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

On a mixed local-nonlocal evolution equation with singular nonlinearity

Kaushik Bal and Stuti Das [email protected] and [email protected] Department of Mathematics and Statistics,
Indian Institute of Technology Kanpur, Uttar Pradesh, 208016, India
Abstract

We will prove several existence and regularity results for the mixed local-nonlocal parabolic equation of the form

utΔu+(Δ)su=f(x,t)uγ(x,t) in ΩT:=Ω×(0,T),u=0 in (n\Ω)×(0,T),u(x,0)=u0(x) in Ω;\displaystyle\displaystyle\begin{split}u_{t}-\Delta u+(-\Delta)^{s}u&=\frac{f(x,t)}{u^{\gamma(x,t)}}\text{ in }\Omega_{T}:=\Omega\times(0,T),\\ u&=0\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u(x,0)&=u_{0}(x)\text{ in }\Omega;\end{split} (0.1)

where

(Δ)su=cn,sP.V.nu(x,t)u(y,t)|xy|n+2s𝑑y.(-\Delta)^{s}u=c_{n,s}\operatorname{P.V.}\int_{\mathbb{R}^{n}}\frac{u(x,t)-u(y,t)}{|x-y|^{n+2s}}dy.

Under the assumptions that γ\displaystyle\gamma is a positive continuous function on Ω¯T\displaystyle\overline{\Omega}_{T} and Ω\displaystyle\Omega is a bounded domain with Lipschitz boundary in n\displaystyle\mathbb{R}^{n}, n>2\displaystyle n>2, s(0,1)\displaystyle s\in(0,1), 0<T<+\displaystyle 0<T<+\infty, f0\displaystyle f\geq 0, u00\displaystyle u_{0}\geq 0, f\displaystyle f and u0\displaystyle u_{0} belongs to suitable Lebesgue spaces. Here cn,s\displaystyle c_{n,s} is a suitable normalization constant, and P.V.\displaystyle\operatorname{P.V.} stands for Cauchy Principal Value.

1 Introduction

In this article, we study the evolution of a mixed local-nonlocal operator under the effect of a singular nonlinearity given by:

utΔu+(Δ)su=f(x,t)uγ(x,t) in ΩT:=Ω×(0,T),u=0 in (n\Ω)×(0,T),u(x,0)=u0(x) in Ω;{}\begin{split}u_{t}-\Delta u+(-\Delta)^{s}u&=\frac{f(x,t)}{u^{\gamma(x,t)}}\;\mbox{ in }\Omega_{T}:=\Omega\times(0,T),\\ u&=0\;\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u(x,0)&=u_{0}(x)\mbox{ in }\Omega;\end{split} (1.1)

where

(Δ)su:=cn,sP.V.nu(x,t)u(y,t)|xy|n+2s𝑑y,(-\Delta)^{s}u:=c_{n,s}\operatorname{P.V.}\int_{\mathbb{R}^{n}}\frac{u(x,t)-u(y,t)}{|x-y|^{n+2s}}dy,

γ\displaystyle\gamma is a positive continuous function on Ω¯T\displaystyle\overline{\Omega}_{T} with Ω\displaystyle\Omega being a bounded domain with Lipschitz boundary in n\displaystyle\mathbb{R}^{n}, n>2\displaystyle n>2, s(0,1)\displaystyle s\in(0,1) and 0<T<+\displaystyle 0<T<+\infty. Assuming that f0\displaystyle f\geq 0 and u00\displaystyle u_{0}\geq 0 with variable summability, in this paper we show several existence and regularity results. To this aim, we start by reviewing the literature concerning our problem.

Singular elliptic problems have been extensively studied in the literature for the past few decades starting with the now classical work of Crandall-Rabinowitz-Tartar [15], who showed that the stationary state of (1.1), under Dirichlet boundary conditions given by

Δu=fuγ(x) in Ω,u>0 in Ω,u=0 in Ω.\displaystyle\displaystyle{}\begin{split}-\Delta u&=\frac{f}{u^{\gamma(x)}}\text{ in }\Omega,\\ u&>0\;\text{ in }\Omega,\\ u&=0\;\text{ in }\partial\Omega.\end{split} (1.2)

admits a unique solution uC2(Ω)C(Ω¯)\displaystyle u\in C^{2}(\Omega)\cap C(\bar{\Omega}) for any γ>0\displaystyle\gamma>0 constant along with the fact that the solution must behave like a distance function near the boundary provided f\displaystyle f is Hölder Continuous. Interestingly enough Lazer-Mckenna [28] showed that the unique solution obtained by [15] is indeed in H01(Ω)\displaystyle H_{0}^{1}(\Omega) iff 0<γ<3\displaystyle 0<\gamma<3. Boccardo-Orsina [9] in a beautiful paper showed that the followings regarding solutions of (1.2)

{uW01,nr(1+γ)nr(1γ)(Ω) if 0<γ<1 and fLr(Ω) with r[1,(2/(1γ))),uH01(Ω) if 0<γ<1 and fLr(Ω) with r=(2/(1γ)),uH01(Ω) if γ=1 and fL1(Ω),u1+γ2H01(Ω) if γ>1 and fL1(Ω),\begin{cases}u\in W_{0}^{1,\frac{nr(1+\gamma)}{n-r(1-\gamma)}}(\Omega)&\text{ if }0<\gamma<1\text{ and }f\in L^{r}(\Omega)\text{ with }r\in\left[1,\left(2^{*}/(1-\gamma)\right)^{\prime}\right),\\ u\in H_{0}^{1}(\Omega)&\text{ if }0<\gamma<1\text{ and }f\in L^{r}(\Omega)\text{ with }r=\left(2^{*}/(1-\gamma)\right)^{\prime},\\ u\in H_{0}^{1}(\Omega)&\text{ if }\gamma=1\text{ and }f\in L^{1}(\Omega),\\ u^{\frac{1+\gamma}{2}}\in H_{0}^{1}(\Omega)&\text{ if }\gamma>1\text{ and }f\in L^{1}(\Omega),\end{cases}

hold, which was extended for variable γ\displaystyle\gamma, introducing certain conditions on its behaviour near the boundary in [14]. The nonlocal variant given by

(Δ)psu=fuγ(x) in Ω,u>0 in Ω,u=0 in Ω.\displaystyle\displaystyle\begin{split}(-\Delta)_{p}^{s}u&=\frac{f}{u^{\gamma(x)}}\text{ in }\Omega,\\ u&>0\text{ in }\Omega,\\ u&=0\text{ in }\partial\Omega.\end{split}

was studied in [7] for p=2\displaystyle p=2 and γ(x)=γ+\displaystyle\gamma(x)=\gamma\in\mathbb{R}_{*}^{+}, the authors proved the existence and uniqueness of positive solutions, according to the range of γ\displaystyle\gamma and summability of f\displaystyle f. For the quasilinear case, we refer [13] for constant γ>0\displaystyle\gamma>0 and for variable singular exponent, the existence results have been obtained in [25]. As for the Mixed local-nonlocal elliptic problem given by

Δpu+(Δ)psu=fuγ(x) in Ω,u>0 in Ω,u=0 in Ω;\displaystyle\displaystyle\begin{split}-\Delta_{p}u+(-\Delta)_{p}^{s}u&=\frac{f}{u^{\gamma(x)}}\text{ in }\Omega,\\ u&>0\text{ in }\Omega,\\ u&=0\text{ in }\partial\Omega;\end{split}

Arora [3] for p=2\displaystyle p=2 and γ(x)=γ+\displaystyle\gamma(x)=\gamma\in\mathbb{R}_{*}^{+}, obtained the existence, uniqueness and regularity properties of the weak solutions by deriving uniform a priori estimates and using the approximation technique. They also obtained some existence and nonexistence results when f\displaystyle f behaves like a distance function. For the case p>1\displaystyle p>1, the constant exponent γ\displaystyle\gamma case has been considered in [26], and the variable exponent can be found in [8]. If one considers the parabolic counterpart i.e, the equation given by:

utΔpu=f(x,t)uγ in ΩT,u=0 in (n\Ω)×(0,T),u(x,0)=u0(x) in Ω.\displaystyle\displaystyle\begin{split}u_{t}-\Delta_{p}u&=\frac{f(x,t)}{u^{\gamma}}\text{ in }\Omega_{T},\\ u&=0\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u(x,0)&=u_{0}(x)\text{ in }\Omega.\end{split}

For such an equation with p2, 0fLm(ΩT)\displaystyle p\geq 2,\;0\leq f\in L^{m}\left(\Omega_{T}\right) with m1\displaystyle m\geq 1 and assuming that ωΩ,dω>0\displaystyle\forall\omega\subset\subset\Omega,\exists\;d_{\omega}>0, such that u0dω\displaystyle u_{0}\geq d_{\omega}, the authors [17] proved the existence of a solution u\displaystyle u of the above problem such that

{uLq0(0,T;W01,q0(Ω)) if 0<γ<1 and fLr(Ω) with r[1,p(n+2)p(n+2)n(1γ)) and q0=m[n(p+γ1)+p(γ+1)]n+2m(1γ),uLp(0,T;W01,p(Ω)) if 0<γ1 and fLm0(ΩT) with m0=p(n+2)p(n+2)n(1γ),uLp(0,T;Wloc1,p(Ω)), if γ>1 and fL1(ΩT).\displaystyle\displaystyle\begin{cases}u\in L^{q_{0}}(0,T;W_{0}^{1,q_{0}}(\Omega))&\text{ if }0<\gamma<1\text{ and }f\in L^{r}(\Omega)\text{ with }r\in\left[1,\frac{p(n+2)}{p(n+2)-n(1-\gamma)}\right)\\ &\text{ and }q_{0}=\frac{m[n(p+\gamma-1)+p(\gamma+1)]}{n+2-m(1-\gamma)},\\ u\in L^{p}(0,T;W^{1,p}_{0}(\Omega))&\text{ if }0<\gamma\leq 1\text{ and }f\in L^{m_{0}}\left(\Omega_{T}\right)\text{ with }m_{0}=\frac{p(n+2)}{p(n+2)-n(1-\gamma)},\\ u\in L^{p}(0,T;W_{\operatorname{loc}}^{1,p}(\Omega)),&\text{ if }\gamma>1\text{ and }f\in L^{1}(\Omega_{T}).\end{cases}

The Nonlocal case (s(0,1))\displaystyle(s\in(0,1)) for the parabolic problem was handled by [1] for γ>0\displaystyle\gamma>0 constant to show existence and uniqueness results along similar lines. If one restricts the range of γ\displaystyle\gamma then various existence, uniqueness and regularity results can be found in Bal-Badra-Giacomoni [4, 5, 6] and Giacomoni-Bougherera [11]. We would also like to mention that the regularity theory of mixed local and non-local operators plays a major role in our problem and we cite the following papers [16, 18, 21, 22, 23, 24, 30, 31] and the references therein.
As for the boundedness of our solutions, we refer Aronson-Serrin [2], where the summability requirement of initial data for boundedness was introduced by Aronson and Serrin, for the case of second-order differential equations without singularity. Outside of the Aronson-Serrin domain, the optimal summability of solutions for the local case without singularity was obtained in Boccardo-Porzio-Primo [10]. These results for the nonlocal case have been obtained in Peral [29], this too for the nonsingular case. For the mixed local-nonlocal operator with singularity, we will be able to get similar types of results here depending on the choice of γ\displaystyle\gamma. We will use suitable approximating problems to get the existence and other summability properties of weak solutions.

Organization of the article

In the next section, we will describe some basic notations and fix some preliminary function spaces to define our solutions, followed by embedding results and other properties regarding those spaces.
Then we will introduce the notion of weak solution for the case γ=0\displaystyle\gamma=0 and show its existence, uniqueness, positivity and other properties. After that, we write about the existence of weak solutions for approximating problems and give definitions of weak solutions for the singular cases, both for constant and variable exponents. We end this section by stating our main theorems regarding the existence and summability of weak solutions and appropriate comments.
The next section contains the proofs of our main results, and we end with another section that gives the asymptotic behaviour of the solutions in a suitable sense.

2 Preliminaries

2.1 Notations

We gather here all the standard notations that will be used throughout the paper.
\displaystyle\bullet We will take n\displaystyle n to be the space dimension and denote by z=(x,t)\displaystyle z=(x,t) to be a point in n×(0,T)\displaystyle\mathbb{R}^{n}\times(0,T), where (0,T)\displaystyle(0,T)\subset\mathbb{R} for some 0<T<\displaystyle 0<T<\infty.
\displaystyle\bullet Let Ω\displaystyle\Omega be an open bounded domain in n\displaystyle\mathbb{R}^{n} with boundary Ω\displaystyle\partial\Omega and for 0<T<\displaystyle 0<T<\infty, let ΩT:=Ω×(0,T)\displaystyle\Omega_{T}:=\Omega\times(0,T).
\displaystyle\bullet We denote the parabolic boundary ΓT\displaystyle\Gamma_{T} by ΓT=(Ω×{t=0})(Ω×(0,T))\displaystyle\Gamma_{T}=(\Omega\times\{t=0\})\cup(\partial\Omega\times(0,T)).
\displaystyle\bullet We define the set (ΩT)δ={(x,t)ΩT:dist((x,t),ΓT)<δ}\displaystyle(\Omega_{T})_{\delta}=\{(x,t)\in\Omega_{T}:\operatorname{dist}((x,t),\Gamma_{T})<\delta\} for δ>0\displaystyle\delta>0 fixed.
\displaystyle\bullet We shall alternately use tg\displaystyle\partial_{t}g or gt or gt\displaystyle\frac{\partial g}{\partial t}\text{ or }g_{t} to denote the time derivative (partial) of a function g\displaystyle g.
\displaystyle\bullet For r>1\displaystyle r>1, the Hölder conjugate exponent of r\displaystyle r will be denoted by r=rr1\displaystyle r^{\prime}=\frac{r}{r-1}.
\displaystyle\bullet The Lebesgue measure of a measurable subset Sn\displaystyle\mathrm{S}\subset\mathbb{R}^{n} will be denoted by |S|\displaystyle|\mathrm{S}|.
\displaystyle\bullet For any open subset Ω\displaystyle\Omega of n\displaystyle\mathbb{R}^{n}, KΩ\displaystyle K\subset\subset\Omega will imply K\displaystyle K is compactly contained in Ω.\displaystyle\Omega.
\displaystyle\bullet \displaystyle\int will denote integration concerning either space or time only, and integration on Ω×Ω\displaystyle\Omega\times\Omega or n×n\displaystyle\mathbb{R}^{n}\times\mathbb{R}^{n} will be denoted by a double integral \displaystyle\iint.
\displaystyle\bullet We will use \displaystyle\iiint to denote integral over n×n×(0,T)\displaystyle\mathbb{R}^{n}\times\mathbb{R}^{n}\times(0,T).
\displaystyle\bullet Average integral will be denoted by \displaystyle\fint.
\displaystyle\bullet The notation ab\displaystyle a\lesssim b will be used for aCb\displaystyle a\leq Cb, where C\displaystyle C is a universal constant which only depends on the dimension n\displaystyle n and sometimes on s\displaystyle s too. C\displaystyle C may vary from line to line or even in the same line.
\displaystyle\bullet .,.\displaystyle\langle.,.\rangle will denote the usual inner product in some associated Hilbert space.
\displaystyle\bullet For any function h\displaystyle h, we denote the positive and negative parts of it by h+=max{h,0}\displaystyle h_{+}=\operatorname{max}\{h,0\} and h=max{h,0}\displaystyle h_{-}=\operatorname{max}\{-h,0\} respectively.
\displaystyle\bullet For k\displaystyle k\in\mathbb{N}, we denote Tk(σ)=max{k,min{k,σ}}\displaystyle T_{k}(\sigma)=\max\{-k,\min\{k,\sigma\}\}, for σ\displaystyle\sigma\in\mathbb{R}.

2.2 Function Spaces

In this section, we present definitions and properties of some function spaces that will be useful for our work. We recall that for En\displaystyle E\subset\mathbb{R}^{n}, the Lebesgue space Lp(E),1p<\displaystyle L^{p}(E),1\leq p<\infty, is defined to be the space of p\displaystyle p-integrable functions u:E\displaystyle u:E\rightarrow\mathbb{R} with the finite norm

uLp(E)=(E|u(x)|p𝑑x)1/p.\|u\|_{L^{p}(E)}=\left(\int_{E}|u(x)|^{p}dx\right)^{1/p}.

By Llocp(E)\displaystyle L_{\operatorname{loc}}^{p}(E) we denote the space of locally p\displaystyle p-integrable functions, which means, uLlocp(E)\displaystyle u\in L_{\operatorname{loc}}^{p}(E) if and only if uLp(F)\displaystyle u\in L^{p}(F) for every FE\displaystyle F\subset\subset E. In the case 0<p<1\displaystyle 0<p<1, we denote by Lp(E)\displaystyle L^{p}(E) a set of measurable functions such that E|u(x)|p𝑑x<\displaystyle\int_{E}|u(x)|^{p}dx<\infty.

Definition 2.1.

The Sobolev space W1,p(Ω)\displaystyle W^{1,p}(\Omega), for 1p<\displaystyle 1\leq p<\infty, is defined as the Banach space of locally integrable weakly differentiable functions u:Ω\displaystyle u:\Omega\rightarrow\mathbb{R} equipped with the following norm

uW1,p(Ω)=uLp(Ω)+uLp(Ω).\|u\|_{W^{1,p}(\Omega)}=\|u\|_{L^{p}(\Omega)}+\|\nabla u\|_{L^{p}(\Omega)}.

The space W01,p(Ω)\displaystyle W_{0}^{1,p}(\Omega) is defined as the closure of the space 𝒞0(Ω)\displaystyle\mathcal{C}_{0}^{\infty}(\Omega), in the norm of the Sobolev space W1,p(Ω)\displaystyle W^{1,p}(\Omega), where 𝒞0(Ω)\displaystyle\mathcal{C}^{\infty}_{0}(\Omega) is the set of all smooth functions whose supports are compactly contained in Ω\displaystyle\Omega.

Definition 2.2.

Let 0<s<1\displaystyle 0<s<1 and Ω\displaystyle\Omega be a open connected subset of n\displaystyle\mathbb{R}^{n}. The fractional Sobolev space Ws,q(Ω)\displaystyle W^{s,q}(\Omega) for any 1q<+\displaystyle 1\leq q<+\infty is defined by

Ws,q(Ω)={uLq(Ω):|u(x)u(y)||xy|nq+sLq(Ω×Ω)},W^{s,q}(\Omega)=\left\{u\in L^{q}(\Omega):\frac{|u(x)-u(y)|}{|x-y|^{\frac{n}{q}+s}}\in L^{q}(\Omega\times\Omega)\right\},

and it is endowed with the norm

uWs,q(Ω)=(Ω|u(x)|q𝑑x+ΩΩ|u(x)u(y)|q|xy|n+qs𝑑x𝑑y)1/q.{}\|u\|_{W^{s,q}(\Omega)}=\left(\int_{\Omega}|u(x)|^{q}dx+\int_{\Omega}\int_{\Omega}\frac{|u(x)-u(y)|^{q}}{|x-y|^{n+qs}}dxdy\right)^{1/q}. (2.1)

It can be treated as an intermediate space between W1,q(Ω)\displaystyle W^{1,q}(\Omega) and Lq(Ω)\displaystyle L^{q}(\Omega). For 0<ss<1\displaystyle 0<s\leq s^{\prime}<1, Ws,q(Ω)\displaystyle W^{s^{\prime},q}(\Omega) is continuously embedded in Ws,q(Ω)\displaystyle W^{s,q}(\Omega), see [19, Proposition 2.1]. The fractional Sobolev space with zero boundary values is defined by

W0s,q(Ω)={uWs,q(n):u=0 in n\Ω}.W_{0}^{s,q}(\Omega)=\left\{u\in W^{s,q}(\mathbb{R}^{n}):u=0\text{ in }\mathbb{R}^{n}\backslash\Omega\right\}.

However W0s,q(Ω)\displaystyle W_{0}^{s,q}(\Omega) can be treated as the closure of 𝒞0(Ω)\displaystyle\mathcal{C}^{\infty}_{0}(\Omega) in Ws,q(Ω)\displaystyle W^{s,q}(\Omega) with respect to the fractional Sobolev norm defined in Eq. 2.1. Both Ws,q(Ω)\displaystyle W^{s,q}(\Omega) and W0s,q(Ω)\displaystyle W_{0}^{s,q}(\Omega) are reflexive Banach spaces, for q>1\displaystyle q>1, for details we refer to the readers [19, Section 2].
The following result asserts that the classical Sobolev space is continuously embedded in the fractional Sobolev space; see [19, Proposition 2.2]. The idea applies an extension property of Ω\displaystyle\Omega so that we can extend functions from W1,q(Ω)\displaystyle W^{1,q}(\Omega) to W1,q(n)\displaystyle W^{1,q}(\mathbb{R}^{n}) and that the extension operator is bounded.

Lemma 2.3.

Let Ω\displaystyle\Omega be a smooth bounded domain in n\displaystyle\mathbb{R}^{n} and 0<s<1\displaystyle 0<s<1. There exists a positive constant C=C(Ω,n,s)\displaystyle C=C(\Omega,n,s) such that

uWs,q(Ω)CuW1,q(Ω),\|u\|_{W^{s,q}(\Omega)}\leq C\|u\|_{W^{1,q}(\Omega)},

for every uW1,q(Ω)\displaystyle u\in W^{1,q}(\Omega).

For the fractional Sobolev spaces with zero boundary value, the next embedding result follows from [12, Lemma 2.1]. The fundamental difference of it compared to Lemma 2.3 is that the result holds for any bounded domain (without any condition of smoothness of the boundary), since for the Sobolev spaces with zero boundary value, we always have a zero extension to the complement.

Lemma 2.4.

Let Ω\displaystyle\Omega be a bounded domain in n\displaystyle\mathbb{R}^{n} and 0<s<1\displaystyle 0<s<1. There exists a positive constant C=C(n,s,Ω)\displaystyle C=C(n,s,\Omega) such that

nn|u(x)u(y)|q|xy|n+qs𝑑x𝑑yCΩ|u|q𝑑x\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{|u(x)-u(y)|^{q}}{|x-y|^{n+qs}}dxdy\leq C\int_{\Omega}|\nabla u|^{q}dx

for every uW01,q(Ω)\displaystyle u\in W_{0}^{1,q}(\Omega). Here, we consider the zero extension of u\displaystyle u to the complement of Ω\displaystyle\Omega.

We now proceed with the basic Poincaré inequality, which can be found in [20, Chapter 5, Section 5.8.1].

Lemma 2.5.

Let Ωn\displaystyle\Omega\subset\mathbb{R}^{n} be a bounded domain with 𝒞1\displaystyle\mathcal{C}^{1} boundary and q1\displaystyle q\geq 1. Then there exist a positive constant C>0\displaystyle C>0 depending only on n\displaystyle n and Ω\displaystyle\Omega, such that

Ω|u|q𝑑xCΩ|u|q𝑑x,uW01,q(Ω).\int_{\Omega}|u|^{q}dx\leq C\int_{\Omega}|\nabla u|^{q}dx,\qquad\forall u\in W^{1,q}_{0}(\Omega).

Specifically if we take Ω=Br\displaystyle\Omega=B_{r}, then we will get for all uW1,q(Br)\displaystyle u\in W^{1,q}(B_{r}),

Br|u(u)Br|q𝑑xcrqBr|u|q𝑑x,\fint_{B_{r}}\left|u-(u)_{B_{r}}\right|^{q}dx\leq cr^{q}\fint_{B_{r}}|\nabla u|^{q}dx,

where c\displaystyle c is a constant depending only on n\displaystyle n, and (u)Br\displaystyle(u)_{B_{r}} denotes the average of u\displaystyle u in Br\displaystyle B_{r}, and Br\displaystyle B_{r} denotes a ball of radius r\displaystyle r centered at x0n\displaystyle x_{0}\in\mathbb{R}^{n}. Here, \displaystyle\fint denotes the average integration.

Using Lemma 2.4, and the above Poincaré inequality, we observe that the following norm on the space W01,q(Ω)\displaystyle W^{1,q}_{0}(\Omega) defined by

uW01,q(Ω)=(Ω|u|q𝑑x+nn|u(x)u(y)|q|xy|n+qs𝑑x𝑑y)1q,\|u\|_{W^{1,q}_{0}(\Omega)}=\left(\int_{\Omega}|\nabla u|^{q}dx+\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{|u(x)-u(y)|^{q}}{|x-y|^{n+qs}}dxdy\right)^{\frac{1}{q}},

is equivalent to the norm

uW01,q(Ω)=(Ω|u|q𝑑x)1q.\|u\|_{W^{1,q}_{0}(\Omega)}=\left(\int_{\Omega}|\nabla u|^{q}dx\right)^{\frac{1}{q}}.

The following is a version of fractional Poincaré.

Lemma 2.6.

Let Ωn\displaystyle\Omega\subset\mathbb{R}^{n} be a bounded domain with 𝒞1\displaystyle\mathcal{C}^{1} boundary and let s(0,1)\displaystyle s\in(0,1) and q1\displaystyle q\geq 1. If uW0s,q(Ω)\displaystyle u\in W^{s,q}_{0}(\Omega), then

Ω|u|q𝑑xcΩΩ|u(x)u(y)|q|xy|n+qs𝑑x𝑑y,\int_{\Omega}|u|^{q}dx\leq c\int_{\Omega}\int_{\Omega}\frac{|u(x)-u(y)|^{q}}{|x-y|^{n+qs}}dxdy,

holds with cc(n,s,Ω)\displaystyle c\equiv c(n,s,\Omega).

In view of Lemma 2.6, we observe that the Banach space W0s,q(Ω)\displaystyle W_{0}^{s,q}(\Omega) can be endowed with the norm

uW0s,q(Ω)=(ΩΩ|u(x)u(y)|q|xy|n+qs𝑑x𝑑y)1q,\|u\|_{W_{0}^{s,q}(\Omega)}=\left(\int_{\Omega}\int_{\Omega}\frac{|u(x)-u(y)|^{q}}{|x-y|^{n+qs}}dxdy\right)^{\frac{1}{q}},

which is equivalent to that of uWs,q(Ω)\displaystyle\|u\|_{W^{s,q}(\Omega)}. For q=2\displaystyle q=2, the space Ws,2(Ω)\displaystyle W^{s,2}(\Omega) enjoys certain special properties and we denote Ws,2(Ω)=Hs(Ω)\displaystyle W^{s,2}(\Omega)=H^{s}(\Omega) and W0s,2(Ω)=H0s(Ω)\displaystyle W_{0}^{s,2}(\Omega)=H_{0}^{s}(\Omega). Endowed with the inner product

u,vH0s(Ω)=ΩΩ(u(x)u(y))(v(x)v(y))|xy|n+2s𝑑x𝑑y,\langle u,v\rangle_{H_{0}^{s}(\Omega)}=\int_{\Omega}\int_{\Omega}\frac{(u(x)-u(y))(v(x)-v(y))}{|x-y|^{n+2s}}dxdy,

we note that (H0s(Ω),H0s(Ω))\displaystyle(H_{0}^{s}(\Omega),\|\cdot\|_{H_{0}^{s}(\Omega)}) is a Hilbert space. Similar thing holds for the space W01,2(Ω)=H01(Ω).\displaystyle W^{1,2}_{0}(\Omega)=H^{1}_{0}(\Omega).

Definition 2.7.

The space X0s(Ω)\displaystyle X_{0}^{s}(\Omega) is defined as

X0s(Ω)={fHs(n) s.t. f=0 a.e. in 𝒞Ω},X_{0}^{s}(\Omega)=\left\{f\in H^{s}(\mathbb{R}^{n})\text{ s.t. }f=0\text{ a.e. in }\mathcal{C}\Omega\right\},

where 𝒞Ω=n\Ω\displaystyle\mathcal{C}\Omega=\mathbb{R}^{n}\backslash\Omega, and is endowed with the norm

uX0s(Ω)=(Q|u(x)u(y)|2|xy|n+2s𝑑x𝑑y)12,\|u\|_{X_{0}^{s}(\Omega)}=\left(\int_{Q}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}dxdy\right)^{\frac{1}{2}},

where Q:=2n\(𝒞Ω×𝒞Ω)\displaystyle Q:=\mathbb{R}^{2n}\backslash(\mathcal{C}\Omega\times\mathcal{C}\Omega).

Moreover W1,q(Ω)\displaystyle W^{-1,q^{\prime}}(\Omega), Ws,q(Ω)\displaystyle W^{-s,q^{\prime}}(\Omega) and Xs(Ω)\displaystyle X^{-s}(\Omega) are defined to be the dual spaces of W01,q(Ω)\displaystyle W_{0}^{1,q}(\Omega), W0s,q(Ω)\displaystyle W_{0}^{s,q}(\Omega) and X0s(Ω)\displaystyle X_{0}^{s}(\Omega) respectively, where q:=qq1\displaystyle q^{\prime}:=\frac{q}{q-1}. Now, we define the local spaces as

Wloc1,q(Ω)={u:Ω:uLq(K),K|u|qdx<, for every KΩ},W_{\operatorname{loc}}^{1,q}(\Omega)=\left\{u:\Omega\rightarrow\mathbb{R}:u\in L^{q}(K),\int_{K}|\nabla u|^{q}dx<\infty,\text{ for every }K\subset\subset\Omega\right\},

and

Wlocs,q(Ω)={u:Ω:uLq(K),KK|u(x)u(y)|q|xy|n+qsdxdy<, for every KΩ}.W_{\operatorname{loc}}^{s,q}(\Omega)=\left\{u:\Omega\rightarrow\mathbb{R}:u\in L^{q}(K),\int_{K}\int_{K}\frac{|u(x)-u(y)|^{q}}{|x-y|^{n+qs}}dxdy<\infty,\text{ for every }K\subset\subset\Omega\right\}.

Now for n>2\displaystyle n>2, we define the critical Sobolev exponent as 2=2nn2\displaystyle 2^{*}=\frac{2n}{n-2}, then we get the following embedding result for any open subset Ω\displaystyle\Omega of n\displaystyle\mathbb{R}^{n}, see for details [20, Chapter 5],

Theorem 2.8.

Let n>2\displaystyle n>2. Then, there exists a constant C\displaystyle C depending only on n\displaystyle n and Ω\displaystyle\Omega, such that for all u𝒞0(Ω)\displaystyle u\in\mathcal{C}_{0}^{\infty}(\Omega)

uL2(Ω)2CΩ|u|2𝑑x.\|u\|_{L^{2^{*}}(\Omega)}^{2}\leq C\int_{\Omega}|\nabla u|^{2}dx.

Similarly, for n>2s\displaystyle n>2s, we define the fractional Sobolev critical exponent as 2s=2nn2s\displaystyle 2_{s}^{*}=\frac{2n}{n-2s}. The following result is a fractional version of the Sobolev inequality(Theorem 2.8) which also implies a continuous embedding of H0s(Ω)\displaystyle H_{0}^{s}(\Omega) in the critical Lebesgue space L2s(Ω)\displaystyle L^{2_{s}^{*}}(\Omega). One can see the proof in [19].

Theorem 2.9.

Let 0<s<1\displaystyle 0<s<1 be such that n>2s\displaystyle n>2s. Then, there exists a constant S(n,s)\displaystyle S(n,s) depending only on n\displaystyle n and s\displaystyle s, such that for all u𝒞0(Ω)\displaystyle u\in\mathcal{C}_{0}^{\infty}(\Omega)

uL2s(Ω)2S(n,s)ΩΩ|u(x)u(y)|2|xy|n+2s𝑑x𝑑y.\|u\|_{L^{2_{s}^{*}}(\Omega)}^{2}\leq S(n,s)\int_{\Omega}\int_{\Omega}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}dxdy.

We now recall the Gagliardo-Nirenberg interpolation inequality that will be useful for proving the boundedness of weak solutions.

Theorem 2.10.

(Gagliardo-Nirenberg) Let 1q<+\displaystyle 1\leq q<+\infty be a positive real number. Let j\displaystyle j and m\displaystyle m be non-negative integers such that j<m\displaystyle j<m. Furthermore, let 1r+\displaystyle 1\leq r\leq+\infty be a positive extended real quantity, p1\displaystyle p\geq 1 be real and θ[0,1]\displaystyle\theta\in[0,1] such that the relations

1p=jn+θ(1rmn)+1θq,jmθ1\frac{1}{p}=\frac{j}{n}+\theta\left(\frac{1}{r}-\frac{m}{n}\right)+\frac{1-\theta}{q},\quad\frac{j}{m}\leq\theta\leq 1

hold. Then,

DjuLp(n)CDmuLr(n)θuLq(n)1θ\left\|D^{j}u\right\|_{L^{p}\left(\mathbb{R}^{n}\right)}\leq C\left\|D^{m}u\right\|_{L^{r}\left(\mathbb{R}^{n}\right)}^{\theta}\|u\|_{L^{q}\left(\mathbb{R}^{n}\right)}^{1-\theta}

for any uLq(n)\displaystyle u\in L^{q}(\mathbb{R}^{n}) such that DmuLr(n)\displaystyle D^{m}u\in L^{r}(\mathbb{R}^{n}). Here, the constant C>0\displaystyle C>0 depends on the parameters j,m,n,q,r,θ\displaystyle j,m,n,q,r,\theta, but not on u\displaystyle u.

The article will extensively use the embedding results and corresponding inequalities. Now, we need to deal with spaces involving time for the parabolic equations, so we introduce them here. As in the classical case, we define the corresponding Bochner spaces as the following

Lq(0,T;W01,q(Ω))={uLq(Ω×(0,T)),uLq(0,T;W01,q(Ω))<},Lq(0,T;W0s,q(Ω))={uLq(Ω×(0,T)),uLq(0,T;W0s,q(Ω))<},L2(0,T;X0s(Ω))={uL2(n×(0,T)),uL2(0,T;X0s(Ω))<},\begin{array}[]{c}L^{q}(0,T;W_{0}^{1,q}(\Omega))=\left\{u\in L^{q}(\Omega\times(0,T)),\|u\|_{L^{q}(0,T;W_{0}^{1,q}(\Omega))}<\infty\right\},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ L^{q}(0,T;W_{0}^{s,q}(\Omega))=\left\{u\in L^{q}(\Omega\times(0,T)),\|u\|_{L^{q}(0,T;W_{0}^{s,q}(\Omega))}<\infty\right\},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ L^{2}(0,T;X_{0}^{s}(\Omega))=\left\{u\in L^{2}(\mathbb{R}^{n}\times(0,T)),\|u\|_{L^{2}(0,T;X^{s}_{0}(\Omega))}<\infty\right\},\end{array}

where

uLq(0,T;W01,q(Ω))=(0TΩ|u|q𝑑x𝑑t)1q,uLq(0,T;W0s,q(Ω))=(0TΩΩ|u(x,t)u(y,t)|q|xy|n+qs𝑑x𝑑y𝑑t)1q,uL2(0,T;X0s(Ω))=(0TQ|u(x,t)u(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t)12,\begin{array}[]{c}\|u\|_{L^{q}(0,T;W_{0}^{1,q}(\Omega))}=\left(\int_{0}^{T}\int_{\Omega}|\nabla u|^{q}dxdt\right)^{\frac{1}{q}},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \|u\|_{L^{q}(0,T;W_{0}^{s,q}(\Omega))}=\left(\int_{0}^{T}\int_{\Omega}\int_{\Omega}\frac{|u(x,t)-u(y,t)|^{q}}{|x-y|^{n+qs}}dxdydt\right)^{\frac{1}{q}},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \|u\|_{L^{2}(0,T;X_{0}^{s}(\Omega))}=\left(\int_{0}^{T}\int_{Q}\frac{|u(x,t)-u(y,t)|^{2}}{|x-y|^{n+2s}}dxdydt\right)^{\frac{1}{2}},\end{array}

with their dual spaces Lq(0,T;W1,q(Ω))\displaystyle L^{q^{\prime}}(0,T;W^{-1,q^{\prime}}(\Omega)), Lq(0,T;Ws,q(Ω))\displaystyle L^{q^{\prime}}(0,T;W^{-s,q^{\prime}}(\Omega)) and L2(0,T;Xs(Ω))\displaystyle L^{2}(0,T;X^{-s}(\Omega)) respectively. Again, the local spaces are defined as

L2(0,T;Hloc1(Ω))={uL2(K×(0,T)):0TK|u|qdxdt<, for every KΩ},L^{2}(0,T;H_{loc}^{1}(\Omega))=\left\{u\in L^{2}(K\times(0,T)):\int_{0}^{T}\int_{K}|\nabla u|^{q}dxdt<\infty,\mbox{ for every }K\subset\subset\Omega\right\},

and

L2(0,T;Hlocs(Ω))={uL2(K×(0,T)):0TKK|u(x,t)u(y,t)|2|xy|n+2sdxdydt<, for every KΩ}.L^{2}(0,T;H_{loc}^{s}(\Omega))=\left\{u\in L^{2}(K\times(0,T)):\int_{0}^{T}\int_{K}\int_{K}\frac{|u(x,t)-u(y,t)|^{2}}{|x-y|^{n+2s}}dxdydt<\infty,\mbox{ for every }K\subset\subset\Omega\right\}.

We now recall the following algebraic inequality that can be found in [1, Lemma 2.22].

Lemma 2.11.
  1. i)

    Let α>0\displaystyle\alpha>0. For every x,y0\displaystyle x,y\geq 0 one has

    (xy)(xαyα)4α(α+1)2(xα+12yα+12)2.(x-y)(x^{\alpha}-y^{\alpha})\geq\frac{4\alpha}{(\alpha+1)^{2}}\left(x^{\frac{\alpha+1}{2}}-y^{\frac{\alpha+1}{2}}\right)^{2}.
  2. ii)

    Let 0<α1\displaystyle 0<\alpha\leq 1. For every x,y0\displaystyle x,y\geq 0 with xy\displaystyle x\neq y one has

    xyxαyα1α(x1α+y1α).\frac{x-y}{x^{\alpha}-y^{\alpha}}\leq\frac{1}{\alpha}(x^{1-\alpha}+y^{1-\alpha}).
  3. iii)

    Let α1\displaystyle\alpha\geq 1. Then there exists a constant Cα\displaystyle C_{\alpha} depending only on α\displaystyle\alpha such that

    |x+y|α1|xy|Cα|xαyα|.|x+y|^{\alpha-1}|x-y|\leq C_{\alpha}\left|x^{\alpha}-y^{\alpha}\right|.

2.3 Weak Solutions

In this subsection, along with the next subsection, we will introduce notions of very weak solutions to our problem and state the main results that we are going to prove. We begin with the definitions of weak solutions for the nonsingular case. We first take f\displaystyle f and u0\displaystyle u_{0} to be in L2\displaystyle L^{2} spaces and then relax the condition. We also state some important properties of the weak solutions that we need to use in the rest of the article. Further, we will introduce suitable approximating problems and properties of their solutions.

Definition 2.12.

Assume (f,u0)L2(ΩT)×L2(Ω)\displaystyle(f,u_{0})\in L^{2}(\Omega_{T})\times L^{2}(\Omega), then we say that u\displaystyle u is an energy solution to problem

{utΔu+(Δ)su=f(x,t) in ΩT,u=0 in (n\Ω)×(0,T),u(x,0)=u0(x) in Ω;{}\begin{array}[]{lcr}\begin{cases}u_{t}-\Delta u+(-\Delta)^{s}u={f(x,t)}&\mbox{ in }\Omega_{T},\\ u=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u(x,0)=u_{0}(x)&\mbox{ in }\Omega;\end{cases}\end{array} (2.2)

if uL2(0,T;H01(Ω))𝒞([0,T],L2(Ω)),utL2(0,T;H1(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega))\cap\mathcal{C}([0,T],L^{2}(\Omega)),u_{t}\in L^{2}(0,T;H^{-1}(\Omega)), and for all ϕL2(0,T;H01(Ω))\displaystyle\phi\in L^{2}(0,T;H_{0}^{1}(\Omega)) we have

0Tut,ϕ𝑑t+0TΩuϕdxdt+120Tnn(u(x,t)u(y,t))(ϕ(x,t)ϕ(y,t))|xy|n+2s𝑑x𝑑y𝑑t=0TΩfϕ𝑑x𝑑t\int_{0}^{T}\left\langle u_{t},\phi\right\rangle dt+\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt+\frac{1}{2}\int_{0}^{T}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}\frac{(u(x,t)-u(y,t))(\phi(x,t)-\phi(y,t))}{|x-y|^{n+2s}}dxdydt=\int_{0}^{T}\int_{\Omega}f\phi dxdt

and u(,t)u0\displaystyle u(\cdot,t)\rightarrow u_{0} strongly in L2(Ω)\displaystyle L^{2}(\Omega), as t0\displaystyle t\rightarrow 0.

We denote

(u(x,t),ϕ(x,t)):=12nn(u(x,t)u(y,t))(ϕ(x,t)ϕ(y,t))×K(x,y,t)𝑑x𝑑y,\mathcal{E}(u(x,t),\phi(x,t)):=\frac{1}{2}\int_{\mathbb{R}^{n}}\int_{\mathbb{R}^{n}}(u(x,t)-u(y,t))(\phi(x,t)-\phi(y,t))\times K(x,y,t)dxdy,

where K(x,y,t)=1|xy|n+2s.\displaystyle K(x,y,t)=\frac{1}{{{\left|x-y\right|}^{n+2s}}}. Following the way for fractional Laplacian in [29], we give the proof of existence for mixed local-nonlocal case for the sake of completeness.

Theorem 2.13.

There exists a solution to problem Eq. 2.2 in the sense of Definition 2.12. Moreover, if f\displaystyle f is also a nonnegative function and u00\displaystyle u_{0}\geq 0, then the solution is also nonnegative.

Proof.

Let us denote 𝒞(Ω×[0,T])\displaystyle\mathcal{C}_{*}^{\infty}(\Omega\times[0,T]) as the 𝒞(Ω×[0,T])\displaystyle\mathcal{C}^{\infty}(\Omega\times[0,T]) functions that vanish in (n\Ω)×[0,T]\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times[0,T] and in Ω×{T}\displaystyle\Omega\times\{T\}. Choosing ϕ𝒞(Ω×[0,T])\displaystyle\phi\in\mathcal{C}_{*}^{\infty}(\Omega\times[0,T]), for uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega)), we define the operator

Lϕ(u):=0TΩuϕtdxdt+0TΩuϕdxdt+0T(u(x,t),ϕ(x,t))𝑑t.L_{\phi}(u):=\int_{0}^{T}\int_{\Omega}-u\phi_{t}dxdt+\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt+\int_{0}^{T}\mathcal{E}(u(x,t),\phi(x,t))dt.

We now observe that u\displaystyle u is an energy solution to Eq. 2.2 with fL2(ΩT)L2(0,T;H1(Ω))\displaystyle f\in L^{2}(\Omega_{T})\subset L^{2}(0,T;H^{-1}(\Omega)) if and only if

Lϕ(u)=0Tf,ϕ𝑑t+Ωu(x,0)ϕ(x,0)𝑑x,L_{\phi}(u)=\int_{0}^{T}\langle f,\phi\rangle dt+\int_{\Omega}u(x,0)\phi(x,0)dx,

where f,ϕ\displaystyle\langle f,\phi\rangle denote the usual inner product of f\displaystyle f and ϕ\displaystyle\phi in L2(Ω)\displaystyle L^{2}(\Omega) or the pairing of f\displaystyle f and ϕ\displaystyle\phi between H1(Ω)\displaystyle H^{-1}(\Omega) and H01(Ω)\displaystyle H^{1}_{0}(\Omega).
We also define the following inner product, for ϕ,φ𝒞(Ω×[0,T])\displaystyle\phi,\varphi\in\mathcal{C}_{*}^{\infty}(\Omega\times[0,T])

φ,ϕ=12φ(x,0),ϕ(x,0)L2(Ω)+0TΩϕφdxdt+0T(φ(x,t),ϕ(x,t))𝑑t,{}\langle\varphi,\phi\rangle_{*}=\frac{1}{2}\langle\varphi(x,0),\phi(x,0)\rangle_{L^{2}(\Omega)}+\int_{0}^{T}\int_{\Omega}\nabla\phi\cdot\nabla\varphi~{}dxdt+\int_{0}^{T}\mathcal{E}(\varphi(x,t),\phi(x,t))dt, (2.3)

and denote by H(Ω×[0,T])\displaystyle H^{*}(\Omega\times[0,T]) the Hilbert space built as the completion of 𝒞(Ω×[0,T])\displaystyle\mathcal{C}_{*}^{\infty}(\Omega\times[0,T]) with the norm ϕ\displaystyle\|\phi\|_{*} induced by the inner product Eq. 2.3.
Now Lϕ\displaystyle L_{\phi} is clearly a linear functional in H(Ω×[0,T])\displaystyle H^{*}(\Omega\times[0,T]) and for any φL2(0,T;H01(Ω))\displaystyle\varphi\in L^{2}(0,T;H_{0}^{1}(\Omega)), by Hölder and Sobolev inequalities, we have

|Lϕ(φ)|cϕ(φL2(0,T;L2(Ω))+φL2(0,T,H01(Ω))+φL2(0,T,H0s(Ω)))c~ϕφ.\left|L_{\phi}(\varphi)\right|\leq c_{\phi}\left(\|\varphi\|_{L^{2}(0,T;L^{2}(\Omega))}+\|\varphi\|_{L^{2}(0,T,H_{0}^{1}(\Omega))}+\|\varphi\|_{L^{2}(0,T,H_{0}^{s}(\Omega))}\right)\leq\tilde{c}_{\phi}\|\varphi\|_{*}.

Therefore, Lϕ\displaystyle L_{\phi} is a bounded linear functional in H(Ω×[0,T])\displaystyle H^{*}(\Omega\times[0,T]), and hence by the Fréchet-Riesz Theorem, there exists 𝒯ϕH(Ω×[0,T])\displaystyle\mathcal{T}\phi\in H^{*}(\Omega\times[0,T]) such that

Lϕ(φ)=φ,𝒯ϕ for all φH(Ω×[0,T]).L_{\phi}(\varphi)=\langle\varphi,\mathcal{T}\phi\rangle_{*}\quad\text{ for all }\varphi\in H^{*}(\Omega\times[0,T]).

It is trivial to show that 𝒯\displaystyle\mathcal{T} is a linear operator in H(Ω×[0,T])\displaystyle H^{*}(\Omega\times[0,T]). Moreover

Lϕ(ϕ)=12Ωϕ2(x,0)𝑑x+0TΩ|ϕ|2𝑑x𝑑t+0T(ϕ(x,t),ϕ(x,t))𝑑t=ϕ2,L_{\phi}(\phi)=\frac{1}{2}\int_{\Omega}\phi^{2}(x,0)dx+\int_{0}^{T}\int_{\Omega}|\nabla\phi|^{2}dxdt+\int_{0}^{T}\mathcal{E}(\phi(x,t),\phi(x,t))dt=\|\phi\|_{*}^{2},

and consequently, ϕ,𝒯ϕ=ϕ2\displaystyle\langle\phi,\mathcal{T}\phi\rangle_{*}=\|\phi\|_{*}^{2}. Then we get by the Cauchy-Schwartz inequality,

ϕ2ϕ𝒯ϕ, i.e., ϕ𝒯ϕ.\|\phi\|_{*}^{2}\leq\|\phi\|_{*}\|\mathcal{T}\phi\|_{*},\quad\text{ i.e., }\quad\|\phi\|_{*}\leq\|\mathcal{T}\phi\|_{*}.

Therefore, this implies that 𝒯\displaystyle\mathcal{T} is injective and hence bijective on its range, and its inverse 𝒯1\displaystyle\mathcal{T}^{-1} has a norm less than or equal to 1\displaystyle 1 and can be extended to the closureM\displaystyle\operatorname{closure}M of Range(𝒯)\displaystyle\operatorname{Range}(\mathcal{T}).
Now, on the other hand, we define

Bu0,f(ϕ):=Ωu0ϕ(x,0)𝑑x+0TΩϕf𝑑x𝑑t.B_{u_{0},f}(\phi):=\int_{\Omega}u_{0}\phi(x,0)dx+\int_{0}^{T}\int_{\Omega}\phi fdxdt.

Denoting ϕ0:=ϕ(x,0)\displaystyle\phi_{0}:=\phi(x,0), we get by Hölder inequality

|Bu0,f(ϕ)|u0L2(Ω)ϕ0L2(Ω)+fL2(0,T;L2(Ω))ϕL2(0,T;L2(Ω))cu0,fϕ,\left|B_{u_{0},f}(\phi)\right|\leq\left\|u_{0}\right\|_{L^{2}(\Omega)}\left\|\phi_{0}\right\|_{L^{2}(\Omega)}+\|f\|_{L^{2}(0,T;L^{2}(\Omega))}\|\phi\|_{L^{2}(0,T;L^{2}(\Omega))}\leq c_{u_{0},f}\|\phi\|_{*},

and thus,

|Bu0,f(𝒯1ψ)|c𝒯1ψcψ.\left|B_{u_{0},f}(\mathcal{T}^{-1}\psi)\right|\leq c\left\|\mathcal{T}^{-1}\psi\right\|_{*}\leq c\|\psi\|_{*}.

Since Bu0,f\displaystyle B_{u_{0},f} and 𝒯1\displaystyle\mathcal{T}^{-1} both are linear, therefore their composition is also so, and by above line, Bu0,f𝒯1\displaystyle B_{u_{0},f}\circ\mathcal{T}^{-1} is bounded too, therefore, by applying the Fréchet-Riesz Theorem again, there exists a unique uM\displaystyle u\in M such that Bu0,f(𝒯1ψ)=ψ,u\displaystyle B_{u_{0},f}(\mathcal{T}^{-1}\psi)=\langle\psi,u\rangle_{*} for every ψM\displaystyle\psi\in M. We denote ϕ=𝒯1ψ\displaystyle\phi=\mathcal{T}^{-1}\psi and so

Bu0,f(ϕ)=𝒯ϕ,u=Lϕ(u)B_{u_{0},f}(\phi)=\langle\mathcal{T}\phi,u\rangle_{*}=L_{\phi}(u)

that is,

0TΩuϕtdxdt+0TΩuϕdxdt+0T(u(x,t),ϕ(x,t))𝑑t=0TΩfϕ𝑑x𝑑t+Ωu(x,0)ϕ(x,0)𝑑x,\begin{array}[]{c}\int_{0}^{T}\int_{\Omega}-u\phi_{t}dxdt+\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt+\int_{0}^{T}\mathcal{E}(u(x,t),\phi(x,t))dt=\int_{0}^{T}\int_{\Omega}f\phi dxdt+\int_{\Omega}u(x,0)\phi(x,0)dx,\end{array}

where ϕL2(0,T;H0s(Ω))L2(0,T;H01(Ω))L2(0,T;H01(Ω))\displaystyle\phi\in L^{2}(0,T;H_{0}^{s}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega))\equiv L^{2}(0,T;H_{0}^{1}(\Omega)) and ϕtL2(0,T;H1(Ω))\displaystyle\phi_{t}\in L^{2}(0,T;H^{-1}(\Omega)). Finally, by a density argument, one can conclude, integrating by parts, that uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega)), utL2(0,T;H1(Ω))\displaystyle u_{t}\in L^{2}(0,T;H^{-1}(\Omega)), and

0TΩutϕ𝑑x𝑑t+0TΩuϕdxdt+0T(u(x,t),ϕ(x,t))𝑑t=0TΩfϕ𝑑x𝑑t.\int_{0}^{T}\int_{\Omega}u_{t}\phi dxdt+\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt+\int_{0}^{T}\mathcal{E}(u(x,t),\phi(x,t))dt=\int_{0}^{T}\int_{\Omega}f\phi dxdt.

Since uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega)), utL2(0,T;H1(Ω))\displaystyle u_{t}\in L^{2}(0,T;H^{-1}(\Omega)) implies that u𝒞([0,T],L2(Ω))\displaystyle u\in\mathcal{C}([0,T],L^{2}(\Omega)) and so we have u(,t)u0\displaystyle u(\cdot,t)\rightarrow u_{0} strongly in L2(Ω)\displaystyle L^{2}(\Omega), as t0\displaystyle t\rightarrow 0. Thus u(x,t)\displaystyle u(x,t) is an energy solution of Eq. 2.2.
Now we show that u0\displaystyle u\geq 0 provided that f\displaystyle f and u0\displaystyle u_{0} are nonnegative. We write u=u+u\displaystyle u=u_{+}-u_{-}, where u+=max{u,0}χΩ\displaystyle u_{+}=\operatorname{max}{\{u,0}\}\chi_{\Omega} and u=max{u,0}χΩ\displaystyle u_{-}=\operatorname{max}{\{-u,0}\}\chi_{\Omega}. We take ϕ=uχ(0,t~),t~>0\displaystyle\phi=u_{-}\chi_{(0,\tilde{t})},\tilde{t}>0, as a test function. Since f0\displaystyle f\geq 0, and ϕ0\displaystyle\phi\geq 0, we have

00TΩfϕ𝑑x𝑑t=0t~Ωfu𝑑x𝑑t.{}0\leq\int_{0}^{T}\int_{\Omega}f\phi dxdt=\int_{0}^{\tilde{t}}\int_{\Omega}fu_{-}dxdt. (2.4)

On the other hand, since u\displaystyle u is 0\displaystyle 0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T), we have that

0T2n(u(x,t)u(y,t))(ϕ(x,t)ϕ(y,t))K(x,y,t)𝑑x𝑑y𝑑t=0t~2n\(Ωc×Ωc)(u(x,t)u(y,t))(u(x,t)u(y,t))K(x,y,t)𝑑x𝑑y𝑑t=0t~ΩΩ(u(x,t)u(y,t))(u(x,t)u(y,t))K(x,y,t)𝑑x𝑑y𝑑t+20t~ΩΩcu(x,t)u(x,t)K(x,y,t)𝑑y𝑑x𝑑t.\begin{array}[]{c}\int_{0}^{T}\iint_{\mathbb{R}^{2n}}(u(x,t)-u(y,t))(\phi(x,t)-\phi(y,t))K(x,y,t)dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ =\int_{0}^{\tilde{t}}\iint_{\mathbb{R}^{2n}\backslash(\Omega^{c}\times\Omega^{c})}(u(x,t)-u(y,t))(u_{-}(x,t)-u_{-}(y,t))K(x,y,t)dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ =\int_{0}^{\tilde{t}}\int_{\Omega}\int_{\Omega}(u(x,t)-u(y,t))(u_{-}(x,t)-u_{-}(y,t))K(x,y,t)dxdydt+2\int_{0}^{\tilde{t}}\int_{\Omega}\int_{\Omega^{c}}u(x,t)u_{-}(x,t)K(x,y,t)dydxdt.\end{array}

Moreover, (u+(x,t)u+(y,t))(u(x,t)u(y,t))0\displaystyle(u_{+}(x,t)-u_{+}(y,t))(u_{-}(x,t)-u_{-}(y,t))\leq 0, and thus

0t~ΩΩ(u(x,t)u(y,t))(u(x,t)u(y,t))K(x,y,t)𝑑x𝑑y𝑑t0t~ΩΩ(u(x,t)u(y,t))2K(x,y,t)𝑑x𝑑y𝑑t0.\begin{array}[]{l}\int_{0}^{\tilde{t}}\int_{\Omega}\int_{\Omega}(u(x,t)-u(y,t))(u_{-}(x,t)-u_{-}(y,t))K(x,y,t)dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq-\int_{0}^{\tilde{t}}\int_{\Omega}\int_{\Omega}(u_{-}(x,t)-u_{-}(y,t))^{2}K(x,y,t)dxdydt\leq 0.\end{array}

Further, we have

0t~ΩΩcu(x,t)u(x,t)K(x,y,t)𝑑y𝑑x𝑑t=0t~ΩΩcu2(x,t)K(x,y,t)𝑑y𝑑x𝑑t0.\int_{0}^{\tilde{t}}\int_{\Omega}\int_{\Omega^{c}}u(x,t)u_{-}(x,t)K(x,y,t)dydxdt=-\int_{0}^{\tilde{t}}\int_{\Omega}\int_{\Omega^{c}}u_{-}^{2}(x,t)K(x,y,t)dydxdt\leq 0.

Therefore, we have shown that

0T2n(u(x,t)u(y,t))(ϕ(x,t)ϕ(y,t))K(x,y,t)𝑑x𝑑y𝑑t0,\int_{0}^{T}\iint_{\mathbb{R}^{2n}}(u(x,t)-u(y,t))(\phi(x,t)-\phi(y,t))K(x,y,t)dxdydt\leq 0,

Similarly since u+u=0\displaystyle\nabla u_{+}\cdot\nabla u_{-}=0, we get

0TΩuϕdxdt=0t~Ωu+udxdt0t~Ω|u|2𝑑x𝑑t0.\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt=\int_{0}^{\tilde{t}}\int_{\Omega}\nabla u_{+}\cdot\nabla u_{-}dxdt-\int_{0}^{\tilde{t}}\int_{\Omega}|\nabla u_{-}|^{2}dxdt\leq 0.

Now since u00\displaystyle u_{0}\geq 0, so u(,0)0\displaystyle u_{-}(\cdot,0)\equiv 0, and we get

0Tut,ϕ𝑑t=0t~Ωutu𝑑x𝑑t=Ωu(x,0)u(x,t~)σ𝑑σ𝑑x=12Ω(u(x,t~))2𝑑x.\begin{array}[]{rcl}\int_{0}^{T}\left\langle u_{t},\phi\right\rangle dt&=&\int_{0}^{\tilde{t}}\int_{\Omega}\frac{\partial u}{\partial t}u_{-}dxdt=\int_{\Omega}\int_{u(x,0)}^{u(x,\tilde{t})}\sigma_{-}d\sigma dx=-\frac{1}{2}\int_{\Omega}(u_{-}(x,{\tilde{t}}))^{2}dx.\end{array}

Combining the above three inequalities, we get from Eq. 2.4 that

00TΩfϕ𝑑x𝑑t=0Tut,ϕ𝑑t+0TΩuϕdxdt+120T2n(u(x,t)u(y,t))(ϕ(x,t)ϕ(y,t))|xy|n+2s𝑑x𝑑y𝑑t12Ω(u(x,t~))2𝑑x0,\begin{array}[]{rcl}0\leq\int_{0}^{T}\int_{\Omega}f\phi dxdt&=&\int_{0}^{T}\left\langle u_{t},\phi\right\rangle dt+\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &&+\frac{1}{2}\int_{0}^{T}\iint_{\mathbb{R}^{2n}}\frac{(u(x,t)-u(y,t))(\phi(x,t)-\phi(y,t))}{|x-y|^{n+2s}}dxdydt\leq-\frac{1}{2}\int_{\Omega}(u_{-}(x,{\tilde{t}}))^{2}dx\leq 0,\end{array}

and this gives that u(,t~)L2(Ω)=0\displaystyle||u_{-}(\cdot,\tilde{t})||_{L^{2}(\Omega)}=0 for each t~>0\displaystyle\tilde{t}>0. Therefore u0\displaystyle u_{-}\equiv 0. So we conclude that u0\displaystyle u\geq 0. We observe that this comparison result also guarantees the uniqueness of energy solution to Eq. 2.2. ∎

Remark 2.14.

Observing that (u(x,t)u(y,t))((u(x,t)k)+(u(y,t)k)+)0\displaystyle(u(x,t)-u(y,t))((u(x,t)-k)_{+}-(u(y,t)-k)_{+})\geq 0, for each k\displaystyle k, we can show that for (f,u0)L(ΩT)×L(Ω)\displaystyle(f,u_{0})\in L^{\infty}(\Omega_{T})\times L^{\infty}(\Omega), the weak solution uL(ΩT)\displaystyle u\in L^{\infty}(\Omega_{T}). The proof will follow exactly similar to that of [32, Theorem 4.2.1].

Now we relax the spaces where f\displaystyle f and u0\displaystyle u_{0} lie. For the case of L1\displaystyle L^{1} data, we consider the set

𝒯:={ϕ:Ω×[0,T], s.t. ϕtΔϕ+(Δ)sϕ=φ in ΩT,φ𝒞0(ΩT),ϕL(Ω×(0,T))𝒞locα,β(Ω×(0,T)),ϕ(x,0)L(Ω),ϕ=0 in (n\Ω)×(0,T],ϕ(x,T)=0 in Ω},\begin{array}[]{rl}\mathcal{T}:=&\{\phi:\Omega\times[0,T]\rightarrow\mathbb{R},\text{ s.t. }-\phi_{t}-\Delta\phi+(-\Delta)^{s}\phi=\varphi\text{ in }\Omega_{T},\\ &\varphi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}),\phi\in L^{\infty}(\Omega\times(0,T))\cap\mathcal{C}_{\operatorname{loc}}^{\alpha,\beta}(\Omega\times(0,T)),\\ &\phi(x,0)\in L^{\infty}(\Omega),\phi=0\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T],\phi(x,T)=0\text{ in }\Omega\},\end{array}

where α,β(0,1)\displaystyle\alpha,\beta\in(0,1).

Definition 2.15.

Let (f,u0)L1(ΩT)×L1(Ω)\displaystyle(f,u_{0})\in L^{1}(\Omega_{T})\times L^{1}(\Omega) be nonnegative functions. Then uL1(ΩT)\displaystyle u\in L^{1}(\Omega_{T}) is a very weak solution to Eq. 2.2 if we have

ΩTu(ϕtΔϕ+(Δ)sϕ)𝑑x𝑑t=ΩTfϕ𝑑x𝑑t+Ωu0(x)ϕ(x,0)𝑑x,ϕ𝒯.\iint_{\Omega_{T}}u\left(-\phi_{t}-\Delta\phi+(-\Delta)^{s}\phi\right)dxdt=\iint_{\Omega_{T}}f\phi dxdt+\int_{\Omega}u_{0}(x)\phi(x,0)dx,\quad\forall\phi\in\mathcal{T}.

The next existence result is following the lines of [29].

Theorem 2.16.

For (f,u0)L1(ΩT)×L1(Ω)\displaystyle(f,u_{0})\in L^{1}(\Omega_{T})\times L^{1}(\Omega) being nonnegative, Eq. 2.2 has a unique nonnegative very weak solution u\displaystyle u in the sense of Definition 2.15.

Proof.

Firstly, we observe that the existence of valid test functions is guaranteed by the result in [16], [21], [30]. We will now obtain the solution as a limit of solutions to approximated problems. Let umL2(0,T;H01(Ω))L(ΩT)\displaystyle u_{m}\in L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(\Omega_{T}) be the solution (exists by Theorem 2.13) to the approximated problem

{(um)tΔum+(Δ)sum=fm in ΩT,um(x,t)=0 in (n\Ω)×(0,T),um(x,0)=u0m(x) in Ω;\displaystyle\displaystyle\begin{cases}(u_{m})_{t}-\Delta u_{m}+(-\Delta)^{s}u_{m}=f_{m}&\text{ in }\Omega_{T},\\ u_{m}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u_{m}(x,0)=u_{0m}(x)&\text{ in }\Omega;\end{cases}

where fm=Tm(f(x,t))\displaystyle f_{m}=T_{m}(f(x,t)) and u0m=Tm(u0(x))\displaystyle u_{0m}=T_{m}(u_{0}(x)) are L\displaystyle L^{\infty} functions. Using (Tk(um))χ(0,t)\displaystyle(T_{k}(u_{m}))\chi_{(0,t)}, for t>0\displaystyle t>0 (admissible by [29, Proposition 3]) as a test function in the approximated problem, it holds that,

ΩLk(um(x,t))𝑑x+0tΩumTk(um)𝑑x𝑑θ+120tQ(Tk(um(x,θ))Tk(um(y,θ)))(um(x,θ)um(y,θ))|xy|n+2s𝑑x𝑑y𝑑θkfL1(ΩT)+C3(k)u0L1(Ω)+C4(k)|Ω|,{}\begin{array}[]{l}\int_{\Omega}L_{k}(u_{m}(x,t))dx+\int_{0}^{t}\int_{\Omega}\nabla u_{m}\cdot\nabla T_{k}(u_{m})dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(T_{k}\left(u_{m}(x,\theta)\right)-T_{k}\left(u_{m}(y,\theta)\right)\right)\left(u_{m}(x,\theta)-u_{m}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq k\|f\|_{L^{1}(\Omega_{T})}+C_{3}(k)\left\|u_{0}\right\|_{L^{1}(\Omega)}+C_{4}(k)|\Omega|,\end{array} (2.5)

where Lk(ρ)=0ρ(Tk(ξ))𝑑ξ.\displaystyle L_{k}(\rho)=\int_{0}^{\rho}\left(T_{k}(\xi)\right)d\xi. Notice that

(Tk(um(x,θ))Tk(um(y,θ)))(um(x,θ)um(y,θ))(Tk(um(x,θ))Tk(um(y,θ)))2\left(T_{k}\left(u_{m}(x,\theta)\right)-T_{k}\left(u_{m}(y,\theta)\right)\right)\left(u_{m}(x,\theta)-u_{m}(y,\theta)\right)\geq\left(T_{k}\left(u_{m}(x,\theta)\right)-T_{k}\left(u_{m}(y,\theta)\right)\right)^{2}

and so

umTk(um)|Tk(um)|2\nabla u_{m}\cdot\nabla T_{k}(u_{m})\geq|\nabla T_{k}(u_{m})|^{2}

and for ρ>0\displaystyle\rho>0 we have

C1(k)ρC2(k)Lk(ρ)C3(k)ρ+C4(k)C_{1}(k)\rho-C_{2}(k)\leq L_{k}(\rho)\leq C_{3}(k)\rho+C_{4}(k)

and

Lk(ρ)C(Tk(ρ))2,L_{k}(\rho)\geq C\left(T_{k}(\rho)\right)^{2},

where C1,C2,C3,C4\displaystyle C_{1},C_{2},C_{3},C_{4} are constants depending only on k\displaystyle k and independent of m\displaystyle m. Therefore taking supremum over t(0,T]\displaystyle t\in(0,T] in Eq. 2.5, we get that {Tk(um)}m\displaystyle\left\{T_{k}(u_{m})\right\}_{m} is bounded in L(0,T;L2(Ω))L2(0,T;H0s(Ω))L2(0,T;H01(Ω))\displaystyle L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{s}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)) and {um}m\displaystyle\left\{u_{m}\right\}_{m} is bounded in L(0,T;L1(Ω))L1(ΩT)\displaystyle L^{\infty}(0,T;L^{1}(\Omega))\subset L^{1}(\Omega_{T}).
Now since by comparison principle proved in Theorem 2.13, {um}m\displaystyle\left\{u_{m}\right\}_{m} is increasing in m\displaystyle m, we get the existence of a measurable function u\displaystyle u such that Tk(u)L(0,T;L2(Ω))L2(0,T;H0s(Ω))L2(0,T;H01(Ω))\displaystyle T_{k}(u)\in L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{s}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)), umu\displaystyle u_{m}\uparrow u strongly in L1(ΩT)\displaystyle L^{1}(\Omega_{T}) and umu\displaystyle u_{m}\uparrow u a.e in ΩT\displaystyle\Omega_{T}. As each um\displaystyle u_{m} is an energy solution, therefore, um𝒞([0,T],L2(Ω))𝒞([0,T],L1(Ω))\displaystyle u_{m}\in\mathcal{C}([0,T],L^{2}(\Omega))\subset\mathcal{C}([0,T],L^{1}(\Omega)) and at each time level t[0,T]\displaystyle t\in[0,T], we have um(,t)L1(Ω)\displaystyle u_{m}(\cdot,t)\in L^{1}(\Omega), this along with the monotonicity of {um}m\displaystyle\{u_{m}\}_{m} in m\displaystyle m allows us to define the pointwise limit (a.e.) u\displaystyle u of {um}m\displaystyle\{u_{m}\}_{m} in Ω\displaystyle\Omega for each time t[0,T]\displaystyle t\in[0,T]. Also u\displaystyle u satisfies u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) in L1\displaystyle L^{1} sense. Again as each um=0\displaystyle u_{m}=0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T), therefore u\displaystyle u also satisfies the same. We now prove that u\displaystyle u is a weak solution to Eq. 2.2 in the sense of Definition 2.15. Let ϕ𝒯\displaystyle\phi\in\mathcal{T}, then as um\displaystyle u_{m} is the energy solution to the approximated problem, we have

ΩT((um)tΔum+(Δ)sum)ϕ𝑑x𝑑t=ΩTfmϕ𝑑x𝑑t.\iint_{\Omega_{T}}\left((u_{m})_{t}-\Delta u_{m}+(-\Delta)^{s}u_{m}\right)\phi dxdt=\iint_{\Omega_{T}}f_{m}\phi dxdt.

Using the fact that umu\displaystyle u_{m}\rightarrow u strongly in L1(ΩT)\displaystyle L^{1}(\Omega_{T}) and um(x,0)=um0(x)u0(x)=u(x,0)\displaystyle u_{m}(x,0)=u_{m0}(x)\rightarrow u_{0}(x)=u(x,0) in L1(Ω)\displaystyle L^{1}(\Omega), we have

ΩT((um)tΔum+(Δ)sum)ϕ𝑑x𝑑t=ΩTum((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑tΩum(x,0)ϕ(x,0)𝑑x=ΩTumφ𝑑x𝑑tΩum(x,0)ϕ(x,0)𝑑xΩTuφ𝑑x𝑑tΩu(x,0)ϕ(x,0)𝑑x=ΩTu((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑tΩu(x,0)ϕ(x,0)𝑑x.\begin{array}[]{rcl}\iint_{\Omega_{T}}\left((u_{m})_{t}-\Delta u_{m}+(-\Delta)^{s}u_{m}\right)\phi dxdt&=&\iint_{\Omega_{T}}u_{m}\left(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi\right)dxdt-\int_{\Omega}u_{m}(x,0)\phi(x,0)dx\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &=&\iint_{\Omega_{T}}u_{m}\varphi dxdt-\int_{\Omega}u_{m}(x,0)\phi(x,0)dx\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\rightarrow&\iint_{\Omega_{T}}u\varphi dxdt-\int_{\Omega}u(x,0)\phi(x,0)dx\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &=&\iint_{\Omega_{T}}u\left(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi\right)dxdt-\int_{\Omega}u(x,0)\phi(x,0)dx.\end{array}

Notice that in the second last line, in order to pass the limit, we have used the facts that ϕ(x,0)L(Ω)\displaystyle\phi(x,0)\in L^{\infty}(\Omega) and φ𝒞0(ΩT)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}). Also, since ϕL(Ω×(0,T))\displaystyle\phi\in L^{\infty}(\Omega\times(0,T)), and fmf\displaystyle f_{m}\rightarrow f in L1(ΩT)\displaystyle L^{1}(\Omega_{T}), we get

ΩTfmϕ𝑑x𝑑tΩTfϕ𝑑x𝑑t as m.\iint_{\Omega_{T}}f_{m}\phi dxdt\rightarrow\iint_{\Omega_{T}}f\phi dxdt\text{ as }m\rightarrow\infty.

Thus

ΩTu((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑t=ΩTfϕ𝑑x𝑑t+Ωu0(x)ϕ(x,0)𝑑x,\iint_{\Omega_{T}}u\left(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi\right)dxdt=\iint_{\Omega_{T}}f\phi dxdt+\int_{\Omega}u_{0}(x)\phi(x,0)dx,

and u\displaystyle u is a weak solution to Eq. 2.2. For the uniqueness let w\displaystyle w be a weak solution of Eq. 2.2 with (f,u0)=(0,0)\displaystyle(f,u_{0})=(0,0), i.e.

{wtΔw+(Δ)sw=0 in ΩT,w=0 in (n\Ω)×(0,T),w(x,0)=0 in Ω;\displaystyle\displaystyle\begin{cases}w_{t}-\Delta w+(-\Delta)^{s}w=0&\text{ in }\Omega_{T},\\ w=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ w(x,0)=0&\text{ in }\Omega;\end{cases}

we want to prove that w0\displaystyle w\equiv 0. For that we take F𝒞0(ΩT)\displaystyle F\in\mathcal{C}_{0}^{\infty}(\Omega_{T}), and let ϕF\displaystyle\phi_{F} be the solution of the backward problem

{(ϕF)tΔϕF+(Δ)sϕF=F in ΩT,ϕF(x,t)=0 in (n\Ω)×(0,T],ϕF(x,T)=0 in Ω.\displaystyle\displaystyle\begin{cases}-(\phi_{F})_{t}-\Delta\phi_{F}+(-\Delta)^{s}\phi_{F}=F&\text{ in }\Omega_{T},\\ \phi_{F}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T],\\ \phi_{F}(x,T)=0&\text{ in }\Omega.\end{cases}

Taking ϕF\displaystyle\phi_{F} as a test function we deduce that for any F𝒞0(ΩT)\displaystyle F\in\mathcal{C}_{0}^{\infty}(\Omega_{T}),

0TΩwF𝑑x𝑑t=0,\int_{0}^{T}\int_{\Omega}wFdxdt=0,

that means, w=0\displaystyle w=0 in 𝒟(ΩT)\displaystyle\mathcal{D}^{\prime}(\Omega_{T}). ∎

Next, we state some important facts regarding the solution of Eq. 2.2.

Proposition 2.17.

Let (f,w0)\displaystyle\left(f,w_{0}\right) are non-negative functions such that (f,w0)L(ΩT)×L(Ω)\displaystyle\left(f,w_{0}\right)\in L^{\infty}(\Omega_{T})\times L^{\infty}(\Omega). Assume that w\displaystyle w be the weak solution to the problem

{wtΔw+(Δ)sw=f in ΩT,w=0 in (n\Ω)×(0,T),w(x,0)=w0(x) in Ω;\displaystyle\displaystyle\begin{cases}w_{t}-\Delta w+(-\Delta)^{s}w=f&\text{ in }\Omega_{T},\\ w=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ w(x,0)=w_{0}(x)&\text{ in }\Omega;\end{cases}

then for all 0<t1<T\displaystyle 0<t_{1}<T and for each ωΩ\displaystyle\omega\subset\subset\Omega fixed, there exists C:=C(t1,n,s,ω)>0\displaystyle C:=C\left(t_{1},n,s,\omega\right)>0 such that

w(x,t)C(t1,n,s,ω) in ω×[t1,T).w(x,t)\geq C\left(t_{1},n,s,\omega\right)\text{ in }\omega\times[t_{1},T).
Proof.

We will use the weak Harnack inequality as in [23]. First, we show the result for any arbitrary t2<T\displaystyle t_{2}<T. As ω¯×[t1,t2]ΩT\displaystyle\bar{\omega}\times[t_{1},t_{2}]\subset\Omega_{T} contains ω×[t1,t2]\displaystyle\omega\times[t_{1},t_{2}], so it is enough to show the result for ω¯×[t1,t2]\displaystyle\bar{\omega}\times[t_{1},t_{2}]. Now since ω¯×[t1,t2]\displaystyle\bar{\omega}\times[t_{1},t_{2}] is compact and hence every open cover admits a finite subcover, it suffices to show the positivity of w\displaystyle w in a uniform neighbourhood of any arbitrary point (x~,t~)ω¯×[t1,t2]\displaystyle(\tilde{x},\tilde{t})\in\bar{\omega}\times[t_{1},t_{2}].
As ω¯×[t1,t2]\displaystyle\bar{\omega}\times[t_{1},t_{2}] is compact, it has a finite and positive distance from the boundary of ΩT\displaystyle\Omega_{T}; let us denote this by D\displaystyle D. We choose 0<r<min{D2,1}\displaystyle 0<r<\operatorname{min}\{\frac{D}{2},1\}, and (x0,t0)ΩT\displaystyle(x_{0},t_{0})\in\Omega_{T} such that t0=t~1316r2 and x0=x~.\displaystyle t_{0}=\tilde{t}-\frac{13}{16}r^{2}\text{ and }x_{0}=\tilde{x}. Now for this (x0,t0)\displaystyle(x_{0},t_{0}), we can choose 0<R=D\displaystyle 0<R=D such that BR(x0)×(t0r2,t0+r2)ΩT\displaystyle B_{R}(x_{0})\times(t_{0}-r^{2},t_{0}+r^{2})\subset\Omega_{T}, and then r\displaystyle r satisfies r<R2\displaystyle r<\frac{R}{2}. Clearly, as f0\displaystyle f\geq 0, w\displaystyle w is a supersolution to the homogeneous problem. Now by using the facts that w0\displaystyle w\geq 0 in Ω×(0,T)\displaystyle\Omega\times(0,T) and w=0\displaystyle w=0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T), we get w0\displaystyle w\geq 0 in n×(0,T)\displaystyle\mathbb{R}^{n}\times(0,T). Then using [23, Theorem 2.8 and Corollary 2.9], and f>0\displaystyle f>0 a.e. in ΩT\displaystyle\Omega_{T}, we get the existence of a positive constant C\displaystyle C depending only on ω,[t1,t2],n\displaystyle\omega,[t_{1},t_{2}],n and s\displaystyle s such that

0<C=V(r2)w(x,t)𝑑x𝑑tessinfV+(r2)w,0<C=\mathchoice{{\vbox{\hbox{$\displaystyle\textstyle\rotatebox[origin={c}]{18.0}{$\displaystyle-\mkern-9.5mu-$}$}}\kern-7.18916pt}}{{\vbox{\hbox{$\displaystyle\scriptstyle\rotatebox[origin={c}]{18.0}{$\displaystyle-\mkern-9.5mu-$}$}}\kern-6.15741pt}}{{\vbox{\hbox{$\displaystyle\scriptscriptstyle\rotatebox[origin={c}]{18.0}{$\displaystyle-\mkern-9.5mu-$}$}}\kern-4.34457pt}}{{\vbox{\hbox{$\displaystyle\scriptscriptstyle\rotatebox[origin={c}]{18.0}{$\displaystyle-\mkern-9.5mu-$}$}}\kern-3.59457pt}}\!\iint_{V^{-}\left(\frac{r}{2}\right)}w(x,t)dxdt\leq\underset{V^{+}\left(\frac{r}{2}\right)}{\operatorname{{ess}\operatorname{inf}}}w,

where V(r2)=Br2(x0)×(t0r2,t034r2)\displaystyle V^{-}\left(\frac{r}{2}\right)=B_{\frac{r}{2}(x_{0})}\times(t_{0}-r^{2},t_{0}-\frac{3}{4}r^{2}) and V+(r2)=Br2(x0)×(t0+34r2,t0+r2)\displaystyle V^{+}\left(\frac{r}{2}\right)=B_{\frac{r}{2}(x_{0})}\times(t_{0}+\frac{3}{4}r^{2},t_{0}+r^{2}). As (t~116r2,t~+116r2)(t0+34r2,t0+r2)\displaystyle(\tilde{t}-\frac{1}{16}r^{2},\tilde{t}+\frac{1}{16}r^{2})\subset(t_{0}+\frac{3}{4}r^{2},t_{0}+r^{2}), so the result follows.
Now, we will extend this up to T\displaystyle T. For this we denote D0=dist{ω,Ω}\displaystyle D_{0}=\operatorname{dist}\{\omega,\partial\Omega\}. As T>0\displaystyle T>0, so r~(0,1]\displaystyle\exists\tilde{r}\in(0,1] and r~<D02\displaystyle\tilde{r}<\frac{D_{0}}{2} such that ω×(T2r~2,T)ΩT\displaystyle\omega\times(T-2\tilde{r}^{2},T)\subset\Omega_{T} with T14r~2>t1\displaystyle T-\frac{1}{4}\tilde{r}^{2}>t_{1}. By the above argument for compact time intervals, we get

w(x,t)C(t1,n,s,ω)>0 in ω×[t1,T14r~2].w(x,t)\geq C\left(t_{1},n,s,\omega\right)>0\text{ in }\omega\times[t_{1},T-\frac{1}{4}\tilde{r}^{2}].

Therefore it just remains to show the positivity of w\displaystyle w in ω×(T14r~2,T)\displaystyle\omega\times(T-\frac{1}{4}\tilde{r}^{2},T). For x0ω¯\displaystyle x_{0}\in\bar{\omega}, Br~2(x0)\displaystyle B_{\frac{\tilde{r}}{2}}(x_{0}) will form an open cover of ω¯\displaystyle\bar{\omega}, and by compactness, it will have a finite subcover. So it is enough to show the positivity of w\displaystyle w in Br~2(x0)×(T14r~2,T)\displaystyle B_{\frac{\tilde{r}}{2}}(x_{0})\times(T-\frac{1}{4}\tilde{r}^{2},T). Now as w\displaystyle w is nonnegative in n×(0,T)\displaystyle\mathbb{R}^{n}\times(0,T) and BD0(x0)×(T2r~2,T)ΩT\displaystyle B_{D_{0}}(x_{0})\times(T-2\tilde{r}^{2},T)\subset\Omega_{T}, so by [23, Theorem 2.8 and Corollary 2.9], we have

0<C=T2r~2T78r~2Br~2(x0)w(x,t)𝑑x𝑑tessinfBr~2(x0)×(T14r~2,T)w,0<C=\fint_{T-2\tilde{r}^{2}}^{T-\frac{7}{8}\tilde{r}^{2}}\fint_{B_{\frac{\tilde{r}}{2}}(x_{0})}w(x,t)dxdt\leq\underset{B_{\frac{\tilde{r}}{2}}(x_{0})\times(T-\frac{1}{4}\tilde{r}^{2},T)}{\operatorname{{ess}\operatorname{inf}}}w,

and hence we conclude.∎

We will use the next parabolic Kato-type inequality to prove comparison results or a priori estimates.

Proposition 2.18.

Let uL2(0,T;H1(Ω))\displaystyle u\in L^{2}(0,T;H^{1}(\Omega)) be a weak solution of

utΔu+(Δ)su=f in ΩT,{}u_{t}-\Delta u+(-\Delta)^{s}u=f\text{ in }\Omega_{T}, (2.6)

with fL1(ΩT)\displaystyle f\in L^{1}(\Omega_{T}) and let Φ𝒞2()\displaystyle\Phi\in\mathcal{C}^{2}(\mathbb{R}) be a convex function such that Φ\displaystyle\Phi^{\prime} is bounded. Then

(Φ(u))tΔΦ(u)+(Δ)sΦ(u)Φ(u)(utΔu+(Δ)su)(\Phi(u))_{t}-\Delta\Phi(u)+(-\Delta)^{s}\Phi(u)\leq\Phi^{\prime}(u)\left(u_{t}-\Delta u+(-\Delta)^{s}u\right)

in the sense that for all ψ𝒞2(ΩT)𝒞([0,T],L2(Ω))\displaystyle\psi\in\mathcal{C}^{2}(\Omega_{T})\cap\mathcal{C}([0,T],L^{2}(\Omega)), with the property that ψ\displaystyle\psi has spatial support compactly contained in Ω\displaystyle\Omega, and ψ0\displaystyle\psi\geq 0 in ΩT\displaystyle\Omega_{T} and ψ(,T)=0\displaystyle\psi(\cdot,T)=0 and ψ(,0)=0\displaystyle\psi(\cdot,0)=0, we have

ΩTΦ(u)(ψtΔψ+(Δ)sψ)𝑑x𝑑tΩTΦ(u)fψ𝑑x𝑑t.\iint_{\Omega_{T}}\Phi(u)\left(-\psi_{t}-\Delta\psi+(-\Delta)^{s}\psi\right)dxdt\leq\iint_{\Omega_{T}}\Phi^{\prime}(u)f\psi dxdt.
Proof.

Assume that u\displaystyle u is smooth enough, otherwise one can use approximation argument. Since Φ\displaystyle\Phi is convex so

Φ(u(x,t))Φ(u(y,t))Φ(u(x,t))(u(x,t)u(y,t)),\Phi(u(x,t))-\Phi(u(y,t))\leq\Phi^{\prime}(u(x,t))(u(x,t)-u(y,t)),

and we get

(Δ)s(Φ(u(x,t)))=n(Φ(u(x,t))Φ(u(y,t)))|xy|n+2s𝑑ynΦ(u(x,t))(u(x,t)u(y,t))|xy|n+2s𝑑y=Φ(u(x,t))(Δ)su(x,t).\begin{array}[]{rcl}(-\Delta)^{s}(\Phi(u(x,t)))&=&\int_{\mathbb{R}^{n}}\frac{\left(\Phi(u(x,t))-\Phi(u(y,t))\right)}{|x-y|^{n+2s}}dy\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\int_{\mathbb{R}^{n}}\Phi^{\prime}(u(x,t))\frac{(u(x,t)-u(y,t))}{|x-y|^{n+2s}}dy\vskip 3.0pt plus 1.0pt minus 1.0pt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &=&\Phi^{\prime}(u(x,t))(-\Delta)^{s}u(x,t).\end{array}

Again, using the fact that the double derivative of a convex function is nonnegative, we get

Δ(Φ(u))=Φ(u)ΔuΦ′′(u)i=1n(uxi)2Φ(u)Δu.-\Delta(\Phi(u))=-\Phi^{\prime}(u)\Delta u-\Phi^{\prime\prime}(u)\sum_{i=1}^{n}\left(\frac{\partial u}{\partial x_{i}}\right)^{2}\leq-\Phi^{\prime}(u)\Delta u.

Now if ψ𝒞0(ΩT)\displaystyle\psi\in\mathcal{C}_{0}^{\infty}\left(\Omega_{T}\right), with ψ0\displaystyle\psi\geq 0 in ΩT\displaystyle\Omega_{T}, we have

ΩTΦ(u)(ψtΔψ+(Δ)sψ)𝑑x𝑑t=ΩT(Φ(u)utΔ(Φ(u))+(Δ)s(Φ(u)))ψ(x,t)𝑑x𝑑tΩTΦ(u)(utΔu+(Δ)su)ψ𝑑x𝑑t=ΩTΦ(u)fψ𝑑x𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}\Phi(u)\left(-\psi_{t}-\Delta\psi+(-\Delta)^{s}\psi\right)dxdt&=&\iint_{\Omega_{T}}\left(\Phi^{\prime}(u)u_{t}-\Delta(\Phi(u))+(-\Delta)^{s}(\Phi(u))\right)\psi(x,t)dxdt\\ &\leq&\iint_{\Omega_{T}}\Phi^{\prime}(u)(u_{t}-\Delta u+(-\Delta)^{s}u)\psi dxdt=\iint_{\Omega_{T}}\Phi^{\prime}(u)f\psi dxdt.\end{array}

Since ψ𝒞0(ΩT)\displaystyle\psi\in\mathcal{C}_{0}^{\infty}\left(\Omega_{T}\right), with ψ0\displaystyle\psi\geq 0 in ΩT\displaystyle\Omega_{T}, can approximate the choice of test functions of our hypothesis, we get the desired result. ∎

Remark 2.19.

We choose the regularization of |u|\displaystyle|u| by

Φϵ(u)=(|u|2+ϵ2)12ϵ.\Phi_{\epsilon}(u)=\left(|u|^{2}+\epsilon^{2}\right)^{\frac{1}{2}}-\epsilon.

for ϵ(0,1)\displaystyle\epsilon\in(0,1). It is then easy to verify that,

  1. 1.

    Φϵ(u)|u|\displaystyle\Phi_{\epsilon}(u)\rightarrow|u| uniformly as ϵ0\displaystyle\epsilon\rightarrow 0,

  2. 2.

    Φϵ(u)\displaystyle\Phi_{\epsilon}^{\prime}(u) is bounded uniformly with respect to ϵ\displaystyle\epsilon.

  3. 3.

    Φϵ(u)\displaystyle\Phi_{\epsilon}(u) is convex.

So, using the above theorem we get

ΩTΦϵ(u)(ψtΔψ+(Δ)sψ)𝑑x𝑑tΩTΦϵ(u)(utΔu+(Δ)su)ψ𝑑x𝑑t=ΩTΦϵ(u)fψ𝑑x𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}\Phi_{\epsilon}(u)\left(-\psi_{t}-\Delta\psi+(-\Delta)^{s}\psi\right)dxdt&\leq&\iint_{\Omega_{T}}\Phi^{\prime}_{\epsilon}(u)(u_{t}-\Delta u+(-\Delta)^{s}u)\psi dxdt=\iint_{\Omega_{T}}\Phi^{\prime}_{\epsilon}(u)f\psi dxdt.\end{array}

Now letting ϵ0\displaystyle\epsilon\rightarrow 0, we get using Eq. 2.6

ΩT|u|(ψtΔψ+(Δ)sψ)𝑑x𝑑tΩTsign(u)fψ𝑑x𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}|u|\left(-\psi_{t}-\Delta\psi+(-\Delta)^{s}\psi\right)dxdt&\leq&\iint_{\Omega_{T}}sign(u)f\psi dxdt.\end{array}

Adding ΩTu(ψtΔψ+(Δ)sψ)𝑑x𝑑t\displaystyle\iint_{\Omega_{T}}u(-\psi_{t}-\Delta\psi+(-\Delta)^{s}\psi)dxdt to both side, we get

ΩTu+(ψtΔψ+(Δ)sψ)𝑑x𝑑tΩTχ{u0}fψ𝑑x𝑑t.\iint_{\Omega_{T}}u_{+}\left(-\psi_{t}-\Delta\psi+(-\Delta)^{s}\psi\right)dxdt\leq\iint_{\Omega_{T}}\chi_{\{u\geq 0\}}f\psi dxdt.

Therefore

u+tΔu++(Δ)su+(utΔu+(Δ)su)sign+u\frac{\partial u_{+}}{\partial t}-\Delta u_{+}+(-\Delta)^{s}u_{+}\leq\left(\frac{\partial u}{\partial t}-\Delta u+(-\Delta)^{s}u\right)\operatorname{sign}^{+}u

in the weak sense.

We now take γ\displaystyle\gamma to be a positive continuous function over Ω¯T\displaystyle\overline{\Omega}_{T} and proceed with the following comparison principle.

Lemma 2.20.

Let s(0,1)\displaystyle s\in(0,1) and a0\displaystyle a\geq 0. Consider (f,u0)L(ΩT)×L(Ω)\displaystyle\left(f,u_{0}\right)\in L^{\infty}(\Omega_{T})\times L^{\infty}(\Omega) to be non-negative bounded functions such that (f,u0)(0,0)\displaystyle\left(f,u_{0}\right)\neq(0,0). Assume that v1,v2\displaystyle v_{1},v_{2} are two non-negative functions with finite energy such that v1,v2L2((0,T);H01(Ω))L(ΩT)\displaystyle v_{1},v_{2}\in L^{2}((0,T);H_{0}^{1}(\Omega))\cap L^{\infty}(\Omega_{T}) with

{(v1)tΔv1+(Δ)sv1f(v1+a)γ(x,t) in ΩT,v1(x,t)=0 in (n\Ω)×(0,T),v1(x,0)u0(x) in Ω;\displaystyle\displaystyle\begin{cases}(v_{1})_{t}-\Delta v_{1}+(-\Delta)^{s}v_{1}\leq\frac{f}{(v_{1}+a)^{\gamma(x,t)}}&\text{ in }\Omega_{T},\\ v_{1}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ v_{1}(x,0)\leq u_{0}(x)&\text{ in }\Omega;\end{cases}

and

{(v2)tΔv2+(Δ)sv2f(v2+a)γ(x,t) in ΩT,v2(x,t)=0 in (n\Ω)×(0,T),v2(x,0)u0(x) in Ω;\displaystyle\displaystyle\begin{cases}(v_{2})_{t}-\Delta v_{2}+(-\Delta)^{s}v_{2}\geq\frac{f}{(v_{2}+a)^{\gamma(x,t)}}&\text{ in }\Omega_{T},\\ v_{2}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ v_{2}(x,0)\geq u_{0}(x)&\text{ in }\Omega;\end{cases}

where γ𝒞(Ω¯T)\displaystyle\gamma\in\mathcal{C}(\overline{\Omega}_{T}) and is positive. Then, v2v1\displaystyle v_{2}\geq v_{1} in ΩT\displaystyle\Omega_{T}.

Proof.

Define u=v1v2\displaystyle u=v_{1}-v_{2}, then uL2(0,T;H01(Ω))L(ΩT)\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(\Omega_{T}). We show that u+=0\displaystyle u^{+}=0. By hypotheses, we get

utΔu+(Δ)suf(1(v1+a)γ(x,t)1(v2+a)γ(x,t)).u_{t}-\Delta u+(-\Delta)^{s}u\leq f\left(\frac{1}{\left(v_{1}+a\right)^{\gamma(x,t)}}-\frac{1}{\left(v_{2}+a\right)^{\gamma(x,t)}}\right).

Now since (1(v1+a)γ(x,t)1(v2+a)γ(x,t))0\displaystyle\left(\frac{1}{\left(v_{1}+a\right)^{\gamma(x,t)}}-\frac{1}{\left(v_{2}+a\right)^{\gamma(x,t)}}\right)\leq 0 in the set {u0}\displaystyle\{u\geq 0\}, then using Propositions 2.18 and 2.19, it holds that

(u+)tΔ(u+)+(Δ)s(u+)0.(u_{+})_{t}-\Delta(u_{+})+(-\Delta)^{s}(u_{+})\leq 0.

Again we notice that u+(x,0)0\displaystyle u_{+}(x,0)\leq 0; therefore by comparison principle as of Theorem 2.13, we have u+0\displaystyle u_{+}\equiv 0, and then we conclude. ∎

Corollary 2.21.

As a consequence of the previous comparison principle we get that for a>0\displaystyle a>0 and γ\displaystyle\gamma fixed, if (f,u0)\displaystyle\left(f,u_{0}\right) are non-negative functions with (f,u0)Lσ(ΩT)×Lσ(Ω)\displaystyle\left(f,u_{0}\right)\in L^{\sigma}(\Omega_{T})\times L^{\sigma}(\Omega), σ2\displaystyle\sigma\geq 2, then the problem

{vtΔv+(Δ)sv=f(v+a)γ(x,t) in ΩT,v(x,t)=0 in (n\Ω)×(0,T),v(x,0)=u0(x) in Ω;{}\begin{array}[]{lcr}\begin{cases}v_{t}-\Delta v+(-\Delta)^{s}v=\frac{f}{(v+a)^{\gamma(x,t)}}&\text{ in }\Omega_{T},\\ v(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ v(x,0)=u_{0}(x)&\text{ in }\Omega;\end{cases}\end{array} (2.7)

has a unique energy solution vaL2(0,T;H01(Ω))\displaystyle v_{a}\in L^{2}(0,T;H_{0}^{1}(\Omega)).
We note that for γ\displaystyle\gamma being a constant, the existence of an energy solution can be shown using a monotonicity argument by observing that v¯=0\displaystyle\underline{v}=0 is a subsolution and v¯\displaystyle\bar{v}, the unique solution to the problem

{v¯tΔv¯+(Δ)sv¯=faγ in ΩT,v¯(x,t)=0 in (n\Ω)×(0,T),v¯(x,0)=u0(x) in Ω;\displaystyle\displaystyle\begin{cases}\bar{v}_{t}-\Delta\bar{v}+(-\Delta)^{s}\bar{v}=\frac{f}{a^{\gamma}}&\text{ in }\Omega_{T},\\ \bar{v}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ \bar{v}(x,0)=u_{0}(x)&\text{ in }\Omega;\end{cases}

is a supersolution. Then Lemma 2.20 allows us to get the existence of a unique solution v\displaystyle v to Eq. 2.7 such that vL2(0,T;H01(Ω))\displaystyle v\in L^{2}(0,T;H_{0}^{1}(\Omega)) and 0vv¯\displaystyle 0\leq v\leq\bar{v} in ΩT\displaystyle\Omega_{T}.
Now if γ\displaystyle\gamma is not constant and belongs to 𝒞(Ω¯T)\displaystyle\mathcal{C}(\overline{\Omega}_{T}), we denote

γ=sup(x,t)ΩTγ(x,t) and γ=inf(x,t)ΩTγ(x,t),\gamma^{*}=\underset{(x,t)\in\Omega_{T}}{\operatorname{sup}}\gamma(x,t)\text{ and }\gamma_{*}=\underset{(x,t)\in\Omega_{T}}{\operatorname{inf}}\gamma(x,t),

and observe that the unique nonnegative solutions to the problems

{v¯tΔv¯+(Δ)sv¯=faγ in ΩT,v¯(x,t)=0 in (n\Ω)×(0,T),v¯(x,0)=u0(x) in Ω;\displaystyle\displaystyle\begin{cases}\bar{v}_{t}-\Delta\bar{v}+(-\Delta)^{s}\bar{v}=\frac{f}{a^{\gamma^{*}}}&\text{ in }\Omega_{T},\\ \bar{v}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ \bar{v}(x,0)=u_{0}(x)&\text{ in }\Omega;\end{cases}

and

{v¯tΔv¯+(Δ)sv¯=faγ in ΩT,v¯(x,t)=0 in (n\Ω)×(0,T),v¯(x,0)=u0(x) in Ω;\displaystyle\displaystyle\begin{cases}\bar{v}_{t}-\Delta\bar{v}+(-\Delta)^{s}\bar{v}=\frac{f}{a^{\gamma_{*}}}&\text{ in }\Omega_{T},\\ \bar{v}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ \bar{v}(x,0)=u_{0}(x)&\text{ in }\Omega;\end{cases}

are sub and supersolution (respectively super and subsolution) of

{v¯tΔv¯+(Δ)sv¯=faγ(x,t) in ΩT,v¯(x,t)=0 in (n\Ω)×(0,T),v¯(x,0)=u0(x) in Ω;{}\begin{array}[]{rcl}\begin{cases}\bar{v}_{t}-\Delta\bar{v}+(-\Delta)^{s}\bar{v}=\frac{f}{a^{\gamma(x,t)}}&\text{ in }\Omega_{T},\\ \bar{v}(x,t)=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ \bar{v}(x,0)=u_{0}(x)&\text{ in }\Omega;\end{cases}\end{array} (2.8)

for a1\displaystyle a\geq 1 (respectively for a<1\displaystyle a<1). Then, by monotonicity argument and comparison principle similar to Lemma 2.20, we get the existence of a unique nonnegative solution to Eq. 2.8, which will be a supersolution to Eq. 2.7 with v¯=0\displaystyle\underline{v}=0 being its subsolution. So Eq. 2.7 has a unique nonnegative solution in this case too. We note that these solutions also satisfy the comparison principle, in the sense that if a1<a2\displaystyle a_{1}<a_{2}, then we have va2va1\displaystyle v_{a_{2}}\leq v_{a_{1}}.

We now list approximated problems for γ\displaystyle\gamma being a constant. Firstly when (f,u0)L(ΩT)×L(Ω)\displaystyle(f,u_{0})\in L^{\infty}(\Omega_{T})\times L^{\infty}(\Omega), we consider the following problem for each k\displaystyle k\in\mathbb{N},

{(uk)tΔuk+(Δ)suk=f(x,t)(uk+1k)γ in ΩT,uk=0 in (n\Ω)×(0,T),uk(x,0)=u0(x) in Ω.{}\begin{array}[]{lcr}\begin{cases}(u_{k})_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k}=\frac{f(x,t)}{(u_{k}+\frac{1}{k})^{\gamma}}&\mbox{ in }\Omega_{T},\\ u_{k}=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u_{k}(x,0)=u_{0}(x)&\mbox{ in }\Omega.\end{cases}\end{array} (2.9)

We now take f\displaystyle f and u0\displaystyle u_{0} may not be bounded and consider

{(uk)tΔuk+(Δ)suk=fk(x,t)(uk+1k)γ in ΩT,uk=0 in (n\Ω)×(0,T),uk(x,0)=u0k(x) in Ω;{}\begin{array}[]{lcr}\begin{cases}(u_{k})_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k}=\frac{f_{k}(x,t)}{(u_{k}+\frac{1}{k})^{\gamma}}&\mbox{ in }\Omega_{T},\\ u_{k}=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u_{k}(x,0)=u_{0k}(x)&\mbox{ in }\Omega;\end{cases}\end{array} (2.10)

where fk(x,t)=Tk(f(x,t))\displaystyle f_{k}(x,t)=T_{k}(f(x,t)) and u0k(x)=Tk(u0(x))\displaystyle u_{0k}(x)=T_{k}(u_{0}(x)).
To treat variable exponent, we take (f,u0)L1(ΩT)×L1(Ω)\displaystyle(f,u_{0})\in L^{1}(\Omega_{T})\times L^{1}(\Omega) and consider the following approximating problem.

{(uk)tΔuk+(Δ)suk=fk(x,t)(uk+1k)γ(x,t) in ΩT,uk=0 in (n\Ω)×(0,T),uk(x,0)=u0k(x) in Ω.{}\begin{array}[]{lcr}\begin{cases}(u_{k})_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k}=\frac{f_{k}(x,t)}{(u_{k}+\frac{1}{k})^{\gamma(x,t)}}&\mbox{ in }\Omega_{T},\\ u_{k}=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u_{k}(x,0)=u_{0k}(x)&\mbox{ in }\Omega.\end{cases}\end{array} (2.11)

We will prescribe conditions on f\displaystyle f and u0\displaystyle u_{0} later on. With these settings, we have the following existence result.

Theorem 2.22.

For each of the above problems Eqs. 2.9, 2.10 and 2.11, there exists a unique nonnegative energy solution uk\displaystyle u_{k} for each k\displaystyle k\in\mathbb{N} satisfying the followings:
(a) each uk\displaystyle u_{k} is bounded,
(b) the sequence {uk}k\displaystyle\{u_{k}\}_{k} is increasing in k\displaystyle k,
(c) for each t0>0\displaystyle t_{0}>0 and ωΩT\displaystyle\omega\subset\subset\Omega_{T}, C(ω,t0,n,s)\displaystyle\exists C(\omega,t_{0},n,s) such that ukC(ω,t0,n,s)\displaystyle u_{k}\geq C(\omega,t_{0},n,s).

Proof.

The existence, uniqueness and nonnegativity of uk\displaystyle u_{k}’s follow from the comparison principle in Lemma 2.20 and Corollary 2.21. Since f\displaystyle f and u0\displaystyle u_{0} are nonnegative, therefore 0fkfk+1\displaystyle 0\leq f_{k}\leq f_{k+1} and 0u0ku0(k+1)\displaystyle 0\leq u_{0k}\leq u_{0(k+1)} for each k\displaystyle k and then using the similar technique as that of Lemma 2.20, we obtain that the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is increasing in k\displaystyle k. Therefore uku1\displaystyle u_{k}\geq u_{1} for all k\displaystyle k. Now for fL(ΩT)\displaystyle f\in L^{\infty}(\Omega_{T}) we have f(u1+1)γf\displaystyle\frac{f}{(u_{1}+1)^{\gamma}}\leq f and for fL(ΩT)\displaystyle f\notin L^{\infty}(\Omega_{T}) we have f1(u1+1)γ(x,t)f1\displaystyle\frac{f_{1}}{(u_{1}+1)^{\gamma(x,t)}}\leq f_{1} and hence f1(u1+1)γ(x,t)L(ΩT)\displaystyle\frac{f_{1}}{(u_{1}+1)^{\gamma(x,t)}}\in L^{\infty}(\Omega_{T}). Therefore u1L(ΩT)\displaystyle u_{1}\in L^{\infty}(\Omega_{T}) and by Proposition 2.17, we obtain that for all ωΩ\displaystyle\omega\subset\subset\Omega and for all t0>0\displaystyle t_{0}>0, u1(x,t)\displaystyle u_{1}(x,t)\geq C(ω,t0,n,s)\displaystyle C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right). Thus for all k1\displaystyle k\geq 1, we have uk(x,t)C(ω,t0,n,s)\displaystyle u_{k}(x,t)\geq C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right). Also, we observe that for each k\displaystyle k, fk(uk+1k)γ(x,t)max{1,kγ}fk,\displaystyle\frac{f_{k}}{(u_{k}+\frac{1}{k})^{\gamma(x,t)}}\leq\operatorname{max}\{1,k^{\gamma^{*}}\}f_{k}, where γ=sup(x,t)ΩTγ(x,t)\displaystyle\gamma^{*}=\underset{(x,t)\in\Omega_{T}}{\operatorname{sup}}\gamma(x,t). Therefore each uk\displaystyle u_{k} is bounded. ∎

Remark 2.23.

As each uk𝒞([0,T],L2(Ω))𝒞([0,T],L1(Ω))\displaystyle u_{k}\in\mathcal{C}([0,T],L^{2}(\Omega))\subset\mathcal{C}([0,T],L^{1}(\Omega)), therefore at each time level t[0,T]\displaystyle t\in[0,T], we have uk(,t)L1(Ω)\displaystyle u_{k}(\cdot,t)\in L^{1}(\Omega), this along with the monotonicity of {uk}k\displaystyle\{u_{k}\}_{k} in k\displaystyle k allows us to define the pointwise limit (a.e.) u\displaystyle u of {uk}k\displaystyle\{u_{k}\}_{k} in Ω\displaystyle\Omega at each time t[0,T]\displaystyle t\in[0,T]. Also u\displaystyle u satisfies u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) in L1\displaystyle L^{1} sense. Again as each uk=0\displaystyle u_{k}=0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T), therefore u\displaystyle u also satisfies the same. Further we have uuk\displaystyle u\geq u_{k} for each k\displaystyle k and hence u(x,t)C(ω,t0,n,s)\displaystyle u(x,t)\geq C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right) for each ωΩ\displaystyle\omega\subset\subset\Omega and t0>0\displaystyle t_{0}>0.

To treat the singular case, we define next as:

Definition 2.24.

Let (f,u0)L1(ΩT)×L1(Ω)\displaystyle\left(f,u_{0}\right)\in L^{1}\left(\Omega_{T}\right)\times L^{1}(\Omega) be a pair of non-negative functions and γ>0\displaystyle\gamma>0 is a constant. We say that u\displaystyle u is a very weak solution to the problem

{utΔu+(Δ)su=f(x,t)uγ in ΩT,u=0 in (n\Ω)×(0,T),u(x,0)=u0(x) in Ω;{}\begin{array}[]{lcr}\begin{cases}u_{t}-\Delta u+(-\Delta)^{s}u=\frac{f(x,t)}{u^{\gamma}}&\mbox{ in }\Omega_{T},\\ u=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u(x,0)=u_{0}(x)&\mbox{ in }\Omega;\end{cases}\end{array} (2.12)

if uL1(ΩT)\displaystyle u\in L^{1}(\Omega_{T}) satisfying u0\displaystyle u\equiv 0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T), ωΩ\displaystyle\forall\omega\subset\subset\Omega and t0>0\displaystyle\forall t_{0}>0, cc(ω,t0,n,s)>0\displaystyle\exists c\equiv c(\omega,t_{0},n,s)>0 such that u(x,t)c>0\displaystyle u(x,t)\geq c>0, in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right), u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot), and for all φ𝒞0(ΩT)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}\left(\Omega_{T}\right), we have

ΩTu(φtΔφ+(Δ)sφ)𝑑x𝑑t=ΩTfφuγ𝑑x𝑑t.\iint_{\Omega_{T}}u\left(-\varphi_{t}-\Delta\varphi+(-\Delta)^{s}\varphi\right)dxdt=\iint_{\Omega_{T}}\frac{f\varphi}{u^{\gamma}}dxdt.
Definition 2.25.

Let (f,u0)\displaystyle\left(f,u_{0}\right) be a pair of non-negative functions with u0L1(Ω)\displaystyle u_{0}\in L^{1}(\Omega), fL1(ΩT)\displaystyle f\in L^{1}(\Omega_{T}) and γ𝒞(Ω¯T)\displaystyle\gamma\in\mathcal{C}(\overline{\Omega}_{T}) be a positive function. Then we say that u\displaystyle u, such that u0\displaystyle u\equiv 0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T), is a very weak solution to the problem

{utΔu+(Δ)su=f(x,t)uγ(x,t) in ΩT,u=0 in (n\Ω)×(0,T),u(x,0)=u0(x), in Ω;{}\begin{array}[]{lcr}\begin{cases}u_{t}-\Delta u+(-\Delta)^{s}u=\frac{f(x,t)}{u^{\gamma(x,t)}}&\mbox{ in }\Omega_{T},\\ u=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u(x,0)=u_{0}(x),&\mbox{ in }\Omega;\end{cases}\end{array} (2.13)

if uL1(ΩT)\displaystyle u\in L^{1}(\Omega_{T}) and ωΩ\displaystyle\forall\omega\subset\subset\Omega and t0>0\displaystyle\forall t_{0}>0, cc(ω,t0,n,s)>0\displaystyle\exists c\equiv c(\omega,t_{0},n,s)>0 such that u(x,t)c>0\displaystyle u(x,t)\geq c>0, in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right), u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot), and for all φ𝒞0(ΩT)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}\left(\Omega_{T}\right), we have

ΩTu(φtΔφ+(Δ)sφ)𝑑x𝑑t=ΩTfφuγ(x,t)𝑑x𝑑t.\iint_{\Omega_{T}}u\left(-\varphi_{t}-\Delta\varphi+(-\Delta)^{s}\varphi\right)dxdt=\iint_{\Omega_{T}}\frac{f\varphi}{u^{\gamma(x,t)}}dxdt.

2.4 Main Results

First, we consider γ\displaystyle\gamma to be a constant and then take it as a positive continuous function. In this section, we state our main results depending on γ\displaystyle\gamma and the regularity of initial conditions.

2.4.1 Existence for bounded data for γ\displaystyle\gamma being a constant

In the case of bounded data, we will have the next existence result.

Theorem 2.26.

Let (f,u0)L(ΩT)×L(Ω)\displaystyle\left(f,u_{0}\right)\in L^{\infty}\left(\Omega_{T}\right)\times L^{\infty}(\Omega) be a pair of non-negative functions and γ>0\displaystyle\gamma>0. Then Eq. 2.12 has a bounded very weak positive solution u\displaystyle u in the sense of Definition 2.24 such that uγ+12L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) and u𝒞([0,T],L2(Ω))\displaystyle u\in\mathcal{C}([0,T],L^{2}(\Omega)). Moreover if γ1\displaystyle\gamma\leq 1 or Supp(f)ΩT\displaystyle\operatorname{Supp}(f)\subset\subset\Omega_{T}, then uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega)).

2.4.2 Existence for general data for γ\displaystyle\gamma being a constant

In this subsection, we will consider general data. According to the regularity of initial conditions, we will consider two cases as u0Lγ+1(Ω)\displaystyle u_{0}\in L^{\gamma+1}(\Omega) and u0L1(Ω)\displaystyle u_{0}\in L^{1}(\Omega). Let us state the following existence results.

Theorem 2.27.

Let (f,u0)L1(ΩT)×Lγ+1(Ω)\displaystyle\left(f,u_{0}\right)\in L^{1}\left(\Omega_{T}\right)\times L^{\gamma+1}(\Omega) be a pair of non-negative functions and γ>0\displaystyle\gamma>0. Then Eq. 2.12 has a very weak positive solution u\displaystyle u in the sense of Definition 2.24 such that uγ+12L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) and uL2(0,T;Lσ(Ω))\displaystyle u\in L^{2}\left(0,T;L^{\sigma}(\Omega)\right), where σ=n(1+γ)n2\displaystyle\sigma=\frac{n(1+\gamma)}{n-2} and σ2\displaystyle\sigma\geq 2 if γ1\displaystyle\gamma\geq 1.

Theorem 2.28.

Let (f,u0)L1(ΩT)×L1(Ω)\displaystyle\left(f,u_{0}\right)\in L^{1}(\Omega_{T})\times L^{1}(\Omega) be a pair of non-negative functions and γ>0\displaystyle\gamma>0. Then Eq. 2.12 has a very weak positive solution u\displaystyle u in the sense of Definition 2.24 such that Tk(u)γ+12L2(0,T;H01(Ω))\displaystyle T_{k}(u)^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) for all k>0\displaystyle k>0.

2.4.3 Improved results for γ\displaystyle\gamma being a constant

We now break γ\displaystyle\gamma in three parts namely 0<γ<1\displaystyle 0<\gamma<1, γ=1\displaystyle\gamma=1 and γ>1\displaystyle\gamma>1. We improve our results to find solutions in better spaces. For this, we take u0Lmax(γ+1,2)(Ω).\displaystyle u_{0}\in L^{\operatorname{max}(\gamma+1,2)}(\Omega). The case γ=1\displaystyle\gamma=1 will be same as that of Theorem 2.27. For γ>1\displaystyle\gamma>1, we cannot find solutions in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)). In fact if we look for L2(0,T;H1(Ω))\displaystyle L^{2}(0,T;H^{1}(\Omega)) estimates, we will only get them in L2(t0,T;Hloc1(Ω))\displaystyle L^{2}(t_{0},T;H_{\operatorname{loc}}^{1}(\Omega)) for each t0>0\displaystyle t_{0}>0. We state our next result for γ>1\displaystyle\gamma>1 as follows:

Theorem 2.29.

Let γ>1\displaystyle\gamma>1. Assume that (f,u0)L1(ΩT)×Lγ+1(Ω)\displaystyle\left(f,u_{0}\right)\in L^{1}\left(\Omega_{T}\right)\times L^{\gamma+1}(\Omega) be a pair of non-negative functions. Then Eq. 2.12 has a very weak positive solution u\displaystyle u in the sense of Definition 2.24 such that uγ+12L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) and uL2(t0,T;Hloc1(Ω))L(0,T;L1+γ(Ω))\displaystyle u\in L^{2}(t_{0},T;H_{\operatorname{loc}}^{1}(\Omega))\cap L^{\infty}(0,T;L^{1+\gamma}(\Omega)) for each t0>0\displaystyle t_{0}>0.

Now we consider the case 0<γ<1\displaystyle 0<\gamma<1, and state our results as follows:

Theorem 2.30.

Let 0<γ<1\displaystyle 0<\gamma<1. Assume that u0L2(Ω)\displaystyle u_{0}\in L^{2}(\Omega) with u00\displaystyle u_{0}\geq 0 and f0\displaystyle f\geq 0 is such that
i)\displaystyle i) fL2γ+1(0,T;L(21γ)(Ω))\displaystyle f\in L^{\frac{2}{\gamma+1}}\left(0,T;L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)\right), or
ii)\displaystyle ii) fLm¯(ΩT)\displaystyle f\in L^{\bar{m}}\left(\Omega_{T}\right) with m¯:=2(n+2)2(n+2)n(1γ)\displaystyle\bar{m}:=\frac{2(n+2)}{2(n+2)-n(1-\gamma)}.
Then Eq. 2.12 has a very weak solution uL2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)) in the sense of Definition 2.24.

Remark 2.31.

As γ<1\displaystyle\gamma<1, so

(21γ)=2n2n(1γ)(n2)<m¯<2γ+1,\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}=\frac{2n}{2n-(1-\gamma)(n-2)}<\bar{m}<\frac{2}{\gamma+1},

and the two spaces L2γ+1(0,T;L(21γ)(Ω))\displaystyle L^{\frac{2}{\gamma+1}}\left(0,T;L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)\right) and Lm¯(ΩT)\displaystyle L^{\bar{m}}\left(\Omega_{T}\right) are not comparable. Also the case γ=1\displaystyle\gamma=1 cannot be considered here since in the proof of Theorem 2.30, we will use Hölder inequality with exponents 2γ+1\displaystyle\frac{2}{\gamma+1} and 21γ\displaystyle\frac{2}{1-\gamma}. If γ1\displaystyle\gamma\to 1, then m¯1\displaystyle\bar{m}\to 1 and 2γ+1\displaystyle\frac{2}{\gamma+1} and (21γ)\displaystyle\left(\frac{2^{*}}{1-\gamma}\right)^{\prime} both tend to 1\displaystyle 1, so that f\displaystyle f will belong to L1(ΩT)\displaystyle L^{1}\left(\Omega_{T}\right).

Now for m<m¯\displaystyle m<\bar{m}, we no longer find solutions in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)) but in a larger space depending on m\displaystyle m.

Theorem 2.32.

Let 0<γ<1\displaystyle 0<\gamma<1. Assume that 0fLm(ΩT)\displaystyle 0\leq f\in L^{m}\left(\Omega_{T}\right), with 1m<m¯\displaystyle 1\leq m<\bar{m}, and that u0L2(Ω)\displaystyle u_{0}\in L^{2}(\Omega) be nonnegative. Then the problem Eq. 2.12 admits a very weak solution uLq¯(0,T;W01,q¯(Ω))L(0,T;L1+γ(Ω))\displaystyle u\in L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega))\cap L^{\infty}(0,T;L^{1+\gamma}(\Omega)), with

q¯=m(γ+1)(n+2)n+2m(1γ).\bar{q}=\frac{m(\gamma+1)(n+2)}{n+2-m(1-\gamma)}.

Moreover uLσ(ΩT)\displaystyle u\in L^{\sigma}\left(\Omega_{T}\right), where

σ=m(γ+1)(n+2)n2(m1).\sigma=\frac{m(\gamma+1)(n+2)}{n-2(m-1)}.
Remark 2.33.

Observe that we can get rid of the fact that uLq¯(0,T;W0s1,q¯(Ω))\displaystyle u\in L^{\bar{q}}(0,T;W_{0}^{s_{1},\bar{q}}(\Omega)) with s1<s\displaystyle s_{1}<s as is done in [1, Theorem 11]. For our case, due to the presence of the leading Laplacian operator, we will get uLq¯(0,T;W0s,q¯(Ω))\displaystyle u\in L^{\bar{q}}(0,T;W_{0}^{s,\bar{q}}(\Omega)).

Remark 2.34.

Clearly q¯m(γ+1)>1\displaystyle\bar{q}\geq m(\gamma+1)>1 and σm(γ+1)>1\displaystyle\sigma\geq m(\gamma+1)>1. Also m<m¯\displaystyle m<\bar{m} is equivalent to q¯<2\displaystyle\bar{q}<2 which implies L2(0,T;H01(Ω))\displaystyle L^{2}\left(0,T;H_{0}^{1}(\Omega)\right)\subset Lq¯(0,T;W01,q¯(Ω))\displaystyle L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega)). In Theorem 2.32 the case γ=1\displaystyle\gamma=1 is not allowed since it yields q¯=2m2\displaystyle\bar{q}=2m\geq 2 which contradicts q¯<2\displaystyle\bar{q}<2.

2.4.4 Further summability for γ\displaystyle\gamma being a constant

Here we state two results regarding the optimal summability of our solutions in terms of summability of the initial data fLr(0,T;Lq(Ω))\displaystyle f\in L^{r}(0,T;L^{q}(\Omega)). For the L\displaystyle L^{\infty}-boundedness, we take u0L(Ω)\displaystyle u_{0}\in L^{\infty}(\Omega), and 1/r\displaystyle 1/r and 1/q\displaystyle 1/q are in the Aronson-Serrin domain, see [2, 29]. We now state the corresponding theorem as:

Theorem 2.35.

Assume that fLr(0,T;Lq(Ω))\displaystyle f\in L^{r}(0,T;L^{q}(\Omega)) with r,q\displaystyle r,q satisfying

1r+n2q<1,\frac{1}{r}+\frac{n}{2q}<1,

and suppose that u0L(Ω)\displaystyle u_{0}\in L^{\infty}(\Omega). Then there exists a positive constant c\displaystyle c such that the unique finite energy solution of Eq. 2.10 satisfies

uk(x,t)L(ΩT)c,\left\|u_{k}(x,t)\right\|_{L^{\infty}\left(\Omega_{T}\right)}\leq c,

and the solution u\displaystyle u obtained as the limit of uk\displaystyle u_{k} satisfies, uL(ΩT)\displaystyle u\in L^{\infty}(\Omega_{T}). Moreover for γ1\displaystyle\gamma\leq 1, uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H^{1}_{0}(\Omega)).

Outside the Aronsom-Serrin zone, the solutions are not expected to be bounded. We divide the region by the straight line 1r=nn21q2n2\displaystyle\frac{1}{r}=\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2} in two parts. Further, we take u00\displaystyle u_{0}\equiv 0. The following summability results we will get for this case.

Theorem 2.36.

Assume u0(x)0\displaystyle u_{0}(x)\equiv 0 and fLr(0,T;Lq(Ω))\displaystyle f\in L^{r}\left(0,T;L^{q}(\Omega)\right), with r>1,q>1\displaystyle r>1,q>1 satisfy

1<1r+n2q.1<\frac{1}{r}+\frac{n}{2q}.

Further, assume that for γ<1\displaystyle\gamma<1,
i)\displaystyle i) if 1r<nn21q2n2\displaystyle\frac{1}{r}<\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2}, then q>(21γ)\displaystyle q>\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}, and
ii)\displaystyle ii) if 1rnn21q2n2\displaystyle\frac{1}{r}\geq\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2}, then r>21+γ\displaystyle r>\frac{2}{1+\gamma}.
Then there exists a positive constant c\displaystyle c such that the sequence of finite energy solutions of Eq. 2.10 satisfies

ukL(0,T;L2σ(Ω))+ukL2σ(0,T;L2σ)c,\left\|u_{k}\right\|_{L^{\infty}\left(0,T;L^{2\sigma}(\Omega)\right)}+\left\|u_{k}\right\|_{L^{2\sigma}\left(0,T;L^{2^{*}\sigma}\right)}\leq c,

where

σ={q(n2)(γ+1)2(n2q) if 1r<nn21q2n2,qrn(γ+1)2(nr+2q2qr) if 1rnn21q2n2.\sigma=\left\{\begin{array}[]{lll}\frac{q(n-2)(\gamma+1)}{2(n-2q)}&\text{ if }\frac{1}{r}<\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2},\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \frac{qrn(\gamma+1)}{2(nr+2q-2qr)}&\text{ if }\frac{1}{r}\geq\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2}.\end{array}\right.

Further, the solution u\displaystyle u obtained as the limit of uk\displaystyle u_{k} satisfies, uL(0,T;L2σ(Ω))L2σ(0,T;L2σ(Ω))\displaystyle u\in L^{\infty}(0,T;L^{2\sigma}(\Omega))\cap L^{2\sigma}(0,T;L^{2^{*}\sigma}(\Omega)).

Remark 2.37.

In Theorem 2.36, we observe that the conditions needed for γ<1\displaystyle\gamma<1 are obvious, as Theorem 2.30 implies that if r=21+γ\displaystyle r=\frac{2}{1+\gamma} and q=(21γ)\displaystyle q=\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}, then by Sobolev embedding, uL2(0,T;L2(Ω))\displaystyle u\in L^{2}(0,T;L^{2^{*}}(\Omega)) which matches with Theorem 2.36 as we get σ=1\displaystyle\sigma=1 in this case.

Remark 2.38.

We note that, for γ1\displaystyle\gamma\geq 1, the singularity allows us to get better summability results as we can go up to r=1\displaystyle r=1 and q=1\displaystyle q=1, which is not allowed for the nonsingular case (see [10, 29]), whereas for γ<1\displaystyle\gamma<1, the singularity gives us better summability for the spatial exponent q\displaystyle q only.

2.4.5 Existence results for γ\displaystyle\gamma being a function

We now consider γ\displaystyle\gamma to be a positive continuous function on Ω¯T\displaystyle\overline{\Omega}_{T}. We will mainly see the behaviour of γ\displaystyle\gamma near the parabolic boundary, and accordingly, we will state two existence results. We recall the strip around the parabolic boundary given by (ΩT)δ={(x,t)ΩT:dist((x,t),ΓT)<δ}\displaystyle(\Omega_{T})_{\delta}=\{(x,t)\in\Omega_{T}:\operatorname{dist}((x,t),\Gamma_{T})<\delta\} for δ>0\displaystyle\delta>0, where ΓT=(Ω×{t=0})(Ω×(0,T))\displaystyle\Gamma_{T}=(\Omega\times\{t=0\})\cup(\partial\Omega\times(0,T)).

Theorem 2.39.

Let δ>0\displaystyle\exists\delta>0 such that γ(x,t)1\displaystyle\gamma(x,t)\leq 1 in (ΩT)δ\displaystyle(\Omega_{T})_{\delta}. Also let u0L2(Ω)\displaystyle u_{0}\in L^{2}(\Omega) with u00\displaystyle u_{0}\geq 0 and f0\displaystyle f\geq 0 satisfies
i)\displaystyle i) fL2(0,T;L(2nn+2)(Ω))\displaystyle f\in L^{2}\left(0,T;L^{\left(\frac{2n}{n+2}\right)}(\Omega)\right), or
ii)\displaystyle ii) fLr¯(ΩT)\displaystyle f\in L^{\bar{r}}\left(\Omega_{T}\right) with r¯:=2(n+2)n+4\displaystyle\bar{r}:=\frac{2(n+2)}{n+4}.
Then Eq. 2.13 has a very weak solution uL2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)) in the sense of Definition 2.25.

Remark 2.40.

Since 2nn+2<2(n+2)n+4=r¯<2\displaystyle\frac{2n}{n+2}<\frac{2(n+2)}{n+4}=\bar{r}<2, so the two spaces L2(0,T;L(2nn+2)(Ω))\displaystyle L^{2}\left(0,T;L^{\left(\frac{2n}{n+2}\right)}(\Omega)\right) and Lr¯(ΩT)\displaystyle L^{\bar{r}}\left(\Omega_{T}\right) are not comparable. Also for the constant case (Theorem 2.30) we got larger possible spaces L2γ+1(0,T;L(21γ)(Ω))\displaystyle L^{\frac{2}{\gamma+1}}\left(0,T;L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)\right) and Lm¯(ΩT)\displaystyle L^{\bar{m}}\left(\Omega_{T}\right) for the belonging of initial data f\displaystyle f, which gives us broader results, this is the cause of considering the constant and nonconstant cases separately.

Theorem 2.41.

Assume that for some γ>1\displaystyle\gamma^{*}>1 and some δ>0\displaystyle\delta>0, we have γL((ΩT)δ)<γ\displaystyle\|\gamma\|_{L^{\infty}((\Omega_{T})_{\delta})}<\gamma^{*}. Assume that u0Lγ+1(Ω)\displaystyle u_{0}\in L^{\gamma^{*}+1}(\Omega) with u00\displaystyle u_{0}\geq 0 and f0\displaystyle f\geq 0 is such that
i)\displaystyle i) fLγ+1(0,T;L(n(γ+1)n+2γ)(Ω))\displaystyle f\in L^{\gamma^{*}+1}\left(0,T;L^{\left(\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}\right)}(\Omega)\right), or
ii)\displaystyle ii) fLr~(ΩT)\displaystyle f\in L^{\tilde{r}}\left(\Omega_{T}\right) with r~:=(n+2)(γ+1)n+2(γ+1)\displaystyle\tilde{r}:=\frac{(n+2)(\gamma^{*}+1)}{n+2(\gamma^{*}+1)}.
Then Eq. 2.13 admits a very weak nonnegative solution u\displaystyle u in the sense of Definition 2.25 such that uγ+12L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma^{*}+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) and uL2(t0,T;Hloc1(Ω))L(0,T;L1+γ(Ω))\displaystyle u\in L^{2}(t_{0},T;H_{\operatorname{loc}}^{1}(\Omega))\cap L^{\infty}(0,T;L^{1+\gamma^{*}}(\Omega)) for each t0>0\displaystyle t_{0}>0.

Remark 2.42.

Since n(γ+1)n+2γ<(n+2)(γ+1)n+2(γ+1)=r~<γ+1,\displaystyle\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}<\frac{(n+2)(\gamma^{*}+1)}{n+2(\gamma^{*}+1)}=\tilde{r}<\gamma^{*}+1, so the two spaces Lγ+1(0,T;L(n(γ+1)n+2γ)(Ω))\displaystyle L^{\gamma^{*}+1}\left(0,T;L^{\left(\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}\right)}(\Omega)\right) and Lr~(ΩT)\displaystyle L^{\tilde{r}}\left(\Omega_{T}\right) are not comparable. Also, for the constant case (Theorem 2.29), we got the largest possible space L1(ΩT)\displaystyle L^{1}(\Omega_{T}) for the belonging of initial data f\displaystyle f.

3 Proof of main results

Proof of Theorem 2.26

We consider the approximated problems Eq. 2.9 in this case and show that {uk}k\displaystyle\left\{u_{k}\right\}_{k} is bounded in L(ΩT)\displaystyle L^{\infty}\left(\Omega_{T}\right). Note that the existence and other properties of {uk}k\displaystyle\{u_{k}\}_{k} follow by Theorem 2.22. We define wk=\displaystyle w_{k}= H(uk)=ukγ+1\displaystyle H\left(u_{k}\right)=u_{k}^{\gamma+1}, and note that as ukL2(0,T;H01(Ω))\displaystyle u_{k}\in L^{2}(0,T;H^{1}_{0}(\Omega)) is bounded, so ukγ+1L2(0,T;H01(Ω))\displaystyle u_{k}^{\gamma+1}\in L^{2}(0,T;H^{1}_{0}(\Omega)). Then, using Kato inequality as in Proposition 2.18, we get

(wk)tΔwk+(Δ)swkH(uk)((uk)tΔuk+(Δ)suk)(γ+1)f in weak sense.\left(w_{k}\right)_{t}-\Delta w_{k}+(-\Delta)^{s}w_{k}\leq H^{\prime}(u_{k})\left(\left(u_{k}\right)_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k}\right)\leq(\gamma+1)f\quad\text{ in weak sense}.

Now let ϑ\displaystyle\vartheta be the unique solution to the problem

{ϑtΔϑ+(Δ)sϑ=(γ+1)f in ΩT,ϑ=0 in (n\Ω)×(0,T),ϑ(x,0)=H(u0(x)) in Ω.\displaystyle\displaystyle\begin{cases}\vartheta_{t}-\Delta\vartheta+(-\Delta)^{s}\vartheta=(\gamma+1)f&\text{ in }\Omega_{T},\\ \vartheta=0&\text{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ \vartheta(x,0)=H(u_{0}(x))&\text{ in }\Omega.\end{cases}

As f\displaystyle f and u0\displaystyle u_{0} are bounded, so by Remark 2.14, ϑL(ΩT)\displaystyle\vartheta\in L^{\infty}(\Omega_{T}) and by comparison principle wkϑ\displaystyle w_{k}\leq\vartheta for all k\displaystyle k. Hence, ukγ+1ϑ\displaystyle u_{k}^{\gamma+1}\leq\vartheta and the claim follows.
Now since ukL(ΩT)L2(0,T;H01(Ω))\displaystyle u_{k}\in L^{\infty}(\Omega_{T})\cap L^{2}(0,T;H^{1}_{0}(\Omega)) and nonnegative, then for any ε>0\displaystyle\varepsilon>0 and θ>0\displaystyle\theta>0, ((uk(x,t)+ε)θεθ)L2(0,T;H01(Ω))\displaystyle(\left(u_{k}(x,t)+\varepsilon\right)^{\theta}-\varepsilon^{\theta})\in L^{2}(0,T;H^{1}_{0}(\Omega)). So choosing 0<ε<1/k\displaystyle 0<\varepsilon<1/k, for t(0,T]\displaystyle t\in(0,T], we take ((uk(x,θ)+ε)γεγ)χ(0,t)\displaystyle(\left(u_{k}(x,\theta)+\varepsilon\right)^{\gamma}-\varepsilon^{\gamma})\chi_{(0,t)} as a test function in Eq. 2.9, and it holds that

0tΩ(uk)t((uk+ε))γεγ)dxdθ+0tΩuk((uk+ε)γεγ)dxdθ+120tQ(uk(x,θ)uk(y,θ))((uk(x,θ)+ε)γ(uk(y,θ)+ε)γ)|xy|n+2s𝑑x𝑑y𝑑θΩtf(x,θ)𝑑x𝑑θ.\begin{array}[]{l}\int_{0}^{t}\int_{\Omega}\left(u_{k}\right)_{t}\left((u_{k}+\varepsilon))^{\gamma}-\varepsilon^{\gamma}\right)dxd\theta+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla\left((u_{k}+\varepsilon)^{\gamma}-\varepsilon^{\gamma}\right)dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{(u_{k}(x,\theta)-u_{k}(y,\theta))((u_{k}(x,\theta)+\varepsilon)^{\gamma}-\left(u_{k}(y,\theta)+\varepsilon\right)^{\gamma})}{|x-y|^{n+2s}}dxdyd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{t}}f(x,\theta)dxd\theta.\end{array}

Now letting ε0\displaystyle\varepsilon\to 0, by Fatou’s lemma, we get for all tT\displaystyle t\leq T

1γ+1Ωukγ+1(x,t)𝑑x+0tΩukukγdxdθ+120tQ(ukγ(x,θ)ukγ(y,θ))(uk(x,θ)uk(y,θ))|xy|n+2s𝑑x𝑑y𝑑θΩtf𝑑x𝑑θ+1γ+1Ωu0γ+1(x)𝑑xfL(ΩT)|ΩT|+1γ+1u0L(Ω)γ+1|Ω|.{}\begin{array}[]{l}\frac{1}{\gamma+1}\int_{\Omega}u_{k}^{\gamma+1}(x,t)dx+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla u_{k}^{\gamma}dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(u_{k}^{\gamma}(x,\theta)-u_{k}^{\gamma}(y,\theta)\right)\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{t}}fdxd\theta+\frac{1}{\gamma+1}\int_{\Omega}u_{0}^{\gamma+1}(x)dx\leq\|f\|_{L^{\infty}(\Omega_{T})}|\Omega_{T}|+\frac{1}{\gamma+1}\|u_{0}\|_{L^{\infty}(\Omega)}^{\gamma+1}|\Omega|.\end{array} (3.1)

Since we know ukukγ=γukγ1|uk|2,\displaystyle\nabla u_{k}\cdot\nabla u_{k}^{\gamma}=\gamma u_{k}^{\gamma-1}|\nabla u_{k}|^{2}, and by Lemma 2.11

(ukγ(x,θ)ukγ(y,θ))(uk(x,θ)uk(y,θ))C(γ)(ukγ+12(x,θ)ukγ+12(y,θ))2,\left(u_{k}^{\gamma}(x,\theta)-u_{k}^{\gamma}(y,\theta)\right)\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)\geq C(\gamma)\left(u_{k}^{\frac{\gamma+1}{2}}(x,\theta)-u_{k}^{\frac{\gamma+1}{2}}(y,\theta)\right)^{2},

we get taking supremum over t(0,T]\displaystyle t\in(0,T] in Eq. 3.1, that the sequence {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma+1}{2}}\right\}_{k} is bounded in L(0,T,L2(Ω))L2(0,T;X0s(Ω))L2(0,T;H01(Ω))L(0,T,L2(Ω))L2(0,T;H01(Ω))\displaystyle L^{\infty}(0,T,L^{2}(\Omega))\cap L^{2}(0,T;X_{0}^{s}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega))\equiv L^{\infty}(0,T,L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)).
Now by Theorem 2.22 and Remark 2.23, as the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is increasing in k\displaystyle k, so the pointwise limit u\displaystyle u of {uk}k\displaystyle\{u_{k}\}_{k} exists; and satisfies u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) and u=0\displaystyle u=0 in (n\Ω)×(0,T)\displaystyle(\mathbb{R}^{n}\backslash\Omega)\times(0,T). Also since {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma+1}{2}}\right\}_{k} is bounded in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)), therefore ukγ+12uγ+12\displaystyle u_{k}^{\frac{\gamma+1}{2}}\rightharpoonup u^{\frac{\gamma+1}{2}} in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H^{1}_{0}(\Omega)). Again by Beppo Levi theorem uku\displaystyle u_{k}\rightarrow u in L1(ΩT)\displaystyle L^{1}(\Omega_{T}). Using Fatou’s Lemma, we have uγ+12L(0,T;L2(Ω))L(ΩT)\displaystyle u^{\frac{\gamma+1}{2}}\in L^{\infty}(0,T;L^{2}(\Omega))\cap L^{\infty}(\Omega_{T}). Also, by Remark 2.23 we get u(x,t)C(ω,t0,n,s)\displaystyle u(x,t)\geq C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right) for any ωΩ\displaystyle\omega\subset\subset\Omega and t0>0\displaystyle t_{0}>0. Using this positivity, it is then easy to show that u\displaystyle u is a very weak solution in the sense of Definition 2.24 and is shown in detail in the proof of Theorem 2.27.
In order to show that u𝒞([0,T],L2(Ω))\displaystyle u\in\mathcal{C}([0,T],L^{2}(\Omega)), we fix kl\displaystyle k\geq l and hence ukul\displaystyle u_{k}\geq u_{l} and note that the function u~=(uk+1k)(ul+1l)L2(0,T;H1(Ω))\displaystyle\tilde{u}=(u_{k}+\frac{1}{k})-(u_{l}+\frac{1}{l})\in L^{2}(0,T;H^{1}(\Omega)) satisfies the equation

{u~tΔu~+(Δ)su~=f(x,t)(uk+1k)γf(x,t)(ul+1l)γ, in ΩT,u~=1k1l, in (n\Ω)×(0,T),u~(x,0)=1k1l, in Ω;\begin{array}[]{lcr}\begin{cases}\tilde{u}_{t}-\Delta\tilde{u}+(-\Delta)^{s}\tilde{u}=\frac{f(x,t)}{(u_{k}+\frac{1}{k})^{\gamma}}-\frac{f(x,t)}{(u_{l}+\frac{1}{l})^{\gamma}},&\mbox{ in }\Omega_{T},\\ \tilde{u}=\frac{1}{k}-\frac{1}{l},&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \tilde{u}(x,0)=\frac{1}{k}-\frac{1}{l},&\mbox{ in }\Omega;\end{cases}\end{array}

Now as (f(uk+1k)γf(ul+1l)γ)0\displaystyle\left(\frac{f}{(u_{k}+\frac{1}{k})^{\gamma}}-\frac{f}{(u_{l}+\frac{1}{l})^{\gamma}}\right)\leq 0 in the set {u~0}\displaystyle\{\tilde{u}\geq 0\}, using Propositions 2.18 and 2.19, it holds that

(u~+)tΔ(u~+)+(Δ)s(u~+)0.(\tilde{u}_{+})_{t}-\Delta(\tilde{u}_{+})+(-\Delta)^{s}(\tilde{u}_{+})\leq 0.

Again we notice that u~+L2(0,T;H01(Ω))\displaystyle\tilde{u}_{+}\in L^{2}(0,T;H^{1}_{0}(\Omega)), has finite energy and u~+(x,0)=0\displaystyle\tilde{u}_{+}(x,0)=0, therefore by comparison principle, it holds that u~+0\displaystyle\tilde{u}_{+}\equiv 0, and then we have (uk+1k)(ul+1l)\displaystyle(u_{k}+\frac{1}{k})\leq(u_{l}+\frac{1}{l}) for kl\displaystyle k\geq l. Therefore we get 0ukul1l1k\displaystyle 0\leq u_{k}-u_{l}\leq\frac{1}{l}-\frac{1}{k}, for kl\displaystyle k\geq l. This implies that the sequence {uk}k\displaystyle\{u_{k}\}_{k} is Cauchy in 𝒞([0,T],L2(Ω))\displaystyle\mathcal{C}([0,T],L^{2}(\Omega)) and hence u𝒞([0,T],L2(Ω))\displaystyle u\in\mathcal{C}([0,T],L^{2}(\Omega)).
We now consider that γ1\displaystyle\gamma\leq 1, then using ukχ(0,t)\displaystyle u_{k}\chi_{(0,t)} as a test function in Eq. 2.9, we get that

12Ωuk2(x,t)𝑑x+0tΩ|uk|2𝑑x𝑑θ+120tQ(uk(x,θ)uk(y,θ))2|xy|n+2s𝑑x𝑑y𝑑θΩtfuk1γ𝑑x𝑑θ+12Ωu02(x)𝑑x.\begin{array}[]{l}\frac{1}{2}\int_{\Omega}u_{k}^{2}(x,t)dx+\int_{0}^{t}\int_{\Omega}|\nabla u_{k}|^{2}dxd\theta+\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)^{2}}{|x-y|^{n+2s}}dxdyd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{\mathrm{t}}}fu_{k}^{1-\gamma}dxd\theta+\frac{1}{2}\int_{\Omega}u_{0}^{2}(x)dx.\end{array}

Since {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L(ΩT)\displaystyle L^{\infty}\left(\Omega_{T}\right), and γ1\displaystyle\gamma\leq 1, so Ωtfuk1γ𝑑x𝑑θCfL(ΩT),\displaystyle\iint_{\Omega_{t}}fu_{k}^{1-\gamma}dxd\theta\leq C\|f\|_{L^{\infty}\left(\Omega_{T}\right)}, and then we conclude that {uk}k\displaystyle\left\{u_{k}\right\}_{k} is bounded in the space L2(0,T;H01(Ω))L(ΩT)\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(\Omega_{T}). We note that the same conclusion holds if Supp(f)ΩT\displaystyle\operatorname{Supp}(f)\subset\subset\Omega_{T} taking into consideration that uku1C\displaystyle u_{k}\geq u_{1}\geq C in Supp(f)\displaystyle\operatorname{Supp}(f) for each k\displaystyle k.

Proof of Theorem 2.27

We consider here the approximating problem Eq. 2.10 and refer Theorem 2.22 for properties of uk\displaystyle u_{k}. As fk\displaystyle f_{k} and u0k\displaystyle u_{0k} are bounded and nonnegative, therefore ukL2(0,T;H01(Ω))\displaystyle u_{k}\in L^{2}(0,T;H^{1}_{0}(\Omega)) is bounded and nonnegative, so we can choose the same test function ((uk(x,θ)+ε)γεγ)χ(0,t),0<ε<1/k,\displaystyle(\left(u_{k}(x,\theta)+\varepsilon\right)^{\gamma}-\varepsilon^{\gamma})\chi_{(0,t)},0<\varepsilon<1/k, in Eq. 2.10 also, and let ε0\displaystyle\varepsilon\to 0 to get

1γ+1Ωukγ+1(x,t)𝑑x+0tΩukukγdxdθ+120tQ(ukγ(x,θ)ukγ(y,θ))(uk(x,θ)uk(y,θ))|xy|n+2s𝑑x𝑑y𝑑θΩtfk𝑑x𝑑θ+1γ+1Ωu0kγ+1(x)𝑑xfL1(ΩT)+1γ+1u0Lγ+1(Ω).\begin{array}[]{l}\quad\frac{1}{\gamma+1}\int_{\Omega}u_{k}^{\gamma+1}(x,t)dx+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla u_{k}^{\gamma}dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(u_{k}^{\gamma}(x,\theta)-u_{k}^{\gamma}(y,\theta)\right)\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{t}}f_{k}dxd\theta+\frac{1}{\gamma+1}\int_{\Omega}u_{0k}^{\gamma+1}(x)dx\leq\|f\|_{L^{1}\left(\Omega_{T}\right)}+\frac{1}{\gamma+1}\left\|u_{0}\right\|_{L^{\gamma+1}\left(\Omega\right)}.\end{array}

Similarly like the proof of Theorem 2.26, it holds that {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma+1}{2}}\right\}_{k} is bounded in L(0,T;L2(Ω))L2(0,T;H01(Ω))\displaystyle L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)).
Now as {uk}k\displaystyle\left\{u_{k}\right\}_{k} is increasing in k\displaystyle k, we get by Fatou’s Lemma and Beppo Levi’s Lemma, that the pointwise limit (as mentioned in Remark 2.23) u\displaystyle u satisfies uγ+12L(0,T;L2(Ω))L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma+1}{2}}\in L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)) and uku\displaystyle u_{k}\uparrow u strongly in L1(ΩT)\displaystyle L^{1}(\Omega_{T}). Since uγ+12L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) and n(1+γ)n2=2(1+γ)2\displaystyle\frac{n(1+\gamma)}{n-2}=\frac{2^{*}(1+\gamma)}{2}, embedding result on uγ+12\displaystyle u^{\frac{\gamma+1}{2}} implies uL2(0,T;L(n(1+γ)n2)(Ω))\displaystyle u\in L^{2}\left(0,T;L^{\left(\frac{n(1+\gamma)}{n-2}\right)}(\Omega)\right). Also, we have u(x,t)\displaystyle u(x,t)\geq C(ω,t0,n,s)\displaystyle C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right) for any ωΩ\displaystyle\omega\subset\subset\Omega and t0>0\displaystyle t_{0}>0. We show that u\displaystyle u is a very weak solution to Eq. 2.12 in the sense of Definition 2.24. Let us take arbitrary ϕ𝒞0(ΩT)\displaystyle\phi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}), we then have

ΩT((uk)tΔuk+(Δ)suk)ϕ𝑑x𝑑t=ΩTfk(uk+1k)γϕ𝑑x𝑑t.\iint_{\Omega_{T}}(\left(u_{k}\right)_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k})\phi\,dxdt=\iint_{\Omega_{T}}\frac{f_{k}}{(u_{k}+\frac{1}{k})^{\gamma}}\phi dxdt.

Now ϕ𝒞0(ΩT)\displaystyle\phi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}) implies ϕt,Δϕ\displaystyle\phi_{t},\Delta\phi and (Δ)sϕ\displaystyle(-\Delta)^{s}\phi all are bounded. Then as uku\displaystyle u_{k}\rightarrow u strongly in L1(ΩT)\displaystyle L^{1}\left(\Omega_{T}\right), we have

ΩT((uk)tΔuk+(Δ)suk)ϕ𝑑x𝑑t=ΩTuk((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑tΩTu((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑t.\begin{array}[]{l}\quad\iint_{\Omega_{T}}(\left(u_{k}\right)_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k})\phi\,dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ =\iint_{\Omega_{T}}u_{k}(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi)dxdt\rightarrow\iint_{\Omega_{T}}u(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi)dxdt.\end{array}

Now since for all ωΩ\displaystyle\omega\subset\subset\Omega and for all t0>0,\displaystyle t_{0}>0, uk(x,t)C(ω,t0,n,s)\displaystyle u_{k}(x,t)\geq C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right), we get the positivity of {uk}k\displaystyle\{u_{k}\}_{k} as uk(x,t)C\displaystyle u_{k}(x,t)\geq C in Suppϕ\displaystyle\operatorname{Supp}\phi for all k1\displaystyle k\geq 1. Hence, we can use the dominated convergence theorem to obtain

ΩTfk(uk+1k)γϕ𝑑x𝑑tΩTfuγϕ𝑑x𝑑t as k.\iint_{\Omega_{T}}\frac{f_{k}}{(u_{k}+\frac{1}{k})^{\gamma}}\phi\,dxdt\rightarrow\iint_{\Omega_{T}}\frac{f}{u^{\gamma}}\phi\,dxdt\text{ as }k\rightarrow\infty.

Thus

ΩT(utΔu+(Δ)su)ϕ𝑑x𝑑t=ΩTfuγϕ𝑑x𝑑t.\iint_{\Omega_{T}}(u_{t}-\Delta u+(-\Delta)^{s}u)\phi\,dxdt=\iint_{\Omega_{T}}\frac{f}{u^{\gamma}}\phi\,dxdt.

Proof of Theorem 2.28

Here we consider Eq. 2.10 but with suffix m\displaystyle m\in\mathbb{N} as

{(um)tΔum+(Δ)sum=fm(x,t)(um+1m)γ in ΩT:=Ω×(0,T),um=0 in (n\Ω)×(0,T),um(x,0)=u0m(x) in Ω.{}\begin{array}[]{lcr}\begin{cases}(u_{m})_{t}-\Delta u_{m}+(-\Delta)^{s}u_{m}=\frac{f_{m}(x,t)}{(u_{m}+\frac{1}{m})^{\gamma}}&\mbox{ in }\Omega_{T}:=\Omega\times(0,T),\\ u_{m}=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u_{m}(x,0)=u_{0m}(x)&\mbox{ in }\Omega.\end{cases}\end{array} (3.2)

The properties of um\displaystyle u_{m}’s follow from Theorem 2.22. As umL2(0,T;H01(Ω))\displaystyle u_{m}\in L^{2}(0,T;H^{1}_{0}(\Omega)) is nonnegative and Tk(um)\displaystyle T_{k}(u_{m}) is bounded, so using ((Tk(um)+ε)γεγ)χ(0,t)\displaystyle\left(\left(T_{k}\left(u_{m}\right)+\varepsilon\right)^{\gamma}-\varepsilon^{\gamma}\right)\chi_{(0,t)} as a test function in Eq. 3.2, we get

0tΩ(um)t((Tk(um)+ε)γεγ)𝑑x𝑑θ+0tΩuk((Tk(um)+ε)γεγ)dxdθ+120tQ(uk(x,θ)uk(y,θ))((Tk(um(x,θ))+ε)γ(Tk(um(y,θ))+ε)γ)|xy|n+2s𝑑x𝑑y𝑑θΩTf(x,θ)𝑑x𝑑θ.\begin{array}[]{l}\quad\int_{0}^{t}\int_{\Omega}\left(u_{m}\right)_{t}\left((T_{k}(u_{m})+\varepsilon)^{\gamma}-\varepsilon^{\gamma}\right)dxd\theta+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla\left((T_{k}(u_{m})+\varepsilon)^{\gamma}-\varepsilon^{\gamma}\right)dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{(u_{k}(x,\theta)-u_{k}(y,\theta))((T_{k}(u_{m}(x,\theta))+\varepsilon)^{\gamma}-\left(T_{k}(u_{m}(y,\theta))+\varepsilon\right)^{\gamma})}{|x-y|^{n+2s}}dxdyd\theta\leq\iint_{\Omega_{T}}f(x,\theta)dxd\theta.\end{array}

We have chosen 0<ε<1/k\displaystyle 0<\varepsilon<1/k arbitrarily in above. Now letting ε0\displaystyle\varepsilon\to 0, by Fatou’s lemma, we get for all tT\displaystyle t\leq T

ΩLk(um(x,t))𝑑x+0tΩumTkγ(um)𝑑x𝑑θ+120tQ(Tkγ(um(x,θ))Tkγ(um(y,θ)))(um(x,θ)um(y,θ))|xy|n+2s𝑑x𝑑y𝑑θfL1(ΩT)+C3(k)u0L1(Ω)+C4(k)|Ω|,\begin{array}[]{l}\quad\int_{\Omega}L_{k}\left(u_{m}(x,t)\right)dx+\int_{0}^{t}\int_{\Omega}\nabla u_{m}\cdot\nabla T_{k}^{\gamma}(u_{m})dxd\theta\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(T_{k}^{\gamma}\left(u_{m}(x,\theta)\right)-T_{k}^{\gamma}\left(u_{m}(y,\theta)\right)\right)\left(u_{m}(x,\theta)-u_{m}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\\ \leq\|f\|_{L^{1}\left(\Omega_{T}\right)}+C_{3}(k)\left\|u_{0}\right\|_{L^{1}(\Omega)}+C_{4}(k)|\Omega|,\end{array}

where Lk(ρ)=0ρ(Tk(ξ))γ𝑑ξ\displaystyle L_{k}(\rho)=\int_{0}^{\rho}\left(T_{k}(\xi)\right)^{\gamma}d\xi. Notice that

(Tkγ(um(x,θ))Tkγ(um(y,θ)))(um(x,θ)um(y,θ))(Tkγ(um(x,θ))Tkγ(um(y,θ)))(Tk(um(x,θ))Tk(um(y,θ)))\begin{array}[]{l}\left(T_{k}^{\gamma}\left(u_{m}(x,\theta)\right)-T_{k}^{\gamma}\left(u_{m}(y,\theta)\right)\right)\left(u_{m}(x,\theta)-u_{m}(y,\theta)\right)\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \geq\left(T_{k}^{\gamma}\left(u_{m}(x,\theta)\right)-T_{k}^{\gamma}\left(u_{m}(y,\theta)\right)\right)\left(T_{k}\left(u_{m}(x,\theta)\right)-T_{k}\left(u_{m}(y,\theta)\right)\right)\end{array}

and so

umTkγ(um)γTkγ1(um)|Tk(um)|2\nabla u_{m}\cdot\nabla T_{k}^{\gamma}(u_{m})\geq\gamma T_{k}^{\gamma-1}(u_{m})|\nabla T_{k}(u_{m})|^{2}

and for ρ>0\displaystyle\rho>0, trivial calculation yields that

C1(k)ρC2(k)Lk(ρ)C3(k)ρ+C4(k) and Lk(ρ)C(Tk(ρ))γ+1,C_{1}(k)\rho-C_{2}(k)\leq L_{k}(\rho)\leq C_{3}(k)\rho+C_{4}(k)\text{ and }L_{k}(\rho)\geq C\left(T_{k}(\rho)\right)^{\gamma+1},

where C1,C2,C3,C4,C\displaystyle C_{1},C_{2},C_{3},C_{4},C are positive constants depending only on k\displaystyle k. Therefore {(Tk(um))γ+12}m\displaystyle\left\{\left(T_{k}\left(u_{m}\right)\right)^{\frac{\gamma+1}{2}}\right\}_{m} is bounded in L(0,T;L2(Ω))L2(0,T;H01(Ω))\displaystyle L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)) and {um}m\displaystyle\left\{u_{m}\right\}_{m} is bounded in L(0,T;L1(Ω))\displaystyle L^{\infty}(0,T;L^{1}(\Omega)).
Now as {um}m\displaystyle\left\{u_{m}\right\}_{m} is increasing in m\displaystyle m, by Fatou’s Lemma and Beppo Levi’s theorem, the pointwise limit u\displaystyle u of {um}m\displaystyle\{u_{m}\}_{m} (as mentioned in Remark 2.23) satisfies (Tk(u))γ+12L(0,T;L2(Ω))L2(0,T;H01(Ω))\displaystyle\left(T_{k}(u)\right)^{\frac{\gamma+1}{2}}\in L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)), umu\displaystyle u_{m}\uparrow u strongly in L1(ΩT)\displaystyle L^{1}(\Omega_{T}). Also u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) in L1\displaystyle L^{1} sense and u\displaystyle u is 0\displaystyle 0 outside Ω\displaystyle\Omega. Then u\displaystyle u can be shown to be a very weak solution of Eq. 2.12 in the sense of Definition 2.24, and the proof is same as that of Theorem 2.27.

Remark 3.1.

In view of Theorem 2.26, if we consider the approximating problem

{(uk)tΔuk+(Δ)suk=fk(x,t)ukγ in ΩT,uk=0 in (n\Ω)×(0,T),uk(x,0)=u0k(x) in Ω;\begin{array}[]{lcr}\begin{cases}(u_{k})_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k}=\frac{f_{k}(x,t)}{u_{k}^{\gamma}}&\mbox{ in }\Omega_{T},\\ u_{k}=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega)\times(0,T),\\ u_{k}(x,0)=u_{0k}(x)&\mbox{ in }\Omega;\end{cases}\end{array}

for γ1\displaystyle\gamma\leq 1, then for each k\displaystyle k, uk\displaystyle u_{k} exists and ukL2(0,T;H01(Ω))𝒞([0,T],L2(Ω))\displaystyle u_{k}\in L^{2}(0,T;H^{1}_{0}(\Omega))\cap\mathcal{C}([0,T],L^{2}(\Omega)). Also, each uk\displaystyle u_{k} is unique (see Remark 3.3). Further, the sequence {uk}k\displaystyle\{u_{k}\}_{k} satisfy the followings:
(a) each uk\displaystyle u_{k} is bounded and nonnegative,
(b) the sequence {uk}k\displaystyle\{u_{k}\}_{k} is increasing in k\displaystyle k (can be shown similarly like Lemma 2.20),
(c) for each t0>0\displaystyle t_{0}>0 and ωΩT\displaystyle\omega\subset\subset\Omega_{T}, C(ω,t0,n,s)\displaystyle\exists C(\omega,t_{0},n,s) such that ukC(ω,t0,n,s)\displaystyle u_{k}\geq C(\omega,t_{0},n,s).
Now for γ1\displaystyle\gamma\leq 1, we have (uk+ε)γεγukγ\displaystyle(u_{k}+\varepsilon)^{\gamma}-\varepsilon^{\gamma}\leq u_{k}^{\gamma}, and hence fk((uk+ε)γεγ)ukγfk\displaystyle\frac{f_{k}((u_{k}+\varepsilon)^{\gamma}-\varepsilon^{\gamma})}{u_{k}^{\gamma}}\leq f_{k}. Therefore we can follow the same procedure as that of Theorem 2.27 to get that the pointwise limit u\displaystyle u of {uk}k\displaystyle\{u_{k}\}_{k} is a very weak solution of Eq. 2.12 and satisfies the corresponding properties of Theorem 2.27. Now let kl\displaystyle k\geq l be fixed, then ukul\displaystyle u_{k}\geq u_{l} and

(ukul)tΔ(ukul)+(Δ)s(ukul)=fk(x,t)ukγfl(x,t)ulγfkflukγ in ΩT.\left(u_{k}-u_{l}\right)_{t}-\Delta(u_{k}-u_{l})+(-\Delta)^{s}\left(u_{k}-u_{l}\right)=\frac{f_{k}(x,t)}{u_{k}^{\gamma}}-\frac{f_{l}(x,t)}{u_{l}^{\gamma}}\leq\frac{f_{k}-f_{l}}{u_{k}^{\gamma}}\text{ in }\Omega_{T}.

Hence using ((ukul+ε)γεγ)χ(0,t)\displaystyle\left((u_{k}-u_{l}+\varepsilon)^{\gamma}-\varepsilon^{\gamma}\right)\chi_{(0,t)}, 0<ε<<1\displaystyle 0<\varepsilon<<1 as a test function in the above inequality and letting ε0\displaystyle\varepsilon\to 0 it holds that

1γ+1Ω(uk(x,t)ul(x,t))γ+1𝑑xΩt(fkfl)𝑑x𝑑θ+1γ+1Ω(u0k(x)u0l(x))γ+1(x)𝑑x.\begin{array}[]{l}\frac{1}{\gamma+1}\int_{\Omega}\left(u_{k}(x,t)-u_{l}(x,t)\right)^{\gamma+1}dx\leq\iint_{\Omega_{t}}\left(f_{k}-f_{l}\right)dxd\theta+\frac{1}{\gamma+1}\int_{\Omega}\left(u_{0k}(x)-u_{0l}(x)\right)^{\gamma+1}(x)dx.\end{array}

Now as by Dominated Convergence Theorem, fkf\displaystyle f_{k}\uparrow f strongly in L1(ΩT)\displaystyle L^{1}(\Omega_{T}) and u0ku0\displaystyle u_{0k}\uparrow u_{0} strongly in Lγ+1(Ω)\displaystyle L^{\gamma+1}(\Omega) if u0Lγ+1(Ω)\displaystyle u_{0}\in L^{\gamma+1}(\Omega), we get that the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is a Cauchy sequence in the space 𝒞([0,T];Lγ+1(Ω))\displaystyle\mathcal{C}([0,T];L^{\gamma+1}(\Omega)) and hence in 𝒞([0,T];L1(Ω))\displaystyle\mathcal{C}([0,T];L^{1}(\Omega)). Therefore this approximation allows us to have u𝒞([0,T],L1(Ω))\displaystyle u\in\mathcal{C}([0,T],L^{1}(\Omega)).
Note that the same approximation technique will not work for γ>1\displaystyle\gamma>1, as in that case we will have ukγ+12L2(0,T;H01(Ω))\displaystyle u_{k}^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)) which may not give us ukL2(0,T;H01(Ω))\displaystyle u_{k}\in L^{2}(0,T;H_{0}^{1}(\Omega)) and the monotonicity of {uk}k\displaystyle\{u_{k}\}_{k} cannot be proven in similar way like Lemma 2.20. However, we will take approximations as Eq. 2.10 for convenience even for γ1\displaystyle\gamma\leq 1.
Further, this approximation technique will not work for u0L1(Ω)\displaystyle u_{0}\in L^{1}(\Omega), as in that case, we need to take test functions like ((Tm(ukul)+ε)γεγ)χ(0,t)\displaystyle\left((T_{m}\left(u_{k}-u_{l})+\varepsilon\right)^{\gamma}-\varepsilon^{\gamma}\right)\chi_{(0,t)} and will end up with ΩLm(uk(x,t)ul(x,t))𝑑x\displaystyle\int_{\Omega}L_{m}\left(u_{k}(x,t)-u_{l}(x,t)\right)dx in the left which will not give us anything desired.

Proof of Theorem 2.29

We consider the approximated problems Eq. 2.10 here and refer Theorem 2.22. Since each ukL2(0,T;H01(Ω))\displaystyle u_{k}\in L^{2}(0,T;H^{1}_{0}(\Omega)) is bounded and γ>1\displaystyle\gamma>1 so we can use ukγχ(0,t)\displaystyle u_{k}^{\gamma}\chi_{(0,t)} as a test function in Eq. 2.10, to get that, for all tT\displaystyle t\leq T,

1γ+1Ωukγ+1(x,t)𝑑x+0tΩukukγdxdθ+120tQ(ukγ(x,θ)ukγ(y,θ))(uk(x,θ)uk(y,θ))|xy|n+2s𝑑x𝑑y𝑑θΩtf𝑑x𝑑θ+1γ+1Ωu0kγ+1(x)𝑑xfL1(ΩT)+1γ+1u0Lγ+1(Ω).{}\begin{array}[]{l}\quad\frac{1}{\gamma+1}\int_{\Omega}u_{k}^{\gamma+1}(x,t)dx+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla u_{k}^{\gamma}dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \quad+\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(u_{k}^{\gamma}(x,\theta)-u_{k}^{\gamma}(y,\theta)\right)\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{t}}fdxd\theta+\frac{1}{\gamma+1}\int_{\Omega}u_{0k}^{\gamma+1}(x)dx\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\|f\|_{L^{1}\left(\Omega_{T}\right)}+\frac{1}{\gamma+1}\left\|u_{0}\right\|_{L^{\gamma+1}\left(\Omega\right)}.\end{array} (3.3)

Using item (i)\displaystyle(i) of Lemma 2.11 and taking supremum over 0<tT\displaystyle 0<t\leq T, we get that {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma+1}{2}}\right\}_{k} is uniformly bounded in L(0,T;L2(Ω))L2(0,T;H01(Ω))\displaystyle L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)) and {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma+1}(\Omega)).
We now show that {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(t0,T;Hlocs(Ω))L2(t0,T;Hloc1(Ω))\displaystyle L^{2}(t_{0},T;H_{loc}^{s}(\Omega))\cap L^{2}(t_{0},T;H_{loc}^{1}(\Omega)) for each t0>0\displaystyle t_{0}>0. Since γ>1\displaystyle\gamma>1, and ΩT\displaystyle\Omega_{T} is bounded, and {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma+1}(\Omega)), we deduce that {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(0,T;L2(Ω))=L2(ΩT)\displaystyle L^{2}(0,T;L^{2}(\Omega))=L^{2}(\Omega_{T}), in particular in L2(K×(t0,T))\displaystyle L^{2}(K\times(t_{0},T)), for every subset K\displaystyle K compactly contained in Ω\displaystyle\Omega and for each t0>0\displaystyle t_{0}>0. Further, as K×KΩ×ΩQ\displaystyle K\times K\subset\Omega\times\Omega\subset Q and all the integrals in the left-hand-side of Eq. 3.3 are positive, hence we have,

t0TKK(uk(x,t)uk(y,t))(ukγ(x,t)ukγ(y,t))|xy|n+2s𝑑x𝑑y𝑑t2fL1(ΩT)+2γ+1u0Lγ+1(Ω),\begin{array}[]{l}\int_{t_{0}}^{T}\int_{K}\int_{K}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)\left(u_{k}^{\gamma}(x,t)-u_{k}^{\gamma}(y,t)\right)}{|x-y|^{n+2s}}dxdydt\leq 2\|f\|_{L^{1}\left(\Omega_{T}\right)}+\frac{2}{\gamma+1}\left\|u_{0}\right\|_{L^{\gamma+1}\left(\Omega\right)},\end{array}

and

t0TKukγ1|uk|2𝑑x𝑑t1γ(fL1(ΩT)+1γ+1u0Lγ+1(Ω)),\int_{t_{0}}^{T}\int_{K}u_{k}^{\gamma-1}|\nabla u_{k}|^{2}dxdt\leq\frac{1}{\gamma}\left(\|f\|_{L^{1}\left(\Omega_{T}\right)}+\frac{1}{\gamma+1}\left\|u_{0}\right\|_{L^{\gamma+1}\left(\Omega\right)}\right),

for every KΩ\displaystyle K\subset\subset\Omega and for each t0>0\displaystyle t_{0}>0.
We now apply the item (iii)\displaystyle(iii) of Lemma 2.11, to get

t0TKK|uk(x,t)uk(y,t)|2|uk(x,t)+uk(y,t)|γ1|xy|n+2s𝑑x𝑑y𝑑t2Cγ(fL1(ΩT)+u0Lγ+1(Ω)).\int_{t_{0}}^{T}\int_{K}\int_{K}\frac{\left|u_{k}(x,t)-u_{k}(y,t)\right|^{2}\left|u_{k}(x,t)+u_{k}(y,t)\right|^{\gamma-1}}{|x-y|^{n+2s}}dxdydt\leq 2C_{\gamma}\left(\|f\|_{L^{1}\left(\Omega_{T}\right)}+\left\|u_{0}\right\|_{L^{\gamma+1}(\Omega)}\right).

Using the positivity of uk\displaystyle u_{k} in ω×[t0,T)\displaystyle\omega\times[t_{0},T) for all k\displaystyle k, we get

t0TKK|uk(x,t)uk(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t22γCγc(K,t0)γ1(fL1(ΩT)+u0Lγ+1(Ω)),{}\int_{t_{0}}^{T}\int_{K}\int_{K}\frac{\left|u_{k}(x,t)-u_{k}(y,t)\right|^{2}}{|x-y|^{n+2s}}dxdydt\leq\frac{2^{2-\gamma}C_{\gamma}}{c_{(K,t_{0})}^{\gamma-1}}\left(\|f\|_{L^{1}\left(\Omega_{T}\right)}+\left\|u_{0}\right\|_{L^{\gamma+1}(\Omega)}\right), (3.4)

and

t0TK|uk|2𝑑x𝑑y𝑑t1γc(K,t0)γ1(fL1(ΩT)+1γ+1u0Lγ+1(Ω)).{}\int_{t_{0}}^{T}\int_{K}|\nabla u_{k}|^{2}dxdydt\leq\frac{1}{\gamma c_{(K,t_{0})}^{\gamma-1}}\left(\|f\|_{L^{1}\left(\Omega_{T}\right)}+\frac{1}{\gamma+1}\left\|u_{0}\right\|_{L^{\gamma+1}(\Omega)}\right). (3.5)

Hence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(t0,T;Hlocs(Ω))L2(t0,T;Hloc1(Ω))L2(t0,T;Hloc1(Ω))\displaystyle L^{2}(t_{0},T;H_{loc}^{s}(\Omega))\cap L^{2}(t_{0},T;H_{loc}^{1}(\Omega))\equiv L^{2}(t_{0},T;H_{loc}^{1}(\Omega)) for each t0>0\displaystyle t_{0}>0.
Since we have {ukγ+12}\displaystyle\left\{u_{k}^{\frac{\gamma+1}{2}}\right\} is uniformly bounded in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega))\subset L2(ΩT)\displaystyle L^{2}\left(\Omega_{T}\right), this implies that the increasing sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L1(ΩT)\displaystyle L^{1}\left(\Omega_{T}\right). Then, there exists a measurable function u\displaystyle u such that uku\displaystyle u_{k}\rightarrow u a.e. in ΩT\displaystyle\Omega_{T} and by Beppo Levi’s theorem uku\displaystyle u_{k}\rightarrow u in L1(ΩT)\displaystyle L^{1}(\Omega_{T}). Since uk=0\displaystyle u_{k}=0 in (n\Ω)×(0,T)\displaystyle\left(\mathbb{R}^{n}\backslash\Omega\right)\times(0,T), extending u\displaystyle u by zero outside of Ω\displaystyle\Omega we conclude that uku\displaystyle u_{k}\rightarrow u a.e. in n×(0,T)\displaystyle\mathbb{R}^{n}\times(0,T) with u=\displaystyle u= 0 in (n\Ω)×(0,T)\displaystyle\left(\mathbb{R}^{n}\backslash\Omega\right)\times(0,T). Now we use Fatou’s lemma in Eqs. 3.3, 3.4 and 3.5, to obtain for each t0>0\displaystyle t_{0}>0, uL2(t0,T;Hloc1(Ω))L(0,T;Lγ+1(Ω))\displaystyle u\in L^{2}(t_{0},T;H_{loc}^{1}(\Omega))\cap L^{\infty}(0,T;L^{\gamma+1}(\Omega)) and uγ+12L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma+1}{2}}\in L^{2}(0,T;H_{0}^{1}(\Omega)). As uk(x,0)=u0k(x)\displaystyle u_{k}(x,0)=u_{0k}(x) for each k\displaystyle k, so u(x,0)=u0(x)\displaystyle u(x,0)=u_{0}(x) in L1\displaystyle L^{1} sense (see Remark 2.23). The rest of the proof will follow similarly as that of Theorem 2.27.

Proof of Theorem 2.30

Here also, we consider the approximating problems Eq. 2.10. Since ukL2(0,T;H01(Ω))\displaystyle u_{k}\in L^{2}(0,T;H^{1}_{0}(\Omega)), we take uk(x,t)χ(0,τ)(t)\displaystyle u_{k}(x,t)\chi_{(0,\tau)}(t) as a test function in Eq. 2.10, to have

sup0τTΩuk2(x,τ)𝑑x+20TΩ|uk|2𝑑x𝑑t+0TQ(uk(x,t)uk(y,t))2|xy|n+2s𝑑x𝑑y𝑑t2ΩTfuk1γ𝑑x𝑑t+u0L2(Ω).{}\begin{array}[]{l}\sup_{0\leq\tau\leq T}\int_{\Omega}u_{k}^{2}(x,\tau)dx+2\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt+\int_{0}^{T}\int_{Q}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)^{2}}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq 2\iint_{\Omega_{T}}fu_{k}^{1-\gamma}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}.\end{array} (3.6)

Case 1: fL2γ+1(0,T;L(21γ)(Ω))\displaystyle f\in L^{\frac{2}{\gamma+1}}\left(0,T;L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)\right)
For this case, since fL2γ+1(0,T;L(21γ)(Ω))\displaystyle f\in L^{\frac{2}{\gamma+1}}\left(0,T;L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)\right), we apply the Hölder inequality two times, first for the space integral and then for the time integral, to obtain

ΩTfuk1γ𝑑x𝑑t0T(Ω|f(x,t)|(21γ)𝑑x)1(21γ)(Ω|uk(x,t)|2𝑑x)1γ2𝑑t=0TfL(21γ)(Ω)ukL2(Ω)1γ𝑑t(0TfL(21γ)(Ω)21+γ𝑑t)1+γ2(0TukL2(Ω)2𝑑t)1γ2.\begin{array}[]{rcl}\iint_{\Omega_{T}}fu_{k}^{1-\gamma}dxdt&\leq&\int_{0}^{T}\left(\int_{\Omega}|f(x,t)|^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}dx\right)^{\frac{1}{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{2^{*}}dx\right)^{\frac{1-\gamma}{2^{*}}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &=&\int_{0}^{T}\|f\|_{L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}{(\Omega)}}\left\|u_{k}\right\|_{L^{2^{*}}(\Omega)}^{1-\gamma}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\left(\int_{0}^{T}\|f\|_{L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)}^{\frac{2}{1+\gamma}}dt\right)^{\frac{1+\gamma}{2}}\left(\int_{0}^{T}\left\|u_{k}\right\|_{L^{2^{*}}(\Omega)}^{2}dt\right)^{\frac{1-\gamma}{2}}.\end{array}

We now apply the Sobolev embedding as of Theorem 2.8 in the last term on the right-hand-side to get

ΩTfuk1γdxdt(C(n))1γ2fL2γ+1(0,T;L(21γ)(Ω))[0TΩ|uk|2dxdt]1γ2.\iint_{\Omega_{T}}fu_{k}^{1-\gamma}dxdt\quad\leq(C(n))^{\frac{1-\gamma}{2}}\|f\|_{L^{\frac{2}{\gamma+1}}\left(0,T;L^{\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}}(\Omega)\right)}\left[\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt\right]^{\frac{1-\gamma}{2}}.

Since γ<1\displaystyle\gamma<1, so we use Young’s inequality to deduce from Eq. 3.6 that

sup0τTΩuk2(x,τ)𝑑x+0TΩ|uk|2𝑑x𝑑t+0TQ(uk(x,t)uk(y,t))2|xy|n+2s𝑑x𝑑y𝑑tC,\begin{array}[]{l}\sup_{0\leq\tau\leq T}\int_{\Omega}u_{k}^{2}(x,\tau)dx+\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt+\int_{0}^{T}\int_{Q}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)^{2}}{|x-y|^{n+2s}}dxdydt\leq C,\end{array}

where C\displaystyle C is a positive constant independent of k\displaystyle k.
Case 2: fLm¯(ΩT)\displaystyle f\in L^{\bar{m}}(\Omega_{T})
For this case, we notice that as fLm¯(ΩT)\displaystyle f\in L^{\bar{m}}\left(\Omega_{T}\right), we can apply the Hölder inequality in the first term on the right-hand-side in Eq. 3.6 with exponents m¯\displaystyle\bar{m} and m¯\displaystyle\bar{m}^{\prime}, to get

sup0τTΩuk2(x,τ)𝑑x+0TΩ|uk|2𝑑x𝑑t+0TQ|uk(x,t)uk(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t2fLm¯(ΩT)[ΩT|uk|(1γ)m¯𝑑x𝑑t]1m¯+u0L2(Ω).{}\begin{array}[]{l}\sup_{0\leq\tau\leq T}\int_{\Omega}u_{k}^{2}(x,\tau)dx+\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt+\int_{0}^{T}\int_{Q}\frac{\left|u_{k}(x,t)-u_{k}(y,t)\right|^{2}}{|x-y|^{n+2s}}dxdydt\\ \leq 2\|f\|_{L^{\bar{m}}\left(\Omega_{T}\right)}\left[\iint_{\Omega_{T}}|u_{k}|^{(1-\gamma)\bar{m}^{\prime}}dxdt\right]^{\frac{1}{\bar{m}^{\prime}}}+\left\|u_{0}\right\|_{L^{2}(\Omega)}.\end{array} (3.7)

We observe that (1γ)m¯=2(n+2)n\displaystyle(1-\gamma)\bar{m}^{\prime}=\frac{2(n+2)}{n} and hence using the Hölder inequality with the exponents nn2\displaystyle\frac{n}{n-2} and n2\displaystyle\frac{n}{2} and by Sobolev embedding (Theorem 2.8), we reach that

ΩT|uk|2(n+2)n𝑑x𝑑t=ΩT|uk|2|uk|4n𝑑x𝑑t0T[Ω|uk(x,t)|2𝑑x]2nukL2(Ω)2𝑑tC(n)[sup0tTΩ|uk(x,t)|2𝑑x]2n0TΩ|uk|2𝑑x𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{2(n+2)}{n}}dxdt&=&\iint_{\Omega_{T}}|u_{k}|^{2}\left|u_{k}\right|^{\frac{4}{n}}dxdt\\ &\leq&\int_{0}^{T}\left[\int_{\Omega}\left|u_{k}(x,t)\right|^{2}dx\right]^{\frac{2}{n}}\left\|u_{k}\right\|_{L^{2^{*}}(\Omega)}^{2}dt\\ &\leq&C(n)\left[\sup_{0\leq t\leq T}\int_{\Omega}\left|u_{k}(x,t)\right|^{2}dx\right]^{\frac{2}{n}}\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt.\end{array}

So using Eq. 3.7 and convexity argument, we get

ΩT|uk|2(n+2)n𝑑x𝑑tC(n)22n((2fLm¯(ΩT))n+2n(ΩT|uk|(1γ)m¯)n+2nm¯+(u0L2(Ω))n+2n).\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{2(n+2)}{n}}dxdt\leq C(n)2^{\frac{2}{n}}\left(\left(2\|f\|_{L^{\bar{m}}\left(\Omega_{T}\right)}\right)^{\frac{n+2}{n}}\left(\iint_{\Omega_{T}}|u_{k}|^{(1-\gamma){\bar{m}^{\prime}}}\right)^{\frac{n+2}{n\bar{m}^{\prime}}}+\left(\left\|u_{0}\right\|_{L^{2}(\Omega)}\right)^{\frac{n+2}{n}}\right).

Since n+2nm¯=1γ2<1\displaystyle\frac{n+2}{n\bar{m}^{\prime}}=\frac{1-\gamma}{2}<1, we use the Young inequality to obtain

ΩT|uk|2(n+2)n𝑑x𝑑tC,\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{2(n+2)}{n}}dxdt\leq C,

where C\displaystyle C is a positive constant independent of k\displaystyle k. Therefore, by Eq. 3.7 we deduce that the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in the space L2(0,T;X0s(Ω))L2(0,T;H01(Ω))L(0,T;L2(Ω))L2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle L^{2}(0,T;X_{0}^{s}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega))\equiv L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)).
Now the rest of the proof follows similarly as that of Theorem 2.27. However, for the sake of completeness, we include it here in a bit different way.
Since the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in the reflexive Banach space L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)), there exist a subsequence of {uk}k\displaystyle\left\{u_{k}\right\}_{k}, still indexed by k\displaystyle k, and a measurable function uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega)) such that uku\displaystyle u_{k}\rightharpoonup u weakly in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)) and uku\displaystyle u_{k}\rightarrow u strongly in L2(ΩT)\displaystyle L^{2}\left(\Omega_{T}\right) and a.e. in Ω×(0,T)\displaystyle\Omega\times(0,T). In addition, since uk=u=0\displaystyle u_{k}=u=0 on 𝒞Ω×(0,T)\displaystyle\mathcal{C}\Omega\times(0,T), extending u0\displaystyle u\equiv 0 outside Ω\displaystyle\Omega, we obtain uku\displaystyle u_{k}\rightarrow u for a.e. (x,t)n×(0,T)\displaystyle(x,t)\in\mathbb{R}^{n}\times(0,T). Again since {uk}\displaystyle\left\{u_{k}\right\} is increasing in k\displaystyle k, and uk(x,0)=u0k(x)\displaystyle u_{k}(x,0)=u_{0k}(x), so we have u(x,0)=u0(x)\displaystyle u(x,0)=u_{0}(x) in L1\displaystyle L^{1} sense (Remark 2.23). Also, by Fatou’s Lemma, we get that uL(0,T;L2(Ω))\displaystyle u\in L^{\infty}(0,T;L^{2}(\Omega)). Hence it follows that

uk(x,t)uk(y,t)|xy|n+2s2u(x,t)u(y,t)|xy|n+2s2 a.e. in Q×(0,T).\frac{u_{k}(x,t)-u_{k}(y,t)}{|x-y|^{\frac{n+2s}{2}}}\rightarrow\frac{u(x,t)-u(y,t)}{|x-y|^{\frac{n+2s}{2}}}\text{ a.e. in }Q\times(0,T).

We take an arbitrary test function φ𝒞0(ΩT)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}) in Eq. 2.10 to get

ΩTukφt𝑑x𝑑t+0TΩuk(Δϕ)𝑑x𝑑t+12QT(uk(x,t)uk(y,t))(φ(x,t)φ(y,t))|xy|n+2s𝑑x𝑑y𝑑t=ΩTfkφ(uk+1k)γ𝑑x𝑑t.\begin{array}[]{l}\quad-\iint_{\Omega_{T}}u_{k}\varphi_{t}dxdt+\int_{0}^{T}\int_{\Omega}u_{k}(-\Delta\phi)dxdt\\ \quad+\frac{1}{2}\iint_{Q_{T}}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)(\varphi(x,t)-\varphi(y,t))}{|x-y|^{n+2s}}dxdydt=\iint_{\Omega_{T}}\frac{f_{k}\varphi}{\left(u_{k}+\frac{1}{k}\right)^{\gamma}}dxdt.\end{array}

Since φ𝒞0(ΩT)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}), therefore φt\displaystyle\varphi_{t} and φ\displaystyle\nabla\varphi both are in L2(ΩT)\displaystyle L^{2}(\Omega_{T}) and since strong convergence implies weak convergence too, so it is clear that

limk0TΩukϕt𝑑x𝑑t=0TΩuϕt𝑑x𝑑t\lim_{k\rightarrow\infty}\int_{0}^{T}\int_{\Omega}u_{k}\phi_{t}dxdt=\int_{0}^{T}\int_{\Omega}u\phi_{t}dxdt

and by weak convergence in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H^{1}_{0}(\Omega)), we get

0TΩuk(Δϕ)𝑑x𝑑t=0TΩukϕdxdt0TΩuϕdxdt=0TΩu(Δϕ)𝑑x𝑑t, as k.\begin{array}[]{rcl}\int_{0}^{T}\int_{\Omega}u_{k}(-\Delta\phi)dxdt&=&\int_{0}^{T}\int_{\Omega}\nabla u_{k}\cdot\nabla\phi dxdt\\ &\to&\int_{0}^{T}\int_{\Omega}\nabla u\cdot\nabla\phi dxdt=\int_{0}^{T}\int_{\Omega}u(-\Delta\phi)dxdt,\text{ as }k\to\infty.\end{array}

We now define

Fk(x,y,t)=uk(x,t)uk(y,t)|xy|n+2s2 and F(x,y,t)=u(x,t)u(y,t)|xy|n+2s2F_{k}(x,y,t)=\frac{u_{k}(x,t)-u_{k}(y,t)}{|x-y|^{\frac{n+2s}{2}}}\text{ and }F(x,y,t)=\frac{u(x,t)-u(y,t)}{|x-y|^{\frac{n+2s}{2}}}\text{. }

Then as FkF\displaystyle F_{k}\rightarrow F a.e. in Q×(0,T)\displaystyle Q\times(0,T) and {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(0,T;X0s(Ω))\displaystyle L^{2}\left(0,T;X_{0}^{s}(\Omega)\right), by weak convergence we reach that

limkQT(uk(x,t)uk(y,t))(φ(x,t)φ(y,t))|xy|n+2s𝑑x𝑑y𝑑t=QT(u(x,t)u(y,t))(φ(x,t)φ(y,t))|xy|n+2s𝑑x𝑑y𝑑t,\begin{array}[]{l}\lim_{k\rightarrow\infty}\iint_{Q_{T}}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)(\varphi(x,t)-\varphi(y,t))}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ =\iint_{Q_{T}}\frac{(u(x,t)-u(y,t))(\varphi(x,t)-\varphi(y,t))}{|x-y|^{n+2s}}dxdydt,\end{array}

for all φ𝒞0(ΩT)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}). Now since for all ωΩ\displaystyle\omega\subset\subset\Omega and for all t0>0,\displaystyle t_{0}>0, we have uk(x,t)C(ω,t0,n,s)\displaystyle u_{k}(x,t)\geq C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right), so we reach that uk(x,t)C\displaystyle u_{k}(x,t)\geq C in Suppϕ\displaystyle\operatorname{Supp}\phi (as support of ϕ\displaystyle\phi is compact in ΩT\displaystyle\Omega_{T}) for all k1\displaystyle k\geq 1. Using this fact, we then have

0|fkφ(uk+1k)γ|φL(ΩT)|f|CγL1(ΩT).0\leq\left|\frac{f_{k}\varphi}{\left(u_{k}+\frac{1}{k}\right)^{\gamma}}\right|\leq\frac{\|\varphi\|_{L^{\infty}\left(\Omega_{T}\right)}|f|}{C^{\gamma}}\in L^{1}\left(\Omega_{T}\right).

So, by the dominated convergence theorem, we get

limkΩTfkφ(uk+1k)γ𝑑x𝑑t=ΩTfφuγ𝑑x𝑑t.\lim_{k\rightarrow\infty}\iint_{\Omega_{T}}\frac{f_{k}\varphi}{\left(u_{k}+\frac{1}{k}\right)^{\gamma}}dxdt=\iint_{\Omega_{T}}\frac{f\varphi}{u^{\gamma}}dxdt.

Finally, we pass to the limit as k\displaystyle k\rightarrow\infty to get

ΩTu(x,t)φt(x,t)𝑑x𝑑t0TΩu(x,t)Δϕ(x,t)𝑑x𝑑t+12QT(u(x,t)u(y,t))(φ(x,t)φ(y,t))|xy|n+2s𝑑x𝑑y𝑑t=ΩTf(x,t)φ(x,t)uγ(x,t)𝑑x𝑑t,\begin{array}[]{l}\quad-\iint_{\Omega_{T}}u(x,t)\varphi_{t}(x,t)dxdt-\int_{0}^{T}\int_{\Omega}u(x,t)\Delta\phi(x,t)dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \quad+\frac{1}{2}\iint_{Q_{T}}\frac{(u(x,t)-u(y,t))(\varphi(x,t)-\varphi(y,t))}{|x-y|^{n+2s}}dxdydt=\iint_{\Omega_{T}}\frac{f(x,t)\varphi(x,t)}{u^{\gamma}(x,t)}dxdt,\end{array}

for all ϕ𝒞0(ΩT)\displaystyle\phi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}). Therefore, u\displaystyle u is a weak solution to Eq. 2.12.

Proof of Theorem 2.32

As uk(0)L(ΩT)L2(0,T;H01(Ω))\displaystyle u_{k}(\geq 0)\in L^{\infty}(\Omega_{T})\cap L^{2}(0,T;H^{1}_{0}(\Omega)), so ((uk(x,t)+ε)θεθ)L2(0,T;H01(Ω))\displaystyle(\left(u_{k}(x,t)+\varepsilon\right)^{\theta}-\varepsilon^{\theta})\in L^{2}(0,T;H^{1}_{0}(\Omega)), for any ε,θ>0\displaystyle\varepsilon,\theta>0. So choosing γθ<1,0<ε<1/k\displaystyle\gamma\leq\theta<1,0<\varepsilon<1/k, we take the test function ((uk(x,t)+ε)θεθ)χ(0,τ)(t)\displaystyle(\left(u_{k}(x,t)+\varepsilon\right)^{\theta}-\varepsilon^{\theta})\chi_{(0,\tau)}(t) in Eq. 2.10, to get for each τT\displaystyle\tau\leq T

0τΩ(uk)t((uk+ε))θεθ)dxdt+0τΩuk((uk+ε)θεθ)dxdt+120τQ(uk(x,t)uk(y,t))((uk(x,t)+ε)θ(uk(y,t)+ε)θ)|xy|n+2s𝑑x𝑑y𝑑tΩTf(x,t)(uk(x,t)+ε)θγ𝑑x𝑑t.\begin{array}[]{l}\quad\int_{0}^{\tau}\int_{\Omega}\left(u_{k}\right)_{t}\left((u_{k}+\varepsilon))^{\theta}-\varepsilon^{\theta}\right)dxdt+\int_{0}^{\tau}\int_{\Omega}\nabla u_{k}\cdot\nabla\left((u_{k}+\varepsilon)^{\theta}-\varepsilon^{\theta}\right)dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{\tau}\int_{Q}\frac{(u_{k}(x,t)-u_{k}(y,t))((u_{k}(x,t)+\varepsilon)^{\theta}-\left(u_{k}(y,t)+\varepsilon\right)^{\theta})}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{T}}f(x,t)\left(u_{k}(x,t)+\varepsilon\right)^{\theta-\gamma}dxdt.\end{array}

Letting ε0\displaystyle\varepsilon\rightarrow 0, integrating the first term, and taking the supremum over τ[0,T]\displaystyle\tau\in[0,T], we get

2θ+1sup0τTΩ|uk(x,τ)|θ+1𝑑x+20TΩukukθdxdt+0TQ(uk(x,t)uk(y,t))(ukθ(x,t)ukθ(y,t))|xy|n+2s𝑑x𝑑y𝑑t2ΩTfukθγ𝑑x𝑑t+2θ+1u0L2(Ω).{}\begin{array}[]{c}\quad\frac{2}{\theta+1}\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{\theta+1}dx+2\int_{0}^{T}\int_{\Omega}\nabla u_{k}\cdot\nabla u_{k}^{\theta}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\int_{0}^{T}\int_{Q}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)\left(u_{k}^{\theta}(x,t)-u_{k}^{\theta}(y,t)\right)}{|x-y|^{n+2s}}dxdydt\leq 2\iint_{\Omega_{T}}fu_{k}^{\theta-\gamma}dxdt+\frac{2}{\theta+1}\left\|u_{0}\right\|_{L^{2}(\Omega)}.\end{array} (3.8)

Since all terms on the left are positive, taking supremum was allowed. Then by item i)\displaystyle i) of Lemma 2.11, we get

θ+12θsup0τTΩ|uk(x,τ)|θ+1𝑑x+(θ+1)220TΩukθ1|uk|2𝑑x𝑑t+0TQ|ukθ+12(x,t)ukθ+12(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t2θ(ΩTfukθγ𝑑x𝑑t+u0L2(Ω)).{}\begin{array}[]{c}\quad\frac{\theta+1}{2\theta}\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{\theta+1}dx+\frac{(\theta+1)^{2}}{2}\int_{0}^{T}\int_{\Omega}u_{k}^{\theta-1}|\nabla u_{k}|^{2}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\int_{0}^{T}\int_{Q}\frac{|u_{k}^{\frac{\theta+1}{2}}(x,t)-u_{k}^{\frac{\theta+1}{2}}(y,t)|^{2}}{|x-y|^{n+2s}}dxdydt\leq\frac{2}{\theta}\left(\iint_{\Omega_{T}}fu_{k}^{\theta-\gamma}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\right).\end{array}

In the previous inequality, we have used the fact that θ+1<2\displaystyle\theta+1<2, now we observe that the term θ+12θ>1\displaystyle\frac{\theta+1}{2\theta}>1 in the left-hand-side can be dropped. So we get

sup0τTΩ|uk(x,τ)|θ+1𝑑x+(θ+1)240τΩukθ1|uk|2𝑑x𝑑t+0TQ|ukθ+12(x,t)ukθ+12(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t2θ(ΩTfukθγ𝑑x𝑑t+u0L2(Ω)).{}\begin{array}[]{c}\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{\theta+1}dx+\frac{(\theta+1)^{2}}{4}\int_{0}^{\tau}\int_{\Omega}u_{k}^{\theta-1}|\nabla u_{k}|^{2}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\int_{0}^{T}\int_{Q}\frac{|u_{k}^{\frac{\theta+1}{2}}(x,t)-u_{k}^{\frac{\theta+1}{2}}(y,t)|^{2}}{|x-y|^{n+2s}}dxdydt\leq\frac{2}{\theta}\left(\iint_{\Omega_{T}}fu_{k}^{\theta-\gamma}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\right).\end{array} (3.9)

Again, using the same tool like Theorem 2.30 and the Hölder inequality with exponent nn2\displaystyle\frac{n}{n-2} and n2\displaystyle\frac{n}{2} we get

ΩT|uk|(θ+1)(n+2)n𝑑x𝑑t=ΩT|uk|2θ+12|uk|4nθ+12𝑑x𝑑t0T(Ω|uk(x,t)|2θ+12𝑑x)n2n(Ω|uk(x,t)|θ+1𝑑x)2n𝑑t=0Tukθ+12L2(Ω)2(Ω|uk(x,t)|θ+1𝑑x)2n𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{(\theta+1)(n+2)}{n}}dxdt&=&\iint_{\Omega_{T}}\left|u_{k}\right|^{2\frac{\theta+1}{2}}\left|u_{k}\right|^{\frac{4}{n}\frac{\theta+1}{2}}dxdt\\ &\leq&\int_{0}^{T}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{2^{*}\frac{\theta+1}{2}}dx\right)^{\frac{n-2}{n}}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{\theta+1}dx\right)^{\frac{2}{n}}dt\\ &=&\int_{0}^{T}\left\|u_{k}^{\frac{\theta+1}{2}}\right\|_{L^{2^{*}}(\Omega)}^{2}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{\theta+1}dx\right)^{\frac{2}{n}}dt.\end{array}

We now take the supremum over t[0,T]\displaystyle t\in[0,T], and apply the Sobolev embedding as that of Theorem 2.8, to get that

ΩT|uk|(θ+1)(n+2)n𝑑x𝑑tC(n)(sup0tTΩ|uk(x,t)|θ+1𝑑x)2nΩT|ukθ+12|2𝑑x𝑑t.\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{(\theta+1)(n+2)}{n}}dxdt\leq C(n)\left(\sup_{0\leq t\leq T}\int_{\Omega}\left|u_{k}(x,t)\right|^{\theta+1}dx\right)^{\frac{2}{n}}\iint_{\Omega_{T}}|\nabla u_{k}^{\frac{\theta+1}{2}}|^{2}dxdt.

Using Eq. 3.9 and convexity argument, we now get

ΩT|uk|(θ+1)(n+2)n𝑑x𝑑tC(n)(2θ)n+2n(ΩTfukθγ𝑑x𝑑t+u0L2(Ω))n+2nC(n)22n(2θ)n+2n((ΩTfukθγ𝑑x𝑑t)n+2n+u0L2(Ω)(n+2)n).{}\begin{array}[]{rcl}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{(\theta+1)(n+2)}{n}}dxdt&\leq&C(n)\left(\frac{2}{\theta}\right)^{\frac{n+2}{n}}\left(\iint_{\Omega_{T}}fu_{k}^{\theta-\gamma}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\right)^{\frac{n+2}{n}}\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&C(n)2^{\frac{2}{n}}\left(\frac{2}{\theta}\right)^{\frac{n+2}{n}}\left(\left(\iint_{\Omega_{T}}fu_{k}^{\theta-\gamma}dxdt\right)^{\frac{n+2}{n}}+\left\|u_{0}\right\|_{L^{2}(\Omega)}^{\frac{(n+2)}{n}}\right).\end{array} (3.10)

Case 1: m=1\displaystyle m=1
Now for the case m=1\displaystyle m=1, we take θ=γ\displaystyle\theta=\gamma in the Eq. 3.10. So we obtain

ΩT|uk|(γ+1)(n+2)n𝑑x𝑑tC(n)22n(2γ)n+2n(fL1(ΩT)n+2n+u0L2(Ω)(n+2)n).{}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{(\gamma+1)(n+2)}{n}}dxdt\leq C(n)2^{\frac{2}{n}}\left(\frac{2}{\gamma}\right)^{\frac{n+2}{n}}\left(\|f\|_{L^{1}(\Omega_{T})}^{\frac{n+2}{n}}+\left\|u_{0}\right\|_{L^{2}(\Omega)}^{\frac{(n+2)}{n}}\right). (3.11)

Also, by Eq. 3.9 we easily have

ukL(0,T;Lγ+1(Ω))2γ[fL1(ΩT)+u0L2(Ω)].{}\left\|u_{k}\right\|_{L^{\infty}(0,T;L^{\gamma+1}(\Omega))}\leq\frac{2}{\gamma}\left[\|f\|_{L^{1}\left(\Omega_{T}\right)}+\left\|u_{0}\right\|_{L^{2}(\Omega)}\right]. (3.12)

Case 2: m>1\displaystyle m>1
For m>1\displaystyle m>1, we take γ<θ<1\displaystyle\gamma<\theta<1. Using Eq. 3.10 and applying Hölder inequality with exponent m\displaystyle m and m\displaystyle m^{\prime}, we have

ΩT|uk|(θ+1)(n+2)n𝑑x𝑑tC1fLm(ΩT)n+2n[ΩT|uk|m(θγ)𝑑x𝑑t]n+2nm+C1u0L2(Ω)(n+2)n,\iint_{\Omega_{T}}|u_{k}|^{\frac{(\theta+1)(n+2)}{n}}dxdt\leq C_{1}\|f\|_{L^{m}\left(\Omega_{T}\right)}^{\frac{n+2}{n}}\left[\iint_{\Omega_{T}}|u_{k}|^{m^{\prime}(\theta-\gamma)}dxdt\right]^{\frac{n+2}{nm^{\prime}}}+C_{1}\left\|u_{0}\right\|_{L^{2}(\Omega)}^{\frac{(n+2)}{n}},

where C1=C(n)22n(2θ)n+2n\displaystyle C_{1}=C(n)2^{\frac{2}{n}}\left(\frac{2}{\theta}\right)^{\frac{n+2}{n}}. We now choose γ<θ<1\displaystyle\gamma<\theta<1 to be such that

(θ+1)(n+2)n=m(θγ), i.e. θ=(n+2)(m1)+nmγn2(m1).\frac{(\theta+1)(n+2)}{n}=m^{\prime}(\theta-\gamma),\text{ i.e. }\theta=\frac{(n+2)(m-1)+nm\gamma}{n-2(m-1)}.

We note that the condition θ<1\displaystyle\theta<1 is equivalent to m<m¯\displaystyle m<\bar{m}; while γ<θ\displaystyle\gamma<\theta is always fulfilled. Since n+2nm<1\displaystyle\frac{n+2}{nm^{\prime}}<1, applying Young’s inequality with ε>0\displaystyle\varepsilon>0, we get

ΩT|uk|(θ+1)(n+2)n𝑑x𝑑tC1fLm(ΩT)n+2n(εΩT|uk|(θ+1)(n+2)n𝑑x𝑑t+C(ε))+C1u0L2(Ω)(n+2)n.\vskip 3.0pt plus 1.0pt minus 1.0pt\iint_{\Omega_{T}}|u_{k}|^{\frac{(\theta+1)(n+2)}{n}}dxdt\leq C_{1}\|f\|_{L^{m}\left(\Omega_{T}\right)}^{\frac{n+2}{n}}\left(\varepsilon\iint_{\Omega_{T}}|u_{k}|^{\frac{(\theta+1)(n+2)}{n}}dxdt+C(\varepsilon)\right)+C_{1}\left\|u_{0}\right\|_{L^{2}(\Omega)}^{\frac{(n+2)}{n}}.

We choose ε\displaystyle\varepsilon small enough such that εC1fLm(ΩT)n+2n=12\displaystyle\varepsilon C_{1}\|f\|_{L^{m}\left(\Omega_{T}\right)}^{\frac{n+2}{n}}=\frac{1}{2} and using the fact that

σ:=m(γ+1)(n+2)n2(m1)=(θ+1)(n+2)n=m(θγ),\sigma:=\frac{m(\gamma+1)(n+2)}{n-2(m-1)}=\frac{(\theta+1)(n+2)}{n}=m^{\prime}(\theta-\gamma),

we get

ΩT|uk|σ𝑑x𝑑tC,{}\iint_{\Omega_{T}}\left|u_{k}\right|^{\sigma}dxdt\leq C, (3.13)

where C\displaystyle C is a positive constant independent of k\displaystyle k. Now in Eq. 3.9 we use Hölder inequality, to get

sup0τTΩukθ+1(x,τ)𝑑x2θ(fLm(ΩT)ukLσ(ΩT)σm+u0L2(Ω)).\sup_{0\leq\tau\leq T}\int_{\Omega}u_{k}^{\theta+1}(x,\tau)dx\leq\frac{2}{\theta}\left(\|f\|_{L^{m}\left(\Omega_{T}\right)}\left\|u_{k}\right\|_{L^{\sigma}\left(\Omega_{T}\right)}^{\frac{\sigma}{m^{\prime}}}+\left\|u_{0}\right\|_{L^{2}(\Omega)}\right).

Since γ<θ\displaystyle\gamma<\theta and by Eq. 3.13 we conclude that the sequence {uk}\displaystyle\left\{u_{k}\right\} is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma+1}(\Omega)), the same thing holds for the case m=1\displaystyle m=1 by Eq. 3.12. Finally, by Eq. 3.11 and Eq. 3.13 we conclude that in both cases, that is 1m<m¯\displaystyle 1\leq m<\bar{m}, the sequence {uk}\displaystyle\left\{u_{k}\right\} is uniformly bounded in Lσ(ΩT)\displaystyle L^{\sigma}(\Omega_{T}), σ:=m(γ+1)(n+2)n2(m1)\displaystyle\sigma:=\frac{m(\gamma+1)(n+2)}{n-2(m-1)}.
Now from Eq. 3.9, again using Hölder inequality, we estimate as

ΩT|ukθ+12|2𝑑x𝑑t2θ(fLm(ΩT)ukLσ(ΩT)σm+u0L2(Ω))C,{}\iint_{\Omega_{T}}|\nabla u_{k}^{\frac{\theta+1}{2}}|^{2}dxdt\leq\frac{2}{\theta}\left(\|f\|_{L^{m}\left(\Omega_{T}\right)}\left\|u_{k}\right\|_{L^{\sigma}\left(\Omega_{T}\right)}^{\frac{\sigma}{m^{\prime}}}+\left\|u_{0}\right\|_{L^{2}(\Omega)}\right)\leq C, (3.14)

where C>0\displaystyle C>0 is a constant independent of k\displaystyle k. Let 1<q¯<2\displaystyle 1<\bar{q}<2 will be specified later. By Hölder inequality, we have

ΩT|uk|q¯𝑑x𝑑t=ΩT|uk|q¯ukθ1ukθ1𝑑x𝑑t(ΩT|uk|2ukθ1𝑑x𝑑t)q¯2×(ΩTukθ1uk2(θ1)2q¯𝑑x𝑑t)2q¯2Eq. 3.14C(ΩTukq¯(1θ)2q¯𝑑x𝑑t)2q¯2.{}\begin{array}[]{rcl}\iint_{\Omega_{T}}|\nabla u_{k}|^{\bar{q}}dxdt=\iint_{\Omega_{T}}\frac{|\nabla u_{k}|^{\bar{q}}u_{k}^{\theta-1}}{u_{k}^{\theta-1}}dxdt&\leq&\left(\iint_{\Omega_{T}}|\nabla u_{k}|^{2}u_{k}^{\theta-1}dxdt\right)^{\frac{\bar{q}}{2}}\times\left(\iint_{\Omega_{T}}\frac{u_{k}^{\theta-1}}{u_{k}^{\frac{2(\theta-1)}{2-\bar{q}}}}dxdt\right)^{\frac{2-\bar{q}}{2}}\\ &\stackrel{{\scriptstyle\lx@cref{creftype~refnum}{eqn25}}}{{\leq}}&C\left(\iint_{\Omega_{T}}u_{k}^{\frac{\bar{q}(1-\theta)}{2-\bar{q}}}dxdt\right)^{\frac{2-\bar{q}}{2}}.\end{array} (3.15)

We now choose q¯\displaystyle\bar{q} to be such that

q¯(1θ)2q¯=σ=m(γ+1)(n+2)n2(m1) i.e. q¯=m(γ+1)(n+2)n+2m(1γ).\frac{\bar{q}(1-\theta)}{2-\bar{q}}=\sigma=\frac{m(\gamma+1)(n+2)}{n-2(m-1)}\text{ i.e. }\bar{q}=\frac{m(\gamma+1)(n+2)}{n+2-m(1-\gamma)}.

We note that q¯<2\displaystyle\bar{q}<2 is equivalent to m<m¯\displaystyle m<\bar{m}; while q¯>1\displaystyle\bar{q}>1 is always fulfilled. Then we get by embedding results and using Eqs. 3.13 and 3.15

0TΩΩ|uk(x,t)uk(y,t)|q¯|xy|n+q¯s𝑑y𝑑x𝑑tC(n,s)ΩT|uk|q¯𝑑x𝑑tC2(ΩTukσ(x,t)𝑑x𝑑t)2q¯2C3,\begin{array}[]{rcl}\int_{0}^{T}\int_{\Omega}\int_{\Omega}\frac{\left|u_{k}(x,t)-u_{k}(y,t)\right|^{\bar{q}}}{|x-y|^{n+\bar{q}s}}dydxdt\leq C(n,s)\iint_{\Omega_{T}}|\nabla u_{k}|^{\bar{q}}dxdt\leq C_{2}\left(\iint_{\Omega_{T}}u_{k}^{\sigma}(x,t)dxdt\right)^{\frac{2-\bar{q}}{2}}\leq C_{3},\end{array}

where C3\displaystyle C_{3} is a positive constant independent of k\displaystyle k. Thus, {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in Lq¯(0,T;W0s,q¯(Ω))Lq¯(0,T;W01,q¯(Ω))Lq¯(0,T;W01,q¯(Ω))\displaystyle L^{\bar{q}}(0,T;W_{0}^{s,\bar{q}}(\Omega))\cap L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega))\equiv L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega)).
Since the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in the reflexive Banach space Lq¯(0,T;W01,q¯(Ω))\displaystyle L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega)), there exist a subsequence of {uk}k\displaystyle\left\{u_{k}\right\}_{k} still indexed by k\displaystyle k and a measurable function uLq¯(0,T;W01,q¯(Ω))\displaystyle u\in L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega)) such that uku\displaystyle u_{k}\rightharpoonup u weakly in Lq¯(0,T;W01,q¯(Ω))\displaystyle L^{\bar{q}}(0,T;W_{0}^{1,\bar{q}}(\Omega)). Also by Fatou’s lemma, we will get uL(0,T;Lγ+1(Ω))\displaystyle u\in L^{\infty}(0,T;L^{\gamma+1}(\Omega)) and uLσ(ΩT)\displaystyle u\in L^{\sigma}(\Omega_{T}), with σ:=m(γ+1)(n+2)n2(m1)\displaystyle\sigma:=\frac{m(\gamma+1)(n+2)}{n-2(m-1)}. As before, using the monotonicity of the sequence {uk}\displaystyle\left\{u_{k}\right\}, we get using Beppo Levi’s theorem that uku\displaystyle u_{k}\rightarrow u strongly in L1(ΩT)\displaystyle L^{1}(\Omega_{T}) and uku\displaystyle u_{k}\rightarrow u a.e in n×(0,T)\displaystyle\mathbb{R}^{n}\times(0,T). By Remark 2.23, this pointwise limit will satisfy u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) in L1\displaystyle L^{1} sense. Then, the rest of the proof will follow from the proof of Theorem 2.27.

Remark 3.2.

We observe that the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in Lr(Ω)\displaystyle L^{r}(\Omega) for every 1rσ\displaystyle 1\leq r\leq\sigma, then following the same lines of the previous proof, we can show that the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in Lq(0,T;W0s,q(Ω))Lq(0,T;W01,q(Ω))Lq(0,T;W01,q(Ω))\displaystyle L^{q}(0,T;W_{0}^{s,q}(\Omega))\cap L^{q}(0,T;W_{0}^{1,q}(\Omega))\equiv L^{q}(0,T;W_{0}^{1,q}(\Omega)) for all 1<qq¯\displaystyle 1<q\leq\bar{q} where 1m<m¯\displaystyle 1\leq m<\bar{m}.

Proof of Theorem 2.35

Let us introduce the following notations: for any measurable function v\displaystyle v we define

Am(v)={(x,t)ΩT:v(x,t)>m}, and for a.e. t(0,T),Amt(v)={xΩ:v(x,t)>m}.\begin{array}[]{c}A_{m}(v)=\left\{(x,t)\in\Omega_{T}:v(x,t)>m\right\},\text{ and for a.e. }t\in(0,T),\quad A_{m}^{t}(v)=\{x\in\Omega:v(x,t)>m\}.\end{array}

We use the idea of the classical proof by D.G. Aronson and J. Serrin, which is to prove a uniform L\displaystyle L^{\infty} bound for uk\displaystyle u_{k} in Ω×(0,τ)\displaystyle\Omega\times(0,\tau), for a positive (small) τ\displaystyle\tau (to be specified later), and then to iterate such an estimate. We consider the approximated problem Eq. 2.10 and take (ukm)+χ(0,t)L2(0,T;H01(Ω))\displaystyle(u_{k}-m)_{+}\chi_{(0,t)}\in L^{2}(0,T;H^{1}_{0}(\Omega)), for m>0,t(0,τ),τ<T\displaystyle m>0,t\in(0,\tau),\tau<T, as a test function to obtain

Ωφm(uk(x,t))dx+0tΩuk(ukm)+dxdθ+120tQ(uk(x,θ)uk(y,θ))((ukm)+(x,θ)(ukm)+(y,θ))|xy|n+2s𝑑x𝑑y𝑑θ0τΩfk(ukm)+(uk+1k)γ𝑑x𝑑θ+Ωφm(uk(x,0))𝑑x,{}\begin{array}[]{c}\int_{\Omega}\varphi_{m}\left(u_{k}(x,t)\right)dx+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla(u_{k}-m)_{+}dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)\left((u_{k}-m)_{+}(x,\theta)-(u_{k}-m)_{+}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\\ \leq\int_{0}^{\tau}\int_{\Omega}\frac{f_{k}(u_{k}-m)_{+}}{(u_{k}+\frac{1}{k})^{\gamma}}dxd\theta+\int_{\Omega}\varphi_{m}\left(u_{k}(x,0)\right)dx,\end{array} (3.16)

where φm(ρ)=0ρ(σm)+𝑑σ=(ρm)+22\displaystyle\varphi_{m}(\rho)=\int_{0}^{\rho}(\sigma-m)_{+}d\sigma=\frac{(\rho-m)_{+}^{2}}{2}. We choose a m>0\displaystyle m>0 large enough such that m>max{u0L(Ω),1}\displaystyle m>\operatorname{max}\{\left\|u_{0}\right\|_{L^{\infty}(\Omega)},1\}, in order to neglect the last term above. Noting that uk(ukm)+=|(ukm)+|2\displaystyle\nabla u_{k}\cdot\nabla(u_{k}-m)_{+}=|\nabla(u_{k}-m)_{+}|^{2}, and

(uk(x,θ)uk(y,θ))((ukm)+(x,θ)(ukm)+(y,θ))((ukm)+(x,θ)(ukm)+(y,θ))2,\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)\left((u_{k}-m)_{+}(x,\theta)-(u_{k}-m)_{+}(y,\theta)\right)\geq\left((u_{k}-m)_{+}(x,\theta)-(u_{k}-m)_{+}(y,\theta)\right)^{2},

we take supremum over t(0,τ]\displaystyle t\in(0,\tau] in Eq. 3.16, to get

(ukm)+L(0,τ;L2(Ω))2+(ukm)+L2(0,τ;X0s(Ω))2+(ukm)+L2(0,τ;H01(Ω))20τAm,ktf(ukm)+𝑑x𝑑t.{}\begin{array}[]{c}\|(u_{k}-m)_{+}\|^{2}_{L^{\infty}(0,\tau;L^{2}(\Omega))}+\|(u_{k}-m)_{+}\|^{2}_{L^{2}(0,\tau;X^{s}_{0}(\Omega))}+\|(u_{k}-m)_{+}\|^{2}_{L^{2}(0,\tau;H^{1}_{0}(\Omega))}\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\int_{0}^{\tau}\int_{A^{t}_{m,k}}f(u_{k}-m)_{+}dxdt.\end{array} (3.17)

Note that, in order to deal with the singularity, we have used the fact that m1\displaystyle m\geq 1 and hence uk>1\displaystyle u_{k}>1 in Am,kt\displaystyle A^{t}_{m,k}, here the subscript k\displaystyle k in Am,kt\displaystyle A^{t}_{m,k} denotes that we are considering the function uk\displaystyle u_{k}. Now the term of the right-hand side above can be estimated as follows,

0τAm,ktf(ukm)+𝑑x𝑑t0τAm,ktf(ukm)+2𝑑x𝑑t+0τAm,ktf𝑑x𝑑t.{}\int_{0}^{\tau}\int_{A_{m,k}^{t}}f(u_{k}-m)_{+}dxdt\leq\int_{0}^{\tau}\int_{A_{m,k}^{t}}f(u_{k}-m)_{+}^{2}dxdt+\int_{0}^{\tau}\int_{A_{m,k}^{t}}fdxdt. (3.18)

We now study each member present in the right-hand side of Eq. 3.18. We first define the followings,

r¯=2r,q¯=2q,η=2η1n,r^=r¯(1+η),q^=q¯(1+η),\bar{r}=2r^{\prime},\bar{q}=2q^{\prime},\eta=\frac{2\eta_{1}}{n},\hat{r}=\bar{r}(1+\eta),\hat{q}=\bar{q}(1+\eta),

where η1=11rn2q\displaystyle\eta_{1}=1-\frac{1}{r}-\frac{n}{2q}. Note that η1(0,1)\displaystyle\eta_{1}\in(0,1) as 0<1r+n2q<1\displaystyle 0<\frac{1}{r}+\frac{n}{2q}<1. Further, simple calculation yields that 1r^+n2q^=n4\displaystyle\frac{1}{\hat{r}}+\frac{n}{2\hat{q}}=\frac{n}{4}. Now, applying Hölder inequality repeatedly, we estimate the first term as

0τAm,ktf(ukm)+2𝑑x𝑑t0τ(Am,kt|f(x,t)|q𝑑x)1q(Am,kt(ukm)+2q𝑑x)1q𝑑t(0τ(Am,kt|f(x,t)|q𝑑x)rq𝑑t)1r(0τ(Am,kt(ukm)+2q𝑑x)rq𝑑t)1r.\begin{array}[]{rcl}\int_{0}^{\tau}\int_{A_{m,k}^{t}}f(u_{k}-m)_{+}^{2}dxdt&\leq&\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}|f(x,t)|^{q}dx\right)^{\frac{1}{q}}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{2q^{\prime}}dx\right)^{\frac{1}{q^{\prime}}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}|f(x,t)|^{q}dx\right)^{\frac{r}{q}}dt\right)^{\frac{1}{r}}\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{2q^{\prime}}dx\right)^{\frac{r^{\prime}}{q^{\prime}}}dt\right)^{\frac{1}{r^{\prime}}}.\end{array}

As again by Hölder inequality with exponent (1+η)\displaystyle(1+\eta), we have

(0τ(Am,kt(ukm)+2q𝑑x)rq𝑑t)1r(0τ(Am,kt(ukm)+2q(1+η)𝑑x)rq(1+η)|Am,kt|rηq(1+η)𝑑t)1r=(0τ(Am,kt(ukm)+q^𝑑x)2rq^|Am,kt|2rηq^𝑑t)1r(0τ(Am,kt(ukm)+q^𝑑x)r^q^𝑑t)1r(1+η)(0τ|Am,kt|r^q^)ηr(1+η),\begin{array}[]{rcl}\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{2q^{\prime}}dx\right)^{\frac{r^{\prime}}{q^{\prime}}}dt\right)^{\frac{1}{r^{\prime}}}&\leq&\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{2q^{\prime}(1+\eta)}dx\right)^{\frac{r^{\prime}}{q^{\prime}(1+\eta)}}|A^{t}_{m,k}|^{\frac{r^{\prime}\eta}{q^{\prime}(1+\eta)}}dt\right)^{\frac{1}{r^{\prime}}}\\ &=&\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{\hat{q}}dx\right)^{\frac{2r^{\prime}}{\hat{q}}}|A^{t}_{m,k}|^{\frac{2r^{\prime}\eta}{\hat{q}}}dt\right)^{\frac{1}{r^{\prime}}}\\ &\leq&\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{\hat{q}}dx\right)^{\frac{\hat{r}}{\hat{q}}}dt\right)^{\frac{1}{r^{\prime}(1+\eta)}}\left(\int_{0}^{\tau}|A^{t}_{m,k}|^{\frac{\hat{r}}{\hat{q}}}\right)^{\frac{\eta}{r^{\prime}(1+\eta)}},\end{array}

therefore

0τAm,ktf(ukm)+2𝑑x𝑑tfLr(0,T;Lq(Ω))(ukm)+Lr^(0,τ;Lq^(Ω))2μk(m)2ηr^,\int_{0}^{\tau}\int_{A_{m,k}^{t}}f(u_{k}-m)_{+}^{2}dxdt\leq\|f\|_{L^{r}(0,T;L^{q}(\Omega))}\|(u_{k}-m)_{+}\|^{2}_{L^{\hat{r}}(0,\tau;L^{\hat{q}}(\Omega))}\mu_{k}(m)^{\frac{2\eta}{\hat{r}}},

where μk(m)=0τ|Am,kt|r^q^𝑑t\displaystyle\mu_{k}(m)=\int_{0}^{\tau}\left|A_{m,k}^{t}\right|^{\frac{\hat{r}}{\hat{q}}}dt. We now use Gagliardo-Nirenberg inequality (Theorem 2.10) for (ukm)+\displaystyle(u_{k}-m)_{+} to get

0τAm,ktf(ukm)+2𝑑x𝑑tcfLr(0,T;Lq(Ω))μk(m)2ηr^(0τ(ukm)+L2(Ω)(1θ)r^(ukm)+L2(Ω)r^θdt)2r^cfμk(m)2ηr^[(ukm)+L(0,τ;L2(Ω))2+(ukm)+L2(0,τ;H01(Ω))2].{}\begin{array}[]{rcl}\int_{0}^{\tau}\int_{A_{m,k}^{t}}f(u_{k}-m)_{+}^{2}dxdt&\leq&c\|f\|_{L^{r}(0,T;L^{q}(\Omega))}\mu_{k}(m)^{\frac{2\eta}{\hat{r}}}\left(\int_{0}^{\tau}\left\|(u_{k}-m)_{+}\right\|_{L^{2}(\Omega)}^{(1-\theta)\hat{r}}\left\|\nabla(u_{k}-m)_{+}\right\|_{L^{2}\left(\Omega\right)}^{\hat{r}\theta}dt\right)^{\frac{2}{\hat{r}}}\\ &\leq&c\|f\|\mu_{k}(m)^{\frac{2\eta}{\hat{r}}}\bigg{[}\left\|(u_{k}-m)_{+}\right\|_{L^{\infty}\left(0,\tau;L^{2}(\Omega)\right)}^{2}+\left\|(u_{k}-m)_{+}\right\|_{L^{2}(0,\tau;H^{1}_{0}(\Omega))}^{2}\bigg{]}.\end{array} (3.19)

where r^θ=2\displaystyle\hat{r}\theta=2, and c\displaystyle c is a constant independent of the choice of k\displaystyle k and m\displaystyle m. Note that we have used the fact that any function in H01(Ω)\displaystyle H^{1}_{0}(\Omega) can be considered as a function in H1(n)\displaystyle H^{1}(\mathbb{R}^{n}). Also, we applied Young’s inequality with exponents r^2\displaystyle\frac{\hat{r}}{2} and its conjugate.
On the other hand, the second term on the right-hand side in Eq. 3.18 can be estimated by Hölder inequality as

0τAm,ktf𝑑x𝑑t0τ(Am,kt|f(x,t)|q𝑑x)1q|Am,kt|1q𝑑t(0τ(Am,kt|f(x,t)|q𝑑x)rq𝑑t)1r(0τ|Am,kt|rq𝑑t)1rfLr(0,T;Lq(Ω))μk(m)2(1+η)r^.{}\begin{array}[]{rcl}\int_{0}^{\tau}\int_{A_{m,k}^{t}}fdxdt&\leq&\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}|f(x,t)|^{q}dx\right)^{\frac{1}{q}}|A^{t}_{m,k}|^{\frac{1}{q^{\prime}}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}|f(x,t)|^{q}dx\right)^{\frac{r}{q}}dt\right)^{\frac{1}{r}}\left(\int_{0}^{\tau}|A^{t}_{m,k}|^{\frac{r^{\prime}}{q^{\prime}}}dt\right)^{\frac{1}{r^{\prime}}}\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\|f\|_{L^{r}(0,T;L^{q}(\Omega))}\mu_{k}(m)^{\frac{2(1+\eta)}{\hat{r}}}.\end{array} (3.20)

Denoting

|(ukm)+|2=(ukm)+L(0,τ;L2(Ω))2+(ukm)+L2(0,τ;H01(Ω))2,|||(u_{k}-m)_{+}|||^{2}=\|(u_{k}-m)_{+}\|_{L^{\infty}(0,\tau;L^{2}(\Omega))}^{2}+\|(u_{k}-m)_{+}\|_{L^{2}(0,\tau;H^{1}_{0}(\Omega))}^{2},

and using Eq. 3.19, Eq. 3.20, we get from Eq. 3.17,

|(ukm)+|2cfLr(0,T;Lq(Ω))[μk(m)2ηr^|(ukm)+|2+μk(m)2(1+η)r^],|||(u_{k}-m)_{+}|||^{2}\leq c\|f\|_{L^{r}(0,T;L^{q}(\Omega))}\left[\mu_{k}(m)^{\frac{2\eta}{\hat{r}}}|||(u_{k}-m)_{+}|||^{2}+\mu_{k}(m)^{\frac{2(1+\eta)}{\hat{r}}}\right],

where c\displaystyle c is a constant which does not depend on k\displaystyle k and m\displaystyle m. Note that μk(m)τ|Ω|r^q^\displaystyle\mu_{k}(m)\leq\tau|\Omega|^{\frac{\hat{r}}{\hat{q}}}, for all k\displaystyle k\in\mathbb{N} and m\displaystyle m, so that we can fix τ\displaystyle\tau, independent of uk\displaystyle u_{k} and m\displaystyle m, suitable small in such a way that cμk(m)2ηr^fLr(0,T;Lq(Ω))=12\displaystyle c\mu_{k}(m)^{\frac{2\eta}{\hat{r}}}\|f\|_{L^{r}(0,T;L^{q}(\Omega))}=\frac{1}{2} and use again the Gagliardo-Nirenberg inequality (see Eq. 3.19), to deduce that

(ukm)+Lr^(0,τ;Lq^(Ω))2c|(ukm)+|2cfLr(0,T;Lq(Ω))μk(m)2(1+η)r^.{}\left\|(u_{k}-m)_{+}\right\|_{L^{\hat{r}}\left(0,\tau;L^{\hat{q}}(\Omega)\right)}^{2}\leq c|||(u_{k}-m)_{+}|||^{2}\leq c\|f\|_{L^{r}(0,T;L^{q}(\Omega))}\mu_{k}(m)^{\frac{2(1+\eta)}{\hat{r}}}. (3.21)

Consider l>m>max{u0L(Ω),1}:=m0\displaystyle l>m>\operatorname{max}\{\left\|u_{0}\right\|_{L^{\infty}(\Omega)},1\}:=m_{0}. Then Al,ktAm,kt\displaystyle A^{t}_{l,k}\subset A^{t}_{m,k}, and using Eq. 3.21, we get

(lm)μk(l)1r^=(0τ((lm)q^|Al,kt|)r^q^𝑑t)1r^(0τ(Al,kt(ukm)+q^𝑑x)r^q^𝑑t)1r^(0τ(Am,kt(ukm)+q^𝑑x)r^q^𝑑t)1r^cμk(m)1+ηr^.\begin{array}[]{rcl}(l-m)\mu_{k}(l)^{\frac{1}{\hat{r}}}=\left(\int_{0}^{\tau}\left((l-m)^{\hat{q}}\left|A_{l,k}^{t}\right|\right)^{\frac{\hat{r}}{\hat{q}}}dt\right)^{\frac{1}{\hat{r}}}&\leq&\left(\int_{0}^{\tau}\left(\int_{A_{l,k}^{t}}(u_{k}-m)_{+}^{\hat{q}}dx\right)^{\frac{\hat{r}}{\hat{q}}}dt\right)^{\frac{1}{\hat{r}}}\\ &\leq&\left(\int_{0}^{\tau}\left(\int_{A_{m,k}^{t}}(u_{k}-m)_{+}^{\hat{q}}dx\right)^{\frac{\hat{r}}{\hat{q}}}dt\right)^{\frac{1}{\hat{r}}}\leq c\mu_{k}(m)^{\frac{1+\eta}{\hat{r}}}.\end{array}

Therefore for all l>m>m0\displaystyle l>m>m_{0}, we have

μk(l)c(lm)r^μk(m)1+η,\mu_{k}(l)\leq\frac{c}{(l-m)^{\hat{r}}}\mu_{k}(m)^{1+\eta},

where c is a constant independent of k\displaystyle k. Now applying [27, Lemma B.1\displaystyle 1], we conclude that μk(m0+dk)=0,\displaystyle\mu_{k}(m_{0}+d_{k})=0, where dk=c[μk(m0)]η2r^(1+η)/η\displaystyle d_{k}=c[\mu_{k}(m_{0})]^{\eta}2^{\hat{r}(1+\eta)/\eta}. As for each k\displaystyle k, dkc(T|Ω|r^q^)η\displaystyle d_{k}\leq c\left(T|\Omega|^{\frac{\hat{r}}{\hat{q}}}\right)^{\eta}, we get

ukL(Ω×[0,τ])d.\left\|u_{k}\right\|_{L^{\infty}(\Omega\times[0,\tau])}\leq d.

We can iterate this procedure in the sets Ω×[τ,2τ],,Ω×[jτ,T]\displaystyle\Omega\times[\tau,2\tau],\cdots,\Omega\times[j\tau,T], where Tjττ\displaystyle T-j\tau\leqslant\tau to conclude that

ukL(ΩT)C uniformly in k.\left\|u_{k}\right\|_{L^{\infty}\left(\Omega_{T}\right)}\leq C\quad\text{ uniformly in }k\in\mathbb{N}.

Now, since the sequence {uk}k\displaystyle\{u_{k}\}_{k} is increasing in k\displaystyle k, we have uL(Ω)C\displaystyle\|u\|_{L^{\infty}(\Omega)}\leq C. Further, if γ1\displaystyle\gamma\leq 1, testing Eq. 2.10 with the function uk\displaystyle u_{k} we can deduce uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H^{1}_{0}(\Omega)) (see Theorem 2.26).

Proof of Theorem 2.36

For γ1\displaystyle\gamma\geq 1, we choose δ>1+γ2\displaystyle\delta>\frac{1+\gamma}{2} and for γ<1\displaystyle\gamma<1, we choose δ>1\displaystyle\delta>1, and take uk2δ1χ(0,t),0<tT\displaystyle u_{k}^{2\delta-1}\chi_{(0,t)},0<t\leq T as test function in Eq. 2.10, with u00\displaystyle u_{0}\equiv 0. We note that, such a test function is admissible as ukL(ΩT)\displaystyle u_{k}\in L^{\infty}(\Omega_{T}). We get tT,\displaystyle\forall t\leq T,

12δΩuk2δ(x,t)𝑑x+0tΩukuk2δ1dxdθ+120tQ(uk2δ1(x,θ)uk2δ1(y,θ))(uk(x,θ)uk(y,θ))|xy|n+2s𝑑x𝑑y𝑑θΩtfuk2δ1γ𝑑x𝑑θ.\begin{array}[]{c}\quad\frac{1}{2\delta}\int_{\Omega}u_{k}^{2\delta}(x,t)dx+\int_{0}^{t}\int_{\Omega}\nabla u_{k}\cdot\nabla u_{k}^{2\delta-1}dxd\theta\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \quad+\frac{1}{2}\int_{0}^{t}\int_{Q}\frac{\left(u_{k}^{2\delta-1}(x,\theta)-u_{k}^{2\delta-1}(y,\theta)\right)\left(u_{k}(x,\theta)-u_{k}(y,\theta)\right)}{|x-y|^{n+2s}}dxdyd\theta\leq\iint_{\Omega_{t}}fu_{k}^{2\delta-1-\gamma}dxd\theta.\end{array}

Taking supremum over t(0,T]\displaystyle t\in(0,T] and using item (i)\displaystyle(i) of Lemma 2.11, by Hölder and Sobolev inequalities, we have

supt[0,T]Ωuk2δ(x,t)𝑑x+2(2δ1)δλ0TukδL2(Ω)2𝑑t+(2δ1)δλs0TukδL2s(Ω)2𝑑t2δfLr(0,T;Lq(Ω))(0T[Ωuk(2δ1γ)q𝑑x]rq𝑑t)1r.{}\begin{array}[]{c}\sup_{t\in[0,T]}\int_{\Omega}u_{k}^{2\delta}(x,t)dx+\frac{2(2\delta-1)}{\delta}\lambda\int_{0}^{T}\left\|u_{k}^{\delta}\right\|_{L^{2^{*}}(\Omega)}^{2}dt+\frac{(2\delta-1)}{\delta}\lambda_{s}\int_{0}^{T}\left\|u_{k}^{\delta}\right\|_{L^{2_{s}^{*}}(\Omega)}^{2}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq 2\delta\left\|f\right\|_{L^{r}\left(0,T;L^{q}(\Omega)\right)}\left(\int_{0}^{T}\left[\int_{\Omega}u_{k}^{(2\delta-1-\gamma)q^{\prime}}dx\right]^{\frac{r^{\prime}}{q^{\prime}}}dt\right)^{\frac{1}{r^{\prime}}}.\end{array} (3.22)

Note that 2δ1δ>1\displaystyle\frac{2\delta-1}{\delta}>1, so we can ignore the constants in left. Denoting by A=(0TukL(2δ1γ)q(Ω)(2δ1γ)r𝑑t)1r,\displaystyle A=\left(\int_{0}^{T}\left\|u_{k}\right\|_{L^{(2\delta-1-\gamma)q^{\prime}}(\Omega)}^{(2\delta-1-\gamma)r^{\prime}}dt\right)^{\frac{1}{r^{\prime}}}, we get from Eq. 3.22,

supt[0,T]Ωuk2δ𝑑x2δfLr(0,T;Lq(Ω))A,{}\sup_{t\in[0,T]}\int_{\Omega}u_{k}^{2\delta}dx\leq 2\delta\left\|f\right\|_{L^{r}\left(0,T;L^{q}(\Omega)\right)}A, (3.23)

and

0T[Ωuk2sδ𝑑x]22s𝑑tc0T[Ωuk2δ𝑑x]22𝑑tcλ2δfLr(0,T;Lq(Ω))A.{}\int_{0}^{T}\left[\int_{\Omega}u_{k}^{2_{s}^{*}\delta}dx\right]^{\frac{2}{2_{s}^{*}}}dt\leq c\int_{0}^{T}\left[\int_{\Omega}u_{k}^{2^{*}\delta}dx\right]^{\frac{2}{2^{*}}}dt\leq\frac{c}{\lambda}2\delta\left\|f\right\|_{L^{r}\left(0,T;L^{q}(\Omega)\right)}A. (3.24)

Case 1: 1<q<nrn+2(r1)\displaystyle 1<q<\frac{nr}{n+2(r-1)} i.e. 1r<nn21q2n2\displaystyle\frac{1}{r}<\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2}
Since, 1<q<nrn+2(r1)\displaystyle 1<q<\frac{nr}{n+2(r-1)} implies q<r\displaystyle q<r and rq<22\displaystyle\frac{r^{\prime}}{q^{\prime}}<\frac{2}{2^{*}}, we apply Hölder inequality with exponent 2q2r\displaystyle\frac{2q^{\prime}}{2^{*}r^{\prime}} to get

AT1r22q[0T(Ωuk(2δ1γ)q𝑑x)22𝑑t]22q.{}A\leq T^{\frac{1}{r^{\prime}}-\frac{2^{*}}{2q^{\prime}}}\left[\int_{0}^{T}\left(\int_{\Omega}u_{k}^{(2\delta-1-\gamma)q^{\prime}}dx\right)^{\frac{2}{2^{*}}}dt\right]^{\frac{2^{*}}{2q^{\prime}}}. (3.25)

Thus from Eq. 3.24, it follows

0T[Ωuk2δ𝑑x]22𝑑t2δcA2δc[0T(Ωuk(2δ1γ)q𝑑x)22𝑑t]22q.{}\int_{0}^{T}\left[\int_{\Omega}u_{k}^{2^{*}\delta}dx\right]^{\frac{2}{2^{*}}}dt\leq 2\delta cA\leq 2\delta c\left[\int_{0}^{T}\left(\int_{\Omega}u_{k}^{(2\delta-1-\gamma)q^{\prime}}dx\right)^{\frac{2}{2^{*}}}dt\right]^{\frac{2^{*}}{2q^{\prime}}}.\vskip 3.0pt plus 1.0pt minus 1.0pt

We now choose 2δ=(2δ1γ)q\displaystyle 2^{*}\delta=(2\delta-1-\gamma)q^{\prime}, that is, δ=12q(n2)(γ+1)(n2q)\displaystyle\delta=\frac{1}{2}\cdot\frac{q(n-2)(\gamma+1)}{(n-2q)}. Note that if γ1\displaystyle\gamma\geq 1, then δ>γ+12\displaystyle\delta>\frac{\gamma+1}{2} if and only if q>1\displaystyle q>1, and for γ<1\displaystyle\gamma<1, δ>1\displaystyle\delta>1 if and only if q>(21γ)\displaystyle q>\left(\frac{2^{*}}{1-\gamma}\right)^{\prime}. Moreover, since q<n2\displaystyle q<\frac{n}{2} it follows that 22q<1\displaystyle\frac{2^{*}}{2q^{\prime}}<1. Thus we get

0T(Ωuk2δ𝑑x)22𝑑t=0T(Ωuk(2δ1γ)q𝑑x)22𝑑tc,{}\int_{0}^{T}\left(\int_{\Omega}u_{k}^{2^{*}\delta}dx\right)^{\frac{2}{2^{*}}}dt=\int_{0}^{T}\left(\int_{\Omega}u_{k}^{\left(2\delta-1-\gamma\right)q^{\prime}}dx\right)^{\frac{2}{2^{*}}}dt\leq c, (3.26)

and therefore from Eq. 3.23, Eq. 3.25 and Eq. 3.26, it follows

ukL(0,T;L2σ(Ω))+ukL2σ(0,T;L2σ(Ω))c, with σ=q(n2)(γ+1)2(n2q)>q2.\left\|u_{k}\right\|_{L^{\infty}\left(0,T;L^{2\sigma}(\Omega)\right)}+\left\|u_{k}\right\|_{L^{2\sigma}\left(0,T;L^{2^{*}\sigma}(\Omega)\right)}\leq c,\text{ with }\sigma=\frac{q(n-2)(\gamma+1)}{2(n-2q)}>\frac{q}{2}.

Case 2: nrn+2(r1)q<n2r\displaystyle\frac{nr}{n+2(r-1)}\leq q<\frac{n}{2}r^{\prime} i.e. 1rnn21q2n2\displaystyle\frac{1}{r}\geq\frac{n}{n-2}\frac{1}{q}-\frac{2}{n-2} and 1r+n2q>1\displaystyle\frac{1}{r}+\frac{n}{2q}>1
In this case we choose δ=12qrn(γ+1)nr2q(r1)\displaystyle\delta=\frac{1}{2}\frac{qrn(\gamma+1)}{nr-2q(r-1)}. Note that for γ1\displaystyle\gamma\geq 1, δ>1+γ2,\displaystyle\delta>\frac{1+\gamma}{2}, if and only if 1r+n2q<1+n2\displaystyle\frac{1}{r}+\frac{n}{2q}<1+\frac{n}{2} which is always satisfied and for γ<1\displaystyle\gamma<1, δ>1\displaystyle\delta>1 if and only if q>2nrnr(γ+1)+4(r1)\displaystyle q>\frac{2nr}{nr(\gamma+1)+4(r-1)} which is satisfied if r>2γ+1\displaystyle r>\frac{2}{\gamma+1}. Now by interpolation, we get that 1(2δ1γ)q=1θ2δ+θ2δ\displaystyle\frac{1}{(2\delta-1-\gamma)q^{\prime}}=\frac{1-\theta}{2\delta}+\frac{\theta}{2^{*}\delta}, where θ=nq(r1)nr(q1)+2q(r1)(0,1]\displaystyle\theta=\frac{nq(r-1)}{nr(q-1)+2q(r-1)}\in(0,1]. Therefore

0TukL(2δ1γ)q(Ω)r(2δ1γ)𝑑tcukL(0,T;L2δ(Ω))2δμ10T(Ωuk2δ𝑑x)22μ2𝑑t,\int_{0}^{T}\left\|u_{k}\right\|_{L^{(2\delta-1-\gamma)q^{\prime}}(\Omega)}^{r^{\prime}(2\delta-1-\gamma)}dt\leq c\left\|u_{k}\right\|_{L^{\infty}\left(0,T;L^{2\delta}(\Omega)\right)}^{2\delta\mu_{1}}\int_{0}^{T}\left(\int_{\Omega}u_{k}^{2^{*}\delta}dx\right)^{\frac{2}{2^{*}}\mu_{2}}dt,

where μ1=(1θ)r(2δ1γ)2δ\displaystyle\mu_{1}=\frac{(1-\theta)r^{\prime}(2\delta-1-\gamma)}{2\delta} and μ2=θr(2δ1γ)2δ\displaystyle\mu_{2}=\frac{\theta r^{\prime}(2\delta-1-\gamma)}{2\delta}. Since μ21\displaystyle\mu_{2}\leq 1, we have that

0TukL(2δ1γ)q(Ω)r(2δ1γ)𝑑tcukL(0,T;L2δ(Ω))2δμ1(0T[Ωuk2δ𝑑x]22𝑑t)μ2,\int_{0}^{T}\left\|u_{k}\right\|_{L^{(2\delta-1-\gamma)q^{\prime}}(\Omega)}^{r^{\prime}(2\delta-1-\gamma)}dt\leq c\left\|u_{k}\right\|_{L^{\infty}\left(0,T;L^{2\delta}(\Omega)\right)}^{2\delta\mu_{1}}\left(\int_{0}^{T}\left[\int_{\Omega}u_{k}^{2^{*}\delta}dx\right]^{\frac{2}{2^{*}}}dt\right)^{\mu_{2}},

and since μ1+μ2<r\displaystyle\mu_{1}+\mu_{2}<r^{\prime}, using Eqs. 3.23 and 3.24, we can conclude the following which will give our final result

0T(Ωuk(2δ1δ)q𝑑x)rq𝑑tc.\int_{0}^{T}\left(\int_{\Omega}u_{k}^{(2\delta-1-\delta)q^{\prime}}dx\right)^{\frac{r^{\prime}}{q^{\prime}}}dt\leq c.

Proof of Theorem 2.39

Consider the approximated problem Eq. 2.11 and corresponding properties of {uk}k\displaystyle\{u_{k}\}_{k}, Theorem 2.22. We denote (ωT~)δ=ΩT\(ΩT)δ\displaystyle(\omega_{\tilde{T}})_{\delta}=\Omega_{T}\backslash(\Omega_{T})_{\delta}, then uk(x)C((ωT~)δ,n,s)>0\displaystyle u_{k}(x)\geq C((\omega_{\tilde{T}})_{\delta},n,s)>0 in (ωT~)δ\displaystyle(\omega_{\tilde{T}})_{\delta} for all k\displaystyle k. Choosing ukχ(0,τ)(t)\displaystyle u_{k}\chi_{(0,\tau)}(t) as test function in Eq. 2.11, we obtain

12Ωuk2(x,τ)𝑑x+0τΩ|uk|2𝑑x𝑑t+120τQ(uk(x,t)uk(y,t))2|xy|n+2s𝑑x𝑑y𝑑tΩTfkuk(uk+1k)γ(x,t)𝑑x𝑑t+12Ωu02(x)𝑑x.\begin{array}[]{l}\quad\frac{1}{2}\int_{\Omega}u_{k}^{2}(x,\tau)dx+\int_{0}^{\tau}\int_{\Omega}|\nabla u_{k}|^{2}dxdt+\frac{1}{2}\int_{0}^{\tau}\int_{Q}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)^{2}}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{T}}\frac{f_{k}u_{k}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\frac{1}{2}\int_{\Omega}u_{0}^{2}(x)dx.\end{array}

Now taking supremum over τ(0,T]\displaystyle\tau\in(0,T], we get

sup0τTΩuk2(x,τ)𝑑x+20TΩ|uk|2𝑑x𝑑t+0TQ(uk(x,t)uk(y,t))2|xy|n+2s𝑑x𝑑y𝑑t2ΩTfkuk(uk+1k)γ(x,t)𝑑x𝑑t+u0L2(Ω)=2(ΩT)δfkuk(uk+1k)γ(x,t)𝑑x𝑑t+2(ωT~)δfkuk(uk+1k)γ(x,t)𝑑x𝑑t+u0L2(Ω)2(ΩT)δfuk1γ(x,t)𝑑x𝑑t+2(ωT~)δfukC(ωT~)δγ(x,t)𝑑x𝑑t+u0L2(Ω)2(ΩT)δ{uk1}fuk1γ(x,t)𝑑x𝑑t+2(ΩT)δ{uk1}fuk1γ(x,t)𝑑x𝑑t+2(ωT~)δfukC(ωT~)δγ(x,t)𝑑x𝑑t+u0L2(Ω)u0L2(Ω)+2fL1(ΩT)+2(1+C(ωT~)δγ()L(ΩT))ΩTfuk𝑑x𝑑t.{}\begin{array}[]{l}\quad\sup_{0\leq\tau\leq T}\int_{\Omega}u_{k}^{2}(x,\tau)dx+2\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt+\int_{0}^{T}\int_{Q}\frac{\left(u_{k}(x,t)-u_{k}(y,t)\right)^{2}}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq 2\iint_{\Omega_{T}}\frac{f_{k}u_{k}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\vskip 3.0pt plus 1.0pt minus 1.0pt\\ =2\iint_{(\Omega_{T})_{\delta}}\frac{f_{k}u_{k}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+2\iint_{(\omega_{\tilde{T}})_{\delta}}\frac{f_{k}u_{k}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq 2\iint_{(\Omega_{T})_{\delta}}fu_{k}^{1-\gamma(x,t)}dxdt+2\iint_{(\omega_{\tilde{T}})_{\delta}}\frac{fu_{k}}{C_{(\omega_{\tilde{T}})_{\delta}}^{\gamma(x,t)}}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq 2\iint_{(\Omega_{T})_{\delta}\cap\left\{u_{k}\leq 1\right\}}fu_{k}^{1-\gamma(x,t)}dxdt+2\iint_{(\Omega_{T})_{\delta}\cap\left\{u_{k}\geq 1\right\}}fu_{k}^{1-\gamma(x,t)}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \quad+2\iint_{(\omega_{\tilde{T}})_{\delta}}\frac{fu_{k}}{C_{(\omega_{\tilde{T}})_{\delta}}^{\gamma(x,t)}}dxdt+\left\|u_{0}\right\|_{L^{2}(\Omega)}\\ \leq\left\|u_{0}\right\|_{L^{2}(\Omega)}+2\|f\|_{L^{1}(\Omega_{T})}+2\left(1+\left\|C_{(\omega_{\tilde{T}})_{\delta}}^{-\gamma(\cdot)}\right\|_{L^{\infty}(\Omega_{T})}\right)\iint_{\Omega_{T}}fu_{k}dxdt.\end{array} (3.27)

Case 1: fL2(0,T;L(2nn+2)(Ω))\displaystyle f\in L^{2}\left(0,T;L^{\left(\frac{2n}{n+2}\right)}(\Omega)\right)
We note that (2)=2nn+2\displaystyle(2^{*})^{\prime}=\frac{2n}{n+2}, then by using Hölder’s and Sobolev’s inequalities, we have

0TΩfuk𝑑x𝑑tC0T(Ω|f(x,t)|(2)𝑑x)1(2)(Ω|uk(x,t)|2𝑑x)12𝑑tC0TfL(2)(Ω)(Ω|uk|2𝑑x)12𝑑tC(0TfL(2)(Ω)2𝑑t)12(0TΩ|uk|2𝑑x𝑑t)12.\begin{array}[]{rcl}\int_{0}^{T}\int_{\Omega}fu_{k}dxdt&\leq&C\int_{0}^{T}\left(\int_{\Omega}|f(x,t)|^{(2^{*})^{\prime}}dx\right)^{\frac{1}{(2^{*})^{\prime}}}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{2^{*}}dx\right)^{\frac{1}{2^{*}}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&C\int_{0}^{T}\|f\|_{L^{\left({2^{*}}\right)^{\prime}}{(\Omega)}}\left(\int_{\Omega}\left|\nabla u_{k}\right|^{2}dx\right)^{\frac{1}{2}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&C\left(\int_{0}^{T}\|f\|_{L^{\left({2^{*}}\right)^{\prime}}{(\Omega)}}^{2}dt\right)^{\frac{1}{2}}\left(\int_{0}^{T}\int_{\Omega}\left|\nabla u_{k}\right|^{2}dxdt\right)^{\frac{1}{2}}.\end{array}

In the above, we use Young’s inequality, and then from Eq. 3.27 we get that

sup0τTΩuk2(x,τ)𝑑x+ukL2(0,T;H01(Ω))2u0L2(Ω)+2fL1(ΩT)+C(1+C(ωT~)δγ()L(ΩT)).\begin{array}[]{l}\quad\sup_{0\leq\tau\leq T}\int_{\Omega}u_{k}^{2}(x,\tau)dx+\left\|u_{k}\right\|_{L^{2}(0,T;H_{0}^{1}(\Omega))}^{2}\leq\left\|u_{0}\right\|_{L^{2}(\Omega)}+2\|f\|_{L^{1}(\Omega_{T})}+C\left(1+\left\|C_{(\omega_{\tilde{T}})_{\delta}}^{-\gamma(\cdot)}\right\|_{L^{\infty}(\Omega_{T})}\right).\end{array}

Therefore the sequence {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)).
Case 2: fLr¯(ΩT)\displaystyle f\in L^{\bar{r}}\left(\Omega_{T}\right)
For this case, we apply Hölder inequality with exponents r¯\displaystyle\bar{r} and r¯\displaystyle\bar{r}^{\prime} to get

0TΩfuk𝑑x𝑑tCfLr¯(ΩT)[ΩT|uk|r¯𝑑x𝑑t]1r¯.{}\begin{array}[]{rcl}\int_{0}^{T}\int_{\Omega}fu_{k}dxdt\leq C\|f\|_{L^{\bar{r}}\left(\Omega_{T}\right)}\left[\iint_{\Omega_{T}}|u_{k}|^{\bar{r}^{\prime}}dxdt\right]^{\frac{1}{\bar{r}^{\prime}}}.\end{array} (3.28)

We observe that r¯=2(n+2)n\displaystyle\bar{r}^{\prime}=\frac{2(n+2)}{n} and hence using the Hölder inequality with exponents nn2\displaystyle\frac{n}{n-2} and n2\displaystyle\frac{n}{2} and by Sobolev embedding (Theorem 2.8), we can write

ΩT|uk|2(n+2)n𝑑x𝑑t=ΩT|uk|2|uk|4n𝑑x𝑑t0T[Ω|uk(x,t)|2𝑑x]2nukL2(Ω)2𝑑tC(n)[sup0τTΩ|uk(x,τ)|2𝑑x]2n0TΩ|uk|2𝑑x𝑑t.{}\begin{array}[]{rcl}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{2(n+2)}{n}}dxdt=\iint_{\Omega_{T}}|u_{k}|^{2}\left|u_{k}\right|^{\frac{4}{n}}dxdt&\leq&\int_{0}^{T}\left[\int_{\Omega}\left|u_{k}(x,t)\right|^{2}dx\right]^{\frac{2}{n}}\left\|u_{k}\right\|_{L^{2^{*}}(\Omega)}^{2}dt\\ &\leq&C(n)\left[\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{2}dx\right]^{\frac{2}{n}}\int_{0}^{T}\int_{\Omega}|\nabla u_{k}|^{2}dxdt.\end{array} (3.29)

So using Eq. 3.28 in Eq. 3.27, we get from Eq. 3.29 by convexity argument

ΩT|uk|2(n+2)n𝑑x𝑑tC(n)22n((C1CfLr¯(ΩT))n+2n(ΩT|uk|r¯)n+2nr¯+C2n+2n),\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{2(n+2)}{n}}dxdt\leq C(n)2^{\frac{2}{n}}\left(\left(C_{1}C\|f\|_{L^{\bar{r}}\left(\Omega_{T}\right)}\right)^{\frac{n+2}{n}}\left(\iint_{\Omega_{T}}|u_{k}|^{{\bar{r}^{\prime}}}\right)^{\frac{n+2}{n\bar{r}^{\prime}}}+C_{2}^{\frac{n+2}{n}}\right),

where C1=2(1+C(ωT~)δγ()L(ΩT))\displaystyle C_{1}=2\left(1+\left\|C_{(\omega_{\tilde{T}})_{\delta}}^{-\gamma(\cdot)}\right\|_{L^{\infty}(\Omega_{T})}\right) and C2=u0L2(Ω)+2fL1(ΩT)\displaystyle C_{2}=\left\|u_{0}\right\|_{L^{2}(\Omega)}+2\|f\|_{L^{1}(\Omega_{T})}. Now since n+2nr¯=12\displaystyle\frac{n+2}{n\bar{r}^{\prime}}=\frac{1}{2}, we use Young’s inequality to obtain

ΩT|uk|r¯𝑑x𝑑t=ΩT|uk|2(n+2)n𝑑x𝑑tC,\iint_{\Omega_{T}}\left|u_{k}\right|^{\bar{r}^{\prime}}dxdt=\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{2(n+2)}{n}}dxdt\leq C,

where C\displaystyle C is a positive constant independent of k\displaystyle k. Therefore, by Eqs. 3.27 and 3.28 we deduce that {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)).
Since {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in the reflexive Banach space L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)), there exist a subsequence of {uk}k\displaystyle\left\{u_{k}\right\}_{k}, still indexed by k\displaystyle k, and a measurable function uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H_{0}^{1}(\Omega)) such that uku\displaystyle u_{k}\rightharpoonup u weakly in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)). Also, since {uk}\displaystyle\left\{u_{k}\right\} is increasing in k\displaystyle k, it holds by Beppo Levi’s theorem that uku\displaystyle u_{k}\rightarrow u strongly in L1(ΩT)\displaystyle L^{1}\left(\Omega_{T}\right) and hence a.e. in ΩT\displaystyle\Omega_{T}. Applying Fatou’s Lemma, we get uL(0,T;L2(Ω))\displaystyle u\in L^{\infty}(0,T;L^{2}(\Omega)). This pointwise limit is actually defined for each t[0,T)\displaystyle t\in[0,T) and satisfies u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) in L1\displaystyle L^{1} sense (see Remark 2.23). We show that this u\displaystyle u is a very weak solution to Eq. 2.13 in the sense of Definition 2.25. Choosing arbitrary ϕ𝒞0(ΩT)\displaystyle\phi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}), we have

ΩT((uk)tΔuk+(Δ)suk)ϕ𝑑x𝑑t=ΩTfkϕ(uk+1k)γ(x,t)𝑑x𝑑t.\iint_{\Omega_{T}}(\left(u_{k}\right)_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k})\phi\,dxdt=\iint_{\Omega_{T}}\frac{f_{k}\phi}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt.

Since ϕ𝒞0(ΩT)\displaystyle\phi\in\mathcal{C}_{0}^{\infty}(\Omega_{T}), therefore ϕt,Δϕ\displaystyle\phi_{t},\Delta\phi and (Δ)sϕ\displaystyle(-\Delta)^{s}\phi all are bounded. As uku\displaystyle u_{k}\rightarrow u strongly in L1(ΩT)\displaystyle L^{1}\left(\Omega_{T}\right), we have

ΩT((uk)tΔuk+(Δ)suk)ϕ𝑑x𝑑t=ΩTuk((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑tΩTu((ϕ)tΔϕ+(Δ)sϕ)𝑑x𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}(\left(u_{k}\right)_{t}-\Delta u_{k}+(-\Delta)^{s}u_{k})\phi\,dxdt&=&\iint_{\Omega_{T}}u_{k}(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi)dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\rightarrow&\iint_{\Omega_{T}}u(-(\phi)_{t}-\Delta\phi+(-\Delta)^{s}\phi)dxdt.\end{array}

Now since for all ωΩ\displaystyle\omega\subset\subset\Omega and for all t0>0,\displaystyle t_{0}>0, uk(x,t)C(ω,t0,n,s)\displaystyle u_{k}(x,t)\geq C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right), we reach that uk(x,t)C\displaystyle u_{k}(x,t)\geq C in Suppϕ\displaystyle\operatorname{Supp}\phi (say ω×[t1,t2]\displaystyle\omega\times[t_{1},t_{2}]) for all k1\displaystyle k\geq 1. Therefore

0|fkϕ(uk+1k)γ(x,t)|ϕCω,t1,t2γ(x,t)L(ΩT)f.0\leq\left|\frac{f_{k}\phi}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}\right|\leq\|\phi C_{\omega,t_{1},t_{2}}^{-\gamma(x,t)}\|_{L^{\infty}(\Omega_{T})}f.

Hence by the dominated convergence theorem

ΩTfkϕ(uk+1k)γ(x,t)𝑑x𝑑tΩTfϕuγ(x,t)𝑑x𝑑t as k.\iint_{\Omega_{T}}\frac{f_{k}\phi}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt\rightarrow\iint_{\Omega_{T}}\frac{f\phi}{u^{\gamma(x,t)}}dxdt\text{ as }k\rightarrow\infty.

Thus, we get the following and conclude

ΩT(utΔu+(Δ)su)ϕ𝑑x𝑑t=ΩTfϕuγ(x,t)𝑑x𝑑t.\iint_{\Omega_{T}}(u_{t}-\Delta u+(-\Delta)^{s}u)\phi\,dxdt=\iint_{\Omega_{T}}\frac{f\phi}{u^{\gamma(x,t)}}dxdt.

Proof of Theorem 2.41

Consider uk\displaystyle u_{k} to be the unique nonnegative solution to Eq. 2.11. Since γ>1\displaystyle\gamma^{*}>1, so we can use ukγχ(0,τ)(t)\displaystyle u_{k}^{\gamma^{*}}\chi_{(0,\tau)}(t) as a test function in Eq. 2.11 to get for all τT\displaystyle\tau\leq T,

1γ+1Ωukγ+1(x,τ)𝑑x+0τΩukukγdxdt+120τQ(ukγ(x,t)ukγ(y,t))(uk(x,t)uk(y,t))|xy|n+2s𝑑x𝑑y𝑑tΩTfkukγ(uk+1k)γ(x,t)𝑑x𝑑t+1γ+1Ωu0γ+1(x)𝑑x.\begin{array}[]{c}\frac{1}{\gamma^{*}+1}\int_{\Omega}u_{k}^{\gamma^{*}+1}(x,\tau)dx+\int_{0}^{\tau}\int_{\Omega}\nabla u_{k}\cdot\nabla u_{k}^{\gamma^{*}}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ +\frac{1}{2}\int_{0}^{\tau}\int_{Q}\frac{(u_{k}^{\gamma^{*}}(x,t)-u_{k}^{\gamma^{*}}(y,t))\left(u_{k}(x,t)-u_{k}(y,t)\right)}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\iint_{\Omega_{T}}\frac{f_{k}u_{k}^{\gamma^{*}}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\frac{1}{\gamma^{*}+1}\int_{\Omega}u_{0}^{\gamma^{*}+1}(x)dx.\end{array}

Now taking supremum over τ(0,T]\displaystyle\tau\in(0,T] and using item i)\displaystyle i) of Lemma 2.11, we get

(γ+1)2γsup0τTΩ|uk(x,τ)|γ+1𝑑x+(γ+1)220TΩukγ1|uk|2𝑑x𝑑t+0TQ|ukγ+12(x,t)ukγ+12(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t(γ+1)22γ(ΩTfkukγ(uk+1k)γ(x,t)𝑑x𝑑t+u0Lγ+1(Ω)).\begin{array}[]{c}\quad\frac{({\gamma^{*}}+1)}{2\gamma^{*}}\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{{\gamma^{*}}+1}dx+\frac{({\gamma^{*}}+1)^{2}}{2}\int_{0}^{T}\int_{\Omega}u_{k}^{{\gamma^{*}}-1}|\nabla u_{k}|^{2}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \quad+\int_{0}^{T}\int_{Q}\frac{|u_{k}^{\frac{{\gamma^{*}}+1}{2}}(x,t)-u_{k}^{\frac{{\gamma^{*}}+1}{2}}(y,t)|^{2}}{|x-y|^{n+2s}}dxdydt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq\frac{({\gamma^{*}}+1)^{2}}{2\gamma^{*}}\left(\iint_{\Omega_{T}}\frac{f_{k}u_{k}^{\gamma^{*}}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\left\|u_{0}\right\|_{L^{\gamma^{*}+1}(\Omega)}\right).\end{array}

We note that 1<γ+1γ<2\displaystyle 1<\frac{\gamma^{*}+1}{\gamma^{*}}<2 and hence it follows

sup0τTΩ|uk(x,τ)|γ+1𝑑x+40TΩ|ukγ+12|2𝑑x𝑑t+20TQ|ukγ+12(x,t)ukγ+12(y,t)|2|xy|n+2s𝑑x𝑑y𝑑t4γ(ΩTfkukγ(uk+1k)γ(x,t)𝑑x𝑑t+u0Lγ+1(Ω)).{}\begin{array}[]{c}\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{{\gamma^{*}}+1}dx+4\int_{0}^{T}\int_{\Omega}|\nabla u_{k}^{\frac{{\gamma^{*}}+1}{2}}|^{2}dxdt\\ +2\int_{0}^{T}\int_{Q}\frac{|u_{k}^{\frac{{\gamma^{*}}+1}{2}}(x,t)-u_{k}^{\frac{{\gamma^{*}}+1}{2}}(y,t)|^{2}}{|x-y|^{n+2s}}dxdydt\leq 4\gamma^{*}\left(\iint_{\Omega_{T}}\frac{f_{k}u_{k}^{\gamma^{*}}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\left\|u_{0}\right\|_{L^{\gamma^{*}+1}(\Omega)}\right).\end{array} (3.30)

Denoting (ωT~)δ=ΩT\(ΩT)δ\displaystyle(\omega_{\tilde{T}})_{\delta}=\Omega_{T}\backslash(\Omega_{T})_{\delta}, we have uk(x,t)C((ωT~)δ,n,s)>0\displaystyle u_{k}(x,t)\geq C((\omega_{\tilde{T}})_{\delta},n,s)>0 in (ωT~)δ\displaystyle(\omega_{\tilde{T}})_{\delta} for all k\displaystyle k. We estimate like Theorem 2.39 as

ΩTfkukγ(uk+1k)γ(x,t)𝑑x𝑑t=(ΩT)δfkukγ(uk+1k)γ(x,t)𝑑x𝑑t+(ωT~)δfkukγ(uk+1k)γ(x,t)𝑑x𝑑t(ΩT)δfukγγ(x,t)𝑑x𝑑t+(ωT~)δfukγC(ωT~)δγ(x,t)𝑑x𝑑t(ΩT)δ{uk1}fukγγ(x,t)𝑑x𝑑t+(ΩT)δ{uk1}fukγγ(x,t)𝑑x𝑑t+(ωT~)δfukγC(ωT~)δγ(x,t)𝑑x𝑑tfL1(ΩT)+(1+C(ωT~)δγ()L(ΩT))ΩTfukγ𝑑x𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}\frac{f_{k}u_{k}^{\gamma^{*}}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt&=&\iint_{(\Omega_{T})_{\delta}}\frac{f_{k}u_{k}^{\gamma^{*}}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt+\iint_{(\omega_{\tilde{T}})_{\delta}}\frac{f_{k}u_{k}^{\gamma^{*}}}{\left(u_{k}+\frac{1}{k}\right)^{\gamma(x,t)}}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\iint_{(\Omega_{T})_{\delta}}fu_{k}^{{\gamma^{*}}-\gamma(x,t)}dxdt+\iint_{(\omega_{\tilde{T}})_{\delta}}\frac{fu_{k}^{\gamma^{*}}}{C_{(\omega_{\tilde{T}})_{\delta}}^{\gamma(x,t)}}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\iint_{(\Omega_{T})_{\delta}\cap\left\{u_{k}\leq 1\right\}}fu_{k}^{{\gamma^{*}}-\gamma(x,t)}dxdt+\iint_{(\Omega_{T})_{\delta}\cap\left\{u_{k}\geq 1\right\}}fu_{k}^{\gamma^{*}-\gamma(x,t)}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &&+\iint_{(\omega_{\tilde{T}})_{\delta}}\frac{fu_{k}^{\gamma^{*}}}{C_{(\omega_{\tilde{T}})_{\delta}}^{\gamma(x,t)}}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\|f\|_{L^{1}(\Omega_{T})}+\left(1+\left\|C_{(\omega_{\tilde{T}})_{\delta}}^{-\gamma(\cdot)}\right\|_{L^{\infty}(\Omega_{T})}\right)\iint_{\Omega_{T}}fu_{k}^{\gamma^{*}}dxdt.\end{array}

Using the above in Eq. 3.30 we get

sup0τTΩ|uk(x,τ)|γ+1𝑑x+20TΩ|ukγ+12|2𝑑x𝑑t4γ(1+C(ωT~)δγ()L(ΩT))ΩTfukγ𝑑x𝑑t+4γ(fL1(ΩT)+u0Lγ+1(Ω)).{}\begin{array}[]{l}\quad\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{{\gamma^{*}}+1}dx+2\int_{0}^{T}\int_{\Omega}|\nabla u_{k}^{\frac{{\gamma^{*}}+1}{2}}|^{2}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq 4\gamma^{*}\left(1+\left\|C_{(\omega_{\tilde{T}})_{\delta}}^{-\gamma(\cdot)}\right\|_{L^{\infty}(\Omega_{T})}\right)\iint_{\Omega_{T}}fu_{k}^{\gamma^{*}}dxdt+4\gamma^{*}\left(\|f\|_{L^{1}(\Omega_{T})}+\left\|u_{0}\right\|_{L^{\gamma^{*}+1}(\Omega)}\right).\end{array} (3.31)

For convenience we take C1=4γ(1+C(ωT~)δγ()L(ΩT))\displaystyle C_{1}=4\gamma^{*}\left(1+\left\|C_{(\omega_{\tilde{T}})_{\delta}}^{-\gamma(\cdot)}\right\|_{L^{\infty}(\Omega_{T})}\right) and C2=4γ(fL1(ΩT)+u0Lγ+1(Ω))\displaystyle C_{2}=4\gamma^{*}\left(\|f\|_{L^{1}(\Omega_{T})}+\left\|u_{0}\right\|_{L^{\gamma^{*}+1}(\Omega)}\right).
Case 1: fLγ+1(0,T;L(n(γ+1)n+2γ)(Ω))\displaystyle f\in L^{\gamma^{*}+1}\left(0,T;L^{\left(\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}\right)}(\Omega)\right)
In this case using Hölder inequality first in the space variable and then by Sobolev inequality and another application of Hölder inequality (in time variable) we get

0TΩfukγ𝑑x𝑑t=0TΩf(ukγ+12)2γγ+1𝑑x𝑑tC0T(Ω|f(x,t)|n(γ+1)n+2γ𝑑x)n+2γn(γ+1)(Ω|ukγ+12(x,t)|2𝑑x)(n2)γn(γ+1)𝑑tC0Tf(,t)Ln(γ+1)n+2γ(Ω)(Ω|ukγ+12|2𝑑x)γγ+1𝑑tC(0Tf(,t)Ln(γ+1)n+2γ(Ω)γ+1𝑑t)1γ+1×(0TΩ|ukγ+12|2𝑑x𝑑t)γγ+1.\begin{array}[]{rcl}\int_{0}^{T}\int_{\Omega}fu_{k}^{\gamma^{*}}dxdt&=&\int_{0}^{T}\int_{\Omega}f\left(u_{k}^{\frac{\gamma^{*}+1}{2}}\right)^{\frac{2\gamma^{*}}{\gamma^{*}+1}}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&C\int_{0}^{T}\left(\int_{\Omega}|f(x,t)|^{\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}}dx\right)^{\frac{n+2\gamma^{*}}{n(\gamma^{*}+1)}}\left(\int_{\Omega}|u_{k}^{\frac{\gamma^{*}+1}{2}}(x,t)|^{2^{*}}dx\right)^{\frac{(n-2)\gamma^{*}}{n(\gamma^{*}+1)}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&C\int_{0}^{T}\|f(\cdot,t)\|_{L^{\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}}{(\Omega)}}\left(\int_{\Omega}|\nabla u_{k}^{\frac{\gamma^{*}+1}{2}}|^{2}dx\right)^{\frac{\gamma^{*}}{\gamma^{*}+1}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&C\left(\int_{0}^{T}\|f(\cdot,t)\|_{L^{\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}}{(\Omega)}}^{\gamma^{*}+1}dt\right)^{\frac{1}{\gamma^{*}+1}}\times\left(\int_{0}^{T}\int_{\Omega}|\nabla u_{k}^{\frac{\gamma^{*}+1}{2}}|^{2}dxdt\right)^{\frac{\gamma^{*}}{\gamma^{*}+1}}.\end{array}

Now since γγ+1<1\displaystyle\frac{\gamma^{*}}{\gamma^{*}+1}<1, so using Young’s inequality in the above estimate, we get from Eq. 3.31 that

sup0τTΩ|uk(x,τ)|γ+1𝑑x+20TΩ|ukγ+12|2𝑑x𝑑tCfLγ+1(0,T;Ln(γ+1)n+2γ(Ω))×(ε0TΩ|ukγ+12|2𝑑x𝑑t+C(ε))+C2.\begin{array}[]{l}\quad\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{{\gamma^{*}}+1}dx+2\int_{0}^{T}\int_{\Omega}|\nabla u_{k}^{\frac{{\gamma^{*}}+1}{2}}|^{2}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ \leq C\left\|f\right\|_{L^{\gamma^{*}+1}\left(0,T;L^{\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}}(\Omega)\right)}\times\left(\varepsilon\int_{0}^{T}\int_{\Omega}|\nabla u_{k}^{\frac{{\gamma^{*}}+1}{2}}|^{2}dxdt+C(\varepsilon)\right)+C_{2}.\end{array}

Choosing ε\displaystyle\varepsilon small enough such that εCfLγ+1(0,T;Ln(γ+1)n+2γ(Ω))=1,\displaystyle\varepsilon C\left\|f\right\|_{L^{\gamma^{*}+1}\left(0,T;L^{\frac{n(\gamma^{*}+1)}{n+2\gamma^{*}}}(\Omega)\right)}=1, we get

sup0τTΩ|uk(x,τ)|γ+1𝑑x+0TΩ|ukγ+12|2𝑑x𝑑tC,\sup_{0\leq\tau\leq T}\int_{\Omega}\left|u_{k}(x,\tau)\right|^{{\gamma^{*}}+1}dx+\int_{0}^{T}\int_{\Omega}|\nabla u_{k}^{\frac{{\gamma^{*}}+1}{2}}|^{2}dxdt\leq C,

which implies that the sequence {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma^{*}+1}{2}}\right\}_{k} is uniformly bounded in L2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)) and {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma^{*}+1}(\Omega)).
Case 2: fLr~(ΩT)\displaystyle f\in L^{\tilde{r}}\left(\Omega_{T}\right)
For this case, we apply Hölder inequality with exponents r~\displaystyle\tilde{r} and r~\displaystyle\tilde{r}^{\prime} to get

0TΩfukγ𝑑x𝑑tCfLr~(ΩT)[ΩT|uk|γr~𝑑x𝑑t]1r~.{}\begin{array}[]{rcl}\int_{0}^{T}\int_{\Omega}fu_{k}^{\gamma^{*}}dxdt\leq C\|f\|_{L^{\tilde{r}}\left(\Omega_{T}\right)}\left[\iint_{\Omega_{T}}|u_{k}|^{\gamma^{*}\tilde{r}^{\prime}}dxdt\right]^{\frac{1}{\tilde{r}^{\prime}}}.\end{array} (3.32)

We note that γr~=(n+2)(γ+1)n\displaystyle\gamma^{*}\tilde{r}^{\prime}=\frac{(n+2)(\gamma^{*}+1)}{n}. Again, we use the same technique as that of Theorem 2.30, Theorem 2.32 and Theorem 2.39. Using the Hölder inequality with exponent nn2\displaystyle\frac{n}{n-2} and n2\displaystyle\frac{n}{2} we get

ΩT|uk|(γ+1)(n+2)n𝑑x𝑑t=ΩT|uk|2γ+12|uk|4nγ+12𝑑x𝑑t0T(Ω|uk(x,t)|2γ+12𝑑x)n2n(Ω|uk(x,t)|γ+1𝑑x)2n𝑑t=0Tukγ+12L2(Ω)2(Ω|uk(x,t)|γ+1𝑑x)2n𝑑t.\begin{array}[]{rcl}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{({\gamma^{*}}+1)(n+2)}{n}}dxdt&=&\iint_{\Omega_{T}}\left|u_{k}\right|^{2\frac{{\gamma^{*}}+1}{2}}\left|u_{k}\right|^{\frac{4}{n}\frac{{\gamma^{*}}+1}{2}}dxdt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &\leq&\int_{0}^{T}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{2^{*}\frac{{\gamma^{*}}+1}{2}}dx\right)^{\frac{n-2}{n}}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{{\gamma^{*}}+1}dx\right)^{\frac{2}{n}}dt\vskip 3.0pt plus 1.0pt minus 1.0pt\\ &=&\int_{0}^{T}\|u_{k}^{\frac{{\gamma^{*}}+1}{2}}\|_{L^{2^{*}}(\Omega)}^{2}\left(\int_{\Omega}\left|u_{k}(x,t)\right|^{{\gamma^{*}}+1}dx\right)^{\frac{2}{n}}dt.\end{array}

Now taking supremum over t[0,T]\displaystyle t\in[0,T] and applying the Sobolev embedding as that of Theorem 2.8, we can write

ΩT|uk|(γ+1)(n+2)n𝑑x𝑑tC(n)(sup0tTΩ|uk(x,t)|γ+1𝑑x)2n(ΩT|ukγ+12|2𝑑x𝑑t).\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{({\gamma^{*}}+1)(n+2)}{n}}dxdt\leq C(n)\left(\sup_{0\leq t\leq T}\int_{\Omega}\left|u_{k}(x,t)\right|^{{\gamma^{*}}+1}dx\right)^{\frac{2}{n}}\left(\iint_{\Omega_{T}}|\nabla u_{k}^{\frac{{\gamma^{*}}+1}{2}}|^{2}dxdt\right).

Using Eq. 3.31 and Eq. 3.32, from the above inequality we get by convexity argument

ΩT|uk|(γ+1)(n+2)n𝑑x𝑑tC(n)(C1ΩTfukγ𝑑x𝑑t+C2)n+2nC(n)22n((C1ΩTfukγ𝑑x𝑑t)n+2n+C2n+2n)C(n)22n((C1CfLr~(ΩT))n+2n(ΩT|uk|γr~𝑑x𝑑t)n+2nr~+C2n+2n).\begin{array}[]{rcl}\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{({\gamma^{*}}+1)(n+2)}{n}}dxdt&\leq&C(n)\left(C_{1}\iint_{\Omega_{T}}fu_{k}^{{\gamma^{*}}}dxdt+C_{2}\right)^{\frac{n+2}{n}}\\ &\leq&C(n)2^{\frac{2}{n}}\left(\left(C_{1}\iint_{\Omega_{T}}fu_{k}^{{\gamma^{*}}}dxdt\right)^{\frac{n+2}{n}}+C_{2}^{\frac{n+2}{n}}\right)\\ &\leq&C(n)2^{\frac{2}{n}}\left(\left(C_{1}C\|f\|_{L^{\tilde{r}}(\Omega_{T})}\right)^{\frac{n+2}{n}}\left(\iint_{\Omega_{T}}|u_{k}|^{\gamma^{*}\tilde{r}^{\prime}}dxdt\right)^{\frac{n+2}{n\tilde{r}^{\prime}}}+C_{2}^{\frac{n+2}{n}}\right).\end{array}

Now as n+2nr~=γγ+1<1\displaystyle\frac{n+2}{n\tilde{r}^{\prime}}=\frac{\gamma^{*}}{\gamma^{*}+1}<1, therefore using Young’s inequality we get

ΩT|uk|γr~𝑑x𝑑t=ΩT|uk|(γ+1)(n+2)n𝑑x𝑑tC,\iint_{\Omega_{T}}\left|u_{k}\right|^{\gamma^{*}\tilde{r}^{\prime}}dxdt=\iint_{\Omega_{T}}\left|u_{k}\right|^{\frac{({\gamma^{*}}+1)(n+2)}{n}}dxdt\leq C,

where C\displaystyle C is a positive constant independent of k\displaystyle k. The above boundedness along with Eqs. 3.32 and 3.31 gives that the sequence {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma^{*}+1}{2}}\right\}_{k} is uniformly bounded in L2(0,T;H01(Ω))L(0,T;L2(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega))\cap L^{\infty}(0,T;L^{2}(\Omega)) and {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma^{*}+1}(\Omega)). We now show {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(t0,T;Hloc1(Ω))\displaystyle L^{2}(t_{0},T;H_{loc}^{1}(\Omega)) for each t0>0\displaystyle t_{0}>0. Since γ>1\displaystyle\gamma^{*}>1, and ΩT\displaystyle\Omega_{T} is bounded and {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma^{*}+1}(\Omega)), we thus have {uk}k\displaystyle\left\{u_{k}\right\}_{k} is uniformly bounded in L2(0,T;L2(Ω))=L2(ΩT)\displaystyle L^{2}(0,T;L^{2}(\Omega))=L^{2}(\Omega_{T}), in particular in L2(K×(t0,T))\displaystyle L^{2}(K\times(t_{0},T)), for every KΩ\displaystyle K\subset\subset\Omega and for each t0>0\displaystyle t_{0}>0. Again as {ukγ+12}k\displaystyle\left\{u_{k}^{\frac{\gamma^{*}+1}{2}}\right\}_{k} is uniformly bounded in L2(0,T;H01(Ω))\displaystyle L^{2}(0,T;H_{0}^{1}(\Omega)), so we have by nonnegativity of uk\displaystyle u_{k}

t0TKukγ1|uk|2𝑑x𝑑t0TΩukγ1|uk|2𝑑x𝑑t4C(γ+1)2,\begin{array}[]{c}\int_{t_{0}}^{T}\int_{K}u_{k}^{\gamma^{*}-1}|\nabla u_{k}|^{2}dxdt\leq\int_{0}^{T}\int_{\Omega}u_{k}^{\gamma^{*}-1}|\nabla u_{k}|^{2}dxdt\leq\frac{4C}{(\gamma^{*}+1)^{2}},\end{array}

for every KΩ\displaystyle K\subset\subset\Omega and for each t0>0\displaystyle t_{0}>0. Using the positivity of uk\displaystyle u_{k} in ω×[t0,T)\displaystyle\omega\times[t_{0},T) for all k\displaystyle k, we conclude

t0TK|uk|2𝑑x𝑑y𝑑tCγc(K,t0)γ1.\int_{t_{0}}^{T}\int_{K}|\nabla u_{k}|^{2}dxdydt\leq\frac{C}{\gamma^{*}c_{(K,t_{0})}^{\gamma^{*}-1}}.

Since {uk}k\displaystyle\left\{u_{k}\right\}_{k} is increasing in k\displaystyle k and is uniformly bounded in L(0,T;Lγ+1(Ω))\displaystyle L^{\infty}(0,T;L^{\gamma^{*}+1}(\Omega)), we get by Beppo Levi’s Lemma uku\displaystyle u_{k}\uparrow u strongly in Lγ+1(ΩT)\displaystyle L^{\gamma^{*}+1}\left(\Omega_{T}\right) and hence in L1(ΩT)\displaystyle L^{1}(\Omega_{T}), where u\displaystyle u is the pointwise limit of uk\displaystyle u_{k} (possibly infinite). Then employing Fatou’s lemma we will get uγ+12L(0,T;L2(Ω))L2(0,T;H01(Ω))\displaystyle u^{\frac{\gamma^{*}+1}{2}}\in L^{\infty}(0,T;L^{2}(\Omega))\cap L^{2}(0,T;H_{0}^{1}(\Omega)) and uL2(t0,T;Hloc1(Ω))\displaystyle u\in L^{2}(t_{0},T;H_{loc}^{1}(\Omega)) for each t0>0\displaystyle t_{0}>0. Remark 2.23 implies that u\displaystyle u satisfies u(,0)=u0()\displaystyle u(\cdot,0)=u_{0}(\cdot) in L1\displaystyle L^{1} sense. Also, we have u(x,t)\displaystyle u(x,t)\geq C(ω,t0,n,s)\displaystyle C\left(\omega,t_{0},n,s\right) in ω×[t0,T)\displaystyle\omega\times\left[t_{0},T\right) for all ωΩ\displaystyle\omega\subset\subset\Omega and for each t0>0\displaystyle t_{0}>0. This u\displaystyle u will be a very weak solution to Eq. 2.13 in the sense of Definition 2.25, and the proof follows exactly similar to Theorem 2.39.

Remark 3.3.

In the above existence results, the cases where uL2(0,T;H01(Ω))\displaystyle u\in L^{2}(0,T;H^{1}_{0}(\Omega)), we can show that the weak solution u\displaystyle u is unique, if it has finite energy. For this we assume that v\displaystyle v is another finite energy solution to Eq. 2.12 or Eq. 2.13. Then, by the construction of u\displaystyle u in each existence result, we get v\displaystyle v is a supersolution to all the approximating problems. Hence by the comparison principle in Lemma 2.20, we deduce that vuk\displaystyle v\geq u_{k} for all k\displaystyle k. Then vu\displaystyle v\geq u. To prove the inverse inequality, let ϕ=vu\displaystyle\phi=v-u, then ϕ0\displaystyle\phi\geq 0 and ϕL2(0,T;H01(Ω))\displaystyle\phi\in L^{2}(0,T;H_{0}^{1}(\Omega)) and hence we can use both ϕ+\displaystyle\phi_{+} or ϕ\displaystyle\phi_{-} as test functions. Also, we note that ϕ\displaystyle\phi solves the problem

{ϕtΔϕ+(Δ)sϕ=f(1vγ(x,t)1uγ(x,t)) in ΩT,ϕ(x,t)=0 in (n\Ω)×(0,T),ϕ(x,0)=0 in Ω.\displaystyle\displaystyle\begin{cases}\phi_{t}-\Delta\phi+(-\Delta)^{s}\phi=f\left(\frac{1}{v^{\gamma(x,t)}}-\frac{1}{u^{\gamma(x,t)}}\right)&\text{ in }\Omega_{T},\\ \phi(x,t)=0&\text{ in }\left(\mathbb{R}^{n}\backslash\Omega\right)\times(0,T),\\ \phi(x,0)=0&\text{ in }\Omega.\end{cases}

Since f(1vγ(x,t)1uγ(x,t))0\displaystyle f\left(\frac{1}{v^{\gamma(x,t)}}-\frac{1}{u^{\gamma(x,t)}}\right)\leq 0 in ΩT\displaystyle\Omega_{T}, by comparison principle, we get ϕ0\displaystyle\phi\leq 0 in ΩT\displaystyle\Omega_{T}. Thus ϕ=0\displaystyle\phi=0 i.e. u=v\displaystyle u=v.

Remark 3.4.

As we can notice, in each of the existence results, we have uL2(t0,T;Wloc1,1(Ω))\displaystyle u\in L^{2}(t_{0},T;W^{1,1}_{\operatorname{loc}}(\Omega)) for each t0>0\displaystyle t_{0}>0 and this gives that our definition of a weak solution is well motivated in the sense that we have derived them integrating by parts twice upon multiplying by a test function.

4 Asymptotic behaviour

This section is devoted to the study of the asymptotic behaviour of finite energy solution to the problem Eq. 2.12, as t\displaystyle t\rightarrow\infty, for the particular case where f\displaystyle f depends only on x\displaystyle x and that f>0\displaystyle f>0 with fLq(Ω)\displaystyle f\in L^{q}(\Omega) and q\displaystyle q will be specified later. We first state the following existence and uniqueness result to the corresponding mixed local-nonlocal elliptic problem, which can be done as a direct application of Theorem 2.27. Consider the problem

{Δw+(Δ)sw=fwγ in Ω,w=0 in (n\Ω).{}\begin{array}[]{lcr}\begin{cases}-\Delta w+(-\Delta)^{s}w=\frac{f}{w^{\gamma}}&\mbox{ in }\Omega,\\ w=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega).\end{cases}\end{array} (4.1)

We define the weak solution to the above problem as:

Definition 4.1.

Suppose that fL1(Ω)\displaystyle f\in L^{1}(\Omega) is a nonnegative function and γ>0\displaystyle\gamma>0. We say that w\displaystyle w is a very weak solution to problem Eq. 4.1 if wL1(Ω)\displaystyle w\in L^{1}(\Omega) satisfies the boundary conditions and ωΩ,\displaystyle\forall\omega\subset\subset\Omega, cω>0\displaystyle\exists c_{\omega}>0 such that w(x)cω>0, a.e. in ω,\displaystyle w(x)\geq c_{\omega}>0,\text{ a.e. in }\omega, and for all φ𝒞0(Ω)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}(\Omega), we have

Ωw(Δφ+(Δ)sφ)𝑑x=Ωfφwγ𝑑x.\int_{\Omega}w(-\Delta\varphi+(-\Delta)^{s}\varphi)dx=\int_{\Omega}\frac{f\varphi}{w^{\gamma}}dx.
Theorem 4.2.

Let fL1(Ω)\displaystyle f\in L^{1}(\Omega) be a non-negative function. Then for all γ>0\displaystyle\gamma>0, the problem Eq. 4.1 has a nonnegative very weak solution w\displaystyle w such that wγ+12H01(Ω)\displaystyle w^{\frac{\gamma+1}{2}}\in H_{0}^{1}(\Omega) in the sense of Definition 4.1.

Proof.

Let wk\displaystyle w_{k} be the unique positive bounded solution to the approximating problem

{Δwk+(Δ)swk=fk(wk+1k)γ in Ω,wk>0 in Ω,wk=0 in (n\Ω);{}\begin{array}[]{lcr}\begin{cases}-\Delta w_{k}+(-\Delta)^{s}w_{k}=\frac{f_{k}}{(w_{k}+\frac{1}{k})^{\gamma}}&\mbox{ in }\Omega,\\ w_{k}>0&\mbox{ in }\Omega,\\ w_{k}=0&\mbox{ in }(\mathbb{R}^{n}\backslash\Omega);\end{cases}\end{array} (4.2)

with fk:=min(k,f)\displaystyle f_{k}:=\min(k,f). We note that the existence of wk\displaystyle w_{k} follows using a simple monotonicity argument. More precisely, one can prove a similar version of existence results and comparison principle as of Theorem 2.13 for the elliptic case also, and using that, one can find the elliptic version of Lemma 2.20 and Corollary 2.21. Again, by the comparison principle, we deduce that the sequence {wk}k\displaystyle\left\{w_{k}\right\}_{k} is increasing in k\displaystyle k. By the Harnack inequality, see [22, Theorem 8.3], it holds that for any set ωΩ\displaystyle\omega\subset\subset\Omega, for all k\displaystyle k we have, wk(x)w1(x)c(ω,n,s) in ω.\displaystyle w_{k}(x)\geq w_{1}(x)\geq c(\omega,n,s)\text{ in }\omega. Also, the details of existence, uniqueness, positivity and monotonicity of wk\displaystyle w_{k}’s can be found in [26, Lemma 3.2].
Now taking ((wk+ε)γεγ),0<ε<1/k,\displaystyle((w_{k}+\varepsilon)^{\gamma}-\varepsilon^{\gamma}),0<\varepsilon<1/k, as a test function in Eq. 4.2 and letting ε0\displaystyle\varepsilon\to 0, it follows that

Ωwkwkγ+122n(wk(x)wk(y))(wkγ(x)wkγ(y))|xy|n+2s𝑑x𝑑yΩf(x)𝑑x.\int_{\Omega}\nabla w_{k}\cdot\nabla w_{k}^{\gamma}+\frac{1}{2}\iint_{\mathbb{R}^{2n}}\frac{\left(w_{k}(x)-w_{k}(y)\right)\left(w_{k}^{\gamma}(x)-w_{k}^{\gamma}(y)\right)}{|x-y|^{n+2s}}dxdy\leq\int_{\Omega}f(x)dx.

By Lemma 2.11, we have

(wk(x)wk(y))(wkγ(x)wkγ(y))C(wkγ+12(x)wkγ+12(y))2,\left(w_{k}(x)-w_{k}(y)\right)\left(w_{k}^{\gamma}(x)-w_{k}^{\gamma}(y)\right)\geq C\left(w_{k}^{\frac{\gamma+1}{2}}(x)-w_{k}^{\frac{\gamma+1}{2}}(y)\right)^{2},

and on the other hand

wkwkγ=γwkγ1|wk|2,\nabla w_{k}\cdot\nabla w_{k}^{\gamma}=\gamma w_{k}^{\gamma-1}|\nabla w_{k}|^{2},

we deduce that the sequence {wkγ+12}k\displaystyle\left\{w_{k}^{\frac{\gamma+1}{2}}\right\}_{k} is bounded in the reflexive Banach space X0s(Ω)H01(Ω)\displaystyle X_{0}^{s}(\Omega)\cap H_{0}^{1}(\Omega). Now by the monotonicity of {wk}k\displaystyle\left\{w_{k}\right\}_{k}, using Beppo-Levi’s theorem, we get the existence of a measurable function w\displaystyle w such that wkw\displaystyle w_{k}\uparrow w a.e. in n,\displaystyle\mathbb{R}^{n}, wkγ+12wγ+12\displaystyle w_{k}^{\frac{\gamma+1}{2}}\rightharpoonup w^{\frac{\gamma+1}{2}} weakly in X0s(Ω)H01(Ω)\displaystyle X_{0}^{s}(\Omega)\cap H_{0}^{1}(\Omega). Clearly wc(ω)\displaystyle w\geq c(\omega) for any set ωΩ\displaystyle\omega\subset\subset\Omega. Then, letting k\displaystyle k\rightarrow\infty in the approximating problems and using the dominated convergence theorem, we can show that w\displaystyle w is a very weak solution to problem Eq. 4.1 with wγ+12H01(Ω)\displaystyle w^{\frac{\gamma+1}{2}}\in H_{0}^{1}(\Omega). We refer [26, Theorem 2.16\displaystyle 2.16] for the boundedness of w\displaystyle w. One can find conditions when wH01(Ω)\displaystyle w\in H^{1}_{0}(\Omega) in [26]. Further, [26, Corollary 5.2\displaystyle 5.2] gives such solution is unique. ∎

We now write the asymptotic behaviour of solutions to problem Eq. 2.12 under suitable conditions on f\displaystyle f and u0\displaystyle u_{0}.

Theorem 4.3.

Suppose that fLq(Ω)\displaystyle f\in L^{q}(\Omega) is a non-negative function depending only on x\displaystyle x, and q\displaystyle q is such that solution to the problem Eq. 4.1 is in H01(Ω)\displaystyle H^{1}_{0}(\Omega) and is unique. Let u\displaystyle u be a finite energy solution to problem Eq. 2.12 with 0u0w\displaystyle 0\leq u_{0}\leq w, where w\displaystyle w is the solution to Eq. 4.1. Then u(,t)w\displaystyle u(\cdot,t)\rightarrow w as t\displaystyle t\rightarrow\infty, in Lσ(Ω)\displaystyle L^{\sigma}(\Omega), where σ<2\displaystyle\sigma<2^{*}.

Proof.

We divide the proof into two cases according to the value of the initial condition u0\displaystyle u_{0}.
The first case u0=0\displaystyle u_{0}=0: In this case using the comparison principle as in Lemma 2.20, we get that uw\displaystyle u\leq w in Ω×(0,T)\displaystyle\Omega\times(0,T) for all T>0\displaystyle T>0. Hence, u\displaystyle u is globally defined. We now show that u(x,)\displaystyle u(x,\cdot) is increasing in t\displaystyle t. Since u\displaystyle u has finite energy, we fix t1>0\displaystyle t_{1}>0 and define v(x,t)=u(x,t1+t),\displaystyle v(x,t)=u(x,t_{1}+t), then v\displaystyle v satisfies the problem

{vtΔv+(Δ)sv=f(x)vγ(x,t) in ΩT,v(x,t)=0 in (n\Ω)×(0,T),v(x,0)=u(x,t1) in Ω.\displaystyle\displaystyle\begin{cases}v_{t}-\Delta v+(-\Delta)^{s}v=\frac{f(x)}{v^{\gamma}(x,t)}&\text{ in }\Omega_{T},\\ v(x,t)=0&\text{ in }\left(\mathbb{R}^{n}\backslash\Omega\right)\times(0,T),\\ v(x,0)=u\left(x,t_{1}\right)&\text{ in }\Omega.\end{cases}

Now as u(x,t1)u00\displaystyle u\left(x,t_{1}\right)\geq u_{0}\equiv 0 in Ω\displaystyle\Omega, using the comparison principle as in Lemma 2.20 again, we have uv\displaystyle u\leq v in ΩT\displaystyle\Omega_{T}. Therefore for all t1>0\displaystyle t_{1}>0, we have u(x,t)u(x,t1+t)\displaystyle u(x,t)\leq u\left(x,t_{1}+t\right) for all xn\displaystyle x\in\mathbb{R}^{n}. Therefore u(,t)\displaystyle u(\cdot,t) is increasing in t\displaystyle t. Since uw\displaystyle u\leq w, so there exists a measurable function u^()=limsuptu(,t)\displaystyle\hat{u}(\cdot)=\underset{t\rightarrow\infty}{\operatorname{lim}\operatorname{sup}}~{}u(\cdot,t). By dominated convergence theorem, u(,t)u^()\displaystyle u(\cdot,t)\rightarrow\hat{u}(\cdot) as t\displaystyle t\rightarrow\infty in Lσ(Ω)\displaystyle L^{\sigma}(\Omega) for σ<2\displaystyle\sigma<2^{*}. Also, we have u^w\displaystyle\hat{u}\leq w. It is now sufficient to show that u^\displaystyle\hat{u} is a very weak solution to problem Eq. 4.1.
Let φ𝒞0(Ω)\displaystyle\varphi\in\mathcal{C}_{0}^{\infty}(\Omega), then using φ\displaystyle\varphi as a test function in Eq. 2.12 and integrating in Ω×(k,k+1)\displaystyle\Omega\times(k,k+1), we get

Ω(u(x,k+1)u(x,k))φ(x)𝑑x+kk+1Ωu(x,t)(Δφ+(Δ)sφ)𝑑x𝑑t=kk+1Ωf(x)uγ(x,t)φ(x)𝑑x𝑑t.\int_{\Omega}(u(x,k+1)-u(x,k))\varphi(x)dx+\int_{k}^{k+1}\int_{\Omega}u(x,t)(-\Delta\varphi+(-\Delta)^{s}\varphi)dxdt=\int_{k}^{k+1}\int_{\Omega}\frac{f(x)}{u^{\gamma}(x,t)}\varphi(x)dxdt.

Now clearly we have

Ω(u(x,k+1)u(x,k))|φ(x)|𝑑x0 as k.\int_{\Omega}(u(x,k+1)-u(x,k))|\varphi(x)|dx\rightarrow 0\text{ as }k\rightarrow\infty.

On the other hand, in the set Ω×(k,k+1)\displaystyle\Omega\times(k,k+1) using the monotonicity of u\displaystyle u in the variable t\displaystyle t, we have

f(x)uγ(x,k+1)|φ(x)|f(x)uγ(x,t)|φ(x)|f(x)uγ(x,k)|φ(x)|,t(k,k+1),\frac{f(x)}{u^{\gamma}(x,k+1)}|\varphi(x)|\leq\frac{f(x)}{u^{\gamma}(x,t)}|\varphi(x)|\leq\frac{f(x)}{u^{\gamma}(x,k)}|\varphi(x)|,\quad\forall t\in(k,k+1),
u(x,k)|Δφ|u(x,t)|Δφ|u(x,k+1)|Δφ|,t(k,k+1),u(x,k)\left|-\Delta\varphi\right|\leq u(x,t)\left|-\Delta\varphi\right|\leq u(x,k+1)\left|-\Delta\varphi\right|,\quad\forall t\in(k,k+1),

and

u(x,k)|(Δ)sφ|u(x,t)|(Δ)sφ|u(x,k+1)|(Δ)sφ|,t(k,k+1).u(x,k)\left|(-\Delta)^{s}\varphi\right|\leq u(x,t)\left|(-\Delta)^{s}\varphi\right|\leq u(x,k+1)\left|(-\Delta)^{s}\varphi\right|,\quad\forall t\in(k,k+1).

Then, since ϕ,Δϕ,(Δ)sϕ\displaystyle\phi,\Delta\phi,(-\Delta)^{s}\phi are bounded, so by the dominated convergence theorem, we will get that, as k\displaystyle k\rightarrow\infty,

kk+1Ωu(x,t)(Δφ+(Δ)sφ)𝑑x𝑑tΩu^(x)(Δφ+(Δ)sφ)𝑑x,\int_{k}^{k+1}\int_{\Omega}u(x,t)(-\Delta\varphi+(-\Delta)^{s}\varphi)dxdt\rightarrow\int_{\Omega}\hat{u}(x)(-\Delta\varphi+(-\Delta)^{s}\varphi)dx,

and

kk+1Ωf(x)uγ(x,t)φ(x)𝑑x𝑑tΩf(x)u^γ(x)φ(x)𝑑x.\int_{k}^{k+1}\int_{\Omega}\frac{f(x)}{u^{\gamma}(x,t)}\varphi(x)dxdt\rightarrow\int_{\Omega}\frac{f(x)}{\hat{u}^{\gamma}(x)}\varphi(x)dx.

Hence, u^\displaystyle\hat{u} is a very weak solution to problem Eq. 4.1. By the uniqueness we then get u^=w\displaystyle\hat{u}=w.
The second case 0<u0w\displaystyle 0<u_{0}\leq w: Let us denote by u^\displaystyle\hat{u} a finite energy solution to problem Eq. 2.12 in this case. Then by the comparison principle in Lemma 2.20, we deduce that uu^w\displaystyle u\leq\hat{u}\leq w, where u\displaystyle u is the weak solution to problem Eq. 2.12 with u0=0\displaystyle u_{0}=0. Hence by the convergence result of the first case, we get that u^(,t)w\displaystyle\hat{u}(\cdot,t)\rightarrow w, as t\displaystyle t\rightarrow\infty, strongly in Lσ(Ω)\displaystyle L^{\sigma}(\Omega) for all σ<2\displaystyle\sigma<2^{*}. ∎

References

  • [1] Boumediene Abdellaoui, María Medina, Ireneo Peral, and Ana Primo. The effect of the Hardy potential in some Calderón-Zygmund properties for the fractional Laplacian. J. Differential Equations, 260(11):8160–8206, 2016.
  • [2] D. G. Aronson and James Serrin. Local behavior of solutions of quasilinear parabolic equations. Arch. Rational Mech. Anal., 25:81–122, 1967.
  • [3] Rakesh Arora and Vicenţiu D. Rădulescu. Combined effects in mixed local–nonlocal stationary problems. Proceedings of the Royal Society of Edinburgh Section A: Mathematics, page 1–47, 2023.
  • [4] Mehdi Badra, Kaushik Bal, and Jacques Giacomoni. Existence results to a quasilinear and singular parabolic equation. Discrete Contin. Dyn. Syst., pages 117–125, 2011.
  • [5] Mehdi Badra, Kaushik Bal, and Jacques Giacomoni. Some results about a quasilinear singular parabolic equation. Differ. Equ. Appl., 3(4):609–627, 2011.
  • [6] Mehdi Badra, Kaushik Bal, and Jacques Giacomoni. A singular parabolic equation: existence, stabilization. J. Differential Equations, 252(9):5042–5075, 2012.
  • [7] Begoña Barrios, Ida De Bonis, María Medina, and Ireneo Peral. Semilinear problems for the fractional Laplacian with a singular nonlinearity. Open Math., 13(1):390–407, 2015.
  • [8] Kheireddine Biroud. Mixed local and nonlocal equation with singular nonlinearity having variable exponent. J. Pseudo-Differ. Oper. Appl., 14(1):Paper No. 13, 24, 2023.
  • [9] Lucio Boccardo and Luigi Orsina. Semilinear elliptic equations with singular nonlinearities. Calc. Var. Partial Differential Equations, 37(3-4):363–380, 2010.
  • [10] Lucio Boccardo, Maria Michaela Porzio, and Ana Primo. Summability and existence results for nonlinear parabolic equations. Nonlinear Anal., 71(3-4):978–990, 2009.
  • [11] Brahim Bougherara and Jacques Giacomoni. Existence of mild solutions for a singular parabolic equation and stabilization. Adv. Nonlinear Anal., 4(2):123–134, 2015.
  • [12] Stefano Buccheri, João Vítor da Silva, and Luís Henrique de Miranda. A system of local-nonlocal p\displaystyle p-Laplacians: the eigenvalue problem and its asymptotic limit as p\displaystyle p\to\infty. Asymptotic Analysis, 128(2):149–181, 2022.
  • [13] Annamaria Canino, Luigi Montoro, Berardino Sciunzi, and Marco Squassina. Nonlocal problems with singular nonlinearity. Bull. Sci. Math., 141(3):223–250, 2017.
  • [14] José Carmona and Pedro J. Martínez-Aparicio. A singular semilinear elliptic equation with a variable exponent. Adv. Nonlinear Stud., 16(3):491–498, 2016.
  • [15] M. G. Crandall, P. H. Rabinowitz, and L. Tartar. On a Dirichlet problem with a singular nonlinearity. Comm. Partial Differential Equations, 2(2):193–222, 1977.
  • [16] Stuti Das. Gradient Hölder regularity in mixed local and nonlocal linear parabolic problem. Journal of Mathematical Analysis and Applications, page 128140, 2024.
  • [17] Ida de Bonis and Linda Maria De Cave. Degenerate parabolic equations with singular lower order terms. Differential Integral Equations, 27(9-10):949–976, 2014.
  • [18] Cristiana De Filippis and Giuseppe Mingione. Gradient regularity in mixed local and nonlocal problems. Mathematische Annalen, pages 1–68, 2022.
  • [19] Eleonora Di Nezza, Giampiero Palatucci, and Enrico Valdinoci. Hitchhiker’s guide to the fractional sobolev spaces. Bulletin des sciences mathématiques, 136(5):521–573, 2012.
  • [20] Lawrence C Evans. Partial Differential Equations, volume 19. American Mathematical Soc., 2010.
  • [21] Yuzhou Fang, Bin Shang, and Chao Zhang. Regularity theory for mixed local and nonlocal parabolic p\displaystyle p-Laplace equations. J. Geom. Anal., 32(1):Paper No. 22, 33, 2022.
  • [22] Prashanta Garain and Juha Kinnunen. On the regularity theory for mixed local and nonlocal quasilinear elliptic equations. Trans. Amer. Math. Soc., 375(8):5393–5423, 2022.
  • [23] Prashanta Garain and Juha Kinnunen. Weak Harnack inequality for a mixed local and nonlocal parabolic equation. Journal of Differential Equations, 360:373–406, 2023.
  • [24] Prashanta Garain and Erik Lindgren. Higher hölder regularity for mixed local and nonlocal degenerate elliptic equations. Calculus of Variations and Partial Differential Equations, 62(2):67, 2023.
  • [25] Prashanta Garain and Tuhina Mukherjee. Quasilinear nonlocal elliptic problems with variable singular exponent. Commun. Pure Appl. Anal., 19(11):5059–5075, 2020.
  • [26] Prashanta Garain and Alexander Ukhlov. Mixed local and nonlocal Sobolev inequalities with extremal and associated quasilinear singular elliptic problems. Nonlinear Anal., 223:Paper No. 113022, 35, 2022.
  • [27] David Kinderlehrer and Guido Stampacchia. An introduction to variational inequalities and their applications. SIAM, 2000.
  • [28] A. C. Lazer and P. J. McKenna. On a singular nonlinear elliptic boundary-value problem. Proc. Amer. Math. Soc., 111(3):721–730, 1991.
  • [29] Tommaso Leonori, Ireneo Peral, Ana Primo, and Fernando Soria. Basic estimates for solutions of a class of nonlocal elliptic and parabolic equations. Discrete Contin. Dyn. Syst., 35(12):6031–6068, 2015.
  • [30] Bin Shang and Chao Zhang. Hölder regularity for mixed local and nonlocal p\displaystyle p-Laplace parabolic equations. Discrete Contin. Dyn. Syst., 42(12):5817–5837, 2022.
  • [31] Bin Shang and Chao Zhang. Harnack inequality for mixed local and nonlocal parabolic p\displaystyle p-Laplace equations. J. Geom. Anal., 33(4):Paper No. 124, 24, 2023.
  • [32] Zhuoqun Wu, Jingxue Yin, and Chunpeng Wang. Elliptic & parabolic equations. World Scientific, 2006.