Brought to you by:
Paper

On existence and regularity of a terminal value problem for the time fractional diffusion equation

, , and

Published 9 April 2020 © 2020 IOP Publishing Ltd
, , Citation Nguyen Huy Tuan et al 2020 Inverse Problems 36 055011 DOI 10.1088/1361-6420/ab730d

0266-5611/36/5/055011

Abstract

In this paper we consider a final value problem for a diffusion equation with time-space fractional differentiation on a bounded domain D of ${\mathbb{R}}^{k}$, k ≥ 1, which includes the fractional power ${\mathcal{L}}^{\beta }$, 0 < β ≤ 1, of a symmetric uniformly elliptic operator $\mathcal{L}$ defined on L2(D). A representation of solutions is given by using the Laplace transform and the spectrum of ${\mathcal{L}}^{\beta }$. We establish some existence and regularity results for our problem in both the linear and nonlinear case.

Export citation and abstract BibTeX RIS

1. Introduction

Nonlinear diffusion equations, an important class of parabolic equations, come from many diffuse phenomena that appear widely in nature. They are proposed as mathematical models of physical problems in many areas, such as filtering, phase transition, biochemistry and dynamics of biological groups. Many new ideas and methods have been developed to consider some various kinds of nonlinear diffusion equations. We list some selected works in recent time, for example Caffarelli et al [1], Duzaar et al [24], Vazquez et al [58] and the references therein.

We present existence and regularity estimates for the solution to a final boundary value problem for a space-time fractional diffusion equation. Let D be an open and bounded domain in ${\mathbb{R}}^{k},\left(k\ge 1\right)$ with boundary ∂D. Given 0 < α < 1 and 0 < β ≤ 1, a forcing (or source) function F, we consider the final value problem for the time fractional diffusion equation

Equation (1.1)

with the boundary condition

Equation (1.2)

and the final condition

Equation (1.3)

where φ is a given function. Here J is the interval (0, T ), the notation ${}^{\mathrm{c}}D_{t}^{\alpha }$ for 0 < α < 1 represents the left Caputo fractional derivative of order α which is defined by

provided that ${I}_{t}^{\alpha }v\left(t\right){:=}{g}_{\alpha }\left(t\right)\star v\left(t\right)$, here ${g}_{\alpha }\left(t\right)=\frac{1}{{\Gamma}\left(\alpha \right)}{t}^{\alpha -1}$, t > 0, ⋆ denotes the convolution. For α = 1, we consider the usual time derivative $\frac{\partial u}{\partial t}$. The fractional power ${\mathcal{L}}^{\beta }\hspace{0.17em}0{< }\beta \le 1$ of the Laplacian operator $\mathcal{L}$ on D is defined by its spectrum. The symmetric uniformly elliptic operator is defined on the space L2(D) by

provided that ${\mathcal{L}}_{ij}\in {C}^{1}\left(\overline{{\Omega}}\right)$, $b\in C\left(\overline{{\Omega}}\right)$, b(x) ≥ 0 for all $x\in \overline{{\Omega}}$, ${\mathcal{L}}_{ij}={\mathcal{L}}_{ji},1\le i,\;j\le k$, and ${\xi }^{T}\left[{\mathcal{L}}_{ij}\left(x\right)\right]\xi \ge {L}_{0}{\vert \xi \vert }^{2}$ for some L0 > 0, $x\in \overline{{\Omega}}$, $\xi =\left({\xi }_{1},{\xi }_{2},\dots ,{\xi }_{k}\right)\in {\mathbb{R}}^{k}$. The equation (1.1) is equipped with $\mathcal{H}v=v\;\mathrm{o}\mathrm{r}\;\mathcal{H}v=\left(\to {L}\nabla v\right).\to {n}$, where $\to {L}={\left[{\mathcal{L}}_{ij}\left(x\right)\right]}_{i,j=1}^{k}$ is a k × k matrix and n is the outer normal vector of ∂D. Then the operator $\mathcal{L}$ is self-adjoint under this impedance boundary condition.

The time fractional reaction diffusion equation arises in describing 'memory' occurring in physics such as plasma turbulence [9] and it was introduced by Nigmatullin [10] to describe diffusion in media with fractal geometry, which is a special type of porous media and is applied in the flow in highly heterogeneous aquifer [11] and single-molecular protein dynamics [12]. In a physical model presented in [13], the fractional diffusion corresponds to a diverging jump length variance in the random walk, and a fractional time derivative arises when the characteristic waiting time diverges.

If the final condition (1.3) is replaced by the initial condition

Equation (1.4)

then problem (1.1), (1.2), (1.4) is called a forward problem (or an initial value problem) for time-space fractional diffusion equations; for applications of this type of equation see [14] and for the abstract form of (1.1)–(1.4) see [15]. Carvalho et al [16] established a local theory of mild solutions for problem (1.1)–(1.4) where ${\mathcal{L}}^{\beta }$ is a sectorial (nonpositive) operator. Guswanto [17] studied the existence and uniqueness of a local mild solution for a class of initial value problems for nonlinear fractional evolution equations and the study of existence of initial value problems was considered by Warma et al [14]. A significant number of papers wew devoted to extend properties holding in the standard setting to the fractional one (see for example [18, 19, 2022]).

Numerical approximation for solutions for problem (1.1)–(1.4) was studied by Jin et al [23, 24] and for other works on fractional diffusion see [2528]. However, the literature on regularity of the initial value problem for fractional diffusion-wave equations is scarce; for the linear case see [26, 2931], and for the nonlinear case see [14, 3234]. Although there are many works on the direct problem, but the results on inverse problem for fractional diffusion are scarce. We can list some papers of Yamamoto and his group see [3537, 38, 39, 40], of Kaltenbacher et al [41, 42], of Rundell et al [43, 44], of Janno see [45, 46], etc.

In practice, initial data of some problems may not be known since many phenomena cannot be measured at the initial time. Phenomena can be observed at a final time t = T, such as, in the image processing area. A picture is not processed at the capturing time t = 0. Instead, one wishes to recover the original information of the picture from its blurry form. Hence, inverse problems or terminal value problems or final value problems (IPs/FVPs), i.e., the fractional differential equations (FDEs) equipped with final value data, have been considered. IPs/FVPs are important in engineering in detecting the previous status of physical fields from its present information. If F = 0 in (1.1), Yamamoto et al [31] showed that problem (1.1)–(1.3) has a unique weak solution when φH2(Ω). If b = 0 and F(u(x, t)) = F(x, t), Tuan et al [47] showed that problem (1.1)–(1.3) has a unique weak solution when φH2(Ω) and FL(0, T ; H2(Ω)), and other works on the homogeneous case for problem (1.1)–(1.3) can be found in [25, 4749]. Wei and her group [6252] studied some regularization methods for homogeneous backward problem and Yamamoto et al [53] considered a backward problem in time for a time-fractional partial differential equation in the one-dimensional case. When α = 1, systems (1.1)–(1.3) are reduced to the backward problem for classical reaction diffusion equations, and were studied in [5456].

To the best of the authors' knowledge this is the first paper that analyzes problem (1.1)–(1.3). We present existence and uniqueness results and derive regularity estimates both in time and space. In what follows, we analyze the difficulties of this problem. By letting$u\left(t,\cdot \right)=\mathcal{O}\left(t\right)\left(\hspace{0.17em}f,\varphi \right)$, the solution operator $\mathcal{O}\left(0\right)$ is really not bounded in L2(D). Hence continuity of mild solutions does not hold at the initial time t = 0. In addition, since the fractional derivative is non-locally defined, if we put v(t) = u(Tt) then ${\left.{}^{\mathrm{c}}D_{t}^{\alpha }u\left(s\right)\right\vert }_{s=T-t}$ does not equal ${}^{\mathrm{c}}D_{t}^{\alpha }v\left(t\right)$, so the problem cannot be changed to an initial value problem. As a result we need new techniques to deal with the FVP (1.1)–(1.3). To the best of our knowledge, the work on the final value problem is still limited.

Our main results in this paper can be split into two parts, linear and nonlinear source functions. Linear models are sometimes good approximations of the real problems under consideration and provides mathematical tools needed to study nonlinear phenomena, especially for semi-linear and quasi-linear equations. In part 1, we consider the regularity property of the solution in the linear case F. We seek to address the following question: if the data is regular, how regular is the solution? Our task in this part is to find a suitable Banach space for the given data (φ, F ) in order to obtain regularity results for the corresponding solution. In part 2, we discuss existence, uniqueness and regularity for the solutions to (1.1)–(1.3) for the nonlinear problem. Our main motivation for deriving regularity results is that one needs it for a rigorous study of a numerical scheme to approximate the solution. To the best of our knowledge, regularity results on inverse initial value problems (final value problems) for fractional diffusion is still unavailable in the literature. For initial value problems, McLean et al [30], Jin et al [24] and Ahmad et al [34] considered existence and regularity results of the solution in C([0, T ]; L2(D)). However it seems the techniques in [24, 34] cannot be applied for our problems (it is impossible to apply some well-known fixed point theorems with some spaces in [24] for establishing unique solutions). To overcome this we need data φ in a suitable space and we will use a Picard iteration argument and then develop some new techniques to obtain existence and regularity of the solution.

The rest of this paper is organized as follows. In section 2, we give basic notations and preliminaries, and we propose a mild solution of our problem. In section 3, we give some regularity results of the linear inhomogeneous problem. Section 4 is devoted to existence and regularity for nonlinear problems and section 5 considers global existence.

2. Notations and preliminaries

2.1. Functional space

In this subsection, we introduce some functional spaces for solutions of FVP (1.1)–(1.3). By ${\left\{{m}_{j}\right\}}_{j\ge 1}$ and ${\left\{{e}_{j}\left(x\right)\right\}}_{j\ge 1}$, we denote the spectrum and sequence of eigenfunctions of $\mathcal{L}$ which satisfy ${e}_{j}\in \left\{v\in {H}^{2}\left(D\right):\mathcal{H}v=0\right\}$, $\mathcal{L}{e}_{j}\left(x\right)={m}_{j}{e}_{j}\left(x\right)$, 0 < m1m2 ≤ ... ≤ mj ≤ ..., and $\underset{j\to \infty }{\mathrm{lim}}{m}_{j}=\infty $. The sequence ${\left\{{e}_{j}\left(x\right)\right\}}_{j\ge 1}$ forms an orthonormal basis of the space L2(D). For a given real number p ≥ 0, the Hilbert scale space H2p(D) is defined by

where (⋅, ⋅) is the usual inner product of L2(D). The fractional power ${\mathcal{L}}^{\beta }$, β ≥ 0, of the Laplacian operator $\mathcal{L}$ on D is defined by

Equation (2.1)

Then, ${\left\{{m}_{j}^{\beta }\right\}}_{j\ge 1}$ is the spectrum of the operator ${\mathcal{L}}^{\beta }$. We denote by Vβ the domain of ${\mathcal{L}}^{\beta }$, and then

where ${\Vert}.{\Vert}$ is the usual norm of L2(D), and Vβ is a Banach space with respect to the norm ${{\Vert}v{\Vert}}_{{\mathbf{V}}_{\beta }}={\Vert}{\mathcal{L}}^{\beta }v{\Vert}$. Moreover, the inclusion VβH2β(D) holds for β > 0. We identify the dual space ${\left({L}^{2}\left(D\right)\right)}^{\prime }={L}^{2}\left(D\right)$ and define the domain ${\mathbf{V}}_{-\beta }{:=}D\left({\mathcal{L}}^{-\beta }\right)$ by the dual space of Vβ, i.e., ${\mathbf{V}}_{-\beta }={\left({\mathbf{V}}_{\beta }\right)}^{\prime }$. Then, Vβ is a Hilbert space endowed with the norm

where (⋅, ⋅)β,β denotes the dual inner product between Vβ and Vβ. We note that the Sobolev embedding VβL2(D) ↪ Vβ holds for 0 < β < 1, and (tilde v,v)β,β = (tilde v, v), for tilde vL2(D), vVβ. Hence, we have

Equation (2.2)

where δij is the Kronecker delta for $i,j\in \mathbb{N}$, i, j ≥ 1. Moreover, for given p1 ≥ 1 and 0 < η < 1, we denote by ${\mathcal{X}}_{{p}_{1},\eta }\left(J{\times}D\right)$ the set of all functions f from J to ${L}^{{p}_{1}}\left(D\right)$ such that

Equation (2.3)

where ${{\Vert}.{\Vert}}_{{p}_{1}}$ is the norm of ${L}^{{p}_{1}}\left(D\right)$. Note that, for fixed t > 0, the Hölder's inequality shows that

In the above inequality, we note that the function $\tau \to {\left(t-\tau \right)}^{\frac{{p}_{2}\left(\eta -1\right)}{{p}_{2}-1}}$ is integrable for ${p}_{2}{ >}\frac{1}{\eta }$. Therefore, if we let ${L}^{{p}_{2}}\left(0,T;{L}^{{p}_{1}}\left(D\right)\right)$, p1, p2 ≥ 1, be the space of all Bochner's measurable functions f from J to ${L}^{{p}_{1}}\left(D\right)$ such that

then the following inclusion holds

Equation (2.4)

and there exists a positive constant C > 0 such that

Equation (2.5)

here, C depends only on p2, η, and T. Moreover, for a given number s such that 0 < s < η, we have ${\mathcal{X}}_{{p}_{1},\eta -s}\left(J{\times}D\right)\subset {\mathcal{X}}_{{p}_{1},\eta }\left(J{\times}D\right)$ since ${\left\vert \left\vert \left\vert \hspace{0.17em},f\right\vert \right\vert \right\vert }_{{\mathcal{X}}_{{p}_{1},\eta }}\le {T}^{s}{\left\vert \left\vert \left\vert \hspace{0.17em}f\right\vert \right\vert \right\vert }_{{\mathcal{X}}_{{p}_{1},\eta -s}}$. Let B be a Banach space, and we denote by C([0, T ], B) the space of all continuous functions from [0, T ] to B endowed with the norm ${{\Vert}v{\Vert}}_{C\left(\left[0,T\right];B\right)}{:=}{\mathrm{sup}}_{0\le t\le T}{{\Vert}v\left(t\right){\Vert}}_{B}$, and by Cθ([0, T ], B), 0 < θ ≤ 1, the subspace of C([0, T ]; B) which includes all Hölder-continuous functions, and is equipped with the norm

In some cases, a given function might not be continuous at t = 0. Hence, it is useful to consider the set $C\left(\left(0,T\right];B\right)$ which consists of all continuous functions from (0, T ] to B. We define by ${C}_{w}^{\rho }\left(\left(0,T\right];B\right)$ the weighted Banach space of all functions v in C((0, T ]; B) such that

Now, we discuss solutions of the FVP for the fractional ordinary equation

Equation (2.6)

where m, vT are given real numbers. Here, we wish to find a representation formula for v in terms of the given function g and the final value data vT. By writing ${}^{\mathrm{c}}D_{t}^{\alpha }={I}_{t}^{1-\alpha }{D}_{t}$, and applying the fractional integral ${I}_{t}^{\alpha }$ on both sides of equation (2.6), we obtain

The Laplace transform yields that

where $\hat{v}$ is the Laplace transform of v. Hence, the inverse Laplace transform implies

Equation (2.7)

Here, Eα,1 and Eα,α are the Mittag-Leffler functions which are generally defined by ${E}_{a,b}\left(z\right)={\sum }_{k=1}^{\infty }\frac{{z}^{k}}{{\Gamma}\left(ak+b\right)}$, for a > 0, $b\in \mathbb{R}$, $z\in \mathbb{C}$. Now, a representation of the solution of FVP (2.6) can be obtained by substituting t = T into (2.7), and using the final value data v(T ) = vT, i.e.,

where

Equation (2.8)

2.2. Mild solutions of FVP (1.1)–(1.3) and unboundedness of solution operators

A representation of solutions and the definition of mild solutions are given in this subsection, and then we analyze the unboundedness of solution operators. By the definition (2.1) of ${\mathcal{L}}^{\beta }$, the identity ${\mathcal{L}}^{\beta }{e}_{j}\left(x\right)={m}_{j}^{\beta }{e}_{j}\left(x\right)$ holds. Hence, in view of the Fourier expansion $u\left(t,x\right)={\sum }_{j=1}^{\infty }{u}_{j}\left(t\right){e}_{j}\left(x\right)$, where uj(t) = (u(t, ⋅), ej), equation (1.1) can be rewritten as

This is equivalent to the equation

By applying the method of solutions of FVPs for fractional ordinary equations in subsection 2.1, and using the final value data (1.3), we derive

Equation (2.9)

where φj = (φ, ej). Therefore, we obtain a spectral representation for u as follows:

For gL2(0, T; L2(D)) and vL2(D), let us denote by ${\mathcal{O}}_{n}$, 1 ≤ n ≤ 3, the following operators

and $\left({\mathcal{O}}_{3}g\right)\left(t\right){:=}-{\mathcal{O}}_{2}\left(t\right)\left({\mathcal{O}}_{1}g\right)\left(T\right)$ on L2(D), for tJ. Then, the solution u can be represented as

Equation (2.10)

for (t, x) ∈ J × D.

One of the most important things, when we consider the well-posedness of a PDE, is the boundedness of solution operators. Corresponding to the initial value problem (1.1), (1.2), (1.4), the solution operators are usually bounded in L2(D); see e.g., [27, 30, 33, 34, 57]. Unfortunately, some solution operators of FVP (1.1)–(1.3) are not bounded on L2(D) at t = 0. For this purpose, we recall that, for 0 < α < 1 and z < 0, there exist positive constants cα, ${\hat{c}}_{\alpha }$ such that

Equation (2.11)

see, for example [58, 59, 60]. Now, let v0 be defined by ${v}_{0,j}{:=}\left({v}_{0},{e}_{j}\right)={j}^{-1/2}{m}_{j}^{-\beta }$, j ≥ 1. Then, it is easy to see that v0 belongs to Vβγ for 0 ≤ γ < 1, and does not for γ ≥ 1. Using the inequalities (2.11), we have

which shows the unboundedness of ${\mathcal{O}}_{2}\left(0\right)$ on L2(D). Similarly, the unboundedness of ${\mathcal{O}}_{3}\left(0\right)$ on L2(D) can be shown.

3. FVP with a linear source

In this section, we study the regularity of mild solutions of FVP (1.1)–(1.3) corresponding to the linear source function F, i.e., F(t, x, u(t, x)) = F(t, x) which does not include u. We will investigate the regularity of the following FVP

Equation (3.1)

where φ, F will be specified later. In order to consider this problem, it it necessary to give a definition of mild solutions based on (2.10) as follows.

Definition 3.1. If a function u belongs to Lp(0, T; Lq(D)), for some p, q ≥ 1, and satisfies the equation

Equation (3.2)

then u is said to be a mild solution of FVP (3.1).

In what follows, we introduce some assumptions on the final value data φ and the linear source function F.

  • (R1) 0 < p, q < 1 such that p + q = 1;
  • (R2) $0{< }r\le \frac{1-\alpha q}{\alpha q}$;
  • (R3) $0{< }s{< }\mathrm{min}\left(\alpha q,1-\alpha q\right)$;
  • (R4) $0{< }{p}^{\prime }\le p-\frac{s}{\alpha },\quad \quad \quad \hspace{0.17em}\hspace{0.17em}\hspace{0.17em}{q}^{\prime }=1-{p}^{\prime },\quad 0{< }r\le \frac{1-\alpha {q}^{\prime }}{\alpha {q}^{\prime }}$;
  • (R5) $0\le \hat{q}\le \mathrm{min}\left(p,q,\frac{s}{\alpha }\right),\quad \hat{p}=1-\hat{q},\quad \hspace{0.17em}\hspace{0.17em}0{< }\hat{r}\le \frac{1-\alpha }{\alpha }$.

In the following lemma, we will show that solutions of FVP (3.1) must be bounded by a power function tαq, for some appropriate number q, i.e.,

for all 0 < tT.

Lemma 3.2. Let p, q be defined by (R1), and u satisfies (3.2). If φVβp, and $F\in {\mathcal{X}}_{2,\alpha q}\left(J{\times}D\right)$, then there exists a constant C0 > 0 such that

Equation (3.3)

Proof. The inequalities (2.11) shows that

Equation (3.4)

Combined with the definition $\left({\mathcal{O}}_{1}F\right)\left(t,x\right)={\sum }_{j=1}^{\infty }{F}_{j}\left(t\right)\star {{\sim}{E}}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{t}^{\alpha }\right){e}_{j}\left(x\right)$, we have that

Equation (3.5)

Hence, we obtain the following estimate

Equation (3.6)

by noting (2.3) and letting ${M}_{1}={\hat{c}}_{\alpha }{m}_{1}^{-\beta p}{T}^{\alpha q}$. In addition, the norm ${\Vert}{\mathcal{O}}_{2}\left(t\right)\varphi {\Vert}$ can be estimated as

The ratio $\frac{1+{m}_{j}^{\beta }{T}^{\alpha }}{1+{m}_{j}^{\beta }{t}^{\alpha }}$ is clearly bounded by both $1+{m}_{j}^{\beta }{T}^{\alpha }$ and $\frac{{T}^{\alpha }}{{t}^{\alpha }}$. Moreover, the increasing property of the sequence ${\left\{{m}_{j}\right\}}_{j\ge 1}$ shows $1\le {m}_{1}^{-\beta }{m}_{j}^{\beta }$. Thus, we have $1+{m}_{j}^{\beta }{T}^{\alpha }\le \left({m}_{1}^{-\beta }+{T}^{\alpha }\right){m}_{j}^{\beta }$. By noting p + q = 1, one can deduce that the ratio is bounded by the product of Tαqtαq and ${\left(1+{m}_{j}^{\beta }{T}^{\alpha }\right)}^{p}$. Bring the above arguments together, and this leads to

Equation (3.7)

where

Now, we proceed to estimate ${\Vert}\left({\mathcal{O}}_{3}F\right)\left(t,\cdot \right){\Vert}$ by using the same techniques as in (3.5) and (3.7). As a consequence of

we can obtain the following estimates

A simple computation shows that

Equation (3.8)

where we let ${M}_{3}={\hat{c}}_{\alpha }{M}_{2}$. Finally, it follows from (3.6)–(3.8), and the identity (3.2) that

The inequality (3.3) is proved by letting ${C}_{0}=\sum _{1\le n\le 3}{M}_{n}.$

Based on lemma 3.2, we consider existence, uniqueness, and regularity of solutions in the following theorem which is divided into two parts. In the first part, we obtain the existence and uniqueness of a mild solution in the space ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$ for some suitable numbers q, r and for the given assumptions on φ and F as in lemma 3.2. In the second part, we improve the smoothness of the mild solution by considering the spatial-fractional derivative ${\mathcal{L}}^{\beta \left(p-{p}^{\prime }\right)}$. It is very important to investigate the continuity of the mild solution. We first show that the mild solution is continuous from (0, T] to L2(D). Moreover, we establish the continuity on the closed interval [0, T] which corresponds to lower spatial-smoothness, Vβq', for a relevant number q'.

Theorem 3.3. 

  • (a)  
    Let p, q, r be defined by (R1), (R2). If φVβp and $F\in {\mathcal{X}}_{2,\alpha q}\left(J{\times}D\right)$, then FVP (3.1) has a unique solution u in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$. Moreover, there exists a positive constant C1 such that
    Equation (3.9)
  • (b)  
    Let p, q, s, r, p', q' be defined by (R1), (R3), (R4). If φVβp, and $F\in {\mathcal{X}}_{2,\alpha q-s}\left(J{\times}D\right)$, then FVP (3.1) has a unique solution u such that
    Moreover, there exists a positive constant C2 such that
    Equation (3.10)

Proof. The proof of part (a) can be easily obtained from lemma 3.2. Indeed, the inequality (3.3) leads to

Since $-\alpha q\left(\frac{1}{\alpha q}-r\right){ >}-1$, the integral in the above inequality exists, i.e., tαq belongs to ${L}^{\frac{1}{\alpha q}-r}\left(0,T;\mathbb{R}\right)$. Hence, FVP (3.1) has a solution u in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$. The uniqueness of u depends on the uniqueness of the ODE (2.6). For the uniqueness of this ODE, see [61] (theorem 3.25). Moreover, the inequality (3.9) is derived by letting ${C}_{1}={C}_{0}{{\Vert}{t}^{-\alpha q}{\Vert}}_{{L}^{\frac{1}{\alpha q}-r}\left(0,T;\mathbb{R}\right)}$. Now, we proceed to prove part (b) which will be presented in the following steps.

Step 1: We prove $u\in {L}^{\frac{1}{\alpha q}-r}\left(0,T;{\mathbf{V}}_{\beta \left(p-{p}^{\prime }\right)}\right)$. Firstly, by the same argument as in the proof of (3.5), we derive the following chain of inequalities

Equation (3.11)

for ${M}_{4}={\hat{c}}_{\alpha }{m}_{1}^{-\beta {p}^{\prime }}{T}^{\alpha {q}^{\prime }+s}$, where the inequality ${\left\vert \left\vert \left\vert F\right\vert \right\vert \right\vert }_{{\mathcal{X}}_{2,\alpha q}}\le {T}^{s}{\left\vert \left\vert \left\vert F\right\vert \right\vert \right\vert }_{{\mathcal{X}}_{2,\alpha q-s}}$ holds. Similarly, from ${{\Vert}{\mathcal{O}}_{2}\left(t\right)\varphi {\Vert}}_{{\mathbf{V}}_{\beta \left(p-{p}^{\prime }\right)}}={\Vert}{\mathcal{L}}^{\beta \left(p-{p}^{\prime }\right)}{\mathcal{O}}_{2}\left(t\right)\varphi {\Vert}$ and the same way as in the proof of (3.7), we have

Equation (3.12)

where we let ${M}_{5}={\hat{c}}_{\alpha }{c}_{\alpha }^{-1}{T}^{\alpha {q}^{\prime }}{\left({m}_{1}^{-\beta }+{T}^{\alpha }\right)}^{{p}^{\prime }}$. Now, we will estimate the norm ${{\Vert}\left({\mathcal{O}}_{3}F\right)\left(t,\cdot \right){\Vert}}_{{\mathbf{V}}_{\beta \left(p-{p}^{\prime }\right)}}$ which will use the same estimate for the fraction $\frac{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{t}^{\alpha }\right)}{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{T}^{\alpha }\right)}$ as in (3.12). Indeed, noting that ${\left(1+{m}_{j}^{\beta }{T}^{\alpha }\right)}^{{p}^{\prime }}\le {\left({m}_{1}^{-\beta }+{T}^{\alpha }\right)}^{{p}^{\prime }}{m}_{j}^{\beta {p}^{\prime }}$, by using (2.11), we see that

Thus, by some simple computations, one can get

Equation (3.13)

with ${M}_{6}={\hat{c}}_{\alpha }{M}_{5}{T}^{s}$, where the inequality ${\left\vert \left\vert \left\vert F\right\vert \right\vert \right\vert }_{{\mathcal{X}}_{2,\alpha q}}\le {T}^{s}{\left\vert \left\vert \left\vert F\right\vert \right\vert \right\vert }_{{\mathcal{X}}_{2,\alpha q-s}}$ has been used. Bring (3.11)–(3.13) and (3.2) together, we have

where M7 = ∑4≤n≤6Mn. Since the function ttαq' is clearly contained in the space ${L}^{\frac{1}{\alpha {q}^{\prime }}-r}\left(0,T;\mathbb{R}\right)$, we can take the ${L}^{\frac{1}{\alpha {q}^{\prime }}-r}\left(0,T;\mathbb{R}\right)$-norm on both sides of the above inequality, namely

Equation (3.14)

where ${M}_{8}={M}_{7}{{\Vert}{t}^{-\alpha {q}^{\prime }}{\Vert}}_{{L}^{\frac{1}{\alpha {q}^{\prime }}-r}\left(0,T;\mathbb{R}\right)}$.

Step 2: We prove $u\in {C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$. Let us consider 0 < t1 < t2T. By (3.2), the difference u(t2, x) − u(t1, x) can be calculated as

By using the differentiation identities

see, for example [58, 59, 60], we have

and

Combining the above arguments gives

Equation (3.15)

Now, we will establish estimates for ${\mathcal{I}}_{j}$, j = 1, 2, 3, 4, and show that ${\mathcal{I}}_{j}$ tends to 0 as t2t1 → 0. Firstly, by the inequality (2.11), we see that the absolute value of ${E}_{\alpha ,\alpha -1}\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right)$ is bounded by ${\hat{c}}_{\alpha }{m}_{j}^{-\beta p}{\omega }^{-\alpha p}$. This implies

Moreover, for 0 < τ < t1, we have

where we note that the estimates ${\left({t}_{2}-\tau \right)}^{1-\alpha q}-{\left({t}_{1}-\tau \right)}^{1-\alpha q}\le {\left({t}_{2}-{t}_{1}\right)}^{1-\alpha q}$ and ${\left({t}_{2}-\tau \right)}^{1-\alpha q}\ge {\left({t}_{1}-\tau \right)}^{s}{\left({t}_{2}-{t}_{1}\right)}^{1-\alpha q-s}$ can be showed easily from 0 < αq < 1, and 1 − αqs > 0. Hence, we deduce

This leads to

Equation (3.16)

where we let ${M}_{9}=\frac{{\hat{c}}_{\alpha }{m}_{1}^{-\beta p}}{1-\alpha q}$. Secondly, an estimate for the term ${\mathcal{I}}_{2}$ can be shown by using (2.11) as follows.

Equation (3.17)

where we let ${M}_{10}={\hat{c}}_{\alpha }{T}^{\alpha p}$. Thirdly, we will estimate the term ${\mathcal{I}}_{3}$. We have

Here, the fraction can be estimated as follows

Equation (3.18)

by using (2.11) and ${{\sim}{E}}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right)={E}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right){\omega }^{\alpha -1}$. Moreover, we can see that

Taking these estimates together, we thus obtain the following chain of the inequalities

which implies that

Equation (3.19)

where ${M}_{11}={\hat{c}}_{\alpha }{c}_{\alpha }^{-1}{T}^{\alpha p}{\left({m}_{1}^{-\beta }+{T}^{\alpha }\right)}^{q}$, and M12 = M11[αq]−1. Fourthly, we proceed to estimate ${\mathcal{I}}_{4}$. According to (3.18), we have $\left\vert \frac{{{\sim}{E}}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right)}{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{T}^{\alpha }\right)}\right\vert \le {M}_{11}{m}_{j}^{-\beta q}{\omega }^{-\alpha q-1}$. Moreover, ${{\sim}{E}}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\left(T-\tau \right)}^{\alpha }\right)\le {\hat{c}}_{\alpha }{m}_{j}^{-\beta p}{\left(T-\tau \right)}^{\alpha q-1}$ can be established by using the inequalities (2.11). Hence, we obtain

and we arrive at

Equation (3.20)

where M13 = M11Ts. We deduce from (3.16), (3.17), (3.19), (3.20) that ${\Vert}{\sum }_{1\le \hspace{0.17em}j\le 4},{\mathcal{I}}_{j}{\Vert}$ tends to 0 as t2t1 tends 0 for 0 < t1 < t2T. Thus, u belongs to the set C((0, T ]; L2(D)). On the other hand, by assumption (R3), we have 0 < αqs < αq and ${\mathcal{X}}_{2,\alpha q-s}\left(J{\times}D\right)\subset {\mathcal{X}}_{2,\alpha q}\left(J{\times}D\right)$. Therefore, the assumptions on φ and F in this theorem also fulfill lemma 3.2. Hence, the inequality (3.3) holds, i.e.,

Equation (3.21)

This implies u belongs to ${C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$. Moreover, by taking the supremum on both sides of (3.21) on (0, T ], we obtain

Equation (3.22)

Step 3: We prove uCs([0, T ]; Vβq'). In this step, we establish the continuity of the solution on the closed interval [0, T ]. Now, we consider 0 ≤ t1 < t2T. If t1 = 0, then ${\mathcal{I}}_{1}=0$. If t1 > 0, then combining (2.2) in the same way as in (3.16) gives

and so

Equation (3.23)

On the other hand, the inequality (3.17) also holds for all 0 ≤ t1 < t2T. Hence, the same way as in the proof (3.17) shows that

Now, we will establish estimates for ${{\Vert}{I}_{3}{\Vert}}_{{\mathbf{V}}_{-\beta {q}^{\prime }}}$ and ${{\Vert}{I}_{4}{\Vert}}_{{\mathbf{V}}_{-\beta {q}^{\prime }}}$. Indeed, we have

Thus, we can derive

Equation (3.24)

where we let

This leads to

Thus, by letting ${M}_{15}=\frac{{M}_{14}}{\alpha \left(p-{p}^{\prime }\right)}{T}^{\alpha \left(p-{p}^{\prime }\right)-s}$, we obtain the estimate

Equation (3.25)

where we have used that

for ${p}^{\prime }\le p-\frac{s}{\alpha }$. By employing (3.24) and following the same way as in (3.25), we have

where

This implies that

Equation (3.26)

where ${M}_{16}={\hat{c}}_{\alpha }{T}^{s}{M}_{15}$. Combining the above arguments guarantees that u belongs to Cs([0, T ]; Vβq'). Moreover, there also exists a positive constant M17 such that

Equation (3.27)

Finally, the inequality (3.10) is obtained by taking the inequality (3.14), (3.22) and (3.27) together. The proof is complete. □

In the next theorem, we will investigate the time-space fractional derivative of the mild solution u. More specifically, we investigate ${}^{\mathrm{c}}D_{t}^{\alpha }{\mathcal{L}}^{-\beta \left(q-\hat{q}\right)}u$, for a suitably chosen number $\hat{q}\le q$. We also establish the continuity of ${}^{\mathrm{c}}D_{t}^{\alpha }{\mathcal{L}}^{-\beta q}u$ on the interval (0, T ] without establishing it at t = 0 since this requires a strong assumption of F, for example, F must be continuous on whole interval [0, T ].

Theorem 3.4. Let $p,q,s,{p}^{\prime },{q}^{\prime },\hat{p},\hat{q},r,\hat{r}$ be defined by (R1), (R3), (R4), (R5).

  • (a)  
    If $\varphi \in {\mathbf{V}}_{\beta \left(p+\hat{q}\right)}$, and $F\in {L}^{\frac{1}{\alpha q-s}+\hat{r}}\left(0,T;{L}^{2}\left(D\right)\right)$, then FVP (3.1) has a unique solution u such that
    Moreover, there exists a positive constant C3 such that
    Equation (3.28)
  • (b)  
    If $\varphi \in {\mathbf{V}}_{\beta \left(p+\hat{q}\right)}$, and $F\in {L}^{\frac{1}{\alpha q-s}+\hat{r}}\left(0,T;{L}^{2}\left(D\right)\right)\cap {C}_{w}^{\alpha }\left(\left(0,T\right];{V}_{-\beta q}\right)$, then FVP (3.1) has a unique solution u such that
    Moreover, there exists a positive constant C4 such that
    Equation (3.29)

Proof. (a) By (R5), we have 0 < αqs < 1, and $\frac{1}{\alpha q-s}+\hat{r}{ >}\frac{1}{\alpha q-s}{ >}1$. Thus, one can deduce from (2.5) that

Equation (3.30)

where we have used the inclusion (2.4). Moreover, the Sobolev embedding ${\mathbf{V}}_{\beta \left(p+\hat{q}\right)}\hookrightarrow {\mathbf{V}}_{\beta p},$ holds. Therefore, the assumptions of this theorem also fulfills part (b) of theorem 3.3. Hence, FVP (3.1) has a unique solution

Now, we prove ${}^{\mathrm{c}}D_{t}^{\alpha }u$ exists and belongs to ${L}^{\frac{1}{\alpha }-\hat{r}}\left(0,T;{\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}\right)\cap {C}_{w}^{\alpha }\left(\left(0,T\right];{\mathbf{V}}_{-\beta q}\right)$. It follows from the identities

see, for example [58, 59, 60], and equation (2.9) that

for all $j\in \mathbb{N}$. Firstly, let us consider the sum ${\sum }_{{n}_{1}\le \hspace{0.17em}j\le {n}_{2}}{\psi }_{j}^{\left(1\right)}\left(t\right){e}_{j}$, for ${n}_{1},{n}_{2}\in \mathbb{N}$, 1 ≤ n1 < n2. By the definition of the dual space ${\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$ of ${\mathbf{V}}_{\beta \left(q-\hat{q}\right)}$, and the identity (2.2) of their dual inner product, we have

Assumption (R5) shows that $0{< }p+\hat{q}{< }1$. Hence, by using the inequalities (2.11), we have $\vert {{\sim}{E}}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\left(t-\tau \right)}^{\alpha }\right)\vert \le {\hat{c}}_{\alpha }{m}_{j}^{-\beta \left(p+\hat{q}\right)}{\left(t-\tau \right)}^{-\alpha \left(p+\hat{q}\right)}{\left(t-\tau \right)}^{\alpha -1}$. This together with the above argument gives

Equation (3.31)

where ${M}_{18}={\hat{c}}_{\alpha }{T}^{\alpha }$. Secondy, we proceed to establish an estimate for the sum ${\sum }_{{n}_{1}\le \hspace{0.17em}j\le {n}_{2}}{\psi }_{j}^{\left(2\right)}\left(t\right){e}_{j}$. Using the inequality (2.11), the absolute value of $\frac{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{t}^{\alpha }\right)}{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{T}^{\alpha }\right)}$ is bounded by ${\hat{c}}_{\alpha }{c}_{\alpha }^{-1}{T}^{\alpha }{t}^{-\alpha }$. Therefore, we derive

which shows that

Equation (3.32)

where ${M}_{19}{:=}{\hat{c}}_{\alpha }{c}_{\alpha }^{-1}{T}^{\alpha }$. Thirdly, we proceed to establish an estimate for the sum ${\sum }_{{n}_{1}\le \hspace{0.17em}j\le {n}_{2}}{\psi }_{j}^{\left(3\right)}\left(t\right){e}_{j}$. By a similar argument as in (3.32), we have

Therefore, we obtain the following estimate

Equation (3.33)

For almost every τ in the interval (0, T ), by (3.30), we have that F(τ, ⋅) belongs to L2(D). This implies ∑1≤jnFj(τ)ej is a Cauchy sequence in L2(D). This together with the embedding

implies that ∑1≤jnFj(τ)ej is also a Cauchy sequence in ${\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$. On the other hand, it follows from $\varphi \in {\mathbf{V}}_{\beta \left(p+\hat{q}\right)}$ that

By (R5), $0\le \hat{q}\le \frac{s}{\alpha }$, and we obtain the inclusion ${\mathcal{X}}_{2,\alpha q-s}\left(J{\times}D\right)\subset {\mathcal{X}}_{2,\alpha \left(q-\hat{q}\right)}\left(J{\times}D\right)$. We deduce that $F\in {\mathcal{X}}_{2,\alpha \left(q-\hat{q}\right)}\left(J{\times}D\right)$, and

Equation (3.34)

by the dominated convergence theorem. Combining these with the estimates (3.31)–(3.33), we have that

Hence ${\sum }_{j=1}^{n}{}^{\mathrm{c}}D_{t}^{\alpha }{u}_{j}\left(t\right){e}_{j}$ is a Cauchy sequence and a convergent sequence in ${\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$. We then conclude that ${}^{\mathrm{c}}D_{t}^{\alpha }u\left(t,\cdot \right)={\sum }_{j=1}^{\infty }{}^{\mathrm{c}}D_{t}^{\alpha }{u}_{j}\left(t\right){e}_{j}$ finitely exists in the space ${\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$. Moreover, by taking the inequalities (3.31)–(3.33), there exists a constant M20 > 0 such that

Equation (3.35)

Now, it follows from $0{< }\hat{r}\le \frac{1-\alpha }{\alpha }$ and 0 < αqs < α that $1\le \frac{1}{\alpha }-\hat{r}{< }\frac{1}{\alpha q-s}+\hat{r}$. This implies the following Sobolev embedding

Moreover, by the assumption (R5), $\hat{q}{< }\frac{s}{\alpha }$, we have $\frac{1}{\alpha q-s}+\hat{r}{ >}\frac{1}{\alpha \left(q-\hat{q}\right)}$. This implies that there exists a constant C* > 0 such that

Equation (3.36)

Hence, we deduce from (3.35) that there exists a constant M21 > 0 satisfying

Equation (3.37)

where we note that ${{\Vert}{t}^{-\alpha }{\Vert}}_{{L}^{\frac{1}{\alpha }-\hat{r}}\left(0,T;\mathbb{R}\right)}{< }\infty $. The inequality (3.28) is proved by letting C3 = M21.

(b) It is clear that the assumptions of this part also satisfy part (a). Therefore, by part (a), it is necessary to prove $u\in {C}_{w}^{\alpha }\left(\left(0,T\right];{\mathbf{V}}_{-\beta q}\right)$, i.e.,

Equation (3.38)

where we note 0 < t1 < t2T. After some simple computations we find that

Equation (3.39)

where ${\mathcal{J}}_{n}=-{\mathcal{L}}^{\beta }{\mathcal{I}}_{n}$, and ${\mathcal{I}}_{n}$ is defined by (3.15). Since F is in ${C}_{w}^{\alpha }\left(\left(0,T\right];{\mathbf{V}}_{-\beta q}\right)$, we have just to prove ${{\Vert}{\mathcal{J}}_{n}{\Vert}}_{{\mathbf{V}}_{-\beta q}}$ approaches 0 as t2t1 approaches 0. Let us first consider ${{\Vert}{\mathcal{J}}_{1}{\Vert}}_{{\mathbf{V}}_{-\beta q}}$. The inequalities (2.11) yields that

We recall that, by (3.30), F belongs to ${\mathcal{X}}_{2,\alpha q-s}\left(J{\times}D\right)$. Thus, we can deduce from the above inequality that

Equation (3.40)

where we have used the same argument as in the estimate (3.16). Let us secondly consider ${{\Vert}{\mathcal{J}}_{2}{\Vert}}_{{\mathbf{V}}_{-\beta q}}$. We have

Equation (3.41)

where (2.11) has been used. Thirdly, we consider the norm ${{\Vert}{\mathcal{J}}_{3}{\Vert}}_{{V}_{-\beta q}}$. It is clear that

Hence, we deduce

By applying (2.11), the absolute value of $\frac{{E}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right)}{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{T}^{\alpha }\right)}$ is bounded by ${\hat{c}}_{\alpha }{c}_{\alpha }^{-1}\frac{1+{m}_{j}^{\beta }{T}^{\alpha }}{1+{\left({m}_{j}^{\beta }{\omega }^{\alpha }\right)}^{2}}$. This is associated with $1+{m}_{j}^{\beta }{T}^{\alpha }\le \left({m}_{1}^{-\beta }+{T}^{\alpha }\right){m}_{j}^{\beta }$ that $\frac{1+{m}_{j}^{\beta }{T}^{\alpha }}{1+{\left({m}_{j}^{\beta }{\omega }^{\alpha }\right)}^{2}}\le \left({m}_{1}^{-\beta }+{T}^{\alpha }\right){m}_{j}^{-\beta }{\omega }^{-2\alpha }$. Thus, we obtain

Equation (3.42)

where ${M}_{23}={\hat{c}}_{\alpha }{c}_{\alpha }^{-1}\left({m}_{1}^{-\beta }+{T}^{\alpha }\right)$. Finally, we can look at ${{\Vert}{\mathcal{J}}_{4}{\Vert}}_{{\mathbf{V}}_{-\beta q}}$ as follows:

where the fraction $\frac{{{\sim}{E}}_{\alpha ,\alpha }\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right)}{{E}_{\alpha ,1}\left(-{m}_{j}^{\beta }{T}^{\alpha }\right)}$ can be estimated in the same way as in the proof of (3.42), and ${M}_{24}={\hat{c}}_{\alpha }{M}_{23}$. This leads to

which shows that

Equation (3.43)

This implies (3.38) by taking (3.39)–(3.43) together. Thus, ${}^{\mathrm{c}}D_{t}^{\alpha }u$ is contained in C((0, T ]; Vβq).

On the other hand, it is easy to see that the estimates (3.31)–(3.33) also hold for $\hat{q}=0$. Hence, we deduce from (3.35) and (3.36) that

Equation (3.44)

Now ${}^{\mathrm{c}}D_{t}^{\alpha }u\in {C}_{w}^{\alpha }\left(\left(0,T\right];{\mathbf{V}}_{-\beta q}\right)$. In addition, there exists a positive constant C' > 0 such that

We can complete the proof by taking (3.28) and the above inequality together.□

4. FVP with a nonlinear source

In this section, we study the existence, uniqueness, and regularity of mild solutions of FVP (1.1)–(1.3) corresponding to the nonlinear source function F(t, x, u(t, x)). It is suitable considering assumptions that u(t, ⋅) and F(t, ⋅, u(t, ⋅)) belong to the same spatial space H. In view of most considerations of PDEs, we let H = L2(D).

We introduce the following assumptions on the numbers $p,q,{p}^{\prime },{q}^{\prime },\hat{p},\hat{q},r,\hat{r}$.

  • (R1b) 0 < q < p < 1 such that p + q = 1;
  • (R4b) $0{< }{p}^{\prime }{< }p, \quad {q}^{\prime }=1-{p}^{\prime },\quad \quad 0{< }r\le \frac{1-\alpha {q}^{\prime }}{\alpha {q}^{\prime }}$;
  • (R4c) $0{< }{p}^{\prime }\le p-q,\quad \quad {q}^{\prime }=1-{p}^{\prime },\quad \quad 0{< }r\le \frac{1-\alpha {q}^{\prime }}{\alpha {q}^{\prime }}$;
  • (R5b) $0\le \hat{q}{< }q, \quad \hat{p}=1-\hat{q},\quad 0{< }\hat{r}\le \frac{1-\alpha }{\alpha }$.

In our work, we will assume on F(t, ⋅, u(t, ⋅)) the following assumptions

  • (A1) F(t, ⋅, 0) = 0, and there exists a constant K > 0 such that, for all v1, v2L2(D) and tJ,
  • (A2) F(t, ⋅, 0) = 0, and there exists a constant K* > 0 such that, for all v1, v2L2(D) and t1, t2J,

Note that the assumption (A1), (A2) imply that, for vL2(D),

Equation (4.1)

We try to develop the ideas of the linear FVP (3.1) to deal with the nonlinear FVP (1.1)–(1.3). In section 3, for the linear function F(t, x) we assume that

Equation (4.2)

where $p,q,s,\hat{r}$ are defined by (R1), (R3), (R5). However, we cannot suppose that the nonlinear source function F(t, x, u(t, x)) satisfies the same assumptions as in (4.2), and then find the solution u. A natural idea might be to combine the idea of lemma 3.2 with the inequality (4.1), i.e., we predict the solution u may be contained in the set

for ρ > 0, 0 < γη < 1.

The prediction will be proved in the next lemma. However, it is necessary to give some useful notes on ${\mathbf{W}}_{\gamma ,\eta }^{\rho }\left(J{\times}D\right)$ as follows. For $w\in {\mathbf{W}}_{\gamma ,\eta }^{\rho }\left(J{\times}D\right)$, we see

The function ττγ(tτ)η−1 is integrable on (0, t) since both numbers −γ and η − 1 are greater than −1. In addition, we have ${\int }_{0}^{t}{\tau }^{-\gamma }{\left(t-\tau \right)}^{\eta -1}\mathrm{d}\tau ={t}^{\eta -\gamma }B\left(\eta ,1-\gamma \right)$, where B(⋅, ⋅) is the beta function see, for example, [58, 59, 60]. Hence, we have

Equation (4.3)

Moreover, if γ < η, then there always exists a real number p such that $1{< }\frac{1}{\eta }{< }p{< }\frac{1}{\gamma }$. This implies that the function tγ belongs to Lp(0, T; L2(D)). Therefore, we can obtain the following inclusions

In the following lemma, we will consider the case γ = η, which we will denote by ${\mathbf{W}}_{\gamma }^{\rho }\left(J{\times}D\right){:=}{\mathbf{W}}_{\gamma ,\eta }^{\rho }\left(J{\times}D\right)$.

Now, the Sobolev embedding VβpL2(D) shows that there exists a positive constant CD depending on D, β, q such that ${\Vert}v{\Vert}\le {C}_{D}{{\Vert}v{\Vert}}_{{\mathbf{V}}_{\beta p}}$ for all vVβp. In this section, we let

and

Lemma 4.1. Let p, q be defined by (R1 ) and ϕ be a function on D. Let $\left\{{w}_{\left(n\right)}\right\}$ be defined by w(0) = ϕ,

Equation (4.4)

If ϕ belongs to Vβp, F satisfies (A1), and k0(T ) < 1, then

Equation (4.5)

where ${\hat{C}}_{0}{:=}{{\sim}{C}}_{0}{{\Vert}\phi {\Vert}}_{{\mathbf{V}}_{\beta p}}$ and ${{\sim}{C}}_{0}=\frac{{M}_{0}}{1-{k}_{0}\left(T\right)}$.

Proof. First, we have

Hence, inequality (4.3) and αq < 1, imply ${w}_{\left(0\right)}\in {\mathbf{W}}_{\alpha q}^{{\hat{C}}_{0}}\left(J{\times}D\right)$. Now, we assume that w(n−1) belongs to ${\mathbf{W}}_{\alpha q}^{{\hat{C}}_{0}}\left(J{\times}D\right)$ for some n ≥ 1. Then, by using (4.3), we have

Equation (4.6)

By induction, the inclusion (4.5) will be proved by showing that w(n) belongs to ${\mathbf{W}}_{\alpha q}^{{\hat{C}}_{0}}\left(J{\times}D\right)$. Indeed, by using the same arguments as in the proof of (3.6), we have

Equation (4.7)

where we have used (4.1), (4.6) and let ${M}_{26}=K{\hat{c}}_{\alpha }{m}_{1}^{-\beta p}B\left(\alpha q,1-\alpha q\right){T}^{\alpha q}$. On the other hand, the norm ${\Vert}{\mathcal{O}}_{2}\left(t\right)\phi {\Vert}$ is estimated by (3.7), i.e.,

where we note that

The norm ${\Vert}\left({\mathcal{O}}_{3}F\left({w}_{\left(n-1\right)}\right)\right)\left(t,\cdot \right){\Vert}$ can be estimated in the same way as in the proof of (3.8). That is,

Equation (4.8)

where (4.1), (4.6) have been used. Here, M27 = KM3B(αq, 1 − αq), ${M}_{3}={\hat{c}}_{\alpha }^{2}{c}_{\alpha }^{-1}{T}^{\alpha q}{\left({m}_{1}^{-\beta }+{T}^{\alpha }\right)}^{p}$.

We deduce from the above arguments and ${w}_{\left(n\right)}\left(t,\cdot \right)=\mathcal{O}\left(t,\cdot \right){w}_{\left(n-1\right)}$ that

Equation (4.9)

by noting that k0(T ) = M26 + M27. Since ${\hat{C}}_{0}=\frac{{M}_{0}}{1-{k}_{0}\left(T\right)}{{\Vert}\phi {\Vert}}_{{\mathbf{V}}_{\beta p}}$, the above inequality implies that ${\Vert}{w}_{\left(n\right)}\left(t,\cdot \right){\Vert}\le {\hat{C}}_{0}{t}^{-\alpha q}$. Therefore, from αq < 1, we obtain the inclusion (4.5).□

Next, it is necessary to give a definition of mild solutions of FVP (1.1)–(1.3).

Definition 4.2. Let F be defined by (A1) or (A2). If a function u belongs to ${L}^{{p}_{2}}\left(0,T;{L}^{{p}_{1}}\left(D\right)\right)$, for some p1, p2 ≥ 1, and satisfies equation (2.10) where F stands for the nonlinear source function F(t, x, u(t, x)), then u is said to be a mild solution of FVP (1.1)–(1.3).

The following theorems presents existence, uniqueness, and regularity of a mild solution of FVP (1.1)–(1.3).

Theorem 4.3. 

  • (a)  
    Let p, q, r, p', q' be defined by (R1), (R4b). If φ belongs to Vβp, F satisfies (A1), and k0(T ) < 1, then FVP (1.1)–(1.3) has a unique solution
    and there exists a positive constant C5 such that
  • (b)  
    Let p, q, r, p', q' be defined by (R1b), (R4c). If φ belongs to Vβp, F satisfies (A1), and k0(T ) < 1, then FVP (1.1)–(1.3) has a unique solution
    and there exists a positive constant C6 such that

Proof. (a) We divide the proof of this part into the following steps.

Step 1: We prove the existence and uniqueness of a mild solution. In order to prove the existence of a mild solution of FVP (1.1)–(1.3), we will construct a convergent sequence in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$ whose limit will be a mild solution of the problem. Here, r is defined by (R2). Let ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}$ be a sequence defined by lemma 4.1 with respect to ϕ = φVβp, then ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}\subset {\mathbf{W}}_{\alpha q}^{{\hat{C}}_{0}}\left(J{\times}D\right)$ where ${\hat{C}}_{0}{:=}\frac{{M}_{2}}{1-{k}_{0}\left(T\right)}{{\Vert}\varphi {\Vert}}_{{\mathbf{V}}_{\beta p}}$. Therefore,

for all n ≥ 1. This together with tαq belonging to ${L}^{\frac{1}{\alpha q}-r}\left(0,T;\mathbb{R}\right)$ implies that ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}$ is a bounded sequence in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$. Now, we will show that ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}$ is convergent by proving that it is also a Cauchy sequence. For fixed n ≥ 1 and k ≥ 1, the definition (4.4) of ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}$ yields that

Since F satisfies lemma 4.1, the latter equation shows that we can apply the same arguments as in lemma 4.1 with ϕ = 0. Hence, one can deduce

where we combined the estimates (4.7), (4.8). From ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}\subset {\mathbf{W}}_{\alpha q}^{{\hat{C}}_{0}}\left(J{\times}D\right)$, we have

Thus, by noting the identity

where a, b > 0 and B(⋅, ⋅) is the beta function, we find that

From the definition of k0(T ), we conclude that

Iterating this method n-times shows

Taking the ${L}^{\frac{1}{\alpha q}-r}\left(0,T;\mathbb{R}\right)$-norm of both sides of the above inequality directly implies

Equation (4.10)

Here, we emphasise that the constants in (4.10) also do not depend on (n, k). Therefore, by letting n go to infinity, we obtain

i.e., ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}$ is a bounded Cauchy sequence in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$. Hence, there exists a function u in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$ such that

and u satisfies equation (2.10), i.e., u is a mild solution of FVP problem (1.1)–(1.3). Moreover, the boundedness (4.9) of ${\left\{{w}_{\left(n\right)}\right\}}_{n\ge 0}$ gives

Equation (4.11)

and so that

where ${M}_{28}={{\sim}{C}}_{0}{{\Vert}{t}^{-\alpha q}{\Vert}}_{{L}^{\frac{1}{\alpha q}-r}\left(0,T;\mathbb{R}\right)}$.

Now, we show the uniqueness of the solution u. Assume that ${\sim}{u}$ is another solution of FVP (1.1)–(1.3). Then, by applying the same argument as in (4.10), we also have

for all $n\in \mathbb{N}$, n ≥ 1. Thus ${{\Vert}u-{\sim}{u}{\Vert}}_{{L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)}=0$ by letting n go to infinity. Hence, $u={\sim}{u}$ in ${L}^{\frac{1}{\alpha q}-r}\left(0,T;{L}^{2}\left(D\right)\right)$.

Step 2: We prove that $u\in {L}^{\frac{1}{\alpha {q}^{\prime }}-r}\left(0,T;{\mathbf{V}}_{\beta \left(p-{p}^{\prime }\right)}\right)$. This will be proved by using the inequality (4.11). We now apply the same arguments as in the proofs of (3.11), and (3.13) to estimate ${{\Vert}u\left(t,\cdot \right){\Vert}}_{{\mathbf{V}}_{\beta \left(p-{p}^{\prime }\right)}}$ as follows. First, we have

Equation (4.12)

where we let

Secondly,

Equation (4.13)

where we let ${M}_{30}={M}_{6}K{{\sim}{C}}_{0}B\left(\alpha q,1-\alpha q\right)$. We recall that ${{\Vert}{\mathcal{O}}_{2}\left(t\right)\varphi {\Vert}}_{{\mathbf{V}}_{\beta \left(p-{p}^{\prime }\right)}}$ have been estimated by (3.12). According to the above arguments, we arrive at the estimate

for M31 = M29 + M5 + M30. By taking the ${L}^{\frac{1}{\alpha {q}^{\prime }}-r}\left(0,T;\mathbb{R}\right)$-norm, then the latter inequalities can be transformed into the following estimate

Equation (4.14)

where ${M}_{32}={M}_{31}{{\Vert}{t}^{-\alpha {q}^{\prime }}{\Vert}}_{{L}^{\frac{1}{\alpha {q}^{\prime }}-r}\left(0,T;\mathbb{R}\right)}$.

Step 3: We prove that $u\in {C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$. Let us consider 0 < t1 < t2T. By the same arguments as in (3.15), we have

Equation (4.15)

Here, by (3.19), ${\Vert}{\mathcal{I}}_{3}{\Vert}$ tends to 0 as t2t1 tends to 0. In what follows, we will establish the convergence for ${\Vert}{\mathcal{I}}_{n}^{N}{\Vert}$, n = 1, 2, 4 which can be treated similarly as in (3.16), (3.17), (3.20) based on the Lipschitzian assumption (A1). We first see that

Equation (4.16)

Note that

Thus, due to the substitution τ = t2μ, we have

Equation (4.17)

As a consequence, $\underset{{t}_{2}-{t}_{1}\to 0}{\mathrm{lim}}{\int }_{0}^{{t}_{1}}{\tau }^{-\alpha q}\left[{\left({t}_{1}-\tau \right)}^{\alpha q-1}-{\left({t}_{2}-\tau \right)}^{\alpha q-1}\right]\mathrm{d}\tau =0$, and so $\underset{{t}_{2}-{t}_{1}\to 0}{\mathrm{lim}}{\Vert}{\mathcal{I}}_{1}^{N}{\Vert}=0$. Secondly we proceed to deal with ${\mathcal{I}}_{2}^{N}$. Now 0 < p0 < p, by (2.11), we have

Equation (4.18)

We deduce the following chain of estimates

Equation (4.19)

This implies $\underset{{t}_{2}-{t}_{1}\to 0}{\mathrm{lim}}{\Vert}{\mathcal{I}}_{2}^{N}{\Vert}=0$. Next, we thirdly proceed to consider ${\mathcal{I}}_{4}^{N}$. The same argument as in (3.20) gives

Equation (4.20)

and we arrive at $\underset{{t}_{2}-{t}_{1}\to 0}{\mathrm{lim}}{\Vert}{\mathcal{I}}_{4}^{N}{\Vert}=0$. The above arguments prove uC((0, T ]; L2(D)). This combines with (4.11) so that $u\in {C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$, and

Equation (4.21)

We complete step 1 by combining the inequalities (4.14) and (4.21).

(b) According to part (a), we have just to prove that uCαq([0, T ]; Vβq'). In this part, we consider 0 ≤ t1 < t2T. Let us first show

Equation (4.22)

for some positive constant M33, where the case t1 = 0 is trivial. It is necessary to prove (4.22) for t1 > 0. From the proof of the estimate (3.23), we have

In addition, the inequalities (2.11) yield that, $\vert {E}_{\alpha ,\alpha -1}\left(-{m}_{j}^{\beta }{\omega }^{\alpha }\right)\vert \le {\hat{c}}_{\alpha }{m}_{j}^{-\beta {p}^{\prime }}{\omega }^{-\alpha {p}^{\prime }}$. This associates with αq' − 2 = αq − 2 + α(pp') so that

where we let

Here, we have used (3.15), (4.17), and α(pp') ≥ αq by (R4b). Secondly, we are going to consider ${\mathcal{I}}_{2}^{N}$. The Sobolev embedding L2(D) ↪ Vβq' yields that there exists a positive constant M34 such that

where we applied (4.19) with respect to 0 < p0 = pp' < p. Thirdly, we now consider ${\mathcal{I}}_{3}^{N}$. By applying the same arguments as in (3.25), one can get

Equation (4.23)

Finally, the arguments proving (3.26) give

Taking the above estimates for ${\mathcal{I}}_{n}^{N}$, 1 ≤ n ≤ 4 together, we conclude that u belongs to Cαq([0, T ]; Vβq'). Moreover, there exists a positive constant M35 such that

By combining this inequality with (4.14), (4.21), we complete the proof.□

Theorem 4.4. Let $p,q,{p}^{\prime },{q}^{\prime },\hat{p},\hat{q},r,\hat{r}$ be defined by (R1), (R4b), (R5b). If φ belongs to ${\mathbf{V}}_{\beta \left(p+\hat{q}\right)}$, F satisfies the assumptions (A2), and k0(T ) < 1, then FVP (1.1)–(1.3) has a unique solution u satisfying that

Moreover, there exists a constant C7 > 0 such that

Equation (4.24)

Proof. Since F satisfies (A2), F also satisfies (A1) with respect to the Lipschitz constant K*. In addition, the Sobolev imbedding ${\mathbf{V}}_{\beta \left(p+\hat{q}\right)}\hookrightarrow {\mathbf{V}}_{\beta p}$ shows that φ belongs to Vβp. Hence, by theorem 4.3, FVP (1.1)–(1.3) has a unique solution

Moreover, the inequality (4.11) also holds. We deduce that, for 0 < tT,

Equation (4.25)

The remainder of this proof falls naturally into two steps as follows.

Step 1: We prove ${}^{\mathrm{c}}D_{t}^{\alpha }u$ finitely exists and belongs to ${L}^{\frac{1}{\alpha }-\hat{r}}\left(0,T;{\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}\right)$. By the same way as in part (a) of theorem 3.4, we have

for all $j\in \mathbb{N}$, j ≥ 1. In view of (4.25), F(t, ⋅, u(t, ⋅)) is contained in L2(D) for 0 < tT. This associates with the Sobolev embedding ${L}^{2}\left(D\right)\hookrightarrow {\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$ that F(t, ⋅, u(t, ⋅)) is contained in ${\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$, namely ${\sum }_{j=1}^{\infty }{F}_{j}\left(t,u\left(t\right)\right){e}_{j}$ is contained in ${\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$. On the other hand, ${\psi }_{j}^{N,2}={\psi }_{j}^{\left(2\right)}$, and the norm ${{\Vert}{\sum }_{j=1}^{\infty }{\psi }_{j}^{N,2}\left(t\right){e}_{j}{\Vert}}_{{\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}}$ exists finitely by (3.32). Now, we consider ${{\Vert}{\sum }_{j=1}^{\infty }{\psi }_{j}^{N,n}\left(t\right){e}_{j}{\Vert}}_{{\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}}$, n = 1, 3. According to the estimates (3.31) and (3.33), the following ones hold:

For 0 < τ < T, we have F(τ, ⋅, u(τ, ⋅)) belonging to L2(D). This follows that the sequence $\left\{{G}_{n}\left(\tau \right)\right\}$, which is defined by ${G}_{n}\left(\tau \right)={\left\{{\sum }_{j\ge n}{F}_{j}^{2}\left(\tau ,u\left(\tau \right)\right)\right\}}^{1/2}$, converges pointwise to 0 as n goes to infinity. Moreover, by (4.25), we have

The function $\tau \to {\left(t-\tau \right)}^{\alpha \left(q-\hat{q}\right)-1}{\tau }^{-\alpha q}$ is integrable on the open interval (0, t), t > 0, since

Therefore, the dominated convergence theorem yields that

This together with ${\left\{{\sum }_{{n}_{1}\le \hspace{0.17em}j\le {n}_{2}}{F}_{j}^{2}\left(\tau ,u\left(\tau \right)\right)\right\}}^{1/2}\le {G}_{{n}_{1}}\left(\tau \right)$ gives

Similarly, we also have

We deduce ${{\Vert}{\sum }_{j=1}^{\infty }{\psi }_{j}^{N,n}\left(t\right){e}_{j}{\Vert}}_{{\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}}$, n = 1, 3 exist finitely. Taking all the above arguments together, we conclude that ${{\Vert}{\sum }_{j=1}^{\infty }{}^{\mathrm{c}}D_{t}^{\alpha }{u}_{j}\left(t\right){e}_{j}{\Vert}}_{{\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}}$ finitely exists. In addition, the Sobolev embedding ${L}^{2}\left(D\right)\hookrightarrow {\mathbf{V}}_{-\beta \left(q-\hat{q}\right)}$ yields that there exists a positive constant M37 such that

Hence,

We now note that

and

This combines with (4.25) and there exists a constant M38 > 0 such that

Equation (4.26)

which leads to

Equation (4.27)

Step 2: We prove ${}^{\mathrm{c}}D_{t}^{\alpha }u\in {C}_{w}^{\alpha }\left(\left(0,T\right];{\mathbf{V}}_{-\beta q}\right)$. We consider 0 < t1 < t2T. A similar argument as in (3.39) yields

where ${\mathcal{J}}_{n}^{N}=-{\mathcal{L}}^{\beta }{\mathcal{I}}_{n}^{N}$ and ${\mathcal{I}}_{n}^{N}$ is defined by (4.15). By applying the Sobolev embedding L2(D) ↪ Vβq, there exists a positive constant M39 such that

where we note that $u\in {C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$. From (3.40) and (4.17), we have

In addition, by

and (4.25), we can obtain the following chain of the inequalities

Equation (4.28)

Finally, the norm ${{\Vert}{\mathcal{J}}_{3}^{N}{\Vert}}_{{V}_{-\beta q}}$ has been estimated by (3.42), and the norm ${{\Vert}{\mathcal{J}}_{4}^{N}{\Vert}}_{{V}_{-\beta q}}$ can be estimated as follows:

It follows from the above arguments that ${}^{\mathrm{c}}D_{t}^{\alpha }u$ belongs to C((0, T]; Vβq). On the other hand, the estimate (4.26) also holds for $\hat{p}=0$ and $\hat{q}=1$, i.e., we have

for 0 < tT. Therefore, ${}^{\mathrm{c}}D_{t}^{\alpha }u\in {C}_{w}^{\alpha }\left(\left(0,T\right];{\mathbf{V}}_{-\beta q}\right)$ there exists a constant M40 > 0 such that

Equation (4.29)

by the Sobolev embedding ${\mathbf{V}}_{\beta \left(p+\hat{q}\right)}\hookrightarrow {\mathbf{V}}_{\beta p}$. The inequality (4.24) is derived by taking the inequality (4.27) and (4.29) together. We finally complete the proof. □

Remark 4.1. At the beginning part of step 1 of the above proof, we recall that ${\psi }_{j}^{N,2}={\psi }_{j}^{\left(2\right)}$. This means that ${\psi }_{j}^{N,2}$ (with respect to the nonlinear case) can be similarly estimated in the same way as in the linear case. More precisely, this term can be estimated as (3.31). Here, the formula of ${\psi }_{j}^{\left(2\right)}$ is given by

Equation (4.30)

see the proof of theorem 3.4. The appearance of the factor ${m}_{j}^{\beta }$ in (4.30) tells that we need $\varphi \in {\mathbf{V}}_{\beta \left(p+\hat{q}\right)}$ to obtain (3.31). In summary, in order to bound the Caputo fractional derivative ${}^{\mathrm{c}}D_{t}^{\alpha }u$, we need the stronger assumption $\varphi \in {\mathbf{V}}_{\beta \left(p+\hat{q}\right)}$ rather than φVβp.

5. Discussion on global existence of solutions

In the previous section, we found a solution u of FVP (1.1)–(1.3) in the set ${\mathbf{W}}_{\alpha q}^{{\hat{C}}_{0}}\left(J{\times}D\right)$. This allows that ${\Vert}u\left(t,\cdot \right){\Vert}\le {\hat{C}}_{0}{t}^{-\alpha q}$ for all 0 < tT. Then, we obtain $u\in {C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$ by establishing the time continuity of u, which corresponds to the boundedness

Equation (5.1)

However, the existence given in theorem 4.3 requires the assumption k0(T ) < 1, which is equivalent to KTαq < M41, where M41 is a constant. This can occur if K or T is small enough.

'Under what conditions is the contractivity condition k0(T ) < 1 satisfied?' This motivates the result in this section.

The purpose of this section is to discuss global existence of solutions, namely, existence of solutions without any assumptions on K and T. To overcome the difficulties of finding solutions in ${C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$, we shall seek solutions in a wider/weaker space than ${C}_{w}^{\alpha q}\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$. The alternative solution space we are going to find is to take inspiration from replacing the supremum (5.1) by the following integral

Equation (5.2)

with suitable parameter b, ρ, and μ. We expect that the mapping $\mathcal{O}$ (formulated by lemma 4.1) on the alternative solution space is contracted as ρ tends to positive infinity.

The above arguments motivate us to denote by ${L}_{\rho ,b}^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$, μ ≥ 1, ρ > 0, b > 0, the weighted Lebesgue space of all functions v: (0, T ) → L2(D) such that

In the next theorem, we present global existence for FVP (1.1)–(1.3) in ${L}_{\rho ,b}^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$. It is helpful to introduce the following special function

which is called the Kummer function or hypergeometric function. We recall the following asymptotic behavior of this function ${}_{1}\mathcal{F}_{1}\left(a,b,z\right){:=}{\Gamma}\left(b\right){\left({\Gamma}\left(a\right)\right)}^{-1}{\mathrm{e}}^{z}{z}^{-\left(b-a\right)}\left(1+O{\vert z\vert }^{-1}\right)$, see [62] (lemma 8) or [63] (chapter 13). Due to a simple integration by substitution, for all t ∈ (0, T ) we observe that

Equation (5.3)

Theorem 5.1. Assume that $\frac{1}{2}{< }\alpha {< }1$. Let p, q be defined by (R1) such that $\frac{1}{2\alpha }{< }q{< }1$. Let μ and b be such that $\frac{1}{\alpha q}{< }\mu {< }2$, $\alpha q-\frac{1}{\mu }{< }b{< }1-\frac{1}{\mu }$. If φ belongs to Vβp, F satisfies (A1), then there exists $\hat{\rho }{ >}0$ such that FVP (1.1)–(1.3) has a unique solution ${u}_{G}\in {L}_{\rho ,b}^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$ with $\rho \ge \hat{\rho }$, and furthermore

Proof. For ${v}_{1},{v}_{2}\in {L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$, we first estimate the ${L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$-norm of ${\mathcal{O}}_{1}{v}_{1}-{\mathcal{O}}_{1}{v}_{2}$. The main idea is splitting the quantity ∥F(τ, ⋅, v1(τ, ⋅)) − F(τ, ⋅, v2(τ, ⋅))∥(tτ)αq−1 into the product of (tτ)αq−1τbeρτ and τbeρτF(τ, ⋅, v1(τ, ⋅)) − F(τ, ⋅, v2(τ, ⋅))∥. Then the ${L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$-norm of ∥v1(τ, ⋅) − v2(τ, ⋅)∥ can be obtained by applying the Hölder inequality and the Lipschitz assumption (A1). Indeed, one can see that

where we denote ${C}_{1}{:=}{\Gamma}\left(\left(\alpha q\mu -1\right)/\left(\mu -1\right)\right){\left(\left(\mu -1\right)/\mu \right)}^{\left(\alpha q\mu -1\right)/\left(\mu -1\right)}$. Here, the asymptotic behavior (5.3) had been used in the second estimate, where we note that

as μ > 1/(αq), and 1 − /(μ − 1) > 0 as b < (μ − 1)/μ. Moreover, by taking ρ large enough we can bound ${\left(T+O\left({\rho }^{-1}\right)\right)}^{\mu -1}$ by a constant independently of x, t. Since μ > 1/(αq), the factor ρ1−αqμ obviously tends to zero as ρ tends to infinity. Furthermore, the latter improper integral is convergent as μ < 2. Summarizing, we can find a constant ρ1 > 0 such that

so we arrive at the following estimate

Equation (5.4)

for all ρρ1, where we have used (4.7) in the first estimate and the Lipschitz assumption (A1) in the second estimate.

Second, we will estimate the ${L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$-norm of the difference ${\mathcal{O}}_{3}{v}_{1}-{\mathcal{O}}_{3}{v}_{2}$. Based on estimating the ${L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$-norm as above, this can be treated by combining the Hölder inequality, the Lipschitz assumption (A1), and the inequality (4.8). Indeed, we see that

where the asymptotic behavior (5.3) was employed. Let us denote by ${\mathcal{I}}_{b,\rho }$ the latter integral, then

Note that (bαq)μ > ((αq − 1/μ) − αq)μ = −1 as b > αq − 1/μ. Hence, using the asymptotic behavior (5.3), we have

where the latter right-hand side tends to zero as ρ tends to infinity. Thus, there exists a ρ2 > 0 such that

Equation (5.5)

for all ρρ2. Taking the estimates (5.4) and (5.5) together gives

for all ρ ≥ max{ρ1; ρ2}, namely, $\mathcal{O}:{L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)\to {L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$ is a contraction mapping. This means that $\mathcal{O}$ has only one fixed point in ${L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$, and so FVP (1.1)–(1.3) has a unique solution uG in the weighted Lebesgue space ${L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$. The desired inequality is obvious. □

Remark 5.1. In fact, we tried to find solutions in the space

Note that the following inclusion holds

Indeed, if $v\in {C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$ then

In order to find solutions in this space, for all ${v}_{1},{v}_{2}\in {C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$, it requires to bound the following quantities

by ${k}_{j}\left(\rho \right){{\Vert}{v}_{1}-{v}_{2}{\Vert}}_{{C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)}$ with kj(ρ), j = 1, 2, tend to zero as ρ tends to infinity. Unfortunately, it does not occur with the term Q2(b, ρ). Indeed, since ${v}_{1},{v}_{2}\in {C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$ we have

which gives

The following conclusions are obvious:

  • Due to the asymptotic behavior (5.3), the supremum of ${\hat{\sigma }}_{b,\rho }\left(t,T\right)$ on (0, T ] tends to infinity as ρ tends to infinity. Hence, $\mathcal{O}:{L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)\to {L}_{b,\rho }^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$ cannot be a contraction mapping for arbitrary T. This is the main reason why we did not find solutions in ${C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$.
  • The idea of using the space ${L}_{\rho ,b}^{\mu }\left(0,T;{L}^{2}\left(D\right)\right)$ fortuitously came when we realized that
    with a suitable number μ ≥ 1. Here, we replaced the supremum (5.1) by the integral (5.2).
  • One can show that Q1(b, ρ) is bounded by ${k}_{1}\left(\rho \right){{\Vert}{v}_{1}-{v}_{2}{\Vert}}_{{C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)}$, where k1(ρ) tends to zero as ρ tends to infinity. This means that we can establish the existence of a mild solution to the initial value problem (1.1), (1.2), (1.4) in ${C}_{w}^{b,\rho }\left(\left(0,T\right];{L}^{2}\left(D\right)\right)$ without any assumptions on K and T.

Acknowledgments

We would like to thank the editor and anonymous referees for their valuable comments and suggestions. This research was supported by Vietnam National Foundation for Science and Technology Development (NAFOSTED) under Grant Number 101.02-2019.09.

Please wait… references are loading.