cond-mat0005297/Lf.tex
1: \documentstyle[aps,manuscript]{revtex}   
2: 
3: \begin{document}
4: 
5: \draft
6: 
7: \title{Fluctuations and the existence of potential in dissipative 
8: semiclassical systems}
9: 
10: \author{Bidhan Chandra Bag and 
11: Deb Shankar Ray{\footnote{pcdsr@mahendra.iacs.res.in}} 
12: }
13: 
14: \address{Indian Association for the Cultivation of Science,
15: Jadavpur, Calcutta 700 032, INDIA.}
16: 
17: \maketitle
18: 
19: \begin{abstract}
20: We examine the weak noise limit of an overdamped dissipative
21: system within a semiclassical description
22: and show how quantization influences the growth
23: and decay of fluctuations of the thermally equilibrated systems. 
24: We trace its origin
25: in a semiclassical counterpart of the generalized potential for the dissipative 
26: system.
27: \end{abstract} 
28: 
29: \vspace{1cm}
30: \pacs{{\bf Keywords} : Dissipative quantum systems, large fluctuations, 
31: generalized potential.\\
32: \\
33: PACS number(s): 05.40.-a, 05.70.Ln}
34: 
35: \newpage
36: 
37: 
38: \section{Introduction}
39: 
40: One of the important issues in nonequilibrium phenomena in the macroscopic
41: nonlinear systems is to understand the 
42: interplay of nonlinearity of the system and the fluctuations of its environment. 
43: The problem is fairly general in the 
44: context of chemical reactions \cite{dykman2},
45: breakdown of electronic devices \cite{kautz}, phase transitions \cite{nature}
46: etc. The essential
47: description of the physical situation rests on the Fokker-Planck equations for the
48: probability distribution functions of the relevant variables of the dynamics.
49: In the weak noise limit the fluctuations have been described 
50: \cite{nature,bray,graham,tel} by appropriate auxiliary
51: Hamiltonian or path integral methods. 
52: The theoretical results have been 
53: corroborated  by remarkable experiments on fluctuations using analogue 
54: electronic circuits \cite{nature} , which allow the phase space 
55: trajectories of fluctuations
56: to be observed directly in a precise manner. These studies have enriched our
57: understanding in several theoretical issues, e. g. , the symmetry between 
58: the growth and the decay of classical fluctuations in equilibrium 
59: and its breakdown under 
60: nonequilibrium conditions \cite{nature} , the existence of a nonequilibrium
61: potential of a dissipative 
62: system \cite{graham,tel} etc.
63: 
64: It is the purpose of this paper to extend the theory to the semiclassical
65: context. The quantization of the system itself adds a new dimension to the
66: interplay of nonlinearity and stochasticity in a dissipative 
67: system. To make a fair comparison with classical theory
68: we adopt the Wigner's phase space distribution function of c-number variables.
69: The weak noise limit can then be appropriately employed to develop an auxiliary 
70: Hamiltonian formulation at the semiclassical level 
71: in terms of these phase space variables. This allows us to realize 
72: the existence of an optimal force of purely quantum origin derivable
73: from the fluctuating field and relate it to the momentum of the auxiliary
74: Hamiltonian. The quantum correction also makes its presence felt in the growth
75: and decay of fluctuations of thermally equilibrated semiclassical 
76: systems keeping the symmetry preserved.
77: 
78: The outline of the paper is as follows: In Sec. II we introduce the general
79: aspects of dynamics of dissipative quantum system 
80: in terms of the Wigner phase space function. 
81: In Sec. III we consider the weak noise 
82: and semiclassical limit under overdamped condition and take resort to the well-known auxiliary 
83: Hamiltonian description. The quantum part of the optimal force is then 
84: explicitly derived. The symmetry between the
85: growth and decay of fluctuations in a thermalized quantum system is 
86: discussed in Sec. IV. The existence of a semiclassical contribution to the potential
87: in a dissipative system is then shown in Sec. V. The paper is concluded in
88: Sec. VI.
89: 
90: 
91: \section{Quantum dynamics of a dissipative system}
92: 
93: We consider a dynamical system characterized by a potential $V(x)$ coupled to an 
94: environment. Evolution of such an open quantum system
95: has been studied over the last several decades under a variety of reasonable 
96: assumptions \cite{louisell,legg,gg1}. Specifically interesting is the semiclassical limit of an Ohmic 
97: environment. The dissipative time evolution of the Wigner distribution
98: function $W(x, p, t)$ for the system with unit mass ($m = 1$) 
99: under the potential $V(x)$ can be described by \cite{zurek,bag1,pat}
100: 
101: \begin{equation}
102: \frac{\partial W}{\partial t} = 
103: -p \frac{\partial W}{\partial x}
104: + \frac{\partial V}{\partial x} \frac{\partial W}{\partial p} 
105: + \epsilon \sum_{n\geq 1} 
106: \frac{ \hbar^{2n}(-1)^n}{ 2^{2n}(2n+1)! }
107: \frac{\partial^{2n+ 1}V}{\partial x^{2n+ 1} }
108: \frac{ \partial^{2n+1} W }{\partial p^{2n+1} } +
109: \gamma \frac{\partial p W}{\partial p}  
110: + D \frac{\partial^2 W}{\partial p^2}  \; \;,
111: \end{equation} 
112: 
113: \noindent
114: where $\gamma$ and $D$ are the dissipation constant and the diffusion 
115: coefficient, respectively. $x$ and $p$ are c-number co-ordinate and 
116: momentum variables. The drift term is a direct consequence \cite{legg} of the 
117: existence of $\gamma$-dependent term in the imaginary part of the exponent
118: in the expression for the propagator for the density operator in the
119: Feynman-Veron theory and has been shown to be responsible for appearance of a
120: damping force in the classical equation of motion for the Brownian particle
121: to ensure quantum-classical correspondence. $\gamma$ and D are related by the
122: fluctuation-dissipation relation, $D = \frac{\gamma}{2} \hbar \omega
123: \coth{\frac{\hbar \omega}{2 k_b T}}$ (in the semiclassical 
124: limit  $D = \gamma k T$ ). $\omega$ is the renormalised linear frequency
125: of the nonlinear system.
126: The quantum correction to classical Liouville motion is contained in the
127: $\hbar$-containing terms in the sum. $\epsilon$ is a parameter (whose value is 1) which is 
128: kept in the equation for bookkeeping the Wigner correction term in our further
129: analysis. We put $\epsilon=1$ at the end of calculation. 
130: 
131: Eq.(1) had been used earlier in several occasions. For example, Zurek and Paz
132: \cite{zurek} and others \cite{pat} have studied some interesting aspects of quantum-classical
133: correspondence in relation to decoherence and chaos. Based on this equation 
134: and its variant chaotic dissipative systems has been studied.
135: \cite{bag1,mil,dit,bag2}. The equation also yields the simplest leading order
136: quantum correction term to classical Kramers' rate \cite{jr}. 
137: The primary reason for choosing Eq.(3) as our starting point is 
138: that it reaches the correct classical limit when $\hbar \rightarrow 0$ so that
139: $D$ becomes a thermal diffusion coefficient ($\gamma k T$) in the high 
140: temperature limit and the Wigner function reduces to the corresponding classical
141: phase space distribution function
142: and we recover the Kramers' equation which describes classical Brownian
143: motion of a particle in phase space.
144: 
145: 
146: \section{Weak noise and semiclassical limit of quantum dissipative dynamics:}
147: 
148: The weak noise limit of a dissipative system within 
149: a semiclassical description can
150: be conveniently described by a ``WKB-like'' ansatz (we refer to ``WKB-like'' since we are 
151: considering more than one dimension. Traditionally WKB refer to one dimension
152: only)
153: of the Eq.(1) for the Wigner function of the form
154: 
155: \begin{equation}
156: W(x, p, t) = Z(x, t) \exp(-\frac{x p}{\hbar}) \exp(-\frac{s}{D_1}) \; \;.
157: \end{equation}
158: 
159: where $D_1 = \frac{D} {\omega^2}$. The weak noise limit is defined \cite{nature}
160: as $D_1 \rightarrow 0$ and semiclassical limit refers to ($\hbar \rightarrow 0$). 
161: $Z(x,t)$ is a prefactor and $s(x, p, t)$ is a classical action which is a
162: function of c-number variables $x$ and $p$ , satisfying the following 
163: Hamilton-Jacobi equation,
164: 
165: \begin{eqnarray}
166: & & \frac{\partial s}{\partial t} + p \frac{\partial s}{\partial x} - 
167: V' \frac{\partial s}{\partial p} - \gamma p \frac{\partial s}
168: {\partial p} 
169: + \omega^2 (\frac
170: {\partial s}{\partial p})^2 \nonumber\\
171: & & + \epsilon
172: \sum_{n\geq 1} \frac{x^{2n} (-1)^{3n+1}}{2^{2n}(2n)!} \frac{\partial^{2n+1} V}
173: {\partial x^{2n+1}} \frac{\partial s}{\partial p} = 0 \; \;.
174: \end{eqnarray}
175: 
176: The derivation of Eq.(3) is based on the following consideration. Since in the
177: weak noise limit $D_1$ is the relevant small parameter one obtains with 
178: ansatz (2) in leading order a term proportional to $\hbar^{2n} \left( \frac{1}
179: {D_1} \frac{\partial s}{\partial p} \right)^{2n+1}$ which is not balanced by
180: any other term of the same order ${D_1}^{-(2n+1)}$. This is because the 
181: highest derivative in Eq.(1) does not have a factor scaling with the 
182: corresponding power of $D_1$. The successive terms next to the leading order are
183: also $\hbar$-containing terms. All of these terms vanishes in the semiclassical
184: limit ($\hbar \rightarrow 0$). Therefore the leading order term that remains gives
185: rise to Eq.(3). It is thus obvious that an ansatz (2) with $\hbar$ finite
186: is not feasible. The semiclassical limit  $\hbar \rightarrow 0$ is a necessary 
187: requirement for the validity of ansatz (2).
188: The above equation can be solved by integrating the Hamiltonian equations of
189: motion,
190: 
191: \begin{eqnarray}
192: \dot{x} & = & p \nonumber\\ 
193: \dot{X} & = & P -  \gamma X \nonumber\\
194: \dot{p} & = & V' + \gamma p - 2  \omega^2 X - \epsilon 
195: \sum_{n\geq 1}  \frac{ (-1)^{3n+1} x^{2n} }{2^{2n}(2n)! } 
196: \frac{\partial^{2n+1} V}{\partial x^{2n+1}}  \nonumber\\
197: \dot{P} & = & \left[V'' - \epsilon 
198: \sum_{n\geq 1} \frac{(-1)^{3n+1}}{2^{2n}(2n)!} 
199: \frac{\partial}{\partial x} \left( x^{2n} 
200: \frac{\partial^{2n+1} V}{\partial x^{2n+1} } \right) \right] X
201: \end{eqnarray}
202: 
203: which are derived from the following effective Hamiltonian $H_{eff}$
204: 
205: \begin{eqnarray}
206: H_{eff} & = & p P - V' X 
207: - \gamma X p 
208: + \omega^2 X^2  \nonumber\\
209: & & + \epsilon
210: \sum_{n\geq 1} \frac{x^{2n} (-1)^{3n+1}}{2^{2n}(2n)!} 
211: \frac{\partial^{2n+1} V}{\partial x^{2n+1}}  X \; \; .
212: \end{eqnarray} 
213: 
214: \noindent
215: Here we have put $\frac{\partial s}{\partial x} = P$ and
216: $\frac{\partial s}{\partial p} = X$.
217: The introduction of additional degree-of-freedom by incorporating the 
218: auxiliary momentum (P) and co-ordinate (X) makes the system an effectively 
219: two-degree-of-freedom system. The origin of these two variables is 
220: the thermal fluctuations of the environment \cite{nature}. 
221: The auxiliary Hamiltonian $H_{eff}$ is not to be 
222: confused with the microscopic Hamiltonian comprising the system, the bath and 
223: their coupling. Thus the phase space 
224: trajectories concern fluctuations of the c-number variables.
225: 
226: 
227: Under overdamped condition ($\ddot{x} \ll \gamma \dot{x} ; \ddot{X} 
228: \ll \gamma \dot{X}$) Eqs. (4) can be easily reduced to the following form ;
229: 
230: \begin{eqnarray}
231: \dot{x} & = & K(x) + \frac{2 \omega^2 X}{\gamma} \nonumber\\ 
232: \dot{X} & = & - \frac{\partial K(x)}{\partial x} X 
233: \end{eqnarray}
234: 
235: \noindent
236: where
237: \begin{equation}
238: K(x) = \frac{1}{ \gamma} \left[-V' + \epsilon 
239: \sum_{n\ge 1} \frac{(-1)^{3n+1} x^{2n}}{2^{2n}(2n)!} 
240: \frac{\partial^{2n+1} V}{\partial x^{2n+1} }  \right] 
241: \end{equation}
242: 
243: It is easy to recognize the quantity $\frac{2 \omega^2 X}{\gamma}$ as a 
244: momentum and redefine it as $p_r$. Therefore Eqs(6) may
245: be rewritten as
246: 
247: \begin{eqnarray}
248: \dot{x} & = & K(x) + p_r  \nonumber\\
249: \dot{p_r} & = & \frac{\partial K(x)}{\partial x} p_r  \; \; .
250: \end{eqnarray}
251: 
252: So the overdamped motion is described by the following effective Hamiltonian,
253: 
254: \begin{equation}
255: H_{od} = \frac{p^2_r}{2} + K(x) p_r  \; \; .
256: \end{equation}
257: 
258: The above auxiliary Hamiltonian description (8, 9) is isomorphic in form
259: to that of Luchinsky and McClintock \cite{nature}, who had considered 
260: an overdamped classical Brownian motion in a force field $K(x)$, driven by a weak 
261: white noise $\zeta(t)$ whose intensity $D_1 \ll 1$ as
262: 
263: \begin{equation}
264: \dot{x} = K(x) + \zeta(t), \; \; \; \langle \zeta \rangle = 0 \; \; ,
265: \; \langle \zeta(t) \zeta(0) \rangle = D_1 \delta(t) \; \; \; .
266: \end{equation}
267: 
268: Equivalently the corresponding Fokker-Planck equation for the probability
269: density $P(x, t)$ is
270: 
271: \begin{equation}
272: \frac{\partial P(x, t)}{\partial t} = - \frac{\partial K(x)}{\partial x} P(x, t)
273: + \frac{D_1}{2} \frac{\partial^2 P(x, t)}{\partial x^2} \; \;.
274: \end{equation}
275: 
276: 
277: The large fluctuations of scale $\gg \sqrt {D_1}$ can therefore be treated 
278: in the weak noise limit $D_1 \rightarrow 0$ \cite{nature,graham,tel}
279: by ``WKB  like'' approximation of the Fokker-Planck equation (11) in the 
280: form 
281: 
282: \begin{equation}
283: P(x, t) = y(x, t) \exp[ \frac{- \phi(x,t)}{D_1}] \; \;.
284: \end{equation}
285: 
286: Here y(x, t) is the prefactor and $\phi(x, t)$ is a ``classical'' action 
287: describing a Hamiltonian-Jacobi equation which can be solved by solving 
288: Hamilton' s equation (8) with $p_r = \frac{\partial \phi}{\partial x}$
289: as the momentum for auxiliary system.
290: 
291: The important distinguishing feature of the above description in which 
292: the system is treated semiclassically is the structure of $K(x)$ which 
293: is given by equation(7) and comprises of two terms ;
294: 
295: \begin{equation}
296: K(x) = K_{cl} + K_{semi}(x)
297: \end{equation}
298: 
299: \noindent
300: where
301: 
302: \begin{equation}
303: K_{cl} = - \frac{V'(x)}{\gamma}
304: \end{equation}
305: 
306: is derivable from purely classical potential $V(x)$ and $K_{semi}(x)$ 
307: does not 
308: explicitly involve $\hbar$,
309: 
310: \begin{equation}
311: K_{semi}(x) = \frac{\epsilon}{\gamma} 
312: \sum_{n\geq 1} \frac{x^{2n} (-1)^{3n+1}}{2^{2n}(2n)!} 
313: \frac{\partial^{2n+1} V}{\partial x^{2n+1}}  
314: \end{equation}
315: 
316: originates from the nonlinearity of the potential $V(x)$ 
317: and quantum nature of the system. The quantum
318: contribution to $K(x)$ is therefore 
319: likely  to influence both the fluctuational
320: and the relaxational paths of the dynamics and is also responsible for the
321: existence of a potential.
322: Our objective is now to explore these aspects in the following two 
323: sections.
324: 
325: \section{Large fluctuations in equilibrated semiclassical systems }
326: 
327: In the thermally equilibrated systems a typical large fluctuation of the 
328: variable $x$ implies a temporary departure from its stable state, $x_s$ to 
329: some remote  state $x_f$. This is followed by a return to $x_s$ as a result 
330: of relaxation in the absence of fluctuations $p_r$. A nonzero value of $p_r$  
331: which results from fluctuations due  
332: to surrounding drives the system away from $x_s$ along a set of trajectories
333: which form the unstable invariant manifold and define the so called 
334: fluctuational paths. On the other hand the system relaxes along the 
335: relaxational return path to $x_s$ under the condition $p_r = 0$, which form 
336: stable invariant manifold. The latter condition implies $\dot{x} = K(x)$.
337: In each case the trajectory represents the optimal paths along which the system 
338: is expected to move with overwhelming probability. Luchinsky and McClintock \cite{nature}
339: have studied these paths in analog electronic circuits and demonstrated the 
340: growth and the decay of {\it classical} fluctuations in equilibrium. We extend
341: this analysis to semiclassical domain using the same model potential,
342: 
343: \begin{equation}
344: V(x) = \frac{1}{4} x^4 - \frac{1}{2} x^2 \; \;.
345: \end{equation}
346: 
347: The quantum contribution to the growth and the decay of fluctuations can be understood
348: by recognizing the $K_{semi}(x)$ term in the dynamics (8). In Fig.1 we 
349: compare both the deterministic fluctuational and relaxational (optimal)
350: paths for quantum and classical thermally equilibrated systems. It is
351: important to note that the maximum possible amplitude of large fluctuations
352: is almost double for the quantum system compared to that for the 
353: corresponding the classical system. This is due to the addition of the 
354: nonlinear force term of quantum origin, $K_{semi}(x)$ in the 
355: c-number equation (8), which is shown to be derivable from the 
356: fluctuating field, and is related to the momentum of the auxiliary Hamiltonian. 
357: 
358: Before leaving this section we make a brief remark on the thermally 
359: nonequilibrated systems. Since the detailed balance is not operative here,
360: the optimal path to a given state not just the time-reversed dynamical path 
361: along which the system moves from this state to the stable state in absence
362: of fluctuations $p_r$. Thus for the driven system, for example, the pattern of 
363: optimal path is generically different from that for the thermally equilibrated 
364: systems. It may display singularities whose topological manifestations as 
365: caustics, switching line, cusps etc have been thoroughly studied 
366: for classical systems. We believe that $K_{semi}(x, t)$ where t 
367: signifies the driving by a periodic force in $V(x, t)$ is likely to play an 
368: important role in their quantum counterparts.
369: 
370: \section{Existence of a  potential for dissipative semiclassical system}
371: 
372: In a significant analysis Graham and coworkers \cite{graham,tel} had  examined the condition for 
373: existence of potential for classical dissipative systems. We now extend this analysis
374: to the present semiclassical context.
375: 
376: The general criterion for a dissipative dynamical system described by 
377: autonomous equations of the form
378: 
379: \begin{equation}
380: \dot{x}^{\nu} = K^\nu (x)
381: \end{equation}
382: 
383: 
384: to have a potential $\phi(x)$ with respect to $Q^{\nu \mu}$ (positive, 
385: semidefinite symmetric matrix, which are considered to be the matrix of
386: transport coefficients) if there exists a single-valued continuously
387: differentiable and globally defined function $\phi(x)$, bounded from 
388: below which is stationary in the limit sets of Eq.(17) and which satisfies
389: 
390: \begin{equation}
391: K^{\nu}(x) = -\frac{1}{2} Q^{\nu \mu} 
392: \frac{\partial \phi(x)}{\partial x^\mu}
393: + r^{\nu} (x)
394: \end{equation}
395: 
396: \noindent
397: with
398: 
399: \begin{equation}
400: r^{\nu} \frac{\partial \phi (x)}{\partial x^{\nu}} = 0 \; \;.
401: \end{equation} 
402: 
403: Here for simplicity $Q^{\nu \mu}$ is assumed to be independent of $x$. The 
404: first and the second terms of Eq.(18) correspond to irreversible and reversible
405: part, respectively.
406: 
407: 
408: The stochastic process $x (D_1, t)$ which involve Eq.(17) and a symmetric 
409: non-negative matrix $Q^{\nu \mu}$ is governed by the Fokker-Planck equation
410: for probability distribution function $P(x, t)$
411: 
412: \begin{equation}
413: \frac{\partial P(x,t)}{\partial t} = - \frac{\partial}{\partial x^{\nu}} 
414: K^{\nu}(x) P + \frac{D_1}{2} 
415: \frac{\partial^2}{\partial x^{\nu} \partial x^{\mu}}
416: Q^{\nu \mu} p \; \;.
417: \end{equation}
418: 
419: For $D_1 = 0$ the above description reduces to (17). For $D_1 \ne 0$ the 
420: steady state distribution defines the function $\phi(x, t)$ by 
421: 
422: \begin{equation}
423: P(x, D_1, t \rightarrow \infty) = N(D_1) \exp [ -\frac{\phi(x, D_1)}{D_1}]
424: \end{equation}
425: 
426: N is the normalization constant.
427: If $\phi(x) = \lim_{D_1 \rightarrow 0}  
428: \phi(x, D_1)$ is a single-valued, continuously differentiable
429: function bounded from below it satisfies
430: 
431: \begin{equation}
432: K^{\nu}(x) \frac{\partial \phi(x)}{\partial x^{\nu}}
433: + \frac{1}{2} Q^{\nu \mu} 
434: \frac{\partial \phi(x)}{\partial x^{\nu}}  
435: \frac{\partial \phi(x)}{\partial x^\mu} = 0 \; \;
436: \end{equation}
437: 
438: Eq. (22) is equivalent to Eqs. (18, 19). Interpreting Eq.(22) as usual as a
439: Hamilton-Jacobi equation by defining $\phi(x)$ as action and $\frac{\partial
440: \phi}{\partial x^{\nu}} = P_{\nu}$ 
441: as the momentum conjugate to $x_{\nu}$, one 
442: can construct the auxiliary Hamiltonian
443: 
444: \begin{equation}
445: H(x,p) = \frac{1}{2} Q^{\nu \mu} p_{\nu} p_{\mu}  + K^{\nu}(x) p_{\nu} \; \;.
446: \end{equation}
447: 
448: Graham and co-workers \cite{graham,tel} have argued that a potential can exist with equation (17) if the 
449: above Hamiltonian is integrable for $H = 0$, because the condition implies 
450: that there exist a smooth separatrix which connect smoothly the stable and unstable
451: manifolds emanating from the hyperbolic fixed points of the dynamical system.
452: 
453: We now turn back to our dissipative semiclassical system described by Eqs(8)
454: and (9) where $K(x)$ is defined by Eq.(7). Recognizing Eq.(9) as Eq.(23) 
455: for a one-degree-of-freedom system we identify 
456: 
457: \begin{eqnarray}
458: Q & = & 1 \nonumber\\
459: K(x) & = & K_{cl} + K_{semi} \; \;.
460: \end{eqnarray}
461: 
462: The potential function $\phi(x)$ can therefore be calculated from Eq.(9) 
463: with $H = 0$ as (since $p_r = \frac{\partial \phi(x)}{\partial x}$)
464: 
465: \begin{equation}
466: \phi(x) = -2 \int K(x) dx \; \;.
467: \end{equation}
468: 
469: The above expression can be made more explicit if we make use of Eq.(7) in (25).
470: We obtain
471: 
472: \begin{equation}
473: \phi(x) = \phi_{cl}(x) + \phi_{semi}(x)
474: \end{equation}
475: 
476: \noindent
477: where
478: 
479: \begin{equation}
480: \phi_{cl} = \frac{2}{\gamma} \int V'(x) dx  = \frac{2}{\gamma} V(x)
481: \end{equation}
482: 
483: \noindent
484: and
485: 
486: \begin{equation}
487: \phi_{semi}(x) = - \frac{2}{\gamma} 
488: \sum_{n\geq 1} \int \frac{x^{2n} (-1)^{3n+1}}{2^{2n}(2n)!} 
489: \frac{\partial^{2n+1} V}{\partial x^{2n+1}}  dx
490: \end{equation}
491: 
492: The existence of a potential for a {\it dissipative, semiclassical} dynamical system
493: is thus ascertained. The method essentially relies on a dynamical definition
494: $p_r$ as a derivative of the potential $\phi(x)$ in a system described by an
495: overdamped quantum Markov process in the weak noise and semiclassical limit. 
496: As elaborated 
497: earlier in Sec.III $p_r$ has a statistical origin which drives the
498: system away from its stable state $x_s$ to a preassigned remote state $x_f$
499: from which the system relaxes in absence of $p_r$. It is thus important
500: to realize that both the dynamical and statistical notions are kept
501: intact in the quantum treatment.
502: 
503: \section{Conclusions}
504: 
505: Keeping in view of the quantum nature of the system 
506: in terms of the Wigner's phase space function we examine the
507: semiclassical dynamics of a dissipative system in an Ohmic environment.
508: The weak noise limit
509: of the stochastic process then allows us to capture the essential
510: features of the dynamics within the framework of an auxiliary 
511: Hamiltonian description at the 
512: semiclassical level. Our results are summarized as follows :
513: 
514: \vspace{0.7cm}
515: \noindent
516: (i) The Wigner's quantum correction to classical Liouville equation gives rise
517: to an optimal force in addition
518: to usual the classical force term. This quantum optimal force is essentially
519: a result of an interplay of nonlinearity of the system and the thermal 
520: fluctuations of its environment and is derivable
521: in terms of an auxiliary Hamiltonian description.
522: 
523: \vspace{0.7cm}
524: \noindent
525: (ii) This term is also responsible for modification of growth and decay
526: of large fluctuations from equilibrium for the  appropriately
527: thermalized quantum system (compared to its classical counterpart).
528: The symmetry of the fluctuational and the relaxational paths, signifying 
529: the detailed balance, however, as expected is kept preserved.
530: 
531: \vspace{0.7cm}
532: \noindent
533: (iii) The quantum correction term implies the existence
534: of a potential for the dissipative semiclassical system.
535: 
536: \vspace{0.5cm}
537: Since the fluctuational and the relaxational paths have been experimentally
538: demonstrated as a part of physical reality by analogue experiments \cite{nature}
539: in the realm of large fluctuations, we believe that the essential 
540: modification of the integrable, nonintegrable and the singular
541: topological features of the dynamics due to  
542: semiclassical correction might be
543: relevant in several contexts. We hope to address some of these issues in a
544: future communication.
545: 
546: \acknowledgments
547: B. C. Bag is indebted to the Council of Scientific and
548: Industrial Research for a fellowship. 
549: 
550: 
551: \begin{thebibliography}{99}
552: 
553: \bibitem{dykman2} M. I. Dykman, E. Mori, J. Ross and M. P. Hunt, J. Chem. Phys. {\bf 100},
554:  5737 (1994).
555: 
556: \bibitem{kautz} 
557: R. L. Kautz, Rep. Prog. Phys. {\bf 59}, 935 (1996).  
558: 
559: \bibitem{nature} 
560: D. G. Luchinsky and P. V. E. McClintock, Nature {\bf 389}, 403 (1997).
561: 
562: \bibitem{bray} 
563: A. J. Bray and A. J. McKane, Phys. Rev. Lett. {\bf 62}, 493 (1989).
564: 
565: \bibitem{graham} 
566: R. Graham, in {\it Noise in Nonlinear Dynamical Systems} Vol. 1 (eds F. Moss
567: and P. V. E. McClintock) 225-278 (Cambridge Univ. Press, 1989).
568: 
569: \bibitem{tel} R. Graham and T. T\'el, Phys. Rev. Lett {\bf 52}, 9 (1984).
570: 
571: \bibitem{louisell} 
572: W. H. Louisell, {\it Quantum Statistical Properties of Radiation} (Wiley, New
573: York, 1973); G. W. Ford, J. T. Lewis and R. J. O'Connell, Ann. Phys.(N.Y.) 
574: {\bf 185}, 270 (1988).
575: 
576: \bibitem{legg}
577: A. J. Leggett and A. O. Caldeira, Physica {\bf A121}, 587 (1983).
578: 
579: \bibitem{gg1} 
580: See, for example, the review by
581: G. Gangopadhay and D. S. Ray, in {\it Advances in Multiphoton Processes and
582: Spectroscopy} Vol 8, edited by S. H. Lin, A. A. Villayes and F. Fujimura,
583: (World Scientific, Singapore, 1993).
584: 
585: \bibitem{zurek} 
586: W. H. Zurek and J. P. Paz, Phys. Rev. Lett. {\bf 72}, 2508 (1994) ;
587: S. Habib, K. Shizume and W. H. Zurek, Phys. Rev. Lett. {\bf 80}, 4361 (1998).
588: 
589: \bibitem{bag1} 
590: B. C. Bag, S. Chaudhuri, J. Ray Chaudhuri and D. S. Ray, Physica D {\bf 125}, 
591: 47 (1999).
592: 
593: \bibitem{pat} A. K. Pattanayak, Phys. Rev. Lett. {\bf 83}, 4526 (1999).
594: 
595: \bibitem{mil} M. E. Goggin, B. Sundaram and P. W. Milonni, Phys. Rev. A {\bf 41},
596: 5705 (1990).
597: 
598: \bibitem{dit} T. Dittrich and R. Graham, Ann. Phys. {\bf 200}, 363 (1990) 
599: 
600: \bibitem{bag2} B. C. Bag and D. S. Ray, J. Stat. Phys. {\bf 96}, 271 (1999). 
601: 
602: \bibitem{jr} J. Ray Chaudhuri, B. C. Bag and D. S. Ray, J. Chem. Phys. 
603: {\bf 111}, 10852 (1999).
604: 
605: \end{thebibliography}
606: 
607: 
608: \begin{figure}
609: \caption{ A plot of system co-ordinate $x$ vs time $t$ signifying the
610: fluctuational and relaxational paths according to Eq.(8) for the
611: model system with $V(x) = \frac{1}{4} x^4 - \frac{1}{2} x^2$ for $(a)$ 
612: semiclassical case ($\epsilon = 1$) and $(b)$ classical case ($\epsilon = 0$), 
613: (units arbitrary).}
614: \end{figure}
615: 
616: 
617: \end{document}
618: 
619: