s_0\sim L^{d+\gamma_s}/E_c$;
2982: \item[\rm(iv)] $\bar s_+\sim L^{d+\gamma_s}$,
2983: \end{enumerate}
2984: where $\gamma_\tau,\gamma_s$ are the same exponents as in Lemma
2985: \ref{lmm} (thus $\gamma_\omega=\lim\limits_{L\to\infty}\frac
2986: {\log\bar\omega}{\log L}=0$).
2987: \end{theorem}
2988: \begin{proof}
2989: The maximal avalanche size is achieved for a certain configuration
2990: of states $V\subset M$, which we can bound as follows: $U_1\subset
2991: V\subset U_d$, where
2992:  $$
2993: U_j=\{x\in M\,|\,\forall i:
2994: \bigl(1-j\tfrac{1-\e}{2d}\bigr)E_c<x_i\le E_c\text{ and }\exists
2995: i_0: x_{i_0}>E_c-1\}.
2996:  $$
2997: Denote also $\tilde U_j=\{x\in M\,|\,\forall i:
2998: \bigl(1-j\tfrac{1-\e}{2d}\bigr)E_c<x_i\le E_c\}\supset U_j$.
2999: 
3000: To estimate the measure of the sites leading to the maximal
3001: avalanche we consider preimages of $\Sigma_N^+\times\tilde U_j$
3002: under the map $F$. It is clear that one needs
3003: $k_\e\in[\frac2{1-\e},\frac{2d}{1-\e}]$ different backwards
3004: iterations $F^{-i_s}$, $s=1,\dots,k_\e$, to cover the spatial
3005: attractor $\mathcal{Y}$. Since the measure $\mu$ is $F$-invariant,
3006: we get for its $\pi_s$-push-forward: $\mu_s(\tilde
3007: U_j)\approx\rho_j/k_\e$, which is $E_c$-independent. Thus we get
3008: the same exponent $(d+\gamma_s)$ for $\bar s_+$ as for
3009: $s_{\text{max}}$.
3010: 
3011: The same arguments yield the asymptotic of $\bar\tau$.
3012: 
3013: To obtain the asymptotic of $\bar\omega$ in $L$ we note that since
3014: the amount of lost energy is $<C L^{\gamma_\kappa+\delta}E_c$
3015: (where $\gamma_\kappa<d$ is the quantity from the proof of Lemma
3016: \ref{lmm}), the average remained energy in a site of configuration
3017: obtained from a maximal one after an avalanche is $(L^dE_c-C
3018: L^{\gamma_\kappa+\delta}E_c)/L^d\approx E_c$. Thus the waiting
3019: time does not grow with $L$.
3020: 
3021: The asymptotic of $\bar\omega$ in $E_c$ is quite different: If $N$
3022: is fixed but $E_c$ grows, then any state from $\partial M$ becomes
3023: at distance $\theta_N\cdot E_c$ after some relatively small number
3024: of iterations.
3025: 
3026: Let us first demonstrate the idea in the simple case $N=2$. For
3027: critical energy $E_c>\e/(1-\e)$ the picture of avalanches is shown
3028: on Figure \ref{pic1}. We see that in a few steps of the dynamics
3029: the configuration becomes far from $\partial M$, i.e. it strongly
3030: contracts in all directions. Actually, it is possible to imagine
3031: the situations when the point is mapped to the vertical strip and
3032: then is shifted horizontally for a long time by excitations of the
3033: first site, but probability of this event exponentially goes to
3034: zero as $E_c\to\infty$.
3035: %%
3036: %%
3037: %%
3038: \noindent
3039: \begin{figure}
3040: \begin{center}
3041: \includegraphics[width=7.5cm]{f13.EPS}
3042: \caption{{\small The figure illustrates how the continuity
3043: domains are mapped under $F_1$ and $F_2$ for $N=2$ and
3044: $E_c>\e/(1-\e)$.}}\label{pic1}
3045: \end{center}
3046: \end{figure}
3047: %%
3048: %%
3049: Thus in a relatively short time the point from $\partial M$ is
3050: mapped into the square $[0,\theta_2E_c]^2$, where the constant
3051: $\theta_2$ is $E_c$-independent. To achieve the boundary $\partial
3052: M$ again it needs $\sim(1-\theta_2)2E_c$ random excitations.
3053: 
3054: The similar picture happens for $N=3$, see Figure \ref{pic2}.
3055: %%
3056: %%
3057: %%
3058: \noindent
3059: \begin{figure}
3060: \begin{center}
3061: \includegraphics[width=7.5cm]{f14.EPS}
3062: \caption{{\small The figure illustrates how the faces
3063: of the continuity domains are mapped under
3064: $F_1$, $F_2$ and $F_3$ for $N=3$ and $E_c>\e/(1-\e)$. The
3065: point of view is at $(E_c,E_c,E_c)$.}} \label{pic2}
3066: \end{center}
3067: \end{figure}
3068: %%
3069: %%
3070: %%
3071: In the general case Theorem \ref{5} insures that after some (few)
3072: number of steps we get strong contraction, so that the point
3073: becomes in the cube $[0,\theta_NE_c]^N$, i.e. far from the
3074: boundary $\partial M$. Thus we need $\sim(1-\theta_N)NE_c$
3075: excitations to make it overcritical and this implies the claim.
3076: Note that the asymptotic for $\bar\omega$ is not for the double
3077: limit $E_c\to\infty$, $L\to\infty$, but for two partial limits
3078: only.
3079: 
3080: To obtain the estimate for $\bar s_0$ we need to estimate the
3081: conditional measure $\mu_s(U_j|\tilde U_j)$, which coincides with
3082: the probability that a randomly chosen configuration $x\in \tilde
3083: U_j$ and site $i$ satisfy: $x_i>E_c-1$. This probability is
3084: $\approx b_1/E_c$. Thus $\mu(\Sigma_N^+\times V)\sim
3085: \sigma_\e/E_c$ and the mean avalanche size is $\langle
3086: s_0\rangle\sim L^{d+\gamma_s}/E_c$.
3087: 
3088: This is however the space-average (the arguments below work also
3089: for $\bar s_+$, $\bar\tau$). We would obtain the same for the
3090: time-average of Lebesgue a.e.-initial condition if we have an
3091: SRB-measure, or for $\mu$-a.e. if we have an ergodic measure
3092: $\mu$. But we cannot guarantee existence of an SRB-measure or an
3093: ergodic measure. However, for a non-ergodic measure, we can
3094: decompose $\mathcal{A}$ into ergodic components, where the
3095: Birkhoff theorem works. By Lemma \ref{lml} each ergodic component
3096: (with non-trivial contribution) intersects in its spatial part the
3097: top-energy cube $B_0$ and so for each of it the same asymptotic of
3098: space-averages holds with universal exponents but maybe different
3099: coefficients. Therefore we obtain the required asymptotic for
3100: $\bar s_0$.
3101: 
3102: Another way to get the last asymptotic is via the formula $\bar
3103: s_0=\frac{\bar\omega\cdot 0+1\cdot\bar s_+}{\bar\omega+1}$.\!\!
3104: \end{proof}
3105: \vspace{0pt}
3106: 
3107: For $d=1$, $E_c\in[\frac{1+\e}{1-\e},\frac2{1-\e}]$ we already
3108: have $\bar\tau\sim N=L^1$ as the theorem states, but $\bar s_0\sim
3109: N=L^1$, $N\to\infty$, which shows in the last respect the critical
3110: energy $E_c=2/(1-\e)$ is small.
3111: 
3112: In \cite{BCK} the estimate $\gamma_\tau>1$ was predicted for all
3113: $d$, while this is a feature of the cases $d>1$. In the latter
3114: cases the analytic calculation of exact values of exponents
3115: $\gamma_\tau,\gamma_s$ is a difficult problem.
3116: 
3117: \begin{rk}\po
3118: The difference between cases $d=1$ and $d>1$ demonstrated in the
3119: theorem is known in the physical literature. The former case is
3120: usually considered as the trivial SOC-model.
3121: \end{rk}
3122: 
3123: 
3124: %%%%%%%%%%%%%%%%%%
3125: %%%%%%%%%%%%%%%%%%
3126: 
3127: \subsection{\hpss Thermodynamic limit}\label{sect13}
3128: 
3129: By thermodynamic limit of an observable $\phi$ we understand the
3130: double limit
3131: $$
3132: [\phi]_\infty=\lim_{\begin{array}{c} \scriptstyle L\to\infty\vspace{-5pt}\\
3133: \scriptstyle E_c\to\infty\end{array}}\phi
3134: $$
3135: if it exists. It is assumed that for physical observables this
3136: limit exists. In \cite{BCK} only limit $L\to\infty$ was
3137: considered, though then the value of energy $E_c$ could serve as
3138: an essential parameter, which is not desirable in the
3139: SOC-paradigm. However it was suggested there that consideration of
3140: $E_c\to\infty$ can be helpful.
3141: 
3142: As an example of non-physical observable we expose $s_0$ (in
3143: \S\ref{Ssec} we called it mathematically relevant): The double
3144: limit does not exists because the repeated limits are different:
3145:  $$
3146: 0=\lim_{L\to\infty}\lim_{E_c\to\infty}\bar s_0\ne
3147: \lim_{E_c\to\infty}\lim_{L\to\infty}\bar s_0=+\infty.
3148:  $$
3149: But $s_+$ and $\tau$ are good physical observables, for the
3150: thermodynamic limits exist:
3151:  $$
3152: [\bar\tau]_\infty=\infty,\quad [\bar s_+ ]_\infty=\infty.
3153:  $$
3154: From \S\ref{305} (use iteration arguments in the topological case)
3155: and \S\ref{evalofentr} we obtain:
3156:  $$
3157: [\h(\hat f)]_\infty=[\h(F)/\bar\tau]_\infty=0 \quad\text{ and }
3158: \quad[\h(F)]_\infty=\infty.
3159:  $$
3160: 
3161: But with $\omega$ the situation is different because the proof
3162: (rather than the vague statement of part (ii)) of Theorem
3163: \ref{tyh} implies:
3164: $\lim_{L\to\infty}\lim_{E_c\to\infty}\bar\omega=\infty$, while
3165: $\lim_{E_c\to\infty}\lim_{L\to\infty}\bar\omega$ is finite. Since
3166: $\omega$ is definitely physically relevant observable one needs
3167: the following reparametrization: $\omega\mapsto\omega/E_c$, which
3168: corresponds to contraction of the waiting time via the following
3169: ansatz:
3170: 
3171: We let the energy quantum added at a unit time to the system equal
3172: $\delta=\hbar$ (instead of 1 as before), but speed up time in the
3173: waiting intervals respectively: $E_c=E_0/\hbar$,
3174: $\omega_\text{new}=\omega\hbar$. Then the thermodynamic limit (the
3175: space part $M$ can be quantized similarly via $L=[l/\hbar]$ with
3176: $l$ a finite length) corresponds to the quasi-classical limit
3177: $\hbar\to0$.
3178: 
3179: The duration of avalanche was suppressed in our definition of
3180: dynamics to length one. So if we want to find the entropy of the
3181: physical system, where each step of avalanche has time-duration
3182: one, we should multiply it by the probability of dropping energy
3183: into the system. For every trajectory this equals
3184: $\frac{\bar\omega}{\bar\omega+\bar\tau}$. This ratio behaves
3185: differently as $L\to\infty$ or $E_c\to\infty$, so we need the
3186: reparametrization described above.
3187: 
3188: In \S \ref{evalofentr} we calculated the entropy of the "return"
3189: Zhang model $\h(F)=\log N$ (this was proved almost surely, but
3190: even with the possible entropy growth for some exceptional
3191: parameters the thermodynamic limit below is unaltered). But after
3192: reparametrization it changes. Denoting by $h^\text{Zhang}$ the
3193: entropy of the reparametrized system we get:
3194:  $$
3195: h_\mu^\text{Zhang}=h_{\mu^u}(\sigma_N^+)\cdot
3196: \langle\frac{\bar\omega_\text{new}}{\bar\omega_\text{new}
3197: +\bar\tau}\rangle,\quad
3198:  \h^\text{Zhang}=d\cdot\log L\cdot
3199: \bigl(\frac{\bar\omega_\text{new}}{\bar\omega_\text{new}
3200: +\bar\tau}\bigr)_{\text{max}}.
3201:  $$
3202: This implies:
3203:  $$
3204: [h_\mu^\text{Zhang}]_\infty=[\h^\text{Zhang}]_\infty=0.
3205:  $$
3206: Therefore the expanding property is lost in the thermodynamic
3207: limit for the original physical system, as was already noticed in
3208: \cite{BCK} for a bit different situation.
3209: 
3210: 
3211: 
3212: \begin{rk}\po
3213: Notice that in the reparametrized system
3214: $\bar\omega_\text{new}\ll\bar\tau$, which is counter-intuitive for
3215: certain SOC-examples (sandpile, earthquakes etc, where one expects
3216: $\bar\omega\gg\bar\tau$). This indicates that the Zhang system
3217: should be modified by introducing the local contraction of time
3218: depending on the avalanche size or speed. We will not consider
3219: such gradient-type models here.
3220:  \end{rk}
3221: 
3222: 
3223: 
3224: %%%%%%%%%%%%%%%%%%
3225: %%%%%%%%%%%%%%%%%%
3226: 
3227: \subsection{\hpss Lyapunov spectrum}\label{L.S.}
3228: 
3229: In \S \ref{hypstr} we showed that the Zhang model is hyperbolic with one
3230: positive exponent and the remainder of the spectrum negative. This hyperbolicity
3231: is lost in the thermodynamic limit.
3232: \begin{prop}\po
3233: For $E_c \geq \e/(1-\e)$ we have
3234: $\sum_{i=1}^N\chi_i^-=\bar s_0\log\e$.
3235: \end{prop}
3236: \begin{proof}
3237: We cannot ensure the existence of an SRB-measure, so
3238:  both the Lyapunov exponents
3239: and $\overline{s}_0$ should be seen as functions on $M$. From the
3240: general theory of Lyapunov exponents we know that
3241: $$
3242: \sum_{i=1}^N \chi_i^-(\hat{x}) =
3243: \lim_{n \rightarrow \infty} \frac1n \sum_{t=1}^{n-1}
3244: \log \det {\mathcal T}(F^t\hat{x})\,.
3245: $$
3246: From Proposition \ref{7} we know that the formula
3247: $\det(L_{ij})=\e^{s_{ij}}$ holds for $E_c \geq \e/(1-\e)$. Hence
3248: \begin{eqnarray*}
3249: \sum_{i=1}^N \chi_i^-(x)&=&
3250: \sum_{i=1}^N \int_{\Sigma_N^+}
3251: \chi_i^-({\bf t},x)
3252: \,d\mu_{\text{{\tiny Ber}}} \\
3253: &=&\int_{\Sigma_N^+} \sum_{i=1}^N
3254: \chi_i^-({\bf t},x)
3255: \,d\mu_{\text{{\tiny Ber}}} \\
3256: &=& \int_{\Sigma_N^+} \lim_{n \rightarrow \infty}
3257: \frac{1}{n} \sum_{t=0}^{n-1} \log (\det {\mathcal T})(F^t({\bf t},x))\,
3258: d\mu_{\text{{\tiny Ber}}} \\
3259: &=& \int_{\Sigma_N^+} \lim_{n \rightarrow \infty}
3260: \frac{1}{n} \sum_{t=0}^{n-1} s({\bf t},x)\,d\mu_{\text{{\tiny Ber}}} \log \e \\
3261: &=& \overline{s}_0(x) \log \e\,.
3262: \end{eqnarray*}
3263: \end{proof}
3264: 
3265: 
3266: 
3267:  \begin{cor}\po
3268: $|\overline{\chi\mathstrut}{^-}|=\frac{\bar s_0}N\log\frac1\e\to
3269: 0$ as $E_c\to\infty$, but $|\overline{\chi\mathstrut}{^-}|\to
3270: \infty$ as $L\to\infty$.
3271: \end{cor}
3272: ~ \\
3273: \noindent
3274: Thus we should study differently the following cases:
3275: ~\\
3276: 
3277: \noindent
3278: {\bf 1. $E_c\to\infty$, but $L$ (and $d$) fixed.} Since
3279: $\lim_{E_c\to\infty}\bar s_0=0$, the negative part of the Lyapunov
3280: spectrum collapses: $\lim_{E_c\to\infty}\chi^-_i=0$ for every
3281: $1\le i\le N$. The hyperbolicity is lost, but the positive
3282: exponent $\chi^+_0=\log N$ survives. In particular, the entropy
3283: does not collapses. \\
3284: 
3285: \noindent
3286: {\bf 2. $L\to\infty$, but $E_c\gg1$ fixed.} Here only a bounded
3287: piece $\chi_1^-,\dots,\chi_k^-$ of the Lyapunov spectrum
3288: collapses, $k=\op{const}$. But the number of elements of this
3289: spectrum grows and in average $|\overline{\chi\mathstrut}{^-}|\to
3290: \infty$. In particular, $|\chi^-|_\text{max}\to \infty$ as
3291: $L\to\infty$. Again, the positive exponent $\chi_0$ and entropy
3292: are preserved and though we loose hyperbolicity there are many
3293: non-degenerate Oscelledec modes. Moreover they prevail over
3294: collapsing modes and so essentially the hyperbolicity is preserved
3295: as well. \\
3296: 
3297: \noindent
3298: {\bf 3. Reparametrized model.}
3299: This was introduced in \S\ref{sect13} and require the
3300: renormalization: multiplication of waiting time by the function
3301: $\frac{\bar\omega}{\bar\omega+\bar\tau}$ along the trajectory. In
3302: this case the Lyapunov spectrum collapses to zero in any limit
3303: $E_c\to\infty$ and $L\to\infty$ and the hyperbolicity is
3304: completely lost. \\
3305: 
3306: Thus the exponential grow of the statistics is suppressed and we
3307: can observe power law statistic as is the basic idea of
3308: SOC-phenomenon. The corresponding SOC-exponents are related to the
3309: asymptotic of the Lyapunov spectrum (as discussed in \cite{BCK}),
3310: but are difficult to calculate analytically.
3311: 
3312: 
3313: %%%%%%%%%%%%%%%%%%%%%%%%
3314: %%%%%%%%%%%%%%%%%%%%%%%%
3315: 
3316: \section{\hps Conclusion}
3317: 
3318: It is of importance to the paradigm of SOC to understand the
3319: mechanisms behind the behavior one observes numerically in the
3320: Zhang model and other sandpile models, and the aim of this paper
3321: is to provide a first step to a rigorous mathematical
3322: understanding of the dynamics of Zhang model.
3323: 
3324: Due to the singularities and non-invertibility of the model there
3325: existed very few applicable results, and hence we had to modify
3326: known results and develop some new methods in order to describe
3327: the dynamical properties of the model.
3328: 
3329: The result of this work is that the singularities play a modest
3330: role in the sense they do not change the main dynamical
3331: characteristics. However the effects of the singularities can be
3332: seen in the rich fractal structure of the spacial attractor.
3333: 
3334: Our analysis allows to take the thermodynamic limit of the main
3335: dynamical quantities, showing that the entropy vanishes and the
3336: Lyapunov-spectrum collapses after re-scaling the model. From a
3337: physical point of view this is interesting. Typically chaotic
3338: dynamics (positive entropy and positive Lyapunov exponents) is an
3339: indication of exponential speeds of mixing (short decay of
3340: correlations) \cite{Ba}, which is not compatible with the
3341: power-law statistics of the SOC-hypothesis. The loss of
3342: hyperbolicity in the thermodynamic limit hence supports the
3343: critical behavior observed numerically \cite{J}, \cite{Z},
3344: \cite{GD}.
3345: 
3346: To conclude: We have shown that the Zhang model is a chaotic
3347: hyperbolic dynamical system, where all the entropy is produced by
3348: the random driving of the system and, due to singularities, the
3349: orbit structure is richer than for a topological Markov chain. The
3350: hyperbolicity indicates that under weak conditions there exists an
3351: SRB-measure (self-organization). In the thermodynamic limit the
3352: hyperbolicity is lost and we may expect power-law statistics
3353: (criticality). In practice the systems that are studied have
3354: finite size and finite critical energy. Hence they are chaotic,
3355: but with small entropy. The SOC-hypothesis is that these weakly
3356: chaotic systems have SRB-measures with the rates of convergence to
3357: these measures being exponential but slow compared to a unit step
3358: in an avalanche, causing the prevailing of power-law statistics.
3359: 
3360: Finally note that modifications of the Zhang model are possible.
3361: For instance, different amounts of energy $\delta_i$ can be added
3362: to different sites $i\in\Lambda$ in the excitation process. This
3363: corresponds to the rectangular form of $\Lambda\subset\Z^d$ and
3364: uniform quantum of energy $\delta=\hbar$. We can also consider
3365: other spacial configurations $\Lambda$. In this way nearly Zhang
3366: models considered in Section \ref{evalofentr} are natural. Another
3367: approach is to use a random amount of energy, i.e. stochastic
3368: $\delta:\Lambda\to\R_+$ (this idea was used in the plasma physics
3369: \cite{KK}), which can also be presented as a skew product with
3370: piecewise affine fibers, but now over a solenoidal system. In all
3371: these theories the ideas from the present paper work well (though
3372: the thermodynamic limit will be sensitive to the form of
3373: $\Lambda$).
3374: 
3375: 
3376: %%%%%%%%%%%%%%%%%%%%%%%%
3377: %%%%%%%%%%%%%%%%%%%%%%%%
3378: 
3379: \appendix
3380: 
3381: 
3382: %%%%%%%%%%%%%%%%%%%%%%%%
3383: %%%%%%%%%%%%%%%%%%%%%%%%
3384: 
3385: \section{Topological entropy of piecewise affine maps}\label{app_A}
3386: 
3387: 
3388: \subsection{\hpss The Buzzi theorem and its generalization}\label{app_A.1}
3389: The statement as it is done in \cite{B1} does not apply to our
3390: situation. In \cite{B2} Buzzi noted that it extends to isometries
3391: and contractions. In fact, the assertion holds always, but since
3392: we cannot make a simple reference we write an adapted proof for
3393: the convenience of the reader.
3394: 
3395:  \begin{theorem} \po \label{buzzi}
3396: Let $X\subset \R^d$ be a bounded polytope and $f: X \rightarrow X$
3397: a piecewise affine map. Then
3398:  $$
3399: \hs(f) \leq \lambda^+(f) + H_{\text{{\em mult}}}(f)\,,
3400:  $$
3401: where
3402:  $$
3403: \lambda^+(f)=\mathop{\overline\lim}\limits_{n \rightarrow \infty}
3404: \sup_{x \in X} \frac{1}{n}\max_k \log \|\Lambda^k d_x f^n\|\,.
3405:  $$
3406:  \end{theorem}
3407: \begin{proof}
3408: Let ${\mathcal P}=\{P_i\}$ be the continuity partition and
3409: $f_i:=f|_{P_i}$ be affine maps. Fix $\e>0$ and
3410: $T=T(\e)\ge\frac{d}\e\log(\sqrt{d}+1)$ such that for $n\ge T$ we
3411: have:
3412:  $$
3413: \text{mult}({\mathcal P}^n)\leq \exp((H_{\text{mult}}(f)+\e)n),\
3414: \|\Lambda df^n\| \leq \exp((\lambda^+(f)+\e)n).
3415:  $$
3416: Take $r=r(\e)$ to be compatible with the partition ${\mathcal
3417: P}^T$ (for $f^T$), i.e. any $r$-ball intersects maximally
3418: $\op{mult}(\mathcal{P}^T)$ partition elements.
3419: 
3420: We will prove that each non-empty cylinder $C({\bf
3421: a})=[P_{a_0}\dots P_{a_{lT-1}}]$ of length $|{\bf a|}=lT$ can be
3422: partitioned into a collection $Q({\bf a})=\{W\}$ satisfying the
3423: following properties:
3424: \begin{enumerate}
3425: \item $\sum_{|{\bf a}|=lT} \text{card}(Q({\bf a}))
3426: \leq C_0 \exp((\lambda^+(f)+H_{\text{mult}}(f)+3\e)lT)$
3427: \item $\text{diam}(f_{a_{lT-1}}\circ \dots \circ f_{a_0}(W)) \leq r\,.$
3428: \end{enumerate}
3429: 
3430: Let us prove this claim by induction assuming it holds for some
3431: $l\ge0$. The base of induction is obvious and $C_0$ is the minimal
3432: cardinality of an $r/2$-ball cover.
3433: 
3434: Take a partition element $W \in Q({\bf a})$ that is used to cover
3435: the cylinder $[P_{a_0}\dots P_{a_{lT-1}}]$. By the induction
3436: hypothesis it has diameter less than $r$, so it can be continued
3437: to cover a non-empty cylinder of length $(l+1)T$ in at most
3438: $\text{mult}({\mathcal P}^T)$ ways. So to cover the cylinders
3439: $[P_{a_0}\dots P_{a_{lT-1}}P_{b_0} \dots P_{b_{T-1}}]$ we make a
3440: division of $W$:
3441:  $$
3442: W=\bigcup_{i=1}^\gamma W'_i\,\,,\gamma \leq \text{mult}({\mathcal
3443: P}^T)\,.
3444:  $$
3445: Let
3446:  $$
3447: W''_i=f_{a_{Tl-1}}\circ\dots \circ f_{a_0}(W'_i)
3448:  $$
3449: and
3450:  $$
3451: W'''_i=f_{b_{T-1}}\circ\dots \circ f_{b_0}(W''_i)\,.
3452:  $$
3453: By the assumption $\text{diam}(W''_i)<r$ for all
3454: $i=1,\dots,\gamma$, but the sets $W'''_i$ may have greater
3455: diameter. We need to divide the sets $W'''_i$ so that they have
3456: diameter less than $r$, and then pull this refinement back to the
3457: partition of sets $W'_i$.
3458: 
3459: Let $L$ be the differential of $f_{b_{T-1}} \circ \dots \circ
3460: f_{b_0}$ on $W_i''$. We can assume that $L$ is symmetric and take
3461: $\{e_k\}$ to be a basis of eigenvectors corresponding to
3462: eigenvalues $\lambda_1,\dots,\lambda_d$. Let $\{v_k\}$ be a basis
3463: in the vector subspace corresponding to $W'''_i$. We can choose
3464: this basis to be orthonormal and triangular with respect to
3465: $\{e_k\}$. Divide $W'''_i$ by the hyperplanes
3466:  $$
3467: \psi_j(x)\stackrel{\text{def}}=\langle v_j,x\rangle=p
3468: \frac{r}{\sqrt{d}},\quad p\in \Z,\quad j=1,\dots,d\,.
3469:  $$
3470: This defines cells $\tilde W$ of diameter less than $r$. Since
3471: $\psi_j(W'''_i)=\psi_j(L(W''_i))$ has $\op{diam}\le|\lambda_i|r$,
3472: the number of cells $\tilde W$ needed to cover $W'''_i$ is less
3473: than or equal to
3474:  $$
3475: (\sqrt{d}+1)^d|\lambda_1|^+ \dots |\lambda_d|^+ \leq
3476: (\sqrt{d}+1)^d \|\Lambda d f^n\| \leq \exp \big{(}(\lambda^+(f) +
3477: 2 \e)T \big{)}\,,
3478:  $$
3479: where $|\lambda_i|^+=\max\{|\l_i|,1\}$.
3480: 
3481: Therefore the total cardinality of the new partition is less than
3482: or equal to
3483:  \begin{eqnarray*}
3484: \op{mult}({\mathcal P}^T)\exp \big{(}(\lambda^+(f)+2\e)T\big{)}
3485: \exp \big{(}(\lambda^+(f)+H_{\text{mult}}(f)+3\e)lT \big{)}  \\
3486: \leq \exp \big{(}(\lambda^+(f)+H_{\op{mult}}(f)+3\e)(l+1)T
3487: \big{)}\,.
3488:  \end{eqnarray*}
3489: This proves the statement.
3490: \end{proof} \\
3491: 
3492: The theorem holds as well for most degenerate piece-wise affine
3493: systems, but there can be problems with $\hm(f)$. Namely the
3494: latter is not defined if the image of continuity domain contains a
3495: boundary face of a continuity domain. But if we assume the
3496: image and the faces always meet transversally, no problems occur
3497: and the above theorem applies literally.
3498: 
3499: In degenerate cases of the Zhang model, the above requirement
3500: holds for most parameters. For instance, if $N=2$ and $\e=0$, then
3501: all $E_c\notin\Z$ satisfy the request. For integer $E_c$ the
3502: theory fails, but one can look just to the whole image set, which
3503: has dimension $<N$: its Poincar\'e return map is piece-wise affine
3504: and it satisfies the requirements. In the above example $N=2$,
3505: $\e=0$ the dynamics is confined to two one dimensional lines,
3506: whence the multiplicity entropy is zero (by dimensional reasons)
3507: and $\h(F)=\log 2$ in this case.
3508: 
3509: %%******************************************************************
3510: %%******************************************************************
3511: 
3512: \subsection{\hpss Entropy of conformal piecewise affine skew-products}\label{app_A.2}
3513: 
3514: 
3515: We say that an affine map is conformal modulo degenerations if the
3516: image on some subspace transversal to the kernel is mapped
3517: conformally to its image. A piecewise affine map is said to be
3518: conformal modulo degenerations if all its affine components are
3519: conformal modulo degenerations. We will assume that degenerations
3520: satisfy the transversality requirement of \ref{app_A.1}.
3521: 
3522: In \cite{KR1} we noticed that $\hm=0$ for piece-wise affine
3523: conformal maps. This easily extends to allow degenerations. Now we
3524: consider a more general situation of skew-product systems of
3525: Zhang's type.
3526: 
3527:  \begin{theorem}\po \label{14}
3528: Let $f_i:X\to X$ be piece-wise affine non-strictly contracting and
3529: conformal modulo degenerations, $i=0,\dots,N-1$. Define
3530: $F:\Sigma_N^+ \times X\to\Sigma_N^+ \times X$ by the formula
3531: $F({\bf t},x)=(\sigma_N^+({\bf t}),f_{t_0}(x))$, ${\bf
3532: t}=t_0t_1\dots\in\Sigma_N^+$, $x\in X\subset \R^d$. Then we have:
3533: $\h(F)=\log N$ (so that a-posteriori the variational principle
3534: holds).
3535:  \end{theorem}
3536: 
3537:  \begin{rk}\po
3538: For $N=2$ and $\e=0$ the affine components have rank 1, and hence
3539: Theorem \ref{14} shows $\h(F)=\log N$. The same holds for
3540: $E_c=\e=1/3$.
3541:  \end{rk}
3542: 
3543: \noindent
3544:  \begin{proof}
3545: It follows from Theorem \ref{buzzi} that it suffices to prove that
3546: $\hm(F)=0$.
3547: 
3548: To achieve the desired equality note that preimages of the
3549: time-like singularity planes $\{t=n/N\}$ never intersect under
3550: inverse iterations of $F(t,x)=(Nt \mod 1,f_{[Nt]}(x))$. Thus
3551: $\hm(F)=\sup_{\bf t}\hm(f_{\bf t})$. We will prove that
3552: $\hm(f_{\bf t})=0$ for all ${\bf t}\in\Sigma_N^+$.
3553: 
3554: A piecewise affine map can be considered as an ordered triple
3555: $(X,{\mathcal P},f)$, where $X$ is a polytope in $\R^d$,
3556: ${\mathcal P}=\{P_i\}$ is a partition of $X$ made up of pairwise
3557: disjoined polytopes (with certain faces of boundary included, so
3558: that the whole boundary is distributed between polytopes) and
3559: $f_i:=f|_{P_i}:P_i \rightarrow X$ are affine maps. We let
3560: $X'=\cup\op{Int}(P_i)$ and $\op{Sing}(f)=X\setminus X'$.
3561: 
3562: For a piecewise affine map $(X,{\mathcal P},f)$ and a point $x \in
3563: X$ construct a piecewise affine map $(X_x,{\mathcal P}_x, f_x)$,
3564: called the differential of $f$ at $x$, by letting
3565: \begin{enumerate}
3566: \item $X_x=\{ y \in \R^d \,|\, \exists \e_0>0 \,\text{s.t.}
3567: \forall \e \in (0,\e_0):\,x+\e y \in X\}\subset\R^d$ is the
3568: tangent cone to $X$.
3569: 
3570: \item ${\mathcal P}_x$ is the partition of $X_x$
3571: consisting of non-empty sets
3572:  $$
3573: P_x=\{ y \in \R^d \,|\, \exists \e_0>0 \,\text{s.t.} \forall \e
3574: \in (0,\e_0):\,x+\e y \in P\}\,,
3575:  $$
3576: where $P \in {\mathcal P}$.
3577: 
3578: \item $f_x:P_x \rightarrow X_{f(x)}$ is the collection of maps
3579:  $$
3580: f_x(y)= \lim_{\e \rightarrow 0^+} \frac{f(x+\e y)- \lim_{\delta
3581: \rightarrow 0^+} f(x+\delta y)}{\e}\,,\quad P_x\in{\mathcal
3582: P}_x\,.
3583:  $$
3584: \end{enumerate}
3585: 
3586: Consider iterated differentials and denote
3587: $f_{x_1,\dots,x_n}:=(\dots (f_{x_1})_{x_2}\dots)_{x_n}$. Note that
3588: $(f^n_{\bf t})_x=(f_{t_{n-1}})_{f_{\bf t}^n(x)}\circ
3589: \dots(f_{t_1})_{f^1_{\bf t}(x)}\circ(f_{t_0})_x$.
3590: 
3591:  % The desired claim will follow from the following analog of
3592:  % Proposition 13 \cite{B2}:
3593: 
3594:  \begin{prop}\po \label{15}
3595: There exists a constant $C\in\R_+$ such that for any subspace
3596: $W\subset\R^d$ and any $(x_1,\dots,x_r)\in X\times W^{r-1}$ we
3597: have:
3598: $$
3599: \op{mult}((f^n_{\bf t}|_W)_{x_1,\dots,x_r})\le
3600: C\,\Bigl(\sup_{V\subset\R^d}\sup_{y_1,\dots,y_{r+1}}\op{mult}
3601: (f_{\bf t}|_V)_{y_1,\dots,y_{r+1}}\Bigr)^n,\qquad \forall n\ge0,
3602: $$
3603: where the collection of points $(y_1,\dots,y_{r+1})$ runs over
3604: $X\times V^r$ with the condition
3605: $\op{rank}(y_2,\dots,y_{r+1})+\op{codim}V
3606: =\op{rank}(x_2,\dots,x_r)+\op{codim}W+1$.
3607:  \end{prop}
3608: 
3609: The theorem follows from this, because for
3610:  $$
3611: \mu(r)=\mathop{\overline\lim}\limits_{n\to\infty}\frac1n
3612: \log\hspace{-25pt} \sup_{\begin{array}{c}
3613: \scriptstyle y_1\in X,y_2\dots,y_r\in V\subset\R^d\\
3614: \scriptstyle \op{rank}(y_2,\dots,y_r)+\op{codim}V=r-1
3615:  \end{array}}\hspace{-25pt}
3616:  \op{mult}(f_{\bf t}^n|_V)_{y_1,\dots,y_r}
3617:  $$
3618: we have: $\hs(f_{\bf t})=\mu(1)\le\mu(2)\le\dots\le\mu(d+1)=0$.
3619: 
3620: To prove the proposition note that when $r=0$ we have (in this
3621: case we do not need $W,V$):
3622: $$
3623: \op{mult}(f_{\bf t}^n)\le\Bigl(\sup_{x\in X} \op{mult}(f_{\bf
3624: t})_x\Bigr)^n=\Bigl(\max_{0\le j<N}\sup_{x\in X}
3625: \op{mult}(f_j)_x\Bigr)^n.
3626: $$
3627: Let $r\ge1$. Consider the continuity partition $\mathcal{P}^{\bf
3628: t}_{x_1,\dots,x_r}$ for $(f_{\bf t})_{x_1,\dots,x_r}$, which is
3629: just the continuity partition
3630: $\mathcal{P}^{(t_0)}_{x_1,\dots,x_r}$ of
3631: $(f_{t_0})_{x_1,\dots,x_r}$ (one iteration), and let
3632: $\mathcal{P}^{n,\bf t}_{x_1,\dots,x_r}$ for $(f^n_{\bf
3633: t})_{x_1,\dots,x_r}$ be the iterated partition. Note that the
3634: latter is the collection of all non-empty intersections
3635: $$
3636: P^{t_0}\cap (f_{t_0})_{P^{t_0}}^{-1}(P^{t_1})\cap
3637: (f_{t_0})_{P^{t_0}}^{-1}(f_{t_1})_{P^{t_1}}^{-1}(P^{t_2})\cap\dots
3638: \cap(f_{t_0})_{P^{t_0}}^{-1}\dots(f_{t_{n-1}})_{P^{t_{n-1}}}^{-1}(P^{t_n}),
3639: $$
3640: where $P^{t_i}$ are elements of
3641: $\mathcal{P}^{(t_i)}_{x_1,\dots,x_r}$ and $f_P$ denotes the
3642: restriction of the (differential of the) map to the corresponding
3643: continuity domain. Every element of these partitions is invariant
3644: under the shift by vectors from $\op{span}(x_2,\dots,x_r)$.
3645: Therefore it intersects the unit sphere $S_1(x_2,\dots,x_r)^\perp$
3646: in the orthogonal complement. Consider the induced partition on
3647: the sphere and refine it so that every element has diameter no
3648: greater than $\varepsilon$. Denote by $n(\mathcal{P}^{n,\bf
3649: t}_{x_1,\dots,x_r}, \varepsilon)$ the minimal cardinality of such
3650: a refinement. Let also $m(\mathcal{P}^{(t_i)}_{x_1,\dots,x_r},
3651: \varepsilon)$ be the maximal number of elements of
3652: $\mathcal{P}^{(t_i)}_{x_1,\dots,x_r}$ that an $\varepsilon$-ball
3653: $B(y,\varepsilon)\cap S_1^\perp$ of $S_1(x_2,\dots,x_r)^\perp$ can
3654: meet.
3655: 
3656: Denote by $n(\mathcal{P}\cap W,\varepsilon)$, $m(\mathcal{P}\cap
3657: W,\varepsilon)$ the corresponding quantities in the subspace $W$.
3658: Then from the above formula for the iterated partition:
3659:  $$
3660: n(\mathcal{P}^{n+1,\bf t}_{x_1,\dots,x_r}\cap W, \varepsilon)\le
3661: n(\mathcal{P}^{n,\bf t}_{x_1,\dots,x_r}\cap W, \varepsilon)\cdot
3662: m(\mathcal{P}^{(t_{n+1})}_{y_1,\dots,y_r}\cap V, \varepsilon),
3663:  $$
3664: where $y_1=f_{\bf t}^n(x_1)$, $y_2=f_{\bf t}^n{}'(x_2),\dots$,
3665: $y_r=f_{\bf t}^n{}'(x_r)$ and $V=f_{\bf t}^n{}'(W)$ with $f_{\bf
3666: t}^n{}'=(f_{\bf t}^n)_{x_1,\dots, x_r}$. Therefore
3667:  $$
3668: n(\mathcal{P}^{n+1,\bf t}_{x_1,\dots,x_r}\cap W, \varepsilon)\le
3669: n(\mathcal{P}^{n,\bf t}_{x_1,\dots,x_r}\cap W, \varepsilon)\cdot
3670: \sup_V\sup_{y_1,\dots,y_r}m(\mathcal{P}^{(t_{n+1})}_{y_1,\dots,y_r}\cap
3671: V,\varepsilon),
3672:  $$
3673: where the supremum is taken over all $V\subset\R^d$ and
3674: $(y_1,\dots,y_r)\in X\times V^{r-1}$ such that codimension of
3675: $\langle y_2\dots,y_r\rangle$ in $V$ equals codimension of
3676: $\langle x_2\dots,x_r\rangle$ in $W$.
3677: 
3678: Since for a fixed $\varepsilon$ the number
3679: $n(\mathcal{P}^{t_0}_{x_1,\dots,x_r}, \varepsilon)$ is finite and
3680: $$
3681: m(\mathcal{P}^{(t_i)}_{y_1,\dots,y_r}\cap V,\varepsilon)\le
3682: \sup_{y\in S_1(y_1,\dots,y_r)^\perp\cap V}
3683: |\mathcal{P}^{(t_i)}_{y_1,\dots,y_r}\cap B(y,\varepsilon)\cap V|,
3684: $$
3685: the claim follows from the following statement. Fix $i\in[0,N)$.
3686: 
3687:  \begin{lem}\po\label{16}
3688: There exists $\varepsilon>0$ (depending only on $i$) such that for
3689: all $V\subset\R^d$ and all $(y_1,\dots,y_r)\in X\times V^{r-1}$,
3690: with $\op{rank}(y_2,\dots,y_r)<\dim V$, and $y\in
3691: S_1(y_2,\dots,y_r)^\perp\cap V$ there exists
3692: $y'\in(y_2,\dots,y_r)^\perp\cap V$ satisfying:
3693:  $$
3694: |\mathcal{P}^{(i)}_{y_1,\dots,y_r}\cap B(y,\varepsilon)\cap V|\le
3695: \op{mult}((f_{i}|_V)_{y_1,\dots,y_r,y'}).
3696:  $$
3697:  \end{lem}
3698: This statement, modulo our notations and restrictions to $V$, is
3699: proved in \cite{B2}. The proposition and hence the theorem follow.
3700: \end{proof}
3701: 
3702: 
3703: %%*************************************************************************
3704: 
3705: \subsection{\hpss Estimates on entropy by angular expansion rates}\label{app_A.3}
3706: 
3707: It is possible to estimate the effect of angular expansion on the
3708: topological entropy of a piecewise affine map $f:X \rightarrow X$
3709: by its spherizations. Define the piecewise smooth map
3710: $d_x^{(s)}\!f:S T_x X \to S T_{f(x)}X$ given at $x\in X'$ by the
3711: formula
3712:  $$
3713: d_x^{(s)}\!f(v)=\frac{d_x f(v)}{\|d_x f(v)\|}\,.
3714:  $$
3715: For $x\in\op{Sing}(f)$ and $v\not\in T_x\op{Sing}(f)$ (the tangent
3716: cone) we let $d_x^{(s)}\!f(v)=\lim\limits_{\e\to+0}d_{x+\e
3717: v}^{(s)} f(v)$. For other $(x,v)\in STX$ the map is not defined.
3718: The angular expansion of $f$ is exactly the expansion in the
3719: fibers of its spherization.
3720: 
3721: If $d_x f$ is degenerate we restrict to the orthogonal component
3722: of its kernel, and consider the map
3723:  $$
3724: d_x f|_{\op{Ker}(d_x f)^\bot}:\op{Ker}(d_x f)^\bot \rightarrow
3725: \op{Im}(d_x f)\,.
3726:  $$
3727: Then the map $S_x(f)=d_x^{(s)} f|_{\op{Ker}(d_x f)^\bot}:S
3728: \op{Ker}(d_x f)^\bot \rightarrow S \op{Im}(d_x f)$ between
3729: $(\op{rank}(d_x f)-1)$-dimensional spheres is given by the formula
3730:  $$
3731: v \mapsto \frac{d_x f|_{\op{Ker}(d_x f)^\bot}(v)}{\|d_x
3732: f|_{\op{Ker}(d_x f)^\bot}(v)\|}.
3733:  $$
3734: 
3735: For $i<d$ we define
3736:  $$
3737: \rho_i(f)=\ls \frac1n \sup_{(x,v)} \max_{0 \leq k \leq i} \log \|
3738: \Lambda^k d_vS_x(f^n) \|.
3739:  $$
3740: Let $m_*=\min_x \op{dim} \op{Ker} (d_x f)$ and $d_*=d-m_*=\max_x
3741: \op{rank} (d_x f)$, where $d$ is the dimension of $X$. The numbers
3742: $\rho_i(f)$ can be non-zero only for $i<d_*$.
3743: 
3744: We have: $\rho_0(f)=0$. The number $\rho_{1}(f)$ measures the
3745: maximal exponential rate with which angles can increase under the
3746: map $f$. The numbers $\rho_i(f)$ for $i<d$ measure the maximal
3747: rate of expansion of the restrictions to $i$-dimensional spheres.
3748: If $f$ is conformal, then $\rho_i(f)=0$ for all $i$.
3749: 
3750: \begin{theorem}\!\!{\bf (\cite{KR2}).} \label{th2}
3751: For piece-wise affine maps $\hm(f)\le\sum_{i=1}^{d_*-1}
3752: \rho_i(f)$.
3753: \end{theorem}
3754: 
3755: We define the maximal expansion rate
3756: $$
3757: \lambda_{\op{max}}(f)=\ls \sup_{x} \frac1n \log \|d_x f^n\|\,,
3758: $$
3759: and the minimal finite expansion rate
3760: $$
3761: \lambda_{\op{min}}(f)=-\ls \sup_{x} \frac{1}{n} \log \| (d_x
3762: f^n|_{\op{Ker}(d_x f^n)^\bot})^{-1}\|\,.
3763: $$
3764: In \cite{KR2} we show that
3765: $$
3766: \rho_i(f)\leq i \Big{(}\lambda_{\op{max}}(f)-\lambda_{\op{min}}(f)
3767: \Big{)}\,.
3768: $$
3769: This gives the following result (the same bound holds for
3770: $\h(f)$):
3771: \begin{theorem}\po \label{thnew}
3772: For a piecewise affine map $f$ it holds:
3773: $$
3774: \hs(f) \leq \lambda^+(f)+\frac{d_* (d_*-1)}{2}
3775: \Big{(}\lambda_{\op{max}}(f)-\lambda_{\op{min}}(f) \Big{)}\,.
3776: $$
3777: \end{theorem}
3778: 
3779: %%%%%%%%%%%%%%%%%%%%%%%%
3780: %%%%%%%%%%%%%%%%%%%%%%%%
3781: 
3782: \section{\hps Generalization of the Moran formula}\label{app_B}
3783: 
3784: Consider an IFS $(M^d,f_1,\dots,f_N)$ on a Riemannian manifold
3785: $M$, where the maps $f_i$ can possess singularities, but we assume
3786: that they are mild in a sense that the number of continuity
3787: domains is finite, any of them has piece-wise smooth boundary and
3788: the map, restricted to any of the domains, smoothly extends to the
3789: adjacent singularities (this is the case of the Zhang model).
3790: 
3791: Remark that the IFS can be interpreted as the dynamical system
3792: $(\Sigma_N^+\times M,F)$, $\hat f({\bf t},x)=(\sigma_N^+{\bf
3793: t},f_{t_0}x)$. Attractor $\mathcal{Y}$ of the IFS can be defined
3794: via the attractor of the extended system $F$, which has the form
3795: $\mathcal{A}=\Sigma_N^+\times\mathcal{Y}$.
3796: 
3797: Let $\|\cdot\|$ be the norm on $TM$ generated by the metric on
3798: $M$. Denote
3799: $$
3800: s_i^+=\max_{x\in M}\|d_xf_i\|,\qquad s_i^-=\bigl(\max_{x\in
3801: M}\|d_xf_i^{-1}\|\bigr)^{-1}.
3802: $$
3803: We assume that the maps are non-degenerate (this is just for
3804: simplicity of arguments) and strictly contracting, so that
3805: $0<s_i^-\le s_i^+<1$.
3806: 
3807: Let $\eta=\max\limits_{x\in M}\#\{i\,|\,x=f_i(y_i)\text{ for some
3808: }y_i\in \mathcal{Y}\}$ be the maximal multiplicity of overlaps and
3809: $\varkappa_i=\max\limits_{x\in
3810: \mathcal{Y}}\#\{y\in\mathcal{Y}\,|\,x=f_i(y)\}$ be the maximal
3811: multiplicity of self-overlaps (we assume it is finite) on the
3812: attractor. Denote also by $\vartheta_i$ the multiplicity of the
3813: continuity partition for $f_i|_\mathcal{Y}$, i.e. the maximal
3814: number of continuity domains intersecting the attractor and
3815: meeting at one point of it.
3816: 
3817: \begin{theorem}\po
3818: Let $\underline{D}=\alpha$, $\overline{D}=\beta$ be the solutions
3819: of the equations
3820: $$
3821: \sum_{i=1}^N\tfrac1{\varkappa_i}|s_i^-|^{\alpha}=\eta,\qquad
3822: \sum_{i=1}^N\vartheta_i|s_i^+|^{\beta}=1.
3823: $$
3824: Then the Hausdorff dimension of the attractor satisfies:
3825: $$
3826: \underline{D}\le \dim_\text{H}(\mathcal{Y})\le\overline{D}.
3827: $$
3828: \end{theorem}
3829: 
3830: In addition to Hausdorff dimension we will need some other
3831: dimensional characteristics (see \cite{P2} for details). Denote by
3832: $\mathfrak{N}(X,\delta)$ the minimal cardinality of covers of $X$
3833: by balls of radius $\delta$. Then the lower and upper box
3834: dimensions are defined by the formula:
3835:  $$
3836: \underline{\dim}_B(X)=\mathop{\underline\lim}\limits_{\delta\to+0}
3837: \log\mathfrak{N}(X,\delta)/\log\tfrac1\delta,\quad
3838: \overline{\dim}_B(X)=\mathop{\overline\lim}\limits _{\delta\to+0}
3839: \log\mathfrak{N}(X,\delta)/\log\tfrac1\delta.
3840:  $$
3841: When these quantities are equal, their value is also called
3842: fractal dimension.
3843: 
3844: Consider a Borel probability measure $\mu\in\mathcal{M}(X)$ (an
3845: SRB-measure on the attractor can be taken in the SOC-context if it
3846: exists). The upper and lower pointwise dimensions are defined then
3847: as
3848:  $$
3849: \underline{d}_\mu(x)=\mathop{\underline\lim}\limits_{\delta\to+0}
3850: \log\mu(B(x,\delta))/\log\delta,\qquad
3851: \overline{d}_\mu(x)=\mathop{\overline\lim}\limits _{\delta\to+0}
3852: \log\mu(B(x,\delta))/\log\delta.
3853:  $$
3854: When they are equal and constant a.e. the measure $\mu$ is called
3855: exact-dimensional. This is precisely the case, when
3856: $\op{supp}\mu=X$ and we have equality in the general chain of
3857: inequalities (together with (\ref{dim-ineq1}) below):
3858:  \begin{equation}\label{dim-ineq}
3859: \op{ess.}\op{inf}\underline{d}_\mu(x)\le \op{dim}_{\text{H}}(X)\le
3860: \underline{\dim}_B(X)\le \overline{\dim}_B(X),
3861:  \end{equation}
3862: where by essential infimum we mean its upper bound taken over all
3863: subsets $U\subset X$ of measure 1 (and similar for
3864: $\op{ess.}\op{sup}\overline{d}_\mu(x)$). The last two inequalities
3865: are known and the first one follows from the inequality
3866: $\op{ess.}\op{inf}\underline{d}_\mu(x)\le
3867: \op{dim}_{\text{H}}(\mu)$ (\cite{P2}), where
3868: $\op{dim}_{\text{H}}(\mu)=\lim_{\delta\to+0}\op{inf}\{
3869: \op{dim}_{\text{H}}(Z)\,|\,\mu(Z)>1-\delta\}$.
3870: 
3871: Other dimensional characteristics of the measure are defined
3872: similarly and satisfy:
3873:  \begin{equation}\label{dim-ineq1}
3874: \underline{\dim}_B(\mu)\le \overline{\dim}_B(\mu)\le
3875: \op{ess.}\op{sup}\overline{d}_\mu(x).
3876:  \end{equation}
3877: Note that $\overline{\dim}_B(\mu)\le \overline{\dim}_B(X)$, while
3878: the quantities $\op{ess.}\op{sup}\overline{d}_\mu(x)$ and
3879: $\overline{\dim}_B(X)$ are in general incomparable.
3880: 
3881: The known formulas for the Hausdorff and other dimensions are
3882: generalizations of Moran's result (\cite{M,H}) and are based on
3883: the Bowen's equation (using the idea of coding); in this case one
3884: usually obtains exact-dimensionality \cite{P2}. In the SOC-context
3885: coding becomes problematic in the presence of singularities
3886: (unless the properties of the SRB-measure are clarified) and thus
3887: we cannot easily establish exact-dimensionality or formula for the
3888: dimension.
3889: 
3890: We prove instead the inequality of the theorem for all the various
3891: dimensions from (\ref{dim-ineq}) and (\ref{dim-ineq1}), which we
3892: denote just by $\dim(\mathcal{Y})$:
3893:  $$
3894: \underline{D}\le \dim(\mathcal{Y})\le\overline{D}.
3895:  $$
3896: 
3897: 
3898: \noindent
3899: \begin{proof}
3900: Let us consider at first the upper box dimension
3901: $\overline{\dim}_B(\mathcal{Y})$. The function $\mathfrak{N}$
3902: satisfies the inequalities:
3903: $$
3904: \tfrac1{\varkappa_i}\mathfrak{N}(\mathcal{Y},\delta/{s_i^-})\le
3905: \mathfrak{N}(f_i(\mathcal{Y}),\delta)\le
3906: \vartheta_i\mathfrak{N}(\mathcal{Y},\delta/{s_i^+}).
3907: $$
3908: The inequality from above is obtained as follows. Let
3909: $\mathcal{S}=\{x_j\}$ be a $\delta$-spanning set, i.e. a
3910: collection of points from $X$ with $U_\delta(\mathcal{S})=X$. Then
3911: $f_i(\mathcal{S})=\{f_i(x_j)\}$ may fail to be a $\delta\cdot
3912: s_i^+$-spanning set thanks to singularities. Whenever
3913: $\delta\ll1$, every $\delta$-ball intersects maximally
3914: $\vartheta_i$ domains of continuity for $f_i$ meeting
3915: $\mathcal{Y}$. Then we need to add maximally $\vartheta_i$ points
3916: for each ball $U_\delta(x_j)$ intersecting singularities. The
3917: inequality from below is proved similarly.
3918: 
3919: Now we have:
3920: $$
3921: \mathcal{Y}=f_1(\mathcal{Y})\cup\dots\cup f_N(\mathcal{Y})
3922: $$
3923: and the same for $U_\delta$-neighborhoods. This implies:
3924: \begin{equation}\label{780}
3925: \mathfrak{N}(\mathcal{Y},\delta)\le
3926: \sum_{i=1}^N\mathfrak{N}(f_i(\mathcal{Y}),\delta)\le
3927: \sum_{i=1}^N\vartheta_i\mathfrak{N}(\mathcal{Y},\delta/s_i^+).
3928: \end{equation}
3929: Denote $\sigma(\delta)=\mathfrak{N}(\mathcal{Y},\delta)
3930: \delta^{\overline{\dim}_B(\mathcal{Y})}$. This functions grows
3931: sub-polynomially:
3932: \begin{equation}\label{848}
3933: \mathop{\overline\lim}\limits_{\delta\to+0}\frac{\log
3934: \sigma(\delta)} {\log 1/\delta}=0.
3935: \end{equation}
3936: 
3937:  \begin{lem}\po\label{l34}
3938: Let $\l_i>1$ be some numbers and $p_i>0$ be some probabilities,
3939: $\sum_{i=1}^Np_i=1$. Then (\ref{848}) implies:
3940:  $$
3941: \mathop{\underline\lim}\limits_{\delta\to+0}\frac{\sum_{i=1}^N
3942: p_i\sigma(\lambda_i\delta)}{\sigma(\delta)}\le1.
3943:  $$
3944:  \end{lem}
3945: 
3946: \noindent
3947: \begin{proof}
3948: Suppose the lower limit is $>\kappa>1$. Then for every
3949: sufficiently small $\delta$ there exists $i\in[1,N]$ such that
3950: $\sigma(\lambda_i\delta)\ge\kappa\sigma(\delta)$.
3951: 
3952: Denote $\bar\lambda=\max_{1\le i\le N}\lambda_i$. Let
3953: $C=\max_{\delta\in[1/\bar\lambda,1]}\sigma(\delta)$. Then:
3954: $$
3955: \sigma(\delta)\le\frac1\kappa\sigma(\lambda_{i_1}\delta)\le
3956: \frac1{\kappa^2}\sigma(\lambda_{i_1}\lambda_{i_2}\delta)\le
3957: \dots\le\frac1{\kappa^{s(\delta)}}
3958: \sigma(\lambda_{i_1}\dots\lambda_{i_{s(\delta)}}\delta),
3959: $$
3960: where $s(\delta)$ is the first number such that
3961: $\lambda_{i_1}\dots\lambda_{i_{s(\delta)}}\delta\in[1/\bar\lambda,1]$.
3962: This number can be estimated as follows:
3963: $s(\delta)\ge-\log\delta/\log\bar\lambda-1$, whence:
3964: $$
3965: \frac{\log\sigma(\delta)}{\log1/\delta}\le \frac{\log
3966: C-s(\delta)\log\kappa}{\log1/\delta}\le
3967: \frac{\log(C\kappa)}{\log1/\delta}-\frac{\log\kappa}{\log\bar\lambda}.
3968: $$
3969: Therefore $\overline\lim_{\delta\to+0}\frac{\log
3970: \sigma(\delta)}{\log1/\delta}\le
3971: -\frac{\log\kappa}{\log\bar\lambda}<0$ and we get a contradiction.
3972: This proves the lemma.
3973: \end{proof} \\
3974: 
3975: Now to obtain the inequality from above for
3976: $\overline{\dim}_B(\mathcal{Y})$ divide (\ref{780}) by
3977: $\mathfrak{N}(\mathcal{Y},\delta)$. Denoting
3978: $\varpi=\sum_{i=1}^N\vartheta_i|s_i^+|^{\overline{\dim}_B(\mathcal{Y})}$,
3979: $\lambda_i=1/{s_i^+}$ and
3980: $p_i=\vartheta_i|s_i^+|^{\overline{\dim}_B(\mathcal{Y})}/\varpi$
3981: we get:
3982:  $$
3983: \frac1\varpi\le\sum_{i=1}^Np_i\frac{\sigma(\lambda_i\delta)}
3984: {\sigma(\delta)}.
3985:  $$
3986: Thus Lemma \ref{l34} implies that $1/\varpi\le 1$ or
3987:  $$
3988: 1\le \sum_{i=1}^N\vartheta_i|s_i^+|^{\dim\mathcal{Y}}
3989:  $$
3990: and the first claim
3991: $\overline{\dim}_B(\mathcal{Y})\le\overline{D}$ follows from the
3992: contraction $|s_i^+|<1$. The same arguments show another statement
3993: that $\overline{d}_\mu(x)\le\overline{D}$.
3994: 
3995: 
3996: 
3997: The inequality from below follows from
3998:  $$
3999: \mathfrak{N}(\mathcal{Y},\delta)\ge
4000: \frac1{\eta}\sum_{i=1}^N\mathfrak{N}(f_i(\mathcal{Y}),\delta)\ge
4001: \frac1{\eta}\sum_{i=1}^N\frac1{\varkappa_i}
4002: \mathfrak{N}(\mathcal{Y},\delta/s_i^-),
4003:  $$
4004: which implies $\sum_{i=1}^N\tfrac1{\varkappa_i}|s_i^-|^
4005: {\underline{\dim}_B(\mathcal{Y})}\le\eta$ and the same for the
4006: lower pointwise dimension: $\underline{d}_\mu(x)\ge\underline{D}$
4007: a.e.
4008: 
4009: 
4010: 
4011: In this case we should define
4012: $$
4013: \sigma(\delta)=\mathfrak{N}(\mathcal{Y},\delta)
4014: \delta^{\underline{\dim}_B(\mathcal{Y})}\text{ or }
4015: \sigma(\delta)=\bigr(\op{ess.}\op{inf}-\log\mu(B(x,\delta))\bigl)
4016: \delta^{\underline{\dim}_B(\mu)}
4017: $$
4018: respectively and use
4019: \begin{lem}\po\label{l34+}
4020: Let $\l_i>1$ be some numbers and $p_i>0$ be some probabilities,
4021: $\sum_{i=1}^Np_i=1$. Then:
4022: $$
4023: \mathop{\underline\lim}\limits_{\delta\to+0}\frac{\log
4024: \sigma(\delta)} {\log 1/\delta}=0 \ \ \Longrightarrow\ \
4025: \mathop{\overline\lim}\limits_{\delta\to+0}\frac{\sum_{i=1}^N
4026: p_i\sigma(\lambda_i\delta)}{\sigma(\delta)}\ge1.
4027: $$
4028: \end{lem}
4029: ~\\
4030: \noindent This is proved similarly to Lemma \ref{l34}. The
4031: inequalities for the Hausdorff dimension follows now from
4032: (\ref{dim-ineq}).
4033:  \end{proof} \\
4034: \abz
4035: 
4036: \noindent {\bf Acknowledgements.} We thank J. Schmeling and Y.
4037: Pesin for several stimulating discussions and references. M.
4038: Rypdal thanks E. Mj\o lhus for some useful questions and comments.
4039: We are grateful to the organizers of the Clay Mathematics
4040: Institute/MSRI Workshop on Recent Progress in Dynamics (2004),
4041: where we finished the final stage of the paper.
4042: 
4043: 
4044: 
4045: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4046: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4047: 
4048: \begin{thebibliography}{WW}
4049: 
4050: \small
4051: 
4052: \bib{AR}
4053: L. Abramov, V. Rokhlin, {\it The entropy of a skew product of
4054: measure-preserving transformations\/}, Amer. Math. Soc. Transl.
4055: Ser. 2, {\bf 48} (1966), 255--265.
4056: 
4057: \bib{Ba}
4058: V. Baladi, {\it Spectrum and Statistical Properties of Chaotic
4059: Dynamics\/}, Proc. 3rd European Congress of Mathematics,
4060: Birkhauser (2001), 203-224.
4061: 
4062: \bib{B1}
4063: J. Buzzi, {\it Intrinsic ergodicity of affine maps in
4064: $[0,1]^d$\/}, Mh. Math. {\bf 124} (1997), 97--118.
4065: 
4066: \bib{B2}
4067: J. Buzzi, {\it Piecewise isometries have zero topological
4068: entropy\/}, Ergod. Th. \& Dynam. Sys. {\bf 21} (2001), 1371--1377.
4069: 
4070: \bib{BTW}
4071: P. Bak, C. Tang, K. Wiesenfeld, {\it Self-Organized Criticality:
4072: An explenation of $1/f$-noise\/}, Phys. Rev. Lett. {\bf 59}, Num 4
4073: (1987), 381--384.
4074: 
4075: P. Bak, C. Tang, K. Wiesenfeld, {\it Self-Organized
4076: Criticality\/}, Phys. Rev. A {\bf 38}, Num 1 (1988), 364--374.
4077: 
4078: \bib{BC}
4079: T. Bogensch\"utz, H. Crauel, {\it The Abramov-Rokhlin Formula\/},
4080: Ergodic theory and related topics, III (Gustrow, 1990), 32--35,
4081: Lecture Notes in Math., {\bf 1514}, Springer, Berlin, 1992.
4082: 
4083: \bib{Bo}
4084: R. Bowen, {\it Entropy for group endomorphisms and homogeneous
4085: spaces\/}, Trans. A.M.S. {\bf 153} (1971), 401--414.
4086: 
4087: \bib{Br}
4088: J.\,R.\ Brown, {\it Ergodic theory and topological dynamics\/},
4089: Pure and Applied Mathematics, {\bf 70}, Academic Press [Harcourt
4090: Brace Jovanovich Publ.], New York-London (1976).
4091: 
4092: \bib{BCK}
4093: Ph. Blanchard, B. Cessac, T. Kr\"uger, {\it What can we learn
4094: about SOC from Dynamical System Theory\/}, J. Statist. Phys. {\bf
4095: 98} (2000), no. 1-2, 375--404.
4096: 
4097: B. Cessac, Ph. Blanchard, T. Kr\"uger, {\it Lyapunov exponents and
4098: transport in the Zhang model of self-organized criticality\/},
4099: Phys. Rev. E {\bf 64} (2001).
4100: 
4101: B. Cessac, Ph. Blanchard, T. Kr\"uger J.\,L. Meunier, {\it
4102: Self-Organized Criticality and Thermodynamic Formalism\/}, J.
4103: Statist. Phys. {\bf 115} (2004), 1283--1326.
4104: 
4105: \bib{DGS}
4106: M. Denker, C. Grillenberger, K. Sigmund, {\it Ergodic theory on
4107: compact spaces\/}, Lecture Notes in Math., {\bf 527.}
4108: Springer-Verlag, Berlin-New York (1976).
4109: 
4110: \bib{DR}
4111: D. Dhar, R. Ramasway, {\it Exactly solved Model of Self-Organized
4112: Critical Phenomena}, Phys. Rew. Let. Vol 63, Num 16 (1089),
4113: 1659--1662.
4114: 
4115: \bib{GD}
4116: A. Giacommetti, A. Dias-Guilera, {\it Dynamical properties of the
4117: Zhang model of self-organized criticality\/}, Phys. Rev. Let. E,
4118: {\bf 58}, no. 1 (1998).
4119: 
4120: \bib{H}
4121: J.\,E. Hutchinson, {\it Fractals and self-similarity\/}, Indiana
4122: University Math. Journal, {\bf 30} no.5 (1981), 713--747.
4123: 
4124: \bib{J}
4125: H.\,J. Jensen, {\it Self-Organized Criticality: Emergent Complex
4126: Behavior in Physical and Biological systems\/}, Cambridge Lecture
4127: Notes in Physics 10, Cambridge University Press (1998).
4128: 
4129: \bib{K}
4130: A. Katok, {\it Lyapunov exponents, entropy and periodic orbits for
4131: diffeomorphisms}, Inst. Hautes Etudes Sci. Publ. Math., {\bf 51}
4132: (1980), 137--173.
4133: 
4134: \bib{KH}
4135: A. Katok, B. Hasselblatt, {\it Introduction to the Modern Theory
4136: of Dynamical Systems\/}, Cambridge University Press (1995).
4137: 
4138: \bib{KS}
4139: A. Katok, J.-M. Strelcyn, {\it Invariant manifolds, Entropy and
4140: Billiards; Smooth Maps with Singularities\/}, Lecture Notes in
4141: Math., {\bf 1222}, Springer, Berlin, 1986.
4142: 
4143: \bib{KK}
4144: B. V. Kozelov, T. V. Kozelova, {\it Cellular automata model of
4145: magnetospheric-ionospheric coupling\/}, Annales Geophysicae, {\bf
4146: 21} (2003), 1931-1938.
4147: 
4148: \bib{KR1}
4149: B. Kruglikov, M. Rypdal, {\it A piece-wise affine contracting map
4150: with positive entropy\/}, ArXiv: math.DS/0504187.
4151: 
4152: \bib{KR2}
4153: B. Kruglikov, M. Rypdal, {\it Entropy via multiplicity\/},
4154: ArXiv: math.DS/0505019.
4155: 
4156: \bib{LY}
4157: F. Ledrappier, L.-S. Young, {\it The metric entropy of
4158: diffeomorphisms}, I-II, Ann. of Math. (2) {\bf 122} (1985), no. 3,
4159: 509--539; 540--574.
4160: 
4161: \bib{M}
4162: P.\,A.\,P. Moran, {\it Additive functions of intervals and
4163: Hausdorff measure\/}, Proc. Camb. Phil. Soc., {\bf 42} (1946),
4164: 15--23.
4165: 
4166: \bib{P1}
4167: Y. Pesin, {\it Dynamical systems with generalized hyperbolic
4168: attractors: hyperbolic, ergodic and topological properties},
4169: Ergod. Th. \& Dynam. Sys. {\bf 12} (1992), 123--151.
4170: 
4171: \bib{P2}
4172: Y. Pesin, {\it Dimension theory in dynamical systems\/}, Chicago
4173: Lect. in Math. Ser., The University of Chicago Press (1997).
4174: 
4175: \bib{Pt}
4176: K. Petersen, {\it Chains, entropy, coding\/}, Ergod. Th. \& Dynam.
4177: Sys. {\bf 6} (1986), no. 3, 415--448.
4178: 
4179: \bib{PY}
4180: M. Pollicott, M. Yuri, {\it Dynamical systems and ergodic
4181: theory\/}, London Mathematical Society Student Texts {\bf 40},
4182: Cambridge University Press (1998).
4183: 
4184: \bib{Ru}
4185: D. Ruelle, {\it Chaotic evolution and strange attractors\/},
4186: Lezioni Lincee, Cambridge University Press (1989).
4187: 
4188: \bib{R}
4189: M. Rypdal, {\it Dynamics of the Zhang model of Self-Organized
4190: Criticality\/}, Master Thesis in mathematics, University of Troms-
4191: 2004, e-printed: http://www.math.uit.no/seminar/preprints.html.
4192: 
4193: \bib{ST}
4194: J. Schmeling, S. Troubetzkoy, {\it Dimension and invertibility of
4195: hyperbolic endomorphisms with singularities\/}, Ergod. Th. \&
4196: Dynam. Sys. {\bf 18} (1998), 1257--1282.
4197: 
4198: \bib{Z}
4199: H.Y. Zhang, {\it Scaling theory of Self-Organized Criticality\/},
4200: Phys. Rev. Lett. {\bf 63} Num 5 (1988), 470--473.
4201: 
4202: \end{thebibliography}
4203: 
4204: 
4205: 
4206: \end{document}
4207: