Abstract

Existence and uniqueness of the solution for a time-fractional diffusion equation with Robin boundary condition on a bounded domain with Lyapunov boundary is proved in the space of continuous functions up to boundary. Since a Green matrix of the problem is known, we may seek the solution as the linear combination of the single-layer potential, the volume potential, and the Poisson integral. Then the original problem may be reduced to a Volterra integral equation of the second kind associated with a compact operator. Classical analysis may be employed to show that the corresponding integral equation has a unique solution if the boundary data is continuous, the initial data is continuously differentiable, and the source term is HΓΆlder continuous in the spatial variable. This in turn proves that the original problem has a unique solution.

1. Introduction

In this paper, we study solvability of the time-fractional diffusion equation (TFDE)πœ•π›Όπ‘‘Ξ¦(π‘₯,𝑑)βˆ’Ξ”π‘₯Ξ¦(π‘₯,𝑑)=𝑓(π‘₯,𝑑),in𝑄𝑇],=Ω×(0,π‘‡πœ•Ξ¦(π‘₯,𝑑)πœ•π‘›(π‘₯)+𝛽(π‘₯,𝑑)Ξ¦(π‘₯,𝑑)=𝑔(π‘₯,𝑑),onΣ𝑇],=Γ×(0,𝑇Φ(π‘₯,0)=πœ“(π‘₯),π‘₯∈Ω,(1.1) where 𝑓,𝑔,πœ“ are any given functions, Ξ©βŠ‚β„π‘›, 𝑛β‰₯2, is a bounded domain with Lyapunov boundary Ξ“βˆˆπ’ž1+πœ†,  0<πœ†<1, andπœ•π›Όπ‘‘1𝑒(𝑑)=ξ‚΅dΞ“(1βˆ’π›Ό)ξ€œd𝑑𝑑0(π‘‘βˆ’πœ)βˆ’π›Όπ‘’(𝜏)dπœβˆ’π‘‘βˆ’π›Όξ‚Άπ‘’(0)(1.2) is the fractional Caputo time derivative of order 0<𝛼<1. Physically fractional diffusion equations describe anomalous diffusion on complex systems like some amorphous semiconductors or strongly porous materials (see [1] and references therein).

As to the mathematical theory of fractional diffusion equations, only the first steps have been taken. In the literature, mainly the Cauchy problems for these equations have been considered until now (see [2–5] and references therein). Existence and uniqueness of a generalized solution for an initial-boundary-value problem for the generalized time-fractional diffusion equation is proved in [6]. However, uniqueness and existence of the classical solution is given only in a special 1-dimensional case.

Our model problem is much simpler than those treated for example, in [3, 5]. However, the boundary integral approach used in this paper can be used in more general situations as well. We decided to concentrate on a simple model instead of the more general ones to clarify the basic idea. Boundary integral approach also allows us to study (TFDE) or its generalizations in weaker spaces such as 𝐿𝑝-spaces or in the scale of anisotropic Sobolev spaces.

The paper is organized as follows. In Preliminaries, we recall the definitions of the potentials and the Poisson integral. We introduce their well-known properties from theory of PDEs of parabolic type, which are needed for proving the existence and uniqueness of the solution. That is, we recall the boundary behavior of the single-layer potential. We show that the volume potential solves the nonhomogeneus TFDE with the zero initial condition. Moreover, we prove that the Poisson integral solves the homogeneous TFDE with a given initial datum. The final section is dedicated to the proof of existence and uniqueness of the solution.

2. Preliminaries

Here we recall the potentials and the Poisson integral and their basic properties. In the sequel we shall assume that the functions appearing in the definitions are smooth enough such that the corresponding integrals exist.

The single-layer potential can be defined as(ξ€œπ‘†πœ‘)(π‘₯,𝑑)=𝑑0ξ€œΞ“πœ•π‘›(𝑦)𝐺(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)πœ‘(𝑦,𝜏)d𝜎(𝑦)d𝜏,π‘₯∈Ω,(2.1) where 𝑛(𝑦) denotes the outward unit normal at π‘¦βˆˆΞ“ and⎧βŽͺ⎨βŽͺβŽ©πœ‹πΊ(π‘₯,𝑑)=βˆ’π‘›/2π‘‘π›Όβˆ’1|π‘₯|βˆ’π‘›π»201214|π‘₯|2π‘‘βˆ’π›Όβˆ£ξƒ­(𝛼,𝛼)(𝑛/2,1),(1,1),π‘₯βˆˆβ„π‘›,𝑑>0,0,π‘₯βˆˆβ„π‘›,𝑑<0,(2.2) is the fundamental solution of (TFDE) [3, 7–9]. Here 𝐻2012 is the Fox 𝐻-function, which is defined via Mellin-Barnes integral representation𝐻2012(𝑧)∢=𝐻2012ξ‚Έξ‚Ή=1π‘§βˆ£(𝛼,𝛼)(𝑛/2,1),(1,1)ξ€œ2πœ‹iπ’žΞ“(𝑛/2+𝑠)Ξ“(1+𝑠)𝑧Γ(𝛼+𝛼𝑠)βˆ’π‘ d𝑠,(2.3) where π’ž is an infinite contour on the complex plane circulating the negative real axis counterclockwise.

The volume potential is defined by(ξ€œπ‘‰πœ‘)(π‘₯,𝑑)=𝑑0ξ€œΞ©πΊ(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)πœ‘(𝑦,𝜏)d𝑦d𝜏,π‘₯∈Ω(2.4) for πœ‘ such that suppπœ‘(β‹…,𝑑)βŠ‚Ξ© for any π‘‘βˆˆ(0,𝑇].

The Poisson integral is defined as(ξ€œπ‘ƒπœ‘)(π‘₯,𝑑)=𝑑0ξ€œΞ©0𝐸(π‘₯βˆ’π‘¦,𝑑)πœ‘(𝑦)d𝑦,π‘₯∈Ω,(2.5) where Ξ©0 is some neighborhood of Ξ© and𝐸(π‘₯,𝑑)=πœ‹βˆ’π‘›/2|π‘₯|βˆ’π‘›ξ‚π»2012ξ‚€14|π‘₯|2π‘‘βˆ’π›Όξ‚(2.6) with𝐻2012(𝑧)=𝐻2012ξ‚Έξ‚Ή=1π‘§βˆ£(1,𝛼)(𝑛/2,1),(1,1)ξ€œ2πœ‹iπ’žΞ“(𝑛/2+𝑠)Ξ“(1+𝑠)𝑧Γ(1+𝛼𝑠)βˆ’π‘ d𝑠(2.7) and π’ž as in the definition of 𝐻2012.

Note that in contrast to classical parabolic partial differential equations, we have a Green matrix {𝐸(π‘₯,𝑑),𝐺(π‘₯,𝑑)} instead of one fundamental solution. We also emphasize that the Green's functions have singularities both in time and spatial variable unlike in the case of classical parabolic PDEs, where singularity occurs only in time.

Let us now state the basic properties of the aforementioned quantities. Since the proofs are strongly based on the detailed analysis of the Fox 𝐻-functions, we shall recall their basic properties. For further details of these functions, we refer to [3, 10, 11].

In order to simplify the notations, we introduce the following function defined for 𝑧>0:𝐻(𝑝)(1𝑧)=ξ€œ2πœ‹iπ’žΞ“(𝑛/2+𝑠)Ξ“(1+𝑠)𝑝Γ(𝛼+𝛼𝑠)Ξ“(𝑠)π‘βˆ’1π‘§βˆ’π‘ d𝑠,𝑝=1,2.(2.8)

Note that, in particular, 𝐻(1)(𝑧)=𝐻2012(𝑧). The following properties of 𝐻(𝑝) are needed.

Lemma 2.1. For the functions 𝐻(𝑝), there holds (i)differentiation formula (d/d𝑧)𝐻(1)(𝑧)=βˆ’π‘§βˆ’1𝐻(2)(z), (ii)the asymptotic behaviour at infinity, ||𝐻(𝑝)||(𝑧)≀𝐢𝑧𝑛/2ξ€·expβˆ’πœŽπ‘§1/(2βˆ’π›Ό)ξ€Έ,𝜎∢=𝛼𝛼/2βˆ’π›Ό(2βˆ’π›Ό),(2.9) for 𝑝=1,2 and 𝑧β‰₯1,(iii)the asymptotic behaviour near zero ||𝐻(𝑝)||⎧βŽͺ⎨βŽͺβŽ©π‘§(𝑧)≀𝐢𝑛/2𝑧if𝑛=2or𝑛=3,2||||𝑧logzif𝑛=4,2if𝑛>4,(2.10) for 𝑝=1,2 and 𝑧≀1. The constants in (𝑖𝑖) and (𝑖𝑖𝑖) can depend on 𝑛, 𝑝, and 𝛼.

Proof. The proofs follow from the Mellin-Barnes integral representation and the analyticity of the functions 𝐻(𝑝) [3, 10, 12].

Remark 2.2. Above and in the sequel, 𝐢 denotes a generic constant, which may depend on various quantities. The only thing that matters is that in our calculations 𝐢 will be independent of π‘₯ and 𝑑.

Let us now concentrate on the properties of the potentials. We start with the single-layer potential π‘†πœ‘. First of all, standard calculations show that π‘†πœ‘ solves the equation (πœ•π›Όπ‘‘βˆ’Ξ”π‘₯)𝑒=0. Moreover, we need to know the boundary behavior of the single-layer potential, which is given in the following result.

Theorem 2.3. Let πœ‘βˆˆπ’ž(Σ𝑇). The single-layer potential defined by (2.1) is continuous in Ω𝑇 with the zero initial value. Moreover, for π‘₯∈Ω and π‘₯0βˆˆΞ“, βˆ‡π‘₯(π‘†πœ‘)(π‘₯,𝑑)⋅𝑛(π‘₯0) has the following limiting value: πœ•π‘›(π‘₯0)ξ€·π‘₯π‘†πœ‘0ξ€Έ,π‘‘βˆΆ=limπ‘₯β†’π‘₯0βˆ‡π‘₯ξ€·π‘₯(π‘†πœ‘)(π‘₯,𝑑)⋅𝑛0ξ€Έ=12πœ‘ξ€·π‘₯0ξ€Έ+ξ€œ,𝑑𝑑0ξ€œΞ“πœ•π‘›(π‘₯0)𝐺π‘₯0ξ€Έπœ‘βˆ’π‘¦,π‘‘βˆ’πœ(𝑦,𝜏)d𝜎(𝑦)d𝜏(2.11) as π‘₯ tends to π‘₯0 nontangentially.

Proof. The proof follows the same lines as in the case of the single-layer potential for the heat equation [13, Chapter 5.2] and is based on a detailed analysis of the kernel 𝐺. Since the proof is rather lengthy, we give only the reference [12, Theorems 1 and 2].

For the volume potential, we have the following result.

Theorem 2.4. Let π‘“βˆˆπ’ž(Σ𝑇) such that 𝑓(β‹…,𝑑) is HΓΆlder continuous uniformly in π‘‘βˆˆ[0,𝑇] and supp𝑓(β‹…,𝑑)βŠ‚Ξ©, π‘‘βˆˆ[0,𝑇]. Then the volume potential 𝑉𝑓 with 𝑉 defined by (2.4) solves πœ•π›Όπ‘‘π‘’βˆ’Ξ”π‘₯𝑒=𝑓 with the zero initial condition.

Proof. The zero initial condition follows since 𝐺 is locally integrable. Indeed, we split the integral 𝑉𝑓 into two parts 𝐼1+𝐼2 depending on whether 𝑧=(1/4)(π‘‘βˆ’πœ)βˆ’π›Ό|π‘₯βˆ’π‘¦|2β‰₯1 or 𝑧≀1.
If 𝑧β‰₯1, we use the fact that 𝑧𝛾exp(βˆ’πœŽπ‘§1/(2βˆ’π›Ό)) is uniformly bounded for any 𝛾,𝜎>0. Then Lemma 2.1 together with the definition of 𝐺 yields ||||𝐺(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)≀𝐢(π‘‘βˆ’πœ)𝛼+π›Όπ›Ύβˆ’1βˆ’π›Όπ‘›/2||||π‘₯βˆ’π‘¦βˆ’2𝛾.(2.12) If we choose 𝑛/2βˆ’1<𝛾<𝑛/2, we see that lim𝑑→0+𝐼1(π‘₯,𝑑)=0.
On the other hand, if 𝑧≀1, then Lemma 2.1 yields ||||⎧βŽͺ⎨βŽͺ⎩𝐺(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)≀𝐢(π‘‘βˆ’πœ)(2βˆ’π‘›)𝛼/2,𝑛=2,3,(π‘‘βˆ’πœ)βˆ’π›Όβˆ’1ξ‚€|||ξ‚€||||logπ‘₯βˆ’π‘¦2(π‘‘βˆ’πœ)βˆ’π›Όξ‚|||+1,𝑛=4,(π‘‘βˆ’πœ)βˆ’π›Όβˆ’1||||π‘₯βˆ’π‘¦4βˆ’π‘›,𝑛>4.(2.13)
Then lim𝑑→0+𝐼2(π‘₯,𝑑)=0 follows immediately for 𝑛=2,3. If 𝑛=4, we may use the fact that 𝑧𝛾|log𝑧| is bounded in (0,1] for any 𝛾>0. If 𝑛>4, we use π‘§βˆ’π›Ύβ‰₯1 for any 𝛾>0. In the preceeding two cases, we obtain ||||𝐺(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)≀𝐢(π‘‘βˆ’πœ)π›Όπ›Ύβˆ’π›Όβˆ’1||||π‘₯βˆ’π‘¦4βˆ’π‘›βˆ’2𝛾.(2.14) If we choose 1<𝛾<2, we see that lim𝑑→0+𝐼2(π‘₯,𝑑)=0.
For the proof of the first claim, we refer to [3, Sections 5.2 and 5.3], where the proof is given in a much more general case of a time-fractional diffusion equation.

Finally, for the Poisson integral there holds the following theorem.

Theorem 2.5. Let πœ“ be a continuous function in Ξ©0. Then the Poisson integral π‘ƒπœ“ with 𝑃 defined by (2.5) solves πœ•π›Όπ‘‘π‘’βˆ’Ξ”π‘₯𝑒=0 with 𝑒(π‘₯,0)=πœ“(π‘₯), π‘₯∈Ω.

Proof. The fact that π‘ƒπœ“ solves πœ•π›Όπ‘‘π‘’βˆ’Ξ”π‘₯𝑒=0 follows from the calculations given in [8]. Note that differentiation inside the integral is allowed because there is no singularity in 𝑑. Therefore, it remains to prove the initial condition.
We proceed as in [13, Proof of Theorem 1.2.1] and consider first the case of constant πœ“. The integral is divided into two parts (π‘ƒπœ“)(π‘₯,𝑑)=𝐼𝑅+𝐼𝑐𝑅, where 𝐼𝑅 is the integral over the ball 𝐡(π‘₯,𝑅) with 𝑅 being so small that 𝐡(π‘₯,𝑅) is contained in Ξ©0 and 𝐼𝑐𝑅 denotes its complementary part. Since 𝑅 is fixed and there is no singularity in the spatial variable, the asymptotic behavior of 𝐺 shows that lim𝑑→0+𝐼𝑐𝑅=0. We need to prove that lim𝑑→0+𝐼𝑅=πœ“.
Introducing spherical coordinates, we get 𝐼𝑅=πœ‹βˆ’π‘›/2πœ”π‘›πœ“ξ€œπ‘…0π‘Ÿβˆ’1𝐻2012ξ‚€14π‘‘βˆ’π›Όπ‘Ÿ2dπ‘Ÿ=πœ‹βˆ’π‘›/2πœ”π‘›πœ“ξ€œπ‘…/2𝑑𝛼/20π‘Ÿβˆ’1𝐻2012ξ€·π‘Ÿ2ξ€Έdπ‘ŸβŸΆπœ‹βˆ’π‘›/2πœ”π‘›πœ“ξ€œβˆž0π‘Ÿβˆ’1𝐻2012ξ€·π‘Ÿ2ξ€Έdπ‘Ÿ,π‘‘βŸΆ0+,(2.15) where πœ”π‘› denotes the surface area of the unit sphere in ℝ𝑛.
To evaluate the last integral denoted by 𝐼, we note that the asymptotic behavior of the integrand guarantees the absolute integrability. Through the change of variables π‘Ÿ2=𝑑, we see that the integral is nothing but half of the Mellin transform of 𝐻2012, 1𝐼=2ℳ𝐻2012(𝑠)=Ξ“(𝑛/2+𝑠)Ξ“(1+𝑠),2Ξ“(1+𝛼𝑠)(2.16) evaluated at the point 𝑠=0.
Therefore, we may conclude that the claim in the case πœ“ is constant. In the case of general πœ“ we may proceed as in [13, Proof of Theorem 1.2.1.]

3. Existence and Uniqueness of the Solution

As it was mentioned in Introduction, we seek the solution in a form of𝑒(π‘₯,𝑑)=(π‘†πœ‘)(π‘₯,𝑑)+(π‘ƒπœ“)(π‘₯,𝑑)+(𝑉𝑓)(π‘₯,𝑑),(3.1) where πœ‘ is to be determined. The density πœ‘ is determined by reducing the original problem to a corresponding integral equation.

We assume that 𝛽 is a continuous function on Σ𝑇. We need to calculate the normal derivative of π‘†πœ‘, 𝑉𝑓, and π‘ƒπœ“. For π‘ƒπœ“, we observe that differentiation inside the integral defining 𝑃 is allowed since there is no singularity in 𝑑. For 𝑉𝑓, the differentiation inside the integral is justified by the calculations given in [8]. Finally, Theorem 2.3 gives the boundary value for the normal derivative of π‘†πœ‘. Then the Robin boundary condition is equivalent withξ‚€12𝐼+π‘Š+π›½π‘†πœ‘=𝐹,(3.2) where(ξ€œπ‘Šπœ‘)(π‘₯,𝑑)=𝑑0ξ€œΞ“πœ•π‘›(π‘₯)𝐺(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)πœ‘(𝑦,𝜏)d𝜎(𝑦)d𝜏(3.3) is the integral in Theorem 2.3 andξ€œπΉ(π‘₯,𝑑)=𝑔(π‘₯,𝑑)βˆ’Ξ©0πœ•π‘›(π‘₯)βˆ’ξ€œπΈ(π‘₯βˆ’π‘¦,𝑑)πœ“(𝑦)d𝑦𝑑0ξ€œΞ©πœ•π‘›(π‘₯)𝐺(π‘₯βˆ’π‘¦,π‘‘βˆ’πœ)𝑓(𝑦,𝜏)d𝑦dπœβˆ’π›½(π‘₯,𝑑)(π‘ƒπœ“)(π‘₯,𝑑)βˆ’π›½(π‘₯,𝑑)(𝑉𝑓)(π‘₯,𝑑).(3.4)

We will prove that (3.2) admits a unique solution for any bounded function 𝐹. Therefore, it is needed to determine the conditions, which guarantee boundedness.

For the second integral on the right-hand side of (3.4), we use the following result.

Lemma 3.1. Let 𝑧=(1/4)|π‘₯βˆ’π‘¦|2π‘‘βˆ’π›Ό with π‘₯βˆˆΞ“ and π‘¦βˆˆΞ©. The following estimates for the normal derivative of 𝐺 hold: (1)if 𝑧β‰₯1, then ||πœ•π‘›(π‘₯)||𝐺(π‘₯βˆ’π‘¦,𝑑)β‰€πΆπ‘‘βˆ’π›Όπ‘›/2βˆ’1||||π‘₯βˆ’π‘¦expβˆ’πœŽπ‘‘βˆ’π›Ό/(2βˆ’π›Ό)||||π‘₯βˆ’π‘¦2/(2βˆ’π›Ό);(3.5)(2)if 𝑧≀1, then ||πœ•π‘›(π‘₯)||𝑑𝐺(π‘₯βˆ’π‘¦,𝑑)β‰€πΆβˆ’π›Όβˆ’1|||||||ξ‚€||||π‘₯βˆ’π‘¦logπ‘₯βˆ’π‘¦2π‘‘βˆ’π›Όξ‚|||𝑑ifn=2,βˆ’π›Όβˆ’1||||π‘₯βˆ’π‘¦3βˆ’π‘›ifnβ‰₯3.(3.6)

Proof. Applying the differentiation formula in Lemma 2.1, we get πœ•π‘›(π‘₯)𝐺(π‘₯βˆ’π‘¦,𝑑)=βˆ’πœ‹βˆ’π‘›/2⟨π‘₯βˆ’π‘¦,𝑛(π‘₯)⟩||||π‘₯βˆ’π‘¦π‘›+2π‘‘π›Όβˆ’1𝑛𝐻(1)(𝑧)+2𝐻(2)ξ€Ύ,(𝑧)(3.7) where 𝑧=(1/4)|π‘₯βˆ’π‘¦|2π‘‘βˆ’π›Ό and βŸ¨β‹…,β‹…βŸ© denotes the inner product in ℝ𝑛.
Using the definition of 𝐻(𝑝) and the property Ξ“(𝑧+1)=𝑧Γ(𝑧) of the Gamma function, it follows that the Mellin transform of 𝑛𝐻(1)+2𝐻(2) is ℳ𝑛𝐻(1)+2𝐻(2)ξ€Έ1(𝑠)=2Ξ“((𝑛+2)/2+𝑠)Ξ“(1+𝑠),Ξ“(𝛼+𝛼𝑠)(3.8) which is nothing but half of the Mellin transform of 𝐻(1) with 𝑛 replaced by 𝑛+2.
Using the estimate (2.9) of Lemma 2.1 with 𝑛+2 instead of 𝑛, we obtain the first estimate for 𝑧β‰₯1.
If 𝑧≀1, we use the estimate (2.10) of Lemma 2.1 with 𝑛+2=4 and 𝑛+2>4 to obtain the second estimate.

Let us return to the estimation of the second integral on r.h.s. of (3.4). Once again we split the integral into two parts 𝐼1 and 𝐼2 depending on whether 𝑧β‰₯1 or 𝑧≀1 with 𝑧=(1/4)|π‘₯βˆ’π‘¦|2(π‘‘βˆ’πœ)βˆ’π›Ό. If 𝑓 is a bounded function and 𝑧β‰₯1, there holds||𝐼1||β‰€πΆβ€–π‘“β€–πΏβˆž(Ω𝑇)ξ€œπ‘‘0(π‘‘βˆ’πœ)𝛼/2βˆ’1ξ€œd𝜏∞4π‘Ÿπ‘›ξ€·expβˆ’πœŽπ‘Ÿ2/(2βˆ’π›Ό)ξ€Έdπ‘Ÿβ‰€πΆβ€–π‘“β€–πΏβˆž(Ω𝑇),(3.9) where we have used the spherical coordinates with π‘Ÿ=(π‘‘βˆ’πœ)βˆ’π›Ό/2|π‘₯βˆ’π‘¦|.

If 𝑧≀1, we have to consider different cases of 𝑛's separately. As an example, let us consider case 𝑛=3. We have||𝐼2||β‰€πΆβ€–π‘“β€–πΏβˆž(Ω𝑇)ξ€œΞ©Γ—(0,𝑑)∩{z≀1}(π‘‘βˆ’πœ)π›Όπ›Ύβˆ’π›Όβˆ’1||||π‘₯βˆ’π‘¦βˆ’2𝛾d𝑦d𝜏,(3.10) where the fact π‘§βˆ’π›Ύβ‰₯1 for any 𝛾>0 is used. Choosing 1<𝛾<3/2, we see that 𝐼2 is bounded.

Using the estimates (2.12) and (2.13) in the proof of Theorem 2.4, we see that 𝑉𝑓 is bounded as well.

For the first integral on the right-hand side, we use the following result [3, Proposition 1].

Lemma 3.2. Let 𝑧=(1/4)|π‘₯|2π‘‘βˆ’π›Ό. For 𝐸 there holds the following: (1)if 𝑧β‰₯1, then ||βˆ‡π‘₯||𝐸(π‘₯,𝑑)β‰€πΆπ‘‘βˆ’π›Ό(𝑛+1)/2ξ€½expβˆ’πœŽπ‘‘βˆ’π›Ό/(2βˆ’π›Ό)|π‘₯|2/(2βˆ’π›Ό)ξ€Ύ;(3.11)(2)if 𝑧≀1, then ||βˆ‡π‘₯||𝐸(π‘₯,𝑑)β‰€πΆπ‘‘βˆ’π›Ό|π‘₯|βˆ’π‘›+1.(3.12)

We split the first integral on right-hand side of (3.4) into two parts 𝐼1 and 𝐼2 depending whether 𝑧=(1/4)|π‘₯βˆ’π‘¦|2π‘‘βˆ’π›Όβ‰₯1 or 𝑧≀1. If πœ“ is a bounded function, then using Lemma 3.2 we have||𝐼1||β‰€πΆπ‘‘βˆ’π›Ό+π›Όπ›Ύβ€–πœ“β€–πΏβˆž(Ξ©0)(3.13) for any 𝛾>0, since 𝑧↦𝑧𝛾exp(βˆ’πœŽπ‘§π›½) is uniformly bounded on [1,∞) for any 𝛽,𝛾,𝜎>0. Similarly, for 𝐼2 there holds||𝐼2||β‰€πΆβ€–πœ“β€–πΏβˆž(Ξ©0)ξ€œΞ©0∩{𝑧≀1}π‘‘βˆ’π›Ό+𝛼𝛾||||π‘₯βˆ’π‘¦βˆ’π‘›+1βˆ’2𝛾d𝑦(3.14) for any 𝛾>0, since π‘§βˆ’π›Ύβ‰₯1 for any 𝛾>0. Choosing 𝛾<1/2, we have||𝐼2||β‰€πΆπ‘‘βˆ’π›Ό+π›Όπ›Ύβ€–πœ“β€–πΏβˆž(Ξ©0).(3.15)

We see that 𝐼2 blows up as 𝑑→0+. Therefore, we have to assume more smoothness on πœ“ to guarantee boundedness. Assume that πœ“ is a continuously differentiable function in Ξ©0. Then integration by parts yields a better kernel 𝐸. Asymptotic behavior of 𝐸 guarantees that the resulting integral is uniformly bounded on π‘₯,𝑑 (see [3, Proposition 1]). We have||||ξ€œΞ©0πœ•π‘›(π‘₯)||||𝐸(π‘₯βˆ’π‘¦,𝑑)πœ“(𝑦)dπ‘¦β‰€πΆβ€–πœ“β€–πΏβˆž(Ξ©0)+β€–βˆ‡πœ“β€–πΏβˆž(Ξ©0).(3.16)

The same reason as above implies that π‘ƒπœ“ is bounded.

Now we are ready to prove that (3.2) has a unique solution.

Theorem 3.3. Let π‘“βˆˆπΏβˆž(Ω𝑇), π‘”βˆˆπ’ž(Σ𝑇), and πœ“βˆˆπ’ž1(Ξ©0). Then the boundary integral equation (3.2) admits a unique bounded, continuous solution πœ‘.

Proof. Using similar estimates as in Lemma 3.1 and in the proof of Theorem 2.4, we see that π‘Š+𝛽𝑆 is an integral operator with a weakly singular kernel. Note that the estimates for the normal derivative given in Lemma 3.1 can be multiplied by |π‘₯βˆ’π‘¦|πœ† in the estimates for π‘Š due to the Lyapunov smoothness of the boundary Ξ“. For details we refer to [12].
We conclude that π‘Š+𝛽𝑆 is a compact operator in π’ž(Σ𝑇) [14, Theorem 2.22]. Moreover, similarly as in [15] we can prove that there exists an integer π‘˜0 such that β€–(2π‘Š+2𝛽𝑆)π‘˜0π‘™πœ‘β€–πΏβˆž(Σ𝑇)≀(𝑀𝑇)𝑙𝑙!β€–πœ‘β€–πΏβˆž(Σ𝑇)(3.17) for some constant 𝑀 and for all π‘™βˆˆβ„•. This, in particular, implies that the homogeneous equation ((1/2)𝐼+π‘Š+𝛽𝑆)πœ‘=0 has a unique solution. Moreover, (2π‘Š+2𝛽𝑆)π‘˜0𝑙 is a contraction for some π‘™βˆˆβ„•. Therefore, (1/2)𝐼+π‘Š+𝛽𝑆 is invertible and the inverse is given by the Neumann series ((1/2)𝐼+π‘Š+𝛽𝑆)βˆ’1=2βˆžξ“π‘˜=0(βˆ’1)π‘˜(2π‘Š+2𝛽𝑆)π‘˜.(3.18)
Since the series is uniformly convergent, we have β€–πœ‘β€–πΏβˆž(Σ𝑇)β‰€πΆβ€–πΉβ€–πΏβˆž(Σ𝑇),(3.19) and continuity of πœ‘ follows from that of 𝐹.

In conclusion, 𝑒 defined by (3.1) solves (TFDE) provided πœ‘ solves (3.2). Combining our results with the results in [3, 8], we have proved our main result, which is stated as follows.

Theorem 3.4. Let π‘”βˆˆπ’ž(Σ𝑇), πœ“βˆˆπ’ž1(Ξ©0), and π‘“βˆˆπ’ž(Σ𝑇) such that 𝑓(β‹…,𝑑) is HΓΆlder continuous uniformly in π‘‘βˆˆ[0,𝑇] and supp𝑓(β‹…,𝑑)βŠ‚Ξ©, π‘‘βˆˆ[0,𝑇]. Then (TFDE) admits a unique classical solution and the solution depends continuously on the data in the following sense: ‖𝑒(π‘₯,𝑑)β€–π’ž(Ω𝑇)ξ‚€β‰€πΆβ€–π‘“β€–π’ž(Ω𝑇)+β€–π‘”β€–π’ž(Σ𝑇)+β€–πœ“β€–π’ž1(Ξ©0).(3.20)

If πœ“ has compact support in Ξ©, we may relax the smoothness assumption on πœ“ and proof of Theorem 3.4 implies the following.

Corollary 3.5. Let 𝑔 and 𝑓 satisfy the assumptions in Theorem 3.4, and let πœ“βˆˆπ’ž(Ξ©) with compact support. Then (TFDE) admits a unique classical solution and the solution depends continuously on the data in the following sense: ‖𝑒(π‘₯,𝑑)β€–π’ž(Ω𝑇)ξ‚€β‰€πΆβ€–π‘“β€–π’ž(Ω𝑇)+β€–π‘”β€–π’ž(Σ𝑇)+β€–πœ“β€–π’ž(Ξ©).(3.21)

Proof. All the arguments are the same as in Theorem 3.4 except now we can choose 𝛾=1 in the estimates for 𝐼1 and 𝐼2 of ξ€œΞ©πœ•π‘›(π‘₯)𝐸(π‘₯βˆ’π‘¦,𝑑)πœ“(𝑦)d𝑦=𝐼1+𝐼2(3.22) in Theorem 3.4. Therefore, we obtain ||||ξ€œΞ©πœ•π‘›(π‘₯)||||𝐸(π‘₯βˆ’π‘¦,𝑑)πœ“(𝑦)d𝑦≀𝐢1+dist(suppπœ“,Ξ“)βˆ’π‘›βˆ’1ξ€Έβ€–πœ“β€–πΏβˆž(Ξ©)(3.23) and the claim follows.

Remark 3.6. The estimates given for 𝐹 reveal that if πœ“ is merely continuous, we have |𝐹(π‘₯,𝑑)|β‰€πΆπ‘‘βˆ’π›½ for some βˆ’1<𝛽<βˆ’π›Ό/2. Then the same technique as in [13, Section 5.3] may be employed to prove that (TFDE) with the initial condition replaced by 𝑒(π‘₯,0)=πœ“(π‘₯), π‘₯∈Ω, has a unique solution, which may not be even continuous and can be unbounded near Γ×{0} and therefore is not a classical solution.

Remark 3.7. The same technique as above may be used for more general time-fractional diffusion equations, where βˆ’Ξ” is replaced by a uniformly elliptic second-order differential operator in nondivergence form with bounded continuous real-valued coefficients depending on π‘₯.

Remark 3.8. In [15], we have proved existence and uniqueness of the solution of TFDE with the zero initial condition and the zero source term with Dirichlet boundary condition. Using the same technique as above, we may also consider the case of nonzero initial condition and nontrivial source term. Indeed, use of the double-layer ansatz leads to a Volterra integral equation of the second kind as in this paper. Then, using the same arguments as above, we can prove uniqueness and existence of a classical solution without any restrictions on 𝑛 or on boundary conditions such as in [6].

Acknowledgment

The author would like to thank the referee's suggestions for the improvement of this paper.