Open Access

On Lr-regularity of global attractors generated by strong solutions of reaction-diffusion equations

   | Sep 16, 2016

Cite

Introduction

This paper is mainly devoted to studying the regularity properties of global attractors for multivalued semiflows generated by strong solutions of reaction-diffusion equations.

The existence and properties of global attractors for dynamical systems generated by reaction-diffusion equations have been studied by many authors over the last thirty years. For equations generating a single-valued semigroup such results are well known since the 80s (see e.g. [5], [6], [7], [8], [24], [34]). Moreover, deep results concerning the structure of the attractors were proved for scalar equations (see [12], [13], [29], [30], [31]). Furthermore, for rather general parabolic equations boundedness of attractors in Sobolev spaces of higher orders were obtained as well (see e.g. [3], [4]).

When uniqueness of the Cauchy problem is not satisfied we have to work with multivalued semiflows rather than semigroups. In this direction, existence and topological properties of global and trajectory attractors have been studied by several authors over the last years (see [2], [9], [10], [11], [14], [16], [20], [22], [23], [25], [26], [36]). However, concerning the structure of the attractor little is known so far in comparison with the single-valued case. Nevertheless, recently the global attractor has been characterised using the unstable manifold of the set of stationary points, generalizing in this way well-known results from the single-valued case (see [17], [18], [19]). In particular, such structure was proved to be true for the global attractor generated by strong solutions of reaction diffusion equations in which the nonlinear term satisfies a critical growth condition.

It is important to point out that there are two different approaches to the study of these equations. One method relies on the construction of weak solutions through Galerkin approximations, whereas the other one makes use of the properties of sectorial operators in order to obtain mild solutions, which are defined by the variation of constants formula. It seems that this has given rise to two separate groups of papers, whose paths have rarely crossed. However, we find it interesting to use the powerful technique of sectorial operators in order to improve the regularity of weak solutions and global attractors in the multivalued case. Indeed, in this paper we prove that under a suitable assumption mild and weak solutions are equivalent, and using this result we are able to improve the regularity of the global attractor generated by strong solutions which was obtained in [17].

This paper is split into three different parts.

In the first section, we prove that the concepts of weak and mild solutions are equivalent provided that an appropriate condition holds.

In the second section, we use the above equivalence in order to show that the global attractor generated by strong solutions of the reaction-diffusion equation is bounded in suitable Lr-spaces in the case where the nonlinear term satisfies a critical growth condition.

Finally, in the third section, considering a supercritical growth condition we define a multivalued semiflow by taking all strong solutions satisfying an energy inequality and then prove that a weak global attractor exists, that is, we construct an attractor which attracts bounded sets of the phase space with respect to the weak topology.

Equivalence of different definitions of solutions

Let Ω ⊂ ℝn, n ≥ 1, be an open bounded subset with sufficiently smooth boundary Ω. We consider the following problem

utΔu+f(t,u)=h(t,x),xΩ,t>τ,u|Ω=0,u(τ,x)=u0(x), xΩ.$$\begin{array}{} \displaystyle \left\{ \begin{array} [c]{l} \dfrac{\partial u}{\partial t}-\Delta u+f(t,u)=h(t,x),\quad x\in \Omega,\, t \gt \tau,\\ u|_{\partial\Omega}=0,\\ u(\tau,x)=u_{0}(x)\text{, }x\in\Omega. \end{array} \right. \end{array}$$

The functions f, h are assumed to satisfy the following conditions:

fC(R×R),$$\begin{array}{} \displaystyle f\in\mathbb{C}(\mathbb{R}\times\mathbb{R}), \end{array}$$

hLloc2(R;L2(Ω)),$$\begin{array}{} \displaystyle h\in L_{loc}^{2}(\mathbb{R};L^{2}(\Omega)), \end{array}$$

|f(t,u)|C1(1+|u|p1),$$\begin{array}{} \displaystyle |f(t,u)|\leq C_{1}(1+|u|^{p-1}), \end{array}$$

f(t,u)uα|u|pC2,$$\begin{array}{} \displaystyle f(t,u)u\geq\alpha|u|^{p}-C_{2}, \end{array}$$

for all (t, u) ∈ ℝ × ℝ, where C1, C2 are positive constants, p ≥ 2 and α > 0 if p > 2 but α ∈ ℝ if p = 2.

Denote F(u)=0uf(s)ds.$\begin{array}{} F(u)=\int_{0}^{u}f(s)ds. \end{array}$ If p > 2 from (4)-(5) we obtain that lim inf|u|f(u)u=+$\begin{array}{} \displaystyle \liminf\limits_{|u|\rightarrow\infty}\frac{f(u)}{u}=+\infty \end{array}$ and that there exists D1, D2, δ > 0 such that

|F(u)|D1(1+up),F(u)δupD2,uR.$$\begin{array}{} \displaystyle |F(u)|\leq D_{1}(1+\left\vert u\right\vert ^{p}),\,\, F(u)\geq\delta\left\vert u\right\vert ^{p}-D_{2},\quad\forall u\in\mathbb{R}. \end{array}$$

If p = 2, then (6) remains valid but with δ ∈ ℝ.

In the sequel we shall denote by H the space L2(Ω) endowed with the norm ∥⋅∥ and the scalar product (⋅, ⋅), and by V the space H01$\begin{array}{} \displaystyle H_{0}^{1} \end{array}$ (Ω) with the norm ∥⋅∥V and the scalar product ((⋅, ⋅)), whereas V = H−1(Ω) is the dual space of V with the norm ∥⋅∥V. The pairing between the space V and V will be denoted by 〈⋅, ⋅〉. Also, putting 1p+1q=1,$\begin{array}{} \displaystyle \frac{1}{p}+\frac{1}{q}=1, \end{array}$ the pairing between the spaces Lp(Ω) and Lq(Ω) will be given by 〈⋅, ⋅〉q,p.

Definition 1

The function uLlocp(τ,+;Lp(Ω))Lloc2(τ,+;V)$\begin{array}{} \displaystyle u\in L_{loc}^{p}(\tau,+\infty;L^{p}(\Omega))\cap L_{loc} ^{2}(\tau,+\infty;V) \end{array}$ is called a weak solution to problem (1) on (τ, +∞) if for all T > τ, vVLp(Ω) one has

ddtut,v+ut,v+f(t,u(t)),vq,p=ht,v.$$\begin{array}{} \displaystyle \frac{d}{dt}\left( u\left( t\right) ,v\right) +\left( \left( u\left( t\right) ,v\right) \right) +\left\langle f(t,u(t)),v\right\rangle _{q,p}=\left( h\left( t\right) ,v\right) . \end{array}$$

in the sense of scalar distributions on (τ, T).

For a weak solution let g(t) = −f(t, u(t)) + h(t) + Δ u(t). From uLlocp(τ,+;Lp(Ω))Lloc2(τ,+;V)$\begin{array}{} \displaystyle |f(t,u)|\leq C_{1}(1+|u|^{p-1}), \end{array}$ and condition (4) it is clear that gLlocq(τ,+;Lq(Ω))+Lloc2(τ,+;V).$\begin{array}{} \displaystyle g\in L_{loc}^{q}(\tau,+\infty;L^{q}(\Omega))+L_{loc}^{2}(\tau,+\infty;V^{\prime}). \end{array}$ Thus, we can rewrite (7) as

ddtut,v=g(t),vV+Lq, for all vVLp(Ω),$$\begin{array}{} \displaystyle \frac{d}{dt}\left( u\left( t\right) ,v\right) =\left\langle g(t),v\right\rangle _{V^{\prime}+L^{q}}\text{, for all }v\in V\cap L^{p}(\Omega)\text{,} \end{array}$$

where 〈⋅, ⋅〉V+Lq denotes pairing between V + Lq(Ω) and VLp(Ω).

It is well-known [10, p.284] that for any uτH there exists at least one weak solution u(⋅). If, moreover, f(t, ⋅) ∈ C1(ℝ) and fu$\begin{array}{} \displaystyle \dfrac{\partial f}{\partial u} \end{array}$ (t, u) ≥ −C for any (t, u), then the solution is unique.

In order to consider an equivalent equality to (7) we recall the following well-known result.

Lemma 1

[33, p.250] LetXbe Banach space with dualX, u, gL1(a, b; X). Then the following statements are equivalent:

u(t)=ξ+atg(s)ds,$\begin{array}{} u(t)=\xi+\int\limits_{a}^{t}g(s)ds, \end{array}$ξXfor a.a. t ∈ (a, b);

abu(t)φ(t)dt=abg(t)φ(t)dt,φC0(a,b);$\begin{array}{} \int\limits_{a}^{b}u(t)\varphi^{\prime}(t)dt=-\int\limits_{a} ^{b}g(t)\varphi(t)dt, \forall\, \varphi\in C_{0}^{\infty}(a,b); \end{array}$

ddtu,ηX,X=g,ηX,X,ηX,$\begin{array}{} \displaystyle \frac{d}{dt}\langle u,\eta\rangle_{X^{\prime},X}=\langle g,\eta \rangle_{X^{\prime},X}, \forall\ \eta\in X^{\prime}, \end{array}$in the sense of scalar distributions on (a, b), where 〈⋅, ⋅〉X,Xdenotes pairing betweenXandX.

If properties 1)-3) hold, then, moreover, u(⋅) is a.e. equivalent to a continuous function from [a, b] inX.

Applying this lemma with X = V + Lq(Ω) we obtain that (7) is equivalent to the equality dudt=g$\begin{array}{} \displaystyle \dfrac{du}{dt}=g \end{array}$ in the sense of X-valued distributions on every interval [τ, T]. Hence,

dudtLlocq(τ,+;Lq(Ω))+Lloc2(τ,+;V)$$\begin{array}{} \displaystyle \frac{du}{dt}\in L_{loc}^{q}(\tau,+\infty;L^{q}(\Omega))+L_{loc}^{2}% (\tau,+\infty;V^{\prime}) \end{array}$$

and (7) is in fact equivalent to the equality

τTdudt,ξ(t)V+Lqdt+τTut,ξ(t)dt+τTf(t,u(t)),ξ(t)q,pdt=τTh(t),ξ(t)dt,$$\begin{array}{} \displaystyle \int_{\tau}^{T}\left\langle \frac{du}{dt},\xi(t)\right\rangle _{V^{\prime }+L^{q}}dt+\int_{\tau}^{T}\left( \left( u\left( t\right) ,\xi(t)\right) \right) dt+\int_{\tau}^{T}\left\langle f(t,u(t)),\xi(t)\right\rangle _{q,p}dt\\ \displaystyle =\int_{\tau}^{T}\left( h(t),\xi(t)\right) dt, \end{array}$$

for any ξLp(τ, T; Lp(Ω)) ∩ L2(τ, T; V).

Property (8) implies that uC([τ, +∞), H) and that the function t ↦ ∥u(t)∥2 is absolutely continuous on [τ, T] and ddtu(t)2=2dudt,uV+Lq$\begin{array}{} \displaystyle \dfrac{d}{dt}\Vert u(t)\Vert^{2}=2\left\langle \dfrac{du}{dt},u\right\rangle _{V^{\prime}+L^{q}} \end{array}$ for a.a. t ∈ (τ, T) [10, p.285]. Hence, the initial condition u(τ) = uτ makes sense.

Also, since the space C0$\begin{array}{} \displaystyle C_{0}^{\infty} \end{array}$ ((τ, T) × Ω) is dense in Lp(τ, T; Lp(Ω)) ∩ L2(τ, T; V), equality (9) is equivalent to

τTΩut,xϕtdxdtτTΩut,x2ϕx2dxdt+τTΩfut,xϕt,xdxdt=τTΩht,xϕt,xdxdt,$$\begin{array}{} \displaystyle -\int_{\tau}^{T}\int_{\Omega}u\left( t,x\right) \frac{\partial\phi }{\partial t}dxdt-\int_{\tau}^{T}\int_{\Omega}u\left( t,x\right) \frac{\partial^{2}\phi}{\partial x^{2}}dxdt+\int_{\tau}^{T}\int_{\Omega }f\left( u\left( t,x\right) \right) \phi\left( t,x\right) dxdt\\ \displaystyle =\int_{\tau}^{T}\int_{\Omega}h\left( t,x\right) \phi\left( t,x\right) dxdt, \end{array}$$

for any ϕC0$\begin{array}{} \displaystyle C_{0}^{\infty} \end{array}$ ((τ, T) × Ω), that is, u(⋅) is a weak solution if and only if it is a solution in the sense of distributions.

Let us consider now another concept of solution. Namely, we will define a mild solution to (1).

It is known that the operator A = Δ:D(A) = H2(Ω) ∩ H01$\begin{array}{} \displaystyle H_{0}^{1} \end{array}$ (Ω) → H is the generator of a strongly continuous semigroup of contractions T : ℝ+ × HH. Moreover, it follows from well-known results [27] that for any xD(A) the function u(t) = T(t) x is the unique classical solution to the problem

dudt=Au(t),t>0,u0=x.$$\begin{array}{} \displaystyle \left\{ \begin{array} [c]{c} \dfrac{du}{dt}=Au(t),\ t \gt 0,\\ u\left( 0\right) =x. \end{array} \right. \end{array}$$

Further, we shall study the inhomogeneous problem

dudt=Au(t)+gt,t>τ,uτ=x,$$\begin{array}{} \displaystyle \left\{ \begin{array} [c]{c} \dfrac{du}{dt}=Au(t)+g\left( t\right) ,\, t \gt \tau,\\ u\left( \tau\right) =x, \end{array} \right. \end{array}$$

where g:ℝ → H.

Definition 2

Let xH and gL1(τ, T; H). Then the function uC([τ, T], H) is called a mild solution to problem (11) on [τ, T] if

ut=Ttτx+τtTtsg(s)ds, τtT.$$\begin{array}{} \displaystyle u\left( t\right) =T\left( t-\tau\right) x+\int_{\tau}^{t}T\left( t-s\right) g(s)ds\text{, }\tau\leq t\leq T. \end{array}$$

If gLloc1$\begin{array}{} \displaystyle L_{loc}^{1} \end{array}$ ([τ, +∞)), then uC([τ, +∞), H) is called a mild solution if (12) is satisfied on every interval [τ, T].

It is called a classical solution on [τ, T] if u(⋅) is continuously differentiable on (τ, T), u(t) ∈ D(A) for any t ∈ (τ, T), u(τ) = x and the equality in (11) is satisfied for every t ∈ (τ, T).

It follows readily from this definition that problem (11) possesses a unique mild solution for every xH. Also, if g is continuously differentiable on [τ, T] and xD(A), then the mild solution is in fact the unique classical solution [27, p.107].

Coming back to our problem (1), let us introduce the concept of mild solution for it.

Definition 3

The function uLlocp(τ,+;Lp(Ω))Lloc2(τ,+;V)C([τ,+),H)$\begin{array}{} \displaystyle u\in L_{loc}^{p}(\tau,+\infty;L^{p}(\Omega))\cap L_{loc} ^{2}(\tau,+\infty;V)\cap C([\tau,+\infty),H) \end{array}$ is called a mild solution to problem (1) on (τ, +∞) if for all T > τ the function g(⋅) = h(⋅) − f(⋅, u(⋅)) belongs to L1(τ, T; H) and the equality (12) holds true.

Our aim now is to show that under an additional assumption the concepts of weak and mild solutions coincide.

Lemma 2

Assume thatu(⋅) is a weak solution to problem (1) with initial datumuτH which satisfies

f,u()L2(τ,T;H).$$\begin{array}{} \displaystyle f\left( \cdot,u(\cdot)\right) \in L^{2}(\tau,T;H). \end{array}$$

Then it is a mild solution as well.

Proof

Let us define the function g(t) = − f(t, u(t)) + h(t), which belongs to L2(τ, T; H) for every τ < T due to (13). We need to prove that the equality (12) holds. For an arbitrary fixed interval [τ, T] we take sequences u0nH2(Ω)H01(Ω)$\begin{array}{} \displaystyle u_{0}^{n}\in H^{2}(\Omega)\cap H_{0}^{1}(\Omega) \end{array}$ , gn(⋅) ∈ C1([τ, T], H), T < T, such that u0nu0$\begin{array}{} \displaystyle u_{0} ^{n}\rightarrow u_{0} \end{array}$ in H, gng in L2(τ, T; H). Let un be the unique classical solution to the problem

dundt=Aun(t)+gnt,t(τ,T¯),unτ=uτn.$$\begin{array}{} \displaystyle \left\{ \begin{array} [c]{c} \dfrac{du^{n}}{dt}=Au^{n}(t)+g^{n}\left( t\right) ,\ t\in(\tau,\overline {T}),\\ u^{n}\left( \tau\right) =u_{\tau}^{n}. \end{array} \right. \end{array}$$

Since uC1([τ + ε, T], H) for any 0 < ε < Tτ, the difference unu satisfies

12ddtunu2+un(t)u(t)V2gn(t)g(t)un(t)u(t) for a.a. tτ+ε,T.$$\begin{array}{} \displaystyle \frac{1}{2}\frac{d}{dt}\left\Vert u^{n}-u\right\Vert ^{2}+\left\Vert u^{n}(t)-u(t)\right\Vert _{V}^{2}\leq\left\Vert g^{n}(t)-g(t)\right\Vert \left\Vert u^{n}(t)-u(t)\right\Vert \text{ for a.a. }t\in\left( \tau+\varepsilon,T\right) . \end{array}$$

Hence, by using Young’s inequality and Gronwall’s lemma it is standard to show that unu in C([τ + ε, T], H).

Further, from the equality

unt=Ttτεun(τ+ε)+τ+εtTtsgn(s)ds, τ+εtT,$$\begin{array}{} \displaystyle u^{n}\left( t\right) =T\left( t-\tau-\varepsilon\right) u^{n} (\tau+\varepsilon)+\int_{\tau+\varepsilon}^{t}T\left( t-s\right) g^{n}(s)ds\text{, }\tau+\varepsilon\leq t\leq T, \end{array}$$

and taking into account that

Ttτεun(τ+ε)Ttτεu(τ+ε),τ+εtTtsgn(s)dsτ+εtTtsg(s)ds,$$\begin{array}{} \displaystyle T\left( t-\tau-\varepsilon\right) u^{n}(\tau+\varepsilon)\rightarrow T\left( t-\tau-\varepsilon\right) u(\tau+\varepsilon),\\ \displaystyle\int_{\tau+\varepsilon}^{t}T\left( t-s\right) g^{n}(s)ds\rightarrow \int_{\tau+\varepsilon}^{t}T\left( t-s\right) g(s)ds, \end{array}$$

where the last convergence follows from Lebesgue’s theorem, we have

ut=Ttτεu(τ+ε)+τ+εtTtsg(s)ds, τ+εtT.$$\begin{array}{} \displaystyle u\left( t\right) =T\left( t-\tau-\varepsilon\right) u(\tau+\varepsilon )+\int_{\tau+\varepsilon}^{t}T\left( t-s\right) g(s)ds\text{, } \tau+\varepsilon\leq t\leq T. \end{array}$$

Passing to the limit as ε → 0 and using that T(tτε) u(τ + ε) → T(tτ) u(τ) we finally obtain that

ut=Ttτu(τ)+τtTtsg(s)ds, τtT,$$\begin{array}{} \displaystyle u\left( t\right) =T\left( t-\tau\right) u(\tau)+\int_{\tau}^{t}T\left( t-s\right) g(s)ds\text{, }\tau\leq t\leq T, \end{array}$$

so we conclude that u(⋅) is a mild solution.

We prove now the converse statement.

Lemma 3

Assume thatu(⋅) is a mild solution to problem (1) with initial datauτHwhich satisfies (13). Then it is a weak solution.

Proof

Since the function g(⋅) = h(⋅) − f(⋅, u(⋅)) belongs to L2(τ, T; H), using standard results [34, p.68] we obtain that the problem

vtΔv=g(t,x),xΩ,t>τ,v|Ω=0,v(τ,x)=uτ(x), xΩ,$$\begin{array}{} \displaystyle \left\{ \begin{array} [c]{l} \dfrac{\partial v}{\partial t}-\Delta v=g(t,x),\quad x\in\Omega,\, t \gt \tau,\\ v|_{\partial\Omega}=0,\\ v(\tau,x)=u_{\tau}(x)\text{, }x\in\Omega, \end{array} \right. \end{array}$$

possesses a unique weak solution v(⋅). However, arguing in the same way as in Lemma 2 we deduce that v(⋅) is a mild solution to problem (11) with x = uτ. Therefore, by uniqueness of the mild solution we have u(⋅) = v(⋅), so our statement follows.

Regularity of the strong global attractor

Let us consider problem (1) in the autonomous case, that is,

utΔu+f(u)=h(x),xΩ,t>τ,u|Ω=0,u(0,x)=u0(x), xΩ,$$\begin{array}{} \displaystyle \left\{ \begin{array} [c]{l} \dfrac{\partial u}{\partial t}-\Delta u+f(u)=h(x),\quad x\in\Omega,\, t \gt \tau,\\ u|_{\partial\Omega}=0,\\ u(0,x)=u_{0}(x)\text{, }x\in\Omega, \end{array} \right. \end{array}$$

and we assume that conditions (2)-(5) are satisfied. In particular, this means that hH.

In [17] the existence of a global attractor for the multivalued semiflow generated by the strong solutions to (14) was proved assuming a critical growth condition on the nonlinear function f. In this section, our aim is to prove that this attractor is bounded in suitable Lr-spaces.

Definition 4

A weak solution u(⋅) to problem (1) is called a strong solution if, additionally, it satisfies

uLloc0,+;VLp(Ω),$$\begin{array}{} \displaystyle u\in L_{loc}^{\infty}\left( 0,+\infty;V\cap L^{p}(\Omega)\right) , \end{array}$$

dudtLloc20,+;H.$$\begin{array}{} \displaystyle \frac{du}{dt}\in L_{loc}^{2}\left( 0,+\infty;H\right) . \end{array}$$

As uL(0, T; VLp(Ω)) ∩ C([0, T]; H), we deduce that uC([0,T];H0w1(Ω)Lwp(Ω))$\begin{array}{} \displaystyle u\in C([0,T];H_{0w}^{1}(\Omega)\cap L_{w}^{p}(\Omega)) \end{array}$ for any T > 0, where H0w1(Ω),Lwp(Ω)$\begin{array}{} \displaystyle H_{0w}^{1}(\Omega),\, L_{w}^{p}(\Omega) \end{array}$ are respectively the spaces H01$\begin{array}{} \displaystyle H_{0}^{1} \end{array}$ (Ω), Lp(Ω) with the weak topology [33, Lemma 1.4, p.263].

Throughout this section we will assume that

p2N2N2$$\begin{array}{} \displaystyle p\leq\frac{2N-2}{N-2} \end{array}$$

if N ≥ 3, that is, f satisfies a critical growth condition. No assumption is imposed if N ≤2.

From (4), (17) and the imbeddings H1(Ω) ⊂ L2NN2$\begin{array}{} \displaystyle L^{\frac{2N}{N-2}} \end{array}$ (Ω), if N ≥ 3, H1(Ω) ⊂ Lq(Ω), for any q ≥ 1 if N ≤2, we get

τTΩfu2dxdtK1τT1+utL2p2(Ω)2p2dtK2τT1+utV2p2dt.$$\begin{array}{} \displaystyle \int_{\tau}^{T}\int_{\Omega}\left\vert f\left( u\right) \right\vert ^{2}dxdt\leq K_{1}\int_{\tau}^{T}\left( 1+\left\Vert u\left( t\right) \right\Vert _{L^{2p-2}(\Omega)}^{2p-2}\right) dt\leq K_{2}\int_{\tau} ^{T}\left( 1+\left\Vert u\left( t\right) \right\Vert _{V}^{2p-2}\right) dt. \end{array}$$

Then the Δu=dudt+fuh$\begin{array}{} \displaystyle \Delta u=\dfrac{du}{dt}+f\left( u\right) -h \end{array}$ and (15)-(16) imply that

uL2τ,T;DA$$\begin{array}{} \displaystyle u\in L^{2}\left(\tau,T;D\left(A\right)\right) \end{array}$$

for any strong solution u(⋅).

Since p2N2N22NN2,$\begin{array}{} \displaystyle p\leq\frac{2N-2}{N-2}\leq\frac{2N}{N-2}, \end{array}$ it is also obvious that VLp(Ω).

By standard results [28, Corollary 7.3] we obtain then that uC([0, +∞), V) and

ddtuV2=2Δu,dudt for a.a. t>0.$$\begin{array}{} \displaystyle \frac{d}{dt}\left\Vert u\right\Vert _{V}^{2}=2\left( -\Delta u,\frac{du} {dt}\right) \text{ for a.a. }t \gt 0. \end{array}$$

It is known [17] that for any u0V there exists at least one strong solution u(⋅) such that u(0) = u0. Moreover, every strong solution satisfies the energy inequality

Eut+2stur2dr=Eus, for all ts0,$$\begin{array}{} \displaystyle E\left( u\left( t\right) \right) +2\int_{s}^{t}\left\Vert u_{r}\right\Vert ^{2}dr=E\left( u\left( s\right) \right) \text{, for all }t\geq s\geq0, \end{array}$$

where E(u(t)) = utV2$\begin{array}{} \displaystyle \left\Vert u\left( t\right) \right\Vert _{V}^{2} \end{array}$ + 2(F(u(t)), 1) − 2(h, u(t)).

Let

Ks+={u:u is a strong solution of (14)}.$$\begin{array}{} \displaystyle K_{s}^{+}=\{u\left(\cdot\right) :u\text{ is a strong solution of (14)}\}. \end{array}$$

We define now the multivalued map G : ℝ+ × VP(V), where P(V) is the set of all non-empty subsets of V, by

G(t,u0)={ut:uKs+ and u0=u0}.$$\begin{array}{} \displaystyle G(t,u_{0})=\{u\left( t\right) :u\in K_{s}^{+}\text{ and }u\left( 0\right) =u_{0}\}. \end{array}$$

We recall now some results proved in [17] about the existence and structure of a global attractor for G. It is worth pointing out that, although in that paper such theorems were proved in the particular three-dimensional case (i.e. N = 3), the proofs in the general N-dimensional setting are identical.

First, we note that G is a strict multivalued semiflow, that is, G(0, x) = x and G(t + s, x) = G(t, G(s, x)) for any t, s ∈ ℝ+, xV.

Moreover, G possesses a global compact invariant attractor 𝓐, that is, 𝓐 is compact in V, it is invariant (i.e. 𝓐 = G(t, 𝓐) for any t ≥ 0) and attracts every bounded set in V, that is,

dist(G(t,B),A)0 as t+, $$\begin{array}{} \displaystyle dist(G(t,B),\mathscr{A)}\rightarrow0\text{ as }t\rightarrow+\infty\text{, } \end{array}$$

for any B set bounded in V, where dist refers to the Hausdorff semidistance between sets given by dist(A, B) = supxA infyBxyV.

The map γ : ℝ → V is called a complete trajectory of Ks+$\begin{array}{} \displaystyle K_{s}^{+} \end{array}$ if γ(⋅ + h)|[0,+∞)Ks+$\begin{array}{} \displaystyle K_{s}^{+} \end{array}$ for any h ∈ ℝ, and this is equivalent to γ being continuous and satisfying

γt+sGst,γs for all sR and t0.$$\begin{array}{} \displaystyle \gamma\left( t+s\right) \in G_{s}\left( t,\gamma\left( s\right) \right) \text{ for all }s\in\mathbb{R}\text{ and }t\geq0. \end{array}$$

The set of all complete trajectories of Ks+$\begin{array}{} \displaystyle K_{s}^{+} \end{array}$ will be denoted by 𝔽s. Let 𝕂s be the set of all complete trajectories which are bounded in V.

The attractor 𝓐 is characterised by the union of all points lying in a bounded complete trajectory, that is,

A={γ0:γKs}=tR{γt:γKs}.$$\begin{array}{} \displaystyle \mathscr{A}=\{\gamma\left( 0\right) :\gamma\left( \cdot\right) \in\mathbb{K}_{s}\}=\cup_{t\in\mathbb{R}}\{\gamma\left( t\right) :\gamma\left( \cdot\right) \in\mathbb{K}_{s}\}. \end{array}$$

A point zX is a fixed point of Ks+$\begin{array}{} \displaystyle K_{s}^{+} \end{array}$ if φ(t) ≡ zKs+$\begin{array}{} \displaystyle K_{s}^{+} \end{array}$ , whereas it is called a fixed point of G if zG(t, z) for all t ≥ 0. In our case these two concepts are equivalent, so we will simply call them fixed points. Moreover, z is a fixed point if and only if zVH2(Ω) and

Δz+f(z)=h in H.$$\begin{array}{} \displaystyle -\Delta z+f(z)=h\text{ in }H. \end{array}$$

The set of all fixed points will be denoted by ℜ.

Finally, in [17] it is proved that the strong global attractor coincides with the unstable manifold of the set of stationary points, and also with the stable one when we consider only bounded complete solutions. This means that

Θs=Ms+(R)=Ms(R),$$\begin{array}{} \displaystyle \Theta_{s}=M_{s}^{+}(\mathfrak{R})=M_{s}^{-}(\mathfrak{R}), \end{array}$$

where

Ms(R)=z:γ()Ks,γ(0)=z,dist(γ(t),R)0,t+,Ms+(R)=z:γ()Fs,γ(0)=z,dist(γ(t),R)0,t.$$\begin{array}{} \displaystyle M_{s}^{-}(\mathfrak{R})=\left\{ z\,:\,\exists\gamma(\cdot)\in\mathbb{K} _{s},\,\,\gamma(0)=z,\,\,\,\, dist(\gamma(t),\mathfrak{R})\rightarrow 0,\,\, t\rightarrow+\infty\right\} ,\\ M_{s}^{+}(\mathfrak{R})=\left\{ z\,:\,\exists\gamma(\cdot)\in\mathbb{F} _{s},\,\, \gamma(0)=z,\,\,\,\,dist(\gamma(t),\mathfrak{R})\rightarrow 0,\,\, t\rightarrow-\infty\right\} . \end{array}$$

Concerning boundedness of the attractor in stronger Lr-spaces than L2NN2$\begin{array}{} \displaystyle L^{\frac{2N}{N-2}} \end{array}$ (Ω) (which follows from the embedding VL2NN2$\begin{array}{} \displaystyle L^{\frac{2N}{N-2}} \end{array}$ (Ω)) in [17] and [18] an estimate in the space L(Ω) is shown to be true if hL(Ω).

Now, using the equivalence between weak and mild solutions from the first section we are able to obtain suitable estimates in Lr-spaces under much weaker assumptions on the function h.

Theorem 4

The global attractor 𝓐 is bounded inLr(Ω), wherer ∈ [1, +∞] ifN ≤3, r ∈ [1, +∞) ifN = 4 and 1 ≤ r < 2NN4$\begin{array}{} \displaystyle \frac{2N}{N-4} \end{array}$ifN ≥ 5.

IfN ≥ 4 andhLq(Ω), for someq > N2$\begin{array}{} \displaystyle \frac{N}{2} \end{array}$ , then 𝓐 is bounded inL(Ω).

Proof

The semigroup T(t) generated by the solutions of problem (10) satisfies the following well-known estimate [32, p.2988]:

T(t)xLr(Ω)MeδttN21q1rxLq(Ω),$$\begin{array}{} \displaystyle \left\Vert T(t)x\right\Vert _{L^{r}(\Omega)}\leq M\frac{e^{\delta t}} {t^{\frac{N}{2}\left( \frac{1}{q}-\frac{1}{r}\right) }}\left\Vert x\right\Vert _{L^{q}(\Omega)}, \end{array}$$

for every 1 ≤ qr ≤∞ and certain M > 0, δ ∈ ℝ.

Take an arbitrary y ∈ 𝓐. In view of (23), there exists a complete trajectory ψ such that ψ(0) = y and ψ(t) ∈ 𝓐 for all t ∈ ℝ. Inequality (18), combined with (15) and Lemma 2, implies that u(⋅) = ψ(⋅ + h) is a mild solution to problem (14) for any h ∈ ℝ. We choose h = −1 and then by the variation of constants formula we have

y=u(1)=T1u0+01T(1s)hf(u(s)ds.$$\begin{array}{} \displaystyle y=u(1)=T\left( 1\right) u\left( 0\right) +\int_{0}^{1}T(1-s)\left( h-f(u(s)\right) ds. \end{array}$$

Hence, applying (26) with q = 2 it follows that

yLr(Ω)Meδu0+01Meδs1sN2121rh+f(u(s)ds$$\begin{array}{} \displaystyle \left\Vert y\right\Vert _{L^{r}(\Omega)}\leq Me^{\delta}\left\Vert u\left( 0\right) \right\Vert +\int_{0}^{1}M\frac{e^{\delta s}}{\left( 1-s\right) ^{\frac{N}{2}\left( \frac{1}{2}-\frac{1}{r}\right) }}\left( \left\Vert h\right\Vert +\left\Vert f(u(s)\right\Vert \right) ds \end{array}$$

for any r ≥ 2. Take an arbitrary value r satisfying r ∈ [2, +∞] if N ≤3, r ∈ [2, +∞) if N = 4 and 2 ≤ r < 2NN4$\begin{array}{} \displaystyle \frac{2N}{N-4} \end{array}$ if N ≥ 5. The inequality ∥f(u(s)∥2R1(1 + utV2p2$\begin{array}{} \displaystyle \left\Vert u\left( t\right) \right\Vert _{V}^{2p-2} \end{array}$ ), together with the boundedness of the attractor in the space V, gives the existence of a constant R2 such that

yLr(Ω)R21+0111sN2121rds=R21+1N21r12+1,$$\begin{array}{} \displaystyle \left\Vert y\right\Vert _{L^{r}(\Omega)}\leq R_{2}\left( 1+\int_{0}^{1} \frac{1}{\left( 1-s\right) ^{\frac{N}{2}\left( \frac{1}{2}-\frac{1} {r}\right) }}ds\right) =R_{2}\left( 1+\frac{1}{\frac{N}{2}\left( \frac {1}{r}-\frac{1}{2}\right) +1}\right) , \end{array}$$

where we have used that due to the conditions imposed on the parameter r, it follows that N21r12+1>0$\begin{array}{} \displaystyle \frac{N}{2}\left( \frac{1}{r}-\frac{1}{2}\right) +1 \gt 0 \end{array}$.

Thus the first statement of the theorem is proved.

In order to prove the second one, we first will prove that if the global attractor is bounded in Ls(Ω), where s satisfies

sN2N>N2,$$\begin{array}{} \displaystyle s\frac{N-2}{N} \gt \frac{N}{2}, \end{array}$$

then it is bounded in fact in L(Ω). By (4) and (17) we get

Ωf(u(s,x)sN2NdxR3(1+Ωu(s,x)sN2Np1dx)R3(1+Ωu(s,x)sdx)R4.$$\begin{array}{} \displaystyle \int_{\Omega}\left\vert f(u(s,x)\right\vert ^{s\frac{N-2}{N}}dx \leq R_{3}(1+\int_{\Omega}\left\vert u(s,x)\right\vert ^{s\frac{N-2}{N}\left( p-1\right) }dx)\\\displaystyle \qquad\qquad\qquad\qquad~~ \leq R_{3}(1+\int_{\Omega}\left\vert u(s,x)\right\vert ^{s}dx)\leq R_{4}. \end{array}$$

Therefore, applying again formula (26) with r = ∞ and q=min{q¯,sN2N}$\begin{array}{} \displaystyle q=\min\{\overline{q},s\frac{N-2}{N}\} \end{array}$ and arguing as before for any y ∈ 𝓐 we have

yL(Ω)Meδu0Lq(Ω)+01Meδs1sN21qhLq(Ω)+f(u(s)Lq(Ω)dsR51+0111sN21qds=R51+1N2q+1,$$\begin{array}{} \displaystyle \left\Vert y\right\Vert _{L^{\infty}(\Omega)} \leq Me^{\delta}\left\Vert u\left( 0\right) \right\Vert _{L^{q}(\Omega)}+\int_{0}^{1}M\frac{e^{\delta s}}{\left( 1-s\right) ^{\frac{N}{2}\frac{1}{q}}}\left( \left\Vert h\right\Vert _{L^{q}(\Omega)}+\left\Vert f(u(s)\right\Vert _{L^{q}(\Omega )}\right) ds\\ \displaystyle\qquad\qquad \leq R_{5}\left( 1+\int_{0}^{1}\frac{1}{\left( 1-s\right) ^{\frac{N} {2}\frac{1}{q}}}ds\right) =R_{5}\left( 1+\frac{1}{-\frac{N}{2q}+1}\right) , \end{array}$$

where we have used that N2q+1>0.$\begin{array}{} \displaystyle -\frac{N}{2q}+1 \gt 0. \end{array}$ Thus, the result follows.

Observe that if N = 4, then the attractor is bounded in Ls(Ω) for an arbitrary s ∈ [2, ∞). Hence, we can pick s such that (27) holds. On the other hand, since we have proved that the attractor is bounded in Ls(Ω) for any s < 2NN4$\begin{array}{} \displaystyle \frac{2N}{N-4} \end{array}$ , then (27) is also satisfied if N = 5, 6.

Further, for N ≥ 7 we will apply the above procedure iteratively so as to achieve (27).

Assume that the attractor is bounded in Ls(Ω) for some s2NN2such thatsN2N<N2.$\begin{array}{} \displaystyle s\geq \frac{2N}{N-2}\,\, \text{such that}\,\, s\frac{N-2}{N} \lt \frac{N}{2}. \end{array}$ Using (28) and applying again (26) with q=sN2Nands¯<s(N2)NN22s(N2)$\begin{array}{} \displaystyle q=s\frac{N-2}{N}\,\, \text{and}\,\, \overline {s} \lt \frac{s(N-2)N}{N^{2}-2s(N-2)} \end{array}$ for any y ∈ 𝓐 we obtain

yLs¯(Ω)R61+0111sN2NrN21s¯ds=R6N21s¯NsN2+1.$$\begin{array}{} \displaystyle \left\Vert y\right\Vert _{L^{\overline{s}}(\Omega)}\leq R_{6}\left( 1+\int_{0}^{1}\frac{1}{\left( 1-s\right) ^{\frac{N}{2}\left( \frac {N}{r\left( N-2\right) }-\frac{1}{\overline{s}}\right) }}ds\right) =\frac{R_{6}}{\frac{N}{2}\left( \frac{1}{\overline{s}}-\frac{N}{s\left( N-2\right) }\right) +1}. \end{array}$$

Therefore, the attractor is bounded in Ls(Ω) as well. For an arbitrary ε > 0 we choose s such that the difference ss satisfies

s¯ss(N2)NN22s(N2)sε.$$\begin{array}{} \displaystyle \overline{s}-s\geq\frac{s(N-2)N}{N^{2}-2s(N-2)}-s-\varepsilon. \end{array}$$

Since s2NN2,$\begin{array}{} \displaystyle s\geq\frac{2N}{N-2}, \end{array}$ we get

s¯ss(N2)NN24N1ε=s2N4ε4NN2N4ε.$$\begin{array}{} \displaystyle \overline{s}-s\geq s\left( \frac{(N-2)N}{N^{2}-4N}-1\right) -\varepsilon =s\frac{2}{N-4}-\varepsilon\geq\frac{4N}{\left( N-2\right) \left( N-4\right) }-\varepsilon. \end{array}$$

There exist ε > 0 and k ∈ ℕ such that

2NN2+k14NN2N4ε<N22(N2)<2NN2+k4NN2N4ε$$\begin{array}{} \displaystyle \frac{2N}{N-2}+\left( k-1\right) \left( \frac{4N}{\left( N-2\right) \left( N-4\right) }-\varepsilon\right) \lt \frac{N^{2}}{2(N-2)} \lt \frac{2N} {N-2}+k\left( \frac{4N}{\left( N-2\right) \left( N-4\right) } -\varepsilon\right) \end{array}$$

and thus proceeding iteratively we obtain in k steps that the global attractor is bounded in Ls(Ω), where s satisfies (27).

Weak attractor for strong solutions in the supercritical case

In this section, our aim is to prove the existence of a weak global attractor for the multivalued semiflow generated by strong solutions to problem (14) satisfying a suitable energy inequality without imposing the assumption (17). In this case, we do not know whether strong solutions belong to the space of continuous functions with values in V, and therefore we are still not able to prove the existence of a strong attractor. Instead, we have to consider a weaker attractor in which the attracting property is satisfied with respect to the weak topology of the space VLp(Ω).

Lemma 5

LetuτVLp(Ω). Then there exists at least one strong solutionuof (1) such thatu(τ) = uτ. Moreover, the energy inequality

E(u(t))+stdudr2drE(u(s))$$\begin{array}{} \displaystyle E(u(t))+\int_{s}^{t}\left\Vert \frac{du}{dr}\right\Vert ^{2}dr\leq E(u(s)) \end{array}$$

holds for allts, a.a. s > 0 ands = 0, whereE(u(t))=utV2+F(u(t)),12h,ut.$\begin{array}{} \displaystyle E(u(t))=\left\Vert u\left( t\right) \right\Vert _{V}^{2}+\left( F(u(t)),1\right) -2\left( h,u\left( t\right) \right) . \end{array}$

Proof

As usual, we take the Galerkin approximations using the basis of eigenfunctions {wj(x), j ∈ ℕ} of −Δ with Dirichlet boundary conditions. Let Xm = span{w1, …, wm} and let Pm be the orthogonal projector from H onto Xm. Then umt,x=j=imaj,mtwjx$\begin{array}{} u_{m}\left( t,x\right) =\sum_{j=i}^{m}a_{j,m}\left( t\right) w_{j}\left( x\right) \end{array}$ will be a solution of the system of ordinary differential equations

dumdt=PmΔumPmfum+Pmh,um0=Pmu0.$$\begin{array}{} \displaystyle \frac{du_{m}}{dt}=P_{m}\Delta u_{m}-P_{m}f\left( u_{m}\right) +P_{m} h,\, u_{m}\left( 0\right) =P_{m}u_{0}. \end{array}$$

It is proved in [10, p.281] that passing to a subsequence um converges to a weak solution u of (1) weakly star in L(0, T; H), weakly in Lp(0, T; Lp(Ω)) and weakly in L2(0, T; V) for all T > 0. Also, u(τ) = uτ and dumdtdudt$\begin{array}{} \displaystyle \dfrac{du_{m}} {dt}\rightarrow\dfrac{du}{dt} \end{array}$ weakly in Lq(0, T; Hs(Ω)) for some s > 0.

Multiplying (30) by dumdt$\begin{array}{} \displaystyle \dfrac{du_{m}}{dt} \end{array}$ we get

ddtumV2+2(F(um),1)2h,um+2dumdt2=0,$$\begin{array}{} \displaystyle \frac{d}{dt}\left( \Vert u_{m}\Vert_{V}^{2}+2(F(u_{m}),1)-2\left( h,u_{m}\right) \right) +2\Vert\frac{du_{m}}{dt}\Vert^{2}=0, \end{array}$$

so (6) implies

um(t)V2+20tddsum(s)2ds+2δutLp(Ω)pum0V2+R1um0Lp(Ω)p+2hum(t)+2hum(0)+R2.$$\begin{array}{} \displaystyle \qquad\qquad~~\Vert u_{m}(t)\Vert_{V}^{2}+2\int\limits_{0}^{t}\Vert\frac{d}{ds}u_{m} (s)\Vert^{2}ds+2\delta\left\Vert u\left( t\right) \right\Vert _{L^{p} (\Omega)}^{p} \\ \leq\Vert u_{m}\left( 0\right) \Vert_{V}^{2}+R_{1}\Vert u_{m}\left( 0\right) \Vert_{L^{p}(\Omega)}^{p}+2\left\Vert h\right\Vert \left\Vert u_{m}(t)\right\Vert +2\left\Vert h\right\Vert \left\Vert u_{m}(0)\right\Vert +R_{2}. \end{array}$$

If p > 2, δ > 0 implies that

umtV2+0tddsum(s)2ds+utLp(Ω)pR3um0V2+um0Lp(Ω)p+R4,$$\begin{array}{} \displaystyle \Vert u_{m}\left( t\right) \Vert_{V}^{2}+\int\limits_{0}^{t}\Vert\frac {d}{ds}u_{m}(s)\Vert^{2}ds+\left\Vert u\left( t\right) \right\Vert _{L^{p}(\Omega)}^{p}\leq R_{3}\left( \Vert u_{m}\left( 0\right) \Vert _{V}^{2}+\Vert u_{m}\left( 0\right) \Vert_{L^{p}(\Omega)}^{p}\right) +R_{4}, \end{array}$$

where Rj > 0. When p = 2 we obtain that

umtV2+0tddsum(s)2dsR3um0V2+R4.$$\begin{array}{} \displaystyle \Vert u_{m}\left( t\right) \Vert_{V}^{2}+\int\limits_{0}^{t}\Vert\frac {d}{ds}u_{m}(s)\Vert^{2}ds\leq R_{3}\Vert u_{m}\left( 0\right) \Vert_{V} ^{2}+R_{4}. \end{array}$$

By the choise of the special basis we have that um(0) → u0 in H01$\begin{array}{} \displaystyle H_{0}^{1} \end{array}$ (Ω) ∩ Lp(Ω).

Hence, umu weakly star in L(0, T; VLp (Ω)) and dumdtdudt$\begin{array}{} \displaystyle \dfrac{du_{m}}{dt}\rightarrow\dfrac{du}{dt} \end{array}$ weakly in L2(0, T; L2(Ω)). Thus from the Ascoli-Arzelà theorem {um} is pre-compact in C([0, T]; H) and then umu in C([0, T]; H). Moreover, it follows that um (t) → u(t) in H0w1(Ω)Lwp(Ω))$\begin{array}{} \displaystyle H_{0w}^{1}(\Omega)\cap L_{w}^{p}(\Omega)) \end{array}$ for any t ∈ [0, T].

Finally, we must check the validity of the energy inequality (29). It is clear from (31) that um satisfy

E(um(t))+stdumdr2drE(um(s))$$\begin{array}{} \displaystyle E(u_{m}(t))+\int_{s}^{t}\left\Vert \frac{du_{m}}{dr}\right\Vert ^{2}dr\leq E(u_{m}(s)) \end{array}$$

for all ts ≥ 0.

Let us define the function

L(u)=f(u)uα|u|p.$$\begin{array}{} \displaystyle L(u)=f(u)u-\alpha|u|^{p}. \end{array}$$

Multiplying the equation in (14) by tum(t) and integrating over (0, T) × Ω we obtain

0TΩtdumdtumdxdt+0Ttum(t)V2dt+0TΩtfumt,xumt,xdxdt0TΩthxumt,xdxdt=0,$$\begin{array}{} \displaystyle \int_{0}^{T}\int_{\Omega}t\frac{du_{m}}{dt}u_{m}dxdt+\int_{0} ^{T}t\left\Vert u_{m}(t)\right\Vert _{V}^{2}dt\\ \displaystyle +\int_{0}^{T}\int_{\Omega}tf\left( u_{m}\left( t,x\right) \right) u_{m}\left( t,x\right) dxdt-\int_{0}^{T}\int_{\Omega}th\left( x\right) u_{m}\left( t,x\right) dxdt=0, \end{array}$$

so

T2umT2+0Ttum(t)V2dt+0TΩtL(um(t,x))dxdt+α0TtumtLp(Ω)pdt=120Tumt2dt+0TΩthxumt,xdxdt.$$\begin{array}{} \displaystyle \frac{T}{2}\left\Vert u_{m}\left( T\right) \right\Vert ^{2}+\int_{0} ^{T}t\left\Vert u_{m}(t)\right\Vert _{V}^{2}dt\\ \displaystyle +\int_{0}^{T}\int_{\Omega}tL(u_{m}(t,x))dxdt+\alpha\int_{0}^{T}t\left\Vert u_{m}\left( t\right) \right\Vert _{L^{p}(\Omega)}^{p}dt\\ \displaystyle =\frac{1}{2}\int_{0}^{T}\left\Vert u_{m}\left( t\right) \right\Vert ^{2}dt+\int_{0}^{T}\int_{\Omega}th\left( x\right) u_{m}\left( t,x\right) dxdt. \end{array}$$

In the same way, for the limit function u we obtain

T2uT2+0Ttu(t)V2dt+0TΩtL(u(t,x))dxdt+α0TtutLp(Ω)pdt=120Tut2dt+0TΩthxut,xdxdt.$$\begin{array}{} \displaystyle \frac{T}{2}\left\Vert u\left( T\right) \right\Vert ^{2}+\int_{0} ^{T}t\left\Vert u(t)\right\Vert _{V}^{2}dt\\ \displaystyle +\int_{0}^{T}\int_{\Omega}tL(u(t,x))dxdt+\alpha\int_{0}^{T}t\left\Vert u\left( t\right) \right\Vert _{L^{p}(\Omega)}^{p}dt\\ \displaystyle =\frac{1}{2}\int_{0}^{T}\left\Vert u\left( t\right) \right\Vert ^{2}dt+\int_{0}^{T}\int_{\Omega}th\left( x\right) u\left( t,x\right) dxdt. \end{array}$$

From the previous convergences it is clear that

0Tumt2dt0Tut2dt,0TΩthxumt,xdxdt0TΩthxut,xdxdt,$$\begin{array}{} \displaystyle \qquad\qquad\int_{0}^{T}\left\Vert u_{m}\left( t\right) \right\Vert ^{2}dt \rightarrow\int_{0}^{T}\left\Vert u\left( t\right) \right\Vert ^{2}dt,\\ \displaystyle\int_{0}^{T}\int_{\Omega}th\left( x\right) u_{m}\left( t,x\right) dxdt \rightarrow\int_{0}^{T}\int_{\Omega}th\left( x\right) u\left( t,x\right)dxdt, \end{array}$$

which implies that the left-hand side of (32) converges to the left-hand side of (33). On the other hand, we have

0TΩtL(u(t,x))dxdtliminf0TΩtL(um(t,x))dxdt,$$\begin{array}{} \displaystyle \int_{0}^{T}\int_{\Omega}tL(u(t,x))dxdt\leq\lim\inf\, \int_{0}^{T}\int_{\Omega }tL(u_{m}(t,x))dxdt, \end{array}$$

which follows from the inequality

L(um(t,x))C2,$$\begin{array}{} \displaystyle L(u_{m}(t,x))\geq-C_{2}, \end{array}$$

the convergence um(t, x) → u(t, x), for a.a. (t, x), and Lebesgue-Fatou’s lemma [35].

Bearing also in mind that

umT2uT2,0Ttu(t)V2dtliminf0Ttum(t)V2dt,0TtutLp(Ω)pdtsliminf0TtumtLp(Ω)pdt,$$\begin{array}{} \displaystyle \qquad\qquad\,\,\left\Vert u_{m}\left( T\right) \right\Vert ^{2} \rightarrow\left\Vert u\left( T\right) \right\Vert ^{2},\\\displaystyle\qquad \int_{0}^{T}t\left\Vert u(t)\right\Vert _{V}^{2}dt \leq\lim\inf\, \int _{0}^{T}t\left\Vert u_{m}(t)\right\Vert _{V}^{2}dt,\\\displaystyle ~\int_{0}^{T}t\left\Vert u\left( t\right) \right\Vert _{L^{p}(\Omega)} ^{p}dts \leq\lim\inf\, \int_{0}^{T}t\left\Vert u_{m}\left( t\right) \right\Vert _{L^{p}(\Omega)}^{p}dt, \end{array}$$

we obtain readily that each term in the the left-hand side of (32) converges to the corresponding term in the left-hand side of (33).

Therefore,

0TtumtLp(Ω)pdt0TtutLp(Ω)pdt,0Ttum(t)V2dt0Ttum(t)V2dt,$$\begin{array}{} \displaystyle \int_{0}^{T}t\left\Vert u_{m}\left( t\right) \right\Vert _{L^{p}(\Omega )}^{p}dt\rightarrow\int_{0}^{T}t\left\Vert u\left( t\right) \right\Vert _{L^{p}(\Omega)}^{p}dt,\\ \displaystyle\quad\,\int_{0}^{T}t\left\Vert u_{m}(t)\right\Vert _{V}^{2}dt\rightarrow\int_{0} ^{T}t\left\Vert u_{m}(t)\right\Vert _{V}^{2}dt, \end{array}$$

and then for any 0 < r < T we get

umustrongly in L2(r,T;V)Lp(r,T;Lp(Ω)),$$\begin{array}{} \displaystyle u_{m}\rightarrow u\, \text{strongly in }L^{2}(r,T;V)\cap L^{p}(r,T;L^{p} (\Omega)), \end{array}$$

so

um(t)ut in VLp(Ω) for a.a. t0,T.$$\begin{array}{} \displaystyle u_{m}(t)\rightarrow u\left( t\right) \text{ in }V\cap L^{p}(\Omega)\text{ for a.a. }t\in\left( 0,T\right) . \end{array}$$

Applying then the dominated convergence theorem we deduce that F(um(t)) → F(u(t)) in L1(Ω) for a.a. t ∈ (0, T). On top of that, in view of umu in C([0, T], H), for any t ∈ [0, T] it follows that F(um(t, x)) → F(u(t, x)) for a.a. x ∈ Ω. Hence, by the inequality

F(um(t,x))D2 if p>2,F(um(t,x))δum(t,x)2D2 if p=2,$$\begin{array}{} \displaystyle F(u_{m}(t,x)) \geq-D_{2}\text{ if }p \gt 2,\\ F(u_{m}(t,x)) \geq\delta\left\vert u_{m}(t,x)\right\vert ^{2}-D_{2}\text{ if }p=2, \end{array}$$

and Lebesgue-Fatou’s lemma [35], we have

ΩF(u(t,x))dxliminfΩF(um(t,x))dx.$$\begin{array}{} \displaystyle \int_{\Omega}F(u(t,x))dx\leq\lim\inf\, \int_{\Omega}F(u_{m}(t,x))dx. \end{array}$$

We know also from um(t) → u(t) weakly in V that

utVliminfum(t)V for all t[0,T].$$\begin{array}{} \displaystyle \left\Vert u\left( t\right) \right\Vert _{V}\leq\lim\inf\, \left\Vert u_{m}(t)\right\Vert _{V}\text{ for all }t\in\lbrack0,T]. \end{array}$$

Thus,

E(u(t))liminfE(um(t)) for all t[0,T].$$\begin{array}{} \displaystyle E(u(t))\leq\lim\inf\, E(u_{m}(t))\text{ for all }t\in\lbrack0,T]. \end{array}$$

Passing to the limit in (21) we obtain (29).

For any u0VLp(Ω) let 𝓡(u0) be the set of strong solutions to problem (14) which satisfy the energy inequality (29) for all ts and a.a. s > 0. This set is non-empty for every initial datum u0VLp(Ω) due to Lemma 5. Denoting X = VLp(Ω) we define now the multivalued mapping G : ℝ+ × XP(X) by

G(t,u0)={y:y=u(t)R(u0)}.$$\begin{array}{} \displaystyle G(t,u_{0})=\{y:y=u(t)\in\mathscr{R}(u_{0})\}. \end{array}$$

It is straightforward to check that G(0, x) = x and G(t + s, x) ⊂ G(t, G(s, x)) for any xX and t, s ∈ ℝ+, that is, G is a multivalued semiflow.

In the sequel, Xw=H0w1ΩLwp(Ω).$\begin{array}{} \displaystyle X_{w}=H_{0w}^{1}\left( \Omega\right) \cap L_{w}^{p}(\Omega). \end{array}$

Definition 5

The set 𝓐 is called a weak global attractor for the multivalued semiflow G if the following properties hold:

𝓐 is bounded in X and compact in Xw.

𝓐 is weakly attracting, that is, for any bounded set B in X and any neighborhood 𝓞 of 𝓐 in Xw there exists T(𝓞) such that

G(t,B)O for all tT.$$\begin{array}{} \displaystyle G(t,B)\subset\mathscr{O}\text{ for all }t\geq T. \end{array}$$

𝓐 is negatively semi-invariant, that is,

AG(t,A) for any t0.$$\begin{array}{} \displaystyle \mathscr{A\subset}G(t,\mathscr{A)}\text{ for any }t\geq0. \end{array}$$

It follows from this definition that if K is a weakly closed set which is weakly attracting, then for any neighborhood 𝓞 of K in Xw there exists T(𝓞) such that

AG(t,A)O for all tT.$$\begin{array}{} \displaystyle \mathscr{A\subset}G(t,\mathscr{A)\subset O}\text{ for all }t\geq T. \end{array}$$

Hence, 𝓐 ⊂ 𝓚 as Xw is a Hausdorff topological space. This means that 𝓐 is the minimal weakly closed set which is weakly attracting.

We recall that the set B0 is said to be absorbing if for any bounded set B in X there exists T(B) such that

G(t,B)B0 for any tT.$$\begin{array}{} \displaystyle G(t,B)\subset B_{0}\text{ for any }t\geq T. \end{array}$$

Lemma 6

The multivalued semiflowGpossesses a bounded absorbing setB0={vX:vV2+vLp(Ω)pR},$\begin{array}{} \displaystyle B_{0}=\{v\in X:\left\Vert v\right\Vert _{V}^{2}+\left\Vert v\right\Vert _{L^{p}(\Omega)}^{p}\leq R\}, \end{array}$whereRis a positive constant.

Remark 1

The set B0 is weakly closed.

Proof

The function E(u(t)) satisfies (29) and multiplying the equation in (14) by u we have

12ddtu2+uV2+αuLp(Ω)pC2+12λ1h2+λ12u2.$$\begin{array}{} \displaystyle \frac{1}{2}\frac{d}{dt}\left\Vert u\right\Vert ^{2}+\left\Vert u\right\Vert _{V}^{2}+\alpha\left\Vert u\right\Vert _{L^{p}(\Omega)}^{p}\leq C_{2}+\frac {1}{2\lambda_{1}}\left\Vert h\right\Vert ^{2}+\frac{\lambda_{1}}{2}\left\Vert u\right\Vert ^{2}. \end{array}$$

Therefore,

u(t)2eλ1tu02+K1, for all t0,$$\begin{array}{} \displaystyle \left\Vert u(t)\right\Vert ^{2}\leq e^{-\lambda_{1}t}\left\Vert u\left( 0\right) \right\Vert ^{2}+K_{1},\text{ for all }t\geq0, \end{array}$$

where K1=2C2λ1+1λ12h2,$\begin{array}{} \displaystyle K_{1}=\frac{2C_{2}}{\lambda_{1}}+\frac{1}{\lambda_{1}^{2}}\left\Vert h\right\Vert ^{2}, \end{array}$ and

tt+ruV2+uLp(Ω)pdtK2(u(t)2+r)K3(eλ1tu02+1+r),$$\begin{array}{} \displaystyle \int_{t}^{t+r}\left( \left\Vert u\right\Vert _{V}^{2}+\left\Vert u\right\Vert _{L^{p}(\Omega)}^{p}\right) dt \leq K_{2}(\left\Vert u(t)\right\Vert ^{2}+r)\\ \displaystyle \qquad\qquad\qquad\qquad\qquad\quad~ \leq K_{3}(e^{-\lambda_{1}t}\left\Vert u\left( 0\right) \right\Vert ^{2}+1+r), \end{array}$$

for some positive constants K2, K3 and any r > 0, t ≥ 0.

Integrating in (29) over (t, t + r) we get

E(t+r)K3(eλ1tu02+1+r)r, for all r>0,t0.$$\begin{array}{} \displaystyle E(t+r)\leq\frac{K_{3}(e^{-\lambda_{1}t}\left\Vert u\left( 0\right) \right\Vert ^{2}+1+r)}{r}\text{, for all }r \gt 0,t\geq0. \end{array}$$

Taking into account that

E(u(t))=utV2+F(u(t)),12h,ututV2+δu(t)Lp(Ω)ph2ut2D2,$$\begin{array}{} \displaystyle E(u(t)) =\left\Vert u\left( t\right) \right\Vert _{V}^{2}+\left( F(u(t)),1\right) -2\left( h,u\left( t\right) \right) \\ \displaystyle\qquad\quad~ \geq\left\Vert u\left( t\right) \right\Vert _{V}^{2}+\delta\left\Vert u(t)\right\Vert _{L^{p}(\Omega)}^{p}-\left\Vert h\right\Vert ^{2}-\left\Vert u\left( t\right) \right\Vert ^{2}-D_{2}, \end{array}$$

it follows easily from the previous estimates the existence of a constant R > 0 such that the bounded set

B0={vX:vV2+vLp(Ω)pR}$$\begin{array}{} \displaystyle B_{0}=\{v\in X:\left\Vert v\right\Vert _{V}^{2}+\left\Vert v\right\Vert _{L^{p}(\Omega)}^{p}\leq R\} \end{array}$$

is absorbing for G.

The theory of global attractors in topological spaces for multivalued semiflows and processes has been developed for example in [1, 15, 21]. However, since the abstract conditions imposed in those papers are not met in our particular situation, we will prove the existence of the weak attractor from scratch.

Theorem 7

The multivalued semiflowGpossesses a weak global attractor 𝓐.

Proof

Take an arbitrary bounded set B. We recall that the ω − limit set of B is given by

ω(B)=s0clXwtsG(t,B).$$\begin{array}{} \displaystyle \omega(B)=\cap_{s\geq0}cl_{X_{w}}\cup_{t\geq s}G(t,B). \end{array}$$

Since the space Xw satisfies the first axiom of countability, an equivalent definition for ω(B) is the following

ω(B)=yX: there exist sequences ynG(tn,xn), tn+, xnXsuch that yny in Xw.$$\begin{array}{} \displaystyle \omega(B)=\left\{ \begin{array} [c]{c} y\in X:\text{ there exist sequences }y_{n}\in G(t_{n},x_{n})\text{, } t_{n}\rightarrow+\infty\text{, }x_{n}\in X\\ \text{such that }y_{n}\rightarrow y\text{ in }X_{w} \end{array} \right\} . \end{array}$$

Therefore, as X is a reflexive Banach space and by Lemma 6 the set ∪tT(B)G(t, B) is bounded for some T(B), any sequence ynG(tn, B) with tn → +∞ has a weakly convergent subsequence. Thus, ω(B) is non-empty. It is obvious that ω(B) belongs to the absorbing set B0, so it is bounded in X, and that it is weakly closed, so it is compact in Xw.

Further, let us check that ω(B) weakly attracts B. Assuming the opposite, there exists a neighborhood 𝓞 of ω(B) and a sequence ynG(tn, B), where tn → +∞, such that yn ∉ 𝓞. But this leads to a contradiction as from {yn} we can extract a converging subsequence whose limit belongs to ω(B).

It remains to prove that ω(B) is negatively semi-invariant.

First of all, let us consider a sequence of strong solutions un(⋅) ∈ 𝓡( u0n$\begin{array}{} \displaystyle u_{0}^{n} \end{array}$ ) such that un(t) ∈ B0 for any t ≥ 0. In view of well-known results (see [20, Lemma 15] and [16, Theorem 3.11]), there exists a weak solution u(⋅) to problem (14) and a subsequence unk(⋅) such that unku strongly in C([0, T], H), for all T > 0 (among other convergences). But un(t) are uniformly bounded in X, so

unku weakly star in L(0,T;VLp(Ω)) for all T>0,unk(t)ut in Xw for any t0.$$\begin{array}{} \displaystyle u^{n_{k}}\rightarrow u\text{ weakly star in }L^{\infty}(0,T;V\cap L^{p} (\Omega))\text{ for all }T \gt 0,\\\displaystyle\qquad\qquad\quad u^{n_{k}}(t)\rightarrow u\left( t\right) \text{ in }X_{w}\text{ for any }t\geq0. \end{array}$$

Also, inequalities (29), (6) and E(un(s)) ≤ C, for any s ≥ 0 and n, imply that dundt$\begin{array}{} \displaystyle \dfrac{du^{n}}{dt} \end{array}$ are bounded in L2(0, T; H) and then

dunkdtdudt weakly in L2(0,T;H).$$\begin{array}{} \displaystyle \frac{du^{n_{k}}}{dt}\rightarrow\frac{du}{dt}\text{ weakly in }L^{2}(0,T;H). \end{array}$$

Therefore, u(⋅) is in fact a strong solution to problem (14). On top of that, arguing in the same way as in Lemma 5 we obtain that

unkustrongly in L2(r,T;V)Lp(r,T;Lp(Ω)), for any 0<r<T,$$\begin{array}{} \displaystyle u^{n_{k}}\rightarrow u\, \text{strongly in }L^{2}(r,T;V)\cap L^{p} (r,T;L^{p}(\Omega)),\text{ for any }0 \lt r \lt T, \end{array}$$

and also that (29) is satisfied for all ts and a.a. s > 0. Thereby, u(⋅) ∈ 𝓡(u0).

Consider now an arbitrary element yω(B) and t > 0. Then there is a sequence ynG(tn, xn) with tn → +∞, xnB such that yny in Xw. In addition, we take N(B) for which G(s, B) ⊂ B0 if stnt and nN. Since G(tn, xn) ⊂ G(t, G(tnt, xn)), there are znG(tnt, xn), un(⋅) ∈ 𝓡(zn) satisfying that un(s) ∈ B0, for any s ≥ 0, and yn = un(t). Passing to a subsequence zn converges in the space Xw to some zω(B). In light of the previous arguments there exists u(⋅) ∈ 𝓡(z) such that u(t) = y. Thus, yG(t, z) ⊂ G(t, ω(B)), which proves that ω(B) is negatively semi-invariant.

Finally, let 𝓐 = ω(B0). Obviously, 𝓐 is bounded in X, compact in Xw and negatively semi-invariant. The weakly attracting property follows from the chain of inclusions

G(t,B)G(tT(B),G(T(B),B))G(tT(B),B0)$$\begin{array}{} \displaystyle G(t,B)\subset G(t-T(B),G(T(B),B))\subset G(t-T(B),B_{0}) \end{array}$$

and the fact that 𝓐 weakly attracts B0.

eISSN:
2444-8656
Language:
English
Publication timeframe:
2 times per year
Journal Subjects:
Life Sciences, other, Mathematics, Applied Mathematics, General Mathematics, Physics