Next Article in Journal
Dynamical Analysis of Rubella Disease Model in the Context of Fractional Piecewise Derivative: Simulations with Real Statistical Data
Next Article in Special Issue
Domains of Quasi Attraction: Why Stable Processes Are Observed in Reality?
Previous Article in Journal
Hotel Comment Emotion Classification Based on the MF-DFA and Partial Differential Equation Classifier
Previous Article in Special Issue
Modeling and Initialization of Nonlinear and Chaotic Fractional Order Systems Based on the Infinite State Representation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Subdiffusion–Superdiffusion Random-Field Transition

Solid State Institute, Technion, Haifa 32000, Israel
Fractal Fract. 2023, 7(10), 745; https://doi.org/10.3390/fractalfract7100745
Submission received: 29 August 2023 / Revised: 3 October 2023 / Accepted: 6 October 2023 / Published: 10 October 2023
(This article belongs to the Special Issue Feature Papers for Mathematical Physics Section)

Abstract

:
A contaminant spreading affected by a random field at boundaries in the comb geometry is considered. The physical effect of the random boundary conditions results in increasing a transport exponent such that the mean squared displacement increases with time from t 1 2 to t 1 2 + 5 α / 2 for real 0 α 1 . This stochastic acceleration due to these space-time-dependent boundary conditions leads to a transition from subdiffusion to superdiffusion. Experimentally, it can be realized by controlling the boundary conditions of 2 D diffusion in the comb geometry.

1. Introduction

Recent experimental investigations have shown that the transport of initial wave packets can be accelerated inside randomly inhomogeneous media [1,2,3]. It also has been shown that this phenomenon results from the generic nature of space-time inhomogeneity of random media, including quantum [3,4,5] and classical transport [4,6,7]. In this paper, diffusion acceleration due to random boundary conditions is studied in the framework of a comb model [7], where the transport can be accelerated by an external random field at the boundaries. This approach is a further generalization of the result obtained in ref. [7]. This phenomenon of diffusion acceleration may have a strong impact on the single-file diffusion of atomic and colloidal systems [8,9], with a general interaction between particles in an out-of-equilibrium situation [10]; see also experimental evidence of anomalous diffusion in these systems [11,12,13], kinetics of surface growth [14,15,16], and fractal elastic properties of material [8,9,16].
A comb model was introduced for understanding anomalous transport in percolating clusters [17,18], and it was considered as a toy model for a porous medium used for the exploration of low-dimensional percolation clusters [17,19] as well. It is a particular example of a non-Markovian phenomenon, which was explained in the framework of continuous-time random walks [18,20,21]. Nowadays, such comb-like models are widely used to describe various experimental applications like transport in low-dimensional composites [22] and the transport of calcium in spiny dendrites [23,24,25,26], as well as the transport of magnesium at the actin polymerization [27]. The branch-shaped models also play an important role in developing an effective comb-shaped configuration of antennas [28], as well as modeling and simulating flows in the cardiovascular and ventilatory systems, which is related to techniques of virtual physiology [29]. See also the extended reviews in refs. [30,31].
Anomalous diffusion on the comb is described by the 2 D probability distribution function (PDF) P = P ( x , y , t ) , and a special behavior is the displacement in the x-direction, which is possible only along the structure axis (x-axis at y = 0 ). An elegant form of the equation, which describes diffusion on the comb-like structure, was introduced in ref. [19] and it reads
t P ( x , y , t ) = D x δ ( y ) 2 x 2 P ( x , y , t ) + D y 2 y 2 P ( x , y , t ) .
Here, D x δ ( y ) is the diffusion coefficient in the x-direction, and D y is the diffusion coefficient in the y-direction. The Dirac δ function in the x component of the diffusion coefficient implies the occurrence of a nonzero current along the backbone at y = 0 only. Thus, this equation describes the transport on the backbone (at y = 0 ), while the fingers play a role of traps. The comb model with infinite length of the fingers describes subdiffusion along the backbone with the mean squared displacement (MSD) increasing with time like t 1 / 2 [17,18,19]. For the finite fingers, subdiffusion is a transient process until time t * , and afterwards, the transport along the backbone corresponds to normal diffusion with the MSD t [27,32].
A generalization of the finite comb model with correlated motion in the fingers affects the diffusive behavior along the backbone in a nontrivial fashion [33,34,35], resulting in quite rich diffusive scenarios of anomalous transport along the backbone. In particular, the MSD is of the order of t μ with 0 < μ < 1 . In this research, we show that nontrivial correlations between the x- and y-directions in the form of a random field at the boundaries of the finite fingers leads to stochastic acceleration/enhancement of the transport, with 1 < μ < 3 . As mentioned above, there are two regimes divided by the time parameter t * , which is by the comb parameters, and it is defined in the ensuing analysis. In this paper, we study the acceleration of subdiffusion when t < t * .

2. A Finite Comb with a Random Field at the Boundaries

In the following, it is convenient to work with the dimensionless comb model equation. Taking into account that the dimension of D x is [ length 3 / time ] and the dimension of D y is [ length 2 / time ] , one obtains the dimensionless time t D y 3 / D x 2 t and coordinates x D y / D x x and y D y / D x y / d , where d is a dimensionless diffusivity along the y-axis. This yields Equation (1) in the dimensionless form
t P ( x , y , t ) = δ ( y ) x 2 P ( x , y , t ) + d y 2 P ( x , y , t ) .
We name it a comb model equation and consider the comb with a finite length of fingers h y h , and the boundary conditions are
d y P ( x , y , t ) | y = h d y P ( x , y , t ) | y = h = W ( x , t ) ,
where W ( x , t ) can be an arbitrary function, which consists of two fluxes at the boundaries y = ± h . We consider it as a random Gaussian process, with δ correlated in space and time
E { W ( x , t ) W ( x , t ) } W ( x , t ) W ( x , t ) = 2 K δ ( x x ) δ ( t t ) .
Here, E { } means the averaging over the Gaussian distribution with the constant variation function K, and the delta correlation functions also reflect the space and time homogeneity. In a general case, K can also be a function of space and time, K = K ( x , t ) . The boundary conditions for the backbone is taken at infinities P ( x = ± , y , t ) = x P ( x = , y , t ) = 0 . Such random reflections at boundaries y = ± h , dependent on time t and the x coordinates, strongly influence the transport along the backbone. This situation with the random boundary conditions (3) can be a realization of the comb structure with the open boundary conditions of the fingers, which are in contact with a random environment (artificial or natural). In particular, it can be relevant to a diffusion problem in neurons, where the transport of ions can be accelerated by an external random field due to synapse fluctuations [7].
Taking into account that the comb transport is a combination of two random processes along the x- and y-directions, one presents the PDF P ( x , y , t ) as a time convolution of these processes. Thus, the PDF in the Laplace domain reads
L P ( x , y , t ) ( s ) = P ˜ ( x , y , s ) = g ˜ ( y , s ) f ˜ ( x , s ) .
Diffusion in the fingers corresponds to the conditions y g ˜ ( y = h ) y g ˜ ( y = h ) 0 (cf. Equation (4)) and g ˜ ( y = 0 , s ) = 1 . Therefore, disregarding (for a moment) the boundary conditions, the solution is of the form
g ˜ ( y , s ) = cosh [ 2 h s / d | y | s / d / d ] / cosh [ 2 h s / d ] ,
which is the fundamental solution of equation s g ˜ = d y 2 g ˜ . An important consequence of the finite comb solution (6) is the existence of two time scales for the transient time t * = h 2 / d . In the dimension time scale, it reads T 0 = h 2 / D y , where t * = T 0 D y 3 D x 2 . In what follows, we assume d = 1 .
Let us consider a coarse-grained function, known as a marginal PDF, which reads
p ( x , t ) = h h P ( x , y , t ) d y .
Integrating Equation (2) with respect to y and taking into account Equation (7) and the boundary conditions (3), one obtains the Langevin equation,
t p ( x , t ) = x 2 P ( x , y = 0 , t ) + W ( x , t ) ,
where W ( x , t ) is the additive space-time-dependent noise. When W ( x , t ) is the additive noise, Equation (8) is known as an Edwards–Wilkinson equation [36]. To obtain this Langevin Edwards–Wilkinson equation in the closed form, one needs to establish a relation between P ( x , y = 0 , t ) and p ( x , t ) . This procedure can be performed in the Laplace domain. Taking into account that g ˜ ( y = 0 , s ) = 1 and P ˜ ( x , y = 0 , s ) = g ˜ ( y , s ) , one obtains
p ˜ ( x , s ) = P ˜ ( x , y = 0 , s ) h h g ˜ ( y , s ) d y .
The integration of Equation (9) with respect to y yields
h h g ˜ ( y , s ) d y = 2 s tanh 2 h s sinh h s cosh 2 h s .
The last expression can be simplified, taking into account the limiting case with h s 1 , which corresponds to the initial time dynamics for t < t * . Note also that t * can be very large.

3. Initial Time Asymptotics

In the case of h s 1 , we have tanh 2 h s 1 , while the second term of the order of e h s in Equation (10) can be neglected. For this limit, when t < t * , one obtains P ˜ ( x , y = 0 , s ) = s 2 p ˜ ( x , s ) , which yields the Langevin Equation (8) in Laplace space as follows
p ˜ ( x , s ) = 1 2 s 1 / 2 x 2 p ˜ ( x , s ) + 1 s W ˜ ( x , s ) + 1 s p 0 .
Performing the Laplace inversion, one obtains the marginal PDF as the space-time-dependent function,
p ( x , t ) = 1 2 Γ ( 1 / 2 ) 0 t ( t t ) 1 / 2 x 2 p ( x , t ) d t + 0 t W ( x , t ) d t + p 0 .
Differentiation with respect to time yields the fractional Edwards–Wilkinson equation,
t p ( x , t ) = 1 2 0 D t 1 2 x 2 p ( x , t ) + W ( x , t ) ,
where 0 D t 1 2 is a so-called fractional Riemann–Liouville derivative; see, e.g, [21,37]. For the real values of 0 < ν < 1 , it reads
0 D t ν f ( t ) = 1 Γ ( 1 ν ) d d t 0 t ( t t ) ν f ( t ) d t .
In our case case, ν = 1 2 . When ν = 0 , Equation (13) reduces to the standard Edwards–Wilkinson equation [36].
An important point of this inferring is the validity of the Laplace transform of the random field W ( x , t ) W ˜ ( x , s ) = L [ W ( x , t ) ] at the condition lim t e s t W ( t ) = 0 , and it can be fulfilled even for strong fluctuations of W ( x , t ) in time.

3.1. Fourier and Laplace Transform of the Marginal PDF

It is also reasonable to perform the Fourier transform with respect to the space coordinates x in Equation (13). Indeed, for an experimental realization of the random-noise acceleration of an initial profile, the random boundary conditions can be constructed by means of the Fourier transform, that is, W ˜ ( x , s ) = F 1 W ˜ ( k , s ) . Taking the limits of the integration at infinity, we suppose that
W ˜ ( x , s ) = e | x | s α / L w ˜ ( x , s ) ,
where L L ( α ) has various physical natures which depend on α , and 0 α 1 . In particular, for α = 0 , it is an effective size of the random-noise influence; when α = 1 , it is a velocity of ballistic motion, while for α = 1 2 , it is the square root of the diffusion coefficient. We also suppose that w ( x , t ) w ( x , t ) = 2 K δ ( x x ) δ ( t t ) is the same as in Equation (4). For α = 0 , it is just when restricted in space that it is convenient to take the boundary conditions at infinity, while the size of an experimental set up is L. Therefore, the random process is tempered in such a way that its influence on the backbone transport is negligible for | x | > L . For α 0 , the random process is a convolution, W ( x , t ) = f ( x , t ) w ( x , t ) , where the delta-correlated process w ( x , t ) is controlled by the Lévy-stable process
f ( x , t ) = 1 2 π i c i c + i e | x | s α / L e s t d s .
In particular, for α = 1 2 , it is the Lévy–Smirnov distribution [31,32]. Some more details of the random process W ( x , t ) are discussed in Appendix A.1 and Appendix A.2.
Then, for the Fourier transforming Equation (11), we have
s p ˜ ( k , s ) = 1 2 s ( k 2 ) p ˜ ( k , s ) + W ^ ( k , s ) + 1 ,
where F [ p 0 ( x ) ] = 1 , and W ^ ( k , s ) is a formal and concise notation of F [ e | x | s α / L w ˜ ( x , s ) ] . Therefore, the marginal PDF in Fourier–Laplace space reads
p ˜ ( k , s ) = 1 + W ^ ( k , s ) s + k 2 s 1 2 / 2 .
This function is random, and its first moment M ˜ 1 ( k , s ) is
M ˜ 1 ( k , s ) = p ˜ ( k , s ) = 1 s + k 2 s 1 2 / 2 ,
where we use that
W ˜ ( k , s ) = 0 e s t e i k x e | x | s α / L w ( x , t ) d x d t = 0 ,
since w ( x , t ) = 0 .
The second moment, M ˜ 2 ( k , s ) , of the marginal PDF in Fourier–Laplace space depends on the second moment W ˜ 2 ( k , s ) , which is
W ˜ 2 ( k , s ) = 0 e s ( t + t ) d t d t e i k ( x + x ) e ( | x | + | x | ) s α / L d x d x w ( x , t ) w ( x , t ) = K s L s α + i k L + L s α i k L = 2 L K s α 1 ( s 2 α + k 2 L 2 )
Therefore, considering the variance in the PDF
Δ p ˜ 2 ( k , s ) = M ˜ 2 ( k , s ) M ˜ 1 2 ( k , s ) ,
we obtain
Δ p ˜ 2 ( k , s ) = s 1 ( s 1 2 + k 2 / 2 ) 2 · 1 + 2 L K s α 1 ( s 2 α + k 2 L 2 ) s 1 ( s 1 2 + k 2 / 2 ) 2 = s 1 ( s 1 2 + k 2 / 2 ) 2 · 2 L K s α 1 ( s 2 α + k 2 L 2 ) .
Taking into account the correlation term (20), we obtain the contribution to the MSD in Laplace space as follows:
Δ x 2 ( s ) = d 2 d k 2 Δ p ˜ 2 ( k , s ) | k = 0 = d 2 d k 2 s 1 ( s 1 2 + k 2 / 2 ) 2 · 2 L K s α 1 ( s 2 α + k 2 L 2 ) k = 0 1 2 = d 2 d k 2 s 1 2 α 1 s 1 2 + k 2 / 2 · 2 L K ( s 2 α + L 2 k 2 ) | k = 0 = 2 L K s 2 1 2 α + 2 L 2 2 L K s ( 3 2 + 5 2 α ) .
Performing the Laplace inversion, we obtain that the contribution to the MSD due to the random boundary conditions leads to superdiffusion,
Δ x 2 ( t ) = 2 L K Γ ( 2 + α ) t ( 1 + α / 2 ) + 2 L 2 2 L K Γ ( 2 + 2 α ) t ( 1 2 + 5 α / 2 ) .
Therefore, for α [ 0 , 1 ] , the random boundary acceleration ranges from normal diffusion with the MSD t to Richardson diffusion with the MSD t 3 .

4. Fokker–Planck Equation

However, this procedure of the straightforward Fourier transform of the Langevin equation is not always feasible, since, in general cases, the random-field W ( x , t ) is not obligatorily absolutely integrable, which is the necessary condition of the Fourier transform. Here, we consider an alternative approach in the framework of a Fokker–Planck Equation (FPE). To this end, we rewrite Equation (11) in such a form that u ( x ) = p ˜ ( x , s ) is an effective coordinate and v ( x ) = d u ( x ) / d x is an effective velocity. In this case, u and v are dynamical variables, and x is an effective time. Note that s plays a role of an arbitrary parameter here. Then, Equation (11) reads
d u ( x ) d x = v ( x ) ,
d v ( x ) d x = 2 s u ( x ) 2 s W ˜ ( x , s ) 2 s p 0 ( x ) .
It is possible to show, following standard recipes (see, e.g., refs. [38,39,40]) that Equation (24) generate a PDF ρ = ρ ( v , u , x ) , which is formally determined by the Gaussian ensemble averaging according to Equation (4),
ρ ( v , u , x ) = E { δ ( v ( x ) v ) δ ( u ( x ) u ) } .
Taking the derivative with respect to x and using standard techniques for the Langevin Equation (24) (see, e.g., ref. [38,39,40]), one obtains the FPE as follows:
x ρ = 2 [ s 1 2 u s 1 2 p 0 ] v ρ v u ρ + 2 K s 2 v 2 ρ .
Some details of the inferring are presented in Appendix A.3.
To estimate the MSD x 2 ( t ) and other fluctuations due to the random-field W ( x , t ) , one looks for the averaged PDF p ( x , t ) and its second moment. It is useful to obtain a system of equations for the moments of u and v from the FPE (26) by introducing a general expression for the moments,
M k , l ( x ) = 0 d u d v u k v l ρ ( u , v , x ) ,
where k , l = 0 , 1 , 2 . Here, we take into account that u is non-negative, while v can be an arbitrary real value. From here, we also obtain the conservation of the total probability, namely d M 0 , 0 ( x ) d x M 0 , 0 = 0 . Substituting u k v l into the FPE (26) and integrating with respect to u and v, one obtains a system of equations for M k , l . For the first moments, these equations are
M 1 , 0 = M 0 , 1 , M 0 , 1 = 2 s 1 2 M 1 , 0 2 s 1 2 p 0 ,
which coincide with averaged Equation (11). If the initial condition is a δ function, p 0 ( x ) = δ ( x ) , the solution for the averaged distribution can be expressed in terms of a Fox H function [21,31,41] in the space-time domain; see Appendix A.4. Note that now the Fourier transform of the averaged values can be performed. The Fourier image of the initial condition is p ^ 0 = F ^ [ p 0 ] = 1 , and the solution in the Laplace domain is
M 1 , 0 ( x , s ) = s 3 4 2 exp | x | 2 s 1 2 ,
M 0 , 1 ( x , s ) = s 1 2 [ 1 2 θ ( x ) ] exp | x | 2 s 1 2 ,
where θ ( x ) is the Heaviside theta function. Taking into account that
M ^ 1 , 0 ( k , s ) = 2 s 1 2 2 s 1 2 + k 2 ,
we obtain the MSD,
x 2 ( t ) = L ^ 1 d 2 M ^ 1 , 0 ( k , s ) d k 2 | k = 0 = t Γ ( 3 / 2 ) .
Next, we estimate the contribution to the MSD from the quadratic fluctuations M 2 , 0 . Since it relates to the PDF p ( x , t ) , our attention here is concentrated on the second moments M 2 , 0 and M 0 , 2 .

Quadratic Fluctuations

Equations for the second moments of u and v from the FPE (26) are as follows:
M 2 , 0 = u 2 v v ρ = 2 M 1 , 1 , M 1 , 1 = 2 s 1 2 u 2 v v ρ + 2 s 1 2 p 0 ( x ) u v v ρ u v 2 u ρ =
2 s 1 2 M 2 , 0 2 s 1 2 M 1 , 0 p 0 ( x ) + M 0 , 2 , M 0 , 2 = 2 s 1 2 u v 2 v ρ + 2 s 1 2 p 0 ( x ) v 2 v ρ + 2 K s 2 v 2 v 2 ρ =
4 s 1 2 M 11 4 s 1 2 p 0 ( x ) M 0 , 1 + 4 K s 2 .
In the matrix form, Equation (31) for the second moments M ( 2 ) = M 2 , 0 , M 1 , 1 , M 0 , 2 T reads
d M ( 2 ) d x = | | A | | M ( 2 ) + f
where f = 0 , 2 s 1 2 M 1 , 0 p 0 ( x ) , 4 s 1 2 M 0 , 1 p 0 ( x ) + 4 K s 2 T , and index T means the vector’s transposition. The matrix | | A | | is determined by the system of Equation (31),
| | A | | = 0 2 0 2 s 0 1 0 4 s 0
Taking into account that p 0 ( x ) = δ ( x ) , it follows that M 1 , 0 ( x , s ) p 0 ( x ) = s 3 4 2 δ ( x ) and M 0 , 1 p 0 ( x ) = 0 . Therefore, the inhomogeneous term is
f ( x ) = 0 , 2 s 5 4 δ ( x ) , 4 K ( x ) s 2 T ,
where it is worth supposing that K is a function of space and time, K = K ( x ) K ( x , s ) . The solution to matrix Equation (32) is
M ( 2 ) ( x , s ) = e | | A | | x 0 x e | | A | | x f ( x ) d x .
A homogeneous solution to Equation (32) is on the order of e λ x and consists of three parts according to the matrix (33), with the characteristic equation λ 3 8 s λ = 0 and the solution
( λ 1 , λ 2 , λ 3 ) = 8 s 1 / 2 , 0 , 8 s 1 / 2 .
According to the boundary conditions for the PDF, M ( 2 ) ( x = ± ) = 0 ; therefore, eigenvalues λ 1 and λ 3 correspond to x > 0 and x < 0 , respectively. The Fourier transform of Equation (31) yields a simple algebraic equation,
| | A ^ | | M ^ ( 2 ) = f ^ ,
where
| | A ^ | | = i k 2 0 2 s i k 1 0 4 s i k , f ^ = 0 2 s 5 4 4 K ^ s 2 .
with the determinant | A ^ | = i k ( k 2 + 8 s ) = i k ( i k + λ 1 ) ( i k + λ 3 ) . The solution for M ^ 2 , 0 ( k , s ) , according to Cramer’s rule [42], reads
M ^ 2 , 0 ( k , s ) = 8 s 5 4 ( k i λ 1 ) ( k i λ 3 ) + 8 K ^ s 2 i k ( i k + λ 1 ) ( i k + λ 3 ) m ^ 1 ( k , s ) + m ^ 2 ( k , s ) .
We study only the second term m ^ 2 ( k , s ) = M ^ 2 , 0 ( k , s ) m ^ 1 ( k , s ) . Performing the Fourier inversion, we consider K ^ ( k ) = K ^ ( k , s ) as follows:
K ^ ( k ) = k 2 ( a + i k ) 3 + ( a i k ) 3 ,
where a = s α / L ; see Section 3.1. This expression has a well-defined cosine transform [43] for K ( x ) , which decays exponentially at infinity, e | x | / L . This is one of many possible realizations of K ( x , t ) (for example, an exotic case is considered in Appendix A.5). Then, we obtain the variation m ^ 2 ( k , s ) as a product of two effective variations: m ^ 2 ( k , s ) = σ ^ 1 ( k , s ) σ ^ 2 ( k ) , where
σ ^ 1 ( k , s ) = 8 i s 2 k k + i 8 s 1 2 k i 8 s 1 2 , σ ^ 2 ( k ) = ( a + i k ) 3 + ( a i k ) 3 .
The Fourier inversion yields
m 2 ( x , s ) = σ 1 ( x x , s ) σ 2 ( x ) d x ,
where σ 2 ( x ) = x 2 e a | x | / 2 , [43], and for σ 1 ( x , s ) , we have [44]
σ 1 ( x , s ) = 1 2 π e i k x σ 1 ( k ) d k = 4 i s 2 π k e i k x d k k + i 8 s 1 2 k i 8 s 1 2 = 2 π i Res e i k z σ 1 ( z ) , z = i 8 s 1 2 ( x > 0 ) + Res e i k z σ 1 ( z ) , z = i 8 s 1 2 ( x < 0 ) = 4 s 2 exp | x | 8 s 1 2 ,
where x = Re ( z ) at the analytic continuation. Thus, the contribution to the MSD in Laplace space reads
Δ x 2 ( s ) = x 2 [ m 2 ( x , s ) ] 1 2 d x .
Taking into account results for σ 1 and σ 2 and denoting b = 8 s 1 2 , we obtain m 2 ( x , s ) in Equation (41) as the result of the following chain of calculations:
1 2 s 2 m 2 ( x , s ) = e b | x x | x 2 e a | x | d x = 2 b 2 a 2 a e b | x | + b e a | x | .
Taking into account that a 1 ( L 1 ) ) , we obtain m 2 ( x , s ) s 3 2 2 e | x | s α / L . Substituting this result in Equation (43), we obtain
Δ x 2 ( s ) = x 2 s 3 4 2 e | x | s α / 2 L d x = 16 2 L 3 s 3 ( α + 1 / 4 ) ,
which eventually yields
Δ x 2 ( t ) = L 1 Δ x 2 ( s ) = 16 2 L 3 Γ ( 3 α + 3 / 4 ) t 3 α 1 4 .
Therefore, for α > 5 12 , transition to superdiffusion takes place due to the random boundary conditions.

5. Conclusions

In this paper, we considered a possible realization of the enhancement of subdiffusion by a space-time random field W ( x , t ) . Experimentally, it can be realized by controlling boundary conditions of the side branches in 2 D diffusion in comb geometry. The physical effect leads to increasing the transport exponent, such that the mean squared displacement (MSD) increases with time from t 1 2 to t 1 2 + 5 α / 2 according to Equation (23) for real 0 α 1 . In this case, the process of acceleration ranges from normal diffusion with the MSD t to Richardson diffusion with the MSD t 3 . It should be admitted that this effect is due to the quadratic fluctuations of the random PDF p ( x , t ) according to the fractional Edwards–Wilkinson Equation (13). This value plays an important role, for example, in the theory of single-file diffusion [8,9], and its time scaling is obtained here in Equation (21). This effect can be also important in studies of calcium transport in spiny dendrites [23,24,25,26], where the interaction between spines and tissues can be modeled by the random boundary conditions in a way that is considered in the present analysis; see also ref. [7]. An important point of this analysis is the validity of the Laplace and Fourier transforms of the random field W ( x , t ) W ˜ ( k , s ) = F L [ W ( x , t ) ] . While the Laplace transform always takes place at the condition lim t e s t W ( t ) = 0 , the situation with the Fourier transform is more sophisticated. As already mentioned, for an experimental realization of the random-noise acceleration of an initial profile, the random boundary conditions can be constructed by means of the Fourier transform, that is, W ˜ ( x , s ) = F 1 W ˜ ( k , s ) . However, in general cases, the randomness is not Fourier-transformable. In this case, the alternative approach in the framework of the effective Fokker–Planck equation is considered in Section 4, such that the MSD x 2 ( t ) and other fluctuations due to the random-field W ( x , t ) are determined by the averaged PDF p ( x , t ) and its second moments according to the system of Equation (31). The physical effect of increasing the transport exponent for the MSD from t 1 2 to t 3 α 1 / 4 is calculated according to Equation (46).
In conclusion, it should be acknowledged that contrary to any experimental realization, where the space-time dependence of the boundary conditions can be controlled, in general (natural) phenomena, the random-noise boundaries are uncontrolled, which results in their unpredictable influence. The acceleration effect considered in Section 4 is just a particular case among any other possible realizations, where this effect either does not take place or can lead to a singular behavior (considered in Appendix A.5.)

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

Appendix A

Appendix A.1. Random Boundary Conditions

The Dirichlet boundary conditions in the fingers are formulated by a tempered random process defined in Equation (15). The tempering is convenient for the analytical treatment of the backbone transport with shifting the boundary conditions at infinity. Therefore, Equation (15) describes a random process in the form of the convolution integral
W ( x , t ) = f ( x , t ) w ( x , t ) ,
where the delta-correlated process w ( x , t ) is controlled by the Lévy-stable process, which is determined by the inverse Laplace transform
f ( x , t ) = L 1 e | x | s α / L = 1 2 π i c i c + i e | x | s α / L e s t d s .
In particular, for α = 1 2 , it is the Lévy–Smirnov distribution [43], f ( x , t ) = L | x | 4 π t 3 e x 2 4 L 2 t . In general case, f ( x , t ) is the Fox H-function (see Appendix A.4) that immediately follows by presenting the exponential in the form of the Fox H-function [30,31] by means of the Mellin–Barnes integral; see Equation (A23). We, however, suggest an alternative calculation. Considering f ( x , t ) as a generalized function, let us present it in a form of series. In this case, we take into account that [45]
L 1 t ν + 1 ( s ) = Γ ( ν ) s ν , Re s > 0 , ν 0 , 1 , 2 , .
Then, from Equations (A2) and (A3), we obtain
f ( x , t ) = n = 0 | x | / L n Γ ( n + 1 ) L 1 s n α = n = 0 lim p n | x | / L p Γ ( p + 1 ) L 1 L t p α 1 Γ ( p α ) = 1 t n = 0 | x | / t α L n Γ ( n + 1 ) Γ ( n α ) , α ( 0 , 1 ) .
The result (A4) can be obtained by the conventional Laplace transform when, instead, (A3) we have
L 1 s ν ( t ) = t ν 1 / Γ ( ν ) , Re ν > 0 .
Then, Equation (A4) becomes [30] (see also the discussion for Equation (A30))
f ( x , t ) = n = 0 | x | / L n Γ ( n + 1 ) L 1 s n α = n = 0 | x | / L n lim p n 1 p Γ ( p + 1 ) L 1 L s p α = n = 0 | x | / L n lim p n Γ ( p + n + 1 ) ( p + n 1 ) p L 1 L s p α = n = 0 | x | L n lim p n ( p + n ) Γ ( p ) · t p α 1 Γ ( p α ) = n = 0 | x | L n lim p n π ( p + n ) ( 1 ) n Γ ( 1 p ) sin [ π ( p + n ) ] · t p α 1 Γ ( p α ) = 1 t n = 0 ( | x | / t α L ) n Γ ( n + 1 ) Γ ( n α ) ,
which coincides exactly with Equation (A4). To some extend, it also proves the validity of Equation (A3).
For α = 0 , we have W ( x , t ) = e | x | / L w ( x , t ) . In this case case, L is the size of the system. For α = 1 , there is the space–time correlation, where f ( x , t ) = δ ( t | x | / L ) .In this case, L is the velocity of the ballistic motion.

Appendix A.2. Random Optical Lattices

We should outline that in the general case of an experimental realization of the random process W ( x , t ) , the latter is characterized by its correlation function,
W ( x , t ) W ( x , t ) = C ( x , t ; x , t ) .
Following the experimental realization of random optical lattices [3], we consider the optical random potential W ( x , t ) as an expansion of a quasiperiodic function. Then, one has [3,4,5]
W ( x , t ) = 1 N m = N N A m exp ( i k m x i ω m t ) + c . c . ,
where the coefficients of the expansion A m are random complex values, while k m and ω m are independent random real values. Denoting averaging over the Gaussian ensemble by , we obtain that A m are controlled by a Gaussian distribution with the averaging property
A m = A m A n = 0 , A m * A n = K δ m , n .
From this property, one obtains for the 1D space-time-dependent potential W ( x , t )
| W ( x . t ) | 2 = 2 K ;
see Equation (4). Note again that this formulation of the random potential is general and corresponds to the experimental setup [3].
Let us define the property of W ( x , t ) by means of the spectral density S ˜ ( k , ω ) of the correlation function C ( x , t ; x , t ) . Following refs. [3,4,5], we present the correlation function (A7) in the following translational invariant in space and time form
C ( x , t ; x , t ) = C ( x x , t t ) = K N m e i k m ( x x ) i ω m ( t t ) + c . c . ,
which, in the limit N , reads
C ( x x , t t ) = d k d ω S ˜ ( k , ω ) e i [ k ( x x ) ω ( t t ) ] + c . c . .
If S ˜ ( k , ω ) = const , then we arrive at Equation (4). To consider the tempered random process, let us consider the spectral density S ˜ ( k , ω ) in the form of Equation (20),
S ˜ ( k , ω ) = S ˜ ( k , s / i ) = 2 L K s α 1 ( s 2 α + k 2 L 2 )
Therefore, the correlation function for, e.g., α = 1 2 , reads
C ( x x , t t ) = K L 2 π 2 i c i c + i s 1 2 e s ( t t ) d k ( s + k 2 L 2 ) e i k ( x x ) = K 2 π i c i c + i s 1 e s ( t t ) e | x x | s 1 2 / L = K · erfc | x x | ( t t ) 1 2 / 2 L K π e ( x x ) 2 4 ( t t ) L 2 ,
which decays according to the Gaussian function of the scaling argument x 2 / t . Note that the form (A13) is not mandatory and, for example, it can be as follows: 2 L K s α s 2 α + k 2 L 2 . Then, the correlation function (A14) behaves according to the Lévy–Smirnov density,
C ( x x , t t ) = K L | x x | 4 π ( t t ) 3 e ( x x ) 2 4 L 2 ( t t ) .

Appendix A.3. Fokker–Planck Equation

Let us differentiate the PDF (25) with respect to x. Then, we have
d d x ρ ( u , v , x ) = { d d u ( x ) δ ( u ( x ) u ) d u ( x ) d x δ ( v ( x ) v ) } + + { δ ( u ( x ) u ) d d v ( x ) δ ( v ( x ) v ) d v ( x ) d x } .
Using properties of the Dirac δ function, we replace derivatives with respect to u ( x ) and v ( x ) by u and v and take them out of the averaging procedure. We also take into account the Langevin Equation (24). Then, Equation (A16) reads
x ρ ( u , v , x ) = u v δ ( u ( x ) u ) δ ( v ( x ) v ) 2 v [ s 1 2 u s 1 2 p 0 ] δ ( u ( x ) u ) δ ( v ( x ) v ) + 2 s 1 2 v δ ( u ( x ) u ) δ ( v ( x ) v ) W ˜ ( x , s ) = v u ρ ( u , v , x ) 2 [ s 1 2 u s 1 2 p 0 ] v ρ ( u , v , x ) + 2 s 1 2 v δ ( u ( x ) u ) δ ( v ( x ) v ) W ˜ ( x , s ) .
The last term in Equation (A17) needs special care, and we present the average value according to the Furutsu–Novikov theorem [46,47]
δ ( u ( x ) u ) δ ( v ( x ) v ) W ˜ ( x , s ) = d x δ ( u ( x ) u ) d d v ( x ) δ ( v ( x ) v ) δ v ( x ) δ W ˜ ( x , s ) × W ˜ ( x , s ) W ˜ ( x , s )
Taking into account Equation (24b), the variational derivative δ v ( x ) δ W ˜ ( x , s ) yields
δ v ( x ) δ W ˜ ( x , s ) = δ δ W ˜ ( x , s ) x 2 s 1 2 u ( x ) 2 s 1 2 W ˜ ( x , s ) 2 s 1 2 p 0 ( x ) d x = 2 s 1 2 x δ W ˜ ( x , s ) δ W ˜ ( x , s ) d x = 2 s 1 2 x δ ( x x ) d x = s 1 2 .
Here, we take into account that the δ function yields 1 2 at the (upper) integration limit. The next step is a calculation of the correlation function in Laplace space W ˜ ( x , s ) W ˜ ( x , s ) , which is defined in the real-time space,
W ( x , t ) W ( x , t ) = 2 K δ ( x x ) δ ( t t ) ,
(cf. Equation (4)). Then, we have
W ˜ ( x , s ) W ˜ ( x , s ) = 0 e s t d t 0 e s t d t W ( x , t ) W ( x , t ) = 2 K δ ( x x ) 0 e s t d t 0 e s t d t δ ( t t ) = K s 1 δ ( x x ) .
Then, substituting the results of Equation (A19) and (A21) in the integral in Equation (A18) yields K s 3 2 . Substituting the obtained result in Equation (A17), one obtains the FPE in Equation (26), which reads
x ρ = 2 [ s 1 2 u s 1 2 p 0 ] v ρ v u ρ + 2 K s 2 v 2 ρ .

Appendix A.4. Fox H-Function

Here, we obtain the solution for M 1 , 0 in terms of the Fox H-function [21,30,31,41]. The Fox H-function is defined by means of the following Mellin–Barnes integral [41,48]
H p , q m , n ( z ) = H p , q m , n z ( a 1 , A 1 ) , . . . , ( a p , A p ) ( b 1 , B 1 ) , . . . , ( b q , B q ) = H p , q m , n z ( a p , A p ) ( b q , B q ) = 1 2 π i Ω ϑ ( s ) z s d s ,
where
ϑ ( s ) = j = 1 m Γ ( b j + B j s ) j = 1 n Γ ( 1 a j A j s ) j = m + 1 q Γ ( 1 b j B j s ) j = n + 1 p Γ ( a j + A j s ) ,
0 n p , 1 m q , a i , b j C , A i , B j R + , i = 1 , . . . , p , j = 1 , . . . , q . The integration contour Ω starts at c i and finishes at c + i , separating the poles of the Gamma function Γ ( b j + B j s ) , j = 1 , . . . , m with those of the Gamma function Γ ( 1 a i A i s ) , i = 1 , . . . , n .
Let us consider Equation (28) for the first moments of the random PDF, which are
M 1 , 0 = M 0 , 1 , M 0 , 1 = 2 s 1 2 M 1 , 0 2 s 1 2 p 0 ,
where prime means the first derivative with respect to x. The initial condition is p 0 ( x ) = δ ( x ) , which ensures the nontrivial solution of the equations. Differentiating the first equation and taking into account the second one, we obtain
M 1 , 0 = 2 s 1 2 M 1 , 0 2 s 1 2 p 0 .
Note that now the Fourier transform of the averaged values can be performed,
F [ M 1 , 0 ( x ) ] ( ) k ) = M 1 , 0 ( k ) .
Therefore, the Fourier transform of Equation (A26) yields
M 1 , 0 ( k , s ) = s 1 2 s 1 2 + k 2 / 2 ,
where F [ p 0 ( x ) ] = 1 . Performing the inverse Fourier transform, we arrive at the solution in Equation (29a), namely
M 1 , 0 ( x , s ) = s 3 4 2 exp | x | 2 s 1 2 ,
which stands for the Laplace inversion. However, we first perform the inverse Laplace transform of Equation (A27), since Equation (A27) is nothing more than the Laplace image of the one-parameter Mittag–Leffler function E α ( z ) with the index α = 1 2 ; see (V. 3) of ref. [49]. Thus, we have
M 1 , 0 ( k , t ) = 1 2 π i c i c + i s 1 2 e s t d s s 1 2 + k 2 / 2 = E 1 2 , 1 k 2 t 1 2 / 2 E 1 2 k 2 t 1 2 / 2 .
To perform the Fourier inverse transform, which is the cosine transformation, the Mittag–Leffler function is presented in the form of the Fox H-function, for which the cosine transformation is well-defined; see, e.g., refs. [31,41]. Then, we have the chain of transformations establishing the relation between the two-parameter Mittag–Leffler function and the Fox H-function,
E α , β ( z ) = n = 0 ( z ) n Γ ( n α + β ) = n = 0 lim ξ n ( ξ + n ) π sin ( π ξ ) Γ ( β α ξ ) z ξ = = 1 2 π i c i c + i Γ ( ξ ) Γ ( 1 ξ ) Γ ( β α ξ ) z ξ d ξ = H 2 , 1 1 , 1 z ( 0 , 1 ) ( 0 , 1 ) , ( 1 β , α )
Here, in the second line, we used the limit
lim ξ n [ ( ξ + n ) π ] / sin ( ξ π ) = lim ξ n ( 1 ) k [ ( ξ + n ) π ] / sin [ ( ξ + n ) π ]
and Euler’s reflection formula Γ ( p ) Γ ( 1 p ) = π sin ( p π ) , [49]. Correspondingly, for the one-parameter Mittag–Leffler function in Equation (A29) with β = 1 and α = 1 2 , we have
E 1 2 k 2 t 1 2 / 2 = H 1 , 2 1 , 1 k 2 t 1 2 / 2 ( 0 , 1 ) ( 0 , 1 ) , ( 0 , 1 2 ) .
Then, taking into account that the validity condition for the the cosine Mellin transform of the Fox H-function is fulfilled, we obtain
0 k ρ 1 cos ( k x ) H p , q m , n a k δ ( a p , A p ) ( b q , B q ) d k = = π x ρ H q + 1 , p + 2 n + 1 , m x δ a ( 1 b q , B q ) , ( 1 + ρ 2 , δ 2 ) ( ρ , δ ) , ( 1 a p , A p ) , ( 1 + ρ 2 , δ 2 ) .
Therefore, the cosine transform of Equation (A31) yields
0 cos ( k | x | ) E 1 2 , 1 k 2 t 1 2 / 2 d k = 0 cos ( k | x | ) H 2 , 1 1 , 1 k 2 t 1 2 / 2 ( 0 , 1 ) ( 0 , 1 ) , ( 0 , 1 2 ) d k = π | x | H 3 , 3 2 , 1 4 x 2 t ( 1 , 1 ) , ( 1 , 1 2 ) , ( 1 , 1 ) ( 1 , 2 ) , ( 1 , 1 ) , ( 1 , 1 ) = π | x | H 1 , 1 1 , 0 4 x 2 t ( 1 , 1 2 ) ( 1 , 2 ) ,
where the reduction formula for the Fox H-function is used in the second line.

Appendix A.5. Giant Fluctuations

We consider specific (exotic) solutions to Equation (36)
| | A ^ | | M ^ ( 2 ) = f ^ ,
where
| | A ^ | | = i k 2 0 2 s i k 1 0 4 s i k , f ^ = 0 2 s 5 4 4 K ^ s 2 .
with the determinant | A ^ | = i k ( k 2 + 8 s ) = i k ( i k + λ 1 ) ( i k + λ 3 ) . The solution for M ^ 2 , 0 ( k , s ) , according to the Cramer’s rule, reads
M ^ 2 , 0 ( k , s ) = 8 s 5 4 ( k i λ 1 ) ( k i λ 3 ) + 8 K ^ s 2 i k ( i k + λ 1 ) ( i k + λ 3 ) m ^ 1 ( k , s ) + m ^ 2 ( k , s ) .
As in the main text, we only study the second term m ^ 2 ( k , s ) = M ^ 2 , 0 ( k , s ) m ^ 1 ( k , s ) . Performing the Fourier inversion, we consider two limiting cases of the diffusion coefficient K: (i) in the first case, K ( x ) is constant, that is K ^ ( k ) = K δ ( k ) ; (ii) in the second case, we consider that K ^ is constant, K ^ = K = const . In the first case, the Fourier inversion yields integration with the Dirac delta function: 1 k δ ( k ) = d δ ( k ) d k ; see, e.g, ref. [50]. Then, from Equation (A36), we have
1 2 π m ^ 2 ( k , s ) e i k x d k = 4 K s 2 i π d d k e i k x ( i k + λ 1 ) ( i k + λ 3 ) | k = 0 = K 2 π x s 5 2 .
Then, from Equation (A37), we obtain m 2 ( x , t ) = K 2 π Γ ( 5 / 2 ) x t 3 2 , which eventually yields the contribution to the MSD,
Δ x 2 ( t ) = 0 L m 2 ( x , t ) x 2 L 7 2 t 3 4 .
The finite length of the backbone L is the gain parameter here. However, this exotic case is considered to demonstrate the alternative Fokker–Planck equation approach for the random boundary acceleration only.
For the second case with K ^ = K , the Fourier transformation yields
m 2 ( x , s ) = 1 2 π m ^ 2 ( k , s ) e i k x d k = 4 K s 2 π e i x k d k i k k + i 8 s 1 2 k i 8 s 1 2 = 2 π i × Res e i k z m 2 ( z , s ) , z = i 8 s 1 2 + e i k z m 2 ( x , s ) , z = i 8 s 1 2 + e i k z m 2 ( z , s ) , z = 0 = K s 5 2 1 + e | x | 8 s 1 2
The first term in square brackets leads to the singular behavior that is in complete analogy with the first case, and its contribution to the MSD is on the order of Δ x 2 ( t ) K 1 2 L t 1 4 , which in the limit L diverges. The contribution to the MSD of the second term in the square brackets is K t 2 Γ ( 3 / 2 ) , which does not contain any physical effect. These two exotic examples can be a crude explanation of giant fluctuations in nature, where the gain parameter is the scale size L.

References

  1. Levi, L.; Rechtsman, M.; Freedman, B.; Schwartz, T.; Manela, O.; Segev, M. Disorder-enhanced transport in photonic quasicrystals. Science 2011, 332, 1541–1544. [Google Scholar] [CrossRef] [PubMed]
  2. Rechtsman, M.; Levi, L.; Freedman, B.; Schwartz, T.; Manela, O.; Segev, M. Disorder enhanced transport in photonic quasicrystals. Opt. Photonics News 2011, 22, 33. [Google Scholar] [CrossRef]
  3. Levi, L.; Krivolapov, Y.; Fishman, S.; Segev, M. Hyper-transport of light and stochastic acceleration by evolving disorder. Nat. Phys. 2012, 8, 912–917. [Google Scholar] [CrossRef]
  4. Krivolapov, Y.; Levi, L.; Fishman, S.; Segev, M.; Wilkinson, M. Super-diffusion in optical realizations of Anderson localization. N. J. Phys. 2012, 14, 043047. [Google Scholar] [CrossRef]
  5. Iomin, A. Hyperdiffusion of quantum waves in random photonic lattices. Phys. Rev. E 2015, 92, 022139. [Google Scholar] [CrossRef]
  6. Arvedson, E.; Wilkinson, M.; Mehlig, B.; Nakamura, K. Staggered Ladder Spectra. Phys. Rev. Lett. 2006, 96, 030601. [Google Scholar] [CrossRef]
  7. Iomin, A. Richardson diffusion in neurons. Phys. Rev. E 2019, 100, 010104. [Google Scholar] [CrossRef]
  8. Kollmann, M. Single-file Diffusion of Atomic and Colloidal Systems: Asymptotic Laws. Phys. Rev. Lett. 2003, 90, 180602. [Google Scholar] [CrossRef]
  9. Lizana, L.; Ambjönsson, T.; Taloni, A.; Barkai, E.; Lomholt, M.A. Foundation of fractional Langevin equation: Harmonization of a many-body problem. It. Phys. Rev. E 2010, 81, 051118. [Google Scholar] [CrossRef]
  10. Poncet, A.; Grabsch, A.; Illien, P.; Bénichou, O. Generalized Correlation Profiles in Single-File Systems. Phys. Rev. Lett. 2021, 127, 220601. [Google Scholar] [CrossRef]
  11. Kukla, V.V.; Kornatowski, J.; Demuth, D.; Girnus, I.I.; Pfeifer, H.; Rees, L.V.C.; Schunk, S.; Unger, K.K.; Karger, J. NMR Studies of Single-File Diffusion in Unidimensional Channel Zeolites. Science 1996, 272, 702–704. [Google Scholar] [CrossRef] [PubMed]
  12. Hahn, K.; Kárger, J.; Kukla, V. Single-File Diffusion Observation. Phys. Rev. Lett. 1996, 76, 2762–2765. [Google Scholar] [CrossRef]
  13. Wei, Q.-H.; Bechinger, C.; Leiderer, P. Single-File Diffusion of Colloids in One-Dimensional Channels. Science 2000, 287, 625–627. [Google Scholar] [CrossRef] [PubMed]
  14. Nattermann, T.; Tang, L.-H. Kinetic surface roughening. I. The Kardar-Parisi-Zhang equation in the weak-coupling regime. Phys. Rev. A 1992, 45, 7156–7161. [Google Scholar] [CrossRef] [PubMed]
  15. Leith, J.R. Fractal scaling of fractional diffusion processes. Signal Process. 2003, 83, 2397–2409. [Google Scholar] [CrossRef]
  16. Taloni, A.; Chechkin, A.; Klafter, J. Correlations in a generalized elastic model: Fractional Langevin equation approach. Phys. Rev. E 2010, 82, 061104. [Google Scholar] [CrossRef]
  17. White, S.R.; Barma, M. Field-induced drift and trapping in percolation networks. J. Phys. A 1984, 17, 2995–3008. [Google Scholar] [CrossRef]
  18. Weiss, G.H.; Havlin, S. Some properties of a random walk on a comb structure. Phys. A 1986, 134, 474–482. [Google Scholar] [CrossRef]
  19. Arkhincheev, V.E.; Baskin, E.M. Anomalous diffusion and drift in the comb model of percolation clusters. Sov. Phys. JETP 1991, 73, 161–165. [Google Scholar]
  20. Montroll, E.W.; Shlesinger, M.F. The wonderful wold of random walks. In Studies in Statistical Mechanics, 11th ed.; Lebowitz, J., Montroll, E.W., Eds.; Noth–Holland: Amsterdam, The Neetherlands, 1984. [Google Scholar]
  21. 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]
  22. Shamiryan, D.; Baklanov, M.R.; Lyons, P.; Beckx, S.; Boullart, W.; Maex, K. Diffusion of solvents in thin porous films. Colloids Surf. A 2007, 300, 111–116. [Google Scholar] [CrossRef]
  23. Yuste, R. Dendritic Spines; MIT Press: Cambridge, MA, USA, 2010. [Google Scholar]
  24. Santamaria, F.; Wils, S.; De Shutter, E.; Augustine, G.J. Anomalous Diffusion in Purkinje Cell Dendrites Caused by Spines. Neuron 2006, 52, 635–638. [Google Scholar] [CrossRef] [PubMed]
  25. Santamaria, F.; Wils, S.; De Shutter, E.; Augustine, G.J. The diffusional properties of dendrites depend on the density of dendritic spines. Eur. J. Neurosci. 2011, 34, 561–568. [Google Scholar] [CrossRef] [PubMed]
  26. Iomin, A.; Méndez, V. Reaction-subdiffusion front propagation in a comblike model of spiny dendrites. Phys. Rev. E 2013, 88, 012706. [Google Scholar] [CrossRef]
  27. Iomin, A.; Zaburdaev, V.; Pfohl, T. Reaction front propagation of actin polymerization in a comb-reaction system. Chaos, Solitons Fractals 2016, 92, 115–122. [Google Scholar] [CrossRef]
  28. Chu, L.C.Y.; Guha, D.; Antar, Y.M.M. Comb-shaped circularly polarised dielectric resonator antenna. IEEE Electron. Lett. 2006, 42, 785–787. [Google Scholar] [CrossRef]
  29. Thiriet, M. Tissue Functioning and Remodeling in the Circulatory and Ventilatory Systems; Springer: New York, NY, USA, 2013. [Google Scholar]
  30. Iomin, A.; Méndez, V.; Horsthemke, W. Fractional Dynamics in Comb-like Structures; World Scientific: Singapore, 2018. [Google Scholar]
  31. Sandev, T.; Iomin, A. Special Functions of Fractional Calculus; World Scientific: Singapore, 2022. [Google Scholar]
  32. Bouchaud, J.-P.; Georges, A. Anomalous Diffusion in Disordered Media: Statistical Mechanisms, Models and Physical Applications. Phys. Rep. 1990, 195, 127–293. [Google Scholar] [CrossRef]
  33. Iomin, A. Subdiffusion on a fractal comb. Phys. Rev. E 2011, 83, 052106. [Google Scholar] [CrossRef]
  34. Méndez, V.; Iomin, A. Comb-like models for transport along spiny dendrites. Chaos Solitons Fractals 2013, 53, 46–51. [Google Scholar] [CrossRef]
  35. Ribeiro, H.V.; Tateishi, A.A.; Alves, L.G.A.; Zola, R.S.; Lenzi, E.K. Investigating the interplay between mechanisms of anomalous diffusion via fractional Brownian walks on a comb-like structure. N. J. Phys. 2014, 16, 093050. [Google Scholar] [CrossRef]
  36. Edwards, S.F.; Wilkinson, D.R. The surface statistics of a granular aggregate. Proc. R. Soc. London Ser. A 1982, 381, 17–31. [Google Scholar]
  37. Podlubny, I. Fractional Differential Equations; Academic Press: San Diego, CA, USA, 1999. [Google Scholar]
  38. Kliatskin, V.I. Stochastic Equations and Waves in Randomly Inhomogeneous Media; Nauka: Moscow, Russia, 1980. (In Russian) [Google Scholar]
  39. Zinn-Justin, J. Quantum Field Theory and Critical Phenomena; Claredon Press: Oxford, UK, 1990. [Google Scholar]
  40. Rytov, S.M.; Kravtsov, Y.A.; Tatarskii, V.I. Principles of Statistical Radiophysics; Springer: Berlin/Heidelberg, Germany, 1987. [Google Scholar]
  41. Haubold, H.J.; Mathai, A.M.; Saxena, R.K. Mittag-Leffler Functions and Their Applications. J. Appl. Math. 2011, 2011, 298628. [Google Scholar] [CrossRef]
  42. Lankaster, P. Theory of Matrices; Academic Press: New York, NY, USA, 1969. [Google Scholar]
  43. Bateman, H.; Erdélyi, A. Tables of Integral Transformations; McGraw-Hill: New York, NY, USA, 1954; Volume I. [Google Scholar]
  44. Sveshnikov, A.; Tikhonov, A. Theory of Functions of a Complex Variable; Nauka: Moscow, Russia, 1974. [Google Scholar]
  45. Brychkov, Y.A.; Prudnikov, A.P. Integral Transforms of Generalised Functions; Nauka: Moscow, Russia, 1977. [Google Scholar]
  46. Furutsu, K. On the theory of radio wave propagation over inhomogeneous earth. J. Res. Nat. Bur. Stand 1963, 67, 39–62. [Google Scholar]
  47. Novikov, E.A. Functionals and the Random-Force Method in Turbulence Theory. Sov. Phys. JETP 1965, 20, 1290–1294. [Google Scholar]
  48. Fox, C. The G and H functions as symmetrical Fourier kernels. Trans. Am. Math. Soc. 1961, 98, 395–429. [Google Scholar]
  49. Bateman, H.; Erdélyi, A. Higher Transcendental Functions [Volumes I–III]; McGraw-Hill: New York, NY, USA, 1953. [Google Scholar]
  50. Madelung, E. Die Mathematischen Hilfsmittel Des Physikers; Springer: Berlin/Göttingen, Germany, 1957. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Iomin, A. Subdiffusion–Superdiffusion Random-Field Transition. Fractal Fract. 2023, 7, 745. https://doi.org/10.3390/fractalfract7100745

AMA Style

Iomin A. Subdiffusion–Superdiffusion Random-Field Transition. Fractal and Fractional. 2023; 7(10):745. https://doi.org/10.3390/fractalfract7100745

Chicago/Turabian Style

Iomin, Alexander. 2023. "Subdiffusion–Superdiffusion Random-Field Transition" Fractal and Fractional 7, no. 10: 745. https://doi.org/10.3390/fractalfract7100745

Article Metrics

Back to TopTop