Fractional calculus with analytic kernels provides a general setting of integral and derivative operators that can be connected to Riemann–Liouville fractional calculus via convergent infinite series. We interpret these operators from an algebraic viewpoint, using Mikusiński's operational calculus, and utilise this algebraic formalism to solve some fractional differential equations.
Citation: Noosheza Rani, Arran Fernandez. An operational calculus formulation of fractional calculus with general analytic kernels[J]. Electronic Research Archive, 2022, 30(12): 4238-4255. doi: 10.3934/era.2022216
[1] | Mehmet Ali Özarslan, Ceren Ustaoğlu . Extended incomplete Riemann-Liouville fractional integral operators and related special functions. Electronic Research Archive, 2022, 30(5): 1723-1747. doi: 10.3934/era.2022087 |
[2] | María Guadalupe Morales, Zuzana Došlá, Francisco J. Mendoza . Riemann-Liouville derivative over the space of integrable distributions. Electronic Research Archive, 2020, 28(2): 567-587. doi: 10.3934/era.2020030 |
[3] | J. Vanterler da C. Sousa, Kishor D. Kucche, E. Capelas de Oliveira . Stability of mild solutions of the fractional nonlinear abstract Cauchy problem. Electronic Research Archive, 2022, 30(1): 272-288. doi: 10.3934/era.2022015 |
[4] | Seda IGRET ARAZ, Mehmet Akif CETIN, Abdon ATANGANA . Existence, uniqueness and numerical solution of stochastic fractional differential equations with integer and non-integer orders. Electronic Research Archive, 2024, 32(2): 733-761. doi: 10.3934/era.2024035 |
[5] | Jin Li, Yongling Cheng . Barycentric rational interpolation method for solving fractional cable equation. Electronic Research Archive, 2023, 31(6): 3649-3665. doi: 10.3934/era.2023185 |
[6] | Jin Li, Yongling Cheng . Barycentric rational interpolation method for solving time-dependent fractional convection-diffusion equation. Electronic Research Archive, 2023, 31(7): 4034-4056. doi: 10.3934/era.2023205 |
[7] | Youming Chen, Weiguo Lyu, Song Yang . A note on the differential calculus of Hochschild theory for $ A_{\infty} $-algebras. Electronic Research Archive, 2022, 30(9): 3211-3237. doi: 10.3934/era.2022163 |
[8] | Ismail T. Huseynov, Arzu Ahmadova, Nazim I. Mahmudov . Perturbation properties of fractional strongly continuous cosine and sine family operators. Electronic Research Archive, 2022, 30(8): 2911-2940. doi: 10.3934/era.2022148 |
[9] | Yaning Li, Mengjun Wang . Well-posedness and blow-up results for a time-space fractional diffusion-wave equation. Electronic Research Archive, 2024, 32(5): 3522-3542. doi: 10.3934/era.2024162 |
[10] | Dewang Li, Meilan Qiu, Jianming Jiang, Shuiping Yang . The application of an optimized fractional order accumulated grey model with variable parameters in the total energy consumption of Jiangsu Province and the consumption level of Chinese residents. Electronic Research Archive, 2022, 30(3): 798-812. doi: 10.3934/era.2022042 |
Fractional calculus with analytic kernels provides a general setting of integral and derivative operators that can be connected to Riemann–Liouville fractional calculus via convergent infinite series. We interpret these operators from an algebraic viewpoint, using Mikusiński's operational calculus, and utilise this algebraic formalism to solve some fractional differential equations.
Fractional calculus is a field of study, like many in mathematics, which arises from a type of generalisation starting from a basic concept: in this case, taking the concepts of differentiation and integration, and generalising them by allowing the orders of derivatives and integrals to take values outside of the set of integers. As well as this purely mathematical motivation, the study of such generalised operators is also useful for scientific applications. The theory of fractional calculus is introduced in textbooks such as [1,2,3,4] while the applications are summarised in survey works such as [5,6].
Surprisingly, perhaps, there are many different definitions of fractional integrals and derivatives in the literature, with no single agreed-upon way to define what (for example) the "derivative to order one-half" should actually mean. Especially in the early 21st century, the number of possible definitions has been growing out of control, including many that are purely mathematical curiosities without any verified applications. In order to put some structure on this rapidly growing field, we return to the mathematical concept of generalisation: it must be possible to define some broader categories [7], with some unifying properties and behaviours, which many of the recently defined operators can be slotted into.
Many of the newly defined fractional integral operators have been created by replacing the power function in the classical Riemann–Liouville integral by a different kernel function, which may have various properties such as singular or nonsingular, etc. For this reason, several researchers have defined general classes of fractional integral operators by taking convolutions with some more general function: for example, Sonine kernels [8,9], analytic kernels [10] and others [11] including some introduced from the viewpoint of applications [12]. We will focus, in this paper, on the general class of operators with analytic kernels, in which the fractional integral operator AIα,β0+ is defined for Re(α)>0 and Re(β)>0 by
AIα,β0+f(x)=∫x0(x−t)α−1A((x−t)β)f(t)dt,x>0, | (1.1) |
where the kernel function A can be any analytic function, with a power series like the following:
A(x)=∞∑r=0ar(α,β)xr. | (1.2) |
Note that the coefficient ar may, in general, depend on the parameters α and β. We believe that this definition is general enough to cover many useful cases with real-world applications, while also specific enough that the solutions can be expressed in a form which is not too abstract to compute easily in those special cases.
A powerful method for solving differential equations is the method of operational calculus [13], formally introduced by Heaviside and later made rigorous by mathematicians including Mikusiński. Briefly, the idea of Mikusiński's operational calculus is to introduce a type of generalised functions (different from those used in distribution theory), so-called "operators" which behave like functions but are not, and which give rise to a unified algebraic structure [14]. The results yielded by Mikusiński's methodology are formally similar to those yielded by Laplace transform methods, but Mikusiński's method has advantages [15] both in how easy it is to make rigorous (the Laplace transform is easy to apply but less easy to justify rigorously) and in how many problems it can apply to (the Laplace transform requires exponential boundedness while Mikusiński's method requires only continuity).
This method has been applied to ordinary differential equations [14], partial differential equations [16], and fractional-order differential equations [17] with fractional derivatives of Riemann–Liouville type [18], Caputo type [19], Hilfer type [20], Erdélyi–Kober type [21,22], Prabhakar type [23,24] and in general classes such as fractional calculus with Sonine kernels [25,26] and fractional calculus with respect to functions [27,28].
In our recent work applying Mikusiński's operational calculus to differential equations of Prabhakar type [23,24], we made use of the series formula for Prabhakar operators [29] to complete the work inside the same function spaces as those used for Riemann–Liouville and Caputo differential equations. The most general setting in which such a series formula can be used is the fractional calculus with general analytic kernels [10], and therefore we wish to focus on this general class and try to find a way to apply Mikusiński's method to differential equations in this model. Such a result would be valuable in providing a greater level of generality to the work on Mikusiński's operational calculus in fractional calculus: by going straight to the general setting, we eliminate the need for multiple papers solving smaller subproblems of this general one.
The structure of this paper is as follows. Section 2 provides key definitions and facts concerning Riemann–Liouville fractional calculus which will be used later. Section 3 is devoted to a development of fractional calculus with analytic kernels in the setting of suitable function spaces, including mapping properties, composition relations, and series formulae for both fractional integral and derivative operators. Section 4 develops a Mikusiński-type operational calculus for these operators, using results of Section 3 and also the general theory of Mikusiński's operational calculus for fractional operators established in previous literature. Section 5 considers applications of the developed theory in solving some initial value problems, and Section 6 concludes the paper with some comparative discussion and pointers towards future work.
The purpose of this section is to briefly introduce the key definitions and some important fundamental properties of Riemann–Liouville and Caputo fractional calculus.
Definition 2.1 ([2,3]). For any α∈C with Re(α)>0, the Riemann–Liouville fractional integral to order α, with constant of integration 0, is the operator RLIα0+ defined by
RLIα0+f(x)=1Γ(α)∫x0(x−t)α−1f(t)dt,x>0, | (2.1) |
where f is any function locally L1 on the positive axis.
For any α∈C with Re(α)≥0, say n−1≤Re(α)<n∈N, the Riemann–Liouville and Caputo fractional derivatives to order α, with constant of integration 0, are the operators RLDα0+ and CDα0+ defined as follows:
RLDα0+f(x)=dndxn(RLIn−α0+f(x)),x>0, | (2.2) |
CDα0+f(x)=RLIn−α0+(dndxnf(x)),x>0, | (2.3) |
where f is any function such that these expressions are defined (for example, a function which is in the ACn space on any compact positive interval).
An important fact about the Riemann–Liouville derivative [2] is that it is precisely the (unique) analytic continuation, in the complex variable α, of the Riemann–Liouville integral, under the convention that
RLIα0+=RLD−α0+,Re(α)<0. | (2.4) |
This means that, in some mathematical sense, the Riemann–Liouville derivative (2.2) is the "natural" fractional derivative arising from the fractional integral (2.1). However, the Caputo derivative (2.3) is nowadays often preferred in applications, because Caputo fractional differential equations require initial conditions of classical type while Riemann–Liouville ones require fractional initial conditions.
There are several important relations involving the Riemann–Liouville integrals and derivatives ("differintegrals" is a word often used in fractional calculus to include both integrals and derivatives), and we now focus on the composition relations as these will be used later in our work below. The composition of two Riemann–Liouville differintegrals, following the convention (2.4) to allow any orders in C, always has a semigroup property if the inner operator is a fractional integral:
RLIα0+(RLIβf(x))=RLIα+β0+f(x),α,β∈C,Reβ>0. | (2.5) |
On the other hand, if the inner operator is a fractional derivative, we have a modified semigroup property involving a finite sum of initial value terms, as follows:
RLDα0+(RLDβf(x))=RLDα+β0+f(x)−n−1∑k=0xk−n−αΓ(k−n−α+1)⋅[RLDβ−n+k0+f(x)]x=c, | (2.6) |
where α,β∈C with Reβ≥0 and the natural number n is defined to be ⌊Reβ⌋+1. For more details on these and other composition relations between Riemann–Liouville differintegrals, the reader may refer to classic works such as [3,§2.7] and [30,§2.3], as well as the recent survey of composition relations [31].
Let us begin by defining the important function spaces to be used throughout this paper. In the following definition, the space Cη comes from Dimovski [32], the space Ωαη from Hadid and Luchko [18], and the space Cnη from Luchko and Gorenflo [19].
Definition 3.1 ([18,19,32]). Let η∈R and α≥0 and m∈Z+0 be fixed.
(a) Cη is the space of all functions expressible as f(x)=xpf1(x) with p>η in R and f1∈C[0,∞).
(b) Ωαη is the space of all functions f∈Cη such that the Riemann–Liouville derivatives RLDν0+f are included in Cη for all real numbers ν with 0≤ν≤α.
(c) Cmη is the space of all m times differentiable functions f such that the mth derivative f(m) is included in Cη.
Clearly, all three of these are vector spaces. An important feature of the Cη space is that, for η≥−1, any function f∈Cη must be continuous on (0,∞) and integrable on any compact [0,X]. They also have the following convolution property: if f∈Cη and g∈Cμ and η,μ≥−1, then f∗g∈Cη+μ+1, where this is the Laplace-type convolution defined by
(f∗g)(x)=∫x0f(x−t)g(t)dt,x>0. | (3.1) |
It is clear that Ω0η=Cη=C0η for any η∈R. We also have the following inclusions:
Cη⊆Cμ⟺η≥μ;Ωαη⊆Ωαν⟺η≥ν;Ωα1η⊆Ωα2η⟺α1≥α2. |
Let us also write down the formal definition of the fractional integral operator with general analytic kernel, while leaving the corresponding derivative operators for the next subsection.
Definition 3.2 ([10]). Let α,β∈C be any parameters with positive real parts, and let A be a function defined by a power series (1.2) with radius of convergence R>XReα. The fractional integral operator with kernel function A and power parameters α and β is defined by the formula (1.1) for any function f on the interval [0,X] such that this integral is well-defined.
One possible function space for this operator is L1[0,X], as proved in [10]. The following theorem shows that C−1 is also a suitable function space for the same operator, assuming that the kernel function A is entire (infinite radius of convergence).
Theorem 3.1. Let α,β∈R+, and let A be a function as in Definition 3.2 which is entire. Then, for any given η≥−1, the operator AIα,β0+ is a linear map of the space Cη into itself:
AIα,β0+:Cη→Cη+α⊆Cη. | (3.2) |
Also, for any function f∈Cη, the infinite series representation
AIα,β0+f(x)=∞∑r=0ar(α,β)Γ(rβ+α)RLIrβ+α0+f(x) | (3.3) |
is locally uniformly convergent for x∈(0,∞).
Proof. The integral with analytic kernel is clearly a linear operator. For the Cη mapping, we write f(x)=xpf1(x) with p>η and f1∈C[0,∞), and set t=xτ in (1.1) to obtain
AIα,β0+f(x)=∫10(x−xτ)α−1A((x−xτ)β)(xτ)pf1(xτ)xdτ=xp+α∫10τp(1−τ)α−1A(xβ(1−τ)β)f1(xτ)dt,=xp+αf2(x), |
where the last integral converges, uniformly with respect to x in any closed interval [0,X], which means f2∈C[0,∞) and AIα,β0+f∈Cη+α. Now
AIα,β0+f(x)=∫x0(x−t)α−1A((x−t)β)f(t)dt=∫x0(x−t)α−1∞∑r=0ar(α,β)(x−t)rβf(t)dt=∫x0∞∑r=0ar(α,β)(x−t)rβ+α−1f(t)dt=∞∑r=0ar(α,β)Γ(rβ+α)⋅1Γ(rβ+α)∫x0(x−t)rβ+α−1f(t)dt=∞∑r=0ar(α,β)Γ(rβ+α)RLIrβ+α0+f(x), |
where interchange of the summation and integration is justified by local uniform convergence of the analytic power series, which also guarantees local uniform convergence of the series formula (3.3).
Proposition 3.1. Let A be an entire kernel function, and fix η≥−1. The fractional integral with kernel A has a semigroup property in the space Cη, of the form
AIα1,β0+∘AIα2,β0+f=AIα1+α2,β0+f,f∈Cη, | (3.4) |
for all α1,α2,β∈C, if and only if
∑r1+r2=kar1(α1,β)ar2(α2,β)B(βr1+α1,βr2+α2)=ak(α1+α2,β),k∈Z+0. |
Proof. Identical to the proof of [10,Theorem 2.10]. Indeed, the result is exactly the same except that here we have specified a function space for f.
Every fractional integral operator should have a corresponding fractional derivative operator. Usually, in line with the classical fundamental theorem of calculus, these operators would have inversion relations as follows: the fractional derivative is a left inverse to the fractional integral, while the fractional integral of the fractional derivative equals the original function minus some initial value terms.
The fractional derivative operators of Riemann–Liouville and Caputo type corresponding to (1.1) were originally defined [10,Definition 2.15] as follows:
ARLDα,β0+f(x)=dmdxm(∫x0(x−t)m−α−1ˉA((x−t)β)f(t)dt,),ACDα,β0+f(x)=∫x0(x−t)m−α−1¯A((x−t)β)dmdtmf(t)dt, |
where m=⌊Reα⌋+1 and f is in some suitable function space in each case, and where ¯A is defined by [10,Eq. (36)]:
¯AΓ⋅AΓ=1, | (3.5) |
where the Γ-transformed version of a function is defined by multiplying each coefficient in its Taylor series by a gamma function multiplier:
A(x)=∞∑n=0anxn⟹AΓ(x)=∞∑n=0anΓ(βn+α)xn. |
The condition (3.5) is, as claimed in [10], sufficient to ensure that the Riemann–Liouville type fractional derivative ARLDα,β0+ is a left inverse to the fractional integral operator AIα,β.
However, there is more subtlety here than was realised in [10]. We must take into account the fact that the coefficients of the power series for A are allowed to depend on α and β. Indeed, a better notation, taking into account the dependence on the parameters, might be
A(α,β)(x)=∞∑n=0an(α,β)xn, |
and then the condition (3.5) should be replaced by
¯A(m−α,β)Γ⋅A(α,β)Γ=1,m=⌊α⌋+1, | (3.6) |
and it is now apparent that this assumption is actually quite complicated and less natural than it originally seemed.
To simplify the discussion, and to focus attention on the case which is most often of interest, let us assume that 0<α<1, so that m=1 and the relation (3.6) does not depend on a variable discrete natural number as well as variable real or complex parameters. We state the inversion relation rigorously, including relevant function spaces, in the following theorem.
Theorem 3.2. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2).
(i) The Riemann–Liouville type fractional derivative operator with kernel A, namely
ARLDα,β0+f(x)=ddx(∫x0(x−t)−αˉA((x−t)β)f(t)dt), |
where ¯A(x)=∑∞n=0¯anxn is defined by the identity
(∞∑r=0ar(α,β)Γ(βr+α)xr)⋅(∞∑n=0¯an(1−α,β)Γ(βn+1−α)xn)=1, | (3.7) |
maps the function space Ωαη into Cη:
ARLDα,β0+:Ωαη→Cη. | (3.8) |
Indeed, for any function f∈Ωαη, the infinite series
∞∑n=0¯an(1−α,β)Γ(nβ+1−α)RLInβ−α0+f(x) | (3.9) |
converges in Cη, locally uniformly for all x∈(0,∞), to the fractional derivative ARLDα,β0+f(x).
(ii) The Riemann–Liouville type fractional derivative operator with kernel A is a left inverse of the fractional integral operator with kernel A, in the sense that, for any f∈Cη, we have AIα,β0+f∈Ωαη and
ARLDα,β0+(AIα,β0+f)=f∈Cη. | (3.10) |
Proof. (ⅰ) We let f∈Ωαη⊂Cη and prove the series formula and function space inclusion for ARLDα,β0+f∈Cη. We can use the series formula for the fractional integral with analytic kernel, by Theorem 3.1:
ARLDα,β0+f(x)=ddx(∞∑n=0¯an(1−α,β)Γ(nβ+1−α)RLInβ+1−α0+f(x)), |
where the series inside the parentheses is locally uniformly convergent. The above series is equal to
∞∑n=0ddx(¯an(1−α,β)Γ(nβ+1−α)RLInβ+1−α0+f(x)), |
or equivalently
∞∑n=0¯an(1−α,β)Γ(nβ+1−α)RLInβ−α0+f(x), |
as long as this new series is also locally uniformly convergent.
Note that the operator RLInβ−α0+ in the series is a fractional derivative if n is small enough that Re(nβ−α)≤0, while it is a fractional integral for all n large enough that Re(nβ−α)>0. So we split the series into two subseries according to the sign of Re(nβ−α). Let N:=⌊α/β⌋ denote the cutoff point, and we have:
ARLDα,β0+f(x)=J1+J2=N∑n=0¯an(1−α,β)Γ(nβ+1−α)RLDα−nβ0+f(x)+∞∑n=N+1¯an(1−α,β)Γ(nβ+1−α)RLInβ−α0+f(x). |
Now, since f∈Ωαη, the first sum J1 is a finite sum of elements of Cη and therefore also in Cη. It remains to prove that the second sum J2 is locally uniformly convergent to an element of Cη.
Since N+1>α/β, we can choose ε>0 so that (N+1)β−α−ε>0. Then, letting g=RL0I(N+1)β−α−εxf∈Cη, we have
J2=∞∑n=0¯an+N+1(1−α,β)Γ((n+N+1)β+1−α)RL0Inβ+εxg(x)=(∞∑n=0Γ(nβ+(N+1)β+1−α)Γ(nβ+ε)⋅¯an+N+1(1−α,β)xnβ+ε−1)∗g, |
where this power series has the same radius of convergence as ¯A due to [1,Chapter Ⅳ Theorem 1], and is therefore locally uniformly convergent and in C−1. Its conjugation with g must then be in Cη, and we have proved both the mapping property of ARLDα,β0+ and the series formula for it.
(ii) Now we let f∈Cη and show that AIα,β0+f∈Ωαη by using the series formula from Theorem 3.1. For any ν with 0≤ν<α, we have:
RLDν0+(AIα,β0+f(x))=RLDν0+(∞∑r=0ar(α,β)Γ(rβ+α)RLIrβ+α0+f(x))=∞∑r=0ar(α,β)Γ(rβ+α)RLDν0+(RLIrβ+α0+f(x))=∞∑r=0ar(α,β)Γ(rβ+α)RLIrβ+α−ν0+f(x)=(∞∑r=0Γ(rβ+α)Γ(rβ+α−ν)⋅ar(α,β)xrβ+α−ν−1)∗f, |
and this is the convolution of f with a function in C−1 (convergence of the power series is again guaranteed by [1]), therefore in Cη as required. Setting ν=α, we get a very similar expression but the first term of the series must be treated differently:
RLDα0+(AIα,β0+f(x))=a0(α,β)Γ(α)f(x)+(∞∑r=1Γ(rβ+α)Γ(rβ)⋅ar(α,β)xrβ−1)∗f, |
which is again in Cη. Now we have proved that AIα,β0+f is in the space Ωαη defined in Definition 3.1(b).
To prove the inversion relation (3.10), the argument is the same as in [10,§2.2], and goes as follows:
ARLDα,β0+(AIα,β0+f(x))=∞∑n=0¯an(1−α,β)Γ(nβ+1−α)RLInβ−α0+(∞∑r=0ar(α,β)Γ(rβ+α)RLIrβ+α0+f(x))=∞∑n=0∞∑r=0¯an(1−α,β)ar(α,β)Γ(nβ+1−α)Γ(rβ+α)RLInβ+rβ0+f(x)=f(x), |
using the inversion relation (3.7) between the AΓ and ¯AΓ series.
Theorem 3.3. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2). Then the Caputo type fractional derivative operator with kernel A, namely
ACDα,β0+f(x)=∫x0(x−t)−αˉA((x−t)β)f′(t)dt, |
where ¯A(x)=∑∞n=0¯anxn is defined by (3.7), is well-defined on the space C1η and maps this space into Cη:
ACDα,β0+:C1η→Cη. | (3.11) |
Further, for any function f∈C1η, the infinite series
∞∑n=0¯an(1−α,β)Γ(nβ+1−α)RLInβ+1−α0+f′(x) | (3.12) |
converges in Cη, locally uniformly for all x∈(0,∞), to the fractional derivative ACDα,β0+f(x).
Proof. Let f∈C1−1. By Definition 3.1(c), f′∈C−1. Then by the definition of the Caputo type operator, it is immediate that ACDα,β0+f=¯AI1−α,β0+f′∈C−1, giving the desired inclusion. The series formula (3.12) is also an immediate consequence of Theorem 3.1.
Theorem 3.4. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2). Then, for any f∈Ωαη, we have the following inversion relation:
AIα,β0+ARLDα,β0+f(x)=f(x)−xα−1A(α,β)(xβ)⋅N∑n=0¯an(1−α,β)Γ(nβ+1−α)RLDα−1−nβ0+f(0), |
where N:=⌊α/β⌋.
Proof. We start by employing the series formula (3.3) and (3.9):
AIα,β0+ARLDα,β0+f(x)=∞∑r=0ar(α,β)Γ(rβ+α)RLIrβ+α0+(∞∑n=0¯an(1−α,β)Γ(nβ+1−α)RL0Inβ−αxf(x))=∞∑r=0∞∑n=0ar(α,β)¯an(1−α,β)Γ(rβ+α)Γ(nβ+1−α)RLD−rβ−α0+RLD−nβ+α0+f(x), |
where we have made use of the notational convention (2.4) that fractional integrals to positive order are fractional derivatives to negative order. The outer operator RLD−rβ−α0+ is a fractional integral for all values of r; the inner operator RLD−nβ+α0+ in the series is a fractional derivative if n is small enough that −nβ+α≥0, while it is a fractional integral for all n large enough that −nβ+α<0.
As discussed in Section 2 above, in Riemann–Liouville fractional calculus, any differintegral of an integral obeys a semigroup property (2.5), while differintegrals of derivatives satisfy instead a relation of the form (2.6). So we have
RLD−rβ−α0+RLD−nβ+α0+f(x)=RLD−rβ−nβ0+f(x) |
if −nβ+α<0, and
RLD−rβ−nβ0+f(x)−Mn−1∑k=0xk−(−rβ−α)−MnΓ(k−(−rβ−α)−Mn+1)RLD−nβ+α−Mn+k0+f(0) |
if −nβ+α≥0, where Mn=⌊−nβ+α⌋+1∈Z+. Splitting the inner series in the above double sum according to the sign of −nβ+α, and letting N:=⌊α/β⌋ denote the cutoff point, we have:
AIα,β0+ARLDα,β0+f(x)=∞∑r=0N∑n=0ar(α,β)¯an(1−α,β)Γ(rβ+α)Γ(nβ+1−α)×(RLIrβ+nβ0+f(x)−Mn−1∑k=0xrβ+α−Mn+kΓ(rβ+α−Mn+k+1)RLD−nβ+α−Mn+k0+f(0))+∞∑r=0∞∑n=N+1ar(α,β)¯an(1−α,β)Γ(rβ+α)Γ(nβ+1−α)RLIrβ+nβ0+f(x)=∞∑r=0∞∑n=0ar(α,β)¯an(1−α,β)Γ(rβ+α)Γ(nβ+1−α)RLI(r+n)β0+f(x)−N∑n=0Mn−1∑k=0¯an(1−α,β)Γ(nβ+1−α)xα−Mn+k×[∞∑r=0ar(α,β)Γ(rβ+α)xrβΓ(rβ+α−Mn+k+1)]RLD−nβ+α−Mn+k0+f(0) |
Now, the double infinite sum reduces to a single term f(x) by using the relation (3.7) between A and ¯A, while the infinite sum over r in square brackets is locally uniformly convergent by using [1,Chapter Ⅳ Theorem 1] again. So we have
AIα,β0+ARLDα,β0+f(x)=f(x)−N∑n=0Mn∑k=1¯an(1−α,β)Γ(nβ+1−α)xα−k[∞∑r=0ar(α,β)Γ(rβ+α)xrβΓ(rβ+α−k+1)]RLD−nβ+α−k0+f(0). |
This looks like the final answer, but it can still be simplified further. As 0<α<1, by definition of N we have 0≤−nβ+α≤α<1 for all n≤N. Therefore, Mn=1 for all n≤N, and the sum over k reduces to just a single term for every value of n, giving:
A0Iα,βxARLDα,βxf(x)=f(x)−N∑n=0¯an(1−α,β)Γ(nβ+1−α)xα−1×[∞∑r=0ar(α,β)Γ(rβ+α)xrβΓ(rβ+α)]RLD−nβ+α−10+f(0)=f(x)−xα−1[∞∑r=0ar(α,β)xrβ]N∑n=0¯an(1−α,β)Γ(nβ+1−α)RLD−nβ+α−10+f(0)=f(x)−xα−1A(xβ)N∑n=0¯an(1−α,β)Γ(nβ+1−α)RLD−nβ+α−10+f(0), |
which completes the proof.
Theorem 3.5. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2). Then, for any f∈C1η, we have the following inversion relation:
(AIα,β0+ACDα,β0+f)(x)=f(x)−f(0),x≥0. |
Proof. By the definition of the Caputo-type fractional derivative with kernel A, and the relation (3.7) between A and ¯A, we have
(AIα,β0+ACDα,β0+f)(x)=RLIα0+AΓ(RLIβ0+)¯AΓ(RLIβ0+)RLI1−α0+f′(x)=RLIα0+RLI1−α0+f′(x)=RLI10+f′(x)=f(x)−f(0), |
where we have used 0<α<1 to get this simple form of the inversion relation with just a single initial value term.
Remark 3.1. It is worth noting that the inversion relation (3.7) defining ¯A bears considerable similarity to some series convolution relations discussed previously in the literature: by Sonine [33] following Eq (5) in his paper of 1883, and by Wick [34] in his paper of 1968, although these works both used ordinary power series, essentially taking β=1 in our formulation, rather than having a fractional power inside the series.
Indeed, under certain assumptions on the parameters involved, the analytic kernel in (1.1) can be seen as a special case of the Sonine kernel. However, fractional calculus with analytic kernels and fractional calculus with Sonine kernels form two separate fields of study, because analytic kernels are not assumed to have a Sonine relationship and Sonine kernels are not assumed to have a power series. One of the purposes of this paper is to show how the series formula can give an advantage in dealing with fractional calculus with general kernels, leading to approaches which work for analytic kernels but would not work in general for Sonine kernels. It should also be noted that the operators considered here have two fractional orders, α and β, and thus it makes sense to speak about properties such as semigroup (Proposition 3.1). The Sonine kernels do not have any parameters (fractional orders), so they have inversion relations but no semigroup properties.
We recall from the classical work of Mikusiński [14], Dimovski [32] and Luchko [17] the following facts. The function space C−1 forms a commutative rng (ring without multiplicative identity) under the operations of addition and convolution in the form of (3.1). Because this rng has no zero divisors, due to a theorem of Titchmarsh, it can be extended to a field M−1, its field of fractions. The elements of this field are abstract algebraic objects, which can be seen as operators or generalised functions, and are not necessarily functions themselves.
The operation of the fractional integral AIα,β0+ with analytic kernel A corresponds to the function Kα,β∈C−1 given by Kα,β(x)=xα−1A(xβ). We let Rα,β denote its algebraic inverse in M−1, a field element which is not itself a function in the function space C−1.
Theorem 4.1. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2). The fractional differential operator with analytic kernel of Riemann–Liouville type may be represented in the field M−1 in the following form:
ARLDα,β0+f=Rα,β∗f−(N∑n=0Γ(nβ+1−α)RL0Dα−1−nβxf(0))I,f∈Ωα−1, | (4.1) |
where N:=⌊α/β⌋ and I denotes the multiplicative identity element in the field M−1.
Proof. Let f∈Ωα−1. From Theorem 3.4, interpreting the result in the notation of the Kα,β function, we have
Kα,β∗ARLDα,β0+f=f−(N∑n=0Γ(nβ+1−α)RLDα−1−nβ0+f(0))Kα,β. |
Upon multiplying both sides by Rα,β, the algebraic inverse of Kα,β, we get the required result.
Theorem 4.2. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2). The fractional differential operator with analytic kernel of Caputo type may be represented in the field M−1 in the following form:
ACDα,β0+f=Rα,β∗f−Rα,β∗[f(0)1],f∈C1−1, | (4.2) |
where f(0)1 denotes the scalar f(0) times the constant function 1.
Proof. Let f∈Cm−1, and use the result of Theorem 3.5 with the notation of Kα,β:
Kα,β∗ARLDα,β0+f=f−f(0)1. |
Upon multiplying both sides by Rα,β, the algebraic inverse of Kα,β, we get the required result.
In the remainder of this paper, we shall assume A is such that the integral operator AIα,β0+ obeys a semigroup property in the form of (3.4). Under this assumption, we can take convolutions of the function Kα,β with itself to find
Knα,β:=Kα,β∗⋯∗Kα,β=Knα,β. | (4.3) |
Following the same notion, we can then define arbitrary positive real powers of Kα,β:
Kλα,β=Kλα,β(x),λ>0. | (4.4) |
Making use also of the algebraic inverse Rα,β, we can define arbitrary real powers:
Kλα,β={Kλα,β,λ>0;identity,λ=0;Rλα,β,λ<0, |
and similarly for arbitrary real powers of Rα,β too. It is clear that these real powers satisfy a semigroup property:
Rλα,β∗Rνα,β=Rλ+να,β,λ,ν∈R. | (4.5) |
Example 4.1. Let 0<α<1 and β>0 and η≥−1, and let A be a function defined by (1.2) and satisfying the semigroup property (3.4). For any constants c∈R and m∈N, we consider the field element
I(Rα,β−cI)m, |
and rewrite it as a sum of negative powers of Rα,β in order to express it as a function in C−1:
I(Rα,β−cI)m=R−mα,βI(I−cRα,β)m=R−mα,β∞∑i=0(m)ii!ciR−iα,β=∞∑i=0(m)ii!ciR−m−iα,β=∞∑i=0(m)ii!ciK(m+i)α,β, |
where
∞∑i=0(m)ii!ciK(m+i)α,β(x)=∞∑i=0(m)ii!cix(m+i)α−1A(xβ)=∞∑i=0ci(m)ii!x(m+i)α−1∞∑r=0ar((m+i)α,β)xrβ=∞∑i=0∞∑r=0ar((m+i)α,β)ci(m)ii!x(m+i)α+rβ−1. | (4.6) |
We mention this here, both because it is an example of doing algebraic manipulations in the abstract field M−1 to obtain a function in C−1, and because it will be useful in the next section when we study some fractional differential equations.
In this section, we shall make use of the mathematical terminology and toolset developed above, to solve some fractional differential equations involving fractional derivatives with general analytic kernels. As in the last part of the preceding section, we shall assume the kernel A is such that the semigroup property (3.4) is always valid; this is necessary to obtain relatively simple forms of the solution functions for the differential equations below.
Example 5.1. We consider the following initial value problem of Riemann–Liouville type, where 0<α<1 and β>0 and λ∈R are fixed while A is a function defined by (1.2) and satisfying the semigroup property (3.4).
ARLDα,β0+y(x)−λy(x)=f(x),x>0; | (5.1) |
RLDα−1−nβ0+y(0)=cn,0≤n≤N, | (5.2) |
where N:=⌊α/β⌋ is a finite non-negative integer and c0,c1,⋯,cN are fixed constants.
Making use of Theorem 4.1, we can rewrite this problem using the algebraic element Rα,β in the field M−1:
Rα,β∗y−λy=f+(N∑n=0Γ(nβ+1−α)RL0Dα−1−nβxf(0))I=f+CI, |
where I denotes the multiplicative identity in the field M−1 and the constant C∈R is given by
C:=N∑n=0Γ(nβ+1−α)cn. |
Now we have converted the differential equation to an algebraic equation, which can be solved in M−1 as follows:
y=IRα,β−λI∗f+CIRα,β−λI. | (5.3) |
Using what we found above (4.6) in the case m=1, we find
IRα,β−λI=∞∑i=0∞∑r=0ar((1+i)α,β)λixiα+rβ+α−1, |
which means (5.3) becomes
y(x)=∫x0∞∑i=0∞∑r=0ar((1+i)α,β)λi(x−t)iα+rβ+α−1f(t)dt+C∞∑i=0∞∑r=0ar((1+i)α,β)λixiα+rβ+α−1, |
So the solution function for the differential equation (5.1)–(5.2) can be written as
y(x)=∫x0∞∑i=0∞∑r=0ar((1+i)α,β)λi(x−t)iα+rβ+α−1f(t)dt+N∑n=0∞∑i=0∞∑r=0cnar((1+i)α,β)Γ(nβ+1−α)λixiα+rβ+α−1, |
this being not just an algebraic abstraction but an actual function in C−1. In fact, our work here has proved that this is the unique solution to (5.1)–(5.2) in the function space C−1.
Example 5.2. We now consider the following initial value problem of Caputo type, where 0<α<1 and β>0 and λ∈R are fixed while A is a function defined by (1.2) and satisfying the semigroup property (3.4).
ACDα,β0+y(x)−λy(x)=f(x),x>0; | (5.4) |
y(0)=c, | (5.5) |
where c0 is a fixed constant.
Making use of Theorem 4.2, we can rewrite this problem using the algebraic element Rα,β in the field M−1:
Rα,β∗y−λy=f+Rα,β∗[y(0)1], |
where 1 is the constant function 1∈C−1.
Now we have converted the differential equation to an algebraic equation, which can be solved in M−1 as follows:
y=IRα,β−λI∗f+Rα,βRα,β−λI∗[y(0)1]. | (5.6) |
We already know from the previous example, the m=1 case of (4.6), that
IRα,β−λI=∞∑i=0∞∑r=0ar((1+i)α,β)λixiα+rβ+α−1, |
therefore
Rα,βRα,β−λI∗[y(0)1]=(I+λIRα,β−λI)∗[y(0)1]=y(0)+λy(0)(∞∑i=0∞∑r=0ar((1+i)α,β)λixiα+rβ+α−1)∗(1)=y(0)+y(0)∞∑i=0∞∑r=0ar((1+i)α,β)λi+1iα+rβ+αxiα+rβ+α. |
Making use of the initial condition too, we now have that the solution of the initial value problem (5.4)–(5.5) is
y(x)=c+∞∑i=0∞∑r=0car((1+i)α,β)λi+1iα+rβ+αxiα+rβ+α+∫x0∞∑i=0∞∑r=0ar((1+i)α,β)λi(x−t)iα+rβ+α−1f(t)dt |
this being not just an algebraic abstraction but an actual function in C−1. In fact, our work here has proved that this is the unique solution to (5.4)–(5.5) in the function space C−1.
This work represents an extension of the theory of Mikusiński's operational calculus to a new generalised setting, namely the setting of fractional integrals and derivatives with analytic kernel functions.
The most comparable previous works are the papers [25,26], where Mikusiński's operational calculus was used to interpret fractional operators of Caputo and Riemann–Liouville types with Sonine kernels and generalised Sonine kernels. However, there are some key differences in the ideas needed. The operators with Sonine kernels do not have any "fractional order", any real or complex number that parametrises them, so the results must be interpreted purely in terms of the arbitrary kernel function. Here, we have two parameters labelling our operators, which means our results depend directly and explicitly on parameters α and β either of which can be interpreted as a fractional order. We also have a series formula, relating our generalised operators to Riemann–Liouville fractional differintegrals, which has been useful in a couple of ways. As we saw above in the proofs of Theorem 3.2 and Theorem 3.4, the series formula was instrumental in allowing us to establish the correct function space mappings for the operators. It also allowed us to obtain a more explicit form for the solutions of the associated fractional differential equations, depending on the analytic function's coefficients rather than on a general kernel function, and this explicit form may be easier to calculate numerically in specific examples. Thus, even though adding the semigroup assumption (3.4) makes the theory of fractional calculus with analytic kernels very similar to the theory of fractional calculus with Sonine kernels previously discussed in [25,26], we do obtain a different theory here with new forms of the results, owing to the series formula which is missing in the general Sonine theory.
As part of our work here, we have also highlighted a sensitive issue in defining the fractional derivatives corresponding to the integral operators with analytic kernels. The formulae written in the original paper [10] are correct but more subtle than was explained there, given the potential dependence of the analytic function coefficients on the parameters. This also indicates some limitations of our results here, as well as potential future directions of extension: we have assumed that the parameter α is in the real interval (0,1) and that the fractional integral operators have a semigroup property, in order to get an elegant form of the results, but it will also be interesting to investigate what happens if these restrictions are removed.
The authors declare there are no conflicts of interest.
[1] | K. S. Miller, B. Ross, An Introduction to the Fractional Calculus and Fractional Differential Equations, John Wiley, New York, 1993. |
[2] | K. B. Oldham, J. Spanier, The Fractional Calculus, Academic Press, New York, 1974. |
[3] | S. G. Samko, A. A. Kilbas, O. I. Marichev, Fractional Integrals and Derivatives: Theory and Applications, Gordon and Breach, Yverdon, 1993. |
[4] | A. A. Kilbas, H. M. Srivastava, J. J. Trujillo, Theory and Applications of Fractional Differential Equations, Elsevier Science B.V., Amsterdam, 2006. |
[5] | R. Hilfer, Applications of Fractional Calculus in Physics, World Scientific, Singapore, 2000. https://doi.org/10.1142/3779 |
[6] |
H. G. Sun, Y. Zhang, D. Baleanu, W. Chen, Y. Q. Chen, A new collection of real world applications of fractional calculus in science and engineering, Commun. Nonlinear Sci. Numer. Simul., 64 (2018), 213–231. https://doi.org/10.1016/j.cnsns.2018.04.019 doi: 10.1016/j.cnsns.2018.04.019
![]() |
[7] |
D. Baleanu, A. Fernandez, On fractional operators and their classifications, Mathematics, 7 (2019), 830. https://doi.org/10.3390/math7090830 doi: 10.3390/math7090830
![]() |
[8] |
A. N. Kochubei, General fractional calculus, evolution equations, and renewal processes, Integr. Equations Oper. Theory, 71 (2011), 583–600. https://doi.org/10.1007/s00020-011-1918-8 doi: 10.1007/s00020-011-1918-8
![]() |
[9] |
Y. Luchko, General fractional integrals and derivatives of arbitrary order, Symmetry, 13 (2021), 755. https://doi.org/10.3390/sym13050755 doi: 10.3390/sym13050755
![]() |
[10] |
A. Fernandez, M. A. Özarslan, D. Baleanu, On fractional calculus with general analytic kernels, Appl. Math. Comput., 354 (2019), 248–265. https://doi.org/10.1016/j.amc.2019.02.045 doi: 10.1016/j.amc.2019.02.045
![]() |
[11] |
M. Jleli, M. Kirane, B. Samet, A derivative concept with respect to an arbitrary kernel and applications to fractional calculus, Math. Methods Appl. Sci., 42 (2019), 137–160. https://doi.org/10.1002/mma.5329 doi: 10.1002/mma.5329
![]() |
[12] |
D. Zhao, M. Luo, Representations of acting processes and memory effects: General fractional derivative and its application to theory of heat conduction with finite wave speeds, Appl. Math. Comput., 346 (2019), 531–544. https://doi.org/10.1016/j.amc.2018.10.037 doi: 10.1016/j.amc.2018.10.037
![]() |
[13] |
L. A. Pipes, The operational calculus Ⅰ, J. Appl. Phys., 10 (1939), 172. https://doi.org/10.1063/1.1707292 doi: 10.1063/1.1707292
![]() |
[14] | J. Mikusiński, Operational Calculus, Pergamon Press, Oxford, 1959. |
[15] |
H. G. Flegg, Mikusinski's operational calculus, Int. J. Math. Educ. Sci. Tech., 5 (1974), 131–137. https://doi.org/10.1080/0020739740050201 doi: 10.1080/0020739740050201
![]() |
[16] |
M. Gutterman, An operational method in partial differential equations, SIAM J. Appl. Math., 17 (1969), 468–493. https://doi.org/10.1137/0117046 doi: 10.1137/0117046
![]() |
[17] | Y. Luchko, Operational method in fractional calculus, Fractional Calc. Appl. Anal., 2 (1999), 463–488. |
[18] | S. B. Hadid, Y. F. Luchko, An operational method for solving fractional differential equations of an arbitrary real order, Panam. Math. J., 6 (1996), 57–73. |
[19] | Y. Luchko, R. Gorenflo, An operational method for solving fractional differential equations, Acta Math. Vietnam., 24 (1999), 207–234. |
[20] | R. Hilfer, Y. F. Luchko, Z. Tomovski, Operational method for the solution of fractional differential equations with generalized Riemann–Liouville fractional derivatives, Fract. Calc. Appl. Anal., 12 (2009), 299–318. |
[21] | Y. Luchko, S. Yakubovich, An operational method for solving some classes of integro-differential equations, Differ. Uravn., 30 (1994), 269–280. |
[22] | S. Yakubovich, Y. Luchko, The Hypergeometric Approach to Integral Transforms and Convolutions, Kluwer Academic Publishers, Amsterdam, 1994. https://doi.org/10.1007/978-94-011-1196-6 |
[23] |
N. Rani, A. Fernandez, Mikusinski's operational calculus for Prabhakar fractional calculus, Integr. Transf. Spec. Funct., (2022), 1–21. https://doi.org/10.1080/10652469.2022.2057970 doi: 10.1080/10652469.2022.2057970
![]() |
[24] |
N. Rani, A. Fernandez, Solving Prabhakar differential equations using Mikusinski's operational calculus, Comput. Appl. Math., 41 (2022), 107. https://doi.org/10.1007/s40314-022-01794-6 doi: 10.1007/s40314-022-01794-6
![]() |
[25] |
Y. Luchko, Operational calculus for the general fractional derivative and its applications, Fract. Calculus Appl. Anal., 24 (2021), 338–375. https://doi.org/10.1515/fca-2021-0016 doi: 10.1515/fca-2021-0016
![]() |
[26] |
Y. Luchko, Fractional differential equations with the general fractional derivatives of arbitrary order in the Riemann–Liouville sense, Mathematics, 10 (2022), 849. https://doi.org/10.3390/math10060849 doi: 10.3390/math10060849
![]() |
[27] |
H. M. Fahad, A. Fernandez, Operational calculus for Riemann–Liouville fractional calculus with respect to functions and the associated fractional differential equation, Fract. Calculus Appl. Anal., 24 (2021), 518–540. https://doi.org/10.1515/fca-2021-0023 doi: 10.1515/fca-2021-0023
![]() |
[28] |
H. M. Fahad, A. Fernandez, Operational calculus for Caputo fractional calculus with respect to functions and the associated fractional differential equation, Appl. Math. Comput., 409 (2021), 126400. https://doi.org/10.1016/j.amc.2021.126400 doi: 10.1016/j.amc.2021.126400
![]() |
[29] |
A. Fernandez, D. Baleanu, H. M. Srivastava, Series representations for fractional-calculus operators involving generalised Mittag-Leffler functions, Commun. Nonlinear Sci. Numer. Simul., 67 (2019), 517–527. https://doi.org/10.1016/j.cnsns.2018.07.035 doi: 10.1016/j.cnsns.2018.07.035
![]() |
[30] | I. Podlubny, Fractional Differential Equations, Academic Press, San Diego, 1998. |
[31] | A. Fernandez, Tables of composition properties of fractional integrals and derivatives, preprint. |
[32] | I. Dimovski, Operational calculus for a class of differential operators, CR Acad. Bulg. Sci., 19 (1966), 1111–1114. |
[33] |
N. Sonine, Sur la généralisation d'une formule d'Abel, Acta Math., 4 (1884), 171–176. https://doi.org/10.1007/BF02418416 doi: 10.1007/BF02418416
![]() |
[34] | J. Wick, Über eine Integralgleichung vom Abelschen Typ, Angew. Math., 48 (1968), T39–T41. |
1. | Marc Jornet, Closed-form solution for a mathematical extension of the multi-term fractional Bateman equations via Mikusiński operational method, 2024, 139, 2190-5444, 10.1140/epjp/s13360-024-05772-1 | |
2. | Zelin Liu, Xiaobin Yu, Yajun Yin, On the Rigorous Correspondence Between Operator Fractional Powers and Fractional Derivatives via the Sonine Kernel, 2024, 8, 2504-3110, 653, 10.3390/fractalfract8110653 | |
3. | Sunday Simon Isah, Arran Fernandez, Mehmet Ali Özarslan, On bivariate fractional calculus with general univariate analytic kernels, 2023, 171, 09600779, 113495, 10.1016/j.chaos.2023.113495 | |
4. | Noosheza Rani, Arran Fernandez, Mikusiński’s operational calculus for multi-dimensional fractional operators with applications to fractional PDEs, 2024, 138, 10075704, 108249, 10.1016/j.cnsns.2024.108249 | |
5. | Arran Fernandez, Noosheza Rani, Mikusiński’s Operational Calculus for Fractional Operators with Different Kernels, 2024, 58, 24058963, 220, 10.1016/j.ifacol.2024.08.193 | |
6. | Marc Jornet, Analysis of the multi-term fractional Bateman equations in radioactive decay by means of Mikusiński algebraic calculus, 2024, 92, 05779073, 623, 10.1016/j.cjph.2024.10.002 | |
7. | Marc Jornet, Fractional Bateman equations in the Atangana-Baleanu sense, 2025, 100, 0031-8949, 025301, 10.1088/1402-4896/ada217 |