1: \documentclass[12pt]{article}
2: \usepackage{hyperref,graphicx}
3:
4: \def\ba{\begin{eqnarray}}
5: \def\ea{\end{eqnarray}}
6: \def\bq{\begin{eqnarray*}}
7: \def\eq{\end{eqnarray*}}
8: \def\be{\begin{equation}}
9: \def\ee{\end{equation}}
10: \def\bm{\begin{math}}
11: \def\me{\end{math}}
12: \def\bi{\bibitem}
13: \def\vr{\vec r}
14: \def\lt{L(t)}
15: \def\al{a(L,T)}
16: \def\hf{1 \over 2}
17: \def\ap{\alpha}
18: \def\lb{\left (}
19: \def\rb{\right )}
20: \def\la{\langle}
21: \def\ra{\rangle}
22: \def\lbr{\left [}
23: \def\rbr{\right ]}
24: \def\del{\partial}
25: \def\grad{\nabla}
26: \def\ul{\underline}
27: \def\etal{{\it et al~}}
28: \def\lra{\leftrightarrow}
29: \def\rar{\rightarrow}
30:
31: \def\prl#1#2#3{{Phys. Rev. Lett.} \textbf{#1}, #2 (#3)}
32: \def\pr#1#2#3{Phys. Rev. \textbf{#1}, #2 (#3)}
33: \def\pra#1#2#3{Phys. Rev. A \textbf{#1}, #2 (#3)}
34: \def\prb#1#2#3{Phys. Rev. B \textbf{#1}, #2 (#3)}
35: \def\pre#1#2#3{Phys. Rev. E \textbf{#1}, #2 (#3)}
36: \def\epl#1#2#3{{Europhys. Lett.} \textbf{#1}, #2 (#3)}
37: \def\ibid#1#2#3{\emph{ibid.} \textbf{#1}, #2 (#3)}
38: \def\jsp#1#2#3{ J. Stat. Phys. \textbf{#1}, #2 (#3)}
39: \def\pla#1#2#3{Phys. Lett. A \textbf{#1}, #2 (#3)}
40: \def\jpa#1#2#3{J. Phys. A \textbf{#1}, #2 (#3)}
41: \def\jcp#1#2#3{J. Chem. Phys. \textbf{#1}, #2 (#3)}
42: \def\jpc#1#2#3{J. Phys. Chem. \textbf{#1}, #2 (#3)}
43: \def\physa#1#2#3{Physica A \textbf{#1}, #2 (#3)}
44: \def\physd#1#2#3{Physica D \textbf{#1}, #2 (#3)}
45: \def\pram#1#2#3{Pramana J. Phys. \textbf{#1}, #2 (#3)}
46: \def\advp#1#2#3{Adv. Phys. \textbf{#1}, #2 (#3)}
47: \def\rmp#1#2#3{Rev. Mod. Phys. \textbf{#1}, #2 (#3)}
48: \def\nat#1#2#3{Nature \textbf{#1}, #2 (#3)}
49: \def\siam#1#2#3{SIAM J. Appl. Math \textbf{#1}, #2 (#3)}
50: \def\ptpo#1#2#3{Prog. Theor. Phys. \textbf{#1}, #2 (#3)}
51:
52: \begin{document}
53:
54: \begin{center}
55:
56: {\Large \bf Phase Separation Driven by Surface Diffusion: A Monte Carlo Study}
57: \vskip1cm
58: S. van Gemmert$^1$, G.T. Barkema$^1$ and Sanjay Puri$^2$ \\
59: \vskip0.25cm
60: {\it $^1$Institute for Theoretical Physics \\
61: Utrecht University, Leuvenlaan 4 \\
62: 3584 CE Utrecht, THE NETHERLANDS.} \\
63: \vskip0.25cm
64: {\it $^2$School of Physical Sciences \\
65: Jawaharlal Nehru University \\
66: New Delhi -- 110067, INDIA.}
67: \end{center}
68:
69: \vskip2cm
70: \begin{abstract}
71: We propose a kinetic Ising model to study phase separation driven by surface
72: diffusion. This model is referred to as \emph{Model S}, and consists of the
73: usual Kawasaki spin-exchange kinetics (\emph{Model B}) in conjunction with
74: a kinetic constraint. We use novel multi-spin coding techniques to develop
75: fast algorithms for Monte Carlo simulations of Models B and S. We use these
76: algorithms to study the late stages of pattern dynamics in these systems.
77: \end{abstract}
78:
79: \newpage
80: \section{Introduction}
81:
82: Consider a homogeneous binary (AB) mixture, which is rendered thermodynamically
83: unstable by a rapid temperature quench below the miscibility gap.
84: The system prefers to be in a phase-separated state at the lower temperature.
85: The far-from-equilibrium evolution
86: of the system from the unstable homogeneous state to the segregated
87: state has received considerable attention \cite{ab94,bf01,ao02,dp03}.
88: This evolution is characterized by the emergence and growth of domains enriched
89: in the components A and B. The terms used to describe this nonequilibrium
90: process are {\it phase ordering dynamics}, {\it domain growth} or {\it coarsening}.
91: A quantitative characterization of phase ordering systems focuses on
92: (a) the domain growth law; (b) the statistical properties of the evolution morphology;
93: and (c) the temporal correlation of pattern dynamics.
94:
95: The equilibrium phase-separated state
96: is uniquely determined by its thermodynamic properties.
97: However, there is a diverse range of kinetic pathways which enable segregation.
98: For example, phase separation in alloys is usually driven by vacancy-mediated
99: diffusion \cite{sp97}. On the other hand, for fluid mixtures, hydrodynamic
100: velocity fields enable convective transport of material along domain
101: boundaries and give rise
102: to novel asymptotic behaviors \cite{es79}. Furthermore, phase separation in
103: mixtures can be frozen (or near-frozen) into mesoscopic states
104: by the presence of quenched disorder \cite{pcp91}, viscoelastic effects
105: \cite{op99,ao02}, etc.
106:
107: In this paper, we present results from a comparative Monte Carlo (MC) study of two
108: kinetic Ising models for phase separation in
109: binary mixtures. The first of these is the usual Kawasaki spin-exchange model
110: \cite{kk72,dp03}, which mimics segregation via diffusion. The second model
111: mimics the case where only surface diffusion is permitted. An important
112: goal of this paper is methodological, viz., the formulation of a kinetic
113: Ising model where bulk diffusion is suppressed. Another important goal is to
114: compare pattern dynamics for phase separation with and without bulk diffusion.
115:
116: This paper is organized as follows. In Sec.~2, we describe the kinetic Ising
117: models studied here and our MC simulation techniques.
118: Our MC approach uses novel multi-spin
119: coding techniques, which enable large-scale and long-time simulations of these
120: models. In Sec.~3, we discuss the domain growth laws which
121: arise from bulk and surface diffusion, and also present detailed numerical results.
122: Finally, Sec.~4 concludes this paper with a summary and discussion of
123: our results.
124:
125: \section{Numerical Methodology}
126:
127: \subsection{Kinetic Ising Models}
128:
129: The standard model for binary mixtures is the Ising model:
130: \ba
131: \label{ham}
132: {\cal H}= -J\sum_{\langle ij \rangle} \sigma_i \sigma_j , \quad \sigma_i = \pm 1,
133: \ea
134: where the spins $\{\sigma_i\}~(i=1 \rightarrow N)$ are located on
135: a discrete lattice. The states $\sigma_i = +1$ or $-1$ denote the presence of an
136: A-atom or B-atom at site $i$, respectively. We consider the case with
137: ferromagnetic ($J>0$) nearest-neighbor interactions, denoted by the subscript
138: $\langle ij \rangle$ in Eq.~(\ref{ham}). The phase diagram for the binary
139: mixture is obtained in an ensemble with fixed temperature
140: $T$ and magnetization $M = \sum_i \sigma_i$.
141:
142: The Ising system does not have an intrinsic dynamics as the Poisson brackets
143: (or commutators) of spin variables are identically zero. Therefore, one
144: introduces stochastic kinetics by placing the system in contact with
145: a heat-bath which induces fluctuations. The Ising model, in conjunction with
146: a physically appropriate spin kinetics, is referred to as a \emph{kinetic
147: Ising model} \cite{dp03,bh02}. An important example is the
148: Kawasaki spin-exchange model \cite{kk72}, which has nearest-neighbor spin
149: exchanges with Metropolis acceptance probabilities. In an MC simulation
150: of this model, a pair of nearest-neighbor sites $i$ and
151: $j$ is randomly selected, and the spins $\sigma_i$ and $\sigma_j$
152: are exchanged. The probability that this exchange is accepted is given by
153: \ba
154: \label{prob}
155: P &=& \min \left[ 1, \exp(-\beta \Delta {\cal H}) \right], \nonumber \\
156: \Delta {\cal H} &=& J(\sigma_i - \sigma_j) \left( \sum_{L_i \neq j} \sigma_{L_i}
157: - \sum_{L_j \neq i} \sigma_{L_j} \right) .
158: \ea
159: Here, $\Delta {\cal H}$ is the energy change due to the proposed
160: spin exchange, and $\beta=(k_BT)^{-1}$ is the inverse temperature, with
161: $k_B$ denoting the Boltzmann constant. In Eq.~(\ref{prob}), $L_i$ denotes
162: the nearest-neighbors of $i$ on the lattice.
163: A single Monte Carlo step (MCS) corresponds
164: to $N$ such attempted exchanges. A large number of
165: MC simulations of the Kawasaki model have been reported in the literature
166: \cite{asm88,mb95}.
167:
168: The phase-separation kinetics in this microscopic model is analogous to that
169: for the coarse-grained Cahn-Hilliard-Cook (CHC) equation,
170: which is obtained as follows:
171: \ba
172: \label{chc}
173: \frac{\partial}{\partial t} \psi (\vec{r},t) &=& - \vec{\nabla} \cdot
174: {\vec{J}}(\vec{r},t) \nonumber \\
175: &=& \vec{\nabla} \cdot \left[ D\vec{\nabla} \mu (\vec{r},t) + \vec{\theta} (\vec{r},t)
176: \right] \nonumber \\
177: &=& \vec{\nabla} \cdot \left[ D\vec{\nabla} \left( \frac{\delta \cal{F}}{\delta \psi}
178: \right) + \vec{\theta} (\vec{r},t) \right] .
179: \ea
180: Here, $\psi (\vec{r},t)$ is the order parameter at space point
181: $\vec{r}$ and time $t$. Typically, $\psi (\vec{r},t)=\rho_A (\vec{r},t)-\rho_B
182: (\vec{r},t)$, where $\rho_A$ and $\rho_B$ denote the local densities of
183: species A and B. In Eq.~(\ref{chc}), the quantities $\vec{J}, D$ and $\mu$ denote the
184: current, diffusion coefficient, and chemical-potential difference between A and B,
185: respectively. The chemical potential is obtained as a functional derivative of the
186: Helmholtz free energy, which is often taken to have the $\psi^4$-form:
187: \ba
188: \label{psi4}
189: {\cal F} [\psi] &=& {\cal H} - TS \nonumber \\
190: &\simeq& \int d\vec{r} \left[ -\frac{1}{2} k_B (T_c-T) \psi^2 +
191: \frac{k_BT_c}{12} \psi^4 + \frac{J}{2} (\vec{\nabla} \psi)^2 \right],
192: \ea
193: where we have identified $\la \sigma_i \ra = \psi (\vec{r}_i)$ in Eq.~(\ref{ham}) and
194: Taylor-expanded various terms. Here, $T_c$ denotes the critical temperature.
195: Finally, the Gaussian white noise term $\vec{\theta} (\vec{r},t)$
196: in Eq.~(\ref{chc}) has zero average
197: and obeys the appropriate fluctuation-dissipation relation. The CHC equation is
198: also known as {\it Model B} in the Hohenberg-Halperin classification scheme
199: for critical dynamics \cite{hh77}. Further, using a master-equation approach, the
200: CHC equation can be motivated from the spin-exchange model \cite{kb74}. Therefore,
201: we will subsequently refer to the Kawasaki model as ``Model B''.
202:
203: Before proceeding, we stress that the general form of the CHC equation
204: contains an order-parameter-dependent mobility \cite{lbm75,ki78,pbd92}:
205: \ba
206: \label{opm}
207: D(\psi) = D_0 \left( 1 - \frac{\psi^2}{{\psi_0}^2} \right) ,
208: \ea
209: where $\psi_0$ is the saturation value of the order parameter at $T=0$. This
210: is not consequential for quenches to moderate temperatures, but plays an
211: important role for deep quenches where $\psi \simeq \pm \psi_0$ in bulk domains.
212: In that case, bulk diffusion is effectively eliminated and domain growth
213: proceeds by surface diffusion \cite{pbl97,cy89,lm92}. In the context of the
214: Kawasaki model, this can be understood by focusing on an interfacial pair
215: with the minimum barrier for interchange: $\sigma_i = +1$ at the periphery of
216: an up-rich domain and having only one neighbor with the same spin value,
217: and $\sigma_j = -1$ in a down-rich domain. At low temperatures, the bulk domains
218: are very pure and the energy barrier to the interchange $\sigma_i \leftrightarrow
219: \sigma_j$ is $\Delta {\cal H} = 4J$. Thus,
220: the time-scale for this interchange $\tau_K \sim \exp (\beta \Delta {\cal H})
221: \rightarrow \infty$ as $T \rightarrow 0$, effectively blocking bulk diffusion.
222: Of course, once an impurity spin is placed inside a bulk domain, there is
223: no further barrier to its diffusion.
224:
225: Apart from this natural blocking of bulk diffusion at $T=0$, there are systems
226: where the bulk mobility diminishes drastically due to physical processes,
227: e.g., one or both of the components may undergo a glass \cite{sj97} or
228: gelation \cite{sb93,sp01} transition. At the phenomenological level, this
229: has been modeled by setting the mobility to zero in regions rich in the
230: glass-phase or gel-phase. At the microscopic level, we propose
231: a kinetic Ising model where bulk diffusion is suppressed by introducing
232: a kinetic constraint. We disallow exchanges $\sigma_i \leftrightarrow
233: \sigma_j$ if the neighboring spins of the pair are
234: all parallel, even though such an exchange would not raise the energy.
235: In this case, segregation is driven primarily by diffusion along domain
236: boundaries, though some bulk transport occurs via impurity $n$-spin clusters.
237: (This bulk diffusion is negligible for moderate to deep quenches.) We will
238: subsequently refer to this model as ``Model S'' \cite{pbl97}. Clearly, Model S
239: can be generalized to the case of reduced (though non-zero) mobility in the bulk
240: domains. This is done by allowing spin-exchanges with different time-scales
241: depending on the number and type of parallel neighbors for a spin pair.
242:
243: \subsection{Numerical Details}
244:
245: All our MC simulations were performed on an $L \times L$ square lattice
246: with periodic boundary conditions. At time $t=0$, the temperature
247: was quenched from $T=\infty$ to $T<T_c$, where
248: $T_c \simeq 2.269$ is the critical temperature of the $d=2$ Ising model.
249: (All energy scales are measured in units of $J$, and we set the
250: Boltzmann constant $k_B = 1$.) The disordered
251: initial state consisted of a uniform mixture of $N_A$ A-atoms
252: and $N_B$ B-atoms with $N=N_A+N_B$. The case with $N_A=N_B$ corresponds
253: to a critical quench.
254:
255: Our MC simulations exploit the technique of \emph{multi-spin coding}.
256: For a general introduction to this technique, see Ref.~\cite{nb99}.
257: The basic idea is that one can exploit the 64-bit computer architecture
258: to undertake a parallel simulation of 64 systems. This is done by storing
259: the spin $\sigma_i$ at site $i$ in the $k^{\rm{th}}$ system
260: in the $k^{\rm{th}}$ bit of a 64-bit word $S[i]$. Recall that, in
261: one elementary move for Model B, we propose to exchange the spins located
262: on nearest-neighbor sites $i$ and $j$. For Model S, we impose the kinetic
263: constraint that a pair of spins surrounded by aligned neighbors is not exchanged.
264:
265: For the $d=2$ square lattice considered here, each site has four
266: nearest-neighbors. Let $n_0$, $n_1$ and $n_2$ be the three nearest-neighbors
267: (other than $j$) of site $i$. Similarly, let $m_0$, $m_1$ and $m_2$ be the
268: three nearest-neighbors (other than $i$) of site $j$. To determine the change
269: in energy resulting from the proposed spin exchanges in all 64 simulations,
270: we first identify which of the six neighbors $(n_0,n_1,n_2,m_0,m_1,m_2)$
271: are antiparallel. This can be done in six operations with the {\it exclusive or}
272: operation $\oplus$:
273: %
274: \begin{eqnarray}
275: A_k &=& S[i] \oplus S[n_k] , \quad k=0 \rightarrow 2,\nonumber \\
276: B_k &=& S[j] \oplus S[m_k] , \quad k=0 \rightarrow 2.
277: \end{eqnarray}
278: %
279: The energy change associated with the spin exchange and (thereby) the
280: acceptance probability is governed by the number of antiparallel
281: spins. In an ordinary program, this would involve a summation over the
282: surrounding spins. With logical operations, it is more convenient to
283: determine the logical variables $P_k$ which tell whether
284: $\sigma_i$ is antiparallel to at least $k$ of its neighbors (other than
285: $\sigma_j$). Note that the Metropolis algorithm only requires
286: $P_1, P_2$ and $P_3$. These can be obtained with six operations:
287: %
288: \begin{eqnarray}
289: P_2 &=& A_0 \land A_1 , \nonumber \\
290: P_1 &=& A_0 \lor A_1 , \nonumber \\
291: P_3 &=& A_2 \land P_2 , \nonumber \\
292: P_2 &=& P_2 \lor ( A_2 \land P_1) , \nonumber \\
293: P_1 &=& P_1 \lor A_2 .
294: \end{eqnarray}
295: %
296: Similarly, the variables $Q_k$ that tell whether $\sigma_j$ is
297: antiparallel to at least $k$ of its neighbors (other than $\sigma_i$)
298: can be obtained with six operations.
299:
300: Finally, the acceptance probability for the proposed spin exchanges is
301: obtained by using \emph{random bit patterns} $R_0$, $R_1$ and
302: $R_2$. These are designed so that the probability for each
303: bit to be 1 is $P_b=\exp(-4\beta J)$. Thus, the following statements
304: comprise the core of our Model S algorithm:
305: %
306: \begin{eqnarray}
307: \rm{Flip} &=& (S[i] \oplus S[j]) \land (P_1 \lor Q_3 \lor R_0) \land (P_2 \lor Q_2 \lor R_1)
308: \land (P_3 \lor Q_1 \lor R_2) , \nonumber \\
309: \rm{Flip} &=& \rm{Flip} \land (P_1 \lor \lnot Q_3)
310: \land (Q_1 \lor \lnot P_3) , \nonumber \\
311: S[i] &=& S[i] \oplus \rm{Flip} , \nonumber \\
312: S[j] &=& S[j] \oplus \rm{Flip} .
313: \end{eqnarray}
314: The implementation of Model B dynamics is simply obtained by omitting the
315: second of the above statements.
316:
317: These 36 operations for Model S (or 30 for Model B) act on all 64 bits and thus
318: perform 64 elementary moves. In conjunction with the required {\it load}
319: and {\it store} operations, and generation of the random
320: bit patterns, our implementation of Model B for a $512^2$ system
321: with multi-spin coding requires 4.6 ns CPU-time per elementary
322: move on an AMD-64 computer with 3 GHz clock frequency.
323: This should be contrasted with a direct implementation
324: of this model, which requires approximately 100 ns CPU-time
325: per elementary move on the same machine.
326:
327: The procedure outlined above, which simulates 64 separate systems,
328: is known as a \emph{synchronous} multi-spin algorithm~\cite{nb99}.
329: The boundaries of these 64 systems can be glued together to yield an
330: \emph{asynchronous} multi-spin algorithm~\cite{nb99}, simulating one
331: system which is 64 times larger. This comes at the cost of (a) more
332: complicated programming; and (b) a small reduction in the program
333: efficiency. The statistical results for domain morphologies
334: presented in Secs.~3.1, 3.2 and 3.3 were obtained by averaging over 150
335: asynchronous simulations with system sizes $L=512$. The results for the
336: autocorrelation function in Sec.~3.4 were obtained by averaging over 64
337: synchronously simulated systems with $L=1024$.
338:
339: \section{Detailed Results}
340:
341: As stated earlier, the initial condition for our MC simulations consists
342: of a random configuration. The temperature is
343: quenched to $T<T_c$ at $t=0$, and the system evolves via either Model B or Model
344: S dynamics towards its new equilibrium state. Figure~\ref{snap} shows the typical
345: \begin{figure}[htb]
346: \centering
347: \includegraphics[width=0.6\textwidth]{fig1.eps}
348: \caption{Evolution pictures for phase separation in a binary (AB) mixture with a
349: critical composition. The component A ($\sigma = +1$) is marked in black,
350: and the component B ($\sigma = -1$) is unmarked. The system was quenched
351: at time $t=0$ from $T=\infty$ to $T=0.63T_c$. The top and bottom panels show
352: snapshots at times $t=10^4$ and $10^6$ MCS, using either Model
353: B dynamics (left), or Model S dynamics (right). The MC simulations were
354: done on square lattices of size $512^2$ with periodic boundary conditions.
355: The details of the simulations are described in the text.
356: \label{snap}}
357: \end{figure}
358: time evolution for a critical composition (50~\% A and 50~\% B)
359: after a quench to $T=0.63T_c$ for Model B (left) and Model S
360: (right). Notice that the evolution morphology has a characteristic
361: domain size, which we denote as $R(t)$.
362: The growth process is substantially slower for S-dynamics,
363: as expected. We will demonstrate shortly that the growth law due to
364: bulk diffusion is $R(t) \sim t^{1/3}$, which is referred to as the
365: Lifshitz-Slyozov (LS) growth law. The corresponding law for segregation
366: via surface diffusion is $R(t) \sim t^{1/4}$.
367: However, we reiterate that bulk diffusion is
368: not eliminated entirely in Model S because of the presence of impurity
369: spin clusters in bulk domains. At high temperatures, there is
370: a reasonable fraction of impurity spins and we expect the S-dynamics
371: to cross over to $t^{1/3}$-growth at late times. The crossover time increases
372: rapidly at lower temperatures where there are very few impurity spins in
373: the bulk as $P_{\rm{imp}} \simeq \left[ 1 + \exp(8 \beta J) \right]^{-1}$
374: in $d=2$.
375:
376: We will study the evolution depicted in Fig.~\ref{snap} using quantities like the
377: correlation function and autocorrelation function.
378:
379: \subsection{Growth Laws}
380:
381: The first relevant property is the growth law governing the
382: segregation process. We computed the typical domain size $R(t)$ as the
383: first zero-crossing of the two-point correlation function:
384: \begin{eqnarray}
385: \label{cor}
386: C(\vec{r},t) & = & {1 \over N} \sum_{i=1}^N \left[ \la \sigma_i (t)
387: \sigma_{i + \vec{r}} (t)\ra - \la \sigma_i (t)\ra \la \sigma_{i + \vec{r}} (t)
388: \ra \right] \\
389: \label{ds}
390: & \equiv & g \left( {r \over R} \right) .
391: \end{eqnarray}
392: Here, $\vec{r}$ denotes the displacement vector, and we consider systems
393: which are translationally invariant and isotropic.
394: The angular brackets in Eq.~(\ref{cor}) denote an averaging over
395: independent initial conditions and noise realizations.
396: Equation~(\ref{ds}) is the dynamical-scaling property of the correlation
397: function \cite{bs74}, and reflects the fact that the morphology is
398: self-similar in time, upto a scale factor (see Fig.~\ref{snap}).
399: One can use other definitions of the length scale also, but these are
400: all equivalent in the scaling regime.
401:
402: At this stage, it is useful to clarify the domain growth laws which arise
403: due to bulk and surface diffusion. A convenient starting point is
404: the CHC equation (\ref{chc}) with an order-parameter-dependent mobility
405: $D(\psi)$. We consider a general situation where
406: the diffusion coefficient at the interface
407: ($\psi = 0$) is $D_s$, and that in the bulk [$\psi = \psi_s (T)$] is $D_b$
408: with $D_b \leq D_s$. This difference in surface and bulk mobilities may
409: be the consequence of low-temperature dynamics or due to physical processes
410: like glass formation or gelation.
411: Then, the simplest functional form which models the mobility is
412: \ba
413: D(\psi) = D_s \left( 1 - \alpha \frac{\psi^2}{\psi_s^2} \right),
414: \quad \alpha = 1 -\frac{D_b}{D_s} ,
415: \ea
416: which is equivalent to Eq.~(\ref{opm}) with $D_0 = D_s$ and $\psi_0^2 =
417: \psi_s^2/\alpha$. We focus on the deterministic case of Eq.~(\ref{chc}):
418: \ba
419: \label{ch}
420: \frac{\partial}{\partial t} \psi (\vec{r},t) = D_s \vec{\nabla} \cdot
421: \left\{ \left( 1 - \alpha \frac{\psi^2}{{\psi_s}^2} \right) \vec{\nabla}
422: \left[ -(T_c - T) \psi + \frac{T_c}{3} \psi^3 - J \nabla^2 \psi \right] \right\} ,
423: \ea
424: where we have used the $\psi^4$-form of the free energy from Eq.~(\ref{psi4}).
425: The saturation value of the order parameter in Eq.~(\ref{ch}) is
426: $\psi_s (T) = \sqrt{3 \left(1-T/T_c\right)}$. Using the natural scales for
427: the order parameter, length and time, we can rewrite Eq.~(\ref{ch})
428: in the dimensionless form:
429: \ba
430: \label{chd}
431: && \frac{\partial}{\partial t} \psi (\vec{r},t) = \vec{\nabla} \cdot
432: \left[ \left( 1 - \alpha \psi^2 \right) \vec{\nabla}
433: \left( -\psi + \psi^3 - \nabla^2 \psi \right) \right] , \nonumber \\
434: && \alpha \in [0,1]~~\mbox{for}~~ D_b \leq D_s .
435: \ea
436:
437: The RHS of Eq.~(\ref{chd}) can be decomposed as \cite{lm92}
438: \ba
439: \label{chsplit}
440: \frac{\partial}{\partial t} \psi (\vec{r},t) &=&
441: (1-\alpha) \nabla^2 (-\psi + \psi^3 - \nabla^2 \psi) + \nonumber \\
442: && \alpha \vec{\nabla} \cdot \left[ \left( 1 - \psi^2 \right) \vec{\nabla}
443: \left( -\psi + \psi^3 - \nabla^2 \psi \right) \right] ,
444: \ea
445: where the first term on the RHS corresponds to bulk diffusion. This term
446: disappears for $\alpha = 1$ or $D_b=0$. The second term on the RHS corresponds to
447: surface diffusion and is only operational at interfaces where $\psi \simeq 0$.
448: Following Ohta \cite{to88}, we can obtain an equation which describes
449: the interfacial dynamics. The location of the interfaces $\vec{r}_i(t)$ is
450: defined by the zeros of the order-parameter field:
451: \ba
452: \psi \left[ \vec{r}_i(t),t \right] = 0 .
453: \ea
454: Focus on a particular interface, and designate the normal coordinate as
455: $n$ (with dimensionality 1) and the interfacial coordinates as $\vec{a}$
456: [with dimensionality $(d-1)$]. Then, the normal velocity $v_n (\vec{a},t)$
457: obeys the integro-differential equation \cite{to88,lm92}:
458: \ba
459: \label{int}
460: 4 \int d \vec{a'} G[\vec{r}_i(\vec{a}), \vec{r}_i(\vec{a'})] v_n (\vec{a'},t)
461: &\simeq& (1-\alpha) \sigma K (\vec{a},t) + \nonumber \\
462: && 4 \alpha \int d \vec{a'}
463: G[\vec{r}_i (\vec{a}), \vec{r}_i (\vec{a'})] \nabla^2 K (\vec{a'},t) ,
464: \ea
465: where $K (\vec{a},t)$ is the local curvature at point $\vec{a}$ on the interface,
466: and $\sigma$ is the surface tension. The Green's function $G(\vec{x},\vec{x'})$ obeys
467: \ba
468: -\nabla^2 G(\vec{x},\vec{x'}) = \delta (\vec{x} - \vec{x'}) .
469: \ea
470:
471: A dimensional analysis of Eq.~(\ref{int}) in the scaling regime yields the
472: growth laws due to surface and bulk diffusion. We identify the scales of
473: various quantities in Eq.~(\ref{int}) as
474: \ba
475: && [d\vec{a}] \sim R^{d-1}, \quad [G] \sim R^{2-d}, \nonumber \\
476: && [v_n] \sim \frac{dR}{dt}, \quad [K] \sim R^{-1}.
477: \ea
478: This yields the crossover behavior of the length scale as
479: \ba
480: R(t) & \sim & (\alpha t)^{1/4} , \quad t \ll t_c , \nonumber \\
481: & \sim & [(1-\alpha) \sigma t]^{1/3}, \quad t \gg t_c ,
482: \ea
483: where the crossover time is
484: \ba
485: \label{tc}
486: t_c \sim \frac{\alpha^3}{(1-\alpha)^4 \sigma^4} .
487: \ea
488:
489: The above scenario applies for both Models B and S, as $D_b < D_s$ in either
490: case. At moderate temperatures, this crossover occurs rapidly for Model B in our
491: simulations. However, in Model S, there is a drastic suppression
492: of bulk diffusion with $D_b \ll D_s$ and $\alpha \simeq 1$. Therefore, the
493: crossover to $t^{1/3}$-growth
494: is strongly delayed and not observed over simulation time-scales. This is
495: seen in Fig.~\ref{domain}, which plots $R$ vs. $t$
496: \begin{figure}[htb]
497: \centering
498: \includegraphics[width=0.8\textwidth]{fig2.ps}
499: \caption{Typical domain size ($R$) as a function of time ($t$) after a quench
500: at $t=0$ from $T=\infty$ to $T=0.63T_c$ (left) and $0.88T_c$ (right). The
501: circles and squares indicate data obtained with Model B and Model S dynamics,
502: respectively. Lines with slope $1/4$ and $1/3$ are also provided on the plots
503: as guides to the eye.
504: \label{domain}}
505: \end{figure}
506: at temperatures $T=0.63 T_c$ and $0.88 T_c$ for Models B and S. The data
507: for Model S is consistent with the growth law $R \sim t^{1/4}$.
508: The early-time data for Model B is also consistent with this
509: growth law, as surface diffusion is dominant at early times. At late
510: times, one sees crossover behavior between the $t^{1/4}$-regime and the
511: asymptotic $t^{1/3}$-regime. Note that the crossover for Model B is delayed
512: at the higher temperature $T=0.88 T_c$ because the decrease in $\alpha$ is
513: more than compensated by the reduction in $\sigma$ due to softening
514: of the interfaces as $T \rightarrow T_c^-$ [see Eq.~(\ref{tc})].
515: More generally, we stress that it has
516: been notoriously difficult to observe the asymptotic $t^{1/3}$-growth in
517: MC simulations of the Kawasaki model \cite{asm88,mb95}.
518: Similar results have been obtained from Langevin studies
519: of coarse-grained models \cite{pbl97,cy89,lm92}.
520:
521: Before proceeding, we remark that we have also studied models where bulk
522: diffusion is more strictly suppressed by imposing additional kinetic
523: constraints which eliminate 2-spin diffusion, 3-spin diffusion, etc. The
524: corresponding results for the growth law are numerically indistinguishable
525: from the Model S results in Fig.~\ref{domain} over the time-scales of
526: our simulation. This underlines the utility of the proposed Model S in
527: the context of phase separation via surface diffusion.
528:
529: It is also relevant to discuss off-critical quenches, where one of the
530: components is present in a larger fraction. In Fig.~\ref{off}, we show
531: evolution pictures for Models B and S for the case with 25\% A and 75\% B.
532: \begin{figure}[htb]
533: \centering
534: \includegraphics[width=0.6\textwidth]{fig3.ps}
535: \vskip1cm
536: \caption[]{Analogous to Fig.~\ref{snap}, but for an off-critical quench
537: with 25\% A and 75\% B. The temperature is $T=0.63T_c$.
538: \label{off}}
539: \end{figure}
540: If the evolution morphology
541: is not bicontinuous, e.g., there are droplets of the
542: minority phase in a matrix of the majority phase, the surface-diffusion
543: mechanism is unable to drive growth. Nevertheless, at high temperatures, growth
544: may still proceed by the Brownian motion of droplets \cite{hf84}. The
545: corresponding growth law depends explicitly on the dimensionality:
546: \ba
547: R(t) \sim (Tt)^{1/(d+2)} .
548: \ea
549: Thus, domain growth through droplet motion obeys the law
550: $R(t) \sim (Tt)^{1/4}$ in $d=2$, which is analogous to the surface-diffusion
551: growth law. The growth kinetics of Models B and S for off-critical mixtures at
552: $T=0.63 T_c$ is shown in Fig.~\ref{offl}. We see that the S-dynamics shows the
553: expected $t^{1/4}$-growth over extended time-regimes. As before, the B-dynamics
554: shows a crossover behavior between the $t^{1/4}$-regime and the asymptotic
555: $t^{1/3}$-regime. At low temperatures, the Brownian
556: mechanism is ineffective and the evolution of Model S freezes into a
557: meso-structure.
558: \begin{figure}[htb]
559: \centering
560: \includegraphics[width=0.8\textwidth]{fig4.ps}
561: \vskip0.5cm
562: \caption[]{Analogous to Fig.~\ref{domain}, but for an off-critical quench
563: with 25\% A and 75\% B. We show results for $T=0.63 T_c$.
564: \label{offl}}
565: \end{figure}
566:
567: \subsection{Correlation Functions}
568:
569: Next, let us study the morphological features of the evolution in
570: Figs.~\ref{snap} and \ref{off}. These are usually characterized by (a) the
571: correlation function defined in Eq.~(\ref{cor}), or (b) its
572: Fourier transform, the structure factor. We have confirmed that these
573: quantities exhibit dynamical scaling for both Models B and S. For the
574: sake of brevity, we do not show these results here.
575:
576: An important theme in this context is a comparison of the morphologies
577: arising from both dynamics. Earlier studies with coarse-grained
578: models \cite{pbl97,cy89,lm92} have found that the correlation functions and
579: structure factors are numerically indistinguishable for growth driven
580: by bulk diffusion or surface diffusion. At the visual level, this also
581: seems to be suggested by the snapshots in Figs.~\ref{snap} and \ref{off}. In
582: Fig.~\ref{grcompare}, we compare the scaling
583: \begin{figure}[htb]
584: \centering
585: \includegraphics[width=0.45\textwidth]{fig5a.ps}
586: \includegraphics[width=0.45\textwidth]{fig5b.ps}
587: \caption{Superposition of scaling functions for Models B (solid line)
588: and S (dashed line) for the evolution depicted in Fig.~\ref{snap}. For Model S,
589: the data set corresponds to $t=10^6$ MCS; for Model B, the data set corresponds to
590: $t=3.4 \times 10^5$ MCS. Both domain sizes coincide at these times.
591: (a) Plot of $C(r,t)$ vs. $r/R$. (b) Plot of $C(r,t) \cdot (r/R)$ vs. $r/R$,
592: so as to magnify the tail behavior.
593: \label{grcompare}}
594: \end{figure}
595: functions for Models B and S for a critical quench with $T=0.63 T_c$.
596: To eliminate finite-size effects, we consider cases with the same
597: typical domain size: $t=10^6$ MCS in Model S and $t=3.4 \times 10^5$ MCS
598: in Model B. In Fig.~\ref{grcompare}(a), we plot $C(r,t)$ vs. $r/R$.
599: The scaling functions superpose on the scale of the plot, in accordance
600: with earlier studies of phenomenological models. In Fig.~\ref{grcompare}(b),
601: we plot $C(r,t) \cdot (r/R)$ vs. $r/R$ so that the large-distance behavior
602: is magnified. Some subtle differences between the two functions are seen at large
603: distances $r/R > 2$. We make some observations in this regard: \\
604: (a) The statistical error in the difference between the curves
605: at the second peak is about four times smaller than the difference,
606: so it cannot be attributed to statistical fluctuations. \\
607: (b) We have also replotted the correlation functions for Models B and S
608: from different times on the scale $C(r,t) \cdot (r/R)$ vs. $r/R$. In that
609: case, the secondary peaks show a much better collapse, suggesting that the
610: discrepancy in Fig.~\ref{grcompare}(b) is not the result of corrections
611: to scaling.
612:
613: Though it is difficult to attribute physical significance to these differences,
614: it is conceptually important to stress the observable differences between the
615: morphologies for Models B and S. Similar scaling plots for the
616: off-critical quench shown in Fig.~\ref{off} are shown in Fig.~\ref{offsc}.
617: \begin{figure}[htb]
618: \centering
619: \includegraphics[width=0.45\textwidth]{fig6a.ps}
620: \includegraphics[width=0.45\textwidth]{fig6b.ps}
621: \caption{Analogous to Fig.~\ref{grcompare}, but for an off-critical quench
622: with 25\% A and 75\% B at $T=0.63 T_c$. In this case, the times for the different
623: data sets are $t = 9.8\times 10^5$ MCS (Model S) and $t= 2.7\times 10^5$ MCS
624: (Model B).
625: \label{offsc}}
626: \end{figure}
627: It is known that the scaling functions for phase-separating systems depend
628: on the degree of off-criticality \cite{sp88}. Notice that the oscillations
629: in the plot of $C(r,t)$ vs. $r/R$ diminish with increase in the off-criticality.
630: Further, the discrepancy between the
631: scaling functions for Models B and S is larger for the off-critical case.
632:
633: \subsection{Island Distribution and Excess Energy}
634:
635: Our subsequent results will focus on the case of a critical quench.
636: An alternative method of describing the domain morphology is the
637: island-size distribution. We define an island as
638: a set of aligned spins, all of whose neighbors are either part
639: of the island, or have an antiparallel spin. Tafa et al. \cite{tpk01} have
640: shown that the domain-size distribution in a phase-separating system
641: $\widetilde{P}(l,t)$ exhibits scaling, and has an exponential decay:
642: \ba
643: \label{fx}
644: \widetilde{P}(l,t) = R^{-1} f\left( \frac{l}{R} \right), \quad
645: f(x) \sim e^{-ax}~~\mbox{for}~~x \rightarrow \infty ,
646: \ea
647: where $l$ is the domain size and $a$ is a constant. The corresponding
648: scaling form for the island-size distribution $P(s,t)$ in $d=2$ is obtained as
649: \ba
650: \label{gx}
651: P(s,t) &=& \int_0^\infty dl \delta (s-bl^2) \widetilde{P}(l,t) \nonumber \\
652: &=& \la s \ra^{-1} ~g\left( \frac{s}{\la s \ra} \right) ,
653: \quad g(x) = \frac{1}{2\sqrt{x}} f(\sqrt{x}) ,
654: \ea
655: where $b$ is a geometric factor, and $\la s \ra$ is the average island size.
656:
657: In Fig.~\ref{histo}, we plot
658: \begin{figure}[htb]
659: \centering
660: \includegraphics[width=0.7\textwidth]{fig7.eps}
661: \caption{Scaled probability distributions for island-sizes in Models B (solid
662: line) and S (dashed line) at $T=0.88 T_c$. The data is shown on a linear-log
663: plot as $\sqrt{s \la s \ra} P(s,t)$ vs. $\sqrt{s/\la s \ra}$, suggested by
664: Eqs.~(\ref{fx})-(\ref{gx}). For Model S, the data set corresponds
665: to $t=10^6$ MCS; for Model B, the data set corresponds to $t=3.4 \times 10^5$ MCS.
666: Both domain sizes coincide at these times.
667: \label{histo}}
668: \end{figure}
669: $\sqrt{s \la s \ra} P(s,t)$ vs. $\sqrt{s/\la s \ra}$ for both
670: Models B and S. We make two observations in this context. First,
671: the data for the two models is numerically indistinguishable on the
672: scale of this plot. The subtle differences in the correlation-function
673: data are not seen in the island-size distribution function.
674: Second, the plot in Fig.~\ref{histo} exhibits an exponential decay,
675: as expected from Eqs.~(\ref{fx})-(\ref{gx}).
676:
677: A macroscopic quantity which depends on the density of small islands is
678: the total energy $E(t)$. The interfacial energy for a domain is $\sigma R^{d-1}$,
679: and the number of domains in the system $\sim N/R^d$. Thus, the overall
680: interfacial energy depends on the length scale as $E(t) - E(\infty)
681: \sim N\sigma/R$. In Fig.~\ref{ener}, we plot
682: \begin{figure}[htb]
683: \centering
684: \includegraphics[width=0.8\textwidth]{fig8.ps}
685: \caption{Plot of the energy per site $E(t)/N$ vs. $R^{-1}$ for Model B
686: at temperatures $T=0.63T_c$ (squares) and $T=0.88T_c$ (circles), and
687: Model S at the same two temperatures (diamonds and crosses).
688: \label{ener}}
689: \end{figure}
690: $E(t)/N$ vs. $R^{-1}$ for both Models B and S at $T=0.63 T_c$ and $0.88 T_c$.
691: We observe a power-law convergence of the excess energy with
692: the slope being proportional to the surface tension $\sigma (T)$.
693: Again, the data sets at the same temperature cannot be distinguished on the
694: scale of the plot.
695:
696: \subsection{Aging of the Autocorrelation Function}
697:
698: The data presented so far has focused on the morphological features
699: of the phase-separating system. Let us next study the temporal correlation
700: of the pattern dynamics in Fig.~\ref{snap}. This is measured by the
701: autocorrelation function:
702: \begin{eqnarray}
703: \label{auto}
704: A(t_w,\tau) & = & {1 \over N} \sum_{i=1}^N \left[ \la \sigma_i (t_w)
705: \sigma_i (t_w+\tau) \ra - \la \sigma_i (t_w) \ra \la \sigma_i (t_w+\tau)
706: \ra \right] ,
707: \end{eqnarray}
708: where the times $t_w$ and $(t_w + \tau)$ are measured after the quench at
709: $t=0$. Here, $t_w$ is the reference time for measurement of the autocorrelation
710: function, and is referred to as the {\it waiting time}.
711: The most general correlation function corresponds to unequal space and
712: time, and combines the definitions in Eqs.~(\ref{cor}) and (\ref{auto}).
713: Equilibrium systems are stationary and the corresponding $A(t_w,\tau)$ only depends
714: upon the time-difference $\tau$. On the other hand, for nonequilibrium systems,
715: $A(t_w,\tau)$ depends on both $t_w$ and $\tau$.
716:
717: There have been some earlier studies of $A(t_w,\tau)$
718: for domain growth in kinetic Ising models. There are two
719: mechanisms which drive the decorrelation process: \\
720: (a) First, there are fluctuations in bulk domains, which
721: give a stationary contribution. Huse and Fisher \cite{hf87} studied
722: decorrelation arising from the appearance of a droplet
723: of (say) down-spins in an up-domain. The probability that
724: a droplet of size $R$ appears via fluctuations
725: is $P_d \propto \exp (-\beta \sigma
726: R^{d-1})$. The lifetime of this droplet is
727: $\tau \sim R^{1/\phi}$, where $\phi$ is the growth exponent.
728: Thus, the corresponding autocorrelation
729: function shows a stretched-exponential behavior:
730: \ba
731: && A_{\rm{eq}}(\tau) \simeq \exp (- \beta \sigma \tau^\theta), \nonumber \\
732: && \theta = (d-1) \phi~\mbox{for}~d<d_c~\mbox{and}~1~\mbox{for}~d>d_c.
733: \ea
734: Here, the critical dimensionality is defined by $(d_c-1) \phi = 1$. \\
735: (b) Second, there is decorrelation due to domain-wall motion. This can be
736: either stochastic (due to thermal fluctuations) or systematic (due to the
737: curvature-reduction mechanism). Consider the $T=0$ case, where there are no
738: fluctuations in the bulk or the surface.
739: The characteristic domain-wall velocity decreases with time, so this mechanism
740: gives a non-stationary or aging ($t_w$-dependent) contribution
741: \cite{bckm97}. Fisher and Huse \cite{fh86} used scaling ideas
742: to argue that the aging contribution to $A(t_w,\tau)$ has a power-law
743: dependence on the length scale:
744: \be
745: \label{power}
746: A_{\rm{age}} (t_w,\tau) = \left[ {R(t_w) \over
747: R(t_w+\tau)} \right]^{\lambda}, \quad R(t_w+\tau) \gg R(t_w) .
748: \ee
749: There have been various studies of the aging exponent
750: $\lambda$ in cases with both spin-flip and
751: spin-exchange kinetics \cite{ab94}. For power-law domain growth,
752: Eq.~(\ref{power}) obeys the scaling form $A_{\rm{age}}(t_w,\tau) = h(\tau/t_w)$,
753: which has been observed in some studies of spin glasses \cite{bckm97}.
754:
755: In Fig.~\ref{age}, we plot $A(t_w,\tau)$ vs. $\tau$ for Models B and S for a
756: critical quench to
757: \begin{figure}[htb]
758: \centering
759: \includegraphics[width=0.8\textwidth]{fig9.eps}
760: \caption{Time-dependence of the autocorrelation function for Models B (solid line)
761: and S (dashed line) at $T=0.63T_c$. The waiting times are $t_w = 10^2, 10^3,
762: 10^4, 10^5$ MCS (from left to right).
763: \label{age}}
764: \end{figure}
765: $T=0.63 T_c$. The solid lines denote data for Model B with waiting times
766: $t_w = 10^2, 10^3, 10^4, 10^5$ (from left to right). The dashed lines denote the
767: corresponding data for Model S. As expected, the autocorrelation function decays
768: more rapidly for Model B. We make the following observations in this regard: \\
769: (a) The quantity $A(t_w,\tau)$ exhibits aging, with an explicit dependence on $t_w$
770: for both Models B and S. In general, the decay is slower for larger $t_w$, i.e.,
771: when the domain size of the reference state is larger.
772: Further, the decay is faster for higher temperatures,
773: where larger fluctuations are present in the system. \\
774: (b) The data in Fig.~\ref{age} is plotted on a log-log scale,
775: and exhibits a continuous curvature for both Models B and S. This is not
776: consistent with the simple power-law decay in Eq.~(\ref{power}). As a matter
777: of fact, the autocorrelation data does not even exhibit $\tau/t_w$-scaling,
778: as we have confirmed. In the case of
779: Model B, this is because the decorrelation process is driven by both bulk fluctuations
780: (with a stationary contribution) and domain-wall motion (with a
781: non-stationary contribution). In the case of Model S, bulk fluctuations
782: have been effectively eliminated and one may naively expect to recover
783: power-law decay. However, this is not the case because interfacial
784: fluctuations also contribute to decorrelation. We believe that the scaling
785: behavior in Eq.~(\ref{power}) is only realized in kinetic Ising models or their
786: coarse-grained analogs at $T=0$. In this limit, coarsening occurs only through
787: the systematic motion of interfaces and the system always reduces its energy.
788: However, the $T=0$ limit is not interesting in the context of kinetic Ising
789: models because the evolving system invariably gets trapped in local free-energy
790: minima. \\
791: (c) In recent work, Puri and Kumar \cite{pk04} have studied the decorrelation
792: process in a spin-1 model using a stochastic model based on the continuous-time
793: random walk formalism. We are currently trying to adapt their modeling
794: to understand the behavior in Fig.~\ref{age}.
795:
796: \section{Summary and Discussion}
797:
798: Let us conclude this paper with a summary and discussion of the results presented
799: here. We have studied phase separation in a kinetic Ising model for phase
800: separation mediated by surface diffusion. This model (referred to as Model S)
801: is obtained by imposing a kinetic constraint on the usual Kawasaki kinetic
802: Ising model (referred to as Model B). In general, the surface diffusion
803: mechanism can drive segregation only when the morphology consists of percolated
804: clusters, i.e., for near-critical quenches. We have undertaken Monte Carlo (MC)
805: simulations of Models B and S using
806: multi-spin coding techniques. These provide accelerated algorithms which
807: enable the simulation of large systems for extended times. Our results
808: show that the major difference between the morphologies of Models B and S
809: lies in the growth dynamics. In this regard, it is relevant to emphasize
810: the following: \\
811: (a) The early-time dynamics ($t \ll t_c^B$) of Model B is also dominated by
812: surface diffusion with the growth law $R(t) \sim t^{1/4}$. For late times
813: ($t \gg t_c^B$), there is a crossover to the $t^{1/3}$-growth regime.
814: In the limit of $T \rightarrow 0$, the crossover time diverges
815: ($t_c^B \rightarrow \infty$).
816: However, the low-temperature dynamics of Model B usually
817: freezes into metastable states. Therefore, it is hard to see an extended
818: regime of $t^{1/4}$-growth in Model B. \\
819: (b) Our kinetic constraint eliminates single-particle bulk diffusion,
820: and we see extended regimes of growth driven by surface diffusion. However,
821: $n$-particle diffusion (with $n \geq 2$) is still possible and is governed
822: by the probability for existence of impurities in bulk domains. Thus, at
823: sufficiently large times ($t \gg t_c^S$), we again expect a crossover to
824: $t^{1/3}$-growth. However, this crossover is extremely delayed,
825: even at moderate temperatures. \\
826: (c) We have also studied kinetic models with constraints which
827: eliminate the diffusion of $n$-spin clusters. The domain growth data obtained
828: from these models is numerically indistinguishable from that for Model S. \\
829: (d) For highly off-critical quenches, the morphology consists of droplets
830: of the minority phase in a matrix of the majority phase. In this case, the
831: surface diffusion mechanism cannot drive phase separation. However, Brownian
832: motion and coalescence of droplets also gives rise to $t^{1/4}$-growth in $d=2$.
833:
834: Apart from growth laws, we have also studied quantitative properties
835: of the evolution morphology like correlation functions and island-size
836: distribution functions. There are subtle differences in the scaled
837: correlation functions for Models B and S, but it is difficult to attribute
838: physical significance to these. Further, these differences are not reflected
839: in the island-size distribution function.
840:
841: Finally, we have studied the aging of the autocorrelation function $A(t_w,\tau)$
842: in Models B and S. In both cases, we find that the decorrelation process is
843: driven by both fluctuations and domain-wall motion. Thus, $A(t_w,\tau)$ does
844: not exhibit a simple power-law decay or scaling behavior. We are presently
845: adapting the continuous-time random walk approach developed by
846: Puri and Kumar \cite{pk04} to study the aging of the autocorrelation functions
847: in Models B and S.
848:
849: \newpage
850:
851: \begin{thebibliography}{99}
852:
853: \bibitem{ab94} A.J. Bray, Adv. Phys. \textbf{43}, 357 (1994).
854:
855: \bibitem{bf01} K. Binder and P. Fratzl, in {\it Phase Transformations in
856: Materials}, edited by G. Kostorz (Wiley-VCH, Weinheim, 2001), p. 409.
857:
858: \bibitem{ao02} A. Onuki, \emph{Phase Transition Dynamics} (Cambridge
859: University Press, Cambridge, 2002).
860:
861: \bibitem{dp03} S. Dattagupta and S. Puri, \emph{Dissipative Phenomena
862: in Condensed Matter: Some Applications}, to be published by Springer-Verlag.
863:
864: \bibitem{sp97} S. Puri, Phys. Rev. E \textbf{55}, 1752 (1997);
865: S. Puri and R. Sharma, Phys. Rev. E \textbf{57}, 1873 (1998).
866:
867: \bibitem{es79} E.D. Siggia, Phys. Rev. A \textbf{20}, 595 (1979);
868: H. Furukawa, \pra{31}{1103}{1985}.
869:
870: \bibitem{pcp91} S. Puri, D. Chowdhury and N. Parekh, J. Phys.
871: A \textbf{24}, L1087 (1991); S. Puri and N. Parekh, J. Phys. A
872: \textbf{25}, 4127 (1992).
873:
874: \bibitem{op99} A. Onuki and S. Puri, Phys. Rev. E \textbf{59}, R1331
875: (1999).
876:
877: \bibitem{kk72} K. Kawasaki, in \emph{Phase Transitions and Critical
878: Phenomena Vol. 2}, ed. by C. Domb and M.S. Green (Academic Press,
879: London, 1972), p. 443.
880:
881: \bibitem{bh02} K. Binder and D.W. Heermann, \emph{Monte Carlo
882: Simulation in Statistical Physics: An Introduction}, Fourth Edition
883: (Springer-Verlag, Berlin, 2002).
884:
885: \bibitem{asm88} J. Amar, F. Sullivan and R.D. Mountain,
886: \prb{37}{196}{1988}; and references therein.
887:
888: \bibitem{mb95} J.F. Marko and G.T. Barkema, \pre{52}{2522}{1995};
889: and references therein.
890:
891: \bibitem{hh77} P.C. Hohenberg and B.I. Halperin, Rev. Mod. Phys.
892: \textbf{49}, 435 (1977).
893:
894: \bibitem{kb74} K. Binder, Z. Phys. \textbf{267}, 313 (1974).
895:
896: \bibitem{lbm75} J.S. Langer, M. Bar-on and H.D. Miller, Phys. Rev. A
897: \textbf{11}, 1417 (1975).
898:
899: \bibitem{ki78} K. Kitahara and M. Imada, Prog. Theor. Phys.
900: Suppl. \textbf{64}, 65 (1978); K. Kitahara, Y. Oono and D. Jasnow, Mod.
901: Phys. Lett. B \textbf{2}, 765 (1988).
902:
903: \bibitem{pbd92} S. Puri, K. Binder and S. Dattagupta, Phys.
904: Rev. B \textbf{46}, 98 (1992); S. Puri, N. Parekh and
905: S. Dattagupta, J. Stat. Phys. \textbf{77}, 839 (1994).
906:
907: \bibitem{pbl97} S. Puri, A.J. Bray and J.L. Lebowitz,
908: Phys. Rev. E \textbf{56}, 758 (1997).
909:
910: \bibitem{cy89} C. Yeung, \emph{Some Problems on Spatial Patterns in
911: Nonequilibrium Systems}, Ph.D. Thesis, Illinois (1989).
912:
913: \bibitem{lm92} A.M. Lacasta, A. Hernandez-Machado, J.M. Sancho and R. Toral,
914: \prb{45}{5276}{1992}; A.M. Lacasta, J.M. Sancho, A. Hernandez-Machado and R. Toral,
915: \prb{48}{6854}{1993}.
916:
917: \bibitem{sj97} D. Sappelt and J. Jackle, Europhys. Lett. {\bf 37}, 13 (1997);
918: Polymer {\bf 39}, 5253 (1998).
919:
920: \bibitem{sb93} F. Sciortino, R. Bansil, H.E. Stanley and P. Alstrom,
921: Phys. Rev. E {\bf 47}, 4615 (1993).
922:
923: \bibitem{sp01} J. Sharma and S. Puri, Phys. Rev. E \textbf{64},
924: 021513 (2001).
925:
926: \bibitem{nb99} M.E.J. Newman and G.T. Barkema, \emph{Monte Carlo Methods
927: in Statistical Physics} (Oxford University Press, Oxford, 1999).
928:
929: \bibitem{bs74} K. Binder and D. Stauffer, Phys. Rev. Lett. \textbf{33}, 1006 (1974).
930:
931: \bibitem{to88} T. Ohta, J. Phys. C \textbf{21}, L361 (1988); K. Kawasaki and
932: T. Ohta, Prog. Theor. Phys. \textbf{68}, 129 (1982).
933:
934: \bibitem{hf84} H. Furukawa, Phys. Rev. A {\bf 29}, 2160 (1984).
935:
936: \bibitem{sp88} S. Puri, Phys. Lett. A \textbf{134}, 205 (1988).
937:
938: \bibitem{tpk01} K. Tafa, S. Puri and D. Kumar, \pre{63}{046115}{2001};
939: Phys. Rev. E \textbf{64}, 056139 (2001).
940:
941: \bi{hf87} D.A. Huse and D.S. Fisher, Phys. Rev. B {\bf 35}, 6841 (1987).
942:
943: \bibitem{bckm97} J.-P. Bouchaud, L.F. Cugliandolo, J. Kurchan and
944: M. Mezard, in \emph{Spin Glasses and Random Fields},
945: ed. by A.P. Young (World Scientific, Singapore 1997), p. 161.
946:
947: \bi{fh86} D.S. Fisher and D.A. Huse, Phys. Rev. Lett. {\bf 56}, 1601 (1986);
948: Phys. Rev. B {\bf 38}, 373, 386 (1988).
949:
950: \bibitem{pk04} S. Puri and D. Kumar, \prl{93}{025701}{2004}; Phys. Rev. E \textbf{70},
951: 051501 (2004).
952:
953: \end{thebibliography}
954:
955: \end{document}
956: