This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Extremal statistics for a one-dimensional Brownian motion with a reflective boundary

Feng Huang1    Hanshuang Chen2 [email protected] 1School of Mathematics and Physics & Key Laboratory of Architectural Acoustic Environment of Anhui Higher Education Institutes, Anhui Jianzhu University, Hefei 230601, China
2School of Physics and Optoelectronic Engineering, Anhui University, Hefei 230601, China
Abstract

In this work, we investigate the extreme value statistics of a one-dimensional Brownian motion (with the diffusion constant DD) during a time interval [0,t]\left[0,t\right] in the presence of a reflective boundary at the origin, when starting from a positive position x0x_{0}. We first obtain the distribution P(M|x0,t)P(M|x_{0},t) of the maximum displacement MM and its expectation M\langle M\rangle. In the short-time limit, i.e., ttdt\ll t_{d} where td=x02/Dt_{d}=x_{0}^{2}/D is the diffusion time from the starting position x0x_{0} to the reflective boundary at the origin, the particle behaves like a free Brownian motion without any boundaries. In the long-time limit, ttdt\gg t_{d}, M\langle M\rangle grows with tt as Mt\langle M\rangle\sim\sqrt{t}, which is similar to the free Brownian motion, but the prefactor is π/2\pi/2 times of the free Brownian motion, embodying the effect of the reflective boundary. By solving the propagator and using a path decomposition technique, we then obtain the joint distribution P(M,tm|x0,t)P(M,t_{m}|x_{0},t) of MM and the time tmt_{m} at which this maximum is achieved, from which the marginal distribution P(tm|x0,t)P(t_{m}|x_{0},t) of tmt_{m} is also obtained. For ttdt\ll t_{d}, P(tm|x0,t)P(t_{m}|x_{0},t) looks like a U-shaped attributed to the arcsine law of free Brownian motion. For tt equal to or larger than order of magnitude of tmt_{m}, P(tm|x0,t)P(t_{m}|x_{0},t) deviates from the U-shaped distribution and becomes asymmetric with respect to t/2t/2. Moreover, we compute the expectation tm\langle t_{m}\rangle of tmt_{m}, and find that tm/t\langle t_{m}\rangle/t is an increasing function of tt. In two limiting cases, tm/t1/2\langle t_{m}\rangle/t\to 1/2 for ttdt\ll t_{d} and tm/t(1+2G)/40.708\langle t_{m}\rangle/t\to(1+2G)/4\approx 0.708 for ttdt\gg t_{d}, where G0.916G\approx 0.916 is the Catalan’s constant. Finally, we analytically compute the statistics of the last time tt_{\ell} the particle crosses the starting position x0x_{0} and the occupation time tot_{o} spent above x0x_{0}. We find that t/t1/2\langle t_{\ell}\rangle/t\to 1/2 in the short-time and long-time limits, and reaches its maximum at an intermediate value of tt. The fraction of the occupation time to/t\langle t_{o}\rangle/t is a monotonic function of tt, and tends towards 1 in the long-time limit. All the theoretical results are validated by numerical simulations.

I Introduction

Despite infrequent occurrences, extreme events are of vital importance as they may bring devastating consequences. Examples include natural calamities, such as earthquake, tsunamis and floods, economic collapses, and outbreak of pandemic [1, 2, 3, 4, 5, 6]. Extreme-value statistics (EVS) has been a branch of statistics which deals with the extreme deviations of a random process from its mean behavior. For independent and identically distributed random variables, it is known that the extreme-value distribution falls into three famous universality classes, namely, Gumbel, Fréchet, and Weibull depending on the tails of the distribution of random variables [7]. The study of EVS has also become extremely important in the field of disordered systems [8, 9], fluctuating interfaces [10, 11], interacting spin systems [12], stochastic transport models [13, 14], random matrices [15, 16, 17], epidemic outbreak [18], and computer search algorithms [19, 20, 21].

In recent years, there is an increasing interest in studying the EVS for strongly correlated stochastic processes [22, 23, 24, 25, 26, 27, 28]. We refer the reader to [29, 30] for two recent reviews on the subject. One of the central goals on this subject is to compute the statistics of extremes, i.e., the maximum MM of a given trajectory x(t)x(t) during an observation time window [0,t]\left[0,t\right], and the time tmt_{m} at which the maximum MM is reached. A paradigmatic example is a one-dimensional Brownian motion for a fixed duration tt starting from the origin, where the joint distribution of MM and tmt_{m} is given by [31]

P0(M,tm|t)=M2πDtm3(ttm)eM2/4Dtm\displaystyle P_{0}(M,t_{m}|t)=\frac{M}{{2\pi D\sqrt{t_{m}^{3}\left({t-{t_{m}}}\right)}}}{e^{-{M^{2}}/4Dt_{m}}} (1)

with the diffusion constant DD. Marginalization of Eq.(1) over tmt_{m}, one gets the one-sided Gaussian distribution of MM,

P0(M|t)=Θ(M)πDteM2/4Dt,\displaystyle P_{0}(M|t)=\frac{\Theta(M)}{\sqrt{\pi Dt}}e^{-M^{2}/4Dt}, (2)

where Θ(z)\Theta(z) is the Heaviside step function such that Θ(z)=1\Theta(z)=1 if z>0z>0 and Θ(z)=0\Theta(z)=0 otherwise. In particular, from Eq.(2) one can easily obtain that the expected maximum M(t)M(t) of a free Brownian motion grows like M(t)=2Dt/π\langle M(t)\rangle=2\sqrt{Dt/\pi}. On the other hand, one can obtain the marginal distribution of tmt_{m} by integrating P0(M,tm|t)P_{0}(M,t_{m}|t) over MM,

P0(tm|t)=1πtm(ttm),0tmt,\displaystyle P_{0}(t_{m}|t)=\frac{1}{\pi\sqrt{t_{m}(t-t_{m})}},\quad 0\leq t_{m}\leq t, (3)

which is often referred to as the “arcsine law” due to P. Lévy [32, 33, 34]. The name stems from the fact that the cumulative distribution of tmt_{m} reads F(z)=0zP0(tm|t)𝑑tm=(2/π)arcsinz/tF(z)=\int_{0}^{z}{P_{0}(t_{m}|t)dt_{m}}=(2/\pi)\arcsin\sqrt{z/t}. A counterintuitive aspect of the U-shaped distribution of Eq.(3) is that its average value tm=t/2\langle t_{m}\rangle=t/2 corresponds to the minimum of the distribution, i.e., the less probable outcome, whereas values close to the extrema tm=0t_{m}=0 and tm=tt_{m}=t are much more likely. In fact, the other two times, i.e., the last time tt_{\ell} the process crosses the origin and the occupation time tot_{o} spent on the positive (or the negative) semi axis, satisfy the same distribution as Eq.(3) as well. The three arcsine laws for Brownian motion play a central role in EVS.

Extreme observables were also studied for variants of Brownian motion. For a class of constrained Brownian motions including Brownian excursions, Brownian meanders, and reflected Brownian bridge, the joint distribution of the maximum MM and the extreme time tmt_{m}, and their marginal distributions have been analytically obtained [35]. For the random acceleration process, which is one of the simplest non-Markov stochastic processes, the distribution of tmt_{m} [36] and the first two moments of the occupation time tot_{o} [37] were studied. For the run-and-tumble motion, a representative model of active particles like E. coli, the distributions of three times, i.e., tmt_{m}, tt_{\ell} and tot_{o} were derived in one dimension [38], and the results were used to compute the statistical properties of the convex hull of a planar run-and-tumble motion [39, 40]. In Refs.[41, 42], the three arcsine laws were generalized to fractional Brownian motion, which is a non-Markovian Gaussian process indexed by the Hurst exponent H(0,1)H\in\left(0,1\right), generalizing standard Brownian motion (H=12H=\frac{1}{2}) to account for anomalous diffusion. Using a perturbative expansion in ϵ=H12\epsilon=H-\frac{1}{2}, the distributions of three times were analytically obtained up to second-order. The distributions of tmt_{m} and tot_{o} are symmetric with respect to half of the duration time t/2t/2. In the leading term, the two distributions are the same, but become distinguishable in the sub-leading term. The distribution of tt_{\ell} is markedly differently from the other two distributions; especially, it is asymmetric with respect to t/2t/2. Recently, Brownian motion and fractional Brownian motion subject to stochastic resetting have received increasing attention due to the emergence of a nonequilibrium steady state and its advantage in accelerating a random search [43, 44, 45, 46]. For a one-dimensional resetting Brownian motion with a fixed duration, the distributions of MM and tmt_{m} have been derived [47, 48], from which statistical properties of the convex hull of a planar resetting Brownian motion were obtained [48]. The distribution of tot_{o} for a one-dimensional resetting Brownian motion is considered, and particular attention was paid to the large deviation property of the distribution [49]. Very recently, extreme statistics and spacing distribution for the positions of a one-dimensional NN Brownian particles under the simultaneous resetting was investigated [28]. EVS has also been studied for general stochastic processes [50, 51, 52, 53, 31, 54, 55, 56, 57]. Extension to study the distribution of the time difference between the minimum and the maximum for stochastic processes has also been made in [58, 59] and to study the EVS before a first passage time through a specified threshold [60, 61, 62, 63, 64]. The distribution of a related observable, i.e., the number of distinct sites visited by a one-dimensional random walker before hitting a target [65], and the joint distribution of the first-passage time to a target and the number of distinct sites visited [66] have been obtained analytically for several representative Markovian processes, such as simple symmetric random walks, biased random walks, persistent random walks, and resetting random walks. Quite remarkably, the statistics of tmt_{m} has found applications in convex hull problems [39, 40, 67, 48] and also in detecting whether a stationary process is equilibrium or not [68, 69].

So far, most of previous studies were paid their attentions to the EVS of the stochastic process without confinement. However, in many practical situations, the stochastic process takes place in a confined geometry. For example, the range of animal foraging can get constrained by natural or human-built obstacles, such as a river, mountains, urban areas, roads, and so on. A planar Brownian motion in the presence of an infinite reflective wall was studied [70, 71]. It was shown that the presence of the wall breaks the isotropy of the process and induces a non-trivial effect on the convex hull of the Brownian motion. Very recently, a Brownian particle in a dd-dimensional ball with radius RR with reflecting boundaries was considered, and the statistics of the maximum displacement Mx(t)M_{x}(t) of along the xx-direction at time tt was studied [72]. It was shown the distribution of the fluctuation (Mx(t)R)/R(M_{x}(t)-R)/R in the long-time limit exhibits a rich variety of behaviors depending on the dimension dd. In a very recent work [73], the authors studied that EVS of a one-dimensional Brownian motion in the presence of a permeable barrier when starting a position infinitely approaching the barrier. They derived the joint distribution of the maximum displacement M(t)M(t) and the time tmt_{m} at the maximum is reached, from which the marginal distributions of M(t)M(t) and tmt_{m} were also obtained. It was also reported that the confinement produces a profound impact on the first-passage properties of the fractional Brownian motion [74] and the ergodicity of heterogeneous diffusion [75].

In this work, we aim to study the EVS of the Brownian motion in one dimension subject to a reflective boundary at the origin when starting from a positive position x0x_{0}, such that the Brownian motion is confined in the positive half-space. From an intuitive perspective of view, when the total time duration tt is much less than the diffusion time td=x02/Dt_{d}=x^{2}_{0}/D from the starting position to the reflective boundary, the reflective boundary has not much impact on the Brownian motion and thus the EVS is similar to that of a free Brownian motion. We should emphasize that tdt_{d} is not confused with the mean first-passage time from x0x_{0} to the origin. It is well-known that the first-passage time tft_{f} follows P(tf)=14πDtf3exp(x02/4Dtf)P(t_{f})=\frac{1}{\sqrt{4\pi Dt_{f}^{3}}}\exp\left({-{x_{0}^{2}}/{4Dt_{f}}}\right). The heavy tail of the distribution leads to the divergence of mean first-passage time. Instead, tdt_{d} can be understood as the typical value of the first-passage time since one can easily check that P(tf)P(t_{f}) attains its unique maximum at tf=tdt_{f}=t_{d}. A special focus should be paid to the case of tt equal to or larger than the order of the magnitude of tdt_{d} where the reflective boundary comes into effect. We first derive the survival probability of the Brownian particle in the presence of a reflective boundary at the origin and an absorbing boundary at x=Mx=M, thus leading to the distribution P(M|x0,t)P(M|x_{0},t) of the maximum displacement MM within a duration tt. We also obtain the expectation M\langle M\rangle of MM as a function of x0x_{0} and tt, from which we find that in the long-time limit MπDt\langle M\rangle\sim\sqrt{\pi Dt}, up to a correction proportional to x02/tx_{0}^{2}/\sqrt{t}. This is qualitatively the same as the case of the free Brownian motion (M2Dt/π\langle M\rangle\sim 2\sqrt{Dt/\pi} for the latter), but with a different prefactor. Moreover, we use a path decomposition technique to obtain the joint distribution P(M,tm|x0,t)P(M,t_{m}|x_{0},t) of the maximum displacement MM and the time tmt_{m} at which MM is reached. Marginalization of P(M,tm|x0,t)P(M,t_{m}|x_{0},t) over MM to obtain the distribution of P(tm|x0,t)P(t_{m}|x_{0},t). For t>tdt>t_{d}, the distribution of tmt_{m} deviates from the so-called arcsine law, and is no longer symmetric with respect to t/2t/2. Interestingly, we find that the ratio of the expectation of tmt_{m} to tt, tm/t\langle t_{m}\rangle/t, increases monotonically with tt, and its value approaches 1/2 in the limit ttdt\ll t_{d} and tends towards another constant (1+2G)/40.708(1+2G)/4\approx 0.708 in the opposite limit ttdt\gg t_{d}, where GG is the Catalan’s constant. At last, we show how the statistics of the other two times, i.e., the last time tt_{\ell} the process passes through the initial position x0x_{0} and the occupation time tot_{o} spent above x0x_{0}, is modified in the presence of the reflective boundary. The expectations of tt_{\ell} and tot_{o} are analytically obtained. In the short-time and the long-time limits, t/t\langle t_{\ell}\rangle/t tends towards 1/21/2. Interestingly, t/t\langle t_{\ell}\rangle/t shows a nonmonotonic dependence on tt, and a maximum t/t\langle t_{\ell}\rangle/t occurs at a moderate value of tt. The fraction of the occupation time to/t\langle t_{o}\rangle/t varies monotonically from 1/21/2 and approaches to 11 in the long-time limit.

The rest of the paper is organized as follows. In Sec.II we define the Brownian motion in one dimension subject to a reflective boundary and propose the extreme-value questions we want to study. In Sec.III and Sec.IV, we derive, both in the time domain and in the Laplace domain, the propagator and survival probability of the Brownian motion in an interval with a reflective end and an absorbing end. In Sec.V, we present the marginal distribution of the maximum displacement MM and its expected value. Using a path decomposition technique, in Sec.VI we obtain the joint distribution of MM and the time tmt_{m} at which the maximum MM is reached. In Sec.VII, we provide the marginal of the extreme time tmt_{m} and make asymptotic analysis for the expectation of tmt_{m} in the short-time and the long-time limits. In Sec.VIII, the statistics of the last time visited the starting position and the occupation time spent above the starting position are analytically given. Finally, the main conclusions and perspective are addressed in Sec.IX.

Refer to caption

Figure 1: A realization of a one-dimensional Brownian motion in the presence of a reflective wall at the origin starting from x0x_{0}. The displacement x(τ)x(\tau) of the particle reaches its maximum MM at the time tmt_{m} during a time interval [0,t]\left[0,t\right].

II Model

Let us consider a one-dimensional Brownian motion starting at x0x_{0}(>0>0) with a reflective boundary at the origin,

x˙=2Dξ(t)\displaystyle\dot{x}=\sqrt{2D}\xi(t) (4)

where DD is the diffusion constant, and ξ(t){\xi}(t) is a Gaussian white noise with zero mean ξ(t)=0\langle{\xi(t)}\rangle=0 and delta correlator ξ(t)ξ(t)=δ(tt)\langle{\xi(t)}{\xi(t^{\prime})}\rangle=\delta({t-t^{\prime}}). During a time interval [0,t]\left[0,t\right], the displacement x(t)x(t) of the Brownian reaches its maximum MM at time tmt_{m}, see Fig.1 for an illustration. We are interested in the joint distribution P(M,tm|x0,t)P(M,t_{m}|x_{0},t) of MM and tmt_{m}, and their marginal distributions of MM and tmt_{m}, P(M|x0,t)P(M|x_{0},t) and P(tm|x0,t)P(t_{m}|x_{0},t).

III Propagator with a reflective wall at x=0x=0 and an absorbing wall at x=Mx=M

We begin with the computation of the probability density function G(x,t|x0)G(x,t|x_{0}) of position of the Brownian particle or the propagator in the presence of a reflective wall at x=0x=0, when starting from an initial position x0[0,M)x_{0}\in\left[0,M\right). On the other hand, the maximum displacement MM of the particle within a time duration [0,t]\left[0,t\right] yields an absorbing boundary at x=Mx=M.

III.1 Propagator in the time domain

The Fokker-Planck equation for governing the time evolution of G(x,t|x0)G(x,t|x_{0}) reads

G(x,t|x0)t=D2G(x,t|x0)x2\displaystyle\frac{{\partial G\left({x,t|{x_{0}}}\right)}}{{\partial t}}=D\frac{{{\partial^{2}}G\left({x,t|{x_{0}}}\right)}}{{\partial{x^{2}}}} (5)

Perfect reflection is imposed by requiring that the probability current vanishes at the origin, such that

G(x,t|x0)x|x=0=0.\displaystyle\frac{\partial G(x,t|x_{0})}{\partial x}|_{x=0}=0. (6)

An absorbing boundary at x=Mx=M requires

G(M,t|x0)=0.\displaystyle G(M,t|x_{0})=0. (7)

Using separation of variables, Eq.(5) can be solved combined with the boundary conditions Eq.(6), Eq.(7), and initial condition G(x,0|x0)=δ(xx0)G(x,0|x_{0})=\delta(x-x_{0}), which gives the series representation of the propagator,

G(x,t|x0)=2Mn=0cos[(n+1/2)πx0M]cos[(n+1/2)πxM]e(n+1/2)2π2DtM2\displaystyle G\left({x,t|{x_{0}}}\right)=\frac{2}{M}\sum\limits_{n=0}^{\infty}{\cos\left[{\frac{{\left({n+1/2}\right)\pi{x_{0}}}}{M}}\right]}\cos\left[{\frac{{\left({n+1/2}\right)\pi x}}{M}}\right]e^{{-\frac{{{{\left({n+1/2}\right)}^{2}}{\pi^{2}}Dt}}{{{M^{2}}}}}} (8)

III.2 Propagator in the Laplace domain

In this subsection, we solve the propagator in the Laplace domain. By the Laplace transformations for Eq.(5), Eq.(6) and Eq.(7), we obtain

sG~(x,s|x0)δ(xx0)=Dd2G~(x,s|x0)dx2,\displaystyle s\tilde{G}({x,s|{x_{0}}})-\delta({x-{x_{0}}})=D\frac{{{d^{2}}\tilde{G}({x,s|{x_{0}}})}}{{d{x^{2}}}}, (9)
G~(M,s|x0)=0,G~(0,s|x0)=0.\displaystyle\tilde{G}({M,s|{x_{0}}})=0,\quad{\tilde{G}^{\prime}}({0,s|{x_{0}}})=0. (10)

Eq.(9) can be solved separately in the region 0<x<x00<x<x_{0} and in the region x0<x<Mx_{0}<x<M. In each region the solutions are

G~(x,s|x0)={A1eαx+B1eαx,0<x<x0,A2eαx+B2eαx,x0<x<M,\displaystyle\tilde{G}({x,s|{x_{0}}})=\left\{\begin{array}[]{llc}A_{1}e^{\alpha x}+B_{1}e^{-\alpha x},&0<x<x_{0},\\ A_{2}e^{\alpha x}+B_{2}e^{-\alpha x},&x_{0}<x<M,\\ \end{array}\right. (13)

where α=s/D\alpha=\sqrt{s/D}.

The boundary conditions in Eq.(10) leads to

{A1αB1α=0,A2eαM+B2eαM=0.\displaystyle\left\{\begin{gathered}{A_{1}}\alpha-{B_{1}}\alpha=0,\hfill\\ {A_{2}}{e^{\alpha M}}+{B_{2}}{e^{-\alpha M}}=0.\hfill\\ \end{gathered}\right. (17)

In addition, we require that G~(x,s|x0)\tilde{G}({x,s|{x_{0}}}) is continuous at x=x0x=x_{0}, i.e.,

A1eαx0+B1eαx0=A2eαx0+B2eαx0.\displaystyle{A_{1}}{e^{\alpha{x_{0}}}}+{B_{1}}{e^{-\alpha{x_{0}}}}={A_{2}}{e^{\alpha{x_{0}}}}+{B_{2}}{e^{-\alpha{x_{0}}}}. (18)

However, due to the presence of the delta function in Eq.(9) the derivative of G~(x,s|x0)\tilde{G}({x,s|{x_{0}}}) is not continuous at x=x0x=x_{0}. To find the matching condition for the derivative we integrate Eq.(9) from x0ϵx_{0}-\epsilon to x0+ϵx_{0}+\epsilon (with ϵ0+\epsilon\to 0^{+}), which yields

dG~(x,s|x0)dx|x=x0+dG~(x,s|x0)dx|x=x0=1D.\displaystyle{\left.{\frac{{d\tilde{G}\left({x,s|{x_{0}}}\right)}}{{dx}}}\right|_{x=x_{0}^{+}}}-{\left.{\frac{{d\tilde{G}\left({x,s|{x_{0}}}\right)}}{{dx}}}\right|_{x=x_{0}^{-}}}=-\frac{1}{D}. (19)

According to Eq.(13), the matching condition in Eq.(19) becomes

(A2αeαx0B2αeαx0)(A1αeαx0B1αeαx0)=1/D.\displaystyle\left({{A_{2}}\alpha{e^{\alpha{x_{0}}}}-{B_{2}}\alpha{e^{-\alpha{x_{0}}}}}\right)-\left({{A_{1}}\alpha{e^{\alpha{x_{0}}}}-{B_{1}}\alpha{e^{-\alpha{x_{0}}}}}\right)=-1/D. (20)

By solving Eq.(17), Eq.(18) and Eq.(20), the coefficients A1A_{1}, B1B_{1}, A2A_{2}, and B2B_{2} are given by

{A1=B1=sinh[α(Mx0)]2αDcosh(αM)A2=cosh(αx0)[tanh(αM)1]2αDB2=cosh(αx0)[tanh(αM)+1]2αD\displaystyle\left\{\begin{gathered}{A_{1}}={B_{1}}=\frac{{\sinh\left[{\alpha\left({M-{x_{0}}}\right)}\right]}}{{2\alpha D\cosh\left({\alpha M}\right)}}\hfill\\ {A_{2}}=\frac{{\cosh\left({\alpha{x_{0}}}\right)\left[{\tanh\left({\alpha M}\right)-1}\right]}}{{2\alpha D}}\hfill\\ {B_{2}}=\frac{{\cosh\left({\alpha{x_{0}}}\right)\left[{\tanh\left({\alpha M}\right)+1}\right]}}{{2\alpha D}}\hfill\\ \end{gathered}\right. (25)

Substituting Eq.(25) into Eq.(13), we obtain the propagator in the Laplace domain,

G~(x,s|x0)=cosh(αx<)sinh[α(Mx>)]αDcosh(αM)\displaystyle\tilde{G}\left({x,s|{x_{0}}}\right)=\frac{{\cosh\left({\alpha{x_{<}}}\right)\sinh\left[{\alpha\left({M-{x_{>}}}\right)}\right]}}{{\alpha D\cosh\left({\alpha M}\right)}} (26)

with x<=min(x,x0)x_{<}=\min(x,x_{0}) and x>=max(x,x0)x_{>}=\max(x,x_{0}).

IV Survival probability with a reflective wall at x=0x=0 and an absorbing wall at x=Mx=M

IV.1 Survival probability in the time domain

It is easy to see that the distribution of the maximum MM over the duration tt in absence of any absorbing barrier is related to the survival probability of the process x(τ)x(\tau) in presence of an absorbing barrier at MM, i.e., Q(x0,t|M)=Prob(x(τ)M|0τt)Q(x_{0},t|M)={\rm{Prob}}(x(\tau)\leq M|0\leq\tau\leq t), where Q(x0,t|M)Q(x_{0},t|M) is the survival probability that the process has not yet hit the absorbing boundary at x=Mx=M up to time tt starting from x0x_{0}, which is given by [76]

Q(x0,t|M)=0M𝑑xG(x,t|x0)=2πn=0(1)nn+1/2cos[(n+1/2)πx0M]e(n+1/2)2π2DtM2\displaystyle Q\left({{x_{0}},t|M}\right)=\int_{0}^{M}dx{G\left({x,t|{x_{0}}}\right)}=\frac{2}{\pi}\sum\limits_{n=0}^{\infty}{\frac{{{{\left({-1}\right)}^{n}}}}{{n+1/2}}\cos\left[{\frac{{\left({n+1/2}\right)\pi{x_{0}}}}{M}}\right]}e^{-\frac{{{{\left({n+1/2}\right)}^{2}}{\pi^{2}}Dt}}{{{M^{2}}}}} (27)

IV.2 Survival probability in the Laplace domain

We will see that in order to obtain the expectation of the maximum displacement MM it is convenient to solve survival probability in the Laplace domain. It is known that the survival probability satisfies the backward equation [76],

Q(x0,t|M)t=D2Q(x0,t|M)x02,\displaystyle\frac{{\partial{Q}\left({{x_{0}},t|M}\right)}}{{\partial t}}=D\frac{{{\partial^{2}}{Q}\left({{x_{0}},t|M}\right)}}{{\partial x_{0}^{2}}}, (28)

subject to the boundary conditions Q(M,t|M)=0Q(M,t|M)=0 and Q(x0,t|M)/x0|x0=0=0\partial Q(x_{0},t|M)/\partial x_{0}|_{x_{0}=0}=0 and initial condition Q(x0,0|M)=1Q(x_{0},0|M)=1 for x0[0,M)x_{0}\in\left[0,M\right). Performing the Lapalce transform, Q~(x0,s|M)=0𝑑testQ(x0,t|M)\tilde{Q}(x_{0},s|M)=\int_{0}^{\infty}dte^{-st}Q(x_{0},t|M), for Eq.(28) to obtain

Dd2Q~(x0,s|M)dx02sQ~(x0,s|M)=1.\displaystyle D\frac{{{d^{2}}{{\tilde{Q}}}\left({{x_{0}},s|M}\right)}}{{dx_{0}^{2}}}-s{{\tilde{Q}}}\left({{x_{0}},s|M}\right)=-1. (29)

Eq.(29) can be solved subject to the boundary conditions, which yields

Q~(x0,s|M)=1scosh(αx0)scosh(αM)\displaystyle\tilde{Q}(x_{0},s|M)=\frac{1}{s}-\frac{\cosh\left(\alpha x_{0}\right)}{s\cosh\left(\alpha M\right)} (30)

with α=s/D\alpha=\sqrt{{s}/{D}} again.

V Marginal distribution P(M|x0,t)P(M|x_{0},t)

Differentiating Eq.(27) with respect to MM gives the probability density function of MM,

P(M|x0,t)=2M2n=0(1)ne(n+1/2)2π2DtM2{x0sin[(n+1/2)πx0M]+2(n+1/2)πDtMcos[(n+1/2)πx0M]}.\displaystyle P\left({M|{x_{0}},t}\right)=\frac{2}{{{M^{2}}}}\sum\limits_{n=0}^{\infty}{{{\left({-1}\right)}^{n}}e^{-\frac{{{{\left({n+1/2}\right)}^{2}}{\pi^{2}}Dt}}{{{M^{2}}}}}\left\{{{x_{0}}\sin\left[{\frac{{\left({n+1/2}\right)\pi{x_{0}}}}{M}}\right]+\frac{{2\left({n+1/2}\right)\pi Dt}}{M}\cos\left[{\frac{{\left({n+1/2}\right)\pi{x_{0}}}}{M}}\right]}\right\}}. (31)

Refer to caption

Figure 2: Marginal distribution P(M|x0,t)P(M|x_{0},t) of maximum displacement MM for different time duration tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. Line and symbols represent the theory and simulation results, respectively.

In Fig.2, we plot the marginal distribution P(M|x0,t)P(M|x_{0},t) for different tt, but with fixed x0=1x_{0}=1 and D=1/2D=1/2. For comparison, we have performed simulations that supports our theory. In all simulations, we have used the time step dt=105dt=10^{-5} and the results are averaged over 10510^{5} different realizations. For tt much less than the diffusion time td=x02/Dt_{d}=x_{0}^{2}/D from the starting position x0x_{0} to the reflective wall at the origin, the diffusing particle is almost unaffected by reflective boundary such that the distribution P(M|x0,t)P(M|x_{0},t) is similar that of freely diffusive particle shown in Eq.(2), from which we can observe that P(M|x0,t)P(M|x_{0},t) decreases monotonically with MM. As tt increases, the reflective boundary comes into effect, and P(M|x0,t)P(M|x_{0},t) is no longer a monotonic function of MM. Interestingly, P(M|x0,t)P(M|x_{0},t) can show a unique maximum at an intermediate level of MM. In the limit ttdt\gg t_{d}, the summation in the right-hand side of Eq.(31) is dominated by only one term for n=0n=0. In this case, P(M|x0,t)P(M|x_{0},t) decays with MM as M3M^{-3} in the large-MM limit.

Taking the derivative of Eq.(30) with respect to MM, we obtain the marginal distribution P(M|x0,t)P(M|x_{0},t) in the Laplace domain,

P~(M|x0,s)=Q~(x0,s|M)M=αsinh(αM)cosh(αx0)scosh2(αM).\displaystyle\tilde{P}(M|x_{0},s)=\frac{\partial\tilde{Q}(x_{0},s|M)}{\partial M}=\frac{\alpha\sinh(\alpha M)\cosh(\alpha x_{0})}{s\cosh^{2}(\alpha M)}. (32)

It turns out that inverting Eq.(32) is not a trivial task. However, we can conveniently compute the expectation of the maximum displacement MM from P~(M|x0,s)\tilde{P}(M|x_{0},s), defined as

M(x0,t)=x0𝑑MMP(M|x0,t).\displaystyle\langle M(x_{0},t)\rangle=\int_{x_{0}}^{\infty}dMMP(M|x_{0},t). (33)

Performing the Laplace transform for Eq.(33) and utilizing Eq.(32), we have

0𝑑testM(x0,t)=x0𝑑MMP~(M|x0,s)=x0s+2arctan(eαx0)cosh(αx0)αs.\displaystyle\int_{0}^{\infty}dte^{-st}\langle M(x_{0},t)\rangle=\int_{x_{0}}^{\infty}dMM\tilde{P}(M|x_{0},s)=\frac{x_{0}}{s}+\frac{2{\rm{\rm{arctan}}}(e^{-\alpha x_{0}})\cosh(\alpha x_{0})}{\alpha s}. (34)

Prior to the inversion of Eq.(34), we first consider two limiting cases. In the short-time limit (ss\to\infty) and in the long-time limit (s0s\to 0), Eq.(34) becomes

0𝑑testM(x0,t)={x0s+Ds3/2,s,πD2s3/2+πx024Ds,s0.\displaystyle\int_{0}^{\infty}dte^{-st}\langle M(x_{0},t)\rangle=\left\{\begin{array}[]{cc}\frac{{{x_{0}}}}{s}+\frac{{\sqrt{D}}}{{{s^{3/2}}}},&s\to\infty,\\ \frac{{\pi\sqrt{D}}}{{2{s^{3/2}}}}+\frac{{\pi x_{0}^{2}}}{{4\sqrt{Ds}}},&s\to 0.\\ \end{array}\right. (37)

Performing the inverse Laplace transform for Eq.(37), we obtain

M(x0,t)={x0+2Dtπ,ttd,πDt+πx024Dt,ttd.\displaystyle\langle M(x_{0},t)\rangle=\left\{\begin{array}[]{cc}x_{0}+\frac{{2\sqrt{Dt}}}{{\sqrt{\pi}}},&t\ll t_{d},\\ \sqrt{\pi Dt}+\frac{{\sqrt{\pi}x_{0}^{2}}}{{4\sqrt{Dt}}},&t\gg t_{d}.\\ \end{array}\right. (40)

In the short-time limit, ttd=x02Dt\ll t_{d}=\frac{x_{0}^{2}}{D}, the result coincides with the counterpart of the free Brownian motion. In the long-time limit, ttdt\gg t_{d}, the expected maximum is weakly related to the initial position x0x_{0}, and grows with t1/2t^{1/2} as tt increases. Interestingly, the expected maximum under the case is π/2\pi/2 times that of the free Brownian motion starting from the origin.

Refer to caption

Figure 3: Expected value M(x0,t)\langle M(x_{0},t)\rangle of the maximum displacement MM as a function of tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. Solid line and symbols represent the theory and simulation results, respectively. For comparison, we also plot the expectation x(t)\langle x(t)\rangle of the displacement x(t)x(t), as indicated by the dotted line.

To invert Eq.(34) for the general case, we intend to rewrite Eq.(34) as series expansions. To the end, let us denote by y=eαx0y=e^{\alpha x_{0}}, and use the equations arctan(1/y)=n=0(1)n2n+1y(2n+1)\arctan({1/y})=\sum_{n=0}^{\infty}{\frac{{{{\left({-1}\right)}^{n}}}}{{2n+1}}{y^{-\left({2n+1}\right)}}} (y1y\geq 1) and cosh(lny)=1+y22y\cosh(\ln y)=\frac{1+y^{2}}{2y} to obtain

0𝑑tM(x0,t)est\displaystyle\int_{0}^{\infty}dt{\langle{M(x_{0},t)}\rangle{e^{-st}}} =\displaystyle= x0s+1αsn=0(1)n2n+1[e2nαx0+e(2n+2)αx0]\displaystyle\frac{{{x_{0}}}}{s}+\frac{1}{{\alpha s}}\sum\limits_{n=0}^{\infty}{\frac{{{{\left({-1}\right)}^{n}}}}{{2n+1}}}\left[{{e^{-2n\alpha{x_{0}}}}+{e^{-(2n+2)\alpha{x_{0}}}}}\right] (41)
=\displaystyle= x0s+1αs2αsn=1(1)n4n21e2nαx0.\displaystyle\frac{{{x_{0}}}}{s}+\frac{1}{{\alpha s}}-\frac{2}{{\alpha s}}\sum\limits_{n=1}^{\infty}{\frac{{{{\left({-1}\right)}^{n}}}}{{4{n^{2}}-1}}}{e^{-2n\alpha{x_{0}}}}.

Performing the Laplace transform inversion for Eq.(41), we obtain

M(x0,t)=x0+2Dtπ4n=1(1)n4n21[Dtπen2x02Dtnx0erfc(nx0Dt)],\displaystyle\langle{M({{x_{0}},t})}\rangle={x_{0}}+2\sqrt{\frac{{Dt}}{\pi}}-4\sum\limits_{n=1}^{\infty}{\frac{{{{\left({-1}\right)}^{n}}}}{{4{n^{2}}-1}}}\left[{\sqrt{\frac{{Dt}}{\pi}}{e^{-\frac{{{n^{2}}x_{0}^{2}}}{{Dt}}}}-n{x_{0}}{\rm{erfc}}\left({\frac{{n{x_{0}}}}{{\sqrt{Dt}}}}\right)}\right], (42)

where erfc(x){\rm{erfc}}(x) is the complementary error function. We note that the result in Eq.(42) was obtained in a previous work as well (see Eq.(5) in [70]). However, in the present work we additionally obtain the distribution of MM and will further derive the statistics of extremal time in the subsequent sections. In Fig.3, we show a log–log plot of M(x0,t)\langle M(x_{0},t)\rangle as a function of tt, where we have fixed x0=1x_{0}=1 and D=1/2D=1/2. The theoretical prediction is in well agreement with simulations. For comparison, we plot in Fig.3 the expectation x(t)\langle x(t)\rangle of the displacement x(t)x(t) as a function of tt as well. The details of derivation of x(t)\langle x(t)\rangle can be found in appendix A. In the short-time limit, x(t)=x0\langle x(t)\rangle=x_{0} keeps unchanged as a free Brownian motion, and in the long-time limit x(t)t\langle x(t)\rangle\sim\sqrt{t}, but the prefactor is smaller than that of the expected maximum displacement.

VI Joint distribution P(M,tm|x0,t)P(M,t_{m}|x_{0},t)

Let us define P(M,tm|x0,t)P(M,t_{m}|x_{0},t) as the joint probability density function that the displacement x(τ)x(\tau) reaches its maximum MM at time tmt_{m} with a duration tt, providing that the Brownian starts from the position x0x_{0} (>0>0). To compute the joint distribution P(M,tm|x0,t)P(M,t_{m}|x_{0},t), we can decompose the trajectory into two parts: a left-hand segment for which 0<τ<tm0<\tau<t_{m}, and a right-hand segment for which tm<τ<tt_{m}<\tau<t, as shown in Fig.1. The statistical weight of the first segment equals to the propagator G(M,tm|x0)G(M,t_{m}|x_{0}). However, it turns out that G(M,tm|x0)=0G(M,t_{m}|x_{0})=0 which implies that the contribution from this part is zero. To circumvent this problem, we compute G(Mϵ,tm|x0)G(M-\epsilon,t_{m}|x_{0}) and later take the limit ϵ0+\epsilon\to 0^{+} [11]. The statistical weight of the second segment is given by the survival probability Q(Mϵ,ttm|M)Q(M-\epsilon,t-t_{m}|M). Due to the Markov property, the joint probability density P(M,tm|x0,t)P(M,t_{m}|x_{0},t) can be written as the product of the statistical weights of two segments [60],

P(M,tm|x0,t)=limϵ0+𝒩G(Mϵ,tm|x0)Q(Mϵ,ttm|M).\displaystyle P\left({M,{t_{m}}|{x_{0}},t}\right)=\mathop{\lim}\limits_{\epsilon\to 0^{+}}\mathcal{N}G\left({M-\epsilon,{t_{m}}|{x_{0}}}\right)Q\left({M-\epsilon,t-{t_{m}}|M}\right). (43)

Here, the normalization factor 𝒩\mathcal{N} may depend on DD and ϵ\epsilon. In principle, 𝒩\mathcal{N} depends also on x0x_{0} and tt. However, here we assume and verify a posteriori that it is indeed independent of x0x_{0} and tt.

Let us integrate P(M,tm|x0,t)P(M,t_{m}|x_{0},t) over tmt_{m} from 0 to tt and over MM from x0x_{0} to \infty, and then perform the Laplace transform with respect to tt, such that the convolution structure of Eq.(43) can be exploited,

1s=0𝑑testx0𝑑M0t𝑑tmP(M,tm|x0,t)=limϵ0+𝒩x0𝑑MG~(Mϵ,s|x0)Q~(Mϵ,s|M).\displaystyle\frac{1}{s}=\int_{0}^{\infty}dt{{e^{-st}}\int_{{x_{0}}}^{\infty}dM{\int_{0}^{t}dt_{m}P({M,{t_{m}}|{x_{0}},t})}}=\mathop{\lim}\limits_{\epsilon\to 0^{+}}\mathcal{N}\int_{x_{0}}^{\infty}dM\tilde{G}({M-\epsilon,s|{x_{0}}})\tilde{Q}({M-\epsilon,s|M}). (44)

Here, G~(Mϵ,s|x0)\tilde{G}({M-\epsilon,s|{x_{0}}}) and Q~(Mϵ,s|M)\tilde{Q}({M-\epsilon,s|M}) can be obtained from Eq.(26) and Eq.(30). Up to the leading order of ϵ\epsilon, they are

G~(Mϵ,s|x0)=cosh(αx0)Dcosh(αM)ϵ,\displaystyle\tilde{G}({M-\epsilon,{s}|{x_{0}}})=\frac{{\cosh\left({\alpha{x_{0}}}\right)}}{{D\cosh\left({\alpha M}\right)}}\epsilon, (45)

and

Q~(Mϵ,s|M)=αtanh(αM)sϵ,\displaystyle\tilde{Q}\left({M-\epsilon,s|M}\right)=\frac{{\alpha\tanh\left({\alpha M}\right)}}{s}\epsilon, (46)

where α=s/D\alpha=\sqrt{s/D} again. By inserting Eq.(45) and Eq.(46) into Eq.(44) and completing the integral over MM on the right-hand side of Eq.(44), we obtain the normalization factor 𝒩\mathcal{N},

𝒩=Dϵ2.\displaystyle\mathcal{N}=\frac{D}{\epsilon^{2}}. (47)

Moreover, we perform double Laplace transformations for P(M,tm|x0,t)P(M,t_{m}|x_{0},t) with respect to tmt_{m} (s\to s) and tt (p\to p), and use Eq.(45), Eq.(46) and Eq.(47), to obtain

0𝑑tept0t𝑑tmestmP(M,tm|x0,t)=cosh((p+s)/Dx0)cosh((p+s)/DM)tanh(p/DM)pD,\displaystyle\int_{0}^{\infty}{dt}{e^{-pt}}\int_{0}^{t}{d{t_{m}}{e^{-s{t_{m}}}}P\left({M,{t_{m}}|{x_{0}},t}\right)}=\frac{{\cosh\left({\sqrt{\left({p+s}\right)/D}{x_{0}}}\right)}}{{\cosh\left({\sqrt{\left({p+s}\right)/D}M}\right)}}\frac{{\tanh\left({\sqrt{p/D}M}\right)}}{{\sqrt{pD}}}, (48)

where we have exchanged the order of integration over tt and tmt_{m}. By using the inverse Laplace transform relations [77],

st1[cosh(νs1/2)sech(as1/2)]\displaystyle\mathcal{L}_{s\to t}^{-1}\left[{\cosh\left({\nu{s^{-1/2}}}\right){\rm{sech}}\left({a{s^{-1/2}}}\right)}\right] =\displaystyle= 1aνθ1(12νa1|ta2),\displaystyle\frac{1}{a}\frac{\partial}{{\partial\nu}}{\theta_{1}}\left({\frac{1}{2}\nu{a^{-1}}|t{a^{-2}}}\right), (49)
st1[s1/2tanh(as1/2)]\displaystyle\mathcal{L}_{s\to t}^{-1}\left[{{s^{-1/2}}\tanh\left({a{s^{-1/2}}}\right)}\right] =\displaystyle= 1aθ2(0|ta2),\displaystyle\frac{1}{a}{\theta_{2}}\left({0|t{a^{-2}}}\right), (50)

where

θ1(z|t)\displaystyle{\theta_{1}}\left({z|t}\right) =\displaystyle= 2n=0(1)ne(n+1/2)2π2tsin[(2n+1)πz],\displaystyle 2\sum\limits_{n=0}^{\infty}{{{\left({-1}\right)}^{n}}e^{-{{\left({n+1/2}\right)}^{2}}{\pi^{2}}t}}\sin\left[{\left({2n+1}\right)\pi z}\right], (51)
θ2(z|t)\displaystyle{\theta_{2}}\left({z|t}\right) =\displaystyle= 2n=0e(n+1/2)2π2tcos[(2n+1)πz],\displaystyle 2\sum\limits_{n=0}^{\infty}{e^{-{{\left({n+1/2}\right)}^{2}}{\pi^{2}}t}}\cos\left[{\left({2n+1}\right)\pi z}\right], (52)

are the Elliptic theta functions, we obtain P(M,tm|x0,t)P(M,t_{m}|x_{0},t) from Eq.(48),

P(M,tm|x0,t)\displaystyle P\left({M,{t_{m}}|{x_{0}},t}\right) =\displaystyle= 1M2νθ1(12νa1|tma2)θ2(0|(ttm)a2)\displaystyle\frac{1}{M^{2}}\frac{\partial}{{\partial\nu}}{\theta_{1}}\left({\frac{1}{2}\nu{a^{-1}}|{t_{m}}{a^{-2}}}\right){\theta_{2}}\left({0|(t-t_{m}){a^{-2}}}\right) (53)
=\displaystyle= 4πDM3n1=0n2=0(1)n1(n1+12)eπ2DM2[(n1+12)2tm+(n2+12)2(ttm)]cos[(n1+1/2)πx0M]\displaystyle\frac{{4\pi D}}{{{M^{3}}}}\sum\limits_{{n_{1}}=0}^{\infty}{\sum\limits_{{n_{2}}=0}^{\infty}{{{\left({-1}\right)}^{{n_{1}}}}\left({{n_{1}}+\frac{1}{2}}\right)}{e^{-\frac{{{\pi^{2}}D}}{{{M^{2}}}}\left[{{{\left({{n_{1}}+\frac{1}{2}}\right)}^{2}}{t_{m}}+{{\left({{n_{2}}+\frac{1}{2}}\right)}^{2}}\left({t-{t_{m}}}\right)}\right]}}}\cos\left[{\frac{{\left({{n_{1}}+1/2}\right)\pi{x_{0}}}}{M}}\right]

where a=M/Da=M/\sqrt{D} and ν=x0/D\nu=x_{0}/\sqrt{D}. In fact, the joint distribution P(M,tm|x0,t)P(M,t_{m}|x_{0},t) in Eq.(53) can be also obtained directedly from the results in the time domain. This can be done more easily by series expansions of G(Mϵ,tm|x0)G(M-\epsilon,t_{m}|x_{0}) and Q(Mϵ,ttm|M)Q(M-\epsilon,t-t_{m}|M) to the first order in terms of Eq.(8) and Eq.(27), and then inserting them into Eq.(43).

VII Marginal distribution P(tm|x0,t)P(t_{m}|x_{0},t)

Integrating P(M,tm|x0,t)P(M,t_{m}|x_{0},t) in Eq.(53) over MM from x0x_{0} to \infty, we get to the marginal distribution P(tm|x0,t)P(t_{m}|x_{0},t) of tmt_{m} at which the maximum displacement MM is achieved,

P(tm|x0,t)\displaystyle P\left({{t_{m}}|{x_{0}},t}\right) =\displaystyle= x0𝑑MP(M,tm|x0,t)\displaystyle\int_{{x_{0}}}^{\infty}{dMP\left({M,{t_{m}}|{x_{0}},t}\right)} (54)
=\displaystyle= 4πDn1=0n2=0(1)n1(n1+12)F(π2D[(n1+12)2tm+(n2+12)2(ttm)],(n1+12)πx0),\displaystyle 4\pi D\sum\limits_{{n_{1}}=0}^{\infty}{\sum\limits_{{n_{2}}=0}^{\infty}{{{\left({-1}\right)}^{{n_{1}}}}\left({{n_{1}}+\frac{1}{2}}\right)}}F\left({{\pi^{2}}D\left[{{{\left({{n_{1}}+\frac{1}{2}}\right)}^{2}}{t_{m}}+{{\left({{n_{2}}+\frac{1}{2}}\right)}^{2}}\left({t-{t_{m}}}\right)}\right],\left({{n_{1}}+\frac{1}{2}}\right)\pi{x_{0}}}\right),

where

F(a,b)\displaystyle F\left({a,b}\right) =\displaystyle= x0𝑑M1M3eaM2cos(bM)\displaystyle\int_{{x_{0}}}^{\infty}dM\frac{1}{{{M^{3}}}}{e^{-\frac{a}{{{M^{2}}}}}}\cos\left({\frac{b}{M}}\right) (55)
=\displaystyle= 12a12aeax02cos(bx0)πb4a3/2eb24aerfi(b2a)+πb8a3/2eb24a[erfi(2ai+bx02ax0)erfi(2aibx02ax0)].\displaystyle\frac{1}{{2a}}-\frac{1}{{2a}}{e^{-\frac{a}{{x_{0}^{2}}}}}\cos\left({\frac{b}{{{x_{0}}}}}\right)-\frac{{\sqrt{\pi}b}}{{4{a^{3/2}}}}{e^{-\frac{{{b^{2}}}}{{4a}}}}{\rm{erfi}}\left({\frac{b}{{2\sqrt{a}}}}\right)+\frac{{\sqrt{\pi}b}}{{8{a^{3/2}}}}{e^{-\frac{{{b^{2}}}}{{4a}}}}\left[{{\rm{erfi}}\left({\frac{{2a{\rm{i}}+b{x_{0}}}}{{2\sqrt{a}{x_{0}}}}}\right)-{\rm{erfi}}\left({\frac{{2a{\rm{i}}-b{x_{0}}}}{{2\sqrt{a}{x_{0}}}}}\right)}\right].

Here erfi(z)=ierf(iz){\rm{erfi}}(z)=-{\rm{i}}{\rm{erf}}({\rm{i}}z) is the imaginary error function, and i{\rm{i}} is the unit imaginary number.

Refer to caption

Figure 4: The distribution P(tm|x0,t)P(t_{m}|x_{0},t) of tmt_{m} at which the displacement reaches its maximum for different values of tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. Lines and symbols represent the theory and simulation results, respectively.

In Fig.4, we show the distribution P(tm|x0,t)P(t_{m}|x_{0},t) of tmt_{m} for several different values of tt. The results from theory and simulations are in agreement well. For ttdt\ll t_{d}, the particle has a very low probability of diffusing the reflective boundary, such that the particle behaves a free diffusion without any boundaries and P(tm|x0,t)P(t_{m}|x_{0},t) is U-shaped as indicated in Eq.(3). For ttdt\gg t_{d}, the reflective boundary comes into effect and P(tm|x0,t)P(t_{m}|x_{0},t) deviates from the U-shaped distribution, and becomes asymmetric with respect to tm=t/2t_{m}=t/2. There is a higher probability of tmt_{m} that occurs near tt.

Moreover, we seek to find the marginal distribution P(tm|x0,t)P(t_{m}|x_{0},t) in the Laplace domain. To the end, we perform double Laplace transforms for P(tm|x0,t)P(t_{m}|x_{0},t), and use Eq.(45), Eq.(46) and Eq.(47) to obtain

0𝑑tept0t𝑑tmestmP(tm|x0,t)=x0𝑑Mcosh((p+s)/Dx0)cosh((p+s)/DM)tanh(p/DM)pD.\displaystyle\int_{0}^{\infty}{dt}{e^{-pt}}\int_{0}^{t}{d{t_{m}}{e^{-s{t_{m}}}}P\left({{t_{m}}|{x_{0}},t}\right)}=\int_{{x_{0}}}^{\infty}{dM\frac{{\cosh\left({\sqrt{\left({p+s}\right)/D}{x_{0}}}\right)}}{{\cosh\left({\sqrt{\left({p+s}\right)/D}M}\right)}}\frac{{\tanh\left({\sqrt{p/D}M}\right)}}{{\sqrt{pD}}}}. (56)

In the short-time limit (pp\to\infty), using cosh(x)ex/2\cosh(x)\sim e^{x}/2 and tanh(x)1\tanh(x)\sim 1 (xx\to\infty), one easily recovers to the result in Eq.(3) from Eq.(56). Furthermore, one observes that Eq.(56) is not invariant under the exchange of p+spp+s\leftrightarrow p. This implies that P(tm|x0,t)P(t_{m}|x_{0},t) is no longer symmetric with respect to t/2t/2 in the presence of the reflective boundary.

Taking the derivative of Eq.(56) with respect to ss and letting s0s\to 0, we obtain the Laplace transformation of the expectation of tmt_{m},

0𝑑tepttm(x0,t)\displaystyle\int_{0}^{\infty}{dt{e^{-pt}}}\langle{{t_{m}}(x_{0},t)}\rangle =\displaystyle= x0𝑑Ms[cosh((p+s)/Dx0)cosh((p+s)/DM)]s0tanh(p/DM)pD\displaystyle-\int_{{x_{0}}}^{\infty}{dM\frac{\partial}{{\partial s}}{{\left[{\frac{{\cosh\left({\sqrt{\left({p+s}\right)/D}{x_{0}}}\right)}}{{\cosh\left({\sqrt{\left({p+s}\right)/D}M}\right)}}}\right]}_{s\to 0}}\frac{{\tanh\left({\sqrt{p/D}M}\right)}}{{\sqrt{pD}}}} (57)
=\displaystyle= 12pDx0𝑑Msinh(p/DM)cosh2(p/DM)[Mcosh(p/Dx0)tanh(p/DM)x0sinh(p/Dx0)]\displaystyle\frac{1}{{2pD}}\int_{{x_{0}}}^{\infty}{dM\frac{{\sinh\left({\sqrt{p/D}M}\right)}}{{{{\cosh}^{2}}\left({\sqrt{p/D}M}\right)}}\left[{M\cosh\left({\sqrt{p/D}{x_{0}}}\right)\tanh\left({\sqrt{p/D}M}\right)-{x_{0}}\sinh\left({\sqrt{p/D}{x_{0}}}\right)}\right]}
=\displaystyle= 14p2x0cosh(p/Dx0)4p3Darg(ep/Dx0iep/Dx0+i)+cosh(p/Dx0)ep/Dx08p2\displaystyle\frac{1}{{4{p^{2}}}}-\frac{{{x_{0}}\cosh\left(\sqrt{p/D}x_{0}\right)}}{{4\sqrt{{p^{3}}D}}}{\rm{arg}}\left({\frac{{{e^{\sqrt{p/D}x_{0}}}-{\rm{i}}}}{{{e^{\sqrt{p/D}x_{0}}}+{\rm{i}}}}}\right)+\frac{{\cosh\left(\sqrt{p/D}x_{0}\right){e^{-\sqrt{p/D}x_{0}}}}}{{8{p^{2}}}}
×Φ(e2p/Dx0,2,12)x04p3Dtanh(p/Dx0),\displaystyle\times\Phi\left({-{e^{-2\sqrt{p/D}x_{0}}},2,\frac{1}{2}}\right)-\frac{{{x_{0}}}}{{4\sqrt{{p^{3}}D}}}\tanh\left(\sqrt{p/D}x_{0}\right),

where arg(z){\rm{arg}}(z) denotes the argument of a complex number zz, i{\rm{i}} is the unit imaginary number, and

Φ(z,2,12)=k=0zk(k+1/2)2,\displaystyle\Phi\left({z,2,\frac{1}{2}}\right)=\sum\limits_{k=0}^{\infty}{\frac{{{z^{k}}}}{{{{\left({k+1/2}\right)}^{2}}}}}, (58)

is the Lerch transcendent function. It turns out that inverting Eq.(57) with respect to pp is a highly nontrivial task. However, in the short-time limit (pp\to\infty) and in the long-time limit (p0p\to 0), Eq.(57) can be simplified to

0𝑑tepttm={12p2,p,1+2G4p2,p0,\displaystyle\int_{0}^{\infty}{dt{e^{-pt}}}\langle{{t_{m}}}\rangle=\left\{\begin{array}[]{cc}\frac{1}{2p^{2}},&p\to\infty,\\ \frac{{1+2G}}{{4{p^{2}}}},&p\to 0,\\ \end{array}\right. (61)

where G=k=0(1)k/(2k+1)20.916G=\sum_{k=0}^{\infty}(-1)^{k}/(2k+1)^{2}\approx 0.916 is the Catalan’s constant. Inverting Eq.(61) with respect to pp, we obtain

tmt={12,ttd,1+2G4,ttd.\displaystyle\frac{{\langle{{t_{m}}}\rangle}}{t}=\left\{\begin{array}[]{cc}\frac{1}{2},&t\ll t_{d},\\ \frac{{1+2G}}{4},&t\gg t_{d}.\\ \end{array}\right. (64)

Refer to caption

Figure 5: tm(x0,t)/t{\langle{t_{m}(x_{0},t)}\rangle}/t as a function of tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. In the short-time limit, ttd=x02Dt\ll t_{d}=\frac{x_{0}^{2}}{D}, tm/t12{\langle{t_{m}}\rangle}/t\to\frac{1}{2}, and in the long-time limit, ttdt\gg t_{d}, tm/t1+2G40.708{\langle{t_{m}}\rangle}/t\to\frac{{1+2G}}{4}\approx 0.708, where G0.916G\approx 0.916 is the Catalan’s constant.

Coincidentally, we realize that the Catalan’s constant GG appears also in the expression of the probability of fractional Brownian motion to first exit at the upper boundary in an interval [74]. However, there is no a direct connection between them at present. In Fig.5, we plot tm/t\langle{t_{m}}\rangle/t as a function of tt for the fixed x0=1x_{0}=1 and D=1/2D=1/2, from which we can see that the ratio tm/t\langle{t_{m}}\rangle/t increases monotonically with tt. tm/t\langle{t_{m}}\rangle/t approaches 1/21/2 in the short-time limit, and tends towards 1+2G40.708\frac{{1+2G}}{4}\approx 0.708 in the long-time limit, which are consistent with our theoretical predictions in Eq.(64).

VIII The statistics of the last time tt_{\ell} the process visits the starting position x0x_{0} and the occupation time tot_{o} spent above x0x_{0}

As mentioned in the Introduction, for a free Brownian motion the last time tt_{\ell} that the process visits the starting position and the occupation time tot_{o} spent above the starting position over a duration tt are also distributed by Eq.(3), called the other two arcsine laws. In this section, we will show how the two arcsine laws are modified in the presence of a reflective wall at the origin.

VIII.1 The statistics of tt_{\ell}

Using the path decomposition technique as done in Sec.V, we can obtain the distribution of the last time tt_{\ell} when a stochastic path visits the starting position x0x_{0}. We decompose the path into two parts: (i) the process starts from x0x_{0}(>0>0) at t=0t=0 and returns to the starting position at tt_{\ell}. The probability density is given by the propagator G0(x0,t|x0)G_{0}(x_{0},t_{\ell}|x_{0}) (here the propagator will be derived only in the presence of a reflective boundary, different from Eq.(8) where a reflective boundary and an absorbing boundary are both present); (ii) In the time interval from tt_{\ell} to tt, the process never returns to the starting position. The probability is given by survival probability Q(x0,tt|x0)Q(x_{0},t-t_{\ell}|x_{0}). Since the process is Markovian, the two parts are statistically independent. Since Q(x0,tt|x0)=0Q(x_{0},t-t_{\ell}|x_{0})=0 and thus the probability of the second part is equal to zero. To overcome this difficulty, we consider the probability as Q(x0+ϵ,tt|x0)+Q(x0ϵ,tt|x0)Q(x_{0}+\epsilon,t-t_{\ell}|x_{0})+Q(x_{0}-\epsilon,t-t_{\ell}|x_{0}) and then take the ϵ0+\epsilon\to 0^{+} limit. Thus, the probability density function of tt_{\ell} can be expressed as

P(t|t)=limϵ0+𝒩G0(x0,t|x0)[Q(x0+ϵ,tt|x0)+Q(x0ϵ,tt|x0)].\displaystyle P(t_{\ell}|t)=\mathop{\lim}\limits_{\epsilon\to 0^{+}}\mathcal{N}G_{0}(x_{0},t_{\ell}|x_{0})\left[Q(x_{0}+\epsilon,t-t_{\ell}|x_{0})+Q(x_{0}-\epsilon,t-t_{\ell}|x_{0})\right]. (65)

The propagator G0(x,t|x0)G_{0}(x,t|x_{0}) in the presence of a reflective wall at x=0x=0 can be easily obtain from the method of image, given by

G0(x,t|x0)=14πDt[e(xx0)24Dt+e(x+x0)24Dt].\displaystyle G_{0}(x,t|x_{0})=\frac{1}{\sqrt{4\pi Dt}}\left[e^{-\frac{(x-x_{0})^{2}}{4Dt}}+e^{-\frac{(x+x_{0})^{2}}{4Dt}}\right]. (66)

The survival probability Q(x0,t|M)Q(x_{0},t|M) should be considered separately for the case M<x0M<x_{0} and for the case M>x0M>x_{0}. When the absorbing boundary MM is below the starting position x0x_{0}, M<x0M<x_{0}, the reflective boundary at the origin does not produce any effect, and thus the survival probability Q(x0,t|M)Q(x_{0},t|M) is a well-known result [76],

Q(x0,t|M<x0)=erf(x0M2Dt).\displaystyle Q(x_{0},t|M<x_{0})={\rm{erf}}\left(\frac{x_{0}-M}{2\sqrt{Dt}}\right). (67)

For M>x0M>x_{0}, the role of the reflective wall at x=0x=0 needs to take into account. In fact, under this case the survival probability Q(x0,t|M>x0)Q(x_{0},t|M>x_{0}) has been obtained in Sec.IV, see Eq.(27).

It is useful to perform the double Laplace transformations for P(t|t)P(t_{\ell}|t) with respect to tt_{\ell} (s\to s) and with respect to tt (p\to p),

0𝑑tept0t𝑑testP(t|t)=limϵ0+𝒩G~0(x0,p+s|x0)[Q~(x0+ϵ,p|x0)+Q~(x0ϵ,p|x0)],\displaystyle\int_{0}^{\infty}dte^{-pt}\int_{0}^{t}dt_{\ell}e^{-st_{\ell}}P(t_{\ell}|t)=\mathop{\lim}\limits_{\epsilon\to 0^{+}}\mathcal{N}\tilde{G}_{0}(x_{0},p+s|x_{0})\left[\tilde{Q}(x_{0}+\epsilon,p|x_{0})+\tilde{Q}(x_{0}-\epsilon,p|x_{0})\right], (68)

where we have exchanged the orders of the integrals over tt and tt_{\ell}, and made the variable substitution t=ttt^{\prime}=t-t_{\ell}. In principle, 𝒩\mathcal{N} can also depend on tt. However, here we assume and verify a posteriori that it is indeed independent of tt. G~0(x0,p|x0)\tilde{G}_{0}(x_{0},p|x_{0}) and Q~(x0+ϵ,p|x0)\tilde{Q}(x_{0}+\epsilon,p|x_{0}) can be obtained from Eq.(66) and Eq.(67), given by

G~(x0,p|x0)=1+e2p/Dx02Dp,\displaystyle\tilde{G}(x_{0},p|x_{0})=\frac{{1+{e^{-2\sqrt{p/D}{x_{0}}}}}}{{2\sqrt{Dp}}}, (69)

and

Q~(x0+ϵ,p|x0)=ϵDp+o(ϵ2).\displaystyle\tilde{Q}(x_{0}+\epsilon,p|x_{0})=\frac{\epsilon}{\sqrt{Dp}}+o(\epsilon^{2}). (70)

Q~(x0ϵ,p|x0)\tilde{Q}(x_{0}-\epsilon,p|x_{0}) can be obtained from Eq.(30),

Q~(x0ϵ,p|x0)=tanh(p/Dx0)Dpϵ+o(ϵ2)\displaystyle\tilde{Q}(x_{0}-\epsilon,p|x_{0})=\frac{{\tanh\left({\sqrt{p/D}{x_{0}}}\right)}}{{\sqrt{Dp}}}\epsilon+o(\epsilon^{2}) (71)

Letting s0s\to 0, the left-hand side of Eq.(68) is just equal to 1/p1/p due to the normalization condition, 0t𝑑tP(t|t)=1\int_{0}^{t}dt_{\ell}P(t_{\ell}|t)=1. We substitute Eq.(69), Eq.(70) and Eq.(71) into the right-hand side of Eq.(68), and then compare both sides of Eq.(68), which yields

𝒩=Dϵ.\displaystyle\mathcal{N}=\frac{D}{\epsilon}. (72)

In the time domain, we have

G(x0,t|x0)=1+ex02Dt4πDt,\displaystyle G(x_{0},t_{\ell}|x_{0})=\frac{{1+{e^{-\frac{{x_{0}^{2}}}{{D{t_{\ell}}}}}}}}{{\sqrt{4\pi D{t_{\ell}}}}}, (73)
Q(x0+ϵ,tt|x0)=ϵπD(tt)+o(ϵ2),\displaystyle Q(x_{0}+\epsilon,t-t_{\ell}|x_{0})={\frac{\epsilon}{{\sqrt{\pi D\left({t-{t_{\ell}}}\right)}}}}+o(\epsilon^{2}), (74)

and

Q(x0ϵ,tt|x0)=ϵx0θ2(0|D(tt)x02)+o(ϵ2)=2ϵx0n=0e(n+1/2)2π2D(tt)x02+o(ϵ2),\displaystyle Q(x_{0}-\epsilon,t-t_{\ell}|x_{0})=\frac{\epsilon}{{{x_{0}}}}{\theta_{2}}\left({0\left|{\frac{{D\left({t-{t_{\ell}}}\right)}}{{x_{0}^{2}}}}\right.}\right)+o(\epsilon^{2})=\frac{2\epsilon}{{{x_{0}}}}\sum\limits_{n=0}^{\infty}{{e^{-\frac{{{{\left({n+1/2}\right)}^{2}}{\pi^{2}}D\left({t-{t_{\ell}}}\right)}}{{x_{0}^{2}}}}}}+o(\epsilon^{2}), (75)

where θ2(z|t){\theta_{2}}({z|t}) is the Elliptic theta function defined in Eq.(51). Therefore, the distribution of tt_{\ell} can be obtained by inserting Eqs.(72), (73), (74) and (75) into Eq.(65),

P(t|t)=D4π1+ex02Dtt[1πD(tt)+1x0θ2(0|D(tt)x02)].\displaystyle P(t_{\ell}|t)=\sqrt{\frac{D}{{4\pi}}}\frac{{1+{e^{-\frac{{x_{0}^{2}}}{{D{t_{\ell}}}}}}}}{{\sqrt{{t_{\ell}}}}}\left[{\frac{1}{{\sqrt{\pi D\left({t-{t_{\ell}}}\right)}}}+\frac{1}{{{x_{0}}}}{\theta_{2}}\left({0\left|{\frac{{D\left({t-{t_{\ell}}}\right)}}{{x_{0}^{2}}}}\right.}\right)}\right]. (76)

By taking the derivative of Eq.(68) with respect to ss and then letting s0s\to 0, we obtain the Laplace transform of the expectation of tt_{\ell}

0𝑑teptt(x0,t)\displaystyle\int_{0}^{\infty}dte^{-pt}\langle t_{\ell}(x_{0},t)\rangle =\displaystyle= 12p2+x0Dp3(1+e2p/Dx0)\displaystyle\frac{1}{{2{p^{2}}}}+\frac{{{x_{0}}}}{{\sqrt{D{p^{3}}}\left({1+{e^{2\sqrt{p/D}{x_{0}}}}}\right)}} (77)
=\displaystyle= 12p2+x0Dp3n=0(1)ne2(n+1)p/Dx0.\displaystyle\frac{1}{{2{p^{2}}}}+\frac{{{x_{0}}}}{{\sqrt{D{p^{3}}}}}\sum\limits_{n=0}^{\infty}{{{\left({-1}\right)}^{n}}{e^{2\left({n+1}\right)\sqrt{p/D}{x_{0}}}}}.

The series representation in the last line of Eq.(77) is to seek conveniently the inverse Laplace transform with respect to pp. For this purpose, we have

t(x0,t)=t2+2x0Dn=0(1)n[Dtπe(n+1)2x02Dt(n+1)x0erfc((n+1)x0Dt)],\displaystyle{\langle t_{\ell}(x_{0},t)\rangle}=\frac{t}{2}+\frac{{2{x_{0}}}}{D}\sum\limits_{n=0}^{\infty}{{{\left({-1}\right)}^{n}}}\left[{\sqrt{\frac{{Dt}}{\pi}}{e^{-\frac{{{{\left({n+1}\right)}^{2}}x_{0}^{2}}}{{Dt}}}}-\left({n+1}\right){x_{0}}{\rm{erfc}}\left({\frac{{\left({n+1}\right){x_{0}}}}{{\sqrt{Dt}}}}\right)}\right], (78)

where erfc(x){\rm{erfc}}(x) is the complementary error function. In Fig.6, we plot t/t\langle t_{\ell}\rangle/t as a function of tt and compared it with the numerical simulations to find an excellent match. Quite interestingly, t/t\langle t_{\ell}\rangle/t exhibits a non-monotonic behaviour with respect to tt. In the limits t0t\to 0 and tt\to\infty, t/t1/2\langle t_{\ell}\rangle/t\to 1/2. At an intermediate value of tt, t3.1216x02/Dt\approx 3.1216x_{0}^{2}/D, t/t\langle t_{\ell}\rangle/t attains its maximum, t/t0.659\langle t_{\ell}\rangle/t\approx 0.659.

Refer to caption

Figure 6: t(x0,t)/t{\langle{t_{\ell}(x_{0},t)}\rangle}/t as a function of tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. Line and symbols correspond to the theoretical and simulation results, respectively.

VIII.2 The statistics of tot_{o}

Let us denote by tot_{o} the occupation time the Brownian process x(τ)x(\tau) spent above the position x0x_{0} during time tt, providing that the process starts from the position y0y_{0} in the presence of a reflective boundary at the origin. The occupation time tot_{o} is conveniently measured by the following the Brownian functional,

to=0tΘ(x(τ)x0)𝑑τ,\displaystyle t_{o}=\int_{0}^{t}\Theta(x(\tau)-x_{0})d\tau, (79)

where Θ(x)\Theta(x) is the Heaviside step function as defined before. Clearly, tot_{o} is a random variable taking different values for different Brownian paths. Let us define P(to|t)P(t_{o}|t) as the probability density function of tot_{o}. Since tot_{o} has only positive support, a natural step is to introduce the Laplace transform of P(to|t)P(t_{o}|t),

Q(y0,s,t)=0testoP(to|t)es0tΘ(x(τ)x0)𝑑τ,\displaystyle Q(y_{0},s,t)=\int_{0}^{t}e^{-st_{o}}P(t_{o}|t)\equiv\langle e^{-s\int_{0}^{t}\Theta(x(\tau)-x_{0})d\tau}\rangle, (80)

where \langle\cdot\rangle denotes the expectation with respect to the process x(τ)x(\tau) with started at y0y_{0}. Q(y0,s,t)Q(y_{0},s,t) evolves according to the backward equation [34],

Q(y0,s,t)t=D2Q(y0,s,t)y02sΘ(y0x0)Q(y0,s,t).\displaystyle\frac{\partial Q(y_{0},s,t)}{\partial t}=D\frac{\partial^{2}Q(y_{0},s,t)}{\partial y_{0}^{2}}-s\Theta(y_{0}-x_{0})Q(y_{0},s,t). (81)

Performing a further Laplace transform for Q(y0,s,t)Q(y_{0},s,t) with respect to tt, Q~(y0,s,p)=0𝑑teptQ(y0,s,t)\tilde{Q}(y_{0},s,p)=\int_{0}^{\infty}dte^{-pt}Q(y_{0},s,t), Eq(81) becomes

Dd2Q~(y0,s,p)dy02[sΘ(y0x0)+p]Q~(y0,s,p)=1.\displaystyle D\frac{d^{2}\tilde{Q}(y_{0},s,p)}{dy_{0}^{2}}-\left[s\Theta(y_{0}-x_{0})+p\right]\tilde{Q}(y_{0},s,p)=-1. (82)

Eq.(82) can be solved for the regions y0<x0y_{0}<x_{0} and y0>x0y_{0}>x_{0}, separately, which yields,

Q~(y0,s,p)={1p+A1ep/Dy0+B1ep/Dy0,y0x0,1p+s+A2e(p+s)/Dy0+B2e(p+s)/Dy0,y0>x0,\displaystyle\tilde{Q}(y_{0},s,p)=\left\{\begin{array}[]{ll}\frac{1}{p}+A_{1}e^{\sqrt{p/D}y_{0}}+B_{1}e^{-\sqrt{p/D}y_{0}},&y_{0}\leq x_{0},\\ \frac{1}{p+s}+A_{2}e^{\sqrt{(p+s)/D}y_{0}}+B_{2}e^{-\sqrt{(p+s)/D}y_{0}},&y_{0}>x_{0},\\ \end{array}\right. (85)

where four unknown coefficients A1A_{1}, B1B_{1}, A2A_{2} and B2B_{2} can be determined as follows. If y0y_{0}\to\infty, the particle will spend all its time above the position x0x_{0}, such that P(to|t)=δ(tot)P(t_{o}|t)=\delta(t_{o}-t) and thus Q~(y0,s,p)=1/(p+s)\tilde{Q}(y_{0},s,p)=1/(p+s). It is clear that this boundary condition leads to A2=0A_{2}=0. The zero-flux boundary condition at the origin requires y0Q~(y0,s,p)|y0=0=0\partial_{y_{0}}\tilde{Q}(y_{0},s,p)|_{y_{0}=0}=0, which leads to A1=B1A_{1}=B_{1}. The matching conditions at y0=x0y_{0}=x_{0} are Q~(x0+,s,p)=Q~(x0,s,p)\tilde{Q}(x_{0}^{+},s,p)=\tilde{Q}(x_{0}^{-},s,p) and y0Q~(y0,s,p)|y0=x0+=y0Q~(y0,s,p)|y0=x0\partial_{y_{0}}\tilde{Q}(y_{0},s,p)|_{y_{0}=x_{0}^{+}}=\partial_{y_{0}}\tilde{Q}(y_{0},s,p)|_{y_{0}=x_{0}^{-}}, which yields

{A1=s2pp+s[p+scosh(p/Dx0)+psinh(p/Dx0)]B2=ssinh(p/Dx0)[sinh((p+s)/Dx0)+cosh((p+s)/Dx0)]p(p+s)[p+scosh(p/Dx0)+psinh(p/Dx0)]\displaystyle\left\{\begin{array}[]{l}{A_{1}}=-\frac{s}{{2p\sqrt{p+s}\left[{\sqrt{p+s}\cosh\left({\sqrt{p/D}{x_{0}}}\right)+\sqrt{p}\sinh\left({\sqrt{p/D}{x_{0}}}\right)}\right]}}\\ {B_{2}}=\frac{{s\sinh\left({\sqrt{p/D}{x_{0}}}\right)\left[{\sinh\left({\sqrt{\left({p+s}\right)/D}{x_{0}}}\right)+\cosh\left({\sqrt{\left({p+s}\right)/D}{x_{0}}}\right)}\right]}}{{\sqrt{p}\left({p+s}\right)\left[{\sqrt{p+s}\cosh\left({\sqrt{p/D}{x_{0}}}\right)+\sqrt{p}\sinh\left({\sqrt{p/D}{x_{0}}}\right)}\right]}}\end{array}\right. (88)

Refer to caption

Figure 7: The fraction of occupation time spent above the starting position x0x_{0}, to(x0,t)/t{\langle{t_{o}(x_{0},t)}\rangle}/t, as a function of tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. Line and symbols correspond to the theoretical and simulation results, respectively.

For the case we are interested in, y0=x0y_{0}=x_{0}, we have

Q~(x0,s,p)=pcosh(p/Dx0)+p+ssinh(p/Dx0)p(p+s)[p+scosh(p/Dx0)+psinh(p/Dx0)].\displaystyle\tilde{Q}(x_{0},s,p)=\frac{{\sqrt{p}\cosh\left({\sqrt{p/D}{x_{0}}}\right)+\sqrt{p+s}\sinh\left({\sqrt{p/D}{x_{0}}}\right)}}{{\sqrt{p\left({p+s}\right)}\left[{\sqrt{p+s}\cosh\left({\sqrt{p/D}{x_{0}}}\right)+\sqrt{p}\sinh\left({\sqrt{p/D}{x_{0}}}\right)}\right]}}. (89)

For x0x_{0}\to\infty, i.e., when the starting point is far away from the reflective boundary, Eq.(89) reduces to Q~(x0,s,p)=1/p(p+s)\tilde{Q}(x_{0}\to\infty,s,p)=1/\sqrt{p(p+s)}, and recovers to the classical arcsine law for the occupation time.

Taking the derivative of Q~(x0,s,p)\tilde{Q}(x_{0},s,p) with respect to ss and then letting s0s\to 0, we obtain the Laplace transform of the mean occupation time,

0𝑑teptto(x0,t)=Q~(x0,s,p)s|s0=cosh(p/Dx0)p2[cosh(p/Dx0)+sinh(p/Dx0)].\displaystyle\int_{0}^{\infty}{dt{e^{-pt}}\langle{{t_{o}}\left({{x_{0}},t}\right)}\rangle}=-{\left.{\frac{{\partial\tilde{Q}\left({{x_{0}},s,p}\right)}}{{\partial s}}}\right|_{s\to 0}}=\frac{{\cosh\left({\sqrt{p/D}{x_{0}}}\right)}}{{{p^{2}}\left[{\cosh\left({\sqrt{p/D}{x_{0}}}\right)+\sinh\left({\sqrt{p/D}{x_{0}}}\right)}\right]}}. (90)

The inversion of Eq.(90) to obtain the mean occupation time,

to(x0,t)=12[t4x02tπDex02Dt+(t+2x02D)erfc(x0Dt)],\displaystyle\langle{{t_{o}}\left({{x_{0}},t}\right)}\rangle=\frac{1}{2}\left[{t-\sqrt{\frac{{4x_{0}^{2}t}}{{\pi D}}}{e^{-\frac{{x_{0}^{2}}}{{Dt}}}}+\left({t+\frac{{2x_{0}^{2}}}{D}}\right){\rm{erfc}}\left({\frac{{{x_{0}}}}{{\sqrt{Dt}}}}\right)}\right], (91)

where erfc(x){\rm{erfc}}(x) is the complementary error function. In Fig.7, we show the fraction of occupation time spent above the starting position x0x_{0}, to(x0,t)/t{\langle{t_{o}(x_{0},t)}\rangle}/t, as a function of tt, where x0=1x_{0}=1 and D=1/2D=1/2 are fixed. The simulation result agrees well with the theoretical formula in Eq.(91). In the short-time limit, ttdt\ll t_{d}, the reflective boundary does not come into effect, and thus to/t1/2{\langle{t_{o}}\rangle}/t\to 1/2 as predicted by the arcsine law. As tt increases, to/t{\langle{t_{o}}\rangle}/t increases monotonically, and tends to 11 in the long-time limit ttdt\gg t_{d}. This is because that under the influence of the reflective wall the process has a higher probability of staying above the starting position x0x_{0} for a longer time tt. The probability can be quantified by integrating the propagator in Eq.(66) over position from x0x_{0} to infinity, x0𝑑xG(x,t|x0)=112erf(x0/Dt)\int_{x_{0}}^{\infty}dxG(x,t|x_{0})=1-\frac{1}{2}{\rm{erf}}(x_{0}/\sqrt{Dt}), from which one can see that the probability is an increasing function of tt.

IX Conclusions

We have shown how the presence of a reflective boundary affects the extremal statistics of a one-dimensional Brownian motion over a fixed duration tt. We analytical obtain the joint and marginal distributions of the maximum displacement MM and the time tmt_{m} at which the maximum MM is reached. As expected, in the short-time limit (ttdt\ll t_{d}), the reflective wall does not produce much impact on the statistics of MM and tmt_{m}, and thus our model is similar to the free Brownian motion. When tt becomes comparable with tdt_{d}, the reflective wall initiates a significant effect on the extremal statistics. P(M|x0,t)P(M|x_{0},t) can display a single maximum at an intermediate level of MM, unlike free Brownian motion where P(M|x0,t)P(M|x_{0},t) decays monotonically with MM. In the long-time limit, the expectation M\langle M\rangle exhibits the same scaling with tt as the free free Brownian motion, M(x0,t)t1/2\langle M(x_{0},t)\rangle\sim t^{1/2}, but the prefactor πD\sqrt{\pi D} is π/2\pi/2 times of free Brownian motion starting from the origin. Regarding to the distribution of tmt_{m}, tmt_{m} has a higher probability occurring near tt, and thus becomes asymmetric with respect to tm=t/2t_{m}=t/2. The ratio tm/t\langle t_{m}\rangle/t increases monotonically from 1/2 as tt increases, and approaches another constant (1+2G)/40.708(1+2G)/4\approx 0.708 in the limit of ttdt\gg t_{d}, where GG is the Catalan’s constant. Moreover, we investigate the statistics of the last time tt_{\ell} visited the initial position x0x_{0} and the cumulative time tot_{o} spent above x0x_{0}. We exactly compute the expectations of tt_{\ell} and tot_{o}. Interestingly, t/t\langle t_{\ell}\rangle/t is a nonmonotonic function of tt, and a maximal t/t\langle t_{\ell}\rangle/t occurs at t3.1216x02/Dt\approx 3.1216x_{0}^{2}/D.

It would be worthy to explore two possible extensions of the present study in the future. These include the study of EVS of Brownian motions in high dimensions in which the boundaries are set to be reflective and otherwise opening. For example, a planar Brownian motion confined in a wedge geometry with reflective boundaries. Another interesting extension is to investigate the EVS of active Brownian motions in the presence of reflective boundary, such as run-and-tumble motions in one dimension confined in a half-axis by a reflective wall. It would be interesting to explore the effect of nonequilibrium noise on the EVS of Brownian motions.

Acknowledgements.
This work was supported by the National Natural Science Foundation of China (Grants No. 11875069) and the Key Scientific Research Fund of Anhui Provincial Education Department (Grants No. 2023AH050116).

Appendix A The expectation x(t)\langle x(t)\rangle of the displacement x(t)x(t) with a reflective wall

We solve the diffusion equation Eq.(9) in the Laplace domain, subject to a reflective boundary at x=0x=0, G~(0,s|x0)=0{\tilde{G}^{\prime}}({0,s|{x_{0}}})=0. The boundary at x=+x=+\infty requires G~(+,s|x0)=0{\tilde{G}}({+\infty,s|{x_{0}}})=0. On the other hand, the continuous condition in Eq.(19) and the matching condition in Eq.(19) are both satisfied. As done in SecIII.2, the solution is

G~(x,s|x0)=eα(x+x0)+eα(x<x>)2αD\displaystyle\tilde{G}\left({x,s|{x_{0}}}\right)=\frac{{{e^{-\alpha\left({x+{x_{0}}}\right)}}+{e^{\alpha\left({{x_{<}}-{x_{>}}}\right)}}}}{{2\alpha D}} (92)

where α=s/D\alpha=\sqrt{s/D}, x<=min(x,x0)x_{<}=\min(x,x_{0}) and x>=max(x,x0)x_{>}=\max(x,x_{0}). Multiplying Eq.(92) by xx and then integrating over xx from 0 to \infty, we have

x(s)=0𝑑xxG~(x,s|x0)=eαx0+αx0α3D\displaystyle\langle{x(s)}\rangle=\int_{0}^{\infty}{dxx\tilde{G}\left({x,s|{x_{0}}}\right)}=\frac{{{e^{-\alpha{x_{0}}}}+\alpha{x_{0}}}}{{{\alpha^{3}}D}} (93)

By Laplace transform inversion for Eq.(93), we obtain the expectation x(t)\langle x(t)\rangle of the displacement x(t)x(t),

x(t)=2ex024DtDtπ+x0erf(x02Dt)\displaystyle\langle{x(t)}\rangle=2{e^{-\frac{{x_{0}^{2}}}{{4Dt}}}}\sqrt{\frac{{Dt}}{\pi}}+{x_{0}}{\rm{erf}}\left({\frac{{{x_{0}}}}{{2\sqrt{Dt}}}}\right) (94)

where erf(x){\rm{erf}}(x) is the error function. Considering the short-time limit and the long-time limit, we have

x(t)={x0,ttd,2Dtπ+x022πDt,ttd.\displaystyle\langle x(t)\rangle=\left\{\begin{array}[]{cc}x_{0},&t\ll t_{d},\\ 2\sqrt{\frac{{Dt}}{\pi}}+\frac{{x_{0}^{2}}}{{2\sqrt{\pi Dt}}},&t\gg t_{d}.\\ \end{array}\right. (97)

References

  • Fisher and Tippett [1928] R. A. Fisher and L. H. C. Tippett, in Mathematical proceedings of the Cambridge philosophical society (Cambridge University Press, 1928), vol. 24, pp. 180–190.
  • Gumbel [1958] E. J. Gumbel, Statistics of extremes (Columbia university press, 1958).
  • Leadbetter et al. [2012] M. R. Leadbetter, G. Lindgren, and H. Rootzén, Extremes and related properties of random sequences and processes (Springer Science & Business Media, 2012).
  • Bouchaud and Mézard [1997] J.-P. Bouchaud and M. Mézard, J. Phys. A: Math. Theor. 30, 7997 (1997).
  • Davison and Huser [2015] A. C. Davison and R. Huser, Annu. Rev. Stat. Appl 2, 203 (2015).
  • Albeverio et al. [2006] S. Albeverio, V. Jentsch, and H. Kantz, Extreme events in nature and society (Springer Science & Business Media, 2006).
  • Gnedenko [1943] B. Gnedenko, Ann. Math. pp. 423–453 (1943).
  • Fyodorov and Bouchaud [2008] Y. V. Fyodorov and J.-P. Bouchaud, J. Phys. A: Math. Theor. 41, 372001 (2008).
  • Fyodorov [2010] Y. V. Fyodorov, Physica A: Statistical Mechanics and its Applications 389, 4229 (2010).
  • Raychaudhuri et al. [2001] S. Raychaudhuri, M. Cranston, C. Przybyla, and Y. Shapir, Phys. Rev. Lett. 87, 136101 (2001).
  • Majumdar and Comtet [2004] S. N. Majumdar and A. Comtet, Phys. Rev. Lett. 92, 225501 (2004).
  • Bar et al. [2016] A. Bar, S. N. Majumdar, G. Schehr, and D. Mukamel, Phys. Rev. E 93, 052130 (2016).
  • Majumdar [2010a] S. Majumdar, Exact Methods in Low-dimensional Statistical Physics and Quantum Computing: Lecture Notes of the Les Houches Summer School: Volume 89, July 2008 p. 407 (2010a).
  • Guillet et al. [2020] A. Guillet, E. Roldán, and F. Jülicher, New J. Phys. 22, 123038 (2020).
  • Dean and Majumdar [2006] D. S. Dean and S. N. Majumdar, Phys. Rev. Lett. 97, 160201 (2006).
  • Majumdar and Vergassola [2009] S. N. Majumdar and M. Vergassola, Phys. Rev. Lett. 102, 060601 (2009).
  • Majumdar and Schehr [2014] S. N. Majumdar and G. Schehr, J. Stat. Mech.: Theo. and Experiment 2014, P01012 (2014).
  • Dumonteil et al. [2013] E. Dumonteil, S. N. Majumdar, A. Rosso, and A. Zoia, Proceedings of the National Academy of Sciences 110, 4239 (2013).
  • Krapivsky and Majumdar [2000] P. Krapivsky and S. N. Majumdar, Phys. Rev. Lett. 85, 5492 (2000).
  • Majumdar and Krapivsky [2002] S. N. Majumdar and P. L. Krapivsky, Phys. Rev. E 65, 036127 (2002).
  • Majumdar [2003] S. N. Majumdar, Phys. Rev. E 68, 026103 (2003).
  • Majumdar [2010b] S. N. Majumdar, Physica A: Statistical Mechanics and its Applications 389, 4299 (2010b).
  • Schehr and Majumdar [2014] G. Schehr and S. N. Majumdar, in First-passage phenomena and their applications (World Scientific, 2014), pp. 226–251.
  • Lacroix-A-Chez-Toine and Mori [2020] B. Lacroix-A-Chez-Toine and F. Mori, J. Phys. A: Math. Theor. 53, 495002 (2020).
  • Perret et al. [2013] A. Perret, A. Comtet, S. N. Majumdar, and G. Schehr, Phys. Rev. Lett. 111, 240601 (2013).
  • Bénichou et al. [2016] O. Bénichou, P. L. Krapivsky, C. Mejía-Monasterio, and G. Oshanin, Phys. Rev. Lett. 117, 080601 (2016).
  • Zarfaty et al. [2022] L. Zarfaty, E. Barkai, and D. A. Kessler, Phys. Rev. Lett. 129, 094101 (2022).
  • Biroli et al. [2023] M. Biroli, H. Larralde, S. N. Majumdar, and G. Schehr, Phys. Rev. Lett. 130, 207101 (2023).
  • Fortin and Clusel [2015] J.-Y. Fortin and M. Clusel, J. Phys. A: Math. Theor. 48, 183001 (2015).
  • Majumdar et al. [2020] S. N. Majumdar, A. Pal, and G. Schehr, Phys. Rep. 840, 1 (2020).
  • Schehr and Le Doussal [2010] G. Schehr and P. Le Doussal, J. Stat. Mech.: Theo. and Experiment 2010, P01009 (2010).
  • Lévy [1940] P. Lévy, Compos. Math. 7, 283 (1940).
  • Feller [1971] W. Feller, Vols. I & II, Wiley I 968 (1971).
  • Majumdar [2007] S. N. Majumdar, in The Legacy Of Albert Einstein: A Collection of Essays in Celebration of the Year of Physics (World Scientific, 2007), pp. 93–129.
  • Majumdar et al. [2008] S. N. Majumdar, J. Randon-Furling, M. J. Kearney, and M. Yor, J. Phys. A: Math. Theor. 41, 365005 (2008).
  • Majumdar et al. [2010] S. N. Majumdar, A. Rosso, and A. Zoia, J. Phys. A: Math. Theor. 43, 115001 (2010).
  • Boutcheng et al. [2016] H. J. O. Boutcheng, T. B. Bouetou, T. W. Burkhardt, A. Rosso, A. Zoia, and K. T. Crepin, J. Stat. Mech: Theo. and Experiment 2016, 053213 (2016).
  • Singh and Kundu [2019] P. Singh and A. Kundu, J. Stat. Mech.: Theo. and Experiment 2019, 083205 (2019).
  • Hartmann et al. [2020] A. K. Hartmann, S. N. Majumdar, H. Schawe, and G. Schehr, J. Stat. Mech.: Theo. and Experiment 2020, 053401 (2020).
  • Singh et al. [2022a] P. Singh, A. Kundu, S. N. Majumdar, and H. Schawe, J. Phys. A: Math. Theor. 55, 225001 (2022a).
  • Sadhu et al. [2018] T. Sadhu, M. Delorme, and K. J. Wiese, Phys. Rev. Lett. 120, 040603 (2018).
  • Sadhu and Wiese [2021] T. Sadhu and K. J. Wiese, Phys. Rev. E 104, 054112 (2021).
  • Evans and Majumdar [2011] M. R. Evans and S. N. Majumdar, Phys. Rev. Lett. 106, 160601 (2011).
  • Evans et al. [2020] M. R. Evans, S. N. Majumdar, and G. Schehr, J. Phys. A: Math. Theor. 53, 193001 (2020).
  • Wang et al. [2021] W. Wang, A. G. Cherstvy, H. Kantz, R. Metzler, and I. M. Sokolov, Phys. Rev. E 104, 024105 (2021).
  • Vinod et al. [2022] D. Vinod, A. G. Cherstvy, W. Wang, R. Metzler, and I. M. Sokolov, Phys. Rev. E 105, L012106 (2022).
  • Singh and Pal [2021] P. Singh and A. Pal, Phys. Rev. E 103, 052119 (2021).
  • Majumdar et al. [2021] S. N. Majumdar, F. Mori, H. Schawe, and G. Schehr, Phys. Rev. E 103, 022135 (2021).
  • Den Hollander et al. [2019] F. Den Hollander, S. N. Majumdar, J. M. Meylahn, and H. Touchette, J. Phys. A: Math. Theor. 52, 175001 (2019).
  • Lamperti [1958] J. Lamperti, Transactions of the American Mathematical Society 88, 380 (1958).
  • Kasahara [1977] Y. Kasahara, Publications of the Research Institute for Mathematical Sciences 12, 801 (1977).
  • Dhar and Majumdar [1999] A. Dhar and S. N. Majumdar, Phys. Rev. E 59, 6413 (1999).
  • Majumdar and Dean [2002] S. N. Majumdar and D. S. Dean, Phys. Rev. E 66, 041102 (2002).
  • Burov and Barkai [2011] S. Burov and E. Barkai, Phys. Rev. Lett. 107, 170601 (2011).
  • Rambeau and Schehr [2011] J. Rambeau and G. Schehr, Phys. Rev. E 83, 061146 (2011).
  • Singh [2022] P. Singh, Phys. Rev. E 105, 024113 (2022).
  • Akimoto et al. [2020] T. Akimoto, T. Sera, K. Yamato, and K. Yano, Phys. Rev. E 102, 032103 (2020).
  • Mori et al. [2019] F. Mori, S. N. Majumdar, and G. Schehr, Phys. Rev. Lett. 123, 200201 (2019).
  • Mori et al. [2020] F. Mori, S. N. Majumdar, and G. Schehr, Phys. Rev. E 101, 052111 (2020).
  • Randon-Furling and Majumdar [2007] J. Randon-Furling and S. N. Majumdar, J. Stat. Mech.: Theo. and Experiment 2007, P10008 (2007).
  • Krapivsky et al. [2010] P. Krapivsky, S. N. Majumdar, and A. Rosso, J. Phys. A: Math. Theor. 43, 315001 (2010).
  • Singh et al. [2022b] P. Singh, S. Santra, and A. Kundu, J. Phys. A: Math. Theor. 55, 465004 (2022b).
  • Neri et al. [2017] I. Neri, E. Roldán, and F. Jülicher, Phys. Rev. X 7, 011019 (2017).
  • Guo et al. [2023] W. Guo, H. Yan, and H. Chen, Phys. Rev. E 108, 044115 (2023).
  • Klinger et al. [2021] J. Klinger, R. Voituriez, and O. Bénichou, Phys. Rev. E 103, 032107 (2021).
  • Klinger et al. [2022] J. Klinger, A. Barbier-Chebbah, R. Voituriez, and O. Bénichou, Phys. Rev. E 105, 034116 (2022).
  • Randon-Furling et al. [2009] J. Randon-Furling, S. N. Majumdar, and A. Comtet, Phys. Rev. Lett. 103, 140602 (2009).
  • Mori et al. [2021] F. Mori, S. N. Majumdar, and G. Schehr, Europhysics Letters 135, 30003 (2021).
  • Mori et al. [2022] F. Mori, S. N. Majumdar, and G. Schehr, Phys. Rev. E 106, 054110 (2022).
  • Chupeau et al. [2015a] M. Chupeau, O. Bénichou, and S. N. Majumdar, Phys. Rev. E 91, 050104 (2015a).
  • Chupeau et al. [2015b] M. Chupeau, O. Bénichou, and S. N. Majumdar, Phys. Rev. E 92, 022145 (2015b).
  • De Bruyne et al. [2022] B. De Bruyne, O. Bénichou, S. N. Majumdar, and G. Schehr, J. Phys. A: Math. Theor. 55, 144002 (2022).
  • Kay and Giuggioli [2023] T. Kay and L. Giuggioli, J. Phys. A: Math. Theor. 56, 345002 (2023).
  • Wiese [2019] K. J. Wiese, Phys. Rev. E 99, 032106 (2019).
  • Cherstvy et al. [2014] A. G. Cherstvy, A. V. Chechkin, and R. Metzler, J. Phys. A: Math. Theor. 47, 485002 (2014).
  • Redner [2001] S. Redner, A guide to first-passage processes (Cambridge University Press, 2001).
  • Oberhettinger and Badii [2012] F. Oberhettinger and L. Badii, Tables of Laplace transforms (Springer Science & Business Media, 2012).