cond-mat0302232/hzs.tex
1: %% This document created by Scientific Word (R) Version 2.0
2: %for good reproduction on cond-mat, include epsf.tex into .tar.gz file
3: \documentclass[12pt]{iopart}
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: %TCIDATA{Created=Wed Aug 01 18:28:30 2001}
6: %TCIDATA{LastRevised=Thu Aug 02 16:39:09 2001}
7: 
8: %\input{tcilatex}
9: \begin{document}
10: 
11: \title[Effects of differential mobility on biased diffusion of two species]
12: {Effects of differential mobility on biased diffusion of two species}
13: \author{R.S. Hipolito, R.K.P. Zia and B. Schmittmann}
14: 
15: \date{\today }
16: 
17: \begin{abstract}
18: Using simulations and a simple mean-field theory, we investigate 
19: jamming transitions in a two-species lattice gas under non-equilibrium
20: steady-state conditions. The two types of particles diffuse with different 
21: mobilities on a square lattice, subject to an excluded volume constraint 
22: and biased in opposite directions. Varying filling fraction, differential 
23: mobility, and drive, we map out the phase diagram, identifying first order and 
24: continuous transitions between a free-flowing disordered and a spatially 
25: inhomogeneous jammed phase. Ordered structures are observed to drift, with
26: a characteristic velocity, in the direction of the more mobile species. 
27: \end{abstract}
28: 
29: \address{Center for Stochastic Processes in Science and Engineering, Physics Department \\ 
30: 	Virginia Polytechnic Institute and State University, \\ 
31: 	Blacksburg, VA 24061-0435, USA}
32: 
33: \jl{1}
34: 
35: %\\Key Words: Driven lattice gases, phase transitions, non-equilibrium.
36: %\\Contact: B. Schmittmann, e-mail: SCHMITTM@vt.edu; +1 540 231-6518, 
37: %	FAX: +1 540 231-7511.\ 
38: 
39: %\submitted
40: %\maketitle
41: 
42: 
43: \section{Introduction}
44: 
45: Since their introduction, driven lattice gases have attracted much attention
46: as some of the simplest systems for exploring non-equilibrium steady states %
47: \cite{KLS,DL17}. Motivated by the physics of fast ionic conductors \cite{FIC}%
48: , the earliest models consisted of a single species of ``charged''
49: particles, diffusing on a periodic lattice, subject to the effects of both a
50: thermal bath and an external (uniform, DC) ``electric'' field. Driven by the
51: field, the system settles into a state with non-trivial current. With no
52: inter-particle interactions other than an excluded volume constraint, the 
53: {\em stationary state} distribution is trivial \cite{Spitzer}, and
54: interesting behavior is displayed only in {\em time-dependent} quantities.
55: However, once interactions are introduced, then a variety of unexpected
56: phenomena arise even in the steady state, such as long-range correlations at 
57: {\em all }temperatures \cite{LRC}.
58: 
59: A natural generalization is to study systems with two species, motivated by,
60: e.g., fast ionic conductors with several mobile species \cite{FIC}. Such
61: systems have been used to model water droplets in microemulsions with
62: distinct charges \cite{water-in-oil}, gel electrophoresis \cite%
63: {gel-electro,Mukamel-gel}, vacancy mediated diffusion in binary alloys \cite%
64: {VMD} and traffic flow \cite{traffic}. Focusing on systems with two species
65: carrying equal but opposite ``charge'', we discovered remarkably complex
66: steady states, even for particles subjected only to the excluded volume
67: constraint \cite{HSZ,BSZ}. For simplicity, the early studies employed only
68: square lattices of $L\times L$, filled with equal numbers of ``positive''
69: and ``negative'' particles. The control parameters in this simple system are
70: just $E$, the strength of the drive, and $\bar{m}$, the overall particle
71: density (regardless of charge). By varying these parameters, we found that
72: this model displays a phase transition \cite{HSZ}, from a disordered state
73: with homogeneous densities at small $E$ or $\bar{m}$, to an ordered state
74: with inhomogeneous densities. Moreover, this line in the $E$-$\bar{m}$ plane
75: consists of a section with discontinuous transitions and another with
76: continuous ones, which may be regarded as analogues of first and second
77: order transitions in equilibrium systems. To understand these phenomena, it
78: is sufficient to adopt a mean-field, continuum approach, based on
79: hydrodynamic like equations for the two conserved densities. A linear
80: stability analysis around the homogeneous state allows us to predict the 
81: {\em existence} of transitions. Fortunately, we were able to find {\em %
82: analytic }solutions for both, homogeneous and inhomogeneous steady states %
83: \cite{VZS}, so that the presence of {\em both types} of transitions can be
84: predicted. Subsequently, such models have been generalized to include
85: rectangular lattices \cite{BSZ}, non-neutral \cite{LZ} or nearly filled \cite%
86: {ST} systems, and ``charge exchange'' dynamics \cite{KSZ97} in
87: one-dimensional \cite{Bonn} and quasi-one-dimensional \cite{KSZ99, MSZ}
88: cases. A dazzling array of surprising phenomena was discovered.
89: 
90: Here, we investigate a further generalization, namely, having species with
91: different mobilities. Indeed, unless there are some underlying symmetry
92: constraints, there is no reason to expect two different species of particles
93: in typical physical systems to have identical mobilities. Using both Monte
94: Carlo techniques and mean-field analyses, we find that, though the phase
95: diagram appears to suffer little change, novel features arise in all regions
96: of parameter space, even at not-so-subtle levels. In the next section, we
97: present some details of how a differential mobility is implemented in
98: simulations, followed by a discussion of our results. Section 3 is devoted
99: to the continuum, mean-field theory and its predictions. We close with a
100: summary of our findings and discuss some possible avenues for future
101: investigations.
102: 
103: \section{Model Specification and Simulation Results}
104: 
105: Following the first studies of biased diffusion of two species, square
106: lattices of size $L\times L\equiv N$ are used. In most of our simulations, $%
107: L=$ $40$. We focus exclusively on neutral systems (i.e., $N_{+}=N_{-}$) at
108: various densities $\bar{m}\equiv \left( N_{+}+N_{-}\right) /N$. A
109: configuration of the system is specified by the set $\left\{ \sigma
110: _{x,y}\right\} $, where $\sigma _{x,y}$ assumes the values $0,\pm 1$ if the
111: site $\left( x,y\right) $ is empty or occupied by a positive ($+$) or
112: negative ($-$) particle. Clearly, $\bar{m}=\sum_{x,y}\sigma _{x,y}^{2}/N$.
113: The diffusive nature of the the particles is modeled by allowing them to
114: jump only to nearest-neighbor empty sites. After a particle and one of its
115: neighboring site are selected, a particle-hole pair is always exchanged,
116: unless it results in a particle jumping against the external ``electric''
117: field. Chosen to point in the positive $y$ axis, the field is parametrized
118: by $\vec{E}=E\hat{y}$, so that its sole effect is to {\em lower} the
119: probability of particle jumps against $\vec{E}$ to $e^{-E}$.
120: 
121: In contrast to previous studies, the two species are endowed with {\em %
122: different mobilities}, modeling say, trucks and cars on a road. Using the
123: self-evident notation $\mu _{\pm }$ for the mobilities, we arbitrarily
124: choose $\mu _{+}\geq \mu _{-}$ and then focus on the ratio $\mu \equiv \mu
125: _{+}/\mu _{-}$. To implement this ratio, we keep a list of the particles and
126: their locations. In one Monte Carlo step (MCS), $\bar{m}N$ particles are
127: chosen from the positive/negative list randomly with the frequency ratio of $%
128: \mu $. In particular, by focusing on integer values of $\mu $, we simply
129: make $\mu $ attempts to move a positive particle for every attempt to move a
130: negative one. By monitoring a few quantities, we found that there is little
131: difference between sequential and random updating in species space,
132: especially since the particle locations were randomly assigned at the
133: beginning. After $\bar{m}N$ attempts, we increment the time by 1 MCS. Below,
134: we will see that it is often convenient to use the ``renormalized'' MCS (1
135: RMCS being defined as $\mu +1$ MCS), during which every slow particle would
136: have had, on the average, one chance to move. Typically, 20K MCS or more are
137: discarded before measurements are taken. In this study, we probe $\mu
138: =1,3,10,30,100,$ and $300$, though most of the data are collected for $\mu
139: =30$, being a compromise between exploring large differentials and dealing
140: with finite computation times.
141: 
142: \subsection{Continuous and discontinuous phase transitions}
143: 
144: Similar to the $\mu =1$ system, ours displays a transition from a disordered
145: homogeneous to an ordered inhomogeneous state, the latter showing a
146: marcoscopic cluster with opposing particles locked in a ``jam.'' To
147: differentiate these phases, we monitor the first few {\em mass }structure
148: factors, defined as 
149: \[
150: S_{n}\equiv \left\langle \left| \tilde{m}_{n}\right| ^{2}\right\rangle , 
151: \]
152: where 
153: \[
154: \tilde{m}_{n}\equiv L^{-2}\sum_{x,y}\sigma _{x,y}^{2}\exp \left[ 2\pi iny/L%
155: \right] 
156: \]
157: is the Fourier tranform of the mass density profile across $y$. Here, $%
158: \left\langle \cdot \right\rangle $ denotes the average over 800
159: configurations, separated from one another by 100 MCS in a typical run of
160: 80K MCS. We also measure the fluctuations in the $S$'s by monitoring the
161: variances: $\left\langle \left[ \left| \tilde{m}_{n}\right| ^{2}-S_{n}\right]
162: ^{2}\right\rangle $. Both continuous and discontinuous transitions are
163: observed, as illustrated by a phase diagram in the $E$-$\bar{m}$ plane for
164: the $\mu =30$ case (Fig.~1). The insets illustrate the nature of the
165: transitions, in two typical parameter domains. 
166: 
167: %===================================
168: \begin{figure}[tbp]
169: \input{epsf}
170: \par
171: \begin{center}
172: \begin{minipage}{.5\textwidth}
173:     \vspace{-1.cm}
174:   \epsfxsize = \textwidth \epsfysize = .9\textwidth \hfill
175:   \epsfbox{./hzs-fig1.eps}
176:     \vspace{-1.cm}
177: \end{minipage}
178: \end{center}
179: \caption{Phase diagram for mobility ratio of 30, with the system being homogeneous
180: and ``jammed'' in the lower-left and upper-right regions, respectively. The
181: solid line marks continuous transitions. For discontinuous transitions, a
182: pair of symbols ($+/-$) are used to denote the jumps in the order parameter
183: when the filling fraction ($\bar{m}$) is increased/decreased. The
184: absence/presence of hysteresis loops is illustrated by the insets, each
185: containing data for sweeps in {\em both }directions (cf. Fig.~2 for further
186: explanations).}
187: \end{figure}
188: %==============================================================
189: 
190: 
191: To provide further detail on how these lines are determined, we expand the
192: insets of Fig.~1 into Fig.~2a,b. Here, we plot the variations in $S_{1}$ as
193: the system is swept {\em back and forth} across the transitions. For the
194: continuous case (Fig.~2a), we show the result from a sweep in $E\in \left[
195: 0,0.5\right] $ while $\bar{m}$ is held at $0.6$. This direction of
196: traversing the transition line is chosen because this line is almost
197: parallel to the $\bar{m}$-axis in this region. Note that there are two sets
198: of data points ($\times /-$), representing the up/down sweeps from a single
199: long run. This run is started with $E=0$ in a random configuration at just
200: over half filling ($\bar{m}=0.6$). Then, the first 20K MCS are discarded,
201: data gathered for the next 80K MCS, $E$ raised by $0.02$, 20K discarded and
202: so on, until $E=1.0$. Thereafter, the process is reversed, decreasing $E$
203: until $E=0$. The figure shows clearly that the transition is continuous with
204: no sign of hysteresis. At the crude level of our investigations, the
205: critical field is then identified through either the largest $\partial
206: S_{1}/\partial E$ or the peak in the variance of $S_{1}$. The transition
207: line is assembled by repeating such runs for $\bar{m}\in \left[ 0.5,0.95%
208: \right] $ in steps of $0.05$. For the discontinuous cases, which are more
209: easily probed by sweeping in $\bar{m}$, we observe the typical hysteresis
210: loops. The inset in Fig.~1 and Fig.~2b show the case of $E=2$. Again, the
211: data represent a single long run, starting with $\bar{m}=0.1$ in a random
212: configuration, discarding the first 20K MCS, gathering data for the next 80K
213: MCS as above, raising $\bar{m}$ by $0.01$ and so on. These data points are
214: represented by crosses ($\times $). After a run at $\bar{m}=0.4$, we
215: continue this process with decreasing $\bar{m}$'s until $\bar{m}=0.1$. This
216: set of data is shown as minus signs ($-$). A hysteresis loop is clearly
217: displayed, and the densities at the jumps are recorded to construct the
218: points plotted in the main figure of Fig.~1. For comparison with the
219: previous studies (i.e., the $\mu =1$ case), we also show the results of such
220: runs as solid lines in Fig.~2a,b. As we see, there is little difference
221: between the $\mu =30$ and $\mu =1$ results, leading us to conclude that,
222: within the accuracy and statistics of our study, there is no significant
223: dependence of the phase diagram on $\mu$.
224: 
225: %===================================
226: \begin{figure}[tbp]
227: \input{epsf}
228: \par
229: \begin{center}
230: \vspace{-7.cm}
231: \begin{minipage}{.9\textwidth}
232:   \epsfxsize = \textwidth \epsfysize = .9\textwidth \hfill
233:   \epsfbox{./hzs-fig2.eps}
234:     \vspace{-1.cm}
235: \end{minipage}
236: \end{center}
237: \caption{Plots of $S_{1}$ as the system is swept across the transitions. (a) $E$ is
238: increased ($\times $) and then decreased ($-$), with $\bar{m}=0.6$ and $\mu
239: =30$. (b) $\bar{m}$ is increased ($\times $) and then decreased ($-$), with $%
240: E=2$ and $\mu =30$. For comparison, the $\mu =1$ cases are shown as solid
241: lines in both panels. With our statistics and accuracy, the effects of
242: mobility differential are relatively minor.}
243: \end{figure}
244: %==============================================================
245: 
246: Before turning to the properties of the inhomogeneous state, let us remark
247: that, of course, there {\em are} non-trivial effects due to $\mu >1$, even
248: in the disordered phase. Since one species is more mobile, there must be a
249: difference between the average velocities, which in turn leads to a non-zero 
250: {\em mass} current even though $N_{+}=N_{-}$. To verify this expectation, we
251: measure the two currents, $J_{\pm }$, by simply keeping track of the number
252: of $+/-$ particles which move up/down. Since $N_{+}=N_{-}$ in our samples,
253: the ratio of the average velocities is just $v_{+}/v_{-}=J_{+}/J_{-}$. With
254: limited observations, our naive expectation: 
255: \[
256: J_{+}/J_{-}=\mu 
257: \]%
258: turns out to be satisfied quite well. As the data in Table 1 show, we have
259: investigated only a few sample points, in both ``arms'' of the disordered
260: region (small $E$ moderate $\bar{m}$, and vice versa).
261: 
262: \begin{center}
263: \begin{tabular}{|r|r|r|r|r|r|}
264: \hline
265: $E$ & $\bar{m}$ & $\mu $ & $J_{+}/J_{-}$ & $J_{+}$ & $J_{-}$ \\ \hline
266: $0.1$ & $0.5$ & $3$ & $5.00$ & $0.22$ & $0.04$ \\ \hline
267: $0.2$ & $0.5$ & $3$ & $2.47$ & $0.39$ & $0.16$ \\ \hline
268: $0.1$ & $0.5$ & $30$ & $34.8$ & $2.35$ & $0.07$ \\ \hline
269: $0.2$ & $0.5$ & $30$ & $31.9$ & $4.44$ & $0.14$ \\ \hline
270: $2.0$ & $0.1$ & $3$ & $3.01$ & $1.13$ & $0.38$ \\ \hline
271: $2.0$ & $0.2$ & $3$ & $2.91$ & $1.81$ & $0.62$ \\ \hline
272: $2.0$ & $0.1$ & $30$ & $31.0$ & $11.1$ & $0.36$ \\ \hline
273: $2.0$ & $0.2$ & $30$ & $28.9$ & $18.4$ & $0.64$ \\ \hline
274: \end{tabular}
275: 
276: \vspace{0.5cm}
277: Table 1. Currents and their ratios for various regions in the homogeneous
278: phase.
279: 
280: \medskip
281: \end{center}
282: 
283: Naturally, we expect other interesting properties associated with the
284: disordered state; yet, these are generally quite subtle. For example, even
285: in the $\mu =1$ case, there are singular structure factors and long-range
286: correlations \cite{KSZ97b}. For more prominent and fascinating features, we
287: turn to the inhomogeneous state.
288: 
289: \subsection{Drifting structures in the ordered state}
290: 
291: Here, the density of particles is high enough so that they ``jam'' into a
292: macroscopic cluster. Consequently, the translational symmetry (in $y$) is
293: spontaneously broken. In neutral systems with $\mu =1$, the cluster performs
294: an {\em unbiased} random walk, and, due to the symmetry under charge
295: ``conjugation'' ($+\Leftrightarrow -$), its charge profile is purely
296: antisymmetric. On the other hand, for charged systems (i.e., $N_{+}\neq
297: N_{-} $) this cluster drifts ``{\em backwards}'' (opposite to the average
298: motion of the majority species)! Though counter-intuitive at first glance,
299: this behavior can be understood \cite{LZ}. Simulating $\mu >1$ systems here,
300: we find the cluster to drift ``{\em forwards},'' i.e., in the direction of
301: the more mobile species. Though this behavior may seem less surprising, it
302: is easy to advance a (wrong) heuristic argument to arrive at the
303: opposite conclusion!
304: 
305: %===================================
306: \begin{figure}[tbp]
307: \input{epsf}
308: \par
309: \begin{center}
310: \vspace{-1.cm}
311: \begin{minipage}{.5\textwidth}
312:   \epsfxsize = \textwidth \epsfysize = .9\textwidth \hfill
313:   \epsfbox{./hzs-fig3.eps}
314:     \vspace{-1.cm}
315: \end{minipage}
316: \end{center}
317: \caption{Drift speed as a function of $\mu $. The inset shows the ``center of mass''
318: of the macroscopic cluster, as a function of time, for three values of $\mu$.}
319: \end{figure}
320: %==============================================================
321: 
322: To characterize the drifting cluster more systematically, we carry out the
323: following analysis. Since the runs are necessarily long, we focus only on
324: neutral lattices with $\bar{m}=0.36$, driven with $E=1$. These parameters
325: put us well in the ordered phase. To study the $\mu $-dependence of the
326: drift speed, we perform simulations with $\mu \in \left[ 2,32\right] $ in
327: steps of $2$. For comparison with earlier studies, we also collect data for $%
328: \mu =1.$ Using a random configuration at the start, the usual first 20K MCS
329: are discarded. To be sure that a macroscopic cluster is present, we first
330: test the magnitude $\left| \tilde{m}_{1}\right| $ and proceed only when it
331: is larger than $0.25$ (c.f., $\left| \tilde{m}_{1}\right| =1\left/ \left[
332: L\sin (\pi /L)\right] \right. \simeq 0.32$ for a fully ordered state).
333: Assuming the center of mass (CM) of the cluster is well characterized by the
334: CM of the entire configuration, we measure the latter at intervals of 500
335: RMCS. Here, the CM of a particular configuration is defined simply by the
336: phase in $\tilde{m}_{1}=|\tilde{m}_{1}|e^{i\theta }$. Also, we find it more
337: convenient to use the RMCS as a unit, since the time scale of the system as
338: a whole is really set by the slow-moving particles. In the inset of Fig. 3,
339: we show how $\theta $ depends on time for three $\mu $'s. Clearly, the
340: dominant behavior is a steady drift. Fitting these data to $v_{\theta
341: }t+const$, we extract the drift speed $v_{\theta }$. Translating $v_{\theta
342: } $ into a velocity, $v$, in units of lattice spacing per 1000 RMCS, we plot 
343: $v\left( \mu \right) $ in Fig. 3. Until we have some understanding of this
344: quantity, we refrain from fitting it to a particular form. Instead, we just
345: note that, interestingly, $v$ seems to saturate at unity. Finally, we have
346: computed the standard deviations from the linear fits naively and found that
347: these are essentially independent of $\mu $. Of course, since we believe the
348: CMs to be performing (biased) random walks, we should expect these
349: deviations to grow as $t^{1/2}$, from which diffusion coefficients can be
350: extracted. Postponing such a detailed study to a later publication, what we
351: may infer from our ``naive'' computation is just that clusters associated
352: with different $\mu $'s execute random walks with the {\em same} diffusion
353: constant but {\em different} biases.
354: 
355: Next, we consider the mass and charge profiles in the jammed state. For each
356: configuration we measure both $\theta $ and 
357: \[
358: m\left( y\right) \equiv \sum_{y}\sigma _{x,y}^{2} \quad and \quad 
359: q\left( y\right) \equiv \sum_{y}\sigma _{x,y}\,\,. 
360: \]%
361: Then we displace the $y$-dependent quantities according to $y\rightarrow
362: u\equiv y-\theta L/2\pi +L/2$, so that the CM is located at $L/2=20$ for
363: convenience. In this co-moving frame, we can now average over the run to
364: produce the profiles: $\left\langle m\left( u\right) \right\rangle $ and $%
365: \left\langle q\left( u\right) \right\rangle $. Plotted in Fig. 4 are these
366: profiles for the parameter set $\left( E,\bar{m},\mu \right) =\left(
367: 1,0.36,30\right) $, and, for comparison with previous results, those for the 
368: $\mu =1$ case as well. Even though the change in the mass profile appears
369: almost imperceptible, there are actually some significant differences. For
370: example, though the densities within (outside) the cluster differ by only
371: about $+0.003\left( -0.003\right) $, this discrepancy translates into a {\em %
372: factor of two} for the densities far from the ``jam.'' Until we have much
373: better statistics, it is unclear if we can attach much meaning to these
374: subtle differences. The changes in the charge profile are more discernable.
375: We may interpret these as an enhancement of the more mobile species' ability
376: to penetrate into the block of the less mobile one. 
377: 
378: 
379: %===================================
380: \begin{figure}[tbp]
381: \input{epsf}
382: \par
383: \begin{center}
384: \vspace{1.cm}
385: \begin{minipage}{.5\textwidth}
386:   \epsfxsize = \textwidth \epsfysize = .9\textwidth \hfill
387:   \epsfbox{./hzs-fig4.eps}
388:     \vspace{-1.cm}
389: \end{minipage}
390: \end{center}
391: \caption{Mass (a) and charge (b) profiles in the ordered state. The dashed line and
392: the open circles correspond to the $\mu =1$ and $\mu =30$ cases,
393: respectively. Note that the charge profile for the latter is not purely
394: antisymmetric about the CM ($L/2=20$).}
395: \end{figure}
396: %==============================================================
397: 
398: 
399: \section{Mean-Field Theory and Comparisons}
400: 
401: To develop a first-step understanding of the large scale phenomena shown
402: above, we exploit the same mean field approach which proved quite successful
403: in the $\mu =1$ case \cite{HSZ,VZS}. In this approach, hydrodynamic like
404: equations of motion are postulated for the densities of the two species, $%
405: \rho _{\pm }\left( \vec{x},t\right) \equiv \left\langle \sigma ^{2}\left( 
406: \vec{x},t\right) \pm \sigma \left( \vec{x},t\right) \right\rangle /2$ : 
407: \[
408: \frac{\partial \rho _{\pm }}{\partial t}=-\vec{\nabla}\cdot \vec{J}_{\pm } 
409: \]
410: where $\vec{J}_{\pm }$ are the associated (density dependent) currents \cite%
411: {HSZ}. Here, we also treat the spatial co-ordinates as continuous variables,
412: in general $d$ dimensions: $\vec{x}=\left( x_{1},...,x_{d}\equiv y\right) $.
413: To model differential mobility, we simply multiply the currents by $\mu
414: _{\pm }$. Absorbing the average mobility into the time scale by defining 
415: \[
416: \mu _{\pm }\equiv \left( 1\pm \delta \right) \,\,, 
417: \]
418: the continuum equations read 
419: \begin{equation}
420: \frac{\partial \rho _{\pm }}{\partial t}=\left( 1\pm \delta \right) \vec{%
421: \nabla}\cdot \left[ \phi \vec{\nabla}\rho _{\pm }-\rho _{\pm }\vec{\nabla}%
422: \phi \mp \rho _{\pm }\phi {\cal E}{\bf \hat{y}}\right] \quad ,  \label{EoM}
423: \end{equation}
424: where 
425: \[
426: \phi \equiv 1-\rho _{+}-\rho _{-} 
427: \]
428: is the hole density and ${\cal E}$ represents the coarse-grained electric
429: field (${\cal E}=2\tanh \left( E/2\right) $ at the most naive level). Of
430: course, these equations can be re-expressed in terms of $\phi $ and the
431: charge density 
432: \[
433: \psi \equiv \rho _{+}-\rho _{-} 
434: \]
435: resulting in a form 
436: \begin{equation}
437: \frac{\partial }{\partial t}\left( 
438: \begin{array}{c}
439: \phi \\ 
440: \psi%
441: \end{array}
442: \right) =\left[ 
443: \begin{array}{cc}
444: 1 & -\delta \\ 
445: -\delta & 1%
446: \end{array}
447: \right] \left( 
448: \begin{array}{c}
449: \nabla \cdot \left[ \nabla \phi +\phi \psi {\cal E}{\bf \hat{y}}\right] \\ 
450: \nabla \cdot \left[ \phi \nabla \psi -\psi \nabla \phi -\phi (1-\phi ){\cal E%
451: }{\bf \hat{y}}\right]%
452: \end{array}
453: \right)  \label{phi-psi EoM}
454: \end{equation}
455: which easily reduces to the one studied previously \cite{HSZ,VZS}.
456: 
457: To find solutions to these equations, we must supply boundary conditions and
458: constraints. For a hypercubic system of volume $L^{d}$, they are 
459: \[
460: \rho _{\pm }\left( \vec{x},t\right) =\rho _{\pm }\left( \vec{x}+L\hat{x}%
461: _{i},t\right) \,\, , i=1,2,...,d 
462: \]
463: and 
464: \[
465: \int \frac{d\vec{x}}{L^{d}}\phi \left( \vec{x},t\right) =1-\bar{m},\quad
466: \int \frac{d\vec{x}}{L^{d}}\psi \left( \vec{x},t\right) =0. 
467: \]
468: We will discuss mainly steady states, i.e., density profiles which are time
469: independent in a {\em co-moving }frame. From Monte Carlo results as well as
470: previous experience with this model, we expect these profile to obey
471: translational invariance in the transverse directions in the whole parameter
472: space. On the other hand, spontaneous breaking of translational invariance
473: in $y$ is manifest for the jammed state. Due to the drift, we will denote
474: these profiles by $\phi \left( u\right) $ and $\psi \left( u\right) $, with $%
475: u=y-vt.$
476: 
477: \subsection{Homogeneous states and linear stability analysis}
478: 
479: Let us consider first the disordered phase, where the steady-state densities
480: are homogeneous in both space and time, i.e., 
481: \[
482: \phi \left( \vec{x},t\right) =1-\bar{m},\quad \psi \left( \vec{x},t\right)
483: =0. 
484: \]%
485: Note that, unlike the $\mu =1$ case, the mass current is no longer zero. The
486: (average), denoted by $C\hat{y}$, is given through 
487: \[
488: C\equiv \hat{y}\cdot \left( \vec{J}_{+}+\vec{J}_{-}\right) =\mu _{+}{\cal E}%
489: (1-\bar{m})\bar{m}/2-\mu _{-}{\cal E}(1-\bar{m})\bar{m}/2=\delta \bar{m}(1-%
490: \bar{m}){\cal E}. 
491: \]%
492: With our normalization of time scales, the charge current is unchanged: 
493: \[
494: J\equiv \hat{y}\cdot \left( \vec{J}_{+}-\vec{J}_{-}\right) =\bar{m}(1-\bar{m}%
495: ){\cal E}. 
496: \]%
497: Some care is needed when direct comparisons of absolute currents with Monte
498: Carlo data are made, since MCS and RMCS differ by a factor of $\mu +1$. One
499: test of this theory which is independent of such details is the ratio of
500: current magnitudes 
501: \[
502: J_{+}/J_{-}=\mu . 
503: \]%
504: As noted above, this prediction agrees well with our limited observations
505: (apart from the first set, in which both currents are very small and noisy).
506: Of course, our crude data should be regarded as merely the first step in a
507: more refined study. That they are consistent with $J_{+}/J_{-}=\mu $ just
508: confirms that this naive way of introducing mobility differentials is a
509: reasonable starting point. Noting that the agreement is somewhat worse for
510: high densities suggests the presence of subtle particle-particle
511: correlations, which can only be uncovered by better statistics.
512: 
513: Beyond these trivial results, a linear stability analysis around this state
514: leads us to the matrix 
515: \[
516: M=\left[ 
517: \begin{array}{cc}
518: 1 & -\delta \\ 
519: -\delta & 1%
520: \end{array}%
521: \right] \left[ 
522: \begin{array}{cc}
523: -k^{2} & ik_{y}{\cal E}(1-\bar{m}) \\ 
524: ik_{y}{\cal E}(2\bar{m}-1) & -k^{2}(1-\bar{m})%
525: \end{array}%
526: \right] 
527: \]%
528: which is purposely written in a form to display the effects of $\delta >0$.
529: In contrast to the $\delta =0$ case, the eigenvalues ($\lambda _{1,2}$) of
530: this $M$ are generically complex. Thus, we should examine their real
531: parts to identify the stability limit. However, since $M$ is
532: real, it can be expressed as $\left| \lambda _{1}\right| ^{2}%
533: %TCIMACRO{\func{Re}}%
534: %BeginExpansion
535: \mathop{\rm Re}%
536: %EndExpansion
537: \lambda _{2}/%
538: %TCIMACRO{\func{Re}}%
539: %BeginExpansion
540: \mathop{\rm Re}%
541: %EndExpansion
542: \lambda _{1}$ (or $1\Leftrightarrow 2$), so that $\det M=0$ is still
543: valid for identifying the stability limit. Since $\det M$ is modified
544: by a simple factor of $\left( 1-\delta ^{2}\right) $, we conclude that
545: differential mobility has {\em no effect} on this line! In other words, the
546: most unstable perturbation is still given by the lowest longitudinal mode: 
547: \[
548: k_{y}=2\pi /L 
549: \]%
550: and the instability occurs at \cite{HSZ} 
551: \[
552: \left( {\cal E}L/2\pi \right) ^{2}(2\bar{m}-1)=1. 
553: \]%
554: Of course we cannot expect good quantitative comparisons with the phase
555: boundaries from Monte Carlo simulations. For example, in the Ising model,
556: mean-field analysis overestimates the critical temperature by nearly a
557: factor of 2. Nevertheless, we are encouraged by one aspect of this
558: prediction, namely, that differential mobility has no effect on the onset of
559: the instability. This seems to be consistent with the observed phase
560: boundaries being quite insensitive to $\mu $.
561: 
562: For completeness, we give the explicit expressions for the eigenvalues: 
563: \begin{eqnarray}
564: \lambda _{1,2}=-k^{2}(1-\bar{\rho})-\delta ik_{y}{\cal E}\bar{\rho} 
565: \pm \sqrt{R}
566: \end{eqnarray}
567: where
568: \begin{eqnarray}
569: R = k^{4}\left[ 
570: \bar{\rho}^{2}+\delta ^{2}(1-\bar{m})\right] &-& (k_{y}{\cal E})^{2}\left[
571: (1-\delta ^{2})(2\bar{m}-1)(1-\bar{m})+\delta ^{2}\bar{\rho}^{2}\right] 
572: \nonumber \\
573:  &+& 2\delta ik^{2}k_{y}{\cal E}\bar{\rho}(1-\bar{\rho})
574: \end{eqnarray}
575: and $\bar{\rho}\equiv \bar{m}/2$. Their complex nature reflects drifts
576: associated with the eigen-perturbations. Given that the species are mobile
577: in different ways, such drifts are to be expected. In principle, it is
578: possible to measure the decay and drift of small perturbations in
579: simulations so that comparisons with these predictions can be made. In
580: practice however, such measurements would require better statistics and
581: longer runs than those in this study.
582: 
583: \subsection{Density profiles in inhomogeneous states}
584: 
585: Turning to the more interesting ordered states, the most immediate question
586: concerns the steady-state profiles. As observed in simulations and expected
587: from symmetry considerations, these profiles are homogeneous in $x$ but
588: depend on $y$. In analogy to the $\mu =1$ case, we seek non-trivial
589: functions $\phi ^{\ast }\left( y\right) $ and $\psi ^{\ast }\left( y\right) $
590: which satisfy Eqn. (\ref{phi-psi EoM}) in a steady state. However, it is
591: clear from Eqn. (\ref{EoM}) that if non-trivial functions lead to zeroes on
592: the right hand side with $\mu =1$, then they also lead to zeroes for any $%
593: \mu $! The conclusion is that our mean-field theory admits inhomogeneous
594: profiles which are (i) stationary 
595: (i.e., do not drift) and (ii) identical to those for $\mu =1$.
596: Of course, we may argue that this obvious discrepancy,
597: when comparing with data, is a ``small'' effect in absolute terms. However,
598: our equations are non-linear and the existence of additional, drifting
599: solutions cannot be ruled out. Unfortunately, we are unable so far, either
600: to find drifting solutions (analytically or numerically), or to prove that
601: our equations admit no such solutions. Nevertheless, for completeness, let
602: us present the analysis which simplifies the problem to a single second
603: order, non-linear, ordinary differential equation.
604: 
605: Following the techniques in \cite{LZ}, we assume the forms 
606: \[
607: \left( 
608: \begin{array}{c}
609: \phi ^{\ast }\left( \vec{x},t\right) \\ 
610: \psi ^{\ast }\left( \vec{x},t\right)%
611: \end{array}%
612: \right) =\left( 
613: \begin{array}{c}
614: \phi ^{\ast }\left( u\right) \\ 
615: \psi ^{\ast }\left( u\right)%
616: \end{array}%
617: \right) 
618: \]%
619: where 
620: \[
621: u\equiv y-vt, 
622: \]%
623: with $v$ being the drift velocity of the macroscopic cluster. Inserting
624: these into Eqn. (\ref{phi-psi EoM}), we obtain 
625: \[
626: -v\frac{d}{du}\left( 
627: \begin{array}{c}
628: \phi ^{\ast } \\ 
629: \psi ^{\ast }%
630: \end{array}%
631: \right) =\frac{d}{du}\left[ 
632: \begin{array}{cc}
633: 1 & -\delta \\ 
634: -\delta & 1%
635: \end{array}%
636: \right] \left( 
637: \begin{array}{c}
638: \left[ \left( d\phi ^{\ast }/du\right) +\phi ^{\ast }\psi ^{\ast }{\cal E}%
639: \right] \\ 
640: \left[ \phi ^{\ast }\left( d\psi ^{\ast }/du\right) -\psi ^{\ast }\left(
641: d\phi ^{\ast }/du\right) -\phi ^{\ast }(1-\phi ^{\ast }){\cal E}\right]%
642: \end{array}%
643: \right) , 
644: \]%
645: which has the form of a vanishing total derivative. So, the first integrals
646: are just constants which are identified with the two currents. 
647: As presented above, we denote the mass current $C$ and
648: the charge-current by $J$.
649: Recognizing that the {\em hole} current must be $-C$, we obtain 
650: \begin{eqnarray}
651: \left( 
652: \begin{array}{c}
653: C \\ 
654: -J%
655: \end{array}%
656: \right) &=& v\left( 
657: \begin{array}{c}
658: \phi ^{\ast } \\ 
659: \psi ^{\ast }%
660: \end{array}%
661: \right) 
662:  \\
663: &+& \left[ 
664: \begin{array}{cc}
665: 1 & -\delta \\ 
666: -\delta & 1%
667: \end{array}%
668: \right] \left( 
669: \begin{array}{c}
670: \left[ \left( d\phi ^{\ast }/du\right) +\phi ^{\ast }\psi ^{\ast }{\cal E}%
671: \right] \\ 
672: \left[ \phi ^{\ast }\left( d\psi ^{\ast }/du\right) -\psi ^{\ast }\left(
673: d\phi ^{\ast }/du\right) -\phi ^{\ast }(1-\phi ^{\ast }){\cal E}\right]%
674: \end{array}%
675: \right) . \nonumber 
676: \end{eqnarray}
677: To simplify these equations, we change variables, as in the $\delta =0$
678: case, to 
679: \[
680: \chi \equiv 1/\phi ^{\ast }\quad and \quad \omega \equiv \psi ^{\ast
681: }/\phi ^{\ast }. 
682: \]%
683: Then, we invert the matrix so that the first derivatives are decoupled: 
684: \[
685: \left( 
686: \begin{array}{c}
687: -d\chi /du+\omega {\cal E} \\ 
688: d\omega /du-(\chi -1){\cal E}%
689: \end{array}%
690: \right) =\frac{-\chi }{1-\delta ^{2}}\left[ 
691: \begin{array}{cc}
692: 1 & \delta \\ 
693: \delta & 1%
694: \end{array}%
695: \right] \left( 
696: \begin{array}{c}
697: -C\chi +v \\ 
698: J\chi +v\omega%
699: \end{array}%
700: \right) . 
701: \]%
702: Being first order differential equations, their full solutions will require
703: two more unknown integration constants. Together with $C$ and $J$, there are
704: four unknowns to be fixed, by the four constraint equations: 
705: \[
706: \chi \left( 0\right) =\chi \left( L\right) \quad and \quad \omega
707: \left( 0\right) =\omega \left( L\right) 
708: \]%
709: from periodic boundary conditions, as well as 
710: \[
711: \int_{0}^{L}\frac{1}{\chi }du=\int \phi ^{\ast }=L\left( 1-\bar{m}\right)
712: \quad and \quad \int \frac{\omega }{\chi }=\int \psi ^{\ast }=0 
713: \]%
714: from the conservation laws. Further simplifications occur when we define $%
715: C,J,v$ with appropriate factors of ${\cal E}$, $\delta $, and $1-\delta
716: ^{2}: $%
717: \begin{eqnarray*}
718: \frac{C}{1-\delta ^{2}} &\equiv &\delta {\cal E}\tilde{C} \\
719: \frac{J}{1-\delta ^{2}} &\equiv &{\cal E}\tilde{J} \\
720: \frac{v}{1-\delta ^{2}} &\equiv &\delta {\cal E}\tilde{v}
721: \end{eqnarray*}%
722: so as to exploit the scaling of $u$ to the variable $u{\cal E}$, and the
723: symmetry of the system under $\delta \Leftrightarrow -\delta $ (with $\tilde{%
724: C},\tilde{J},\tilde{v}$ even in $\delta $). Finally, denoting 
725: $d/d(u{\cal E})$ by prime ($^{\prime }$), we arrive at 
726: \begin{eqnarray}
727: \chi ^{\prime } &=&\omega +\delta \left( \tilde{J}-\tilde{C}\right) \chi
728: ^{2}+\delta \tilde{v}\chi \left( 1+\delta \omega \right)  \label{DEchi} \\
729: \omega ^{\prime } &=&(\chi -1)-\left( \tilde{J}-\delta ^{2}\tilde{C}\right)
730: \chi ^{2}-\delta \tilde{v}\chi \left( \delta +\omega \right) .
731: \label{DEomega}
732: \end{eqnarray}%
733: One advantage of this form lies in that, in the limit $\delta \rightarrow 0$%
734: , it easily reduces to the previous case: $\chi ^{\prime \prime }=(\chi -1)-%
735: \tilde{J}\chi ^{2}$. The other is the clear separation of $\chi ,\omega $
736: into even/odd functions of $\delta $. Note that, under this ``parity''
737: operation, $u{\cal E}$ is odd.
738: 
739: As in the $\delta =0$ case, the first equation can be trivially solved for $%
740: \omega $ and the resultant inserted into the second, leading us to a single
741: (though quite complicated) equation for $\chi .$ As pointed out above, there
742: exists at least {\em one} solution for any $\delta $, corresponding to a
743: stationary ($v=0$) state and involving $\tilde{C}=\tilde{J}$ (i.e., $%
744: J=\delta C$).
745: 
746: Although we have not found non-trivial drifting solutions analytically, we
747: are able to test the consistency of these equations against the measured
748: profiles and $v$. For example, we can begin with Eqn.(\ref{DEchi}) and
749: integrate over $u$ to get relationships between $v$ and averages of the
750: profiles. Indeed, there are infinitely many similar such relations, obtained
751: from first multiplying this equation by functions of $\chi $ (or $\phi
752: ^{\ast }$) before integration. To minimize possible large errors, we choose
753: to consider two equations so that the averages involve no higher power of $%
754: \chi $ than unity. Specifically, we {\em divide} Eqn.(\ref{DEchi}) by $\chi $
755: and $\chi ^{2}$ before integration. The results are 
756: \begin{eqnarray*}
757: 0 &=&\left\langle \psi \phi \right\rangle +\delta \left( \tilde{J}-\tilde{C}%
758: \right) +\delta \tilde{v}\left( 1-\bar{m}\right) \\
759: 0 &=&\delta \left( \tilde{J}-\tilde{C}\right) \left\langle \chi
760: \right\rangle +\delta \tilde{v}\left( 1+\delta \left\langle \omega
761: \right\rangle \right)
762: \end{eqnarray*}%
763: where the average is defined via 
764: \[
765: \left\langle \bullet \right\rangle \equiv \frac{1}{L}
766: \int_{0}^{L} \bullet \, du 
767: \]%
768: Eliminating the unknown currents, we arrive at 
769: \[
770: \frac{v/{\cal E}}{1-\delta ^{2}}=\frac{\left\langle \psi \phi \right\rangle
771: \left\langle \chi \right\rangle }{1-\left\langle \phi \right\rangle
772: \left\langle \chi \right\rangle +\delta \left\langle \psi ^{\ast }\chi
773: \right\rangle } 
774: \]%
775: To compare with data, we retrace the rescaling of time and re-introduce the
776: two mobilities, so that 
777: \[
778: v=\frac{2\mu _{+}\mu _{-}\left\langle \chi \right\rangle \left\langle \psi
779: ^{\ast }\phi ^{\ast }\right\rangle {\cal E}}{\left( \mu _{+}+\mu _{-}\right) %
780: \left[ 1-\left( 1-\bar{m}\right) \left\langle \chi \right\rangle \right]
781: +\left( \mu _{+}-\mu _{-}\right) \left\langle \psi ^{\ast }\chi
782: \right\rangle }. 
783: \]%
784: Setting $\mu _{+}=30$ and $\mu _{-}=1$ would correspond to using RMCS as a
785: unit of time. Using the naive ${\cal E}=2\tanh \left( E/2\right) $ with $E=1$
786: and computing the appropriate averages from the measured profiles (Fig. 4),
787: this relation predicts $v\approx 0.004$. Being both of the right sign and
788: order of magnitude, this is an encouraging result.
789: 
790: There is an alternative approach, based on a {\em discrete} version of a
791: simple hopping model. Without delving into the details, we only state that
792: it can be obtained as an approximation from the full master equation. By
793: considering $\left\langle \sigma \left( \vec{x},t\right) \right\rangle $ and
794: ignoring all correlations, the result is a discrete equation for the average
795: densities $\rho _{\pm }\left( \vec{x},t\right) $. Focusing only on the $y$
796: co-ordinate (i.e., averaging the densities over $x$), our starting point is 
797: \begin{equation}
798: \rho _{\pm }\left( y,t+1\right) -\rho _{\pm }\left( y,t\right) =\frac{1}{4}%
799: %TCIMACRO{\binom{\mu }{1}}%
800: %BeginExpansion
801: {\mu  \choose 1}%
802: %EndExpansion
803: \left\{ J_{in}-J_{out}\right\} ,  \label{start}
804: \end{equation}%
805: where the jumps into/out-of site $y$ are given by 
806: \begin{eqnarray*}
807: J_{in} &\equiv &\rho _{\pm }\left( y\mp 1,t\right) \phi \left( y,t\right)
808: +e^{-E}\rho _{\pm }\left( y\pm 1,t\right) \phi \left( y,t\right) \\
809: J_{out} &\equiv &\rho _{\pm }\left( y,t\right) \phi \left( y\pm 1,t\right)
810: +e^{-E}\rho _{\pm }\left( y,t\right) \phi \left( y\mp 1,t\right) .
811: \end{eqnarray*}%
812: Here the time unit is a RMCS, while the factor $1/4$ accounts for half of
813: jumps being transverse, so that each term in the $J$'s represents jumps with 
814: $1/4$ probability. Of course, periodic boundary conditions are imposed: $%
815: \rho _{\pm }\left( y,t\right) =\rho _{\pm }\left( y+L,t\right) $. For an
816: inhomogeneous drifting steady state, we seek functions of the form $\rho
817: _{\pm }\left( y,t\right) =\rho _{\pm }^{\ast }\left( u\right) $, so that we
818: would write {\em naively} 
819: %==================
820: \begin{eqnarray*}
821: \rho _{\pm }^{\ast }\left( u-v\right) -\rho _{\pm }^{\ast }\left( u\right) 
822: &=&\frac{1}{4}%
823: %TCIMACRO{\binom{\mu }{1}}%
824: %BeginExpansion
825: {\mu  \choose 1}%
826: %EndExpansion
827: \left\{ \rho _{\pm }^{\ast }\left( u\mp 1\right) \phi ^{\ast }\left(
828: u\right) -\rho _{\pm }^{\ast }\left( u\right) \phi ^{\ast }\left( u\pm
829: 1\right) \right.  \\
830: &&+\left. e^{-E}\left[ \rho _{\pm }^{\ast }\left( u\pm 1\right) \phi ^{\ast
831: }\left( u\right) -\rho _{\pm }^{\ast }\left( u\right) \phi ^{\ast }\left(
832: u\mp 1\right) \right] \right\} .
833: \end{eqnarray*}%
834: Expecting $v$ to be of the order of $10^{-3}$, the first term on the left
835: must be modified if we wish to apply this equation to data analysis. We
836: believe a good approximation is to interpolate the left hand side: 
837: \[
838: \rho _{\pm }^{\ast }\left( u-v\right) -\rho _{\pm }^{\ast }\left( u\right)
839: \simeq v\left[ \rho _{\pm }^{\ast }\left( u-1\right) -\rho _{\pm }^{\ast
840: }\left( u\right) \right] 
841: \]%
842: Rearranging the result, we have 
843: \begin{eqnarray*}
844: 4v\left[ \rho _{\pm }^{\ast
845: }\left( u-1\right)- \rho _{\pm }^{\ast }\left( u\right) \right] &=& %
846: %TCIMACRO{\binom{\mu }{1}}%
847: %BeginExpansion
848: {\mu  \choose 1}%
849: %EndExpansion
850: \left\{ e^{-E}\rho _{\pm }^{\ast }\left( u\pm 1\right) \phi ^{\ast }\left(
851: u\right)-\rho _{\pm }^{\ast }\left( u\right) \phi ^{\ast }\left( u\pm
852: 1\right) \right.  \\
853: &+& \left. \rho _{\pm }^{\ast }\left( u\mp 1\right) \phi ^{\ast }\left(
854: u\right) -e^{-E}\rho _{\pm }^{\ast }\left( u\right) \phi ^{\ast }\left( u\mp
855: 1\right) \right\} 
856: \end{eqnarray*}%
857: which can be ``integrated'' once, as usual. The constants are just the
858: steady state currents ($J_{\pm }^{\ast }\equiv \hat{y}\cdot \vec{J}_{\pm }$%
859: ), so that our basic equations read 
860: \begin{eqnarray}
861: 4J_{+}^{\ast } &=&\mu \left[ \rho _{+}^{\ast }\left( u\right) \phi ^{\ast
862: }\left( u+1\right) -e^{-E}\rho _{+}^{\ast }\left( u+1\right) \phi ^{\ast
863: }\left( u\right) \right] -4v\rho _{+}^{\ast }\left( u\right)  \label{J+} \\
864: 4J_{-}^{\ast } &=&\left[ e^{-E}\rho _{-}^{\ast }\left( u\right) \phi ^{\ast
865: }\left( u+1\right) -\rho _{-}^{\ast }\left( u+1\right) \phi ^{\ast }\left(
866: u\right) \right] -4v\rho _{-}^{\ast }\left( u\right) \quad . \label{J-} 
867: \end{eqnarray}%
868: In principle, these are recursive relations for the profiles and, imposing
869: constraints ($\sum \rho _{\pm }^{\ast }=N_{\pm }$), the unknown profiles and
870: constants ($J_{\pm }^{\ast }$ and $v$) can be found. In practice, this
871: method is not simple. Here, let us simply test these relations against the
872: profiles from the data. By regarding $e^{-E}$ also as an unknown, we can
873: write four {\em linear }equations for these constants and see how well the
874: agreement is. For simplicity, in a manner similar to the continuum study
875: above, we choose to sum Eqns. (\ref{J+},\ref{J-}) for two equations. For the
876: other pair, we first divide Eqns. (\ref{J+},\ref{J-}) by $\rho _{\pm }^{\ast
877: }\left( u\right) $ respectively and then perform the sum. The results are$%
878: \allowbreak $%
879: \[
880: \left[ 
881: \begin{array}{c}
882: J_{+}^{\ast } \\ 
883: J_{-}^{\ast } \\ 
884: v \\ 
885: e^{-E}%
886: \end{array}%
887: \right] \simeq \left[ 
888: \begin{array}{c}
889: 0.99\times 10^{-2} \\ 
890: -0.65\times 10^{-3} \\ 
891: 0.63\times 10^{-3} \\ 
892: 0.39%
893: \end{array}%
894: \right] 
895: \]%
896: Again, we are encouraged by how well this crude scheme functions, since the
897: measured $v$ is about $0.9\times 10^{-3}$ and $e^{-1}\simeq 0.37$. As for
898: the currents, their small values are typical of jammed states. Though our
899: data for them are too noisy for a meaningful comparison, we are quite
900: satisfied by the relative magnitudes and signs.
901: 
902: \section{Summary and Outlook}
903: 
904: We have reported simulation results and mean-field arguments for jamming
905: transitions in a three-state lattice gas under non-equilibrium conditions.
906: In our simple model, positive and negative particles diffuse on a lattice,
907: subject to an excluded volume constraint and an external drive which biases
908: their motion in opposite directions. Generalizing earlier studies of this
909: model\cite{HSZ}, we focus here on the effect of differential mobility,
910: allowing one species to be more mobile than the other by a factor of $\mu $.
911: With equal numbers of the two species in the system, our key findings are as
912: follows: The previously observed phase transition persists for the range $%
913: \mu \in \left[ 1,300\right] $, i.e., a free-flowing state, for low densities
914: ($\bar{m}$) and drive ($E$), giving way to a jammed phase at higher $\bar{m}$
915: and $E$. Moreover, there is little quantitative change to the $\bar{m}$-$E$
916: phase diagram. Crossing the transition line at low $\bar{m}$ and high $E$,
917: we observe hysteresis loops in the density structure factor, consistent with
918: first order transitions. Across the high $\bar{m}$ and low $E$ portion of
919: the line, the transitions are continuous. While the phase diagram is
920: essentially unaffected by $\mu >1$, both disordered and ordered phases
921: exhibit a systematic drift, i.e., non-vanishing mass current. Focusing on
922: ordered structures, we find that the cluster drifts in the direction of the
923: more mobile species. According to the data, hole and charge density profiles
924: depend in subtle ways on $\mu $, and their drift velocity appears to
925: saturate as $\mu $ increases. To shed more light on these findings, we
926: present a mean-field theory, in the form of coupled, nonlinear partial
927: differential equations of motion for the hole and charge densities.
928: Homogeneous solutions, corresponding to disordered phases, follow trivially
929: from the conservation laws for the two particle numbers. A linear stability
930: analysis demonstrates the presence of an instability, as the overall density
931: increases. Unfortunately, we were not able to find inhomogeneous \emph{%
932: drifting} solutions, so that we can only offer several consistency checks
933: between equations and data, in order to build confidence in our theoretical
934: description.
935: 
936: Many questions remain open for further study. First of all, a better
937: understanding of the drift velocity would be desirable. Are there
938: inhomogeneous solutions to our mean-field equations with non-trivial drift?
939: If a proof of their absence can be established, then we should include
940: noise terms (to promote the mean-field equations to Langevin equations) and
941: study the effects of fluctuations and correlations. Perhaps these will
942: renormalize the ``bare'' co-efficients in Eqn. (\ref{EoM}) so as to admit
943: drifting structures. Such analysis, along with further simulations, should
944: also settle the question whether the velocity continues to grow or
945: approaches a saturation value as $\mu $ increases. In addition, the noise
946: terms will allow us to investigate particle correlations and structure
947: factors. As shown in preceding studies \cite{KSZ97b}, these can be extremely
948: interesting, even in the disordered phase. Going beyond \emph{equal-time}
949: structure factors, a study of \emph{dynamic} correlations would provide
950: considerable insight into currents and drifts. Finally, we note that, \emph{%
951: at }the first order line, we should find \emph{coexisting} phases. Since the
952: mass current is significant in the disordered region but quite small in the
953: jammed region, these states will undoubtedly display a rich variety of
954: behavior.
955: 
956: Nearly all the simulations performed here are based on a $40\times 40$
957: lattice. For fixed $E$, the observed transitions cannot possibly persist as the
958: longitudinal size ($L_y$) goes to infinity. At the mean-field level, the
959: inhomogeneous solutions effectively depend only on the product $\mathcal{E}%
960: L_y$. Thus, performing simulations with a range of $E$ and $L_y$ would be
961: useful for establishing the presence of ``$EL_y$ scaling'' or proving its
962: absence. In general, explorations of finite size effects are clearly crucial
963: before reliable conclusions on the nature of the phases and transitions can
964: be drawn.
965: 
966: A natural expansion of parameter space is to allow for unequal numbers of
967: the species. Such systems exhibit drifting structures even in the \emph{%
968: absence} of differential mobility\cite{LZ}. As a consequence, one might
969: wonder whether it is possible to adjust these particle numbers and $\mu $
970: such that the jam becomes stationary. Viewed as traffic problems, involving
971: cars and trucks at different densities and mobilities, the answers to these
972: questions may shed some light on the essence of traffic jams.
973: 
974: 
975: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
976: \emph{Acknowledgements.}
977: We thank A. Vasudevan for helpful discussions.  This research was supported in part by 
978: a grant from the US National Science Foundation through the Division of Materials Research.
979: 
980: 
981: 
982: 
983: 
984: 
985: \section*{References}
986: 
987: \begin{thebibliography}{99}
988: \bibitem{KLS} Katz S, Lebowitz J L, and Spohn H 1983 Phys. Rev. B 
989: {\bf 28}, 1655; and 1984 J. Stat. Phys. {\bf 34}, 497.
990: 
991: \bibitem{DL17} Schmittmann B and Zia R K P 1995 {\em Phase Transitions and
992: Critical Phenomena}, Vol. 17, eds. Domb C and Lebowitz J L (New York, Academic
993: Press).
994: 
995: \bibitem{FIC} Chandra S 1981 {\em Superionic Solids. Principles and
996: Applications} (Amsterdam, North Holland).
997: 
998: \bibitem{Spitzer} Spitzer F 1970 Adv. Math. {\bf 5}, 247. 
999: 
1000: \bibitem{LRC} Zhang M Q, Wang J S, Lebowitz J L, and Vall\`{e}s J L 1988
1001: J. Stat. Phys. {\bf 52}, 1461; Garrido P L, Lebowitz J L, Maes C,
1002: and Spohn H 1990 Phys. Rev. A {\bf 42}, 1954.
1003: 
1004: \bibitem{water-in-oil} Aertsens M and Naudts J 1990 J. Stat. Phys. {\bf 62},
1005: 609.
1006: 
1007: \bibitem{gel-electro} Rubinstein M 1987 Phys. Rev. Lett. {\bf 59}, 1946;
1008: Duke T A J 1989 Phys. Rev. Lett. {\bf 62}, 2877; Schnidman Y 1991 {\em %
1009: Mathematics in Industrial Problems IV}, ed. Friedman A (Berlin, Springer);
1010: Widom B, Viovy J L, and Desfontaines A D 1991 J. Phys I (France) {\bf 1}%
1011: , 1759.
1012: 
1013: \bibitem{Mukamel-gel} Alon U and Mukamel D 1997 Phys. Rev. E {\bf 55}, 1783.
1014: 
1015: \bibitem{VMD} Schmittmann B, Zia R K P, and Triampo T 2000 Braz. J. Phys. 
1016: {\bf 30} 139 (cond-mat/9912135) and references therein.  
1017: 
1018: \bibitem{traffic} Biham O, Middleton A A, and Levine D 1992 Phys. Rev. {\bf A
1019: 46}, R6128; Leung K-t 1994 Phys. Rev. Lett. {\bf 73}, 2386.
1020: 
1021: \bibitem{HSZ} Schmittmann B, Hwang K, and Zia R K P 1992 Europhys. Lett. {\bf %
1022: 19}, 19.
1023: 
1024: \bibitem{BSZ} Bassler K E, Schmittmann B, and Zia R K P 1993 Europhys. Lett. 
1025: {\bf 24}, 115.
1026: 
1027: \bibitem{VZS} Vilfan I, Zia R K P, and Schmittmann B 1994 Phys. Rev. Lett. 
1028: {\bf 73}, 2071. 
1029: 
1030: \bibitem{LZ} Leung K-t and Zia R K P 1997 Phys. Rev. E {\bf 56}, 308.
1031: 
1032: \bibitem{ST} Schmittmann B and Thies M 2002 Europhys. Lett. {\bf 57}, 178.
1033: 
1034: \bibitem{KSZ97} Korniss G, Schmittmann B, and Zia R K P 1995 Europhys. Lett. 
1035: {\bf 32}, 49; and 1997 J. Stat. Phys. {\bf 86}, 721.
1036: 
1037: \bibitem{Bonn} Arndt P F, Heinzel T, and Rittenberg V 1999 J. Stat. Phys. {\bf %
1038: 97}, 1; Arndt P F and Rittenberg V 2002 J. Stat. Phys. {\bf 107}, 989.
1039: 
1040: \bibitem{KSZ99} Korniss G, Schmittmann B, and Zia R K P 1999 Europhys. Lett. 
1041: {\bf 45}, 431.
1042: 
1043: \bibitem{MSZ} Mettetal J, Schmittmann B, and Zia R K P 2002 Europhys. Lett. 
1044: {\bf 58}, 653.
1045: 
1046: \bibitem{KSZ97b} Korniss G, Schmittmann B, and Zia R K P 1997 Physica {\bf A239%
1047: }, 111; and 1997 J. Phys. {\bf A30}, 3837.
1048: 
1049: \end{thebibliography}
1050: 
1051: \end{document}
1052: 
1053: 
1054: