arxiv:cond-mat/ v1 [cond-mat.stat-mech] 19 Jan 2006

Similar documents
16. Working with the Langevin and Fokker-Planck equations

Weak Ergodicity Breaking WCHAOS 2011

Weak Ergodicity Breaking

This is a Gaussian probability centered around m = 0 (the most probable and mean position is the origin) and the mean square displacement m 2 = n,or

Maximal distance travelled by N vicious walkers till their survival arxiv: v1 [cond-mat.stat-mech] 17 Feb 2014

Anomalous Lévy diffusion: From the flight of an albatross to optical lattices. Eric Lutz Abteilung für Quantenphysik, Universität Ulm

Kolmogorov Equations and Markov Processes

CHAPTER V. Brownian motion. V.1 Langevin dynamics

Theory of fractional Lévy diffusion of cold atoms in optical lattices

Uncertainty quantification and systemic risk

Reflected Brownian Motion

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 23 Feb 1998

Stationary distributions of non Gaussian Ornstein Uhlenbeck processes for beam halos

Poisson Jumps in Credit Risk Modeling: a Partial Integro-differential Equation Formulation

Smoluchowski Diffusion Equation

State Space Representation of Gaussian Processes

Separation of Variables in Linear PDE: One-Dimensional Problems

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 1 Aug 2001

Local vs. Nonlocal Diffusions A Tale of Two Laplacians

arxiv:cond-mat/ v2 [cond-mat.stat-mech] 2 Dec 2004

ODE Homework Solutions of Linear Homogeneous Equations; the Wronskian

PROBABILITY: LIMIT THEOREMS II, SPRING HOMEWORK PROBLEMS

Perturbed Gaussian Copula

Nonlinear Autonomous Systems of Differential

Weak Ergodicity Breaking. Manchester 2016

Math 308 Exam I Practice Problems

New Physical Principle for Monte-Carlo simulations

Fokker-Planck Equation with Detailed Balance

The Kramers problem and first passage times.

Lecture 12: Detailed balance and Eigenfunction methods

Brownian motion and the Central Limit Theorem

1 Elementary probability

VIII.B Equilibrium Dynamics of a Field

F n = F n 1 + F n 2. F(z) = n= z z 2. (A.3)

Stochastic Volatility and Correction to the Heat Equation

The dynamics of small particles whose size is roughly 1 µmt or. smaller, in a fluid at room temperature, is extremely erratic, and is

An efficient approach to stochastic optimal control. Bert Kappen SNN Radboud University Nijmegen the Netherlands

8 Example 1: The van der Pol oscillator (Strogatz Chapter 7)

Mathematical Methods for Neurosciences. ENS - Master MVA Paris 6 - Master Maths-Bio ( )

ELEMENTS OF PROBABILITY THEORY

Physics 116C The Distribution of the Sum of Random Variables

Statistical Mechanics of Active Matter

A new approach for investment performance measurement. 3rd WCMF, Santa Barbara November 2009

The method of lines (MOL) for the diffusion equation

Week 9 Generators, duality, change of measure

Lecture 12: Detailed balance and Eigenfunction methods

Diffusion in a Logarithmic Potential Anomalies, Aging & Ergodicity Breaking

9 More on the 1D Heat Equation

I forgot to mention last time: in the Ito formula for two standard processes, putting

Semiclassical analysis and a new result for Poisson - Lévy excursion measures

Gillespie s Algorithm and its Approximations. Des Higham Department of Mathematics and Statistics University of Strathclyde

arxiv: v1 [cond-mat.stat-mech] 31 Mar 2008

A path integral approach to the Langevin equation

MATH3203 Lecture 1 Mathematical Modelling and ODEs

ENGI 9420 Lecture Notes 1 - ODEs Page 1.01

UNDERSTANDING BOLTZMANN S ANALYSIS VIA. Contents SOLVABLE MODELS

Basic Concepts and Tools in Statistical Physics

Existence Theory: Green s Functions

LANGEVIN THEORY OF BROWNIAN MOTION. Contents. 1 Langevin theory. 1 Langevin theory 1. 2 The Ornstein-Uhlenbeck process 8

HW2 Solutions. MATH 20D Fall 2013 Prof: Sun Hui TA: Zezhou Zhang (David) October 14, Checklist: Section 2.6: 1, 3, 6, 8, 10, 15, [20, 22]

Path integrals for classical Markov processes

LANGEVIN EQUATION AND THERMODYNAMICS

Real roots of random polynomials and zero crossing properties of diffusion equation

Dynamics of a tagged monomer: Effects of elastic pinning and harmonic absorption. Shamik Gupta

Let's transfer our results for conditional probability for events into conditional probabilities for random variables.

e (x y)2 /4kt φ(y) dy, for t > 0. (4)

1. Stochastic Process

1 2 predators competing for 1 prey

5 Applying the Fokker-Planck equation

Short-time expansions for close-to-the-money options under a Lévy jump model with stochastic volatility

Handbook of Stochastic Methods

Effective dynamics for the (overdamped) Langevin equation

The Laplace Transform

2012 NCTS Workshop on Dynamical Systems

Math Applied Differential Equations

Final Exam May 4, 2016

1 Review of di erential calculus

Nonconstant Coefficients

Math 308 Exam I Practice Problems

Stochastic process. X, a series of random variables indexed by t

Dynamical Systems. August 13, 2013

6. Brownian Motion. Q(A) = P [ ω : x(, ω) A )

Partial Differential Equations

Examples include: (a) the Lorenz system for climate and weather modeling (b) the Hodgkin-Huxley system for neuron modeling

laplace s method for ordinary differential equations

Introduction. Statistical physics: microscopic foundation of thermodynamics degrees of freedom 2 3 state variables!

Multiple Random Variables

MATH 425, FINAL EXAM SOLUTIONS

BMIR Lecture Series on Probability and Statistics Fall 2015 Discrete RVs

Large deviations of the top eigenvalue of random matrices and applications in statistical physics

Errata for Stochastic Calculus for Finance II Continuous-Time Models September 2006

Math 215/255 Final Exam (Dec 2005)

Stochastic continuity equation and related processes

p 1 ( Y p dp) 1/p ( X p dp) 1 1 p

arxiv: v1 [math.pr] 18 Nov 2018

Stochastic (intermittent) Spikes and Strong Noise Limit of SDEs.

Math 216 Final Exam 24 April, 2017

Example 4.1 Let X be a random variable and f(t) a given function of time. Then. Y (t) = f(t)x. Y (t) = X sin(ωt + δ)

HJB equations. Seminar in Stochastic Modelling in Economics and Finance January 10, 2011

Lecture 4: Introduction to stochastic processes and stochastic calculus

Transcription:

Statistical Properties of Functionals of the Paths of a Particle Diffusing in a One-Dimensional Random Potential arxiv:cond-mat/6455v cond-mat.stat-mech] 9 Jan 6 Sanjib Sabhapandit,, Satya N. Majumdar, and Alain Comtet, Laboratoire de Physique Théorique et Modèles Statistiques, Université Paris-Sud, Bâtiment, 945 Orsay Cedex, France Université Pierre et Marie Curie - Paris 6, Institut Henri Poincaré, rue Pierre et Marie Curie, Paris, F-755, France Dated: June 6, 7 We present a formalism for obtaining the statistical properties of functionals and inverse functionals of the paths of a particle diffusing in a one-dimensional quenched random potential. We demonstrate the implementation of the formalism in two specific examples: where the functional corresponds to the local time spent by the particle around the origin and where the functional corresponds to the occupation time spent by the particle on the positive side of the origin, within an observation time window of size t. We compute the disorder average distributions of the local time, the inverse local time, the occupation time and the inverse occupation time, and show that in many cases disorder modifies the behavior drastically. PACS numbers: 5.4.-a,.5.-r, 46.65.+g I. INTRODUCTION The statistical properties of functionals of a one dimensional Brownian motion have been extensively studied and have found numerous applications in diverse fields ranging from probability theory,, 3], finance 4, 5, 6], mesocopic physics 7], computer science 8], and in understanding weather records 9]. The position xτ of a one dimensional Brownian motion evolves with time τ via the Langevin equation dx dτ = ητ, starting from x = x, where ητ is a thermal Gaussian white noise with mean ητ = and a correlator ητητ = δτ τ. A functional T is simply the integral up to time t T = t V xτ dτ, where V x is a prescribed non-negative function whose choice depends on the specific observable of interest. For a fixed initial position x of the Brownian motion and a fixed observation time t, the value of T varies from one history or realization of the Brownian path {xτ} to another see Fig. and a natural question is: what is the probability density function pdf PT t, x? Following the path integral methods devised by Feynman ], Kac showed, ] that the calculation of the pdf PT t, x can essentially be reduced to a quantum mechanics problem, namely solving a single particle Shrödinger equation in an external potential V x. This formalism is known in the literature as the celebrated Feynman-Kac formula. Subsequently, this method has been widely used to calculate the pdf of T with different choices of V x as demanded by specific applications. This has been reviewed recently in Ref. 8]. In particular, two most popular applications correspond respectively to the choices, V x = δx a and V x = θx, where δx is the Dirac s delta function and θx is the Heaviside step function. In the former T t t t 3 t 4 R R R 3 R 4 FIG. : Schematic plots of T defined by Eq. as a function of t, corresponding to four different realizations of the paths {xτ}, for τ t] denoted by R, R, R 3 and R 4 respectively. For fixed t t or t or t 3 or t 4, shown by vertical dashed lines, T takes different value for different realizations. On the other hand for a fixed T horizontal dashed line corresponding t is different for different realizations: t for R, t for R, t 3 for R 3 and t 4 for R 4. case, the corresponding functional Ta = t δxτ adτ has the following physical meaning: Tada is just the time spent by the particle in the vicinity of the point a in space, i.e., in the region a, a + da], out of the total observation time t. Note that, by definition, Tada = t. The functional Ta is known as the local time density in the literature. In the second case V x = θx, the functional T = t θxτdτ measures the time spent by the particle on the positive side of the origin out of the total time t and is known as the occupation time. The probability distribution of the occupation time was originally computed by Lévy ], t

T T/t PT t, dt = π arcsin, and is known as the arcsine law of Lévy. Since then, the local and the occupation times for pure diffusion, have been studied extensively by mathematicians, 3, 4, 5, 6, 7]. Recently, the study of the occupation time has seen a revival in physics literature and has been used in understanding the dynamics out of equilibrium in coarsening systems 8, 9], ergodicity properties in anomalously diffusive processes, ], in renewal processes ], in models related to spin glasses 3], in simple models of blinking quantum dots 4], and also in the context of persistence 5, 6]. Local and Occupation times have been also studied in the context of stochastic ergodicity breaking 7], first passage time 8], diffusion controlled reactions activated by catalytic sites 9] and diffusion on graphs 3, 3]. In polymer science, a long flexible polymer of length t is often modeled by a Brownian path up to time t. In this context, the local time at a position r is proportional to the concentration of monomers at r in a polymer of length t. A natural and important question is how to generalize the Feynman-Kac formalism to calculate the statistical properties of the functionals of the type in Eq. when xτ is not just a pure diffusion process, but it represents the position of a particle in an external random medium. While various properties of diffusion in random media have been widely studied in the past 3, 33, 34, 35], the study of statistical properties of functionals in random media is yet to receive its much deserved attention. In this paper we undertake this task. More precisely, we are interested in calculating the pdf PT t, x of a functional T as in Eq. where xτ now evolves via the Langevin equation dx dτ = F xτ + ητ 3 where ητ represents the thermal noise as in Eq. and Fx = du/dx represents the external force, the derivative of the potential Ux, felt by the particle. Most generally, the external potential consists of two parts, Ux = U d x + U r x, a deterministic part U d x and a random part U r x. The random part of the potential U r x is quenched in the sense that it does not change during the time evolution of the particle, but fluctuates from one sample to another according to some prescribed probability distribution. Consequently, the pdf PT t, x will also fluctuate from one sample of the random potential to another and the goal is to compute the disorder averaged pdf PT t, x where the... denote the average over the distribution of the random potential. A popular model for the random potential is the celebrated Sinai model 36], where various disorder averaged physical quantities can be computed analytically 3, 33, 37, 38, 39, 4, 4, 4], and yet the results exhibit rich and nontrivial behaviors and also capture many of the qualitative behaviors of more complex realistic disordered systems. The Sinai model assumes that U r x = Bx where Bx represents a Brownian motion in space, i.e., db dx = ξx 4 where ξx is a Gaussian white noise with mean ξx = and a correlator ξxξx = δx x. The constant represents the strength of the random potential. In this paper, we first present a generalization of the Feynman-Kac formalism to calculate the pdf PT t, x in presence of an arbitrary external potential Ux. To obtain explicit results using this formalism, we next assume that the random part of the potential is as in the Sinai model, i.e., our external potential is of the form Ux = U d x + Bx, where Bx is a Brownian motion in space and U d x is the non-random deterministic part of the potential. It turns out that the asymptotic behavior of the disorder averaged pdf PT t, x, quite generically, has three different qualitative behavior depending on the curvature of the deterministic potential U d x, i.e., whether U d x has a convex concave upward shape representing a stable potential i.e., attractive force towards the origin, a concave concave downward shape representing unstable potential a repulsive force away from the origin or just flat indicating the absence of any external potential. To facilitate explicit calculation, we model the deterministic potential simply by, U d x = µ x, so that µ < represents a stable potential, µ > represents an unstable potential and µ = represents a flat potential. This specific choice facilitates explicit calculation, but the results are qualitatively similar if one chooses another form of this potential. Thus, in our model, we will consider the external potential as Ux = µ x + Bx 5 where Bx = x ξx dx is the trajectory of a Brownian motion in space see Fig.. Note that the case µ = corresponds to the pure Sinai model. Figure shows typical realization of potentials for µ =, µ > and µ <. The corresponding force in Eq. 3 is simply given by Fx = µ signx + ξx. 6 We will demonstrate how to calculate explicitly, using our generalized Feynman-Kac formalism, the disorder averaged pdf PT t, x when the external potential is of the form given by Eq. 5. Despite the simplicity of the choice of the external potential, a variety of rich and interesting behavior can be obtained by tuning the parameter µ/, as shown in this paper. We will present detailed results for the two functionals, namely for the local time and the occupation time corresponding to the choices V x = δx and V x = θx respectively in Eq.. Also, to keep the discussion simple, we will present our final results for x = corresponding to the particle starting at the origin. However, our method is not limited only to this specific choice. Some of these results were briefly announced in a previous letter 43]. In addition, in this paper we also introduce the notion of inverse functional, which is defined as follows. If V x in Eq. is non-negative, then for each path {xτ}, T is a non-decreasing function of t, which we formally denote by T = g t {xτ}, x. Therefore for a given realization of path {xτ} and given T there is a unique value of t see

3 Ux µ = Ux µ > x Ux µ < x x FIG. : A classical particle represented by diffusing in a typical realization of the potential Ux = µ x + Bx, where Bx represents the trajectory of a Brownian motion in space with B =. The three figures are for µ =, µ > and µ < respectively. The dash lines show the potential for =. Fig., which we formally write as the inverse of the functional g 49] t = g T {xτ}, x. 7 This inverse time t physically means the observation time that is required for any given path {xτ} in order to produce a prescribed value of T. Of course, for the same value T, for a different path {xτ}, the value of t will be different. Thus, t is a random variable for a fixed T, which takes different values for different realizations of paths and we would like to compute its pdf, which we denote by It T, x and by definition It T, dt =. Clearly, this pdf will also differ from sample to sample of the external potential in Eq. 5 and our goal is to obtain the disorder averaged distribution It T, x. In this paper, we present detailed results for It T, again for the two choices of V x = δx and V x = θx corresponding to the local time and the occupation time respectively. The inverse local and occupation times are useful for experimentalists as they provide an estimate on the required measurement time. For example, in the context of polymers, the inverse local time is the typical length of a polymer required to obtain a certain monomer concentration. The rest of the paper is organized as follows. In Sec. II, we present our general approach for computing the pdf PT t, x of the functional T defined by Eq. for a given t, and the pdf It T, x of the inverse functional defined by Eq. 7 for a given T, for a given sample of the random potential, for arbitrary starting position of the particle x = x and for arbitrary but non-negative V x. After this section we consider the specific examples of local time and occupation time by setting V x = δx and V x = θx respectively. We will use different notations for the pdfs in the two examples to avoid any misunderstanding. In the first example, where T is the local time, we denote the pdf of the local time PT t, for a given t by P loc T t, and the pdf of the inverse local time It T, for a given T by I loc t T. In the second example, where T is the occupation time we denote the pdf of the occupation time PT t, for a given t by P occ T t and the pdf of the inverse occupation time It T, for a given T by I occ t T. While our final goal is to obtain the disorder averaged distributions P loc T t, I loc t T, P occ T t and I occ t T, it is however, instructive to study the pure case first, before tackling the problem with disorder which is obviously harder. In the same spirit, we have presented the detailed discussions on the local time, inverse local time, occupation time and inverse occupation time for the pure case = in Secs. III, V, VII and IX respectively, before computing their disorder average in Secs. IV, VI, VIII and X respectively. Sec. XI contains some concluding remarks. Some of the details are relegated to the appendixes. The results are summarized in Tables I, II, and III. II. GENERAL APPROACH In this section we will show how to compute the pdfs PT t, x and It T, x for arbitrary non-negative V x and arbitrary starting position x = x, for each realization of random force Fx, by using a backward Fokker-Planck equation approach. In the following discussion we will denote the functional defined in Eq. by g t {xτ}, x, and use T as the value of the functional for a given path {xτ}, for τ t]. Since V x is considered to be non-negative, T defined by Eq. has only positive support. Therefore, a natural step is to introduce the Laplace transform of the pdf PT t, x with respect to T Q p x, t = PT t, xe pt dt = e pg t {xτ},x = { exp p t x=x V xt ] dt } x=x, 8 where x=x denotes the average over all paths that start at the position x = x and propagate up to time t. Our aim is to derive a backward Fokker-Planck equation for Q p x, t with respect to the initial position x = x. We consider a particle starting from the initial position x, evolves via Eq. 3 up to time t + t. Now we split the time interval, t + t] into two parts: an infinitesimal interval, t], over which the particle experiences an infinitesimal displacement x from its initial position x and the remaining interval t, t + t] in which the particle evolves

4 from a starting position x + x. Since x = x, one gets t V xt ] dt = V x t + O t ]. Therefore using Eq. 8, and splitting the integral over t into the above two time intervals we obtain { } t+ t Q p x, t + t = exp p V xt ] dt x=x = e Q pv x t p x + x, t, 9 x where x denotes the average with respect to all possible displacements x. Now, in the limit t, integrating Eq. 3 one gets, x = Fx t + t ητdτ + O t ]. Hence, using the zero mean and the uncorrelated properties of the noise we get x x lim = Fx and lim =, t t t t Therefore, by Taylor expanding Eq. 9 for small x, and taking the limit t, one arrives at the backward Fokker- Planck equation, Q p t = Q p x + Fx Q p x pv xq p, with the initial condition Q p x, =, which is easily checked by Eq. 8. The advantage of the above equation over the usual Feynman-Kac formalism, ] is that, in the later case one has a forward Fokker-Planck equation spatial derivative with respect to the final position where after obtaining the solution of the differential equation, one has to again perform an additional step of integration over the final position. In contrast, Eq. involves the spatial derivatives with respect to the initial position of the particle, and hence no additional step of integration over the final position is required. The standard practice of attacking the partial differential equations of above type is by using the method of Laplace transform. We define the Laplace transform of Q p x, t with respect to t ux = = Q p x, te αt dt dt e αt dt e pt PT t, x, 3 where for notational convenience, we have suppressed the α and p dependence of ux. Now by taking Laplace transform of Eq. with respect to t we obtain the ordinary differential equation u x + Fxu x α + pv x]ux =, 4 where u x = du/dx. The appropriate boundary conditions ux ± are to be derived from the observation that if the particle starts at x ± it will never cross the origin in finite time. Note that Eq. 4 is valid for each sample of the quenched random force Fx. Thus in principle, from the solution ux one obtains PT t, x by inverting the double Laplace transform in Eq. 3 for each sample of quenched random potential, and then takes the average over the disorder. Our next goal in this section, is to show how to compute the pdf It T, x for a given sample of the quenched random force Fx. It turns out that It T, x is related to the pdf PT t, x in their Laplace space as shown below. By definition we have, It T, x = δ t g T {xτ}, x x. 5 However, it is elementary that for each realization of path {xτ} δ t g T {xτ}, x = δ T g t {xτ}, x dt dt, 6 where dt/dt is the usual Jacobian of the transformation, which is simply dt/dt as both T and t have only positive support. It immediately follows from the above two equations that It T, x = δ T g dt t {xτ}, x. 7 dt x Therefore, Laplace transform of It T, x with respect to T reads dt e pt It T, x = e t {xτ},x dg pg dt x = p t Q px, t, 8 where Q p x, t is given by Eq. 8. Now taking a further Laplace transform in Eq. 8 with respect to t, it is straightforward to obtain dt e αt dt e pt It T, x = αux. 9 p Thus, we have established via Eqs. 8 and 9, the relationships between the Laplace transforms of the pdf of the functional T defined by Eq. and the pdf of the inverse functional defined by Eq. 7. Hence, again in principle, from the solution ux of the ordinary differential equation 4, one obtains It T, x by inverting the double Laplace transform in Eq. 9 for each sample of quenched random potential, and then takes the average over the disorder. Note that putting α = in Eq. 9 and inverting the Laplace transform with respect to p immediately verifies the normalization condition It T, xdt =. In the rest of the paper, we will demonstrate how to implement this formalism for the particular examples of the local time corresponding to the choice V x = δx and the occupation time corresponding to the choice V x = θx. Since in these examples we consider the starting position of the particle to be the origin, we need to only find the solution u of

5 the differential equation 4. In each example, we will consider the pure cases = first, which help us anticipate the general features of the results in the disordered case > studied later. III. LOCAL TIME WITHOUT DISORDER = In this case V x = δx, corresponds to T in Eq. being the local time in the vicinity of the origin and PT t, in Eq. 3 being P loc T t the pdf of the local time T for a given observation time window of size t and the starting position of the particle x =. For our purpose we only need the solution u of the differential equation 4, which corresponds to the starting position of the particle being the origin. However, to obtain u we have to solve Eq. 4 in the entire region of x with the boundary conditions ux ± which are derived from the following observation. If the initial position x ±, the particle can not reach the origin in finite time, which means that the local time T =. Therefore, by substituting PT t, x ± δt in Eq. 3 one obtains the boundary conditions ux ± = α. We have to obtain the solutions ux = u + x for x > and ux = u x for x < by solving Eq. 4 separately in the respective two regions, u ± x + Fxu ± x αu ±x =, with the boundary conditions u + x = /α and u x = /α, and then matching the two solutions u + x and u x at x =. The matching conditions are u + = u = u, u + u = pu. The first condition follows from the fact that the solution must be continuous at x = and the second one is derived by integrating Eq. 4 across x =. By making a constant shift u ± x = /α+a ± y ± x, from Eq. one finds that y ± x satisfy the p-independent homogeneous equation y ±x + Fxy ±x αy ± x = 3 with the boundary conditions y + x and y x. The constants A ± are determined by the matching conditions given in Eq., which can be rewritten as A + y + = A y = u α, A + y + A y = pu. 4b 4a Eliminating the constants A ± from Eq. 4, we obtain the Laplace transform u, defined by Eq. 3 with PT t, P loc T t, as u = = where λα is simply given by λα = z z + dt e αt dt e pt P loc T t λα αp + λα]. 5 with z ± x = y ±x y ± x. 6 Note that putting p = in Eq. 5 and then inverting the Laplace transform with respect to α readily verifies the normalization condition P loc T tdt =. 7 Since λα is independent of p, inverting the Laplace transform in Eq. 5 with respect to p yields Gα = = λα α dt e αt P loc T t exp λαt], 8 which is valid for any arbitrary force Fx. In the following subsections we will consider qualitatively different types of deterministic potentials to derive more explicit results. A. Flat potential We first consider the simple Brownian motion without any external force, Fx =. In this case the solutions of Eq. 3 are obtained as y ± x = y ± exp x ] α. 9 Using the solutions in Eq. 6 one gets λα = α and hence the Laplace transform Gα in Eq. 8 becomes Gα = dt e αt P loc T t = e αt. 3 α Now inverting the Laplace transform with respect to α, one finds that the distribution of the local time is Gaussian for all T and t, P loc T t = πt exp T B. Unstable potential t ]. 3 Now we consider the case of a Brownian particle moving in an unstable potential Ux such that Ux ±. The corresponding repulsive force Fx drives the particle

6 eventually either to + or to. The pdf of the local time P loc T t in the case of an unstable potential tends to a steady distribution P loc T as t, which can be computed explicitly. To see this consider the function Gα in Eq. 8. By making a change of variable τ = αt, it follows from Eq. 8, Gα = α dτ P loc T τ. 3 α Assuming P loc T t = P loc T, we find form the above equation that Gα P loc T/α as α. Comparing this behavior with Eq. 8 gives P loc T = λexp λt], 33 provided λ is a finite positive number. Thus generically for all repulsive force Fx, the local time distribution has a universal Poisson distribution in the limit t. The only dependence on the precise form of the force Fx is through the rate constant λ. The rate constant λ can be expressed in terms of the force Fx in a more explicit manner. Putting α = in Eq. 3 and solving the resulting equation with the boundary conditions y + x and y x we get x y + x = y + ψ ydy, x >, 34 ψ ydy x y x = y ψ ydy, x <, 35 ψ ydy where ψy = exp y Fxdx]. Substituting these results in Eq. 6 gives the rate constant as ] λ = ψ ydy +. 36 ψ ydy Let us now consider a simple example where the potential Ux = µ x with µ >, corresponding to the repulsive force Fx = µ signx from the origin. In this case ψy = exp µ y ] and hence from Eq. 36 we get λ = µ. C. Stable potential We now turn our attention to the complementary situation when the potential Ux is stable, i.e., Ux ±. In this case the force Fx is attractive towards the origin so that the system eventually reaches a well defined stationary state. The stationary probability distribution px for the position of the particle is given by the Gibbs measure px = e Ux, 37 Z where Ux = x Fx dx and Z is the partition function, Z = e Ux dx. 38 In this case the Laplace transform Gα of the pdf of the local time P loc T t is still given by Eq. 8. However, unlike the unstable potential in the previous section, the distribution P loc T t does not approach a steady state as t. Instead it has a rather different asymptotic behavior. To deduce this asymptotic behavior, let us first consider the average local time T = t δxt ] dt. For large t, the average δxt ] approaches its stationary value δxt ] p, where p = /Z from Eq. 37. Hence as t the ratio T/t approaches the limit T t Z, 39 where Z is given by Eq. 38. Thus for large t, the average local time scales linearly with time t, which indicates that the natural scaling limit in this case is when t, T but keeping the ratio r = T/t fixed. We will see that in this scaling limit the local time distribution P loc T t tends to the following asymptotic form ] T P loc T t exp tφ, 4 t where Φr is a large deviation function. To compute the large deviation function we first substitute this presumed asymptotic form of P loc T t given by Eq. 4 in the Laplace transform Gα = e αt P loc T tdt and then make a change of variable r = T/t in the integration. The resulting integral can be evaluated in the large T limit by the method of steepest descent, which gives Gα exp TWα] where Wα = min r {α + Φr}/r]. Comparing this result with Eq. 8 gives min r α + Φr r ] = λα, 4 where λα is given by Eq. 6. Thus λα is just the Legendre transform of Φr. Inversion of this transform gives the exact large deviation function Φr = max α α + rλα], 4 with λα given by Eq. 6. This is a general result valid for any confining potential Ux. We will now explicitly compute the large deviation Φr for the particular potential given by Eq. 5 with µ < and =. Substituting the corresponding force Fx = µ signx in Eq. 3 and solving the resulting differential equations with the boundary conditions y + x and y x we get y ± x = y ± exp µ + ] µ + α x. 43 Substituting these results in Eq. 6 we get λα = µ + µ + α. From Eq. 4 one then gets the large deviation function Φr = r µ. 44

7 It turns out that for this particular form of the force Fx = µ signx, the Laplace transform in Eq. 8 can be inverted to get the local time distribution P loc T t exactly for all T and t. The calculations are presented in appendix A. We find that in the large t limit, the distribution reduces to the asymptotic form P loc T t exp t πt ] T t µ, 45 near the mean T = µ t, which verifies the result obtained above by the large deviation function calculation. In fact, the limiting Gaussian form of the distribution of the local time near its mean value is quite generic for any stable potentials where the system eventually becomes ergodic and is just the manifestation of the central limit theorem. From the definition, T T = t {δxt ] δxt ] } dt, it follows that when T T, the random variables δxt ] δxt ] at different times t become only weakly correlated. Then in the limit when t is much larger than the correlation time between these variables, one expects the central limit theorem to hold which predicts a Gaussian form for T near its mean value T. IV. LOCAL TIME WITH DISORDER > So far we have considered the case where the random part of the potential was not present. In this section we will study the effect of the randomness by adding a random part to the potential. In particular, we will consider the diffusive motion of the particle when the force Fx is given by Eq. 6 with >. Equation 8 still remains valid for each realization of the force Fx, i.e., for each realization of {ξx}. Our aim is to compute the average of the pdf of the local time P loc T t over the noise history {ξx}. From Eq. 8, one needs to know the distribution of λα = z z + ]/, which is now a random variable since Fx is random. The variables z + and z are independent of each other and therefore their joint probability distribution factorizes to the individual distributions. The calculations of these distributions are presented in appendix B. Using the distributions of z ± from Eqs. B9 and B respectively with a ± = α, one gets with qt = exp λαt] = ] qt, 46 q w µ/ exp { w + T + α }] dw w 47 α + T = α µ/ + T µ/ K µ/, 48 where K ν x is the modified Bessel function of order ν 44] and K ν x = K ν x. Averaging Eq. 8 over disorder we finally get the exact formula dt e αt P loc T t = d q αq T ]. 49 dt However, it is not an easy task to invert the Laplace transform to get the exact distribution P loc T t for all T and t. In the following subsections we will extract the asymptotic behaviors of P loc T t, for the three cases, when the deterministic part of the potential is: i flat corresponding to µ =, ii unstable corresponding to µ > and iii stable corresponding to µ <. A. Flat potential µ = Sinai model We first consider a particle diffusing in the continuous Sinai potential, i.e., µ = in Eq. 5. Our aim is to find out how this random potential modifies the behavior of the local time. In this case substituting qt and q from Eq. 48 with µ = in Eq. 49 we get the Laplace transform of the disorder averaged local time distribution as dt e αt P loc T t = αk α/ T K α + T. 5 We will now consider the interesting limit where both t and T are large, but the ratio y = T/t is kept fixed. This corresponds to taking the limit α with αt = s keeping fixed. In this limit α K log α 5 and α + T s K K. 5 Therefore substituting t = s/αy and T = s/α in Eq. 5, in the limit α we get dy e s/y s αy P loc s α ] s = αy 4 log α s s K. 53 The above equation suggests that, in the limit t and T, while their ratio T/t is kept fixed, P loc T t should have the scaling form P loc T t = t log t f T/t. 54

½ ݵ 8 Now substituting the above form in Eq. 53 and making the change of variable ỹ = /y, we obtain after straightforward simplification ] dỹ e sỹ f /ỹ s ỹ = 4K. 55 Note that the right hand side of the above equation is simply the Laplace transform of the function f /ỹ/ỹ. Therefore by using the identity e a /4ω ω e sω dω = K a s 56 and the convolution property of Laplace transform, we can invert the Laplace transform in Eq. 55 with respect to s. Inverting the Laplace transform and after simplification we finally get / dx f /ỹ = ỹ x x exp ỹx x Therefore the scaling function f y is simply given by f y = y / dx x x exp By substituting x x = /z gives f y = y 4 y x x ]. 57 ]. 58 dz exp y zz 4 z, 59 where the integral can be evaluated exactly 44], which finally gives the scaling function in Eq. 54 as f y = y e y/ K y/. 6 However the scaling given by Eq. 54 breaks down for very small y very small T when y. The scaling function is displayed in Fig. 3. In the large y limit, using the asymptotic the behavior K ν x π/xe x from Eq. 6 we find that f y π y 3/ e y/. B. Unstable potential µ > In this case the behavior in the limit t can be obtained by either setting α = in the integral in Eq. 47 or taking the α limit in K ν. in Eq. 48, which gives qt Γµ/ µ/ + T µ/, 6 where Γx is the Gamma function 45]. Substituting qt and q in Eq. 49 and inverting the Laplace transform with respect to α gives P loc T t = µ + T µ/, 6.5 3 Ý FIG. 3: The scaling function f y in Eq. 54. The solid line is plotted by using Eq. 6, and the dash line is plotted by using the limiting form f y π y 3/ e y/ as y. i.e., in the limit t, the distribution P loc T t tends to a steady state distribution P loc T for all T. The disorder averaged local time distribution has a broad power law distribution even though for each sample the local time has a narrow exponential distribution see Eq.33 in Sec. III B. This indicates wide sample to sample fluctuations and lack of self-averaging. C. Stable potential µ < In this case substituting qt and q from Eq. 48 in Eq. 49 and denoting ν = µ / we get dt e αt P loc T t = αk ν α/ ] α + T + T ν/ K ν. 63 T We consider the scaling limit where both t and T are large, but their ratio y = T/t is kept fixed. This corresponds to taking the limit α with keeping αt = s fixed, which gives the following limiting forms + T s α, 64 α K ν Γν ν, 65 α K ν α + T K ν s. 66

9 Substituting the above limits on the right hand side of Eq. 63 and making change of variables t = s/αy and T = s/α on the left hand side, it is straightforward to get dy e s/y s s αy P ] s loc α = αy in the limit α. P loc T t 4 ν Γ ν s s ν/ K ν s ], 67 This suggests the limiting form for P loc T t t f T/t, 68 fy.5.4.3.. ւ ν < ւ ν > in the scaling limit t and T with a fixed ratio y = T/t. To compute the scaling function we substitute the above scaling form in Eq. 67, and make the change of variable ỹ = /y. Then Eq. 67 simplifies to the Laplace transform dỹ e sỹ f /ỹ ỹ ] = 4 ν Γ ν s ν/ K ν s ], 69 which can be inverted with respect to s, by using the identity a ν e a /4ω ω ν+ e sω dω = s ν/ K ν a s 7 and the convolution property of the Laplace transform. After simplification, the inverse Laplace transform gives / f y = yν dx ν Γ ν x ν+ x ν+ ] y exp. 7 x x By making a change of variable x x = /z in the above integral, it can be presented in the form f y = yν ν Γ ν 4 z ν / z 4 / exp y z dz, 7 which now can be expressed in more elegant forms 45] as ] π y ν f y = e y/ Γ U/, + ν, y/, ν 73 where Ua, b, x is the Confluent Hypergeometric Function of the Second Kind also known as Kummer s function of the second kind 45], which has the following limiting behaviors: U/, + ν, x Γν π x ν for small x, 74a U/, + ν, x x for large x. 74b 3 4 5 y/ FIG. 4: The scaling function f y in Eq. 68, plotted by using Eq. 73. ν = µ /. The scaling function f y is displayed in Fig. 4. Using the limiting behaviors from Eq. 74, one finds that the scaling function decays as f y y 4ν 3/ e y/ for large y. For small y, the scaling function behaves as f y y ν, which increases with y for ν >, however, diverges when y for ν <, a behavior qualitatively similar to the Sinai case see Fig. 3. For ν <, the disorder wins over the strength of the stable potential. In that situation when the particle gets trapped in the wells of the random potential, the weak external deterministic force often can not lift it out of the well and send towards the origin. Therefore, the scaling function f y carries very large weight near y = which corresponds to very small local time T for a given observation time t. Note that, for the particular value ν = /, the scaling function has a simple form f y = /πy exp y/. V. INVERSE LOCAL TIME WITHOUT DISORDER = The inverse local time means how long one has to observe the particle until the total time spent in the infinitesimal neighborhood of the origin is T. The double Laplace transform of the pdf of the inverse local time is obtained by simply putting x = in Eq. 9. Corresponding u in Eq. 9, which is nothing but the double Laplace transform of the pdf of local time, has already been evaluated in Sec. III and is given by Eq. 5. Substituting u and replacing It T, with the pdf of the inverse local time I loc t T, after straightforward simplification, for x = Eq. 9 reads dt e αt dt e pt I loc t T = p + λα, 75 where λα is given by Eq. 6, which depends on the force Fx through Eq. 3. Inverting the Laplace transform with

T I loct T.4.3.. STABLE POTENTIAL µ < FLAT POTENTIAL µ = it decays exponentially I loc t T exp µ t/. On the other hand when the potential is unstable, µ >, as we see from Eq. 79, the distribution I loc t T is not normalized to unity. In this case the particle escapes to ± with probability e µt and Eq. 78 gives the distribution only for those events where the particle does not escape to ±. Therefore for µ >, it is appropriate to represent the full normalized distribution as I loc t T = T ] exp T + µt πt 3 t + e µt δt. 8 UNSTABLE POTENTIAL µ > ւ Note that the second term does not show up in the Laplace transform of I loc t T with respect to t. 4 6 8 t/t VI. INVERSE LOCAL TIME WITH DISORDER > FIG. 5: The pdfs of the inverse local time for stable µ = /, flat µ = and unstable µ = / potentials, plotted using Eq. 78 and T =. respect to p gives the general formula dt e αt I loc t T = exp λαt], 76 valid for arbitrary force Fx, a result known in the mathematics literature 7, 46]. We first consider the pure case where the force given by Eq. 6 with =. Substituting solutions of Eq. 3 for Fx = µ signx in Eq. 6 we obtain λα = µ + µ + α. Now using this λα in Eq. 76 and making a change of the parameter α = β µ / we get ] dt e βt e µt/ I loc t T = e µt e βt, 77 where the right hand side is simply the Laplace transform of e µt/ I loc t T with respect to t. The Laplace transform can be inverted to obtain the exact pdf of the inverse local time I loc t T = T exp πt 3 T + µt t ]. 78 with the normalization condition { for µ, I loc t Tdt = e µ+ µ T = e µt for µ >, 79 which is simply obtained by putting α = in Eq. 76. As we infer from Eq. 78, although in the limit t the inverse local time distribution I loc t T exp T /t is independent of µ, for large t it depends on the nature of the potential, as shown in Fig 5. While in the absence of any force, i.e., µ = the inverse local time distribution has a power-law tail I loc t T t 3/, for the stable potential, i.e., µ <, In this section, we switch on the disorder by considering > in the force given by Eq. 6. In the presence of disorder, taking the disorder average of Eq. 76 gives dt e αt I loc t T = exp λαt], 8 with λα = z z + ]/, where z + and z are independent random variables, whose distributions are given by Eqs. B9 and B respectively with a ± = α. The object exp λαt] on the right hand side of Eq. 8, has already been evaluated in Sec. IV, which is given by Eq. 46. In the following subsections we will determine the behavior of I loc t T in the scaling limit t, T, while keeping their ratio x = t/t fixed, for the three qualitatively different cases: i µ =, ii µ > and iii µ <. A. Flat potential µ = Sinai model Following the analysis of Sec. IV A, in the limit α with keeping αt = s fixed, exp λαt] 4 s log α K. 8 Therefore, substituting t = xs/α and T = s/α, in the limit α, Eq. 8 reads s sx dxe sx α I loc s ] = 4 s α α log α K. 83 This suggest the scaling form I loc t T = T log T g t/t, 84 in the limit t, T but keeping x = t/t fixed. Substituting this scaling form in Eq. 83, after straightforward

½ ܵ.7.6.5.4.3.. 5 5 5 3 FIG. 6: The scaling function g y in Eq. 84. The solid line is plotted by using Eq. 86, and the dash line is plotted by using the limiting forms: g x πx / exp /x for small x and g x logx/x for large x. simplification one obtains Ü dxe sx g x = 4K s. 85 Now direct comparison of the above equation with Eq. 55 gives g x = f /x/x, where f x is given by Eq. 6. Substituting f /x one obtains the scaling function g x as g x = x e /x K /x, 86 which is displayed in Fig. 6. The scaling function increases as g x πx / exp /x for small x and decays as g x logx/x at large x. B. Unstable potential µ > In this case the right hand side of Eq. 8 is given by α + T/ exp λαt] = +T ν K ν K ν α/, 87 with ν = µ/. Putting α = in the above equation gives the normalization condition, I loc t Tdt = + T ν, which implies that for the unstable potential, where the force is repulsive from the origin, the particle escapes to ± with probability + T ν, and the disorder averaged pdf I loc t T obtained by inverting the Laplace transform in Eq. 8 represents only those events where the particle does not escape to ±. Now in the limit of α with αt = s keeping fixed, one gets α K ν Γν α + T K ν Therefore Eq. 87 becomes exp λαt] T ν/ s ν, 88 s K ν. 89 ] ν 4T s 3 ν Γ s ν/ K ν. 9 ν In the corresponding limit T, t, but keeping their ratio x = t/t fixed, using the scaling form I loc t T = T ν+ g t/t 9 in Eq. 8 one finally arrives at the Laplace transform ] e sx 4 s g xdx = 3 ν Γ s ν/ K ν. ν 9 The Laplace transform can be inverted with respect to s to obtain the scaling function g x and in fact the inversion has already been done in Sec. IV C. Comparing the above equation with Eq. 69 readily gives g x = ν f /x/x where f x is given by Eq. 73. Substituting f /x gives g x = π ] e /x ν Γ ν U/, + ν, /x, 93 x ν+ where Ua, b, x is the Confluent Hypergeometric Function of the Second Kind, whose its small and large x behaviors are given in Eq. 74. The scaling function g x is displayed in Fig. 7. The scaling function increases as g x exp /x for small x and eventually decreases for large x as g x /x ν. In particular, for ν = / it has a very simple form g x = /π 3 x 3/ exp /x. C. Stable potential µ < Following the analysis of Sec. IV C, in the limit α, keeping αt = s fixed one gets exp λαt] = ] 4 s ν Γ s ν/ K ν, 94 ν with ν = µ /. On the other hand, in the corresponding limit T, t, but keeping t/t = x fixed, using the scaling form I loc t T = T g 3t/T, 95

ν gx..9.6.3 ν = 3 4 x ւ ν = FIG. 7: The scaling functions g x in Eq. 9 plotted by using Eq. 93. ν = µ /. one gets dt e αt I loc t T = dxe sx g 3 x, 96 with s = αt. Therefore, in this scaling limit Eq. 8 becomes ] dxe sx 4 s g 3 x = ν Γ s ν/ K ν. ν 97 Now comparing the above equation with Eq. 9 one gets g 3 x = ν g x, 98 where the scaling function g x is given by Eq. 93 and displayed in Fig. 7. While I loc t T has the same scaling function up to a multiplicative factor of ν for both stable and unstable potential, the physical behaviors, however, are quite different. For the stable potential, I loc t T is normalized to unity. Note that the scaling function g 3 x becomes narrower as one increases ν, as expected since the particle becomes more localized near the origin. For the unstable potential, on the other hand, the weight of I loc t T decreases as T ν, as one increases ν, as expected since when the repulsive force from the origin becomes stronger, the particle escapes to ± with a higher probability. VII. OCCUPATION TIME WITHOUT DISORDER = In this case V x = θx, corresponds to T in Eq. being the occupation time in the region x > and PT t, in Eq. 3 being P occ T t the pdf of the occupation time for a given observation time window of size t and the initial position of the particle x =. Again as before, we need to solve the differential equation 4 for x > and x < separately and then match the solutions at x =. The matching condition for the slope of the solutions is obtained by integrating Eq. 4 across x =. Thus the matching conditions are u + = u = u, and u + = u, 99 where u ± x satisfy the following differential equations: u +x + Fxu +x α + pu + x =, for x >, and u x + Fxu x αu x =, for x <. The boundary conditions of u ± x when x ± are obtained from the fact that if the starting position goes to ±, the particle will never cross the origin in finite time: PT t, x = δt T and PT t, x = δt, and hence from Eq. 3 u + = α + p, and u = α. Writing u + x = /α + p + B + y + x and u x = /α + B y x, we obtain the homogeneous differential equations for y ± x as y + x + Fxy + x α + py +x =, 3 for x >, and y x + Fxy x αy x =, 4 for x <, with the boundary conditions y + x = and y x =. The constants B ± are determined by the matching conditions given in Eq. 99, which can be rewritten as α + p + B +y + = α + B y = u, B + y + = B y. 5a 5b Eliminating the constants from Eq. 5, we obtain the double Laplace transform of the pdf of the occupation time where u = = l α, p α t dt e αt dt e pt P occ T t + l α, p α + p, 6 ] z l α, p =, 7 z z + ] z + l α, p =, 8 z z +

3 and z ± x = y ± x/y ±x. Note that l α, p + l α, p =. 9 Putting p =, in Eq. 6 gives u = /α, and hence inverting the Laplace transform with respect to α readily verifies the normalization t P occ T tdt =. For any symmetric deterministic potential the distribution of the occupation time is symmetric about its mean T = t/, i.e., P occ T t = P occ t T t. Then, it follows from this symmetry that l α + p, p = l α, p. In other words, the double-integral in Eq. 6 remains invariant under the following simultaneous replacements: α + p α and α α + p. Thus under these replacements one must have z + z and vice versa, which also implies that z + = z for p =. Equivalently, l α, = l α, = /, which also directly follows from Eqs. 9 and. Therefore if one splits the distribution function into two parts: P occ T t = R L T t + R R T t such that dt e αt t dt e αt t dt e pt R L T t = l α, p, α dt e pt R R T t = l α, p α + p, 3 then it follows from the above discussion that R L t T t = R R T t. This symmetry of the distribution will come handy later. Moreover, putting p = and inverting the Laplace transforms with respect to α gives the normalization for each part separately t R L T tdt = t R R T tdt =. 4 As an example, we first consider the pure case: = in the force given by Eq. 6. For Fx = µ signx, the solutions of Eqs. 3 and 4 are obtained as y + x = y + exp for x >, and, y x = y exp µ + µ + α + p ] x, 5 µ + ] µ + α x, 6 for x <. These give the expressions for z ± = y ±/y ± as z + = µ + ] µ + α + p, z = µ + ] 7 µ + α. In the following subsections we will consider the three different cases: i µ =, ii µ > and iii µ <. A. Flat potential µ = For µ =, using z + = α + p and z = α from Eq. 6 we get t dt e αt dt e pt P occ T t = αα + p. 8 Inverting the double Laplace transform with respect to p and then with respect to α finally reproduces the well known Lévy s arcsine law ] for the pdf of the occupation time of an ordinary Brownian motion, P occ T t = π, < T < t. 9 Tt T The distribution P occ T t diverges on both ends T = and T = t, which indicates that the Brownian particle tends to stay on one side of the origin. B. Unstable potential µ > Since for µ >, the force is repulsive from the origin x =, one would expect the occupation time distribution to be convex concave upward, with minimum at T = t/. Now in the limit of large window size t, the part of the distribution P occ T t to the left of the midpoint T = t/ approaches R L T t, as the midpoint itself goes to. By making a change of variable z = αt, it follows from Eq. z/α dz e z dt e pt R L T z/α = l α, p. Now the large t limit of R L T t can be obtained by taking α in the above equation, where one realizes that R L T t approaches a steady t independent distribution, R L T t R L T, whose Laplace transform is given by dt e pt R L T = l, p, where l, p is obtained from Eq. 7, by using z ± from Eq. 7, which gives l, p = µ 3µ + µ + p. The above Laplace transform can be inverted with respect to p, which gives R L T = µ e µ T/ 3µ exp πt 9µ T ] 3µ erfc T. 3 with the normalization R L TdT = l, = /.

4 The limiting behavior of the distribution is given by R L T µ πt, 4 for small T and decays exponentially for large T, R L T 9µ e µ T/. 5 π T 3/ C. Stable potential µ < As we discussed earlier in Sec. III C in the context of the local time, for generic stable potential Ux the system eventually becomes ergodic at large t and hence the average θxt] approaches its stationary value θxt] Z + /Z, where Z = e Ux dx is the equilibrium partition function and Z + = e Ux dx is the restricted partition function. Therefore, for large t the average occupation time T = t θxt ] dt scales linearly with t Z+ T t. 6 Z Note that when the potential Ux is symmetric about zero, the average occupation time T = t/ for all t. From the definition, T T = t {θxt ] θxt ] } dt, it follows that when T T, the random variables θxt ] θxt ] at different times t become only weekly correlated. Then in the limit when t is much larger than the correlation time between these variables, one expects the central limit theorem to hold, which predicts a Gaussian form for the distribution of the occupation time T near the mean value T, ] T T P occ T t exp, 7 where the variance = T T can be obtained from the Laplace transform of the moments T n e αt dt = n n u p n, 8 p= with u given by Eq. 6. For the particular attractive force Fx = µ signx, using z ± from Eq. 7 in Eq. 6 and taking derivatives with respect to p in we get u p u p = p= α, 9 = p= α 3 + 4µ α + O. 3 α Therefore inverting the Laplace transform in Eq. 8 with respect to α immediately gives T = t/ for all t, and T = t /4 + t/4µ for large t which gives = t/4µ. VIII. OCCUPATION TIME WITH DISORDER > Now we consider the occupation time when the disorder is switched on: > in Eq. 6. Our aim is to calculate the disorder averaged P occ T t. As one realizes from Eqs. 6, 7 and 8, to calculate P occ T t one needs the distribution of z + and z, which are given by Eqs. B9 and B with a + = α + p and a = α respectively. In the following subsections, we will consider the three cases: i µ =, ii µ > and iii µ <. A. Flat potential µ = Sinai model We first consider the diffusive motion of a particle in a continuous Sinai potential, where the potential itself is a Brownian motion in space. In the limit of large window size t the left half of the disorder averaged pdf of the occupation time R L T t for T t/ is obtained by taking the disorder average in Eq.. The right half of the distribution for t/ T t is just the symmetric reflection of the left part. The detailed calculations for R L T t are presented in appendix C. We find that R L T t has a large t behavior R L T t RT, 3 log t where the function RT is independent of t. The limiting behaviors of RT are given by RT, 3 πt as T and for large T. RT T, 33 B. Unstable potential µ > For µ >, we find that disorder does not change the asymptotic behavior of the distribution for the pure case qualitatively. The calculations are presented in appendix D. We find that in the limit t the left half of the disorder averaged occupation time distribution tends to a t independent form R L T t = R L T. 34 In fact the small T limit of R L T remains same as in the pure case R L T µ πt. 35 For large T, the distribution R L T still decays exponentially R L T e bt, 36

Ó Ýµ 5 where the decay coefficient b is, however, different from the pure case see Eq. D7. C. Stable potential µ <.5 ½ This particular situation, where one finds the interplay between two competing processes, is a very interesting one. On one hand, as we discussed in Sec. VII C, the stable potential in the absence of the disordered potential makes the system ergodic in the large t limit, and as a result the pdf of the occupation time is peaked at = t/ and decays fast away from it. On the other hand, as we discussed in Sec. VIII A, without any underlying deterministic potential the disorder averaged pdf of the occupation time is convex concave upward with a minimum at T = t/ and diverges at the both ends T and T t. Therefore, if both the stable potential and disordered potential are included, as their relative strength ν = µ / is varied, one expects a phase transition at some critical value ν c where the system looses ergodicity. In the scaling limit where both t and T, but their ratio y = T/t is kept fixed, we find that the disorder averaged pdf of the occupation time has a scaling form P occ T t = t f ot/t. 37 The calculation of the scaling function f o y is presented in appendix E, where we find the Beta law f o y = Bν, ν y y]ν, y, 38 where ν = µ / and Bν, ν is the Beta function 44]. Now if one tunes the parameter ν by varying either µ or the disorder strength, the distribution P occ T t exhibits a phase transition in the ergodicity of the particle position at ν c = Fig. 8. For ν < ν c the distribution f o y in Eq. 37 is convex with a minimum at y = / and diverges at the two ends y =,. This means that particle tends to stay on one side of the origin such that T is close to either or t. In other words the paths with small number of zero crossings carry more weight than the ones that cross many times. For ν > ν c the scenario is exactly opposite, where f o y is maximum at the mean value y = / indicating that particle tends to spend equal times on both sides of the origin x =, such that paths with large number of zero crossings, for which T is closer to t/ carry larger weight. Similar phase transition in the ergodicity properties of a stochastic process as one changes a parameter, was first noted in the context of diffusion equation 9], and later found for a class of Gaussian Markov processes ] and in simple models of coarsening 47, 48]. A very interesting observation about Eq. 38 is that for ν = /, the result is same as Lévy s result for the onedimensional Brownian motion given by Eq. 9. It seems as if the attractive force cancels the effect of disorder exactly at ν = /. However, this is no more true in the context of the local time..5 ½.5 FIG. 8: The scaling functions f ox in Eq. 37 plotted by using Eq. 38. IX. INVERSE OCCUPATION TIME WITHOUT DISORDER = In this case It T, in Eq. 9 is replaced with I occ t T, which is the distribution of the time t needed to observe the particle with a starting position x =, until the total amount of time spent on the positive side x > is T. Corresponding u in Eq. 9 with x =, which is the double Laplace transform of the pdf of the occupation time, has already been evaluated in Sec. VII, which is given by Eq. 6. Substituting u in Eq. 9 gives dt e pt T Ý dt e αt I occ t T = l α, p α + p, 39 where l α, p is given by Eq. 8. Comparing the above equation with Eq. 3, one can infer that I occ t T and R R T t have the same functional form, i.e., I occ t T = R R T t and especially for symmetric deterministic potential I occ t T = R R T t = R L t T t. It is useful to present the above equation in the following form, dz e z dτ e ατ I occ τ + z z β = l β, α β, β 4 which has been obtained by substituting p = β α in Eq. 39 and subsequently making the change of variables βt = z and τ = t T. On the right hand side, we have substituted l α, β α = l β, α β, using Eq. and l α, p is given by Eq. 7. Now by taking the limit β in Eq. 4, one obtains the large T behavior of I occ t T. For the pure case, = in Eq. 6, we have already obtained z ± in Sec. VII, which are given by Eq. 7 and hence we can evaluate l α, p and l α, p by using