cond-mat0211453/mf.tex
1: \documentstyle[prl,aps,epsf,multicol]{revtex}
2: %\topmargin 0.2cm
3: %\tightenlines
4: \begin{document}
5: %\tightenlines
6: \draft
7: 
8: \title{Persistence in $q$-state Potts model: A Mean-Field approach}
9: 
10: \author{G. Manoj}
11: 
12: \address{ Department of Physics and Center for Stochastic Processes in Science and Engineering,\\ 
13: Virginia Polytechnic Institute and State University, Blacksburg, 
14: VA 24061, USA.}
15: 
16: \date{\today}
17: 
18: \maketitle
19: 
20: \begin{abstract}
21: 
22: We study the Persistence properties of the $T=0$ coarsening dynamics of one dimensional
23: $q$-state Potts model using a modified mean-field approximation (MMFA). In this approximation,
24: the spatial correlations between the interfaces separating spins with different
25: Potts states is ignored, but the correct time dependence of the mean density $P(t)$
26: of persistent spins is imposed. For this model, it is known that $P(t)$ follows a power-law
27: decay with time, $P(t)\sim t^{-\theta(q)}$ where $\theta(q)$ is the $q$-dependent
28: persistence exponent.
29: We study the spatial structure of the persistent region within the 
30: MMFA. We show that the persistent site pair correlation function $P_{2}(r,t)$ has the
31: scaling form $P_{2}(r,t)=P(t)^{2}f(r/t^{\frac{1}{2}})$ for all values of 
32: the persistence exponent $\theta(q)$.
33: The scaling function has the limiting behaviour
34: $f(x)\sim x^{-2\theta}$ ($x\ll 1$) and $f(x)\to 1$ ($x\gg 1$). 
35: We then show within the Independent Interval Approximation (IIA) that 
36: the distribution $n(k,t)$ of separation $k$ between two consecutive persistent spins
37: at time $t$ has the asymptotic scaling form $n(k,t)=t^{-2\phi}g(t,\frac{k}{t^{\phi}})$ where the
38: dynamical exponent has the form $\phi$=max($\frac{1}{2},\theta$). The behaviour of the scaling
39: function for large and small values of the arguments is found analytically. We find that 
40: for small separations $k\ll t^{\phi}, n(k,t)\sim P(t)k^{-\tau}$ where $\tau$=max($2(1-\theta),2\theta$),
41: while for large separations $k\gg t^{\phi}$, $g(t,x)$ decays exponentially with $x$.
42: The unusual dynamical scaling form and the behaviour of the scaling function is supported by
43: numerical simulations.
44: 
45: \end{abstract}
46: 
47: \pacs{}
48: 
49: \vspace{1cm}
50: \begin{multicols}{2}
51: \section{Introduction}
52: 
53: In recent times, the notions of Persistence and the associated Persistence exponent
54: has become one of the active topics of research in Non-equilibrium Physics\cite{SATYA}.
55: In general, the persistence probability $P(t)$ is the probability that a stochastic
56: variable $\phi(t)$ remains above or below a certain arbitrary value (say, its initial value)
57: for a time interval $[0:t]$. The idea of Persistence is particularly relevant 
58: in coarsening systems where $P(t)$ has the natural interpretation of being the fraction
59: of space in the system which remains in the same phase, starting from a random initial
60: configuration. In these systems (as in some other systems also) $P(t)$ typically
61: decays as a power-law at large $t$, with an exponent that is non-trivial to compute
62: and does not appear to be simply related to other known exponents that characterize the
63: process.
64: 
65: The coarsening dynamics of the one dimensional $q$-state Potts model at 
66: zero temperature is one of the few cases where the Persistence exponent $\theta(q)$ is 
67: known exactly. The solution  was provided by Derrida {\it et.al} through a mapping
68: of the process to a coagulation model in steady state\cite{DERRIDA}. It was shown that at late times $t$, well beyond
69: the time scale of equilibration, the fraction of persistent spins left in a finite system
70: of linear dimension $L$ scale as $P_{q}(t\gg L^{2}, L)\sim L^{-2\theta(q)}$, 
71: where $\theta(q)$ is given by the non-trivial expression
72: 
73: \begin{equation}
74: \theta(q)=-\frac{1}{8}+\frac{2}{\pi^2}\left[cos^{-1}\left(\frac{2-q}{q\sqrt{2}}\right)\right]^{2}.
75: \label{eq:EQ1}
76: \end{equation}
77: 
78: For times $t\ll L^{2}$, it follows that $P_{q}(t)\sim t^{-\theta(q)}$. An interesting question in this
79: context is about the spatial distribution of spins which are persistent up to any given time
80: $t$. Clearly, in a many-body process like the time evolution of Potts model, the probability
81: that a given spin is persistent is closely linked to the state of other spins. This inter-dependence
82: of spins is crucial in that it makes the time evolution at any single site strongly
83: non-Markovian, which makes the computation of the Persistence exponent highly non-trivial. 
84: A study of the spatial aspects of the Persistence problem is therefore important from the point
85: of view developing an intuitive understanding of the phenomenon, and also illuminates
86: the interplay between Persistence decay and the underlying domain coarsening process.
87: 
88: The spatial distribution of persistent sites and its time evolution have been studied
89: previously through numerical simulations in one dimensional
90: diffusion equation\cite{ZANETTE}, $q$-state Potts model\cite{MANOJ0,MANOJ1,MANOJ2,STEVE1} ,
91: two dimensional Ising model\cite{JAIN} and one dimensional Ising model with parallel dynamics\cite{PURU}. 
92: An analytic study using a rate equation approach under the Independent Interval Approximation(IIA)
93: has been carried out for one dimensional $A+A\to\emptyset$ model which is closely
94: related to 1D Ising model\cite{MANOJ2}. It is now generally understood from physical arguments and 
95: simulations that for a coarsening process
96: in $d$ dimensions where characteristic length scale has the power-law growth
97: $L(t)\sim t^{1/z}$, the set of persistent sites forms a fractal structure with
98: fractal dimension $d_{f}=d-z\theta$ over length scales $r\ll t^{1/z}$\cite{STEVE1,MANOJ3}. The distribution is
99: homogeneous beyond this length scale. Furthermore, since $d_{f}\geq 0$, the distribution is
100: expected to be homogeneous over all length scales if $\theta > d/z$. This has important
101: consequences for systems like the Potts model where $\theta$ changes with the Potts state $q$.
102: In particular, for Potts model in $d=1$, Bray and O'Donoghue\cite{STEVE1} argued that 
103: a transition from fractal to homogeneous distribution occurs as $\theta$ crosses $\frac{1}{2}$.
104: This transition is also marked by an abrupt
105: change in the dynamical exponent characterizing the separation between persistent sites. 
106: The characteristic length scale  was conjectured to have the unusual dynamical scaling
107: form ${\cal L}(t)\sim t^{\phi}$ with $\phi=max(\frac{1}{2},\theta)$.
108: This conjecture based on physical arguments was supported by numerical simulations\cite{STEVE1}. 
109: 
110: %In spite of this, a good understanding of this transition is still lacking. In particular,
111: %the precise asymptotic behaviour of quantities like pair correlation and empty interval distribution
112: %when $q\geq 3$ is not known.
113: In this paper, we use a Mean-Field approach to address the problem of spatial
114: distribution of persistent sites in $q$-state Potts model. 
115: The essential idea behind this approach is as follows. 
116: It is well-known that the $T=0$ coarsening dynamics of the $q$-state Potts model 
117: can be mapped to a reaction-diffusion process. In this process, the interfaces between different
118: species of Potts spins are represented by diffusing particles $A$, which annihilate or coagulate
119: upon meeting with $q$-dependent probabilities. In the
120: mean field approach, these diffusing particles are treated as homogeneously distributed, with (time-dependent) 
121: density equal to the average density in the original reaction-diffusion problem. This approach has been
122: discussed in some earlier works as a  heuristic argument \cite{CARDY} and as 
123: a toy model for persistence \cite{STEVE2}.  We argue that this approach
124: yields a lower bound for the Persistence exponent in the Potts model. We then construct an
125: artificial model which is devoid of spatial correlations among diffusing particles, but with persistence
126: exponent tuned to be exactly equal to the Potts model value. We refer to this model as the
127: Modified Mean-Field Approximation (MMFA) and use this approximation to study the spatial
128: distribution of persistent sites in $q$-state Potts model. 
129: 
130: We outline our main results at this point. Within the MMFA, we show analytically
131: that the correlation length for the spatial distribution of persistent spins scales
132: as $\xi(t)\sim t^{\frac{1}{2}}$ and the equal time pair correlation function
133: $P_{2}({\bf r},t)$ (defined as the probability
134: that the spin at origin and the point ${\bf r}$ are persistent at time $t$) has the scaling
135: form $P_{2}({\bf r},t)=P(t)^{2}f(r/\sqrt{t})$ for any value of $\theta$. This shows that the persistent
136: spins have a fractal distribution with $d_{f}=1-2\theta$ over length scales $r\ll t^{\frac{1}{2}}$
137: when $\theta < \frac{1}{2}$, but $d_{f}=0$ when $\theta\geq \frac{1}{2}$. 
138: We find that the characteristic length scale of the spatial distribution of persistent spins has 
139: the unusual scaling form ${\cal L}(t)\sim t^{\phi}$ where $\phi$=max($\frac{1}{2},\theta$), in
140: agreement with the conjecture in \cite{STEVE1}. 
141: The empty interval
142: distribution itself has the scaling form $n(k,t)={\cal L}(t)^{-2}g[t,k/{\cal L}(t)]$, where
143: $g(t,x)\sim t^{-\psi}x^{-\tau}$ for $x\ll 1$ and decays exponentially with $x$ when $x\gg 1$.
144: The exponents $\psi$ and $\tau$ depends on $\theta$ through the relations $\psi=\theta (2\theta-1)H(\theta-
145: \frac{1}{2})$ and $\tau$=max[$2(1-\theta),2\theta$], where $H(x)$ is the Heaviside step function.
146: We support these results with numerical simulations.
147: 
148: The paper is arranged as follows. In the next section, we outline the mean field
149: approach. In Section III, we introduce the MMFA and compute the pair correlation and Empty
150: Interval Distribution of persistent sites to characterize their spatial distribution. These
151: predictions are compared with the results of numerical simulations
152: in the $q$-state Potts model in Sec. IV. We summarize our results and present our
153: conclusions in Sec. V. 
154: 
155: \section{The Mean-Field approximation}
156: 
157: In the zero temperature coarsening dynamics of $q$-state Potts model in $d=1$, 
158: the interfacial points between different species of Potts spins perform independent random
159: walks on the lattice and annihilate each other with probability $\frac{1}{q-1}$
160: ($\frac{q-2}{q-1}$), or coagulate with probability $\frac{q-2}{q-1}$. In the process,
161: persistent sites are `wiped out', and the surviving random walkers build up spatial
162: correlations among themselves. The distribution of intervals between the surviving random
163: walkers at any (sufficiently late) time $t$ is described by a (stationary) scaling function
164: which is known exactly for all values of $q$\cite{DERRIDA2}. The average density $n(t)$ at time $t$ decays as
165: 
166: \begin{equation}
167: n(t)\simeq \frac{q-1}{q}\frac{1}{\sqrt{2\pi t}}
168: \label{eq:EQ2}
169: \end{equation}
170: 
171: asymptotically\cite{DERRIDA2}. The essential idea behind the mean field (MF) approximation is to treat the 
172: random walkers
173: as forming a homogeneous background of average density $n(t)$, as far as the persistent sites
174: are concerned. We define the Persistence probability $P(t)$ as 
175: the probability that the site at the origin is 
176: unvisited by any walker till time $t$. Then, the probability that the site 
177: the probability that the site is visited by a walker between time
178: $t$ and $t+dt$ is $-\frac{\partial P(t)}{\partial t}$.
179: Let $R(x,t)$ be the probability that the site at origin is visited by
180: a walker for the first time at time $t$, whose initial position was 
181: $x$ at $t=0$. Within the MF approximation, any walker would survive
182: with probability $n(t)$, and the probability that it will make a first
183: visit to origin at time $t$ is given by 
184: $q(x,t)=\sqrt{\frac{2}{\pi}}\frac{x}{t^{3/2}}e^{-\frac{x^2}{2t}}$\cite{FELLER1}.
185: It follows that $R(x,t)=n(t)q(x,t)$.
186: We now integrate $R(x,t)$ over all initial positions $x$ and multiply
187: by the probability that the origin is persistent at time $t$, which is simply $P(t)$. 
188: So we find
189: 
190: \begin{equation}
191: \frac{\partial P(t)}{\partial t}=-P(t)n(t)K_{1}(t)~~~~~~~~d=1
192: \label{eq:EQ3}
193: \end{equation}
194: 
195: where $K_{1}(t)=\int_{-\infty}^{\infty}q(x,t)dx$ is the Smoluchowski constant\cite{SMOL} in $d=1$.
196: After substituting for $q(x,t)$ and $n(t)$, we find $\frac{\partial P(t)}{\partial t}=-\frac{\theta^{*}}{t}P(t)$,
197: so that $P(t)\sim t^{-\theta^{*}(q)}$, and $\theta^{*}(q)=\frac{\sqrt{2}}{\pi}\frac{q-1}{q}$ is the Persistence
198: exponent within the MF model\cite{STEVE2}. It is interesting to compare the mean field prediction for $\theta$ 
199: with the exact value
200: of the exponent. For $q=2$, $\theta^{*}(2)\simeq 0.225$, while the exact value from Eq.\ref{eq:EQ1} is $\frac{3}{8}$.
201: For the $q=\infty$ case, the MF model predicts $\theta^{*}(\infty)\simeq 0.45$, which is to be compared with the exact
202: value $\theta(\infty)=1$. Upon extending the comparison to the entire range of values of $q$, it is clear that
203: the mean field treatment consistently under-estimates the value of $\theta$. 
204: 
205: We now argue that $\theta^{*}(q)$ is, in fact, a lower bound for $\theta(q)$.
206: In the mean field approach discussed so far, it is assumed that the random walkers disappear from the 
207: lattice at random at such a rate so  that their average density falls as $n(t)$. The actual reaction-diffusion process is quite
208: different, because only walkers which come very close to another walker are likely to be removed.
209: Clearly in regions of space where walkers come close to each other, they are likely to visit
210: the same site again and again. This effect is much more within the mean field approach, where the walkers actually
211: pass through each other, possibly several times before disappearing. Thus, it is plausible that for the
212: same average density of walkers, a larger number of persistent sites will be visited in the actual reaction-diffusion 
213: model, compared to its mean field analogue.
214: Since this is true for all times, the average density of persistent sites in mean field theory
215: will be higher than the same in the actual Potts model dynamics. This would naturally imply that
216: 
217: \begin{equation}
218: \theta^{*}(q)\leq \theta(q)
219: \label{eq:EQ4}
220: \end{equation}
221: 
222: Interestingly, we show now that the Mean Field argument presented above yields the correct
223: value for the Persistence exponent for $A+A\to\emptyset$ model in $d=2$. This is not surprising,
224: since for this model, the upper critical dimension is $d_{c}=2$, and the mean field treatment becomes exact above
225: this dimension. It can be shown that in $d=2$, the probability that a random walker starting at an
226: an arbitrary point crosses a circle of radius $a$ around the origin for the first
227: time at $t$ is given by the expression\cite{SMOL2D}
228: 
229: \begin{equation}
230: K_{2}(t)\simeq \frac{4\pi D}{log(4Dt/a^{2})} ~~~~~~~t\gg a^{2}/D
231: \label{eq:EQ5}
232: \end{equation}
233: 
234: which is the analogue of Smoluchowski constant in $d=2$. 
235: The asymptotic particle density decay for $A+A\to\emptyset$ model in $d=2$ has the form 
236: $n(t)\simeq \frac{1}{8\pi}\frac{log(Dt)}{Dt}$\cite{BPLEE}. Upon extending the MF arguments presented
237: previously, we find that 
238: 
239: \begin{equation}
240: \frac{\partial P(t)}{\partial t}=-P(t)n(t)K_{2}(t)~~~~~~~d=2
241: \end{equation}
242: 
243: After substituting for $n(t)$ and $K_{2}(t)$, and taking the limit $a\to 0$, we find that 
244: $P(t)\sim t^{-\frac{1}{2}}$ so that $\theta^{*}=1/2$ in $d=2$. This result is exact, 
245: as has  been shown by a rigorous field-theoretic calculation\cite{CARDY}.
246: 
247: We thus observe that while the mean field approach, in general, gives only a lower bound for the 
248: persistence exponent, it correctly identifies the essential features that brings about this power-law
249: decay, ie., the diffusive motion of interfacial points and the $\frac{1}{\sqrt{t}}$ decay in
250: their over-all density. In the following sections, we use a slightly modified version of this treatment
251: to study the spatial distribution of persistent sites in $q$-state Potts model.
252: 
253: \section{The Modified Mean Field Approximation (MMFA)}
254: 
255: Our purpose is to use the mean field approach to study the spatial distribution of persistent spins
256: in the $q$-state Potts model, and in particular, to understand the transition from fractal to
257: homogeneous distribution as $\theta$ crosses $\frac{1}{2}$. However, it may be noted
258: that in the mean field approximation to the dynamics of the Potts model, the largest value of $\theta$
259: (attained at $q=\infty$) is $\frac{2}{\sqrt{\pi}}\simeq 0.45$ which is less than the transition
260: value $\frac{1}{2}$. This problem is circumvented by defining an artificial model where we define the 
261: diffusing particles as non-interacting random walkers, who can pass through each other. The model
262: also allows for multiple occupancy of lattice sites. The dynamics consists of random walkers being
263: picked at random and taken out of the lattice at a time-dependent rate, which is tuned to produce 
264: power-law decay $P(t)\sim t^{-{\theta}^{\prime}}$ with any arbitrary ${\theta}^{\prime}$. From
265: the arguments presented in the previous section, it is obvious that this would be the case 
266: if the average density were to 
267: decay as $n(t)\sim \sqrt{\frac{\pi}{2}}\frac{{\theta}^{\prime}}{\sqrt{t}}$ asymptotically. 
268: By construction, this model is devoid of spatial correlations among reacting particles 
269: (ie., it is still mean field) but ${\theta}^{\prime}$ is now arbitrary. If we now choose 
270: ${\theta}^{\prime}=\theta(q)$, this model is an approximation to the $q$-state Potts model, with
271: the simplifying feature that the spatial correlation between interfacial points is now absent.
272: We shall refer to this model as the Modified Mean Field Approximation (MMFA) for the original
273: Potts model.
274: 
275: \subsection{Pair correlation for persistent spins within the MMFA}
276: 
277: In this section, we compute the equal time pair correlation function for persistent
278: spins in $q$-state Potts model under the MMFA.  We define $P_{2}(r,t)$ as the probability that 
279: both the site at origin and the site at $r>0$ are persistent at time $t$. Our purpose is
280: to compute $P_{2}(r,t)$ for various $r$.
281: 
282: The generalization of Eq.\ref{eq:EQ3} to this case is
283: 
284: \begin{equation}
285: -\frac{\partial P_{2}(r,t)}{\partial t}=2P_{2}(r,t)\int_{-\infty}^{r}R_{r}(x,t)dx
286: \label{eq:EQ6}
287: \end{equation}
288: 
289: where $R_{r}(x,t)$ is the probability that a particle with initial position
290: $x (-\infty < x < r)$ will make a first visit to the origin at time $t$, {\it without
291: ever crossing $r$ in the interval [0:t]}.
292: The factor 2 in front takes into account the probability that either of the sites
293: could be reached by one of the diffusing particles. Unlike the first case,
294: $R_{r}(x,t)$ is now different for $x<0$ and $0\leq x<r$. For $x<0$, the constraint
295: of no crossing at $r$ is irrelevant for the computation of $R_{r}(x,t)$ , since to 
296: reach $r$, the particle would have to cross the origin first. So $R_{r}(x,t)=R(x,t)$
297: simply for $x<0$, and so 
298: 
299: \begin{equation}
300: \int_{-\infty}^{0}R_{r}(x,t)dx=\frac{\theta}{2t}
301: \label{eq:EQ7}
302: \end{equation}
303: 
304: For $x>0$, this is no longer true, and $R_{r}(x,t)$ needs to be computed
305: separately. The quantity that we need here is $q_{r}(x,t)$, the probability
306: that a diffusing particle whose position at $t=0$ is $x$, will reach the origin
307: for the first time at $t$, without ever crossing the point $r$ in between.
308: Then $R_{r}(x,t)=n(t)q_{r}(x,t)$.
309: To find $q_{r}(x,t)$, let us use the following standard method. Consider a random
310: walk starting from $0< x < r$ at $t=0$ with absorbing barriers at $0$ and $r$.
311: If the probability distribution of the position $z$ of the walker at time $t$ is
312: $u_{x}(z,t)$, then 
313: 
314: \begin{equation}
315: q_{r}(x,t)=\frac{\partial u_{x}(z,t)}{\partial z}|_{z=0}
316: \label{eq:EQ8}
317: \end{equation}
318: 
319: The expression for $u_{x}(z,t)$ is known exactly, and the asymptotic form at
320: large $t$\cite{FELLER1} is
321: 
322: \begin{equation}
323: u_{x}(z,t)=\frac{1}{\sqrt{2\pi t}}\sum_{k=-\infty}^{\infty}e^{-\frac{(z-x-2kr)^2}{2t}}
324: -e^{-\frac{(z+x-2kr)^2}{2t}}
325: \label{eq:EQ9}
326: \end{equation}
327: 
328: from which, we find
329: 
330: \begin{eqnarray}
331: q_{r}(x,t)=\frac{1}{\sqrt{2\pi t}}\sum_{k=-\infty}^{\infty}
332: \frac{x+2kr}{t}e^{-\frac{(z-x-2kr)^2}{2t}}-\nonumber \\
333: \frac{x-2kr}{t}e^{-\frac{(z+x-2kr)^2}{2t}}
334: \label{eq:EQ10}
335: \end{eqnarray}
336: 
337: We note that for $r\gg t^{\frac{1}{2}}$, the $k=0$ mode is the dominant term in
338: the sum, and this gives $q_{r}(x,t)\approx \sqrt{\frac{2}{\pi t^3}}xe^{-\frac{x^2}{2t}}$+
339: smaller terms that vanish as $\frac{r}{t^{1/2}}\to\infty$. Clearly in this limit, we
340: recover the $r=\infty$ term, as we should. It then follows that
341: 
342: \begin{equation}
343: \int_{0}^{r}R_{r}(x,t)dx=\frac{\theta}{2t}G\left(\frac{r}{\sqrt{t}}\right)
344: \label{eq:EQ11}
345: \end{equation}
346: 
347: where
348: 
349: \begin{equation}
350: G(x)=1-\eta+\sum_{k=1}^{\infty}2\eta^{4k^2}-\eta^{(1+2k)^2}-\eta^{(1-2k)^2}
351: \label{eq:EQ12}
352: \end{equation}
353: 
354: with $\eta=e^{-\frac{x^2}{2}}$. After substitution of Eq.\ref{eq:EQ7} and Eq.\ref{eq:EQ11} 
355:  in Eq. \ref{eq:EQ6}, we find 
356: 
357: \begin{figure}
358: \epsfxsize=2.3in
359: \epsfbox{MFfig1.ps}
360: \caption{Plot of $G(x)$ vs $x$}
361: \end{figure}
362: 
363: \begin{equation}
364: \frac{\partial P_{2}(r,t)}{\partial t}=-2P_{2}(r,t)\frac{\theta}{2t}\left[1+G(\frac{r}{\sqrt{t}})\right]
365: \label{eq:EQ13}
366: \end{equation}
367: 
368: which admits a scaling solution of the form
369: 
370: \begin{equation}
371: P_{2}(r,t)=P(t)^{2}f(\frac{r}{\sqrt{t}})
372: \label{eq:EQ14}
373: \end{equation}
374: 
375: and the scaling function $f(x)$ is given by the following expression.
376: 
377: \begin{equation}
378: \frac{x}{2}\frac{\partial f}{\partial x}=-\theta f(x)\left[1-G(x)\right]
379: \label{eq:EQ15}
380: \end{equation}
381: 
382: Let us now consider the limiting behaviour of the scaling function for $x\ll 1$ and
383: $x\gg 1$. In the first case, it is clear from Fig.1 that $G(x)\approx 0$, and so
384: $\frac{x}{2}\frac{\partial f}{\partial x}=-\theta f(x)$, which implies that
385: $f(x)\sim x^{-2\theta}$ as $x\to 0$. In the opposite extreme $G(x)\to 1$ as $x\to\infty$,
386: and so $\frac{x}{2}\frac{\partial f}{\partial x}\approx 0$, which means that
387: $f(x)$ approaches a constant value in this limit. It is further clear that, from the definition of the scaling
388: function as given by the expression Eq.\ref{eq:EQ14}, this constant is unity, since
389: we expect $P_{2}(r,t)\to P(t)^{2}$ as $r\to\infty$. 
390: For convenience of later calculations, we approximate the scaling function as 
391: 
392: \begin{eqnarray}
393: f(x)= a^{2\theta}x^{-2\theta}~~~~~~:x\leq a\nonumber \\
394: f(x)= 1~~~~~~~~~~:x> a
395: \label{eq:EQ16}
396: \end{eqnarray}
397: 
398: where $a$ is a number, of order unity.
399: 
400: We see that under the MMFA, the pair correlation function has a scaling form which is same for 
401: all values of $\theta$, with power-law decay $P_{2}(r,t)\sim P(t)r^{-2\theta}$ for short
402: distances $r\ll t^{\frac{1}{2}}$.
403: As is well-known, power law decay of pair correlation function points to the underlying
404: scale invariance of the spatial distribution of the persistent spins. This is characteristic
405: of a fractal distribution under some circumstances. To see this, let us first define
406: $C(r,t)=P(t)^{-1}P_{2}(r,t)$, which is the probability of finding a persistent spin at
407: a distance $r$ from another persistent spin. Now, the integral $M(R,t)=\int_{1}^{R}C(r,t)dr$
408: is the total number of persistent spins within a radius $R$ of a persistent spin. Clearly,
409: from the scaling form described above, $M(R,t)\sim R^{d_{f}}$ for $R\ll t^{\frac{1}{2}}$,
410: where $d_{f}=$max$(1-2\theta,0)$. For $R\gg t^{\frac{1}{2}}$, we find that $M(R,t)\simeq RP(t)$,
411: which is simply a homogeneous distribution. Thus, if we look at length scales
412: $R\ll t^{\frac{1}{2}}$, there is a fractal structure when $\theta < \frac{1}{2}$. However, 
413: when $\theta\geq \frac{1}{2}$, this scale-invariant structure is replaced by a few isolated sites, 
414: whose number does not grow with the length scale of observation. 
415: 
416: Clearly, the spatial distribution of persistent
417: spins undergoes a transition as $\theta$ crosses $\frac{1}{2}$. Indeed, if we consider time
418: scales beyond equilibration time $t\gg L^{2}$, for $\theta < \frac{1}{2}$ the total
419: number of persistent spins left in the system scales as $L^{1-2\theta}$, whereas for 
420: $\theta\geq \frac{1}{2}$, there are only a finite number of persistent spins left.
421: This important difference is not adequately reflected in the pair correlation function, which
422: has the same scaling form for all values of $\theta$. In the next section, we study another
423: quantity to characterize the spatial distribution which undergoes a rather significant
424: change in its scaling properties across the transition. This quantity is the empty interval
425: distribution, which is one of the standard tools in the study of one-dimensional reaction-diffusion
426: processes.
427: 
428: \subsection{The Empty Interval Distribution}
429: 
430: An empty interval, in our convention, is the separation between two consecutive persistent
431: sites. The empty interval distribution (EID) $n(k,t)$ is defined as the number of such intervals
432: of length $k$ at time $t$. For convenience, we also divide this quantity with the system
433: size $N$ so that $n(k,t)$ satisfies the following normalization conditions.
434: 
435: \begin{equation}
436: \int_{1}^{\infty}n(k,t)dk=P(t)\hspace{0.3cm};\int_{1}^{\infty}kn(k,t)dk=1
437: \label{eq:EQ17}
438: \end{equation}
439: 
440: Computing the EID directly, even under the mean field approximation, is non-trivial.
441: Instead we shall compute it from the pair correlation function using the Independent
442: Interval Approximation (IIA), where the lengths of successive empty intervals are
443: considered as independent random variables. The IIA has been a valuable tool in the study
444: of one-dimensional problems, and has been successfully applied to study spatial distribution
445: of persistent spins in $A+A\to\emptyset$ model.
446: Under the IIA, the relation between $n(k,t)$ and $P_{2}(r,t)$ is 
447: 
448: \begin{equation}
449: P_{2}(r,t)=n(r,t)+P(t)^{-1}\int_{1}^{r}dx n(x,t)P_{2}(r-x,t)
450: \label{eq:EQ18}
451: \end{equation}
452: 
453: It is convenient to express this relation in terms of the Laplace
454: transforms ${\tilde C}(p,t)=\int_{1}^{\infty}C(r,t)e^{-pr}dr$ and ${\tilde n}(p,t)=
455: \int_{1}^{\infty}n(s,t)e^{-ps}ds$, where $C(r,t)$ was defined in the previous
456: section. Under these transformations, Eq.\ref{eq:EQ18} maybe expressed in the form
457: 
458: \begin{equation}
459: {\tilde n}(p,t)=\frac{P(t){\tilde C}(p,t)}{1+{\tilde C}(p,t)}
460: \label{eq:EQ19}
461: \end{equation}
462: 
463: From the scaling form for $P_{2}(r,t)$ given by Eq.\ref{eq:EQ14} , we find that
464: 
465: \begin{equation}
466: {\tilde C}(p,t)=P(t)\sqrt{t}I(q,t)
467: \label{eq:EQ20}
468: \end{equation}
469: 
470: where $q=p\sqrt{t}$, and $I(q,t)=\int_{t^{-\frac{1}{2}}}^{\infty}f(x)e^{-qx}dx$
471: The lower limit is put as $t^{-\frac{1}{2}}$ instead of zero to
472: take care of possible small argument divergence in the scaling function.
473: 
474: Let us first consider the case where $\theta< \frac{1}{2}$:
475: In this case the scaling function $f(x)$ is integrable, so we put the lower limit
476: in the previous equation as zero. Using Eq.\ref{eq:EQ16}, we find that
477: 
478: \begin{equation}
479: I(q,t)=a^{2\theta}q^{2\theta-1}\gamma(1-2\theta,qa)+\frac{1}{q}e^{-qa}~~~~~; {\theta} < \frac{1}{2}
480: \label{eq:EQ21}
481: \end{equation}
482: 
483: where $\gamma(\alpha,x)=\int_{0}^{x}e^{-t}t^{\alpha-1}dt$ is the incomplete Gamma function.
484: After substituting in Eq.\ref{eq:EQ19} and Eq. \ref{eq:EQ20}, and taking the 
485: $t\to\infty$ limit (keeping $q$ fixed), we find
486: 
487: %After substituting in Eq.20 and Eq.19, we find that
488: 
489: %\begin{eqnarray}
490: %{\tilde n}(p,t)=P(t)-\nonumber \\t^{-\frac{1}{2}}\frac{q}{qt^{-\frac{1}{2}}P(t)^{-1}+a^{2\theta}q^{2\theta}\gamma(1-2\theta,qa)+e^{-qa}}
491: %\end{eqnarray}
492: 
493: %Now we take the $t\to\infty$ limit, then the first term in the denominator of the second
494: %term drops out, and we find
495: 
496: \begin{equation}
497: {\tilde n}(p,t)=t^{-\frac{1}{2}}\left[P(t)t^{\frac{1}{2}}-
498: \frac{q}{a^{2\theta}q^{2\theta}\gamma(1-2\theta,qa)+e^{-qa}}\right]
499: \label{eq:EQ22}
500: \end{equation}
501: 
502: It follows that 
503: 
504: \begin{equation}
505: n(k,t)=t^{-1}h\left(\frac{k}{t^{\frac{1}{2}}}\right)
506: \label{eq:EQ23}
507: \end{equation}
508: 
509: so that 
510: 
511: \begin{equation}
512: {\tilde n}(p,t)=t^{-\frac{1}{2}}[t^{\frac{1}{2}}P(t)-h_{1}(q)]
513: \label{eq:EQ24}
514: \end{equation}
515: 
516: which has the same form as Eq.\ref{eq:EQ22}, and 
517: $h_{1}(q)=\int_{0}^{\infty}h(x)[1-e^{-qx}]dx$. After integrating by parts, we find
518: 
519: \begin{equation}
520: h_{1}(q)=-G(\infty)+q\left[\lim_{x\to 0}xG(x)+{\tilde G}(q)\right]
521: \label{eq:EQ25}
522: \end{equation}
523: 
524: where $G(x)=\int_{x}^{\infty}h(y)dy$ and ${\tilde G}(q)=\int_{0}^{\infty}G(x)e^{-qx}dx$.
525: We assume that $G(x)$ is integrable, so that $\lim_{x\to 0}xG(x)=0$ and $G(\infty)=0$.
526: Finally we have 
527: 
528: \begin{equation}
529: {\tilde G}(q)=\frac{1}{a^{2\theta}q^{2\theta}\gamma(1-2\theta,qa)+e^{-qa}}
530: \label{eq:EQ26}
531: \end{equation}
532: 
533: Now we may try to deduce the behaviour of the function $G(x)$ at large and small
534: arguments from its Laplace transform, given by the previous equation. To find
535: the behaviour of $G(x)$ near $x=0$, we use the standard formula\cite{LAPLACE}
536: 
537: \begin{equation}
538: lim_{t\to 0}t^{-\rho}g(t)=lim_{s\to\infty}\frac{s^{\rho+1}{\tilde g(s)}}{\rho!}~~~~;\rho > -1
539: \label{eq:EQ27}
540: \end{equation}
541: 
542: where ${\tilde g(s)}$ is the L.T of $g(t)$. Now,
543: for large $q$, $\gamma(1-2\theta,qa)\simeq \Gamma(1-2\theta)$, so that 
544: ${\tilde G}(q)=\frac{a^{-2\theta}q^{-2\theta}}{\Gamma(1-2\theta)}$ as $q\to\infty$. 
545: It follows that $lim_{x\to 0}x^{1-2\theta}G(x)=\frac{a^{-2\theta}}{(2\theta-1)!\Gamma(1-2\theta)}$
546: from which we find
547: 
548: \begin{equation}
549: G(x)\sim \frac{a^{-2\theta}x^{2\theta-1}}{(2\theta-1)!\Gamma(1-2\theta)}~~~~x\to 0
550: \label{eq:EQ28}
551: \end{equation}
552: 
553: and, after using the relation $h(x)=-\frac{\partial G}{\partial x}$, 
554: 
555: \begin{equation}
556: h(x)\sim x^{-2(1-\theta)}~~~~; x\to 0
557: \label{eq:EQ29}
558: \end{equation}
559: 
560: 
561: 
562: 
563: The behaviour of $G(x)$ at large $x$, one has to look for the singularities of
564: ${\tilde G}(q)$ in Eq. \ref{eq:EQ26}. If ${\tilde G}(q)$ has a 
565: singularity of the form ${\tilde G}(q)\sim (q-q^{*})^{-\nu}$, then, upon inversion
566: of the L.T, it follows that $G(x)\sim x^{\nu-1}e^{q^{*}x}$ as $x\to\infty$\cite{LAPLACE}.
567: In order to find the singularity, we plotted the denominator of 
568: Eq. \ref{eq:EQ26} against its argument (Fig. 2). We find that the function
569: crosses zero at one point in the negative $q$ axis. By careful numerical analysis using
570: bisection method, we
571: have determined this crossing point to be at $q^{*}a=-\lambda$, where the numerical 
572: constant $\lambda\simeq 0.32$ for $\theta=\frac{3}{8}$ and $\lambda\simeq 0.85$ for
573: $\theta=\frac{1}{2}$. This implies that the leading term in the decay of $G(x)$ at large $x$ is
574: exponential, ie., $G(x)\sim exp(-\frac{\lambda}{a}x)$ as $x\to \infty$, with a possible 
575: power-law prefactor. Consequently, the scaling function $h(x)$ also has similar exponential
576: decay at large $x$. 
577: 
578: \begin{equation}
579: h(x)\sim e^{-\frac{\lambda}{a}x}~~~~~~; x\gg 1
580: \label{eq:EQ30}
581: \end{equation}
582:  
583: 
584: \begin{figure}
585: \epsfxsize=2.3in
586: \epsfbox{MFfig2.ps}
587: \caption{The figure shows the inverse of ${\tilde G}(q)$ plotted against $qa$.}
588: \end{figure}
589: 
590: We also determine the characteristic length scales of the distribution using
591: the scaling form for $n(k,t)$. These may be defined as the ratios of moments
592: of the distribution.
593: 
594: \begin{equation}
595: L_{m}(t)=I_{m}(t)/I_{m-1}(t)~~~~~; m=0,1,2,......
596: \label{eq:EQ31}
597: \end{equation}
598: 
599: where $I_{m}(t)=\sum_{k}n(k,t)k^{m}$ are the moments of $n(k,t)$.
600: Clearly, $L_{1}(t)=P(t)^{-1}\sim t^{\theta}$ by definition, whereas all higher
601: order length scales 
602: 
603: \begin{equation}
604: L_{j}(t)\sim t^{\frac{1}{2}}~~~~~; j >1
605: \label{eq:EQ32}
606: \end{equation}
607: 
608: which follows from the dynamic scaling form given by Eq.\ref{eq:EQ23}, \ref{eq:EQ29} and \ref{eq:EQ30} 
609: for $n(k,t)$.
610: 
611: We now continue our study of empty interval distribution for the case where $\theta\geq \frac{1}{2}$.
612: Although the MMFA allows us to study arbitrarily large values of $\theta$, we restrict ourselves
613: to the regime $\theta < 1$, since our basic aim is to study the persistence in $q$-state Potts model
614: where $\theta(q)\leq 1$. Furthermore, for $q=\infty$ where $\theta=1$, $n(k,t)$ can be
615: found exactly\cite{STEVE1} and is known to be a pure exponential.
616: 
617: For $\frac{1}{2}< \theta < 1$, the scaling function $f(x)$ has a non-integrable $x^{-2\theta}$ 
618: singularity near $x=0$ (We do not study explicitly the logarithmic singularity occurring
619: for $\theta = \frac{1}{2}$). We integrate by parts and find
620: 
621: \begin{eqnarray}
622: I(q,t)=\frac{a^{2\theta}}{2\theta-1}[t^{\theta-\frac{1}{2}}e^{-qt^{-\frac{1}{2}}}-a^{1-2
623: \theta}e^{-qa}-\nonumber \\
624: q^{2\theta-1}\gamma(2-2\theta,qa)]+\frac{1}{q}e^{-qa}~~~~~; {\theta}\geq \frac{1}{2}
625: \label{eq:EQ33}
626: \end{eqnarray}
627: 
628: Let us now define $\lambda=\frac{p}{P(t)}=qt^{-\frac{1}{2}}P(t)^{-1}$. For $t\to\infty$ and
629: finite $\lambda$, we have
630: 
631: \begin{equation}
632: I(q,t)=\frac{a^{2\theta}}{2\theta-1} t^{\theta-\frac{1}{2}}+\frac{1}{\lambda t^{\frac{1}{2}}P(t)}
633: \label{eq:EQ34}
634: \end{equation}
635: 
636: After substitution in Eq.\ref{eq:EQ20} we find
637: 
638: \begin{equation}
639: {\tilde C}(p,t)=P(t)\left[\frac{a^{2\theta}}{2\theta-1}t^{\theta}+\frac{1}{\lambda P(t)}\right]
640: \label{eq:EQ35}
641: \end{equation}
642: 
643: We define the constant $\beta=\frac{t^{\theta}P(t)a^{2\theta}}{2\theta-1}$, in terms of which
644: ${\tilde C}(p,t)=\beta+\frac{1}{\lambda}$. Now we substitute in Eq.\ref{eq:EQ19} and find
645: 
646: \begin{equation}
647: {\tilde n}(p,t)=P(t)\frac{\beta}{1+\beta}\left[\frac{\lambda+\beta^{-1}}{\lambda+(1+\beta)^{-1}}\right]
648: \label{eq:EQ36}
649: \end{equation}
650: 
651: Upon inversion of the L.T, we find that 
652: 
653: \begin{equation}
654: n(k,t)=P(t)^{2}\phi[kP(t)]
655: \label{eq:EQ36A}
656: \end{equation}
657: 
658: where 
659: 
660: \begin{equation}
661: \phi(x)=\frac{\beta}{1+\beta}\left[\delta(x)+\frac{1}{\beta(1+\beta)}e^{-x(1+\beta)^{-1}}\right]
662: \label{eq:EQ36B}
663: \end{equation}
664: 
665: The scaling function has a rather unnatural $\delta$-function singularity at the origin. 
666: However, a more careful analysis show that for any finite (but still large) time $t$, the
667: divergence at origin is only power-law, but with a different exponent than the
668: previous case ($\theta < \frac{1}{2}$). We start with the expression given by
669: Eq.\ref{eq:EQ20} and Eq.\ref{eq:EQ33} for ${\tilde C}(p,t)$.
670: After keeping the leading finite $t$ correction, we find that
671: 
672: \begin{equation}
673: {\tilde C}(p,t)=\beta+\frac{1}{\lambda}-
674: {\beta}t^{-\theta(2\theta-1)}\lambda^{2\theta-1}
675: \label{eq:EQ37}
676: \end{equation}
677: 
678: For purposes that will be clear later, let us define $m(k,t)=kn(k,t)$ so that
679: $\sum_{k}m(k,t)=1$. We also define the associated Laplace transform ${\tilde m}(p,t)$.
680: The Laplace transforms are related through
681: 
682: \begin{equation}
683: {\tilde m}(p,t)=-\frac{\partial {\tilde n}(p,t)}{\partial p}
684: \label{eq:EQ38}
685: \end{equation}
686: 
687: Using the expression Eq.\ref{eq:EQ19} for ${\tilde n}(p,t)$, we find that
688: 
689: \begin{equation}
690: {\tilde m}(p,t)=-\frac{P(t){\tilde C}^{\prime}(p,t)}{[1+{\tilde C}(p,t)]^{2}}
691: \label{eq:EQ39}
692: \end{equation}
693: 
694: where ${\tilde C}^{\prime}(p,t)=\frac{\partial {\tilde C}(p,t)}{\partial p}$, and is given by
695: the expression
696: 
697: \begin{equation}
698: {\tilde C}^{\prime}(p,t)=-\frac{1}{P(t)}
699: \left[\frac{1}{\lambda^{2}}+(2\theta-1)\beta t^{-\theta(2\theta-1)}\lambda^{2(\theta-1)}\right]
700: \label{eq:EQ40}
701: \end{equation}
702: 
703: After substitution in Eq.\ref{eq:EQ38} and taking the limit $t\to\infty$, we find
704: 
705: \begin{equation}
706: {\tilde m}(p,t)=\frac{1+(2\theta-1)\beta t^{-\theta(2\theta-1)}\lambda^{2\theta}}
707: {[1+\lambda(1+\beta)]^{2}}
708: \label{eq:EQ41}
709: \end{equation}
710: 
711: which gives the scaling forms 
712: 
713: \begin{equation}
714: m(k,t)=P(t)\psi(t,kP(t))~~~;~~~n(k,t)=P(t)^{2}\Phi(t,kP(t))
715: \label{eq:EQ42}
716: \end{equation}
717: 
718: where 
719: 
720: \begin{equation}
721: x\Phi(t,x)=\psi(t,x)
722: \label{eq:EQ43}
723: \end{equation}
724: 
725: by definition. The Laplace transform of the scaling function $\psi(t,x)$ is
726: 
727: \begin{equation}
728: {\tilde \psi}(t,\lambda)=\frac{1+(2\theta-1)\beta t^{-\theta(2\theta-1)}\lambda^{2\theta}}
729: {[1+\lambda(1+\beta)]^{2}}
730: \label{eq:EQ44}
731: \end{equation}
732: 
733: We notice that if the finite $t$ correction term is not included, 
734: $lim_{\lambda\to\infty}\lambda^{2}{\tilde \psi}(t,\lambda)$ is finite, and in that case, the
735: small argument divergence of $\psi(t,x)$ will be sharper than any power-law. This is what
736: is reflected in the appearance of the $\delta$-function in Eq.\ref{eq:EQ38}. 
737: However, when this term is
738: included, the multiplying factor has to be $\lambda^{2-2\theta}$ in order to make the resulting
739: expression finite as $\lambda\to\infty$. This implies that the small $x$ divergence for $\psi(t,x)$
740: has the power-law form $\psi(t,x)\sim t^{-\theta(2\theta-1)}x^{1-2\theta}$ for small $x$. From
741: Eq. \ref{eq:EQ45}, we find a similar power-law divergence in $\Phi(t,x)$ also.
742: 
743: \begin{equation}
744: \Phi(t,x)\sim t^{-\theta(2\theta-1)}x^{-2\theta}~~~~~; x\ll 1
745: \label{eq:EQ45}
746: \end{equation}
747: 
748: In the large $x$ limit, $\Phi(t,x)$ becomes time independent, and decays exponentially
749: with $x$ as in Eq. \ref{eq:EQ36B}, ie., 
750: 
751: \begin{equation}
752: \Phi(t,x)\simeq \frac{1}{(1+\beta)^2}e^{-\frac{x}{1+\beta}}~~~~; x\gg 1
753: \label{eq:EQ46}
754: \end{equation}
755: 
756: The characteristic length scales are easy to compute. From the scaling form, it
757: follows that all the characteristic lengths have identical asymptotic scaling
758: behaviour.
759: 
760: \begin{equation}
761: L_{j}(t)\sim t^{\theta}~~~~~; j=1,2,...
762: \label{eq:EQ47}
763: \end{equation}
764: 
765: The difference in the asymptotic scaling behaviour of the characteristic length
766: scale as $\theta$ crosses $\frac{1}{2}$ may be seen as a competition between two length scales, 
767: the diffusive length scale ${\cal L}_{D}(t)\sim \sqrt{Dt}$ which gives the 
768: mean separation between two random walkers, and the persistence scale 
769: ${\cal L}_{p}(t)=P(t)^{-1}\sim t^{\theta}$ which is the mean separation between two
770: persistent spins. The characteristic length scale is dominated by
771: the larger of the two, ie., we may write 
772: 
773: \begin{equation}
774: {\cal L}(t)\sim t^{\phi}~~~~; \phi=$max$(\frac{1}{2},\theta)
775: \label{eq:EQ48}
776: \end{equation}
777: 
778: where ${\cal L}(t)$ is defined through the dynamical scaling form for $n(k,t)$.
779: 
780: \begin{equation}
781: n(k,t)={\cal L}(t)^{-2}g(t,\frac{k}{{\cal L}(t)})
782: \label{eq:EQ49}
783: \end{equation}
784: 
785: The scaling function $g(t,x)=h(x)$ when $\theta < \frac{1}{2}$ and $g(t,x)=\Phi(t,x)$ when
786: $\theta \geq \frac{1}{2}$. In general, the small argument behaviour of $g(t,x)$ has the
787: power-law form
788: 
789: \begin{equation}
790: g(t,x)\sim t^{-\psi}x^{-\tau}~~~~; x\to 0
791: \label{eq:EQ50}
792: \end{equation}
793: 
794: where the exponents $\psi$ and $\tau$ are given by
795: 
796: \begin{equation}
797: \psi=\theta(2\theta-1)H(\theta-\frac{1}{2})~~;~~\tau=$max$[2\theta,2(1-\theta)],
798: \label{eq:EQ51}
799: \end{equation}
800: 
801: and $H(x)$ is the Heaviside step function. For large $x$, the scaling function is time-independent 
802: and decays exponentially with $x$.
803: We also note from the scaling form that over small distances $k\ll t^{\phi}$, 
804: 
805: \begin{equation}
806: n(k,t)\sim P(t)k^{-\tau}~~~~;k\ll t^{\phi}
807: \label{eq:EQ52}
808: \end{equation}
809: 
810: where $\tau$ is given by Eq.\ref{eq:EQ51}. 
811: 
812: \section{NUMERICAL RESULTS}
813: 
814: We studied the quantities $P_{2}(r,t)$ and $n(k,t)$ numerically by simulating the kinetics
815: of $q$-state Potts model with random initial conditions. The time evolution of spin
816: configurations via Glauber dynamics is implemented using the mapping of this dynamics
817: to a reaction-diffusion problem, as mentioned in the introduction. In this procedure,
818: a set of diffusing particles $A$ are initially distributed at random on the lattice
819: with a certain average initial density $n_{0}$ (which we fix as $\frac{1}{2}$). When two
820: diffusing particles meet, they annihilate each other or coagulate with probability
821: $\frac{1}{q-1}$ and $\frac{q-2}{q-1}$ respectively. We count one MC step in the simulation
822: after the position of every particle in the lattice has been updated once. Persistent
823: spins(sites) at any time $t$ are those sites which have not been touched by a random
824: walker till that time. All the simulations were done on a lattice with $2^{17}$ sites, and 
825: the results were averaged over 100 different starting configurations. In order to check the
826: different dynamic scaling behaviour for $\theta <\frac{1}{2}$ and $\theta\geq \frac{1}{2}$,
827: we did our simulations for three different values of $q$-2,5 and 10. For later reference, we
828: note that from Eq.\ref{eq:EQ1}, the corresponding values of the persistence exponent are
829: $\theta(2)=3/8=0.375, \theta(5)\approx 0.6928$ and $\theta(10)\approx 0.8310$. 
830: In Figs. 3-5, and later in Figs. 7-10, we have employed
831: logarithmic binning of the data in intervals of size $1.5^{n}$ ($n=1,2,....$). 
832: since the statistical noise was considerable. However, for all exponent measurements, we have
833: used only the bare(not binned) data.
834: 
835: In Fig. 3-5, we have plotted the scaling function $f(x)$ for the pair correlation function
836: $P_{2}(r,t)$ against  the scaling variable $r/\sqrt{t}$ for three $q$-values, $q=2,5$ and 10. 
837: We find excellent scaling collapse for all three values of $q$, which is in agreement with
838: the dynamic scaling picture provided by the MMFA in Eq.\ref{eq:EQ14}. 
839: In the figures, we find power-law decay of $f(x)$ for small $x$, with a sharp cross-over to 
840: the flat long-distance behaviour, which is also in agreement with the assumption we made
841: in Eq.\ref{eq:EQ16}. We also note that the constant $a$ introduced in Eq.\ref{eq:EQ16} 
842: is in fact very close to 1. 
843: 
844: \begin{figure}
845: \epsfxsize=1.8in
846: \epsfbox{MFfig3.ps}
847: \caption{The scaled pair correlation $f(x)=P(t)^{-2}P_{2}(r,t)$ is plotted against
848: the dimensionless scaling variable $x=r/\sqrt{t}$ for $q=2$ on a logarithmic scale. The straight line
849: is a guide to eye, and has slope $2\theta(2)=0.75$ , which is the MMFA prediction. The time $t$ is
850: measured in MC steps and distance $r$ is measured in units of lattice spacing.}
851: \end{figure}
852: 
853: \begin{figure}
854: \epsfxsize=1.8in
855: \epsfbox{MFfig4.ps}
856: \caption{Same as Fig.3, for $q=5$. The slope of the straight line is $2\theta(5)\simeq 1.38$,
857: which is the MMFA prediction.}
858: \end{figure}
859: 
860: \begin{figure}
861: \epsfxsize=1.8in
862: \epsfbox{MFfig5.ps}
863: \caption{Same as Fig.3, for $q=10$.The slope of the straight line is $2\theta(10)\simeq 1.67$,
864: which is the MMFA prediction.}
865: \end{figure}
866: 
867: 
868: In Fig. 6, we plot the characteristic length scale $L_{2}(t)$ against time $t$ for $q=2,5$
869: and 10 and measure the dynamical exponent $\phi$. 
870: The observed slopes of the lines are systematically higher than the theoretical
871: prediction in Eq.\ref{eq:EQ50} by around 0.05, while the statistical error in all the
872: three cases was only $\sim 10^{-4}$ or smaller. The observed deviation could be possibly
873: due to the fact that the asymptotic behaviour is not fully reflected over the time scales
874: which we used. The number of persistent spins left in the system falls rapidly with time for
875: high values of $q$, and so we were forced to restrict ourselves to times $t\leq 32000$. In fact,
876: even for $q=2$ case, previous simulations over longer time scales have shown the presence
877: of an additive power-law correction to the asymptotic scaling behaviour\cite{MANOJ2}.
878: 
879: 
880: \begin{figure}
881: \epsfxsize=1.8in
882: \epsfbox{MFfig6.ps}
883: \caption{The figure shows the characteristic length scale $L_{2}(t)$ (measured in units
884: of lattice spacing, definition
885: in text) plotted against time $t$ (measured as number of MC steps) on a logarithmic scale 
886: for three Potts values $q=2,5$ and 10. The measured slopes of
887: the lines are respectively 0.5507, 0.7391 and 0.8672. The corresponding 
888: theoretical predictions are, to the same accuracy, 0.5000, 0.6928 and 0.8310. }
889: \end{figure}
890: 
891: \begin{figure}
892: \epsfxsize=1.8in
893: \epsfbox{MFfig7.ps}
894: \caption{The scaled EID $h(x)=tn(k,t)$ is plotted against 
895: the dimensionless scaled separation $x=k/\sqrt{t}$ for $q=2$ Potts model. The excellent scaling collapse
896: validates the scaling form given in Eq. \ref{eq:EQ23}. The straight line has slope 
897: $\tau=2(1-\theta(2))=5/4$, which is the MMFA-IIA prediction.The time $t$ is
898: measured in MC steps and distance $k$ is measured in units of lattice spacing.}
899: \end{figure}
900: 
901: In Fig. 7-9, we check the dynamic scaling form Eq.\ref{eq:EQ51} for $n(k,t)$ against the scaling
902: variable $x=k/t^{\phi}$ for three values of $q$--2,5 and 10. We find that for
903: $q=2$, excellent scaling collapse is obtained with $\phi=\frac{1}{2}$ (Fig. 7). For
904: small $x$, we find power-law decay of the scaling function, which crosses
905: over to fast exponential decay at large $x$. For higher values of $q$ 
906: (where $\theta(q) > \frac{1}{2}$), we find that with
907: $\phi=\theta$, we find very good scaling collapse for $x\gg 1$. But for $x\ll 1$,
908: we find systematic deviation from scaling collapse, which was also observed earlier 
909: by Bray and O'Donoghue\cite{STEVE1}. 
910: This observation supports the theoretical prediction based on MMFA, and shows that in this regime, the scaling 
911: function has explicit time dependence. To show this more clearly, and to verify
912: the predicted time-dependence, we plotted the quantity $n(k,t)/P(t)$ against $k$
913: for three widely spaced values of $t$ for all three $q$ values studied in Fig.9.
914: We see that in all three cases, a simple power-law decay with $k$ is observed for $k\ll t^{\phi}$, 
915: thus validating Eq.\ref{eq:EQ52}. The measurement of the exponent $\tau$ gives
916: values in reasonable agreement with theoretical prediction, although for $q=10$, the
917: statistical error is significant.   
918: 
919: 
920: 
921: \begin{figure}
922: \epsfxsize=1.8in
923: \epsfbox{MFfig8.ps}
924: \caption{The scaled EID $\Phi(t,x)=tn(k,t)$ is plotted against 
925: the scaled separation $x=kP(t)$ on a logarithmic scale for $q=5$ Potts model. We see that the 
926: scaling function is explicitly time-dependent for small $k$, but is time-independent for large $k$. 
927: The straight line in the figure gives the theoretical prediction $\tau=2\theta(5)\simeq 1.38$
928: for power-law decay at small $x$ (see discussion in text). The time $t$ is
929: measured in MC steps and distance $k$ is measured in units of lattice spacing.}
930: \end{figure}
931: 
932: \begin{figure}
933: \epsfxsize=1.8in
934: \epsfbox{MFfig9.ps}
935: \caption{Same as Fig. 6, for $q=10$. The straight line in the figure gives the 
936: theoretical prediction $\tau=2\theta(10)\simeq 1.67$
937: for power-law decay at small $x$ (see discussion in text). The time $t$ is
938: measured in MC steps and distance $k$ is measured in units of lattice spacing.}
939: \end{figure}
940: 
941: \begin{figure}
942: \epsfxsize=1.8in
943: \epsfbox{MFfig10.ps}
944: \caption{In the figure, $n(k,t)/P(t)$ is plotted against $k$ for two widely separated values of
945: $t$ for each value of $q=2,5$ and 10 
946:  (top to bottom). In all the cases, the function is independent of $t$ for $k\ll t^{\phi}$,
947: and shows the power law decay $\sim k^{-\tau}$. We measure $\tau\simeq 1.32\pm 0.03$, $1.41\pm 0.04$ and
948: $1.61\pm 0.22$ for $q=2$, $q=5$ and $q=10$ respectively. 
949: The corresponding MMFA-IIA predictions are, to the same accuracy, 1.25, 1.38 and
950: 1.66.}
951: \end{figure}
952: 
953: 
954: 
955: \section{Conclusions}
956: In this paper, we have studied the spatial aspects of persistence in one dimensional $q$-state
957: Potts model using a mean field approximation. We have computed the pair correlation function 
958: for persistent spins under this
959: approximation, and used it to compute the empty interval distribution under the independent interval
960: approximation. We find dynamical scaling behaviour in 
961: both these quantities. The time dependence of the characteristic length scale and the behaviour of the
962: scaling function was found in both cases. We showed analytically within the mean field
963: approximation the transition from fractal to homogeneous distribution of persistent spins 
964: as the persistence exponent crosses $\frac{1}{2}$. We support our results by numerical simulations
965: in the kinetic $q$-state Potts model.
966: 
967: \section{Acknowledgments}
968: This research was supported in part by a grant (DMR 0088451) from the U.S National Science Foundation. The author
969: would like to thank P. Ray for a critical reading of the manuscript and suggestions for improvement.
970: 
971: 
972: \begin{references}
973: \bibitem{SATYA} For a review, S. N. Majumdar, Curr. Sci. (India) {\bf 77}, 370 (1999)
974: (e-print cond-mat/9907407).
975: \bibitem{DERRIDA} B. Derrida, V. Hakim and V. Pasquier, Phys. Rev. Lett. {\bf 75}, 751 (1995);
976: J. Stat. Phys. {\bf 85}, 763 (1996).
977: \bibitem{ZANETTE} D. H. Zanette, Phys. Rev. E. {\bf 55}, 2462 (1997).
978: \bibitem{MANOJ0} G. Manoj and P. Ray, e-print cond-mat/9901130.
979: \bibitem{MANOJ1} G. Manoj and P. Ray, J. Phys. A {\bf 33,} L109 (2000).
980: \bibitem{MANOJ2} G. Manoj and P. Ray, J. Phys. A {\bf 33}, 5489 (2000).
981: \bibitem{STEVE1} A. J. Bray and S. J. O'Donoghue, Phys. Rev. E {\bf 62}, 3366 (2000).
982: \bibitem{JAIN} S. Jain and H. Flynn, J. Phys. A {\bf 33}, 8383 (2000).
983: \bibitem{PURU} G. I. Menon, P. Ray and P. Shukla, Phys. Rev. E {\bf 64}, 046102 (2001).
984: \bibitem{MANOJ3} G. Manoj and P. Ray, Phys. Rev. E {\bf 62}, 7755 (2000).
985: \bibitem{CARDY} J. Cardy, J. Phys. A {\bf 28}, L19 (1995).
986: \bibitem{STEVE2} S. J. O'Donoghue and A. J. Bray, Phys. Rev. E {\bf 64}, 041105 (2001).
987: \bibitem{DERRIDA2} B. Derrida and R. Zeitak, Phys. Rev. E {\bf 54}, 2513 (1996).
988: \bibitem{FELLER1} W. Feller, {\it Introduction to Probability Theory and Applications}
989: (Wiley, New York, 1966) Vol. 1.
990: \bibitem{SMOL} M. von Smoluchowski, Z. Phys. Chem. {\bf 92}, 129 (1917).
991: \bibitem{SMOL2D} D. C. Torney and H. M. McConnell, Proc. R. Soc. (London) Ser. A {\bf 387}, 147 (1983).
992: \bibitem{BPLEE} B. P. Lee, J. Phys. A {\bf 27}, 2633 (1994).
993: \bibitem{LAPLACE} M. G. Smith, {\it Laplace Transform Theory} 
994: (Van Nostrand and Company, London, 1966).
995: \end{references}
996: \end{multicols}
997: 
998: \end{document}
999: