nlin0011044/fsle.tex
1: %%%%%%%   19 gennaio 2001 - ROMA
2: \documentstyle[pre,aps,epsfig,multicol]{revtex}
3: %\documentstyle[preprint,aps,epsfig]{revtex}
4: %\tightenlines
5: 
6: \begin{document}
7: \title{Linear and nonlinear 
8: information flow \\ in spatially extended systems}
9: \author{Massimo Cencini $^{(a)}$ and Alessandro Torcini$^{(b),(c)}$}
10: \address{$(a)$ Max-Planck-Institut f\"ur Physik komplexer Systeme,
11: N\"othnitzer  Str. 38, D-01187 Dresden, Germany  } 
12: \address{$(b)$ Physics Department, Universit\'a ``La Sapienza'',
13: piazzale Aldo Moro 2, I-00185 Roma, Italy} 
14: \address{$(c)$ Istituto Nazionale di Fisica della Materia, UdR Firenze,
15: L.go E. Fermi, 3 - I-50125 Firenze, Italy}
16: 
17: \draft
18: 
19: \maketitle
20: 
21: \begin{abstract}
22: Infinitesimal and finite amplitude error propagation in 
23: spatially extended 
24: systems are numerically and theoretically investigated.
25: The information transport in these systems can be characterized
26: in terms of the propagation velocity of perturbations $V_p$.
27: A linear stability analysis is sufficient to capture all
28: the relevant aspects associated to propagation of 
29: infinitesimal disturbances. In particular, this 
30: analysis gives the propagation velocity $V_L$ of 
31: infinitesimal errors. If linear mechanisms prevail
32: on the nonlinear ones $V_p = V_L$. On the contrary, if 
33: nonlinear effects are predominant finite amplitude 
34: disturbances can eventually propagate faster than infinitesimal 
35: ones (i.e. $V_p > V_L$). The finite size Lyapunov exponent
36: can be successfully employed to discriminate the linear
37: or nonlinear origin of information flow.
38: A generalization of finite size Lyapunov exponent
39: to a comoving reference frame allows to state a
40: marginal stability criterion able to provide $V_p$
41: both in the linear and in the nonlinear case.
42: Strong analogies are found
43: between information spreading and propagation 
44: of fronts connecting steady states 
45: in reaction-diffusion systems. The analysis of
46: the common characteristics of these two phenomena 
47: leads to a better understanding of the role played
48: by linear and nonlinear mechanisms for the flow of 
49: information in spatially extended systems.
50: \end{abstract}
51: %
52: \pacs{PACS numbers: 05.45.-a,68.10.Gw,05.45.Ra,05.45.Pq}
53: %
54: % explanation of PACS numbers:
55: % ----------------------------
56: %
57: % 47.20.Ky Nonlinearity (including bifurcation theory)
58: % 05.45.Ra Coupled map lattices
59: % 05.45.-a Nonlinear dynamics and nonlinear dynamical
60: % systems (see also 45 Classical mechanics of
61: % discrete systems)
62: % 05.45.Pq Numerical simulations of chaotic models
63: % 68.10.Gw Interface activity, spreading
64: % 47.54.+r Pattern selection; pattern formation
65: % 02.30.Jr Partial differential equations
66: % 05.40.-a Fluctuation phenomena, random processes, noise, and Brownian motion
67: 
68: 
69: \begin{multicols}{2}
70: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
71: \section{Introduction}
72: \label{sec:1}
73: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
74: It is well recognized that chaotic dynamics generates a flow of
75: information in {\it bit space}: due to the sensitive dependence on initial
76: conditions one has an information flow from ``insignificant'' digits
77: towards ``significant'' ones \cite{Shaw}.  
78: In spatially distributed systems, due to the spatial coupling, one has an 
79: information flow both in {\it bit space} and in {\it real space}.
80: The flow in {\it bit space} is typically characterized in 
81: terms of the maximal growth $\lambda$ rate of infinitesimal disturbances
82: (i.e. of the maximal Lyapunov exponent), while the spatial information flow 
83: can be measured in terms of the maximal velocity of disturbance propagation
84: $V_p$ \cite{K86,BR87,grass89,PV94}. 
85: 
86: The evolution of a typical infinitesimal disturbance in low dimensional systems 
87: is fully determined once the maximal Lyapunov exponent is known. 
88: The situation is more complicated in spatio-temporal chaotic systems,
89: where infinitesimal perturbations can evolve both in time 
90: and in space. In this case a complete description of the dynamics
91: in the tangent space requires the introduction of other indicators,
92: e.g.: the comoving Lyapunov exponents \cite{DK87} and
93: the spatial and the specific Lyapunov spectra \cite{lepri}.
94: 
95: Nevertheless, the complete knowledge of  these Lyapunov spectra is not 
96: sufficient to fully characterize 
97: the irregular behaviors emerging in dynamical systems, this is
98: particularly true when the evolution of finite perturbations
99: is concerned.  Indeed, finite disturbances, which  are not confined 
100: in the tangent space, but are governed by the complete nonlinear 
101: dynamics, play a fundamental
102: role in the erratic behaviors observed in some high dimensional
103: system \cite{CK88,PLOK93,TP94,TGP95,CFVV99}. A rather 
104: intriguing phenomenon, termed {\it stable chaos}, 
105: has been reported in \cite{PLOK93}: the authors 
106: observed that even a linearly stable system 
107: (i.e. with $\lambda < 0$) can display an erratic behavior with $V_p>0$.
108: 
109: The first attempts to 
110: describe nonlinear perturbation evolution have been reported in
111: \cite{dressler92}. However, in these studies the analysis
112: was limited to the temporal growth rate associated with 
113: second order derivatives of one dimensional maps.
114: A considerable improvement 
115: along this direction has been recently achieved with the introduction of
116: the
117: finite size Lyapunov exponent (FSLE) \cite{ABCPV96}:
118: a generalization of the maximal Lyapunov exponent able to 
119: describe also finite amplitude perturbation evolution. 
120: In particular, the FSLE has been already demonstrated 
121: useful in investigating high dimensional systems \cite{CFVV99}.
122: 
123: The aim of this paper is to fully characterize the infinitesimal 
124: and finite amplitude perturbation evolution in spatio-temporal
125: chaotic systems. Coupled map lattices (CML's) \cite{cml} 
126: are employed to mimic spatially extended chaotic systems.
127: The FSLE is successfully applied to discriminate 
128: the linear or nonlinear origin of information propagation 
129: in CML's. Moreover, a generalization
130: of the FSLE to comoving reference frame (finite size comoving Lyapunov exponent)
131: allows to state
132: a marginal stability criterion able to predict $V_p$
133: in both cases: linear or nonlinear propagation.
134: A parallel with front propagation in reaction-diffusion \cite{KPP,wim2}
135: (non-chaotic) systems is worked out. The 
136: analogies between the two phenomena authorize to draw a 
137: correspondence between ``pulled'' (``pushed'') fronts 
138: and linear (nonlinear) information spreading.
139: 
140: The paper is organized as follows.
141: In Sect. II the FSLE is introduced
142: and applied to low dimensional systems (i.e. to single chaotic maps).
143: Sect. III is devoted to the description and 
144: comparison of linear and nonlinear 
145: disturbance propagation observed in different CML models.
146: The finite size comoving Lyapunov exponent is introduced in Sect. IV
147: and employed to introduce a generalized marginal stability criterion
148: for the determination of $V_p$. A discussion on information
149: propagation in non chaotic systems conclude Sect. IV.
150: The analogies between disturbance propagation in chaotic
151: systems and front propagation connecting steady states 
152: are analyzed in Sect. V.  The Appendix is devoted to the
153: estimation of finite time corrections for the computation of the 
154: FSLE in extended systems. Finally, some conclusive
155: remarks are reported in Sect. VI.
156: 
157: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
158: \section{Finite Size Lyapunov Exponent: \\
159: Low Dimensional Models}
160: \label{sec:2}
161: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
162: %introduzione dell'fsle
163: 
164: Let us introduce the FSLE by considering the
165: dynamical evolution of the state variable 
166: ${\bf x}={\bf x}(t)$ ruled by
167: $$
168: {\dot {\bf x}}(t)={\bf f}[{\bf x}(t)] \, ,
169: $$
170: where ${\bf f}$ represents a chaotic flow in the phase space.
171: In order to evaluate the growth rate of non infinitesimal perturbations 
172: one can proceed as follows: a reference, ${\bf x}(t)$, and 
173: a perturbed, ${\bf x}^{'}(t)$, trajectories are considered.
174: The two orbits are initially placed at a distance 
175: $\delta(0)=\delta_{min}$, with $\delta_{min}\ll 1$,
176: assuming a certain norm  $\delta(t)=||{\bf x}^{'}(t)- {\bf x}(t)||$. 
177: In order to ensure that the perturbed orbit relaxes
178: on the attractor a first scratch integration is performed
179: for both the orbits until their distance has grown from $\delta_{min}$ 
180: to $\delta_0$ (where $1 >> \delta_0 >> \delta_{min})$. 
181: This transient ensures also the 
182: alignment of the perturbation along the direction of maximal expansion. 
183: Then the two trajectories are let to evolve and 
184: the growth of their distance $\delta(t)$ through different pre-assigned 
185: thresholds ($\delta_n=\delta_0 r^{n}$, with $n=0,\dots,N$ and typically
186: $1 < r \le 2$) is analyzed.
187: 
188: After the first threshold, $\delta_0$, is attained  
189: the times $\tau(\delta_n,r)$ required for 
190: $\delta(t)$ to grow from $\delta_n$ up to $\delta_{n+1}$
191: are registered.
192: When the largest threshold $\delta_N$ (which should be obviously chosen
193: smaller than the attractor size) is reached, the perturbed
194: trajectory is rescaled to the initial distance 
195: $\delta_{min}$ from the reference one.
196: 
197: By repeating the above procedure ${\cal N}$ times, 
198: for each threshold $\delta_n$, one obtains
199: a set of ``doubling'' times (this terminology is strictly
200: speaking correct only if $r=2$) 
201: $\{\tau_i(\delta_n,r)\}_{i=1,\dots,{\cal N}}$
202: and one can define the average of any observable
203: $A = A(t)$ on this set of doubling times as:
204: $$
205:  \langle A\rangle_e \equiv \frac{1}{\cal N} \sum_{i=1}^{\cal N} A_i\,,
206: $$
207: where $ A_i = A(\tau_i(\delta_n,r))$. 
208: The average $\langle \cdot \rangle_e$
209: does not coincide with an usual time average $\langle \cdot \rangle_t$
210: along a considered trajectory in the phase space, since
211: the doubling times typically depend on the 
212: considered point along the trajectory and on the threshold
213: $\delta_n$.  The two averages are linked 
214: (at least in the continuous case) 
215: via the following straightforward relationship~\cite{cencrep}
216: \begin{equation}
217: \langle A(t) \rangle_t = \frac{1}{T} \int_0^T dt A(t) = 
218: \frac{ \langle A \tau \rangle_e}{\langle \tau \rangle_e}
219: \label{eq:2.1}
220: \end{equation}
221: where  $T = \sum_{i=1,{\cal N}} \tau_i(\delta_n,r)$ and
222: $\langle \tau(\delta_n,r)\rangle_{e} = {T/\cal{N}}$.
223: 
224: A natural definition of the Finite Size Lyapunov Exponent
225: $\lambda(\delta_n)$ is the following~\cite{ABCPV96}:
226: \begin{equation}
227: \lambda(\delta_n) \equiv \left\langle{1 \over \tau(\delta_n,r)}\right\rangle_{t} \ln r \equiv
228: {1 \over \langle \tau(\delta_n,r) \rangle_{e}} \ln r \, .
229: \label{eq:2.2}
230: \end{equation}
231: The last equality stems from the relationship among the two
232: averages reported in (\ref{eq:2.1}).
233: 
234: In the limit of infinitesimal perturbation $\delta_n$
235: and of infinite $T$ (or $\cal N$) the FSLE converges to the
236: usual maximal Lyapunov exponent
237: \begin{equation}
238: \lim_{{\cal N} \to \infty}
239: \lim_{\delta_n \to 0} \lambda(\delta_n) = \lambda \, .
240: \label{eq:2.3}
241: \end{equation}
242: In practice, at small enough $\delta_n$,
243: $\lambda(\delta_n)$ displays a plateau
244: $\sim \lambda$.
245: Moreover, one can verify that $\lambda(\delta_n)$ is
246: independent of $r$, at least for not too large 
247: $r$~\cite{ABCPV96}.
248: 
249: In Eq.~(\ref{eq:2.2}) continuous time has been assumed, 
250: but discrete time is the most natural choice when 
251: experimental data sets (typically sampled at fixed intervals) 
252: are considered. In order to generalize the FSLE's definition
253: to the case of discrete time dynamical systems, let us consider 
254: the following map
255: $$ {\bf x}(t+1)={\bf F}({\bf x}(t)) \,,$$ 
256: where ${\bf x}$ is a continuous 
257: variable, and $t$ assumes integer values.
258: In this case $\tau(\delta_n,r)=\tau$ has simply to be interpreted as the minimum 
259: ``integer'' time such that $\delta(\tau) \ge  \delta_{n+1}$, and, 
260: since now $\delta(\tau)/\delta_n$ is a fluctuating quantity, 
261: the following definition is obtained
262: \begin{equation}
263: \lambda(\delta_n) \equiv {1 \over \langle \tau(\delta_n,r) \rangle_{e}}
264: \left\langle \ln \left( \delta(\tau) \over \delta_n
265:  \right) 
266: \right\rangle_{e} \, .
267: \label{eq:2.5}
268: \end{equation}
269: A theoretical estimation of (\ref{eq:2.5}) is rarely possible,
270: and in most cases, one can only rely on a numerical
271: computation of $\lambda(\delta_n)$. However, 
272: in the following we will report two simple cases 
273: for which an approximate analytic expression for
274: the FSLE can be worked out.
275: 
276: Let us first consider the tent map 
277: $$ F(x)= 1- 2\left|x-\frac{1}{2}\right|$$
278: where $x \in [0:1]$. This is a one dimensional chaotic map,
279: since  $\lambda= \ln 2$ is positive.
280: 
281: Due to the simplicity of this map, one can estimate the expression
282: (\ref{eq:2.5}) analytically obtaining the following
283: approximation 
284: \begin{equation}
285: \lambda(\delta) \simeq \ln 2 -\delta\,,
286: \label{eq:fsle_tent}
287: \end{equation}
288: valid for not too large $\delta$ values.
289: The maximal Lyapunov exponent is correctly recovered
290: in the limit $\delta \to 0$ and the above expression 
291: reproduces quite well the numerical estimate of the FSLE
292: (see Fig.~\ref{fig:maps} (a) ). An important point to stress
293: is that for this map the finite amplitude perturbations grow 
294: with the same rate or slower than the infinitesimal ones. 
295: The contraction of perturbations at large scales is
296: due to saturation effects related to the attractor
297: size. A similar dependence of the FSLE on the considered
298: scale is observed for the majority of the chaotic maps 
299: (logistic, cubic, etc.), as we have verified. 
300: %fig1
301: %%%%%%%%%%%%%%%%%%%%%%%%
302: \begin{figure}
303: \epsfxsize=8truecm
304: \epsfysize=6truecm
305: \epsfbox{Fig1a.eps}
306: \epsfxsize=8truecm
307: \epsfysize=6truecm
308: \epsfbox{Fig1b.eps}
309: %\narrowtext
310: \caption{$\lambda(\delta)$ versus $\delta$ for the tent map (a) and the 
311: shift map with $\beta=2$ (b). The continuous lines are the analytically 
312: computed FSLE and the boxes the numerically evaluated one. 
313: The two maps are displayed in the insets.  
314: }
315: \label{fig:maps}
316: \end{figure}
317: %%%%%%%%%%%%%%%%%%%%%%%%
318: One can wonder if there are systems for which, at variance with 
319: the behavior (\ref{eq:fsle_tent}), the finite size
320: corrections leads to an enhancement of the growth rate at 
321: large scales. As shown in \cite{TGP95}, the shift map
322: $F(x)= \beta x \; {\mbox {mod}} \; 1$ represents a good candidate. 
323: Also in this case it is possible to obtain
324: an analytical expression for the FSLE, when
325: $\delta < [1/(r+\beta)]$,
326: \begin{equation}
327: \lambda(\delta)=\frac{1}{1-\delta} \left[
328: (1-2\delta) \ln \beta+
329: \delta \ln \left({1-\beta\delta \over \delta}\right)
330: \right]\,,
331: \label{eq:lambda_shift}  
332: \end{equation}
333: which again correctly reproduces the numerical data
334: (see Fig.~\ref{fig:maps} (b) ) and in the limit
335: $\delta \to 0$ reduces to the corresponding maximal
336: Lyapunov exponent $\lambda=\ln \beta$.
337: As expected, finite amplitude disturbances can grow
338: faster than infinitesimal ones : 
339: \begin{equation}
340: \lambda(\delta) > \lambda(\delta \to 0)=\lambda \qquad {\mbox {for}} 
341: \:\: 0 < \delta \le \delta^{sat}\,,
342: \label{eq:nonlin}
343: \end{equation}
344: where $\delta^{sat}$ indicates the threshold at which saturation
345: effects set in. An even more interesting situation
346: is represented by the circle map 
347: $F(x)= \alpha + x \:\; {\mbox {mod}} \; 1$. This
348: map is marginally stable (i.e. $\lambda\!=\!\!0$),
349: but it is unstable at finite scales.
350: Indeed, the FSLE is given by 
351: $\lambda(\delta)\!\!=\!\!\delta/(1-\delta) \ln[(1-\delta)/\delta]$,
352: which is positive for $0 < \delta <1/2$. Therefore at small, but
353: finite, perturbations a positive growth rate is observed in spite 
354: of the (marginal) stability against infinitesimal perturbations. 
355: As a consequence, the circle map can exhibit behaviors that can
356: be hardly distinguished from chaos under the influence of
357: noise, since small perturbations may be occasionally driven 
358: into the nonlinear (unstable) regime and therefore amplified.
359: Of course, the role of noise can be played by coupling with
360: other maps, e.g. it has been found that coupled circle maps display 
361: behavior
362: resembling (for some aspects) that of a chaotic system \cite{TGP95}.
363: This phenomenon becomes even more striking in certain coupled stable maps where, 
364: even if the maximal Lyapunov exponent  is
365: negative \cite{PLOK93}, one can have a strong sensitivity to non infinitesimal
366: perturbations \cite{Kantz} (see  Sect.~\ref{sec:4.2} for a detailed discussion).
367: 
368: The two maps here examined for which (\ref{eq:nonlin}) holds
369: have a common characteristic: they are discontinuous.
370: However, in order to observe similar 
371: strong nonlinear effects, it is sufficient to consider a continuous map 
372: with high, but finite, first-derivative $|F^\prime|$ values 
373: \cite{TP94}. In this respect a simple example, that 
374: will be examined more in detail in Sect. IV, is represented 
375: by the map:
376: \begin{equation}
377: F(x)=\left \{ \begin{array}{cc}
378: b x & \qquad 0 \leq x < 1/b\\
379: 1-c(1-q)(x-1/b) & \qquad 1/b \leq x < \frac{b+c}{bc}\\
380: q+d(x-\frac{b+c}{bc}) & \qquad \frac{b+c}{bc} \leq x \leq 1\,;
381: \end{array}
382: \right.
383: \label{eq:stable}
384: \end{equation}                                                 
385: with $b=2.7$, $d=0.1$, $q=0.07$ and $c=500$.
386: For $c \to \infty$ the map (\ref{eq:stable}) reduces to the one
387: studied in \cite{PLOK93}. For the map (\ref{eq:stable}) the FSLE 
388: dependence on $\delta$ is similar to that observed 
389: for the shift map.
390: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
391: \section{Information Spreading in Spatially Distributed Systems}
392: \label{sec:3}
393: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
394: In this section we will examine the mechanisms behind the
395: information flow in spatially distributed systems. 
396: In particular, the influence of
397: linear and nonlinear effects on information (error) spreading
398: will be analyzed. As a prototype of spatially 
399: distributed system Coupled Map Lattices (CML's) \cite{cml}
400: are considered :
401: \begin{eqnarray}
402: x_i(t+1)&=&F(\tilde{x}_i(t)) \nonumber \\
403: \tilde{x}_i(t)=(1-\varepsilon) x_i(t) &+& {\varepsilon \over 2} (x_{i-1}(t)+x_{i+1}(t))
404: \label{eq:cml}
405: \end{eqnarray}
406: where $t$ and $i$ are the discrete temporal and spatial indices, 
407: $L$ is the lattice size ($i=-L/2,\dots,L/2$), $x_i(t)$ the
408: state variable, and $\varepsilon \in [0:1]$ measures the
409: strength of the diffusive coupling. 
410: $F(x)$ is a nonlinear map of the interval 
411: ruling the local dynamics.
412: 
413: In order to understand how the information spreads
414: along the chain, let us consider two replicas of the same 
415: system, ${\bf x}(t)=\{x_i(t)\}$  and
416: ${\bf x}^\prime(t)=\{x_i^\prime(t)\}$, 
417: that initially differ only in
418: a single site of the lattice (e.g. $i=0$) of a quantity $d_0$, i.e.
419: \begin{equation}
420: |x^\prime_i(0)-x_i(0)|=\Delta x_i(0)=d_0 \delta_{i,0}\,,
421: \label{eq:tzero}
422: \end{equation}
423: where $\delta_{i,0}$ is the Kronecker's delta. 
424: In a chaotic system the perturbation will typically
425: grow locally and spread along the chain.
426: These phenomena can be studied by considering the
427: difference field
428: \begin{equation}
429: \Delta x_i(t)= |x^\prime_i(t)-x_i(t)|=
430: |F(\tilde{x}^{'}_i(t-1))-F(\tilde{x}_i(t-1))|\,. 
431: \label{eq:dyn}
432: \end{equation}
433: It has to be stressed that the full nonlinear dynamics 
434: contributes to the evolution of $\Delta x_i(t)$.
435: 
436: The spreading of this initially localized 
437: disturbance can be characterized in terms of
438: the velocity of information
439: propagation $V_P$ \cite{K86,grass89}.
440: As shown in Fig.~\ref{fig:torc.1}, $\Delta x_i(t)$ 
441: can grow only within a light-cone, determined by  $V_p$. 
442: For velocities higher than
443: $V_p$ the disturbance is instead damped. This individuates
444: a sort of predictability ``horizon'' in space-time: 
445: i.e. an interface separating 
446: the perturbed from the unperturbed region.
447: % fig 2
448: %%%%%%%%%%%%%%%%%%%%%%%%
449: \begin{figure}
450: \epsfxsize=8.truecm
451: \epsfysize=6truecm
452: \epsfbox{Fig2.eps}
453: %\narrowtext
454: \caption{Evolution of $\Delta x_i(t)$, 
455: for a chain of coupled tent map lattices
456: with a coupling $\varepsilon=2/3$.
457: The initial perturbation
458: is taken as in (\ref{eq:tzero}) with $d_0=10^{-8}$. 
459: }
460: \label{fig:torc.1}
461: \end{figure}
462: %%%%%%%%%%%%%%%%%%%%%%%% 
463: The velocity $V_p$ can be directly measured 
464: by detecting the leftmost, $i_l(t)$, and the rightmost,
465: $i_r(t)$, sites for which at time $t$ the perturbation 
466: $\Delta x_i(t)$ exceeds a preassigned
467: threshold. The definition of $V_p$ is the following
468: \begin{equation}
469: V_p=\lim_{t\to \infty}\lim_{L\to \infty} {i_r(t) -i_l(t)\over 2t}\,,
470: \label{eq:vel1}
471: \end{equation}
472: where  the limit $L\to \infty$ has to be taken first to avoid boundaries 
473: effects. The velocity (\ref{eq:vel1}) does not depend
474: on the chosen threshold values \cite{K86,grass89,TP94}.
475: 
476: Since the dynamics of the difference field 
477: (\ref{eq:dyn}) is not confined in the tangent space, 
478: non linearities can play a crucial role in the information propagation.
479: Indeed, we will see that the evolution of the disturbances strongly depend 
480: on the considered map $F(x)$ and in particular on the
481: shape of $\lambda(\delta)$. In the next subsection 
482: propagation in CML's with local chaotic
483: maps for which $\lambda(\delta)\leq \lambda \quad \forall \delta$
484: is discussed.  Local maps for which the condition 
485: (\ref{eq:nonlin}) holds will be the subject of 
486: Sect. \ref{sec:3.2}.
487: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
488: \subsection{Linear Mechanisms}
489: \label{sec:3.1}
490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
491: Since in this subsection we consider 
492: CML's for which the local instabilities are 
493: essentially dominated by the behavior
494: of infinitesimal perturbations, 
495: most of the features can be understood
496: by limiting the analysis to the tangent space.
497: 
498: The evolution in tangent space is
499: obtained by linearizing 
500: Eq.~(\ref{eq:cml}), i.e.
501: \begin{eqnarray}
502: \delta x_{i}(t+1)&=&F^\prime(\tilde{x}_i(t)) [ \delta x_i(t)+
503: \nonumber\\
504: \frac{\varepsilon}{2}
505: (\delta x_{i+1}(t)&-&2 \delta x_{i}(t)+\delta x_{i-1}(t)) ]\,,
506: \label{eq:tang}
507: \end{eqnarray}
508: where $F^\prime$ is the first derivative of a one dimensional
509: chaotic map. Let us again consider as initial condition for
510: the evolution of (\ref{eq:tang}) a localized perturbation
511: as  (\ref{eq:tzero}) with $d_0$ infinitesimal.
512: The spatio-temporal dynamics of the 
513: tangent vector $\{ \delta x_i(t) \}$ is
514: determined by the interaction and competition of 
515: two different mechanisms present in Eq.~(\ref{eq:tang}):
516: the chaotic instability and the spatial diffusion.
517: 
518: As a first approximation, the effects of the 
519: two mechanisms can be treated as independent.
520: The chaotic instability leads to an average 
521: exponential growth of the infinitesimal disturbance:
522: $|\delta x_0(t)| \approx {d_0 \exp[\lambda t]}$.
523: On the other hand, the spatial diffusion, due to the
524: coupling, approximately leads to a 
525: spatial Gaussian spreading of the disturbance :
526: $|\delta x_i(t)| \approx 
527: |\delta x_0(t)| 
528: {/ \sqrt{4\pi Dt}} \exp( -{i^2 / 4Dt})$ 
529: where $D=\varepsilon/2$.
530: Combining these two effects one obtains
531: \begin{equation}
532: |\delta x_i(t)| \approx d_0 
533: {1 \over \sqrt{2\pi \varepsilon t}} \exp \left(\lambda t 
534: -{i^2 \over 2 \varepsilon t}\right)\,.
535: \label{eq:shape}
536: \end{equation}
537: Since the chaotic nature of the phenomenon
538: will typically  induces fluctuations, Eq.~(\ref{eq:shape}) can only describe 
539: the average shape of the disturbance. 
540: Moreover, Eq.~(\ref{eq:shape}) holds only 
541: when the perturbation is infinitesimal, since when
542: the disturbance reaches finite values a saturation mechanism 
543: (due to the nonlinearities) sets in preventing the divergence
544: of $|\delta x_i(t)|$. 
545: 
546: To verify the  validity of (\ref{eq:shape}), we studied the evolution
547: of localized perturbations of a homogeneous spatio-temporal 
548: chaotic state, in particular coupled logistic and tent maps
549: have been considered in the regime of ``fully developed turbulence''
550: \cite{K86}. Firstly, the system is randomly initialized and 
551: let to relax for a relatively long transient. 
552: At this stage two replicas of the same system are
553: generated and to one of the two a localized perturbation 
554: (like (\ref{eq:tzero})) is added. The evolution of the 
555: difference field (\ref{eq:dyn}) is then monitored at 
556: successive times.  In order to wash out the fluctuations,
557: the shape of the disturbance is obtained
558: averaging over many distinct realizations.
559: %fig 3
560: %%%%%%%%%%%%%%%%%%%%%%%%
561: \begin{figure}
562: \epsfxsize=8.truecm
563: \epsfysize=6truecm
564: \epsfbox{Fig3.eps}
565: %\narrowtext
566: \caption{Average evolution of perturbations for a CML of logistic maps
567: ($f(x)=4x(1-x)$) for (a) $\varepsilon=1/3$ and (b) $\varepsilon=1/10$. 
568: $\langle \Delta x_i(t)\rangle$ is reported as a function of $i^2$ in a
569: lin-log scale at different 
570: times (from bottom to top $t=10,20,30,40$). Deviations from 
571: a straight line correspond to deviation from the Gaussian shape. 
572: $\langle \Delta x_i(t)\rangle$ is obtained as an average over $10^3$ realizations, 
573: for each one $\Delta x_i(0)$ has been chosen as (\ref{eq:tzero}) 
574: with $d_0=10^{-7}$.
575: For comparison the prediction (\ref{eq:shape})
576: is also reported (dashed lines).}
577: \label{fig:fronte}
578: \end{figure}
579: %%%%%%%%%%%%%%%%%%%%%%%% 
580: As one can see from Fig.~\ref{fig:fronte} Eq.~(\ref{eq:shape}) is fairly well 
581: verified for large enough coupling while it fails at small $\varepsilon$ \cite{nota}. 
582: These discrepancies are due to the finite spatial resolution
583: (that in CML's is always fixed to $1$), since
584: for small diffusivity constant the discretization of the Laplacian 
585: becomes inappropriate.
586: The expression (\ref{eq:shape}) for disturbance evolution
587: has been already proposed in Ref.~\cite{WB93}
588: for CML's in two dimensions. Deviations from
589: (\ref{eq:shape}) have been observed also in 
590: \cite{WB93}, but attributed to anomalous diffusive behaviors.
591: It has to be remarked that expression  (\ref{eq:shape})
592: is valid only at short times, since asymptotically ($t \to \infty$) 
593: the infinitesimal leading edge of the propagating front 
594: $|\delta x_i(t)|$ assumes
595: an exponential profile \cite{TGP95}.
596: 
597: For what concerns the propagation velocity,  
598: an estimation of $V_p$ can be obtained for infinitesimal
599: perturbations by the evaluation of the so-called maximal 
600: comoving Lyapunov exponents $\Lambda(v)$ \cite{DK87}. 
601: The time evolution of an intially localized (infinitesimal) 
602: disturbance (\ref{eq:tzero}) in a reference frame moving 
603: with velocity $v$ can be expressed as 
604: \begin{equation}
605: |\delta x_i(t)| \sim d_0 e^{\Lambda(v)t}\,,
606: \label{eq:growth}
607: \end{equation}
608: by following the perturbation along the world line $i=v t$
609: one can easily measures the corresponding
610: comoving Lyapunov exponent $\Lambda(v)$ (for more
611:  details see \cite{DK87,TP92}).
612: The information propagation velocity is the maximal
613: velocity for which a disturbance still propagates 
614: without being damped. Therefore it can be defined through
615: the following marginal stability criterion \cite{DK87}:
616: \begin{equation}
617: \Lambda(V_L) \equiv 0\,,
618: \label{eq:lvzero}
619: \end{equation}
620: where the velocity has been now indicated with $V_L$ in order
621: to stress that it has been obtained via 
622: a linear analysis. For the maps considered
623: in this section the identity $V_p = V_L$ is always fulfilled.
624: % fig 4
625: 
626: %%%%%%%%%%%%%%%%%%%%%%%%
627: \begin{figure}
628: \epsfxsize=8.truecm
629: \epsfysize=6truecm
630: \epsfbox{Fig4.eps}
631: %\narrowtext
632: \caption{Comparison between the directly measured propagation 
633: velocities $V_p = V_L$ (circles)
634: and the prediction (\ref{eq:velocity}) (boxes) for a CML of 
635: logistic maps with (a)
636: $a=3.9$ and (b) $a=4$, and of tent maps with (c) $a=2$ (d) $a=1.8$.
637: Lattices of $4\cdot 10^4$ maps has been used.}
638: \label{fig:vel}
639: \end{figure}
640: %%%%%%%%%%%%%%%%%%%%%%%%                                       
641: As shown in \cite{DK87}, in a closed system with symmetric coupling
642: $\Lambda=\Lambda(v)$ has typically a concave shape, 
643: with the maximum located at $v=0$ (in particular
644: $\lambda = \Lambda(v=0)$). An approximate expression 
645: can be obtained for $\Lambda(v)$, by substituting 
646: $i=vt$ in (\ref{eq:shape}) and by comparing
647: it with (\ref{eq:growth}) :
648: \begin{equation}
649: \Lambda(v)=\lambda-{v^2 / 2 \varepsilon} \quad .
650: \label{eq:comapr}
651: \end{equation} 
652: This parabolic expression for
653: $\Lambda(v)$ suffers of the same limits mentioned for 
654: the Gaussian approximation (\ref{eq:shape}) for the
655: disturbance evolution. 
656: Anyway, from (\ref{eq:comapr}) an analytical
657: prediction can be obtained for $V_L$ :
658: \begin{equation}
659: V_A=\sqrt{2\varepsilon \lambda}\,,
660: \label{eq:velocity}
661: \end{equation} 
662: which, as shown in  Fig.~\ref{fig:vel}, is indeed very good 
663: apart from some deviations for 
664: $\varepsilon \approx 0$ and $\varepsilon \approx 1$.
665: In Sect.~\ref{sec:5} we will rederive (\ref{eq:velocity}) 
666: by assuming that the chaotic perturbation 
667: behaves as a front connecting a stable and 
668: an unstable (metastable) fixed point
669: in a non-chaotic reaction diffusion system.
670: 
671: Let us briefly recall that another method (not suffering for
672: boundary problems) to determine the comoving Lyapunov exponent
673: has been introduced in~\cite{TP92}. The method relies on
674: the computation of {\it specific} Lyapunov exponents $\lambda(\mu)$ 
675: associated to an exponentially decaying perturbation (with spatial
676: decay rate $\mu$). In other words
677: one assumes that the spatio-temporal evolution of an infinitesimal
678: disturbance can be written as 
679: \begin{equation}
680: |\delta x_i(t)| \sim d_0 e^{\lambda(\mu)t-\mu i}\,.
681: \label{eq:specific}
682: \end{equation}
683: Since the asymptotic leading edge of the front separating perturbed 
684: from unperturbed part of the chain has an exponential shape,
685: the above assumption (\ref{eq:specific}) is appropriate to
686: study its evolution.
687: 
688: It is straightforward to show that the comoving Lyapunov
689: exponents are related to the specific ones via a Legendre
690: transform~\cite{TP92}, all the data concerning comoving exponents
691: reported in this paper have been obtained with such a method. Moreover,
692: a further results concerns the linear velocity $V_L$, it
693: can be shown \cite{TGP95} that
694: its value corresponds to the minimal propagation velocity
695: $V(\mu) = \lambda(\mu) / \mu$ associated with perturbations
696: of the form (\ref{eq:specific}), i.e.
697: \begin{equation}
698: V_L = \min_\mu \frac{\lambda(\mu)}{\mu} \equiv \frac{\lambda_L}
699: {\mu_L}
700: \label{eq:vmu}
701: \end{equation}
702: where $\mu_L$ and $\lambda_L=\lambda(\mu_L)$ represent the
703: spatial decay rate and the temporal growth rate of the leading
704: edge, respectively. The expression (\ref{eq:vmu}) for the linear
705: velocity is identical to the one derived for propagation of fronts
706: connecting a stable to an unstable steady state~\cite{wim1}. 
707: 
708: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
709: \subsection{Non Linear Mechanisms}
710: \label{sec:3.2}                   
711: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
712: In this section we investigate the case of coupled maps for which  
713: $\lambda(\delta) > \lambda(0)$ in some interval of $\delta$.
714: As noticed in Sect.~\ref{sec:2}, 
715: this behavior can be observed in chaotic (absolutely unstable) maps,
716: as well as in stable and marginally stable maps. Let us first
717: analyze chaotic maps, non-chaotic ones will be discussed
718: in Sect. \ref{sec:4.2}.
719: 
720: For these systems it is possible to have $V_p >  V_L$, 
721: this means that the disturbance can still propagate 
722: also in the velocity range $[V_L,V_p]$, 
723: even if the corresponding comoving Lyapunov exponents 
724: are negative. Therefore,
725: the linear marginal stability 
726: criterion (\ref{eq:lvzero}) does not hold anymore. 
727: We want to stress that the condition (\ref{eq:nonlin})
728: is necessary, but not sufficient to ensure that $V_p > V_L$,
729: since all the details of the coupled model play also an important role.
730: 
731: In Fig.~\ref{fig:nonlinear} the spatio-temporal evolution of an
732: initially localized disturbance of a chain of coupled shift
733: maps is reported.
734: As shown in \cite{isola}, when the coupling $\varepsilon \le 1/2$ the
735: maximal Lyapunov exponent for such model coincides with that
736: of the single map (namely, $\lambda=\ln(\beta)$) and if $\beta > 1$
737: the system is chaotic. Initially the perturbation, that is still
738: infinitesimal, spreads with the linear velocity $V_L$
739: above defined. At later time it
740: begins to propagate faster with a velocity $V_P > V_L$.
741: Comparing  Fig.~\ref{fig:nonlinear} with
742: Fig.~\ref{fig:torc.1}, one can see that the second stage of
743: the propagation sets in when the bulk of the perturbed region 
744: reaches sufficiently high values. As a matter of fact the initial 
745: stage of propagation disappears if we initialize the two replicas
746: with a disturbance of amplitude ${\cal O}(1)$. From these facts it is
747: evident that the origin of the information propagation characterized
748: by $V_p > V_L$ should be due to the strong nonlinear effects present
749: in this type of CML.
750: % fig 5
751: 
752: %%%%%%%%%%%%%%%%%%%%%%%%
753: \begin{figure}
754: \epsfxsize=8.truecm
755: \epsfysize=6truecm
756: \epsfbox{Fig5.eps}
757: %\narrowtext
758: \caption{The same of Fig.~\ref{fig:torc.1} for a lattice of coupled shift maps
759: with $\beta=1.03$,  $\varepsilon=1/3$.}
760: \label{fig:nonlinear}
761: \end{figure}
762: %%%%%%%%%%%%%%%%%%%%%%%% 
763: The behavior at long times can be understood by considering
764: the dependence of $\lambda(\delta)$ on the disturbance amplitude 
765: as shown in Fig.~\ref{fig:maps}b:
766: actually  the figure refers to the single map with $\beta=2$, but the 
767: shape of $\lambda(\delta)$ is qualitatively the same also for 
768: the coupled system and for other $\beta$-values.
769: 
770: Until the perturbation is infinitesimal $\lambda(\delta) \simeq \lambda$
771: and the linear analysis applies, when the disturbance becomes
772: bigger than a certain amplitude, $\delta^{NL}$, 
773: the growth will be faster, since
774: now $\lambda(\delta) > \lambda$. As it can be clearly seen in
775: Fig.~\ref{fig:invasion}, the perturbation is well reproduced by 
776: the linear approximation (\ref{eq:shape}) until 
777: the amplitude of the perturbation reaches a critical value
778: $\delta^{NL} \sim O(10^{-4})$ above which the nonlinear effects set in. 
779: At this stage
780: the nonlinear instabilities begin to push the front leading to
781: an increase of its velocity and deforming the profile of the perturbation.
782: This becomes exponential at much shorter times than in the linear
783: situation discussed in previous subsection. Moreover,
784: when the propagation is dominated by nonlinear mechanisms
785: the spatial decay rate $\mu_{NL}$ of the asymptotic leading edge
786: will be greater of the linear expected value $\mu_L$: this
787: result can be explained again invoking the analogy with
788: propagation of fronts connecting steady states \cite{TGP95}.
789: %fig 6
790: %%%%%%%%%%%%%%%%%%%%%%%%
791: \begin{figure}
792: \epsfxsize=8.truecm
793: \epsfysize=6truecm
794: \epsfbox{Fig6.eps}
795: %\narrowtext
796: \caption{Evolution of the perturbation $\Delta x_i(t)$ for a CML 
797: of  shift maps with
798: $\varepsilon=1/3$ and $\beta=1.04$ at four different times, $t=250,450,650,1000$.
799: The solid lines are the expected Gaussian  shape (\ref{eq:shape}).
800: The decay rate of the asymptotic exponential profile is 
801: $\mu_{NL} \sim 1.47$, noticeably greater than the linear 
802: value $\mu_L = 0.42$.}
803: \label{fig:invasion}
804: \end{figure}
805: %%%%%%%%%%%%%%%%%%%%%%%%
806: %fig 7
807: %%%%%%%%%%%%%%%%%%%%%%%%
808: \begin{figure}
809: \epsfxsize=8.truecm
810: \epsfysize=6truecm
811: \epsfbox{Fig7.eps}
812: %\narrowtext
813: \caption{Time evolution of $\Delta x_0(t)$ (solid line) and $\Delta
814: x_i(t)$ (dashed line), for $i=V_2 t$ with $V_2=1/3$. 
815: In this last case the data for two different chain configurations are reported.
816: The map used is the shift map with $\beta=1.1$, $\varepsilon=1/3$ and
817: $L=2\cdot 10^3$. For these parameters one has $V_L=0.250$ and $V_p=0.342$.
818: Note that between $t=200$ and $t=500$ for $\Delta x_i(t)$
819: the numerical precision is reached.}
820: \label{fig:bho}
821: \end{figure}
822: %%%%%%%%%%%%%%%%%%%%%%%% 
823: To better clarify the difference between the linear and non linear mechanisms
824: we show the behavior of $\Delta x_i(t)$ along  the world line $i=V t$ 
825: for two different velocities:  $V_1=0$ and $V_L<V_2<V_p$.
826: From Fig.~\ref{fig:bho}, one observes  for the zero-velocity situation
827: an exponential increase of the perturbation with rate $\lambda$
828: until it eventually saturates. For the case corresponding to
829: velocity $V_2$ an initial exponential decay with rate $\Lambda(V_2)$
830: is seen, followed at later times by a resurgence of the disturbance.
831: The successive evolution of the perturbation is no more exponential and 
832: exhibits strong fluctuations. 
833: 
834: These features suggest that in order to generalize the 
835: criterion (\ref{eq:lvzero}) to nonlinear driven information spreading
836: the growth of finite amplitude perturbations in a moving reference frame
837: should be analyzed.
838: 
839: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
840: \section{Finite Size Lyapunov Exponent: \\
841: Extended Systems}
842: \label{sec:4}
843: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
844: In this section we  introduce the 
845: Finite Size Comoving Lyapunov Exponent (FSCLE) that is
846: a generalization of the  FSLE to a moving reference frame.
847: First of all let us define the FSLE for an extended
848: system: in this case exactly the same definition
849: given in Sect.~\ref{sec:2} applies, apart from
850: some ambiguities in the choice of the
851: norm  to employ for measuring the distance
852: $\delta(t)$ of the perturbed, 
853: ${\bf x}^\prime(t)=\{x_i^\prime(t)\}$, 
854: from the unperturbed replica, ${\bf x}(t)=\{x_i(t)\}$. 
855: A natural choice could be to perturb randomly
856: ${\bf x}(t)$ and to look for the doubling times associated
857: to the evolution of the distance
858: \begin{equation}
859: {\tilde \Delta} (t) = \frac{1}{L} \sum_{i=-L/2}^{L/2} |x_i^\prime(t)-x_i(t)| 
860: \quad ;
861: \label{norm1}
862: \end{equation}
863: an alternative choice consists in
864: perturbing a single site of the chain, let's say $i=0$, 
865: at time $t=0$ and to evaluate the ``single site'' norm
866: \begin{equation}
867: \Delta x_0 (t) = |x_0^\prime(t)-x_0(t)| 
868: \quad .
869: \label{norm2}
870: \end{equation}
871: We have verified that the two norms (\ref{norm1}) and
872: (\ref{norm2}) give equivalent results for what concerns
873: the evaluation of $\lambda(\delta)$. In this paper
874: we will limit to consider the norm (\ref{norm2}).
875: 
876: In order to measure the FSCLE in a reference frame moving
877: with velocity $v$, we have simply measured
878: the difference (\ref{norm2}) along a world line $i=v t$;
879: i.e.
880: \begin{equation}
881: \Delta x_{i_c+[vt]} (t)  = |x_{i_c+[vt]}^\prime(t)-x_{i_c+[vt]}(t)| 
882: \quad .
883: \label{normv}
884: \end{equation}
885: where $[ \, \cdot \,  ]$ denotes the integer part and 
886: $i_c$ is introduced below.
887: 
888: The FSCLE is then estimated as in Sect.~\ref{sec:2}. Once 
889: a set of thresholds $\delta_n = r^n \delta_0$, with
890: $n=0,\dots,N$, is chosen and the perturbation is initialized 
891: as $\Delta x_i(0)=
892: \delta_{min} \, \delta_{i,0}$ with $\delta_{min}\ll \delta_0$. 
893: A preliminary transient evolution is performed in order to
894: allow to the perturbed orbit to relax on the attractor. 
895: At the end of this short transient the position $i_c$
896: where the perturbation reaches its maximum is detected,
897: this point is taken as the vertex of the light cone from
898: which all the considered world lines $i = v \, t$ depart.
899: 
900: The FSCLE is then defined as
901: \begin{equation}
902: \Lambda(\delta_n,v)={1 \over \langle \tau(\delta_n)\rangle_e} \left\langle 
903: \log\left({{\Delta} \over {\delta}_n}\right)
904: \right\rangle_e\,,
905: \end{equation}
906: where the dependence on the velocity $v$ derives from the employed 
907: norm (\ref{normv}). Note that we have used Eq.~(\ref{eq:2.5})
908: due to the discreteness of the temporal evolution.
909: 
910: In the limit of very small perturbation the FSCLE reduces to the comoving
911: Lyapunov exponent 
912: \begin{equation}
913: \lim_{\delta \to 0} \Lambda(\delta,v) =\Lambda(v)\,,
914: \label{eq:fslim}
915: \end{equation}
916: and, for $v=0$ one has the FSLE $\lambda(\delta)$. 
917:  
918: Actually there are finite time effects which prevents the limit 
919: (\ref{eq:fslim}) to be correctly attained.
920: This is related to the fact that the FSCLE's can be obtained 
921: only via finite time measurements. In the Appendix
922: we show how one can include such finite time corrections.
923: 
924: In the next subsections we will give evidences that 
925: the marginal {\it linear} stability criterion
926: (\ref{eq:lvzero}) can be generalized with the aid
927: of the FSCLE in the following way:
928: \begin{equation}
929: \max_{\delta} \{ \Lambda(\delta,v) \} =0\qquad {\mbox {for}} \, v \ge V_p\,,
930: \label{eq:crit}
931: \end{equation}
932: where $V_p$ can be either $V_L$, if the information
933: propagation is due to linear mechanisms, or greater
934: than $V_L$, when nonlinear mechanisms prevail on the
935: linear ones. For $v > V_p$ one has $\Lambda(\delta,v)=0$, 
936: since due to the definition of the
937: FSCLE negative growth rate appear to be $0$.
938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
939: \subsection{Chaotic Systems}
940: \label{sec:4.1}
941: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
942: Let us first consider chaotic systems for which
943: $V_p \equiv V_L$, in this case 
944: $$\max_{\delta} \{ \Lambda(\delta,v) \} = \Lambda(v)$$
945:  and the generalized criterion (\ref{eq:crit}) reduces to the linear
946: one (\ref{eq:lvzero}). 
947: 
948: As already stressed in Sect. \ref{sec:3.2}, shift coupled
949: maps represent a prototype of the class of chaotic models 
950: for which $V_p$ can be eventually bigger than $V_L$.
951: For this model the behavior of $\Lambda(\delta,v)$ for various 
952: velocities is reported in Fig.~\ref{fig:fscle1}.
953: 
954: For $v<V_L$, we observe that $\Lambda(\delta,v) \sim \Lambda(v)$ up to a 
955: certain value of the disturbance amplitude $\delta^{NL}$, above which
956: the FSCLE increases and exhibits a clear peak at some higher $\delta$-value.
957: From Fig.~\ref{fig:fscle1} (a) it is clear that
958: $\delta^{NL}$ (which denotes the set in of the regime dominated
959: by nonlinear mechanisms) decreases for increasing velocities and finally
960: vanishes at $v = V_L$. This behavior is reasonable since, as shown in 
961: Fig.~\ref{fig:nonlinear}, initially the disturbance evolves
962: along the chain following the linear mechanism 
963: characterized by $\Lambda(v)$, but as soon as in the central site
964: $\delta > \delta^{NL}$ the nonlinear mechanism begins to be active. 
965: Thus a nonlinear front is excited and this invades the linear region
966: propagating with a velocity higher than $V_L$, therefore the higher is $v$ 
967: the smaller is the scale at which nonlinear effects are observed.
968: 
969: In the interval $V_L \leq v \leq V_p$ (see Fig.~\ref{fig:fscle1}(b)), 
970: $\Lambda(\delta,v)$ is still 
971: positive and its maximal value decreases for increasing $v$. 
972: The FSCLE vanishes for $v=V_p$ as Fig.~\ref{fig:vlvnl} (a) shows. 
973: In this velocity range $\Lambda(v)$ is always negative,
974: therefore the instabilities observed in a reference frame
975: moving with velocity higher than $V_L$ have a fully nonlinear
976: origin. The residual fluctuations present in $\Lambda(\delta,v)$ are 
977: essentially due to the limited statistics.
978: Indeed the nonlinear growth as shown in
979: Fig.~\ref{fig:bho} is extremely fluctuating.
980: %fig 8
981: 
982: %%%%%%%%%%%%%%%%%%%%%%%%
983: \begin{figure}
984: \epsfxsize=8.truecm
985: \epsfysize=6truecm
986: \epsfbox{Fig8a.eps}
987: \epsfxsize=8.truecm
988: \epsfysize=6truecm
989: \epsfbox{Fig8b.eps}
990: %\narrowtext
991: \caption{$\Lambda(\delta,v)$ as a function $\delta$ 
992: for various velocities. The data
993: refer to the coupled shift maps with  $\beta=1.1$ and 
994: $\varepsilon=1/3$, with these parameters
995: $V_L \sim 0.250$ and $V_p \sim 0.342$.
996: In (a) results for velocities $v < V_L$ are reported,
997: namely (from top to bottom) $v=0$, $v=0.10$ and $v=0.16$. 
998: The straight lines indicate $\Lambda(v)$. 
999: In (b) velocities in the range $[V_L : V_p]$ are reported, from
1000: the top $v=0.25$, $v=0.286$, $v=0.30$, $v=0.33$, $v=0.34$. 
1001: Chains of lengths from $L=10^4$ up to $L=10^5$
1002: have been employed and the statistics is 
1003: over $2 \cdot 10^3$ doubling times.}
1004: \label{fig:fscle1}
1005: \end{figure}
1006: %%%%%%%%%%%%%%%%%%%%%%%% 
1007: In Figure~\ref{fig:vlvnl}(a) the dependence of 
1008: $\max_{\delta}\{\Lambda(\delta,v)\}$ on $v$ for coupled shift maps is 
1009: reported, as expected it vanishes exactly for $v=V_p$.  
1010: For $v < V_L$ a non-monotonous behavior of
1011: $\max_{\delta}\{\Lambda(\delta,v)\}$ is observable,
1012: this is probably due to the complex interplay of linear and nonlinear
1013: effects. For $v > V_L$, 
1014: the behavior is smoother and a monotonous decrease is observed.
1015: As discussed in Sect.~\ref{sec:2}, the discontinuity present
1016: in the shift map is not necessary in order to observe
1017: nonlinear mechanisms prevailing on linear ones.
1018: 
1019: As an example of continuous map exhibiting an information
1020: propagation velocity $V_p > V_L$
1021: the map (\ref{eq:stable}) is considered. 
1022: This type of CML has been already studied in \cite{TP94}:
1023: it has been observed that in a certain parameter range
1024: $V_p$ can be finite even if the map is non chaotic. 
1025: Moreover, 
1026: also in the chaotic regime there is a window of parameters
1027: where $V_p > V_L$, with our parameters choice $\lambda \approx 0.182>0$.
1028: 
1029: %fig 9
1030: 
1031: %%%%%%%%%%%%%%%%%%%%%%%%
1032: \begin{figure}
1033: \epsfxsize=8.truecm
1034: \epsfysize=6truecm
1035: \epsfbox{Fig9a.eps}
1036: \epsfxsize=8.truecm
1037: \epsfysize=6truecm
1038: \epsfbox{Fig9b.eps}
1039: %\narrowtext
1040: \caption{$\max_{\delta}\Lambda(\delta,v)$ 
1041: (dashed line with points) versus $v$ 
1042: compared with $\Lambda(v)$ (continuous line).
1043: (a) Results for coupled shift maps with the same parameters as Fig.~\ref{fig:fscle1},  
1044: the vertical lines indicates  $V_L \approx 0.250$ and the 
1045: measured (\ref{eq:vel1}) propagation velocity $V_p \approx 0.342$. 
1046: (b) Data for the coupled maps (\ref{eq:stable}) with parameters 
1047: $b=2.7$, $d=0.1$, $q=0.07$ 
1048: and $c=500$ and with coupling $\varepsilon=2/3$,
1049: the vertical lines indicates $V_L
1050: \approx 0.39$ and $V_p \approx 0.59$. 
1051: In this second case chains of length $L=2\cdot 10^4$ have been considered
1052: and the statistics is over $2 \cdot 10^4$ doubling times.                            
1053: The reported values for $\max_{\delta}\Lambda(\delta,v)$ 
1054: refer to an average over five values of $\delta$ around the 
1055: peak position of $\Lambda(\delta,v)$.
1056: }
1057: \label{fig:vlvnl}  
1058: \end{figure}
1059: %%%%%%%%%%%%%%%%%%%%%%%%
1060: Also in the present case we observe an overall behavior
1061: of the FSCLE resembling that of the coupled shift maps.
1062: The main point that we want to remark is that in the limit
1063: $v \to V_p$  $\Lambda(\delta,v) \to 0$,
1064: as shown in Fig.~\ref{fig:vlvnl} (b). 
1065: At variance with the shift map (that is an everywhere
1066: expanding map) the considered map shows contracting 
1067: and expanding intervals. Therefore the disturbances
1068: during their spatio-temporal evolution can be
1069: alternatively expanded or contracted. This leads
1070: to strong fluctuations in the FSCLE-values, that are 
1071: difficult to remove. 
1072: 
1073: As a matter of fact we observed
1074: that even for velocities slightly larger than $V_p$
1075: $\Lambda(\delta,v)$ can be non zero. But the statistics
1076: of these anomalous fluctuations is extremely low,
1077: referring to the parameters reported in Fig.~\ref{fig:vlvnl} (b)
1078: for $v=0.6 > V_p = 0.59$ we observed an expansion (instead
1079: of the expected contraction) of disturbances in the 
1080: $0.5 \%$ of the studied
1081: cases. For higher velocities $\Lambda(\delta,v)$ is
1082: zero for any considered $\delta$.
1083: 
1084: 
1085: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1086: \subsection{Non Chaotic Systems}
1087: \label{sec:4.2}
1088: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1089: On the basis of the linear analysis discussed in Sect.~\ref{sec:3.1}
1090: the propagation of disturbances should be present only
1091: in chaotic systems. 
1092: For non chaotic ones $\lambda \leq 0$ and from Eq.~(\ref{eq:lvzero}) 
1093: one trivially obtains $V_L=0$. 
1094: On the other hand, in the class of systems for which 
1095: $\lambda(\delta)\geq \lambda$ 
1096: there are also stable and marginally stable systems. Therefore,
1097: a propagation due solely to non linear terms can
1098: still be present \cite{PLOK93,TP94,TGP95}. 
1099: In this section we want to discuss 
1100: the possible employ of the FSLE  to characterize these maps.
1101: 
1102: Before entering into the description of the FSLE computation in such systems,
1103: it is of interest to recall an important phenomenon which appears in 
1104: stable systems: the so-called ``stable chaos'' \cite{PLOK93}. 
1105: Stable systems asymptotically  evolve towards 
1106: trivial attractors (i.e. fixed points, periodic or quasi-periodic orbits).
1107: However, in spatially extended systems it may happen that the time needed 
1108: to reach the asymptotic state is very long: it has been found that in certain 
1109: stable CML the transient time diverges exponentially with the number of 
1110: elements of the chain \cite{CK88}. 
1111: Moreover, this transient is characterized by a quasi-stationary
1112: behavior allowing for the investigation of the properties of the 
1113: model with statistical consistency. In Ref. \cite{PLOK93}
1114: it has been shown that a chain of coupled maps of the type (\ref{eq:stable}),
1115: considered in their discontinuous limit (i.e. for $c \to \infty$),
1116: is non chaotic but still exhibits erratic behaviors. This
1117: is associated to a non-zero information spreading within
1118: the system.
1119: 
1120: 
1121: As far as the computation of the FSLE for these systems is concerned,
1122: some remarks are worth to be done. The definition of the
1123: FSLE in term of error doubling times cannot be used 
1124: in a straightforward manner to determine negative expansion rates.
1125: Another important point is that, at variance with the case 
1126: of chaotic maps, finite perturbations should be now considered
1127: in order to observe an expansion.
1128: 
1129: These two points impede a straightforward implementation of our method to
1130: study these systems. Indeed, if one starts with too small perturbations  
1131: the propagation does not manifest, while if one initializes the system with a
1132: finite perturbation 
1133: $\lambda(\delta)$ cannot be estimated with the required accuracy. 
1134: Indeed, the only way to have an independence of $\lambda(\delta)$ from 
1135: the initial conditions is to initialize the system with infinitesimal
1136: perturbations and then to follow them until they become finite due to
1137: the dynamics of the system (see Ref.~\cite{ABCPV96} for a detailed discussion).
1138: 
1139: The coupled (\ref{eq:stable}) maps for $c\to \infty$ have been recently
1140: analyzed by Letz \& Kantz \cite{Kantz} in term of 
1141: an indicator similar to the FSLE (i.e. able to quantify 
1142: the growth rate of non infinitesimal perturbations). This
1143: indicator turns out to be negative for infinitesimal 
1144: perturbations and becomes positive for finite perturbations.
1145: This means that a finite perturbation of sufficient amplitude 
1146: can propagate along the system due to nonlinear effects.
1147: This confirms what was previously observed
1148: in Ref.~\cite{TGP95} for marginally stable systems.
1149: %fig 10
1150: 
1151: %%%%%%%%%%%%%%%%%%%
1152: \begin{figure}
1153: \epsfxsize=8.truecm
1154: \epsfysize=6truecm
1155: \epsfbox{Fig10.eps}
1156: %\narrowtext
1157: \caption{$I(\delta)$ as a
1158: function of the amplitude perturbation in
1159: a lin-log scale for the discontinuous coupled 
1160: maps (\ref{eq:stable}), in the limit $c \to \infty$, with
1161: $\varepsilon=1/3$ and $L=35$. The map is reported in the inset.
1162: The negative maximal Lyapunov $\lambda = -0.105$ is recovered
1163: at small scales. For the computation details see the text.
1164: In the inset the single map is shown.}
1165: \label{fig:stable}                                      
1166: \end{figure}
1167: %%%%%%%%%%%%%%%%%%%
1168: In Fig.~\ref{fig:stable} we show the behavior of 
1169: a quantity $I(\delta_n)$ similar to $\lambda(\delta_n)$ which 
1170: has been obtained as follows.
1171: We considered two trajectories at an initial distance $\delta_n$,
1172: after one time step evolution the distance $\delta$ 
1173: between the trajectory is measured. Then one of the
1174: two trajectory is rescaled at a distance 
1175: $\delta_n$ from the other, keeping the direction of the perturbation unchanged,
1176: and the procedure is repeated several times and for several
1177: values of $\delta_n$.
1178: Then we averaged $\ln(\delta/\delta_n)$ over many different initial conditions
1179: obtaining $I(\delta_n)$. For $\delta_n \to 0$, this is nothing 
1180: but  the usual algorithm for computing the maximal Lyapunov exponent~\cite{benettin}.
1181: 
1182: As discussed in Ref.~\cite{ABCPV96} this method suffers of the problem that 
1183: when $\delta_n$ is finite one is not sure to sample correctly the measure on the 
1184: statistically stationary state. Indeed a finite perturbation will generically bring 
1185: the trajectory out from the ``attractor''. Nevertheless the result is in good agreement 
1186: with the one obtained in \cite{Kantz} and confirms that at the origin 
1187: of the perturbation propagation in this system there should be a mechanism very 
1188: similar to the one discussed for the shift map. Further studies related to the
1189: ``stable chaos'' phenomenon have been recently performed~\cite{ginelli}.
1190: 
1191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1192: \section{Analogies with front propagation in reaction diffusions systems}
1193: \label{sec:5}
1194: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1195: The perturbation evolution in spatially distributed systems 
1196: can be described as the motion of an interface separating 
1197: perturbed from unperturbed regions. 
1198: In this spirit, one can wonder if and to what extent it is possible to draw
1199: an analogy between the evolution of this kind of interface and the propagation 
1200: of fronts connecting steady states in reaction diffusion systems. 
1201: As already noticed in Ref.~\cite{TGP95}, the two phenomena display many similarities.
1202: In the following we will discuss the similarities and differences, in particular 
1203: we will introduce a simple phenomenological model which can help us in highlighting
1204: the analogies.
1205: 
1206: Let us start by recalling the basic features of fronts propagation in 
1207: reaction diffusion systems  with reference to
1208: the Fisher-Kolmogorov-Petrovsky-Piscounov (FKPP) equation \cite{KPP}
1209: \begin{equation}
1210: \partial_t \theta(z,t) = D\partial^2_{zz}\theta(z,t) + G(\theta(z,t))\,,
1211: \label{eq:kpp}
1212: \end{equation}
1213: where $\theta(z,t)$ represents the concentration of a reactants which diffuses and 
1214: reacts, the chemical kinetics is given by $G(\theta)$. Typically the function 
1215: $G(\theta)\in C^1[0,1]$ (with $G(0)=G(1)=0$) exhibits
1216: one stable ($\theta=1$) and one unstable ($\theta=0$) fixed point.
1217: Once the system is prepared on the stable state ($\theta(z)=0 \, \forall z$),
1218: an initial (sufficiently steep) perturbation (e.g. a step function
1219: $\theta(z,t=0) = \Theta{(z-z_0)}$ ) will give rise to a smooth front 
1220: moving with a velocity $V_p$, that will connect
1221: the unstable and the stable fixed points:
1222: as a result the stable state will invade the unstable one.
1223: This equation admits many different traveling solutions that
1224: are typically characterized by their propagation velocities,
1225: however for a sufficiently steep initial perturbation
1226: of the unstable state the selected front is unique
1227: and its velocity $V_p$ is bounded in the interval $[V_{min}, V_{max}]$,
1228: where
1229: \begin{equation}
1230: V_{min}=2\sqrt{D G^{'}(0)}
1231: \label{eq:vmin}
1232: \end{equation}
1233: and
1234: \begin{equation}
1235: V_{max}=2\sqrt{D
1236: \sup_{0<\theta<1}\left\{ {G(\theta) \over \theta} \right\}}\,.
1237: \label{eq:vmax}
1238: \end{equation}  
1239: If $G(\theta)$ is concave $\sup_{0<\theta<1}\{G(\theta)/\theta\}$ is attained
1240: at $\theta=0$ (i.e. $\sup_{0<\theta<1}\{G(\theta)/\theta\}=G^{'}(0)$) and the
1241: selected velocity is always the minimum one: the front is ``pulled'' by the 
1242: growth of infinitesimal perturbations of the unstable state.
1243: Otherwise, if $G(\theta)$ is convex one can observe a velocity
1244: $V_p > V_{min}$: the front is now ``pushed'' by the growth 
1245: of finite amplitude perturbations. These
1246: results are known as the Aronson \& Weinberger theorem
1247: \cite{ara}
1248: (for more details see Ref. \cite{wim2}).
1249: The velocity $V_{min}$ can be easily obtained by performing a
1250: linear analysis of (\ref{eq:kpp}) and by employing a marginal
1251: stability criterion \cite{wim1}.
1252: 
1253: The subject of our analysis is the spreading of perturbations 
1254: in a chaotic media. In order to compare our case with
1255: the FKPP, one we should consider the time evolution of the difference 
1256: of two chaotic trajectories $\{\Delta x_i(t)\}_{i=1,L}$. 
1257: However, the nature of the two phases separated by the front is now different from 
1258: the FKPP case. 
1259: The interface separates an unstable ($\Delta x_i=0$) state from a ``statistically
1260: stable'' one. With the term ``statistically stable'' we mean that behind the front,
1261: in the bulk of the perturbed region, $\Delta x_i(t)$ does not converge to a stable 
1262: fixed point (as for the FKPP) but it fluctuates in a stationary way
1263: around an average value. This suggests that a good model for reproducing
1264: this dynamical evolution would be a FKPP with a stochastic kinetics~\cite{armero}.
1265: However, many similarities with FKPP can be
1266: established by neglecting the chaotic fluctuations and
1267: considering the average shape of the front
1268: \cite{TGP95}. In practice, this can be done by averaging
1269: the perturbation evolution over many different initial conditions.
1270: 
1271: Once the chaotic fluctuations are neglected, one can 
1272: express the average perturbation growth in any site of the 
1273: chain via the following mean-field approximation:
1274: \begin{equation}
1275: u_i(t+1)=\exp[\lambda(\tilde{u}_i(t))]\, \tilde{u}_i(t)\,,
1276: \label{eq:model}
1277: \end{equation}
1278: where $i$, $t$ are the discrete space and time indices and 
1279: $\tilde{u}_i=(1-\varepsilon)u_i+\varepsilon/2(u_{i+1}+u_{i-1})$. Here, 
1280: $u_i(t)$ indicates the ``average'' 
1281: $\Delta x_i(t)$ and  $\lambda(u)$ the corresponding
1282: FSLE (at least for positive growth rates).
1283: In the limit $u\to 0$  $\lambda(u)\to \lambda$, while,
1284: for $u \sim {\cal O}(1)$, $\lambda(u)$ should reflect the 
1285: saturation effects associated to the nonlinear map. 
1286: In the infinitesimal limit ($u \to 0$)
1287: one essentially recovers the model discussed in Ref.~\cite{PP98}.
1288: 
1289: By passing to continuous variables is possible to
1290: show that Eq.~(\ref{eq:model}) can be 
1291: reduced to Eq.~(\ref{eq:kpp}) (at least at the
1292: leading order).
1293: In order to transform Eq.~(\ref{eq:model}) in its
1294: continuous version,
1295: let us introduce infinitesimal spatial and temporal 
1296: resolutions $dx$ and $dt$, and  assume that the
1297: diffusive scaling holds, i.e. $dx^2=dt$. 
1298: Limiting to the first order expansion in $dt$
1299: (second order in $dx$), one easily shows that
1300: the continuous counterpart of Eq.~(\ref{eq:model}) is
1301: \begin{equation}
1302: \partial_t u= \lambda(u)u+\frac{\varepsilon}{2} \partial^2_{xx}u\,,
1303: \label{eq:model_kpp} 
1304: \end{equation}
1305: which is nothing but (\ref{eq:kpp}) with 
1306: \begin{equation}
1307: \lambda(u)={G(u)\over u}\,,
1308: \label{eq:mapping}
1309: \end{equation}
1310: and $D=\varepsilon/2$. 
1311: As a consequence of the identity (\ref{eq:mapping}),
1312: one should observe a pulled dynamics for the
1313: chaotic front if $\lambda \ge \lambda(u) \quad \forall u \ge 0$,
1314: and a pushed one could occur
1315: only if $\max_{u}\{\lambda(u)\} > \lambda$.
1316: This can be considered as a reformulation of the 
1317: Aronson \& Weinberger theorem \cite{ara} in the
1318: context of information propagation.
1319: The velocity bounds (\ref{eq:vmin}),
1320: (\ref{eq:vmax}) can be now identified with 
1321: $V_a = \sqrt{2\varepsilon \lambda}$, i.e. Eq.~(\ref{eq:velocity}), 
1322: the first one and with
1323: \begin{equation}
1324: V_s=\sqrt{2\varepsilon \max_{u}\{\lambda(u)\}}
1325: \end{equation}
1326: the second one. We stress again that the lower velocity
1327: bound is indeed represented by $V_L$ and that it
1328: coincides with $V_a$ only for sufficiently strong 
1329: diffusive coupling $\varepsilon$.
1330: 
1331: Another interesting point that we can investigate via a 
1332: numerical simulation of the mean field effective 
1333: equation (\ref{eq:model})
1334: is the dependence of the propagation properties on the
1335: specific shape of $\lambda(u)$. 
1336: For instance, as previously conjectured, we expect that 
1337: if a monotonous decreasing $\lambda(u)$ is considered
1338: one should observe linear propagation, only.
1339: Indeed, numerical integrations of the model (\ref{eq:model}), 
1340: with the choice $\lambda(u)=A-Bu$, being $A$ and $B$ positive
1341: constants, show that $V_p =V_a= \sqrt{2 \varepsilon A}$,
1342: where $A$ corresponds to the maximal Lyapunov of the effective model.
1343: A generalization of this simple model would require to 
1344: consider $A$ and $B$ as fluctuating quantities generated by
1345: suitable stochastic processes. But while for the choice 
1346: of $A$ one can have some hint \cite{PP98}, this is not 
1347: the case for $B$. 
1348: %fig 11
1349: 
1350: %%%%%%%%%%%%%%%%%%%%%%%%
1351: \begin{figure}
1352: \epsfxsize=8.truecm
1353: \epsfysize=6truecm
1354: \epsfbox{Fig11.eps}
1355: %\narrowtext
1356: \caption{Propagation velocities 
1357: (symbols) for the model (\ref{eq:model}) 
1358: with $\lambda(u)$ given by
1359: (\ref{eq:mod1}) as a function of $B = \ln(r_2)$,
1360: with $\varepsilon=1/3$,  $A=\ln{1.1}$ and
1361: $C=A$. The threshold  values are fixed to
1362: $\delta^{NL}= 10^{-3}$ and $\delta^{sat}=0.53$
1363: and a chain of $4 \cdot 10^4$ sites has been used.
1364: The curves refer to $V_{a}=\sqrt{2\varepsilon A}$ (solid line)
1365: and $V_{s}=\sqrt{2\varepsilon B}$ (dashed line).
1366: Note that in the linear case, $B<A$,
1367: there is a perfect agreement between the measured velocity and the
1368: prediction. 
1369: }
1370: \label{fig:vlvnl_model}  
1371: \end{figure}
1372: %%%%%%%%%%%%%%%%%%%%%%%%
1373: Let us now consider the non linear propagation case, as we have
1374: previously seen a necessary condition in order
1375: to have $ V_p > V_L$ is that $\max_{u}\{\lambda(u)\} > \lambda$
1376: for some finite $u$. For what concerns the actual value of the
1377: selected velocity, this will depend on the value of the diffusive
1378: coupling and on the specific shape of $\lambda(u)$. 
1379: In the following we will examine the dependence of 
1380: $V_p$ on two quantities that characterize $\lambda(u)$:
1381: the difference $|\max_{\delta} \lambda(\delta) - \lambda|$ and the
1382: scale $\delta^{NL}$ at which the non linear effects set in.
1383: 
1384: As a first example let us consider $\lambda(u)$ as a step function:
1385: \begin{equation}
1386: \lambda(u)=\left\{ \begin{array}{ll}
1387: A & 0<u<\delta^{NL}\\ 
1388: B & \delta^{NL}\leq u <\delta^{sat}\\ 
1389: -C & u\geq \delta^{sat}\,,
1390: \end{array}
1391: \right.
1392: \label{eq:mod1}
1393: \end{equation}
1394: where $A$, $B$ and $C$ are positive quantities, with $B > A$,
1395: while $\delta^{NL}$ and $\delta^{sat}$ are amplitude
1396: thresholds. The parameter $A$ is nothing but the Lyapunov exponent, 
1397: $B$ mimics the non linear terms leading to
1398: an enhancement of the growth rate, and the last term mimics the 
1399: damping of the perturbation due to the saturation effects.
1400: In Fig.~\ref{fig:vlvnl_model} it is reported 
1401: the behavior of the propagation velocity
1402: for the model (\ref{eq:model}) with $\lambda(u)$ given by
1403: (\ref{eq:mod1}) for various values of $B$, once $A$ and $C$ are
1404: fixed. From the figure is evident that if $B < A$ then 
1405: $V_p \equiv V_a$ (i.e. we are in the linear regime), 
1406: while as soon as $B > A$ an increase of $V_p$
1407: with respect to $V_a$ is observed.
1408: % fig 12
1409: 
1410: %%%%%%%%%%%%%%%%%%%%%%%%
1411: \begin{figure}
1412: \epsfxsize=8.truecm
1413: \epsfysize=6truecm
1414: \epsfbox{Fig12.eps}
1415: %\narrowtext
1416: \caption{Propagation velocities $V_p$ as a function of $\delta^{NL}$
1417: for the model (\ref{eq:model}), (\ref{eq:mod1}) with $\varepsilon=1/3$, 
1418: $A=\ln(1.1)$, $B=\ln(1.3)$, $C=A$, $\delta^{sat}_0=0.53$. 
1419: The crosses refer to the case $\delta^{sat}=\delta^{sat}_0$,
1420: while the boxes to the case $\delta^{sat}=\delta_0^{sat}+\delta^{NL}$.
1421: The two lines correspond to $V_{a}=\sqrt{2\varepsilon A}$ and 
1422: $V_{s}=\sqrt{2\varepsilon B}$.}
1423: \label{fig:step}  
1424: \end{figure}
1425: %%%%%%%%%%%%%%%%%%%%%%%%
1426:  In the whole examined parameter 
1427: range $V_p$ is always reasonably approximated by $V_s$, but smaller.
1428: By increasing $B$ one observes an increase of the difference
1429: between the measured velocity and the linear prediction.
1430: These results confirm that the condition (\ref{eq:nonlin}) is indeed a
1431: necessary condition in order to observe nonlinear propagation
1432: of information. 
1433: 
1434: We will now investigate the role of $\delta^{NL}$ 
1435: in determining the propagation velocity. It is quite obvious that
1436: modeling the dynamics via the Eqs.~(\ref{eq:model}) and
1437: (\ref{eq:mod1}) the nonlinear effects will disappear in
1438: the limit $\delta_{NL}\to \delta^{sat}$ and this is indeed
1439: confirmed by the simulations (see Fig. \ref{fig:step}).
1440: In order to examine a less obvious situation,
1441: a modification of the expression (\ref{eq:mod1})
1442: is also considered. In this second case 
1443: $\delta^{sat}=\delta^{sat}_0+\delta^{NL}$
1444: is assumed, therefore by varying $\delta^{NL}$
1445: the extension of the amplitude interval over which the nonlinear
1446: mechanism is active will not be modified (being fixed to $\delta^{sat}_0$).
1447: But also in this second case $V_p \to V_a$ for increasing
1448: $\delta^{NL}$, this indicates that the smaller are the perturbation
1449: amplitudes affected by the nonlinear mechanisms
1450: the stronger will be the nonlinear effect on the velocity :
1451: $V_p \to V_s$ for $\delta^{NL} \to 0$. 
1452: 
1453: Typically, in generic CML's by varying a control parameter 
1454: both the difference $|\max_{\delta} \lambda(\delta) - \lambda|$  
1455: as well as $\delta^{NL}$ will change. Moreover,
1456: even the definition of $\delta^{NL}$ for a continuous 
1457: $\lambda(\delta)$ is not obvious. In order to understand
1458: the validity of the mean-field approximation
1459: (\ref{eq:mod1}) in a more realistic case we consider
1460: the shift map. In particular, for $\lambda(u)$ we
1461: employed the analytical expression (\ref{eq:lambda_shift})
1462: valid for the single uncoupled map.
1463: %fig 13
1464: 
1465: %%%%%%%%%%%%%%%%%%%%%%%%
1466: \begin{figure}
1467: \epsfxsize=8.truecm
1468: \epsfysize=6truecm
1469: \epsfbox{Fig13.eps}
1470: %\narrowtext
1471: \caption{Information propagation velocities for the shift map with 
1472: $\varepsilon=1/3$: (boxes) linear velocities $V_L$ and (crosses) directly
1473: measured nonlinear ones. The two lines correspond to
1474: $V_{a}$ (\ref{eq:velocity}) (dotted line) 
1475: and to the propagation velocity for the model 
1476: (\ref{eq:model}) with $\lambda(u)$ 
1477: given by (\ref{eq:lambda_shift}) (solid line). The dashed curve with asterisks is 
1478:  $V_{s}=\sqrt{2 \varepsilon \max_{\delta}\{\lambda(\delta)\}}$. }
1479: \label{fig:vlvnl_shift}  
1480: \end{figure}
1481: %%%%%%%%%%%%%%%%%%%%%%%%
1482: 
1483: The FSLE for coupled shift maps is actually different from 
1484: (\ref{eq:lambda_shift}), but we are neglecting correlations 
1485: among different sites and effects due to the specific measure 
1486: associated to the system. 
1487: Nevertheless, a numerical integration of Eq.~(\ref{eq:model})
1488: equipped with (\ref{eq:lambda_shift}) reproduces 
1489: semi-quantitatively the 
1490: features observed for a lattice of coupled shift maps 
1491: (see Fig.~\ref{fig:vlvnl_shift}). In particular, also the
1492: simple mean-field model (\ref{eq:model}) with the choice
1493: (\ref{eq:lambda_shift}) is able to forecast the
1494: observed transition
1495: from nonlinear to linear behavior for $\beta \to 2$.
1496: 
1497: Concluding this section we can safely affirm that (\ref{eq:model}) is a 
1498: reasonable 
1499: model to mimic the perturbation evolution at a mean field level, neglecting
1500: the spatio-temporal fluctuations and correlations. In the same fashion
1501: $\lambda(u)$ can be considered as an ``effective'' non linear kinetics 
1502: for the perturbation evolution.
1503: 
1504: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1505: \section{Final remarks}
1506: \label{sec:6}
1507: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1508: 
1509: In this paper information (error) propagation in extended chaotic
1510: systems has been studied in details. In particular, we have analyzed
1511: the relevance of linear and nonlinear mechanisms 
1512: for the propagation phenomena in spatio-temporal chaotic
1513: coupled map lattices. Linear stability analysis is
1514: not always able to fully characterize disturbance propagation.
1515: This is particularly true for (marginally) stable systems
1516: with strong nonlinearities,
1517: where finite size perturbations are responsible for information
1518: spreading in the system. When the nonlinear effects prevail
1519: on the linear ones the propagation velocity of information
1520: $V_p$ can be higher than the linear velocity $V_L$. A
1521: necessary condition for the occurrence of information
1522: spreading induced by nonlinear mechanisms has been
1523: expressed in terms of the Finite Size Lyapunov Exponent (FSLE).
1524: We have also shown the existence of strong analogies between
1525: error propagation and front propagation in reaction-diffusion
1526: models. In particular, the above mentioned necessary condition
1527: is analogous to the Aronson \& Weinberger theorem \cite{ara} for 
1528: front connecting stable and unstable steady states.
1529: In the linear and nonlinear case, the propagation velocity 
1530: $V_p$ can be identified via an unique marginal stability criterion
1531: involving Finite Size Lyapunov Exponents defined in a
1532: moving reference frame. This result generalizes the 
1533: corresponding linear criterion expressed in terms of 
1534: the maximal comoving Lyapunov exponents
1535: for the identification of $V_L$ \cite{DK87}. 
1536: 
1537: These results can be of some interest for
1538: the synchronization and the control of extended 
1539: systems. It has been recently shown that
1540: the synchronization of coupled extended 
1541: systems is strongly influenced
1542: by nonlinear effects. In particular, 
1543: the synchronization time exponentially
1544: diverges with the system size, even in non-chaotic
1545: situations, provided that $V_p > 0$ \cite{noi}.
1546: We believe that in these systems the appropriate
1547: indicator to characterize such transition
1548: would be the ``transverse'' Lyapunov 
1549: exponent~\cite{pecora}, once extended to finite scales.
1550: As far as control schemes are concerned, since
1551: they rely mainly on linear analysis~\cite{ogy},
1552: new nonlinear methods (e.g. based on the concept of FSLEs)
1553: should be introduced in order to control the erratic 
1554: behaviors due to fully nonlinear mechanisms.
1555: 
1556: A further aspect that should be addressed in future work
1557: concerns the extension of the applicability of the FSLE
1558: also to linearly stable systems: a candidate in this
1559: respect could be the indicator recently introduced
1560: in \cite{Kantz}. 
1561: 
1562: Finally, it is reasonable to expect that the
1563: present analysis is not limited to discrete
1564: models but that it can be applied to continuous 
1565: extended systems described in terms of PDE's,
1566: e.g. to the complex Ginzburg-Landau equation.
1567: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
1568: \acknowledgments
1569: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1570: Stimulating interactions with F.~Cecconi, R.~Livi, K.~Kaneko,
1571: A.~Pikovsky, A.~Politi, and A.~Vulpiani are gratefully acknowledged.
1572: Part of this work has been developed at the Institute of Scientific 
1573: Interchange in Torino, during the workshop on ``Complexity and
1574: Chaos'' in October 1999.
1575: We acknowledge CINECA in Bologna and INFM for providing us access to the 
1576: parallel CRAY T3E computer under the grant ``Iniziativa Calcolo
1577: Parallelo''.
1578: 
1579: 
1580: \begin{appendix}
1581: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
1582: \section{}
1583: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
1584: 
1585: In any numerical computation of the Lyapunov exponents one is forced to
1586: use a finite time approximation for an infinite time limit. 
1587: Nevertheless, provided that the convergence to the asymptotic value 
1588: is fast enough, this is not a dramatic problem.
1589: As a matter of fact, in low dimensional systems
1590: very fast convergence to the asymptotic value is
1591: usually observed.
1592: However, this problem manifest more
1593: dramatically in high dimensional systems, where the time to align along 
1594: the direction of maximal expansion could be very long \cite{PP98,P93,Orszag}.
1595: 
1596: In the present case this problem is
1597: complicated by the fact that the FSLE is intrinsically a finite time 
1598: indicator. Indeed the time a perturbation takes to grow from a
1599: value $\delta$ to $r \delta$ is finite unless $\delta \to 0$.
1600: However, for the CML models here analyzed and for initially
1601: localized perturbation (\ref{eq:tzero}) it is possible to evaluate 
1602: the corrections to apply to the FSCLE, estimated at
1603: finite time, in order to recover, for sufficiently small $\delta$,  
1604: the expected limit $\Lambda(v)$. 
1605: 
1606: These corrections allow for a faster convergence
1607: of the FSCLE to its asymptotic values.
1608: 
1609: For the sake of simplicity we consider maps with constant slope, 
1610: e.g. the shift map $F(x)=\beta\, x\;\;{\mbox {mod}}\,1$,
1611: and the CML defined in Eq.~(\ref{eq:cml}).
1612: In this case (for $\varepsilon < 1/2$) the maximal Lyapunov exponent
1613: of the CML coincides with that of the single map \cite{isola};
1614: i.e.  $\lambda=\ln \beta$. We will limit to the case
1615: $v=0$ (i.e to the FSLE), since the extension to generic $v$ is 
1616: straightforward.
1617: As shown in ~\cite{P93,lyapcorr} the finite time evolution
1618: of an infinitesimal perturbation $d_0$ initially
1619: localized in $i=0$ can be expressed at time $T$ as the sum of the 
1620: contributions $M(m,T)$ associated to all the paths connecting
1621: the space-time point $(i=0,t=0)$  to the point $(i=0,t=T)$,
1622: i.e.
1623: \begin{equation}
1624: \frac{\delta x_{i=0}(T)}{d_0} = \beta^T \sum_{m=0}^T M(m,T)  \quad .
1625: \label{path}
1626: \end{equation}
1627: where $m$ is the number of ``diagonal links'' connecting
1628: $(i,t)$ to $(i \pm 1,t+1)$ present in the path of length $T$.
1629: Each path $[m,T]$ of length $T$ with $m$ diagonal links contributes
1630: to the above sum with a term
1631: \begin{equation}
1632: M(m,T) = N(m,T)
1633: \left(\frac{\varepsilon}{2}\right)^m (1-\varepsilon)^{T-m}
1634: \end{equation}
1635: where $N(m,T)$ is the multiplicity associated to each path 
1636: $[m,T]$ (for more details see \cite{lyapcorr}).
1637: As can see from Eq.~(\ref{path}) the {\it finite time Lyapunov
1638: exponent} $\lambda_T$ will be given by
1639: \begin{equation}
1640: \lambda_T=
1641: \frac{1}{T}\ln\left|\frac{\delta x_{0}(T)}{d_0}\right| = \ln \beta + 
1642: \frac{1}{T}\ln\left|\sum_{m=0}^T M(m,T)\right|
1643: \quad .
1644: \label{lyt}
1645: \end{equation}
1646: the first contribution is the asymptotic one, while 
1647: the second one will vanish in the limit $T \to \infty$
1648: and it is the finite-time correction to evaluate.
1649: This second contribution can be numerically estimated
1650: by considering the finite time evolution of the Lyapunov
1651: eigenvector $\{W_i(t)\}$ associated to the maximal Lyapunov exponent,
1652: once it is initialized as $W_i(0) = d_0 \, \delta_{i,0}$.
1653: The evolution of $\{W_i(t)\}$ in the tangent space 
1654: is ruled by the following equation 
1655: \begin{equation}
1656: W_i(t+1)=\beta \left[(1-\varepsilon) W_i(t)+
1657: {\varepsilon \over 2} (W_{i-1}(t)+W_{i+1}(t))\right]\,.
1658: \label{eq:ac.tangent}
1659: \end{equation}
1660: In order to evaluate the finite time corrections,
1661: one should iterate at the same time 
1662: Eq.~(\ref{eq:ac.tangent}) (with $\beta$
1663: fixed to one) and the two replicas 
1664: required for computing $\lambda(\delta)$ (see Sect.~\ref{sec:2}).
1665: Then the estimation of the FSLE should modified in the following way:
1666: \begin{eqnarray} 
1667: \lambda(\delta)={1 \over \langle \tau(\delta_n,r)\rangle_e }
1668:  \left\langle \ln \left( { |\delta x_i(t+\tau(\delta_n,r))| \over |\delta
1669:  x_i(t)| }  \right) \right. \nonumber \\
1670: \left.  - 
1671: \ln \left( { |W_i(t+\tau(\delta_n,r))| \over |W_i(t)|}
1672: \right) \right\rangle_e \,.  
1673: \label{eq:FSLE_discrete_new}
1674: \end{eqnarray}
1675: The case of maps with non constant slope is computationally much
1676: heavier. Since for each different path the local multiplier
1677: $F^{'}(\tilde{x_i})$ will be different and they will depend
1678: on the particular trajectory under consideration \cite{P93,lyapcorr}.
1679: As a matter of fact we have observed that if the multipliers are
1680: equally distributed among positive and negative values, the finite
1681: time corrections essentially cancel out.
1682: \end{appendix}
1683: 
1684: 
1685: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1686: \begin{thebibliography}{99}
1687: \bibitem{Shaw} R.~Shaw, {\it Z. Naturf.} {\bf 36a}, 80 (1981).
1688: 
1689: \bibitem{K86} K.~Kaneko,  Physica D {\bf 23},  436 (1986).
1690: 
1691: \bibitem {BR87} T.~Bohr and D.~Rand,  Physica D {\bf 52}, 532 (1987).
1692: 
1693: \bibitem {grass89} P.~Grassberger,  Physica Scripta {\bf 40}, 346 (1989).
1694: 
1695: \bibitem{PV94} 
1696: G.~Paladin and A.~Vulpiani, J. Phys. {\bf A27}, 4911 (1994).
1697: 
1698: \bibitem{DK87} R.J.~Deissler and K.~Kaneko, Phys. Lett. {\bf A119}, 397 (1987).
1699: 
1700: \bibitem {lepri} S.~Lepri, A.~Politi and A.~Torcini, 
1701: J. Stat. Phys., {\bf 82} (1996) 1429; 
1702: J. Stat. Phys., {\bf 88}  (1997) 31; 
1703: Chaos, {\bf 7} (1997) 701.
1704: 
1705: \bibitem {CK88} J.P.~Crutchfield  and  K.~Kaneko, 
1706: \prl {\bf 60}, 2715 (1988).
1707: 
1708: \bibitem{PLOK93} A.~Politi, R.~Livi, G.L.~Oppo and R.~Kapral,
1709: Europhys. Lett. {\bf 22}, 571 (1993).
1710: 
1711: \bibitem{TP94} A.~Politi and A.~Torcini, Europhys. Lett. {\bf 28}, 545 (1994).
1712: 
1713: \bibitem{TGP95} A.~Torcini, P.~Grassberger and A.~Politi,  
1714: J. Phys. {\bf A27}, 4533 (1995).
1715: 
1716: \bibitem{CFVV99} M.~Cencini, M.~Falcioni, D.~Vergni and A.~Vulpiani,
1717:  Physica D {\bf 130}, 58 (1999).
1718: 
1719: \bibitem{dressler92} U.~Dressler and J.D.~Farmer,  
1720: Physica D {\bf 59}, 365 (1992); T.J.~Taylor, Nonlinearity {\bf 6}, 369 (1993).
1721: 
1722: \bibitem{ABCPV96} E.~Aurell, G.~Boffetta, A.~Crisanti, G.~Paladin and
1723: A.~Vulpiani, \prl {\bf 77}, 1262 (1996);
1724: J. Phys. {\bf A30}, 1 (1997).
1725: 
1726: \bibitem{cml} I.~Waller and R.~Kapral, \pra {\bf 30} , 2047
1727: (1984); K. Kaneko, Prog. Theor. Phys. {\bf 72}, 980 (1984).
1728: 
1729: \bibitem {KPP} R.A.~Fisher, Ann. Eugenics {\bf 7}, 355
1730: (1937); A.N.~Kolmogorov, I.~Petrovsky and N.~Piscounov,
1731: Bull. Univ. Moscow, Ser. Int.  {\bf A1}, 1 (1937).
1732: 
1733: \bibitem{wim2} U.~Ebert and W.~van~Saarloos, Physica D {\bf 146}, 1 (2000).
1734: 
1735: \bibitem{cencrep} A recent review on the subject is
1736: G.~Boffetta, M.~Cencini, M.~Falcioni and A.~Vulpiani,
1737: unpublished (nlin.CD/0101029).
1738: 
1739: \bibitem {Kantz} T.~Letz and H.~Kantz, \pre {\bf 61},  2533 (2000).
1740: 
1741: 
1742: \bibitem{nota}Note that also for $\varepsilon \to 1$ deviations are observed. 
1743: This is due to the fact that for $\varepsilon=1$ the system reduces to
1744: two uncoupled subchains, therefore also for $\varepsilon \sim 1$
1745: the coupling among nearest neighbors is actually very small.
1746: 
1747: \bibitem {WB93} W.~van~de~Water and T.~Bohr, Chaos {\bf 3}, 747 (1993).
1748: 
1749: \bibitem{TP92} A.~Politi and A.~Torcini,  Chaos {\bf 2}, 293 (1992).
1750: 
1751: \bibitem{wim1} W.~van~Saarloos, \pra {\bf 37}, 211 (1988);
1752: {\it ibidem}, {\bf 39}, 6367 (1989).
1753: 
1754: \bibitem {isola} S.~Isola, A.~Politi, S.~Ruffo, and A.~Torcini,
1755: Phys. Lett. A, {\bf 143}, 365 (1990).
1756: 
1757: \bibitem{benettin} I.~Shimada and T.~Nagashima, Prog. Theor. Phys.
1758: {\bf 61}, 1605 (1979); G.~Benettin, L.~Galgani, A.~Giorgilli and
1759: J.M.~Strelcyn, Meccanica, March 15 and 21 (1980).
1760: 
1761: \bibitem{ginelli} F. Ginelli, R. Livi, and A. Politi,
1762: private communication.
1763: 
1764: \bibitem{ara} D.G.~Aronson and H.F.~Weinberger,
1765: Adv. Phys. {\bf 30}, 33 (1978).
1766: 
1767: \bibitem{armero} J. Armero, J. Casademunt, L. Ram{\'i}rez-Piscina,
1768: and J.M. Sancho, \pre {\bf 58}, 5494 (1998);
1769: A. Rocco, U. Ebert, and W. van Saarloos, \pre {\bf 62}, R13 (2000).
1770: 
1771: \bibitem{PP98} 
1772: A.S.~Pikovsky and J.~Kurths, \pre {\bf 49},  898 (1994).
1773: 
1774: \bibitem{noi} L.~Baroni, R.~Livi, and A.~Torcini, accepted for
1775: publication on \pre, it will appear in the volume of March 2001
1776: (nlin.CD/0003010). 
1777: 
1778: \bibitem{pecora} L.M.~Pecora and T.L.~Carroll,
1779: \prl {\bf 64} 821 (1990).   
1780: 
1781: \bibitem{ogy} E.~Ott, C.~Grebogi, and J.A.~Yorke,
1782: \prl {\bf 64}, 1196 (1990).
1783: 
1784: \bibitem{P93} A.~Pikovsky, Chaos {\bf 3}, 225 (1993).
1785: 
1786: \bibitem{Orszag} I.~Goldhirsh, P.L.~Sulem and S.A.~Orszag,  Physica D {\bf 27},
1787: 311 (1987).
1788: 
1789: \bibitem{lyapcorr} R.~Livi, S.~Ruffo, and A.~Politi,
1790: J. Phys. A {\bf 25}, 4813 (1992);
1791: A.~Torcini, R.~Livi, S.~Ruffo, and A.~Politi,
1792: \prl {\bf 78}, 1391 (1997).
1793: \end{thebibliography}
1794: \end{multicols}
1795: 
1796: \end{document}
1797: 
1798: 
1799: 
1800: