1: \oddsidemargin=0cm
2: \evensidemargin=0cm
3: %\flushbottom
4: %\documentstyle[preprint,aps,psfig]{revtex}
5: \documentstyle[prl,aps,epsf,multicol]{revtex}
6: \newcommand{\be}{\begin{equation}}
7: \newcommand{\ee}{\end{equation}}
8: \newcommand{\bea}{\begin{eqnarray}}
9: \newcommand{\eea}{\end{eqnarray}}
10: \begin{document}
11:
12:
13: \title{Exact Tagged Particle Correlations in the Random Average Process}
14: \author{R. Rajesh$^1$ and Satya N. Majumdar$^{1,2}$}
15: \address{ {\small 1. Department of Theoretical Physics, Tata Institute of
16: Fundamental Research, Homi Bhabha Road, Mumbai 400005, India.}\\
17: {\small 2. Laboratoire de Physique Quantique (UMR C5626 du CNRS),
18: Universit\'e Paul Sabatier, 31062 Toulouse Cedex, France.}}
19: \date{\today}
20: \maketitle
21: \widetext
22: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
23:
24: \begin{abstract}
25:
26: We study analytically the correlations between the positions of tagged
27: particles in the random average process, an interacting particle system in
28: one dimension. We show that in the steady state the mean squared
29: auto-fluctuation of a tracer particle grows subdiffusively
30: $\sigma_0^2(t)\sim t^{1/2}$ for large time $t$ in the absence of external
31: bias but grows diffusively $\sigma_0^2(t)\sim t$ in the presence of a
32: nonzero bias. The prefactors of the subdiffusive and diffusive growths as
33: well as the universal scaling function describing the crossover between
34: them are computed exactly. We also compute $\sigma_r^2(t)$, the mean
35: squared fluctuation in the position difference of two tagged particles
36: separated by a fixed tag shift $r$ in the steady state and show that the
37: external bias has a dramatic effect in the time dependence of
38: $\sigma_r^2(t)$. For fixed $r$, $\sigma_r^2(t)$ increases monotonically
39: with $t$ in absence of bias but has a non-monotonic dependence on $t$ in
40: presence of bias. Similarities and differences with the simple exclusion
41: process are also discussed.
42:
43:
44: \vskip 5mm \noindent PACS numbers: 64.60.-i, 05.70.Ln \end{abstract}
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: \begin{multicols}{2}
47:
48: \section{Introduction}
49:
50: Interacting particle systems in one dimension are amongst the simplest
51: examples of many body systems that are far from equilibrium\cite{liggett}.
52: One of the most studied examples is the simple exclusion process in one
53: dimension. In this system, each site of a one dimensional lattice is either
54: occupied by a hardcore particle or it is empty. In a small time interval $dt$,
55: each particle attempts to hop to the neighboring lattice site on the
56: right with probability $pdt$, to the left neighboring site with
57: probability $qdt$ and stays at the original site with probability
58: $1-(p+q)dt$. An attempted hop is completed provided the target site is
59: empty. A wealth of results are known for this system\cite{liggett,spohn,DE}.
60:
61: Another interacting particle system in one dimension that has attracted
62: recent interest is the random average process (RAP)\cite{KG,RM1}. In the
63: RAP, particles are located on a real line as opposed to a lattice in the
64: simple exclusion process. Let $x_i(t)$ be the position of the $i$-th
65: particle at time $t$ (see Fig. 1).
66: \begin{figure}
67: \narrowtext\centerline{\epsfxsize\columnwidth \epsfbox{fig1.eps}}
68: \caption{The stochastic moves in the RAP.}
69: \end{figure}
70: \noindent
71: In a small time interval $dt$, each particle jumps to the right with
72: probability $pdt$ by an amount $r_i^{+}(x_{i+1}-x_i)$, to the left with
73: probability $qdt$ by an amount $r_i^{-}(x_i-x_{i-1})$ and stays at its
74: original location with probability $1-(p+q)dt$. Here $r_i^{+}$ and
75: $r_i^{-}$ are independent random variables drawn from the interval $[0,1]$
76: with identical probability density function (pdf) $f(r)$. Thus the jumps
77: in either direction is a random fraction of the gap to the nearest
78: particle in that direction. For convenience, we have defined the RAP with
79: random sequential dynamics, though it has been studied with parallel
80: dynamics as well\cite{KG,RM1}. The detailed study of the RAP is important
81: since it has shown up either directly or in disguise in a variety of
82: problems including traffic models\cite{KG}, models of mass
83: transport\cite{RM1}, models of force fluctuation in bead packs\cite{SC},
84: models of voting systems\cite{melzak,FF}, models of wealth
85: distribution\cite{IKR} and the generalized Hammersley process\cite{AD}.
86: Like the simple exclusion process, some aspects of the RAP are
87: analytically tractable\cite{KG,RM1,ZS}. In this paper, we derive some new
88: exact results on the tracer fluctuations in the RAP where the dynamics of
89: tagged particles are followed.
90:
91: The tracer diffusion has been studied in detail for the simple exclusion
92: process and many interesting results are known\cite{liggett}. In the
93: exclusion process, the combined effect of hardcore interaction and the
94: external bias $(p-q)$ shows up rather dramatically in the asymptotic long
95: time behavior of the mean squared auto-fluctuation in the position of a
96: tracer particle in the steady state. If $\zeta_i(t)=x_i(t)-\langle
97: x_i(t)\rangle$ denotes the deviation in the position $x_i(t)$ of the
98: $i$-th particle from its average value, then the mean squared
99: auto-fluctuation is defined as, $\sigma_0^2(t_0,t_0+t)=\langle
100: [\zeta_i(t_0+t)-\zeta_i(t_0)]^2\rangle$, where $t_0$ is the waiting time
101: after which one starts measuring the fluctuations. In the steady state
102: $t_0\to \infty$, the asymptotic behavior of $\sigma_0^2(t)=\lim_{t_0\to
103: \infty}\sigma_0^2(t_0,t_0+t)$ for large $t$ is known\cite{liggett}. In
104: absence of external bias ($p=q=1/2$), i.e. for the symmetric exclusion
105: process (SEP), $\sigma_0^2(t)\sim A t^{1/2}$ for large $t$ where the
106: constant $A=(2/\pi)^{1/2}(1-\rho)/\rho$ is known exactly in terms of the
107: density $\rho$ of the particles\cite{harris,arratia,AP}. This slow
108: subdiffusive growth is due to the caging effect arising from hard core
109: exclusion in one dimension where a particle is always hemmed in by its
110: neighbors. However in the asymmetric case (ASEP) when a nonzero bias
111: $p-q>0$ is switched on, one finds, somewhat unexpectedly,
112: $\sigma_0^2(t)\sim D t$ for large $t$ where the tracer diffusion
113: coefficient $D=(p-q)(1-\rho)$\cite{DF,KvB}. The crossover from the
114: subdiffusive to diffusive behavior of $\sigma_0^2(t)$, as an infinitesimal
115: bias is switched on, was understood in a physically transparent way via a
116: rather unusual mapping of the exclusion process to a $(1+1)$-dimensional
117: interface model\cite{MB1,MB2}. This mapping also established that an
118: appropriately defined sliding tagged-particle correlation function varies
119: anomalously as $t^{2/3}$\cite{MB1}. This anomalous $t^{2/3}$ growth
120: also shows up in the mean square fluctuation of the center of mass of the
121: particles when viewed from a special moving frame\cite{vBKS}.
122:
123: A question then arises naturally: what are the corresponding results on
124: the tracer diffusion for the RAP? The only known result is for the fully
125: asymmetric RAP with $q=0$ (and time rescaled by $p$) where the particles
126: move only to the right. In this limit, $\sigma_0^2(t)$ was
127: computed by Krug and Garcia using a phenomenological hydrodynamic Langevin
128: equation based on heuristic arguments as well as using an independent jump
129: approximation\cite{KG}. Their result shows that $\sigma_0^2(t)\sim
130: D_{1}t$ for large $t$ with $D_{1}={\rho}^{-2}\mu_1\mu_2/(\mu_1-\mu_2)$
131: where $\rho$ is the density of the particles and $\mu_k=\int_0^{1}dr r^k
132: f(r)$ is the $k$-th moment of the pdf $f(r)$. Later, Sch\"utz attempted to
133: derive this result rigorously\cite{schutz} by writing down the exact
134: equation of evolution of the equal time correlation function $G_r(t)=\langle
135: \zeta_0(t)\zeta_r(t)\rangle$ and then using a chain of arguments. Note
136: that the definition $\sigma_0^2(t_0,t_0+t)= \langle
137: [\zeta_i(t_0+t)-\zeta_i(t_0)]^2\rangle$ involves both the variance
138: $\langle \zeta_i^2(t)\rangle$ which is an equal time observable as well as
139: the unequal time correlation $\langle \zeta_i(t_0)\zeta_i(t_0+t)\rangle$.
140: Thus a proper approach, as followed in this paper, would be to compute
141: these correlation functions exactly and then take the steady state $t_0\to
142: \infty$ limit.
143:
144: The main results of this paper can be summarized as follows:
145: \begin{enumerate}
146: \item We compute exactly the mean squared auto-fluctuation in the
147: displacement of a single tracer particle, $\sigma_0^2(t_0,t_0+t)=\langle
148: [\zeta_i(t_0+t)-\zeta_i(t_0)]^2\rangle$ for large $t_0$ and $t$ for all
149: values of $p$ and $q$ in the RAP. In the steady state $t_0\to \infty$, we
150: show that $\sigma_0^2(t)=\lim_{t_0\to \infty}\sigma_0^2(t_0,t_0+t)\sim
151: A_{SRAP} t^{1/2}$ for large $t$ for the symmetric RAP (SRAP) with $p=q$.
152: For the asymmetric RAP (ARAP) where $p>q$, we find $\sigma_0^2(t)\sim
153: D_{ARAP}t$ for large $t$. The constants
154: $A_{SRAP}=2{\rho}^{-2}(p\mu_1/\pi)^{1/2}\mu_2/(\mu_1-\mu_2)$ and
155: $D_{ARAP}={\rho}^{-2}(p-q){\mu_1\mu_2}/{(\mu_1-\mu_2)}$ are
156: computed exactly. For the special case $q=0$ and $p=1$, $D_{ARAP}$ reduces
157: to $D_1$ computed earlier in Refs. \cite{KG,schutz}.
158: \item We compute exactly the universal scaling function that
159: describes the crossover behavior of $\sigma_0^2(t)$ from the subdiffusive
160: $t^{1/2}$ growth to the diffusive $t$ growth as one switches on an
161: infinitesimal bias $(p-q)$.
162: \item We generalize the single tracer particle fluctuation
163: $\sigma_0^2(t_0,t_0+t)$ to the fluctuation in the position difference of
164: two tagged particles defined as $\sigma_r^2(t_0,t_0+t)=\langle
165: [\zeta_{i+r}(t_0+t)-\zeta_i(t_0)]^2\rangle$. We show that in the steady
166: state $\sigma_r^2(t)=\lim_{t_0\to \infty}\sigma_r^2(t_0,t_0+t)$ grows
167: monotonically with $t$ for a fixed tag shift $r$ for the SRAP. For the
168: ARAP on the other hand, it grows with $t$ in a non-monotonic fashion with a
169: single minimum at a characteristic time $t^{*}=r/{\mu_1(p-q)}$.
170: \item We also compute various scaling functions that describe the
171: crossover of the tracer fluctuations from their non-steady state behavior
172: to the steady state behavior as the waiting time $t_0\to \infty$.
173: \end{enumerate}
174:
175: The paper is organized as follows. In Sec. II, we define the model
176: precisely and set up our notations. In Sec. III, we calculate the equal
177: time correlation function for the RAP for all $p$ and $q$. Sec. IV
178: contains the exact calculation of the unequal time correlation function.
179: In Sec. V we compute the mean squared fluctuation in the displacement
180: of a single tracer particle. The Secs. VA and VB contain
181: respectively the discussions on the SRAP and the ARAP, while the crossover
182: between them is discussed in Sec. VC. The Sec. VI contains the
183: generalization to the two-tag correlation functions. Finally we conclude
184: with a summary and discussion in Sec. VII.
185:
186: \section{The Model and Preliminaries}
187:
188: We consider a system of particles of average density $\rho$ located on a
189: real line. Let $x_i(t)$ denote the position of the $i$-th particle at time
190: $t$ (see Fig. 1). In an infinitesimal time interval $dt$, each particle
191: jumps with probability $pdt$ to the right, with probability $qdt$ to the
192: left and with probability $1-(p+q)dt$ it rests at its original location.
193: The actual amount by which a particle jumps (either to the right or to the
194: left) is a random fraction of the gap between the particle and its
195: neighboring particle (to the right or to the left). For example, the jump
196: to the right is by an amount $r_i^{+}(x_{i+1}-x_i)$ and to the left by
197: $r_i^{-}(x_i-x_{i-1})$. The random variables $r_i^{\pm}$ are independently
198: drawn from the interval $[0,1]$ and each is distributed according to the
199: same pdf $f(r)$ which is arbitrary. We start from an arbitrary but fixed
200: initial condition at $t=0$ and averaging of physical quantities is done
201: over all histories of evolution keeping the initial condition fixed. The
202: time evolution of the positions $x_i(t)$'s can be represented by the exact
203: Langevin equation
204: \be
205: x_i (t+dt)=x_i (t) + \gamma_i(t),
206: \label{one}
207: \ee
208: where $\gamma_i (t)$ are random variables given by
209: \be
210: \gamma_i(t)=\cases
211: {r_i^{+} (x_{i+ 1}(t)-x_i(t)) \quad\mbox{with prob $pdt$},\cr
212: r_i^{-} (x_{i-1}(t)-x_i(t)) \quad\mbox{with prob $qdt$},\cr
213: 0 \mbox{\hskip16.2mm with prob $1-(p+q)dt$}.\cr}
214: \label{two}
215: \ee
216: The random variables $r_i^{\pm}$ are independent and each is distributed
217: over the interval $[0,1]$ with the same pdf $f(r)$. The $k$-th moment of
218: the pdf is denoted by $\mu_k=\int_0^1 dr r^k f(r)$. Note that since $0\leq
219: r\leq 1$ and $f(r)\geq 0$, $\mu_1\geq \mu_2$.
220:
221: We define a new random variable $\zeta_i(t)$ which measures the deviation
222: of $x_i(t)$ from its mean value
223: \be
224: \zeta_i (t)=x_i(t)-\langle x_i(t) \rangle.
225: \label{three}
226: \ee
227: From Eqs. (\ref{one}) and (\ref{two}), one can easily derive the
228: evolution rules for the $\zeta_i$ variables. We find
229: \be
230: \zeta_i (t+dt)=\zeta_i (t) -(p-q) \frac{\mu_1}{\rho} dt +\eta_i(t),
231: \label{four}
232: \ee
233: where $\eta_i(t)$ is given by
234: \be
235: \eta_i (t)=\cases
236: {r_i^{+} (\zeta_{i+1}(t)-\zeta_i(t)+ {\rho}^{-1}) \quad \mbox{with prob
237: $pdt$},\cr
238: r_i^{-} (\zeta_{i-1}(t)-\zeta_i(t)-{\rho}^{-1}) \quad \mbox{with prob
239: $qdt$},\cr
240: 0 \quad \mbox{\hskip21.2mm with prob $1-(p+q)dt$}.\cr}
241: \label{five}
242: \ee
243: By definition, $\langle \zeta_i(t)\rangle =0$. Also from Eq.
244: (\ref{five}), it follows that $\langle \eta_i(t)\rangle= (p-q)
245: \frac{\mu_1}{\rho} dt$.
246:
247: In this paper, we will focus on the mean squared displacement of a tagged
248: particle. It turns out that the asymptotic behavior of the mean squared
249: displacement depends crucially on whether one starts measuring these
250: fluctuations after some finite waiting time $t_0$ or if one first waits
251: for an infinite time and then starts measuring the statistics. The latter
252: corresponds to measuring the fluctuations in the steady state. This is
253: similar to the `approach to stationary' versus `stationary' regimes found
254: in various interface models\cite{pers}. This can be quantified precisely
255: in terms of the following correlation function,
256: \bea
257: \sigma_0^2(t_0,t_0+t) &=&\langle \left( \zeta_i(t+t_0)-\zeta_i(t_0)
258: \right)^2 \rangle,
259: \label{six} \\
260: &=& G_0(t\!+\!t_0)+ G_0(t_0) - 2 C_0(t_0, t_0\!+\!t)
261: \label{seven},
262: \eea
263: where $G_r(t)=\langle \zeta_i(t)\zeta_{i+r}(t)\rangle$ is the equal time
264: correlation function and $C_r(t_0,t_0+t)=\langle
265: \zeta_i(t_0)\zeta_{i+r}(t_0+t)$ with $t>0$ denotes the unequal time
266: correlation function. For $t=0$, the unequal time correlation function
267: reduces to the equal time correlation function, $C_r(t_0,t_0)=G_r(t_0)$.
268: Note that we have assumed an infinite system so that the translational
269: invariance holds. In the next two sections we calculate analytically the
270: correlation functions $G_r(t)$ and $C_r(t_0,t_0+t)$ respectively.
271:
272: \section{Equal Time Correlation Function}
273:
274: In this section, we calculate the equal time correlation function
275: $G_r(t)=\langle \zeta_i(t)\zeta_{i+r}(t)\rangle$ exactly for the RAP for
276: all $p$ and $q$. Our starting point is Eq. (\ref{four}) in conjunction
277: with Eq. (\ref{five}) describing the evolution of the $\zeta_i$ variables
278: with time. We consider the evolution equations (Eq. (\ref{four})) for both
279: $\zeta_i(t+dt)$ and $\zeta_{i+r}(t+dt)$, multiply them and then take the
280: average $\langle \rangle$ over all histories, keeping terms only upto $O(dt)$.
281: This yields, in the limit $dt\to 0$, the exact evolution equation of the
282: correlation function $G_r(t)$ and we obtain,
283: \bea
284: \frac{d}{dt} G_r(t)&=&\mu_1(p+q) \left[ G_{r+1}(t) +G_{r-1}(t)-2 G_r(t)\right]
285: \nonumber \\
286: &+&\delta_{r, 0}\, \mu_2 (p+q) \left[ \rho^{-2} +2
287: \left(G_0(t)-G_1(t)\right)\right].
288: \label{eight}
289: \eea
290: The Eq. (\ref{eight}) is valid for all positive and negative integers $r$
291: including $r=0$ and clearly $G_r(t)=G_{-r}(t)$. Thus the equation of
292: evolution for the two point correlations involve only two point correlations
293: and not higher order correlations. This closure property is crucial for
294: obtaining an exact solution for the correlation functions. The key reason
295: behind this closure lies in the fact that the random fractions
296: $r_i^{\pm}$'s at time $t$ are independent of the $\zeta_i(t)$. One
297: noteworthy fact about Eq. (\ref{eight}) is that the rates $p$ and $q$
298: make their appearance only as an overall multiplicative factor $(p+q)$.
299: We could absorb this factor into the time by doing a suitable rescaling, and
300: hence, the equal time correlation function
301: $G_r(t)$ is same for both the ARAP and the SRAP.
302:
303: We note that this equation was also derived in Ref.\cite{schutz} by a
304: rather lengthy method, but was left unsolved. In this section, we derive
305: an exact solution of Eq. (\ref{eight}). Note that even though Eq.
306: (\ref{eight}) represents the diffusion equation (in discrete space) with a
307: source term at the origin $r=0$, its solution is nontrivial due to the
308: fact that the source term depends on $G_0(t)$ and $G_1(t)$ which need to
309: be determined self-consistently. Similar diffusion equations with source
310: term for the correlation functions have also appeared recently in the
311: context of aggregation models with injection\cite{RM2}. Before proceeding to
312: solve Eq. (\ref{eight}), we first set up our notations. We define the
313: standard Fourier transform
314: \be
315: {\bar G}(k,t)=\sum_{r=-\infty}^{\infty} G_r(t) e^{ikr},
316: \label{fourier}
317: \ee
318: the Laplace transform
319: \be
320: {\tilde G}_r(s)=\int_0^{\infty} G_r(t) e^{-st} dt,
321: \label{laplace}
322: \ee
323: and the joint Fourier-Laplace transform,
324: \be
325: F(k,s)=\int_0^{\infty}{\bar G}(k,t)e^{-st}dt=\sum_{r=-\infty}^{\infty}
326: {\tilde G}_r(s) e^{ikr}.
327: \label{jointfl}
328: \ee
329:
330: Taking the joint Fourier-Laplace transform of Eq. (\ref{eight}) we obtain
331: \be
332: F(k,s)=\frac{\mu_2 (p+q) \left[ {\rho}^{-2}+2s\left({\tilde G}_0(s)-{\tilde
333: G}_1(s)\right)\right]}{s\left[s+2\mu_1 (p+q)(1-\cos k)\right]},
334: \label{fl1}
335: \ee
336: where we have assumed that initially $G_r(0)=0$ which is true for any
337: fixed initial condition. For random initial condition, $F(k,s)$ will
338: contain additional terms arising from the initial condition, but one can
339: show that they do not contribute to the asymptotic large time properties
340: of $G_r(t)$ as long as the initial condition has only short ranged
341: correlations. We therefore use $G_r(0)=0$ without any loss of generality.
342:
343: The Eq. (\ref{fl1}) contains two unknowns ${\tilde G}_0(s)$ and ${\tilde
344: G}_1(s)$. One of them, say ${\tilde G}_1(s)$ can however be expressed in
345: terms of ${\tilde G}_0(s)$ by taking directly the Laplace transform of Eq.
346: (\ref{eight}) for $r=0$ and using $G_1(t)=G_{-1}(t)$. This gives the
347: relation
348: \be
349: s{\tilde G}_0(s)=(p\!+\!q)\!\left[\frac{\mu_2 \rho^{-2}}{s}-2(\mu_1\!-\!\mu_2)
350: \left({\tilde G}_0(s)- {\tilde G}_1(s)\right)\right].
351: \label{g0g1}
352: \ee
353: Substituting Eq. (\ref{g0g1}) in Eq. (\ref{fl1}) we obtain
354: \be
355: F(k,s)=\frac{\mu_2}{(\mu_1-\mu_2)}{ {\left[
356: \mu_1 (p+q){\rho}^{-2}-s^2{\tilde G}_0(s)\right]}\over {s\left[ s+2\mu_1
357: (p+q)(1-\cos k)\right]}}.
358: \label{fl2}
359: \ee
360:
361: We now have to determine ${\tilde G}_0(s)$ self-consistently. This can be
362: done by using the inverse Fourier transform
363: \be
364: {\tilde G}_r(s)= {1\over {2\pi}}\int_{-\pi}^{\pi} F(k,s)e^{-ikr}dk.
365: \label{invg0}
366: \ee
367: Substituting the expression of $F(k,s)$ from Eq. (\ref{fl2}) in Eq.
368: (\ref{invg0}) at $r=0$, we obtain the exact ${\tilde G}_0(s)$
369: \be
370: {\tilde G}_0(s)={ {\mu_1\mu_2 (p+q)}\over {(\mu_1-\mu_2)}}{
371: {{\rho}^{-2}I(0,s)}\over {s\left[1+{ {\mu_2}\over
372: {(\mu_1-\mu_2)}}sI(0,s)\right]}},
373: \label{g0s}
374: \ee
375: where $I(r,s)$ is given by the integral
376: \bea
377: I(r,s)&=&{1\over {2\pi}}\int_{-\pi}^{\pi} {{e^{-ikr}dk}\over {
378: \left[s+2\mu_1^\prime (1-\cos k)\right]} } \nonumber \\
379: &=&\frac{1}{\sqrt{s^2+4\mu_1^\prime s}}
380: \!\left(\!\frac{2 \mu_1^\prime +s-\sqrt{s^2+4\mu_1^\prime s}} {2
381: \mu_1^\prime }\right)^{|r|}\!\!,
382: \label{irs}
383: \eea
384: where $\mu_1^\prime=\mu_1 (p+q)$.
385: Knowing ${\tilde G}_0(s)$ determines $F(k,s)$ completely by Eq.
386: (\ref{fl2}) and hence ${\tilde G}_r(s)$ for all $r$ by the Fourier
387: inversion formula in Eq. (\ref{invg0}). We obtain
388: \be
389: {\tilde G}_r(s)={ {\mu_1\mu_2 (p+q)}\over {(\mu_1-\mu_2)} } {
390: {{\rho}^{-2}I(r,s)}\over { s\left[1+ { {\mu_2}\over {(\mu_1-\mu_2)} }
391: sI(0,s)\right]} },
392: \label{grs}
393: \ee
394: where $I(r,s)$ is given by Eq. (\ref{irs}).
395:
396:
397: To obtain $G_r(t)$ we need to perform the inverse Laplace transform
398: $G_r(t)={\cal L}^{-1}[{\tilde G}_r(s)]$ with respect to $s$. In general
399: for arbitrary $t$ this is difficult. However, for large $t$, this inverse
400: can be obtained in closed form. For large $t$, one needs to consider the
401: small $s$ behavior of ${\tilde G}_r(s)$ in Eq. (\ref{grs}). Let us first
402: consider the case $r=0$. Putting $r=0$ in Eq. (\ref{irs}) and taking the
403: $s\to 0$ limit we find to leading order,
404: \be
405: I(0,s)\sim {1\over {2\sqrt{\mu_1 (p+q) s}} }.
406: \label{i0s}
407: \ee
408: Substituting this small $s$ expression of $I(0,s)$ in Eq. (\ref{g0s}) and
409: taking the inverse Laplace transform we find that to leading order for
410: large $t$,
411: \be
412: G_0(t)=\frac{\sqrt{\mu_1(p+q)} \mu_2 \rho^{-2}}{(\mu_1-\mu_2) \sqrt{\pi}}
413: \sqrt{t}.
414: \label{g0t1}
415: \ee
416:
417: Next we consider the behavior of $G_r(t)$ for $|r|>0$. From Eq. (\ref{irs})
418: it is clear that the appropriate scaling limit consists of taking
419: the limit $s\to 0$, $|r|\to \infty$ but
420: keeping $|r|\sqrt{s}$ fixed. In this scaling limit,
421: Eq. (\ref{irs}) yields,
422: \be
423: I(r,s)= {1\over {2 \sqrt{\mu_1(p+q) s}}}\exp\left(\frac{-|r|\sqrt{s}}
424: {\sqrt{\mu_1 (p+q)}} \right).
425: \label{irss}
426: \ee
427: We note that the formula for $I(r,s)$ in Eq. (\ref{irss}) reduces to Eq.
428: (\ref{i0s}) for $|r|=0$. This indicates that even though Eq. (\ref{irss})
429: was derived in the scaling limit, it continues to hold even for $r=0$.
430:
431: Substituting this small $s$ expression of $I(r,s)$ in Eq. (\ref{grs}) and
432: taking the inverse Laplace transform we obtain for large $t$,
433: \be
434: G_r(t)= \frac{\sqrt{\mu_1 (p+q)} \mu_2 \rho^{-2}}{2(\mu_1-\mu_2)}{\cal
435: L}^{-1}\left[s^{-3/2}e^{-|r| \sqrt{s/[\mu_1(p+q)]}}\right].
436: \label{grt1}
437: \ee
438: Fortunately the inverse Laplace transform in Eq. (\ref{grt1}) can be done
439: in closed form which gives us the following asymptotic scaling behavior of
440: the equal time correlation function $G_r(t)$,
441: \be
442: G_r(t)=\frac{\sqrt{\mu_1 (p+q)} \mu_2 \rho^{-2}}{(\mu_1-\mu_2) \sqrt{\pi}}
443: \sqrt{t} f_1\left(\frac{|r|}{2\sqrt{\mu_1 (p+q) t}}\right).
444: \label{grt2}
445: \ee
446: Here $f_1(y)$ is a universal scaling function independent of the model
447: parameters such as $p$, $q$ and the moments $\mu_k$ of the pdf $f(r)$ and
448: is given by
449: \be
450: f_1(y)= e^{-y^2}-{\sqrt{\pi}}~y \mbox{~erfc}(y),
451: \label{f1}
452: \ee
453: where ${\rm erfc}(y)=2/\sqrt{\pi}\int_y^{\infty}e^{-u^2}du$ is the
454: standard complimentary error function. This scaling function has the
455: asymptotic behaviors, $f_1(y)\sim 1- \sqrt{\pi}y$ as $y\to 0$ and $\sim
456: y^{-2}e^{-y^2}/2$ for $y\to \infty$.
457:
458: As a final remark, we note again that if one puts $|r|=0$ in the formula
459: for $G_r(t)$ in Eq. (\ref{grt2}) one recovers the correct $G_0(t)$ as
460: given by Eq. (\ref{g0t1}). Thus the scaling range includes even the $r=0$
461: point. The Eq. (\ref{g0t1}) thus provides us the exact behavior of the
462: first two terms in the expression for $\sigma_0^2(t_0,t_0+t)$ in Eq.
463: (\ref{seven}). The remaining task is to evaluate the third term in Eq.
464: (\ref{seven}) which involves the unequal time correlation function and
465: this is done in the next section.
466:
467: \section{Unequal time correlations}
468:
469: In this section we compute the two time tag-tag correlation function
470: $C_r(t_0,t_0+t)=\langle \zeta_i(t_0)\zeta_{i+r}(t_0+t)\rangle$ for the
471: RAP. We start at time $t_0$ and then evolve the $\zeta_{i+r}$ variables by
472: Eq. (\ref{four}) for all subsequent time. Let us first rewrite the Eq.
473: (\ref{four}) at time $t_0+t+dt$,
474: \be
475: \zeta_{i+r} (t_0\!+\!t\!+\!dt)=\zeta_{i+r} (t_0\!+\!t) -
476: (p\!-\!q) \frac{\mu_1}{\rho} dt +\!\eta_{i+r}(t_0\!+\!t).
477: \label{four1}
478: \ee
479: We then multiply both sides of Eq. (\ref{four1}) by $\zeta_{i}(t_0)$
480: and average over the noise keeping terms only upto $O(dt)$. In the limit
481: $dt\to 0$, we obtain the exact evolution equation of the two time correlation
482: function,
483: \bea
484: \frac{dC_r(t_0,t_0\!+\!t)}{dt}&=&\mu_1\left[ p
485: C_{r+1}(t_0,t_0\!+\!t)+qC_{r-1}(t_0,t_0\!+\!t)\right. \nonumber \\
486: &-& \left. (p+q)C_r(t_0,t_0+t)\right] \mbox{\hskip3mm for~} t \geq 0 .
487: \label{crteqn}
488: \eea
489: Note that at $t=0$, the unequal time correlation function reduces to the
490: equal time correlation function $C_r(t_0,t_0)=G_r(t_0)$. Thus starting at
491: $t=0$ with the initial condition $C_r(t_0,t_0)=G_r(t_0)$, the function
492: $C_r(t_0,t_0+t)$ evolves with time $t$ according to the Eq.
493: (\ref{crteqn}).
494:
495: As in the preceding section, we define the Fourier transform ${\bar
496: C}(k,t_0,t_0+t)=\sum_{r=-\infty}^{\infty}C_r(t_0,t_0+t)e^{ikr}$. Taking
497: the Fourier transform of Eq. (\ref{crteqn}) we obtain
498: \be
499: {\bar C}(k,t_0,t_0+t)={\bar G}(k,t_0)e^{-\mu_1 \alpha(k) t},
500: \label{ckt}
501: \ee
502: where $\alpha(k)=p+q-(pe^{-ik}+qe^{ik})$ and ${\bar G}(k,t_0)$ is the
503: Fourier transform of the equal time correlation function as defined by Eq.
504: (\ref{fourier}). Taking further the Laplace transform
505: $H(k,s,t)=\int_0^{\infty}{\bar C}(k,t_0,t_0+t)e^{-st_0}dt_0$ of Eq.
506: (\ref{ckt}) we obtain
507: \be
508: H(k,s,t)=F(k,s)e^{-\mu_1 \alpha(k) t},
509: \label{hkst}
510: \ee
511: where $F(k,s)$ is given exactly by Eq. (\ref{fl2}) with ${\tilde G}_0(s)$
512: determined from Eq. (\ref{g0s}).
513:
514: Proceeding as in the previous section, the Laplace transform ${\tilde
515: C}_r(s,t)= \int_0^{\infty}C_r(t_0, t_0+t)e^{-st_0}dt_0$ can then be
516: determined from the joint Fourier-Laplace transform $H(k,s,t)$ by the
517: inversion formula
518: \be
519: {\tilde C}_r(s,t)= {1\over {2\pi}}\int_{-\pi}^{\pi}H(k,s,t)e^{-ikr}dk,
520: \label{crst1}
521: \ee
522: where $H(k,s,t)$ is given by Eq. (\ref{hkst}). Substituting in Eq.
523: (\ref{crst1}) the exact expression of $F(k,s)$ from Eq. (\ref{fl2}) and
524: that of ${\tilde G}_0(s)$ from Eq. (\ref{g0s}), we obtain the following
525: final expression of the Laplace transform
526: \bea
527: {\tilde C}_r(s,t)&=&{ {\mu_1\mu_2(p+q)}\over {(\mu_1-\mu_2)}}{
528: {{\rho}^{-2}}\over {s\left[1+{ {\mu_2}\over
529: {(\mu_1-\mu_2)}}sI(0,s)\right]}}\nonumber \\
530: &\mbox{x}& \frac{1}{2\pi} \int_{-\pi}^{\pi} {{
531: e^{-ikr-\mu_1\alpha(k)t}dk}\over { \left[s+2\mu_1 (p+q)(1-\cos k)\right]} }.
532: \label{crst2}
533: \eea
534: Note that for $t=0$, ${\tilde C}_r(s,t)$ as given by Eq. (\ref{crst2})
535: reduces to ${\tilde G}_r(s)$ given by Eq. (\ref{grs}) as expected. The
536: equation (\ref{crst2}) is central to our subsequent analysis for various
537: limiting behaviors.
538:
539:
540: \section{Mean Squared Tracer Auto-fluctuation}
541:
542: In this section we calculate $\sigma_0^2(t_0,t_0+t)$ in the RAP using the
543: exact results for the equal time and two time correlation functions
544: obtained in the previous sections. We consider first the symmetric case
545: SRAP with $p=q$ in subsection A followed by the derivation for the
546: asymmetric case ARAP with $p>q$ in subsection B. In subsection C, we show
547: how the steady state fluctuation $\sigma_0^2(t)= lim_{t_0\to
548: \infty}\sigma_0^2(t_0,t_0+t)$ crosses over from the subdiffusive behavior
549: to the diffusive behavior as one switches on an infinitesimal bias and we
550: calculate the crossover scaling function exactly.
551:
552: \subsection{SRAP}
553:
554: Here we consider the symmetric case $p=q$. For the calculation of
555: $\sigma_0^2(t_0,t_0+t)$ we only need the asymptotic behavior of
556: $C_r(t_0,t_0+t)$ for $r=0$ as evident from Eq. (\ref{seven}). To obtain
557: $C_0(t_0,t_0+t)$ we need to invert the Laplace transform in Eq.
558: (\ref{crst2}) for $r=0$ and $p=q$. As before, this inversion is difficult
559: in general for all $t_0$. However the finite but large $t_0$ limit can be
560: worked out by analyzing the small $s$ behavior of Eq. (\ref{crst2}). It
561: turns out that the appropriate scaling limit in this case involves taking
562: $s\to 0$, $t\to \infty$ but keeping $st$ fixed. In this scaling limit, the
563: integration in Eq. (\ref{crst2}) can be carried out in closed form and we
564: obtain (with $p=q$),
565: \be
566: {\tilde C}_0(s,t)=\frac{\sqrt{2 p \mu_1} \mu_2
567: \rho^{-2}}{2(\mu_1-\mu_2)s^{3/2}}e^{st/2}{\rm erfc}\left(\sqrt{st/2}\right).
568: \label{c0sts}
569: \ee
570: We then need to invert the Laplace transform in Eq. (\ref{c0sts}) with
571: respect to $s$ to obtain the asymptotic behavior of $C_0(t_0,t_0+t)$ for
572: large $t_0$. Fortunately this inversion can be done in closed form and we
573: obtain
574: \be
575: C_0(t_0,t_0+t)=\frac{\sqrt{2 p \mu_1} \mu_2
576: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}\sqrt{t_0}f_2\left({ {t}\over
577: {2 t_0}}\right),
578: \label{c0t1}
579: \ee
580: where the scaling function $f_2(y)$ is again universal and is given by,
581: \be
582: f_2(y)=\sqrt{1+y}-\sqrt{y} .
583: \label{sc2}
584: \ee
585:
586: We are now ready to compute $\sigma_0^2(t_0,t_0+t)$ from Eq.
587: (\ref{seven}). Using the result for the equal time correlation in Eq.
588: (\ref{g0t1}) and the one for the two time correlation in Eq. (\ref{c0t1}),
589: we obtain from Eq. (\ref{seven}) our main result
590: \bea
591: \lefteqn{ \sigma_0^2(t_0,t_0+t)=}\quad \nonumber \\
592: & &\frac{\sqrt{2 p \mu_1} \mu_2 \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}
593: \left[\sqrt{t_0+t}+\sqrt{t_0}-2\sqrt{t_0}f_2\left({ {t}\over
594: {2 t_0}}\right)\right],
595: \label{smsf1}
596: \eea
597: where $f_2(y)$ is given by Eq. (\ref{sc2}). Note that this result in Eq.
598: (\ref{smsf1}) is derived in the scaling limit when both $t_0$ and $t$ are
599: large with their ratio $t/t_0$ kept fixed.
600:
601: We now discuss two different limits of Eq. (\ref{smsf1}). First we
602: consider the steady state limit $t_0\to \infty$ with $t$ large but fixed.
603: In this limit, Eq. (\ref{smsf1}) yields
604: \be
605: \sigma_0^2(t)={\rm lim}_{t_0\to
606: \infty}\sigma_0^2(t_0,t_0+t)=\frac{2\sqrt{p\mu_1 } \mu_2
607: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}\sqrt{t}.
608: \label{smsf2}
609: \ee
610: In the opposite limit, when the waiting time $t_0$ is finite (away from
611: the steady state) but the evolved time $t$ goes to infinity, we obtain from
612: Eq. (\ref{smsf1},
613: \be
614: {\rm lim}_{t\to \infty}\sigma_0^2(t_0,t_0+t)=\frac{\sqrt{ 2 p \mu_1 } \mu_2
615: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}\sqrt{t}.
616: \label{smsf3}
617: \ee
618: Thus the mean squared auto-fluctuation in these two opposing limits differ
619: by a factor $\sqrt{2}$. The Eqs. (\ref{smsf1}), (\ref{smsf2}) and
620: (\ref{smsf3}) are amongst the important new results of this paper.
621:
622:
623: \subsection{ARAP}
624:
625: In this subsection we calculate $\sigma_0^2(t_0,t_0+t)$ in the asymmetric
626: case when $p>q$. Once again we have to invert the Laplace transform in Eq.
627: (\ref{crst2}) for $r=0$ but now with $p>q$. In this case it turns out the
628: appropriate scaling limit consists of taking $s\to 0$, $t\to \infty$ as in
629: the SRAP but keeping $\sqrt{s} t$ instead of the scaling variable $st$ in
630: the SRAP. In this scaling limit, the integration in Eq. (\ref{crst2}) with
631: $r=0$ yields
632: \be
633: {\tilde C}_0(s,t)=\frac{\sqrt{\mu_1 (p+q)} \mu_2
634: \rho^{-2}}{2(\mu_1-\mu_2)s^{3/2}}e^{-(p-q)\sqrt{\mu_1 s/(p+q)}\,t}.
635: \label{c0sta}
636: \ee
637: The Laplace transform in Eq. (\ref{c0sta}) can be inverted as in Eq.
638: (\ref{grt1}) and we obtain
639: \be
640: C_0(t_0,t_0+t)= \frac{\sqrt{\mu_1(p+q)} \mu_2
641: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}\sqrt{t_0}f_1\! \left[ {
642: {\sqrt{\mu_1}(p-q)t}\over {2\sqrt{(p+q) t_0}}}\right]\!,
643: \label{c0t2}
644: \ee
645: where the universal scaling function $f_1(y)=e^{-y^2}-\sqrt{\pi}y~ {\rm
646: erfc}(y)$ is the same as in Eq. (\ref{f1}).
647:
648: Substituting the results in Eq. (\ref{c0t2}) and Eq. (\ref{g0t1}) in Eq.
649: (\ref{seven}) we obtain
650: \bea
651: \lefteqn{\sigma_0^2(t_0,t_0+t)= \frac{\sqrt{\mu_1 (p+q)} \mu_2
652: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}}\qquad\nonumber \\
653: &&\mbox{x}\left[\sqrt{t_0+t}+\sqrt{t_0}-2\sqrt{t_0}f_1 \left( {
654: {\sqrt{\mu_1}(p-q)t}\over {2\sqrt{(p+q)t_0}}}\right)\right].
655: \label{amsf1}
656: \eea
657:
658: As in the SRAP we now discuss the two different limits. In the steady
659: state $t_0\to \infty$ with fixed large $t$ we obtain from Eq. (\ref{amsf1}),
660: \be
661: \sigma_0^2(t)={\rm lim}_{t_0\to \infty}\sigma_0^2(t_0,t_0+t)=\frac{ \mu_1
662: \mu_2 \rho^{-2}(p-q)}{(\mu_1-\mu_2)}t.
663: \label{amsf2}
664: \ee
665: Thus in this case $\sigma_0^2(t)$ grows diffusively for large $t$,
666: $\sigma_0^2(t)=D_{ARAP}t$ where the diffusion constant,
667: \be
668: D_{ARAP}={\rho}^{-2}(p-q){{\mu_1\mu_2}\over {(\mu_1-\mu_2)}},
669: \label{dc}
670: \ee
671: depends explicitly on $p$ and $q$. For $q=0$ and $p=1$, it reduces to
672: the expression $D_1={\rho}^{-2}\mu_1\mu_2/(\mu_1-\mu_2)$ derived by Krug
673: and Garcia using the independent jump approximation\cite{KG} and later
674: rederived by Sch\"utz\cite{schutz} using a different approach.
675:
676: We make a brief comment here on the approach used in Ref.\cite{schutz} in
677: deriving the diffusion constant $D_1$. In his approach, Sch\"utz started
678: with the evolution equation (\ref{eight}) for the equal time correlation
679: function and then used a chain of arguments to derive the diffusion
680: constant $D_1$. His approach didn't require any knowledge of the two time
681: correlation function or even the solution of the equal time correlation
682: function. As evident from the definition in Eq. (\ref{seven}) that
683: $\sigma_0^2(t_0,t_0+t)$ requires the knowledge of both the equal and the
684: two time correlation functions. Thus it was rather remarkable that the
685: correct value of the diffusion constant for $q=0$ and $p=1$ was recovered
686: in Ref. \cite{schutz}. However this turns out to be purely fortuitous.
687: Note that the evolution equation (\ref{eight}) is independent of the bias in
688: the system.
689: Thus the approach of Sch\"utz would predict that the diffusion
690: constant is also completely independent of the bias $(p-q)$ and is always given
691: by $D_1$ (provided $t$ is scaled by $(p+q)$).
692: This is clearly wrong as evident from the exact expression in
693: Eq. (\ref{dc}). In particular for the symmetric case $p=q=1/2$, the arguments
694: of Ref. \cite{schutz} would predict a diffusive growth of $\sigma_0^2(t)$
695: with the diffusion constant $D_1$. This is again incorrect since for $p=q$
696: the diffusion constant is $0$ from Eq. (\ref{dc}) which is consistent with
697: the correct asymptotic subdiffusive growth of $\sigma_0^2(t)$ as given
698: exactly by Eq. (\ref{smsf2}). The problem in the derivation of Sch\"utz
699: can be traced back to the fact that his arguments only used equal time
700: correlations (which involve only $(p+q)$) and not the two time
701: correlations. The dependence on the bias $(p-q)$ of the diffusion constant
702: $D_{ARAP}$ comes only from the two time correlations. The derivation of
703: Ref.\cite{schutz} misses this important fact and is rather fortuitous to
704: obtain the correct value $D_1$ of the diffusion constant for the special
705: case when $p=1$ and $q=0$.
706:
707: We end this subsection by discussing the other limit when the system is
708: away from the steady state, i.e. when $t_0$ is large but finite and $t\to
709: \infty$. In this limit, we obtain from Eq. (\ref{amsf1})
710: \be
711: {\rm lim}_{t\to \infty}\sigma_0^2(t_0,t_0+t)=\frac{\sqrt{\mu_1 (p+q) } \mu_2
712: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}\sqrt{t},
713: \label{amsf3}
714: \ee
715: the same result as in the SRAP in this limit [Eq. (\ref{smsf3})]. Thus
716: away from the steady state the tracer particle doesn't sense the presence
717: of bias. The exact result in Eq. (\ref{amsf3}) is consistent with that of
718: Krug and Garcia using a phenomenological hydrodynamic equation\cite{KG}.
719:
720: \subsection{Crossover Between SRAP and ARAP}
721:
722: In the previous subsections, we have seen that the asymptotic large $t$
723: behavior of $\sigma_0^2(t_0,t_0+t)$ does not depend on the bias $(p-q)$, when
724: the system is away from the steady state (finite $t_0$). However, in the
725: steady state ($t_0\to \infty$) it behaves rather differently in the
726: symmetric and asymmetric cases. In the steady state of the SRAP ($p=q$),
727: $\sigma_0^2(t)\sim t^{1/2}$ while for the ARAP ($p>q$), $\sigma_0^2(t)\sim
728: t$. Thus a natural question is: How does the behavior of $\sigma_0^2(t)$
729: crosses over from the subdiffusive growth for $p=q$ to the diffusive
730: growth as one switches on an infinitesimal bias $(p-q)$? In this
731: subsection we compute exactly the universal scaling function that
732: describes this crossover behavior of $\sigma_0^2(t)$.
733:
734: To calculate the crossover behavior we return to our central equation
735: (\ref{crst2}) with $r=0$. We have seen in the previous subsections that in
736: the scaling limit $s\to 0$ and $t\to \infty$ of the Eq. (\ref{crst2}), the
737: appropriate scaling variable that is kept fixed is $st$ for $p=q$, where
738: as, it is $\sqrt{s}t$ for $p>q$. Thus, to compute the crossover behavior,
739: we need to keep the leading order terms in both of these scaling variables
740: fixed while expanding Eq. (\ref{crst2}) for small $s$ and large $t$. This
741: makes the calculation of the crossover behavior somewhat delicate. To
742: leading order, we find after elementary algebra
743: \bea
744: \lefteqn{ {\tilde C}_0(s,t)=
745: \frac{\sqrt{\mu_1(p+q)} \mu_2 \rho^{-2}}{(\mu_1-\mu_2)s^{3/2}}}\qquad
746: \nonumber\\
747: &&\mbox{x}{1\over
748: {2\pi}}\int_{-\infty}^{\infty} { {
749: e^{-i(p-q)\sqrt{\mu_1s/(p+q)}\,tz-stz^2/2}}\over {1+z^2}}dz.
750: \label{crst3}
751: \eea
752: Note that for the symmetric case $p=q$, the integral in Eq. (\ref{crst3})
753: can be done and we get back Eq. (\ref{c0sts}) of Sec. VA. Similarly,
754: for the asymmetric case $p>q$, in the limit $s\to 0$ keeping the scaling
755: variable $\sqrt{s}t$ fixed, one drops the second term in the exponential
756: in the integrand of Eq. (\ref{crst3}) and performing the resulting
757: integral we recover the Eq. (\ref{c0sta}) of Sec. VB.
758:
759: To compute the crossover behavior we need to keep both the terms inside
760: the exponential in the integrand of Eq. (\ref{crst3}) and perform the
761: integral. Fortunately this integral can be done in closed form using the
762: standard convolution theorem. We omit the details here and present only
763: the final result,
764: \bea
765: \lefteqn{ {\tilde C}_0(s,t)= \frac{\sqrt{\mu_1 (p+q)} \mu_2
766: \rho^{-2}}{4(\mu_1-\mu_2)s^{3/2}}}\qquad\nonumber\\
767: &&\mbox{x}\left[e^{u-v}{\rm erfc}\left( {{2u-v}\over {2\sqrt{u}} }\right)
768: +e^{u+v}{\rm erfc}\left( {{2u+v}\over {2\sqrt{u}} }\right)\right],
769: \label{cr1}
770: \eea
771: where $u=st/2$ and $v=(p-q)\sqrt{\mu_1 s/(p+q)}\,t$. We then expand the Eq.
772: (\ref{cr1}) further for small $s$ to obtain the steady state $t_0\to
773: \infty$ behavior. Note that we needed to first do the integral in Eq.
774: (\ref{crst3}) and then take the $s\to 0$ limit. The reverse order
775: unfortunately does not work. Expanding Eq. (\ref{cr1}) for small $s$,
776: keeping only the leading order terms in $s$ and finally inverting the
777: Laplace transform of the resulting expression we obtain for large $t_0$
778: \bea
779: \lefteqn{C_0(t_0,t_0+t)=\frac{\sqrt{\mu_1(p+q)} \mu_2
780: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}} \quad\nonumber \\
781: &&\mbox{x}\left[\sqrt{t_0}-\sqrt{ {{t}\over {2}}}e^{-w^2(t)}
782: -{{\sqrt{\pi \mu_1}(p-q)t}\over {2\sqrt{p+q}}}{\rm
783: erf}\left(w(t)\right) \right]\!,
784: \label{cr2}
785: \eea
786: where $w(t)=(p-q)\sqrt{\mu_1 t/ \left[2(p+q)\right]}$.
787:
788: We now use the results from Eq. (\ref{cr2}) and Eq. (\ref{g0t1}) in Eq.
789: (\ref{seven}) and eventually take the strict $t_0\to \infty$ limit to
790: obtain the final form of the steady state auto-fluctuation
791: \bea
792: \sigma_0^2(t)&=&{\rm lim}_{t_0\to \infty}\sigma_0^2(t_0,t_0+t)
793: \nonumber\\
794: &=&\frac{ \mu_1 \mu_2 (p-q)\rho^{-2}}{(\mu_1-\mu_2)}t Y\left[ {
795: {(p-q)\sqrt{\mu_1 t}}\over {\sqrt{2(p+q)}}}\right],
796: \label{cr3}
797: \eea
798: where $Y(y)$ is a universal crossover scaling function given by
799: \be
800: Y(y)={\rm erf}(y)+ {1\over {\sqrt {\pi}}}{e^{-y^2}\over {y}}.
801: \label{crsc}
802: \ee
803: The scaling function has the asymptotic behavior $Y(y)\sim
804: 1/(\sqrt{\pi}y)$ as $y\to 0$ and $Y(y)\to 1$ and $y\to \infty$. Note that
805: for fixed $p-q>0$, if we take the limit $t\to \infty$ in Eq. (\ref{cr3})
806: (which corresponds to $y\to\infty$ in the scaling function in Eq.
807: (\ref{crsc})) we recover the result of Eq. (\ref{amsf2}). Similarly if we
808: take the $p-q\to 0$ limit for fixed $t$ in Eq. (\ref{cr3}) (corresponding
809: to taking $y\to 0$ limit in the scaling function $Y(y)$), we recover, as
810: expected, the result of Eq. (\ref{smsf2}) of the symmetric case. Thus the
811: Eq. (\ref{cr3}) and the associated scaling function $Y(y)$ in Eq.
812: (\ref{crsc}) describes the crossover behavior from the subdiffusive to
813: diffusive growth as one switches on an infinitesimal bias.
814:
815: \section{Generalization to the two-tag correlation Function}
816:
817: So far in this paper we have concentrated only on the mean squared
818: auto-fluctuation of a tracer particle, $\sigma_0^2(t_0,t_0+t)= \langle
819: \left( \zeta_i(t+t_0)-\zeta_i(t_0) \right)^2 \rangle$. A natural
820: generalization of the auto-fluctuation would be to study the two-tag
821: correlation function defined as
822: \bea
823: \sigma_r^2(t_0,t_0+t) &=&\langle \left( \zeta_{i+r}(t+t_0)-\zeta_{i}(t_0)
824: \right)^2 \rangle,
825: \label{tt1} \\
826: &=& G_0(t\!+\!t_0)+ G_0(t_0) - 2 C_r(t_0, t_0\!+\!t)
827: \label{tt2},
828: \eea
829: where $G_r(t)$ and $C_r(t_0,t_0+t)$ are the usual equal time time and the
830: two time correlation functions already defined and derived in the previous
831: sections. Note that for $r=0$, the two-tag correlation in Eq. (\ref{tt1})
832: reduces to the single tag function $\sigma_0^2(t_0,t_0+t)$.
833:
834: Of particular interest would be to compute the two-tag correlation
835: function in the steady state, i.e. $\sigma_r^2(t)=\lim_{t_0\to
836: \infty}\sigma_r^2(t_0,t_0+t)$. For the exclusion process this two-tag
837: correlation function was first introduced in Ref. \cite{MB2} and the
838: presence of bias was found to have a dramatic effect on the time
839: dependence of $\sigma_r^2(t)$ for a fixed $r$. It was found numerically
840: that while in the SEP $\sigma_r^2(t)$ increases monotonically with $t$ for
841: a fixed tag-shift $r$, in the ASEP $\sigma_r^2(t)$ has a non-monotonic
842: dependence on $t$\cite{MB2}. In the ASEP $\sigma_r^2(t)$ first decreases
843: with time $t$, becomes a minimum at some characteristic time $t^{*}$ and
844: then starts increasing again. A harmonic model was proposed in Ref.
845: \cite{MB2} for which $\sigma_r^2(t)$ could be computed analytically and
846: was found to be in qualitative agreement with the numerical results of the
847: exclusion process. But to the best of our knowledge, exact calculation of
848: $\sigma_r^2(t)$ for the exclusion process is still an unsolved problem.
849: However it turns out that for the RAP, it is possible to compute this
850: function $\sigma_r^2(t)$ exactly for large $t$. The exact solution of
851: $\sigma_r^2(t)$ in the RAP, as shown below, shares the similar features as
852: in the exclusion process.
853:
854: From Eq. (\ref{tt2}) it is evident that we just need to compute the large
855: $t_0$ behavior of the two time correlation function $C_r(t_0,t_0+t)$ for
856: fixed nonzero $r$. In the previous sections we have analyzed in detail the
857: $r=0$ case. It turns out that the analysis for $r\neq 0$ proceeds more or
858: less in the same manner as in the $r=0$ case. We start, once again, from
859: the central equation (\ref{crst2}). To avoid separate calculations for the
860: SRAP and the ARAP, we take the line of approach used to calculate the
861: crossover behavior in subsection VC. For $r\neq 0$, it turns out that the
862: equation (\ref{crst3}) gets replaced by a similar looking equation,
863: \bea
864: \lefteqn{ {\tilde C}_r(s,t)=
865: \frac{\sqrt{\mu_1(p+q)} \mu_2 \rho^{-2}}{(\mu_1-\mu_2)s^{3/2}}}\qquad \nonumber\\
866: &&\mbox{x}{1\over
867: {2\pi}}\int_{-\infty}^{\infty} { {
868: e^{-iz\sqrt{s/[\mu_1(p+q)]}R-stz^2/2}}\over {1+z^2}}dz,
869: \label{tt3}
870: \eea
871: where $R=r+\mu_1(p-q)t$ signifies the drift of the particles to the right
872: with average velocity $\mu_1(p-q)$ for $p>q$. Clearly for $r=0$, Eq.
873: (\ref{tt3}) reduces to Eq. (\ref{crst3}). Starting with Eq. (\ref{tt3}) we
874: then follow exactly the same steps as used in subsection VC. Since the
875: steps are identical we skip all the details and present only the final
876: result. In the strict steady state limit $t_0\to \infty$, we finally obtain
877: the following scaling form
878: \bea
879: \sigma_r^2(t)&=&{\rm lim}_{t_0\to \infty}\sigma_r^2(t_0,t_0+t)\nonumber
880: \\
881: &=& \frac{\sqrt{2\mu_1(p\!+\!q)}\mu_2
882: \rho^{-2}}{(\mu_1-\mu_2)\sqrt{\pi}}\sqrt{t}W\!\left( {R\over
883: {\sqrt{2\mu_1(p+q)t}}}\right)\!,
884: \label{tt4}
885: \eea
886: where $R=r+\mu_1(p-q)t$ and $W(y)$ is again a universal scaling function
887: given by,
888: \be
889: W(y)=e^{-y^2}+\sqrt{\pi}y\,{\rm erf}(y).
890: \label{tt5}
891: \ee
892: Clearly $\mu_1(p-q)t$ represents the average drift while
893: $l(t)=\sqrt{2\mu_1(p+q)t}$ represents the diffusive length scale.
894:
895: We note that the scaling function $W(y)$ is a symmetric function of $y$
896: about $y=0$ with a minimum at $y=0$. For the SRAP, $p=q$ and hence $R=r$.
897: Thus for a fixed $r$, it follows from Eq. (\ref{tt4}) that $\sigma_r^2(t)$
898: increases monotonically with $t$. For the ARAP on the other hand, $p>q$
899: and $R=r+\mu_1(p-q)t$. If one fixes $r$ to a negative value and increases
900: $t$, the variable $R$ remains negative till the characteristic time
901: $t=t^*=r/\mu_1(p-q)$, beyond which it becomes positive. The scaling
902: variable $y=R/\sqrt{2\mu_1(p+q)t}$ behaves in the same way. Thus
903: $\sigma_r^2(t)$ in Eq. (\ref{tt4}) first decreases with time, becomes a
904: minimum at $t=t^*=-r/\mu_1(p-q)$ and then starts increasing again. In Fig.
905: (2) we plot the function $\sigma_r^2(t)$ in Eq. (\ref{tt4}) for both the
906: SRAP (with $p=q=1/2$) and the ARAP (with $p=1$ and $q=0$) for the same
907: value of $r=-2$ and choosing the parameter values $\mu_1=1/2$, $\mu_2=1/4$,
908: $\rho=1$. These features in the RAP, derived here exactly, are
909: qualitatively similar to those in the exclusion process studied in Ref.
910: \cite{MB2}.
911: \begin{figure}
912: \narrowtext\centerline{\epsfxsize\columnwidth \epsfbox{fig2.eps}}
913: \caption{The steady state two-tag correlation function $\sigma_r^2(t)$ in
914: Eq. (\ref{tt4}) plotted as a function of $t$ for fixed $r=-2$ for parameter
915: values $\mu_1=1/2$, $\mu_2=1/4$ and $\rho=1$. The solid line shows the
916: monotonic growth of $\sigma_r^2(t)$ with $t$ for the SRAP ($p=q=1/2$)
917: while the dashed line shows the non-monotonic growth for the ARAP ($p=1$
918: and $q=0$).}
919: \end{figure}
920:
921: \section{Conclusions}
922: In this paper we have studied analytically the mean squared fluctuations
923: in the diffusion of both a single tagged particle and two tagged particles
924: in the random average process (RAP) for all values of the hopping rates
925: $p$ and $q$ in one dimension. We have shown that in the steady state, the
926: auto-fluctuation of a tagged particle grows subdiffusively as
927: $\sigma_0^2(t)\sim A_{SRAP}t^{1/2}$ for $p=q$ and diffusively
928: $\sigma_0^2(t)\sim D_{ARAP}t$ for $p>q$ where
929: $A_{SRAP}=2{\rho}^{-2}(p\mu_1/\pi)^{1/2}\mu_2/(\mu_1-\mu_2)$ and
930: $D_{ARAP}={\rho}^{-2}(p-q){\mu_1\mu_2}/{(\mu_1-\mu_2)}$. These behaviors
931: of $\sigma_0^2(t)$ are similar to those in the simple exclusion process
932: except the prefactors
933: $A=(2/\pi)^{1/2}(1-\rho)/\rho$\cite{harris,arratia,AP} and
934: $D=(p-q)(1-\rho)$\cite{DF,KvB} are different in the exclusion process.
935: Besides the steady state mean squared two-tag fluctuation $\sigma_r^2(t)$
936: in the RAP grows monotonically with $t$ for $p=q$ and non-monotonically
937: for $p>q$, in much the same way as in the exclusion process.
938:
939: These findings raise the question whether or not the RAP is in the same
940: universality class as the simple exclusion process in one dimension.
941: Perhaps the RAP is just a coarse grained version of the exclusion process
942: in one dimension? The answer to this question seems to be in the negative
943: due to a very crucial difference between the two processes. In the
944: exclusion process for $p>q$, it is well known that there exists an
945: anomalous $t^{2/3}$ growth hidden in the problem apart from the usual
946: $t^{1/2}$ and $t$ growth\cite{vBKS,MB1}. This anomalous growth shows up
947: either in the mean squared fluctuation of the center of mass of the
948: particles when viewed from a special moving frame\cite{vBKS} or
949: alternately in the two-tag correlation function $\sigma_r^2(t)$ if one
950: chooses the tag shift $r$ to be sliding with time with a special velocity
951: $r=-\rho^2(p-q)$\cite{MB1}. It turns out that the prefactor of this
952: $t^{2/3}$ growth is proportional to $\propto {{d^2j(\rho)}\over
953: {d\rho^2}}$ where $j(\rho)$ is the current density in a hydrodynamical
954: description\cite{MB1}. For the exclusion process, $j(\rho)=\rho(1-\rho)$
955: and hence the prefactor is nonzero. For the RAP on the other hand,
956: $j(\rho)=\mu_1(p-q)$ and is independent of $\rho$. This is because
957: $j(\rho)=\rho \langle v\rangle$ where the average velocity $\langle
958: v\rangle=\mu_1(p-q)/\rho$ as can be easily derived from Eqs. (\ref{one})
959: and (\ref{two}). As a result, for the RAP, the anomalous $t^{2/3}$ growth
960: is absent which puts it in a different universality class than the simple
961: exclusion process. In this sense the RAP seems to be closer to the
962: harmonic model studied in Ref. \cite{MB2}.
963:
964: In this paper we have considered the RAP only in one dimension. An obvious
965: generalization would be to higher dimensions. A natural way to generalize
966: the model to higher dimensions would be as follows. One considers
967: particles located in the continuous $d$-dimensional space. In a small time
968: interval $dt$, each particle makes a list of all its nearest neighbors in
969: various directions in space, chooses one of them at random and jumps in
970: the corresponding direction by a random fraction of the Euclidean distance
971: to that neighbor. This is an isotropic version, a generalization of the
972: SRAP. Similarly one can define an anisotropic version as well. To the best
973: of our knowledge, the RAP has not been studied so far in higher
974: dimensions. The question of tracer diffusion in higher dimensions,
975: especially in two dimensions where one may expect a logarithmic
976: correction, also remains completely open.
977:
978:
979:
980:
981:
982: \begin{thebibliography}{999}
983:
984: \bibitem{liggett} T. M. Liggett, {\it Interacting Particle Systems}
985: (Springer-Verlag, New York, 1985).
986:
987: \bibitem{spohn} H. Spohn, {\it Large Scale Dynamics of Interacting
988: Particles} (Springer-Verlag, Berlin Heidelberg, 1991).
989:
990: \bibitem{DE} B. Derrida and M.R. Evans, p-277 in {\it Nonequilibrium
991: Statistical Mechanics in one dimension}, edited by V. Privman (Cambridge
992: University Press, Cambridge, 1997).
993:
994: \bibitem{KG} J. Krug and J. Garcia, J. Stat. Phys. {\bf 99}, 31 (2000).
995:
996: \bibitem{RM1} R. Rajesh and S. N. Majumdar, J. Stat. Phys. {\bf 99}, 943
997: (2000).
998:
999: \bibitem{SC} S. N. Coppersmith, C.-h. Liu, S. N. Majumdar, O. Narayan and
1000: T. A. Witten, Phys. Rev. E {\bf 53}, 4673 (1996).
1001:
1002: \bibitem{melzak} Z. A. Melzak, {\it Mathematical Ideas, Modeling and
1003: Applications, Vol II of Companion to Concrete Mathematics} (Wiley, New
1004: York, 1976), p.271.
1005:
1006: \bibitem{FF} P. A. Ferrari and L. R. G. Fontes, El. J. Prob. {\bf 3},
1007: Paper no. 6 (1998).
1008:
1009: \bibitem{IKR} S. Ispolatov, P. L. Krapivsky and S. Redner, Eur. Phys. J.
1010: {\bf B2}, 267 (1998).
1011:
1012: \bibitem{AD} D. Aldous and P. Diaconis, Probablity Theory and Related
1013: Fields, {\bf 103}, 199 (1995).
1014:
1015: \bibitem{ZS} F. Zielen and A. Schadschneider, cond-mat/0104298.
1016:
1017: \bibitem{harris} T. Harris, J. Appl. Prob., {\bf 2}, 323 (1965)
1018:
1019: \bibitem{arratia} R. Arratia, Ann. Prob., {\bf 11}, 362 (1983).
1020:
1021: \bibitem{AP} S. Alexander and P. Pincus, Phys. Rev. B. {\bf 18}, 2011
1022: (1978).
1023:
1024: \bibitem{DF} A. Demasi and P. A. Ferrari, J. Stat. Phys., {\bf 38}, 603
1025: (1985).
1026:
1027: \bibitem{KvB} R. Kutner and H. van Beijeren, J. Stat. Phys., {\bf 39}, 317
1028: (1985).
1029:
1030: \bibitem{MB1} S. N. Majumdar and M. Barma, Physica A. {\bf 177}, 366
1031: (1991).
1032:
1033: \bibitem{MB2} S. N. Majumdar and M. Barma, Phys. Rev. B. {\bf 44}, 5306
1034: (1991).
1035:
1036: \bibitem{vBKS} H. van Beijeren, R. Kutner, and H. Spohn, Phys. Rev. Lett.
1037: {\bf 54}, 2026 (1985).
1038:
1039: \bibitem{schutz} G. M. Sch\"{u}tz, J. Stat. Phys. {\bf 99}, 1045 (2000).
1040:
1041: \bibitem{pers} J. Krug, H. Kallabis, S.N. Majumdar, S.J. Cornell, A.J.
1042: Bray, and C. Sire, Phys. Rev. E {\bf 56}, 2702 (1997); S.N. Majumdar and
1043: A.J. Bray, cond-mat/0009439, to appear in Phys. Rev. Lett.
1044:
1045: \bibitem{RM2} R. Rajesh and S. N. Majumdar, Phys. Rev. E. {\bf 62}, 3186
1046: (2000).
1047:
1048:
1049: \end{thebibliography}
1050:
1051: \end{multicols}
1052:
1053: \end{document}
1054: