Skip to content
Publicly Available Published by De Gruyter January 29, 2021

On fractional heat equation

  • Anatoly N. Kochubei EMAIL logo , Yuri Kondratiev and José Luís da Silva

Abstract

In this paper, the long-time behavior of the Cesaro mean of the fundamental solution for fractional Heat equation corresponding to random time changes in the Brownian motion is studied. We consider both stable subordinators leading to equations with the Caputo-Djrbashian fractional derivative and more general cases corresponding to differential-convolution operators, in particular, distributed order derivatives.

1 Introduction

Let {Xt, t ≥ 0; Px, xE} be a strong Markov process in the phase space ℝd. Denote by Tt its transition semigroup (in an appropriate Banach space) and by L the generator of this semigroup. Let St, t ≥ 0 be a subordinator (i.e., a non-decreasing real-valued Lévy process) with S0 = 0 and the Laplace exponent Φ:

E[eλSt]=etΦ(λ)t,λ>0.

We assume that St is independent of Xt.

Denote by Et, t > 0 the inverse subordinator, and introduce the time-changed process Yt=XEt. Our goal is to analyze the properties of Yt depending on the given Markov process Xt and the particular choice of subordinator St, see Theorem 3.1 and Theorem 4.1 below. There is a lot of interest on this kind of problem in diverse disciplines. In addition to purely stochastic motivations, the transform of the Markov process Xt in the non-Markov one Yt implies the presence of effects in the corresponding dynamics. This feature is important in a number of physical models. In particular, progress in the understanding of this process may lead to the realization of useful models of biological time in the evolution of species and ecological systems. Currently, there exist rather complete studies of such problems in the case of so-called θ-stable subordinators (0 < θ < 1) [6, 15] and in special examples of initial processes Xt (see, e.g., [17], [12], [13]).

As a basic characteristic of the new process Yt, we may study the time evolution

u(t,x)=Ex[f(Yt)]

for a given initial data f.

As it was pointed out in several works, see e.g. [20], [7] and references therein, u(t, x) is the unique strong solution (in some appropriate sense) to the following Cauchy problem

(1.1)Dt(k)u(t,x)=Lu(t,x)u(0,x)=f(x).

Here we have a generalized fractional derivative (GFD for short)

Dt(k)ϕ(t)=ddt0tk(ts)(ϕ(s)ϕ(0))ds

with a kernel k uniquely defined by Φ.

Let u0(t, x) be the solution to a similar Cauchy problem but with ordinary time derivative

(1.2)tu(t,x)=Lu(t,x)u(0,x)=f(x).

In stochastic terminology, it is the solution to the forward Kolmogorov equation corresponding to the process Xt. Under quite general assumptions there is a convenient and essentially obvious relation between these evolutions that is known as the subordination principle:

u(t,x)=0u0(τ,x)Gt(τ)dτ,

where Gt(τ) is the density of Et.

A similar relation holds for fundamental solutions (or heat kernels in another terminology) v(x, t) and vE(x, t) of equations (1.2) and (1.1), respectively. For certain classes of a priori bounds for fundamental solutions v(x, t), the properties of the subordinated kernels were studied in [8]. The main technical tool used in this work involves a scaling property assumed for Φ [8] that is a global condition on the Lévy characteristic Φ(λ). It is nevertheless difficult to give an interpretation of this scaling assumption in terms of the subordinator.

The aim of the present work is to extend the class of random times for which one may obtain information about the time asymptotic of vE(x, t). We consider the following three classes of admissible kernels kLloc1(+), characterized in terms of the Laplace transforms K(λ) as λ → 0 (i.e., as local conditions):

(C1)K(λ)~λθ1,0<θ<1.
(C2)K(λ)~λ1L(1λ),L(y):=μ(0)log(y)1.
(C3)K(λ)~λ1L(1λ),L(y):=Clog(y)1s,s>0,C>0.

We would like to emphasize that these classes of kernels leads to differential-convolution operators, in particular, the Caputo-Djrbashian fractional derivative (C1) and distributed order derivatives (C2),(C3). For each kernel of this type, we establish the asymptotic properties of the subordinated heat kernels from different classes of a priory bounds. It is important to stress that in working with much more general random times (i.e., without the scaling property), a price must be paid for such an extension, namely the replacement of vE(x, t) by its Cesaro mean. This is the key technical observation that underlies the analysis of several different model situations.

2 Preliminaries

Let S = {S(t), t ≥ 0} be a subordinator, that is a process with stationary and independent non-negative increments starting from 0. They form a special class of Lévy processes taking values in [0, ∞) and their sample paths are non-decreasing. In addition we assume that S has no drift (see [3] for more details). The infinite divisibility of the law of S implies that its Laplace transform can be expressed in the form

E(eλS(t))=etΦ(λ),λ0,

where Φ:[0, ∞) → [0, ∞), called the Laplace exponent (or cumulant), is a Bernstein function. The associated Lévy measure σ has support in [0, ∞) and fulfills

(2.1)(0,)(1τ)dσ(τ)<

such that the Laplace exponent Φ can be expressed as

(2.2)Φ(λ)=(0,)(1eλτ)dσ(τ),

which is known as the Lévy-Khintchine formula for the subordinator S. In addition we assume that the Lévy measure σ satisfy

(2.3)σ(0,)=.

For the given Lévy measure σ, we define the function k by

(2.4)k:(0,)(0,),tk(t):=σ((t,))

and denote its Laplace transform by K; i.e., for any λ ≥ 0 one has

(2.5)K(λ):=0eλtk(t)dt.

We note that by the Fubini theorem, the function K is given in terms of the Laplace exponent. Specifically,

K(λ)=0eλt(t,)dσ(s)dt=(0,)0seλtdtdσ(s)=1λΦ(λ),

i.e.,

(2.6)Φ(λ)=λK(λ),λ0.

Given the inverse process E of the subordinator S, namely

(2.7)E(t):=inf{s0:S(s)t}=sup{s0:S(s)t},

the marginal density of E(t) will be denoted by Gt(τ), t, τ ≥ 0, more explicitly

Gt(τ)dτ=τP(E(t)τ)=τP(S(τ)t)=τP(S(τ)<t).

Example 1

θ-stable subordinator and Gamma processes.

  1. Let S be a θ-stable subordinator θ ∈ (0, 1) with Laplace exponent

    Φθ(λ)=λθ=θΓ(1θ)0(1eλτ)τ1θdτ,

    from which it follows that the Lévy measure σ is given by

    dσ(τ)=θΓ(1θ)τ1θdτ.

    The restriction θ ∈ (0, 1) and not θ ∈ (0, 2) is due to the requirement (2.1). The boundary θ = 1 corresponds to a degenerate case since S(t) = t.

    We have K(λ)=λθ1 and k(t) = tθ/Γ(1 − θ). The corresponding GFD Dt(k) is the Caputo-Djrbashian fractional derivativeDt(θ) of order θ ∈ (0, 1).

  2. The Gamma process Y(a, b) with parameters a, b > 0 is given by its Laplace exponent Φ(a, b) as

    Φ(a,b)(λ)=alog(1+λb)=0(1eλτ)aτ1ebτdτ,

    where the second equality is known as the Frullani integral [1]. The corresponding Lévy measure is given by

    dσ(τ)=aτ1ebτdτ.

    We have K(λ)=λ1alog(1+λb) and k(t) = aΓ(0, bt). The corresponding GFD is given by

    (Dt(a,b)f)(t)=ddt0tΓ(0,b(ts))(f(s)f(0))ds.

An important characteristic of the density Gt(τ) is given by its Laplace transform. More precisely, does the τ-Laplace (or t-Laplace) transform of Gt(τ) are known for an arbitrary subordinator? Thus, we would like to compute the following integrals

0eλτGt(τ)dτor0eλtGt(τ)dt.

The answer for the t-Laplace transform is affirmative and the result is given in (2.14) below. On the other hand, for the τ-Laplace transform a partial answer has been given for the class of θ-stable processes; namely

Example 2

(cf. Prop. 1(a) in [5]).

If S is a θ-stable process, then the inverse process E(t) has the Mittag-Leffler distribution, as follows,

(2.8)E(eλE(t))=n=0(λtθ)nΓ(nθ+1)=Eθ(λtθ).

It follows from the asymptotic behavior of the Mittag-Leffler function Eθ that

E(eλE(t))~Ctθ,ast.

In addition, using the fact that

(2.9)Eθ(x)=0exτMθ(τ)dτ,x0,

where Mθ is the so-called M-Wright (cf. [14] for more details and properties), it follows that

E(eλE(t))=0eλtθτMθ(τ)dτ=0eλτtθMθ(τtθ)dτ

from which we obtain the density of E(t) explicitly as

(2.10)Gt(τ)=tθMθ(τtθ).

For a general subordinator, the following lemma determines the t-Laplace transform of Gt(τ), with k and K given in (2.4) and (2.5), respectively.

Lemma 2.1

Thet-Laplace transform of the densityGt(τ) is given by

(2.11)0eλtGt(τ)dt=K(λ)eτλK(λ).

In addition, the double (τ, t)-Laplace transform ofGt(τ) is given by

00epτeλtGt(τ)dtdτ=K(λ)λK(λ)+p.

Proof

For any τ ≥ 0 let ητ be the distribution of S(τ), that is

(2.12)E(eλS(τ))=eτΦ(λ)=0eλsdητ(s).

Defining

(2.13)g(λ,τ):=K(λ)eτΦ(λ),τ,λ>0

under assumption (2.3), for all t > 0 it follows from Theorem 3.1 in [16] that the density Gt(τ) of the random variable E(t) if given by

Gt(τ)=0tk(ts)dητ(s).

It follows then that

(2.14)0eλtGt(τ)dt=g(λ,τ)=K(λ)eτΦ(λ).

In fact, by the Fubini’s theorem we obtain

0eλtGt(τ)dt=0eλt0tk(ts)dητ(s)dt=0seλtk(ts)dtdητ(s)=K(λ)0eλsdητ(s)=g(λ,τ).

In addition, it follows easily from (2.13) that

0g(λ,τ)dτ=1λ

so that (2.14) may be written as

0eλtdt0Gt(τ)dτ=1λ

which implies that Gt(τ) is a τ-density on ℝ+:

0Gt(τ)dτ=1.

Finally, the double (τ, t)-Laplace transform follows from

(2.15)00epτeλtGt(τ)dtdτ=0epτg(λ,τ)dτ=K(λ)0epτeτλK(λ)dτ=K(λ)λK(λ)+p.

3 Subordinated heat kernel

In this section, we investigate the long-time behavior of the fundamental solutions for fractional evolution equations corresponding to random time changes in the Brownian motion by the inverse process Et, t ≥ 0. We consider three classes of time change, namely those corresponding to the θ-stable subordinator, 0 < θ < 1, the distributed order derivative, and the class of Stieltjes functions. Henceforth L will always denotes a slowly varying function at infinity (SVF), that is,

limxL(λx)L(x)=1,

see for instance [4] and [19]) for more details, while C, C′ are constants whose values are unimportant, and which may change from line to line.

Let v(x, t) be the fundamental solution of the heat equation

(3.1){u(x,t)t=12Δu(x,t)u(x,0)=δ(x),

where Δ denotes the Laplacian in ℝd. It is well known that the solution v(x, t) of (3.1), called heat kernel (also known as Green function), is given by

(3.2)v(x,t)=1(2πt)d2e|x|22t

and the associated stochastic process is the classical Brownian motion in ℝd. Notice that the heat kernel v(x, t) has the following long-time behavior

v(x,t)~Ctd/2,ast.

We are interested in studying the long-time behavior of the subordination of the solution v(x, t) by the density Gt(τ), that is,

(3.3)vE(x,t):=0v(x,τ)Gt(τ)dτ=1(2π)d20τd2e|x|22τGt(τ)dτ.

Then vE(x, t) is the fundamental solution of the general fractional time differential equation, that is,

(3.4){Dt(k)u(x,t)=12Δu(x,t)u(x,0)=δ(x).

Here Dt(k) are differential-convolution operators defined, for any nonnegative kernel kLloc1(+), by

(3.5)(Dt(k)u)(t):=ddt0tk(tτ)u(τ)dτk(t)u(0),t>0.

(See [11] for more details and examples.)

In order to study the time evolution of vE(x, t), one possibility is to define its Cesaro mean

Mt(vE(x,t)):=1t0tvE(x,s)ds,

which may be written as

(3.6)Mt(vE(x,t))=0v(x,τ)Mt(Gt(τ))dτ.

The long-time behavior of the Cesaro mean Mt(vE(x,t)) was investigated in [10, Sec. 3] for the three classes of admissible kernels and d ≥ 3. The method was based on the ratio Tauberian theorem from [12]. More precisely, the following theorem was shown.

Theorem 3.1

LetvE(x, t) be the subordination ofv(x, t) by the kernelGt(τ). Then the long-time behavior of the Cesaro mean ofvE(x, t) ast → ∞ is given by

Mt(vE(x,t))~{Ctθ,k(C1),Clog(t)1,k(C2),Clog(t)1s,k(C3).

In the next section we use an alternative method to find the long-time behavior of the Cesaro mean Mt(vE(x,t)).

4 Alternative method for subordinated heat kernel

The Laplace transform method is based on the result of Lemma 2.1 wherein the t-Laplace transform of the subordination vE(x, t) is explicitly given by

(4.1)(vE(x,))(λ)=C0τd2e|x|22τ(G(τ))(λ)dτ=CK(λ)0τd/2e|x|22ττλK(λ)dτ.

The integral in (4.1) is computed according to the formula,

0τd/2eaτbτdτ={πe2abb,d=1,2K0(2ab),d=2,2(ab)(2d)/4Kd/21(2ab),d3,

(see for instance [9, page 146, eqs. (27), (29)]), where a=|x|22, b=λK(λ), and Kν(z) is the modified Bessel function of the second kind [2, Sec. 9.6]. The asymptotic of the Bessel function Kν(z) as z → 0 is well known (e.g., see [2, Eqs. (9.6.8) and (9.6.9)]) and is given by

(4.2)K0(z)~ln(z),
(4.3)Kν(z)~12Γ(ν)(z2)ν~Czν,(ν)>0.

With these explicit formulas, we study each class (C1), (C2), and (C3) separately which constitutes the main contribution of this paper.

Theorem 4.1

Letv(x, t) be the fundamental solution of the heat equation (3.1) andvE(x, t) its subordination by the densityGt(τ). Then the long-time asymptotic ofvE(x, t) is given according the admissible classes of kernelskby

  1. Mt(vE(x,t))~{Ctθ/2,d=1,Ctθlog(2|x|tθ/2),d=2,C|x|(θ+1)(2d)/2tθ,d3.
  2. Mt(vE(x,t))~{Clog(t)1/2e2μ(0)|x|log(t)1/2,d=1,Clog(t)1ln(2μ(0)|x|log(t)1/2),d=2,C|x|2dlog(t)1,d3.
  3. Mt(vE(x,t))~{Clog(t)(1+s)/2eC2|x|log(t)(1+s)/2,d=1,Clog(t)1sln(C|x|log(t)(1+s)/2),d=2,C|x|2dlog(t)1s,d3.

Proof

  1. For this class, K(λ)=λθ1, 0 < θ < 1.

    1. For d = 1, we obtain

      (vE(x,))(λ)=Cλ1+θ/2e2|x|λθ/2=λ(1θ/2)L(1λ),

      where L(y)=Ce2|x|yθ/2 is a SVF. An applicationof the Karamata Tauberian theorem (see for example [18, Sec. 2.2] or [4, Sec. 1.7]) gives

      Mt(vE(x,t))~Ctθ/2e2|x|tθ/2~Ctθ/2,t.
    2. For d = 2, we have

      (vE(x,))(λ)=Cλ(1θ)K0(2|x|λθ/2)=λ(1θ)L(1λ),

      where L(y)=CK0(2|x|yθ/2) is a SVF. Invokingthe Karamata Tauberian theorem and (4.2) yields, for t → ∞,

      Mt(vE(x,t))~CtθK0(2|x|tθ/2)~Ctθln(2|x|tθ/2).
    3. For d ≥ 3, the Laplace transform of vE(x, t) has the form

      (vE(x,))(λ)=C|x|(2d)/2λ(1θ)(1λ)θ(2d)/4Kd21(2|x|λθ/2)=λ(1θ)L(1λ),

      where L(y)=C|x|(2d)/2yθ(2d)/4Kd21(2|x|yθ/2) is a SVF. It follows from the Karamata Tauberian theorem and (4.3) that

      Mt(vE(x,t))~CtθL(t)~C|x|(θ+1)(2d)/2tθ,t.

  2. Here we have K(λ)~λ1L(λ1) as λ → 0, where L(y) = μ(0)log(y)−1, μ(0) ≠ 0.Again we study the three different cases d = 1, d = 2 and d ≥ 3.

    1. For d = 1, the t-Laplace transform of vE(x, t) can be written, for λ → 0, as

      (vE(x,))(λ)=Cλ1log(λ1)1/2e2μ(0)|x|log(λ1)1/2=λ1L(1λ),

      where L(y)=Clog(y)1/2e2μ(0)|x|log(y)1/2 is a SVF. An application of the Karamata Tauberian theorem gives

      Mt(vE(x,t))~CL(t)~Clog(t)1/2e2μ(0)|x|log(t)1/2,t.
    2. If d = 2, we have

      (vE(x,))(λ)=Cλ1log(λ1)1K0(2μ(0)|x|log(λ1)1/2)=λ1L(1λ),

      where L(y)=Clog(y)1K0(2μ(0)|x|log(y)1/2) is a SVF. As t → ∞ then by the Karamata Tauberian theorem and (4.2) we obtain

      Mt(vE(x,t))~CL(t)~Clog(t)1ln(2μ(0)|x|log(t)1/2).
    3. For d ≥ 3, it follows that, as λ → 0,

      (vE(x,))(λ)=C|x|(2d)/2λ1log(λ1)1+(2d)/4×Kd21(C|x|log(λ1)1/2)=λ1L(1λ),

      where

      L(y)=C|x|(2d)/2log(y)1+(2d)/4Kd21(C|x|log(y)1/2)

      is a SVF. To verify that L(y) is a SVF, one may note that log(y)−1+(2−d)/4 as well as is Kd21(C|x|log(y)1/2) according to (4.3); the stated result then followsfrom Prop. 1.3.6 in [4]. It follows from the KaramataTauberian theorem and (4.3) that

      Mt(vE(x,t))~CL(t)~C|x|2dlog(t)1,t.

  3. We now have K(λ)~Cλ1L(λ1)1s, as λ → 0 and s > 0, C > 0.

    1. For d = 1, the t-Laplace transform of vE(x, t) can be written, for λ → 0, as

      (vE(x,))(λ)=Cλ1log(λ1)(1+s)/2eC2|x|log(λ1)(1+s)/2=λ1L(1λ),

      where L(y)=Clog(y)(1+s)/2eC2|x|log(y)(1+s)/2 is a SVF, as is easily seen. An application of the Karamata Tauberian theorem gives, as t → ∞,

      Mt(vE(x,t))~CL(t)~Clog(t)(1+s)/2eC2|x|log(t)(1+s)/2.
    2. For d = 2, we have

      (vE(x,))(λ)=Cλ1log(λ1)1sK0(C|x|log(λ1)(1+s)/2)=λ1L(1λ),

      where

      L(y)=Clog(y)1sK0(C|x|log(y)(1+s)/2)

      is a SVF. Use the Karamata Tauberian theorem and (4.2) now yield the behavior as t → ∞

      Mt(vE(x,t))~CL(t)~Clog(t)1sln(C|x|log(t)(1+s)/2).
    3. For d ≥ 3, it follows that

      (vE(x,))(λ)=C|x|(2d)/2λ1log(λ1)(1+s)(1(2d)/4)×Kd21(C2|x|log(λ1)(1+s)/2)=λ1L(1λ),

      as t → ∞, where

      L(y)=C|x|(2d)/2log(y)(1+s)(2+d)/4Kd21(C2|x|log(y)(1+s)/2)

      is a SVF. We note that L(y) is the product of two SVF’s which isa SVF (see Prop. 1.3.6 in [4]). It the follows from the Karamata Tauberian theorem and (4.3) that

      Mt(vE(x,t))~CL(t)~C|x|2dlog(t)1s,t.

Remark

(Gaussian convolution kernel).

We consider the nonlocal operator on functions u:d defined in integral form by

(4.4)(u)(x):=(a*u)(x)u(x)=da(xy)[u(y)u(x)]dy,

where the convolution kernel a is non-negative, symmetric, bounded, and integrable, i.e.,

(4.5)a(x)0,  a(x)=a(x),  a(x)L(d)L1(d).

In addition, the kernel a is a density in ℝd with finite second moment; explicitly

(4.6)a:=da(x)dx=1,d|x|2a(x)dx<.

Since is a bounded operator in L2(ℝd), its heat semigroup et can be easily computed by using the exponential series according to

et=eteta=etk=0tkakk!=etId+etk=1tkakk!.

By removing the singular part etId of the heat semigroup, we obtain the regularized heat kernel

(4.7)v(x,t)=etk=1tkak(x)k!

with the source at the origin. In other words, for any fL2(ℝd), a solution to the nonlocal Cauchy problem

(4.8){u(x,t)t=u(x,t),u(x,0)=f(x),

has the form u(x, t) = etf(x)+(v * f)(x, t) with v given by (4.7). In particular, the fundamental solution of the problem (4.8) is

u(x,t)=etδ(x)+v(x,t).

If we denote by vE(x, t) the subordination of v(x, t) (the regular part of u(x, t)) by the density Gt(τ), then it turns out that the Cesaro mean of vE(x, t) long time behavior depends crucially on the ratio between |x| and t. The details of this investigation we postpone for a forthcoming paper.

Acknowledgement

This work has been partially supported by Center for Research in Mathematics and Applications (CIMA) related with the Statistics, Stochastic Processes and Applications (SSPA) group, through the grant UIDB/MAT/04674/2020 of FCT-Fundação para a Ciência e a Tecnologia, Portugal. The work of the first-named author was funded in part under the research work “Markov evolutions in real and p-adic spaces” of the Dragomanov National Pedagogical University of Ukraine.

References

[1] J. Arias-de Reyna, On the theorem of Frullani. Proc. Amer. Math. Soc. 109, No 1 (1990), 165–175; DOI: 10.1090/S0002-9939-1990-1007485-4.Search in Google Scholar

[2] M. Abramowitz and I. A. Stegun (Eds.), Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. Dover Publications Inc., New York, 1992; Reprint of the 1972 Edition.Search in Google Scholar

[3] J. Bertoin, Lévy Processes, Vol. 121 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1996.Search in Google Scholar

[4] N. H. Bingham, C. M. Goldie, and J. L. Teugels, Regular Variation, Vol. 27 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, 1987.10.1017/CBO9780511721434Search in Google Scholar

[5] N. H. Bingham, Limit theorems for occupation times of Markov processes. Z. Wahrsch. Verw. Gebiete17 (1971), 1–22; DOI: 10.1007/BF00538470.Search in Google Scholar

[6] B. Baeumer and M. M. Meerschaert, Stochastic solutions for fractional Cauchy problems. Fract. Calc. Appl. Anal. 4, No 4 (2001), 481–500.Search in Google Scholar

[7] Z.-Q. Chen, Time fractional equations and probabilistic representation. Chaos Solitons Fractals102 (2017), 168–174; DOI: 10.1016/j.chaos.2017.04.029.Search in Google Scholar

[8] Z.-Q. Chen, P. Kim, T. Kumagai, and J. Wang, Heat kernel estimates for time fractional equations. Forum Math. 30, No 5 (2018), 1163–1192; DOI: 10.1515/forum-2017-0192.Search in Google Scholar

[9] A. Erdélyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi (Eds.), Tables of Integral Transforms, Vol. I. McGraw-Hill Book Company, Inc., New York-Toronto-London, 1954.Search in Google Scholar

[10] A. N. Kochubei, Yu. G. Kondratiev, and J. L. da Silva, Random time change and related evolution equations. Time asymptotic behavior. Stochastics and Dynamics4 (2019), Art. 2050034, 24 pp.; DOI: 10.1142/S0219493720500343.Search in Google Scholar

[11] A. N. Kochubei, General fractional calculus, evolution equations, and renewal processes. Integral Equations Operator Theory71, No 4 (2011), 583–600; DOI: 10.1007/s00020-011-1918-8.Search in Google Scholar

[12] Y.-C. Li, R. Sato, and S.-Y. Shaw, Ratio Tauberian theorems for positive functions and sequences in Banach lattices. Positivity11, No 3 (2007), 433–447; DOI: 10.1007/s11117-007-2085-7.Search in Google Scholar

[13] A. Mimica, Heat kernel estimates for subordinate Brownian motions. Proc. Lond. Math. Soc. (3)113, No 5 (2016), 627–648; DOI: 10.1112/plms/pdw043.Search in Google Scholar

[14] F. Mainardi, A. Mura, and G. Pagnini, The M-Wright function in time-fractional diffusion processes: A tutorial survey. Int. J. Differential Equ. 2010 (2010), Art. 104505, 29 pp.; DOI: 10.1155/2010/104505.Search in Google Scholar

[15] M. M. Meerschaert and H.-P. Scheffler, Limit theorems for continuous-time random walks with infinite mean waiting times. J. Appl. Probab. 41, No 3 (2004), 623–638; DOI: 10.1239/jap/1091543414.Search in Google Scholar

[16] M. M. Meerschaert and H.-P. Scheffler, Triangular array limits for continuous time random walks. Stochastic Process. Appl. 118, No 9 (2008), 1606–1633; DOI: 10.1016/j.spa.2007.10.005.Search in Google Scholar

[17] M. Magdziarz and R. L. Schilling, Asymptotic properties of Brownian motion delayed by inverse subordinators. Proc. Amer. Math. Soc. 143, No 10 (2015), 4485–4501; DOI: 10.1090/proc/12588.Search in Google Scholar

[18] E. Seneta, Regularly Varying Functions, Vol. 508 of Lect. Notes Math., Springer, 1976.10.1007/BFb0079658Search in Google Scholar

[19] R. L. Schilling, R. Song, and Z. Vondraček, Bernstein Functions: Theory and Applications. De Gruyter Studies in Math., De Gruyter, Berlin, 2 Edition, 2012.10.1515/9783110269338Search in Google Scholar

[20] B. Toaldo, Convolution-type derivatives, hitting-times of subordinators and time-changed C0-semigroups. Potential Anal. 42, No 1 (2015), 115–140; DOI: 10.1007/s11118-014-9426-5.Search in Google Scholar

Received: 2020-08-04
Published Online: 2021-01-29
Published in Print: 2021-02-23

© 2021 Diogenes Co., Sofia

Downloaded on 19.4.2024 from https://www.degruyter.com/document/doi/10.1515/fca-2021-0004/html
Scroll to top button