14
Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to confinement ´ Edgar Rold´ an *1 and Shamik Gupta 2 1 Max-Planck Institute for the Physics of Complex Systems, cfAED and GISC, N¨ othnitzer Straße 38, 01187 Dresden, Germany 2 Department of Physics, Ramakrishna Mission Vivekananda University, Belur Math, Howrah 711 202, West Bengal, India (Dated: July 27, 2017) We study the dynamics of overdamped Brownian particles diffusing in conservative force fields and undergoing stochastic resetting to a given location with a generic space-dependent rate of resetting. We present a systematic approach involving path integrals and elements of renewal theory that allows to derive analytical expressions for a variety of statistics of the dynamics such as (i) the propagator prior to first reset; (ii) the distribution of the first-reset time, and (iii) the spatial distribution of the particle at long times. We apply our approach to several representative and hitherto unexplored examples of resetting dynamics. A particularly interesting example for which we find analytical expressions for the statistics of resetting is that of a Brownian particle trapped in a harmonic potential with a rate of resetting that depends on the instantaneous energy of the particle. We find that using energy-dependent resetting processes is more effective in achieving spatial confinement of Brownian particles on a faster timescale than by performing quenches of parameters of the harmonic potential. PACS numbers: 05.40.-a, 02.50.-r, 05.70.Ln I. INTRODUCTION Changes are inevitable in nature, and those that are most dramatic with often drastic consequences are the ones that occur all of a sudden. A particular class of such changes comprises those in which the system dur- ing its temporal evolution makes a sudden jump (a “re- set”) to a fixed state or configuration. Many nonequilib- rium processes are encountered across disciplines, e.g., in physics, biology, and information processing, which in- volve sudden transitions between different states or con- figurations. The erasure of a bit of information [1, 2] by mesoscopic machines may be thought of as a physical pro- cess in which a memory device that is strongly affected by thermal fluctuations resets its state (0 or 1) to a pre- scribed erasure state [3–7]. In biology, resetting plays an important role inter alia in sensing of extracellular lig- ands by single cells [8], and in transcription of genetic in- formation by macromolecular enzymes called RNA poly- merases [9]. During RNA transcription, the recovery of RNA polymerases from inactive transcriptional pauses is a result of a kinetic competition between diffusion and resetting of the polymerase to an active state via RNA cleavage [9], as has been recently tested in high-resolution single-molecule experiments [10]. Also, there are ample examples of biochemical processes that initiate (i.e., re- set) at random so-called stopping times [11–13], with the initiation at each instance occurring in different regions of space [14]. In addition, interactions play a key role in determining when and where a chemical reaction oc- * Corresponding author: [email protected] curs [11], a fact that affects the statistics of the resetting process. For instance, in the above mentioned example of recovery of RNA polymerase by the process of resetting, the interaction of the hybrid DNA-RNA may alter the time that a polymerase takes to recover from its inactive state [15]. It is therefore quite pertinent and timely to study resetting of mesoscopic systems that evolve under the influence of external or conservative force fields. Simple diffusion subject to resetting to a given loca- tion at random times has emerged in recent years as a convenient theoretical framework to discuss the phe- nomenon of stochastic resetting [16–21]. The framework has later been generalized to consider different choices of the resetting position [22, 23], resetting of continuous- time random walks [24, 25], L´ evy [26] and exponential constant-speed flights [27], time-dependent resetting of a Brownian particle [28], and in discussing memory ef- fects [29] and phase transitions in reset processes [30]. Stochastic resetting has also been invoked in the con- text of many-body dynamics, e.g., in reaction-diffusion models [31], fluctuating interfaces [32, 33], interacting Brownian motion [34], and in discussing optimal search times in a crowded environment [35–38]. However, lit- tle is known about the statistics of stochastic resetting of Brownian particles that diffuse under the influence of force fields [39], and that too in presence of a rate of resetting that varies in space. In this paper, we study the dynamics of overdamped Brownian particles immersed in a thermal environment, which diffuse under the influence of a force field, and whose position may be stochastically reset to a given spa- tial location with a rate of resetting that has an essential dependence on space. We use an approach that allows to obtain exact expressions for the transition probability arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017

arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

Path-integral formalism for stochastic resetting:Exactly solved examples and shortcuts to confinement

Edgar Roldan∗1 and Shamik Gupta2

1Max-Planck Institute for the Physics of Complex Systems,cfAED and GISC, Nothnitzer Straße 38, 01187 Dresden, Germany

2Department of Physics, Ramakrishna Mission Vivekananda University,Belur Math, Howrah 711 202, West Bengal, India

(Dated: July 27, 2017)

We study the dynamics of overdamped Brownian particles diffusing in conservative force fields andundergoing stochastic resetting to a given location with a generic space-dependent rate of resetting.We present a systematic approach involving path integrals and elements of renewal theory that allowsto derive analytical expressions for a variety of statistics of the dynamics such as (i) the propagatorprior to first reset; (ii) the distribution of the first-reset time, and (iii) the spatial distribution ofthe particle at long times. We apply our approach to several representative and hitherto unexploredexamples of resetting dynamics. A particularly interesting example for which we find analyticalexpressions for the statistics of resetting is that of a Brownian particle trapped in a harmonicpotential with a rate of resetting that depends on the instantaneous energy of the particle. We findthat using energy-dependent resetting processes is more effective in achieving spatial confinement ofBrownian particles on a faster timescale than by performing quenches of parameters of the harmonicpotential.

PACS numbers: 05.40.-a, 02.50.-r, 05.70.Ln

I. INTRODUCTION

Changes are inevitable in nature, and those that aremost dramatic with often drastic consequences are theones that occur all of a sudden. A particular class ofsuch changes comprises those in which the system dur-ing its temporal evolution makes a sudden jump (a “re-set”) to a fixed state or configuration. Many nonequilib-rium processes are encountered across disciplines, e.g., inphysics, biology, and information processing, which in-volve sudden transitions between different states or con-figurations. The erasure of a bit of information [1, 2] bymesoscopic machines may be thought of as a physical pro-cess in which a memory device that is strongly affectedby thermal fluctuations resets its state (0 or 1) to a pre-scribed erasure state [3–7]. In biology, resetting plays animportant role inter alia in sensing of extracellular lig-ands by single cells [8], and in transcription of genetic in-formation by macromolecular enzymes called RNA poly-merases [9]. During RNA transcription, the recovery ofRNA polymerases from inactive transcriptional pauses isa result of a kinetic competition between diffusion andresetting of the polymerase to an active state via RNAcleavage [9], as has been recently tested in high-resolutionsingle-molecule experiments [10]. Also, there are ampleexamples of biochemical processes that initiate (i.e., re-set) at random so-called stopping times [11–13], with theinitiation at each instance occurring in different regionsof space [14]. In addition, interactions play a key rolein determining when and where a chemical reaction oc-

∗Corresponding author: [email protected]

curs [11], a fact that affects the statistics of the resettingprocess. For instance, in the above mentioned example ofrecovery of RNA polymerase by the process of resetting,the interaction of the hybrid DNA-RNA may alter thetime that a polymerase takes to recover from its inactivestate [15]. It is therefore quite pertinent and timely tostudy resetting of mesoscopic systems that evolve underthe influence of external or conservative force fields.

Simple diffusion subject to resetting to a given loca-tion at random times has emerged in recent years asa convenient theoretical framework to discuss the phe-nomenon of stochastic resetting [16–21]. The frameworkhas later been generalized to consider different choicesof the resetting position [22, 23], resetting of continuous-time random walks [24, 25], Levy [26] and exponentialconstant-speed flights [27], time-dependent resetting ofa Brownian particle [28], and in discussing memory ef-fects [29] and phase transitions in reset processes [30].Stochastic resetting has also been invoked in the con-text of many-body dynamics, e.g., in reaction-diffusionmodels [31], fluctuating interfaces [32, 33], interactingBrownian motion [34], and in discussing optimal searchtimes in a crowded environment [35–38]. However, lit-tle is known about the statistics of stochastic resettingof Brownian particles that diffuse under the influence offorce fields [39], and that too in presence of a rate ofresetting that varies in space.

In this paper, we study the dynamics of overdampedBrownian particles immersed in a thermal environment,which diffuse under the influence of a force field, andwhose position may be stochastically reset to a given spa-tial location with a rate of resetting that has an essentialdependence on space. We use an approach that allowsto obtain exact expressions for the transition probability

arX

iv:1

703.

1061

5v2

[co

nd-m

at.s

tat-

mec

h] 2

6 Ju

l 201

7

Page 2: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

2

a) b) c)

reset

reset

reset

FIG. 1: Stochastic resetting of Brownian particles moving in force fields and subject to stochastic resetting witha rate that may depend on space. Illustration of the motion of a single overdamped Brownian particle (grey sphere) thatdiffuses in two dimensions (x, y) in presence of a conservative potential V (x, y). The particle traces a stochastic trajectory (blackcurve) in the two-dimensional space until its position is reset at a random instant of time to its initial value (orange arrow);subsequent to the reset, the particle resumes its stochastic motion until the next reset happens. The rate of resetting r(x, y)(left colorbar) may depend on the position of the particle. We study three different scenarios: a) Resetting of free Brownianparticles under a space-independent rate of resetting; b) Resetting of free Brownian particles under a space-dependent rate ofresetting; c) Resetting of Brownian particles moving in force fields under a space-dependent rate of resetting.

prior to the first reset, the first reset-time distribution,and, most importantly, the stationary spatial distribu-tion of the particle. The approach is based on a combina-tion of the theory of renewals [40] and the Feynman-Kacpath-integral formalism of treating stochastic processes[41–44], and consists in a mapping of the dynamics of theBrownian resetting problem to a suitable quantum me-chanical evolution in imaginary time. We note that theFeynman-Kac formalism has been applied extensively inthe past to discuss dynamical processes involving diffu-sion [45], and has to the best of our knowledge not beenapplied to discuss stochastic resetting. To demonstratethe utility of the approach, we consider several differentstochastic resetting problems, see Fig. 1: i) Free Brow-nian particles subject to a space-independent rate of re-setting (Fig. 1a)); ii) Free Brownian particles subject toresetting with a rate that depends quadratically on thedistance to the origin (Fig. 1b)); and iii) Brownian parti-cles trapped in a harmonic potential and undergoing resetevents with a rate that depends on the energy of the par-ticle (Fig. 1c)). In this paper, we consider for purposesof illustration the corresponding scenarios in one dimen-sion, although our general approach may be extended tohigher dimensions. Remarkably, we obtain exact analyt-ical expressions in all cases, and, notably, in cases ii) andiii), where a standard treatment of analytic solution byusing the Fokker-Planck approach may appear daunting,and whose relevance in physics may be explored in thecontext of, e.g., optically-trapped colloidal particles andhopping processes in glasses and gels. We further ex-plore the dynamical properties of case iii), and comparethe relaxation properties of dynamics corresponding topotential energy quenches and due to sudden activationof space-dependent stochastic resetting.

II. GENERAL FORMALISM

A. Model of study: resetting of Brownian particlesdiffusing in force fields

Consider an overdamped Brownian particle diffusingin one dimension x in presence of a time-independentforce field F (x) = −∂xV (x), with V (x) denoting a po-tential energy landscape. The dynamics of the particle isdescribed by a Langevin equation of the form

dx

dt= µF (x) + η(t), (1)

where µ is the mobility of the particle, defined as thevelocity per unit force. In Eq. (1), η(t) is a Gaussianwhite noise, with the properties

〈η(t)〉 = 0, 〈η(t)η(t′)〉 = 2Dδ(t− t′), (2)

where 〈 · 〉 denotes average over noise realizations, andD>0 is the diffusion coefficient of the particle, with thedimension of length-squared over time. We assume thatthe Einstein relation holds: D = kBTµ, with T being thetemperature of the environment, and with kB being theBoltzmann constant. In addition to the dynamics (1),the particle is subject to a stochastic resetting dynam-ics with a space-dependent resetting rate r(x), whereby,while at position x at time t, the particle in the ensuinginfinitesimal time interval dt either follows the dynamics(1) with probability 1 − r(x)dt, or resets to a given re-set destination x(r) with probability r(x)dt. Our analysisholds for any arbitrary reset function r(x), with the onlyobvious constraint r(x) ≥ 0 ∀x; moreover, the formalismmay be generalized to higher dimensions. In the follow-ing, we consider the reset location to be the same as theinitial location x0 of the particle, that is, x(r) = x0.

A quantity of obvious interest and relevance is the spa-tial distribution of the particle: What is the probability

Page 3: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

3

P (x, t|x0, 0) that the particle is at position x at time t,given that its initial location is x0? From the dynamicsgiven in the preceding paragraph, it is straightforward towrite the time evolution equation of P (x, t|x0, 0):

∂P (x, t|x0, 0)

∂t= −µ∂(F (x)P (x, t|x0, 0))

∂x+D

∂2P (x, t|x0, 0)

∂x2

−r(x)P (x, t|x0, 0) +

∫dy r(y)P (y, t|x0, 0)δ(x− x0), (3)

where the first two terms on the right hand side accountfor the contribution from the diffusion of the particle inthe force field F (x), while the last two terms stand for thecontribution owing to the resetting of the particle: thethird term represents the loss in probability arising fromthe resetting of the particle to x0, while the fourth termdenotes the gain in probability at the location x0 owingto resetting from all locations x 6= x0. When exists, thestationary distribution Pst(x|x0) satisfies

0 = −µ∂(F (x)Pst(x|x0))

∂x+D

∂2Pst(x|x0)

∂x2

−r(x)Pst(x|x0) +

∫dy r(y)Pst(y|x0)δ(x− x0). (4)

It is evident that solving for either the time-dependentdistribution P (x, t|x0, 0) or the stationary distributionPst(x|x0) from Eqs. (3) and (4), respectively, is aformidable task even with F = 0, unless the functionr(x) has simple forms. For example, in Ref. [17], theauthors considered a solvable example with F (x) = 0,where the function r(x) is zero in a window around x0

and is constant outside the window.In this work, we employ a different approach to solve

for the stationary spatial distribution, by invoking thepath integral formalism of quantum mechanics and byusing elements of the theory of renewals. In this ap-proach, we compute Pst(x|x0), the stationary distribu-tion in presence of reset events, in terms of suitably-defined functions that take into account the occurrenceof trajectories that evolve without undergoing any resetevents in a given time, see Eq. (30) below. This ap-proach provides a viable alternative to obtaining the sta-tionary spatial distribution by solving the Fokker-Planckequation (4) that explicitly takes into account the occur-rence of trajectories that evolve while undergoing resetevents in a given time. As we will demonstrate below,the method allows to obtain exact expressions even incases with nontrivial forms of F (x) and r(x).

B. Path-integral approach to stochastic resetting

Here, we invoke the well-established path-integral ap-proach based on the Feynman-Kac formalism to discussstochastic resetting. To proceed, let us first consider arepresentation of the dynamics in discrete times ti = i∆t,with i = 0, 1, 2, . . ., and ∆t > 0 being a small time step.The dynamics in discrete times involves the particle at

position xi at time ti to either reset and be at x(r) at thenext time step ti+1 with probability r(xi)∆t or follow thedynamics given by Eq. (1) with probability 1− r(xi)∆t.The position of the particle at time ti is thus given by

xi =

{xi−1 + ∆t

(µF (xi) + ηi

)with prob. 1− r(xi−1)∆t,

x(r) with prob. r(xi−1)∆t,(5)

where we have defined F (xi) ≡ (F (xi−1) + F (xi))/2,and have used the Stratonovich rule in discretizing thedynamics (1), and where the time-discretized Gaussian,white noise ηi satisfies

〈ηiηj〉 = σ2δij , (6)

with σ2 a positive constant with the dimension of length-squared over time-squared. In particular, the joint prob-ability distribution of occurrence of a given realization{ηi}1≤i≤N of the noise, with N being a positive integer,is given by

P [{ηi}] =

(1

2πσ2

)N/2exp

(− 1

2σ2

N∑i=1

η2i

). (7)

In the absence of any resetting and forces, the displace-ment of the particle at time t ≡ N∆t from the initial

location is given by ∆x ≡ xN −x0 = ∆t∑Ni=1 ηi, so that

the mean-squared displacement is 〈(∆x)2〉 = σ2N(∆t)2.In the continuous-time limit, N → ∞,∆t → 0, keep-ing the product N∆t fixed and finite and equal to t,the mean-squared displacement becomes 〈(∆x)2〉 = 2Dt,with D ≡ limσ→∞,∆t→0 σ

2∆t/2.

1. The propagator prior to first reset.

What is the probability of occurrence of particle tra-jectories that start at position x0 and end at a givenlocation x at time t = N∆t without having under-gone any reset event? From the discrete-time dynam-ics given by Eq. (5) and the joint distribution (7), theprobability of occurrence of a given particle trajectory{xi}0≤i≤N ≡ {x0, x1, x2, . . . , xN−1, xN = x} is given by

Pno res[{xi}] = det(J )

(1

2πσ2

)N/2×

N∏i=1

exp

(− (xi − xi−1 − µF (xi)∆t)

2

2σ2(∆t)2

)N−1∏i=0

(1− r(xi)∆t

).

(8)

Here, the factor∏N−1i=0 (1− r(xi)∆t) enforces the condi-

tion that the particle has not reset at any of the instantsti, i = 0, 1, 2, . . . , N − 1, while J is the Jacobian matrixfor the transformation {ηi} → {xi}, which is obtained

Page 4: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

4

from Eq. (5) as J1≤i,j≤N ≡(∂ηi∂xj

)or equivalently

J =

1

∆t −µF ′(x1)

2 0 0 . . .

− 1∆t −

µF ′(x1)2

1∆t −

µF ′(x2)2 0 . . .

......

......

N×N

,

(9)with primes denoting derivative with respect to x. Onethus has

det(J ) =

(1

∆t

)N N∏i=1

(1− ∆tµF ′(xi)

2

)

'(

1

∆t

)Nexp

(−

N∑i=1

∆tµF ′(xi)

2

), (10)

where in obtaining the last step, we have used the small-ness of ∆t. Thus, for small ∆t, we get

Pno res[{xi}] =

(1

2πσ2(∆t)2

)N/2×

N∏i=1

exp

(− (xi − xi−1 − µF (xi)∆t)

2

2σ2(∆t)2− ∆tµF ′(xi)

2

)

×N−1∏i=0

exp(− r(xi)∆t

)=

(1

2πσ2(∆t)2

)N/2exp

(∆t[r(xN )− r(x0)

])× exp

(−∆t

N∑i=1

[ [(xi − xi−1 − µF (xi)∆t)/∆t]2

2σ2∆t

+µF ′(xi)

2+ r(xi)

]).

(11)

From Eq. (11), it follows by considering all possibletrajectories that the probability density that the particlewhile starting at position x0 ends at a given location x attime t = N∆t without having undergone any reset eventis given by

Pno res(x, t|x0, 0) =

(1

2πσ2(∆t)2

)N/2exp

(∆t[r(x)− r(x0)

])×N−1∏i=1

∫ ∞−∞

dxi exp(−∆t

N∑i=1

[ [(xi − xi−1 − µF (xi)∆t)/∆t]2

2σ2∆t

+µF ′(xi)

2+ r(xi)

]). (12)

In the limit of continuous time, defining Dx(t) ≡

limN→∞

(1

4πD∆t

)N/2∏N−1i=1

∫∞−∞ dxi, one gets the exact

expression for the corresponding probability density asthe following path integral:

Pno res(x, t|x0, 0) =

∫ x(t)=x

x(0)=x0

Dx(t) exp (−Sres[{x(t)}]) ,

(13)

where on the right hand side of Eq. (13), we have intro-duced the resetting action as

Sres[{x(t)}] =

∫ t

0

dt

[[(dx/dt)− µF (x))2

4D+µF ′(x)

2+ r(x)

].

(14)Invoking the Feynman-Kac formalism, we identify thepath integral on the right hand side of Eq. (13) with thepropagator of a quantum mechanical evolution in (nega-tive) imaginary time due to a quantum Hamiltonian Hq

(see Appendix), to get

Pno res(x, t|x0, 0) = exp

2D

∫ x

x0

F (x) dx

)Gq(x,−it|x0, 0),

(15)with

Gq(x,−it|x0, 0) ≡ 〈x| exp(−Hqt)|x0〉, (16)

where the quantum Hamiltonian is

Hq ≡ −1

2mq

∂2

∂x2+ Vq(x), (17)

the mass in the equivalent quantum problem is

mq ≡1

2D, (18)

and the quantum potential is given by

Vq(x) ≡ µ2(F (x))2

4D+µF ′(x)

2+ r(x). (19)

Note that in the quantum propagator in Eq. (16), thePlanck’s constant has been set to unity, h = 1, while thetime τ of propagation is imaginary: τ = −it [46]. Sincethe Hamiltonian contains no explicit time dependence,the propagator Gq(x,−it|x0, 0) is effectively a functionof the time t to propagate from the initial location x0

to the final location x, and not individually of the initialand final times. Let us note that on using D = kBTµ,the prefactor equals exp (−Q(t)/2kBT ), where Q(t) ≡∫ xx0∂xV (x) dx is the heat absorbed by the particle from

the environment along the trajectory {x(t)} [47, 48].

2. Distribution of the first-reset time

Let us now ask for the probability of occurrence oftrajectories that start at position x0 and reset for thefirst time at time t. In terms of Pno res(x, t|x0, 0), onegets this probability density as

Pres(t|x0) =

∫ ∞−∞

dy r(y)Pno res(y, t|x0, 0), (20)

since by the very definition of Pres(t|x0), a reset has tohappen only at the final time t when the particle hasreached the location y, where y may in principle take anyvalue in the interval [−∞,∞]. The probability densityPres(t|x0) is normalized as

∫∞0

dt Pres(t|x0) = 1.

Page 5: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

5

3. Spatial time-dependent probability distribution

Using renewal theory, we now show that knowingPno res(x, t|x0, 0) and Pres(t|x0) is sufficient to obtain thespatial distribution of the particle at any time t. Theprobability density that the particle is at x at time twhile starting from x0 is given by

P (x, t|x0, 0) = Pno res(x, t|x0, 0)

+

∫ t

0

∫ ∞−∞

dy r(y)P (y, t− τ |x0, 0)Pno res(x, t|x0, t− τ)

= Pno res(x, t|x0, 0)

+

∫ t

0

dτ R(t− τ |x0)Pno res(x, t|x0, t− τ), (21)

where we have defined the probability density to reset attime t as

R(t|x0) ≡∫ ∞−∞

dy r(y)P (y, t|x0, 0). (22)

One may easily understand Eq. (21) by invoking the the-ory of renewals [40] and realizing that the dynamics isrenewed each time the particle resets to x0. This may beseen as follows. The particle while starting from x0 mayreach x at time t by experiencing not a single reset; thecorresponding contribution to the spatial distribution isgiven by the first term on the right hand side of Eq. (21).The particle may also reach x at time t by experienc-ing the last reset event (i.e., the last renewal) at timeinstant t− τ , with τ ∈ [0, t], and then propagating fromthe reset location x(r) = x0 to x without experiencingany further reset, where the last reset may take placewith rate r(y) from any location y ∈ [−∞,∞] where theparticle happened to be at time t− τ ; such contributionsare represented by the second term on the right handside of Eq. (21). The spatial distribution is normalizedas∫∞−∞ dx P (x, t|x0, 0) = 1 for all possible values of x0

and t.Multiplying both sides of Eq. (21) by r(x), and then

integrating over x, we get

R(t|x0) =

∫ ∞−∞

dx r(x)Pno res(x, t|x0, 0)

+

∫ t

0

[∫ ∞−∞

dx r(x)Pno res(x, t|x0, t− τ)

]R(t− τ |x0).

(23)

The square-bracketed quantity on the right hand side isnothing but Pres(τ |x0), so that we get

R(t|x0) = Pres(t|x0)+

∫ t

0

dτ Pres(τ |x0)R(t−τ |x0). (24)

Taking the Laplace transform on both sides of Eq. (24),we get

R(s|x0) = Pres(s|x0) + Pres(s|x0)R(s|x0), (25)

where R(s|x0) and Pres(s|x0) are respectively the Laplace

transforms of R(t|x0) and Pres(t|x0). Solving for R(s|x0)from Eq. (25) yields

R(s|x0) =Pres(s|x0)

1− Pres(s|x0). (26)

Next, taking the Laplace transform with respect to timeon both sides of Eq. (21), we obtain

P (x, s|x0, 0) =(

1 + R(s|x0))Pno res(x, s|x0)

=Pno res(x, s|x0)

1− Pres(s|x0), (27)

where we have used Eq. (26) to obtain the last equality.An inverse Laplace transform of Eq. (27) yields the time-dependent spatial distribution P (x, t|x0, 0).

4. Stationary spatial distribution

On applying the final value theorem, one may obtainthe stationary spatial distribution as

Pst(x|x0) = lims→0

sP (x, s|x0, 0) = lims→0

sPno res(x, s|x0)

1− Pres(s|x0),

(28)provided the stationary distribution (i.e.,limt→∞ P (x, t|x0, 0)) exists. Now, since Pres(t|x0)is normalized to unity,

∫∞0

dt Pres(t|x0) = 1, wemay expand its Laplace transform to leading orders

in s as Pres(s|x0) ≡∫∞

0dt exp(−st)Pres(t|x0) =

1 − s〈t〉res + O(s2), provided that the mean first-resettime 〈t〉res, defined as

〈t〉res ≡∫ ∞

0

dt tPres(t|x0), (29)

is finite. Similarly, we may expand Pno res(x, s|x0, 0)

to leading orders in s as Pno res(x, s|x0, 0) =∫∞0

dt Pno res(x, t|x0, 0) − s∫∞

0dt tPno res(x, t|x0, 0) +

O(s2), provided that∫∞

0dt tPno res(x, t|x0, 0) is finite.

From Eq. (28), we thus find the stationary spatial dis-tribution to be given by the integral over all times ofthe propagator prior to first reset divided by the meanfirst-reset time:

Pst(x|x0) =1

〈t〉res

∫ ∞0

dt Pno res(x, t|x0, 0). (30)

III. EXACTLY SOLVED EXAMPLES

A. Free particle with space-independent resetting

Let us first consider the simplest case of free diffu-sion with a space-independent rate of resetting r(x) = r,

Page 6: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

6

with r a positive constant having the dimension of in-verse time. Here, on using Eq. (15) with F (x) = 0, wehave

Pno res(x, t|x0, 0) = Gq(x,−it|x0, 0)

= 〈x| exp(−Hqt)|x0〉, (31)

where the quantum Hamiltonian is in this case, followingEqs. (17-19), given by

Hq = − 1

2mq

∂2

∂x2+ r; mq =

1

2D, h = 1. (32)

Since in the present situation, the effective quantum po-tential Vq(x) = r is space independent, we may rewriteEq. (31) as:

Pno res(x, t|x0, 0) = exp(−rt)Gq(x,−it|x0, 0), (33)

with

Gq(x,−it|x0, 0) ≡ 〈x| exp(−Hqt)|x0〉, (34)

where the quantum Hamiltonian is now that of a freeparticle:

Hq ≡ −1

2mq

∂2

∂x2; mq =

1

2D, h = 1. (35)

Therefore, the statistics of resetting of a free particle un-der a space-independent rate of resetting may be foundfrom the quantum propagator of a free particle, which isgiven by [42]

Gq(x, τ |x0, 0) =

√mq

2πhiτexp

(−mq(x− x0)2

2hiτ

). (36)

Plugging in Eq. (36) the parameters in Eq. (35) togetherwith τ = −it, we have

Gq(x,−it|x0, 0) =1√

4πDtexp

(− (x− x0)2

4Dt

). (37)

Using Eq. (37) in Eq. (33), we thus obtain

Pno res(x, t|x0, 0) =exp(−rt)√

4πDtexp

(− (x− x0)2

4Dt

), (38)

and hence, the distribution of the first-reset time may befound on using Eq. (20):

Pres(t|x0) = r exp(−rt) 1√4πDt

∫ ∞−∞

dx exp

(− (x− x0)2

4Dt

)= r exp(−rt), (39)

which is normalized to unity:∫∞

0dt Pres(t|x0) = 1, as

expected.

Using Eq. (39), we get Pres(s|x0) = r/(s+ r), so that

Eq. (26) yields R(s|x0) = r/s. An inverse Laplace trans-form yields R(t|x0) = r, as also follows from Eq. (22)

by substituting r(y) = r and noting that P (y, t|x0, 0) isnormalized with respect to y.

Next, the probability density that the particle is at xat time t, while starting from x0, is obtained on usingEq. (21) as

P (x, t|x0, 0) =exp(−rt)√

4πDτexp(−(x− x0)2/(4Dt))

+ r

∫ t

0

dτexp(−rτ)√

4πDτexp(−(x− x0)2/(4Dτ)).

(40)

Taking the limit t→∞, we obtain the stationary spatialdistribution as

Pst(x|x0) = r

∫ ∞0

dτ exp(−rτ)exp(−(x− y)2/(4Dτ))√

4πDτ,

(41)

which may also be obtained by using Eqs. (30) and (38),and also Eq. (39) that implies that 〈t〉res = 1/r. FromEq. (40), we obtain an exact expression for the time-dependent spatial distribution as

P (x, t|x0, 0) =exp(−rt) exp(−(x− x0)2/4Dt)√

4πDt

+

exp

(− |x−x0|√

D/r

)erfc

(|x−x0|√

4Dt−√rt)

√4D/r

−exp

(|x−x0|√D/r

)erfc

(|x−x0|√

4Dt+√rt)

√4Dt

, (42)

while Eq. (41) yields the exact stationary distribution as

Pst(x|x0) =1

2√D/r

exp

(−|x− x0|√

D/r

), (43)

where erfc(x) ≡ (2/√π)∫∞x

dt exp(−t2) is the comple-mentary error function. The stationary distribution (43)may be put in the scaling form

Pst(x|x0) =1

2√D/rR

(|x− x0|√D/r

), (44)

where the scaling function is given by R(y) ≡ exp(−y).For the particular case x0 = 0, Eq. (43) matches withthe result derived in Ref. [16]. Note that the steady statedistribution (44) exhibits a cusp at the resetting locationx0. Since the resetting location is taken to be the sameas the initial location, the particle visits repeatedly intime the initial location, thereby keeping a memory ofthe latter that makes an explicit appearance even in thelong-time stationary state.

Page 7: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

7

B. Free particle with “parabolic” resetting

We now study the dynamics of a free Brownian parti-cle whose position is reset to the initial position x0 witha rate of resetting that is proportional to the square ofthe current position of the particle. In this case, wehave r(x) = αx2, with α > 0 having the dimension of1/((Length)2Time). From Eqs. (15) and (16), and giventhat in this case F (x) = 0, we get

Pno res(x, t|x0, 0) = Gq(x,−it|x0, 0) = 〈x| exp(−Hqt)|x0〉,(45)

with the Hamiltonian obtained from Eq. (17) by settingVq(x) = αx2:

Hq = − 1

2mq

∂2

∂x2+ αx2; mq =

1

2D, h = 1. (46)

We thus see that the statistics of resetting of a free par-ticle subject to a “parabolic” rate of resetting may befound from the propagator of a quantum harmonic os-cillator. Following Schulman [42], a quantum harmonicoscillator with the Hamiltonian given by

Hq = − 1

2mq

∂2

∂x2+

1

2mqω

2qx

2, (47)

with mq and ωq being the mass and the frequency of theoscillator, has the quantum propagator

Gq(x, τ |x0, 0) =√mqωq

2iπh sinωqτexp

(iωq

2h sinωqτ[(x2 + x2

0) cosωqτ − 2xx0]

).

(48)

Using the parameters given in Eq. (46), and substituting

τ = −it and ωq =√

4Dα in Eq. (48), we have

Gq(x,−it|x0, 0) =(α/D)1/4√

2π sinh(t√

4Dα)

× exp

(−

√α/D

2 sinh(t√

4Dα)[(x2

0 + y2) cosh(t√

4Dα)− 2xx0]

).

(49)

We may now derive the statistics of resetting by us-ing the propagator (49). Equation (45) together withEq. (49) imply

Pno res(x, t|x0, 0) =(α/D)1/4√

2π sinh(t√

4Dα)

× exp

(−

√α/D

2 sinh(t√

4Dα)[(x2

0 + x2) cosh(t√

4Dα)− 2x0x]

).

(50)

Integrating Eq. (50) over x, we get the distribution of thefirst-reset time as

Pres(t|x0) =

∫ ∞−∞

dy r(y)Pno res(y, t|x0, 0)

=α(α/D)1/4√sinh(t

√4Dα)

exp

(−x2

0

√α/D

2tanh(t

√4Dα)

)

×√α/D coth(t

√4Dα) + x2

0(α/D)cosech2(t√

4Dα)

(α/D)5/4 coth5/2(t√

4Dα).

(51)

For the case x0 = 0, Eqs. (50) and (51) reduce tosimpler expressions:

Pno res(x, t|x0 = 0, 0) =

(α/D)1/4√2π sinh(t

√4Dα)

exp

(−x2√α/D coth(t

√4Dα)

2

),

(52)

and

Pres(t|x0 = 0) =√Dα

(tanh(t√

4Dα))3/2√sinh(t

√4Dα)

. (53)

Equation (53) may be put in the scaling form

Pres(t|x0 = 0) =√DαG

(t√

4Dα), (54)

with G(y) = tanh(y)3/2/√

sinh(y). Equation (53) yieldsthe mean first-reset time 〈t〉res for x0 = 0 to be given by

〈t〉res =(Γ(1/4))2

4√

2πDα, (55)

where Γ is the Gamma function. Equations (53) and (55)yield the stationary spatial distribution on using Eq. (30):

Pst(x|x0 = 0) =4√Dα(α/D)1/4

(Γ(1/4))2

×∫ ∞

0

dt√sinh(t

√4Dα)

exp

(−x2√α/D coth(t

√4Dα)

2

).

=23/4(α/D)1/4

√πΓ(1/4)

(x2√α/D

2

)1/4

K1/4

(x2√α/D

2

),(56)

where Kn(x) is the n−th order modified Bessel functionof the second kind. Equation (56) implies that the sta-tionary distribution is symmetric around x = 0, whichis expected since the resetting rate is symmetric aroundx0 = 0. The stationary distribution (56) may be put inthe scaling form

Pst(x|x0 = 0) =23/4(α/D)1/4

√πΓ(1/4)

R(

x

(D/α)1/4

), (57)

Page 8: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

8

0

0.1

0.2

0.3

0.4

0.5

-5 -3 -1 0 1 3 5

SimulationsTheory

FIG. 2: Theory versus simulation results for the sta-tionary spatial distribution of a free Brownian parti-cle undergoing parabolic resetting. The points denotesimulation results, while the line stands for the exact analyt-ical expression given in Eq. (56). The numerical results wereobtained from 104 independent simulations of the Langevindynamics described in Section II A. The error bars associatedwith the data points are smaller than the symbol size. Theparameter values are D = 1.0, α = 0.5.

where the scaling function is given by R(y) =(y2/2)1/4K1/4(y2/2).

The result (56) is checked in simulations in Fig. 2. Thesimulations involved numerically integrating the dynam-ics described in Section II A, with integration timestep

equal to 0.01. Using∫∞

0dt tµ−1Kν(t) = 2µ−2Γ

(µ2 −

ν2

)Γ(µ2 + ν

2

)for |Re ν| < |Re µ| [49], we find that

Pst(x|x0 = 0) given by Eq. (56) is correctly normalized tounity. Moreover, using the results that as x→ 0, we have

Kν(x) = Γ(ν)2

(x2

)−νfor Re ν > 0 and that as x → ∞,

we have Kν(x) =(π2x

)1/2

exp(−x) for real x [49], we get

Pst(x|x0 = 0) ∼

{(α/D)1/4√

πfor x→ 0,

exp(−x2√α/D/2) for |x| → ∞.

(58)

Using Eq. (56) and the result dK1/4(x)/dx =

−(1/2)(K3/4(x) +K5/4(x)

), it may be easily shown

that as x → 0±, one has dPst(x|x0 = 0))/dx =

∓√α/DΓ(3/4)/(

√πΓ(1/4)), implying thereby that the

first derivative of Pst(x|x0 = 0) is discontinuous atx = 0. We thus conclude that the spatial distributionPst(x|x0 = 0) exhibits a cusp singularity at x = 0. Thisfeature of cusp singularity at the resetting location x0 = 0is also seen in the stationary distribution (43), and isa signature of the steady state being a nonequilibriumone [16, 21, 32, 33]. Note the existence of faster-than-exponential tails suggested by Eq. (58) in comparison tothe exponential tails observed in the case of resetting ata constant rate, see Eq. (43). This is consistent withthe fact that with respect to the case of resetting at aspace-independent rate, a parabolic rate of resetting im-plies that the further the particle is from x0 = 0, the

more enhanced is the probability that a resetting eventtakes place, and, hence, a smaller probability of findingthe particle far away from the resetting location.

Let us consider the case of an overdamped Brownianparticle that is trapped in a harmonic potential V (x) =(1/2)κx2, with κ > 0, and is undergoing the Langevindynamics (1). At equilibrium, the distribution of theposition of the particle is given by the Boltzmann-Gibbsdistribution

Peq(x) = exp(−κx2/2kBT )/Z, (59)

with Z =√

2πkBT/κ being the partition function. Com-paring Eqs. (58) and (59), we see that using a harmonicpotential with a suitable κ, the stationary distribution ofa free Brownian particle undergoing parabolic resettingmay be made to match in the tails with the stationarydistribution of a Brownian particle trapped in the har-monic potential and evolving in the absence of any re-setting. On the other hand, the cusp singularity in theformer cannot be achieved with the Langevin dynamicsin any harmonic potential without the inclusion of reset-ting events.

Let us note that the stationary states (43) and (56)are entirely induced by the dynamics of resetting. In-deed, in the absence of any resetting, the dynamics of afree diffusing particle does not allow for a long-time sta-tionary state, since in the absence of a force, there is noway in which the motion of the particle can be boundedin space. On the other hand, in presence of resetting,the dynamics of repeated relocation to a given positionin space can effectively compete with the inherent ten-dency of the particle to spread out in space, leading to abounded motion, and, hence, a relaxation to a stationaryspatial distribution at long times. In the next section, weconsider the situation where the particle even in the ab-sence of any resetting has a localized stationary spatialdistribution, and investigate the change in the nature ofthe spatial distribution of the particle owing to the in-clusion of resetting events.

C. Particle trapped in a harmonic potential withenergy-dependent resetting

We now introduce a resetting problem that is relevantin physics: an overdamped Brownian particle immersedin a thermal bath at temperature T and trapped witha harmonic potential centered at the origin: V (x) =(1/2)κx2, where κ > 0 is the stiffness constant of the har-monic potential. The particle, initially located at x0 = 0,may be reset at any time t to the origin with a probabil-ity that depends on the energy of the particle at time t.The dynamics is shown schematically in Fig. 3. For pur-poses of illustration of the nontrivial effects of resetting,we consider the following space-dependent reseting rate:

r(x) =3

2τc

V (x)

kBT=

3

4

µ2κ2

Dx2, (60)

Page 9: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

9

reset

FIG. 3: Illustration of the energy-dependent resettingof a Brownian particle moving in a harmonic poten-tial. A Brownian particle (grey circle) immersed in a ther-mal bath at temperature T moves with diffusion coefficientD, with its motion being confined by a harmonic potentialV (x) = κx2/2 (green), with κ being the stiffness constant.Here, x is the position of the particle with respect to the cen-ter of the potential. The particle, initially located in the trapcenter (t′ = 0, left panel), diffuses at subsequent times inthe energy landscape (t′ < t, middle panel), until a resettingevent occurs at time t′ = t (right panel). The black curverepresents the history of the particle from t′ = 0 up to thetime corresponding to each snapshot. The rate of resetting(right colorbar) is proportional to the instantaneous energyof the particle, and, therefore, a reset is more likely to takeplace as the particle climbs up the potential.

where we use D = kBTµ in obtaining the second equality.Note that the resetting rate is proportional to the energyof the particle (in units of kBT ) divided by the timescaleτc ≡ 1/µκ that characterizes the relaxation of the particlein the harmonic potential in the absence of any resetting.In this way, it is ensured that the rate of resetting (60)has units of inverse time. Note also that in the absenceof any resetting, the particle relaxes to an equilibriumstationary state with a spatial distribution given by theusual Boltzmann-Gibbs form:

Pr(x)=0st (x) =

√κ

2πkBTexp

(− κx2

2kBT

). (61)

Using F (x) = −∂xV (x) = −κx and the expression (60)for the resetting rate in Eq. (19), we find that the poten-tial of the corresponding quantum mechanical problem isgiven by

Vq(x) =µ2(F (x))2

4D+µF ′(x)

2+ r(x) =

µ2κ2x2

D− µκ

2,

(62)where we have used F ′(x) = −κ. From Eqs. (15)and (16), we obtain

Pno res(x, t|x0 = 0, 0) = exp

2D

∫ x

x0

F (x) dx

)× exp

(µκt

2

)〈x| exp(−Hqt)|x0 = 0〉

= exp

(− x2

4Dτc

)exp

(t

2τc

)〈x| exp(−Hqt)|x0 = 0〉,

(63)

where the quantum Hamiltonian is given by

Hq = − 1

2mq

∂2

∂x2+µ2κ2x2

D; mq =

1

2D, h = 1. (64)

We thus find that the propagator 〈x| exp(−Hqt)|x0 =0〉 is given by the propagator of a quantum harmonicoscillator, which has been calculated in Sec. III B. In fact,the Hamiltonian given by Eq. (64) is identical to that inEq. (46) with the identification α = µ2κ2/D = 1/Dτ2

c , sothat by substituting x0 = 0 and α = 1/(Dτ2

c ) in Eq. (50),we obtain

〈x| exp(−Hqt)|x0 = 0〉 =

1√2πDτc sinh(2t/τc)

exp

(− x2

2Dτccoth(2t/τc)

).

(65)

From Eqs. (63) and (65), we obtain

Pno res(x, t|x0 = 0, 0) =

exp(t/2τc)√2πDτc sinh(2t/τc)

exp

(− x2

2Dτc

[1

2+ coth(2t/τc)

]).

(66)

Following Eq. (20), we may now calculate the proba-bility of the first-reset time by using Eq. (66) to get

Pres(t|x0 = 0) =exp(t/2τc)√

2πDτc sinh(2t/τc)

×∫ ∞−∞

dx3x2

4Dτ2c

exp

(− x2

2Dτc

[1

2+ coth(2t/τc)

])=

3 exp(t/2τc)

4τc

√1

sinh(2t/τc)(1/2 + coth(2t/τc))3,

(67)

which may be checked to be normalized:∫∞0

dt Pres(t|x0 = 0) = 1. The first-reset timedistribution (67) may be written in the scaling form

Pres(t|x0 = 0) =3

4τcG(

2t

τc

), (68)

with the scaling function given by G(y) =exp(y/4) sinh(y)−1/2(1/2 + coth(y))−3/2.

The mean first-reset time, given by 〈t〉res ≡∫∞0

dtPres(t|x0 = 0), equals

〈t〉res =4τc√

32F1

(1

8,

1

2;

9

8;−1

3

), (69)

where pFq(a1, a2, . . . , ap; b1, b2, . . . , bq;x) is the general-ized hypergeometric function. Introducing the variablez ≡ 2t/τc, and using Eq. (66), we get

Pst(x|x0 = 0) =1

〈t〉res

√τc

8πDexp

(− x2

4Dτc

)

Page 10: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

10

0

0.2

0.4

0.6

0.8

-3 -2 -1 0 1 2 3

SimulationsTheory

FIG. 4: Theory versus simulation results for thestationary spatial distribution of a Brownian parti-cle trapped in a harmonic potential and undergoingenergy-dependent resetting. The points denote simula-tion results, while the line stands for the exact analyticalexpression given in Eq. (71). The numerical results wereobtained from 104 independent simulations of the Langevindynamics described in Section II A. The error bars associatedwith the data points are smaller than the symbol size. Theparameter values are D = 1.0, τc = 0.5.

×∫ ∞

0

dzexp(z/4)√

sinh(z)exp

(−x

2 coth z

2Dτc

)=

1

〈t〉res

√τc

8πDexp

(− x2

4Dτc

)×1

2

(x2

4Dτc

)−1/4

Γ(1/8)W1/8,1/4

(x2

Dτc

), (70)

where Wµ,ν is Whittaker’s W function.Using Eq. (69) in Eq. (70), we obtain

Pst(x|x0 = 0) =(1/8)Γ(1/8)

2F1

(18 ,

12 ; 9

8 ;− 13

)×√

3

8πDτcexp

(− x2

4Dτc

)(4Dτcx2

)1/4

W1/8,1/4

(x2

Dτc

),

(71)

which may be checked to be normalized to unity. Wemay write the stationary distribution in terms of a scaledposition variable as

Pst(x|x0 = 0) =(1/8)Γ(1/8)

2F1

(18 ,

12 ; 9

8 ;− 13

)√ 3

8πDτcR(

x√2Dτc

),

(72)

with R(y) ≡ exp(−y2/2)(2/y2)1/4W1/8,1/4(2y2) beingthe scaling function. The expression (71) is checked insimulations in Fig. 4. The simulations involved numeri-cally integrating the dynamics described in Section II A,with integration timestep equal to 0.01. Using the results

that as x→ 0, we have Wk,µ(x) = Γ(2µ)Γ(1/2+µ−k)x

1/2−µ for

0 ≤ Re µ < 1/2, µ 6= 0 and that as x → ∞, we have

Wk,µ(x) ∼ e−x/2xk for real x [49], we get

Pst(x|x0 = 0) ∼

(1/8)Γ(1/8)

2F1(1/8,1/2;9/8;−1/3)

√3/8Dτc

Γ(5/8) for x→ 0,

exp(− 3x2

4Dτc

)for |x| → ∞.

We see again the existence of a cusp at the resettinglocation x0 = 0, similar to all other cases we studied inthis paper.

The results of this subsection could inspire future ex-perimental studies using optical tweezers in which theresetting protocol could be effectively implemented by us-ing feedback control [3, 5, 6]. Interestingly, colloidal andmolecular gelly and glassy systems show hopping motionof their constituent particles between potential traps or“cages,” the latter originating from the interaction of theparticles with their neighbors [50]. Such a phenomenon isalso exhibited by out-of-equilibrium glasses and gels dur-ing the process of aging [51]. Our results in this sectioncould provide valuable insights into the aforementioneddynamics, since the emergent potential cages may be wellapproximated by harmonic traps and the hopping processas a resetting event.

IV. SHORTCUTS TO CONFINEMENT

A hallmark of the examples solved exactly in Sec. III byusing our path-integral formalism is the existence of sta-tionary distributions with prominent cusp singularities(see Figs. 2 and 4). These examples demonstrate thatthe particle can be confined around a prescribed locationby using appropriate space-dependent rates of resetting.

In physics and nanotechnology, the issue of achievingan accurate control of fluctuations of small-sized parti-cles is nowadays attracting considerable attention [52–54]. For instance, using optical tweezers and noisy elec-trostatic fields, it is now possible to control accuratelythe amplitude of fluctuations of the position of a Brow-nian particle [55, 58, 59]. Such fluctuations may becharacterized by an effective temperature. Experimentshave reported effective temperatures of a colloidal par-ticle in water up to 3000K [52], and have recently beenused to design colloidal heat engines at the mesoscopicscale [55, 60]. Effective confinement of small systems isof paramount importance for success of quantum-basedcomputations with, e.g., cold atoms [56, 57].

Does stochastic resetting provide an efficient way to re-duce the amplitude of fluctuations of a Brownian particle,thereby providing a technique to reduce the associated ef-fective temperature? We now provide some insights intothis question.

Consider the following example of a nonequilib-rium protocol: i) a Brownian particle is initially con-fined in a harmonic trap with a potential V (x) =(1/2)κx2 for a sufficiently long time such that itis in an equilibrium state with spatial distributionPeq(x) = exp(−κx2/2kBT )/Z, with Z =

√2πkBT/κ;

Page 11: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

11

ii) a space-dependent (parabolic) resetting rate r(x) =(3/2τc)V (x)/kBT , with τc = (1/µκ), is suddenlyswitched on by an external agent. In order words,the rate of resetting is instantaneously quenched fromr(x) = 0 to r(x) = (3/4)(µ2κ2/D)x2; iii) the particle islet to relax to a new stationary state in the presence ofthe trapping potential and parabolic resetting. At theend of the protocol, the particle relaxes to the stationarydistribution Pst(x) given by Eq. (72).

0 0.5 1 1.5 2

0.6

0.7

0.8

0.9

1

-5 0 50

5

10

-5 0 50

5

10

-5 0 50

5

10

Time

FIG. 5: Shortcut to particle confinement. Confinementof a Brownian particle using harmonic potentials and space-dependent rates of resetting. The symbols represent simu-lation results for relaxation processes of 105 non-interactingBrownian particles that are initially in equilibrium in a har-monic potential V (x) = (1/2)κx2. The blue circles repre-sent the mean-squared displacement of the particle in unitsof 〈x2(0)〉 after an instantaneous quench to a harmonic po-tential with stiffness κ′ = 1.7κ (the blue arrow in the in-set). The yellow circles on the other hand represent themean-squared displacement of the particle in units of 〈x2(0)〉after an instantaneous quench of the rate of resetting fromr(x) = 0 to r(x) = (3/4)(µ2κ2/D)x2 (the orange arrowin the inset). For the case in which the stiffness of thepotential is quenched from κ to κ′, it is easily seen that〈x2(t)〉 = 〈x2(0)〉 exp(−2µk′t) + (D/(µk′))[1 − exp(−2µk′t)],thereby implying a relaxation timescale τquench = 1/(2µk′)and yielding the corresponding curve in the figure. Note thatthe time in the x-axis is measured in units of τc = 1/µκ.The other curve depicting the process of relaxation in pres-ence of resetting may be fitted to a good approximation toA + Be−t/τreset . One observes that τreset ≈ τquench/3. Theparameter values are D = 1, κ = 1 and µ = 10.

We first note that before the sudden switching on ofthe resetting dynamics, which we assume to happen at areference time instant t = 0, the mean-squared displace-ment of the particle is given by

〈x2(0)〉 =kBT

κ, (73)

which follows from the equilibrium distribution beforethe resetting is switched on, and is in agreement withthe equipartition theorem κ〈x2(0)〉/2 = kBT/2. Afterthe sudden switching on of the space-dependent rate of

resetting, the variance of the position of the particle re-laxes at long times to the stationary value

〈x2(∞)〉 =kBT/κ√

3 2F1

(18 ,

12 ; 9

8 ;− 13

) ' 0.59kBT

κ, (74)

as follows from Eq. (72). The resetting dynamics in-duces in this case a reduction by about 40% of the vari-ance of the position of the particle with respect to itsinitial value. We note that such a reduction of the am-plitude of fluctuations of the particle could also have beenachieved by performing a sudden quench of the stiffnessof the harmonic potential by increasing its value from κto κ′ ' (1/0.59)κ ' 1.7κ, without the need for switchingon of resetting events. To understand the difference be-tween the two scenarios, it is instructive to compare thetime evolution of the mean-squared displacement 〈x2(t)〉towards the stationary value in the two cases, see insetin Fig. 5. We observe that resetting leads to the samedegree of confinement in a shorter time. For the case inwhich the stiffness of the potential is quenched from κ toκ′, it is easily seen that 〈x2(t)〉 = 〈x2(0)〉 exp(−2µk′t) +(D/(µk′))[1 − exp(−2µk′t)], thereby implying a relax-ation timescale τquench = 1/(2µk′) and yielding the cor-responding curve in Fig. 5. The other curve depictingthe process of relaxation in presence of resetting may befitted to a good approximation to A + Be−t/τreset . Oneobserves that τreset ≈ τquench/3. Thus, for the exampleat hand, we may conclude that a sudden quench of re-setting profiles provides a shortcut to confinement of theposition of the particle to a desired degree with respectto a potential quench. Similar conclusions were arrivedat for mean first-passage times of resetting processes andequivalent equilibrium dynamics [61].

It may be noted that the confinement protocol bya sudden quench of resetting profiles introduced aboveis amenable to experimental realization. Using micro-scopic particles trapped with optical tweezers [55, 59]or feedback traps [58, 62, 63], it is now possible tomeasure and control the position of a Brownian parti-cle with subnanometric precision. Recent experimen-tal setups allow to exert random forces to trappedparticles, with a user-defined statistics for the randomforce [52, 53, 55, 64]. The shortcut protocol using reset-ting could be explored in the laboratory by designing afeedback-controlled experiment with optical tweezers andby employing random-force generators according to, e.g.,the protocol sketched in Fig. 3.

V. CONCLUSIONS AND OUTLOOK

In this paper, we addressed the fundamental questionof what happens when a continuously evolving stochasticprocess is repeatedly interrupted at random times by asudden reset of the state of the process to a given fixedstate. To this end, we studied the dynamics of an over-damped Brownian particle diffusing in force fields and

Page 12: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

12

resetting to a given spatial location with a rate that hasan essential dependence on space, namely, the probabilitywith which the particle resets is a function of the currentlocation of the particle.

To address stochastic resetting in the aforementionedscenario, we employed a path-integral approach, dis-cussed in detail in Eqs. (15)-(19) in Sec. II B 1. Invok-ing the Feynman-Kac formalism, we obtained an equalitythat relates the probability of transition between differ-ent spatial locations of the particle before it encountersany reset to the quantum propagator of a suitable quan-tum mechanical problem (see Sec. II B 1). Using this for-malism and elements from renewal theory, we obtainedclosed-form analytical expressions for a number of statis-tics of the dynamics, e.g., the probability distribution ofthe first-reset time (Sec. II B 2), the time-dependent spa-tial distribution (Sec. II B 3), and the stationary spatialdistribution (Sec. II B 4).

We applied the method to a number of representativeexamples, including in particular those involving nontriv-ial spatial dependence of the rate of resetting. Remark-ably, we obtained the exact distributions of the afore-mentioned dynamical quantifiers for two non-trivial prob-lems: the resetting of a free Brownian particle under“parabolic” resetting (Sec. III B) and the resetting of aBrownian particle moving in a harmonic potential witha resetting rate that depends on the energy of the parti-cle (Sec. III C). For the latter case, we showed that usinginstantaneous quenching of resetting profiles allows to re-strict the mean-squared displacement of a Brownian par-ticle to a desired value on a faster timescale than by using

instantaneous potential quenches. We expect that sucha shortcut to confinement would provide novel insights inongoing research on, e.g., engineered-swift-equilibrationprotocols [65, 66] and shortcuts to adiabaticity [67–69].

Our work may also be extended to treat systems ofinteracting particles, with the advantage that the cor-responding quantum mechanical system can be treatedeffectively by using tools of quantum physics and many-body quantum theory. Our approach also provides a vi-able method to calculate path probabilities of complexstochastic processes. Such calculations are of particu-lar interest in many contexts, e.g., in stochastic ther-modynamics [70–73], and in the study of several bio-logical systems such as molecular motors [14, 74], ac-tive gels [75], genetic switches [76, 77], etc. As a spe-cific application in this direction, our approach allows toexplore the physics of Brownian tunnelling [78], an in-teresting stochastic resetting version of the well-knownphenomenon of quantum tunneling, which serves to un-veil the subtle effects resulting from stochastic resettingin, e.g., transport through nanopores [79].

VI. ACKNOWLEDGEMENTS

ER thanks Ana Lisica and Stephan Grill for initial dis-cussions, Ken Sekimoto and Luca Peliti for discussions onpath integrals, and Domingo Sanchez and Juan M. Torresfor discussions on quantum mechanics.

[1] Landauer R 1961 Irreversibility and heat generation inthe computing process IBM J. Res. Develop. 5 183

[2] Bennett C H 1973 Logical reversibility of computationIBM J. Res. Develop. 17 525

[3] Berut A, Arakelyan A, Petrosyan A, Ciliberto S, Dillen-schneider R, and Lutz E 2012 Experimental verificationof Landauer’s principle linking information and thermo-dynamics Nature 483 187

[4] Mandal D and Jarzynski C 2012 Work and informationprocessing in a solvable model of Maxwell’s demon Proc.Natl. Acad. Sci. 109 11641

[5] Roldan E, Martınez I A, Parrondo J M R, and PetrovD 2014 Universal features in the energetics of symmetrybreaking Nature Phys. 10 457

[6] Koski J V, Maisi V F, Pekola J P, and Averin D V 2014Experimental realization of a Szilard engine with a singleelectron Proc. Natl. Acad. Sci. 111 13786

[7] Fuchs J, Goldt G, and Seifert U 2016 Stochastic thermo-dynamics of resetting Europhys. Lett. 113 60009

[8] Mora T 2015 Physical Limit to Concentration SensingAmid Spurious Ligands Phys. Rev. Lett. 115 038102

[9] Roldan E, Lisica A, Sanchez-Taltavull D, Grill S W 2016Stochastic resetting in backtrack recovery by RNA poly-merases Phys. Rev. E 93 062411

[10] Lisica A, Engel C, Jahnel M, Roldan E, Galburt E A,Cramer P, and Grill S W 2016 Mechanisms of backtrack

recovery by RNA polymerases I and II Proc. Natl. Acad.Sci. USA 113 2946

[11] Gillespie D T, Seitaridou E, and Gillespie C A 2014 Thesmall-voxel tracking algorithm for simulating chemical re-actions among diffusing molecules J. Chem. Phys. 14112649

[12] Hanggi P, Talkner P, and Borkovec M 1990 Reaction-ratetheory: fifty years after Kramers Rev. Mod. Phys. 62 251

[13] Neri I, Roldan E, and Julicher F 2017 Statistics of Infimaand Stopping Times of Entropy production and Appli-cations to Active Molecular Processes Phys. Rev. X 7011019

[14] Julicher F, Ajdari A, and Prost J 1997 Modeling molec-ular motors Rev. Mod. Phys. 69 1269

[15] Zamft B, Bintu L, Ishibashi T, and Bustamante C J2012 Nascent RNA structure modulates the transcrip-tional dynamics of RNA polymerases Proc. Natl. Acad.Sci. 109 8948

[16] Evans M R and Majumdar S N 2011 Diffusion withstochastic resetting Phys. Rev. Lett. 106 160601

[17] Evans M R and Majumdar S N 2011 Diffusion with op-timal resetting J. Phys. A: Math. Theor. 44 435001

[18] Evans M R and Majumdar S N 2014 Diffusion with re-setting in arbitrary spatial dimension J. Phys. A: Math.Theor. 47 285001

[19] Christou C and Schadschneider A 2015 Diffusion with

Page 13: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

13

resetting in bounded domains J. Phys. A: Math. Theor.48 285003

[20] Eule S and Metzger J J 2016 Non-equilibrium steadystates of stochastic processes with intermittent resettingNew J. Phys. 18 033006

[21] Nagar A and Gupta S 2016 Diffusion with stochastic re-setting at power-law times Phys. Rev. E 93 060102(R)

[22] Boyer D and Solis-Salas C 2014 Random walks with pref-erential relocations to places visited in the past and theirapplication to biology Phys. Rev. Lett. 112 240601

[23] Majumdar S N, Sabhapandit S, and Schehr G 2015 Ran-dom walk with random resetting to the maximum posi-tion Phys. Rev. E 92 052126

[24] Montero M and Villarroel J 2013 Monotonic continuous-time random walks with drift and stochastic reset eventsMiquel Montero Phys. Rev. E 87 012116

[25] Mendez V and Campos D 2016 Characterization of sta-tionary states in random walks with stochastic resettingPhys. Rev. E 93 022106

[26] Kusmierz L, Majumdar S N, Sabhapandit S, and SchehrG 2014 First order transition for the optimal search timeof Levy flights with resetting Phys. Rev. Lett. 113 220602

[27] Campos D and Mendez V 2015 Phase transitions in op-timal search times: How random walkers should combineresetting and flight scales Phys. Rev. E 92 062115

[28] Pal A, Kundu A, and Evans M R 2016 Diffusion undertime-dependent resetting J. Phys. A: Math. Theor. 49225001

[29] Boyer D, Evans M R, and Majumdar S N 2017 Longtime scaling behaviour for diffusion with resetting andmemory J. Stat. Mech.: Theory Exp. 023208

[30] Harris R J and Touchette H 2017 Phase transitions inlarge deviations of reset processes J. Phys. A: Math.Theor. 50 10LT01

[31] Durang X, Henkel M, and Park H 2014 The statisti-cal mechanics of the coagulationdiffusion process with astochastic reset J. Phys. A: Math. Theor. 47 045002

[32] Gupta S, Majumdar S N, and Schehr G 2014 Fluctuatinginterfaces subject to stochastic resetting Phys. Rev. Lett.112 220601

[33] Gupta S and Nagar A 2016 Resetting of fluctuating in-terfaces at power-law times J. Phys. A: Math. Theor. 49445001

[34] Falcao R and Evans M R 2017 Interacting Brownian mo-tion with resetting J. Stat. Mech.: Theory Exp. 023204

[35] Kusmierz L and Gudowska-Nowak E 2015 Optimal first-arrival times in Lvy flights with resetting Phys. Rev. E92 052127

[36] Reuveni S 2016 Optimal stochastic restart renders fluc-tuations in first passage times universal Phys. Rev. Lett.116 170601

[37] Bhat U, De Bacco C, and Redner S 2016 Stochasticsearch with Poisson and deterministic resetting J. Stat.Mech.: Theory Exp. 083401

[38] Pal A and Reuveni S 2017 First passage under restartPhys. Rev. Lett. 118 030603

[39] Pal A 2015 Diffusion in a potential landscape withstochastic resetting Phys. Rev. E 91 012113

[40] Cox D 1962 Renewal Theory (London: Methuen)[41] Feynman R P and Hibbs A R 2010 Quantum Mechanics

and Path Integrals (New York: McGraw-Hill Companies,Inc.)

[42] Schulman L S 1981 Techniques and Applications of PathIntegration (UK: John Wiley & Sons)

[43] Kac M 1949 On distribution of certain Wiener functionalsTrans. Am. Math. Soc. 65 1

[44] Kac M 1951 On some connections between probabil-ity theory and differential and integral equations, inProc. Second Berkeley Symp. on Math. Statist. and Prob.(Berkeley: University of California Press)

[45] Majumdar S N 2005 Brownian Functionals in Physicsand Computer Science Curr. Sci. 89 2076

[46] In quantum mechanics, this time transformation is oftencalled the Wick’s rotation in honor of Gian-Carlo Wick

[47] Sekimoto K 1998 Langevin equation and thermodynam-ics Prog. Theor. Phys. Suppl. 130 17

[48] Sekimoto K 2000 Stochastic Energetics (Berlin: Springer)[49] Olver F W J, Daalhuis A B O, Lozier D W, Schnei-

der B I, Boisvert R F, Clark C W, Miller B R, andSaunders B V eds. NIST Digital Library of Math-ematical Functions, http://dlmf.nist.gov/, Release

1.0.14 of 2016-12-21

[50] Roldan-Vargas S, Rovigatti L, and Sciortino F 2017 Con-nectivity, dynamics, and structure in a tetrahedral net-work liquid Soft Matter 13 514

[51] Berthier L, and Biroli G 2011 Theoretical perspective onthe glass transition and amorphous materials Rev. Mod.Phys. 83 587

[52] Martınez I A, Roldan E, Parrondo J M R, and PetrovD 2013 Effective heating to several thousand kelvins ofan optically trapped sphere in a liquid Phys. Rev. E 87032159

[53] Berut A, Petrosyan A, and Ciliberto S 2015 Energy flowbetween two hydrodynamically coupled particles kept atdifferent effective temperatures EPL (Europhys. Lett.)107 (6) 60004.

[54] Dieterich E., Camunas-Soler J., Ribezzi-Crivellari M.,Seifert U, and Ritort, F 2015 Single-molecule measure-ment of the effective temperature in non-equilibriumsteady states Nature Phys. 11 (11), 97

[55] Martınez I A, Roldan E, Dinis L and Rica R A 2017Colloidal heat engines: a review Soft matter 13 (1) 22.

[56] Cirac J I, and Zoller P 1995 Quantum computations withcold trapped ions Phys. Rev. Lett. 74 (20) 4091.

[57] Bloch I 2005 Ultracold quantum gases in optical latticesNature Phys. 1 (1) 23.

[58] Gavrilov M, Chetrite R and Bechhoefer J 2017 Di-rect measurement of nonequilibrium system entropyis consistent with Gibbs-Shannon form arXiv preprintarXiv:1703.07601.

[59] Ciliberto S 2017 Experiments in Stochastic Thermody-namics: Short History and Perspectives Phys. Rev. X 7021051.

[60] Martınez I A, Roldan E, Dinis L, Petrov D, Parrondo JM R, and Rica R A 2016 Brownian carnot engine NaturePhys. 12 (1) 67.

[61] Evans M R, Majumdar S N and Mallick K 2013 Optimaldiffusive search: nonequilibrium resetting versus equilib-rium dynamics J. Phys. A 46 (18) 185001.

[62] Gavrilov M, Jun Y and Bechhoefer J, 2014 Real-time cal-ibration of a feedback trap Rev. Sci. Instr. 85 (9) 095102.

[63] Jun Y and Bechhoefer J 2012.Virtual potentials for feed-back traps Phys. Rev. E 86 (6) 061106.

[64] Martınez I A, Roldan E, Dinis L, Petrov D and RicaR A 2015 Adiabatic processes realized with a trappedBrownian particle Phys. Rev. Lett. 114 (12) 120601.

[65] Martınez I A, Petrosyan A, Guery-Odelin D, Trizac E

Page 14: arXiv:1703.10615v2 [cond-mat.stat-mech] 26 Jul 2017 · 2019-02-12 · Path-integral formalism for stochastic resetting: Exactly solved examples and shortcuts to con nement Edgar Rold

14

and Ciliberto S 2016 Engineered swift equilibration of aBrownian particle Nature Phys. 12 (9) 843.

[66] Granger L, Dinis L, Horowitz J M and Parrondo J M R2016. Reversible feedback confinement EPL (Europhys.Lett.) 115 (5) 50007.

[67] Deffner S, Jarzynski C and del Campo A 2014 Classicaland quantum shortcuts to adiabaticity for scale-invariantdriving Phys. Rev. X 4 (2) 021013.

[68] Deng J, Wang Q H, Liu Z, Hanggi P and Gong J 2013.Boosting work characteristics and overall heat-engineperformance via shortcuts to adiabaticity: Quantum andclassical systems Phys. Rev. E 88 (6) 062122.

[69] Tu Z C 2014 Stochastic heat engine with the considera-tion of inertial effects and shortcuts to adiabaticity Phys.Rev. E 89 (5) 052148.

[70] Jarzynski C 2011 Equalities and inequalities: irre-versibility and the second law of thermodynamics at thenanoscale Annu. Rev. Condens. Matt. Phys. 2 329

[71] Seifert U 2012 Stochastic thermodynamics, fluctuationtheorems and molecular machines Rep. Prog. Phys. 7512

[72] Celani A, Bo S, Eichhorn R and Aurell E 2012 Anomalousthermodynamics at the microscale Phys. Rev. Lett. 109

(26) 260603.[73] Bo S and Celani A 2017 Stochastic processes on multi-

ple scales: averaging, decimation and beyond Bull. Am.Phys. Soc. 62 1.

[74] Guerin T, Prost J, and Joanny J-F 2011 Motion reversalof molecular motor assemblies due to weak noise Phys.Rev. Lett. 106 068101

[75] Basu A, Joanny J-F, Julicher F, and Prost J 2008 Ther-mal and non-thermal fluctuations in active polar gelsEur. Phys. J. E 27 149

[76] Perez-Carrasco R, Guerrero P, Briscoe J, and Page KM2016 Intrinsic noise profoundly alters the dynamics andsteady state of morphogen-controlled bistable geneticswitches PLoS Comput. Biol. 12 e1005154

[77] Schultz D, Walczak A M, Onuchic J N, Wolynes P G2008 Extinction and resurrection in gene networks Proc.Natl. Acad. Sci. 105 19165

[78] Roldan E and Gupta S (in preparation)[79] Trepagnier E H, Radenovic A, Sivak D, Geissler P, and

Liphardt L 2007 Controlling DNA capture and propaga-tion through artificial nanopores Nano Lett. 7 2824