This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

11institutetext: Dian Feng1,∗ 22institutetext: 1. School of Mathematical Sciences, Fudan University, Shanghai 200433, China
22email: [email protected] corresponding author
33institutetext: Masahiro Yamamoto2,3 44institutetext: 2. Graduate School of Mathematical Sciences, The University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8914, Japan 55institutetext: 3. Department of Mathematics, Faculty of Science, Zonguldak Bülent Ecevit University, Zonguldak 67100, Türkiye
55email: [email protected]

Global and local existence of solutions for nonlinear systems of time-fractional diffusion equations

Dian Feng1    Masahiro Yamamoto2,3
(Received: XX 2024 / Revised: … / Accepted: ……)
Abstract

In this paper, we consider initial-boundary value problems for two-component nonlinear systems of time-fractional diffusion equations with the homogeneous Neumann boundary condition and non-negative initial values. The main results are the existence of solutions global in time and the blow-up. Our approach involves the truncation of the nonlinear terms, which enables us to handle all local Lipschitz continuous nonlinear terms, provided their sum is less than or equal to zero. By employing a comparison principle for the corresponding linear system, we establish also the non-negativity of the nonlinear system.

Keywords:
Nonlinear time-fractional system weak solutionglobal existenceblow-up
MSC:
26A33 (primary) 33E12 34A08 34K37 35R11 60G22 …
journal: Fract. Calc. Appl. Anal.

1 Introduction

Let tα\partial_{t}^{\alpha} be the fractional derivative of order α(0,1)\alpha\in(0,1) defined on the fractional Sobolev spaces, which is defined as the closure of the classical Caputo derivative

dtαv(t):=1Γ(1α)0t(ts)αdvds(s)𝑑s, for vC1[0,T] satisfying v(0)=0d_{t}^{\alpha}v(t):=\frac{1}{\Gamma(1-\alpha)}\int^{t}_{0}(t-s)^{-\alpha}\frac{dv}{ds}(s)ds,\quad\mbox{ for }v\in C^{1}[0,T]\mbox{ satisfying }v(0)=0

(see Section 2 for the details). Let Ωd,d=1,2,3\Omega\subset\mathbb{R}^{d},d=1,2,3 be a bounded domain with smooth boundary Ω\partial\Omega. Moreover by ν\nu we denote the unit outward normal vector to Ω\partial\Omega at xΩx\in\partial\Omega.

We consider an initial-boundary value problem for the following nonlinear system of time-fractional equations with the homogeneous Neumann boundary condition

{tα(ua)=Δu+f(u,v),xΩ,0<t<T,tα(vb)=Δv+g(u,v),xΩ,0<t<T,νu=νv=0, on Ω×(0,T),\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u-a)=\Delta u+f(u,v),\quad x\in\Omega,0<t<T,\\ \partial_{t}^{\alpha}(v-b)=\Delta v+g(u,v),\quad x\in\Omega,0<t<T,\\ \partial_{\nu}u=\partial_{\nu}v=0,\quad\text{ on }\partial\Omega\times(0,T),\end{array}\right. (1.1)

where we assume some conditions on the nonlinear terms ff and gg, as Assumption 1 in Section 2 describes.

When α=1\alpha=1, the system is reduced to a classical reaction-diffusion system, and can be regarded as a special case of the Klausmeier-Gray-Scott model Kla ; WS . This system is considered to be a model equation for the vegetation pattern formation, which describes the self-organization of vegetation spatial patterns resulting from the interaction between water source distribution and plant growth. In Pi and Pi2 , it was proved that a weak solution in L1L^{1} exists in the case where the coefficients in front of the spatial diffusion terms are not necessarily equal. However for the our fractional derivative case 0<α<10<\alpha<1, we can not apply the same technique as for α=1\alpha=1, so that we have to assume that the coefficients of spatial diffusion terms are equal, that is, 11. It should be a future work to discuss more general elliptic operators in (1.1).

For 0<α<10<\alpha<1, in contrast to the traditional reaction-diffusion process, we introduce a fractional order in the time derivative, indicating that the interaction between vegetation growth and water is influenced by soil medium heterogeneity. The fractional derivatives enable the representation of time memory effects, characterized by their nonlocal properties. Although there has been some research on various nonlinear fractional-order equation or coupled linear fractional-order equation systems, as seen in BP ; LHL ; FLY , to the best of the authors’ knowledge, there has been no investigation concerning nonlinear fractional-order systems.

Our approach involves the truncation of the nonlinear terms, which is inspired by Pierre Pi ; Pi2 . Through a comparison principle and a compact mapping property, we establish a convergent sequence of non-negative functions for the linear case. Moreover, u+vu+v has the energy estimation in the sense of L1L^{1}-norm, implying the convergence of the sequence in the L1L^{1} sense. Ultimately, we obtain weak solutions in the L1L^{1} norm.

The rest of this paper is organized as follows. In Section 2, we present the result regarding the global existence of weak solutions to the initial-boundary value problems for the nonlinear system of time-fractional diffusion equations. Section 3 is devoted to proving two key lemmas, which serve as the basis for the proofs of the main results. These lemmas establish the non-negativity of solution to the linear case, provided that the initial value and source term are non-negative, and demonstrate that the mapping from initial value and non-homogeneous term to solution of the corresponding linear system, is compact with suitable norms. Section 4 completes the proof of the first main result. Section 5 discusses the blow-up of solutions under other conditions on the nonlinear terms, which rejects the global existence in time, in general. Section 6 provides some conclusions and remarks.

2 Well-posedness results

In this section, we deal with the following initial-boundary value problem for the nonlinear system of time-fractional equations (1.1) with the time-fractional derivative of order α(0,1)\alpha\in(0,1)

{tα(ua)=Δu+f(u,v),xΩ, 0<t<T,tα(vb)=Δv+g(u,v),xΩ, 0<t<T,νu=νv=0, on Ω×(0,T),\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u-a)=\Delta u+f(u,v),\quad x\in\Omega,\ 0<t<T,\\ \partial_{t}^{\alpha}(v-b)=\Delta v+g(u,v),\quad x\in\Omega,\ 0<t<T,\\ \partial_{\nu}u=\partial_{\nu}v=0,\quad\text{ on }\partial\Omega\times(0,T),\end{array}\right. (2.1)

along with the initial condition (2.2) formulated below.

For α(0,1)\alpha\in(0,1), let dtα\mathrm{d}^{\alpha}_{t} denote the classical Caputo derivative:

dtαw(t):=0t(ts)αΓ(1α)dwds(s)ds,wW1,1(0,T).\mathrm{d}^{\alpha}_{t}w(t):=\int_{0}^{t}\frac{(t-s)^{-\alpha}}{\Gamma(1-\alpha)}\frac{\mathrm{d}w}{\mathrm{d}s}(s)\mathrm{d}s,\ w\in W^{1,1}(0,T).

Here Γ()\Gamma(\cdot) denotes the gamma function. For a consistent treatment of nonlinear time-fractional diffusion equations, we extend the classical Caputo derivative dtα\mathrm{d}^{\alpha}_{t} as follows. First of all, we define the Sobolev-Slobodeckij space Hα(0,T)H^{\alpha}(0,T) with the norm Hα(0,T)\|\cdot\|_{H^{\alpha}(0,T)} for 0<α<10<\alpha<1:

wHα(0,T):=(wL2(0,T)+0T0T|w(t)w(s)|2|ts|1+2αdtds)12\|w\|_{H^{\alpha}(0,T)}:=\left(\|w\|_{L^{2}(0,T)}+\int_{0}^{T}\int_{0}^{T}\frac{|w(t)-w(s)|^{2}}{|t-s|^{1+2\alpha}}\mathrm{d}t\mathrm{d}s\right)^{\frac{1}{2}}

(e.g., AdamsAd ). Furthermore, we set H0(0,T):=L2(0,T)H^{0}(0,T):=L^{2}(0,T) and

Hα(0,T):={Hα(0,T),0<α<12,{wH12(0,T);0T|w(t)|2tdt<},α=12,{wHα(0,T);w(0)=0},12<α1H_{\alpha}(0,T):=\left\{\begin{array}[]{ll}H^{\alpha}(0,T),&0<\alpha<\frac{1}{2},\\ \left\{w\in H^{\frac{1}{2}}(0,T);\int_{0}^{T}\frac{|w(t)|^{2}}{t}\mathrm{~{}d}t<\infty\right\},&\alpha=\frac{1}{2},\\ \left\{w\in H^{\alpha}(0,T);w(0)=0\right\},&\frac{1}{2}<\alpha\leq 1\end{array}\right.

with the norms defined by

wHα(0,T):={wHα(0,T),α12,(wH12(0,T)2+0T|w(t)|2tdt)12,α=12.\|w\|_{H_{\alpha}(0,T)}:=\left\{\begin{array}[]{ll}\|w\|_{H^{\alpha}(0,T)},&\alpha\neq\frac{1}{2},\\ \left(\|w\|_{H^{\frac{1}{2}(0,T)}}^{2}+\int_{0}^{T}\frac{|w(t)|^{2}}{t}\mathrm{~{}d}t\right)^{\frac{1}{2}},&\alpha=\frac{1}{2}.\end{array}\right.

Moreover, for β>0\beta>0, we set

Jβw(t):=0t(ts)β1Γ(β)w(s)ds, 0<t<T,wL1(0,T).J^{\beta}w(t):=\int^{t}_{0}\frac{(t-s)^{\beta-1}}{\Gamma(\beta)}w(s)\mathrm{d}s,\ 0<t<T,w\in L^{1}(0,T).

It was proved, for instance, in Gorenflo, Luchko, and YamamotoGLY , that the operator Jα:L2(0,T)Hα(0,T)J^{\alpha}:L^{2}(0,T)\to H_{\alpha}(0,T) is an isomorphism for α(0,1)\alpha\in(0,1).

Now we are ready to give the definition of the extended Caputo derivative

tα:=(Jα)1,𝒟(tα)=Hα(0,T).\partial_{t}^{\alpha}:=\left(J^{\alpha}\right)^{-1},\ \mathcal{D}(\partial_{t}^{\alpha})=H_{\alpha}(0,T).

Henceforth 𝒟()\mathcal{D}(\cdot) represents the domain of an operator under consideration. It has been demonstrated that tα\partial_{t}^{\alpha} represents the minimal closed extension of dtα\mathrm{d}^{\alpha}_{t}, where 𝒟(dtα):={vC1[0,T];v(0)=0}\mathcal{D}(\mathrm{d}^{\alpha}_{t}):=\{v\in C^{1}[0,T];v(0)=0\}. The relation tαw=dtαw\partial_{t}^{\alpha}w=\mathrm{d}^{\alpha}_{t}w is maintained for wC1[0,T]w\in C^{1}[0,T] with w(0)=0w(0)=0. For further details, we can refer to Gorenflo et al. GLY and YamamotoY4 .

Now we will define initial condition for problem (2.1) as follows:

u(x,)a(x)Hα(0,T),v(x,)b(x),Hα(0,T) for almost all xΩu(x,\cdot)-a(x)\in H_{\alpha}(0,T),v(x,\cdot)-b(x),\in H_{\alpha}(0,T)\ \text{ for almost all }x\in\Omega (2.2)

and write down a complete formulation of an initial-boundary value problem for the nonlinear system of time-fractional equations (1.1):

{tα(ua)=Δu+f(u,v),xΩ,0<t<T,tα(vb)=Δv+g(u,v),xΩ,0<t<T,νu=νv=0, on Ω×(0,T),u(x,)a(x),v(x,)b(x)Hα(0,T), for almost all xΩ.\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u-a)=\Delta u+f(u,v),\quad x\in\Omega,0<t<T,\\ \partial_{t}^{\alpha}(v-b)=\Delta v+g(u,v),\quad x\in\Omega,0<t<T,\\ \partial_{\nu}u=\partial_{\nu}v=0,\quad\text{ on }\partial\Omega\times(0,T),\\ u(x,\cdot)-a(x),v(x,\cdot)-b(x)\in H_{\alpha}(0,T),\quad\text{ for almost all }x\in\Omega.\end{array}\right. (2.3)

It is worth mentioning that the terms tα(ua)\partial_{t}^{\alpha}(u-a) and tα(vb)\partial_{t}^{\alpha}(v-b) in the first two lines of (2.3) are well-defined for almost all xΩx\in\Omega and 0<t<T0<t<T, due to the inclusion formulated in the last line of (2.3). Especially for 12<α<1,\frac{1}{2}<\alpha<1, noting that wHα(0,T)w\in H_{\alpha}(0,T) implies w(0)=0w(0)=0 by the trace theorem, we can understand that the left-hand side means that u(x,0)=a(x)u(x,0)=a(x) and v(x,0)=b(x)v(x,0)=b(x) in the trace sense with respect to tt. While for α<12\alpha<\frac{1}{2}, we do not need any initial conditions.

Now we are well prepared to investigate the initial-boundary value problem (2.3). We first provide the definition of weak solutions for equations (2.3). Henceforth we set

Q:=Ω×(0,T)Q:=\Omega\times(0,T)

and

(tα)ψ(x,s):=1Γ(1α)sT(ts)αψt(x,t)dt(\partial_{t}^{\alpha})^{*}\psi(x,s):=\frac{-1}{\Gamma(1-\alpha)}\int^{T}_{s}(t-s)^{-\alpha}\frac{\partial\psi}{\partial t}(x,t)\mathrm{d}t

for ψC1([0,T];L1(Ω))\psi\in C^{1}([0,T];L^{1}(\Omega)). Then:

Definition 1

We call (u,v)L1(0,T;L1(Ω))2(u,v)\in L^{1}(0,T;L^{1}(\Omega))^{2} a weak solution to (2.3) if

{Q(ua)(tα)ψdxdt=QuΔψdxdt+Qf(u,v)ψdxdt,Q(vb)(tα)ψdxdt=QvΔψdxdt+Qg(u,v)ψdxdt,\left\{\begin{array}[]{l}\int_{Q}(u-a)(\partial_{t}^{\alpha})^{*}\psi\,\mathrm{d}x\mathrm{d}t=\int_{Q}u\Delta\psi\,\mathrm{d}x\mathrm{d}t+\int_{Q}f(u,v)\psi\,\mathrm{d}x\mathrm{d}t,\cr\\ \int_{Q}(v-b)(\partial_{t}^{\alpha})^{*}\psi\,\mathrm{d}x\mathrm{d}t=\int_{Q}v\Delta\psi\,\mathrm{d}x\mathrm{d}t+\int_{Q}g(u,v)\psi\,\mathrm{d}x\mathrm{d}t,\end{array}\right. (2.4)

for all ψ(x,t)C(Q¯)\psi(x,t)\in C^{\infty}(\overline{Q}) satisfying νψ|Ω×(0,T)=0\partial_{\nu}\psi|_{\partial\Omega\times(0,T)}=0 and ψ(,T)=0\psi(\cdot,T)=0 in Ω\Omega.

If a weak solution (u,v)L1(0,T;L1(Ω))2(u,v)\in L^{1}(0,T;L^{1}(\Omega))^{2} is sufficiently smooth, then we can verify by the definition that (u,v)(u,v) satisfies (1.1) pointwise in the classical sense.

Based on the definition, we can give a lemma immediately.

Lemma 1

If (2.3) has a solution (u,v)L2(0,T;H2(Ω))2(u,v)\in L^{2}(0,T;H^{2}(\Omega))^{2} such that ua,vbHα(0,T;L2(Ω))u-a,v-b\in H_{\alpha}(0,T;L^{2}(\Omega)), then (u,v)(u,v) is a weak solution.

Proof

Since (u,v)(u,v) is a solution of (2.3), we have

{tα(ua)=Δu+f(u,v),tα(vb)=Δv+g(u,v).\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u-a)=\Delta u+f(u,v),\\ \partial_{t}^{\alpha}(v-b)=\Delta v+g(u,v).\end{array}\right. (2.5)

Multiply both sides of (2.5) by ψ(x,t)C(Q¯)\psi(x,t)\in C^{\infty}(\overline{Q}), taking into account the boundary condition of ψ(x,t)\psi(x,t) in Definition 1, and integrate by parts. This leads us to the same formula as (2.4), thus completing the proof.

Assumption 1

Let f,g:2f,g:\mathbb{R}^{2}\longrightarrow\mathbb{R} be local Lipschitz continuous. More precisely, for arbitrarily given M>0M>0, there exists a constant C=CM>0C=C_{M}>0 such that

|f(ξ,η)f(ξ^,η^)|+|g(ξ,η)g(ξ^,η^)|C(|ξξ^|+|ηη^|).|f(\xi,\eta)-f(\widehat{\xi},\widehat{\eta})|+|g(\xi,\eta)-g(\widehat{\xi},\widehat{\eta})|\leq C(|\xi-\widehat{\xi}|+|\eta-\widehat{\eta}|). (2.6)

for all ξ,η,ξ^,η^[M,M]\xi,\eta,\widehat{\xi},\widehat{\eta}\in[-M,M]. Moreover we assume

f(0,η)=g(ξ,0)=0, for all ξ0 and η0.f(0,\eta)=g(\xi,0)=0,\quad\mbox{ for all }\xi\geq 0\mbox{ and }\eta\geq 0.

Finally we assume that we can find a constant λ>0\lambda>0 such that

f(ξ,η)+λg(ξ,η)0 for all ξ,η0.f(\xi,\eta)+\lambda g(\xi,\eta)\leq 0\quad\mbox{ for all }\xi,\eta\geq 0.

Here, compared to the conditions required in Pi , we impose stronger requirements on ff and gg, which are crucial for the convergence of the truncated non-linear terms in Sections 4. Unlike Pi , we no longer restrict λ\lambda to be less than or equal to 1.

Now we state the first main result in this article, which validates the well-posedness for given T>0T>0 of the initial-boundary value problem (2.3).

Theorem 2.1

In (2.3), we assume a,b0a,b\geq 0, a,bH1(Ω)L(Ω)a,b\in H^{1}(\Omega)\cap L^{\infty}(\Omega), with f,gf,g satisfy Assumption 1. Then, for arbitrarily given T>0T>0, there exists at least one weak solution (u,v)L1(Ω×(0,T))2(u,v)\in L^{1}(\Omega\times(0,T))^{2} such that u,v0u,v\geq 0 in QQ.

We do not know the uniqueness of weak solution under our assumption, which is the same as for the case α=1\alpha=1.

Before starting with the proof of Theorem 2.1, we introduce some notations and derive several results necessary for the proof.

For an arbitrary constant p0>0p_{0}>0, define an elliptic operator AA as follows:

{(Av)(x):=Δv(x)p0v(x),xΩ,𝒟(A)={vH2(Ω);νv=0 on Ω}.\left\{\begin{array}[]{l}\left(-Av\right)(x):=\Delta v(x)-p_{0}v(x),\quad x\in\Omega,\\ \mathcal{D}(A)=\left\{v\in H^{2}(\Omega);\partial_{\nu}v=0\text{ on }\partial\Omega\right\}.\end{array}\right. (2.7)

Henceforth, by \|\cdot\| and (,)(\cdot,\cdot) we denote the standard norm and the scalar product in L2(Ω)L^{2}(\Omega), respectively. It is well-konwn that the operator AA is self-adjoint in L2(Ω)L^{2}(\Omega). Moreover, for a sufficiently large constant p0>0p_{0}>0, we can verify that AA is positive definiteLY1 . Therefore, by choosing a constant p0p_{0} large enough, the spectrum of the operator AA is comprised of discrete positive eigenvalues, herein represented as 0<λ1λ20<\lambda_{1}\leq\lambda_{2}\leq\cdots, each uniquely designated by its multiplicity. Additionally, λn\lambda_{n}\to\infty as nn\to\infty. Set φn\varphi_{n} be the eigenfunction corresponding to the eigenvalue λn\lambda_{n} such that Aφn=λnφnA\varphi_{n}=\lambda_{n}\varphi_{n}, and (φi,φj)=δij(\varphi_{i},\varphi_{j})=\delta_{ij}. Then the sequence {φn}n\{\varphi_{n}\}_{n\in\mathbb{N}} is orthonormal basis in L2(Ω)L^{2}(\Omega). For any γ>0\gamma>0, we can define the fractional power AγA^{\gamma} of the operator AA by the following relation (see, e.g.,Pa ):

Aγv=n=1λnγ(v,φn)φn,A^{\gamma}v=\sum_{n=1}^{\infty}\lambda^{\gamma}_{n}(v,\varphi_{n})\varphi_{n},

where

v𝒟(Aγ):={vL2(Ω):n=1λn2γ(v,φn)2<}v\in\mathcal{D}(A^{\gamma}):=\left\{v\in L^{2}(\Omega):\sum_{n=1}^{\infty}\lambda_{n}^{2\gamma}(v,\varphi_{n})^{2}<\infty\right\}

and

Aγv=(n=1λn2γ(v,φn)2)12.\|A^{\gamma}v\|=\left(\sum_{n=1}^{\infty}\lambda_{n}^{2\gamma}(v,\varphi_{n})^{2}\right)^{\frac{1}{2}}.

We have 𝒟(Aγ)H2γ(Ω)\mathcal{D}(A^{\gamma})\subset H^{2\gamma}(\Omega) for γ>0\gamma>0. Since 𝒟(Aγ)L2(Ω)\mathcal{D}(A^{\gamma})\subset L^{2}(\Omega), identifying the dual (L2(Ω))(L^{2}(\Omega))^{\prime} with itself, we have 𝒟(Aγ)L2(Ω)𝒟(Aγ)\mathcal{D}(A^{\gamma})\subset L^{2}(\Omega)\subset\mathcal{D}(A^{\gamma})^{\prime}. Henceforth we set 𝒟(Aγ)=𝒟(Aγ)\mathcal{D}(A^{-\gamma})=\mathcal{D}(A^{\gamma})^{\prime}, which consists of bounded linear functionals on 𝒟(Aγ)\mathcal{D}(A^{\gamma}). For w𝒟(Aγ)w\in\mathcal{D}(A^{-\gamma}) and v𝒟(Aγ)v\in\mathcal{D}(A^{\gamma}), by w,vγγ{}_{-\gamma}\langle w,v\rangle_{\gamma}, we denote the value which is obtained by operating ww to vv. We note that 𝒟(Aγ)\mathcal{D}(A^{-\gamma}) is a Hilbert space with norm:

w𝒟(Aγ)={n=1λn2γ|w,vγγ|2}12.\|w\|_{\mathcal{D}(A^{-\gamma})}=\left\{\sum_{n=1}^{\infty}\lambda^{-2\gamma}_{n}\left|{}_{-\gamma}\langle w,v\rangle_{\gamma}\right|^{2}\right\}^{\frac{1}{2}}.

We further note that

w,vγγ=(w,v) if wL2(Ω) and v𝒟(Aγ).{}_{-\gamma}\langle w,v\rangle_{\gamma}=(w,v)\text{ if }w\in L^{2}(\Omega)\text{ and }v\in\mathcal{D}(A^{\gamma}).

Moreover we define two other operators S(t)S(t) and K(t)K(t) as:

S(t)a=n=1Eα,1(λntα)(a,φn)φn,aL2(Ω),t>0S(t)a=\sum_{n=1}^{\infty}E_{\alpha,1}(-\lambda_{n}t^{\alpha})(a,\varphi_{n})\varphi_{n},\ a\in L^{2}(\Omega),t>0

and

K(t)a=n=1tα1Eα,α(λntα)(a,φn)φn,aL2(Ω),t>0,K(t)a=\sum_{n=1}^{\infty}t^{\alpha-1}E_{\alpha,\alpha}(-\lambda_{n}t^{\alpha})(a,\varphi_{n})\varphi_{n},\ a\in L^{2}(\Omega),t>0,

where Eα,β(z)E_{\alpha,\beta}(z) denotes the Mittag-Leffler function defined by a convergent series as follows:

Eα,β(z)=k=0zkΓ(αk+β),α>0,β,z.E_{\alpha,\beta}(z)=\sum_{k=0}^{\infty}\frac{z^{k}}{\Gamma(\alpha k+\beta)},\ \alpha>0,\beta\in\mathbb{C},z\in\mathbb{C}.

It follows directly from the definitions given above that AγK(t)a=K(t)AγaA^{\gamma}K(t)a=K(t)A^{\gamma}a and AγS(t)a=S(t)AγaA^{\gamma}S(t)a=S(t)A^{\gamma}a for a𝒟(Aγ)a\in\mathcal{D}(A^{\gamma}). Moreover, the following norm estimates were proved inSY :

AγS(t)a\displaystyle\|A^{\gamma}S(t)a\| Ctαγa,\displaystyle\leq Ct^{-\alpha\gamma}\|a\|, (2.8)
AγK(t)a\displaystyle\|A^{\gamma}K(t)a\| Ctα(1γ)1a,aL2(Ω),t>0,0γ1.\displaystyle\leq Ct^{\alpha(1-\gamma)-1}\|a\|,\ a\in L^{2}(\Omega),t>0,0\leq\gamma\leq 1.

3 Key Lemma

For the proof of Theorem 2.1, we show several lemmas.

Lemma 2

(Non-negativity) We assume that cL(Q)c\in L^{\infty}(Q), aH1(Ω)a\in H^{1}(\Omega), a0a\geq 0 in Ω\Omega and FL2(Q)F\in L^{2}(Q), F0F\geq 0 in QQ. Let uL2(0,T;H2(Ω))u\in L^{2}(0,T;H^{2}(\Omega)) satisfy (3.1) and uaHα(0,T;L2(Ω))u-a\in H_{\alpha}(0,T;L^{2}(\Omega)):

{tα(ua)=Δu+c(x,t)u+F(x,t),xΩ,0<t<T,νu=0, on Ω,0<t<T,u(x,)a(x)Hα(0,T), for almost all xΩ.\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u-a)=\Delta u+c(x,t)u+F(x,t),\quad x\in\Omega,0<t<T,\\ \partial_{\nu}u=0,\quad\text{ on }\partial\Omega,0<t<T,\\ u(x,\cdot)-a(x)\in H_{\alpha}(0,T),\quad\text{ for almost all }x\in\Omega.\end{array}\right. (3.1)

Then u0u\geq 0 in QQ.

The proof is simialr to Luchko and Yamamoto LY ; LY1 where the coefficient cc is assumed to be more smooth than L(Q)L^{\infty}(Q). For the case cL(Q)c\in L^{\infty}(Q), approximating cc by functions in C0(Q)C^{\infty}_{0}(Q) in the space Lκ(Q)L^{\kappa}(Q) with arbitrarily fixed κ>1\kappa>1. For completeness, we provide the proof in Appendix.

Here and henceforth we set A=Δ+p0A=-\Delta+p_{0}. For w0H1(Ω)w_{0}\in H^{1}(\Omega) and FL2(Q)F\in L^{2}(Q), we consider

{tα(ww0)+Aw=F(x,t),xΩ,0<t<T,νw=0,xΩ,0<t<T,w(x,)w0(x)Hα(0,T), for almost all xΩ.\left\{\begin{array}[]{l}\partial_{t}^{\alpha}\left(w-w_{0}\right)+Aw=F(x,t),\quad x\in\Omega,0<t<T,\\ \partial_{\nu}w=0,\quad x\in\partial\Omega,0<t<T,\\ w(x,\cdot)-w_{0}(x)\in H_{\alpha}(0,T),\quad\text{ for almost all }x\in\Omega.\end{array}\right. (3.2)

Then we know (e.g., KRY ) that there exists a unique solution wL2(0,T;H2(Ω))w\in L^{2}(0,T;H^{2}(\Omega)) to (3.2) satisfying ww0Hα(0,T;L2(Ω))w-w_{0}\in H_{\alpha}(0,T;L^{2}(\Omega)). By w(w0,F)w(w_{0},F) we denote the solution.

We can prove a compactness property:

Lemma 3

Let M1>0M_{1}>0 be arbitrarily chosen constant. Then a set

{w(w0n,Fn)|(w0n,Fn)H1(Ω)×L2(Q),w0nL1(Ω)+FnL1(Q)M1}\left\{w(w_{0}^{n},F_{n})|(w_{0}^{n},F_{n})\in H^{1}(\Omega)\times L^{2}(Q),\ \|w_{0}^{n}\|_{L^{1}(\Omega)}+\|F_{n}\|_{L^{1}(Q)}\leq M_{1}\right\}

contains a subsequence which converges in L1(Q)L^{1}(Q).

Proof

We introduce an operator τh(f)\tau_{h}(f) for h>0h>0 and fL1(0,T;X)f\in L^{1}(0,T;X) with Banach space XX by τh(f)(t)=f(t+h)\tau_{h}(f)(t)=f(t+h).

We show two lemmas for the proof of Lemma 3.

Lemma 4

(SimonSi ) Let XBYX\subset B\subset Y be Banach spaces and the embedding XYX\to Y be compact. We assume

vBCvX1θvYθ,vXY, where 0θ1,\|v\|_{B}\leq C\|v\|^{1-\theta}_{X}\|v\|^{\theta}_{Y},\quad\forall v\in X\cap Y,\quad\text{ where }0\leq\theta\leq 1,
 a set U is bounded in L1(0,T;X)\mbox{ a set }U\mbox{ is bounded in }L^{1}(0,T;X)

and

τhuuL1(0,Th;Y)0 as h0 uniformly for uU.\|\tau_{h}u-u\|_{L^{1}(0,T-h;Y)}\to 0\mbox{ as }h\to 0\mbox{ uniformly for }u\in U.

Then UU is relatively compact in L1(0,T;B)L^{1}(0,T;B).

Lemma 5

Let Ωn\Omega\in\mathbb{R}^{n}, if γ>d4\gamma>\frac{d}{4}, then we have

AγwL2(Ω)CwL1(Ω).\|A^{-\gamma}w\|_{L^{2}(\Omega)}\leq C\|w\|_{L^{1}(\Omega)}.

Proof of Lemma 5

Since 𝒟(Aγ)H2γ(Ω)L(Ω)\mathcal{D}(A^{\gamma})\subset H^{2\gamma}(\Omega)\subset L^{\infty}(\Omega), we can get vL(Ω)CAγvL2(Ω)\|v\|_{L^{\infty}(\Omega)}\leq C\|A^{\gamma}v\|_{L^{2}(\Omega)}. Moreover, we have

AγwL2(Ω)\displaystyle\|A^{-\gamma}w\|_{L^{2}(\Omega)} =supφL2(Ω)=1|(Aγw,φ)L2(Ω)|\displaystyle=\sup_{\|\varphi\|_{L^{2}(\Omega)}=1}|(A^{-\gamma}w,\varphi)_{L^{2}(\Omega)}|
=supφL2(Ω)=1|(w,Aγφ)L2(Ω)|\displaystyle=\sup_{\|\varphi\|_{L^{2}(\Omega)}=1}|(w,A^{-\gamma}\varphi)_{L^{2}(\Omega)}|
CwL1(Ω)AγφL(Ω)\displaystyle\leq C\|w\|_{L^{1}(\Omega)}\|A^{-\gamma}\varphi\|_{L^{\infty}(\Omega)}
CwL1(Ω)AγAγφL2(Ω)\displaystyle\leq C\|w\|_{L^{1}(\Omega)}\|A^{\gamma}A^{-\gamma}\varphi\|_{L^{2}(\Omega)}
CwL1(Ω).\displaystyle\leq C\|w\|_{L^{1}(\Omega)}.

Then we finished the proof. \blacksquare

Now we proceed to the proof of Lemma 3. In Lemma 4, set X=𝒟(A1δ),Y=𝒟(Aδ)X=\mathcal{D}(A^{1-\delta}),Y=\mathcal{D}(A^{-\delta}) and B=L1(Ω)B=L^{1}(\Omega). The intermediate spaces BL between XX and YY are given by

[X,Y]θ=𝒟(A1δθ),θ[0,1].[X,Y]_{\theta}=\mathcal{D}(A^{1-\delta-\theta}),\quad\forall\theta\in[0,1].

If we choose θ=1δ\theta=1-\delta, we obtain [X,Y]1δ=𝒟(A0)=L2(Ω)[X,Y]_{1-\delta}=\mathcal{D}(A^{0})=L^{2}(\Omega). Referring to the interpolation result of Sobolev spaces, this implies

wL2(Ω)Cw𝒟(A1δ)δw𝒟(Aδ)1δ.\|w\|_{L^{2}(\Omega)}\leq C\|w\|_{\mathcal{D}(A^{1-\delta})}^{\delta}\|w\|_{\mathcal{D}(A^{-\delta})}^{1-\delta}.

Applying the embedding theory of Lp(Ω)L^{p}(\Omega) for p[1,]p\in[1,\infty], we have wL1(Ω)CwL2(Ω)\|w\|_{L^{1}(\Omega)}\leq C\|w\|_{L^{2}(\Omega)}, which means

wL1(Ω)Cw𝒟(A1δ)δw𝒟(Aδ)1δ.\|w\|_{L^{1}(\Omega)}\leq C\|w\|_{\mathcal{D}(A^{1-\delta})}^{\delta}\|w\|_{\mathcal{D}(A^{-\delta})}^{1-\delta}.

In order to apply Lemma 4, we still need to prove the following inclusion relation:

𝒟(A1δ)L1(Ω)𝒟(Aδ).\mathcal{D}(A^{1-\delta})\subset L^{1}(\Omega)\subset\mathcal{D}(A^{-\delta}). (3.3)

To establish (3.3), we start by considering the following relationship:

𝒟(Aδ)H2δ(Ω)H2δ(Ω)L(Ω),\mathcal{D}(A^{\delta})\subset H^{2\delta}(\Omega)\subset H^{2\delta^{\prime}}(\Omega)\subset L^{\infty}(\Omega),

where d4<δ<δ\frac{d}{4}<\delta^{\prime}<\delta. The Kondrachov embedding theorem implies H2δ(Ω)H2δ(Ω)H^{2\delta}(\Omega)\subset H^{2\delta^{\prime}}(\Omega) is compact. Consequently, we obtain that 𝒟(Aδ)L(Ω)\mathcal{D}(A^{\delta})\subset L^{\infty}(\Omega) is compact. By the definition, L1(Ω)(L(Ω))(𝒟(Aδ))=𝒟(Aδ)L^{1}(\Omega)\subset\left(L^{\infty}(\Omega)\right)^{\prime}\subset\left(\mathcal{D}(A^{\delta})\right)^{\prime}=\mathcal{D}(A^{-\delta}). Here XX^{\prime} denotes the dual space of a Banach space XX. Hence, we have proved (3.3), and the mapping 𝒟(A1δ)𝒟(Aδ)\mathcal{D}(A^{1-\delta})\to\mathcal{D}(A^{-\delta}) is compact.

We set wn:=w(w0n,Fn)w_{n}:=w(w_{0}^{n},F_{n}) for each nn\in\mathbb{N}. In order to prove Lemma 3, in terms of Lemma 4, it is sufficient to show that wnL1(0,T;𝒟(A1δ))\|w_{n}\|_{L^{1}(0,T;\mathcal{D}(A^{1-\delta}))} is bounded and τhwnwnL1(0,T;𝒟(Aδ))0\|\tau_{h}w_{n}-w_{n}\|_{L^{1}(0,T;\mathcal{D}(A^{-\delta}))}\to 0 as h0h\to 0 uniformly for each nn\in\mathbb{N}.

Using the operator S(t)S(t) and K(t)K(t), the solution can be expressed by the formulaLY1 ; LY2 :

wn(t)=S(t)w0n+0tK(ts)Fn(s)ds.w_{n}(t)=S(t)w_{0}^{n}+\int_{0}^{t}K(t-s)F_{n}(s)\mathrm{d}s. (3.4)

Writing (3.4) as

wn(t)=S(t)w0n+0tAγK(ts)AγFn(s)ds.w_{n}(t)=S(t)w_{0}^{n}+\int_{0}^{t}A^{\gamma}K(t-s)A^{-\gamma}F_{n}(s)\mathrm{d}s.

Then, multiplying both sides by A1δA^{1-\delta}, we obtain

A1δwn(t)=AS(t)Aδw0n+0tA1δ+γK(ts)AγFn(s)ds,A^{1-\delta}w_{n}(t)=AS(t)A^{-\delta}w_{0}^{n}+\int_{0}^{t}A^{1-\delta+\gamma}K(t-s)A^{-\gamma}F_{n}(s)\mathrm{d}s, (3.5)

where δ>γ>d4\delta>\gamma>\frac{d}{4}. Applying the norm estimates (2.8), we have A1δ+γK(ts)(ts)α(δγ)1\|A^{1-\delta+\gamma}K(t-s)\|\leq(t-s)^{\alpha(\delta-\gamma)-1} and AS(t)tα\|AS(t)\|\leq t^{-\alpha}. By combining this with Lemma 5, we can derive

A1δwn(t)L2(Ω)C0tsαw0nL1(Ω)ds+C0t(ts)α(δγ)1Fn(s)L1(Ω)ds.\left\|A^{1-\delta}w_{n}(t)\right\|_{L^{2}(\Omega)}\leq C\int_{0}^{t}s^{-\alpha}\|w_{0}^{n}\|_{L^{1}(\Omega)}\mathrm{d}s+C\int_{0}^{t}(t-s)^{\alpha(\delta-\gamma)-1}\|F_{n}(s)\|_{L^{1}(\Omega)}\mathrm{d}s.

Furthermore, applying Young’s inequality, we reach

A1δwnL1(0,T;L2(Ω))C(w0nL1(Ω)+FnL1(0,T;L1(Ω))).\left\|A^{1-\delta}w_{n}\right\|_{L^{1}(0,T;L^{2}(\Omega))}\leq C\left(\|w_{0}^{n}\|_{L^{1}(\Omega)}+\|F_{n}\|_{L^{1}(0,T;L^{1}(\Omega))}\right).

Multiply (3.2) by AδA^{-\delta}, and we have

Aδtα(wnw0n)=A1δwn+AδFn.A^{-\delta}\partial_{t}^{\alpha}(w_{n}-w_{0}^{n})=-A^{1-\delta}w_{n}+A^{-\delta}F_{n}.

Applying (3.5) and Lemma 5 again, we can obtain

Aδtα(wnw0n)L1(0,T;L2(Ω))\displaystyle\|A^{-\delta}\partial_{t}^{\alpha}(w_{n}-w_{0}^{n})\|_{L^{1}(0,T;L^{2}(\Omega))}
\displaystyle\leq C(w0nL1(Ω)+FnL1(0,T;L1(Ω)))+AδFnL1(0,T;L2(Ω))\displaystyle C\left(\|w_{0}^{n}\|_{L^{1}(\Omega)}+\|F_{n}\|_{L^{1}(0,T;L^{1}(\Omega))}\right)+\|A^{-\delta}F_{n}\|_{L^{1}(0,T;L^{2}(\Omega))}
\displaystyle\leq C(w0nL1(Ω)+FnL1(0,T;L1(Ω))).\displaystyle C\left(\|w_{0}^{n}\|_{L^{1}(\Omega)}+\|F_{n}\|_{L^{1}(0,T;L^{1}(\Omega))}\right).

Here we have obtained

A1δwnL1(0,T;L2(Ω))\displaystyle\|A^{1-\delta}w_{n}\|_{L^{1}(0,T;L^{2}(\Omega))} C(w0nL1(Ω)+FnL1(0,T;L1(Ω))),\displaystyle\leq C\left(\|w_{0}^{n}\|_{L^{1}(\Omega)}+\|F_{n}\|_{L^{1}(0,T;L^{1}(\Omega))}\right), (3.6)
Aδtα(wnw0n)L1(0,T;L2(Ω))\displaystyle\|A^{-\delta}\partial_{t}^{\alpha}(w_{n}-w_{0}^{n})\|_{L^{1}(0,T;L^{2}(\Omega))} C(w0nL1(Ω)+FnL1(0,T;L1(Ω))).\displaystyle\leq C\left(\|w_{0}^{n}\|_{L^{1}(\Omega)}+\|F_{n}\|_{L^{1}(0,T;L^{1}(\Omega))}\right).

To utilize the Lemma 4, we now need to prove:

wn(t+h)wn(t)L1(0,Th;𝒟(Aδ))0 as h0.\|w_{n}(t+h)-w_{n}(t)\|_{L^{1}(0,T-h;\mathcal{D}(A^{-\delta}))}\longrightarrow 0\quad\mbox{ as }h\to 0.

Denote w~n:=tα(Aδ(wnw0n))L1(0,T;L2(Ω))\widetilde{w}_{n}:=\partial_{t}^{\alpha}(A^{-\delta}(w_{n}-w_{0}^{n}))\in L^{1}(0,T;L^{2}(\Omega)). Applying the fractional integral JαJ^{\alpha} to both sides, we have w~n:=Aδ(wnw0n)=Jαw~n\widetilde{w}_{n}:=A^{-\delta}(w_{n}-w_{0}^{n})=J^{\alpha}\widetilde{w}_{n}. Noticing that wn(t+h)wn(t)=(wn(t+h)w0n)(wn(t)w0n)w_{n}(t+h)-w_{n}(t)=(w_{n}(t+h)-w_{0}^{n})-(w_{n}(t)-w_{0}^{n}), and

wn(t+h)wn(t)L1(0,Th;𝒟(Aδ))=Jαw~n(t+h)Jαw~n(t)L1(0,Th;L2(Ω)).\|w_{n}(t+h)-w_{n}(t)\|_{L^{1}(0,T-h;\mathcal{D}(A^{-\delta}))}=\|J^{\alpha}\widetilde{w}_{n}(t+h)-J^{\alpha}\widetilde{w}_{n}(t)\|_{L^{1}(0,T-h;L^{2}(\Omega))}.

Now, we represent Jαw~n(t+h)Jαw~n(t)J^{\alpha}\widetilde{w}_{n}(t+h)-J^{\alpha}\widetilde{w}_{n}(t):

Jαw~n(t+h)Jαw~n(t)\displaystyle J^{\alpha}\widetilde{w}_{n}(t+h)-J^{\alpha}\widetilde{w}_{n}(t) (3.7)
=\displaystyle= 1Γ(α)0t(t+hs)α1w~n(s)ds1Γ(α)0t(ts)α1w~n(s)ds\displaystyle\frac{1}{\Gamma(\alpha)}\int_{0}^{t}(t+h-s)^{\alpha-1}\widetilde{w}_{n}(s)\mathrm{d}s-\frac{1}{\Gamma(\alpha)}\int_{0}^{t}(t-s)^{\alpha-1}\widetilde{w}_{n}(s)\mathrm{d}s
+\displaystyle+ tt+h(t+hs)α1w~n(s)ds.\displaystyle\int_{t}^{t+h}(t+h-s)^{\alpha-1}\widetilde{w}_{n}(s)\mathrm{d}s.

Let

I(t):=0t((t+hs)α1(ts)α1)w~n(s)ds,J(t):=tt+h(t+hs)α1w~n(s)ds.I(t):=\int_{0}^{t}((t+h-s)^{\alpha-1}-(t-s)^{\alpha-1})\widetilde{w}_{n}(s)\mathrm{d}s,\quad J(t):=\int_{t}^{t+h}(t+h-s)^{\alpha-1}\widetilde{w}_{n}(s)\mathrm{d}s.

As for the first term concerning I(t)I(t), we have

I(t)𝒟(Aδ)0t((ts)α1(t+hs)α1)w~n(s)𝒟(Aδ)ds.\|I(t)\|_{\mathcal{D}(A^{-\delta})}\leq\int_{0}^{t}\left((t-s)^{\alpha-1}-(t+h-s)^{\alpha-1}\right)\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}s.

Integrating both sides with respect to the time variable tt, we obtain the following estimate:

0ThI(t)𝒟(Aδ)dt\displaystyle\int_{0}^{T-h}\|I(t)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}t (3.8)
\displaystyle\leq C0Th((Ths)α(Ts)α+hα)w~n(s)𝒟(Aδ)ds\displaystyle C\int_{0}^{T-h}((T-h-s)^{\alpha}-(T-s)^{\alpha}+h^{\alpha})\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}s
\displaystyle\leq C0Thhαw~n(s)𝒟(Aδ)ds\displaystyle C\int_{0}^{T-h}h^{\alpha}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}s
\displaystyle\leq Chαw~nL1(0,T;𝒟(Aδ)).\displaystyle Ch^{\alpha}\|\widetilde{w}_{n}\|_{L^{1}(0,T;\mathcal{D}(A^{-\delta}))}.

After that, we start the estimate of J(t)J(t):

0ThJ(t)𝒟(Aδ)dt\displaystyle\int^{T-h}_{0}\|J(t)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}t (3.9)
\displaystyle\leq 0Thtt+h(t+hs)α1w~n(s)𝒟(Aδ)dsdt.\displaystyle\int^{T-h}_{0}\int_{t}^{t+h}(t+h-s)^{\alpha-1}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}s\mathrm{d}t.

Let the integrand (t+hs)α1w~n(s)𝒟(Aδ)(t+h-s)^{\alpha-1}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})} be denoted as b(s,t)b(s,t). Then, by Fubini’s theorem, we can exchange the orders of integration:

0Thtt+hb(s,t)dsdt=0h0sb(s,t)dtds+hThshsb(s,t)dtds+ThTshThb(s,t)dtds.\int^{T-h}_{0}\int_{t}^{t+h}b(s,t)\mathrm{d}s\mathrm{d}t=\int_{0}^{h}\int_{0}^{s}b(s,t)\mathrm{d}t\mathrm{d}s+\int_{h}^{T-h}\int_{s-h}^{s}b(s,t)\mathrm{d}t\mathrm{d}s+\int_{T-h}^{T}\int_{s-h}^{T-h}b(s,t)\mathrm{d}t\mathrm{d}s.

Then the estimate for J(t)J(t) can be divided into the following three parts:

J1\displaystyle J_{1} =0h0s(t+hs)α1w~n(s)𝒟(Aδ)dtds\displaystyle=\int_{0}^{h}\int_{0}^{s}(t+h-s)^{\alpha-1}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}t\mathrm{d}s (3.10)
=0hw~n(s)𝒟(Aδ)(0s(t+hs)α1dt)ds\displaystyle=\int_{0}^{h}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\left(\int_{0}^{s}(t+h-s)^{\alpha-1}\mathrm{d}t\right)\mathrm{d}s
Chαw~nL1((0,T);𝒟(Aδ))\displaystyle\leq Ch^{\alpha}\|\widetilde{w}_{n}\|_{L^{1}((0,T);\mathcal{D}(A^{-\delta}))}

and

J2\displaystyle J_{2} =hThshs(t+hs)α1w~(s)𝒟(Aδ)dtds\displaystyle=\int_{h}^{T-h}\int_{s-h}^{s}(t+h-s)^{\alpha-1}\|\widetilde{w}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}t\mathrm{d}s (3.11)
=hThw~n(s)𝒟(Aδ)(shs(t+hs)α1dt)ds\displaystyle=\int_{h}^{T-h}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\left(\int_{s-h}^{s}(t+h-s)^{\alpha-1}\mathrm{d}t\right)\mathrm{d}s
Chαw~nL1((0,T);𝒟(Aδ))\displaystyle\leq Ch^{\alpha}\|\widetilde{w}_{n}\|_{L^{1}((0,T);\mathcal{D}(A^{-\delta}))}

and

J3\displaystyle J_{3} =ThTshTh(t+hs)α1w~n(s)𝒟(Aδ)dtds\displaystyle=\int^{T}_{T-h}\int_{s-h}^{T-h}(t+h-s)^{\alpha-1}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\mathrm{d}t\mathrm{d}s (3.12)
=ThTw~n(s)𝒟(Aδ)(shTh(t+hs)α1dt)ds\displaystyle=\int_{T-h}^{T}\|\widetilde{w}_{n}(s)\|_{\mathcal{D}(A^{-\delta})}\left(\int_{s-h}^{T-h}(t+h-s)^{\alpha-1}\mathrm{d}t\right)\mathrm{d}s
Chαw~nL1((0,T);𝒟(Aδ)).\displaystyle\leq Ch^{\alpha}\|\widetilde{w}_{n}\|_{L^{1}((0,T);\mathcal{D}(A^{-\delta}))}.

Consequently,

τhwnwnL1(0,T;𝒟(Aδ))O(hα).\|\tau_{h}w_{n}-w_{n}\|_{L^{1}(0,T;\mathcal{D}(A^{-\delta}))}\leq O(h^{\alpha}).

Thus we have completed the proof of Lemma 3.

4 Completion of the proof of Theorem 2.1

Proof

Henceforth C>0,C1>0C>0,C_{1}>0, etc. denote generic constants which are independent of nn\in\mathbb{N} and the choice of variables ξ,η\xi,\eta. Henceforth when we write C=C0(n)C=C_{0}(n), etc., we understand that the constant depends on nn such as C0(n)C_{0}(n) above.

We begin by constructing approximating solutions unu_{n} and vnv_{n}. In order to handle the nonlinear terms ff and gg, we introduce the truncated functions below. We choose ψ1C0(2)\psi_{1}\in C_{0}^{\infty}(\mathbb{R}^{2}) such that 0ψ1(ξ,η)10\leq\psi_{1}(\xi,\eta)\leq 1 for all ξ,η\xi,\eta\in\mathbb{R} and

ψ1(ξ,η)={1, if |ξ|,|η|1,0, if |ξ|2 or |η|2.\psi_{1}(\xi,\eta)=\left\{\begin{array}[]{l}1,\quad\mbox{ if }|\xi|,|\eta|\leq 1,\\ 0,\quad\mbox{ if }|\xi|\geq 2\mbox{ or }|\eta|\geq 2.\end{array}\right.

Next, we set ψn(ξ,η)=ψ1(ξn,ηn)\psi_{n}(\xi,\eta)=\psi_{1}\left(\frac{\xi}{n},\frac{\eta}{n}\right). With this choice, 0ψn10\leq\psi_{n}\leq 1 for all nn\in\mathbb{N}, and ψn\psi_{n} tends pointwise to 1 as nn tends to \infty.

We define truncated nonlinear terms as follows

{fn(ξ,η)=ψn(ξ,η)f(ξ,η),ξ,η,gn(ξ,η)=ψn(ξ,η)g(ξ,η),ξ,η.\left\{\begin{array}[]{l}f_{n}(\xi,\eta)=\psi_{n}(\xi,\eta)f(\xi,\eta),\quad\xi,\eta\in\mathbb{R},\\ g_{n}(\xi,\eta)=\psi_{n}(\xi,\eta)g(\xi,\eta),\quad\xi,\eta\in\mathbb{R}.\end{array}\right.

Then we can verify that there exists a constant C0(n)>0C_{0}(n)>0 such that

{fnL(2),gnL(2)C0(n),kfnL(2),kgnL(2)C0(n),k=1,2\left\{\begin{array}[]{l}\|f_{n}\|_{L^{\infty}(\mathbb{R}^{2})},\,\|g_{n}\|_{L^{\infty}(\mathbb{R}^{2})}\leq C_{0}(n),\\ \|\partial_{k}f_{n}\|_{L^{\infty}(\mathbb{R}^{2})},\,\|\partial_{k}g_{n}\|_{L^{\infty}(\mathbb{R}^{2})}\leq C_{0}(n),\quad k=1,2\end{array}\right. (4.1)

and

fn(0,η)=gn(ξ,0)=0 for all ξ,η.f_{n}(0,\eta)=g_{n}(\xi,0)=0\quad\mbox{ for all }\xi,\eta\in\mathbb{R}. (4.2)

Indeed, (4.2) is verified directly. Moreover

|fn(r,s)|+|gn(r,s)|sup|ξ|,|η|2n|f(ξ,η)|+sup|ξ|,|η|2n|g(ξ,η)|,|f_{n}(r,s)|+|g_{n}(r,s)|\leq\sup_{|\xi|,|\eta|\leq 2n}|f(\xi,\eta)|+\sup_{|\xi|,|\eta|\leq 2n}|g(\xi,\eta)|,

for all r,sr,s\in\mathbb{R}, which means the first estimate in (4.1). The rest estimates in (4.1) follow from the definition of fn,gnf_{n},g_{n}. Furthermore we note that fn,gnf_{n},g_{n} satisfy Assumption 1.

A truncated form of equation (2.3) is given by:

{tα(una)=Aun+Fn(un(x,t),vn(x,t)),xΩ,0<t<T,tα(vnb)=Avn+Gn(un(x,t),vn(x,t)),xΩ,0<t<T,νun=νvn=0, on Ω×(0,T),un(x,)a(x)Hα(0,T),vn(x,)b(x)Hα(0,T), for almost all xΩ.\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u_{n}-a)=-Au_{n}+F_{n}(u_{n}(x,t),v_{n}(x,t)),\quad x\in\Omega,0<t<T,\\ \partial_{t}^{\alpha}(v_{n}-b)=-Av_{n}+G_{n}(u_{n}(x,t),v_{n}(x,t)),\quad x\in\Omega,0<t<T,\\ \partial_{\nu}u_{n}=\partial_{\nu}v_{n}=0,\quad\text{ on }\partial\Omega\times(0,T),\\ u_{n}(x,\cdot)-a(x)\in H_{\alpha}(0,T),\ v_{n}(x,\cdot)-b(x)\in H_{\alpha}(0,T),\quad\mbox{ for almost all }x\in\Omega.\end{array}\right. (4.3)

Here we recall that Aw=Δwp0w-Aw=\Delta w-p_{0}w, Q=Ω×(0,T)Q=\Omega\times(0,T), and for nn\in\mathbb{N} we set

Fn(un,vn):=fn(un,vn)+p0un,Gn(un,vn):=gn(un,vn)+p0vn.F_{n}(u_{n},v_{n}):=f_{n}(u_{n},v_{n})+p_{0}u_{n},\ G_{n}(u_{n},v_{n}):=g_{n}(u_{n},v_{n})+p_{0}v_{n}.

Now we can prove:

Lemma 6

Let a,bH1(Ω)L(Ω)a,b\in H^{1}(\Omega)\cap L^{\infty}(\Omega) and a,b0a,b\geq 0 in Ω\Omega.
(i) For nn\in\mathbb{N}, there exists a unique solution (un,vn)L2(0,T;H2(Ω))2(u_{n},v_{n})\in L^{2}(0,T;H^{2}(\Omega))^{2} such that una,vnbHα(0,T;L2(Ω))u_{n}-a,v_{n}-b\in H_{\alpha}(0,T;L^{2}(\Omega)).
(ii) un,vn0u_{n},v_{n}\geq 0 in QQ for each nn\in\mathbb{N}.

Proof of Lemma 6.

Since (4.1) implies the global Lipschitz contintuity of fn,gnf_{n},g_{n} over 2\mathbb{R}^{2} for each nn, in terms of a,bH1(Ω)L(Ω)a,b\in H^{1}(\Omega)\cap L^{\infty}(\Omega), we can apply a usual iteration method to establish the unique existence of (un,vn)L2(0,T;H2(Ω))2(u_{n},v_{n})\in L^{2}(0,T;H^{2}(\Omega))^{2} and una,vnbHα(0,T;L2(Ω))u_{n}-a,v_{n}-b\in H_{\alpha}(0,T;L^{2}(\Omega)), which complete the proof of Lemma 6 (i). Further details can be found, for example, in FLY ; LY1 ; LY2 .

Next we will prove (ii). First we can find functions Cn1(x,t),Cn2(x,t)L(Q)C^{1}_{n}(x,t),C^{2}_{n}(x,t)\in L^{\infty}(Q) such that

|fn(un(x,t),vn(x,t))|Cn1(x,t)un(x,t),\displaystyle|f_{n}(u_{n}(x,t),v_{n}(x,t))|\leq C^{1}_{n}(x,t)u_{n}(x,t), (4.4)
|gn(un(x,t),vn(x,t))|Cn1(x,t)vn(x,t), for almost all (x,t)Q.\displaystyle|g_{n}(u_{n}(x,t),v_{n}(x,t))|\leq C^{1}_{n}(x,t)v_{n}(x,t),\quad\mbox{ for almost all }(x,t)\in Q.

Actually, we can set Cn1(x,t)=sgn(un(x,t))1fnL(Q)C^{1}_{n}(x,t)=\operatorname{sgn}(u_{n}(x,t))\|\partial_{1}f_{n}\|_{L^{\infty}(Q)} and Cn2(x,t)=sgn(vn(x,t))2fnL(Q)C^{2}_{n}(x,t)=\operatorname{sgn}(v_{n}(x,t))\|\partial_{2}f_{n}\|_{L^{\infty}(Q)}, where

sgn(ξ):={1, if |ξ|0,1, if |ξ|<0.\operatorname{sgn}(\xi):=\left\{\begin{array}[]{r}1,\quad\mbox{ if }|\xi|\geq 0,\\ -1,\quad\mbox{ if }|\xi|<0.\end{array}\right.

Indeed, by (4.1) and (4.2), we apply the mean value theorem and obtain

|fn(un(x,t),vn(x,t))|=|fn(un(x,t),vn(x,t))fn(0,vn(x,t))|\displaystyle|f_{n}(u_{n}(x,t),v_{n}(x,t))|=|f_{n}(u_{n}(x,t),v_{n}(x,t))-f_{n}(0,v_{n}(x,t))|
\displaystyle\leq 1fnL(2)|un(x,t)|.\displaystyle\|\partial_{1}f_{n}\|_{L^{\infty}(\mathbb{R}^{2})}|u_{n}(x,t)|.

The proof for gng_{n} is similar and so the verification of (4.4) is complete.

Then we can rewrite (4.3) as

{tα(una)=ΔunCn1(x,t)un(x,t)+Wn(x,t),(x,t)Q,tα(vnb)=ΔvnCn2(x,t)vn(x,t)+Vn(x,t),(x,t)Q,νun=νvn=0, on Ω×(0,T),un(x,)a(x)Hα(0,T),vn(x,)b(x)Hα(0,T),\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(u_{n}-a)=\Delta u_{n}-C^{1}_{n}(x,t)u_{n}(x,t)+W_{n}(x,t),\quad(x,t)\in Q,\\ \partial_{t}^{\alpha}(v_{n}-b)=\Delta v_{n}-C_{n}^{2}(x,t)v_{n}(x,t)+V_{n}(x,t),\quad(x,t)\in Q,\\ \partial_{\nu}u_{n}=\partial_{\nu}v_{n}=0,\quad\text{ on }\partial\Omega\times(0,T),\\ u_{n}(x,\cdot)-a(x)\in H_{\alpha}(0,T),\quad v_{n}(x,\cdot)-b(x)\in H_{\alpha}(0,T),\end{array}\right. (4.5)

where we set

Wn(x,t):=Cn1(x,t)un(x,t)+fn(un(x,t),vn(x,t)),\displaystyle W_{n}(x,t):=C_{n}^{1}(x,t)u_{n}(x,t)+f_{n}(u_{n}(x,t),v_{n}(x,t)),
Vn(x,t):=Cn2(x,t)vn(x,t)+gn(un(x,t),vn(x,t)).\displaystyle V_{n}(x,t):=C_{n}^{2}(x,t)v_{n}(x,t)+g_{n}(u_{n}(x,t),v_{n}(x,t)).

We note that Wn,Vn0W_{n},V_{n}\geq 0 in QQ by (4.4). Therefore we apply Lemma 2 to unu_{n} and vnv_{n}, so that un,vn0u_{n},v_{n}\geq 0 in QQ. Thus the proof of Lemma 6 is complete. \blacksquare

Next, integrating the first two equations in (4.3) with respect to xx, we obtain

tαΩ(una)(x,t)dx\displaystyle\partial^{\alpha}_{t}\int_{\Omega}(u_{n}-a)(x,t)\mathrm{d}x =Ωfn(un(x,t),vn(x,t))dx,\displaystyle=\int_{\Omega}f_{n}(u_{n}(x,t),v_{n}(x,t))\mathrm{d}x, (4.6)
tαΩ(vnb)(x,t)dx\displaystyle\partial^{\alpha}_{t}\int_{\Omega}(v_{n}-b)(x,t)\mathrm{d}x =Ωgn(un(x,t),vn(x,t))dx.\displaystyle=\int_{\Omega}g_{n}(u_{n}(x,t),v_{n}(x,t))\mathrm{d}x.

Here we use the fact that ΩΔun(x,t)dx=ΩΔvn(x,t)dx=0\int_{\Omega}\Delta u_{n}(x,t)\mathrm{d}x=\int_{\Omega}\Delta v_{n}(x,t)\mathrm{d}x=0 because νun=νvn=0\partial_{\nu}u_{n}=\partial_{\nu}v_{n}=0 on Ω×(0,T)\partial\Omega\times(0,T). Summing up the above equations in unu_{n} and λvn\lambda v_{n}, we can obtain

tαΩ(una+λvnλb)(x,t)dx=Ωfn(un,vn)+λgn(un,vn)dx0.\partial_{t}^{\alpha}\int_{\Omega}(u_{n}-a+\lambda v_{n}-\lambda b)(x,t)\mathrm{d}x=\int_{\Omega}f_{n}(u_{n},v_{n})+\lambda g_{n}(u_{n},v_{n})\mathrm{d}x\leq 0. (4.7)

By applying the fractional integral to both sides, by un,vn0u_{n},v_{n}\geq 0 in QQ, we obtain

un(,t)L1(Ω)+vn(,t)L1(Ω)C(aL1(Ω)+bL1(Ω)).\|u_{n}(\cdot,t)\|_{L^{1}(\Omega)}+\|v_{n}(\cdot,t)\|_{L^{1}(\Omega)}\leq C(\|a\|_{L^{1}(\Omega)}+\|b\|_{L^{1}(\Omega)}). (4.8)

Integrating each equation of (4.3) with respect to xx and tt and using νun=νvn=0\partial_{\nu}u_{n}=\partial_{\nu}v_{n}=0 on Ω×(0,T)\partial\Omega\times(0,T), we derive

fn(un,vn)L1(0,T;L1(Ω))=J1α(una)L1(0,T;L1(Ω)),\displaystyle\|f_{n}(u_{n},v_{n})\|_{L^{1}(0,T;L^{1}(\Omega))}=\|J^{1-\alpha}(u_{n}-a)\|_{L^{1}(0,T;L^{1}(\Omega))}, (4.9)
gn(un,vn)L1(0,T;L1(Ω))=J1α(vnb)L1(0,T;L1(Ω)).\displaystyle\|g_{n}(u_{n},v_{n})\|_{L^{1}(0,T;L^{1}(\Omega))}=\|J^{1-\alpha}(v_{n}-b)\|_{L^{1}(0,T;L^{1}(\Omega))}.

Since J1αw(x,)L1(0,T)Cw(x,)L1(0,T)\|J^{1-\alpha}w(x,\cdot)\|_{L^{1}(0,T)}\leq C\|w(x,\cdot)\|_{L^{1}(0,T)} for wL1(0,T;L1(Ω))w\in L^{1}(0,T;L^{1}(\Omega)) by Young’s inequality, we see

J1α(una)L1(0,T;L1(Ω))+J1α(vnb)L1(0,T;L1(Ω))\displaystyle\|J^{1-\alpha}(u_{n}-a)\|_{L^{1}(0,T;L^{1}(\Omega))}+\|J^{1-\alpha}(v_{n}-b)\|_{L^{1}(0,T;L^{1}(\Omega))} (4.10)
\displaystyle\leq C(unaL1(0,T;L1(Ω))+vnbL1(0,T;L1(Ω))).\displaystyle C(\|u_{n}-a\|_{L^{1}(0,T;L^{1}(\Omega))}+\|v_{n}-b\|_{L^{1}(0,T;L^{1}(\Omega))}).

Therefore, in terms of (4.8) - (4.10), we reach

un(,t)L1(Ω)+vn(,t)L1(Ω)+fn(un,vn)L1(0,T;L1(Ω))+gn(un,vn)L1(0,T;L1(Ω))\displaystyle\|u_{n}(\cdot,t)\|_{L^{1}(\Omega)}+\|v_{n}(\cdot,t)\|_{L^{1}(\Omega)}+\|f_{n}(u_{n},v_{n})\|_{L^{1}(0,T;L^{1}(\Omega))}+\|g_{n}(u_{n},v_{n})\|_{L^{1}(0,T;L^{1}(\Omega))} (4.11)
\displaystyle\leq C(aL1(Ω)+bL1(Ω)),\displaystyle C(\|a\|_{L^{1}(\Omega)}+\|b\|_{L^{1}(\Omega)}),

for all nn\in\mathbb{N}. We emphasize that the constant C>0C>0 is independent of nn\in\mathbb{N}. By (4.11), we can obtain

fn(un,vn)L1(Q)+gn(un,vn)L1(Q)C\|f_{n}(u_{n},v_{n})\|_{L^{1}(Q)}+\|g_{n}(u_{n},v_{n})\|_{L^{1}(Q)}\leq C (4.12)

for all nn\in\mathbb{N}.

Next we will establish the convergence of fn(un,vn)f_{n}(u_{n},v_{n}) and gn(un,vn)g_{n}(u_{n},v_{n}). We set a0:=a+λbL(Ω)a_{0}:=\|a+\lambda b\|_{L^{\infty}(\Omega)} and W:=un+λvnW:=u_{n}+\lambda v_{n} in QQ. Hence (4.3) implies

{tα(Waλb)=ΔW+(fn(un,vn)+λgn(un,vn)), in Q,νW=0, on Ω×(0,T).\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(W-a-\lambda b)=\Delta W+(f_{n}(u_{n},v_{n})+\lambda g_{n}(u_{n},v_{n})),\quad\mbox{ in }Q,\\ \partial_{\nu}W=0,\quad\mbox{ on }\partial\Omega\times(0,T).\end{array}\right. (4.13)

By the definition of fn,gnf_{n},g_{n}, we have

fn(ξ,η)+λgn(ξ,η)=ψn(ξ,η)(f(ξ,η)+λg(ξ,η))0 for all ξ,η.f_{n}(\xi,\eta)+\lambda g_{n}(\xi,\eta)=\psi_{n}(\xi,\eta)(f(\xi,\eta)+\lambda g(\xi,\eta))\leq 0\quad\mbox{ for all }\xi,\eta\in\mathbb{R}. (4.14)

On the other hand, W~(x,t):=a+λbL(Ω)\widetilde{W}(x,t):=\|a+\lambda b\|_{L^{\infty}(\Omega)} satisfies

{tα(W~a+λbL(Ω))=ΔW~, in Q,νW~=0, on Ω×(0,T).\left\{\begin{array}[]{l}\partial_{t}^{\alpha}(\widetilde{W}-\|a+\lambda b\|_{L^{\infty}(\Omega)})=\Delta\widetilde{W},\quad\mbox{ in }Q,\\ \partial_{\nu}\widetilde{W}=0,\quad\mbox{ on }\partial\Omega\times(0,T).\end{array}\right.

Applying the comparison principle(LY1, , (Corollary 1)) to WW and W~\widetilde{W}, in terms of (4.14), we obtain

W(x,t)W~(x,t),(x,t)Q.W(x,t)\leq\widetilde{W}(x,t),\quad(x,t)\in Q.

By Lemma 6 (ii), we have W(x,t)=un(x,t)+λvn(x,t)0W(x,t)=u_{n}(x,t)+\lambda v_{n}(x,t)\geq 0 for (x,t)Q(x,t)\in Q. Therefore,

0un(x,t)+λvn(x,t)a+λbL(Ω),(x,t)Q.0\leq u_{n}(x,t)+\lambda v_{n}(x,t)\leq\|a+\lambda b\|_{L^{\infty}(\Omega)},\quad(x,t)\in Q.

Since ξ+ηmax{1,1λ}(ξ+λη)\xi+\eta\leq\max\left\{1,\,\frac{1}{\lambda}\right\}(\xi+\lambda\eta) for ξ,η0\xi,\eta\geq 0, then we have

0\displaystyle 0 un+vnmax{1,1λ}(un+λvn)\displaystyle\leq u_{n}+v_{n}\leq\max\left\{1,\,\frac{1}{\lambda}\right\}(u_{n}+\lambda v_{n}) (4.15)
Ca+λbL(Ω)=:C1, almost everywhere in Q.\displaystyle\leq C\|a+\lambda b\|_{L^{\infty}(\Omega)}=:C_{1},\quad\mbox{ almost everywhere in }Q.

Because un,vnu_{n},v_{n} are both non-negative, we deduce un,vnC1u_{n},v_{n}\leq C_{1}. Therefore

|fn(un(x,t),vn(x,t))|+|gn(un(x,t),vn(x,t))|\displaystyle|f_{n}(u_{n}(x,t),v_{n}(x,t))|+|g_{n}(u_{n}(x,t),v_{n}(x,t))| (4.16)
\displaystyle\leq sup|ξ|,|η|C1|fn(ξ,η)|+sup|ξ|,|η|C1|gn(ξ,η)|\displaystyle\sup_{|\xi|,|\eta|\leq C_{1}}|f_{n}(\xi,\eta)|+\sup_{|\xi|,|\eta|\leq C_{1}}|g_{n}(\xi,\eta)|
\displaystyle\leq sup|ξ|,|η|C1|f(ξ,η)|+sup|ξ|,|η|C1|g(ξ,η)|=:C2,\displaystyle\sup_{|\xi|,|\eta|\leq C_{1}}|f(\xi,\eta)|+\sup_{|\xi|,|\eta|\leq C_{1}}|g(\xi,\eta)|=:C_{2},

for all nn\in\mathbb{N} and almost all (x,t)Q(x,t)\in Q. We note that C2C_{2} does not depend on nn.

Then we can choose a subsequence nn^{\prime}\in\mathbb{N} such that

limnfn(un(x,t),vn(x,t))=f(u(x,t),v(x,t)), for almost all (x,t)Q,\displaystyle\lim_{n^{\prime}\to\infty}f_{n^{\prime}}(u_{n^{\prime}}(x,t),\,v_{n^{\prime}}(x,t))=f(u(x,t),\,v(x,t)),\quad\mbox{ for almost all }(x,t)\in Q, (4.17)
limngn(un(x,t),vn(x,t))=g(u(x,t),v(x,t)), for almost all (x,t)Q.\displaystyle\lim_{n^{\prime}\to\infty}g_{n^{\prime}}(u_{n^{\prime}}(x,t),\,v_{n^{\prime}}(x,t))=g(u(x,t),\,v(x,t)),\quad\mbox{ for almost all }(x,t)\in Q.

In order to prove this, we can assume that nC12n\geq\frac{C_{1}}{2}. Then, by the definition of fnf_{n} and gng_{n}, we deduce

fn(un(x,t),vn(x,t))=f(un(x,t),vn(x,t)),\displaystyle f_{n}(u_{n}(x,t),\,v_{n}(x,t))=f(u_{n}(x,t),\,v_{n}(x,t)), (4.18)
gn(un(x,t),vn(x,t))=g(un(x,t),vn(x,t)),\displaystyle g_{n}(u_{n}(x,t),\,v_{n}(x,t))=g(u_{n}(x,t),\,v_{n}(x,t)),

for almost all (x,t)Q:=Ω×(0,T)(x,t)\in Q:=\Omega\times(0,T) if nC12n\geq\frac{C_{1}}{2}. On the other hand, in terms of (4.11) we apply Lemma 3 to see that (un,vn),n(u_{n},v_{n}),n\in\mathbb{N} contains a convergent subsequence in L1(Q)L^{1}(Q). Therefore, we can find a subsequence {n}\{n^{\prime}\}\subset\mathbb{N} and some (u,v)(L1(Q))2(u,v)\in(L^{1}(Q))^{2} such that (un,vn)(u,v)(u_{n^{\prime}},v_{n^{\prime}})\longrightarrow(u,v) in L1(Q)2L^{1}(Q)^{2} as nn^{\prime}\to\infty. Hence, choosing further a subsequence of {n}\{n^{\prime}\}, denoted by the same notations, such that

limnun(x,t)=u(x,t),\displaystyle\lim_{n^{\prime}\to\infty}u_{n^{\prime}}(x,t)=u(x,t), (4.19)
limnvn(x,t)=v(x,t), for almost all (x,t)Q.\displaystyle\lim_{n^{\prime}\to\infty}v_{n^{\prime}}(x,t)=v(x,t),\quad\mbox{ for almost all }(x,t)\in Q.

Hence, by f,gC(2)f,g\in C(\mathbb{R}^{2}) and (4.18)-(4.19), we verified (4.17).

Finally considering (4.16) and (4.17), we apply the Lebesgue convergence theorem to verify

limnfn(un,vn)f(u,v)L1(Q)=0,\displaystyle\lim_{n\to\infty}\|f_{n}(u_{n},v_{n})-f(u,v)\|_{L^{1}(Q)}=0,
limngn(un,vn)g(u,v)L1(Q)=0.\displaystyle\lim_{n\to\infty}\|g_{n}(u_{n},v_{n})-g(u,v)\|_{L^{1}(Q)}=0.

Multiplying both sides of (4.3) with ψ\psi, after integrating, we obtain

(una,(tα)ψ)\displaystyle(u_{n}-a,(\partial_{t}^{\alpha})^{*}\psi) =(un,Aψ)+(Fn,ψ),\displaystyle=(u_{n},-A\psi)+(F_{n},\psi), (4.20)
(vnb,(tα)ψ)\displaystyle(v_{n}-b,(\partial_{t}^{\alpha})^{*}\psi) =(vn,Aψ)+(Gn,ψ),\displaystyle=(v_{n},-A\psi)+(G_{n},\psi),

for all ψC(Q¯)\psi\in C^{\infty}(\overline{Q}) satisfying νψ=0\partial_{\nu}\psi=0 in Ω×(0,T)\partial\Omega\times(0,T) and ψ(,T)=0\psi(\cdot,T)=0 in Ω\Omega. Passing nn in (4.20) to the limit, we can see

(ua,(tα)ψ)\displaystyle(u-a,(\partial_{t}^{\alpha})^{*}\psi) =(u,Δψ)+(f(u,v),ψ),\displaystyle=(u,\Delta\psi)+(f(u,v),\psi), (4.21)
(vb,(tα)ψ)\displaystyle(v-b,(\partial_{t}^{\alpha})^{*}\psi) =(v,Δψ)+(g(u,v),ψ),\displaystyle=(v,\Delta\psi)+(g(u,v),\psi),

for all ψC(Q¯)\psi\in C^{\infty}(\overline{Q}) satisfying νψ=0\partial_{\nu}\psi=0 in Ω×(0,T)\partial\Omega\times(0,T) and ψ(,T)=0\psi(\cdot,T)=0 in Ω\Omega. This implies that (u,v)(u,v) is a weak solution to equation (2.3), we have completed the proof of Theorem 2.1.

5 Blow up for a system for diffusion equation with convex nonlinear terms

In this section, we show that the global existence does not hold true for some nonlinear terms, which is our second main result. Here, we do not require the coefficients in front of the spatial diffusion terms to be equal. Specifically, for the second line in (1.1), we consider

tα(vb)=dΔv+g(u,v).\partial_{t}^{\alpha}(v-b)=d\Delta v+g(u,v).

For the statement, we introduce the following assumption for the nonlinear terms ff and gg.

Assumption 2

For f,gC(2)f,g\in C^{\infty}(\mathbb{R}^{2}), there exist constants λ>0,p>1\lambda>0,p>1 and C=Cp,λ>0C=C_{p,\lambda}>0 such that

f(ξ,η)+λg(ξ,η)C(ξp+ηp), for all ξ,η0.f(\xi,\eta)+\lambda g(\xi,\eta)\geq C(\xi^{p}+\eta^{p}),\quad\text{ for all }\xi,\eta\geq 0. (5.1)

Now we are ready to state our main result on the blow-up with upper bound of the blow-up time for p>1p>1.

Theorem 5.1

Let f,gf,g satisfy the Assumption 2 and a,bH2γ(Ω)a,b\in H^{2\gamma}(\Omega), 34<γ1\frac{3}{4}<\gamma\leq 1, satisfy νa=νb=0\partial_{\nu}a=\partial_{\nu}b=0 on Ω\partial\Omega and a,b0,0a,b\geq 0,\not\equiv 0 in Ω\Omega. Then there exists some T=Tα,p,a,b>0T=T_{\alpha,p,a,b}>0 such that the solutions of (2.3) satisfying

u,vC([0,T];H2(Ω)),ua,vbHα(0,T;L2(Ω)),u,v\in C([0,T];H^{2}(\Omega)),\quad u-a,v-b\in H_{\alpha}(0,T;L^{2}(\Omega)),

exist for 0<t<Tα,p,a,b0<t<T_{\alpha,p,a,b} and

limtTα,p,a,b(u(,t)L1(Ω)+v(,t)L1(Ω))=\lim_{t\uparrow T_{\alpha,p,a,b}}\left(\|u(\cdot,t)\|_{L^{1}(\Omega)}+\|v(\cdot,t)\|_{L^{1}(\Omega)}\right)=\infty

holds. Moreover, we can bound Tα,p,a,bT_{\alpha,p,a,b} from above as

Tα,p,a,b(1(p1)Γ(2α)Cp,λ1(1|Ω|Ωa+λbdx)p1)1α=:T(α,p,a,b).T_{\alpha,p,a,b}\leq\left(\frac{1}{(p-1)\Gamma(2-\alpha)C_{p,\lambda}^{-1}\left(\frac{1}{|\Omega|}\int_{\Omega}a+\lambda b\mathrm{~{}d}x\right)^{p-1}}\right)^{\frac{1}{\alpha}}=:T^{*}(\alpha,p,a,b).
Proof

Step 1. Our proof is similar to the one for a single equation (FLY ). We show the following two lemmas from FLY .

Lemma 7

Let hL2(0,T)h\in L^{2}(0,T) and cC[0,T]c\in C[0,T]. Then there exists a unique solution yHα(0,T)y\in H_{\alpha}(0,T) to

tαyc(t)y=h,0<t<T.\partial_{t}^{\alpha}y-c(t)y=h,\quad 0<t<T.

Moreover, if h0h\geq 0 in (0,T)(0,T), then y0y\geq 0 in (0,T)(0,T).

Lemma 8

Let c0>0,a00,p>1c_{0}>0,a_{0}\geq 0,p>1 be constants and ya0,za0Hα(0,T)C[0,T]y-a_{0},z-a_{0}\in H_{\alpha}(0,T)\cap C[0,T] satisfy

tα(ya0)c0yp,tα(za0)c0zp, in (0,T).\partial_{t}^{\alpha}(y-a_{0})\geq c_{0}y^{p},\quad\partial_{t}^{\alpha}(z-a_{0})\leq c_{0}z^{p},\quad\text{ in }(0,T).

Then yzy\geq z in (0,T)(0,T).

Step 2. We set

θ(t):=Ω(u(x,t)+λv(x,t))dx=Ω((ua)+λ(vb))(x,t)dx+m0\theta(t):=\int_{\Omega}\left(u(x,t)+\lambda v(x,t)\right)\mathrm{d}x=\int_{\Omega}\left((u-a)+\lambda(v-b)\right)(x,t)\mathrm{d}x+m_{0}

where m0:=Ω(a+λb)dxm_{0}:=\int_{\Omega}(a+\lambda b)\mathrm{d}x. Here we see that m0>0m_{0}>0 because a,b0,0a,b\geq 0,\not\equiv 0 in Ω\Omega and λ>0\lambda>0 by the assumption of Theorem 5.1. Then by (1.1) we have

tα(θ(t)m0)\displaystyle\partial_{t}^{\alpha}(\theta(t)-m_{0}) =Ω(Δu+λdΔv)(x,t)dx+Ω(f(u,v)+λg(u,v))(x,t)dx\displaystyle=\int_{\Omega}(\Delta u+\lambda d\Delta v)(x,t)\mathrm{d}x+\int_{\Omega}(f(u,v)+\lambda g(u,v))(x,t)\mathrm{d}x
=Ω(f(u,v)+λg(u,v))(x,t)dx.\displaystyle=\int_{\Omega}(f(u,v)+\lambda g(u,v))(x,t)\mathrm{d}x.

On the other hand, we have

θ(t)=Ω(u+λv)dx(Ωdx)1p(Ω(u+λv)pdx)1p=|Ω|1p(Ω(u+λv)pdx)1p\theta(t)=\int_{\Omega}(u+\lambda v)\mathrm{d}x\leq\left(\int_{\Omega}\mathrm{d}x\right)^{\frac{1}{p^{\prime}}}\left(\int_{\Omega}(u+\lambda v)^{p}\mathrm{d}x\right)^{\frac{1}{p}}=|\Omega|^{\frac{1}{p^{\prime}}}\left(\int_{\Omega}(u+\lambda v)^{p}\mathrm{d}x\right)^{\frac{1}{p}}

with 1p+1p=1\frac{1}{p^{\prime}}+\frac{1}{p}=1, and (u+λv)pCp,λ(up+vp)(u+\lambda v)^{p}\leq C_{p,\lambda}(u^{p}+v^{p}) for u,v0u,v\geq 0. Hence, (5.1) yields

θp(t)\displaystyle\theta^{p}(t) |Ω|ppΩ(u+λv)pdxCp,λ|Ω|ppΩ(up+vp)dx\displaystyle\leq|\Omega|^{\frac{p}{p^{\prime}}}\int_{\Omega}(u+\lambda v)^{p}\mathrm{d}x\leq C_{p,\lambda}|\Omega|^{\frac{p}{p^{\prime}}}\int_{\Omega}(u^{p}+v^{p})\mathrm{d}x
Cp,λ|Ω|ppΩ(f(u,v)+λg(u,v))(x,t)dx.\displaystyle\leq C_{p,\lambda}|\Omega|^{\frac{p}{p^{\prime}}}\int_{\Omega}(f(u,v)+\lambda g(u,v))(x,t)\mathrm{d}x.

So we obtain

tα(θ(t)m0)Cp,λ1|Ω|ppθp(t),0<t<T.\partial_{t}^{\alpha}(\theta(t)-m_{0})\geq C_{p,\lambda}^{-1}|\Omega|^{-\frac{p}{p^{\prime}}}\theta^{p}(t),\quad 0<t<T.

Step 3. This step is devoted to the construction of a lower solution θ¯(t)\underline{\theta}(t) satisfying

tα(θ¯(t)m0)C0θ¯p(t),0<t<Tϵ,limtTθ¯(t)=.\partial_{t}^{\alpha}(\underline{\theta}(t)-m_{0})\leq C_{0}\underline{\theta}^{p}(t),\quad 0<t<T-\epsilon,\quad\lim_{t\uparrow T}\underline{\theta}(t)=\infty. (5.2)

Here we set C0:=Cp,λ1|Ω|ppC_{0}:=C_{p,\lambda}^{-1}|\Omega|^{-\frac{p}{p^{\prime}}}. The construction of θ¯(t)\underline{\theta}(t) is same as the proof in FLY . As a possible lower solution, we consider

θ¯(t):=m0(TTt)m,m.\underline{\theta}(t):=m_{0}\left(\frac{T}{T-t}\right)^{m},\ m\in\mathbb{N}.

By the definition tα(θ¯(t)a0)(t)=dtαθ¯(t)\partial_{t}^{\alpha}(\underline{\theta}(t)-a_{0})(t)=\mathrm{d}^{\alpha}_{t}\underline{\theta}(t). We can get

tα(θ¯(t)m0)(t)=dtαθ¯(t)=m0Tmdtα(1(Tt)m)m0Tm+1αmΓ(2α)1(Tt)m+1.\partial_{t}^{\alpha}(\underline{\theta}(t)-m_{0})(t)=\mathrm{d}^{\alpha}_{t}\underline{\theta}(t)=m_{0}T^{m}\mathrm{d}^{\alpha}_{t}\left(\frac{1}{(T-t)^{m}}\right)\leq\frac{m_{0}T^{m+1-\alpha}m}{\Gamma(2-\alpha)}\frac{1}{(T-t)^{m+1}}.

This equation holds for arbitrary m,T>0m\in\mathbb{N},T>0 and 0<t<Tϵ0<t<T-\epsilon.

Finally we claim that for any p>1p>1 and m0>0m_{0}>0, there exist constants mm\in\mathbb{N} and T>0T>0 such that

m0Tm+1αmΓ(2α)1(Tt)m+1C0θ¯p(t)=C0m0pTmp(Tt)mp, 0<t<Tϵ.\frac{m_{0}T^{m+1-\alpha}m}{\Gamma(2-\alpha)}\frac{1}{(T-t)^{m+1}}\leq C_{0}\underline{\theta}^{p}(t)=\frac{C_{0}m_{0}^{p}T^{mp}}{(T-t)^{mp}},\ 0<t<T-\epsilon. (5.3)

Therefore, if

T\displaystyle T (mΓ(2α)C0m0p1)1α(1(p1)Γ(2α)C0m0p1)1α\displaystyle\geq\left(\frac{m}{\Gamma(2-\alpha)C_{0}m_{0}^{p-1}}\right)^{\frac{1}{\alpha}}\geq\left(\frac{1}{(p-1)\Gamma(2-\alpha)C_{0}m_{0}^{p-1}}\right)^{\frac{1}{\alpha}} (5.4)
={(p1)Γ(2α)Cp,λ1(1|Ω|Ω(a+λb)dx)p1}1α=:T(α,p,a,b),\displaystyle=\left\{(p-1)\Gamma(2-\alpha)C_{p,\lambda}^{-1}\left(\frac{1}{|\Omega|}\int_{\Omega}(a+\lambda b)\mathrm{~{}d}x\right)^{p-1}\right\}^{-\frac{1}{\alpha}}=:T^{*}(\alpha,p,a,b),

then (5.3) is satisfied.

Consequently, we can obtain

θ¯(t):=m0(T(α,p,a,b)T(α,p,a,b)t)m\underline{\theta}(t):=m_{0}\left(\frac{T^{*}(\alpha,p,a,b)}{T^{*}(\alpha,p,a,b)-t}\right)^{m}

satisfies (5.2).

Now it suffices to apply Lemma 8 to get

θ(t)θ¯(t),0tT(α,p,a,b)ϵ.\theta(t)\geq\underline{\theta}(t),\quad 0\leq t\leq T^{*}(\alpha,p,a,b)-\epsilon.

Since ϵ>0\epsilon>0 was arbitrarily chosen, we obtain

Ω(u(x,t)+λv(x,t))dx=θ(t)θ¯(t)=m0(T(α,p,a,b)T(α,p,a,b)t)m.\int_{\Omega}\left(u(x,t)+\lambda v(x,t)\right)\mathrm{d}x=\theta(t)\geq\underline{\theta}(t)=m_{0}\left(\frac{T^{*}(\alpha,p,a,b)}{T^{*}(\alpha,p,a,b)-t}\right)^{m}.

This means that the solution u+vu+v cannot exist for t>T(α,p,a,b)t>T^{*}(\alpha,p,a,b). Hence, the blow-up time T(α,p,a,b)T(α,p,a,b)T(\alpha,p,a,b)\leq T^{*}(\alpha,p,a,b). The proof of Theorem 5.1 is complete.

6 Concluding remarks

It’s worth mentioning that the conditions on ff and gg in the Assumption 1 are f(0,η)=g(ξ,0)=0f(0,\eta)=g(\xi,0)=0 for all ξ0\xi\geq 0 and η0\eta\geq 0, and f,gf,g are local Lipschitz continuous. If we improve the regularity of ff and gg such that f,gC1(2)f,g\in C^{1}(\mathbb{R}^{2}), we can obtain the non-negativity by comparison principle. For simplicity, we choose an initial value a=0a=0, and denote u~\widetilde{u} as the solution of

tαu~=Δu~+f(u~,v~)f(0,v~)\partial_{t}^{\alpha}\widetilde{u}=\Delta\widetilde{u}+f(\widetilde{u},\widetilde{v})-f(0,\widetilde{v})

Noticing f(u~,v~)f(0,v~)=1f(ξ,v~)u~f(\widetilde{u},\widetilde{v})-f(0,\widetilde{v})=\partial_{1}f(\xi,\widetilde{v})\widetilde{u}, we can rewrite the above equation as

tαu~=Δu~+1f(ξ,v~)u~.\partial_{t}^{\alpha}\widetilde{u}=\Delta\widetilde{u}+\partial_{1}f(\xi,\widetilde{v})\widetilde{u}.

Here ξ\xi is some number between 0 and u~(x,t)\widetilde{u}(x,t). Similarly, we obtain the equation for v~\widetilde{v} as

tαv~=Δv~+2g(u~,η)v~,\partial_{t}^{\alpha}\widetilde{v}=\Delta\widetilde{v}+\partial_{2}g(\widetilde{u},\eta)\widetilde{v},

where η\eta is some number between 0 and v~\widetilde{v}. Therefore, by the comparison principleLY2 and Lemma 2, we can derive the non-negativity of the solutions u~\widetilde{u} and v~\widetilde{v}.

Moreover, it can be observed that limited in L1(Ω×(0,T))L^{1}(\Omega\times(0,T)), we can derive the global existence of solutions. For the uniqueness of solutions, we need to impose stronger priori assumptions on the right-hand side term and the initial values. Similarly to LY2 , we impose the following requirements on the nonlinear term g(u,v)g(u,v). For g:𝒟(A0γ×A0γ)L2(Ω)g:\mathcal{D}(A_{0}^{\gamma}\times A_{0}^{\gamma})\to L^{2}(\Omega), we assume that for some constant m>0m>0, there exists a constant Cg=Cg(m)>0C_{g}=C_{g}(m)>0 such that

{ (i) g(u,)+g(,v)Cg,g(u1,)g(u2,)Cgu1u2𝒟(A0γ),g(,v1)g(,v2)Cgv1v2𝒟(A0γ) if u𝒟(A0γ),v𝒟(A0γ),u1𝒟(A0γ),u2𝒟(A0γ),v1𝒟(A0γ),v2𝒟(A0γ)m and  (ii) there exists a constant ε(0,34) such that g(u,)H2ε(Ω)+g(,v)H2ε(Ω)Cg(m) if u𝒟(A0γ),v𝒟(A0γ)m,\left\{\begin{array}[]{l}\text{ (i) }\|g(u,\cdot)\|+\|g(\cdot,v)\|\leq C_{g},\quad\left\|g\left(u_{1},\cdot\right)-g\left(u_{2},\cdot\right)\right\|\leq C_{g}\left\|u_{1}-u_{2}\right\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\\ \left\|g\left(\cdot,v_{1}\right)-g\left(\cdot,v_{2}\right)\right\|\leq C_{g}\left\|v_{1}-v_{2}\right\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)}\\ \text{ if }\|u\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\|v\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\left\|u_{1}\right\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\left\|u_{2}\right\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\left\|v_{1}\right\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\left\|v_{2}\right\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)}\leq m\\ \text{ and }\\ \text{ (ii) there exists a constant }\varepsilon\in\left(0,\frac{3}{4}\right)\text{ such that }\\ \|g(u,\cdot)\|_{H^{2\varepsilon}(\Omega)}+\|g(\cdot,v)\|_{H^{2\varepsilon}(\Omega)}\leq C_{g}(m)\quad\text{ if }\|u\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\|v\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)}\leq m,\end{array}\right.

along with initial values a,ba,b satisfying a𝒟(A0γ),b𝒟(A0γ)m\|a\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)},\|b\|_{\mathcal{D}\left(A_{0}^{\gamma}\right)}\leq m. After expressing solutions u(t)u(t) and v(t)v(t) as

u(t)=S(t)a+0tK(ts)(p0ug(u(s),v(s)))ds,\displaystyle u(t)=S(t)a+\int_{0}^{t}K(t-s)\left(p_{0}u-g(u(s),v(s))\right)\mathrm{d}s,
v(t)=S(t)b+0tK(ts)(p0v+g(u(s),v(s)))ds,\displaystyle v(t)=S(t)b+\int_{0}^{t}K(t-s)\left(p_{0}v+g(u(s),v(s))\right)\mathrm{d}s,

we can apply the contraction theorem similarly to the approach demonstrated in LY2 . In this situation, there exists a constant T=T(m)>0T=T(m)>0 such that the solutions to the system (2.3) in L2(0,T;H2(Ω))C([0,T];𝒟(A0γ))L^{2}(0,T;H^{2}(\Omega))\cap C([0,T];\mathcal{D}(A^{\gamma}_{0})) are unique.

7 Appendix. Proof of Lemma 2

Since 1<κ<1<\kappa<\infty, the embedding L(Q)Lκ(Q)L^{\infty}(Q)\hookrightarrow L^{\kappa}(Q) holds, thus we have cLκ(Q)c\in L^{\kappa}(Q). Furthermore, given that C0(Q)C_{0}^{\infty}(Q) is dense in Lκ(Q)L^{\kappa}(Q), there exist functions {cn}C0(Q)\{c_{n}\}\in C_{0}^{\infty}(Q) such that limncn=c\lim_{n\to\infty}c_{n}=c in Lκ(Q)L^{\kappa}(Q).

We define the operator AnA_{n} regarding cn(x,t)c_{n}(x,t) in L2(Ω)L^{2}(\Omega) similar to LY1 as

(Anv)(x)=Δv(x)cn(x,t)v(x),xΩ.(A_{n}v)(x)=-\Delta v(x)-c_{n}(x,t)v(x),\quad x\in\Omega.

Correspondingly, the operators S(t)S(t) and K(t)K(t) related to (A0v)(x):=Δv(x)c0v(x)\left(-A_{0}v\right)(x):=\Delta v(x)-c_{0}v(x) with an arbitrary constant c0>0c_{0}>0 remain the same

S(t)a=n=1Eα,1(μntα)(a,φn)φn,aL2(Ω),t>0,S(t)a=\sum_{n=1}^{\infty}E_{\alpha,1}\left(-\mu_{n}t^{\alpha}\right)\left(a,\varphi_{n}\right)\varphi_{n},\quad a\in L^{2}(\Omega),t>0,\\

and

K(t)a=n=1tα1Eα,α(μntα)(a,φn)φn,aL2(Ω),t>0.K(t)a=\sum_{n=1}^{\infty}t^{\alpha-1}E_{\alpha,\alpha}\left(-\mu_{n}t^{\alpha}\right)\left(a,\varphi_{n}\right)\varphi_{n},\quad a\in L^{2}(\Omega),t>0.

Additionally, the operators in (LY1, , (2.14)) are updated to

G(t):=0tK(ts)F(s)𝑑s+S(t)a,Qnv(t)=Qn(t)v(t):=(c0+cn(,t))v(t).\begin{array}[]{l}G(t):=\int_{0}^{t}K(t-s)F(s)ds+S(t)a,\\ Q_{n}v(t)=Q_{n}(t)v(t):=\left(c_{0}+c_{n}(\cdot,t)\right)v(t).\end{array}

In accordance on the previous modifications, the solution presented in (LY1, , (4.3)) has been updated to reflect the changes in the operators AnA_{n} and QnQ_{n} as defined above. The updated solution is given by

un(F,a)(t)=G(t)+0tK(ts)Qnun(s)ds,0<t<T.u_{n}(F,a)(t)=G(t)+\int_{0}^{t}K(t-s)Q_{n}u_{n}(s)\mathrm{d}s,\quad 0<t<T.

Since cnC0(Q)c_{n}\in C_{0}^{\infty}(Q), according to (LY1, , Theorem 2), we know that un(F,a)(x,t)0u_{n}(F,a)(x,t)\geq 0 in Ω×(0,T)\Omega\times(0,T) and that un(F,a)L2(0,T;H2(Ω))u_{n}(F,a)\in L^{2}(0,T;H^{2}(\Omega)) as well as un(F,a)aHα(0,T;L2(Ω))u_{n}(F,a)-a\in H_{\alpha}(0,T;L^{2}(\Omega)). Following a similar approach as in LY1 , we can extract a subsequence un(F,a){u_{n^{\prime}}(F,a)} from un(F,a){u_{n}(F,a)} such that un(F,a)u(F,a){u_{n^{\prime}}(F,a)}\to{u(F,a)} in L2(0,T;H2(Ω))L^{2}(0,T;H^{2}(\Omega)) and un(F,a)au(F,a)a{u_{n^{\prime}}(F,a)}-a\to{u(F,a)}-a in Hα(0,T;L2(Ω))H_{\alpha}(0,T;L^{2}(\Omega)). Finally, un(F,a)(x,t)0u_{n}(F,a)(x,t)\geq 0 implies that u(F,a)0u(F,a)\geq 0 in QQ.

Acknowledgements.
D. Feng is supported by Key-Area Research and Development Program of Guangdong Province (No.2021B0101190003) and Science and Technology Commission of Shanghai Municipality (23JC1400501). M. Yamamoto is supported partly by Grant-in-Aid for Scientific Research (A) 20H00117 and Grant-in-Aid for Challenging Research (Pioneering) 21K18142 of Japan Society for the Promotion of Science.

Conflict of interest

The authors declare that they have no conflict of interest.

References

  • (1) Adams R A 1975 Sobolev Spaces (New York: Academic Press)
  • (2) Bergh J and Löfström J 1976 Interpolation Spaces: An Introduction (Springer-Verlag: Berlin)
  • (3) Baranwal V K, Pandey R K, Tripathi M P and Singh O P 2012 An analytic algorithm for time fractional nonlinear reaction-diffusion equation based on a new iterative method Communications in Nonlinear Science and Numerical Simulation 17(10) 3906-3921
  • (4) Floridia G, Liu Y and Yamamoto M 2023 Blow-up in L1(Ω)L^{1}(\Omega)-norm and global existence for time-fractional diffusion equations with polynomial semilinear terms Advances in Nonlinear Analysis 12(1) 20230121
  • (5) Gorenflo R, Luchko Y and Yamamoto M 2015 Time-fractional diffusion equation in the fractional Sobolev spaces Fract. Calc. Appl. Anal. 18 799-820
  • (6) Klausmeier C A 1999 Regular and irregular patterns in semiarid vegetation Science 284(5421) 1826-1828
  • (7) Kubica A, Ryszewska K and Yamamoto M 2020 Theory of Time-fractional Differential Equations An Introduction (Tokyo: Springer)
  • (8) Li Z, Huang X and Liu Y 2023 Initial-boundary value problems for coupled systems of time-fractional diffusion equations Fractional Calculus and Applied Analysis 26(2) 533-566
  • (9) Luchko Y and Yamamoto M 2017 On the maximum principle for a time-fractional diffusion equation Fract. Calc. Appl. Anal. 20 1131-1145
  • (10) Luchko Y and Yamamoto M 2023 Comparison principles for the time-fractional diffusion equations with the Robin boundary conditions. Part I: Linear equations Fract. Calc. Appl. Anal. 26 (2023) no.4 1504-1544, correction: Fract. Calc. Appl. Anal. 26 no. 6, 2959-2960
  • (11) Luchko Y and Yamamoto M 2022 Comparison principles for the linear and semiliniar time-fractional diffusion equations with the Robin boundary condition arXiv preprint arXiv:2208.04606
  • (12) Pazy A 1983 Semigroups of Linear Operators and Applications to Partial Differential Equations (Berlin: Springer)
  • (13) Pierre M 2003 Weak solutions and supersolutions in L1L^{1} for reaction-diffusion systems J. Evol. Equ. 3 153-168
  • (14) Pierre M 2010 Global existence in reaction-diffusion systems with control of mass: a survey Milan Journal of Mathematics 78 417-455
  • (15) Sakamoto K and Yamamoto M 2011 Initial value/boundary value problems for fractional diffusion-wave equations and applications to some inverse problems J. Math. Anal. Appl. 382 426-447
  • (16) Simon J 1986 Compact sets in the space Lp(0,T;B)L^{p}(0,T;B) Annali di Matematica pura ed applicata 146 65-96
  • (17) Wang X, Shi J and Zhang G 2021 Bifurcation and pattern formation in diffusive Klausmeier-Gray-Scott model of water-plant interaction. J. Math. Anal. Appl. 497(1) 124860
  • (18) Yamamoto M 2022 Fractional calculus and time-fractional differential equations: revisit and construction of a theory Mathematics 10 698

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.