nlin0004030/fd.tex
1: \documentstyle[aps,pre,manuscript]{revtex}   
2: 
3: \begin{document}
4: 
5: \draft
6: \title{Fluctuation-dissipation relationship in chaotic dynamics}
7: 
8: \author{Bidhan Chandra Bag and 
9: Deb Shankar Ray {\footnote {e-mail : pcdsr@mahendra.iacs.res.in} }
10: }
11: 
12: \address{Indian Association for the Cultivation of Science,
13: Jadavpur, Calcutta 700 032, INDIA.}
14: 
15: \maketitle
16: 
17: \begin{abstract}
18: We consider a general N-degree-of-freedom dissipative 
19: system which admits of chaotic behaviour. 
20: Based on a Fokker-Planck description associated with the dynamics
21: we establish that the drift and the diffusion coefficients
22: can be related through a set of stochastic parameters which
23: characterize the steady state of the dynamical system in a way similar to
24: fluctuation-dissipation relation in non-equilibrium statistical mechanics.
25: The proposed relationship is verified by numerical experiments on a
26: driven double well system.
27: \end{abstract}
28: 
29: \vspace{2.5cm}
30: 
31: \pacs{PACS number(s) : 05.45.-a, 05.70.Ln, 05.20.-y} 
32: 
33: \newpage
34: 
35: 
36: \section{Introduction}
37: 
38: Although deterministic in principle classically chaotic motion is stochastic
39: in nature. Ever since the early numerical study of Chirikov mapping \cite{cas}
40: revealed that the motion of a phase space variable can be characterized by
41: a simple random walk diffusion equation, attempts have been made to describe 
42: the chaotic motion in terms of Langevin or Fokker-Planck equations \cite{cas,lic}. 
43: It is therefore
44: easy to comprehend a close connection between classical chaos and statistical
45: mechanics. Two distinct situations arise in this context. The first one
46: concerns whether classical chaos may serve as a basis for classical 
47: statistical  mechanics since the ultimate justification of the postulates
48: of statistical mechanics like Boltzmann hypothesis of molecular chaos, ergodicity or
49: the postulate of equal a priori probability rests on the dynamics of 
50: each particle \cite{jkb,ma,gas}. The second one concerns the following : Given that the classical 
51: chaotic motion is stochastic how and to what extent one can realize the
52: formulation of statistical mechanics  
53: for useful description of classical chaos 
54: \cite{kai,ono,wid,ber,gras,bon,bon1,zur,pat,cohen,pat1,sc1,sc2,sc3,bb1,bb2}
55: keeping in mind that one essentially deals here with a few-degree-of-freedom 
56: system. The present paper addresses the second issue.
57: 
58: The emergence of stochastic behaviour of the classically chaotic system is 
59: due to the loss of correlation of initially nearby trajectories. This is
60: reflected in the nature of the largest Lyapunov exponent \cite{ben} whose calculation rests
61: on the linear equation of motion for the separation of these trajectories.
62: When chaos has fully set in, the time dependence of the linear stability matrix
63: or Jacobian of the system \cite{fox} in the equation of motion in the tangent space can be
64: described as a stochastic process since the phase space variables behave as 
65: stochastic variables. In a number of recent studies we have shown 
66: \cite{sc1,sc2,sc3,bb1,bb2} that this 
67: fluctuation of the Jacobian  is amenable to a theoretical description in
68: terms of the theory of multiplicative noise. This allows us to realize a number 
69: of important results of nonequilibrium statistical mechanics, like
70: Kubo relation \cite{sc1}, fluctuation-decoherence relation \cite{sc2}, exponential divergence
71: of quantum fluctuations \cite{sc3,bb1,bb2}, thermodynamically inspired quantities, e. g. ,
72: entropy production in chaotic dynamics. Based on a Fokker-Planck description in the tangent space
73: where the drift and the diffusion coefficients explicitly depend on the phase space
74: variables or dynamical properties of the system, we show that a connection
75: between the two moments in terms of the stochastic parameters which 
76: characterize the long time limit of the dynamical system can be established
77: in the spirit of fluctuation-dissipation relation. 
78: We verify the theoretical
79: proposition by numerical experiments on a simple dissipative 
80: system.
81: 
82: The rest of the paper is organized as follows : In Sec. II we introduce a
83: Fokker-Planck description of the dynamical system in the tangent space
84: and identify the drift and diffusion coefficients which are the functions
85: of fluctuations of the phase space variables. This is followed by solving the 
86: Fokker-Planck equation for the steady state distribution required for the calculation
87: of long time averages in Sec III. In Sec. IV the dynamical stochastic parameters
88: which characterize the long time behaviour of the system are introduced. The first one of them is
89: a well-known stochastic parameter closely related to Kolmogorov entropy. 
90: With the help of these stochastic parameters we 
91: establish a connection between the drift and diffusion coefficients of the Fokker-Planck
92: equation in the spirit of
93: fluctuation-dissipation relation in nonequilibrium statistical mechanics.
94: In Sec. V we illustrate the general method by an explicit numerical
95: example to verify the
96: theoretical proposition. The paper is concluded in Sec. VI.
97: 
98: 
99: \section{A Fokker-Planck equation for dissipative 
100: chaotic dynamics}
101: 
102: We are concerned here with 
103: a general N-degree-of-freedom system whose Hamiltonian is
104: given by
105: \begin{equation}
106: H = \sum_{i=1}^N \frac{p_i^2}{2 m_i} +V(\{q_i\}, t) \; \;,
107: \; \; i = 1 \cdots N  \label{e1}
108: \end{equation}
109: where $\{q_i, p_i\}$ are the co-ordinate and 
110: momentum of the i-th degree-of-freedom, respectively, 
111: which satisfy the generic form of equations 
112: 
113: \begin{equation}
114: \dot {q}_i = \frac{\partial H}{\partial p_i}  \; \; {\rm and} \; \; 
115: \dot {p}_i = -\frac{\partial H}{\partial q_i}  \; \;.
116: \label{e2}
117: \end{equation}
118: 
119: We now make the Hamiltonian system dissipative by introducing $-\gamma p_i$
120: on the right hand side of 
121: the second of Eqs.(2). 
122: For simplicity we assume $\gamma$ to be the
123: same for all the N degrees of freedom. 
124: By invoking the 
125: symplectic structure of the 
126: Hamiltonian dynamics as
127: \begin{eqnarray*}
128: z_i =
129: \left\{ \begin{array}{ll}
130: q_i & \; \; {\rm for} \; i=1 \cdots N  \; \;, \\
131: p_{i-N} & \; \; {\rm for} \; i=N+1, \cdots 2N \; \;.
132: \end{array}
133: \right.
134: \end{eqnarray*}
135: and defining I as 
136: \begin{eqnarray*}
137: I =
138: \left[
139: \begin{array}{cc}
140: 0 & E \\
141: -E & -\gamma E
142: \end{array}
143: \right]
144: \end{eqnarray*}
145: where E is an $N\times N$ unit matrix, and $0$ is an  $N\times N$ null matrix,
146: the equation of motion for the dissipative system can be written 
147: as
148: \begin{equation}
149: \dot{z}_i = \sum_{j=1}^{2N} 
150: I_{ij} \frac{\partial H}{\partial z_j} \hspace{0.1cm}. \label{e3}
151: \end{equation}
152: 
153: We now consider two nearby trajectories, $z_i, \dot{z}_i$ and
154: $z_i+ X_i$, $\dot{z}_i+ \dot{X}_i$ at the same time $t$
155: in $2N$ dimensional phase space. The time evolution of separation of these 
156: trajectories is then determined by
157: \begin{equation}
158: \dot{X}_i = \sum_{j=1}^{2N} J_{ij}(t) X_j \label{e4}
159: \end{equation}
160: in the tangent space $\{X_i\}$, where
161: \begin{eqnarray}
162: J_{ij}= \sum_k I_{ik} \frac{\partial^2 H}{\partial z_k \partial z_j}\; \; .
163: \label{e5}
164: \end{eqnarray}
165: Therefore the $2N \times 2N$ linear stability 
166: matrix $\underline{J}$ assumes the following form
167: \begin{equation}
168: \underline{J} =
169: \left[
170: \begin{array}{cc}
171: 0 & E \\
172: \underline{M(t)} & -\gamma E
173: \end{array}
174: \right]
175: \label{e6}
176: \end{equation}
177: where \underline{M} is an $N\times N$ matrix. Note that the time dependence of 
178: stability matrix $\underline{J}(t)$ is due to the second derivative 
179: $\frac{\partial^2 H}{\partial z_k \partial z_j}$ which is determined \cite{fox} by 
180: the equation of motion (3). 
181: The procedure for calculation of $X_i$
182: and the related quantities is to solve the 
183: trajectory equation (3) simultaneously
184: with Eq.(4). Thus when the dissipative 
185: system described by 
186: (3) is chaotic, \underline{J}(t) becomes 
187: (deterministically) stochastic 
188: due to the fact that $z_i$-s behave as stochastic variables
189: and the equation of motion (\ref{e4}) in the tangent space
190: can be interpreted as a stochastic equation \cite{sc1,sc2,sc3,bb1,bb2}.
191: 
192: In the next step we shall be concerned with a stochastic description of 
193: $\underline{J}(t)$ or $\underline{M(t)}$. For convenience we split up 
194: \underline{M} into two parts as
195: \begin{equation}
196: \underline{M} = \underline{M_0} + \underline{M_1(t)} \label{e7}
197: \end{equation}
198: where $\underline{M_0}$ is independent of variables $\{z_i\}$ 
199: and therefore behaves a sure or constant part
200: and $\underline{M_1}$ is determined by the
201: variables $\{z_i\}$ for $i = 1 \cdots 2N$. 
202: $\underline{M_1}$ refers to the fluctuating part.
203: We now rewrite the equation of motion (\ref{e4}) in 
204: tangent space as
205: \begin{eqnarray}
206: \dot{X} & = & \underline{J} X  \nonumber \\
207:        & = & L\left(\left\{X_i\right\}, \left\{z_i\right\}\right)
208:        \label{e8}
209: \end{eqnarray}
210: where $X$ and $L$ are the vectors with $2N$ components. Corresponding to 
211: (\ref{e7}) $L$ in (\ref{e8}) can be split up again to yield
212: 
213: \begin{eqnarray}
214: \dot{X} = L^0(X) + L^1(X, \{z_i(t)\})   
215: %\nonumber \\
216: %{\rm or} \; \; \; \; \; 
217: %\dot {X}_i & = & {L_i}^0 (\{X_i\})
218: %+L^1_i(\{X_i\},\{z_i\}) 
219: \; \; , \; \; \; \; i = 1 \cdots 2N \; \; .
220: \label{e9}
221: \end{eqnarray}
222: 
223: Eq.(\ref{e4}) indicates that Eq.(\ref{e8}) 
224: is linear in $\{X_i\}$. Eqs.(\ref{e4}), (\ref{e5}) and (\ref{e6})
225: express the fact the first $N$ components of $L^1$ are zero and the last $N$ 
226: components of $L^1$ are the functions of $\{X_i\}$ for $i = 1 \cdots N$.
227: The fluctuation in $L_i^1$ is caused by the chaotic variables 
228: $\left\{z_i\right\}$-s.
229: This allows us to write the following relation (which will be used later on),
230: \begin{equation}
231: \nabla_X \cdot L^1 \phi(\{X_i\}) = L^1 \cdot \nabla_X \phi(\{X_i\}) \; \;
232: \label{e10}
233: \end{equation}
234: where $\phi(\{X_i\}$) is any function of $\{X_i\}$. $\nabla_X$ refers to
235: differentiation with respect to components $\{X_i\}$ (explicitly 
236: $X_i = \Delta q_i$ for $i = 1 \cdots N$ and  $X_i = \Delta p_i$ for 
237: $i = N+1 \cdots 2N$).
238: 
239: Note that Eq.(\ref{e9}) by virtue of (\ref{e8}) 
240: is a linear stochastic differential equation
241: with multiplicative noise where the noise is due to $\{z_i\}$ determined by 
242: equation of motion (\ref{e3}). 
243: This is the starting point of our further analysis.
244: 
245: Eq.(\ref{e9}) determines a stochastic process with 
246: some given initial conditions
247: $\{X_i(0)\}$. We now consider the motion of a representative point $X$ in
248: $2N$ dimensional tangent space ($X_1 \cdots X_{2N}$) as governed by 
249: Eq.(\ref{e9}).
250: The equation of continuity which expresses the conservation of points
251: determines the variation of density function $\phi(X, t)$ in time as given by
252: \begin{equation}
253: \frac{\partial \phi(X, t)}{\partial t} = - {\bf \nabla_X} \cdot L(t) 
254: \phi(X, t) \; \;.
255: \label{e11}
256: \end{equation}
257: 
258: Expressing $A_0$ and $A_1$ as
259: \begin{equation}
260: A_0  =  - \nabla_X \cdot L^0 \; \; {\rm and} \; \; 
261: A_1 =  - \nabla_X \cdot L^1 
262: \label{e12}
263: \end{equation}
264: we may rewrite the equation of continuity as
265: \begin{equation}
266: \frac{\partial \phi(X, t)}{\partial t} = [A_0 + \alpha A_1(t)] 
267: \phi(X, t) \; \;.
268: \label{e13}
269: \end{equation}
270: 
271: It is easy to recognize that while $A_0$ denotes the sure part $A_1$ contains the 
272: multiplicative fluctuations through $\{z_i(t)\}$. $\alpha$ is a parameter
273: introduced from outside to keep track of the order of fluctuations in the 
274: calculations. At the end we put $\alpha =1$. 
275: 
276: One of the main results for the linear equations of the form with 
277: multiplicative noise may now be in order \cite{van}. The average equation of 
278: $\langle \phi \rangle$ obeys [ $P(x, t) \equiv \langle \phi \rangle$],
279: \begin{equation}
280: {\dot P} = \left \{ A_0 + \alpha \langle A_1 \rangle
281: + \alpha^2 \int_0^\infty d\tau
282: \langle \langle A_1(t) \exp(\tau A_0) A_1(t-\tau) 
283: \rangle \rangle \exp (-\tau A_0) \right \} P(x, t) \; \; .
284: \label{e14}
285: \end{equation}
286: 
287: The above result is based on second order cumulant expansion and is 
288: valid when fluctuations are small but rapid and the correlation 
289: time $\tau_c$ is short but finite or more precisely
290: 
291: \begin{equation}
292: \langle \langle A_1(t) A_1(t') \rangle \rangle = 0 \; \; 
293: \; \; \; {\rm for} \; \; 
294: \left|t-t'\right| \; > \; \tau_c
295: \label{e15}
296: \end{equation}
297: We have, in general, $\langle A_1 \rangle \ne 0$. Here  
298: $\langle \langle \cdots \rangle \rangle$ implies
299: $\langle \langle \zeta_i \zeta_j \rangle \rangle 
300: = \langle \zeta_i \zeta_j \rangle -\langle \zeta_i \rangle \langle \zeta_j 
301: \rangle $.
302: 
303: The Eq.(\ref{e14}) is exact in the limit $\tau_c \rightarrow 0$. 
304: Making use of relation (\ref{e12}) in (\ref{e14}) we obtain 
305: \begin{eqnarray}
306: \frac{\partial P}{\partial t} & = & 
307: \left\{ -\nabla \cdot  L^0  -\alpha \langle \nabla \cdot  L^1 \rangle
308: + \alpha^2 \int_0^\infty d\tau
309: \langle \langle \nabla \cdot L^1(t)   \exp ( -\tau \nabla \cdot L^0)  
310: \right.\nonumber\\
311: & & \left. \nabla \cdot L^1(t-\tau) \rangle \rangle
312: \exp ( \tau \nabla \cdot L^0 )  \right\} P \; \; .
313: \label{e16}
314: \end{eqnarray}
315: The above equation can be transformed into the following Fokker-Planck
316: equation ($\alpha = 1$) for probability density function $P(X,t)$, 
317: (the details are given in the Appendix A);
318: \begin{equation}
319: \frac{\partial P(X,t)}{\partial t} = -\nabla . F P +
320: \sum_{i,j} {\cal D}_{ij} \frac{\partial^2 P}{\partial X_i \partial X_j}
321: \label{e17}
322: \end{equation}
323: where,
324: \begin{equation}
325: F=L^0 + \langle L^1 \rangle + Q
326: \label{e18}
327: \end{equation}
328: and $Q$ is a $2N$-dimensional vector whose components are
329: defined by
330: \begin{eqnarray}
331: Q_j=-\int_0^\infty \langle \langle R'_j \rangle \rangle
332: d \tau d_1(\tau) d_2(\tau)
333: \label{e19}
334: \end{eqnarray}
335: Here the determinants $Det_1(\tau)$, $Det_2(\tau)$ and $R'_j$ are given by
336: \begin{eqnarray}
337: Det_1(\tau)& = & \left|\frac{d X^{-\tau}}{d X} \right|
338: \; \; \; \; {\rm and} \; \; Det_2(\tau)=
339: \left|\frac{d X}{d X^{-\tau}} \right|
340: \nonumber \\
341: {\rm and} \; \; \; R'_j & = & \sum_i L_i^1(X,t) \frac{\partial}{\partial X_i}
342: \sum_k L_k^1(X^{-\tau},t-\tau) \frac{\partial X_j}{\partial X_k^{-\tau}}
343: \; \; .
344: \label{a20}
345: \end{eqnarray}
346: 
347: It is easy to recognize $F$ as an evolution operator. Because 
348: of the dissipative perturbation we note that div $F < 0$.
349: 
350: The diffusion coefficient ${\cal D}_{ij}$ in Eq.(\ref{e17}) is defined as
351: \begin{equation}
352: {\cal D}_{ij}=\int_0^\infty \sum_k \langle \langle L_i^1(X,t) 
353: L_k^1(X^{-\tau},t-\tau)\frac{dX_j}{dX_k^{-\tau}}
354: \rangle \rangle Det_1(\tau) Det_2(\tau) d\tau
355: \label{e21}
356: \end{equation}
357: We have followed closely van Kampen's approach \cite{van} to generalized Fokker-Planck
358: equation (\ref{e17}). Before concluding this section several critical
359: remarks regarding this derivation need attention:
360: 
361: First, the stochastic process $\underline{M_1(t)}$ determined by $\{z_i\}$ is
362: obtained {\it exactly} by solving equations of motion (\ref{e3}) for the
363: chaotic motion of the system. It is therefore necessary to emphasize that we have 
364: {\it not assumed} any special property of noise, such as, $\underline{M_1(t)}$
365: is Gaussian or $\delta$-correlated. We reiterate Van Kampen's
366: emphasis in this approach.
367: 
368: Second, the only assumption made about the noise is that its
369: correlation time $\tau_c$ is short but finite compared to the
370: coarse-grained timescale over which the average quantities evolve.
371: 
372: Third, we take care of fluctuations upto second order which implies 
373: that the deterministic noise is not too strong.
374: 
375: Eq.(\ref{e17}) is the required Fokker-Planck equation in the tangent space
376: $\{X_i\}$. Note that the drift and diffusion coefficients are
377: determined by the phase space $\{z_i\}$ properties of the chaotic system and
378: directly depend on the correlation functions of the fluctuations of the
379: second derivatives of the Hamiltonian (\ref{e5}).
380: 
381: 
382: \section{The Steady state distribution and the calculation of averages}
383: 
384: In what follows we shall be concerned with the long time limit of the 
385: dynamical system. Thus the steady state distribution of the tangent space 
386: co-ordinates
387: $X_i(i = 1\cdots 2N)$ are specially relevant for the present
388: purpose. To make all these co-ordinates  dimensionless 
389: we use the following transformations in Eq.(17)
390: 
391: \begin{eqnarray*}
392: \tau^\prime & = & \omega' t  \; \;, \nonumber \\
393: y_i & = & \frac{X_i}{d_0} \; \; \; {\rm for} \; i = 1 \cdots N  \; \;,
394: \end{eqnarray*}
395: 
396: \begin{equation}
397: y_i = \frac{X_i}{\omega' d_0} \; \; \; \; \; 
398: {\rm for} \; i = N+1 \cdots 2N \; \;,
399: \end{equation}
400: 
401: \noindent
402: where $\omega'$ is a 
403: scaling constant having dimension of reciprocal of time 
404: (a possible choice is the linearized frequency of the dynamical system) and
405: $\tau'$ becomes a dimensionless variable. 
406: $d_0$ is a constant (to be specified later) having the dimension of length.
407: The resulting Fokker-Planck Eq.(17) reduces to
408: \begin{equation}
409: \frac{\partial P (y, \tau')}{\partial \tau'} = - \nabla . F' (y) P
410: +  \sum_{i,j} {\cal D'}_{ij} (y)  \frac{\partial^2 P}
411: {\partial y_i \partial y_j} \; \; .
412: \end{equation}
413: 
414: Note that Eq.(23) is independent of $d_0$ since $F(X)$ is linear in 
415: $\{X_i\}$ and ${\cal D}$($X$) is quadratic in $\{X_i\}$. 
416: Next we consider the stationary state of the system ($\frac{\partial P}
417: {\partial \tau'} = 0$) and make use of the following linear transformation
418: ( with $\alpha_{2N} = 1$ ) 
419: \begin{equation}
420: U = \sum_{i=1}^{2N} \alpha_i y_i
421: \end{equation}
422: \noindent
423: 
424: \noindent
425: in Eq.(23) to obtain the equation for steady state probability distribution
426: $P_s(U)$ :
427: 
428: \begin{equation}
429: \frac{\partial}{\partial U} \lambda U P_s (U) + {\cal D}_s \frac{\partial^2 P_s}
430: {\partial U^2} = 0 \; \; .
431: \end{equation}
432: 
433: \noindent
434: $\alpha_i$-s ($i = 1 \cdots 2N-1$) are the constants to be determined.
435: \noindent
436: Here 
437: \begin{equation}
438: \lambda U = - \sum_i \alpha_i F'_{i} (y)
439: \end{equation}
440: 
441: \noindent
442: and
443: \begin{equation}
444: {\cal D}_s = \sum_{i, j} {\cal D'}_{ij} \alpha_i \alpha_j \; \; ,
445: \end{equation}
446: and we disregard the time dependence of ${\cal D'}$ under weak noise
447: approximation, to treat ${\cal D'}$ as a constant in the usual way.
448: 
449: Putting (24) in (26) and comparing the coefficients of $y_i$ on both sides we
450: obtain $2N$ algebraic equation (for $\alpha_i......\alpha_{2N-1}$ and
451: $\lambda$). The set $\{ \alpha_i\}$ and $\lambda$ are therefore known.
452: 
453: The exact steady state solution, $P_s$ has the well known Gaussian form
454: which is given by
455: 
456: %\begin{eqnarray*}
457: %P_s = N \exp\left (-{\frac{\lambda U^2}{2 {\cal D}_s}}\right)
458: %\end{eqnarray*}
459: %\noindent
460: %or
461: \begin{equation}
462: P_s(\{y_i\}) = N \exp \left (-\frac{\lambda}{2 {\cal D}_s} \sum_{i, j}
463: \alpha_i \alpha_j y_i y_j \right ) \; \;,
464: \end{equation}
465: 
466: \noindent
467: where $N$ is the normalization constant.
468: Eq.(28) expresses the probability distribution of 
469: tangent space co-ordinates of 
470: the dynamical system in the long time limit. The important relevant quantity
471: which measures the separation of initially nearby trajectories when the 
472: system has attained the stationary state can be computed by calculating the
473: average of $\sum_{i=1}^{2N} y^2_i$. Making use of 
474: the distribution (28) we obtain
475: \begin{equation}
476: \left \langle \sum_{i =1}^{2N} y^2_i \right \rangle = \frac{{\cal D}_s}{\lambda}
477: \sum_{i = 1}^{2N} \frac{1}{\alpha^2_i}
478: \end{equation}
479: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
480: Note that the average as calculated above is a function of $D_s$, $\lambda$ and
481: $\alpha_i$-s which are dependent on the phase space properties of the dynamical 
482: system.
483: 
484: 
485: \section{Stochastic parameters, connection between ${\cal D}_s$ and $\lambda$ ;
486: Fluctuation-dissipation relation}
487: 
488: Eq.(25) is a steady state Fokker-Planck equation in tangent space
489: with linear drift and constant
490: diffusion coefficients where the co-ordinates have been 
491: expressed as dimensionless variables $\{y_i\}$. $\lambda$ and ${\cal D}_s$
492: are the first and second moments, respectively, of the underlying stochastic
493: process. Our objective here is seek for a connection between the two moments. In
494: standard nonequilibrium statistical mechanics this connection is expressed by the 
495: fluctuation-dissipation relation through temperature, an equilibrium parameter 
496: characterizing the equilibrium state. Our approach here is to
497: follow a somewhat similar procedure.
498: This implies that we search for the stochastic parameters which characterize 
499: the long time limit of the nonlinear dynamical system. We show that an 
500: appropriate relation between ${\cal D}_s$ and $\lambda$ can be established 
501: through these parameters.
502: 
503: An important parameter proposed many years ago by Casartelli et. al.\cite{tell}
504: (a precursor for the largest Lyapunov exponent used as a measure of 
505: regularity or chaoticity of a nonlinear dynamical system) is the long time
506: average of $\ln \frac{d(t)}{d_0}$ where $d_0$ is the separation of the 
507: two initially nearby  trajectories and $d(t)$ is the corresponding 
508: separation at some time $t$. To express $d(t)$ (having dimension of length)
509: we write $d(t) = [\sum_{i=1}^{N} (X_i)^2 + \sum_{i = N+1}^{2N} (\frac{X_i}{\omega'})^2]^{\frac{1}{2}}$.
510: $d(t)$ is determined by solving numerically Eqs. (3) and (4) 
511: simultaneously or their appropriately transformed version 
512: for the initial condition $z_0$ corresponding to Eq.(3). In going from j-th
513: to j+1-th step of iteration in course of time evolution 
514: any of the components of $X$ say $X_i$ has
515: to be initialized as $X_i^{j0} = \frac{X_i^j}{d_j} d_0$.
516: This initialization implies that at each step iteration starts with
517: same magnitude of $d_0$ but the direction of $d_0$ for step j+1 is that of
518: $d(t)$ for j-th step (considered in terms of the ratio $\frac{X_i^j}{d_j}$). For
519: a pictorial illustration we refer to Fig.1 of Ref. \cite{ben}.
520: j-th time of iteration implies $t = j T$ ($j = 1,2 \cdots \infty)$ and $T$ is 
521: the characteristic time which corresponds to the shortest ensemble averaged period
522: of nonlinear dynamical system.  Thus following Casartelli et. al. \cite{tell}
523: a stochastic parameter can be defined by the following time average
524: of $\ln \frac{d_j}{d_0}$ as
525: 
526: \begin{equation}
527: \sigma_n (t, z_0, d_0) = \frac{1}{n}\sum_j^n \ln \frac{d_j}{d_0}
528: \end{equation}
529: 
530: It has been shown \cite{tell} that as $n\rightarrow \infty$ $\sigma_n$ has a 
531: definite value. For the disordered system it is positive and for the regular 
532: system it is zero. The difference of $\sigma_n$ from the largest Lyapunov 
533: exponent is also noteworthy. Our object here is to generalize (30)
534: by defining the other higher  order moments (higher than the first
535: $\sigma_{n \rightarrow \infty}$).
536: To express these quantities we define first
537: 
538: \begin{equation}
539: \sigma' = {\rm ln} \frac{d(t)}{d_0}
540: \end{equation}
541: 
542: We now make use  of the
543: transformation (22) to express $d(t)$ as a dimensionless quantity in terms
544: of $\sigma'$ as follows:
545: 
546: \begin{equation}
547: {\rm ln} \sum_{i = 1}^{2N} {y^2} = 2 \sigma' \; \; .
548: \end{equation}
549: 
550: 
551: The method of cumulant expansion on the other hand tells us that the average of the sum of 
552: $y^2_i$ can be written as 
553: 
554: \begin{equation}
555: \left \langle \sum_{i =1}^{2N} y^2_i \right \rangle = 
556: \exp\left (\sum_m A_m \right ) \; \; \; \; \; \; \; m = 1, 2, 3, \cdots
557: \end{equation}
558:  
559: where $A_m$-s result from 
560: cumulants of the stochastic quantity $2 \sigma'$. $A_m$-s are 
561: calculated dynamically from the following relations
562: \begin{eqnarray}
563: A_1 & = & m_1  \; \; , \; \; 
564: A_2  =  \frac{1}{2!} [m_2 - m^2_1] \; \; , \; \; 
565: A_3  =  \frac{1}{3!} [m_3 - 3 m_1 m_2 + 2 m_1^3] \nonumber\\
566: A_4 & = & \frac{1}{4!} [m_4 - 3 m_2^2 - 4 m_1 m_3 + 12 m_1^2 m_2 - 6 m_1^4]  
567: \; \; {\rm etc.}
568: \end{eqnarray}
569: 
570: \noindent
571: where $m_k = \frac{2^k}{n} \sum_{j=1}^n \left ({\rm ln} \frac{d_j}{d_0}
572: \right )^k$
573: [$k = 1, 2, 3, 4, ...$]. 
574: In the spirit of Ref. \cite{tell} we enquire, whether these moments/cumulants
575: reach their steady state values in the long time limit.
576: We have numerically examined the dependence of $m_k$-s
577: on various parameters. The parameters are $n$, the time, $d_0$, the measure
578: of initial separation, the characteristic time $T$ (j-th time of iteration 
579: implies $t = j T, j = 1, 2, \cdots \infty$). Our observation is that the limit
580: $m_k$ or limit $A_m$ as $n\rightarrow \infty$ seems to exist in all cases. 
581: We have examined \cite{bb3} these limits for a number of test cases, e. g. ,
582: for Lorentz system, Henon-Heiles system and others.
583: In Fig.(1) we exhibit a typical representative long time behaviour of the 
584: cumulants $A_m (m = 1 \; {\rm to} \; 4)$ 
585: for a driven double well potential system discussed in the next section. 
586: It is apparent that they attain their long time
587: limits as $n \rightarrow \infty$. Secondly, the first two cumulants are
588: much
589: higher compared to others The first moment is the stochastic parameter 
590: defined by Casartelli et. al. \cite{tell} as a quantity closely related to Kolmogorov
591: entropy. We are therefore led to believe that the quantities $A_m$-s 
592: characterize the long time limit or the steady state of a dynamical system.
593: 
594: 
595: The relations (33) and (29) can now be combined to give
596: \begin{equation}
597: {\cal D}_s = \frac{\lambda}{\sum_{i = 1}^{2n}}\frac{1}{\alpha_i^2} 
598: \exp \left (\sum_m A_m \right ) \; \; .
599: \end{equation}
600: 
601: The above relation is the central result of this paper. This establishes a 
602: connection between the drift and the diffusion coefficients of the Fokker-Planck 
603: equation (25) through the stochastic  parameters characterizing long time behaviour
604: of the nonlinear dynamical system. It must be emphasized that both the 
605: drift $\lambda$ and the diffusion ${\cal D}_s$ coefficients arise from 
606: the deterministic stochasticity implied in the dynamical equation
607: of motion (3). The relation (35) is therefore reminiscent of the familiar
608: fluctuation-dissipation relation.
609: 
610: A few points regarding the relation (35) are in order. It is  important to 
611: note that the fluctuation-dissipation relation in conventional nonequilibrium
612: statistical mechanics is valid for a stochastic system for which the noise 
613: is internal. The spiritual root of this relation lies at the dynamic balance 
614: between the input of energy into the system from the fluctuations of the 
615: surrounding and the output of energy from the system due to its dissipation
616: into the surrounding. The system-reservoir model \cite{loui,gg} developed over the last few
617: decades suggests that the coupling between the system and the reservoir is 
618: responsible for a common origin of drift and diffusion. In the present theory 
619: this common mechanism is the fluctuations of the phase space variables (or
620: second derivative of the Hamiltonian) inherent in both the drift $\lambda$
621: and the diffusion $D_s$ coefficients of the Fokker-Planck equation. We
622: point out that the relation is still valid for the pure Hamiltonian system 
623: ($\gamma = 0$). For this reason the relation (35) is somewhat formal in
624: contrast to the standard fluctuation-dissipation relation.
625: 
626: 
627: \section{An example and numerical verification}
628: 
629: 
630: To illustrate the theory developed above, we now choose a driven
631: double-well oscillator system with Hamiltonian
632: \begin{equation}
633: H=\frac{p_1^2}{2}+aq_1^4-bq_1^2+ \epsilon q_1\cos\Omega t
634: %\label{e39}
635: \end{equation}
636: where $p_1$ and $q_1$ are the momentum and position variables of the 
637: system. $a$ and $b$ are the constants characterizing the potential.
638: $\epsilon$ includes the effect of coupling constant and the
639: driving strength of the external field with frequency $\Omega$. This model
640: has been extensively used in recent years for the study of 
641: chaotic dynamics \cite{sc1,sc2,lin}.
642: 
643: The dissipative equations of motion for the tangent space variables $X_1$ 
644: and $X_2$ corresponding to $q_1$ and $p_1$ (Eq.8) read as follows:
645: \begin{equation}
646: \frac{d}{d t} \left[
647: \begin{array}{c}
648: X_1 \\
649: X_2
650: \end{array} \right]
651: = \underline{J} \left[
652: \begin{array}{c}
653: X_1 \\
654: X_2
655: \end{array} \right] \; \; , \; \;
656: \left \{
657: \begin{array}{c}
658: \Delta q_1 = X_1 \\
659: \Delta p_1 = X_2
660: \end{array} \right \} \; \; .
661: \label{e40}
662: \end{equation}
663: where $\underline{J}$ as expressed in our earlier notation 
664: $z_1 = q_1$ and $z_2 = p_1$ is given by
665: \begin{eqnarray*}
666: \left(
667: \begin{array}{cc}
668: 0 & 1 \\
669: \zeta (t)+2b \;  & \; -\gamma
670: \end{array}
671: \right) \; \; ,
672: \end{eqnarray*}
673: where $\zeta(t)=-12 a z_1^2$. 
674: Eq.(\ref{e40}) is thus rewritten as 
675: \begin{equation}
676: \frac{d}{d t} \left(
677: \begin{array}{c}
678: X_1 \\
679: X_2
680: \end{array}
681: \right) = L^0+L^1
682: \label{e41}
683: \end{equation}
684: with
685: \begin{eqnarray*}
686: L^0=\left(
687: \begin{array}{c}
688: X_2\\
689: 2bX_1-\gamma X_2
690: \end{array}
691: \right)
692: \; \; \; \; {\rm and} \; \;
693: L^1=\left(
694: \begin{array}{c}
695: 0\\
696: \zeta (t) X_1
697: \end{array}
698: \right) \; \; ,
699: \end{eqnarray*}
700: where $L^0$ and $L^1$ are the constant and the 
701: fluctuating parts(vectors), respectively.
702: The fluctuation in $L^1$, i.e., in $\zeta(t)$ is due to stochasticity of the
703: following chaotic dissipative dynamical equations of motion;
704: \begin{equation}
705: \dot{z}_1 = z_2 \; \; {\rm and} \; \; 
706: \dot{z}_2  =  -az_1^3+2bz_1-\epsilon \cos\Omega t -\gamma z_2 \; \; .
707: \label{42}
708: \end{equation}
709: The result of Eq.(\ref{a5}) can then be applied and after some algebra
710: the Fokker-Planck equation (17) for the dissipative driven double-well
711: oscillator assumes the following form:
712: \begin{eqnarray}
713: \frac{\partial P}{\partial t} & = & -X_2 \frac{\partial P}{\partial X_1}
714: -\omega^2 X_1 \frac{\partial P}{\partial X_2} + \gamma 
715: \frac{\partial}{\partial X_2} (X_2 P) +{\cal D}_{21} \frac{\partial^2 P}
716: {\partial X_2 \partial X_1} + {\cal D}_{22} \frac{\partial^2 P}
717: {\partial X_2^2}
718: \end{eqnarray}
719: \noindent
720: where 
721: 
722: \begin{eqnarray*}
723: {\cal D}_{21}=
724: X_1^2 \int_0^\infty \langle\langle \zeta(t) \zeta(t-\tau) \rangle \rangle
725: \tau e^{-\gamma \tau}d\tau
726: \end{eqnarray*}
727: and
728: \begin{equation}
729: {\cal D}_{22}=
730: X_1^2 \int_0^\infty \langle\langle \zeta(t) \zeta(t-\tau) \rangle \rangle
731: e^{-\gamma \tau}d\tau 
732: - X_1 X_2 \int_0^\infty \langle\langle \zeta(t) \zeta(t-\tau) 
733: \rangle \rangle
734: \tau e^{-\gamma \tau}d\tau  \; \; 
735: \end{equation}
736: 
737: \noindent
738: with
739: \begin{equation}
740: \omega^2 =  2b+c+c_2 \; \; , \; \; 
741: c_2  =  \int_0^\infty \langle\langle \zeta(t)\zeta(t-\tau) \rangle\rangle  
742: \tau e^{-\gamma \tau} d\tau \; \; {\rm and} \; \;
743: c  =  \langle \zeta \rangle \; \; .
744: \end{equation}
745: 
746: \noindent
747: The similarity of the equation (40) to generalized Kramers' equation can not
748: be overlooked. This suggests a clear interplay of chaotic diffusive motion
749: and dissipation in the dynamics.
750: 
751: Using the transformation (22) Eq.(40) can be written as
752: \begin{eqnarray}
753: \frac{\partial P}{\partial \tau'} & = & -y_2 \frac{\partial P}{\partial y_1}
754: -\overline{\omega}^2 y_1 \frac{\partial P}{\partial y_2} + \overline{\gamma} 
755: \frac{\partial}{\partial y_2} (y_2 P) +
756: {\cal D'}_{21} \frac{\partial^2 P}
757: {\partial y_2 \partial y_1} + {\cal D'}_{22} \frac{\partial^2 P}
758: {\partial y_2^2} 
759: \end{eqnarray}
760: 
761: \noindent
762: where 
763: \begin{eqnarray*}
764: \overline{\omega}^2 = \frac{\omega^2}{\omega'^2} \; \; , 
765: \; \; \overline{\gamma} = \frac{\gamma}{\omega'} \; \; , \; \; 
766: {\cal D'}_{21}= \frac
767: {y_1^2 (0)}{{\omega'}^2}\int_0^\infty \langle\langle \zeta(\tau') 
768: \zeta(\tau'-\tau) \rangle \rangle
769: \tau e^{-\gamma \tau}d\tau \; \; {\rm and}
770: \end{eqnarray*}
771: 
772: \begin{equation}
773: {\cal D'}_{22}= \frac
774: {y_1^2 (0)}{{\omega'}^2} \int_0^\infty \langle\langle \zeta(\tau') 
775: \zeta(\tau'-\tau) \rangle \rangle
776: e^{-\gamma \tau}d\tau 
777: - \frac{y_1(0) y_2 (0)}{\omega'}\int_0^\infty \langle\langle \zeta(\tau') 
778: \zeta(\tau'-\tau) \rangle \rangle
779: \tau e^{-\gamma \tau}d\tau  \; \; 
780: \end{equation}
781: 
782: \noindent
783: and the time dependence of $y_1$ and $y_2$ in the diffusion coefficients 
784: have been frozen under weak noise approximation.
785: 
786: Now using the linear transformation (24) in Eq.(43) we obtain in the 
787: stationary state
788: \begin{equation}
789: \frac{\partial}{\partial U} \lambda U P_s +{\cal D}_s \frac{\partial^2 P_s}
790: {\partial U^2} = 0
791: \end{equation}
792: 
793: \noindent
794: where
795: \begin{equation}
796: U =  \alpha_1 y_1 + y_2 \; \; {\rm and} \; \;
797: \lambda U  =  - \alpha_1 y_2 -{\overline{\omega}}^2 y_1 
798: + \overline{\gamma} y_2
799: \end{equation}
800: 
801: \noindent
802: and
803: \begin{equation}
804: {\cal D}_s = {\cal D'}_{22}
805: \end{equation}
806: 
807: where for simplicity it has been assumed that ${\cal D'}_{21}$
808: is much small compared to the Markovian contribution ${\cal D'}_{22}$.
809: 
810: Comparing the coefficients of $y_1$ and $y_2$ on both sides of Eq.(46) we obtain
811: \begin{eqnarray*}
812: \lambda \alpha_1 = - {\overline{\omega}}^2
813: \; \; {\rm and} \; \; 
814: \lambda = - \alpha_1 + \overline{\gamma}
815: \end{eqnarray*}
816: 
817: \noindent
818: Therefore we have
819: \begin{equation}
820: \alpha_1 = \frac{-\overline{\gamma} - \sqrt{{\overline{\gamma}}^2 + 4 
821: {\overline{\omega}}^2}}{2} \; \; {\rm and} \; \;
822: \lambda = \frac{\overline{\gamma} + \sqrt{{\overline{\gamma}}^2 + 4 
823: {\overline{\omega}}^2}}{2} \; \; .
824: \end{equation}
825: 
826: Here the negative value of $\alpha_1$ is taken to make $\lambda$ positive
827: for a physically allowed solution of the steady state distribution (49).
828: The solution of Eq.(45) is given by
829: \begin{equation}
830: P_s  =  N \exp\left(- \frac{\lambda}{2 {\cal D}_s} ({\alpha_1}^2 y_1^2 + 
831: 2 \alpha_1 y_1 y_2 + y_2^2)\right) \; \; .
832: %N \exp\left (- \frac{\lambda U^2}{ 2 {\cal D}_s}\right ) \nonumber\\
833: \end{equation}
834: 
835: With the help of above distribution the average quantities in
836: tangent space can be calculated. Thus we have
837: \begin{equation}
838: \left \langle y_1^2 + y_2^2 \right \rangle = 
839: \frac{{\cal D}_s}{\lambda} \left ( \frac{1}
840: {\alpha_1^2} + 1 \right ) \; \; .
841: \end{equation} 
842: 
843: The fluctuation dissipation relation (35) can then be obtained by combining (50) with 
844: (33) as follows ;
845: \begin{equation}
846: {\cal D}_s = \frac{\lambda}{(\frac{1}{\alpha_1^2} +1 )} 
847: \exp \left (\sum_m A_m \right) \; \; .
848: \end{equation}
849: 
850: $\lambda$ and $\alpha_1$ are to be calculated using (48). For these we require 
851: explicit numerical evaluation of $\overline{\omega}^2$ as defined in (43) and 
852: (44). 
853: The dissipative chaotic motion is governed by equations (37) and (39). We
854: choose the following values of the parameters \cite{lin} $a = 0.5$, $b = 10$, 
855: $\epsilon = 10$,  
856: $\Omega = 6.07$ and $\gamma=0.001$. The coupling-cum-field strength $\epsilon$
857: has been varied from set to set. We choose the initial conditions $z_1(0) = -3.5$ and
858: $z_2(0) = 0$ which ensures strong global chaos. Note that $c_2$ as expressed
859: in (42) and in the diffusion coefficients are the integrals over the 
860: correlations of $\zeta(t)$ ($\zeta(t)$ is the fluctuating part of the second 
861: derivative of the potential $V(z)$ and is given by $\zeta(t) = -12 az_1^2$). 
862: To calculate the correlation function $\langle \langle\zeta(t)\zeta(t-\tau)
863: \rangle\rangle$ and the average $\langle\zeta(t)\rangle$ it is necessary 
864: to determine long time series in $\zeta(t)$ by numerically solving the classical
865: equation of motion (39). The next step is to carry out the averaging over
866: the time series.
867: For further details of the numerical procedure we refer to the earlier
868: work \cite{sc3,bb1,bb2}. On the other hand the cumulants $ A_m (m = 1, 2, 3, 4)$ (as defined in (34)
869: and (35)) are calculated from Eqs.(37) and (39) directly. The method has 
870: already been outlined in Sec.(IV) and in Ref. \cite{tell}.  We then plot the  theoretically
871: calculated values of ${\cal D}_s$ from the evaluation of $\lambda$, $\alpha_1$
872: and the cumulants for several values of the coupling constant $\epsilon$
873: (Eq.36) and compare them with the diffusion coefficients obtained from 
874: the direct numerical integration of Eqs.(39) and (37) with the appropriate 
875: transformation (22) for the corresponding values of $\epsilon$.
876: The result is shown in Fig. 2. It may be noted that the theoretical and 
877: numerical results are in good agreement. The validity of the 
878: fluctuation-dissipation relation as proposed in Eq.(35) is therefore
879: reasonably satisfactory.
880: 
881: \section{Conclusions}
882: 
883: The crucial question of instability of classical motion essentially rests on
884: the linear stability matrix or Jacobian matrix associated with the 
885: equations of motion. While the linear stability analysis
886: around the fixed points is based on the assumption of constancy of this
887: matrix we take full account of the time dependence of the quantity in the chaotic regime by 
888: considering it to be a stochastic process, since the  phase variables 
889: behave stochastically. Based on a Fokker-Planck description in the tangent space
890: we trace the origin of chaotic diffusion and drift in the correlation of 
891: fluctuations of the linear stability matrix.
892: 
893: The main conclusions of this study are the following :
894: 
895: (i) We show that a class of dynamical stochastic parameters which attain their 
896: steady state values in the long time limit of the dynamical system may be
897: used to characterize the dynamical steady state of the system. The first one
898: of them which was proposed by Casartelli et. al. \cite{tell} several years ago as a
899: measure of the chaoticity of the system is closely related to Kolmogorov
900: entropy.
901: 
902: (ii) We establish a connection between the drift and the diffusion coefficients of
903: the Fokker-Planck equation and the dynamical stochastic parameters in the spirit
904: of fluctuation-dissipation relation. The realization of this relation in chaotic 
905: dynamics therefore carries the message that although
906: comprising a few degrees of freedom a chaotic system may behave as a statistical
907: mechanical system (although in a somewhat different sense).
908: 
909: The theoretical relations proposed here are generic for N-degree-of-freedom 
910: chaotic Hamiltonian system with or without dissipation and have been verified
911: by numerical analysis of a driven nonlinear dissipative system. We hope that the present
912: approach will find useful application in searching for the related thermodynamically
913: inspired quantities in few-degree-of-freedom systems.
914: 
915: 
916: \acknowledgments
917: B. C. Bag is indebted to the Council of Scientific and
918: Industrial Research (C.S.I.R.), Government of India, for a fellowship. 
919: 
920: \newpage
921: 
922: \begin{appendix}
923: 
924: \section{The derivation of the Fokker-Planck equation}
925: 
926: We first note that the
927: operator $e (-\tau \nabla \cdot  L^0)$ provides the solution 
928: of the equation [Eq.(\ref{e13}), $\alpha = 0$]
929: \begin{equation}
930: \frac{\partial f(X, t)}{\partial t} = -\nabla_X \cdot  L^0 f(X,t)
931: \label{a1}
932: \end{equation}
933: $f$ signifies the ``unperturbed'' part of $P$ which can be found 
934: explicitly in terms of characteristic curves. The equation
935: \begin{equation}
936: \dot X = L^0 (X) 
937: \label{a2}
938: \end{equation}
939: determines for a fixed $t$  a mapping from $X(\tau=0)$ to $X(\tau)$, i. e., 
940: $X \rightarrow X^\tau$ with inverse $(X^\tau)^{-\tau}=X$ .
941: The solution of (\ref{a1}) is
942: \begin{equation}
943: f(X,t)= f(X^{-t}, 0) \left | \frac{d X^{-t}}{d X} \right | = 
944: e \left [ -t \nabla \cdot F_0  \right ]  f(X, 0 ).
945: \label{a3}
946: \end{equation}
947: $\left | \frac{d(X^{-t})}{d(X)} \right |$ being a Jacobian determinant. The
948: effect of $e(-t \nabla \cdot  L^0)$ on $f(X)$ is as 
949: \begin{equation}
950: e(-t \nabla \cdot L^0) f(X,0) = f(X^{-t}, 0) \left|\frac
951: {d X^{-t}}{d X} \right| \; \;.
952: \label{a4}
953: \end{equation}
954: 
955: This simplification in Eq.(\ref{e16}) yields
956: \begin{eqnarray}
957: \frac{\partial P}{\partial t} & = &
958: \left\{ -\nabla \cdot  L^0 
959: -\alpha \langle \nabla \cdot L^1 \rangle + \alpha^2 \int_0^\infty
960: d\tau \left| \frac{d X^{-\tau}}{dX}
961: \right| \right. \nonumber \\
962: & & \langle \langle \nabla \cdot  L^1(X,t) 
963: {\bf \nabla}_{-\tau} \cdot L^1({\bf x}^{-\tau}, t-\tau) \rangle 
964: \rangle \left. \left| \frac{dX}{d X^{-\tau}}
965: \right|  \right\} P \; \;.
966: \label{a5}
967: \end{eqnarray}
968: 
969: Now to express the Jacobian, ${X}^{-\tau}$ and $\nabla_{-\tau}$
970: in terms of $\nabla$ and $X$ we solve Eq.(\ref{a2}) 
971: for short time (this is
972: consistent with the assumption that the fluctuations are rapid \cite{van}). 
973: 
974: We now write the solution of Eq.(\ref{a2}) [using Eqs.(4-6)] as follows ;
975: 
976: \begin{equation}
977: \left(
978: \begin{array}{c}
979: X_1^{-\tau} \\
980: \vdots  \\
981: X_N^{-\tau}
982: \end{array}
983: \right)
984: = -\tau \left(
985: \begin{array}{c}
986: X_{N+1}\\
987: \vdots  \\
988: X_{2N}
989: \end{array}
990: \right)+\left(
991: \begin{array}{c}
992: X_1\\
993: \vdots  \\
994: X_N
995: \end{array}
996: \right)
997: =\left(
998: \begin{array}{c}
999: \bar{G}_1(X)\\
1000: \vdots  \\
1001: \bar{G}_N(X)
1002: \end{array}
1003: \right)
1004: \label{a6}
1005: \end{equation}
1006: and
1007: \begin{equation}
1008: \left(
1009: \begin{array}{c}
1010: X_{N+1}^{-\tau} \\
1011: \vdots  \\
1012: X_{2N}^{-\tau}
1013: \end{array}
1014: \right)
1015: = e^{\gamma \tau} \left(
1016: \begin{array}{c}
1017: X_{N+1}\\
1018: \vdots  \\
1019: X_{2N}
1020: \end{array}
1021: \right)-\tau\left(
1022: \begin{array}{c}
1023: G_{N+1}(X)\\
1024: \vdots  \\
1025: G_{2N}(X)
1026: \end{array}
1027: \right)
1028: =\left(
1029: \begin{array}{c}
1030: \bar{G}_{N+1}(X)\\
1031: \vdots  \\
1032: \bar{G}_{2N}(X)
1033: \end{array}
1034: \right)
1035: \label{a7}
1036: \end{equation}
1037: Here the terms of $O(\tau^2)$ are neglected. Since the 
1038: vector $X^{-\tau}$ is expressible as a function of $X$ we write
1039: \begin{equation}
1040: X^{-\tau} = \bar{G}(X) \; \; ,
1041: \label{a8}
1042: \end{equation}
1043: and the following simplification holds good;
1044: \begin{eqnarray}
1045: L^1(X^{-\tau}, t-\tau).\nabla_{-\tau} & = &
1046: L^1(\bar{G}(X),t-\tau) . \nabla_{-\tau} \nonumber \\
1047: & = & \sum_k L_k^1(\bar{G}(X),t-\tau)\frac{\partial}{\partial X_k^{-\tau}}
1048: \nonumber \\
1049: & = & \sum_j \sum_k L_k^1(\bar{G}(X),t-\tau)g_{jk}
1050: \frac{\partial}{\partial X_j} \; \; \; \; \; 
1051: \; \; \; ; j,k = 1 \cdots 2N
1052: \label{a9}
1053: \end{eqnarray}
1054: where 
1055: \begin{equation}
1056: g_{jk}=\frac{\partial X_j}{\partial X_k^{-\tau}}
1057: \label{a10}
1058: \end{equation}
1059: 
1060: In view of Eqs.(\ref{a6}) and (\ref{a7}) we note:
1061: \begin{eqnarray*}
1062: {\rm if} \; \; \; j=k \; \; {\rm then} \; \; g_{jk} & = & 1, \; \; k=1 \cdots N \\
1063: & = & e^{-\gamma \tau}, \; \; k=N+1 \cdots 2N\\
1064: \\
1065: {\rm if} \; \; \; j \ne k \; \; {\rm then} 
1066: \; \; g_{jk} & \propto & -\tau e^{-\gamma \tau} \; \; {\rm or} \; \;  0
1067: \end{eqnarray*}
1068: 
1069: \noindent
1070: Thus $g_{jk}$ is a function of $\tau$ only. Let 
1071: \begin{equation}
1072: R_j = \sum_k L_k^1(\bar{G}(X),t-\tau)g_{jk}
1073: \label{a11}
1074: \end{equation}
1075: From Eqs.(\ref{e8}), (\ref{e9}) and (\ref{a8}) we write
1076: \begin{equation}
1077: L_i^1(X^{-\tau},t-\tau)
1078: =L_i^1(\bar{G}(X),t-\tau) = 0 \; \; \; \; {\rm for} \; 
1079: i=1 \cdots N
1080: \label{a12}
1081: \end{equation}
1082: So the conditions (\ref{a11}), (\ref{a12}) and (\ref{a6}) imply that
1083: \begin{eqnarray}
1084: R_j(X, t-\tau) & = & R_j(X_1 \cdots X_N, t-\tau) 
1085: \; \; \; {\rm for} \; j=1 \cdots N \nonumber \\
1086: R_j(X, t-\tau) & = & R_j(X_1 \cdots X_{2N}, t-\tau) 
1087: \; \; \; {\rm for} \; j=N+1 \cdots 2N
1088: \label{a13}
1089: \end{eqnarray}
1090: We next carry out the following simplifications of 
1091: $\alpha^2$-term in Eq.(\ref{a5}).
1092: We make use of the relation (\ref{e10}) to obtain
1093: \begin{eqnarray}
1094: L^1(X,t).\nabla\sum_jR_j\frac{\partial}{\partial X_j} P(X,t)
1095: & = & \sum_i L_i^1(X,t)\frac{\partial}{\partial X_i}\sum_j R_j
1096: \frac{\partial}{\partial X_j}P(X,t)
1097: \nonumber \\
1098: & = & \sum_{i,j} L_i^1(X,t)
1099: R_j \frac{\partial^2}{\partial X_i \partial X_j} P(X,t)\nonumber \\
1100: &&+\sum_j
1101: R'_j
1102: \frac{\partial}{\partial X_j}P(X,t)
1103: \label{a14}
1104: \end{eqnarray}
1105: where
1106: \begin{equation}
1107: R'_j=\sum_i L_i^1 (X,t) \frac{\partial}{\partial X_i} R_j
1108: \label{a15}
1109: \end{equation}
1110: The conditions (\ref{a12}) and (\ref{a13}) imply that
1111: \begin{eqnarray}
1112: R'_j & = & 0 \; \; \; \; \; {\rm for} \; j=1 \cdots N \nonumber \\
1113: R'_j & = & 
1114: R'_j(X_1 \cdots X_N, t-\tau)  \ne 0 ;\; \; \; \; \; {\rm for} 
1115: \; j=N+1 \cdots 2N 
1116: \label{a16}
1117: \end{eqnarray}
1118: By (\ref{a16}) one has
1119: \begin{equation}
1120: R' . \nabla P(X,t) = \nabla . R' P(X,t)
1121: \label{a17}
1122: \end{equation}
1123: 
1124: Making use of Eqs.(\ref{e10}), (\ref{a9}), 
1125: (\ref{a14}) and (\ref{a17}) in Eq.(\ref{a5}) we obtain the
1126: Fokker-Planck equation (\ref{e17}).
1127: 
1128: \end{appendix}
1129: 
1130: \begin{thebibliography}{99}
1131: 
1132: \bibitem{cas}
1133: G. Casati, B. V. Chirikov, F. M. Izrailev and J. Ford 
1134: in {\it Stochastic Behaviour in Classical and Hamiltonian Systems},
1135: Lecture notes in Physics, Vol.~{83},
1136: edited by G. Casati and J. Ford (Springer, Berlin, 1979).
1137: 
1138: \bibitem{lic}
1139: A. J. Lichtenberg and M. Lieberman, {\it Regular and Stochastic Motion}
1140: (Springer-Verlag, New York, 1983).
1141: 
1142: \bibitem{jkb}
1143: See, for example, J. K. Bhattacharjee, {\it Statistical Physics : Equilibrium
1144: and Non-Equilibrium Aspects} (Allied Publishers Limited, New Delhi, 1997).
1145: 
1146: \bibitem{ma}
1147: S. K. Ma, {\it Statistical Mechanics} (World Scientific, Singapore, 1982).
1148: 
1149: \bibitem{gas}
1150: P. Gaspard, {\it Chaos, Scattering and Statistical Mechanics} 
1151: (Cambridge University Press, New York, 1998).
1152: 
1153: \bibitem{kai}
1154: T. Kai and Tomita, Prog. Theo. Phys. {\bf 64} 1532 (1980).
1155: 
1156: \bibitem{ono}
1157: Y. Takahasi and Y. Oono, Prog. Theo. Phys. {\bf 71} 851 (1984).
1158: 
1159: \bibitem{wid}
1160: M. Widom, D. Densimon, L. P. Kadanoff and S. J. Shenkar, J. Stat. Phys.
1161: {\bf 32} 443 (1983).
1162: 
1163: \bibitem{ber}
1164: V. L. Berdichevsky and M. V. Alberti, Phys. Rev. A {\bf 44} 858 (1991).
1165: 
1166: \bibitem{gras}
1167: P. Grassberger and I. Procaccia, Physica D {\bf 13} 34 (1984).
1168: 
1169: \bibitem{bon}
1170: M. Bianucci, L. Bonci, G. Trefan, and  B. J. West P. Grigolini,
1171: Phys. Lett. A. {\bf 174} 337 (1993).
1172: 
1173: \bibitem{bon1}
1174: M. Bianucci, R. Mannella, X. Fan, P. Grigolini and  B. J. West, 
1175: Phys. Rev. E {\bf 47} 1510 (1993).
1176: 
1177: \bibitem{zur}
1178: W. H. Zurek and J. P. Paz, Phys. Rev. Lett. {\bf 72} 2508 (1994).
1179: 
1180: \bibitem{pat}
1181: A. K. Pattanayak and P. Brumer, Phys. Rev. Lett. {\bf 79} 4131 (1997).
1182: 
1183: \bibitem{cohen}
1184: D. Cohen, Phys. Rev. Lett. {\bf 78} 2878 (1997).
1185: 
1186: \bibitem{pat1}
1187: A. K. Pattanayak, Phys. Rev. Lett. {\bf 83} 4526 (1999).
1188: 
1189: \bibitem{sc1}
1190: S. Chaudhuri, G. Gangopadhyay and D. S. Ray, Phys. Rev. E {\bf 47} 311 (1993).
1191: 
1192: \bibitem{sc2}
1193: S. Chaudhuri, G. Gangopadhyay and D. S. Ray, Phys. Rev. E {\bf 52} 2262 (1995).
1194: 
1195: \bibitem{sc3}
1196: S. Chaudhuri, G. Gangopadhyay and D. S. Ray, Phys. Rev. E {\bf 54} 2359 (1996).
1197: 
1198: \bibitem{bb1}
1199: B. C. Bag, S. Chaudhuri, J. Ray Chaudhuri and D. S. Ray, Physica D {\bf 125}
1200: 47 (1999)
1201: 
1202: \bibitem{bb2}
1203: B. C. Bag and D. S. Ray, J. Stat. Phys. {\bf 96} 271 (1999).
1204: 
1205: \bibitem{ben}
1206: G. Benettin, L. Galgani and J. Strelcyn, Phys. Rev. A {\bf 14} 2338 (1976).
1207: 
1208: \bibitem{fox}
1209: R. F. Fox and T. C. Elston, Phys. Rev. E {\bf 49} 3683 (1994).
1210: 
1211: \bibitem{tell}
1212: M. Casartelli, E. Diana, L. Galgani and A. Scotti, Phys. Rev. A {\bf 13} 
1213: 1921 (1976).
1214: 
1215: \bibitem{van}
1216: N. G. van Kampen, Phys. Rep. {\bf 24} 171 (1976).
1217: 
1218: \bibitem{bb3}
1219: B. C. Bag and D. S. Ray, Unpublished results.
1220: 
1221: \bibitem{loui}
1222: W. H. Louisell, {\it Quantum Statistical Properties of Radiation} 
1223: (Wiley, New York, 1973).
1224: 
1225: \bibitem{gg}
1226: G. Gangopadhyay and D. S. Ray in {\it Advances in Multiphoton Processes and
1227: Spectroscopy}, Vol.~{8}, edited by S. H. Lin, A. A. Villayes and F. Fujimura
1228: (World Scientific, Singapore, 1993).
1229: 
1230: \bibitem{lin}
1231: W. A. Lin and L. E. Ballentine, Phys. Rev. Lett. {\bf 65} 2927 (1990).
1232: 
1233: \end{thebibliography}
1234: 
1235: \newpage
1236: 
1237: \begin{figure}
1238: \caption{
1239: The first four dimentionless cumulants $A_1, A_2, A_3$ and $A_4$ are plotted 
1240: against dimentionless time for the
1241: dynamical system described by Eq.(39).
1242: }
1243: \end{figure}
1244: 
1245: \begin{figure}
1246: \caption{
1247: The diffusion coefficients calculated numerically (marked as dark 
1248: squares) using Eqs. (38) and (39) after transformation (22) are compared
1249: with theoretically obtained values (marked as circles) using Eq.(51) for several values of
1250: the coupling-cum-external field strength $\epsilon$ (units are arbitrary).
1251: }
1252: \end{figure}
1253: 
1254: 
1255: \end{document}
1256: