cond-mat0008360/xxx.tex
1: %\documentclass{epl}
2: \documentstyle[pre,aps,tighten,epsf,floats]{revtex}
3: \begin{document}
4: %\usepackage{graphics}     
5: %\usepackage{psfrag}
6: \title{Multiple current reversals in forced inhomogeneous ratchets} 
7: \author{\large Debasis Dan,$^{1,*}$ Mangal C. Mahato,$^{2}$ andA. M. Jayannavar$^{1,\dagger }$ \vspace{0.2in}}\address{\large $^1$ Institute of Physics, Sachivalaya Marg,
8: Bhubaneswar 751005, India}
9: \address{\large $^2$ Department of Physics, Guru Ghasidas University, Bilaspur 495009, India}\maketitle
10: \maketitle
11: \begin{abstract}
12:   
13:   Transport properties of overdamped Brownian paricles in a rocked 
14: thermal ratchet with space dependent
15: friction coefficient is studied. By tuning the parameters, the
16: direction of current exhibit multiple reversals, both as a function
17: of the thermal noise strength as well as the amplitude of rocking
18: force. Current reversals also occur under
19: deterministic conditions and exhibits intriguing structure. All these
20: results arise due to mutual interplay between potential asymmetry,
21: noise, driving frequency and inhomogeneous friction.
22: 
23: \end{abstract}
24: \vspace{1.0in}
25: %\newpage 
26:       Fluctuation induced transport in ratchet systems has been an 
27: active field of research over the last decade. In these systems in the
28: absence of any net microscopic forces, asymmetric potential can be 
29: used to induce a unidirectional particle flow when subjected to an external
30: nonthermal fluctuations. These studies have been motivated in 
31: part by the attempt
32: to understand the mechanism of movement of protein motors in biological 
33: systems~\cite{Motor}. To this effect several physical  models have been proposed under the 
34: name of rocking ratchets~\cite{Rock,Rock2,Main}, flashing
35: ratchets~\cite{Flash,Flash2},
36: diffusion ratchets~\cite{Diffu}, correlation
37: ratchets~\cite{Corr}, etc. In all these studies the potential is taken 
38: to be asymmetric
39: in space. It has also been shown that one can  obtain  unidirectional 
40: current 
41: in the presence of spatially symmetric potentials. For these nonequilibrium 
42: systems external random force should be time asymmetric~\cite{Ma}  or the presence
43: of space dependent mobility is
44: required~\cite{IntMod,Mill,Jay,But,Kam,Spc_mob,Dan3}. By suitably tuning the system 
45: parameters such as temperature, friction coefficient, mass, etc, one can 
46: even change the direction of the current. Indeed, the study of current 
47: reversal phenomena has given rise to a research activity on its own. 
48: The motivation  being the possibility of new particle separation
49: devices superior to existing methods  such as electrophoretic
50: method for particles of micrometer scale~\cite{Kep}. 
51: 
52:       Bartussek \textit{et. al.}~\cite{Main} showed the occurrence of
53: current reversal in
54: rocked thermal ratchet with both amplitude of rocking force as well as 
55: the temperature of thermal bath.  They attributed this current reversal to the 
56: ``mutual interplay between noise and finite-frequency driving''.
57: Multiple current reversals
58: have also been shown in the deterministic limit of these ratchets when
59: the inertial term is taken into account~\cite{Rock2,Under}. However 
60: these multiple current reversals in inertial ratchets are not robust 
61: in the presence of noise. Beside rocking ratchets 
62: current reversals have also been observed in flashing
63: ratchets~\cite{Reim,Bier,Flash2}. We in this work report that
64: that multiple reversals can be achieved even in  rocked \textit{overdamped} ratchet
65: in the presence of space dependent mobility, as a function of the noise strength
66: and amplitude of the rocking force. More over these systems 
67: show current reversals when rocked adiabatically.
68: In the deterministic overdamped case we get current reversal as a function 
69: of the amplitude of rocking force. Most of our results are attributed to the 
70: presence of space dependent mobility.
71: We have studied the same system as due to Bartussek \textit{et
72:   al}~\cite{Main}, except the presence 
73: of space dependent friction term. In earlier work the spatial asymmetry of 
74: the potential is responsible for unidirectional currents and their reversal 
75: as function of frequency.  As mentioned previouly one can get  
76: unidirectional currents in the presence of symmetric potentials, but
77: for this space dependent friction is
78: required~\cite{IntMod,Mill,Jay,But,Kam,Spc_mob,Dan3}. In these systems transport
79: direction is given by the phase shift between mobility and periodic
80: potential. This phase shift induces spatial symmetry breaking as required for 
81: the directed motion.
82: Appropriately choosing the phase shift leads to current reversal. We would like to 
83: emphasize that space dependent friction does not alter the equilibrium 
84: property of the system, however when the system is driven out of equilibrium 
85: not trivial dynamical effects arise due to space dependent friction~\cite{Dan3,PLA,PRE}. Naturally
86: in our present system we expect additional effects arising due to combination 
87: of spatial asymmetry and position dependence of friction coefficient.
88: It is to be noted that systems with space dependent friction are not uncommon.
89: Brownian motion in confined geometries show space dependent friction~\cite{Conf}. 
90: Particles diffusing close to surface have space dependent 
91: friction coefficient~\cite{Conf,Surf}. It is believed that molecular motor 
92: proteins move 
93: close along the periodic structure of microtubules and will therefore 
94: experience a position dependent mobility~\cite{Spc_mob}. Frictional inhomogeneities are 
95: common in super lattice structures and Josephson
96: junctions~\cite{Falco} also.
97: 
98:           We consider an overdamped Brownian particle moving in a
99: asymmetric potential V(x) with space dependent friction coefficient 
100: $\eta$(x) under the influence of external force field $F$(t) at
101: temperature $T$. 
102: Throughout our analysis we take the ratchet potential
103: V(x) = $-\frac{1}{2 \pi} (\sin (2\pi x) + \frac{\mu}{4} \sin (4 \pi
104: x))$. Here $\mu$ is the asymmetry parameter with values taken in the
105: range $0 <\mu <1 $,  friction coefficient $\eta (x) = \eta_{0}(1-\lambda \sin
106: (2\pi x + \phi))$, $|\lambda| < 1$. $\phi$ determines the relative
107: phase shift between  friction coefficient and potential . 
108: The forcing term is 
109: taken to be F(t) = A$\sin (w t + \theta)$, ($w = \frac{2
110:   \pi}{\tau}$, where $\tau$ is period of force) . Without any 
111: loss of  generality $\theta$ is taken to be zero. 
112:  The correct Langevin equation for this system in the overdamped limit
113: is given by~\cite{IntMod,Pram,Sancho} 
114: 
115: %
116: \begin{equation}
117:  \dot{x} =  -\frac{(V'(x) - F(t))}{\eta (x)} - k_{B}T \frac{\eta '(x)}{(\eta 
118:    (x))^{2}} + \sqrt{\frac{k_{B}T}{\eta (x)}}\xi (t) ,
119: \label{Langvn}
120: \end{equation} 
121: %
122: where $\xi(t)$ is a zero mean thermal Gaussian noise with correlation 
123: $<\xi (t) \xi (t')> = 2 \delta (t-t')$
124:   The equation of motion is equivalently described by the Fokker
125: Planck Equation ( FPE)~\cite{IntMod} 
126: 
127: %
128: \begin{equation}
129:  \frac{\partial P(x,t)}{\partial t} = \frac{\partial}{\partial x}
130:  \frac{1}{\eta (x)} [k_{B}T \frac{\partial}{\partial x} + (V'(x) -
131:  F(t))] P(x,t) .
132:  \label{FPE}
133: \end{equation} 
134: %
135: where $P(x,t)$ is the probability density at position $x$ at time
136: $t$. Equation~(\ref{FPE}), in the form of a continuity equation 
137: 
138: \begin{equation}
139:  \frac{\partial P(x,t)}{\partial t} = - \frac{\partial
140:    J(x,t)}{\partial x} ,
141:  \label{Conti}
142: \end{equation}
143: 
144: where 
145: \begin{equation}
146:   J(x,t) = -\frac{1}{\eta(x)} [(V'(x)-F(t)) + k_{B}T 
147:   \frac{\partial}{\partial x}]P(x,t),
148:  \label{Current}
149: \end{equation}
150: is the probability current. Since the potential and the driving force
151: have spatial and temporal periodicity respectively, therefore $J(x,t)
152: = J(x + 1,t + \tau)$, ~\cite{Risk,Main}. The average current $j$
153: in the system is given by 
154: \begin{equation}
155:  j = \lim_{t \rightarrow \infty} \frac{1}{\tau} \int_{t}^{t+\tau} dt
156:  \int_{0}^{1} J(x,t) dx .
157:  \label{Avg_curr}
158: \end{equation}
159: %
160: It should be noted that for symmetric potential and $\lambda = 0$, $j
161: = 0$. Rectification of current is possible if the potential is either
162: asymmetric or $\lambda \neq 0$ with $\phi \neq 0, \pi$~\cite{Dan3}. $j$ is 
163: independent of the initial phase $\theta$ of the driving force.
164: 
165:  We explore various parameter regimes of the problem 
166: extensively by solving the FPE, Eq.~(\ref{FPE}) numerically with 
167: finite difference method. 
168: Throughout we have set current $j$ and all other physical quantities 
169: such as $T$, $A$, $w$ in dimensionless form .
170: 
171:  In the Fig.~\ref{j-T}, the average current $j$ is plotted 
172: \textit{vs} $T$ for various values of $w$. Here $\lambda = 0.1 , \phi =
173: 0.2 \pi$, $A = 0.5$ and $\mu = 1$. Unlike Bartussek et al. ~\cite{Main}, 
174: where they show that
175: current reversal is not possible under adiabatic conditions,
176: however current reversal even in adiabatic condition can be obtained
177: in the present case for
178: some values of $\phi$~\cite{Dan3}. For moderately high frequency $w = 4.0$ 
179: the current reverses its sign \textit{twice} at $T=0.1$ and
180: $T=0.22$ and asymptotically goes to zero for higher values of $T$ as shown 
181: in the figure. This phenomena of twice current reversal with
182: temperature $T$ is the foremost feature of our system, previously    
183: unseen in any overdamped system. 
184: The inset shows the zero contour plot 
185: of the current $j$ versus $T$ and $\phi$ for three values of
186: $w$. Crossing this zero contour line implies current reversal. 
187: It can be seen that twice current reversal occurs only for a very narrow
188: range of $\phi$. For frequencies higher than certain 
189: critical frequency $w_{c}(\phi)$ current flows in only one direction for 
190: all temperatures, i.e no current reversal occurs. The plot
191: $w_{c}$ \textit{vs} $\phi$ is shown in Fig.~\ref{wc_phi}. 
192: The adiabatic curve  $w < 1$ is not shown in the figure as it
193: goes much beyond the scale of the graph. However it has a similar qualitative
194: shape as for $w = 3.0$ curve. As mentioned before, currents are 
195: due to the combined effect of phase shift $\phi$ coming from space dependent
196: friction and asymmetry parameter $\mu$. In the regime $\phi = 0.2 \pi$
197: and $A = 0.5 $ and in the absence of asymmetry current flows in the negative
198: direction for all values of $T$. The asymmetric case ( $\mu = 1.0$) in the
199: absence of space dependent friction give current in the positive direction
200: only as a function of temperature. Separately in both these cases 
201: absolute value of current
202: exhibits a maxima as a function of $T$, reminiscent of stochastic 
203: resonance phenomena. In a purely asymmetric case ($\lambda = 0$) current 
204: vanishes rapidly when $T$ exceeds the temperature associated with the 
205: barrier height. Whereas, in the symmetric case due to space dependent friction
206: absolute values of currents are significantly higher and decay slowly to 
207: zero in the large temperature regime. Naturally in the presence of 
208: both asymmetry and space dependent friction for the case under study 
209: the low temperature regime is dominated by the effect of asymmetry while the 
210: high temperature regime is dominated by space dependent friction. From this
211: , one can qualitatively explain the current reversals from positive 
212: to negative side as a function of temperature even in the adiabatic limit. 
213: In the regime $\phi > \pi$ current due to space dependent friction and 
214: potential asymmetry flows in same direction and hence no reversal is possible
215: as a function of $T$.  For
216: frequencies higher than the interwell frequency $w_{0}$, 
217: the low temperature scenario
218: changes. The direction of current in this regime is more of a
219: interplay between potential asymmetry and $w$ than that of
220: $\lambda$. Due to higher frequency the Brownian particles do not get
221: enough time to cross the right barrier which is at a larger distance 
222: from the minima. Number of particles moving about the 
223: potential minima increases. This fact is amply reflected in Fig.~\ref{P_av}, 
224: where the time averaged probability curve 
225: $ P_{av}(x) = \frac{1}{\tau} \int_{0}^{\tau} P(x,t) dt$
226: is plotted as function of $x$ for various values of
227: $w$ with $T = 0.05$. It is to be noted that the distribution is
228: independent of $\eta (x)$. Figure~\ref{P_av} shows that the 
229: probability of finding the
230: Brownian particles near the minima of the potential well increases with
231: increasing frequency, consequently the probable number of particles
232: near the potential barrier decreases. Since the distance from a
233: potential minima to the basin of attraction of next minima is less
234: from the steeper side than from the slanted side, hence in one period
235: the particles get enough time to climb the potential barrier from the
236: steeper side than from the slanted side, resulting in a negative
237: current. On increasing  the temperature , the particles get kicks of
238: larger intensity and hence they easily cross the slanted barrier,
239: resulting in a current reversal and positive current. On further
240: increasing the temperature, the effect of $\lambda$ dominates as
241: mentioned previously  and as a result the current
242: again flows in the negative direction, implying a second current
243: reversal as shown in dotted curve of Fig~\ref{j-T}. It is obvious from 
244: the above argument that for higher frequency the first reversal
245: temperature will be higher as shown in the Fig.~\ref{j-T-w}. The
246: dotted line in the base of Fig.~\ref{j-T-w} shows the contour of zero
247: current. But as mentioned previously, since the effect of $\lambda$ dominates 
248: for higher temperature, therefore the second reversal temperature
249: decreases with increasing frequency. Beyond the critical frequency $w_{c}$
250: there is no reversal of current as shown in
251: Fig.(~\ref{j-T-w},~\ref{wc_phi}). In the high frequency regime 
252: the effect of space dependent friction dominates the nature of current. 
253: We would like to emphasize here that in absence of asymmetry current reversals does
254: not take place. 
255: 
256: Multiple current reversals can also be seen when the amplitude ( $A$) of the
257: forcing term is varied in a suitable parameter regime of our system.
258: In Fig.~\ref{j-A}, the plot of $j$ versus $A$ is shown for
259: different values of $w$, keeping $\lambda$, $\phi$ and $T$ fixed at
260: 0.1, $0.88\pi$ and $0.05$ respectively. For $w = 4.0$ curve, we can see as
261: many as four current reversals. For very large value of $A$, the current
262: asymptotically goes to a constant value depending on the value of
263: $\phi$, as was previously shown for the adiabatic
264: case~\cite{Dan3}. This value was shown to be $-\frac{\lambda}{2} \sin
265: (\phi)$.  This is special to the space dependent friction
266: ratchet where the currents saturate to a finite value in the large $A$ 
267: limit. In the absence of space dependent friction it is to be
268: noted that currents decay to zero in the same asymptotic
269: regime. As the asymptotic value depends on the phase $\phi$, so we can 
270: choose it appropriately to make the asymptotic current positive or
271: negative. In the present case $\phi$ is chosen such that the
272: asymptotic current is negative which guarantees at least one current
273: reversal irrespective of frequency. The oscillatory behavior in the $j-A$
274: characteristics is the reminiscent of the deterministic dynamics
275: \cite{Main,Lich} which will be discussed later. The inset in Fig.~\ref{j-A} shows the
276: zero contour of $j$ versus $\phi$ and $A$ for $w = 4.0$. For $\phi > \pi$
277: only two current reversals can be seen. For $\pi > \phi > \pi - \epsilon$ 
278: (where $\epsilon$ is a small number ) or $0 < \phi < \epsilon$, four 
279: or more current reversals can occur. The value of $\epsilon$ depends 
280: critically on $T$, large for small $T$ and vice-versa. It is also to be noted
281: that we have even number of reversals for finite frequency driving. 
282: These oscillatory features along with their associated current reversals 
283: disappear in the high temperature regime as expected. 
284: Figure~\ref{Deter} is for $w = 0.25$ and shows that
285: current reversal is also possible
286: in the deterministic regime of the overdamped system. This deterministic
287: reversal of current cannot be attributed to the chaotic motion of the
288: system like in an forced underdamped oscillator~\cite{Under}. It solely arises due the 
289: presence of space dependent friction. As shown in Ref.~\cite{Main} here too 
290: the deterministic current shows  quantization and phase locking
291: behaviour. However all these complex features are not robust in  the 
292: presence of noise as discussed in earlier literature.
293: 
294:         In conclusion, we have studied the transport properties of
295: overdamped Brownian particles moving in an asymmetric potential 
296:  with space dependent 
297: friction coefficient and rocked by periodic force. We observe several novel
298: and complex features arising due to the interplay between asymmetry and 
299: inhomogeneous friction.
300: Currents in the low temperature regime is mostly influenced by the asymmetry
301: of the potential. At higher temperatures it is controlled by the modulation
302: parameter $\lambda$ of the friction coefficient. We find current reversal
303: with temperature even when the forcing is adiabatic. In the presence of 
304: finite frequency, twice current reversals are seen. As function of
305: amplitude of the forcing term we observe multiple current reversals.
306: Current even
307: reverses its sign in the adiabatic deterministic regime. 
308: All the above results can be understood in a qualitative manner.
309: We expect that our analysis should be applicable for the motion of particle
310: in porous media and for molecular motors where space dependent friction can
311: arise due to the confinement of particles.
312: 
313:    MCM acknowledges partial financial support and hospitality
314: from the Institute of Physics, Bhubaneswar. MCM and AMJ acknowledgepartial financial support from the Board of Research in Nuclear
315: Sciences, DAE, India.
316: 
317: %
318: % FIG 1
319: %
320: %\begin{figure}
321: %\centerline{
322: %\psfig{file=fig1.ps,width=1.8in,%
323: %bbllx=109pt,bblly=107pt,bburx=529pt,bbury=934pt}}
324: %\vspace{0.1in}
325: %\setlength{\columnwidth}{3.2in}
326: %\centerline{\caption{
327: %$I$-$V$ characteristic given by the PT with $\gamma \approx 2$
328: %(solid lines). The dashed line is a conjecture.
329: %\label{Fig_I_V}
330: %}}
331: %\end{figure}
332: % 
333: 
334: 
335: 
336: \begin{thebibliography}{0}
337: 
338:  \bibitem{Motor}
339:    S. Leibler, Nature {\bf370},(1994),412;
340:    J. Maddox, ibid {\bf365}(1993),203;
341:    ibid,{\bf368},(1994),287;
342:    ibid,{\bf368},(1994),287.
343: 
344:  \bibitem{Rock}
345:    M.O. Magnasco, Phys. Rev. Lett. {\bf 71}, 1477, (1993);
346:    I. Derenyi and T. Vicsek, Phys. Rev. Lett. {\bf75},(1995),374.
347: 
348:   \bibitem{Rock2}
349:    P. Jung, J. G. Kissner and P. H\"{a}nggi, Phys. Rev. Lett,
350:    {\bf76}(1996),3436.
351: 
352:  \bibitem{Main}  R. Bartussek, P. Hanggi, and J.G. Kissner, Europhys. Lett.
353:  {\bf 28}, 459, (1994).
354: 
355: 
356:  \bibitem{Flash}
357:    J. Prost et al, Phys. Rev. Lett. {\bf72},(1994),2652;
358:    J. F. Chauwin, A. Ajdari and J. Prost,
359:    Europhys. Lett. {\bf27},(1994), 421;
360:    J. Rousselet et al, Nature, {\bf370}, (1994), 446;
361:    C. Van den Broeck et al, Lecture Notes in Physics:
362:    ``\bf{Statistical Mechanics of Biocomplexity}'', Vol{527},
363:    (Springer-Verlag Berlin, Heidelberg, New York), (1999), Page 93;
364:    P. Hanggi and R. Bartussek, \emph{Nonlinear Physics of Complex
365:      Systems - Current Status and Future Trends}, Lecture Notes in Physics,
366:    Vol. 476, ed. by J. Parisi, S.C. Mueller, and W. Zimmermann (Springer,
367:    Berlin, 1996), pp. 294-308.
368: 
369:   \bibitem{Flash2}
370:     A. Ajdari and J. Prost, Europhys. Lett. {\bf32}, (1995), 373.
371:     
372: 
373:  \bibitem{Diffu}
374:    P. Reimann, R. Bartussek, R. H\"{a}uss ler and P. H\"{a}nggi,
375:    Phys. Lett. A, {\bf215}, (1996), 26.
376:    
377:   \bibitem{Corr} C.R. Doering, W. Horsthemke, and J. Riordan, Phys. Rev.
378:     Lett. {\bf 72}, 2984, (1994).
379: 
380:   \bibitem{Ma}A. Ajdari, D. Mukamel, L. Peliti, and J. Prost,
381:  J. Phys. (Paris) {\bf 14}, 1551, (1994);
382:  M.C. Mahato, and A.M. Jayannavar, Phys. Lett.
383:  A{\bf 209}, 21, (1995);
384:  D.R. Chialvo, and M.M. Millonas, Phys. Lett.
385:  A{\bf 209}, 26, (1995).
386: 
387: 
388:   \bibitem{Mill} M. M. Millonas, Phys. Rev. Lett {\bf 74}, 10 (1995).
389: 
390: 
391:   \bibitem{Jay}  A. M. Jayannavar, Phys. Rev. E {\bf 53}, 2957 (1996).
392: 
393:   \bibitem{IntMod} M. C. Mahato, T. P. Pareek and A. M. Jayannavar, Int. J. Mod. Phys
394:     B {\bf 10}, 3857 (1996). Cond-Mat/9603103.
395: 
396:   \bibitem{But} M. B\"{u}ttiker, Z. Phys. B {\bf 68}, 161 (1987).
397: 
398:   \bibitem{Kam} N. G. van Kampen, IBM. J. Res. Develop {\bf 32}, 107 (1988).
399: 
400:   \bibitem{Spc_mob} Rolf H. Luchsinger, Phys. Rev. E, {\bf62}, (2000) 272.
401: 
402:   \bibitem{Dan3} D. Dan, M. C. Mahato and A. M. Jayannavar,
403:     Int. J. Mod. Phys. B, (in press);
404:     Cond-Mat, (0006244)
405: 
406:   \bibitem{Kep} C. Kettner et al,
407:     Phys. Rev. E, {\bf61}, (2000), 312.
408: 
409:   \bibitem{PLA} D. Dan, M. C. Mahato and A. M. Jayannavar,
410:   Phys. Lett. A {\bf 258}, 217 (1999).
411: 
412:   \bibitem{PRE} D. Dan, M. C. Mahato and A. M. Jayannavar, Phys. Rev. E
413:   {\bf 60}, 6421, (1999).
414: 
415:   \bibitem{Under} Jose L. Mateos,
416:     Phys. Rev. Lett, {\bf84}, (2000), 258.
417: 
418:   \bibitem{Reim}
419:     P. Reimann
420:     Phys. Rep., {\bf290}, (1997), 149.
421: 
422:   \bibitem{Bier}
423:     M. Bier and R. D. Astumian
424:     Phys. Rev. Lett., {\bf32}, (1995), 373.
425: 
426:   \bibitem{Conf}
427:     Luc P. Faucheux and A. J. Libchaberm,
428:     Phy. Rev. E, {\bf49}, (1994), 5158. 
429: 
430:   \bibitem{Surf} H. Brenner,
431:     Chem. Eng. Sc., {\bf16}, (1962), 242.
432: 
433:   \bibitem{Falco} C. M. Falco,
434:     Am. J. Phys., {\bf44}, (1976), 733. 
435: 
436:   \bibitem{Pram} A. M. Jayannavar and M. C. Mahato, Pramana- J. Phys {\bf 45},
437:     369 (1995).
438: 
439: 
440:   \bibitem{Sancho} J.M. Sancho, M. San Miguel, and D. Duerr, J. Stat. Phys.
441:  {\bf 28}, 291, (1982).
442:     
443:   \bibitem{Risk}  H. Risken, {\em The Fokker Planck Equation}
444:     (Springer Verlag, Berlin
445:     , 1984).
446: 
447: 
448:  \bibitem{Lich} T. E. Dialynas, K. Lindenberg and G. P. Tsironis,
449:    Phys. Rev. E, {\bf56}, (1997), 3976.
450: 
451: %  \bibitem{PLA}
452: %    \Name{D. Dan, M. C. Mahato \and A. M. Jayannavar}
453: %    \REVIEW{Phys. Lett. A}{258}{1999}{217}.
454: 
455: %  \bibitem{PRE}
456: %    \Name{D. Dan, M. C. Mahato \and A. M. Jayannavar}
457: %    \REVIEW{Phys. Rev. E}{60}{1999}{6421}.
458: 
459: \end{thebibliography}
460: 
461: \newpage
462: 
463: \begin{figure}  
464: \centerline{\epsfbox{fig1.eps}}
465: \vspace{.0in}
466: \caption{The mean current $j$ \textit{vs} temperature $T$ for $\phi =
467:   0.2 \pi$, $A=0.5$ and $\lambda = 0.1$. The driving frequencies are $w=
468:   3.0, 4.0$ and $5.0$. The inset shows the contour of zero current for 
469:   same values of $T, \lambda$. Regions enclosed on the right hand side 
470:   of a given contour is the negative current region and vice
471:   versa. Note that we have only one reversal for all values of $\phi > 
472:   \pi$.}
473: \label{j-T}
474: \end{figure}
475: 
476: \begin{figure}  
477: \centerline{\epsfbox{w_c.eps}}
478: \caption{The critical $w$ above which no current reversal with
479:   $T$ occurs \textit{vs} $\phi$ for $A=0.5$ and $\lambda=0.1$.}
480: \label{wc_phi}
481: \end{figure}
482: 
483: \begin{figure}  
484: \centerline{\epsfbox{prob.eps}}
485: \vspace{.0in}
486: \caption{The time averaged probability $P_{av}$ \textit{vs} x for
487:   various values of $w$ and $T = 0.05$ and $A=0.5$. The
488:   distribution is independent of $\phi$ and $\lambda$.}
489: \label{P_av}
490: \end{figure}
491: 
492: \begin{figure} 
493: \centerline{\epsfbox{3D.ps}} 
494: \vspace{-1.1cm}
495: \caption{The mean current $j$ \textit{vs} $w$ and $T$ for $A = 0.5, 
496:   \phi=0.2 \pi$ and $\lambda = 0.1$. The contour on the base shows the 
497:   zero current.}
498: \label{j-T-w}
499: \end{figure}
500: 
501: 
502: \begin{figure}
503: \centerline{\epsfbox{fig22.eps}}
504: \vspace{.0in}
505: \caption{The mean current $j$ with amplitude $A$ of the forcing term
506:   for $\phi = 0.88 \pi,T=0.05$ and $\lambda=0.1$ with $w=3.0,
507:   4.0, 5.0$. The inset shows the contour of zero current for $T=0.05$
508:   for $w = 4.0$. The dots in the inset shows the four values of
509:   $A$ for $\phi = 0.9 \pi$ where the current reversal occurs. }
510: \label{j-A}
511: \end{figure}
512: 
513: \begin{figure}
514: \centerline{\epsfbox{deter.ps}}
515: \vspace{.0in}
516: \caption{The deterministic current $j$ \textit{vs} amplitude of
517:   forcing $A$ for $w=0.25$ and $\phi=0.2 \pi$ and $\lambda = 0.1$.}
518: \label{Deter}
519: \end{figure}
520: 
521: %\begin{figure}
522: %  \onefigure{3D.eps}
523: %  \vspace{0.0in}
524: %  \caption{}
525: %  \label{Deter}  
526: %\end{figure}
527: 
528: 
529: \end{document}
530: 
531: