Home Fractional Fokker-Planck-Kolmogorov equations associated with SDES on a bounded domain
Article Publicly Available

Fractional Fokker-Planck-Kolmogorov equations associated with SDES on a bounded domain

  • Sabir Umarov EMAIL logo
Published/Copyright: October 31, 2017

Abstract

This paper is devoted to the fractional generalization of the Fokker-Planck equation associated with a nonlinear stochastic differential equation on a bounded domain. The driving process of the stochastic differential equation is a Lévy process subordinated to the inverse of Lévy’s mixed stable subordinators. The Fokker-Planck equation is given through the general Waldenfels operator, while the boundary condition is given through the general Wentcel’s boundary condition. As a fractional operator a distributed order differential operator with a Borel mixing measure is considered. In the paper fractional generalizations of the Fokker-Planck equation are derived and the existence of a unique solution of the corresponding initial-boundary value problems is proved.

1 Introduction

The Fokker-Planck equation, including fractional versions, play an important role in the modeling of various random processes. Its applications can be found in statistical physics [6, 13, 35], quantum mechanics [6, 29], biology [7, 26], finance [23, 35], just to mention a few. The most of these applications assume an idealistic model in which a random quantity may take values in the whole d-dimensional space ℝd. Fractional generalizations of the Fokker-Planck equation in this case are obtained in works [9, 12, 14, 15, 24, 37]; see also references therein. The connection of a wide class of fractional Fokker-Planck equations with their associated stochastic differential equation driven by time-changed Lévy processes was studied in paper [14].

What concerns stochastic processes in a bounded domain, they are also spread out broadly, an example of which is a diffusion in a bounded region. The essential difference of the stochastic process in a bounded domain from the case of stochastic processes in ℝd is the influence of the boundary or near boundary processes to the whole picture. There is a rich literature on the stochastic processes and the associated (non fractional) Fokker-Planck equations on a bounded domain; see recent monograph by Taira [30].

In the present paper we will discuss fractional generalizations of Fokker-Planck-Kolmogorov (FPK for short) equations associated with nonlinear stochastic differential equations on a bounded domain with a smooth boundary. The fractional diffusion and associated FPK equation on a bounded domain was studied in papers [1, 8, 10, 21, 32] in various particular cases. Our goal in this paper is to study general initial-boundary value problems describing stochastic processes undergoing inside a bounded region, as well as on its boundary. We also discuss initial-boundary value problems for fractional FPK equations associated with stochastic differential equations driven by fractional Brownian motion on a bounded domain. Note that fractional FPK equation in the whole space ℝd associated with stochastic processes driven by fractional Brownian motion was considered in papers [2, 5, 15, 16].

For reader’s convenience we start the discussion with a well known particular cases of the influence of the boundary and related boundary conditions. Consider a nonlinear stochastic differential equation

dXt=b(t,Xt)dt+σ(t,Xt)dBt,X0=x,(1.1)

where x ∈ ℝd is a random vector independent of m-dimensional Brownian motion Bt, and the mappings

b:(0,)×RdRd,σ:(0,)×RdRd×m

satisfy Lipschitz and linear growth conditions. Namely, there exist positive constants C1 and C2 such that for all x, y ∈ ℝd the inequalities

k=1d|bk(t,x)bk(t,y)|+k=1d=1m|σk,(t,x)σk,(t,y)|C1|xy|,(1.2)

and

k=1d|bk(t,x)|+k=1d=1m|σk,(t,x)|C2(1+|x|)(1.3)

hold. The stochastic differential equation (1.1) is understood as

Xt=X0+0tb(s,Xs)ds+0tσ(s,Xs)dBs,(1.4)

with the second integral in the sense of Itô.

The backward Kolmogorov equation associated with SDE (1.1) in terms of the density function u(t, x) of Xtx = (Xt|X0 = x) is

u(t,x)t=A(t)u(t,x),t>0,xRd,(1.5)

where t-dependent differential operator A(t) is defined in the form

A(t)=k=1dbk(t,x)xk+j,k=1daj,k(t,y)2xjxk,(1.6)

with (d × d)-matrix-function {aj,k(t, x), j, k = 1 … d}, coinciding with the matrix-function 12σ(t, x) σT(t, x). Here σT(t, x) is the matrix transposed to the matrix-function σ(t, x). The corresponding Fokker-Planck equation, or forward Kolmogorov equation, has the form

u(t,x)t=A(t)u(t,x),t>0,xRd,(1.7)

where A* is the formally adjoint operator to A(t). Both equations (1.5) and (1.7) are accompanied with the initial condition

u(0,x)=u0(x),xRd,(1.8)

where u0(x) is the density function of the initial vector X0. Unifying the terminology used in these two equations we call the pair of equations (1.5) and (1.7)the FPK equation, the term used throughout the current paper.

If the solution process Xt of the stochastic differential equation in (1.1) is allowed to change only in a bounded region Ω ⊂ ℝd with a smooth boundary Ω, then the probability ℙ(Xt ∈ ℝd \ Ω) of Xt being out of the region Ω is zero. The associated FPK equation in this case needs to be supplemented by boundary conditions. In order to see in what form the boundary conditions emerge, we need to introduce the notion of probability current, a d-dimensional vector field Φ(t, x), components of which are defined by

Φk(t,x)=bk(t,x)p(t,x)j=1dyj[aj,k(t,x)p(t,x)],k=1,,d.

Using the probability current, one can write forward FPK equation (1.7) in the form of a conservation law:

p(t,x)t+k=1dΦk(t,x)xk=0,

or, the same,

p(t,x)t+Φ(t,x)=0,

where =(x1,,xd) is the gradient operator and the symbol “” means the dot product of two vector objects.

Let S be a (d−1)-dimensional hyper-surface in Ω, and nx, xS, be an outward normal to S at the point xS. Then the total flow of probabilities through the hyper-surface S can be calculated by the surface integral

SΦ(t,x)nxdS.

Two boundary conditions are common in the study of stochastic processes in a bounded region:

  1. reflecting boundary condition, and

  2. absorbing boundary condition.

The reflecting boundary condition means that there is no probability flow across the boundary, leading to the condition

Φ(t,x)nx=0,t>0,xΩ.(1.9)

In this case Xt will stay in the region Ω forever and be reflected when Xt reaches the boundary Ω. The absorbing boundary means that Xt will be absorbed by boundary as soon as Xt reaches the boundary, leading to the condition

p(t,x)=0.t>0,xΩ.(1.10)

Let a boundary operator 𝓑 be defined in the form

Bφ(x)=μ(x)φ(x)+ν(x)φnx,t>0,xΩ,(1.11)

where the functions μ(x) and ν(x) are continuous on the boundary Ω, satisfying some physical conditions discussed below, and nx is the outward normal at point xΩ. In order to define operators that will be used in the FPK equations associated with the SDEs in a bounded domain we introduce the following spaces:

CB2(Ω):={φC2(Ω):Bφ(x)=0,xΩ},

with the boundary operator 𝓑 defined in (1.11). Introduce the operator A𝓑(t) on the space CB2(Ω) :

AB(t):={A(t)φ(x)=k=1dbk(t,x)φ(x)xk+j,k=1daj,k(t,x)2φ(x)xjxk,xΩ}.

In other words the operator A𝓑(t) formally is the same as the operator A(t) defined in (1.6), but with the domain Dom(A𝓑(t)) = CB2(Ω) for each fixed t > 0. The operator A𝓑 is linear and maps the space CB2(Ω) to C(Ω).

Both boundary conditions for stochastic processes in a domain with absorbing or reflecting boundaries discussed above can be reduced to the boundary condition

Bp(t,x)=0,t>0,xΩ.

Therefore, the forward and backward FPK equations in the case of bounded domain in terms of the density function u(t, x) of Xtx = (Xt|X0 = x) can be formulated with the help of the operator A𝓑. Namely, the backward FPK equation is given by

u(t,x)t=AB(t)u(t,x),t>0,xΩ,

and the forward FPK equation is given by

u(t,x)t=AB(t)u(t,x),t>0,xΩ,

where AB is the formally adjoint operator to A𝓑.

Thus, the explicit form of the FPK equation associated with a SDE with drift and diffusion coefficients independent of the time variable t is given by the initial-boundary value problem

u(t,x)t=j=1dbj(x)u(t,x)xj+i,j=1dai,j(x)2u(t,x)xixjt>0,xΩ,(1.12)
Bu(t,x)μ(x)u(t,x)n+ν(x)u(t,x)=0,t>0,xΩ,(1.13)
u(0,x)=δ0(x),xΩ,(1.14)

where bj(x), j = 1, …, d, are drift coefficients and ai,j(x), i, j = 1, …, d, are diffusion coefficients; the functions μ(x′) and ν(x′) are continuous functions defined on the boundary Ω. It is well known that the stochastic process solving SDE corresponding to the FPK equation in (1.12)-(1.14) is continuous. Such a stochastic process represents a diffusion process of a Markovian particle.

The paper is organized as follows. Section 2 provides necessary preliminary information on general (integer order) FPK equations through the Waldenfels operator and Wentcel’s boundary conditions. The corresponding stochastic processes represent solutions of associated stochastic differential equations on a bounded domain with combined continuous diffusion, jump, and other relevant phenomena. The main results of the paper is in Section 3. In this section we derive initial-boundary value problems for fractional order FPK equations. The associated stochastic processes represent solutions of stochastic differential equations in a bounded domain with an appropriate time-changed driving processes. In this section we also prove the existence of a unique solution to the obtained initial-boundary value problems.

2 Preliminaries and auxiliaries

FPK equation (1.12)-(1.14) does not take into account jumping events and some specific phenomena (diffusion on the boundary, jumps on the boundary or into the domain, viscosity, etc.) which may occur on the boundary. The general case with jump and viscosity effects involve pseudo-differential operators in the FPK equation and in the boundary condition. The theory of pseudo-differential operators was first developed by Kohn and Nirenberg [20] and Hörmander [17]. An operator A is called pseudo-differential with a symbol σA(x, ξ), if

Af(x)=1(2π)dRdσA(x,ξ)eixξf~(ξ)dξ,(2.1)

where f~ (ξ) is the Fourier transform of f:

f~(ξ)=Rdf(x)eixξdx.

The symbols of pseudo-differential operators in [20, 17] are smooth and satisfy some growth conditions as |ξ| → ∞. However, symbols of pseudo-differential operators considered in the present paper are not smooth. We refer the reader to works [3, 18, 33] where appropriate classes of symbols are studied.

For simplicity we assume that Ω ⊂ ℝd is a bounded domain with a smooth boundary Ω ⊂ ℝd−1. In the general case the (backward) FPK equation can be formulated as follows:

u(t,x)t=A(x,D)u(t,x),t>0,xΩ,(2.2)
W(x,D)u(t,x)=0,t>0,xΩ,(2.3)
u(0,x)=u0(x),xΩ,(2.4)

where 𝓐(x, D) is a second order Waldenfels operator acting on the space of twice differentiable functions defined on Ω, that is

A(x,D)φ(x)=c0(x)φ(x)+i=1dbi(x)φ(x)xi+ij=1daij(x)2φ(x)xixj+Ωφ(y)ϕ(x,y)φ(x)i=1d(yixi)φ(x)yiν(x,dy),xΩ,(2.5)

and 𝓦(x′, D) is a boundary pseudo-differential operator defined on the space of twice differentiable functions defined on Ω through the local coordinates x′ = (x1, …, xd−1) ∈ Ω (see [30]) :

W(x,D)=Q(x,D)+μ(x)nδ(x)A(x,D)+Γ1(x,D)+Γ2(x,D),(2.6)

with (pseudo)-differential operators

Q(x,D)φ(x)=j,k=1d1αj,k(x)2φ(x)xjxk+j=1d1β(x)φ(x)xj+γ(x)φ(x),(2.7)
Γ1(x,D)φ(x)=Ωφ(y)τ1(x,y)φ(x)+k=1d1(ykxk)φ(x)xkν1(x,dy),(2.8)

and

Γ2(x,D)φ(x)=Ωφ(y)τ2(x,y)φ(x)+k=1d1(ykxk)φ(x)xkν2(x,dy).(2.9)

Here τj(x′, y′), j = 1, 2, are local unit functions and νj(x′, dy′), j = 1, 2, are Lévy kernels satisfying some conditions indicated below, and u0 is the density function of the initial state X0. If the initial state X0 = 0, then u0(x) = δ0(x). The boundary condition (2.3) is called a second order Wentcel’s boundary condition, to credit Wentcel’s contribution [36] to the theory of diffusion processes.

Obviously, initial-boundary value problem (2.2)-(2.4) recovers (1.12)-(1.14) if ν(x, ⋅) = 0, Q(x′, D) φ(x′)= γ(x′) φ(x′), δ(x′) = 0, and νj(x′, ⋅) = 0, j = 1, 2. Moreover, one can work with coefficients aij(x), i.j = 1.…, n, bj(x), j = 1, …, d, c0(x), and the symbol σA(x, ξ) of the operator A(x, D) in (2.5) satisfying mild conditions similar to Lipschitz (1.2) and linear growth condition (1.3). However, in this paper, for simplicity, we assume the following conditions:

  1. aij(x) ∈ C(Ω) ∩ C(Ω) and aij(x) = aij(x) for all i, j = 1, …, d, and xΩ, and there exists a constant a0 > 0 such that

    i,j=1daij(x)ξiξja0|ξ|2,xΩ¯,ξRd;
  2. bj(x) ∈ C(Ω) ∩ C(Ω), j = 1, …, d;

  3. c0(x) ∈ C(Ω) ∩ C(Ω), and c0(x) ≤ 0 in Ω;

  4. the local unity function ϕ(x, y) and the kernel function ν(x, dy) are such that the symbol σA(x, ξ) of the operator 𝓐(x, D) belongs to the class of symbols S(Ω × ℝd) = C(Ω × ℝd) ∩ C(Ω × (ℝd \ G)), where G ∈ ℝd is a set of d-dimensional measure zero; see [33] for the theory of pseudo-differential operators with such a nonregular symbols.

Equation (2.2) describes a diffusion process accompanied by jumps in Ω with the drift vector (b1(x), …, bn(x)) and diffusion coefficient defined by the matrix-function ai,j(x), and jumps governed by the Lévy measure ν(x, ⋅). We assume additionally that the condition

  1. A(x,D)[1(x)]=c0(x)+Ω[1ϕ(x,y)]ν(x,dy)0,xΩ,

is fulfilled to ensure that the jump phenomenon from x ∈ Ω to the the outside of a neighborhood of x is “dominated” by the absorption phenomenon at x (see [30] for details).

The coefficients αij(x′), i, j = 1.…, d, βj(x′), j = 1, …, d, and γ(x′) of the operator Q(x, D) in (2.5) satisfy the following conditions:

  1. αij(x′) ∈ C(Ω) and αij(x′) = αij(x′) for all i, j = 1, …, d−1, and x′ ∈ Ω, and there exists a constant α0 > 0 such that

    i,j=1d1αij(x)ξiξjα0|ξ|2,xΩ,ξRd1;
  2. βj(x′) ∈ C(Ω), j = 1, …, d;

  3. γ(x′) ∈ C(Ω), and γ(x′) ≤ 0 in Ω.

    The symbols σΓ1(x′,ξ) and σΓ2(x′,ξ) of the boundary pseudo-differential operators Γ1(x′, D) and Γ2(x′, D) in equations (2.8) and (2.9) satisfy the following condition:

  4. the local unity functions ϕk(x, y), k = 1, 2, and the kernel function νk(x, dy), k = 1, 2, are such that the symbols σΓk(x′,ξ) of operators Γk(x′, D), k = 1, 2, belong to the class of symbols S(Ω, ℝd) = C(Ω × (ℝdG0)), where G0 ∈ ℝd is a set of d-dimensional measure zero.

The boundary condition (2.3) with the operator 𝓦(x, D) defined in equations (2.6)-(2.9) describes a combination of continuous diffusion and jumping processes undergoing on the boundary, as well as jumps from the boundary into the region, and the viscosity phenomenon near the boundary. Namely, the term

j,k=1d1αj,k(x)2u(t,x)xjxk+j=1d1β(x)u(t,x)xj+b(x)u(t,x)xj

governs the diffusion process on the boundary, the term γ(x′) u(t, x′) is responsible for the absorption phenomenon at the point x′ ∈ Ω, the term μ(x)nx expresses the reflexion phenomenon at x′ ∈ Ω, the term δ(x′) A(x′, D) u(t, x′) expresses the viscosity near x′ ∈ Ω, and the terms Γ1 (x′, D) u(t, x′) and Γ2 (x′, D) u(t, x′) govern jump processes on the boundary and jump processes from the boundary into the region, respectively. We assume that the condition

  1. W(x,D)[1(x)]=γ(x)+Ω[1τ1(x,y)]ν1(x,dy)+Ω[1τ2(x,y)]ν2(x,dy)0,xΩ,

is fulfilled to ensure that the jump phenomenon from x′ ∈ Ω to the outside of a neighborhood of x′ on the boundary Ω or inward to the region Ω is “dominated” by the absorption phenomenon at x′.

We also assume the following transversality condition of the boundary operator 𝓦:

  1. Ων2(x,dy)=,if μ(x)=δ(x)=0.

Fractional order FPK equations use fractional derivatives, and in more general case, fractional distributed order operators. Below we briefly recall the definitions and some properties of these operators referring the reader to the sources [19, 22, 33, 34] for details. By definition, the Caputo-Djrbashian derivative of order 0 < β < 1 is defined by

τDβg(t)=1Γ(1β)τtg(s)ds(ts)β,τ<t,(2.10)

where τ ≥ 0 is an initial point and Γ(z), z ∈ ℂ \ {…, −2, −1, 0}, is Euler’s gamma function. We assume that the function g(t), t ≥ 0, is a differentiable function. We write Dβ when τ = 0. Making use of the fractional integration operator of order α > 0

τJαf(t)=1Γ(α)τt(ts)α1f(s)ds,

we can write τDβ also in the operator form

τDβ=τJ1βddt.

In our further analysis, without loss of generality, we assume that τ = 0.

Let μ be a Borel measure defined on the Borel sets of the interval (0, 1). The distributed order differential operator with the mixing measure μ, by definition, is

Dμ,tg(t)=01Dβg(t)dμ(β).

Two functions 𝓚μ(t) and Φμ(s) defined with the help of the measure μ respectively by

Kμ(t)=01tβΓ(1β)dμ(β),t>0,(2.11)
Φμ(s)=01sβdμ(β),s>0.(2.12)

play an important role in our further analysis. We denote by g~(s) the Laplace transform of a function g, that is

g~(s)=0g(t)estdt.

One can verify using the Laplace transform formula for fractional derivatives,

[Dβg]~(s)=sβg~(s)sβ1g(0+)(2.13)

that the formula

[Dμ,tg]~(s)=Φμ(s)g~(s)g(0+)Φμ(s)s,s>0,(2.14)

holds. Moreover, using the formula [tβ]~(s)=Γ(1β)sβ1 (see [33]), one can also verify that the relation

Kμ~(s)=Φμ(s)s,s>0,(2.15)

holds.

The widest class of processes for which the Itô calculus can be extended is the class of semimartingales [27]. A semimartingale is a càdlàg process (the processes with paths which are right continuous and have left limits) which has the representation (see details in [27]) Zt = Z0 + Vt + Mt, where Z0 is a d-dimensional random vector, Vt is an adapted finite variation process, and Mt is a local martingale. An example of a semimartingale is a Lévy process defined as follows. A stochastic process Lt ∈ ℝd, t ≥ 0, is called a Lévy process if the following three conditions are verified: (1) L0 = 0; (2) Lt has independent stationary increments; (3) for all ϵ, t > 0, limst ℙ(|LtLs | > ϵ) = 0. Lévy processes have càdlàg modifications. Lévy processes are fully described by three parameters: a vector b ∈ ℝd, a nonnegative definite matrix Σ, and a measure ν defined on ℝd \ {0} such that ∫ min (1, |x|2) < ∞, called its Lévy measure. The Lévy-Khintchine formula provides a characterization of Lévy process through its characteristic function Φt(ξ) = etΨ(ξ), with the Lévy symbol (see, e.g. [28])

Ψ(ξ)=i(b,ξ)(Σξ,ξ)+Rd{0}(ei(x,ξ)1i(x,ξ)1(|x|1))ν(dx).(2.16)

A Lévy subordinator is a nonnegative and nondecreasing Lévy process. A Lévy subordinator with the Lévy symbol Ψ(s) = sβ, s > 0, is called a β-stable Lévy subordinator. The composition of two Lévy processes is again a Lévy process. In particular, a time-changed process LDt with a stable Lévy subordinator Dt is again a Lévy process. Transition probabilities of such a time-changed process satisfy FPK type equation with a pseudo-differential operator on the right hand side, whose symbol is continuous but not smooth. However, the composition LWt, where Wt = inf{τ ≥ 0: Dτ > t}, the inverse to the stable subordinator Dt is not a Lévy process, but still is a semimartingale. This drastically changes the associated FPK equation: now it is a time-fractional pseudo-differential equation [14, 15], implying non-Markovian behaviour of the process.

The properties of the density function of the inverse process Wt are listed below in the more general case when Wt represents the inverse of a process which belongs to the class 𝓢μ of mixtures of an arbitrary number of independent stable subordinators with a mixing measure μ. Namely, a non-negative increasing Lévy stochastic process Dt belongs to the class 𝓢μ if

lnEesDt=tΦμ(s).

It is known (see e.g. [28]) that mixtures of independent stable processes of different indices are no longer stable. For a stochastic process in 𝓢μ we use the notation Dμ,t showing the dependence on the measure μ, and for their inverses the notation Wμ,t. The density function of Wμ,t is denoted by ftμ(τ) . Note that such mixtures model complex diffusions and other types of stochastic processes with several simultaneous diffusion modes. Their associated FPK type equations are distributed order differential equations (see e.g. [19, 22, 34]).

Lemma 2.1

The density functionftμ(τ)possesses the following properties:

  1. limt→+0ftμ(τ) = δ0(τ), τ ≥ 0;

  2. limτ→+0ftμ(τ) = 𝓚μ(t), t > 0;

  3. limτ→∞ftμ(τ) = 0, t ≥ 0;

  4. Lts[ftμ(τ)](s)=Φμ(s)seτΦμ(s),s>0,τ0.

Lemma 2.2

The functionftμ(τ)satisfies for each t > 0 the equation

Dμ,tftμ(τ)=τftμ(τ)δ0(τ)Kμ(t),(2.17)

in the sense of distributions.

For the proofs of these lemmas we refer the reader to [33].

3 Main results: fractional FPK equations associated with SDEs on a bounded domain

In this section we derive fractional order FPK equations associated with SDEs on a bounded domain as initial-boundary value problems. We also prove existence of a unique solution to obtained initial-boundary value problems.

Theorem 3.1

Let Xt be a stochastic process associated with the FPK equation(2.2)-(2.4)and Wμ,tbe the inverse to a Lévy’s subordinator Dμ,t ∈ 𝓢μwith a mixing measure μ. Then the FPK equation associated with the time changed stochastic process XWμ,thas the form

Dμ,tv(t,x)=A(x,D)v(t,x),t>0,xΩ,(3.1)
W(x,D)v(t,x)=0,t>0,xΩ,(3.2)
v(0,x)=u0(x),xΩ.(3.3)

where u0CW2(Ω) ≡ {φC2(Ω) : 𝓦(x′, D) φ(x′) = 0}.

Proof

Let Tt be the semigroup with the infinitesimal generator 𝔸 = 𝓐(x, D) : C(Ω) → C(Ω), with the domain

D(A)={ϕC2(Ω):W(x,D)ϕ(x)=0,xΩ}.

Then the unique solution of problem (2.2)-(2.4) has the form

u(t,x)=Ttu0(x)=E[u0(Xt)|X0=x],

indicating connection of the solution u(t, x) of the FPK equation in (2.2)-(2.4) with the stochastic process Xt. Now, consider the function v(t, x) obtained from the latter replacing Xt by XWμ,t, that is

v(t,x)=E[u0(XWμ,t)|X0=x]=0E[u0(XWμ,t)|Wμ,t=τ|X0=0]dτ=0u(τ,x)fWμ,t(τ)dτ=0fWμ,t(τ)Tτu0(x)dτ.(3.4)

We will show that v(t, x) defined above satisfies the initial-boundary value problem for fractional FPK equation in (3.1)-(3.3). First, we show that v(t, x) satisfies equation (3.1). Indeed, one can readily see that

v(t,x)=0τ{JfDμ,τ(t)}(Tτu0(x))dτ.

Now it follows from the definition of Dμ,t that the Laplace transform of v(t, x) satisfies

v~(s,x)=01sβdμ(β)s0eτ01sβdμ(β)(Tτu0(x))dτ=Φμ(s)su~(Φμ(s),x),s>ω¯,(3.5)

where ũ(s, x) is the Laplace transform of u(t, x), the function Φμ(s) is defined in (2.12), and ω > 0 is a number such that s > ω if Φμ(s) > ω (ω is uniquely defined, since as is seen from Definition (2.12), the function Φμ(s) is a strictly increasing function). The function ũ(s, x) satisfies the equation

su~(s,x)A(x,D)u~(s,x)=u0(x),xΩ,(3.6)

Indeed, applying the Laplace transform to both sides of equation (2.2) and taking into account the initial condition (2.4), we obtain equation (3.6). It follows from equations (3.5) and (3.6) that the composite function

u~(Φμ(s),x)=sv~(s,x)/Φμ(s)

satisfies the equation

(Φμ(s)A(x,D))sv~(s,x)Φμ(s)=u0(x),s>ω¯,xΩ,

or, the same, the equation

(Φμ(s)A(x,D))v~(s,x)=u0(x)Φμ(s)s,s>ω¯,xΩ.

We rewrite the latter in the form

Φμ(s)v~(s,x)Φμ(s)sv(0+,x)=A(x,D)v~(s,x),s>ω¯,xΩ.(3.7)

We notice that the left hand side of the latter equation is the Laplace transform of the expression 𝓓μ,tv(t, x) due to formula (2.13). Therefore, equation (3.7) is equivalent to equation (3.1).

Further, using (2.3), we have

W(x,D)v(t,x)=W(x,D)0u(τ,x)fWμ,t(τ)dτ=0W(x,D)u(τ,x)fWμ,t(τ)dτ=0

since W(x′, D) u(τ, x′) = 0 for all τ > 0 due to boundary condition (2.3).

Finally, making use of the dominated convergence theorem,

v(0,x)=limt0+E[u0(XWμ,t)|XWμ,0=x]=E[u0(XWμ,t)|XWμ,0=x]=E[u0(X0)|X0=x]=E[u0(x)]=u0(x).

Hence, v(t, x) defined in (3.4) satisfies the initial-boundary value problem in (3.1)-(3.3) for the fractional order FPK equation.□

In the particular case of Wt being the inverse of a single Lévy’s subordinator Dt with the stability index β ∈ (0, 1) this theorem implies the following result:

Corollary 3.2

Let Xt be a stochastic process associated with the FPK equation(2.2)-(2.4)and Wt be the inverse to the Lévy’s stable subordinator with a stability index 0 < β < 1. Then the FPK equation associated with the time changed stochastic process XWthas the form

Dβv(t,x)=A(x,D)v(t,x),t>0,xΩ,(3.8)
W(x,D)v(t,x)=0,t>0,xΩ,(3.9)
v(0,x)=u0(x),xΩ.(3.10)

An important question is the existence of a unique solution of the initial-boundary value problem in equations (3.13)-(3.15).

Theorem 3.3

Let the conditions (i) – (iv), (a) – (d), and (C1) – (C3) are verified. Then initial-boundary value problem(3.13)-(3.15)for fractional order FPK equation has a unique solution v(t, x) in the space C((0, ∞) × Ω) ∩ C1(t > 0; CW2(Ω).

Remark 3.1

Here C1(t > 0; CW2(Ω) is the space of vector-functions differentiable in t and belonging to CW2(Ω) for each fixed t > 0.

Proof

Introduce the operator 𝔘𝓦 as follows: 𝔘𝓦 = 𝓐(x, D) with the domain Dom(𝔘𝓦) = {ϕC(Ω) : 𝓐(x, D) ϕC(Ω), 𝓦(x′, D) ϕ (x′) = 0, ∀ x′ ∈ Ω}. As is proved in [31] if the conditions of the theorem are verified, then there exists a Feller semigroup {Tt}t≥0 (nonnegative and contractive) on Ω generated by 𝔘𝓦. That is for any φC(Ω) such that 0 ≤ φ(x) ≤ 1 on Ω one has 0 ≤ Ttφ(x) ≤ 1 on Ω. Moreover, it follows from the general semigroup theory (see, e.g. [30]) that the equation Ttφ(x) = et𝔘𝓦φ(x) holds. Hence, for an arbitrary u0C(Ω) the function u(t, x) = Ttu0 (x), t ≥ 0, x ∈ Ω, exists and solves the following initial value problem for a differential-operator equation

u(t,x)t=UWu(t,x),t0,xΩ,(3.11)
limt0u(t,x)=u0(x),xΩ.(3.12)

Since the operator 𝓐(x, D) is elliptic with the spectrum in the negative real axis, it follows from the smoothness of a solution to parabolic equations that u(t, x) has all the derivatives if t > 0 and x ∈ Ω. Thus, in particular, this function belongs to the space C((0, ∞) × Ω) ∩ C1(t > 0; CW2(Ω). The existence and uniqueness of a solution of initial-boundary value problem (3.1)-(3.3) immediately follows from representation (3.4). Also, it follows from this representation that v(t, x) has all derivatives if t > 0 and that the estimate

|v(t,x)|0ftμ(τ)|Tτu0(x)|dτsupt0Ttu0(x)C(Ω¯)=supt0u(t,x)C(Ω¯),t0,xΩ¯,

holds. Thus, the function v(t, x) inherits all the properties of u(t, x), including the property C((0, ∞) × Ω) ∩ C1(t > 0; CW2(Ω).□

Now consider the following initial-boundary value problem (2.2)-(2.4) for t-dependent generalization of the FPK equation

u(t,x)t=B(x,D)u(t,x)+(γ+1)tγ2A(x,D)u(t,x),t>0,xΩ,(3.13)
W(x,D)u(t,x)=0,t>0,xΩ,(3.14)
u(0,x)=φ(x),xΩ,(3.15)

where 𝓑(x, D) is a pseudo-differential operator whose order is strictly less than the order of the operator 𝓐(x, D). We assume that 𝓐(x, D) is an elliptic Waldenfels operator defined in (2.5) and 𝓦(x′, D) is a Wentcel’s boundary pseudo-differential operator defined in (2.6). The parameter γ runs in the interval (−1, 1). Equation (3.13) is a parabolic equation: if 0 < γ <1, then it is degenerate; if −1 < γ < 0, then it is singular. Obviously, the initial-boundary value problem in (3.13)-(3.15) recovers problem (2.2)-(2.4), if 𝓑(x, D) = 0 and γ = 0.

Here is the motivating example: if 𝓑(x, D) = 0, γ = 2H − 1 and 𝓐(x, D) = Δ, Laplace operator, then equation (3.13) reduces to

u(t,x)t=Ht2H1Δu(t,x),t>0,xΩ,

which represents the FPK equation associated with the fractional Brownian motion with the Hurst parameter H ∈ (0, 1) (see, e.g. [5, 16]). Therefore, initial-boundary value problem (3.13)-(3.15) can be considered as a FPK equation of a stochastic process driven by fractional Brownian motion. By definition, the fractional Brownian motion with the Hurst parameter H is a stochastic process BtH, Gaussian with mean zero for each fixed t ≥ 0, with continuous paths and the following covariance function:

RH(s,t)=E(BtHBsH)=12(t2H+s2H|ts|2H).

Fractional Brownian motion is not Markovian and is not a semimartingale. Therefore, a stochastic integral and stochastic differential equation driven by a fractional Brownian motion can not be defined in the Itô sense. Nevertheless, one can define stochastic differential equations driven by fractional Brownian motion in a meaningful sense [4, 11, 25]. Thus, the class of initial-boundary value problems of the form (3.13)-(3.15) describes stochastic processes in a bounded region Ω driven by not only Lévy processes, but also fractional Brownian motion. Below we derive the fractional FPK equation associated with such a stochastic process with a time-changed driving process.

Theorem 3.4

Let Xt be a stochastic process associated with the FPK equation(3.13)-(3.15)and Wμ,tbe the inverse to a Lévy’s subordinator Dμ,t ∈ 𝓢μwith a mixing measure μ. Then the FPK equation associated with the time changed stochastic process XWμ,thas the form

Dμ,tv(t,x)=B(x,D)v(t,x)+γ+12Gμ,γ,tA(x,D)v(t,x),t>0,xΩ,(3.16)
W(x,D)v(t,x)=0,t>0,xΩ,(3.17)
v(0,x)=u0(x),xΩ.(3.18)

where the operator Gμ,γ,tis defined as

Gμ,γ,tv(t,x)=Φμ(t)Lst1[Γ(γ+1)2πiCiC+imμ(z)v~(z,x)(ρ(s)ρ(z))γ+1dz](t),(3.19)

where * denotes the convolution operation, the symbolLst1means the inverse Laplace transform, 0 < C < s, the functionsρ(z) and mμ(z) are defined by

ρ(z)=01eβLn(z)dμ(β),mμ(z)=01βzβdμ(β)ρ(z),

the function Φμ(t) is defined in(2.12), and u0CW2(Ω) ≡ {ϕC2(Ω) : 𝓦(x′, D) ϕ (x′) = 0}.

Proof

Let u(t, x) and v(t, x) be density functions of Xt and XWμ,t, respectively. Then due to the formula for the conditional density, we have

v(t,x)=fXWμ,t(x)=0f(XWμ,t|Wμ,t=τ)P(Wμ,tdτ)=0u(τ,x)P(Wμ,tdτ)=0ftμ(τ)u(τ,x)dτ,t0,xΩ.(3.20)

By definition of Xt the function u(t, x) solves problem (3.13)-(3.15). We will show that v(t, x) satisfies problem (3.16)-(3.18). To show that v(t, x) satisfies equation (3.16), we compute

Dμ,tv(t,x)=0Dμ,tftμ(τ)u(τ,x)dτ.

Here the change of the order of Dμ,t and the integral is valid thanks to the estimate obtained in [15] for the density function ftμ(τ) of a mixture of stable subordinators having mixing measure μ with suppμ ⊂ (0, 1). It follows from Lemma 2.2 that

Dμ,tv(t,x)=0ftμ(τ)τu(τ)dτKμ(t)0δ0(τ)u(τ)dτ.(3.21)

Integration by parts in the first integral, we have

0ftμ(τ)τu(τ)dτ=0ftμ(τ)u(τ)τdτlimτftμ(τ)u(τ)+limτ0ftμ(τ)u(τ).(3.22)

The first limit on the right hand side is zero due to part (c) of Lemma 2.1. Due to part (b) of Lemma 2.1 the second limit on the right hand side of (3.22) has the same value as the second integral on the right side of (3.21), but with the opposite sign. Hence, it follows that

Dμ,tv(t,x)=0ftμ(τ)τu(τ,x)dτ.

Now using equation (3.13), we have

Dμ,tv(t,x)=0ftμ(τ)B(x,D)u(τ,x)+(γ+1)τγ2A(x,D)u(τ,x)dτ=B(x,D)0ftμ(τ)u(τ,x)dτ+γ+12A(x,D)0ftμ(τ)τγu(τ,x)dτ=B(x,D)v(t,x)+γ+12A(x,D)0ftμ(τ)τγu(τ,x)dτ=B(x,D)v(t,x)+γ+12A(x,D)Gμ,γ,tv(t,x),

where

Gμ,γ,tv(t,x)=0ftμ(τ)τγu(τ,x)dτ.(3.23)

The fact that the operator Gμ,γ,t has the representation (3.19) is proved in [16].

Further, we show that 𝓦(x′, D)v(t, x′) = 0 if x′ ∈ Ω. Indeed, using 𝓦(x′, D) u(τ, x′) = 0 for all τ ≥ 0, we have

W(x,D)v(t,x)=0W(x,D)u(τ,x)fWμ,t(τ)dτ=0,xΩ,

for any fixed t ≥ 0.

Finally, making use of part (a) of Lemma 2.1 and the dominated convergence theorem,

limt0+v(t,x)=limt0+0ftμ(τ)u(τ,x)dτ=0limt0+ftμ(τ)u(τ,x)dτ=0δ0(τ)u(τ,x)dτ=u(0,x)=u0(x).

Hence, v(t, x) defined in (3.20) satisfies the initial-boundary value problem in (3.16)-(3.18) for time dependent fractional order FPK equation.□

Remark 3.2

The properties of the operator Gγ,t are studied in paper [16], including the fact that the family {Gγ,t, −1 < γ < 1} possesses the semigroup property. Namely, for any γ, δ ∈ (−1, 1), γ+δ ∈ (−1, 1), one has Gγ,tμGδ,tμ=Gγ+δ,tμ=Gδ,tμGγ,tμ, where “∘” denotes the composition of two operators.

Similarly to Corollary 3.2, in the particular case of Wt being the inverse of a single Lévy’s subordinator Dt with the stability index β ∈ (0, 1) Theorem 3.4 implies the following result:

Corollary 3.5

Let Xt be a stochastic process associated with the FPK equation(3.13)-(3.15)and Wt be the inverse to the Lévy’s stable subordinator with a stability index 0 < β < 1. Then the FPK equation associated with the time changed stochastic process XWthas the form

Dβv(t,x)=B(x,D)v(t,x)+γ+12Gμ,γ,tA(x,D)v(t,x),t>0,xΩ,W(x,D)v(t,x)=0,t>0,xΩ,v(0,x)=u0(x),xΩ.

Theorem 3.6

Let the operators 𝓐(x, D), 𝓑(x, D), and 𝓦(x′, D) in problem(3.16)-(3.18)satisfy the following conditions:

  1. The pseudo-differential operator 𝓐(x, D) is an elliptic Waldenfels operator defined in(2.5)with the conditions (i) – (iv) and (C1) satisfied;

  2. The operator 𝓑(x, D) is a pseudo-differential operator whose order is strictly less then the order of 𝓐(x, D);

  3. The pseudo-differential operator 𝓦(x′, D) is a Wentcel’s boundary operator defined in(2.6)with the conditions (a) – (d), (C2), and (C3) satisfied.

Then for an arbitrary u0CW2 (Ω) initial-boundary value problem(3.16)-(3.18)has a unique solution v(t, x) in the spaceC([0, ∞) × Ω) ∩ C1(t > 0; CW2(Ω)).

Proof

The arguments used in the proof of Theorem 3.3 do not work in the case of problem (3.16)-(3.18), since the solution u(t, x) of problem (3.13)-(3.15) does not have a representation through the Feller semigroup. To prove the theorem we will use the properties of pseudo-differential operators.

Consider the operator

AW(t)=tγ2A(x,D)+B(x,D)

with the domain Dom(𝓐𝓦(t)) = {ϕC2(Ω) : 𝓦(x′, D) ϕ(x′)=0, x′ ∈ Ω}. This operator is a pseudo-differential operator with the symbol

σ(t,x,ξ)=tγ2σA(x,ξ)+σB(x,ξ),t0,xΩ,ξRd,(3.24)

where σ𝓐(x, ξ) and σ𝓑(x, ξ) are symbols of the operators 𝓐(x, D) and 𝓑(x, D), respectively. Due to conditions (A) and (B) of the theorem, for each fixed t > 0 the symbol σ(t, x, ξ) satisfies the following ellipticity estimate

σ(t,x,ξ)κt|ξ|δ,|ξ|C,(3.25)

with some constants κt > 0, C > 0, and δ>0.

One can see that the solution of problem (3.13)-(3.15) has a formal representation

u(t,x)=S(t,x,D)u0(x),t0,xΩ,(3.26)

where the solution pseudo-differential operator 𝓢(t, x, D) has the symbol

s(t,x,ξ)=etσ(t,x,ξ),t0,ξRd.(3.27)

The fact that u(t, x) satisfies equation (3.13) can be verified by direct calculation. To show this fact let us extend u0(x) for all x ∈ ℝd \ Ω by zero, and denote the extended function again by u0(x). Let ũ0(ξ) is the Fourier transform of u0(x). Then, we have

u(t,x)t=S(t,x,D)u0(x)t=1(2π)dtRdetσ(t,x,ξ)eixξu~0(ξ)dξ=1(2π)dRdσ(t,x,ξ)+tσ(t,x,ξ)tetσ(t,x,ξ)eixξu~0(ξ)dξ=1(2π)dRd(γ+1)tγ2σA(x,ξ)+σB(x,ξ)eixξetσ(t,x,ξ)u~0(ξ)dξ=(γ+1)tγ2A(x,D)+B(x,D)w(t,x),t>0,xΩ,

where w(t, x) has the Fourier transform

w~(t,ξ)=etσ(t,x,ξ)u~0(ξ),t>0,ξRd.

Changing the differentiation and integration operators in the above calculation is legitimate due to estimate (3.25). Now calculating the inverse Fourier transform of w~(t, ξ), and using the definition (2.1) of pseudo-differential operators, we have

w(t,x)=1(2π)dRdetσ(t,x,ξ)eixξu~0(ξ)dξ=1(2π)dRds(t,x,ξ)eixξu~0(ξ)dξ=S(t,x,D)u0(x),t>0,xΩ.

Thus, w(t, x) = u(t, x), and hence u(t, x) defined by (3.26) satisfies equation (3.13). Further, since 𝓦(x′, D)u0(x′) = 0 for x′ ∈ Ω, it follows from (3.26) that

W(x,D)u(t,x)=0,t>0xΩ.

Moreover, it follows from (3.27) that 𝓢(0, x, D) = I, the identity operator, implying u(0, x) = u0(x). Finally, estimate (3.25) and representation (3.26) of the solution imply the inclusion of the solution to the space C([0, ∞) × Ω) ∩ C1(t > 0; CW2(Ω)).

The existence and uniqueness of a solution to initial-boundary value problem (3.16)-(3.18) immediately follows from representation (3.20), namely

v(t,x)=0ft(τ)u(τ,x)dτ,

where u(t, x) is the solution of problem (3.13)-(3.15). The arguments here are similar to the proof of Theorem 3.3. Also, similar to Theorem 3.3, the function v(t, x) inherits all the properties of u(t, x), including the property C((0, ∞) × Ω) ∩ C1(t > 0; CW2(Ω).□

Remark 3.3

The technique used to prove Theorem 3.6 is applicable for FPK equations in the whole space ℝd obtained in papers [15, 16].


Dedicated to Professor Virginia Kiryakova on the occasion of her 65th birthday and the 20th anniversary of FCAA


References

[1] O.P. Agrawal, Solution for a fractional diffusion wave equation in a bounded domain. Nonlinear Dynamics29 (2002), 145–155.10.1023/A:1016539022492Search in Google Scholar

[2] F. Baudoin, L. Coutin, Operators associated with a stochastic differential equations driven by fractional Brownian motion. Stoch. Process. Appl. 17, No 5 (2007), 550–574.10.1016/j.spa.2006.09.004Search in Google Scholar

[3] O.E. Barndorff-Nielsen, T. Mikosch, S. Resnick (Eds), Lévy Processes: Theory and Applications. Birkhäuser (2001).10.1007/978-1-4612-0197-7Search in Google Scholar

[4] C. Bender, An Itô formula for generalized functionals of a fractional Brownian motion with arbitrary Hurst parameter. Stoch. Process. Appl. 104, No 1 (2003), 81–106.10.1016/S0304-4149(02)00212-0Search in Google Scholar

[5] F. Biagini, Y. Hu, B. Oksendal, T. Zhang, Stochastic Calculus for Fractional Brownian Motion and Application. Springer (2008).10.1007/978-1-84628-797-8Search in Google Scholar

[6] N.N. Bogolyubov, D.P. Sankovich, N.N. Bogolyubov and statistical mechanics. Russian Math. Surveys49, No 5 (1994), 19–49.10.1070/RM1994v049n05ABEH002419Search in Google Scholar

[7] P.C. Bressloff, Stochastic Processes in Cell Biology. Springer (2014).10.1007/978-3-319-08488-6Search in Google Scholar

[8] N. Burch, R. Lehouch, Classical, nonlocal, and fractional diffusion equations on bounded domains. Internat. J. for Multiscale Computational Engineering9, No 6 (2011), 661–674.10.1615/IntJMultCompEng.2011002402Search in Google Scholar

[9] A.V. Chechkin, J. Klafter, I.M. Sokolov, Fractional Fokker-Planck equation for ultraslow kinetics. Europhysics Letters63, No 3 (2003), 326–332.10.1209/epl/i2003-00539-0Search in Google Scholar

[10] Zh.-Q. Chen, M. Meerschaert, E. Nane, Space-time fractional diffusion on bounded domain. J. Math. Anal. Appl. 393 (2012), 479–488.10.1016/j.jmaa.2012.04.032Search in Google Scholar

[11] L. Decreusefond, A.S. Üstünel, Stochastic analysis of the fractional Brownian motion. Potential Analysis10, No 2 (1998), 177–214.10.1023/A:1008634027843Search in Google Scholar

[12] J. Gajda, Fractional Fokker-Planck equation with space dependent drift and diffusion: the case of tempered α-stable waiting times. Acta Physica Polonica B44 (2013), 1149–1161.10.5506/APhysPolB.44.1149Search in Google Scholar

[13] C.W. Gardiner, Handbook of Stochastic Methods: for Physics, Chemistry and the Natural Sciences. Springer (1985).10.1007/978-3-662-02452-2Search in Google Scholar

[14] M. Hahn, K. Kobayashi, S. Umarov, SDEs driven by a time-changed Lévy process and their time-fractional order pseudo-differential equations. J. Theoret. Probab. 25, No 1 (2012), 262–279.10.1007/s10959-010-0289-4Search in Google Scholar

[15] M. Hahn, S. Umarov, Fractional Fokker-Planck-Kolmogorov type equations and their associated stochastic differential equations. Fract. Calc. Appl. Anal. 14, No 1 (2011), 56–79; 10.2478/s13540-011-0005-9; https://www.degruyter.com/view/j/fca.2011.14.issue-1/issue-files/fca.2011.14.issue-1.xml.Search in Google Scholar

[16] M. Hahn, K. Kobayashi, S. Umarov, Fokker-Planck-Kolmogorov equations associated with time-changed fractional Brownian motion. Proc. Amer. Math. Soc. 139, No 2 (2011), 691–705.10.1090/S0002-9939-2010-10527-0Search in Google Scholar

[17] L. Hörmander, Pseudo-differential operators. Comm. Pure Appl. Math. 18 (1965), 501–517.10.1007/978-3-540-49938-1_3Search in Google Scholar

[18] N. Jacob, Pseudo-differential Operators and Markov Processes. Imperial College Press, London (2001).10.1142/p245Search in Google Scholar

[19] A. Kochubei, Distributed order calculus and equations of ultraslow diffusion. J. Math. Anal. Appl. 340, No 1 (2008), 252–281.10.1016/j.jmaa.2007.08.024Search in Google Scholar

[20] J.J. Kohn, L. Nirenberg, An algebra of pseudo-differential operators. Comm. Pure Appl. Math. 18 (1965), 269–305.10.1002/cpa.3160180121Search in Google Scholar

[21] M. Meerschaert, E. Nane, P. Vellaysamy, Fractional Cauchy problems on bounded domain. Annals of Probability37, No 3 (2009), 979–1007.10.1214/08-AOP426Search in Google Scholar

[22] M. Meerschaert, H.-P. Scheffler, Stochastic model for ultraslow diffusion. Stoch. Proces. Appl. 116, No 9 (2006), 1215–1235.10.1016/j.spa.2006.01.006Search in Google Scholar

[23] A.V. Melnikov, Financial Markets: Stochastic Analysis and the Pricing of Derivative Securities. Translations of Mathematical Monographs184, Amer. Math. Sci. (1999).Search in Google Scholar

[24] R. Metzler, J. Klafter, The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Phys. Rep. 339, No 1 (2000), 1–77.10.1016/S0370-1573(00)00070-3Search in Google Scholar

[25] D. Nualart, The Malliavin Calculus and Related Topics. 2nd Ed., Springer (2006).Search in Google Scholar

[26] S. Pressé, A data-driven alternative to the fractional Fokker-Planck equation. J. of Statistical Mechanics: Theory and Experiment2015 (2015), P07009.10.1088/1742-5468/2015/07/P07009Search in Google Scholar

[27] P. Protter, Stochastic Integration and Differential Equations. Springer-Verlag (1991).10.1007/978-3-662-02619-9Search in Google Scholar

[28] K-I. Sato, Lévy Processes and Infinitely Divisible Distributions. Cambridge University Press (1999).Search in Google Scholar

[29] B. Shizgal, Spectral Methods in Chemistry and Physics: Applications to Kinetic Theory and Quantum Mechanics. Springer (2015)10.1007/978-94-017-9454-1Search in Google Scholar

[30] K. Taira, Semigroups, Boundary Value Problems and Markov Processes. Springer (2014).10.1007/978-3-662-43696-7Search in Google Scholar

[31] K. Taira, On the Existence of Feller Semigroups with Boundary Conditions. Ser. Memoirs of the Amer. Math. Sci. 475 (1992).10.1090/memo/0475Search in Google Scholar

[32] Z. Tomovski, T. Sandev, Exact solutions for fractional diffusion equation in a bounded domain with different boundary conditions. Nonlinear Dynamics71, No 4 (2013) 671–683.10.1007/s11071-012-0710-xSearch in Google Scholar

[33] S. Umarov, Introduction to Fractional and Pseudo-Differential Equations with Singular Symbols. Springer (2015).10.1007/978-3-319-20771-1Search in Google Scholar

[34] S. Umarov, R. Gorenflo, The Cauchy and multi-point problem for distributed order fractional differential equations. Zeitschrift für Analysis und ihre Anwendungen24 (2005), 449–466.10.4171/ZAA/1250Search in Google Scholar

[35] N.C. Van Kampen, Stochastic Processes in Physics and Chemistry. Elsevier (1997).Search in Google Scholar

[36] A.D. Wentcel, On boundary conditions for multidimensional diffusion processes. Theory of Probability and its Applications4 (1959), 164–177.10.1137/1104014Search in Google Scholar

[37] K. Weron, M. Magdziarz, Modeling of subdiffusion in space-time-dependent force fields beyond the fractional Fokker-Planck equation. Physical Review E77 (2008), Id # 036704.10.1103/PhysRevE.77.036704Search in Google Scholar PubMed

Received: 2017-05-24
Published Online: 2017-10-31
Published in Print: 2017-10-26

© 2017 Diogenes Co., Sofia

Downloaded on 16.11.2025 from https://www.degruyterbrill.com/document/doi/10.1515/fca-2017-0067/html
Scroll to top button