Next Article in Journal
Iterative Variable Selection for High-Dimensional Data: Prediction of Pathological Response in Triple-Negative Breast Cancer
Next Article in Special Issue
Non-Debye Relaxations: Two Types of Memories and Their Stieltjes Character
Previous Article in Journal
On a Retarded Nonlocal Ordinary Differential System with Discrete Diffusion Modeling Life Tables
Previous Article in Special Issue
Higher-Order Symmetries of a Time-Fractional Anomalous Diffusion Equation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Diffusion–Advection Equations on a Comb: Resetting and Random Search

1
Research Center for Computer Science and Information Technologies, Macedonian Academy of Sciences and Arts, Bul. Krste Misirkov 2, 1000 Skopje, Macedonia
2
Institute of Physics & Astronomy, University of Potsdam, D-14776 Potsdam-Golm, Germany
3
Institute of Physics, Faculty of Natural Sciences and Mathematics, Ss. Cyril and Methodius University, Arhimedova 3, 1000 Skopje, Macedonia
4
Department of Physics, Technion, Haifa 32000, Israel
5
Faculty of Computer Science and Engineering, Ss. Cyril and Methodius University, P.O. Box 393, 1000 Skopje, Macedonia
*
Author to whom correspondence should be addressed.
Mathematics 2021, 9(3), 221; https://doi.org/10.3390/math9030221
Submission received: 25 December 2020 / Revised: 17 January 2021 / Accepted: 20 January 2021 / Published: 22 January 2021
(This article belongs to the Special Issue Fractional Calculus in Anomalous Transport Theory)

Abstract

:
This review addresses issues of various drift–diffusion and inhomogeneous advection problems with and without resetting on comblike structures. Both a Brownian diffusion search with drift and an inhomogeneous advection search on the comb structures are analyzed. The analytical results are verified by numerical simulations in terms of coupled Langevin equations for the comb structure. The subordination approach is one of the main technical methods used here, and we demonstrated how it can be effective in the study of various random search problems with and without resetting.

1. Introduction

In the standard theory of Brownian motion, the probability density function (PDF) P 0 ( x , t ) for finding a particle at position x at time t has a Gaussian form,
P 0 ( x , t ) = 1 4 π D t e ( x x 0 ) 2 4 D t ,
for the initial conditions given by the Dirac delta function P 0 ( x , t = 0 ) = δ ( x x 0 ) and for the natural (vanishing) boundary conditions at infinity, P 0 ( ± , t ) = 0 and x P ( x , t ) | x = ± = 0 . Brownian motion is characterized by linear growth of the mean squared displacement (MSD), x 2 ( t ) = 2 D t , which means normal diffusion. However, in a large variety of transport phenomena in randomly inhomogeneous media, one observes deviation from this linear growth in time such that the MSD has a power-law dependence on time, x 2 ( t ) t α , which is a signature of anomalous diffusion (see, for example, [1]). When 0 < α < 1 , it corresponds to subdiffusion, while α > 1 corresponds to superdiffusion.
One of the well-known examples of anomalous diffusion is Brownian motion on a comb, which is governed by the Fokker–Planck equation [2,3]
t P ( x , y , t ) = δ ( y ) D x x 2 P ( x , y , t ) + D y y 2 P ( x , y , t ) ,
with the initial condition P ( x , y , t = 0 ) = δ ( x x 0 ) δ ( y ) , and the boundary conditions for P ( x , y , t ) and q P ( x , y , t ) , q = { x , y } , are set to zero at infinity, x = ± , y = ± . Here, D x δ ( y ) and D y are diffusion coefficients along the x and y directions, respectively. The δ -function in the Fokker–Planck operator
L FP = δ ( y ) D x x 2 + D y y 2
means that diffusion along the x direction (the so-called backbone) is allowed only at y = 0 . Along the y direction (the so-called fingers), the particle performs normal diffusion. The MSD along the backbone has a power-law dependence on time x 2 ( t ) = D x D y t 1 / 2 Γ ( 3 / 2 ) , i.e., subdiffusion is observed in the system [2,3], while normal diffusion takes place along the fingers. Different generalizations of the comb geometry have been considered. For example, various diffusion processes have been considered in a comb with a finite finger length [4,5,6], diffusion on cylindrical [7,8] and circular combs [9,10,11], more complex branched structures [12], random comb models [13], and a comb with ramified teeth [14], as well as the problem of first encounters for two workers [15]. Diffusion processes in fractal mesh and grid structures have been considered as well: In this case, anomalous diffusion of a particle is affected by the fractal structure of the infinite numbers of backbones and fingers [16]. It has been shown that these models are useful for description of anomalous transport through porous solid pellets with various porous geometries [17]. Comb models are also applicable for describing diffusion in percolation clusters [2,18,19], anomalous transport of inert compounds in spiny dendrites [20,21,22], modeling electron transport in disordered nanostructured semiconductors [23,24], dispersive transport of charge carriers in two-layer polymers [25], percolative phonon-assisted hopping in two-dimensional disordered systems [26,27], and anomalous diffusion of fluorescence recovery after photobleaching in a random-comb model [13]. Another interesting realization is that turbulent diffusion in a comb appears to be due to multiplicative noise [28,29].
Nowadays, one of the most explored problems in stochastic processes is the problem of stochastic resetting, meaning that a particle is reset to the initial (or any other) position from time to time. The one-dimensional Brownian motion with Markovian resetting with a constant resetting rate r was introduced by Evans and Majumdar [30]. It was shown that the solution for the PDF approaches a non-equilibrium steady state and, in the long-time limit, its MSD is saturated, x 2 ( t ) 1 / r (also see the review paper [31] for more details). Moreover, Brownian motion in a two-dimensional comb in the presence of stochastic (Markovian) resetting can be solved analytically [32,33,34]. The marginal PDFs along both the backbone and fingers approach non-equilibrium steady states, and the MSDs are saturated according to the resetting rate: x 2 ( t ) 1 / r and y 2 ( t ) 1 / r [32,33,34]. These models have been extended to diffusion processes with non-static resetting [34]. Stochastic resetting is a natural mechanism in various search processes, such as foraging [35], population dynamics [36], Michaelis–Menten enzymatic reactions [37], and human behavior of finding resources [38], to mention but a few. Resetting may also affect the first-passage properties and completion of the process. In particular, in the case of the one-dimensional Brownian search, the mean arrival time at the absorbing boundary becomes finite in the presence of resetting [30], while it is infinite in the absence of resetting [39]. The resetting dynamics of a Brownian particle under external potentials have been analyzed in detail, as well [40,41,42,43,44,45]. This issue can also be employed to understand resetting in molecular reaction systems.
Another important topic in stochastic processes is the random search problem. Many studies on random searches in foraging theory with incomplete information have employed a Brownian search as a default strategy [46], while others have proposed Lévy flights as an efficient strategy for searching for sufficiently sparse targets [47], stating that the Lévy process is one of the most natural and optimal search strategies [48,49,50,51,52]. Various search strategies have been introduced and proposed, including different combinations of search processes [53,54,55,56,57].
The corresponding Fokker–Planck equation of a Brownian random search process for the non-normalized density function f ( x , t ) with a δ -sink of strength P fa reads [53,54,55]
t f ( x , t ) = D x 2 f ( x , t ) P fa ( t ) δ ( x X ) ,
where D is a diffusion coefficient. One assumes here that the initial position is given at x = x 0 by f ( x , t = 0 ) = δ ( x x 0 ) . The δ -sink means that the random searcher positioned at the beginning at x = x 0 will be removed at the first arrival at x = X , i.e., f ( x = X , t ) = 0 . Therefore, P fa ( t ) represents the first arrival time distribution (FATD) [53,54,55], which is obtained from Equation (3):
P fa ( t ) = d d t f ( x , t ) d x = d d t S ( t ) ,
which is a negative time derivative of the survival probability S ( t ) = f ( x , t ) d x . The FATD for the Brownian search is described by the Lévy–Smirnov density:
P fa ( t ) = | X x 0 | 4 π D t 3 × e ( X x 0 ) 2 4 D t
with the long-time asymptotics, P fa ( t ) | X x 0 | t 3 / 2 . Other important characteristics of searching are the search reliability and the efficiency. The search reliability is considered as the cumulative arrival probability [54]:
P = 0 P fa ( t ) d t = P fa ( s = 0 ) ,
which, for the Brownian search, is P = 1 (the searcher will find the target with the probability of one), while the search efficiency [54],
E = 1 t = 0 P fa ( s ) d s ,
represents the averaged inverse search time. Here, P fa ( s ) = L P fa ( t ) ( s ) = 0 fa ( t ) e s t d t is the Laplace image of P fa ( t ) . For a one-dimensional Brownian search, it is given by [54]:
E = 2 D ( X x 0 ) 2 .
In a similar way, a Brownian random search on a comblike structure has also been considered. In this case, the initial position of the searcher is located at the backbone at ( x , y ) = ( x 0 , 0 ) , and the target is also located at the backbone at ( x , y ) = ( X , 0 ) , with a δ -sink δ ( x X ) δ ( y ) of the strength fa ( t ) [58]. The FATD is given in terms of the Fox H-function with the long-time asymptotics reducing to the power law, fa ( t ) | X x 0 | t 5 / 4 , while the search reliability equals one, and the efficiency becomes [58]:
E = 24 D x 2 D y 2 ( X x 0 ) 4 .
The inhomogeneous advection on the comb, where the motion along the backbone is interrupted by Brownian motion in the fingers, can be described by the Fokker–Planck operator
L FP = v δ ( y ) x | x | + D y y 2 .
It results in turbulent diffusion, which is characterized by the log-normal distribution and exponential growth of the MSD in time [29]. This behavior is analogous to one-dimensional geometric Brownian motion [29], which is used in the Black–Scholes model for option pricing [59,60]. The FATD is the Lévy–Smirnov distribution, and the process is suitable for searching for long-distance targets. Turbulent diffusion occurs due to a multiplicative noise, in contrast to the additive noise in Brownian diffusion with a drift. The impact of the resetting mechanism on turbulent diffusion is one of the main issues in this paper.
This paper is organized as follows. In Section 2, the one-dimensional Brownian motion in the presence of a drift and resetting to the initial position of the particle is considered. As the main characteristics of the process, the PDF and the MSD are obtained. The analytical treatments of the corresponding Brownian search problem with the drift in the cases of a single target and two targets are presented. The main features of a subordination approach—as the analytical tool used throughout the analysis—are described as well. The analytical results obtained for the FATD, the search reliability, and the efficiency are verified by numerical simulations. Brownian motion with drift in the presence of stochastic resetting on a two-dimensional comb is investigated in Section 3. Both analytical and numerical results are presented. The problem of a Brownian search with a drift on a two-dimensional comb in the cases of a single target and two targets is analyzed in detail. Section 4 is devoted to inhomogeneous advection with stochastic resetting on the comb. It is shown that three different scenarios for the MSD, depending on the resetting parameter, can be observed. These are: (i) exponential growth of the MSD in time, (ii) linear growth of the MSD in time, and (iii) saturation of the MSD. The results for the FATD, the search reliability, and the efficiency are presented as well. A generalization of the inhomogeneous advection search is also considered. A summary of the analysis is provided in Section 5.

2. One-Dimensional Brownian Motion with Drift

In this section, we consider resetting and search problems in the framework of a one-dimensional diffusion process with a drift. We show that this “simple” addition of a drift term in the corresponding equations leads to new physical effects, which are also based on well-known results of the one-dimensional diffusion–advection equation.
Therefore, to set the stage for the clear presentation of the analysis, we first offer a short overview of the results related to the one-dimensional diffusion–advection equation without stochastic resetting, which will be used later to find the corresponding results for more general problems. The corresponding Fokker–Planck equation with a constant velocity V reads [1]
t P 0 ( x , t ) = D x 2 V x P 0 ( x , t ) .
The initial condition P 0 ( x , t = 0 ) = δ ( x x 0 ) and vanishing boundary conditions at infinity, P 0 ( ± , t ) = 0 and x P 0 ( ± , t ) = 0 , are imposed. In Laplace space, it reads
s P 0 ( x , s ) δ ( x x 0 ) = D x 2 V x P 0 ( x , s ) .
The solution of Equation (11) is the Galilei-shifted Gaussian (see, for example, Ref. [1]),
P 0 ( x , t ) = 1 4 π D t e ( x x 0 V t ) 2 4 D t ,
and in the following, we shall need its Laplace image, which reads
P 0 ( x , s ) = 1 2 D 1 s D + V 2 4 D 2 e V 2 D ( x x 0 ) s D + V 2 4 D 2 | x x 0 | .
From here, we find that the PDF is normalized, since
x 0 ( t ) 0 = P 0 ( x , t ) d x = 1 4 π D t e ( x x 0 V t ) 2 4 D t d x = 2 4 π D t 0 e y 2 4 D t d y = 1 ,
and, respectively,
x 0 ( s ) 0 = P 0 ( x , s ) d x = 1 s .
The MSD is given by
x 2 ( s ) 0 = x 2 P 0 ( x , s ) d x = x 0 2 s + 2 ( D + x 0 V ) s 2 + 2 V 2 s 3 ,
x 2 ( t ) 0 = x 2 P 0 ( x , t ) d x = 2 D t + ( x 0 + V t ) 2 ,
which means that the short-time diffusive behavior ( x 2 ( t ) 0 t ) turns into ballistic motion in the long-time limit x 2 ( t ) 0 t 2 .

2.1. One-Dimensional Diffusion–Advection Equation with Stochastic Resetting

Now, we consider the diffusion–advection equation in the presence of stochastic resetting. The Fokker–Planck equation reads
t P r ( x , t ) = D x x 2 V x P r ( x , t ) r P r ( x , t ) + r δ ( x x 0 )
with the same initial condition P r ( x , t = 0 ) = δ ( x x 0 ) . Here, r is the rate of resetting to the initial position x 0 . The last two terms of the equation represent the loss of the probability from position x due to the reset to the initial position and the gain of the probability at x 0 due to resetting from all other positions, respectively. This equation means that between any two consecutive resetting events, the particle undergoes diffusion with a constant drift.
From the Laplace transform of Equation (19), one finds
s P r ( x , s ) δ ( x x 0 ) = s s + r D x 2 V x P r ( x , s ) .
Then, the inverse Laplace transform yields Equation (19) in the equivalent form
t P r ( x , t ) = d d t 0 t η ( t t ) D x 2 V x P r ( x , t ) d t ,
where η ( t ) = e r t and η ( s ) = 1 s + r .
This equation can be solved by using a subordination approach [1,59,61,62,63]. Equation (21) in Laplace space reads
s P r ( x , s ) δ ( x x 0 ) = s η ( s ) D x 2 V x P r ( x , s ) .
Let us present the solution of Equation (21) in the subordination form with the integral
P r ( x , t ) = 0 P 0 ( x , u ) h ( u , t ) d u ,
where P 0 ( x , t ) is the solution in Equation (13). Here, h ( u , t ) is the so-called subordination function. The latter is the PDF, which subordinates the process governed by Equation (21) to the process governed by Equation (13). By the Laplace transform of Equation (23), and by using the subordination function
h ( u , s ) = 1 s η ( s ) e u / η ( s ) ,
we find
P r ( x , s ) = 0 P 0 ( x , u ) h ( u , s ) d u = 1 s η ( s ) 0 P 0 ( x , u ) e u / η ( s ) d u = 1 s η ( s ) P 0 ( x , 1 / η ( s ) ) .
Performing the variable change s 1 / η ( s ) in Equation (12), we have
1 η ( s ) P 0 ( x , 1 / η ( s ) ) δ ( x x 0 ) = D x 2 V x P 0 ( x , 1 / η ( s ) ) .
Therefore, from Equations (25) and (26), we obtain Equation (22). Eventually, from Equations (14) and (25), we obtain the PDF in the presence of resetting,
P r ( x , s ) = s + r s 1 2 D 1 s + r D + V 2 4 D 2 e V 2 D ( x x 0 ) s + r D + V 2 4 D 2 | x x 0 | .
From here, one finds
P r ( x , t ) = e r t P 0 ( x , t ) + r 0 t e r t P 0 ( x , t ) d t ,
where P 0 ( x , t ) is defined by Equation (13), which is the solution of the corresponding Fokker–Planck equation without resetting. We also note that the case without the drift ( V = 0 ) yields the known result for free diffusion with stochastic resetting [31].
From Equations (27) and (16), we find that the PDF P r ( x , t ) is normalized ( x 0 ( t ) r = 1 ), since
x 0 ( s ) r = P r ( x , s ) d x = 1 s η ( s ) P 0 ( x , 1 / η ( s ) ) d x = 1 s η ( s ) x 0 ( 1 / η ( s ) ) 0 = η ( s ) s η ( s ) = 1 s .
From Equation (29) for the MSD, we find
x 2 ( t ) r = x 2 P r ( x , s ) d x = 1 s η ( s ) x 2 P 0 ( x , 1 / η ( s ) ) d x = 1 s η ( s ) x 2 ( 1 / η ( s ) ) 0 = x 0 2 η ( s ) + 2 ( D + x 0 V ) η 2 ( s ) + 2 V 2 η 3 ( s ) s η ( s ) = x 0 2 s + 2 ( D + x 0 V ) η ( s ) s + 2 V 2 η 2 ( s ) s ,
which results in
x 2 ( t ) r = L 1 x 0 2 s + 2 ( D + x 0 V ) s ( s + r ) + 2 V 2 s ( s + r ) 2 = x 0 2 + 2 ( D + x 0 V ) ( 1 e r t ) r + 2 V 2 ( 1 e r t r t e r t ) r 2 .
Then, the long-time limit yields saturation of the MSD,
x 2 ( t ) r x 0 2 + 2 ( D + x 0 V ) r + 2 V 2 r 2 ,
while the short-time limit corresponds to the result without resetting, Equation (18). In the absence of the drift, the MSD reads x 2 ( t ) r = x 0 2 + 2 D r ( 1 e r t ) [31].

Langevin Equation

We compare the analytical results against the ones obtained from direct numerical simulation of the dynamics by considering a Langevin equation in the presence of stochastic resetting to the initial position [44],
x ( t + Δ t ) = x ( 0 ) , with probability r Δ t , x ( t ) + V Δ t + 2 D Δ t ζ ( t ) , with probability ( 1 r Δ t ) ,
where ζ ( t ) is a zero-mean Gaussian noise and V, D , and r are parameters that are used equivalently in the analytical case. Regarding the temporal evolution of the variance, ensembles of 10 4 particle positions were simulated considering a time step of Δ t = 0.01 across a time span of 10 3 in order to observe convergence of the processes.
A graphical representation of the PDF is given in Figure 1 (left panel). The numerical results for the MSD, represented by dots, triangles, and squares in Figure 1 (right panel), show excellent agreement with the analytical results, represented by lines. A typical trajectory of a particle is shown in Figure 2.

2.2. One-Dimensional Brownian Search with Drift

The random Brownian search with drift in one dimension is described by the Fokker–Planck equation [54]
t f ( x , t ) = D x 2 V x f ( x , t ) P fa ( t ) δ ( x X ) ,
where f ( x , t ) is the non-normalized density function, and the last term in the equation is a δ -sink of strength P fa ( t ) , which is considered as the FATD. From the condition f ( x = X , s ) = 0 , one finds the FATD [39,54]
P fa ( s ) = e V 2 D ( X x 0 ) s D + V 2 4 D 2 | X x 0 | P fa ( t ) = | X x 0 | 4 π D t 3 e ( X x 0 V t ) 2 4 D t ,
which is the exponentially truncated Lévy–Smirnov probability density. The search reliability is [54,55]
P = P fa ( s = 0 ) = e V ( X x 0 ) 2 D | V ( X x 0 ) | 2 D = 1 , for V ( X x 0 ) > 0 , e V ( x 0 X ) D , for V ( X x 0 ) < 0 ,
while the search efficiency has the form [54,55]
E = 2 D + | V ( X x 0 ) | ( X x 0 ) 2 × 1 , for V ( X x 0 ) > 0 , e V ( x 0 X ) D , for V ( X x 0 ) < 0 .
For V = 0 , one recovers the known result for the random Brownian search (8). The FATD and the search efficiency are depicted in Figure 3. In Figure 4, we present the survival probability obtained by numerical simulations (dots, triangles, and squares) in the framework of the Langevin equation approach and the numerical inverse Laplace transform in MATHEMATICA [64].
Next, we extend our analysis to the case of the Brownian search with drift in the presence of two sinks located at x = X 1 and x = X 2 , where X 1 < X 2 . The problem is governed by the Fokker–Planck equation
t f ( x , t ) = D x 2 V x f ( x , t ) P fa , 1 ( t ) δ ( x X 1 ) P fa , 2 ( t ) δ ( x X 2 ) ,
where f ( x = X 1 , t ) = f ( x = X 2 , t ) = 0 and P fa , 1 ( t ) + P fa , 2 ( t ) = P fa ( t ) is the FATD. Without loss of generality, we consider V > 0 . By the Fourier–Laplace transformation, we obtain
f ( k , s ) = e ı k x 0 P fa , 1 ( s ) e ı k X 1 P fa , 2 ( s ) e ı k X 2 s + D k 2 + ı V k .
By the inverse Fourier transform, we find
f ( x , s ) = 1 2 D 1 s D + V 2 4 D 2 e V 2 D ( x x 0 ) e s D + V 2 4 D 2 | x x 0 | P fa , 1 ( s ) e V 2 D ( x X 1 ) e s D + V 2 4 D 2 | x X 1 | P fa , 2 ( s ) e V 2 D ( x X 2 ) e s D + V 2 4 D 2 | x X 2 | .
Three cases of the initial position of the searcher are considered. These are: (i) x 0 < X 1 < X 2 , (ii) X 1 < x 0 < X 2 , and (iii) X 1 < X 2 < x 0 . Following the same approach as is used in the case of one sink, we find the FATD in the Laplace space as follows:
(i)
x 0 < X 1 < X 2 :
P fa ( s ) = e s D + V 2 4 D 2 V 2 D ( X 1 x 0 ) ,
(ii)
X 1 < x 0 < X 2
P fa ( s ) = e s D + V 2 4 D 2 + V 2 D ( x 0 X 1 ) 1 e s D + V 2 4 D 2 V 2 D ( X 2 X 1 ) 1 e 2 s D + V 2 4 D 2 ( X 2 X 1 ) + e s D + V 2 4 D 2 V 2 D ( X 2 x 0 ) 1 e s D + V 2 4 D 2 + V 2 D ( X 2 X 1 ) 1 e 2 s D + V 2 4 D 2 ( X 2 X 1 ) ,
(iii)
X 1 < X 2 < x 0
P fa ( s ) = e s D + V 2 4 D 2 + V 2 D ( x 0 X 2 ) .
Then, the search reliability reads
P = P fa ( s = 0 ) = 1 for x 0 < X 1 < X 2 , 1 for X 1 < x 0 < X 2 , e V D ( x 0 X 2 ) for X 1 < X 2 < x 0 .
The search efficiency for the case X 1 < x 0 < X 2 can be analyzed numerically, while the other two cases can be calculated exactly and correspond to the single-target problem.

3. Brownian Motion with Drift on a Comb: Stochastic Resetting and Random Search Problem

3.1. Diffusion–Advection Equation on a Comb with Stochastic Resetting

The problem of the diffusion–advection equation on a comblike structure was introduced by Arkhincheev and Baskin in Ref. [2], and was also studied in Refs. [65,66] in terms of random walks with anisotropy that appear due to the presence of an external electrical field. We extend the model with stochastic resetting, which results in the Fokker–Planck equation
t P r ( x , y , t ) = δ ( y ) D x x 2 v x P r ( x , y , t ) + D y y 2 P r ( x , y , t ) r P r ( x , y , t ) + r δ ( x x 0 ) δ ( y )
with the initial condition P r ( x , y , t = 0 ) = δ ( x x 0 ) δ ( y ) , where v is a constant velocity. For the Laplace image, one looks for the solution in the form
P r ( x , y , s ) = g ( x , s ) e s + r D y | y | ,
which yields the following marginal PDF:
p 1 , r ( x , s ) = P r ( x , y , s ) d y = 2 D y s + r g ( x , s ) .
Performing the Laplace transform of Equation (44) and taking into account Equation (46), one obtains
s p 1 , r ( x , s ) δ ( x x 0 ) = s ( s + r ) 1 / 2 2 D y D x x 2 v x p 1 , r ( x , s ) .
By the inverse Laplace transform, we find the equation
t p 1 , r ( x , t ) = 1 2 D y d d t 0 t η r ( t t ) D x x 2 v x p 1 , r ( x , t ) d t ,
where η r ( t ) = L 1 ( s + r ) 1 / 2 = e r t t 1 / 2 Γ ( 1 / 2 ) (for more details on tempered operators, see Ref. [67]). Disregarding resetting, and taking into account that η r = 0 ( t ) = t 1 / 2 / Γ ( 1 / 2 ) , one obtains
t p 1 , 0 ( x , t ) = 1 2 D y t 1 / 2 D x x 2 v x p 1 , 0 ( x , t ) ,
where t α f ( t ) = 1 Γ ( 1 α ) d d t 0 t ( t t ) α f ( t ) d t is the Riemann–Liouville fractional derivative of order 0 < α < 1 [68].
In the framework of the subordination approach, the Laplace image of the solution to Equation (49) is presented as follows:
p 1 , r ( x , s ) = 1 s η r ( s ) P 0 ( x , 1 / η r ( s ) ) = ( s + r ) 1 / 2 s P 0 ( x , ( s + r ) 1 / 2 ) ,
where P 0 ( x , u ) is the PDF (13). Then, the PDF (50) reads
p 1 , r ( x , s ) = e v 2 D x ( x x 0 ) 2 D x 2 D y 1 s η r ( s ) e 2 D y D x 1 η r ( s ) + v 2 4 D x 2 | x x 0 | 2 D y D x 1 η r ( s ) + v 2 4 D x 2 1 / 2 ,
which is
p 1 , r ( x , s ) = e v 2 D x ( x x 0 ) 2 D x 2 D y ( s + r ) 1 / 2 s e 2 D y D x ( s + r ) 1 / 2 + v 2 4 D x 2 | x x 0 | 2 D y D x ( s + r ) 1 / 2 + v 2 4 D x 2 1 / 2 .
The case without resetting ( r = 0 ) yields the PDF
p 1 , 0 ( x , s ) = e v 2 D x ( x x 0 ) 2 D x 2 D y s 1 / 2 e 2 D y D x s 1 / 2 + v 2 4 D x 2 | x x 0 | 2 D y D x s 1 / 2 + v 2 4 D x 2 1 / 2 .
Eventually, the solution to Equation (48) reads
p 1 , r ( x , t ) = e r t p 1 , 0 ( x , t ) + r 0 t e r t p 1 , 0 ( x , t ) d t .
From Equations (50) and (16), one can easily check that the PDF is normalized, since
x 0 ( s ) r = ( s + r ) 1 / 2 s P 0 ( x , ( s + r ) 1 / 2 ) d x = ( s + r ) 1 / 2 s x 0 ( ( s + r ) 1 / 2 ) 0 = 1 s ,
that is, x 0 ( t ) r = 1 . For the MSD in the case of resetting, we find
x 2 ( t ) r = x 0 2 + D x D y + x 0 v D y erf ( r t ) r + v 2 2 D y 1 e r t r ,
which, in the absence of resetting, reduces to
x 2 ( t ) 0 = x 0 2 + 2 ( D + x 0 V ) t 1 / 2 Γ ( 3 / 2 ) + 2 V 2 t = x 0 2 + D x D y + x 0 v D y t 1 / 2 Γ ( 3 / 2 ) + v 2 2 D y t .
Therefore, in the long-time limit, saturation of the MSD is due to the resetting,
x 2 ( t ) x 0 2 + D x D y 1 r + v 2 2 D y 1 r ,
while in the short-time limit (when erf ( x ) 2 z / π and e z 1 z for z 1 ), we recover the result obtained for the resetting-free case in Equation (57).

Langevin Equation

To analyze the diffusion dynamics numerically, we use a system of Langevin equations in the presence of drift and stochastic resetting to the initial position [33,44]:
x ( t + Δ t ) = x ( 0 ) , with probability r Δ t , x ( t ) + A ( y ) v Δ t + 2 D 1 Δ t ζ 1 ( t ) , with probability ( 1 r Δ t ) ,
y ( t + Δ t ) = y ( 0 ) , with probability r Δ t , y ( t ) + 2 D 2 Δ t ζ 2 ( t ) , with probability ( 1 r Δ t ) ,
where A ( y ) is a function that mimics the Dirac δ -function. To simulate A ( y ) , diffusion across the x directions is permitted in a narrow strip with a width of 2 ϵ along the y axis such that the value of ϵ is of the same order of magnitude as the diffusion coefficients [69]. A graphical representation of the PDF is depicted in Figure 5 (left panel), while the MSD is presented in Figure 5 (right panel). A good agreement between the analytical and numerical results for the MSDs is obtained. The individual trajectory along the backbone in the presence of stochastic resetting is shown in Figure 6. The case without resetting is shown in Figure 7 and Figure 8.

3.2. Brownian Search with Drift on a Comb

In this section, we consider a Brownian random search process on a comblike structure in the presence of an external drift along the backbone (constant external force). The Fokker–Planck equation then reads
t f ( x , y , t ) = δ ( y ) D x x 2 v x f ( x , y , t ) + D y y 2 f ( x , y , t ) fa ( t ) δ ( x X ) δ ( y )
with the initial condition f ( x , y , t = 0 ) = δ ( x x 0 ) δ ( y ) . The δ -sink means that the searcher is annihilated by reaching the point ( x , y ) = ( X , 0 ) . According to the definition in Equation (4), the FATD results from double integration
fa ( t ) = d d t f ( x , y , t ) d x d y = d d t S ( t ) ,
where S ( t ) = f ( x , y , t ) d x d y is the survival probability.
By the Laplace transform of Equation (61), we find
s f ( x , y , t ) δ ( x x 0 ) δ ( y ) = δ ( y ) D x x 2 v x f ( x , y , s ) + D y y 2 f ( x , y , s ) fa ( s ) δ ( x X ) δ ( y ) .
Again, using the substitution
f ( x , y , s ) = g ( x , s ) e s D y | y | f 1 ( x , s ) = f ( x , y , s ) d y = 2 D y s g ( x , s ) ,
we find
s f 1 ( x , s ) δ ( x x 0 ) = 1 2 D y s × s 1 / 2 D x x 2 v x f 1 ( x , s ) fa ( s ) δ ( x X ) .
By the Fourier transform with respect to x, we find
s f 1 ( k , s ) e ı k x 0 = 1 2 D y s × s 1 / 2 D x k 2 ı v k f 1 ( k , s ) fa ( s ) e ı k X ,
and thus,
f 1 ( k , s ) = s 1 / 2 s 1 / 2 + D x 2 D y k 2 + ı v 2 D y k e ı k x 0 fa ( s ) e ı k X ,
which is the solution of the equation
t f 1 ( x , t ) = 1 2 D y 1 / 2 t 1 / 2 D x x 2 v x f 1 ( x , t ) fa ( t ) δ ( x X ) .
The inverse Fourier transform of Equation (65) with respect to k gives
f 1 ( x , s ) = 1 2 D s 1 / 2 s 1 / 2 + V 2 4 D 1 / 2 e V 2 D ( x x 0 ) s 1 / 2 D + V 2 4 D 2 | x x 0 | fa ( s ) e V 2 D ( x X ) s 1 / 2 D + V 2 4 D 2 | x X | ,
where D = D x 2 D y and V = v 2 D y . From the condition
f ( x = X , y = 0 , t ) = g ( x = X , t ) = s 1 / 2 2 D y f 1 ( x = X , t ) = 0 ,
we conclude that it corresponds to f 1 ( x = X , t ) = 0 , i.e., f 1 ( x = X , s ) = 0 . Thus, for the FATD in Laplace space, we have
fa ( s ) = e v 2 D x ( X x 0 ) 2 D y D x s 1 / 2 + v 2 4 D x 2 | X x 0 | .
We also note that the FATD (68) can be directly obtained from the FATD for the one-dimensional search with drift (see Equation (34)). Then, we have fa ( s ) = P fa ( s 1 / 2 ) , where we use D = D x 2 D y and V = v 2 D y . The search reliability becomes
P = fa ( s = 0 ) = e v ( X x 0 ) 2 D x | v ( X x 0 ) | 2 D x = 1 , for v ( X x 0 ) > 0 , e v ( x 0 X ) D x , for v ( X x 0 ) < 0 ,
while the efficiency is given by
E = 4 D x 2 D y ( X x 0 ) 4 e v ( X x 0 ) 2 D x | v ( X x 0 ) | 2 D x 6 D x 2 D y 1 + | v ( X x 0 ) | 2 D x + v 2 ( X x 0 ) 2 2 D y = 4 D x 2 D y ( X x 0 ) 4 × 6 D x 2 D y 1 + v ( X x 0 ) 2 D x + v 2 ( X x 0 ) 2 2 D y , for v ( X x 0 ) > 0 , e v ( X x 0 ) D x 6 D x 2 D y 1 v ( X x 0 ) 2 D x + v 2 ( X x 0 ) 2 2 D y , for v ( X x 0 ) < 0 .
For v = 0 , we recover the known result (9). The FATD and the efficiency for the random search with drift on the comb are depicted in Figure 9. In Figure 10, we give the survival probability obtained by numerical simulations of the Langevin equation and the numerical inverse Laplace transform obtained by MATHEMATICA [64].
One can also consider the Brownian search with drift on a comb in the presence of two sinks located on the backbone at x = X 1 and x = X 2 , where X 1 < X 2 . The corresponding Fokker–Planck equation is
t f ( x , y , t ) = δ ( y ) D x x 2 v x f ( x , y , t ) + D y y 2 f ( x , y , t ) fa , 1 ( t ) δ ( x X 1 ) + fa , 2 ( t ) δ ( x X 2 ) δ ( y )
with the initial condition f ( x , y , t = 0 ) = δ ( x x 0 ) δ ( y ) . We consider v > 0 , while the case for v < 0 can be treated in a similar way. Following the same approach suggested for the one-sink problem, we consider the marginal PDF f 1 ( x , t ) = f ( x , y , t ) . Then, its Laplace image reads
f 1 ( k , s ) = s 1 / 2 s 1 / 2 + D x 2 D y k 2 + ı v 2 D y k e ı k x 0 fa , 1 ( s ) e ı k X 1 fa , 2 ( s ) e ı k X 2 .
Due to the presence of two sinks, the absorbing condition is f ( x = X 1 , y = 0 , t ) = f ( x = X 2 , y = 0 , t ) = 0 , and correspondingly, f 1 ( x = X 1 , t ) = f 1 ( x = X 2 , t ) = 0 . We will also consider three cases of the initial positions of the searcher: (i) x 0 < X 1 < X 2 , (ii) X 1 < x 0 < X 2 , and (iii) X 1 < X 2 < x 0 . For the FATD fa ( t ) = fa , 1 ( t ) + fa , 2 ( t ) , we obtain:
(i)
x 0 < X 1 < X 2 :
fa ( s ) = e 2 D x D x s 1 / 2 + v 2 4 D x 2 v 2 D x ( X 1 x 0 ) ,
(ii)
X 1 < x 0 < X 2
fa ( s ) = e 2 D x D x s 1 / 2 + v 2 4 D x 2 + v 2 D x ( x 0 X 1 ) 1 e 2 D x D x s 1 / 2 + v 2 4 D x 2 v 2 D x ( X 2 X 1 ) 1 e 2 2 D x D x s 1 / 2 + v 2 4 D x 2 ( X 2 X 1 ) + e 2 D x D x s 1 / 2 + v 2 4 D x 2 v 2 D x ( X 2 x 0 ) 1 e 2 D x D x s 1 / 2 + v 2 4 D x 2 + v 2 D x ( X 2 X 1 ) 1 e 2 2 D x D x s 1 / 2 + v 2 4 D x 2 ( X 2 X 1 ) ,
(iii)
X 1 < X 2 < x 0
fa ( s ) = e 2 D x D x s 1 / 2 + v 2 4 D x 2 + v 2 D x ( x 0 X 2 ) .
Therefore, the search reliability is
P = 1 for x 0 < X 1 < X 2 , 1 for X 1 < x 0 < X 2 , e v D x ( x 0 X 2 ) for X 1 < X 2 < x 0 .
The search efficiency for the case X 1 < x 0 < X 2 can be analyzed numerically, while the other two cases can be calculated exactly, and the results are the same as for the single-target problem on the comb.

4. Inhomogeneous Advection on a Comb

4.1. Inhomogeneous Advection with Stochastic Resetting

Inhomogeneous advection on a comb in the presence of stochastic resetting is described by the following equation:
t P ( x , y , t ) = v δ ( y ) x { x P ( x , y , t ) } + D y 2 P ( x , y , t ) r P ( x , y , t ) + r δ ( x x 0 ) δ ( y )
with the initial condition P ( x , y , t = 0 ) = δ ( x x 0 ) δ ( y ) . The case without resetting was considered in Ref. [29]. By the Laplace transform, we have
s P ( x , y , s ) δ ( x x 0 ) δ ( y ) = v s s + r δ ( y ) x { x P ( x , y , s ) } + D s s + r y 2 P ( x , y , s ) .
The solution can be represented in the form P ( x , y , s ) = g ( x , s ) e s + r D | y | . Then, the marginal PDF along the backbone is p 1 ( x , s ) = P ( x , y , s ) d y = 2 D s + r g ( x , s ) , and the equation for the marginal PDF reads
s p 1 ( x , s ) δ ( x x 0 ) = v 2 D s s + r x { x p 1 ( x , s ) } .
The inverse Laplace transform yields
t p 1 ( x , t ) = v 2 D d d t 0 t η ( t t ) x { x p 1 ( x , t ) } d t ,
where η ( t ) = L 1 ( s + r ) 1 / 2 = e r t t 1 / 2 Γ ( 1 / 2 ) . In order to find the solution of this equation, we use the subordination approach. Let us consider the standard inhomogeneous advection equation
t P 0 ( x , t ) = V x { x P 0 ( x , t ) } ,
which, in Laplace space, reads
s P 0 ( x , s ) δ ( x x 0 ) = V x { x P 0 ( x , s ) } .
The solution of the equation for x > x 0 in Laplace space is [29]
P 0 ( x , s ) = θ ( x x 0 ) V x e s V log x x 0 .
From the subordination approach, we obtain
p 1 ( x , s ) = 1 s η ( s ) P 0 ( x , 1 / η ( s ) ) = ( s + r ) 1 / 2 s P 0 ( x , ( s + r ) 1 / 2 ) = 2 D v θ ( x x 0 ) x ( s + r ) 1 / 2 s e 2 D v ( s + r ) 1 / 2 log x x 0 = s + r s 2 D v θ ( x x 0 ) x ( s + r ) 1 / 2 e 2 D v ( s + r ) 1 / 2 log x x 0 ,
where we used V v 2 D . The inverse Laplace transform yields the solution
p 1 ( x , t ) = e r t P 1 ( x , t ) + r 0 t e r t P 1 ( x , t ) d t ,
where
P 1 ( x , t ) = 2 D v θ ( x x 0 ) x L 1 s 1 / 2 e 2 D v s 1 / 2 log x x 0 = 2 θ ( x x 0 ) x 4 π v 2 D 2 t × exp log 2 x x 0 4 v 2 D 2 t
is the solution of Equation (77) without resetting [29].
From Equation (84), the Laplace image of the MSD reads
x 2 ( s ) = x 0 2 s + r s ( s + r ) 1 / 2 ( s + r ) 1 / 2 v D ,
and the inverse Laplace transform yields
x 2 ( t ) = x 0 2 e r t E 1 / 2 v D t 1 / 2 + r 0 t e r t E 1 / 2 v D t 1 / 2 d t .
Here, E α ( z ) = n = 0 z n Γ ( α n + 1 ) is the one-parameter Mittag–Leffler function, and its Laplace image is L E α ( a t α ) = s α 1 s α a [68]. For the large argument, the Mittag–Leffler function reads E 1 / 2 v D t 1 / 2 e v 2 D t (see Ref. [68]). Therefore, this yields the long-time limit of the MSD, which has three different regimes: (i) exponential growth in time for r < v 2 / D , (ii) linear growth with time for r = v 2 / D , and (iii) saturation for r > v 2 / D .

4.2. Langevin Equation

In the case of inhomogeneous advection on a comb, completely different results are obtained due to the multiplicative noise. The microscopic approach to the process without resetting is described by the Langevin equation:
x ˙ ( t ) = v A ( y ) x ( t ) ,
y ˙ ( t ) = 2 D ζ ( t ) ,
where ζ ( t ) is a Gaussian noise. Therefore, y ( t ) is a random Brownian motion, so A ( y ) is a function of the random variable. Without any restriction of generality, we may use A ( y ) = δ ( y ) , which corresponds to the inhomogeneous advection motion on the comb.
To validate the analytical results, an ensemble of 10 6 particles was simulated based on the system of Langevin Equations (89) and (90), which, in the presence of resetting, are
x ( t + Δ t ) = x ( 0 ) , with probability r Δ t , x ( t ) + v A ( y ) x ( t ) Δ t , with probability ( 1 r Δ t ) ,
y ( t + Δ t ) = y ( 0 ) , with probability r Δ t , y ( t ) + 2 D Δ t ζ ( t ) , with probability ( 1 r Δ t ) ,
where the time step is Δ t = 0.01 . It should be admitted that for the multiplicative noise, one cannot use the numerical approximation of the Dirac δ -function considered in Section 3. For numerical purposes, one can use another approximation of the Dirac δ -function (see Ref. [4]). Consequently, we use a zero-mean Gaussian function A ( y ) = 1 2 π σ exp ( y 2 / [ 2 σ 2 ] ) in the limit σ 0 .
A graphical representation of the PDF is shown in Figure 11 (left panel), while the MSD obtained analytically and by numerical simulations is shown in Figure 11 (right panel). A typical trajectory along the backbone in the presence of stochastic resetting is shown in Figure 12.

4.3. Inhomogeneous Advection Search on a Comb

For the completeness of the analysis, we consider an inhomogeneous advection search on the comb, or the so-called turbulent diffusion search. The corresponding Fokker–Planck equation reads [29]
t F ( x , y , t ) = v δ ( y ) x { x F ( x , y , t ) } + D y 2 F ( x , y , t ) fa ( t ) δ ( x X ) δ ( y ) .
Following the standard procedure described in Section 3.2, we find the FATD in Laplace space as follows:
fa ( s ) = e 2 D v s log X x 0 .
The inverse Laplace transform of Equation (94) for X > x 0 gives the Lévy–Smirnov distribution [29]:
fa ( t ) = log X x 0 4 π v 2 D 2 t 3 × exp log 2 X x 0 4 v 2 D 2 t .
The search reliability equals one ( P = 1 ), while the efficiency is
E = 0 e 2 D v s log X x 0 d s = 2 v 2 D 2 log 2 X x 0 .
Therefore, the turbulent diffusion search is more efficient than the Brownian search for long-distance targets, but the searcher should have a prior knowledge of the direction of the target, which is not the case for the Brownian search. The FATD and the efficiency are shown in Figure 13. The survival probability is shown in Figure 14.
Remark 1.
As observed above in Equations (95) and (96), the corresponding inhomogeneous advection search problem cannot be considered for the initial position of the searcher at the origin ( x 0 = 0 ). To consider a random search with the initial condition at x 0 = 0 , we add some constant advection v 0 = v ϵ , where ϵ > 0 . In this case, x 0 = 0 is no longer a singular point for the space derivative. The search equation reads
t F ( x , y , t ) = v δ ( y ) x ( x + ϵ ) F ( x , y , t ) + D y 2 F ( x , y , t ) fa ( t ) δ ( x X ) δ ( y ) .
Following the same procedure for the FATD described above, we obtain
fa ( s ) = e 2 D v s log X + ϵ x 0 + ϵ .
For x 0 = 0 , it reads fa ( s ) = e 2 D v s log X + ϵ ϵ .
The reliability becomes P = 1 , and the efficiency becomes
E = 2 v 2 D 2 log 2 X + ϵ x 0 + ϵ ,
which is also valid for x 0 = 0 . However, the searcher moves only in the advection direction.
Remark 2.
For completeness of the analysis, we also consider a search model with inhomogeneous advection of the form x | x | λ , 0 < λ < 1 [28]. Note that for λ = 1 , it corresponds to the inhomogeneous advection search problem considered above. The search Equation (93) now reads
t F ( x , y , t ) = v δ ( y ) sgn ( x ) λ | x | λ 1 + | x | λ x F ( x , y , t ) + D y 2 F ( x , y , t ) fa ( t ) δ ( x X ) δ ( y ) ,
where
sgn ( x ) = x | x | = 2 θ ( x ) 1 = 1 , for x > 0 , 0 , for x = 0 , 1 , for x < 0 .
Following the same procedure for the FATD at X > x 0 > 0 , we obtain
fa ( s ) = e 2 D v s X 1 λ x 0 1 λ 1 λ .
The search reliability is P = fa ( s = 0 ) = 1 . The inverse Laplace transform of Equation (101) yields
fa ( t ) = X 1 λ x 0 1 λ 4 π ( 1 λ ) 2 v 2 D 2 t 3 × exp X 1 λ x 0 1 λ 2 4 ( 1 λ ) 2 v 2 D 2 t .
Therefore, the efficiency has the form
E = 2 ( 1 λ ) 2 v 2 D 2 X 1 λ x 0 1 λ 2 .
Remark 3.
We note that in the limit λ 1 , the FATD (101) and the efficiency (103) become
lim λ 1 fa ( s ) = e 2 D v s lim λ 1 X 1 λ x 0 1 λ 1 λ = e 2 D v s log X x 0 ,
lim λ 1 E = lim λ 1 2 ( 1 λ ) 2 v 2 D 2 X 1 λ x 0 1 λ 2 = lim λ 1 2 ( 1 λ ) v 2 D 2 X 1 λ x 0 1 λ log X log x 0 = 2 v 2 D 2 log 2 X x 0 ,
where we use L’Hôpital’s rule, and thus, we recover the corresponding results ((94) and (96)) for the inhomogeneous advection search.
Remark 4.
For λ = 0 , in Equation (101), for the FATD, we obtain fa ( s ) = e 2 D v s ( X x 0 ) . This case corresponds to homogeneous advection on the comb, and thus, the FATD can be obtained from Equation (68) for D x 0 . Namely, we have
lim D x 0 fa ( s ) = lim D x 0 e v 2 D x ( X x 0 ) 2 D y D x s 1 / 2 + v 2 4 D x 2 | X x 0 | = lim D x 0 e v 2 D x ( X x 0 ) v 2 D x 1 + 8 D y D x v 2 s 1 / 2 | X x 0 | = lim D x 0 e v 2 D x ( X x 0 ) v 2 D x 1 + 4 D y D x v 2 s 1 / 2 | X x 0 | = e 2 D v s ( X x 0 ) for X > x 0 .
The efficiency reads
E = 2 v 2 D 2 X x 0 2 ,
which has the same behavior as the efficiency for the one-dimensional Brownian search (see Equation (8)).

5. Summary

We present an overview of various drift–diffusion and inhomogeneous advection problems with and without resetting on comblike structures. Both the Brownian diffusion search with drift and the inhomogeneous advection search on the comb structures were analyzed. The analytical results were verified by numerical simulations in terms of coupled Langevin equations for the comb structures. The subordination approach was one of the main technical methods used here, and we demonstrated how it can be effective in the study of various random search problems with and without resetting.
In conclusion, consideration of the comb model (or comb geometry) is an important issue for investigation of the interplay between diffusion, drift, and geometry. For example, further modification of the comb model, like a fractal tartan [16,70,71,72], can be an interesting task for understanding of the impact of fractal geometry on the fractional transport, as well as for the experimental implementation and the technological design and development of sparse sensor arrays [73,74].

Author Contributions

Conceptualization, T.S., A.I. and L.K.; Formal analysis, V. D.; Investigation, T.S., V.D. and A.I.; Methodology, T.S., V.D. and A.I.; Supervision, L.K.; Visualization, T.S. and V.D.; Writing—original draft, T.S., V.D. and A.I.; Writing—review and editing, A.I. and L.K. All authors contributed to each part of this work equally. All authors have read and agreed to the published version of the manuscript.

Funding

T.S. was supported by the Alexander von Humboldt Foundation.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PDFprobability density function
MSDmean squared displacement
FATDfirst arrival time distribution

References

  1. Metzler, R.; Klafter, J. The random walk’s guide to anomalous diffusion: A fractional dynamics approach. Phys. Rep. 2000, 339, 1–77. [Google Scholar] [CrossRef]
  2. Arkhincheev, V.; Baskin, E. Anomalous diffusion and drift in a comb model of percolation clusters. Soviet Phys. JETP 1991, 73, 161–300. [Google Scholar]
  3. Iomin, A.; Mendéz, V.; Horsthemke, W. Fractional Dynamics in Comb-Like Structures; World Scientific: Singapore, 2018. [Google Scholar]
  4. Méndez, V.; Iomin, A.; Horsthemke, W.; Campos, D. Langevin dynamics for ramified structures. J. Stat. Mech. 2017, 2017, 063205. [Google Scholar] [CrossRef] [Green Version]
  5. Liu, L.; Zheng, L.; Liu, F. Time fractional Cattaneo-Christov anomalous diffusion in comb frame with finite length of fingers. J. Mol. Liquids 2017, 233, 326–333. [Google Scholar] [CrossRef] [Green Version]
  6. Liu, L.; Zheng, L.; Liu, F.; Zhang, X. Anomalous diffusion in finite length fingers comb frame with the effects of time and space Riesz fractional Cattaneo-Christov flux and Poiseuille flow. J. Comput. Math. 2018, 36, 563–578. [Google Scholar] [CrossRef]
  7. Iomin, A.; Méndez, V. Does ultra-slow diffusion survive in a three dimensional cylindrical comb? Chaos Solitons Fractals 2016, 82, 142–147. [Google Scholar] [CrossRef]
  8. Dzhanoev, A.; Sokolov, I. The effect of the junction model on the anomalous diffusion in the 3D comb structure. Chaos Solitons Fractals 2018, 106, 330–336. [Google Scholar] [CrossRef] [Green Version]
  9. Liu, C.; Fan, Y.; Lin, P. Numerical investigation of a fractional diffusion model on circular comb-inward structure. Appl. Math. Lett. 2020, 100, 106053. [Google Scholar] [CrossRef]
  10. Fan, Y.; Liu, L.; Zheng, L.; Li, X. Subdiffusions on circular branching structures. Commun. Nonlin. Sci. Numer. Simul. 2019, 77, 225–238. [Google Scholar] [CrossRef]
  11. Iomin, A. Anomalous diffusion in umbrella comb. Chaos Solitons Fractals 2021, 142, 110488. [Google Scholar] [CrossRef]
  12. Suleiman, K.; Zheng, L.; Liu, C.; Zhang, X.; Wang, E. The Effect of Geometry on the Diffusion: Branched Archimedean spiral. Int. Commun. Heat Mass Transf. 2020, 117, 104733. [Google Scholar] [CrossRef]
  13. Yuste, S.B.; Abad, E.; Baumgaertner, A. Anomalous diffusion and dynamics of fluorescence recovery after photobleaching in the random-comb model. Phys. Rev. E 2016, 94, 012118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Rebenshtok, A.; Barkai, E. Occupation times on a comb with ramified teeth. Phys. Rev. E 2013, 88, 052126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Peng, J.; Agliari, E. First encounters on combs. Phys. Rev. E 2019, 100, 062310. [Google Scholar] [CrossRef] [PubMed]
  16. Sandev, T.; Iomin, A.; Kantz, H. Anomalous diffusion on a fractal mesh. Phys. Rev. E 2017, 95, 052107. [Google Scholar] [CrossRef] [Green Version]
  17. Zhokh, A.; Trypolskyi, A.; Strizhak, P. Relationship between the anomalous diffusion and the fractal dimension of the environment. Chem. Phys. 2018, 503, 71–76. [Google Scholar] [CrossRef]
  18. Matan, O.; Havlin, S.; Stauffer, D. Scaling properties of diffusion on comb-like structures. J. Phys. A Math. Gen. 1989, 22, 2867. [Google Scholar] [CrossRef]
  19. Sibatov, R.T.; Uchaikin, V.V. Fractional kinetics of charge carriers in supercapacitors. Appl. Eng. Life Soc. Sci. 2019, 87, 87–116. [Google Scholar]
  20. Méndez, V.; Iomin, A. Comb-like models for transport along spiny dendrites. Chaos Solitons Fractals 2013, 53, 46–51. [Google Scholar] [CrossRef] [Green Version]
  21. Iomin, A.; Méndez, V. Reaction-subdiffusion front propagation in a comblike model of spiny dendrites. Phys. Rev. E 2013, 88, 012706. [Google Scholar] [CrossRef] [Green Version]
  22. Wang, Z.; Zheng, L.; Sun, Y. The double fractional Cattaneo model on anomalous transport of compounds in spiny dendrites structure. J. Stat. Mech. 2020, 2020, 093203. [Google Scholar] [CrossRef]
  23. Uchaikin, V.V.; Sibatov, R. Fractional Kinetics in Solids: Anomalous Charge Transport in Semiconductors, Dielectrics, and Nanosystems; World Scientific: Singapore, 2013. [Google Scholar]
  24. Sibatov, R.; Morozova, E. Multiple trapping on a comb structure as a model of electron transport in disordered nanostructured semiconductors. J. Exper. Theor. Phys. 2015, 120, 860–870. [Google Scholar] [CrossRef]
  25. Sibatov, R.; Morozova, E. On Theory of Dispersive Transport in a Two-Layer Polymer Structure. Russian Phys. J. 2016, 59, 722–728. [Google Scholar] [CrossRef]
  26. Sibatov, R.; Shulezhko, V.; Svetukhin, V. Fractional Derivative Phenomenology of Percolative Phonon-Assisted Hopping in Two-Dimensional Disordered Systems. Entropy 2017, 19, 463. [Google Scholar] [CrossRef]
  27. Sibatov, R.T. Fractal Generalization of the Scher–Montroll Model for Anomalous Transit-Time Dispersion in Disordered Solids. Mathematics 2020, 8, 1991. [Google Scholar] [CrossRef]
  28. Baskin, E.; Iomin, A. Superdiffusion on a comb structure. Phys. Rev. Lett. 2004, 93, 120603. [Google Scholar] [CrossRef]
  29. Sandev, T.; Iomin, A.; Kocarev, L. Hitting times in turbulent diffusion due to multiplicative noise. Phys. Rev. E 2020, 102, 042109. [Google Scholar] [CrossRef]
  30. Evans, M.R.; Majumdar, S.N. Diffusion with stochastic resetting. Phys. Rev. Lett. 2011, 106, 160601. [Google Scholar] [CrossRef] [Green Version]
  31. Evans, M.R.; Majumdar, S.N.; Schehr, G. Stochastic resetting and applications. J. Phys. A: Math. Theor. 2020, 53, 193001. [Google Scholar] [CrossRef]
  32. Tateishi, A.; Ribeiro, H.; Sandev, T.; Petreska, I.; Lenzi, E. Quenched and annealed disorder mechanisms in comb models with fractional operators. Phys. Rev. E 2020, 101, 022135. [Google Scholar] [CrossRef] [Green Version]
  33. Domazetoski, V.; Masó-Puigdellosas, A.; Sandev, T.; Méndez, V.; Iomin, A.; Kocarev, L. Stochastic resetting on comblike structures. Phys. Rev. Res. 2020, 2, 033027. [Google Scholar] [CrossRef]
  34. Antonio Faustino dos Santos, M. Comb Model with Non-Static Stochastic Resetting and Anomalous Diffusion. Fractal Fract. 2020, 4, 28. [Google Scholar] [CrossRef]
  35. Bartumeus, F.; Catalan, J. Optimal search behavior and classic foraging theory. J. Phys. A Math. Theor. 2009, 42, 434002. [Google Scholar] [CrossRef] [Green Version]
  36. Visco, P.; Allen, R.J.; Majumdar, S.N.; Evans, M.R. Switching and growth for microbial populations in catastrophic responsive environments. Biophys. J. 2010, 98, 1099–1108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Reuveni, S.; Urbakh, M.; Klafter, J. Role of substrate unbinding in Michaelis–Menten enzymatic reactions. Proc. Natl. Acad. Sci. USA 2014, 111, 4391–4396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Bell, W. The Behavioural Ecology of Finding Resources; Springer: Dordrecht, The Netherlands, 1990. [Google Scholar]
  39. Redner, S. A Guide to First-Passage Processes; Cambridge University Press: Cambridge, UK, 2001. [Google Scholar]
  40. Ray, S.; Mondal, D.; Reuveni, S. Péclet number governs transition to acceleratory restart in drift-diffusion. J. Phys. A Math. Theor. 2019, 52, 255002. [Google Scholar] [CrossRef] [Green Version]
  41. Ray, S.; Reuveni, S. Diffusion with resetting in a logarithmic potential. J. Chem. Phys. 2020, 152, 234110. [Google Scholar] [CrossRef]
  42. Masó-Puigdellosas, A.; Campos, D.; Méndez, V. Transport properties and first-arrival statistics of random motion with stochastic reset times. Phys. Rev. E 2019, 99, 012141. [Google Scholar] [CrossRef] [Green Version]
  43. Ahmad, S.; Nayak, I.; Bansal, A.; Nandi, A.; Das, D. First passage of a particle in a potential under stochastic resetting: A vanishing transition of optimal resetting rate. Phys. Rev. E 2019, 99, 022130. [Google Scholar] [CrossRef] [Green Version]
  44. Pal, A. Diffusion in a potential landscape with stochastic resetting. Phys. Rev. E 2015, 91, 012113. [Google Scholar] [CrossRef] [Green Version]
  45. Singh, R.; Metzler, R.; Sandev, T. Resetting dynamics in a confining potential. J. Phys. A: Math. Theor. 2020, 53, 505003. [Google Scholar] [CrossRef]
  46. Stephens, D.W.; Krebs, J.R. Foraging Theory; Princeton University Press: Princeton, NJ, USA, 1986; Volume 1. [Google Scholar]
  47. Shlesinger, M.F.; Klafter, J. Lévy walks versus Lévy flights. In On Growth and Form; Springer: Dordrecht, The Netherlands, 1986; pp. 279–283. [Google Scholar]
  48. Dybiec, B.; Gudowska-Nowak, E.; Chechkin, A. To hit or to pass it over—Remarkable transient behavior of first arrivals and passages for Lévy flights in finite domains. J. Phys. A Math. Theor. 2016, 49, 504001. [Google Scholar] [CrossRef]
  49. Lomholt, M.A.; Tal, K.; Metzler, R.; Joseph, K. Lévy strategies in intermittent search processes are advantageous. Proc. Natl. Acad. Sci. USA 2008, 105, 11055–11059. [Google Scholar] [CrossRef] [Green Version]
  50. Lomholt, M.A.; Ambjörnsson, T.; Metzler, R. Optimal target search on a fast-folding polymer chain with volume exchange. Phys. Rev. Lett. 2005, 95, 260603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Bartumeus, F.; Levin, S.A. Fractal reorientation clocks: Linking animal behavior to statistical patterns of search. Proc. Natl. Acad. Sci. USA 2008, 105, 19072–19077. [Google Scholar] [CrossRef] [Green Version]
  52. Reynolds, A. Optimal random Lévy-loop searching: New insights into the searching behaviours of central-place foragers. EPL (Europhys. Lett.) 2008, 82, 20001. [Google Scholar] [CrossRef]
  53. Chechkin, A.V.; Metzler, R.; Gonchar, V.Y.; Klafter, J.; Tanatarov, L.V. First passage and arrival time densities for Lévy flights and the failure of the method of images. J. Phys. A Math. Gen. 2003, 36, L537. [Google Scholar] [CrossRef]
  54. Palyulin, V.V.; Chechkin, A.V.; Metzler, R. Lévy flights do not always optimize random blind search for sparse targets. Proc. Natl. Acad. Sci. USA 2014, 111, 2931–2936. [Google Scholar] [CrossRef] [Green Version]
  55. Palyulin, V.V.; Chechkin, A.V.; Metzler, R. Space-fractional Fokker–Planck equation and optimization of random search processes in the presence of an external bias. J. Stat. Mech. 2014, 2014, P11031. [Google Scholar] [CrossRef] [Green Version]
  56. Palyulin, V.V.; Chechkin, A.V.; Klages, R.; Metzler, R. Search reliability and search efficiency of combined Lévy–Brownian motion: Long relocations mingled with thorough local exploration. J. Phys. A Math. Theor. 2016, 49, 394002. [Google Scholar] [CrossRef]
  57. Palyulin, V.V.; Mantsevich, V.N.; Klages, R.; Metzler, R.; Chechkin, A.V. Comparison of pure and combined search strategies for single and multiple targets. Eur. Phys. J. B 2017, 90, 170. [Google Scholar] [CrossRef] [Green Version]
  58. Sandev, T.; Iomin, A.; Kocarev, L. Random search on comb. J. Phys. A Math. Theor. 2019, 52, 465001. [Google Scholar] [CrossRef]
  59. Magdziarz, M. Black-Scholes formula in subdiffusive regime. J. Stat. Phys. 2009, 136, 553–564. [Google Scholar] [CrossRef]
  60. Stojkoski, V.; Sandev, T.; Basnarkov, L.; Kocarev, L.; Metzler, R. Generalised geometric Brownian motion: Theory and applications to option pricing. Entropy 2020, 22, 1432. [Google Scholar] [CrossRef] [PubMed]
  61. Barkai, E. Fractional Fokker-Planck equation, solution, and application. Phys. Rev. E 2001, 63, 046118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Meerschaert, M.M.; Benson, D.A.; Scheffler, H.P.; Baeumer, B. Stochastic solution of space-time fractional diffusion equations. Phys. Rev. E 2002, 65, 041103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Bazhlekova, E.; Bazhlekov, I. Subordination approach to space-time fractional diffusion. Mathematics 2019, 7, 415. [Google Scholar] [CrossRef] [Green Version]
  64. Mallet, A. Numerical Inversion of Laplace Transform, 2000. Wolfram Library Archive, 2015. Available online: http://library.wolfram.com/infocenter/MathSource/2691 (accessed on 24 December 2020).
  65. Arkhincheev, V. Anomalous diffusion and charge relaxation on comb model: Exact solutions. Phys. A 2000, 280, 304–314. [Google Scholar] [CrossRef]
  66. Lenzi, E.; da Silva, L.; Tateishi, A.; Lenzi, M.; Ribeiro, H. Diffusive process on a backbone structure with drift terms. Phys. Rev. E 2013, 87, 012121. [Google Scholar] [CrossRef] [Green Version]
  67. Sandev, T. Generalized Langevin equation and the Prabhakar derivative. Mathematics 2017, 5, 66. [Google Scholar] [CrossRef] [Green Version]
  68. Mainardi, F. Fractional Calculus and Waves in Linear Viscoelasticity: An Introduction to Mathematical Models; World Scientific: Singapore, 2010. [Google Scholar]
  69. Ribeiro, H.V.; Tateishi, A.A.; Alves, L.G.; Zola, R.S.; Lenzi, E.K. Investigating the interplay between mechanisms of anomalous diffusion via fractional Brownian walks on a comb-like structure. New J. Phys. 2014, 16, 093050. [Google Scholar] [CrossRef] [Green Version]
  70. Golmankhaneh, A.K.; Fernandez, A. Fractal calculus of functions on cantor tartan spaces. Fractal Fract. 2018, 2, 30. [Google Scholar] [CrossRef] [Green Version]
  71. Balankin, A.S.; Golmankhaneh, A.K.; Patiño-Ortiz, J.; Patiño-Ortiz, M. Noteworthy fractal features and transport properties of Cantor tartans. Phys. Lett. A 2018, 382, 1534–1539. [Google Scholar] [CrossRef]
  72. Sandev, T.; Iomin, A.; Kantz, H. Fractional diffusion on a fractal grid comb. Phys. Rev. E 2015, 91, 032108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Fang, H.; Qiu, Z.; O’Leary, R.; Gachagan, A.; Mulholland, A. Linear Ultrasonic Array Development incorporating Cantor Set Fractal Geometry. In Proceedings of the 2018 IEEE International Ultrasonics Symposium (IUS), Kobe, Japan, 22–25 October 2018; pp. 1–4. [Google Scholar]
  74. Cohen, R.; Eldar, Y.C. Sparse Array Design via Fractal Geometries. arXiv 2020, arXiv:2001.01217. [Google Scholar] [CrossRef]
Figure 1. Left panel: Probability density function (PDF) (27) at t = 1 ; right panel: Mean squared displacement (MSD) (30) for D = 1 , x 0 = 0 , V = 1 and r = 0 (blue solid line), r = 1 (red dashed line), and r = 2 (black dot-dashed line).
Figure 1. Left panel: Probability density function (PDF) (27) at t = 1 ; right panel: Mean squared displacement (MSD) (30) for D = 1 , x 0 = 0 , V = 1 and r = 0 (blue solid line), r = 1 (red dashed line), and r = 2 (black dot-dashed line).
Mathematics 09 00221 g001
Figure 2. A typical trajectory of particles in the presence of stochastic resetting to the initial position x 0 = 0 , for r = 1 , D = 1 , and V = 1 . The resetting events are represented by black dots. Dashed regions are introduced for these resetting events to be more visible.
Figure 2. A typical trajectory of particles in the presence of stochastic resetting to the initial position x 0 = 0 , for r = 1 , D = 1 , and V = 1 . The resetting events are represented by black dots. Dashed regions are introduced for these resetting events to be more visible.
Mathematics 09 00221 g002
Figure 3. Left panel: First arrival time distribution (FATD) (34) for D = 1 , V = 0 , and X x 0 = 2 (blue solid line), V = 1 and X x 0 = 2 (red dashed line), and V = 1 and X x 0 = 2 (black dot-dashed line). Right panel: Efficiency (36) for D = 1 and V = 0 (blue solid line), v = 10 and X > x 0 (red dashed line), and v = 10 and X < x 0 (black dot-dashed line).
Figure 3. Left panel: First arrival time distribution (FATD) (34) for D = 1 , V = 0 , and X x 0 = 2 (blue solid line), V = 1 and X x 0 = 2 (red dashed line), and V = 1 and X x 0 = 2 (black dot-dashed line). Right panel: Efficiency (36) for D = 1 and V = 0 (blue solid line), v = 10 and X > x 0 (red dashed line), and v = 10 and X < x 0 (black dot-dashed line).
Mathematics 09 00221 g003
Figure 4. The survival probability for the one-dimensional search with drift for D = 1 , V = 0 , and X x 0 = 2 (blue solid line), V = 1 and X x 0 = 2 (red dashed line), and V = 1 and X x 0 = 2 (black dot-dashed line). The numerical results are represented by dots, triangles, and squares.
Figure 4. The survival probability for the one-dimensional search with drift for D = 1 , V = 0 , and X x 0 = 2 (blue solid line), V = 1 and X x 0 = 2 (red dashed line), and V = 1 and X x 0 = 2 (black dot-dashed line). The numerical results are represented by dots, triangles, and squares.
Mathematics 09 00221 g004
Figure 5. Left panel: PDF (51) at t = 1 ; right panel: MSD (56) for D x = 1 , D y = 1 , x 0 = 0 , v = 2 , and r = 0 (blue solid line), r = 0.5 (red dashed line), and r = 2 (black dot-dashed line).
Figure 5. Left panel: PDF (51) at t = 1 ; right panel: MSD (56) for D x = 1 , D y = 1 , x 0 = 0 , v = 2 , and r = 0 (blue solid line), r = 0.5 (red dashed line), and r = 2 (black dot-dashed line).
Mathematics 09 00221 g005
Figure 6. A typical trajectory of particles in the presence of stochastic resetting to the initial position ( x 0 , y 0 ) = ( 0 , 0 ) for r = 0.5 , D x = 1 , D y = 1 , and v = 2 . The resetting events are represented by black dots. Dashed regions are introduced for these resetting events to be more visible.
Figure 6. A typical trajectory of particles in the presence of stochastic resetting to the initial position ( x 0 , y 0 ) = ( 0 , 0 ) for r = 0.5 , D x = 1 , D y = 1 , and v = 2 . The resetting events are represented by black dots. Dashed regions are introduced for these resetting events to be more visible.
Mathematics 09 00221 g006
Figure 7. Left panel: PDF (53) at t = 1 ; right panel: MSD (57) for D x = 1 , D y = 1 , x 0 = 0 , and v = 0 (blue solid line), v = 2 (red dashed line), and v = 5 (black dot-dashed line).
Figure 7. Left panel: PDF (53) at t = 1 ; right panel: MSD (57) for D x = 1 , D y = 1 , x 0 = 0 , and v = 0 (blue solid line), v = 2 (red dashed line), and v = 5 (black dot-dashed line).
Mathematics 09 00221 g007
Figure 8. A typical trajectory of particles in the absence of stochastic resetting for D x = 1 , D y = 1 , and v = 2 .
Figure 8. A typical trajectory of particles in the absence of stochastic resetting for D x = 1 , D y = 1 , and v = 2 .
Mathematics 09 00221 g008
Figure 9. Left panel: FATD (68) for D x = 1 , D y = 1 , v = 0 , and X x 0 = 2 (blue solid line), v = 1 and X x 0 = 2 (red dashed line), and v = 1 and X x 0 = 2 (black dot-dashed line). Right panel: Efficiency (70) for D x = 1 , D y = 1 , and v = 0 (blue solid line), v = 10 and X > x 0 (red dashed line), and v = 10 and X < x 0 (black dot-dashed line).
Figure 9. Left panel: FATD (68) for D x = 1 , D y = 1 , v = 0 , and X x 0 = 2 (blue solid line), v = 1 and X x 0 = 2 (red dashed line), and v = 1 and X x 0 = 2 (black dot-dashed line). Right panel: Efficiency (70) for D x = 1 , D y = 1 , and v = 0 (blue solid line), v = 10 and X > x 0 (red dashed line), and v = 10 and X < x 0 (black dot-dashed line).
Mathematics 09 00221 g009
Figure 10. The survival probability for the random search with the drift on the comb for D x = 1 , D y = 1 , v = 0 , and X x 0 = 2 (blue solid line), v = 1 and X x 0 = 2 (red dashed line), and v = 1 and X x 0 = 2 (black dot-dashed line). The numerical results are represented by dots, triangles, and squares.
Figure 10. The survival probability for the random search with the drift on the comb for D x = 1 , D y = 1 , v = 0 , and X x 0 = 2 (blue solid line), v = 1 and X x 0 = 2 (red dashed line), and v = 1 and X x 0 = 2 (black dot-dashed line). The numerical results are represented by dots, triangles, and squares.
Mathematics 09 00221 g010
Figure 11. Left panel: PDF (85) at t = 2 ; right panel: MSD (88) for D = 1 , x 0 = 1 , v = 1 , and r = 0.7 (blue solid line), r = 1 (red dashed line), and r = 2 (black dot-dashed line).
Figure 11. Left panel: PDF (85) at t = 2 ; right panel: MSD (88) for D = 1 , x 0 = 1 , v = 1 , and r = 0.7 (blue solid line), r = 1 (red dashed line), and r = 2 (black dot-dashed line).
Mathematics 09 00221 g011
Figure 12. A typical trajectory of particles in the presence of stochastic resetting to the initial position ( x 0 , y 0 ) = ( 1 , 0 ) for r = 1 , D = 1 , and v = 1 . The resetting events are represented by black dots. Dashed regions are introduced for these resetting events to be more visible.
Figure 12. A typical trajectory of particles in the presence of stochastic resetting to the initial position ( x 0 , y 0 ) = ( 1 , 0 ) for r = 1 , D = 1 , and v = 1 . The resetting events are represented by black dots. Dashed regions are introduced for these resetting events to be more visible.
Mathematics 09 00221 g012
Figure 13. Left panel: FATD (95) for D = 1 , v = 1 , and X / x 0 = 2.5 (blue solid line), X / x 0 = 5 (red dashed line), and X / x 0 = 7.5 (black dot-dashed line). Right panel: Efficiency (96) for D = 1 , V = 1 (blue solid line), v = 5 (red dashed line), and v = 10 (black dot-dashed line).
Figure 13. Left panel: FATD (95) for D = 1 , v = 1 , and X / x 0 = 2.5 (blue solid line), X / x 0 = 5 (red dashed line), and X / x 0 = 7.5 (black dot-dashed line). Right panel: Efficiency (96) for D = 1 , V = 1 (blue solid line), v = 5 (red dashed line), and v = 10 (black dot-dashed line).
Mathematics 09 00221 g013
Figure 14. The survival probability for the inhomogeneous advection search on a comb for D = 1 , v = 1 , and X / x 0 = 2 (blue solid line), X / x 0 = 5 (red dashed line), and X / x 0 = 10 (black dot-dashed line).The numerical results are represented by dots, triangles, and squares.
Figure 14. The survival probability for the inhomogeneous advection search on a comb for D = 1 , v = 1 , and X / x 0 = 2 (blue solid line), X / x 0 = 5 (red dashed line), and X / x 0 = 10 (black dot-dashed line).The numerical results are represented by dots, triangles, and squares.
Mathematics 09 00221 g014
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sandev, T.; Domazetoski, V.; Iomin, A.; Kocarev, L. Diffusion–Advection Equations on a Comb: Resetting and Random Search. Mathematics 2021, 9, 221. https://doi.org/10.3390/math9030221

AMA Style

Sandev T, Domazetoski V, Iomin A, Kocarev L. Diffusion–Advection Equations on a Comb: Resetting and Random Search. Mathematics. 2021; 9(3):221. https://doi.org/10.3390/math9030221

Chicago/Turabian Style

Sandev, Trifce, Viktor Domazetoski, Alexander Iomin, and Ljupco Kocarev. 2021. "Diffusion–Advection Equations on a Comb: Resetting and Random Search" Mathematics 9, no. 3: 221. https://doi.org/10.3390/math9030221

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop