cond-mat0006217/2c.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%
2: % This is a REVTeX file %
3: %%%%%%%%%%%%%%%%%%%%%%%%%
4: 
5: % To get 1-column output,
6: % comment out those ending with "%c2",  and enable those with "%c1".
7: % To get 2-column output, do the reverse.
8: 
9: \documentstyle[prd,aps,epsf,epsfig]{revtex} %c2
10: %\documentstyle[preprint,prd,aps,epsf,epsfig]{revtex}%c1
11: \begin{document}
12: \draft
13: \twocolumn[\hsize\textwidth\columnwidth\hsize\csname@twocolumnfalse\endcsname%c2
14: 
15: %:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
16: \title{Heuristic derivation of continuum kinetic equations 
17: from microscopic dynamics}
18: %:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
19: 
20: \author
21: {{Kwan-tai Leung}\cite{email_ktl}}
22: \address{
23: Institute of Physics, Academia Sinica,
24: Taipei, Taiwan 11529, R.O.C.
25: }
26: 
27: \maketitle
28: \centerline{\small (Last revised \today)}
29: 
30: \begin{abstract}
31: We present an approximate and heuristic scheme for 
32: the derivation of continuum kinetic equations
33: from microscopic dynamics for stochastic, interacting systems.  
34: The method consists of a mean-field type, decoupled approximation 
35: of the master equation followed by the `naive' continuum limit.
36: The Ising model and driven diffusive systems are used as illustrations.
37: The equations derived are in agreement with other approaches,
38: and consequences of the microscopic dependences 
39: of coarse-grained parameters compare favorably with exact 
40: or high-temperature expansions.
41: The method is valuable when more systematic and rigorous
42: approaches fail, and when microscopic inputs in
43: the continuum theory are desirable.
44: \end{abstract}
45: 
46: \pacs{PACS numbers: 05.50.+q, 05.10.Gg, 64.60.Cn}
47: \vspace{2pc}
48: ]%c2
49: 
50: %-------------------------------------
51: \section{Introduction}\label{INTRO}
52: %-------------------------------------
53: 
54: Ever since its introduction
55: in a classic treatment of the Brownian motion\cite{langevin},  
56: the Langevin equation has been playing an important 
57: role in modern statistical physics.
58: It provides a mathematical framework and a physical basis
59: for studying stochastic processes in statistical, mechanical systems.
60: Applications are wide-ranging\cite{vankampen}, including
61: chemical reactions, laser physics, diffusive processes,
62: and modern theories of dynamical critical phenomena\cite{hh}.
63: Recent topics such as surface growth\cite{surface}
64: and pattern formation\cite{pattern} 
65: also rely heavily on the Langevin equation.
66: 
67: It is fair to say, however, that despite its popularity, 
68: the Langevin equation for a specific problem 
69: is seldom derived from the corresponding microscopics.
70: It is often postulated on grounds of symmetry and physical reasoning.
71: Only rarely in simple circumstances is it derived,
72: for example, in reasonable details from the more
73: fundamental master equation.
74: In this article, we shall present an approximate 
75: scheme for deriving the Langevin equation, starting from 
76: the microscopic specification of the dynamics.
77: Our goal is not to provide a formal derivation, 
78: but rather to propose a simple and heuristic means that can be 
79: applied generally to many stochastic systems, 
80: several of which shall be discussed below.
81: 
82: Expositions of the historical, philosophical and technical 
83: aspects of the Langevin equation are beyond the scope of this article, 
84: interested readers are thus referred to the 
85: relevant literature\cite{vankampen}.
86: This paper is organized as follows:
87: An elementary recapitulation of the 
88: master equation and Langevin equation
89: is presented in Section \ref{TDGL}.
90: Section \ref{DERIVATION} contains several examples 
91: as illustrations of the method, as well as assessment
92: of the quality of the approximation involved.
93: Conclusion is given in Section \ref{CONCLUSION}.
94: A discussion of the noise correlation is given in the Appendix.
95: 
96: 
97: %-------------------------------------
98: \section{master equation and time-dependent 
99: Ginzburg-Landau equation}\label{TDGL}
100: %-------------------------------------
101: 
102: In statistical physics, one of the most important applications 
103: of the Langevin equation is 
104: in the theories of dynamical critical phenomena\cite{hh}.
105: Therefore, our discussion shall be cast in that language,
106: although it should be obvious that the method itself is not limited
107: to systems exhibiting those phenomena.
108: For concreteness, consider the kinetic Ising model 
109: that obeys Glauber (i.e., spin-flip) dynamics\cite{Glauber}.  
110: At the classical, microscopic level of description,
111: the system consists of $N$ spins $\sigma_i$ 
112: %(or particles through the lattice-gas correspondence)
113: interacting {\em via\/} the Hamiltonian
114: \begin{equation}
115: H=-J \sum_{\langle i,j \rangle} \sigma_i \sigma_j,
116: \label{ham}
117: \end{equation}
118: where $J$ is the coupling constant, and 
119: $\langle i,j \rangle$ denotes a sum over nearest-neighbor pairs.
120: The time evolution of the system is governed by a master equation
121: \begin{eqnarray}
122: && P({\vec\sigma};t+1)-P({\vec\sigma};t)= 
123: \nonumber\\ &&\qquad 
124: \sum_{\vec\sigma'}
125:          \left[ w({\vec\sigma'}\to {\vec\sigma})P({\vec\sigma'};t)
126:               -w({\vec\sigma}\to {\vec\sigma'})P({\vec\sigma};t) \right],
127: \label{master}
128: \end{eqnarray}
129: where $P({\vec\sigma};t)$ is the joint probability of finding the
130: system in the spin configuration 
131: $\vec\sigma\equiv\{\sigma_1,\sigma_2,\cdots,\sigma_N\}$
132: at time $t$, and $w$'s are the transition rates
133: between two configurations that differ only by one spin flip.
134: There is a great deal of freedom in the choice of $w$,
135: as long as the following detailed balance condition 
136: is satisfied to ensure the same equilibrium distribution 
137: $P_{\rm eq}({\vec\sigma})\sim e^{-\beta H(\vec\sigma)}$
138: \begin{equation}
139:  {w({\vec\sigma'}\to {\vec\sigma}) \over
140:   w({\vec\sigma}\to {\vec\sigma'})      }
141: ={P_{\rm eq}({\vec\sigma}) \over P_{\rm eq}({\vec\sigma'}) }
142: =e^{-\beta[H(\vec\sigma)-H(\vec\sigma')]},
143: \label{db}
144: \end{equation}
145: with $\beta=1/k_B T$.
146: In practice, the choice is largely dictated by mathematical convenience.
147: %(hereafter the Boltzmann constant $k_B \equiv 1$).
148: The most common choices of $W$ are the Metropolis rate in Monte Carlo
149: simulations for its ease of implementation, 
150: and the heat-bath rate (also known as the Kawasaki
151: rate\cite{kawasaki}) in analytic calculations for its analyticity.
152: In this paper, we confine our attention to the latter choice.
153: It is given by
154: \begin{equation}
155: w({\vec\sigma}\to {\vec\sigma'})=
156:        { 1 \over 1+e^{-\beta[H(\vec\sigma)-H(\vec\sigma')]} }.
157: \label{heatbath}
158: \end{equation}
159: 
160: Since the master equation is not very convenient for 
161: analytic purposes such as a renormalization group analysis, 
162: one often turns to a mesoscopic, continuum representation.
163: For an Ising system with Glauber dynamics, 
164: the relevant continuum field is 
165: the local magnetization density $\phi(\vec{r},t)$, which
166: obeys a kinetic equation
167: \begin{eqnarray}
168: {\partial\phi\over \partial t}
169: &=& -\Gamma {\delta {\cal H}\over \delta \phi}+\zeta,
170: \label{tdgl}\\
171: {\cal H} &=&
172: \int d^d r\,\left\{ {1\over 2}(\nabla\phi)^2 + V(\phi) \right\},
173: \label{hamcg}\\
174: V(\phi) &=&
175:  {u\over 2}\phi^2 + {g\over 4!}\phi^4 + \cdots.
176: \label{Vcg}
177: \end{eqnarray}
178: This is an example of the time-dependent Ginzburg-Landau (TDGL) 
179: kinetic equation.
180: In (\ref{hamcg}), $d$ is the dimensionality of the system,
181: and $\cal H$ is a coarse-grained Hamiltonian. 
182: %with $V$ the
183: %usual single- or double-well potential depending on
184: %whether $u$ is positive or negative.
185: For (\ref{tdgl}) to describe a stochastic process, the noise term
186: $\zeta(\vec{r},t)$ is needed, which 
187: accounts for the effect of thermal fluctuations and prevents
188: the system from trapping in metastable states.
189: For mathematical convenience, it is often taken to be Gaussian 
190: with zero mean:
191: \begin{eqnarray}
192: \langle \zeta(\vec{r},t) \rangle&=&0, \\
193: \langle \zeta(\vec{r},t)\zeta(\vec{r}',t') \rangle&=&
194: 2{D} \delta(\vec{r}-\vec{r}')\delta(t-t').
195: \label{noisecorrel}
196: \end{eqnarray}
197: For equilibrium systems, 
198: the correlation $D$ in (\ref{noisecorrel}) 
199: has to be chosen to ensure that
200: the stationary solution of Eq.~(\ref{tdgl}) is consistent with
201: the Boltzmann weight ${\cal P}_{\rm eq}\sim e^{-{\cal H}}$ (cf. Appendix).
202: For a system as simple as the Ising model,
203: the static continuum Hamiltonian $\cal H$ 
204: can actually be derived from the microscopic $H$ 
205: {\em via\/} the partition function
206: by means of the Hubbard-Stratonovich transformation\cite{amit},
207: which is a trick based on the Gaussian integral.
208: %$\int_{-\infty}^{\infty} e^{-ax^2}=\sqrt{\pi/a}$.
209: For the dynamics, the TDGL equation can be derived by 
210: coarse graining the master equation\cite{langer71,omegaexp}, 
211: which in principle yields expressions of the coarse-grained
212: parameters $\Gamma$, $u$ and $g$ in (\ref{tdgl})-(\ref{Vcg}) 
213: as functions of microscopic ones in (\ref{master}).
214: 
215: However, for more complicated $H$,
216: %(such as when the interaction is not quadratic), 
217: the Hubbard-Stratonovich transformation fails and
218: the coarse graining cannot be done explicitly.
219: Moreover, these methods rely on the existence of a Hamiltonian
220: $H$ and the associated equilibrium Boltzmann weight $e^{-\beta H}$,
221: which is not valid in generic non-equilibrium situations
222: defined only by the dynamics\cite{ness}.
223: For these reasons, it is highly desirable to
224: have a way with which a continuum description can be
225: extracted directly from the dynamics\cite{intuitive}.
226: %In the next section, we shall present such a procedure.
227: 
228: 
229: %-------------------------------------
230: \section{factorization and naive continuum expansion}\label{DERIVATION}
231: %-------------------------------------
232: 
233: Our method is very simple. It consists of two steps:
234: a mean-field type factorization of joint probabilities 
235: into singlet ones in the master equation, 
236: followed by a `naive' continuum expansion. 
237: The result is a continuum kinetic equation
238: with full knowledge of the microscopic dependence of 
239: the coarse-grained parameters.  Since the input is the master equation, 
240: whether the system is an equilibrium one or not\cite{ness} 
241: is irrelevant.  To illustrate, we now discuss several examples
242: in increasing order of sophistication.
243: 
244: 
245: 
246: \subsection{1D Ising model}
247: %-------------------------------------
248: 
249: Focusing on a spin at position $x$
250: in a one-dimensional (1D) Ising model, it is easy to find by 
251: integrating out all other spins in Eq.~(\ref{master}) that
252: \begin{eqnarray}
253:  P_+&&(x;t+1)-P_+(x;t)= 
254:      P_{---}w_{---}
255:     +P_{--+}w_{--+}
256: \nonumber\\
257: &&
258:     +P_{+-+}w_{+-+}
259:     +P_{+--}w_{+--}
260:     -P_{-+-}w_{-+-}
261: \nonumber\\
262: &&
263:     -P_{-++}w_{-++}
264:     -P_{++-}w_{++-}
265:     -P_{+++}w_{+++},
266: \end{eqnarray}
267: where $P_+(x;t)$ denotes the singlet probability of finding the spin up
268: at site $x$ at time $t$, and
269: $P_{+++}(x;t)$ denotes the joint probability of finding three spins up
270: at site $x-1$, $x$, $x+1$ respectively, and so on.
271: From (\ref{heatbath}), the heat-bath transition rates are given by
272: $w_{---}=w_{+++}=W_4$,
273: $w_{--+}=w_{+--}=w_{-++}=w_{++-}=W_0$,
274: $w_{+-+}=w_{-+-}=W_{-4}$, where 
275: \begin{equation}
276: W_n\equiv { 1 \over 1+e^{n\beta J} }.
277: \label{w_n}
278: \end{equation}
279: Adopting a mean-field approximation, the joint probabilities
280: are replaced by their factorizations, e.g.,
281: $P_{++-}(x;t)\to P_+(x-1;t)P_+(x;t)P_-(x+1;t)$.
282: Since $\sum_\sigma \sigma P_\sigma(x;t)=P_+(x;t)-P_-(x;t)
283: =\langle \sigma\rangle$,
284: and $\sum_\sigma P_\sigma(x;t)=P_+(x;t)+P_-(x;t)=1$,
285: in the spirit of coarse graining 
286: we proceed to make the identification 
287: \begin{equation}
288: P_\pm(x;t) \leftrightarrow {1\pm \phi(x;t)\over 2},
289: \label{P2phi}
290: \end{equation}
291: where $\phi$ is the local magnetization density.
292: By using $\phi$ instead of spin number densities, we take
293: advantage of symmetries anticipated in the final kinetic equation.
294: Since a spin flip depends on a total of $z+1$ spins in (\ref{master}),
295: where $z=2d$ for hyper-cubic lattices,
296: the factorization effectively produces a power series expansion
297: in $\phi$ up to $\phi^{z+1}$.
298: After replacing $P$'s by $\phi$'s,  
299: we make the transition to the continuum by `naively'
300: expanding about $x$, such as:
301: \begin{equation}
302: \phi(x\pm 1;t)\rightarrow \phi(x;t)
303:               \pm {\partial \phi(x;t)\over \partial x}
304:               +{1\over2}{\partial^2 \phi(x;t)\over \partial x^2} +\cdots.
305: \label{naive}
306: \end{equation}
307: For most applications, we are only interested in the
308: long-distance behavior, hence it suffices to stop at the lowest
309: derivatives as shown.
310: This procedure results in 
311: a deterministic kinetic equation for $\phi$ in precisely the form of 
312: (\ref{tdgl})-(\ref{Vcg}), barring the noise term $\zeta$:
313: \[
314: {\partial \phi\over \partial t}
315: =-\Gamma\left(-{\partial^2\phi\over\partial x^2}+r\phi
316:               +{ g\over 6}\phi^3\right),
317: \]
318: where the coefficients are given by
319: \begin{eqnarray}
320: \Gamma&=&{1\over 2} (W_{-4} - W_4),
321: \nonumber\\
322: r&=&{ 1\over 2\Gamma}\left( 3W_4-W_{-4}+2W_0 \right),
323: \nonumber\\
324: g&=&{ 3\over \Gamma} \left( W_4+W_{-4}-2W_0 \right).
325: \label{1DIsing}
326: \end{eqnarray}
327: 
328: Several remarks are in order:
329: \begin{enumerate}
330: \item
331: Symmetries in the resulting 
332: continuum equation (with respect to $\phi$, $x$ and $t$)
333: are as expected, because the approximations respect those symmetries
334: and leave them intact.
335: \item
336: There are explicit temperature dependences in the coefficients 
337: which cannot be deduced by symmetry or physical reasoning.
338: Such dependences are specific to the choice of jump rates
339: which manifests through the approximations used.
340: \item
341: Noting that $W_{-n}=1-W_n$ for any $n$,  we find
342: $\Gamma={1\over 2}-W_4>0$ and $g=0$ at any $T$\cite{truncate}, and
343: $r=2 W_4/\Gamma$ has one zero, at $T=0$.
344: %While $\Gamma$ must be positive for stability reason,
345: This is consistent with the absence of 
346: phase transition in the 1D Ising model at any finite temperature, 
347: an improvement over the usual mean-field result 
348: $T_c^{\rm MF}=2 J/k_B$.
349: There is no stability problem arising from $g=0$
350: because the quadratic coefficient is positive for $T>0$.
351: \item
352: In the presence of an external magnetic field $h$, the degeneracies
353: in jump rates are lifted 
354: (e.g., $W_{+\pm-}=(1+e^{\pm 2\beta h})^{-1}$). 
355: To $O(h)$, the kinetic equation acquires a new term 
356: $\Gamma\mu h$ on the right hand side, where 
357: \begin{equation}
358: \mu={2\beta \over \Gamma} (W_0^2 + W_4 W_{-4}).
359: \end{equation}
360: Linear response then determines that the susceptibility
361: is $\chi=\mu/r=\beta (1-\gamma^2/2)/(1-\gamma)$, 
362: where $\gamma\equiv \tanh 2\beta J$. 
363: In the Appendix we show that $\mu$ is 
364: needed to fix the noise correlation.
365: \item
366: Besides capturing the correct symmetries,
367: our results compare quite well 
368: with exact results.  From (\ref{tdgl}), the
369: relaxation time can be read off easily as
370: $\tau=1/\Gamma r=1/(1-\gamma)$.
371: This turns out to be exact\cite{Glauber}.
372: For the susceptibility, deviation from
373: the exact result $\beta e^{2\beta J}$\cite{Glauber} 
374: shows up only at $O((\beta J)^5)$ when expanded in $\beta J$.
375: Hence, our method has the advantage that 
376: it embodies a refined mean-field theory, 
377: as already applied to studies of stochastic resonance 
378: in Ising systems\cite{leungneda}.
379: \item
380: Finally, due to the factorizations only the deterministic terms
381: in (\ref{tdgl}) can be derived. The noise term has 
382: to be deduced separately (see Appendix).
383: The result for the noise correlation $D$ in (\ref{noisecorrel})
384: is ${D}=k_B T \mu \Gamma$.
385: \end{enumerate}
386: 
387: Having gone through the details of our method, we now turn 
388: to a few less trivial examples.
389: 
390: 
391: 
392: \subsection{2D Ising model}
393: %-------------------------------------
394: 
395: The same procedure can be applied to the 2D Ising model 
396: with Glauber dynamics.
397: Again we obtain (\ref{tdgl}), with the parameters given by
398: \begin{eqnarray}
399: \Gamma&=&{1\over 8}  (      -2W_4 + 2W_{-4} -  W_8 + W_{-8}),
400: \label{Tdep_G2d}\\
401: r&=&{1\over 8 \Gamma}( 6W_0+12W_4 - 4W_{-4} + 5W_8 -3W_{-8}),
402: \label{Tdep_u2d}\\
403: g&=&{3\over 2 \Gamma}(-6W_0-4 W_4 + 4W_{-4} + 5W_8 + W_{-8}),
404: \label{Tdep_g2d}\\
405: \mu&=&{\beta \over 2\Gamma} (3W_0^2+4W_4 W_{-4} + W_8 W_{-8}).
406: \label{Tdep_mu2d}
407: \end{eqnarray}
408: It is worth noting that $\Gamma$, $g$ and $\mu$ 
409: are positive definite for all $T>0$, whereas $r$ has one zero at
410: $T_c^{\rm GL} \approx 3.0898 J/k_B \approx 1.3616 T_c$, 
411: again an improvement over the mean-field prediction 
412: $T_c^{\rm MF}=4 J/k_B$, 
413: where $T_c=-2J/k_B\ln(\sqrt{2}-1)\approx2.2692J/k_B$ 
414: is the exact critical temperature.
415: As expected, there is no $\phi^5$ and higher order term\cite{truncate}.
416: 
417: The results of $\tau$ and $\chi$ for the Gaussian case
418: ($g=0$) are quite satisfactory.  They differ from high-temperature
419: series expansions\cite{hte} at order $O((\beta J)^5)$ and
420: $O((\beta J)^4)$, respectively,  whereas the usual mean-field
421: results are worse, at $O((\beta J)^3)$ and $O((\beta J)^2)$.
422: 
423: 
424: 
425: \subsection{3D Ising model}
426: %-------------------------------------
427: 
428: Despite being more tedious (128 terms on the right-hand side 
429: of the master equation),
430: we also derive the kinetic equation 
431: for the 3D Ising model with Glauber dynamics. The results are:
432: \begin{eqnarray}
433: \Gamma&=&{1\over 32}  
434:          (W_{-12} +4W_{-8}+5W_{-4}-5W_4 -
435: \nonumber \\ &&\qquad 
436:           4 W_8- W_{12}),
437: \label{Tdep_G3d}\\
438: r&=&{1\over 32 \Gamma}  
439:          (-5W_{-12} -18W_{-8}-15W_{-4}+20W_{0}+
440: \nonumber \\ &&\qquad 
441:           45W_4+30W_8+7W_{12}),
442: \label{Tdep_u3d}\\
443: g&=&{15\over 16\Gamma}
444:          (-W_{-12} +6W_{-8}+9W_{-4}-12W_{0}-
445: \nonumber \\ &&\qquad 
446:          15W_4+6W_8+7W_{12}),
447: \label{Tdep_g3d}\\
448: \mu&=&{\beta \over 8\Gamma} 
449:          (10W_{0}^2+15W_{-4}W_4+6W_{-8}W_8+
450: \nonumber \\ &&\qquad 
451:           W_{-12}W_{12}),
452: \label{Tdep_mu3d}
453: \end{eqnarray}
454: As for 2D Ising,  $\Gamma$, $g$ and $\mu$ 
455: are positive definite for all $T>0$, and $r$ has one zero at
456: $T_c^{\rm GL} \approx 5.0733 J/k_B $, 12\% higher than 
457: the best estimate\cite{landau},
458: compared to $T_c^{\rm MF}=6 J/k_B$. 
459: %The coefficient of $\phi^5$ is negative but small, about
460: %an order of magnitude smaller than that of $\phi^3$, 
461: %whereas the coefficient of $\phi^7$ vanishes\cite{rates}.
462: %--N.B. bibitem{rates} is disabled too.
463: 
464: 
465: 
466: 
467: \subsection{1D driven lattice gas}
468: %-------------------------------------
469: 
470: In many generic non-equilibrium systems\cite{ness},
471: the free energy does not exist and one has to start from
472: the dynamics, such as described by the master equation.
473: A notable example is the driven diffusive system\cite{dds}, which
474: is regarded as a paradigm of spatially extensive 
475: interacting systems that exhibit cooperative phenomena 
476: in steady-state non-equilibrium situations.
477: In its standard form, it models an Ising-like lattice gas of particles
478: whose motion along a certain direction 
479: is biased by an external drive denoted by $E$.
480: For $E=0$, the model reduces to the ordinary kinetic Ising model
481: with Kawasaki, or spin-exchange, dynamics (model B in \cite{hh}).
482: 
483: A question subject to recent debate concerns the form of 
484: nonlinearities associated with $E$\cite{debate1,debate2}.
485: That is an important issue because the nonlinearities decide 
486: to which universality class of critical behavior the system belongs.
487: It is interesting to see what the present method says about that.
488: First, we consider a one-dimensional, simplified version in which
489: the particles are not interacting except being hard-core, 
490: but their hoppings to nearest neighbors are biased by having
491: different jump rates, $p$ and $q$, to the right and left respectively.
492: Hence, the master equation reads
493: \begin{eqnarray}
494: && P_+(x;t+1)-P_+(x;t)=
495: \nonumber\\ &&\qquad  
496: p P_{+-}(x-1;t) + q P_{-+}(x;t)
497: \nonumber\\ &&\qquad  
498: -p P_{+-}(x;t) -q P_{-+}(x-1;t),
499: \label{1Dddsmaster}
500: \end{eqnarray}
501: where as usual an up(down) spin corresponds to the occupation 
502: of a particle(hole), and joint probabilities such as 
503: $P_{+-}(x-1;t)$ means the probability
504: of finding a particle-hole pair at site $x-1$ and $x$.
505: After factorizations and applications of
506: (\ref{P2phi}) and (\ref{naive}), we readily find
507: \begin{equation}
508: {\partial \phi \over \partial t}=
509: {\cal D} {\partial^2 \phi \over \partial x^2} 
510: + {{\cal E}\over 2} {\partial \phi^2 \over \partial x}
511: \label{1Dddstdgl}
512: \end{equation}
513: where the diffusion coefficient is ${\cal D}=(p+q)/2$, as expected,
514: and the coefficient of driving is ${\cal E}=(p-q)$.
515: The nonlinear term is the same as 
516: in the `standard' field theoretic model\cite{ddsft} which was
517: proposed on grounds of symmetries.
518: As side remarks, note that we obtain the diffusion equation for $p=q$, 
519: and that $\cal E$ is smooth in the 
520: `infinite' drive limit ($p=1$, $q=0$) which is used in most
521: Monte Carlo simulations of driven diffusive systems. 
522: 
523: 
524: 
525: \subsection{2D driven lattice gas}
526: %-------------------------------------
527: 
528: Generalization of the previous result to
529: the 2D interacting driven lattice gas is immediate, 
530: despite the unpleasant fact that there are altogether 
531: 512 terms in the master equation.
532: In the presence of a drive $E$ along the $+y$ direction and 
533: attractive ($J>0$ in (\ref{ham})) interaction between particles,
534: the heat-bath rates for hoppings of particles
535: along and against the drive take the form
536: \begin{equation}
537: W_{n,\pm E}\equiv { 1 \over 1+e^{n\beta J \mp E \beta J} },
538: \label{w_ne}
539: \end{equation}
540: where the dimensionless $E$ ($0\leq E < \infty$)
541: represents the ``work done'' on the particle by the field. 
542: Obviously, the rates for hoppings perpendicular to $E$ are $W_{n,0}=W_n$.
543: 
544: Going through the same procedure as above,
545: we eventually obtain a kinetic equation which is in complete
546: agreement with the standard field theory of 
547: the driven diffusive system\cite{ddsft}: 
548: \begin{eqnarray}
549: {\partial \phi\over \partial t}
550: &=& 
551: -\left( \alpha_x    {\partial^4\over \partial x^4}
552:        +\alpha_{xy} {\partial^4\over \partial x^2\partial y^2}
553:        +\alpha_{y}  {\partial^4\over \partial y^4} \right)\phi
554: \nonumber \\ && 
555: +\left( r_x {\partial^2\over \partial x^2}
556:        +r_y {\partial^2\over \partial y^2}\right)\phi
557: +{1\over 6} \left( g_x{\partial^2\over \partial x^2}
558:                   +g_y{\partial^2\over \partial y^2}\right) \phi^3
559: \nonumber \\ &&
560: +{{\cal E}\over 2} {\partial \phi^2 \over \partial y}.
561: \label{dds2d}
562: \end{eqnarray}
563: The anisotropies are generated by the drive.
564: Excluding the last term, this is the anisotropic generalization
565: of the deterministic TDGL equation 
566: with conserved magnetization, i.e., model B\cite{hh}:
567: \begin{equation}
568: {\partial \phi\over \partial t} = 
569: \nabla^2\left( -\alpha \nabla^2\phi + r \phi +{g\over 6} \phi^3\right).
570: \label{modelB}
571: \end{equation}
572: All coefficients are determined:
573: \begin{eqnarray}
574: \alpha_x &=& 
575: {1\over 384}(69 - 85 W_4 - 68 W_8 - 17 W_{12}),
576: \\
577: \alpha_{xy} &=& 
578: {1\over 256}
579: (20 - 20\,W_4 - 16\,W_8 - 4\,W_{12} + W_{-12,-E} + 
580: \nonumber \\ &&
581: %\quad
582: W_{-12,E} + 4\,W_{-8,-E} + 4\,W_{-8,E} + 
583: \nonumber \\ &&
584: %\quad
585: 5\,W_{-4,-E} + 5\,W_{-4,E} - 5\,W_{4,-E} - 5\,W_{4,E} - 
586: \nonumber \\ &&
587: %\quad
588: 4\,W_{8,-E} - 4\,W_{8,E} - W_{12,-E} - W_{12,E}),
589: \label{dds2d_alpxy}
590: \\
591: \alpha_{y} &=& 
592: {1\over 768}
593: (8\,W_{-12,-E} + 8\,W_{-12,E} + 31\,W_{-8,-E} + 
594: \nonumber \\ &&
595: %\quad
596: 31\,W_{-8,E} + 35\,W_{-4,-E} + 35\,W_{-4,E} - 
597: \nonumber \\ &&
598: %\quad
599: 10\,W_{0,-E} - 10\,W_{0,E} - 50\,W_{4,-E} - 50\,W_{4,E} - 
600: \nonumber \\ &&
601: %\quad
602: 37\,W_{8,-E} - 37\,W_{8,E} - 9\,W_{12,-E} - 9\,W_{12,E}),
603: \label{dds2d_alpy}
604: \\
605: r_x &=& 
606: {1\over 32}(-9 + 25\,W_4 + 20\,W_8 + 5\,W_{12}),
607: \label{dds2d_rx}
608: \\
609: r_y &=& 
610: {1\over 64}
611: (-2\,W_{-12,-E} - 2\,W_{-12,E} - 7\,W_{-8,-E} - 
612: \nonumber \\ &&
613: %\quad
614: 7\,W_{-8,E} - 5\,W_{-4,-E} - 5\,W_{-4,E} + 10\,W_{0,-E} + 
615: \nonumber \\ &&
616: %\quad
617: 10\,W_{0,E} + 20\,W_{4,-E} + 20\,W_{4,E} + 13\,W_{8,-E} + 
618: \nonumber \\ &&
619: %\quad
620: 13\,W_{8,E} + 3\,W_{12,-E} + 3\,W_{12,E}),
621: \label{dds2d_ry}
622: \\
623: g_x &=& 
624: {5\over 16} ( 3 + W_4 - 4\,W_8 - 3\,W_{12} ),
625: \label{dds2d_gx}
626: \\
627: g_y &=& 
628: {1\over 32}
629: (6\,W_{-12,-E} + 6\,W_{-12,E} + 7\,W_{-8,-E} + 
630: \nonumber \\ &&
631: %\quad
632: 7\,W_{-8,E} - W_{-4,-E} - W_{-4,E} + 6\,W_{0,-E} + 
633: \nonumber \\ &&
634: %\quad
635: 6\,W_{0,E} + 4\,W_{4,-E} + 4\,W_{4,E} - 13\,W_{8,-E} - 
636: \nonumber \\ &&
637: %\quad
638: 13\,W_{8,E} - 9\,W_{12,-E} - 9\,W_{12,E})
639: \label{dds2d_gy}
640: \\
641: {\cal E}&=& 
642: {1\over 16}
643: (-W_{-12,-E} + W_{-12,E} - 3\,W_{-8,-E} + 
644: \nonumber \\ &&
645: %\quad
646: 3\,W_{-8,E} - 3\,W_{-4,-E} + 3\,W_{-4,E} - 
647: \nonumber \\ &&
648: %\quad
649: 2\,W_{0,-E} + 2\,W_{0,E} - 3\,W_{4,-E} + 3\,W_{4,E} - 
650: \nonumber \\ &&
651: %\quad
652: 3\,W_{8,-E} + 3\,W_{8,E} - W_{12,-E} + W_{12,E}).
653: \label{dds2d_e}
654: \end{eqnarray}
655: They have these important properties:
656: \begin{enumerate}
657: \item
658: All but $\cal E$ are even in $E$,  consistent with the
659: invariance of the dynamics under $\{E\to -E, y\to -y\}$.
660: \item
661: The quadratic coefficient $r_x$ is independent of $E$. 
662: It has one zero at $T_c^{\rm GL}=3.86143 J/k_B$, 
663: as shown in Fig.~\ref{fig:rx}.
664: In contrast, $r_y$ depends on $E$. 
665: Fig.~\ref{fig:ry} displays the behavior 
666: of $r_y(T,E)$ versus $r_x(T)$ as $T$ is
667: lowered from above $T_c^{\rm GL}$ at fixed $E$,
668: as well as $r_y(T=T_c^{\rm GL},E)$ versus $E$.
669: It shows that for any $E>0$, 
670: $r_x$ always vanishes before $r_y$ does when $T$ is decreased.
671: For small $E$, $r_y\approx r_x + c E^2$ where $c>0$.
672: Consequently, at the critical temperature, the dominant
673: derivatives come from the $r_y$ and $\alpha_x$ terms,
674: leading to the identification of an intrinsically anisotropic 
675: critical theory with scaling of momenta $k_y\sim k_x^2$. 
676: This agrees with a previous perturbative argument\cite{ddsft}. 
677: \item
678: The coefficient $\cal E$ of the leading nonlinearity induced
679: by the drive vanishes linearly in $E$ at small $E$, and saturates
680: to a constant at $E=\infty$. This dependence, exhibited in 
681: Fig.~\ref{fig:E}, is already anticipated above from the 1D model 
682: and argued previously\cite{debate2},
683: but at odds with the claims in \cite{debate1}.
684: \item
685: At $E=0$: we find $\alpha_x=\alpha_y \neq \alpha_{xy}$--
686: the continuum model derived is not rotationally invariant.
687: This is not surprising because the initial lattice model is not.
688: However, it turns out that $\alpha_{xy}=\alpha_x$ at $T_c^{\rm GL}$.
689: At present, we are not sure whether this acquirement
690: of higher symmetry at the critical point is general for
691: the underlying lattice model or specific to the method.
692: \end{enumerate}
693: 
694: 
695: 
696: 
697: \subsection{two-species driven lattice gas}
698: %-------------------------------------
699: 
700: In all the above examples, each site has only two local states
701: (spin up or down).
702: In the two-species driven lattice gas model\cite{shz}, 
703: motivated in part by multi-ionic conductors and traffic flow problems,
704: there are three possibilities: 
705: as a hole, or either of two types of particles.
706: The two types of particles are driven in opposite directions
707: as if they were oppositely charged and driven by an electric field,
708: with local particle densities denoted by $\rho_+$ and $\rho_-$.
709: Due to the extra local state, it is not easy to write
710: down the correct set of equations by symmetry and intuition alone.
711: One way to proceed\cite{shz} 
712: is to express the entropy in terms of 
713: $\rho_+$ and $\rho_-$, and obtain the diffusion terms by 
714: functional differentiations. 
715: The driving terms can then be added to the kinetic equations
716: by generalizing that for the one-species model.
717: 
718: Another way is to apply the current method\cite{leungzia97}.
719: The equations derived are the same as in \cite{shz},
720: with the advantage of tractable microscopic origins 
721: in each coefficients.
722: Hence, this model further testifies the usefulness of the present 
723: approach when symmetry and intuition are not very helpful.
724: 
725: We end this section with a final remark.
726: Since the method begins with factorization of joint probabilities, 
727: the ensuing equation is deterministic, all information about
728: correlations seem to have lost. 
729: The situation can be remedied, however, by introducing a noise term 
730: to restore a probabilistic description. 
731: Correlations can then be computed by averaging over the noise
732: by means of standard field-theoretic techniques\cite{msr}.
733: For equilibrium systems, the noise can be fixed by 
734: requirements such as the fluctuation-dissipation theorem.
735: For nonequilibrium systems, there is no general rule. 
736: One usually has to extrapolate by analogy to equilibrium.
737: 
738: 
739: %-------------------------------------
740: \section{Conclusion}\label{CONCLUSION}
741: %-------------------------------------
742: 
743: We have presented in details a very simple and straight forward
744: method to derive the deterministic 
745: kinetic equations from known microscopic dynamics 
746: for stochastic, interacting systems.
747: The method has a mean-field flavor. 
748: It preserves the underlying symmetries of the dynamics
749: and is in line with the spirit of coarse graining.
750: The resulting equations are in good agreement with other
751: either more or less rigorous approaches,
752: as demonstrated explicitly {\em via\/} several examples.
753: Hence, despite the approximate and heuristic nature of our approach, 
754: it proves to be a useful and convenient means to obtain
755: a correct continuum theory, especially 
756: (i) when other more rigorous approaches do not apply;
757: (ii) when symmetries of the system is not intuitively obvious; and
758: (iii) when microscopic dependences of the continuum parameters are wanted.
759: 
760: 
761: 
762: \vspace{0.5cm}
763: \noindent{\bf{Acknowledgments}:}
764: This work is supported by the National Science Council of R.O.C.
765: under grant number NSC89-2112-M-001-015,
766: and a Main-Theme Grant of the Academia Sinica.
767: 
768: 
769: %-------------------------------------------
770: \appendix
771: \section{noise correlation}
772: %-------------------------------------------
773: 
774: Since the present method only gives
775: the deterministic part of the kinetic equation,
776: the noise term has to be considered separately.  
777: Here we follow the common practice to assume that 
778: the noise $\zeta$ is Gaussianly distributed 
779: and correlated over negligible ranges.
780: Then the only question is to determine
781: its correlation $D$ in (\ref{noisecorrel}).
782: 
783: There are two ways to do it. The first makes use of the 
784: correspondence between the Langevin equation 
785: \begin{eqnarray}
786: {\partial \phi\over \partial t}&=&
787: -\Gamma {\delta {\cal H}\over \delta \phi} + \zeta
788: \label{app:tdgl} \\
789: \langle \zeta(\vec{r},t)\zeta(\vec{r}',t') \rangle&=&
790: 2{D} \delta(\vec{r}-\vec{r}')\delta(t-t'),
791: \label{app:noise}
792: \end{eqnarray}
793: and the Fokker-Planck equation
794: \begin{equation}
795: {\partial {\cal P}\over \partial t}=
796: -\int \, d^dx {\delta \over \delta \phi} 
797: \left( -\Gamma {\delta {\cal H} \over \delta \phi} {\cal P} 
798: -{D} {\delta {\cal P}\over \delta \phi} \right),
799: \label{app:FP}
800: \end{equation}
801: which is a continuity equation.
802: A stationary solution of the Fokker-Planck equation is obtained by
803: setting zero the probability current, i.e., $(\cdots )=0$, which gives
804: ${\cal P} \propto e^{-\Gamma {\cal H}/{D}}$.
805: Since the free energy 
806: ${\cal F}[h]$ in the presence of an external field $h$
807: is of the form ${\cal F}[h]={\cal F}[0] - h \phi$,
808: it differs from $\cal H$ by a factor of $\mu$.
809: Hence, by matching $e^{-{\cal F}[h]/k_BT}$
810: and $e^{-\Gamma {\cal H}[h]/{D}}$,
811: we deduce that
812: \begin{equation}
813: {D}=k_B T\mu\Gamma.
814: \label{app:N}
815: \end{equation}
816: In passing, it is worth noting that 
817: the kinetic coefficient defined in 
818: \begin{equation}
819: {\partial \phi\over \partial t}=
820: -\lambda {\delta {\cal F}\over \delta \phi} + \zeta
821: \label{app:tdgl2}
822: \end{equation}
823: is $\lambda={D}/k_B T$: the Einstein relation.
824: 
825: An alternative way to determine $D$ 
826: is to use the fluctuation-dissipation theorem, 
827: which in momentum-frequency space takes the form 
828: \begin{equation}
829: { 2 k_B T\over \omega} {\rm Im} \,\chi (k,\omega) = G (k,\omega).
830: \label{app:FDT}
831: \end{equation}
832: Although neither the susceptibility $\chi$ nor the two-point correlation
833: function $G$ can be calculated in closed form for general
834: $\cal H$, (\ref{app:FDT}) holds order by order so that we only
835: need to consider the Gaussian model in the case of $g=0$.
836: Thus, by Fourier transforms, we obtain
837: $\chi(k,\omega)=\Gamma \mu/[-i\omega + \Gamma (k^2+r)]$
838: and $G(k,\omega)=2{D}/[\omega^2 + \Gamma^2 (k^2+r)^2]$,
839: which by virtue of (\ref{app:FDT}) also gives (\ref{app:N}).
840: 
841: %Since the deductions require the knowledge of 
842: %the equilibrium distribution,
843: %they do not apply to nonequilibrium
844: %situations\cite{ness}.  
845: %In those cases, 
846: %there is no general rule as to how the noise should be fixed.
847: %Sometimes an extrapolation from equilibrium analogy is all one can do.
848: 
849: 
850: 
851: %:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
852: % References:
853: %:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
854: \begin{references}
855: 
856: \bibitem[*]{email_ktl} E-mail: leungkt@phys.sinica.edu.tw
857: 
858: \bibitem
859: {langevin}
860: P. Langevin, Comptes Rendus de l'Acad{\'e}mie des Sciences (Paris)
861: {\bf 146}, 530 (1908).
862: 
863: \bibitem
864: {vankampen}
865: See, e.g.,
866: N.G. van Kampen, {\it Stochastic Processes in Physics and Chemistry},
867: (North-Holland, Amsterdam, 1981).
868: 
869: \bibitem
870: {hh}
871: For a review, see 
872: P.C. Hohenberg and B.I. Halperin, Rev. Mod. Phys. {\bf 49}, 435 (1977).
873: 
874: \bibitem
875: {surface}
876: M. Kardar, G. Parisi, and Y.C. Zhang, Phys. Rev. Lett. {\bf 56}, 889 (1986).
877: 
878: \bibitem
879: {pattern}
880: J.S. Langer, in Les Houches XLVI 1986, {\it Chance and Matter},
881: ed. by J.Souleti, J. Anninemus, and R. Stora, 
882: Elsevier Science Pub. B.V.(1987);
883: M.C. Cross and P.C. Hohenberg, 
884: %``Pattern formation outside of equilibrium'', 
885: Reviews of Modern Physics, {\bf 65}, Issue 3, pp. 851-1112 (1993).
886: 
887: \bibitem
888: {Glauber}
889: R.J. Glauber, J. Math. Phys. {\bf 4}, 294 (1963).
890: 
891: \bibitem
892: {kawasaki}
893: K. Kawasaki, in {\em Phase Transitions and Critical Phenomena} Vol.~2, 
894: eds. C. Domb and M. S. Green, (Academic, NY, 1972).
895: 
896: \bibitem
897: {amit}
898: See e.g., D.J. Amit, {\it Field Theory, the Renormalization Group,
899: and Critical Phenomena}, 2nd ed. (World Scientific, Singapore, 1984).
900: 
901: \bibitem
902: {langer71}
903: J.S. Langer,  Ann. Phys. (NY) {\bf 65}, 53 (1971).
904: 
905: \bibitem
906: {omegaexp}
907: The Langevin equation can also be obtained by correspondence
908: with the Fokker-Planck equation which
909: can be derived from an expansion of the master equation 
910: in inverse powers of the system volume.  
911: However,  we are not aware of such a self-contained and explicit
912: derivation which gives the Ginzburg-Landau form without making
913: any assumption.
914: 
915: \bibitem
916: {ness}
917: By `generic' we refer to systems which have nonzero probability current
918: in configuration space (the terms enclosed by brackets in (\ref{app:FP})) 
919: even in stationary states. In the master equation,
920: this corresponds to the breakdown of detailed balance, i.e., 
921: the right-hand side of (\ref{master}) does not vanish 
922: term-by-term in stationary states but only after summation.
923: Usually this is also accompanied by a breakdown 
924: of the fluctuation-dissipation theorem, and
925: the lack of a well-defined free energy.
926: 
927: \bibitem
928: {intuitive}
929: For critical properties, this is not a serious problem because
930: the relevant physics is in the long-wavelength, long-time
931: limits, where by virtue of universality, 
932: a kinetic equation postulated solely by 
933: symmetry considerations is sufficient.  
934: However, this often requires physical intuition and insight
935: whereas the present method does not.
936: 
937: \bibitem
938: {truncate}
939: Due to factorizations, terms of $O(\phi^n)$ and higher are absent
940: where $n$ is the number of spins involved in the energetics in $w$
941: of (\ref{master}). For Glauber dynamics, $n=2d+1$.
942: 
943: \bibitem
944: {leungneda}
945: K.-t. Leung and Z. N{\'e}da,  Phys. Lett. A {\bf 246}, 505 (1998);
946: Phys. Rev. E {\bf 59}, 2730 (1999).
947: 
948: %\bibitem
949: %{rates}
950: %Temperature dependences of the parameters are specific to the choice
951: %of the jump rate.  Thus, if Metropolis rate is used, we find
952: %$g$ has a zero for 2D Ising model, but the coefficient of $\phi^5$ 
953: %is positive throughout. Similarly for 3D Ising model.
954: %While there is no problem of stability, the
955: %critical temperature becomes less accurate.
956: %As in most analytic schemes, the 
957: %heat-bath rate is preferable for its analyticity in its argument.  
958: 
959: \bibitem
960: {hte}
961: See, e.g., B. Dammann and J.D. Reger, Europhys. Lett. {\bf 21} 157 (1993);
962: M.F. Sykes, D.S. Gaunt, P.D. Roberts and J.A. Wyles,
963: J. Phys. A {\bf 5} 624 (1972).
964: 
965: \bibitem
966: {landau}
967: See, e.g., D.P. Landau, Physica A {\bf 205}, 41 (1994).
968: 
969: \bibitem
970: {dds}
971: S. Katz, J.L. Lebowitz, and H. Spohn, Phys. Rev. B {\bf28}, 1655 (1983);
972: B. Schmittmann and R.K.P. Zia, in: {\em Phase Transitions
973: and Critical Phenomena} Vol. 17, eds. C. Domb and J.L. Lebowitz, 
974: (Academic Press, N.Y., 1995), and refs. therein.
975: 
976: \bibitem
977: {debate1}
978: P.L. Garrido, F. de los Santos, and M.A. Mu\~{n}oz, Phys.
979: Rev. E{\bf 57}, 752 (1998);
980: F. de los Santos and P.L. Garrido, J. Stat. Phys. {\bf 96},
981: 303 (1999);
982: F. de los Santos and M.A. Mu\~{n}oz, Phys. Rev.
983: E {\bf 61}, 1161 (2000).
984: 
985: \bibitem
986: {debate2}
987: B. Schmittmann, H.K. Janssen, U.C. T\"{a}uber, R.K.P. Zia,
988: K.-t. Leung, and J.L. Cardy,
989: Phys. Rev. E {\bf 61}, 5977 (2000).
990: 
991: \bibitem
992: {ddsft}
993: K.-t.~Leung and J.L. Cardy, J. Stat. Phys. {\bf44}, 567 (1986)
994: and {\bf 45}, 1087 (1986);
995: H.K. Janssen and B. Schmittmann, Z. Phys. {\bf64}, 503 (1986).
996: 
997: \bibitem
998: {shz}
999: B. Schmittmann, K. Hwang and R.K.P. Zia, Europhys. Lett. {\bf 19}, 19 (1992).
1000: 
1001: \bibitem
1002: {leungzia97}
1003: K.-t. Leung and R.K.P. Zia, Phys. Rev. E {\bf 56}, 308 (1997).
1004: 
1005: \bibitem
1006: {msr}
1007: H. K. Janssen, Z. Phys. B {\bf23}, 377 (1976);
1008: C. DeDominicis, J. Physique {\bf37}, C1-247 (1976).
1009: 
1010: 
1011: \end{references}
1012: 
1013: %:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
1014: % Figures Captions:
1015: %:::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::::
1016: %\onecolumn
1017: %\vspace{2cm}
1018: 
1019: \begin{figure}[htp]
1020: \epsfig{figure=fig1.eps,width=3.0in,angle=-0}
1021: \caption{
1022: Quadratic coefficient in the transverse direction $r_x$ 
1023: plotted vs. temperature. 
1024: Its zero locates the critical temperature $T_c^{\rm GL}$.
1025: }
1026: \label{fig:rx}
1027: \end{figure}
1028: 
1029: 
1030: \begin{figure}[htp]
1031: \epsfig{figure=fig2a.eps,width=3.0in,angle=-0}
1032: \epsfig{figure=fig2b.eps,width=3.0in,angle=-0}
1033: \caption{
1034: (a) Trends of $r_y$ vs. $r_x$ as $T$ is varied across $T_c^{\rm GL}$
1035: at fixed $E$. From bottom to top: $E=0$, 2, 4, 6, 10, 14, 20, and 50.
1036: (b) Intercept $r_y(r_x=0)$ in (a) plotted vs. $E$.
1037: }
1038: \label{fig:ry}
1039: \end{figure}
1040: 
1041: 
1042: \begin{figure}[htp]
1043: \epsfig{figure=fig3.eps,width=3.0in,angle=-0}
1044: \caption{
1045: The coefficient of the leading nonlinearity, $\cal E$ vs. 
1046: the microscopic drive $E$ at
1047: different temperatures. From top to bottom:
1048: $k_B T/J=1$, 3, $3.86143(=T_c^{\rm GL}$), 5, 10, 20 and 50. 
1049: }
1050: \label{fig:E}
1051: \end{figure}
1052: 
1053: \end{document}
1054: 
1055: