1: \documentclass[12pt]{iopart}
2:
3: \usepackage{graphicx}
4: \usepackage{bm}% bold math
5:
6: \newcommand{\bd}{\begin{document}}
7: \newcommand{\ed}{\end{document}}
8: \newcommand{\beq}{\begin{equation}}
9: \newcommand{\eeq}{\end{equation}}
10: \newcommand{\nid}{\noindent}
11: \newcommand{\su}{\section}
12: \newcommand{\ssu}{\subsection}
13: \newcommand{\sssu}{\subsubsection}
14: \newcommand{\baR}{\begin{array}}
15: \newcommand{\eaR}{\end{array}}
16: \newcommand{\ben}{\begin{enumerate}}
17: \newcommand{\een}{\end{enumerate}}
18:
19: \newcommand\bbbc{{\sf I\!\!C}}
20:
21: \newcommand\cB{{\cal B}}
22: \newcommand\cH{{\cal H}}
23: \newcommand\cM{{\cal M}}
24: \newcommand\cL{{\cal L}}
25:
26: \bd
27: \title[]{Does the complex susceptibility of the H\'enon map have a pole
28: in the upper-half plane ? A numerical investigation.}
29: %\titlerunning{}
30:
31: \author{B. Cessac}
32: \address{INRIA, 2004 Route des Lucioles, 06902 Sophia-Antipolis, France.\\
33: Institut Non Linéaire de Nice, 1361 route des Lucioles, 06560 Valbonne,
34: France.\\
35: Universit\'e de Nice, Parc Valrose, 06000 Nice, France.}
36:
37: \begin{abstract}
38: It has been rigorously shown in \cite{CMP} that the complex susceptibility of
39: chaotic maps of the interval can have a pole in the upper-half complex plane.
40: We develop a numerical procedure allowing to exhibit this pole from
41: time series.
42: We then apply the same analysis to the H\'enon map and
43: conjecture that the complex susceptibility has also a pole
44: in the upper-half complex plane.
45: \end{abstract}
46:
47: \pacs{02.70.-c,05.10.-a,05.45.-a}
48: %\noindent{ \it Keywords\/ Linear response, chaotic systems.\\}
49: \submitto{\NL}
50: \maketitle
51:
52: Consider a dynamical system $f : \cM \to \cM$, on a manifold $·\cM$
53: endowed with the Lebesgue measure. Assume that there is a unique $f$-ergodic
54: probability measure $\rho$ so that $\lim_{T \to \infty}\frac{1}{T} \sum_{t=1}^{T} \phi(f^n(x)) = \int \phi d\rho$
55: for continuous functions $\phi$ and for Lebesgue-almost every $x \in \cM$. Then,
56: $\rho$ is called the SRB measure or physical measure (see \cite{Young} for a discussion
57: on these two terminologies). Assume that when $f$ is perturbed in its neighborhood, the
58: average value $\rho(A)$ of a given smooth observable $A$ varies differentiably with $f$.
59: Then the derivative operator is called the \textit{linear response function}.
60: In the case of uniformly hyperbolic systems, and provided that $f$ and its perturbations
61: are sufficiently smooth, then the linear response is given by \cite{Ru99}:
62:
63: \beq\label{drt}
64: \delta \rho(A) = \sum_{n=0}^\infty \rho(X.\nabla(A\circ f^n)),
65: \eeq
66:
67:
68: \nid where $X$ is a (time-independent) perturbation. Ruelle has indeed shown that this series is
69: convergent for suitable perturbations. In the case of a time-dependent
70: perturbation $X_t$ this formula extends formally to:
71:
72: \beq\label{drtt}
73: \delta \rho_t(A) = \sum_{n=0}^\infty (\kappa_{n} X_{t-n})A.
74: \eeq
75:
76: \nid The operator $\kappa_n : \cH \to \cB^\ast$, where $\cH$ is a suitable space of vector fields and $\cB^\ast$
77: the dual of a suitable space of functions, is defined by:
78:
79: \beq \label{kappa}
80: (\kappa_n X_{t-n})A = \rho(X_{t-n}.\nabla(A\circ f^n)); \qquad n \geq 0,
81: \eeq
82:
83: \nid while, for $n<0$, ($\kappa_n X_{t-n})A=0$ (causality).
84: It is called the ``response function''. In this case, one recovers the
85: standard definition of the linear response in non equilibrium statistical
86: mechanics ((\ref{drtt}) is a convolution product), and classical results such as fluctuation-dissipation relations
87: or Onsager theory can be recovered \cite{Ru99}. Moreover, this result does not require the assumption
88: of closeness to equilibrium.
89:
90:
91: However, one can find examples of non uniformly hyperbolic systems where
92: the series (\ref{drt}) does not converge, even for maps of the interval with absolutely
93: continuous invariant measures \cite{CMP}. In this case, one considers
94: the \textit{complex susceptibility} function, $\chi(\omega)=(\hat{\kappa}_\omega X)A$
95: where $\hat{\kappa}_\omega$ is the Fourier transform
96: of the response function. It is also convenient to use the variable $\lambda=e^{i \omega}$.
97: Then, $\chi(\omega)=\Psi(\lambda)$.
98: Under some conditions on $X$ and $f$
99: (interval maps), Ruelle proved that $\Psi(\lambda)$ extends to a meromorphic
100: extension with no pole at $\lambda=1$. He also showed that
101: this function has a pole inside the unit disk $|\lambda| <1 $ (i.e. in the upper-half complex plane for the frequencies
102: $\omega$) and other poles, corresponding to
103: the eigenvalues of the Perron-Frobenius operator (Ruelle-Pollicott resonances \cite{Pollicott,RuelleP}), outside the unit disk
104: (in the lower-half complex plane).
105: The existence of a
106: pole of the complex susceptibility in the upper-half complex plane
107: must not be misinterpreted.
108: This is not a ``violation of causality'', which has no meaning
109: in a dynamical system, causal by construction. This simply expresses
110: that an arbitrary small perturbation of $f$ does not give
111: a response proportional to the perturbation. According
112: to Ruelle: ``this might be expressed by saying that $\rho$ is not
113: linearly stable''.
114:
115: This result is quite interesting and intriguing
116: for a physicist and two natural questions arise. Does this property exist in larger dimensional
117: chaotic dynamical systems ? And is it possible to have an experimental/numerical
118: characterization of the effects of this pole ? The first question is difficult to address from
119: a mathematical point of view and is still unresolved even for classical
120: examples such as the H\'enon map. The second one is the main concern of the present paper.
121: We indeed propose a numerical experimentation protocol allowing to
122: numerically approximate the complex susceptibility (\ref{susc}) and to investigate the numerically
123: observable effects of having a pole in the upper-half complex plane.
124:
125: The paper is organized as follows. After a brief summary of Ruelle's results
126: for the logistic map (section \ref{RuelleTh}), we describe our numerical methods
127: in section \ref{NumMet} and apply it to the logistic map where one can use the guidelines of
128: Ruelle's mathematical results as
129: a validation (section \ref{Log}). Then, we apply it to the H\'enon map for the standard parameter
130: values and conjecture that this map could also exhibit a pole in the
131: upper-half complex plane (section \ref{Henon}).
132:
133:
134: \su{Reminder of Ruelle's results.}\label{RuelleTh}
135:
136: The dynamical system defined by the logistic map:
137:
138: \beq\label{f}
139: f(x)=1-2x^2, \quad x \in [-1,1],
140: \eeq
141:
142: \nid has an absolutely continuous ergodic measure with density :
143:
144: \beq\label{rhoinv}
145: \rho(x)=\frac{1}{\pi\sqrt{1-x^2}}.
146: \eeq
147:
148: The linear response (\ref{drt}) reads, in this case:
149:
150: \beq\label{drt1D}
151: \delta \rho(A) = \sum_{n=0}^\infty \int X(x) \frac{d}{dx}A(f^nx) \rho(dx),
152: \eeq
153:
154: \nid while the complex susceptibility reads:
155:
156: \beq\label{susc}
157: \chi(\omega)=\Psi(\lambda)=
158: \sum_{n=0}^\infty \lambda^n \int A'(f^n x) (f^n)'(x)X(x)\rho(dx)
159: \eeq
160:
161: \nid with $\lambda=e^{i\omega}$.
162: Ruelle has shown the following results \cite{CMP}.
163: Consider the variable
164: change $x=\sin(\frac{\pi}{2} y)=\varpi(y)$
165: which maps $f$ on $g(y)=1-2|y|$ [tent map]. Then the density $\rho$
166: is mapped onto the density $\sigma_0(y)=\frac{1}{2}$.
167: In this variable, we have :
168:
169: \beq\label{PsiL0}
170: \Psi(\lambda)=\sum_{n=0}^\infty \lambda^n \int_{-1}^1 (\cL^n_0 Y)(s)B'(s)ds
171: \eeq
172:
173: \nid
174: where $B=A \circ \varpi$, $Y(y)=\sigma_0(y) \frac{X(\varpi(y))}{\varpi'(y)}$
175: and $\cL_0\Phi(y)=\Phi\left(\frac{y-1}{2}\right)-\Phi\left(\frac{1-y}{2}\right)$.
176:
177: Consider first a square integrable holomorphic perturbation $X \equiv X_0$ that \textit{vanishes} at $\left\{-1,1\right\}$.
178: Then, the series
179: (\ref{PsiL0}) has a meromorphic extension in $\bbbc$ and has no pole inside the unit disc.
180: Moreover, it is holomorphic at $\lambda=1$ (resp. $\omega=0$ ).
181: Indeed, one has, using integration by parts:
182:
183: \beq \label{PsioL}
184: \Psi(\lambda)=
185: -\sum_{n=0}^\infty \lambda^n \int_{-1}^1 (\cL^n Y'_0)(s)B(s)ds,
186: \eeq
187:
188: \nid where $\cL$ is the Perron-Frobenius operator:
189:
190: \beq\label{PF}
191: \cL\Phi(y)=\frac{1}{2}\Phi\left(\frac{y-1}{2}\right)+\frac{1}{2}\Phi\left(\frac{1-y}{2}\right)
192: \eeq
193:
194: In the space of analytic functions, and in the basis of monomials $\left\{y^m\right\}$, $m=0,1,2, \dots$ this operator
195: is represented by an infinite triangular matrix. The eigenvalues are on the diagonal.
196: They are of the form $1/4^n, n=0,1 \dots $.
197: The corresponding right eigenfunctions are (Bernoulli) polynomials
198: such that $\sigma_0(y)=\frac{1}{2}$ (invariant density corresponding to the eigenvalue $1$);
199:
200:
201: \beq \label{sigma1}
202: \sigma_1(y)=-\frac{1}{3}-2y +y^2
203: \eeq
204:
205: \nid (eigenvalue $\frac{1}{4}$), etc... This simple argument can be found
206: in \cite{Gaspard} pp 123-125. This result is also a direct
207: consequence of \cite{BaladiRugh}.
208:
209: Therefore, for a perturbation $Y'_0(y)=\sum_{k=1}^\infty C_k \sigma_k(y)$,
210: and for an observable $A$ such that the integral $\int_{-1}^1 Y'_0(s)B(s)ds$
211: is finite, one finds easily that the susceptibility
212: has poles $\lambda_k^{-1}, k \geq 1$, outside the unit disk.
213: For example, if $Y'_0(y)=\sigma_1(y)$, the series (\ref{PsioL})
214: reads:
215:
216: \beq\label{psi0}
217: \Psi(\lambda) =- \sum_{n=1}^\infty \left(\frac{\lambda}{4}\right)^n \int_{-1}^1 \sigma_1(s)B(s)ds
218: \eeq
219: %
220: %
221:
222: This case
223: is used as a benchmark for the numerical procedure
224: developed in the next section.
225: The corresponding perturbation $X_0$ reads, in the variable $x$:
226:
227: \beq\label{X0num}
228: X_0(x)=\pi\sqrt{1-x^2}\left[1-\frac{2}{3\pi}\arcsin(x) - \frac{4}{\pi^2}\arcsin^2(x)
229: + \frac{8}{3\pi^3}\arcsin^3(x)\right]
230: \eeq
231:
232: \nid where $X_0(\pm 1)=0$.
233: \\
234:
235:
236: Ruelle has also shown that there is a
237: meromorphic function :
238:
239:
240: \beq\label{Phimoins}
241: \Phi_-(z)=\frac{1}{z+1}+w(z),
242: \eeq
243:
244: \nid where $w(z)$ is of order $O(z+1)$,
245: and where $\Phi_-(\pm 1)=0$ such that $\cL_0\Phi_-=\sqrt{f'(-1)}\Phi_-$ ($=2 \Phi_-$ in our case).
246: The function $w$ in (\ref{Phimoins}) is therefore a solution of:
247:
248: \begin{equation}\label{Wz}
249: w\left(\frac{z-1}{2}\right)-w\left(\frac{1-z}{2}\right)-2w(z)=\frac{2}{3-z}.
250: \end{equation}
251:
252: \nid $w$ is not an analytic function. It can be obtained via the recursion:
253:
254: \beq\label{wrec}
255: w(z) = \left\{\begin{array}{ccc}
256: &&w(2z+1)+\frac{1}{2-2z} ; \qquad z <0;\\
257: &&-w(1-2z)-\frac{1}{2+2z} ; \qquad z >0;
258: \end{array}
259: \right.
260: \eeq
261:
262: \nid reminiscent of the Tagaki function discussed by Gaspard in the context
263: of the multi-baker map \cite{Gaspard}. Returning to the variable
264: $x$, the function $\Phi_-$ corresponds to :
265:
266:
267: \beq\label{Xm}
268: X_-(x)=\pi\sqrt{1-x^2}\left[\frac{1}{\frac{2}{\pi}\arcsin(x)+1} + w(\frac{2}{\pi}\arcsin(x)+1)\right].
269: \eeq
270:
271: The corresponding series $\Psi_-(\lambda)=
272: \sum_{n=0}^\infty \lambda^n \int \rho(x) A'(f^{n} x) (f^{n})'(x)X_-(x)dx$,
273: converges if $|\lambda|\sqrt{|f'(-1)|} < 1$,
274: and, in its domain of convergence, has the form $\frac{G_A}{1-\lambda|\sqrt{|f'(-1)|} }$
275: where $G_A$ is a constant depending on $A$. The complex susceptibility
276: has therefore a pole at $\lambda=\frac{1}{\sqrt{f'(-1)}}=\frac{1}{2}$, inside the unit disk
277: (resp. $\omega=i\log(2)$).
278: Therefore, the series $\Psi_-$ diverges for real frequencies $\omega$.
279:
280:
281: \su{The numerical method.}\label{NumMet}
282:
283: The numerical method used here has been introduced by Reick in \cite{Reick} and independently in
284: \cite{JAB1,JAB2,JAB3}. It is purely heuristic and there is no mathematical proof
285: of its validity, though a discussion with some analytic developments
286: can be found in \cite{Reick} and \cite{JAB3}. Thus,
287: in this section, we shall use formal expressions.
288:
289: Consider a time-dependent perturbation of $f$ of type\footnote{Note that this way of perturbing avoids
290: the introduction of $f^{-1}$ in the computation
291: of the response function as found e.g. in \cite{Ru99}. It allows one to write convolution formulae
292: and the Fourier transform in the standard form they have for continuous time systems.
293: This was the main motivation for this choice in \cite{JAB3} and
294: we pursue along these lines in the present paper.} :
295:
296: \beq\label{SDP}
297: x_{t+1}=f\left[x_t+\epsilon X_t(x_t)\right].
298: \eeq
299:
300: The variation of the average value of $A$, at time $t$, is formally
301: given by:
302:
303: \beq\label{suscgen}
304: \Delta\rho_t(A)=\epsilon \sum_{n=-\infty}^t \int A'(f^{t-n}x)(f^{t-n})'(x)X_n(x) \rho(dx)
305: + O(\epsilon^2).
306: \eeq
307:
308: \nid which corresponds to (\ref{drtt}) to the linear order and up
309: to a factor $\epsilon$. Note that the term $O(\epsilon^2)$ is not uniform in $t$.\\
310:
311: Consider now perturbations of type $X_t(x) \equiv X(x)e^{-i\omega t}$ where $\omega$
312: is real. Then:
313:
314:
315: \beq\label{RepLin}
316: \Delta\rho_t(A)=\epsilon e^{-i\omega t} \chi(\omega)+ O(\epsilon^2),
317: \eeq
318:
319: \nid where $\chi(\omega)$, the complex susceptibility for $X(x)$, is given by eq. (\ref{susc}).
320: Then, at the linear order in $\epsilon$:
321:
322: \beq
323: \chi(\omega)=\frac{\Delta\rho_t(A)e^{i\omega t}}{\epsilon}.
324: \eeq
325:
326:
327: Since $\chi(\omega)$ does not depend on time one may write:
328:
329: \beq\label{chiT}
330: \chi(\omega)=\frac{1}{\epsilon T} \sum_{t=1}^T \Delta\rho_t(A)e^{i\omega t}
331: \eeq
332:
333: The time-dependent SRB state $\rho_t$
334: involves an average over initial conditions in
335: the distant past \cite{Ru99}. One can argue (see \cite{Reick} Appendix A)
336: that the above
337: average over $t$ makes the average over initial
338: conditions unnecessary. The idea is therefore to replace $ \Delta\rho_t(A)$ by $A(x'(t))-A(x(t))$
339: where $x'(t)$ is a typical trajectory of the perturbed system,
340: with a perturbation $X(x)e^{-i\omega t}$
341: and $x(t)$ is a typical trajectory of the unperturbed system. Then, one
342: computes:
343:
344: \beq \label{chinum}
345: S_T(\omega,x) = \frac{1}{\epsilon T} \sum_{t=1}^T \left[ A(x'(t))-A(x(t)) \right]e^{i \omega t}
346: \eeq
347:
348: The hope is that $S_T(\omega,x) \to \chi(\omega)$ as $T \to \infty$, where the limit is independent of $x(0)$.
349: There is no mathematical proof of the above statement.
350: Note that $T$ must be taken so large that roundoff errors play their role of selecting the limit to be the SRB state.
351: On practical grounds, as discussed in \cite{Reick}, this procedure is numerically
352: reliable provided that
353: $\epsilon T$ is sufficiently large and $\omega T >> 1$ (basically
354: one uses $\epsilon T \omega >>1$). These conditions have been checked in the simulations performed in the present paper.
355: Since $S_T(\omega,x)$
356: is expected to give a fairly good approximation of $\chi(\omega)$, the Fourier
357: transform of the linear response, when $T$ is sufficiently large,
358: one also expects that the inverse Fourier series of $S_T(\omega,x)$, called $R(t)$ (where
359: we drop the $T$ and $x$ for simplicity)
360: gives a fairly good approximation of the linear response (\ref{drtt}).\\
361:
362:
363: \nid\textbf{Remarks}
364:
365: \ben
366:
367: \item The quantity obtained by
368: this procedure is, \textit{stricto sensu}, not the \textit{linear} response but
369: the \textit{total} response since one can apply it for arbitrary large $\epsilon$.
370: As a test, one must a posteriori check that the susceptibility is independent
371: of $\epsilon$ on a certain range of small $\epsilon$ values [e.g.
372: it does not vary if one replaces $\epsilon$ by $2\epsilon$].
373:
374: \item In the case of the logistic map,
375: any typical trajectory of (\ref{f}) approaches the
376: points $\pm 1$ within a distance of order $\epsilon$
377: with a characteristic time of order $\frac{1}{\sqrt{\epsilon}}$.
378: In a numerical simulation where $\epsilon$ is small
379: but finite this arises often, especially if one respects the condition
380: $\epsilon T\omega >> 1$. However,
381: in this case, a perturbation $X(x)e^{-i\omega t}$ can push
382: the trajectory out of the interval $[-1,1]$ leading
383: to an exponential divergence.
384: To avoid this effect
385: we have used a perturbation $X(x,t)=X(x)(1 + e^{- i \omega t})$
386: where $X$ is positive about $x=-1$
387: and negative about $x=+1$, so that the perturbation is always
388: directed inside the interval $[-1,1]$ whenever $x=\pm 1$.
389:
390: \een
391:
392: \su{Application to the logistic map.}\label{Log}
393:
394:
395: \ssu{A benchmark.}
396:
397: Let us first consider the case where the complex susceptibility is well defined
398: on the real axis (for the frequency $\omega$) by the series (\ref{susc}).
399: More precisely, we consider a perturbation, corresponding to the Ruelle-Pollicott resonance $\frac{1}{4}$.
400: In the variable $x$ this perturbation is given by eq. (\ref{X0num}).
401: For the observable $A$ one can use any choice such that the integral $\int_{-1}^1 \sigma_1(s)B(s)ds$ in ( \ref{psi0})
402: is finite. We choose $A=X_0$. The complex susceptibility reads $\Psi(\lambda)=-\frac{C}{4-\lambda}$
403: where $C>0$ is a constant.\\
404:
405: In the figure \ref{Fres1}a the complex susceptibility, computed
406: with this algorithm and compared to the theoretical value, is drawn.
407: To generate the real and imaginary parts of the susceptibility
408: we run two independent simulations. Starting from the same
409: mother trajectory, we apply in the first case a perturbation $\epsilon \cos(\omega t)$
410: and compute the corresponding susceptibility, which gives $Re(\chi(\omega))$.
411: In the second case we apply a perturbation $-\epsilon \sin(\omega t)$, which gives $Im(\chi(\omega))$.
412: Note that the perturbations are therefore real and the trajectories stay on the real axis.
413: The amplitude of the perturbation
414: was fixed to $\epsilon=10^{-2},5 \times 10^{-3},10^{-3}$. Averages are performed by computing
415: the time average $S_T(\omega,x)$ (\ref{chinum}) for
416: $N$ initial conditions $x_n, n=1 \dots N$ randomly chosen and then computing
417: $\frac{1}{N}\sum_{n=1}^N S_T(\omega,x_n)$. This procedure allows us to reduce the
418: numerical noise and to compute error bars.
419: We took $T=10^6,N=10$ for $\epsilon=10^{-2}$;
420: $T=2.10^6,N=20$ for $\epsilon=5 \times 10^{-3}$ and $T=10^7, N=10$
421: for $\epsilon=10^{-3}$.
422: Note that numerical noise is large and
423: that the resonance curve is
424: very flat, requiring to have small error bars. This requires
425: an average over a very long time $T$, roughly given by the condition $\epsilon T \omega >> 1$,
426: and limits the range of
427: $\epsilon$ values that one can reach. Note that the condition $\epsilon T \omega >> 1$ is always
428: violated as $\omega$ approaches zero, whatever the (finite) value of $T$.
429: This explains the discrepancy observed for small frequencies.
430: If one excludes this, the agreement is quite good.
431: Note also that the experimental curve has a magnitude
432: that does not depend of $\epsilon$,
433: in this range of $\epsilon$ values, as expected.
434:
435: The (approximate) linear response $R(t)$ can be easily obtained by computing the inverse Fourier series
436: of $\chi(\omega)$ (e.g. by a direct summation). It is drawn
437: in Figure \ref{Fres1}b (in log scale for the $y$ axis). Note that
438: up to numerical noise $R(t)$ is real, as expected. One observes an exponential decay very close
439: to the theoretically expected decay $4^{-t}$, corresponding to the
440: mixing rate given by the first Ruelle-Pollicott resonance
441: (and also to the characteristic decay rate toward equilibrium, in agreement with the fluctuation-dissipation theorem).
442: Note that the range of validity for the interpolation is very thin ($\sim [0,4]$). Indeed, rapidly
443: the perturbation becomes so weak that one measures only the numerical noise.\\
444:
445:
446: %
447: %%%%%%%%%%%%%%%%%%%% Susceptibilité res 1
448: \begin{figure}[ht]
449: \begin{center}
450: \includegraphics[height=8cm,width=7cm,clip=false]{Susceptibilitesig1.eps}
451: \hspace{1cm}
452: \includegraphics[height=8cm,width=7cm,clip=false]{InterpolReponseres1.eps}
453: \vspace{0.5cm}
454: \caption{(a) Susceptibility for a perturbation (\ref{X0num}) and
455: an observable also given by (\ref{X0num}), for $\epsilon=10^{-2},5 \times 10^{-3},10^{-3}$.
456: (b) Corresponding approximate linear response [denoted by R(t)], in log scale
457: and theoretically expected curve $y=a-log(4)t$.
458: \label{Fres1}}
459: \end{center}
460: \end{figure}
461: %%%%%%%%%%%%%%%%%%%%%%%%%
462: %
463:
464: \ssu{Perturbation $X_-$.}
465:
466: As we saw, the perturbation (\ref{Xm}) leads to a diverging series (\ref{susc})
467: for real frequencies. However, when computing the expression (\ref{chinum}) one is not dealing
468: with a series, but with a finite sum, which diverges
469: as $T$ grows. The divergence rate provides useful hints on the presence of the pole.
470:
471: Consider indeed the formal expression
472: (\ref{suscgen}) of $\Delta \rho_t(A)$ in the case of the logistic map.
473: $f^{t-n}$ has $t-n$ zeros and oscillates rapidly (for large $t-n$) with
474: a period $\sim 4^{n-t}$. One can decompose the integral
475: $\int_{-1}^{+1} \rho(x) A'(f^{t-n} x) (f^{t-n})'(x)X(x)dx$ over intervals
476: delimited by the zeros of
477: $f^{t-n}$. The density $\rho$ has strong variations about $\pm 1$ and
478: small variations in the central part. The contribution
479: of the first interval (containing $-1$) is given, for large $t-n$, by :
480:
481:
482: \beq\label{ExpPole}
483: \baR{ccc}
484: &\int_{-1}^{ -1+4^{n-t}} \rho(x) A'(f^{t-n} x) (f^{t-n})'(x)X(x)dx \\
485: \sim &\frac{4^{t-n}}{\pi} \int_{0}^{4^{n-t}} \frac{A'(f^{t-n}(-1+u))X(-1+u)}{\sqrt{u}}du
486: \eaR
487: \eeq
488:
489: \nid where $x=-1+u$. If $A'(f^{t-n}(-1+u))X(-1+u)$ \textit{does not vanish about $u=0$ ($x=-1$)} and has small variations
490: on this interval, this contribution is $\sim 4^{t-n} \int_{0}^{4^{n-t}} \frac{1}{\sqrt{u}}du =
491: 2^{t-n}$, which diverges\footnote{If one considers now the contribution of the ``bulk'' [inner intervals] to the integral (\ref{ExpPole})
492: one can figure out that it does not diverge, essentially because the derivative
493: $(f^{t-n})'(x)$ has alternating signs and because the density $\rho$
494: has small variations in the bulk. More precisely, the contribution of the bulk
495: is provided by a decomposition on the Ruelle-Pollicott eigenfunctions $\sigma_n$ and decays exponentially
496: due to mixing. Thus, the existence of a pole in the upper-half complex plane is
497: due to a boundary effect obtained when the perturbation weights the points $\left\{-1,1\right\}$
498: (where the density diverges). } when $t-n \to \infty$, with a rate exactly given by the pole
499: $\lambda=\frac{1}{2}$.
500: Therefore, another good benchmark of our method, in the case where the complex susceptibility
501: has a pole in the upper-half plane, consists in computing (\ref{chinum}),
502: and taking the inverse Fourier series, providing in this way an approximation for $\Delta \rho_t(A)$.
503: One should see then the exponential divergence with a rate $\log(2)$ (which
504: is exactly the value of the Lyapunov exponent $\lambda$).
505:
506: However, there is another effect that must be taken into account in the numerics.
507: When the amplitude of the response, $\sim \epsilon 2^t$, is small
508: one computes numerically the first term of the Taylor expansion (\ref{suscgen}).
509: But when $\epsilon 2^t$ becomes too large the numerical computation
510: includes also the nonlinear terms and one computes in fact
511: the complete susceptibility, including nonlinear terms. This one does not diverge
512: because the trajectories of the perturbed system remain in the interval $[-1,1]$,
513: but it does not
514: depend linearly on $\epsilon$.\\
515:
516: Therefore, one expects the following. There is a time cut-off
517: $t_m(\epsilon) \sim - \frac{\log(\epsilon)}{\log(2)}$
518: beyond which one does not compute the \textit{linear} response but
519: the complete response,
520: including nonlinear effects which saturate the growth of the perturbation.
521: Then one should observe an exponential growth $2^t$ up to $t_m(\epsilon)$ in the linear response
522: of the
523: form $2^t H_{[0,t_m(\epsilon)]}(t)$,
524: where $H_I()$ is the characteristic function of the interval~$I$. The corresponding susceptibility
525: is:
526:
527: $$\sum_{n=0}^{t_m} (2\lambda)^n=
528: \frac{1-(2\lambda)^{t_m}}{1-2\lambda}$$
529:
530: \nid which reads, using the frequency $\omega$:
531:
532: \beq\label{Explth}
533: \baR{ccc}
534: &\frac{(1-2\cos(\omega))(1-2^{t_m}\cos(\omega t_m))+2^{t_m+1}\sin(\omega)\sin(\omega t_m)}{5-4cos(\omega)}&\\
535: +i \ &\frac{2 \sin(\omega)(1-2^{t_m} \cos(\omega t_m))-2^{t_m} \sin(\omega t_m)(1-2\cos(\omega)))}{5-4cos(\omega)}&
536: \eaR
537: \eeq
538:
539: \nid Therefore, it exhibits oscillations due to the cut off $t_m$.
540: For longer time, mixing should lead to an exponential decay of the response.
541:
542: To check this we have first computed (\ref{chinum}) for the perturbation (\ref{Xm}) and for the observable
543: $A(x)=x-\frac{x^2}{2}$ (the derivative of $A$ gives
544: a maximal value for the term $A'(f^{t-n}(-1+u))$ in equation (\ref{ExpPole})).
545: The function $X_-$ has been computed with the recursion (\ref{wrec}) up to the
546: order $4$. The $\epsilon$ values $10^{-2},10^{-3}$,$10^{-4} (T=10^5,N=400)$,$10^{-5} (T=10^6,N=400)$,$10^{-6} (T=4\times 10^6,N=1600)$ have been considered
547: (but only $10^{-4},10^{-5},10^{-6}$ are represented in Fig. \ref{Fres2}a,b for the legibility
548: of the figure.). We have
549: used a fit procedure to compare the experimental data with the form (\ref{Explth}).
550: We have also computed the approximate linear response. The results are
551: presented in Fig. \ref{Fres2}a,b.
552:
553: %
554: %
555: %
556: %
557: %%%%%%%%%%%%%%%%%%%% Susceptibilité
558: \begin{figure}[ht]
559: \begin{center}
560: \includegraphics[height=7cm,width=7cm,clip=false]{ExplicSusceptibiliteXPhimordre4obsxmx2.eps}
561: \hspace{1cm}
562: \includegraphics[height=7cm,width=7cm,clip=false]{Reponsemap1DpertPhimordre4obsxmx2.eps}
563: \vspace{0.5cm}
564: \caption{(a) Susceptibility for the perturbation
565: $X_-(x)$, where $w$ was computed up to order $4$,
566: and observable $A(x)=x-\frac{x^2}{2}$. The parameter $\epsilon$ takes
567: the value $10^{-5},10^{-6}$. In full line are drawn the fitting curves obtained
568: from (\ref{Explth}).
569: (b)Linear response.
570: \label{Fres2}}
571: \end{center}
572: \end{figure}
573: %%%%%%%%%%%%%%%%%%%%%%%%%
574: %
575: %%
576: %
577: %
578: %
579:
580: One observes indeed a clear dependency with $\epsilon$. Moreover, the real and imaginary part
581: exhibit the expected oscillations due to the cut-off, with a perfect agreement with the
582: form (\ref{Explth}). Taking the inverse Fourier series one obtains the linear response
583: in Fig. \ref{Fres2}b. One sees clearly the exponential growth $2^t$ and the $\epsilon$
584: dependent cut-off. Performing a fit in log scale one obtains (fig. \ref{FInterpolReponseunebossecutoff})
585: an exponential increase with a rate $0.65$ very close to the expect value $log(2)=0.693$.
586: After this there is an exponential decay with an approximate rate $-0.24$. (We have only represented
587: the fit for $\epsilon=10^{-6}$ but one sees easily in Fig.\ref{FInterpolReponseunebossecutoff}
588: that the decay rate is similar for $\epsilon=10^{-4},10^{-5}$). This rate
589: is slower than the first Ruelle Pollicott resonance $-\log(4)=-1.386$. It might be that we are observing a crossover regime
590: where exponential amplification and exponential mixing are competing.
591:
592: %
593: %
594: %
595: %
596: %%%%%%%%%%%%%%%%%%%% Fit
597: \begin{figure}[ht]
598: \begin{center}
599: \includegraphics[height=7cm,width=7cm,clip=false]{InterpolReponseunebossecutoff.eps}
600: \vspace{0.5cm}
601: \caption{Fit of the linear response corresponding to Fig. \ref{Fres2}b.
602: \label{FInterpolReponseunebossecutoff}}
603: \end{center}
604: \end{figure}
605: %%%%%%%%%%%%%%%%%%%%%%%%%
606: %
607: %%
608: %
609: %
610: %
611:
612:
613:
614: \su{The H\'enon map.}\label{Henon}
615:
616: The H\'enon map:
617:
618: \beq
619: \left\{
620: \baR{ccc}
621: x_{t+1}&=& 1-ax_t^2 + y_t\\
622: y_{t+1}&=& bx_t
623: \eaR
624: \right.
625: \eeq
626:
627: \nid with $a=1.4,b=0.3$ has a strange attractor with a fractal
628: structure. Moreover, numerical estimates of the largest Lyapunov exponent,
629: using e.g. Eckmann-Ruelle algorithm \cite{ER} gives a value $\sim 0.422$.
630: But the H\'enon map is not uniformly hyperbolic. There are points with tangencies between
631: stable and unstable manifolds. These points may be responsible for phenomena analogous
632: to those described in the previous section,
633: for specific perturbations weighting those points. From this observation,
634: it is conjectured that there may also exist
635: a pole in the upper-half complex plane (and possibly more complex
636: singularities \cite{RCom}). However, there is no mathematical result for
637: this and the form of the perturbations/observable leading to
638: such a singularity is not known.
639:
640: Using the same method as in the previous section, a natural empirical approach
641: consists in applying a perturbation $\epsilon e^{i\omega t}$ in one of the directions $(x,y)$
642: and investigating the effect on the linear response for one of the variables $(x,y)$.
643: As an example, we have numerically computed the response of the observable $A(x,y)=x$ to a perturbation
644: $\epsilon e^{i\omega t}$ in the direction $y$. Denote by $\chi_{xy}(\omega)$ and $R_{xy}(t)$ the corresponding susceptibility and
645: approximate response. They are drawn in Fig. \ref{SusHenon}a,b
646: respectively, for $\epsilon=10^{-3} (T=10^5,N=100),10^{-4} (T=10^5,N=100),2 \times 10^{-5} (T=10^6,N=400),10^{-5} (T=10^6,N=1600)$. Note that essentially the same curves are obtained,
647: up to a phase factor, when perturbing direction $y$ and computing the response of $y$. In particular $|\chi_{xy}(\omega)|=|\chi_{yy}(\omega)|$.
648:
649:
650: %
651: %
652: %%%%%%%%%%%%%%%%%%%% Susceptibilité pour une perturbation constante en x et observable y, modèle de Hénon
653: \begin{figure}[ht]
654: \begin{center}
655: \includegraphics[height=9cm,width=7cm,clip=false]{SusceptibiliteHenonpertunobsX0.eps}
656: \hspace{1cm}
657: \includegraphics[height=9cm,width=7cm,clip=false]{ReponseHenonpertunobsX0.eps}
658: \vspace{0.5cm}
659: \caption{(a) Susceptibility for a constant perturbation $X=\epsilon$
660: in the direction $y$ for the observable $A(x,y)=x$.
661: (b) Corresponding linear response.}\label{SusHenon}
662: \end{center}
663: \end{figure}
664: %%%%%%%%%%%%%%%%%%%%%%%%%
665: %
666: %
667: %
668: %
669: %
670: %
671: %
672:
673: One observes similar effects as in the previous section. The amplitude
674: of the complex susceptibility increases with $\epsilon$ (and this effect is not stabilized, neither
675: by increasing the time $T$ of average well beyond the criterion $\epsilon\omega T >>1$,
676: nor by increasing the number of sample trajectories). The linear response has three parts.
677: For short times ($t<20$) there is a small bump for
678: the modulus of the response: its height does not seem to depend on $\epsilon$.
679: The perturbation/observables have a nonzero projection onto
680: the Ruelle-Pollicott modes. It might be that this bump comes from
681: this part, having a well behaved linear response.
682: The second part exhibits an exponential increase with a time cut-off depending
683: on $\epsilon$. This is very similar to the observations made for the
684: logistic map: exponential amplification of the linear response
685: until nonlinearities induce saturation of this effect.
686: Then, mixing leads to an exponential decay (third part of the curve).
687:
688: The exponential instability is well fitted by the curve $e^\alpha e^{\lambda t}$ with $\alpha =-6.80 \pm 0.87$
689: and $\lambda=0.455 \pm 0.038$ (see Fig. \ref{Finterpol}).
690: Note that the positive Lyapunov exponent of the H\'enon map
691: is $0.42(2)$ for these values of $a,b$.
692: The decay in the third part is well interpolated by a sum
693: of exponential terms.
694: The dominant term is $\beta e^{\gamma t}$
695: with $\gamma=-0.129\pm 0.015$. The value of $\gamma$
696: is in agreement with the exponential decay of the
697: correlation function
698: $C_{xy}(t)=<f^t(x) y>-<x><y>$,
699: where $<g>=\int g(X) \rho(dX), \ X=(x,y)$ and $\rho$ is the SRB measure ( Fig. \ref{Finterpol}). Thus the last part apparently
700: obeys the fluctuation-dissipation theorem.
701: Correlation function has been computed by the standard
702: Fast Fourier Transform method (see \cite{NR}, chapter
703: 13.2).
704:
705:
706: Let us also remark that there are no thin peak in the susceptibility
707: (the frequency resolution is $0.00612$) and the thickness of the peaks
708: does not change with $\epsilon$.
709: Since the width\footnote{Consider the complex susceptibility in the neighbourhood of a simple
710: pole $\omega_0=x_0 + i y_0$. Then $\chi(\omega) \sim\frac{A}{\omega-\omega_0}$. Assume that this pole is close to
711: the real axis. Set $\omega=x+iy$. On the real axis $|\chi(\omega)|^2 \sim \frac{|A|^2}{(x-x_0)^2+y_0^2}$.
712: The modulus is maximal at $x=x_0$, and has value $\frac{|A|}{y_0}$ (resonance peak).
713: The width of the resonance peak is $2|x_1-x_0|$ where $x_1$ is such
714: that $[\chi(x_1)|=\frac{|A|}{2 y_0}$. Thus, the width is $2y_0$. } of the resonance peak is given by the imaginary part
715: of the pole, this suggests that the imaginary part of the poles is bounded away from
716: zero. This leads us to conjecture that there is no pole on the real axis
717: (for the frequency $\omega$). This conjecture is compatible\footnote{I thank
718: one of the reviewers for this remark.} with
719: the available results (in particular exponential decay of correlations
720: for many parameters $a$ and $b$ (see \cite{BeYou1,BeYou2,You,WaYou})).
721:
722:
723: %
724: %%%%%%%%%%%%%%%%%%%% Interpolation exponentielle, modèle de Hénon
725: \begin{figure}[ht]
726: \begin{center}
727: \includegraphics[height=7cm,width=10cm,clip=false]{resonancesHenon.eps}
728: \vspace{0.5cm}
729: \caption{(a) Fit of the linear response $|R_{xy}(t)|$, $\epsilon=10^{-6}$ for the H\'enon map.
730: (b) Fit of the correlation function $C_{xy}(t)$ for the H\'enon map.}\label{Finterpol}
731: \end{center}
732: \end{figure}
733: %%%%%%%%%%%%%%%%%%%%%%%%%
734: %
735: %
736: %
737: %
738: %
739: %
740: %
741:
742:
743: \su{Conclusion.}
744:
745:
746:
747: There is currently an intensive research activity in mathematics dealing with linear response
748: theory for interval maps (see e.g. \cite{BA1,BA2} and reference therein).
749: In the case of non uniformly hyperbolic maps one can find examples where the susceptibility has an arbitrary large number of poles
750: in the upper-half complex plane \cite{JiRu} or where the average of a smooth function
751: with respect to the SRB measure of the perturbed system $f + \epsilon X$ is not even Lipschitz
752: at $\epsilon=0$ \cite{BA1}. These results are intriguing for a physicist and raise some natural
753: questions: can we measure some effects related to these situations?
754: What is the ``genericity'' of these examples? Do these effects arise in dynamical systems modeling physical
755: systems ? do they occur in larger dimensions? Answering these questions at the level of
756: rigor of mathematics is probably a too formidable task at the current state of research in this field.
757: One hope is that some hints can be provided by numerical approaches.
758:
759: In this spirit, we have presented in this work a numerical procedure allowing to
760: compute the complex susceptibility/linear response. For this, we use an approximation
761: of the complex susceptibility (\ref{susc}) by a finite sum, which
762: can be easily computed. In the cases where the series (\ref{susc}) converges
763: this sum gives a fairly good approximation. When there is a pole in the upper-half
764: plane a correct treatment of this sum allows us to recover the pole location.
765: Applying this procedure to the H\'enon model we conjecture that such
766: a pole could also be present.
767:
768: There are several possible developments of the present work.
769: On mathematical grounds it could be interesting to apply
770: this procedure to the cases discussed in e.g. \cite{JiRu,BA1,BA2}
771: that can be addressed by rigorous methods. (In particular the cases in
772: \cite{BA1,BA2} should provide examples where the linear response
773: does not decay exponentially, but does not grow exponentially either.)
774: An intermediate case between these mathematically tractable cases and
775: the difficult H\'enon model would be Collet-Eckmann maps.
776: Another question is: ``what is there beyond the pole'' ?
777: In some sense, the pole in the upper-half complex plane may ``hide'' more complex
778: singularities lying behind it such as cuts.
779: Is it possible to have any (numerical) idea
780: of which type of singularities are there ? A natural
781: way of doing this is to remove the exponential instability by
782: multiplying the perturbation by a damping factor (or equivalently
783: to use a complex frequency $\omega$). This is natural from a mathematical
784: point of view, but tricky in the numerics, because the damping factor
785: is either smaller than the exponential instability (and the response grows rapidly inducing nonlinear effects, as we saw)
786: or it is bigger (and the perturbation becomes rapidly numerically $0$, then we are measuring short time transients).
787: One has thus to make small variations of the damping factor around the pole and look
788: at the changes in the susceptibility/response curve.
789:
790: On more physical grounds, a natural continuation
791: of the present work, could be to consider a lattice of coupled logistic maps,
792: to apply an harmonic perturbation at some point and look at the induced effects,
793: in the spirit of this work. This would be one step
794: towards the investigation the effects of large dimension (``thermodynamic limit'')
795: on the singularity induced by the microscopic dynamics.
796:
797:
798: \ack
799: I warmly thank David Ruelle for suggesting to me the
800: present work and for helpful remarks and suggestions.
801: I thank Jean-Louis Meunier for illuminating advices and Jacques-Alexandre Sepulchre
802: for helpful discussions.
803: I think that this paper has been widely improved due to the remarks of the referees
804: and their positive criticism. I greatly acknowledge them. \\
805:
806:
807: \begin{thebibliography}{99}
808:
809: \bibitem{BA1} Baladi V., ``On the susceptibility function of piecewise expanding interval maps'',
810: Preprint arxiv.org (v1: 2006, v2, v3:2007), to appear in Comm. Math. Phys. (2007).
811:
812: \bibitem{BA2} Baladi V., Smana D., ``Linear response formula for piecewise expanding maps'', Preprint arxiv.org, submitted for publication, (2007).
813:
814: \bibitem{BaladiRugh} Baladi V., Rugh H. H., Jiang Y., ``Dynamical determinants via dynamical conjugacies for postcritically finite
815: polynomials'', J. Stat. Phys., 108, 973-993 (2002)
816:
817: \bibitem{BeYou1} Benedicks M. Young L.S., ``Sinai-Bowen-Ruelle measures for certain Henon maps'', Invent. Math, 112, (1993), 541-576;
818:
819: \bibitem{BeYou2} Benedicks M. Young L.S.,``Markov extensions and decay of correlations for certain Henon maps'', Asterisque 261, (2000), 13-56.
820:
821: \bibitem{ER}Eckmann J.P., Ruelle D., ``Ergodic Theory of Strange attractors''
822: Rev. of Mod. Physics, 57, 617,(1985).
823:
824:
825: \bibitem{JAB1} Cessac B., Sepulchre J.A., ``Stable resonances and signal propagation in a chaotic network of coupled units'', Phys. Rev. E 70, 056111 (2004).
826:
827: \bibitem{JAB2} Cessac B., Sepulchre J.A., "Transmitting a signal by amplitude modulation in a chaotic network'", Chaos 16, 013104 (2006).
828:
829: \bibitem{JAB3} Cessac B., Sepulchre J.A., "Linear response, susceptibility and resonances in chaotic toy models", Physica D, Volume 225, Issue 1 , Pages 13-28, (2007).
830:
831: \bibitem{Gaspard} Gaspard P. ``Chaos, scattering and statistical mechanics'',
832: Cambridge Non-Linear Science series 9, (1998)
833:
834: \bibitem{NR} Press W, Flannery P., Teukolsky S., Vetterling W., ``Numerical Recipes in C'', Cambridge University
835: Press (1988)
836:
837: \bibitem{Pollicott} Pollicott M., {\em Invent. Math., 81, 413-426 (1985)}, Ruelle D., {\em J. Differential Geometry, 25 (1987), 99-116}.
838:
839: \bibitem{Reick} Reick C.H., ``Linear response of the Lorentz system'', Phys. Rev. E, 66, 036103, (2002).
840:
841: \bibitem{RuelleP} Ruelle D., {\em J. Differential Geometry, 25 (1987), 99-116}.
842:
843: \bibitem{Ru97} D. Ruelle, Differentiation of SRB states, {\it Com.
844: Math. Phys.} {\bf 187} (1997) 227-241.
845:
846: \bibitem{Ru03} D. Ruelle, Differentiation of SRB states: Corrections
847: and Complements, {\it Com. Math. Phys.} {\bf 234} (2003) 185-190.
848:
849: \bibitem{Ru98} D. Ruelle, General linear response formula in
850: statistical mechanics, and the fluctuation-dissipation theorem far from
851: equilibrium, {\it Phys. Lett. A} {\bf 245} (1998) 220-224.
852: \bibitem{CMP} Ruelle D., ``Differentiating the absolutely continuous invariant measure
853: of an interval map $f$ with respect to $f$'', Commun. Math. Phys., 258, 445-453 (2005).
854:
855:
856: \bibitem{Ru99} D. Ruelle, Smooth Dynamics and new theoretical ideas in
857: nonequilibrium statistical mechanics, {\em J. of Stat. Phys.} {\bf
858: 95} (1999) 393-468.
859:
860: \bibitem{RCom} Ruelle D. Private communication.
861:
862: \bibitem{JiRu} Jiang Y., Ruelle D.,
863: "Analyticity of the susceptibility function for unimodal Markovian maps of the interval." Nonlinearity 18, 2447-2453 (2005).
864:
865: \bibitem{WaYou} Wang Q.D., Young L.S., ``Strange attractors with one direction of instability'', Commun. Math. Phys. 218, (2001), 1--97.
866:
867: \bibitem{You} Young L.S. ``Statistical properties of dynamical systems with some hyperbolicity'', Annals of Math., (1998), 585-650.
868: \bibitem{Young} Young L.S. ``What are SRB measures and which dynamical systems have them ?'', J. Stat. Phys., 108,
869: (2002), 733-754.
870:
871: \end{thebibliography}
872:
873: \ed
874:
875:
876: