cond-mat0503324/7.tex
1: \documentclass[twocolumn,showpacs]{revtex4}
2: %\documentclass[preprint,showpacs]{revtex4}
3: 
4: \usepackage{graphicx}
5: 
6: \begin{document}
7: 
8: 
9: \title{Nonequilibrium Reweighting on the Driven Diffusive Lattice Gas}
10: %\title{Title}
11: 
12: \author{Hwee Kuan Lee and Yutaka Okabe}
13: 
14: \affiliation{Department of Physics, Tokyo Metropolitan University, Hachioji, Tokyo 192-0397, Japan}
15: 
16: 
17: 
18: \begin{abstract}
19: The nonequilibrium reweighting technique, which was 
20: recently developed by the present authors, is used 
21: for the study of the nonequilibrium steady states. 
22: The renewed formulation of the nonequlibrium reweighting 
23: enables us to use the very efficient multi-spin coding. 
24: We apply the nonequilibrium reweighting to 
25: the driven diffusive lattice gas model. 
26: Combining with the dynamical finite-size scaling theory, 
27: we estimate the critical temperature $T_c$ and the dynamical exponent $z$. 
28: We also argue that this technique
29: has an interesting feature that enables explicit calculation of 
30: derivatives of thermodynamic quantities  without resorting to numerical
31: differences.
32: \end{abstract}
33: 
34: \pacs{75.40.Gb, 05.10.Ln, 66.30.Hs}
35: 
36: \maketitle
37: 
38: %\section{Background}
39: 
40: Most phenomena occurring in nature are in nonequilibrium states. 
41: Nonequilibrium systems, such as epidemic~\cite{henkel}, 
42: vehicular traffic~\cite{chowdhury}, biological network~\cite{berry}, 
43: and catalysis~\cite{ziff}, have captured a lot of attention.  
44: %
45: Monte Carlo simulation has become a standard tool in scientific computing, 
46: and advanced simulation methods, such as cluster 
47: algorithms~\cite{swendsen,hklee} and generalized ensemble 
48: methods~\cite{berg,lee,oliveira,wang1}, have been developed.  
49: However, many advanced Monte Carlo methods are not applicable 
50: to nonequilibrium systems.  Efficient Monte Carlo algorithms 
51: for nonequilibrium simulation are highly demanded.
52: 
53: 
54: Quite recently, the present authors~\cite{hklee1} have developed
55: a reweighting method for nonequilibrium systems based on the Sequential
56: Importance Sampling (SIS)~\cite{doucet,liubook}. With nonequilibrium 
57: reweighting, only simulation at a single temperature is required to obtain
58: information for a range of temperatures. 
59: %
60: The nonequilibrium reweighting method differs conceptually from 
61: conventional Monte Carlo methods. In many Monte Carlo methods, a
62: sequence of micro-states is sampled using the Boltzmann distribution.
63: One can interpret this as sampling over a ``path" generated by the
64: associated Monte Carlo updates.
65: Thermodynamic quantities are then averaged over this path. In 
66: nonequilibrium reweighting, many paths are first 
67: sampled with a trial distribution that is not necessarily equal to the 
68: Boltzmann distribution. Then thermodynamic quantities are calculated 
69: based on the relative probability between the trial distribution and the
70: Boltzmann distribution. The relative probability is called ``weights" in
71: literature, which we shall use hereafter. The advantage of this is that
72: one could sample many paths at one temperature and 
73: then calculate required thermodynamic quantities for a range of 
74: temperatures.
75: 
76: 
77: 
78: Moreover, Saracco and Albano~\cite{saracco,albano} have proposed 
79: an effective analysis of nonequilibrium phase transitions, 
80: in the study of the driven diffusive lattice gas model~\cite{kartz}, 
81: using a dynamical finite-size scaling theory.  
82: The behavior of nonequilibrium phase transitions can be extracted 
83: from short time dynamics~\cite{luo,zheng,okano}. 
84: If we combine the advantages of dynamical finite-size scaling and 
85: nonequilibrium reweighting, we can achieve an effective way of 
86: simulation for nonequilibrium systems.
87: %
88: 
89: In this paper, we apply the nonequilibrium reweighting method~\cite{hklee1} 
90: to the study of the nonequilibrium steady states~\cite{dickman,zia}. 
91: We illustrate our method on the driven diffusive lattice 
92: gas model~\cite{kartz}. 
93: We reformulate the nonequilibrium reweighting method and implement very 
94: efficient multi-spin coding~\cite{bahnot,roland}.   We also make modifications 
95: to the dynamical finite-size scaling relation, which was originally proposed 
96: by Saracco and Albano~\cite{saracco,albano}, so that the advantages of 
97: reweighting and dynamical finite-size scaling can be combined.
98: 
99: 
100: %\section{Model}
101: 
102: Let us start with explaining the driven diffusive lattice gas model 
103: proposed by Katz, Lebowitz and Spohn (KLS)~\cite{kartz}. This
104: system is one of the most well known nonequilibrium models exhibiting 
105: a nonequilibrium steady state. It was first proposed as
106: a model for super-ionic conductors, and attained its popularity due to 
107: its complex collective behavior. It is constructed as a $L_x \times L_y$
108: square lattice with {\it half-filled} lattice sites having periodic 
109: boundary conditions. Its Hamiltonian is given by 
110: %
111: \begin{equation}
112: {\mathcal H} = -4 \sum_{\langle ij, i'j' \rangle} n_{ij} n_{i'j'},
113: \label{eq:klsham}
114: \end{equation}
115: %
116: where the summation is over nearest lattice sites. The variable 
117: $n_{ij} = 1$ when the site is filled and $n_{ij}=0$ otherwise. Attempts for
118: each particle to jump to an empty nearest neighbor site are given by the
119: Metropolis rate~\cite{metropolis},
120: %
121: \begin{equation}
122: T_{\beta,E}(\sigma'|\sigma) = \min \left[ 1, 
123: \exp ( - \beta (\Delta {\mathcal H} - \epsilon  E ) )\right],
124: \label{eq:rate}
125: \end{equation}
126: %
127: where $\sigma$ and $\sigma'$ are the system configurations before and 
128: after the jump, $\Delta \mathcal H$ represents the change in energy due
129: to the jump, $E$ is a constant driving force, $\epsilon = -1$, $0$ or 
130: $1$ depending on whether the jump is against, orthogonal or along the 
131: direction of the drive, and $\beta = 1/T$ is the inverse temperature 
132: of the thermal bath.  The $L_y$ direction is taken as 
133: the direction of the drive. 
134: The KLS model exhibits an 
135: order-disorder second order phase transition. In the ordered phase, 
136: strips of high- and low-density domains are formed along the direction 
137: of the drive. In the final steady state, the particles are condensed 
138: into a single strip parallel to the direction of the drive~\cite{hurtado}.
139: Hence the order parameter can be defined as the density 
140: profile along the direction of the drive~\cite{saracco}, and moments of
141: the order parameters are given by
142: %
143: \begin{equation}
144: \rho^{k} = \frac{1}{(L_x/2)} \sum_{j=1}^{L_x} \left| 
145: \frac{1}{ L_y}\sum_{i=1}^{L_y} n_{ij} - \frac{1}{2} \right|^{k},
146: %\rho^{k} = (L_x/2)^{-1} \sum_{j=1}^{L_x} \left| 
147: %(L_y)^{-1} \sum_{i=1}^{L_y} n_{ij} - \frac{1}{2} \right|^{k},
148: \end{equation}
149: %
150: where $n_{ij} = 0$ or $1$ as defined in Eq.~(\ref{eq:klsham}), and 
151: $k = 1, 2, 4$ represents the first, second and fourth moments of the order
152: parameter, respectively.
153: 
154: 
155: %\section{Method}
156: 
157: 
158: We briefly review the nonequilibrium reweighting based on SIS, 
159: and show the implementation on the KLS model.  Define a
160: path $\vec{x}_t$, a sequence of points in phase space $\sigma_i$ which
161: were visited during the course of simulation, as
162: %
163: \begin{equation}
164: \vec{x}_t = (\sigma_1, \sigma_2,\cdots \sigma_t).
165: \end{equation}
166: %
167: This path can be sampled using the Monte Carlo method at an inverse temperature
168: $\beta$ and a constant drive $E$. The objective is to calculate 
169: the appropriate weights for computing 
170: the thermal average of a quantity $Q$ at another inverse temperature $\beta'$ 
171: and another drive $E'$,
172: %
173: \begin{equation}
174: \langle Q(t) \rangle_{\beta',E'} = \sum_{j=1}^n w^j_t Q(\vec{x}^j_t) / 
175: \sum_{j=1}^n w^j_t,
176: \end{equation}
177: %
178: where the sum is over all sampled paths indexed by $j$ and $w^j_t$ are 
179: the weights.  The number of paths is denoted by $n$.  
180: To calculate the weights, the following steps are 
181: implemented,
182: %
183: %
184: \begin{enumerate}
185: %
186: %
187: \item Suppose $\vec{x}^j_t = (\sigma_1^j,\cdots,\sigma_t^j)$ 
188: up to time $t$ is sampled 
189: from a simulation at the inverse temperature $\beta$ and drive $E$.
190: %
191: \item To go from $t$, choose a pair of neighboring lattice sites at 
192: random. If one of the two sites is empty, move the particle to the empty
193: site with the rate, $T_{\beta,E}({\sigma'}^j|\sigma_t^j)$, 
194: which is the Kawasaki spin exchange process.  ${\sigma'}^j$
195: denotes the system configuration after the move.
196: %
197: \item An incremental weight $\delta w^j$ has different 
198: values according to two possible outcomes,
199: %
200: \begin{enumerate}
201: %
202: \item  If the move is accepted; $\sigma_{t+1}^j = {\sigma'}^j $ and 
203: $\delta w^j = T_{\beta',E'}({\sigma'}^j|\sigma^j_t) / 
204: T_{\beta,E}({\sigma'}^j|\sigma^j_t)$.
205: %
206: \item  If the move is rejected; $\sigma_{t+1}^j = \sigma_t^j$ and
207: $\delta w^j = [1- T_{\beta',E'}({\sigma'}^j|\sigma^j_t)] / 
208: [1-T_{\beta,E}({\sigma'}^j|\sigma^j_t)] $.
209: \end{enumerate}
210: %
211: The weights at $t+1$ is given by this incremental weight 
212: through the relation $w^j_{t+1} = \delta w^j \times w^j_t$ 
213: with $w^j_1=1$.
214: \end{enumerate}
215: %
216: %
217: For each path $j \in \{ 1, \cdots, n\}$, these steps are repeated until
218: some predetermined maximum Monte Carlo time is reached.  
219:  
220: 
221: We make a comment on the technical detail of calculating the weights. 
222: For case of infinite drive ($E=\infty$), possible values of 
223: incremental weights $\delta {w_i}$ are,
224: %
225: \begin{equation}
226: \begin{array}{rcl}
227: \delta {w_0}&=&1,        \\
228: \delta {w_1}&=&\exp(-12 (\beta'-\beta)),  \\
229: \delta {w_2}&=&\exp(-8 (\beta'-\beta)),   \\
230: \delta {w_3}&=&\exp(-4 (\beta'-\beta))   \\
231: \delta {w_4}&=& ( 1 - \exp(-12 \beta')) / ( 1 - \exp(-12 \beta))),  \\
232: \delta {w_5}&=& ( 1 - \exp(-8 \beta') ) / ( 1 - \exp(-8 \beta)) ),  \\
233: \delta {w_6}&=& ( 1 - \exp(-4 \beta') ) / ( 1 - \exp(-4 \beta)) ). 
234: \end{array}
235: \end{equation}
236: %
237: The weights can then be written as a product of incremental weights,
238: %
239: \begin{equation}
240: w^j_t = (\delta {w_1})^{h_1^j(t)} (\delta {w_2})^{h_2^j(t)} \cdots
241: 	(\delta {w_6})^{h_6^j(t)}.
242: \label{eq:msw}
243: \end{equation}
244: %
245: where $h^j_1(t) \cdots h^j_6(t) $ are the number of hits on the 
246: incremental weights $ \delta w_1 \cdots \delta w_6$ during the course of
247: simulation from time $1$ to $t$. Note that $\delta w_0$ is irrelevant in
248: Eq.~(\ref{eq:msw}). Generalization of this counting method to the case 
249: of finite $E$ is trivial.  Since the calculation of weights has been reduced
250: to accumulating a histogram, the multi-spin coding technique~\cite{bahnot} 
251: can be implemented not only for the spin update process but also for 
252: the calculation of histogram of incremental weights.  For system configuration updates, 
253: we follow the multi-spin coding technique similar to 
254: the case of the Kawasaki spin exchange model~\cite{roland}. 
255: Once the histogram $h^j_1(t) \cdots h^j_6(t)$ is obtained, 
256: using Eq.~(\ref{eq:msw}) allows us to reweight to a large number 
257: of temperatures (drives) with negligible extra computational
258: efforts.  A large increase of efficiency has been obtained 
259: by a new formulation of the nonequilibrium reweighting. 
260: The details of the multi-spin coding for the nonequilibrium 
261: reweighting will be given elsewhere. 
262: 
263: 
264: 
265: %
266: For the dynamical finite-size scaling, we use the following equation,
267: %
268: \begin{equation}
269: \rho^{k} = b^{-\frac{k \beta}{\nu_{\|}}} 
270: \rho^{* (k)}( b^{-z} \tau, b^{\frac{1}{\nu_{\|}}} \epsilon, b^{-1} L_y,
271: b^{-\frac{ \nu_{\bot} }{ \nu_{\|}} } L_x, b^{x_0} \rho_0),
272: \end{equation}
273: %
274: where $k$ is the $k$th moment of the order parameter, $\rho^{*(k)}$ is 
275: the scaling function for the $k$th moment, $b$ is the spatial rescaling
276: factor, $\epsilon = (T - T_c)/T_c$, $\beta$ is the critical exponent for
277: the order parameter (it should not be confused with 
278: the inverse temperature), $\nu_{\|}$ and $\nu_{\bot}$ are the critical 
279: exponents for the correlation length parallel and orthogonal to the 
280: drive, respectively. $z$ is the dynamical exponent and $\tau$ is Monte 
281: Carlo steps per site. In addition to the original scaling form of 
282: Saracco and Albano~\cite{saracco,albano}, our scaling form has a term 
283: $b^{x_0} \rho_0$ to reflect the initial system 
284: configuration~\cite{okano,zheng}. $x_0$ is an independent exponent and
285: $\rho_0 \ll 1$ is the order parameter of the initial configuration.
286: Simulations have to be started with a chosen value of $\rho_0$ for all 
287: samples. We prepare our initial configuration with $\rho_0=0$ by 
288: inserting $L_y/2$ particles for each column in the lattice and then 
289: shuffling each column independently. Letting $b = \tau^{1/z}$, we have 
290: %
291: \begin{equation}
292: \rho^{k} = \tau^{-\frac{k \beta}{\nu_{\|} z}} 
293: \rho^{* (k)}( \tau^{\frac{1}{\nu_{\|}z}} \epsilon, 
294: \tau^{-\frac{1}{z}} L_y,
295: \tau^{-\frac{\nu_{\bot} }{ \nu_{\|}z} } L_x,
296: \tau^{\frac{x_0}{z}} \rho_0).
297: \end{equation}
298: %
299: In the limit of $L_x \rightarrow \infty$ at the critical temperature 
300: ($\epsilon=0$) with $\rho_0=0$, the ratio-of-moments reduces 
301: to a scaling function with a single argument,
302: %
303: \begin{equation}
304: \frac{\langle \rho^4 \rangle }{\langle \rho^2 \rangle^2} = 
305:       g(\tau^{-1/z} L_y)
306:       \mbox{ with } \rho_0 = 0, \epsilon 
307:       = 0, L_x \rightarrow \infty.
308: \label{eq:bscale}
309: \end{equation}
310: %
311: By plotting the ratio-of-moments versus $\tau L_y^{-z}$ at $T_c$ and 
312: $\rho_0 = 0$, neglecting corrections to scaling, the curves for 
313: different system sizes $L_y$ will collapse into a single curve. 
314: A measure of goodness-of-fit can be 
315: defined for the ``curve-collapse" as 
316: %
317: \begin{equation}
318: \eta =  \frac{1}{x_{\max} - x_{\min}} \int_{x_{\min}}^{x_{\max}}
319: \left|
320:  g_{L_{y1}} (x) - g_{L_{y2}} (x)
321: \right| d x,
322: \label{eq:fit}
323: \end{equation}
324: %
325: where $g_{L_{y1}}(x) = g( \tau L_{y1}^{-z})$
326: and   $g_{L_{y2}}(x) = g( \tau L_{y2}^{-z})$.
327: Our task is to choose $T_c$ and $z$ which minimize $\eta$. 
328: In using the relation Eq.~(\ref{eq:bscale}), we should check 
329: that the system size $L_x$ orthogonal to the drive 
330: is large enough. At this point, we should mention that 
331: Leung~\cite{leung} has studied the KLS model using finite size
332: scaling at nonequilibrium steady states. While we focus on
333: dynamical behaviors, his finite size scaling method was developed
334: for analysis at stead states.
335: 
336: \begin{figure}
337: \begin{picture}(0,180)(0,0)
338: \put(-130,0){\scalebox{.3}{
339: %\includegraphics{./L64x32Einfinite_m.pdf}
340: \includegraphics{./L64x32Einfinite_m.eps}
341: }}
342: \end{picture}
343: \caption{%
344: Plot of order parameter with infinite drive for $64 \times 32$ lattice 
345: with actual simulation performed at $T = 3.160$ shown with a bold line.
346: From top to bottom values of $T$ are 3.150, 3.155, 3.160, 3.165, 3.170.
347: Averages were taken over $4.096 \times 10^6$ samples.%
348: }
349: \label{fig:rhoInf}
350: \end{figure}
351: 
352: \begin{figure}
353: \begin{picture}(0,180)(0,0)
354: \put(-130,0){\scalebox{.3}{
355: %\includegraphics{./L32x32Efinite_m.pdf}
356: \includegraphics{./L32x32Efinite_m.eps}
357: }}
358: \end{picture}
359: \caption{%
360: Plot of order parameter with finite drive for $32 \times 32$ lattice 
361: with actual simulation performed at $(T,E) = (2.765, 0.515)$ and 
362: $(T,E)=(2.780,0.500)$. Reweighted data are combined using weighted mean.
363: >From top to bottom values of $T$ and $E$ are $(T,E)$ = (2.760,0.520), 
364: (2.765,0.515), (2.770,0.510), (2.775,0.505), (2.780,0.500), 
365: (2.785,0.495). Averages were taken over $2.048 \times 10^6$ samples for
366: each simulation. 
367: }
368: \label{fig:rhoE05}
369: \end{figure}
370: 
371: 
372: \begin{figure}
373: \begin{picture}(0,180)(0,0)
374: \put(-130,0){\scalebox{.3}{
375: %\includegraphics{./fit_z.pdf}
376: \includegraphics{./fit_z.eps}
377: }}
378: \end{picture}
379: \caption{%
380: Plots showing the goodness-of-fit $\eta$ and corresponding values of
381: dynamical exponent $z$ for various temperatures. Data are generated by
382: fitting ratio-of-moments for $L_y=64$ and $L_y=128$ between the range 
383: $\langle \rho^4 \rangle / \langle \rho^2 \rangle^2 = 
384: 1.2$ and $1.405$. $T_c$ is estimated from the temperature with the 
385: best fit ($T_c=3.175 \pm 0.002$). Error bars were generated by fitting
386: several ranges of ratio-of-moments.%
387: }
388: \label{fig:fit}
389: \end{figure}
390: 
391: 
392: %\section{Results}
393: 
394: We now show the results of the Monte Carlo simulation 
395: for the KLS model. 
396: We first illustrate the reweighting for the order parameter,
397: and then show how reweighting can be combined with dynamical finite-size 
398: scaling (Eq.~(\ref{eq:bscale})) to calculate the critical temperature and
399: dynamical exponent. 
400: Figure \ref{fig:rhoInf} shows how data over a range
401: of temperatures can be extracted from simulations at a single
402: temperature.  The temporal evolution of the order parameter $\rho$ 
403: for the infinite drive ($E=\infty$) was investigated 
404: for $64 \times 32$ lattice. 
405: Simulations were performed at $T=3.160$, 
406: and data were reweighted to nearby temperatures, $T=3.150, 3.155, 
407: 3.165, 3.170$ (from top to bottom). Averages were taken over $4.096 
408: \times 10^6$ samples. We made independent calculations directly at 
409: $T=3.150$, for example, to check the effectiveness 
410: of the reweighting. The deviation of the data between the reweighted
411: ones from $T=3.160$ and the direct ones at $T=3.150$ are 
412: found to be the same within statistical errors.
413: 
414: We also made simulations for the finite drive ($E \approx 0.5$).
415: We illustrate the reweighting over both $E$ and $T$. 
416: We performed two simulations at $(T,E)$ = $(2.765,0.515)$ and 
417: $(2.780,0.500)$ for $32 \times 32$ lattice.  
418: The reweighting of the order parameter is made 
419: by using $\bar{\rho}  = ( \mbox{$\sum_{k=1}^2$} \rho_k / 
420: \Delta^2_k  )/( \mbox{$\sum_{k=1}^2$} 1 / \Delta^2_k  )$, where 
421: $\rho_{1,2}$ and $\Delta_{1,2}$ are the order parameter and error 
422: estimates from the first and second simulations, respectively. 
423: Figure \ref{fig:rhoE05} shows the temporal evolution of the order
424: parameter for several temperatures and drives. 
425: Data was reweighted to several values at $(T,E)$ = $(2.760,0.520)$, 
426: $(2.770,0.510)$, $(2.775,0.505)$, $(2.785,0.495)$. 
427: Averages were taken over $2.048 \times 
428: 10^6$ samples for each simulation. Generally, we found that reweighting
429: is effective when the distributions $P_{\beta,E}(\vec{x}_t^j)$ and 
430: $P_{\beta',E'}(\vec{x}_t^j)$ have sufficient overlaps. Error bars and 
431: fluctuations of weights~\cite{liubook} can also be used as quantitative 
432: measures on the effective range of reweighting.
433: 
434: 
435: \begin{figure}
436: \begin{picture}(0,180)(0,0)
437: \put(-130,0){\scalebox{.3}{
438: %\includegraphics{./curve_collapse.pdf}
439: \includegraphics{./curve_collapse.eps}
440: }}
441: \end{picture}
442: \caption{%
443: Scaling plot of $\langle \rho^4 \rangle/ \langle \rho^2 \rangle^2$
444: versus $\tau L_y^{-z}$ for $z = 2.09$, $L_y=64$ (solid line) and 
445: $L_y=128$ (dotted line) at $T=3.175$. Initial system configurations 
446: were prepared with $\rho_0=0$.%
447: }
448: \label{fig:scaling}
449: \end{figure}
450: 
451: To determine $T_c$, we use the dynamical finite-size scaling 
452: of the ratio-of-moments 
453: (Eq.~(\ref{eq:bscale})).  Here we concentrate on the infinite 
454: drive ($E=\infty$). 
455: We simulated $64 \times 64$ and $64 \times 128$ lattices, 
456: and calculated the ratio of the moments, $\langle \rho^4 \rangle/\langle \rho^2 \rangle^2$.  
457: Before going into the discussion of the fitting, we make 
458: a comment on the system size $L_x$ whether we can consider 
459: as $L_x \rightarrow \infty$. 
460: We performed simulations for both $L_x=64$ and $L_x=128$, 
461: and confirmed that the ratio of moments for $64 \times L_y$ and 
462: $128 \times L_y$ coincided with each other to within statistical 
463: fluctuations. 
464: Thus, we may regard that $L_x=64$ is large enough. 
465: Since $\nu_{\|} > \nu_{\bot}$ for the KLS model, 
466: the correlation length orthogonal to the drive, $\xi_{\bot}$, 
467: develops slowly; hence $L_x=64$ is large enough to use 
468: the scaling relation Eq.~(\ref{eq:bscale}).
469: Now we show the fitting procedure.  
470: Fitting was performed for several temperatures
471: near $T_c$, which were reweighted from the data obtained 
472: at a single temperature, and 
473: for each temperature we adjusted the value of $z$ such 
474: that the goodness-of-fit $\eta$, Eq.~(\ref{eq:fit}), becomes minimum. 
475: Figure \ref{fig:fit} shows $\eta$ 
476: for several temperatures and the values of $z$ used to calculate $\eta$.
477: The best fit occurs at $T_c = 3.175 \pm 0.002$; the error bar on $T_c$
478: is estimated by including all neighboring temperatures where the mean 
479: values of $\eta$ are within two standard deviations of $\eta$ at 
480: $T = 3.175$. The value of $z$ within $T=3.175\pm0.002$ is 
481: $z=2.09\pm 0.01$, and we use this value as our estimate of the dynamical
482: exponent. Figure \ref{fig:scaling} shows the scaling plot of
483: $\langle \rho^4 \rangle/ \langle \rho^2 \rangle^2$ as a function of 
484: $\tau L_y^{-z}$ for $64\times 64$ (solid line) and 
485: $64 \times 128$ (dotted line) lattice sizes at $T = 3.175$ and $z = 2.09$. 
486: The curves are almost 
487: indistinguishable at this scale although some corrections to scaling can
488: be observed below $\tau L_y^{-z} = 0.02$. To study the corrections to 
489: scaling, the goodness-of-fit for ratio-of-moments for smaller sizes, 
490: that is, $64\times 32$ and $64 \times 64$ lattices, 
491: was also calculated using a similar procedure. 
492: The best fit occurs at $T = 3.155 \pm 0.005$ with $z = 2.23 \pm 0.03$. 
493: The estimate for $T_c$ increases with the system size, whereas 
494: that for $z$ decreases.  Our estimates of $T_c$ and $z$ are 
495: compatible with the recent estimates for infinite lattice,  
496: $T_c=3.1980\pm0.0002$~\cite{caracciolo}, $T_c=3.200 
497: \pm 0.010$~\cite{albano}, $z=2.016\pm0.040$~\cite{albano}. 
498: A more systematic analysis of the corrections to scaling 
499: to get a precise estimate of $T_c$ and several critical 
500: exponents for infinite lattice will be left to a separate 
501: publication.  Before closing we show the actual procedure 
502: of the reweighting for each system size.  For $64 \times 32$ lattice, 
503: $4.096\times 10^6$ samples were used for the simulation
504: at $T=3.16$. For $64 \times 64$ lattice, $8.19 \times 10^5$ samples 
505: were used for each simulation at $T = 3.174$ and $3.180$.  
506: Results were then reweighted to other temperatures and 
507: combined using weighted mean, $\bar{r}  = ( \mbox{$\sum_{k=1}^2$} r_k / 
508: \Delta^2_k  )/( \mbox{$\sum_{k=1}^2$} 1 / \Delta^2_k  )$.  Here 
509: $r_{1,2}$ and $\Delta_{1,2}$ are the ratio-of-moments and error 
510: estimates from the first and second simulations, respectively.
511: For $64 \times 128$ lattice size, $1.64 \times 10^5$ samples 
512: were used for each simulation at $T = 3.174, 3.177$ and $3.180$, 
513: and reweighted results were combined using the same procedure.
514: 
515: 
516: 
517: 
518: %
519: %\section{Conclusion}
520: %
521: 
522: To summarize, we have studied the use of nonequilibrium reweighting 
523: based on SIS for the nonequilibrium steady states. 
524: We have reformulated the nonequilibrium reweighting method, 
525: which is convenient for the multi-spin coding.  
526: As a result, a large increase of efficiency has been achieved 
527: for the performance of simulations. 
528: We have applied the nonequilibrium reweighting to 
529: the driven diffusive lattice gas model (the KLS model). 
530: Combining with dynamical finite-size scaling theory, we have 
531: estimated $T_c$ and 
532: the dynamical exponent $z$. 
533: 
534: 
535: Finally, we make a remark on possible applications. 
536: The nonequilibrium reweighting method is very general and 
537: has some very interesting properties.  For example, the 
538: fluctuation-dissipation theorem does not hold for nonequilibrium systems
539: and derivatives of thermodynamic quantities had been estimated using 
540: finite differences~\cite{valles1987}. With reweighting, 
541: derivatives can be calculated directly by differentiating the 
542: weights explicitly, that is, 
543: %
544: \begin{equation}
545: \frac{d \langle Q(t) \rangle_{\beta'}}{d \beta'} = 
546: \frac{ \sum_{j=1}^n Q(\vec{x}^j_t) \frac{d w^j_t}{d\beta'}  }{
547: \sum_{j=1}^n w^j_t } - 
548: \langle Q(t) \rangle_{\beta'} 
549: \frac{ \sum_{j=1}^n \frac{d w^j_t}{d \beta'}}{ \sum_{j=1}^n w^j_t}.
550: \end{equation}
551: %
552: Here, $d w^j_t / d \beta'$ can be obtained by differentiating 
553: Eq.~(\ref{eq:msw}) with respect to $\beta'$. 
554: We believe that the nonequilibrium reweighting method would have 
555: several directions for applications.
556: %We believe that the 
557: %nonequilibrium reweighting method would be a good alternative to doing 
558: %simulations. 
559: 
560: This work is supported by a Grant-in-Aid for Scientific Research from
561: the Japan Society for the Promotion of Science. The computation of
562: this work has been done using computer facilities of the Supercomputer
563: Center, Institute of Solid State Physics, University of Tokyo.
564: 
565: \begin{thebibliography}{99}
566: \bibitem{henkel}      M. Henkel and H. Hinrichsen, J. Phys. A {\bf 37}, R117 (2004). 
567: \bibitem{chowdhury}   D. Chowdhury, L. Santen, and A Schadschneider, Phys. Rep. {\bf 329}, 199 (2000). 
568: \bibitem{berry}       H. Berry, Phys. Rev. E {\bf 67}, 031907 (2003). 
569: \bibitem{ziff}        R. M. Ziff, E. Gulari, and Y. Barshad, Phys. Rev. Lett. {\bf 56}, 2553 (1986). 
570: \bibitem{swendsen}    R. H. Swendsen and J. S. Wang, Phys. Rev. Lett. {\bf 58}, 86 (1987). 
571: \bibitem{hklee}       H. K. Lee and R. H. Swendsen, Phys. Rev. B {\bf 64}, 214102 (2001). 
572: \bibitem{berg}        B. A. Berg and T. Neuhaus, Phys. Lett. B {\bf 267}, 249 (1991).
573: \bibitem{lee}         J. Lee, Phys. Rev. Lett. {\bf 71}, 211 (1993).
574: \bibitem{oliveira}    P. M. C. de Oliveira, T. J. P. Penna, and H. J. Herrmann, Eur. Phys. J. B {\bf 1}, 205 (1998).
575: \bibitem{wang1}       F. Wang and D. P. Landau, Phys. Rev. Lett. {\bf 86}, 2050 (2001).
576: 
577: \bibitem{hklee1}      H. K. Lee and Y. Okabe, Phys. Rev. E {\bf 71}, 015102(R) (2005). 
578: \bibitem{doucet}      A. Doucet, N. De Freitas and N. Gordon, {\it Sequential Monte Carlo Methods in Practice}, (Springer, New York, 2001).
579: \bibitem{liubook}     J. S. Liu, {\it Monte Carlo Strategies in Scientific Computing}, (Springer, 2001).
580: 
581: \bibitem{saracco}     G. P. Saracco and E. V. Albano, J. Chem. Phys. {\bf 118}, 4157 (2003). 
582: \bibitem{albano}      E. V. Albano and G. P. Saracco, Phys. Rev. Lett. {\bf 88}, 145701 (2002). 
583: 
584: \bibitem{kartz}       S. Kartz, J. L. Lebowitz, and H. Spohn, Phys. Rev. B {\bf 28}, 1655 (1983). 
585: 
586: \bibitem{luo}         H. J. Luo, L. Sch\"{u}lke, and B. Zheng, Phys. Rev. Lett. {\bf 81}, 180 (1998). 
587: \bibitem{zheng}       B. Zheng, M. Schulz, and S. Trimper, Phys. Rev. E {\bf 59}, R1351 (1999). 
588: \bibitem{okano}       K. Okano, L. Sch\"{u}lke, K. Yamagishi, and B. Zheng, Nucl. Phys. B {\bf 485}, 727 (1997). 
589: %\bibitem{ito}         N. Ito, K. Hukushima, K. Ogawa and Y. Ozeki, J. Phys. Soc. Jpn. {\bf 69}, 1931 (2000). 
590: \bibitem{dickman}     J. Marro and R. Dickman, {\it Nonequilibrium Phase Transitions in Lattice Models}, (Cambridge University Press, 1999).
591: \bibitem{zia}         B. Schmittmann and R. K. P. Zia, {\it Statistical Mechanics of Driven Diffusive Systems}, (Academic Press, 1995).
592: 
593: \bibitem{bahnot}      G. Bahnot, D. Duke, and R. Salvador, Phys. Rev. B {\bf 33}, 7841 (1986).
594: \bibitem{roland}      C. Roland and M. Grant, Phys. Rev. Lett. {\bf 60}, 2657 (1988).
595: \bibitem{metropolis}  N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. M. Teller, and E. Teller, J. Chem. Phys. {\bf 21}, 1087 (1953). 
596: \bibitem{hurtado}      P. I. Hurtado, J. Marro, P. L. Garrido and E. V. Albano, Phys. Rev. B, {\bf 67}, 014206 (2003).
597: \bibitem{leung}        K-t. Leung, Phys. Rev. Lett. {\bf 66}, 453, (1991).
598: 
599: %\bibitem{caria}       M. Caria, {\it Measurement Analysis}, (Imperial College Press, 2000).
600: \bibitem{caracciolo}  S. Caracciolo, A. Gambassi, M. Gubinelli, and A. Pelissetto, J. Phys. A {\bf 36}, L315 (2003).
601: \bibitem{valles1987}  J. L. Vall\'{e}s and  J. Marro, J. Stat. Phys. {\bf 49}, 89 (1987).
602: \end{thebibliography}
603: 
604: 
605: %\bibitem{ferrenberg}  A. M. Ferrenberg, R. H. Swendsen, Phys. Rev. Lett., {\bf 61}, 2635, (1988). 
606: 
607: \end{document}
608: 
609: