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

Short-time large deviations of first-passage functionals for high-order stochastic processes

Lulu Tian    Hanshuang Chen [email protected]    Guofeng Li School of Physics and Optoelectronic Engineering, Anhui University, Hefei 230601, China
Abstract

We consider high-order stochastic processes x(t)x(t) described by the Langevin equation dmx(t)dtm=2Dξ(t)\frac{{{d^{m}}x\left(t\right)}}{{d{t^{m}}}}=\sqrt{2D}\xi(t), where ξ(t)\xi(t) is a delta-correlated Gaussian noise with zero mean, and DD is the strength of noise. We focus on the short-time statistics of the first-passage functionals A=0T[x(t)]n𝑑tA=\int_{0}^{T}\left[x(t)\right]^{n}dt along the trajectories starting from x(0)=Lx(0)=L and terminating whenever passing through the origin for the first-time at t=Tt=T. Using the optimal fluctuation method, we analytically obtain the most likely realizations of the first-passage processes for a given constraint AA with n=0n=0 and 1, corresponding to the first-passage time itself and the area swept by the first-passage trajectory, respectively. The tail of the distribution of AA shows an essential singularity at A0A\to 0, Pm,n(A|L)exp(αm,nL2mnn+2DA2m1)P_{m,n}(A|L)\sim\exp\left(-\frac{\alpha_{m,n}L^{2mn-n+2}}{DA^{2m-1}}\right), where the explicit expressions for the exponents αm,0\alpha_{m,0} and αm,1\alpha_{m,1} for arbitrary mm are obtained.

I Introduction

From fluctuation–dissipation relation in equilibrium systems [1] to fluctuation theorem in nonequilibrium systems [2], Brownian motion plays a central role in statistical mechanics and other fields. As one of the simplest stochastic processes, Brownian motion is described by the Langevin equation

dxdt=2Dξ(t),\displaystyle\frac{dx}{dt}=\sqrt{2D}\xi(t), (1)

where ξ(t)\xi(t) is the Gaussian white noise with ξ(t)=0\langle{\xi(t)}\rangle=0 and ξ(t)ξ(t)=δ(tt)\langle{\xi(t)}{\xi(t^{\prime})}\rangle=\delta({t-t^{\prime}}), and DD is the strength of noise. Considering a Brownian trajectory with a fixed duration tt, a functional can be constructed along the trajectory,

A=0tZ[x(τ)]𝑑τ,\displaystyle A=\int_{0}^{t}Z\left[x(\tau)\right]d\tau, (2)

where Z(x)Z(x) can be an arbitrary function of xx. Due to the randomness of Brownian motion, the functional AA is a well-defined random variable, and thus one of main concerns is statistical properties of the functional AA.

Brownian functionals appear across various disciplines including physics, chemistry, biology, computer science and mathematics [3]. Since the celebrated Feynman–Kac formula [4], the statistical properties of the Brownian functionals with a fixed time interval in one dimension have been studied extensively. Two commonly studied Brownian functionals in the literature are the local time (Z(x)=δ(xx)Z(x)=\delta(x-x_{\ell})) and residence or occupation time (Z(x)=𝕀[a,b](x)Z(x)=\mathbb{I}_{\left[a,b\right]}(x), where 𝕀[a,b](x)=1\mathbb{I}_{\left[a,b\right]}(x)=1 for x[a,b]x\in\left[a,b\right] and zero otherwise), which measures the time the particle spent in the vicinity of a given location xx_{\ell} and the time spent a given spatial range [a,b]\left[a,b\right], respectively. These functionals have found applications in chemical reactions, catalytic processes, and fluorescent imaging in molecular biology [5, 6, 7, 8, 9]. For one-dimensional Brownian motions starting from the origin, the cumulative distribution of the residence time spent on a half-axis was known as the celebrated arcsine law of Lévy [10, 11, 3]. Recent studies led to many extensions with different scenarios, such as diffusion under confinement [12, 13, 14, 15], in the presence of a potential landscape [16, 17], heterogeneous diffusion processes [18], random acceleration process [19, 20, 21], fractional Brownian motion [22, 23], run-and-tumble motion [24, 25], resetting stochastic processes [26, 27, 28, 29, 30], and multi-particles systems [31, 32, 33, 34, 35]. Another form of functionals is Z(x)=xnZ(x)=x^{n}, where nn is a real number, and particularly n=1n=1 corresponds to the area swept by the Brownian motion. Area distribution under a Brownian excursion or a Bessel excursion over a fixed time interval were also studied and found diverse applications in physics [36, 37].

In recent years, the study of first-passage Brownian functionals have received growing attention [3]. One main reason is that first-passage problem underlies many important processes in nature, such as diffusion-limited reactions, neural firing, and animal foraging, etc [38, 39]. In this context, the duration is defined as the time of Brownian motion from an initial position to a prescribed position, e.g. the origin, for the first time. Different from the functionals with a fixed time interval, the first-passage time is not fixed but random. In the literature, a commonly employed functionals along a first-passage trajectory is Z(x)=xnZ(x)=x^{n}, see also the definition in Eq.(4), where n=0n=0 simply represents the first-passage time itself, and n=1n=1 corresponds to the area swept by the first-passage trajectory. The first-passage area has been investigated in the context of Brownian motion [40, 41, 42], drifted Brownian motion [43, 44], in the presence of a harmonic potential (Ornstein-Uhlenbeck process) [45, 46] or a logarithmic potential (Bessel process) [47], and stochastic resetting [48, 49, 50].

Although encouraging progress has been made, in most of cases, however, the exact results for the distributions of Brownian functionals or first-passage Brownian functionals are unavailable. Fortunately, one can resort the large deviation theory to obtain desirable results in the long-time or short-time limit [51]. Large deviation theory is beyond the central limit theorem, which quantifies not only the Gaussian fluctuations about the typical value, but also the relative likelihood of atypical fluctuations [52]. The latter is of particular interest since atypical fluctuations can lead to many striking effects on the underlying dynamical systems, such as escape from a metastable state [53], dynamical phase transition [54, 55, 56, 57], and anomalous scaling [58, 59, 60, 61, 62]. In the long-time limit, the large deviations can be determined by the rate function, which is the Legendre-Fenchel transform of the scaled cumulant generating function of time-averaged observables [63]. In the short-time limit (or the weak-noise limit), large deviations can accurately described by the optimal fluctuation method (OFM), which traces back to early literature [64, 65, 66]. This OFM calculation is based on the path integral representations for stochastic systems subject to certain constraints, and then determine, via the variation of action functional, the optimal, i.e., the most likely path which possesses a minimum action. The OFM has found numerous applications in different areas of physics. When applied to Brownian motion, the OFM becomes geometrical optics [67]. Recently, the OFM has been developed and applied in a number of studies of large deviations of Brownian motion with various constraints [68, 51, 69, 42, 58, 62, 70] and other stochastic systems, such as fractional Brownian motion [72, 73] and random acceleration process [74, 75], active Brownian motion [76] and others [71].

In the present work, we consider high-order stochastic processes x(t)x(t), introduced first in the mathematics and statistics literature [77, 78], indexed by an integer m1m\geq 1,

x(m)(t)=2Dξ(t),\displaystyle x^{(m)}(t)=\sqrt{2D}\xi(t), (3)

where the superscript (m)(m) denotes the mmth derivative with respect to time. In particular, for m=1m=1 the process x(t)x(t) corresponds to standard Brownian motion, and for m=2m=2 it represents the random acceleration process. Due the non-Markov properties of high-order processes, exact results for m2m\geq 2 are rare. The persistence properties of the processes for general nn were initiated in Ref.[79], and were calculated approximately for small values of mm in Ref.[80].

We here focus on the short-time large deviations of the first-passage functionals AA of the high-order processes x(t)x(t) with the form (see Fig.1 for an illustration)

A=0T[x(t)]n𝑑t,\displaystyle{A}=\int_{0}^{T}\left[x(t)\right]^{n}dt, (4)

where n0n\geq 0 is an integer, and TT is the first-passage time of the process starting from the initial conditions

x(0)=L,x(0)==x(m1)(0)=0,\displaystyle x(0)=L,\quad x^{\prime}(0)=\cdots=x^{(m-1)}(0)=0, (5)

and terminating whenever hitting the origin, i.e.,

x(T)=0.\displaystyle x(T)=0. (6)

The main concern in the present work is placed on the A0A\to 0 tail of the probability distribution Pm,n(A|L)P_{m,n}(A|L) of the functional AA for a set of given mm, nn and LL. Thanks to the OFM, the large-deviation tail can be obtained by the determination of the optimal path, that is the most likely realization of the process x(t)x(t), conditioned on the specified value of A0A\to 0. The calculation shows that Pm,n(A|L)P_{m,n}(A|L) has an essential singularity at A0A\to 0, Pm,n(A|L)exp(αm,nL2mnn+2DA2m1)P_{m,n}(A|L)\sim\exp\left(-\frac{\alpha_{m,n}L^{2mn-n+2}}{DA^{2m-1}}\right), where the exponent αm,n\alpha_{m,n} is highly nontrivial, depending on the model parameters mm and nn. We obtain analytically the nontrivial exponents αm,n\alpha_{m,n} for n=0,1n=0,1 and for arbitrary mm, see Eqs.(60) and (90), respectively. Our results are consistent with two recent works in Brownian motion (m=1m=1) [42] and random acceleration process (m=2m=2) [74]. Therefore, our work is an extension of the OFM in higher-order processes.

Refer to caption

Figure 1: First-passage trajectory x(t)x(t) for the high-order process with m=3m=3. The process starts from x=L>0x=L>0 and terminates whenever x(t)x(t) passes through the origin at a random time t=Tt=T. First-passage functionals along the stochastic trajectory is defined as A=0Txn(t)𝑑t{A}=\int_{0}^{T}x^{n}(t)dt.

II Optimal fluctuation method for high-order processes

The dimensional analysis suggests the following scaling behavior of Pm,n(A|L)P_{m,n}(A|L),

Pm,n(A|L)=D1/(2m1)Ln+2/(2m1)Fm,n(D1/(2m1)ALn+2/(2m1)),\displaystyle P_{m,n}(A|L)=\frac{{{D^{1/\left({2m-1}\right)}}}}{{{L^{n+2/\left({2m-1}\right)}}}}{F_{m,n}}\left({\frac{{{D^{1/\left({2m-1}\right)}}A}}{{{L^{n+2/\left({2m-1}\right)}}}}}\right), (7)

where Fm,n(z)F_{m,n}(z) is a dimensionless scaling function of the dimensionless argument z=D1/(2m1)A/Ln+2/(2m1)z={D^{1/\left({2m-1}\right)}}A/{L^{n+2/\left({2m-1}\right)}}. To the best of our knowledge, the scaling function Fm,n(z)F_{m,n}(z) is only known for m=1m=1 where the process describes Brownian motion [42], which is a Markov process. While for m2m\geq 2, the dynamics is non-Markovian and the scaling function Fm,n(z)F_{m,n}(z) is unknown at present. On the physical grounds Fm,n(z)F_{m,n}(z) is expected to have a single maximum, of order 1, at z1z\sim 1, which indicates that the probability distribution Pm,n(A|L)P_{m,n}(A|L) is expected to have its maximum, of order D1/(2m1)/Ln+2/(2m1){D^{1/\left({2m-1}\right)}}/{L^{n+2/\left({2m-1}\right)}}, at ALn+2/(2m1)/D1/(2m1)A\sim{L^{n+2/\left({2m-1}\right)}}/{D^{1/\left({2m-1}\right)}}. Instead of attempting to determine the entire scaling function, in the present work we focus on the leading-order of Fm,n(z)F_{m,n}(z) in the limit of z0z\to 0. This asymptotic corresponds to the ALn+2/(2m1)/D1/(2m1)A\ll{L^{n+2/\left({2m-1}\right)}}/{D^{1/\left({2m-1}\right)}} tail of the distribution Pm,n(A|L)P_{m,n}(A|L).

Using the property of the Gaussian noise in the Langevin equation (3), it is clear that the probability of any path {x(t)}\left\{x(t)\right\} with 0tT0\leq t\leq T can be written as [3]

P[{x(t)}]eS,\displaystyle P[\left\{x(t)\right\}]\propto e^{-{S}}, (8)

where S{S} is called the action, given by

S=14D0T[x(m)]2𝑑t.\displaystyle{S}=\frac{1}{4D}\int_{0}^{T}\left[x^{(m)}\right]^{2}dt. (9)

In the presence of constraints, the process is pushed into a large-deviation regime, and the action defined in Eq.(9) becomes very large in the short-time limit (or equivalently in the weak-noise limit D0D\to 0). Therefore, the dominating contribution to the probability distribution comes from the optimal path: a single deterministic trajectory x(t)x^{*}(t) which minimizes the action functional Eq.(9) subject to the specified additional constraints. The minimization procedure leads to the Euler-Lagrange equation for the optimal path. Once the optimal path x(t)x^{*}(t) is determined, one can evaluate the probability distribution of the specific large deviation up to a pre-exponential factor,

lnPS[x(t)],\displaystyle-\ln P\simeq S[x^{*}(t)], (10)

by plugging the optimal path into the action functional S[x(t)]S[x(t)] given by Eq.(9).

We begin with a modified action functionals subject to the specified additional constraint Eq.(4),

𝒮λ=12D0T[12(x(m))2λxn(t)]𝑑t,\displaystyle\mathcal{S}_{\lambda}=\frac{1}{2D}\int_{0}^{T}\left[\frac{1}{2}\left(x^{(m)}\right)^{2}-\lambda x^{n}(t)\right]dt, (11)

where λ\lambda is the Lagrange multiplier that enforces the constraint 𝒜=A\mathcal{A}=A. The minimization of 𝒮λ\mathcal{S}_{\lambda} can be done via the variation of the constrained action functionals, which results in the Euler-Lagrange equation (see appendix A for details)

x(2m)(t)λnxn1(t)=0.\displaystyle{x^{\left(2m\right)}}\left(t\right)-\lambda n{x^{n-1}}\left(t\right)=0. (12)

To solve Eq.(12), the boundary conditions Eq.(5) and Eq.(6) must be taken into account. However, these boundary conditions are not enough. The others can be obtained from the variation of action functionals (see appendix A for the derivation),

x(m)(T)==x(2m2)(T)=0.\displaystyle\quad x^{(m)}(T)=\cdots=x^{(2m-2)}(T)=0. (13)

In addition, the action should be minimized with respect to TT, which leads to the optimal value of the first-passage time TT_{*}. As shown in the appendix A, this additional minimization brings about an extra condition

x(2m1)(T)=0.\displaystyle x^{(2m-1)}(T_{*})=0. (14)

Once the optimal path x(t)x^{*}(t) and the optimal first-passage time TT_{*} are found, we can express the Lagrange multiplier λ\lambda in terms of AA from Eq.(4). Then, using the action defined in Eq.(9) we obtain, up to a preexponential factor, the A0A\to 0 tail of Pm,n(A|L)P_{m,n}(A|L). It scales as

lnPm,n(A|L)S[x(t)]=αm,nL2mnn+2DA2m1.\displaystyle-\ln P_{m,n}(A|L)\simeq S\left[x^{*}(t)\right]=\frac{\alpha_{m,n}L^{2mn-n+2}}{DA^{2m-1}}. (15)

Equation (15) indicates that the distribution Pm,n(A|L)P_{m,n}(A|L) has a universal essential singularity exp(A12m)\sim\exp\left(-A^{1-2m}\right) for the A0A\to 0 tail. The exponents αm,n\alpha_{m,n} are highly nontrivial depending on the order mm of the stochastic process and the constrained parameter nn on path observables. For m=1m=1, corresponding to standard Brownian motion, the full distribution of AA were exactly derived by Majumdar and Meerson [42] and the exponents α1,n\alpha_{1,n} were known, which are α1,n=1/(n+2)2\alpha_{1,n}=1/{(n+2)^{2}} for all n>2n>-2. For m2m\geq 2, the high-processes are non-Markov, and the exact expressions for the full distribution of AA are unknown. However, for m=2m=2, which is the random acceleration process, Meerson has recently used the OFM to derive the first three exponents [74], α2,0=3/4\alpha_{2,0}=3/4, α2,1=108/625\alpha_{2,1}=108/625, α2,2=(27/256)tanh4(π/2)=0.074625\alpha_{2,2}=(27/256)\tanh^{4}(\pi/2)=0.074625\cdots. The main purpose of the present work is to deduce the exponents αm,n\alpha_{m,n} for higher-order processes, i.e., for m3m\geq 3. Using the OFM, we will obtain analytically the exponents αm,n\alpha_{m,n} for n=1,2n=1,2, and for arbitrary m1m\geq 1, given by Eqs.(60) and (90), respectively.

III Large deviation exponents for A0A\to 0 tail

III.1 n=0n=0

For n=0n=0, the functional 𝒜=T\mathcal{A}=T in Eq.(4) is the first-passage time itself. The Euler-Lagrange equation (12) becomes very simple: x(2m)(t)=0{x^{\left(2m\right)}}\left(t\right)=0, whose general solution reads

x(t)=c0+c1t+c2t2++c2m1t2m1.\displaystyle x(t)=c_{0}+c_{1}t+c_{2}t^{2}+\cdots+c_{2m-1}t^{2m-1}. (16)

Using the intial conditions in Eq.(5), we obtain

c0=L,c1==cm1=0.\displaystyle c_{0}=L,\quad c_{1}=\cdots=c_{m-1}=0. (17)

Inserting further the boundary conditions in Eqs.(6) and (13) to Eq.(16), we obtain the following linear equations,

{Tmcm+Tm+1cm+1++T2m2c2m2+T2m1c2m1=Lm!cm+(m+1)!1!cm+1T++(2m2)!(m2)!c2m2Tm2+(2m1)!(m1)!c2m1Tm1=0(2m3)!c2m3+(2m2)!1!c2m2T+(2m1)!2!c2m1T2=0(2m2)!c2m2+(2m1)!1!c2m1T=0\displaystyle\left\{\begin{gathered}{T^{m}}{c_{m}}+{T^{m+1}}{c_{m+1}}+\cdots+{T^{2m-2}}{c_{2m-2}}+{T^{2m-1}}{c_{2m-1}}=-L\hfill\\ m!{c_{m}}+\frac{{\left({m+1}\right)!}}{{1!}}{c_{m+1}}T+\cdots+\frac{{\left({2m-2}\right)!}}{{\left({m-2}\right)!}}{c_{2m-2}}{T^{m-2}}+\frac{{\left({2m-1}\right)!}}{{\left({m-1}\right)!}}{c_{2m-1}}{T^{m-1}}=0\hfill\\ \vdots\hfill\\ \left({2m-3}\right)!{c_{2m-3}}+\frac{{\left({2m-2}\right)!}}{{1!}}{c_{2m-2}}T+\frac{{\left({2m-1}\right)!}}{{2!}}{c_{2m-1}}{T^{2}}=0\hfill\\ \left({2m-2}\right)!{c_{2m-2}}+\frac{{\left({2m-1}\right)!}}{{1!}}{c_{2m-1}}T=0\hfill\\ \end{gathered}\right. (24)

We first let c2m1c_{2m-1} be a parameter, and rewrite the last (m1)(m-1) equations in Eq.(24) in the matrix form,

Ac=b\displaystyle\textbf{A}\textbf{c}=\textbf{b} (25)

where

A=[1Cm+11TC2m3m3Tm3C2m2m2Tm201C2m3m4Tm4C2m2m3Tm3001C2m21T0001],\displaystyle\textbf{A}=\left[{\begin{array}[]{*{20}{c}}1&{C_{m+1}^{1}T}&\cdots&{C_{2m-3}^{m-3}{T^{m-3}}}&{C_{2m-2}^{m-2}{T^{m-2}}}\\ 0&1&\cdots&{C_{2m-3}^{m-4}{T^{m-4}}}&{C_{2m-2}^{m-3}{T^{m-3}}}\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&\cdots&1&{C_{2m-2}^{1}T}\\ 0&0&\cdots&0&1\end{array}}\right], (31)
c=[cmcm+1c2m3c2m2],\displaystyle\textbf{c}=\left[{\begin{array}[]{*{20}{c}}{{c_{m}}}\\ {{c_{m+1}}}\\ \vdots\\ {{c_{2m-3}}}\\ {{c_{2m-2}}}\end{array}}\right], (37)

and

b=c2m1[C2m1m1Tm1C2m1m2Tm2C2m12T2C2m11T]\displaystyle\textbf{b}=-{c_{2m-1}}\left[{\begin{array}[]{*{20}{c}}{C_{2m-1}^{m-1}{T^{m-1}}}\\ {C_{2m-1}^{m-2}{T^{m-2}}}\\ \vdots\\ {C_{2m-1}^{2}{T^{2}}}\\ {C_{2m-1}^{1}T}\end{array}}\right] (43)

with Cmn=m!/[n!(mn)!]C_{m}^{n}={{m!}\mathord{\left/{\vphantom{{m!}{\left[{n!\left({m-n}\right)!}\right]}}}\right.\kern-1.2pt}{\left[{n!\left({m-n}\right)!}\right]}} is binominal coefficient. Here, the coefficient matrix A is an upper triangular matrix and its inverse can be easily obtained, given by

A1=[1Cm+11T(1)m3C2m3m3Tm3(1)m2C2m2m2Tm201(1)m4C2m3m4Tm4(1)m3C2m2m3Tm3001C2m21T0001]\displaystyle\textbf{A}^{-1}=\left[{\begin{array}[]{*{20}{c}}1&{-C_{m+1}^{1}T}&\cdots&{{{\left({-1}\right)}^{m-3}}C_{2m-3}^{m-3}{T^{m-3}}}&{{{\left({-1}\right)}^{m-2}}C_{2m-2}^{m-2}{T^{m-2}}}\\ 0&1&\cdots&{{{\left({-1}\right)}^{m-4}}C_{2m-3}^{m-4}{T^{m-4}}}&{{{\left({-1}\right)}^{m-3}}C_{2m-2}^{m-3}{T^{m-3}}}\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&\cdots&1&{-C_{2m-2}^{1}T}\\ 0&0&\cdots&0&1\end{array}}\right] (49)

or

(A1)ij={0,fori>j,1,fori=j,(1)jiCm+j1jiTjifori<j.\displaystyle\left(\textbf{A}^{-1}\right)_{ij}=\left\{\begin{array}[]{lll}0,&{\rm{for}}\quad i>j,\\ 1,&{\rm{for}}\quad i=j,\\ {\left({-1}\right)^{j-i}}C_{m+j-1}^{j-i}{T^{j-i}}&{\rm{for}}\quad i<j.\end{array}\right. (53)

It follows that c=A1b\textbf{c}=\textbf{A}^{-1}\textbf{b},

cm+i1\displaystyle{c_{m+i-1}} =\displaystyle= j=1m1(A1)ijbj\displaystyle\sum\limits_{j=1}^{m-1}\left(\textbf{A}^{-1}\right)_{ij}{b_{j}} (54)
=\displaystyle= c2m1Tmij=im1(1)jiCm+j1jiC2m1mj\displaystyle-{c_{2m-1}}{T^{m-i}}\sum\limits_{j=i}^{m-1}{{{\left({-1}\right)}^{j-i}}C_{m+j-1}^{j-i}}C_{2m-1}^{m-j}
=\displaystyle= (1)miTmiC2m1mic2m1,i=1,,m\displaystyle{\left({-1}\right)^{m-i}}{T^{m-i}}C_{2m-1}^{m-i}{c_{2m-1}},\quad i=1,\cdots,m

Substituting Eq.(54) into the first equation of Eq.(24), we have

L\displaystyle-L =\displaystyle= i=1mcm+i1Tm+i1\displaystyle\sum\limits_{i=1}^{m}{{c_{m+i-1}}{T^{m+i-1}}} (55)
=\displaystyle= (1)mc2m1T2m1i=1m(1)iC2m1mi\displaystyle{\left({-1}\right)^{m}}{c_{2m-1}}{T^{2m-1}}\sum\limits_{i=1}^{m}{{\left({-1}\right)}^{i}}C_{2m-1}^{m-i}
=\displaystyle= c2m1(1)m+1mC2m1m12m1T2m1.\displaystyle{c_{2m-1}}\frac{{{{\left({-1}\right)}^{m+1}}mC_{2m-1}^{m-1}}}{{2m-1}}{T^{2m-1}}.

From Eq.(55), we obtain

c2m1=(1)m(2m1)mC2m1m1LT2m1.\displaystyle{c_{2m-1}}=\frac{{{{\left({-1}\right)}^{m}}\left({2m-1}\right)}}{{mC_{2m-1}^{m-1}}}\frac{L}{{{T^{2m-1}}}}. (56)

Substituting Eq.(56) into Eq.(54), we obtain

cm+i1=(1)i(2m1)C2m1mimC2m1m1LTm+i1,i=1,,m.\displaystyle{c_{m+i-1}}={\left({-1}\right)^{i}}\frac{{\left({2m-1}\right)C_{2m-1}^{m-i}}}{{mC_{2m-1}^{m-1}}}\frac{L}{{{T^{m+i-1}}}},\quad i=1,\cdots,m. (57)

Refer to caption

Figure 2: Rescaled optimal path x(t)/Lx(t)/L as a function of t/Tt/T for the high-order processes with the first five values of mm, where the first-passage time TT is fixed.

Using Eqs.(17) and (57), we obtain the optimal path x(t)x(t) for the high-order processes with a specified value of TT,

x~(t~)=1+i=1mc~m+i1t~m+i1,\displaystyle\tilde{x}\left({\tilde{t}}\right)=1+\sum\limits_{i=1}^{m}{{{\tilde{c}}_{m+i-1}}{{\tilde{t}}^{m+i-1}}}, (58)

where x~=x/L\tilde{x}=x/L, t~=t/T\tilde{t}=t/T, and c~m+i1=cm+i1Tm+i1/L=(1)i(2m1)C2m1mi/(mC2m1m1){\tilde{c}}_{m+i-1}={c_{m+i-1}}T^{m+i-1}/L={\left({-1}\right)^{i}}\left({2m-1}\right)C_{2m-1}^{m-i}/\left({mC_{2m-1}^{m-1}}\right). In Fig.2, we show the optimal path x~(t)\tilde{x}(t) for the first five values of mm. For m=1m=1 (standard Brownian motion), the optimal path is a straight line, and thus the OFM becomes geometrical optics. For m2m\geq 2, the optimal path is no longer a straight line, and is now a polynomial of tt of order 2m12m-1.

Using the optimal path in Eq.(58), we calculate the action defined in Eq.(9)

lnPm,0(T)S=αm,0L2DT2m1,\displaystyle-\ln P_{m,0}(T)\simeq S=\frac{{{\alpha_{m,0}}{L^{2}}}}{{D{T^{2m-1}}}}, (59)

where the general expression of the exponent αm,0\alpha_{m,0} is given explicitly

αm,0=(2m1)[(m1)!]24.\displaystyle{\alpha_{m,0}}=\frac{{\left({2m-1}\right){{\left[{\left({m-1}\right)!}\right]}^{2}}}}{4}. (60)

For several small mm, we have

α1,0=14,α2,0=34,α3,0=5,α4,0=63,α5,0=1296.\displaystyle{\alpha_{1,0}}=\frac{1}{4},\quad{\alpha_{2,0}}=\frac{3}{4},\quad{\alpha_{3,0}}=5,\quad{\alpha_{4,0}}=63,\quad{\alpha_{5,0}}=1296. (61)

Obviously, for m=1m=1 (Brownian motion) and m=2m=2 (random acceleration process), our results are consistent with previous findings [42, 74].

III.2 n=1n=1

For n=1n=1, the functional 𝒜=T\mathcal{A}=T defined in Eq.(4) measures the area swept by the first-passage process. The Euler-Lagrange equation (12) becomes: x(2m)(t)=λ{x^{\left(2m\right)}}\left(t\right)=\lambda, whose solution is given by

x(t)=L+cmtm++c2m1t2m1+λ(2m)!t2m,\displaystyle x(t)=L+c_{m}t^{m}+\cdots+c_{2m-1}t^{2m-1}+\frac{\lambda}{(2m)!}t^{2m}, (62)

where we have used the boundary conditions in Eq.(5). Substituting the boundary conditions in Eqs.(6) and (13) to Eq.(62), we obtain the following linear equations,

{Tmcm+Tm+1cm+1++T2m2c2m2+T2m1c2m1=Lλ(2m)!T2mm!cm+(m+1)!1!cm+1T++(2m2)!(m2)!c2m2Tm2+(2m1)!(m1)!c2m1Tm1=λm!Tm(2m3)!c2m3+(2m2)!1!c2m2T+(2m1)!2!c2m1T2=λ3!T3(2m2)!c2m2+(2m1)!1!c2m1T=λ2!T2\displaystyle\left\{\begin{gathered}{T^{m}}{c_{m}}+{T^{m+1}}{c_{m+1}}+\cdots+{T^{2m-2}}{c_{2m-2}}+{T^{2m-1}}{c_{2m-1}}=-L-\frac{\lambda}{{\left({2m}\right)!}}{T^{2m}}\hfill\\ m!{c_{m}}+\frac{{\left({m+1}\right)!}}{{1!}}{c_{m+1}}T+\cdots+\frac{{\left({2m-2}\right)!}}{{\left({m-2}\right)!}}{c_{2m-2}}{T^{m-2}}+\frac{{\left({2m-1}\right)!}}{{\left({m-1}\right)!}}{c_{2m-1}}{T^{m-1}}=-\frac{\lambda}{{m!}}{T^{m}}\hfill\\ \vdots\hfill\\ \left({2m-3}\right)!{c_{2m-3}}+\frac{{\left({2m-2}\right)!}}{{1!}}{c_{2m-2}}T+\frac{{\left({2m-1}\right)!}}{{2!}}{c_{2m-1}}{T^{2}}=-\frac{\lambda}{{3!}}{T^{3}}\hfill\\ \left({2m-2}\right)!{c_{2m-2}}+\frac{{\left({2m-1}\right)!}}{{1!}}{c_{2m-1}}T=-\frac{\lambda}{{2!}}{T^{2}}\hfill\\ \end{gathered}\right. (69)

As done in the last subsection, we first let c2m1c_{2m-1} be a parameter, and rewrite the last m1m-1 equations in Eq.(69) in the matrix form,

Ac=d,\displaystyle\textbf{A}\textbf{c}=\textbf{d}, (70)

where A and c are given in Eq.(31) and Eq.(37), respectively. The vector d in Eq.(70) is given by

d=bλ[1m!m!Tm1(m1)(m+1)!Tm113!(2m3)!T312!(2m2)!T2],\displaystyle\textbf{d}=\textbf{b}-\lambda\left[{\begin{array}[]{*{20}{c}}{\frac{1}{{m!m!}}{T^{m}}}\\ {\frac{1}{{\left({m-1}\right)\left({m+1}\right)!}}{T^{m-1}}}\\ \vdots\\ {\frac{1}{{3!\left({2m-3}\right)!}}{T^{3}}}\\ {\frac{1}{{2!\left({2m-2}\right)!}}{T^{2}}}\end{array}}\right], (76)

where b is given in Eq.(43). The solution of Eq.(70), c=A1d\textbf{c}=\textbf{A}^{-1}\textbf{d}, reads

cm+i1=(1)miTmiC2m1mic2m1+(1)miλTm+1i(mi)(mi+1)!(m+i1)!,i=1,,m.\displaystyle{c_{m+i-1}}={\left({-1}\right)^{m-i}}{T^{m-i}}C_{2m-1}^{m-i}{c_{2m-1}}+{\left({-1}\right)^{m-i}}\frac{{\lambda{T^{m+1-i}}\left({m-i}\right)}}{{\left({m-i+1}\right)!\left({m+i-1}\right)!}},\quad i=1,\cdots,m. (77)

Substituting Eq.(77) into the first line of Eq.(69), we obtain

c2m1=(1)m(2m1)(m!)2L+λT2m(mm21/2)mC2m1m1(m!)2T2m1.\displaystyle{c_{2m-1}}=\frac{{{{\left({-1}\right)}^{m}}\left({2m-1}\right){{\left({m!}\right)}^{2}}L+\lambda{T^{2m}}\left({m-{m^{2}}-1/2}\right)}}{{mC_{2m-1}^{m-1}{{\left({m!}\right)}^{2}}{T^{2m-1}}}}. (78)

Using Eq.(78), Eq.(77) can be rewritten as

cm+i1=(1)i(2m1)C2m1mimC2m1m1Tm+i1[L+(1)mλT2m(mm21/2)(2m1)(m!)2]+(1)miλTm+1i(mi)(mi+1)!(m+i1)!,i=1,,m.\displaystyle{c_{m+i-1}}=\frac{{{{\left({-1}\right)}^{i}}\left({2m-1}\right)C_{2m-1}^{m-i}}}{{mC_{2m-1}^{m-1}{T^{m+i-1}}}}\left[{L+\frac{{{{\left({-1}\right)}^{m}}\lambda{T^{2m}}\left({m-{m^{2}}-1/2}\right)}}{{\left({2m-1}\right){{\left({m!}\right)}^{2}}}}}\right]+\frac{{{{\left({-1}\right)}^{m-i}}\lambda{T^{m+1-i}}\left({m-i}\right)}}{{\left({m-i+1}\right)!\left({m+i-1}\right)!}},\quad i=1,\cdots,m.

Using the constraint A=0Tx(t)𝑑tA=\int_{0}^{T}x(t)dt, we obtain the relation between AA and λ\lambda,

λ=(1)m+12(2m+1)(m!)2T2m+1[(2m2+12m)LT2m2A].\displaystyle\lambda=\frac{{{{\left({-1}\right)}^{m+1}}2\left({2m+1}\right){{\left({m!}\right)}^{2}}}}{{{T^{2m+1}}}}\left[{\left({2{m^{2}}+1-2m}\right)LT-2{m^{2}}A}\right]. (80)

The optimal value of first-passage time TT_{*} can be obtained from the boundary condition in Eq.(14), which leads to

(2m1)!c2m1+λT=0.\displaystyle\left({2m-1}\right)!{c_{2m-1}}+\lambda T=0. (81)

Eq.(81) can be simplified by using Eq.(78) and Eq.(80), which yields the optimal first-passage time TT_{*}

T=κmAL,\displaystyle T_{*}=\frac{\kappa_{m}A}{L}, (82)

with

κm=2m+12m1.\displaystyle\kappa_{m}=\frac{2m+1}{2m-1}. (83)

Utilizing Eq.(82), Eq.(80) and Eq.(III.2) can be respectively rewritten as

λ=2(1)m+1(m!)2κm2mL2m+1A2m,\displaystyle\lambda=\frac{{2{{\left({-1}\right)}^{m+1}}{{\left({m!}\right)}^{2}}}}{{\kappa_{m}^{2m}}}\frac{{{L^{2m+1}}}}{{{A^{2m}}}}, (84)

and

cm+i1=2(1)iCmmi+1κmm+i1Cm+i1mLm+iAm+i1,i=1,,m.\displaystyle{c_{m+i-1}}=\frac{{2{{\left({-1}\right)}^{i}}C_{m}^{m-i+1}}}{{\kappa_{m}^{m+i-1}C_{m+i-1}^{m}}}\frac{{{L^{m+i}}}}{{{A^{m+i-1}}}},\quad i=1,\cdots,m. (85)

Using the rescaled coordinates, x~=x/L\tilde{x}=x/L and t~=t/T\tilde{t}=t/T^{*}, the optimal path can be rewritten as

x~(t~)=1+i=0mc~m+it~m+i,\displaystyle\tilde{x}\left({\tilde{t}}\right)=1+\sum\limits_{i=0}^{m}{{{\tilde{c}}_{m+i}}{{\tilde{t}}^{m+i}}}, (86)

where c~m+i=cm+iTm+i/L{{{\tilde{c}}_{m+i}}}=c_{m+i}T_{*}^{m+i}/L is given by

c~m+i=2(1)i+1CmmiCm+im,i=0,,m.\displaystyle{{{\tilde{c}}_{m+i}}}=\frac{{2{{\left({-1}\right)}^{i+1}}C_{m}^{m-i}}}{{C_{m+i}^{m}}},\quad i=0,\cdots,m. (87)

Refer to caption

Figure 3: Rescaled optimal path x(t)/Lx(t)/L as a function of t/Tt/T_{*} for the high-order processes with the first five values of mm, where the first-passage area A=LT/κmA=LT_{*}/\kappa_{m} is fixed with κm=(2m+1)/(2m1)\kappa_{m}=(2m+1)/(2m-1).

In Fig.3, we plot the optimal path in the rescaled coordinates for the first five values of mm with a specified first-passage area A=LT/κmA=LT_{*}/\kappa_{m}. The action in Eq.(9) along the optimal path Eq.(86) can be computed as

S\displaystyle S =\displaystyle= λ2T2m+14D(m!)2(2m+1)+λ2Dm!i=1m(m+i1)!(i1)!(m+i)cm+i1Tm+i\displaystyle\frac{{{\lambda^{2}}{T^{2m+1}}}}{4D{{{\left({m!}\right)}^{2}}\left({2m+1}\right)}}+\frac{{\lambda}}{2D{m!}}\sum\limits_{i=1}^{m}{\frac{{\left({m+i-1}\right)!}}{{\left({i-1}\right)!\left({m+i}\right)}}{c_{m+i-1}}{T^{m+i}}} (88)
+\displaystyle+ 14Di=1mj=1m(m+i1)!(m+j1)!(i1)!(j1)!(i+j1)cm+i1cm+j1Ti+j1\displaystyle\frac{1}{4D}\sum\limits_{i=1}^{m}{\sum\limits_{j=1}^{m}{\frac{{\left({m+i-1}\right)!\left({m+j-1}\right)!}}{{\left({i-1}\right)!\left({j-1}\right)!\left({i+j-1}\right)}}{c_{m+i-1}}{c_{m+j-1}}{T^{i+j-1}}}}

Substituting Eq.(82), Eq.(84) and Eq.(85) into Eq.(88), we obtain

lnPm,1(A)S=αm,1L2m+1DA2m1,\displaystyle-\ln P_{m,1}(A)\simeq S=\frac{{{\alpha_{m,1}}{L^{2m+1}}}}{{D{A^{2m-1}}}}, (89)

where the exponent αm,1{\alpha_{m,1}} is obtained explicitly

αm,1\displaystyle{\alpha_{m,1}} =\displaystyle= (m!)2κm2m(2m1)=(m!)2(2m1)2m1(2m+1)2m.\displaystyle\frac{{{{\left({m!}\right)}^{2}}}}{{{\kappa_{m}^{2m}}\left({2m-1}\right)}}=\frac{{{{\left({m!}\right)}^{2}}{{\left({2m-1}\right)}^{2m-1}}}}{{{{\left({2m+1}\right)}^{2m}}}}. (90)

For several small mm, we have

α1,1=19,α2,1=108625,α3,1=112500117649,α4,1=527067524782969,α5,1=557885504160025937424601.\displaystyle{\alpha_{1,1}}=\frac{1}{9},\quad{\alpha_{2,1}}=\frac{108}{625},\quad{\alpha_{3,1}}=\frac{112500}{117649},\quad{\alpha_{4,1}}=\frac{52706752}{4782969},\quad{\alpha_{5,1}}=\frac{5578855041600}{25937424601}. (91)

One can compare the results in Eq.(91) with previous works for m=1m=1 (Brownian motion) and for m=2m=2 (random acceleration process) [42, 74], and find that they are exactly the same.

IV Conclusions

In conclusion, we have employed the OFM to study the short-time statistics of the first-passage functionals A=0Txn(t)𝑑tA=\int_{0}^{T}x^{n}(t)dt for the high-order processes x(t)x(t) indexed by a positive integer mm. Via the variation of a modified action functional, we obtain the equation of motion for the most likely realization of x(t)x(t). Combining boundary conditions, the equation of motion can be solved analytically for n=0,1n=0,1 and the optimal path x(t)x(t) can be correspondingly obtained. The action along the optimal path can be also obtained analytically. As a result, the distribution of AA displays an essential singularity at the tail of A0A\to 0, Pm,n(A|L)exp(αm,nL2mnn+2DA2m1)P_{m,n}(A|L)\sim\exp\left(-\frac{\alpha_{m,n}L^{2mn-n+2}}{DA^{2m-1}}\right), where the exponent αm,n\alpha_{m,n} is highly nontrivial. The main contribution in the present work is that we have obtained the explicit expressions of αm,n\alpha_{m,n} for n=0,1n=0,1 and for arbitrary mm (see Eqs.(60) and (90)), which enables us to evaluate the exact A0A\to 0 tail of the distributions of the first-passage time itself and the area swept by the high-order processes till the first-passage time.

For n=2n=2, the Euler-Lagrange equation (12) can be analytically solved as well, but the solution is no longer a polynomial. This leads to the difficulty in obtaining the coefficients by the boundary conditions Eqs.(5), (6) and (13), such that the determination of the exponents αm,2\alpha_{m,2} for general mm is still a challenging task. However, for m=1m=1 (Brownian motion) and m=2m=2 (random acceleration process), the exponents αm,2\alpha_{m,2} are known [42, 74]. For n3n\geq 3, the Euler-Lagrange equation (12) cannot be solved analytically, and thus the exponents αm,n\alpha_{m,n} for n3n\geq 3 can be only determined by numerical methods (except for m=1m=1).

Appendix A Derivations of Euler-Lagrange equation and boundary conditions

We perform a linear variation of the constrained action functional defined in Eq.(11),

δSλ=0T(x(m)δx(m)λnxn1δx)𝑑t+TT+δT(12[x(m)]2λxn)𝑑t.\displaystyle\delta{S_{\lambda}}=\int_{0}^{T}{\left({{x^{\left(m\right)}}\delta{x^{\left(m\right)}}-\lambda n{x^{n-1}}\delta x}\right)dt}+\int_{T}^{T+\delta T}{\left({\frac{1}{2}{{\left[{{x^{\left(m\right)}}}\right]}^{2}}-\lambda{x^{n}}}\right)dt}. (92)

Performing mm times integrations in parts in the first integral in Eq.(92), and evaluating the second integral in the limit of δT0\delta T\to 0, we obtain

δSλ\displaystyle\delta{S_{\lambda}} =\displaystyle= (1)m0T(x(2m)λnxn1)δx𝑑t+x(m)δx(m1)|0Tx(m+1)δx(m2)|0T++x(2m1)δx|0T\displaystyle{\left({-1}\right)^{m}}\int_{0}^{T}{\left({{x^{\left({2m}\right)}}-\lambda n{x^{n-1}}}\right)\delta xdt}+\left.{{x^{\left(m\right)}}\delta{x^{\left({m-1}\right)}}}\right|_{0}^{T}-\left.{{x^{\left({m+1}\right)}}\delta{x^{\left({m-2}\right)}}}\right|_{0}^{T}+\cdots+\left.{{x^{\left({2m-1}\right)}}\delta x}\right|_{0}^{T} (93)
+\displaystyle+ [12[x(m)(T)]2λxn(T)x(2m1)(T)x(m1)(T)].\displaystyle\left[{\frac{1}{2}{{\left[{{x^{\left(m\right)}}\left(T\right)}\right]}^{2}}-\lambda{x^{n}}\left(T\right)-{x^{\left({2m-1}\right)}}\left(T\right){x^{\left({m-1}\right)}}\left(T\right)}\right].

Each of the three terms in the variation must vanish independently for arbitrary δx\delta x and δT\delta T. The first term yields the Euler-Lagrange equation x(2m)(t)λnxn1(t)=0x^{(2m)}(t)-\lambda nx^{n-1}(t)=0. According to boundary conditions at initial stage in Eq.(5) and one at final stage x(T)=0x(T)=0, we obtain other boundary conditions at final stage: x(m)(T)==x(2m2)(T)=0x^{(m)}(T)=\cdots=x^{(2m-2)}(T)=0. The last term leads to the condition: x(2m1)(T)x(m1)(T)=0x^{(2m-1)}(T)x^{(m-1)}(T)=0. One has to select one of two options: x(2m1)(T)=0x^{(2m-1)}(T)=0 or x(m1)(T)x^{(m-1)}(T) (they cannot hold simultaneously, as the problem then would be overdetermined). As one can check, the condition x(2m1)(T)=0x^{(2m-1)}(T)=0 would give a local maximum of the action functionals, whereas the condition x(m1)(T)=0x^{(m-1)}(T)=0 yields the desired minimum.

Acknowledgements.
This work was supported by the National Natural Science Foundation of China (11875069), the Key Scientific Research Fund of Anhui Provincial Education Department (2023AH050116), and Anhui Project (Grant No. 2022AH020009)

References

  • Zwanzig [2001] R. Zwanzig, Nonequilibrium statistical mechanics (Oxford university press, 2001).
  • Seifert [2012] U. Seifert, Reports on progress in physics 75, 126001 (2012).
  • 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.
  • Kac [1949] M. Kac, Transactions of the American Mathematical Society 65, 1 (1949).
  • Agmon [1984] N. Agmon, The Journal of chemical physics 81, 3644 (1984).
  • Berezhkovskii et al. [1998] A. M. Berezhkovskii, V. Zaloj, and N. Agmon, Phys. Rev. E 57, 3937 (1998).
  • Nguyen and Grebenkov [2010] B. T. Nguyen and D. S. Grebenkov, Journal of Statistical Physics 141, 532 (2010).
  • Barato et al. [2018] A. C. Barato, E. Roldán, I. A. Martínez, and S. Pigolotti, Phys. Rev. Lett. 121, 090601 (2018).
  • Agmon [2011] N. Agmon, The Journal of Physical Chemistry A 115, 5838 (2011).
  • Lévy [1940] P. Lévy, Compos. Math. 7, 283 (1940).
  • Feller [1971] W. Feller, Vols. I & II, Wiley I 968 (1971).
  • Grebenkov [2007] D. S. Grebenkov, Phys. Rev. E 76, 041139 (2007).
  • Kaldasch and Engel [2022] S. Kaldasch and A. Engel, Phys. Rev. E 105, 034132 (2022).
  • Kay and Giuggioli [2023] T. Kay and L. Giuggioli, Journal of Physics A: Mathematical and Theoretical 56, 345002 (2023).
  • Huang and Chen [2024] F. Huang and H. Chen, Physica A: Statistical Mechanics and its Applications 633, 129389 (2024).
  • Majumdar and Comtet [2002] S. N. Majumdar and A. Comtet, Phys. Rev. Lett. 89, 060601 (2002).
  • Sabhapandit et al. [2006] S. Sabhapandit, S. N. Majumdar, and A. Comtet, Phys. Rev. E 73, 051102 (2006).
  • Singh [2022] P. Singh, Phys. Rev. E 105, 024113 (2022).
  • 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).
  • Reymbaut et al. [2011] A. Reymbaut, S. N. Majumdar, and A. Rosso, Journal of Physics A: Mathematical and Theoretical 44, 415001 (2011).
  • 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).
  • Singh and Kundu [2021] P. Singh and A. Kundu, Phys. Rev. E 103, 042119 (2021).
  • Singh and Kundu [2019] P. Singh and A. Kundu, J. Stat. Mech.: Theo. and Experiment 2019, 083205 (2019).
  • Pal et al. [2019] A. Pal, R. Chatterjee, S. Reuveni, and A. Kundu, Journal of Physics A: Mathematical and Theoretical 52, 264002 (2019).
  • Singh and Pal [2021] P. Singh and A. Pal, Phys. Rev. E 103, 052119 (2021).
  • Den Hollander et al. [2019] F. Den Hollander, S. N. Majumdar, J. M. Meylahn, and H. Touchette, Journal of Physics A: Mathematical and Theoretical 52, 175001 (2019).
  • Bressloff [2020] P. C. Bressloff, Phys. Rev. E 102, 042135 (2020).
  • Yan and Chen [2023] H. Yan and H. Chen, Physica Scripta 98, 125226 (2023).
  • Godreche and Luck [2001] C. Godreche and J. Luck, Journal of Statistical Physics 104, 489 (2001).
  • Burov and Barkai [2011] S. Burov and E. Barkai, Phys. Rev. Lett. 107, 170601 (2011).
  • Burenev et al. [2023] I. N. Burenev, S. N. Majumdar, and A. Rosso, Phys. Rev. E 108, 064113 (2023).
  • Burenev et al. [2024] I. N. Burenev, S. N. Majumdar, and A. Rosso, Phys. Rev. E 109, 044150 (2024).
  • Smith and Meerson [2024] N. R. Smith and B. Meerson, Physica A: Statistical Mechanics and its Applications 639, 129616 (2024).
  • Majumdar and Comtet [2004] S. N. Majumdar and A. Comtet, Phys. Rev. Lett. 92, 225501 (2004).
  • Barkai et al. [2014] E. Barkai, E. Aghion, and D. A. Kessler, Phys. Rev. X 4, 021036 (2014).
  • Redner [2001] S. Redner, A guide to first-passage processes (Cambridge University Press, 2001).
  • Bray et al. [2013] A. J. Bray, S. N. Majumdar, and G. Schehr, Adv. Phys. 62, 225 (2013).
  • Kearney and Majumdar [2005] M. J. Kearney and S. N. Majumdar, Journal of Physics A: Mathematical and General 38, 4097 (2005).
  • Kearney and Martin [2016] M. J. Kearney and R. J. Martin, Journal of Physics A: Mathematical and Theoretical 49, 195001 (2016).
  • Majumdar and Meerson [2020a] S. N. Majumdar and B. Meerson, Journal of Statistical Mechanics: Theory and Experiment 2020, 023202 (2020a).
  • Kearney et al. [2007] M. J. Kearney, S. N. Majumdar, and R. J. Martin, Journal of Physics A: Mathematical and Theoretical 40, F863 (2007).
  • Abundo and Vescovo [2017] M. Abundo and D. D. Vescovo, Methodology and Computing in Applied Probability 19, 985 (2017).
  • Kearney and Martin [2021] M. J. Kearney and R. J. Martin, Journal of Physics A: Mathematical and Theoretical 54, 055002 (2021).
  • Abundo [2023a] M. Abundo, Stochastic Analysis and Applications 41, 358 (2023a).
  • Radice [2023] M. Radice, Phys. Rev. E 108, 044151 (2023).
  • Singh and Pal [2022] P. Singh and A. Pal, J. Phys. A: Math. Theor. 55, 234001 (2022).
  • Dubey and Pal [2023] A. Dubey and A. Pal, Journal of Physics A: Mathematical and Theoretical 56, 435002 (2023).
  • Abundo [2023b] M. Abundo, Methodology and Computing in Applied Probability 25, 92 (2023b).
  • Agranov et al. [2020] T. Agranov, P. Zilber, N. R. Smith, T. Admon, Y. Roichman, and B. Meerson, Physical Review Research 2, 013174 (2020).
  • Touchette [2009] H. Touchette, Physics Reports 478, 1 (2009).
  • Assaf and Meerson [2017] M. Assaf and B. Meerson, J. Phys. A: Math. Theor. 50, 263001 (2017).
  • Nyawo and Touchette [2017] P. T. Nyawo and H. Touchette, Europhysics Letters 116, 50009 (2017).
  • Nyawo and Touchette [2018] P. T. Nyawo and H. Touchette, Phys. Rev. E 98, 052103 (2018).
  • Mukherjee and Smith [2023] S. Mukherjee and N. R. Smith, Phys. Rev. E 107, 064133 (2023).
  • Mukherjee et al. [2024] S. Mukherjee, P. Le Doussal, and N. R. Smith, Phys. Rev. E 110, 024107 (2024).
  • Nickelsen and Touchette [2018] D. Nickelsen and H. Touchette, Physical review letters 121, 090602 (2018).
  • Meerson [2019a] B. Meerson, Phys. Rev. E 100, 042135 (2019a).
  • Smith [2022] N. R. Smith, Phys. Rev. E 105, 014120 (2022).
  • Nickelsen and Touchette [2022] D. Nickelsen and H. Touchette, Phys. Rev. E 105, 064102 (2022).
  • Smith [2024] N. R. Smith, Physica A: Statistical Mechanics and its Applications 650, 129987 (2024).
  • Touchette [2018] H. Touchette, Physica A: Statistical Mechanics and its Applications 504, 5 (2018).
  • Halperin and Lax [1966] B. Halperin and M. Lax, Physical Review 148, 722 (1966).
  • Zittartz and Langer [1966] J. Zittartz and J. Langer, Physical Review 148, 741 (1966).
  • Lifshitz [1968] I. Lifshitz, Sov. Phys. JETP 26, 012110 (1968).
  • Meerson and Smith [2019] B. Meerson and N. R. Smith, Journal of Physics A: Mathematical and Theoretical 52, 415001 (2019).
  • Meerson [2019b] B. Meerson, Journal of Statistical Mechanics: Theory and Experiment 2019, 013210 (2019b).
  • Meerson [2020] B. Meerson, Journal of Statistical Mechanics: Theory and Experiment 2020, 103208 (2020).
  • Bar and Meerson [2023] T. Bar and B. Meerson, Journal of Statistical Mechanics: Theory and Experiment 2023, 093301 (2023).
  • Meerson et al. [2016] B. Meerson, E. Katzav, and A. Vilenkin, Physical review letters 116, 070601 (2016).
  • Meerson and Oshanin [2022] B. Meerson and G. Oshanin, Physical Review E 105, 064137 (2022).
  • Hartmann and Meerson [2024] A. K. Hartmann and B. Meerson, Phys. Rev. E 109, 014146 (2024).
  • Meerson [2023] B. Meerson, Physical Review E 107, 064122 (2023).
  • Chen et al. [2024] H. Chen, L. Tian, and G. Li, arXiv preprint arXiv:2407.00302 (2024).
  • Majumdar and Meerson [2020b] S. N. Majumdar and B. Meerson, Physical Review E 102, 022113 (2020b).
  • Shepp [1966] L. A. Shepp, The Annals of Mathematical Statistics pp. 321–354 (1966).
  • Wahba [1978] G. Wahba, Journal of the Royal Statistical Society Series B: Statistical Methodology 40, 364 (1978).
  • Majumdar et al. [1996] S. N. Majumdar, C. Sire, A. J. Bray, and S. J. Cornell, Phys. Rev. Lett. 77, 2867 (1996).
  • Ehrhardt et al. [2004] G. C. M. A. Ehrhardt, S. N. Majumdar, and A. J. Bray, Phys. Rev. E 69, 016106 (2004).