Open Access

Non-autonomous perturbations of a non-classical non-autonomous parabolic equation with subcritical nonlinearity


Cite

Introduction
Asymptotic behavior of non-autonomous equations

What is the difference between the asymptotic behavior of an autonomous and a non-autonomous equation? This may be, at first glance, a simple question to answer. Let us discuss this problem a little. Consider a general non-autonomous differential equation given by

{u˙=F(t,u),ts,u(s)=u0X,$$\begin{array}{} \displaystyle \left\{ \begin{array}{*{20}{l}} {\dot u = F(t,u),\,t \geqslant s,}\\ {u(s) = {u_0} \in X,} \end{array}\right. \end{array}$$

where X is a Banach space and F is in some metric space ℭ of functions. Assume also that for each u0X, s ∈ ℝ and F ∈ ℭ, the problem (1) has a uniquely defined solution u(t, s, F,u0) for all times ts and the map (t, s, u0) ↦ u(t, s, F, u0) is continuous for each F ∈ ℭ.

We know that, when F is independent of time, u(t, s, F, u0) = u(ts, 0, F, u0), that is, the dependence of t and s in u are artificial, and in fact u depends only on the elapsed time ts. Hence, the asymptotic behavior (the behavior for large times) can be obtained making t — ∞ or s → — ∞, indistinctly.

However, if F is time dependent, the dependence of t and s of the solution is explicit and the scenarios arising from making t → ∞ and s → — ∞ may be completely different. This is not so surprising once we realize that in the autonomous case (F independent of time) we have only one vector field, namely F, driving the solutions but, in the non-autonomous case we have infinitely many vector fields (F(t, ·) for each t) driving the solutions, and their behavior may be completely different for t → ∞ and s → —∞. This clarifies a little the understanding of asymptotic behavior for non-autonomous equations, and shows that it is not an easy task to study this subject. First, one must define which behavior will be treated: the forward attract/on (when t → ∞) or the pullback attract/on (when s → — ∞). Once the framework is set, we reach another problem: in any of them it is clear that the asymptotic behavior of the solutions of (1) are related with the behavior of the vector fields F(t, ·) when t → ±∞, which can be unrelated with the behavior of each F(t, ·). This could have been said in another way: the translates θsF alone are not enough, in general, to describe the asymptotic dynamics of (1) (for more details on this subject we refer to [7] or [9, 16, 17]). Thus, the next question comes quite naturally: how do we introduce the limiting vector fields of F(t, ·) in the study?

Consider ℭ the space of all functions H: ℝ × XX, that are bounded in sets of the form ℝ × B, where B is a bounded set of X. Consider also the shift operatorθt: ℭ → ℭ given by

θtH(,)=H(t+,), for each t.$$\begin{array}{} \displaystyle \theta_tH(\cdot,\cdot)=H(t+\cdot,\cdot), \hbox{ for each }t\in \mathbb{R}. \end{array}$$

Now define Σ0 = {θtF}t∈ℝ, which is the set of all translations of F and let

Σ= closure of Σ0 in ,$$\begin{array}{} \displaystyle \Sigma=\hbox{ closure of } \Sigma_0 \hbox{ in } \mathfrak{C}, \end{array}$$

which is known as the hull of F.

Using our assumptions for the problem (1), we know that each problem

{u˙=H(t,u), for tu(0)=u0X,$$\begin{array}{} \displaystyle \left\{ \begin{array}{*{20}{l}} {\dot u = H(t,u),{\rm{\;for\;}}t \in \mathbb{R}}\\ {u(0) = {u_0} \in X,} \end{array}\right. \end{array}$$

has a uniquely defined solution φ(t, H)u0 for each t ⩾ 0, u0X and H ∈ Σ.

Note that we are now dealing with all the solutions of the problem (1) but also with all the solutions of the limiting vector fields of F(t, ·).

Remark 1.

To obtain problem (1), just consider H = θsF. The solutions ξ of (1) and ψ of (2) in this case are related by

ψ()=ξ(+s).$$\begin{array}{} \displaystyle \psi ( \cdot ) = \xi ( \cdot + s). \end{array}$$

These objects, namely the solutions φ(t, H)u0 in X and the shift operator θt in Σ

Clearly we can consider the restriction of the shift operator θt to Σ.

, give rise to what we call a non-autonomous dynamical system

See Definition 5.

. Several authors have studied this object, in the pursuit of fully understanding of the asymptotic dynamics of equation (1), and there are two distinct branches: the pullback approach (see, for example [16,22]), which deals with pullback attraction, and the uniform approacch (see, for example [28]). Each group has achieved several interesting results concerning asymptotic behavior of non-autonomous equations, and up until recently, these approaches seemed unrelated. In [5,7] the authors unify these results, presenting relations between these frameworks. To this end and to reach the full extent of this theory, they transform the non-autonomous dynamical system defined by φ and θt in an autonomous one, non-trivially, by defining

Π(t)(u0,H)=(φ(t,H)u0,θtH),$$\begin{array}{} \displaystyle \Pi (t)({u_0},H) = (\varphi (t,H){u_0},{\theta _t}H), \end{array}$$

which is called the skew-product semiflow, and study the autonomous semiflow defined by Π to obtain results for φ simply analysing the canonical projection in the first coordinate.

To be a little more precise, inside the study of non-autonomous equations such as (1) we can distinct at least four different notions of attractors, namely:

(i) the global attractor for the skew-product semiflow;

(ii) the pullback attractor for the evolution process.

(iii) the cocyle attractor for the non-autonomous dynamical system and

(iv) the uniform attractor for the non-autonomous system.

We will give a detailed description of each one of these objects in Section 2, as well as the relationships between these concepts, as done in [7], to describe the non-autonomous problems (1) in a very complete way.

Small perturbations

Imagine now that we have not only a single F(t, ·) but a family {Hε(t, ·)}ε∈[0,1] such that Hε is close (in some sense) to F as ε → 0. Are we able to obtain results on the asymptotic behavior of the problems

{u˙=Hϵ(t,u), for t>0u(s)=u0X,$$\begin{array}{} \displaystyle \left\{ \begin{array}{*{20}{l}} {\dot u = {H_\varepsilon }\left( {t,u} \right),{\rm{ for }}t \gt 0} \hfill \\ {u\left( S \right) = {u_0} \in X,} \hfill \\ \end{array}\right. \end{array}$$

for ε sufficiently small, given that we know the behavior of (1)?

Note that to study each problem one must perform all the previous discussion; that is, each ε will generate a different non-autonomous dynamical system and a skew-product semiflow. The question is: can we obtain results of continuity of the different types of asymptotic behavior as ε → 0?

This question, theoretical as is sounds, has a meaning in applications. Models in the real world are always approximations, due to data collection, empirical laws and simplifications, and thus, it is crucial that we are able to transfer properties from an equation to some small perturbations. Without this property, we have no guarantee whatsoever that the real phenomena will have a behavior close to our model.

In [5,7], the authors provide and extensive study on this topic, giving a detailed study of non-autonomous dynamical systems, different scenarios of asymptotic behavior and relationships among them, extracting informations from the skew-product semiflow and transporting them to the non-autonomous dynamical system. Also, the reader can find a deep study of continuity of small perturbations of non-autonomous system, but arising from autonomous equations (see for instance [7, 15, 16]).

The study of non-autonomous perturbations of non-autonomous dynamical systems directly is still an almost blank page, and in this paper we give some steps in this direction, by studying non-autonomous perturbations of a non-autonomous equation, to provide results of continuity of the asymptotic behavior using the framework discussed above.

It is not known, so far, how to do the general theory when we consider non-autonomous perturbations of a non-autonomous system. In examples, we can see however that few steps are clear: first one must be able to prove the global existence and uniqueness of solutions not only for the equation in question, but also for all the limiting vector fields associated with the non-linearity - this step is the key for the development of the following results - to be able to construct a non-autonomous system. Then we must be able, with some uniformity on the vector fields, to obtain an uniform estimate that allow us to find a compact set that attracts all the solutions, independently of the vector field. Once this is done, we can find such an attractor for the associated skew-product semiflow and with this object at hand, we can try to understand all the asymptotic behaviors of our equation. Dealing with perturbations adds a difficulty to this process, since we must be able to do such study for each ε small and obtain the result with uniformity in this parameter.

Non-autonomous non-classical parabolic equations

As mentioned before, a general theory for the study of non-autonomous perturbations of non-autonomous models is not available at this point, so we will put our best efforts to understand in some elaborated examples, in order to obtain a deep understanding of such perturbations. The problem we will deal with in this paper is to study non-autonomous perturbations of some non-autonomous parabolic equations.

Non-classical parabolic equations arise as models describing physical phenomena such as non-Newtonian flow, soil mechanics, heat conduction, etc. (see, for instance, [11,8,21,23,25,29,30] and references therein). We will focus our study in non-autonomous perturbations of the following non-classical non-autonomous parabolic equation

{utγ(t)ΔutΔu=f(u), in Ωu=0, on Ω$$\begin{array}{} \displaystyle \left\{ \begin{array}{*{20}{l}} {{u_t} - \gamma (t)\Delta {u_t} - \Delta u = f(u),{\rm{\;in\;}}\Omega }\\ {u = 0,{\rm{\;on\;}}\partial \Omega } \end{array}\right. \end{array}$$

where Ω ⊂ ℝn is a smooth bounded domain, for some n ⩾ 3, with f and γ satisfying some suitable conditions. More specifically, we will deal with perturbations of the form

{utγ(t)ΔutΔu=gϵ(t,u), in Ωu=0, on Ω$$\begin{array}{} \displaystyle \left\{ \begin{array}{*{20}{l}} {{u_t} - \gamma \left( t \right)\Delta {u_t} - \Delta u = {g_\varepsilon }\left( {t,u} \right),{\;\text{in}\;}\Omega } \hfill \\ {u = 0,{\;\text{on}\;}\partial \Omega } \hfill \\ \end{array}\right. \end{array}$$

where {gε}ε∈[0, 1] is a family of non-autonomous functions satisfying some continuity conditions.

In the work of Aifantis et al., [1-3] we can find a quite general approach to deduce these equations in the autonomous case without delay. In the aforementioned papers, it is pointed out that the classical reaction-diffusion equation

utΔu=g(u)$$\begin{array}{} \displaystyle {u_t} - \Delta u = g(u) \end{array}$$

does not contain each aspect of the reaction-diffusion problem, and it neglects viscidity, elasticity, and pressure of medium in the process of solid diffusion. The authors obtained a diffusion theory similar to Fick’s classical model for solute in an undisturbed solid matrix, obtaining a hyperbolic equation

ut+D1utt=D2Δu,$$\begin{array}{} \displaystyle {u_t} + {D_1}{u_{tt}} = {D_2}\Delta u, \end{array}$$

where D1 and D2 are positive constants. Assign viscosity to the diffusing substance, they arrived to que following equation

ut+D1utt=D2Δu+D3Δut,$$\begin{array}{} \displaystyle {u_t} + {D_1}{u_{tt}} = {D_2}\Delta u + {D_3}\Delta {u_t}, \end{array}$$

and neglecting the inertia term, finally obtained the non-classical parabolic equation

ut=D2Δu+D3Δut,$$\begin{array}{} \displaystyle {u_t} = {D_2}\Delta u + {D_3}\Delta {u_t}, \end{array}$$

where D3 is also a positive constant.

The asymptotic behavior of the model without delay terms and with constant coefficients

utμΔutΔu+g(u)=f(x),  μ[0,1]$$\begin{array}{} \displaystyle {u_t} - \mu \Delta {u_t} - \Delta u + g(u) = f(x),\qquad \mu \in [0,1] \end{array}$$

is studied in [31], where, in particular, it is shown the well-posedness of the problem and the existence of the global attractor either in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$ or in H2(Ω), depending on the regularity of the initial data. They also showed the continuity of the global attractor in Hausdorff semidistance when μ → 0 in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$).

The introduction of a time dependence in coefficient γ(t) represents the variability of viscosity in time due to, for example, external environmental temperatures. This time dependence provides the system with a non-autonomous nature.

The study of a non-autonomous case with delay appeared in [12] for the first time, where it was established the well-posedness of the problem when γ(t) ≡ γ is constant.

In [26], Rivero studied the existence of the pullback attractor and its continuity under non-autonomous perturbations, showing the existence of a concrete structure under some assumptions on the non-linearity and giving a first approach to the study of perturbations in non-autonomous problems.

Remark 2.

This example is understood by us as a good starting point to the study of non-autonomous perturbations. Mainly because (2) is a non-autonomous equation, but term that causes this phenomena (the function γ) has no effect on the equilibria, which are the equilibria of the elliptic equation - Δu = f(u).

Novelties

In this paper we give a step towards understanding non-autonomous perturbations of non-autonomous equations. We first study the problem (6) for each ε ∈ [0, 1] using the ideas presented in Subsection 1.1, but always having the discussion of Subsection 1.2 in mind, that is, not only we will deal with each equation separately for each ε, but also we have to take into account that we must be able to obtain the results with uniformity for ε ∈ [0, 1]. Using this, we will be able to obtain continuity results for the family of equations given by (6). In the next section, we will present a detailed description of our main results.

Remark 3.

One important thing to stress out is that, even that f does not depend on t, the function γ makes problem (5) non-autonomous.

Description of the main results

To describe the contents of our work and to state the main results, we first make some assumptions on the functions γ, f and gε as follows: suppose that γ: ℝ → (0, ∞) is a uniformly continuous function which satisfies 0 < γ0γ(t) ⩽ γ1 < ∞ and the family {gε}εε[0, 1] of continuously differentiable functions from ℝ2 to ℝ with g0(t, s) = f (s) for all t, s ε ℝ, that satisfies

|gϵ(t,s1)gϵ(t,s2)|α|s1s2|(1+|s1|ρ1+|s2|ρ1),$$\begin{array}{} \displaystyle |g_\epsilon(t,s_1)-g_\epsilon(t,s_2)|\leqslant \alpha|s_1-s_2|(1+|s_1|^{\rho-1}+|s_2|^{\rho-1}), \end{array}$$

limsup|s|gϵ(t,s)sδ<λ1,$$\begin{array}{} \displaystyle \limsup_{|s|\to \infty} \frac{g_\epsilon(t,s)}{s} \leqslant \delta < \lambda_1, \end{array}$$

0tgϵ(t,s)ds<$$\begin{array}{} \displaystyle \int_{0}^\infty \frac{\partial}{\partial t} g_\epsilon(t,s)ds<\infty \end{array}$$

Also we assume that there exists a bounded function β defined in the interval [0, 1] with β (δ) → 0, as δ → 0+, and satisfying

supt|gϵ(t,s)f(s)|β(ϵ)(1+|s|ρ1), for all s and ϵ[0,1],$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}|g_\epsilon(t,s)-f(s)|\leqslant \beta(\epsilon)(1+|s|^{\rho-1}), \hbox{ for all }s\in \mathbb{R} \hbox{ and } \epsilon\in [0,1], \end{array}$$

and

supt|sgϵ(t,s)f(s)|β(ϵ)(1+|s|ρ1), for all s and ϵ[0,1],$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}|\partial_sg_\epsilon(t,s)-f'(s)|\leqslant \beta(\epsilon)(1+|s|^{\rho-1}), \hbox{ for all }s\in \mathbb{R} \hbox{ and } \epsilon\in [0,1], \end{array}$$

where λ1 > 0 is the first eigenvalue of the negative Laplacian A = —Δ with Dirichlet boundary condition, for some α > 0 and 1ρ<n+2n2$\begin{array}{} \displaystyle 1\leqslant \rho< \frac{n+2}{n-2} \end{array}$, with (H1), (H2) uniformly for t ∈ ℝ and ε ∈ [0, 1] and (H3) uniformly for ε ∈ [0, 1].

In Section 2 we present a brief summary about the theory of autonomous and non-autonomous systems, with their respective attractors. In Section 3 we will describe the precise spaces, along with the required topologies, to fit the family of non-autonomous non-classical equations in the framework described in Subsection 1.1. Sections 4 and 5 are devoted to prove that each equation (6) generates a non-autonomous dynamical system and prove the existence of the several types of attractors described in Section 2, respectively.

In Sections 6 and 7, motivated by the discussion in Subsection 1.2, we study the upper semicontinuity and topological structural stability of each kind of global attractor found in Section 5, respectively. With all this work, we provide a complete study for the various scenarios of asymptotic behavior for non-autonomous dynamical systems described in Subsection 1.1.

Preliminaries: Asymptotic dynamics of non-autonomous equations

We will briefly present the theory described in [7], which studies non-autonomous differential equations in different frameworks and gives relations between these dynamics.

Semigroups

First of all, we define the notions of semigroups and their global attractors (the reader may see [18] for more details of this theory).

Let (X, d) be a metric space and 𝒞(X) the set of all continuous maps from X into itself. A semigroup in X is a one parameter family {T(t): t ⩾ 0} such that

(a)T (0) = IdX, with IX being the identity in X,

(b)T(t)T(s) = T(t + s), for all t, s ⩾ 0 and

(c) the map [0, ∞) × X ∋ (t, x) ↦ T(t)xX is continuous.

From now on we are going to denote by dH(·, ·) the Hausdorff semidistance between two subsets of X, that is, for any A, BX:

dH(A,B)=supaAinfbBd(a,b).$$\begin{array}{} \displaystyle {\rm d}_H(A,B)=\displaystyle{\sup_{a\in A}\inf_{b\in B}} \ d(a,b). \end{array}$$

Definition 1.

A compact set 𝒜 is called a global attractor of {T(t): t ⩾ 0} if satisfies:

(i)𝒜 is invariant for {T(t): t ⩾ 0}; that is, T(t)𝒜 = 𝒜, for all t ⩾ 0.

(ii)𝒜attracts bounded subsets under the action of {T(t): t ⩾ 0}; that is, for each bounded subset B of X we have

limtdH(T(t)B,A)=0.$$\begin{array}{} \displaystyle \lim _{t\to \infty } {\rm d}_H(T(t)B, \mathscr{A})=0. \end{array}$$

The global attractor of a semigroup describes the asymptotic behavior of the semigroup. To be more precise, we define a global solution of {T(t): t ⩾ 0} as a function ξ : ℝ → X such that T(t)ξ (s) = ξ (t + s) for all t ⩾ 0 and s ∈ ℝ. Then, we know that if 𝒜 is the global attractor of {T(t): t ⩾ 0} we have

A={xX:x=ξ(0) for some bounded global solution ξ of {T(t):t0}}.$$\begin{array}{} \displaystyle {\scr A} = \{ x \in X:x = \xi (0){\;\text{for some bounded global solution }}\; \xi \;{\text{of}}\; \{ T(t):t \geqslant 0\} \} . \end{array}$$

This characterization means that not only the global attractor attracts all positive orbits {T(t)x: t ⩾ 0} (xX) but it actually consists of all bounded globally defined solutions. Moreover, the global attractor for a semigroup is unique.

To obtain existence of global attractors for semigroups, we will need some definitions.

Definition 2.

Let B, CX. We say that BabsorbsC under the action of {T(t):t ⩾ 0} if there exists T ⩾ 0 such that

T(t)CB, for all tT.$$\begin{array}{} \displaystyle T(t)C\subset B, \hbox{ for all } t\geqslant T. \end{array}$$

Definition 3.

We say that a semigroup {T(t): t ⩾ 0} is asymptotically compact if given sequences tn → ∞ and {xn}n∈𝔹 bounded in X such that {T(tn)xn}n∈ℕ is bounded, then {T(tn)xn}n∈ℕ is precompact in X.

With these definitions we are able to state the main result about existence of global attractors, that we be needed later.

Theorem 1.

(Theorem 3.4 in [18]). Let {T(t): t ⩾ 0} be an asymptotically compact semigroup. Assume that there exists a bounded set BX such that B absorbs all bounded subsets of X under the action of {T(t): t ⩾ 0}. Then {T(t): t ⩾ 0} has a global attractor 𝒜 and 𝒜 ⊂ B.

Evolutions processes

Now we are going to define evolution processes and their pullback atractors (see [9,16] for more details). These concepts appear in the literature as natural generalizations for semigroups and global attractors, respectively.

Again, let (X, d) be a metric space. An evolution process in X is a two parameter family {T(t, s): ts} in 𝒞(X) such that

(a)T(t, t) = IdX,

(b)T(t, s)T(s, τ) = T(t, τ), for all ts ⩾ τ and

(c) the map 𝒫 × X ∋ (t, s, x) ↦ T(t, s)xX is continuous, where 𝒫 = {(t, s) ∈ ℝ2: ts}.

Definition 4.

A family of compact sets {A(t)}t∈ℝ is called a pullback attractor of T(t,s) if satisfies:

(i) {A(t)}t∈ℝ is invariant; that is, T(t,s)A(s) = A(t), for all ts.

(ii) {A(t)}t∈ℝpullback attracts bounded subsets; that is, for each bounded subset B of X and t ∈ ℝ, we have

limsdH(T(t,s)B,A(t))=0.$$\begin{array}{} \displaystyle \mathop {\lim }\limits_{s \to - \infty } {{\rm{d}}_H}(T(t,s)B,A\left( t \right)) = 0. \end{array}$$

(iii) {A(t)}t∈ℝ is the minimal family of closed sets with property (ii).

Remark 4.

We note that when T(t,s) = S(ts), the family {S(t): t ⩾ 0} is a semigroup in X and Definition 4 reduces to the definition of global attractors. The first difference that appears is item (iii) in Definition 4 and it ensures the uniqueness of the pullback attractor, since it does not follows directly from (i) and (ii) as in the autonomous case.

Non-autonomous dynamical systems

Now, we will introduce the concept of non- autonomous dynamical systems, which is a general method that provides a way to form the base space for a given non-autonomous differential equation. The idea of this method is to consider the family of non-linearities as a base flow driven by the time shift.

Definition 5.

A non-autonomous dynamical system (NDS) is a quadruple (φ, θ)(X,Σ), where X, Σ are a metric spaces with metrics dX and dΣ, respectively; θ ≐ (θt: t ⩾ 0} is a semigroup in Σ, called the shift operator (or driving semigroup), and φ: ℝ+ × Σ × XX is a map

Note that we use the notation φ(t, σ, x) = φ(t, σ)x for all (t, σ, x) ∈ ℝ+ × Σ × X.

that verifies

(i) φ (0, σ) = IdX for all σ ∈ Σ;

(ii)+ × Σ ∋ (t, σ) ↦ φ(t, σ)uX is continuous, and

(iii) φ(t + s, σ) = φ(t, θ, σ)φ(s, σ), for all t, s ⩾ 0 and σ ∈ Σ.

The map φ is called the cocycle semiflow and property (iii) is know as the cocycle property.

To define the cocycle attractor and the uniform attractor for a NDS (φ, θ)(X,Σ), we first must define the concepts of non-autonomous set, invariance and pullback attraction in this framework:

Definition 6.

A non-autonomous set is a family {D(σ)}σ∈Σ of subsets of X indexed in Σ. We say that (D(σ)}σ∈Σ is an open (closed, compact) non-autonomous set if each fiberD(σ) is an open (closed, compact) subset of X.

Definition 7.

A non-autonomous set {D(σ)}σ∈Σ is invariant under the NDS (φ, θ)(X,Σ) if

φ(t,σ)D(σ)=D(θtσ), for all t0 and σΣ.$$\begin{array}{} \displaystyle \varphi (t,\sigma )D(\sigma ) = D({\theta _t}\sigma ), \;{\text{for all }}\;t \geqslant 0\;{\rm{and}}\; \sigma \in \Sigma . \end{array}$$

To define the concept of pullback attraction, we must ask some additional properties on Σ, and from now on, we are going to assume that Σ, is compact and invariant for the driving semigroup (θ:t ⩾ 0}, and also that (θt: t ⩾ 0} is a group over Σ; that is, θt is invertible, and we denote θt1=θt$\begin{array}{} \displaystyle \theta _t^{ - 1} = {\theta _{ - t}} \end{array}$.

Remark 5.

Actually, these assumptions can be dropped. We can obtain the same results requiring only that {θt : t ⩾ 0} possess a global attractor in Σ, with virtually no additional work, but with a more difficult notation. So, for simplicity, we shall assume all the hypotheses above.

Definition 8.

A compact non-autonomous set {A(σ)}σ∈Σ is called a cocycle attractor of (φ, θ)(X,Σ) if

(i) {A(σ)}σ∈Σ is invariant under (φ, θ)(X,Σ);

(ii) {A(σ)}σ∈Σ pullback attracts all bounded subsets BX, i.e.

limt+dH(φ(t,θtσ)B,A(σ))=0.$$\begin{array}{} \displaystyle \lim _{t\to +\infty }{\rm d }_H (\varphi (t,\theta _{-t}\sigma)B,A(\sigma))=0. \end{array}$$

(iii) {A(σ)}σ∈Σ is the minimal among the closed non-autonomous sets with property (ii).

We can also deal with the uniform attraction for a NDS - in this framework, the attraction do not depend on the chosen σ ∈ Σ, - that is, we say that the subset KX is uniform attracting for the NDS (φ, σ)(X,Σ) if for each BX bounded,

limtsupσΣdH(φ(t,σ)B,K)=0.$$\begin{array}{} \displaystyle \lim_{t\to\infty}\sup_{\sigma\in\Sigma}{\rm d}_H(\varphi(t,\sigma)B,K) = 0. \end{array}$$

Definition 9.

A compact subset 𝒜 ⊂ X is called the uniform attractor for the NDS (φ, σ)(X,Σ) if it is the minimal closed subset of X that uniform attracts all bounded subsets of X.

Theorem 2.

A NDS (φ, σ)(XΣ)has a uniform attractor if and only if there exists a compact uniform attracting set K.

Skew-product semiflows

When dealing with non-autonomous dynamical systems, it is worthwhile to question if we can transform them into an autonomous one; and that is precisely the case: for a given NDS (φ, θ)(X,Σ) we can define a semigroup {Π(t): t ⩾ 0} in the product space X=X×Σ$\begin{array}{} \displaystyle \mathbb X=X\times\Sigma \end{array}$ as (see [27,28] for more details)

Π(t)(u,σ)=(φ(t,σ)u,θtσ),$$\begin{array}{} \displaystyle \Pi(t) (u,\sigma) = (\varphi(t,\sigma)u,\theta_t\sigma), \end{array}$$

which is called skew-product semiflow associated with (φ, θ)(X,Σ).

So far we obtained three different objects:

the evolution process {T(t, s): ts};

the non-autonomous dynamical system (φ, θ)(X,Σ) and

the skew-product semiflow {Π(t): t ⩾ 0},

and the four different notions of ‘attractors’ listed in the Introduction:

(i) the pullback attractor {A(t)}t∈ℝ of {T(t, s): ts};

(ii) the cocycle attractor {A(σ)}σ∈Σ of (φ, θ)(X,Σ);

(iii) the uniform attractor 𝒜 of (φ, θ)(X,Σ) and

(iv) the global attractor 𝔸 of {Π(t): t ⩾ 0},

and now we presente briefly the relationships between these objects (as in [7]). We begin with the relation between the global attractor of {Π(t): t ⩾ 0} and the cocycle attractor of (φ, θ)(X,Σ).

Theorem 3.

(Propositions 3.30 and 3.31 in [22], or Theorem 3.4 in [14]). Let (φ, θ)(X,Σ)be a non-autonomous dynamical system and let {Π(t): t ⩾ 0} be the associated skew product semiflow on X × Σ with a global attractor 𝔸. Then {A(σ)}σ∈Σwith A(σ) = {xX : (x, σ) ∈ 𝔸} is the cocycle attractor of (φ, θ)(X,Σ).

The following theorem shows the relationship between the global attractor of a skew product semiflow and the pullback attractors of the evolution processes it may contain.

Theorem 4.

(Theorem 2.7 in [5]). Assume that the skew product semiflow {Π(t): t ⩾ 0} possesses a global attractor 𝔸. Then the evolution process {Tσ(t, s) : ts} given by

Tσ(t,s)u=φ(ts,θsσ)u,uX,$$\begin{array}{} \displaystyle T_{\sigma}(t,s)u= \varphi(t-s,\theta_s\sigma)u, \ u\in X, \end{array}$$

possesses a pullback attractor {Aσ(t)}t∈ℝ. Moreover

A=σΣ[tAσ(t)×{σ}].$$\begin{array}{} \displaystyle \mathbb{A}=\bigcup_{\sigma\in\Sigma}\left[\bigcup_{t\in \mathbb{R}}{A}_\sigma(t)\times \{\sigma\}\right]. \end{array}$$

Therefore, if 𝔸 is the global attractor of the associated semigroup {Π(t) : t ∈ ℝ} of (φ, σ)(X,Σ), then the uniform attractor 𝒜 is the projection on X of the global attractor, that is, 𝒜 = πX(𝔸), where πX : X × Σ → X is the projection over X.

Now, the relationship between the uniform attractor and the pullback attractor is clear.

Theorem 5.

The NDS (φ, σ)(X,Σ)has a uniform attractor 𝒜 if and only if the associated skew-product semiflow {Π(t): t ⩾ 0} has a global attractor 𝔸 and

A=πX(A)=σΣtAσ(t)$$\begin{array}{} \displaystyle \mathscr A = \pi_X(\mathbb A) = \bigcup_{\sigma\in\Sigma}\bigcup_{t\in \mathbb {R}}{A}_\sigma(t) \end{array}$$

The following result shows us the required assumptions in order to obtain the existence of the global attractoi for the skew-product semiflow {Π(t) : t ⩾ 0} associated with the NDS (φ, σ)(X,Σ) based on the existence of it cocycle attractor (see [14,22] for more details).

Theorem 6.

Suppose that {A(σ)}σ∈Σis the cocycle attractor of (φ, θ)(X,Σ), {Π(t) : t ⩾ 0} is the associated skew-product semiflow. Assume that {A(σ)}σ∈Σis uniformly attracting, i.e.,

limt+supσΣdist(φ(t,θtσ)D,A(σ))=0,$$\begin{array}{} \displaystyle \mathop {\lim }\limits_{t \to + \infty } \mathop {\sup }\limits_{\sigma \in \Sigma } {\rm{dist}}(\varphi (t,{\theta _{ - t}}\sigma )D,A(\sigma )) = 0, \end{array}$$

and thatσ∈ΣA(σ) is precompact in X. Then the set 𝔸 associated with {A(σ)}σ∈Σ, given by

A=σΣA(σ)×{σ},$$\begin{array}{} \displaystyle \mathbb{A}=\bigcup_{\sigma\in \Sigma}A(\sigma)\times \{\sigma\}, \end{array}$$

is the global attractor of {Π(t) : t ⩾ 0}.

With these results we complete the relations between the four different asymptotic dynamics we presented. In this paper, as we said before, we will try to deal with these four dynamics, to obtain as much information as we can of equation (6).

Driving semigroups of translations for (6)

In this section we will put the family of equations (6) in the framework described on Subsection 1.1. To this end, let A = −Δ : D(A) ⊂ XX be the negative Laplacian operator with Dirichlet boundary condition, defined in D(A)=H01(Ω)H2(Ω)$\begin{array}{} \displaystyle D(A) = H_0^1(\Omega ) \cap {H^2}(\Omega ) \end{array}$, where X = L2(Ω). Consider the fractional power scale Xα, with α ∈ ℝ, generated by (X,A). Also, consider the Nemytskii operators gϵe(,)$\begin{array}{} \displaystyle g^e_\epsilon(\cdot,\cdot) \end{array}$ defined as gϵe(t,u)(x)=gϵ(t,u(x))$\begin{array}{} \displaystyle g^e_\epsilon(t,u)(x)=g_\epsilon(t,u(x)) \end{array}$ for each (t, x) ∈ ℝ × Ω and u : Ω → ℝ.

Following the ideas in [26], we define the operators

Bγ(t)=(I+γ(t)A)1 and A˜γ(t)=ABγ(t),$$\begin{array}{} \displaystyle {B_\gamma }(t) = {(I + \gamma (t)A)^{ - 1}}{\rm{ and }}{\tilde A_\gamma }(t) = A{B_\gamma }(t), \end{array}$$

and the function gϵ,γ(t,u)=Bγ(t)gϵe(t,u)$\begin{array}{} \displaystyle g_{\epsilon,\gamma}(t,u)=B_\gamma(t)g_{\epsilon}^e(t,u) \end{array}$, for each ε ∈ [0,1], we can write problem (6) as

ut=Fϵ(t,u),$$\begin{array}{} \displaystyle u_t=F_\epsilon(t,u), \end{array}$$

where Fε(t, u) = −Ãγ(t)u + gε,γ(t, u). The domain of the operators Ãγ(t) does not depend on time and the operators ℝ ∋ tBγ(t) and ℝ ∋ t ↦ Ãγ(t) are absolutely continuous functions.

To study (6), for each ε ∈ [0,1], we must be able to study the hull of the function Fε(·,·) in a suitable space. This is our task in the next subsection.

Driving groups of translations

We begin considering the sets

𝒞1 = Cb(ℝ, ℝ) of continuous bounded functions from ℝ to itself with metric

d1(λ1,λ2)=supt|λ1(t)λ2(t)|;$$\begin{array}{} \displaystyle d_1(\lambda_1,\lambda_2)=\sup_{t\in \mathbb{R}}|\lambda_1(t)-\lambda_2(t)|; \end{array}$$

𝒞2 of continuous functions from ℝ2 to ℝ, which satisfies: h ∈ 𝒞2 if there exists constants γ,ε ⩾ 0 such that

supt|h(t,s1)h(t,s2)|γ|s1s2|(1+|s1|ρ1+|s2|ρ1), for all s1,s2,$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}|h(t,s_1)-h(t,s_2)|\leqslant \gamma|s_1-s_2|(1+|s_1|^{\rho-1}+|s_2|^{\rho-1}), \hbox{ for all }s_1,s_2\in \mathbb{R}, \end{array}$$

and

supt|h(t,0)|ε.$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}|h(t,0)|\leqslant \omega. \end{array}$$

In 𝒞2 we introduce the norm

hC2=supt,ss0|h(t,s)h(t,0)||s|(1+|s|ρ1)+supt|h(t,0)|,$$\begin{array}{} \displaystyle \|h\|_{\mathcal{C}_2}=\sup_{\underset{s\neq 0}{t,s\in \mathbb{R}}}\frac{|h(t,s)-h(t,0)|}{|s|(1+|s|^{\rho-1})}+\sup_{t\in \mathbb{R}}|h(t,0)|, \end{array}$$

and the distance

d2(h1,h2)=h1h2C2.$$\begin{array}{} \displaystyle {d_2}({h_1},{h_2}) = \|{h_1} - {h_2}\|{_{{{\cal C}_2}}}. \end{array}$$

Remark 6.

We have that 𝒞1 and 𝒞2 are Banach spaces with norms ||λ||𝒞1 = supt∊ℝ |λ(t)| and ||·||𝒞2, respectively.

Clearly we have that γ𝒞1 and gε𝒞2, for all ε ∊ [0,1]. We define the group of translations in both

Here we denote both groups the same, since there will be no confusion of notation.

𝒞1 and 𝒞2, {θt: t ∊ ℝ}, by

θtλ(s)=λ(t+s), for all λC1 and t,s;$$\begin{array}{} \displaystyle \theta_t\lambda (s)=\lambda (t+s), \hbox{ for all }\lambda \in \mathcal{C}_1 \hbox{ and }t,s\in \mathbb{R}; \end{array}$$

and

θth(r,s)=h(t+r,s), for all hC2 and t,r,s.$$\begin{array}{} \displaystyle \theta_th(r,s)=h(t+r,s), \hbox{ for all } h\in \mathcal{C}_2 \hbox{ and } t,r,s\in \mathbb{R}. \end{array}$$

Also, let

Γ be the hull of γ in 𝒞1; that is,

Γ= closure of {θtγ}t in (C1,d1).$$\begin{array}{} \displaystyle \Gamma=\hbox{ closure of } \{\theta_t\gamma\}_{t\in \mathbb{R}} \hbox{ in } (\mathcal{C}_1,d_1). \end{array}$$

𝒢ε be the hull of gε in 𝒞2; that is,

Gϵ= closure of {θtgϵ}t in (C2,d2).$$\begin{array}{} \displaystyle \mathscr{G}_\epsilon= \hbox{ closure of } \{\theta_tg_\epsilon\}_{t\in \mathbb{R}} \hbox{ in } (\mathscr{C}_2,d_2). \end{array}$$

Remark 7.

Since γ is bounded and uniformly continuous on ℝ, the set Γ is compact in (𝒞1, d1), using the Arzelá-Ascoli Theorem.

Note that, by simple computations, we have that there exists a constant C > 0 such that

suptBλ1(t)Bλ2(t)L(H1,H01(Ω))Cλ1λ2C1$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}\|B_{\lambda_1}(t)-B_{\lambda_2}(t)\|_{\mathscr{L}(H^{-1},H^1_0(\Omega))}\leqslant C\|\lambda_1-\lambda_2\|_{\mathscr{C}_1} \end{array}$$

and

suptA˜λ1(t)A˜λ2(t)L(H01(Ω))Cλ1λ2C1,$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}\|\tilde{A}_{\lambda_1}(t)-\tilde{A}_{\lambda_2}(t)\|_{\mathscr{L}(H^1_0(\Omega))}\leqslant C\|\lambda_1-\lambda_2\|_{\mathscr{C}_1}, \end{array}$$

for all λ1, λ2 ∊ Γ, where H—1 is the dual space of H01(Ω)$\begin{array}{} \displaystyle H^1_0(\Omega) \end{array}$.

Since g0(t, s) = f(s) for all t, s ∊ ℝ, 𝒢0 = {f} and hence it is compact in (𝒢2, d2).

Since hypotheses (H1)-(H4) are uniform for t ∊ ∨, they all are satisfied by every function in 𝒢ε.

Using items 1 and 2 from the previous remark, we will make one additional assumption on the family {𝒢ε}:

Gϵ is compact in (C2,d2) for each ϵ(0,1].$$\begin{array}{} \displaystyle \mathscr{G}_\epsilon \hbox{ is compact in } (\mathscr{C}_2,d_2) \hbox{ for each } \epsilon\in (0,1]. \end{array}$$

Remark 8.

Condition (C) is verified, for instance, when each gε is time-independent, or periodic in t, or almost-periodic in t (for the latter, see for instance [20, Appendix - Theorem 11]).

Now we can see that each function hε𝒢ε defines a Nemytskii operator from ×Xs2$\begin{array}{} \displaystyle \mathbb{R}\times X^{\frac s2} \end{array}$ into L2nn+2r(Ω)$\begin{array}{} \displaystyle L^{\frac{2n}{n+2r}}(\Omega) \end{array}$, for suitable s and r.

Lemma 7.

Assume that the family {𝒢ε}ε∊[0,1]satisfies (H1), A is the negative Dirichlet Laplacian in X with domainX1=H2(Ω)H01(Ω)$\begin{array}{} \displaystyle X^1=H^2(\Omega)\cap H^{1}_0(\Omega) \end{array}$and consider its closed extension to Hr=(Xr2)$\begin{array}{} \displaystyle H^{-r}=(X^{\frac{r}{2}})' \end{array}$, the dual space of Xr2$\begin{array}{} \displaystyle X^{\frac r2} \end{array}$, (in particular,H1=H01(Ω))$\begin{array}{} \displaystyle H^{-1}=H^1_0(\Omega)') \end{array}$. Then the Nemytskii operators{hϵe}ϵ[0,1]$\begin{array}{} \displaystyle \{h^e_\epsilon\}_{\epsilon\in [0,1]} \end{array}$are well defined from×Xs2$\begin{array}{} \displaystyle \mathbb{R}\times X^{\frac s2} \end{array}$intoL2nn+2r(Ω)$\begin{array}{} \displaystyle L^{\frac{2n}{n+2r}}(\Omega) \end{array}$provided thatr[(ρ1)(n2)4,1],s[r,1][n22ρ1,1]$\begin{array}{} xs\displaystyle r\in \left[\frac{(\rho-1)(n-2)}{4},1\right], s \in [r,1]\cap \left[\frac{n}{2}-\frac{2}{\rho-1},1\right] \end{array}$, for each hε𝒢ε. If B is a bounded subset ofXs2$\begin{array}{} \displaystyle X^{\frac s2} \end{array}$then there exists a constant C = C(B) > 0 such that

supthϵe(t,u1)hϵe(t,u2)L2nn+2r(Ω)Cu1u2Xs2,forall ϵ[0,1].$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}\|h^e_\epsilon(t,u_1)-h^e_\epsilon(t,u_2)\|_{L^{\frac{2n}{n+2r}}(\Omega)}\leqslant C\|u_1-u_2\|_{X^{\frac s2}}, \hbox{ for all }\epsilon \in [0,1]. \end{array}$$

Moreover, if r can be taken strictly less than 1 andJ ⊂ ℝ is an arbitrary subset, hεe$\begin{array}{} \displaystyle h_\varepsilon ^e \end{array}$takes J × B in a precompact set of H−1, for eachε ∊ [0,1].

Proof. Following [10], using hypothesis (H1) we have that

supthϵe(t,u)hϵe(t,v)L2nn+2r(Ω)c[Ω[|uv|(1+|u|ρ1+|v|ρ1)]2nn+2r]n+2r2nc˜uvL2nn2r(Ω)(1+uLn(ρ1)2r(Ω)ρ1+vLn(ρ1)2r(Ω)ρ1)c¯uvXs2(1+uXs2ρ1+vXs2ρ1).$$\begin{array}{} \displaystyle \begin{equation} \begin{split} \sup_{t\in\mathbb{R}}\|h_\epsilon^e(t,u)-h_\epsilon^e(t,v)&\|_{L^{\frac{2n}{n+2r}}(\Omega)}\leqslant c\left[ \int_\Omega \left[|u-v|(1+|u|^{\rho-1}+|v|^{\rho-1})\right]^{\frac{2n}{n+2r}}\right]^{\frac{n+2r}{2n}}\\ &\leqslant\tilde c\|u-v\|_{L^{\frac{2n}{n-2r}}(\Omega)}\left(1+\|u\|_{L^{\frac{n(\rho-1)}{2r}}(\Omega)}^{\rho-1}+\|v\|_{L^{\frac{n(\rho-1)}{2r}}(\Omega)}^{\rho-1}\right)\\ &\leqslant\bar c\|u-v\|_{X^{\frac s2}}\left(1+\|u\|_{X^{\frac s2}}^{\rho-1}+\|v\|_{X^{\frac s2}}^{\rho-1}\right). \end{split} \end{equation} \end{array}$$

or any s[r,1][n22ρ1,1]$\begin{array}{} \displaystyle s \in [r,1]\cap \left[\frac{n}{2}-\frac{2}{\rho-1},1\right] \end{array}$. The last statement holds since Hr is compact embedded in H−1 and L2nn+2r(Ω)Hr$\begin{array}{} \displaystyle {L^{\frac{{2n}}{{n + 2r}}}}(\Omega )\hookrightarrow{H^{ - r}} \end{array}$Hr for r < 1.

We define now

C2e=Cb0(×H01(Ω),L2nn+2(Ω))$\begin{array}{} \displaystyle \mathscr{C}^e_2=C_b^0(\mathbb{R}\times H^1_0(\Omega),L^{\frac{2n}{n+2}}(\Omega)) \end{array}$ as the set of all continuous functions from ×H01(Ω)$\begin{array}{} \displaystyle \mathbb{R}\times H^1_0(\Omega) \end{array}$ taking values on L2nn+2(Ω)$\begin{array}{} \displaystyle L^{\frac{2n}{n+2}}(\Omega) \end{array}$ that are bounded on sets ℝ × B, where B is a bounded set of H01(Ω)$\begin{array}{} \displaystyle H^1_0(\Omega) \end{array}$ with metric

d2e(σ1,σ2)=k=12kσ1σ2ke1+σ1σ2ke,$$\begin{array}{} \displaystyle d_2^e(\sigma_1,\sigma_2)=\sum_{k=1}^\infty 2^{-k}\frac{\|\sigma_1-\sigma_2\|_k^e}{1+\|\sigma_1-\sigma_2\|_k^e}, \end{array}$$

where if Bk={uH01(Ω):uH01(Ω)k}$\begin{array}{} \displaystyle B^k=\{u\in H^1_0(\Omega)\colon \|u\|_{H^1_0(\Omega)}\leqslant k\} \end{array}$ we have

σ1σ2ke=suptsupuBkσ1(t,u)σ2(t,u)L2nn+2(Ω).$$ \|\sigma_1-\sigma_2\|_k^e=\sup\limits_{t\in \mathbb{R}}\sup\limits_{u\in B^k}\|\sigma_1(t,u)-\sigma_2(t,u)\|_{L^{\frac{2n}{n+2}}(\Omega)}. $$

Remark 9.

Here we have that C2e$\begin{array}{} \displaystyle \mathscr{C}_2^e \end{array}$ with the metric d2e$\begin{array}{} \displaystyle d_2^e \end{array}$ is a Frechét space, and a sequence {σk}k∊ℕ converges in C2e$\begin{array}{} \displaystyle C_2^e \end{array}$ if and only if it converges in each seminorm ne$\begin{array}{} \displaystyle \|\cdot\|_n^e \end{array}$.

Define the group of translations

Again, since there will be no confusion, we denote the group of translations the same.

{θt : t ∊ ℝ} in C2e$\begin{array}{} \displaystyle \mathscr{C}_2^e \end{array}$ by

θth(s,v)=h(t+s,v), for all hC2e,t,s and vH01(Ω).$$\begin{array}{} \displaystyle \theta_t h(s,v) = h(t+s,v), \hbox{ for all } h\in \mathscr{C}_2^e, \ t,s\in \mathbb{R} \hbox{ and }v\in H^1_0(\Omega). \end{array}$$

Since Lemma 7 implies that gϵeC2e$\begin{array}{} \displaystyle g_\epsilon^e\in \mathscr{C}_2^e \end{array}$, for all ε ∊ [0,1], let also

Gεe$\begin{array}{} \displaystyle \mathscr{G}^e_\epsilon \end{array}$ be the hull of gϵe$\begin{array}{} \displaystyle g^e_\epsilon \end{array}$ in C2e$\begin{array}{} \displaystyle \mathscr{C}_2^e \end{array}$; that is,

Gϵe= closure of {θtgϵe}t in (C2e,d2e).$$\begin{array}{} \displaystyle \mathscr{G}^e_\epsilon=\hbox{ closure of } \{\theta_tg^e_\epsilon\}_{t\in \mathbb{R}} \hbox{ in } (\mathscr{C}_2^e,d_2^e). \end{array}$$

Now, in order to have a better understanding of the set Gϵe$\begin{array}{} \displaystyle \mathscr{G}_\epsilon^e \end{array}$, we present the following results:

Lemma 8.

If h𝒢ε, thenheGϵe$\begin{array}{} \displaystyle h^e\in \mathscr{G}_\epsilon^e \end{array}$.

Proof. This follows easily by Lemma 7.

Lemma 9.

There exists a constant L ⩾ 0, such that for all h1,h2𝒢ε, we have that

d2e(h1e,h2e)Lh1h2C2.$$\begin{array}{} \displaystyle d_2^e(h^e_1,h^e_2)\leqslant L\|h_1-h_2\|_{\mathscr{C}_2}. \end{array}$$

Proof. Fix k ∊ ℕ. We have that

h1eh2eke=suptsupuBkh1e(t,u)h2e(t,u)L2nn+2(Ω)=suptsupuBk(Ω|h1(t,u(x))h2(t,u(x))|2nn+2dx)n+22nc˜h1h2C2supuBk(Ω|u(x)(1+|u(x)|ρ1)|2nn+2dx)n+22n,$$\begin{array}{l} \displaystyle \|h_1^e-&h_2^e\|_k^e=\sup\limits_{t\in \mathbb{R}}\sup\limits_{u\in B^k}\|h_1^e(t,u)-h_2^e(t,u)\|_{L^{\frac{2n}{n+2}}(\Omega)}\\ &=\sup\limits_{t\in \mathbb{R}}\sup\limits_{u\in B^k} \left(\int_\Omega|h_1(t,u(x))-h_2(t,u(x))|^{\frac{2n}{n+2}}dx\right)^{\frac{n+2}{2n}}\\ &\leqslant \tilde{c} \|h_1-h_2\|_{\mathscr{C}_2}\cdot \sup\limits_{u\in B^k}\left(\int_\Omega|u(x)(1+|u(x)|^{\rho-1})|^{\frac{2n}{n+2}}dx\right)^{\frac{n+2}{2n}}, \end{array}$$

where c˜$\begin{array}{} \displaystyle \tilde{c} \end{array}$ does not depend on h1,h2 and hence, arguing as in Lemma 7, we obtain

h1eh2ekec¯h1h2C2supuBk[uH01(Ω)(1+uH01(Ω)ρ1)]c^h1h2C2[k(1+kρ1)],$$\begin{array}{} \displaystyle \|h_1^e-h_2^e\|_k^e & \leqslant \overline{c} \|h_1-h_2\|_{\mathscr{C}_2}\cdot \sup_{u\in B^k} \big[\|u\|_{H^1_0(\Omega)}(1+\|u\|_{H^1_0(\Omega)}^{\rho -1})\big] \\ & \leqslant \hat{c} \|h_1-h_2\|_{\mathscr{C}_2} [k(1+k^{\rho-1})], \end{array}$$

where ĉ does not depend on h1,h2.

Hence

d2e(h1e,h2e)c^h1h2C2k=12k+1kρ,$$\begin{array}{} \displaystyle d_2^e(h_1^e,h_2^e)\leqslant \hat{c}\|h_1-h_2\|_{\mathscr{C}_2}\sum_{k=1}^{\infty} 2^{-k+1}k^{\rho}, \end{array}$$

and the result follows since k=12k+1kρ$ \sum _{k = 1}^\infty {2^{ - k + 1}}{k^\rho }$ is a convergent series.

Proposition 10.

If condition (C) holds thenσϵGϵe$\begin{array}{} \displaystyle \sigma_\epsilon \in \mathscr{G}_\epsilon^e \end{array}$if and only if there exists hε𝒢εsuch thatσϵ=hϵe$\begin{array}{} \displaystyle \sigma_\epsilon=h_\epsilon^e \end{array}$, for eachε ∊ [0,1].

Proof. The result is trivial if ε = 0. Assume that ε ∊ (0,1]. One inclusion follows from Lemma 8. Now if σϵGϵe$\begin{array}{} \displaystyle \sigma_\epsilon\in \mathscr{G}_\epsilon^e \end{array}$, then exists a sequence {tn}n∊ℕ ⊂ ℝ such that θtngϵeσϵ$\begin{array}{} \displaystyle \theta_{t_n}g^e_\epsilon \to \sigma_\epsilon \end{array}$ in C2e$\begin{array}{} \displaystyle \mathscr{C}_2^e \end{array}$, by definition. Consider the sequence {θtngε}n∈ℕ in 𝒞2. Since 𝒢ε is compact, we can assume without loss of generality, that θtngεhε𝒢ε in 𝒞2. Thus, using Lemma 9 we have that

d2e(σϵ,hϵe)d2e(σϵ,θtngϵe)+d2e(θtngϵe,hϵe)d2e(σϵ,θtngϵe)+LθtngϵhϵC2,$$\begin{array}{} \displaystyle d_2^e(\sigma_\epsilon,h_\epsilon^e)&\leqslant d_2^e(\sigma_\epsilon,\theta_{t_n}g^e_\epsilon)+d_2^e(\theta_{t_n}g_\epsilon^e,h_\epsilon^e)\\ &\leqslant d_2^e(\sigma_\epsilon,\theta_{t_n}g^e_\epsilon) + L\|\theta_{t_n}g_\epsilon-h_\epsilon\|_{\mathscr{C}_2}, \end{array}$$

and making n → ∞ we obtain that σϵ=hϵe$\begin{array}{} \displaystyle \sigma_\epsilon=h_\epsilon^e \end{array}$.

Corollary 11.

If condition (C) holds, thenGϵe$\begin{array}{} \displaystyle \mathscr{G}_\epsilon^e \end{array}$ is compact in C2e$\begin{array}{} \displaystyle \mathscr{C}_2^e \end{array}$, for eachε ∊ [0,1].

Now we are finally in condition to define the space that will be suitable for our study of (6). Define

C*=Cb0(×H01(Ω),H01(Ω))$\begin{array}{} \displaystyle \mathscr{C}_*=C^0_b(\mathbb{R}\times H^1_0(\Omega),H^1_0(\Omega)) \end{array}$ the set of continuous functions which are bounded in sets of the form ℝ × B, where B is a bounded subset of H01(Ω)$\begin{array}{} \displaystyle H^1_0(\Omega) \end{array}$, with distance defined as

d*(σ1,σ2)=n=12nσ1σ2n*1+σ1σ2n*,$$\begin{array}{} \displaystyle d_*(\sigma_1,\sigma_2)=\sum_{n=1}^\infty 2^{-n}\frac{\|\sigma_1-\sigma_2\|_n^*}{1+\|\sigma_1-\sigma_2\|_n^*}, \end{array}$$

where if Bn={uH01(Ω):uH01(Ω)n}$\begin{array}{} \displaystyle B^n=\{u\in H^1_0(\Omega)\colon \|u\|_{H^1_0(\Omega)}\leqslant n\} \end{array}$ we have

σ1σ2n*=suptsupuBnσ1(t,u)σ2(t,u)H01(Ω).$$\begin{array}{} \displaystyle \|\sigma_1-\sigma_2\|_n^*=\sup_{t\in \mathbb{R}}\sup_{u\in B^n}\|\sigma_1(t,u)-\sigma_2(t,u)\|_{H^1_0(\Omega)}. \end{array}$$

Now let Fϵ(t,u)=A˜γ(t)u+Bγ(t)ge(t,u)$\begin{array}{} \displaystyle F_\epsilon(t,u)=-\tilde{A}_\gamma(t)u+B_\gamma(t)g^e(t,u) \end{array}$ be given as in (8). It is simple to see, recalling the definitions of A˜γ(t)$\begin{array}{} \displaystyle \tilde{A}_\gamma(t) \end{array}$ and Bγ(t) and the fact that gϵeC2e$\begin{array}{} \displaystyle g_\epsilon^e\in \mathscr{C}_2^e \end{array}$, that Fε𝒞* for each ε ∊ [0,1]. Again, we can define the group

a We once again denote the same

{θt: t ∊ ℝ} in 𝒞* by

θth(s,v)=h(t+s,v), for all hC*,t,s and vH01(Ω).$$\begin{array}{} \displaystyle \theta_t h(s,v)=h(t+s,v), \hbox{ for all } h\in \mathscr{C}_*, \ t,s\in \mathbb{R} \hbox{ and } v\in H^1_0(\Omega). \end{array}$$

Definition 10

With the notations above, we set Σε as the hull of Fε in 𝒞*; that is,

Σϵ= closure of {θtFϵ}t in (C*,d*).$$\begin{array}{} \displaystyle \Sigma_\epsilon= \hbox{ closure of } \{\theta_tF_\epsilon\}_{t\in \mathbb{R}} \hbox{ in } (\mathscr{C}_*,d_*). \end{array}$$

Before we proceed with the study of (6), we will need some characterization result for Σε.

Lemma 12

We have that

(a)Σ0={BλfeA˜λ}λΓ$\begin{array}{} \displaystyle {\Sigma _0} = {\{ {B_\lambda }{f^e} - {\tilde A_\lambda }\} _{\lambda \in \Gamma }} \end{array}$and it is compact in (𝒞*,d*);

if(C)holds true, then for each ε > 0 we have

Σϵ{BλhϵeA˜λ}λΓ,hϵGϵ,$$\begin{array}{} \displaystyle \Sigma_\epsilon\subseteq \{B_\lambda h_\epsilon^e - \tilde{A}_\lambda\}_{\lambda\in \Gamma,h\epsilon\in \mathscr{G}_\epsilon}, \end{array}$$

and Σε is compact in (𝒞*,d*).

where Bλ(t) = (I + λ(t)A)–1andA˜λ(t)=ABλ(t)$\begin{array}{} \displaystyle {\tilde A_\lambda }(t) = A{B_\lambda }(t) \end{array}$, for each λ ∈ Γ.

Proof. Since g0(t,s) = f(s) for all t,s ∈ ℝ and Γ is compact, item (a) follows immediately. Now, fix ε > 0 and let Hε ∈ Σε. Then, by definition, there exists a real sequence {tn}n∈ℕ such that Hϵ=C*limnθtn(BγgϵeA˜γ)$\begin{array}{} \displaystyle H_\epsilon=\mathscr{C}_*-\displaystyle \lim_{n\to \infty}\theta_{t_n}(B_\gamma g^e_\epsilon - \tilde{A}_\gamma) \end{array}$ that is, the sequence θtn(BγgϵeA˜γ)$\begin{array}{} \displaystyle \theta_{t_n}(B_\gamma g^e_\epsilon - \tilde{A}_\gamma) \end{array}$ converges to Hε in the metric of 𝒞* defined above.

We can extract a subsequence of {tn}n∈ℕ, which we shall denote the same, and elements λ ∈ Γ and hε ∈ 𝒢ε such that γ=Climnθsnγ$\begin{array}{} \displaystyle \gamma = {\cal C} - \mathop {\lim }\limits_{n \to \infty } {\theta _{{s_n}}}\gamma \end{array}$ and hϵe=C2elimnθtngϵe$\begin{array}{} \displaystyle h_\epsilon^e=\mathscr{C}_2^e-\displaystyle\lim_{n\to \infty}\theta_{t_n}g_\epsilon^e \end{array}$, by the compactness of Γ and 𝒢ε and Proposition 10. Thus Hϵ=BλhϵeA˜λ$\begin{array}{} \displaystyle H_\epsilon=B_\lambda h_\epsilon^e -\tilde{A}_\lambda \end{array}$.

The last statement is clear from (a) and (b), which concludes the result.

Thus our problem in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$ takes the form

{u˙=Hϵ(t,u),t>0u(0)=u0H01(Ω),$$\left\{\begin{array}{l} &\dot{u}=H_\epsilon(t,u), \quad t \gt 0\\ &u(0)=u_0\in H^1_0(\Omega), \end{array}\right.$$

for each Hε ∈ Σε, which is precisely equation (4). As a matter of fact we have, for each fixed ε ∈ [0,1], a problem equals to (2). Our first task is to find, for each ε ∈ [0,1] and Hε ∈ Σε, a solution t ↦ φ(t, Hε)u0 of (10). But we will solve this problem in a slightly different way, considering all possible functions in {BλhϵeA˜λ}λΓ,hϵGϵ$\begin{array}{} \displaystyle \{B_\lambda h^e_\epsilon-\tilde{A}_\lambda\}_{\lambda\in \Gamma,h_\epsilon\in \mathscr{G}_\epsilon} \end{array}$ and therefore, we will denote this space by Γ ⋄ 𝒢ε; that is,

ΓGϵ={BλhϵeA˜λ}λΓ,hϵGϵ.$$\begin{array}{} \displaystyle \Gamma\diamond \mathscr{G}_\epsilon=\{B_\lambda h^e_\epsilon-\tilde{A}_\lambda\}_{\lambda\in \Gamma,h_\epsilon\in \mathscr{G}_\epsilon}. \end{array}$$

Remark 10.

It is clear that the maps ℝ ∋ tBλ(t) and tA˜λ(t)$\begin{array}{} \displaystyle \mathbb{R}\ni t\mapsto \tilde{A}_\lambda(t) \end{array}$ are absolutely continuous, for each λ ∈ Γ.

Non-autonomous dynamical systems and skew-product semiflows for (10)

In this section we will show that equation (10) generates a non-autonomous dynamical system (φϵ,θ)(H01(Ω),Σϵ)$\begin{array}{} \displaystyle (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma_\epsilon)} \end{array}$, for each ε ∈ [0,1].

Local existence and uniqueness of solutions

Using Remark 10, Lemma 7 and the results on [24, Chapter 5] we have the following theorem of local existence and uniqueness of solutions.

Theorem 13.

Assume that hypotheses (H1) and (C) are satisfied. Then, for each bounded subsetBH01(Ω)$\begin{array}{} \displaystyle B \subset H_0^1(\Omega ) \end{array}$, ε ∈ [0,1] andHε ∈ Γ ⋄ 𝒢εthere exists ε = ε(B, ε, Hε) > 0 such that for each u0B there exists a unique function

[0,ε]tφϵ(t,Hϵ)u0,$$\begin{array}{} \displaystyle [0,\omega]\ni t\mapsto \varphi_\epsilon(t,H_\epsilon)u_0, \end{array}$$

withφϵ(,Hϵ)u0C1([0,ε],H01(Ω))$\begin{array}{} \displaystyle \varphi_\epsilon(\cdot,H_\epsilon)u_0\in C^1([0,\omega],H^1_0(\Omega)) \end{array}$satisfying (10); that is, φε(0,Hε)u0 = u0 and

ddtφϵ(t,Hϵ)u0=Hϵ(t,φϵ(t,Hϵ)u0),for0<t<ω.$$\begin{array}{l} \frac{d}{dt}\varphi_\epsilon(t,H_\epsilon)u_0 = H_\epsilon(t,\varphi_\epsilon(t,H_\epsilon)u_0), \;\text{for} \;0 \lt t \lt \omega. \end{array}$$

Global existence of solutions

Assume that (H2) holds true. Following the ideas of [19,26], for any vH01(Ω)$\begin{array}{} \displaystyle v\in H^1_0(\Omega) \end{array}$, hε ∈ 𝒢ε and each δ > 0 there exists a constant Kδ > 0 such that

Ωhϵe(t,v)vδvL2(Ω)2+Kδ,ΩΦϵ(t,v)δvL2(Ω)2+Kδ$$\begin{array}{} \displaystyle &\int_\Omega h^e_\epsilon(t,v) v \leqslant \delta\|v\|_{L^2(\Omega)}^2 + K_\delta,\\ &\int_\Omega \Phi_\epsilon(t,v) \leqslant \delta\|v\|_{L^2(\Omega)}^2 + K_\delta \end{array}$$

with Φϵ(t,r)=0rhϵ(t,θ)dθ$\begin{array}{} \displaystyle \Phi_\epsilon(t,r)=\int_0^r h_\epsilon(t,\theta)d\theta \end{array}$, uniformly in t ∈ ℝ and ε ∈ [0,1].

Now for each vH01(Ω)$\begin{array}{} \displaystyle v \in H_0^1(\Omega ) \end{array}$, we define the energy functional Lb,ε(t, v) as

Lb,ϵ(t,v)=12(vL2(Ω)2+bvH01(Ω)2)bΩΦϵ(t,v),$$\begin{array}{} \displaystyle L_{b,\epsilon}(t,v)=\frac12\left(\|v\|^2_{L^2(\Omega)}+b\|v\|^2_{H^1_0(\Omega)}\right)-b\int_\Omega \Phi_\epsilon(t,v), \end{array}$$

with b > 0. It is easy to prove that for δ12b$\begin{array}{} \displaystyle \delta \leqslant \frac{1}{{2b}} \end{array}$,

Lb,ϵ(t,v)b2vH01(Ω)2bKδ$$\begin{array}{} \displaystyle {L_{b,}}(t,v)\geqslant \frac{b}{2}v_{H_0^1(\Omega )}^2 - b{K_\delta } \end{array}$$

and for any δ > 0,

Lb,ϵ(t,v)bλ1+2(1+bδ)2λ1vH01(Ω)2+bKδ,$$\begin{array}{} \displaystyle {L_{b,}}(t,v)\leqslant \frac{{b{\lambda _1} + 2(1 + b\delta )}}{{2{\lambda _1}}}\| v \|_{H_0^1(\Omega )}^2 + b{K_\delta }, \end{array}$$

uniformly in time t ∈ ℝ, with λ1 > 0 the first eigenvalue of A = −Δ with Dirichlet boundary conditions.

Now, assuming that (H3) holds true and using (12), for a solution u(t) = φε(t, Hε)u0 of (10), where Hε ∈ Γ ⋄ 𝒢ε, we have

ddtLb,ϵ(t,u)λ(t)(u,ut)H01(Ω)uH01(Ω)2+(u,hϵe(t,u))L2(Ω)butL2(Ω)2+λ(t)utH01(Ω)2+C1λ1γ1η+2δ2λ1uH01(Ω)2+λ(t)12ηb2ηutH01(Ω)2+C,$$\begin{array}{l} \frac{d}{dt}L_{b,\epsilon}(t,u) &\leqslant-\lambda(t)(u,u_t)_{H_0^1(\Omega)}-\|u\|^2_{H_0^1(\Omega)}\\&\qquad +(u,h^e_\epsilon(t,u))_{L^2(\Omega)}-b\left(\|u_t\|^2_{L^2(\Omega)}+\lambda(t)\|u_t\|^2_{H_0^1(\Omega)} \right)+C\\ &\leqslant -\left(1-\frac{\lambda_1\gamma_1\eta+2\delta}{2\lambda_1} \right)\|u\|^2_{H_0^1(\Omega)}+\lambda(t)\frac{1-2\eta b}{2\eta}\|u_t\|^2_{H_0^1(\Omega)}+C, \end{array}$$

for δ, η > 0 and taking δλ1γ12,λ1,η<2(λ1δ)λ1γ1andb>12η$\delta \in \left( {{\lambda _1} - \frac{{{\gamma _1}}}{2},{\lambda _1}} \right),\eta \lt \frac{{2({\lambda _1} - \delta )}}{{{\lambda _1}{\gamma _1}}}\;{\text{and}\; b} \gt \frac{1}{{2\eta }}$, we have

ddtLb,ϵ(t,u)kLb,ϵ(t,u)+C,$$\begin{array}{} \displaystyle \frac{d}{{dt}}{L_{b,}}(t,u) \leqslant - k{L_{b,}}(t,u) + C, \end{array}$$

with k, C > 0 that do not depend neither on time t ∈ ℝ nor on ε ∈ [0,1]. Therefore,

u(t)H01(Ω)2Ku0H01(Ω)2ekt+C,$$\begin{equation} \|u(t)\|^2_{H_0^1(\Omega)}\leqslant K\|u_0\|^2_{H_0^1(\Omega)}e^{-kt}+C, \end{equation}$$

for certain constants k > 0 and K, C ⩾ 0 which do not depend on time, ε and Hϵ ∈ Γ ⋄ 𝒢ϵ, and thus we have the global existence of solutions for (10). To summarise, we have so far the following result

Theorem 14

Assume that conditions(H1)-(H3)and(C)are satisfied. Then, for each ε ∈ [0,1], Hϵ ∈ Γ ⋄ 𝒢ϵandu0H01(Ω)$\begin{array}{} \displaystyle {u_0} \in H_0^1(\Omega ) \end{array}$, equation (10) has a solution φ (·, Hε)u0 defined for all t ⩾ 0. Moreover, there exists abounded subset B0 ofH01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$, independent of ε ∈ [0,1], such that for each bounded subsetBH01(Ω)$\begin{array}{} \displaystyle B \subset H_0^1(\Omega ) \end{array}$, there exists T = T (B) ⩾ 0 such that for tT we have

φϵ(t,Hϵ)BB0.$$\begin{array}{} \displaystyle \varphi_\epsilon(t,H_\epsilon)B\subset B_0. \end{array}$$

Proof. It is simple to see that, using (16), each solution given by Theorem 13 exists for all t ⩾ 0. Now define

B0={vH01(Ω):vH01(Ω)22C},$$ B_0=\{v\in H^1_0(\Omega)\colon \|v\|^2_{H^1_0(\Omega)}\leqslant 2C\}, $$

where C is given in (16). Given B a bounded subset of H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$, set B=supvBvH01(Ω)$\|B\|=\sup_{v\in B}\|v\|_{H^1_0(\Omega)}$ and choose

T=1kln(KB2C),$$ T=\frac1k\ln\left(\frac{K\|B\|^2}{C}\right), $$

where k and K are given in (16).

Since Σε ⊂ Γ ⋄ 𝒢ε and Σε is a compact invariant set for the group of translations {θt: t ∈ ℝ} in 𝒞*, a simple consequence of Theorem 14 is the following:

Theorem 15

Assume that conditions(H1)-(H3)and(C)are satisfied. Then, for each ε ∈ [0,1], equation (10) generates a non-autonomous dynamical systemε, θ) in(H01(Ω),Σϵ)$\begin{array}{} \displaystyle (H^1_0(\Omega),\Sigma_\epsilon) \end{array}$, where for each ε ∈ [0,1], Hε ∈ Σ andu0H01(Ω)$\begin{array}{} \displaystyle {u_0} \in H_0^1(\Omega ) \end{array}$, the function+tφε (t, Hε)u0is the unique solution of (10).

Skew-product semiflows for (10)

Using (3), we are able to define, for each ε ∈ [0,1], the skew-product semiflow

{Πϵ(t):t0} in Xϵ=H01(Ω)×Σϵ$$\begin{array}{} \displaystyle \{\Pi_\epsilon(t)\colon t \geqslant 0\} \hbox{ in } \mathbb{X}_\epsilon= H^1_0(\Omega)\times \Sigma_\epsilon \end{array}$$

associated with (φϵ,θ)(H01(Ω),Σϵ)$\begin{array}{} \displaystyle (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma_\epsilon)} \end{array}$, by setting

Πϵ(t)(u0,Hϵ)=(φϵ(t,Hϵ)u0,θtHϵ),$$\begin{array}{} \displaystyle \Pi_\epsilon(t)(u_0,H_\epsilon)=(\varphi_\epsilon(t,H_\epsilon)u_0,\theta_tH_\epsilon), \end{array}$$

for all t ⩾ 0, (u0, Hε) ∊ ℝε and ε ∈ [0,1].

Different attractors for (10)

In this section we will use the result of Section 2 to obtain different attractors for equation (10), for the different frameworks described in Subsection 1.1.

Global attractors for skew-product semiflows

In this section, we will use the theory of autonomous equations to find a global attractor for each one of the skew-product semiflows defined by (17).

We first will see a very simple result, that follows from Theorem 14.

Proposition 16.

Assume that(H1)-(H3)and(C)are satisfied and fix ε ∈ [0,1]. Then there exists a bounded setBϵXϵ$\begin{array}{} \displaystyle \mathbb{B}_\epsilon\subset \mathbb{X}_\epsilon \end{array}$such that given a bounded subsetBXϵ$\begin{array}{} \displaystyle \mathbb{B}\subset \mathbb{X}_\epsilon \end{array}$there exists T = T (B) ⩾ 0 such that

Πϵ(t)BBϵ,foralltT.$$\begin{array}{l} \displaystyle \Pi_\epsilon(t)\mathbb{B}\subset \mathbb{B}_\epsilon, {for \; all\; }t \geqslant T. \end{array}$$

Proof. Let B0H01(Ω)$\begin{array}{} \displaystyle {B_0} \subset H_0^1(\Omega ) \end{array}$ be as in Theorem (14) and define BϵB0×Σϵ$\begin{array}{} \displaystyle \mathbb{B}_\epsilon\doteq B_0\times \Sigma_\epsilon \end{array}$. Since B0 is bounded in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$ and Σ is compact space, we have that 𝔹ε is bounded in 𝕏ε. Moreover, if 𝔹 is a bounded subset of 𝕏ε, we have that BB×Σϵ$\begin{array}{} \displaystyle \mathbb{B}\subset B\times \Sigma_\epsilon \end{array}$, for some bounded set BH01(Ω)$\begin{array}{} \displaystyle B \subset H_0^1(\Omega ) \end{array}$. Hence, if T is as in Theorem 14 we obtain the result.

Asymptotical compactness

In order to obtain a global attractor for each skew-product semiflow, we must prove that the semigroups {Πε(t): t ⩾ 0} are asymptotically compact. That is our goal for the next few results.

Lemma 17

Define Hλ(t,v)=A˜λ(t)vC*$\begin{array}{} \displaystyle {H_\lambda }(t,v) = - {\tilde A_\lambda }(t)v \in {{\cal C}_*} \end{array}$. Then the problem

{u˙=Hλ(t,u),t>0u(0)=u0H01(Ω)$$ \left\{\begin{array}{*{20}{l}} {\dot u = {H_\lambda }(t,u),\;t \gt 0}\\ {u(0) = {u_0} \in H_0^1(\Omega )} \end{array}\right.$$

has a unique solution φχ(t) defined for allt ⩾ 0 given by

φλ(t)u0=u00tA~λ(s)φλ(s)u0ds,forallt0.$$\begin{array}{l} \displaystyle {\varphi _\lambda }(t){u_0} = {u_0} - \int_0^t {{{\tilde A}_\lambda }} (s){\varphi _\lambda }(s){u_0}\;ds,{\;{ for \;all }\;}t \geqslant 0. \end{array}$$

and also, there exists constants K, k > 0 which do not depend on λ such that

φλ(t)u0H01(Ω)Ku0H01(Ω)ekt,forallt0.$$\begin{equation} \|\varphi_\lambda(t)u_0\|_{H^1_0(\Omega)}\leqslant K\|u_0\|_{H^1_0(\Omega)}e^{-kt}, \;{ for \;all }\; t\geqslant 0. \end{equation}$$

Proof. The local existence and uniqueness of φ* follows from Theorem 13. Proceeding as in Subsection 4.2 with hϵ ≡ 0, we can see that we can take C = 0 in (16) and gives us the global existence of φ* and (19). Equation (18) is a simple consequence of the theory of ordinary differential equations, since A˜λ(t)$\begin{array}{} \displaystyle - {\tilde A_\lambda }(t) \end{array}$ is a uniformly bounded operator of H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$, and the bounds do not depend on λ ∈ Γ.

Lemma 18

IfHϵ=BλhϵeA˜λΣϵ$\begin{array}{} \displaystyle H_\epsilon=B_\lambda h^e_\epsilon-\tilde{A}_\lambda \in \Sigma_\epsilon \end{array}$ and u0H01(Ω)$\begin{array}{} \displaystyle u_0\in H^1_0(\Omega) \end{array}$then

φϵ(t,Hϵ)u0=φλ(t)u0+ψ(t)(u0,Hϵ),foreacht0,$$\begin{array}{} \displaystyle \varphi_\epsilon(t,H_\epsilon)u_0=\varphi_\lambda(t)u_0+\psi(t)(u_0,H_\epsilon), \;{ for \;each }\; t \geqslant 0, \end{array}$$

whereψ(t)(u0,Hϵ)0tBλ(s)hϵe(s,φϵ(s,Hϵ)u0)ds$\begin{array}{} \displaystyle \psi(t)(u_0,H_\epsilon)\doteq \int_0^tB_{\lambda}(s)h_\epsilon^e(s,\varphi_\epsilon(s,H_\epsilon)u_0)ds \end{array}$. Moreover, themap ψ(t) is a compact map fromH01(Ω)×Σϵ$\begin{array}{} \displaystyle H^1_0(\Omega)\times \Sigma_\epsilon \end{array}$intoH01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$.

Proof. Clearly the right side of the equation is a solution of (10), thus the uniqueness shows the equality. Now let B be a bounded set of H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$. By (16) the set B{φϵ(s,Hϵ)B:s[0,t],HϵΣϵ}$\begin{array}{} \displaystyle \mathscr{B}\doteq \{\varphi_\epsilon(s,H_\epsilon)B\colon s\in[0,t], \ H_\epsilon\in \Sigma_\epsilon\} \end{array}$ is bounded and thus Lemma 7 ensures that hϵGϵhϵe([0,t],B)$\begin{array}{} \displaystyle \cup_{h_\epsilon\in \mathscr{G}_\epsilon}h_\epsilon^e([0,t],\mathscr{B} \end{array}$ is a precompact set of H—1. The fact that Bλ (t) is a uniformly bounded bounded linear operator for t ∈ ℝ and λ ∈ Γ concludes the proof.

Using these two lemmas, we are able to prove the asymptotical compactness for the skew-product semiflows.

Proposition 19

The skew-product semiflowϵ(t): t ⩾ 0} is asymptotically compact, for each ε ∈ [0,1].

Proof. Let {un}n∈ℕ be a bounded sequence in H01(Ω),{Hϵ,n}n$\begin{array}{} \displaystyle H_0^1(\Omega ),{\{ {H_{,n}}\} _{n \in \mathbb{N}}} \end{array}$ a bounded sequence in Σε and {tn}n∈ℕ be a real sequence with tn → ∞ as n → ∞ and such that the sequence {Πε(tn)(un, Hε,n)} is bounded. We have

Πϵ(tn)(un,Hϵ,n)=(φϵ(tn,Hϵ,n)un,θtnHϵ,n), for all nN.$$ \Pi_\epsilon(t_n)(u_n,H_{\epsilon,n})=(\varphi_\epsilon(t_n,H_{\epsilon,n})u_n,\theta_{t_n}H_{\epsilon,n}), \hbox{ for all } n\in \mathbb{N}. $$

Since Σ is compact, we can assume, up to a subsequence, that there exists Hε ε Σε such that Hϵ=C*limnθtnHϵ,n$\begin{array}{} \displaystyle H_{\epsilon}=\mathscr{C}_*-\displaystyle \lim_{n\to \infty}\theta_{t_n}H_{\epsilon,n} \end{array}$.

Now using Lemma 18 we can write

φϵ(tn,Hϵ,n)un=φ*(tn)un+ψ(tn)(un,Hϵ,n), for each n.$$\begin{array}{} \displaystyle \varphi_\epsilon(t_n,H_{\epsilon,n})u_n=\varphi_*(t_n)u_n+\psi(t_n)(u_n,H_{\epsilon,n}), \hbox{ for each } n\in \mathbb{N}. \end{array}$$

Since {un}n∈ℕ is bounded in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$, Lemma 17 implies that φ*(tn)un → 0 as n → ∞. It is now simple to see that the sequence {φε(tn,Hε,n)un}n∈ℕ is precompact in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$, which proves that the sequence {Πε(tn)(un, Hε,n)} has a convergent subsequence in 𝕏ε and concludes the proof.

Now we can join the results of Propositions 16 and 19, together with Theorem 1, to obtain the next theorem.

Theorem 20

(Existence of the global attractor). Assume that(H1)-(H3)and(C)hold. Then the skew-product semiflowε(t): t ⩾ 0} associated with the non-autonomous dynamical system(φϵ,θ)(H01(Ω),Σ)$\begin{array}{} \displaystyle (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma)} \end{array}$has a global attractor 𝔸εin 𝕏εfor each ε ∈ [0,1]. Moreover

AϵB0×Σϵ,foreachε[0,1],$$\begin{array}{} \displaystyle \mathbb{A}_\epsilon\subset B_0\times \Sigma_\epsilon,\;{for\;each}\; \varepsilon\in [0,1], \end{array}$$

where B0 is the bounded set given in Theorem 14.

We know that the attractors 𝔸ε can be characterized by

Aϵ={(u0,Hϵ)Xϵ: there exists a bounded global solution ηϵ of {Πϵ(t):t0} through (u0,Hϵ)}.$$\begin{array}{} \displaystyle \mathbb{A}_\epsilon=\{(u_0,H_\epsilon)\in \mathbb{X}_\epsilon\colon &\hbox{ there exists a bounded global solution } \eta_\epsilon \\& \hbox{ of } \{\Pi_\epsilon(t)\colon t \geqslant 0\} \hbox{ through } (u_0,H_\epsilon)\}. \end{array}$$

It is not difficult to see that we can write ηε as

ηϵ(t)=(ξϵ(t),θtHϵ),$$\begin{array}{} \displaystyle \eta_\epsilon(t)=(\xi_\epsilon(t),\theta_tH_\epsilon), \end{array}$$

where ξ is a global solution of (10); that is, φε(ts, θsHεε(s) = ξε(t) for all ts, and ξ (0) = u0. Using Lemma 18 we can write

ξϵ(t)=φϵ(ts,θsHϵ)ξϵ(s)=φλ(ts)ξϵ(s)+ψ(ts)(ξϵ(s),θsHϵ),$$\begin{array}{} \displaystyle \xi_\epsilon(t)=\varphi_\epsilon(t-s,\theta_sH_\epsilon)\xi_\epsilon(s)=\varphi_\lambda(t-s)\xi_\epsilon(s)+\psi(t-s)(\xi_\epsilon(s),\theta_sH_\epsilon), \end{array}$$

where

ψ(ts)(ξϵ(s),θsHϵ)=0tsBθsλ(r)θshϵe(r,φ(r,θsHϵ)ξϵ(s))dr=0tsBλ(r+s)hϵe(r+s,ξϵ(r+s))dr=stBλ(r)hϵe(r,ξϵ(r))dr.$$\begin{array}{} \displaystyle \psi(t-s)(\xi_\epsilon(s),\theta_sH_\epsilon)&=\int_0^{t-s}B_{\theta_s\lambda}(r)\theta_sh_\epsilon^e(r,\varphi(r,\theta_sH_\epsilon)\xi_\epsilon(s))dr \\&=\int_0^{t-s}B_\lambda(r+s)h^e_\epsilon(r+s,\xi_\epsilon(r+s))dr\\ &=\int_s^tB_\lambda(r)h_\epsilon^e(r,\xi_\epsilon(r))dr. \end{array}$$

Now, making s → — ∞ in (20), since {ξε(s)}s∈ℝ is bounded in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$ we obtain, using (19), that

ξ(t)=tBλ(r)hϵe(r,ξ(r))dr.$$\begin{array}{} \displaystyle \xi(t)=\int_{-\infty}^tB_\lambda(r)h_\epsilon^e(r,\xi(r))dr. \end{array}$$

Using Lemma 7 we can show (following the ideas of [26]) that the solution ξε is bounded in H2(Ω)H01(Ω)$\begin{array}{} \displaystyle {H^2}(\Omega ) \cap H_0^1(\Omega ) \end{array}$, and the bound does not depend on the particular ξε (neither it does on ε). Hence we obtain that there exists a bounded subset D0 of H2(Ω)H01(Ω)$\begin{array}{} \displaystyle {H^2}(\Omega ) \cap H_0^1(\Omega ) \end{array}$ such that

AϵD0×Σϵ, for each ϵ[0,1].$$\begin{array}{} \displaystyle \mathbb{A}_\epsilon \subset D_0\times \Sigma_\epsilon, \hbox{ for each } \epsilon\in[0,1]. \end{array}$$

Other attractors

Now using Theorem 20 we are able to obtain other attractors for equation (10).

Assume that(H1)-(H3)and(C)hold true. Then we have, for each ε ∈ [0,1], that

(a)the non-autonomous dynamical system (φϵ,θ)(H01(Ω),Σϵ)$\begin{array}{} \displaystyle (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma_\epsilon)} \end{array}$ has a uniform attractor 𝒜 and

Aϵ=πH01(Ω)AϵD0;$$\begin{array}{} \displaystyle \mathscr{A}_\epsilon =\pi_{H^1_0(\Omega)}\mathbb{A}_\epsilon \subset D_0; \end{array}$$

(b)the the non-autonomous dynamical system(φϵ,θ)(H01(Ω),Σϵ)$\begin{array}{} (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma_\epsilon)} \end{array}$ has a cocycle attractor {A(H)}H∈Σε with ∪Hε∈ ΣεA(Hε) ⊂ D0, and

A(Hϵ)={vH01(Ω):(v,Hϵ)Aϵ};$$\begin{array}{} \displaystyle A(H_\epsilon)=\{v\in H^1_0(\Omega)\colon (v,H_\epsilon)\in \mathbb{A}_\epsilon\}; \end{array}$$

(c)for each Hε ∈ Σε, the evolution process {THe (t, s): ts} given by

THε (t, s) = φε (t — s, θsHε), for all t ⩾ s,

has a pullback attractor {AHε(t)}t ∈ ℝ with ∪t ∈ ℝAHe(t) ⊂ D0, and

Aϵ=HϵΣϵ[tAHϵ(t)×{Hϵ}].$$\begin{array}{} \displaystyle \mathbb{A}_\epsilon=\bigcup_{H_\epsilon\in \Sigma_\epsilon}\left[\bigcup_{t\in \mathbb{R}}A_{H_\epsilon}(t)\times \{H_\epsilon\}\right]. \end{array}$$

where D0is given in (21).

Proof. Using Theorem 20, we have easily that item (a) follows from Theorem 5, item (b) from Theorem 3 and item (c) from Theorem 4.

Upper semicontinuity of attractors

This section is devoted to study the upper semicontinuity of the semigroups {Πε(t): t ⩾ 0} as perturbations of {Π0(t): t ⩾ 0}, as a part of the study described in Subsection 1.2. So far, we have treated the family of equations (6) (and equivalently, (10)) individually for each ε ∈ [0,1], but now it is time to look at all these equations together at once.

Assuming that (H1)-(H3) and (C) hold, we obtained so far a family of semigroups {Πε(t): t ⩾ 0} in Xϵ=H01(Ω)×Σϵ$\begin{array}{} \displaystyle \mathbb{X}_\epsilon=H^1_0(\Omega)\times \Sigma_\epsilon \end{array}$, and for each ε, a global attractor åε.

Definition 11.

We say that a family {Kε}ε∈[01] is upper semicontinuous at 0 in a metric space (X, d) if given sequences n}n∈ℕ (0,1] and xn ∈ Kεn, with εn ∈ 0+ as n → ∞, there exists a convergent subsequence of {xn}n ∈ ℕ with limit belonging to the closure of K0 in (X, d).

The previous definition is equivalent to the following: a family {Kε }ε∈[0,1] is upper semicontinuous at 0 in a metric space (X, d) if

limϵ0+dH(Kϵ,K0)=0.$$\begin{array}{} \displaystyle \lim_{\epsilon \to 0^+}{\rm d}_H(K_\epsilon,K_0)=0. \end{array}$$

To prove the upper semicontinuity of {Aϵ}ϵ[0,1]$\begin{array}{} \displaystyle \{\mathbb{A}_\epsilon\}_{\epsilon\in [0,1]} \end{array}$, we set the base space as H01(Ω)×L$\begin{array}{} \displaystyle H_0^1(\Omega ) \times {{\scr L}_ * } \end{array}$, with a metric ∂ defined by

d[(u1,H1),(u2,H2)]=u1u2H01(Ω)+d(H1,H2),$$\begin{array}{} \displaystyle \mathfrak{d}[(u_1,H_1),(u_2,H_2)]=\|u_1-u_2\|_{H^1_0(\Omega)}+d_\ast(H_1,H_2), \end{array}$$

for all (u1,H1),(u2,H2)H01(Ω)×L$\begin{array}{} \displaystyle ({u_1},{H_1}),({u_2},{H_2}) \in H_0^1(\Omega ) \times {{\scr L}_ * } \end{array}$.

Before studying the upper semicontinuity of the family of attractors {åε}ε∈[0,1], we will need some convergence results.

Lemma 22.

If(H4)holds and hε𝒢ε, we have

supt|hϵ(t,s)f(s)|β(ϵ)(1+|s|ρ1),$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}|h_\epsilon(t,s)-f(s)|\leqslant \beta(\epsilon)(1+|s|^{\rho-1}), \end{array}$$

for all s ∈ ℝ.

Proof. Let hε𝒢ε and (tm} be a real sequence such that

limmd2(hϵ,θtmgϵ)=limmsupt,s|hϵ(t,s)θtmgϵ(t,s)|=0,$$\begin{array}{} \displaystyle \lim_{m\to \infty}d_2(h_\epsilon,\theta_{t_m}g_\epsilon)=\lim_{m\to \infty}\sup_{t,s\in \mathbb{R}}|h_\epsilon(t,s)-\theta_{t_m}g_\epsilon(t,s)|=0, \end{array}$$

where d2 is as in Subsection 3.1. Hence

suptR|hϵ(t,s)f(s)|suptR|hϵ(t,s)gϵ(t+tm,s)|+suptR|gϵ(t+tm,s)f(s)|d2(hϵ,θtmgϵ)+suptR|gϵ(t,s)f(s)|d2(hϵ,θtmgϵ)+β(ϵ)(1+|s|ρ1),$$\begin{equation} \begin{split} \sup\limits_{t\in \mathbb{R}}|h_\epsilon(t,s)&-f(s)|\leqslant \sup\limits_{t\in \mathbb{R}}|h_\epsilon(t,s)-g_{\epsilon}(t+t_m,s)|+\sup\limits_{t\in \mathbb{R}}|g_{\epsilon}(t+t_m,s)-f(s)| \\&\leqslant d_2(h_\epsilon,\theta_{t_m}g_\epsilon)+\sup\limits_{t\in \mathbb{R}}|g_{\epsilon}(t,s)-f(s)|\\& \leqslant d_2(h_\epsilon,\theta_{t_m}g_\epsilon) + \beta(\epsilon)(1+|s|^{\rho-1}), \end{split} \end{equation}$$

and making m → ∞ we obtain the result.

Lemma 23.

Assume that(H4)and(C)hold. Then there exists a constant c > 0 such that

suphϵGϵsupthϵe(t,u)fe(u)L2nn+2(Ω)cβ(ϵ)(1+uH01(Ω)ρ1),$$\begin{array}{} \displaystyle \sup_{h_\epsilon\in \mathscr{G}_\epsilon}\sup_{t\in \mathbb{R}}\|h_\epsilon^e(t,u)-f^e(u)\|_{L^{\frac{2n}{n+2}}(\Omega)}\leqslant c\beta(\epsilon)(1+\|u\|^{\rho-1}_{H^1_0(\Omega)}), \end{array}$$

for alluH01(Ω)$\begin{array}{} \displaystyle u \in H_0^1(\Omega ) \end{array}$.

Proof. Proceeding as in Lemma 7 and using Lemma 22 we have

suptRhϵe(t,u)fe(u)L2nn+2(Ω)=Ω|hϵe(t,u)fe(u)|2nn+2n+22nβ(ϵ)Ω(1+|u|ρ1)2nn+2n+22nc~β(ϵ)(1+uLn(ρ1)2(Ω)ρ1)cβ(ϵ)(1+uH01(Ω)ρ1),$$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}\|h_\epsilon^e(t,u)&-f^e(u)\|_{L^{\frac{2n}{n+2}}(\Omega)}=\left[\int_\Omega |h^e_\epsilon(t,u)-f^e(u)|^{\frac{2n}{n+2}}\right]^{\frac{n+2}{2n}}\\ &\leqslant \beta(\epsilon)\left[\int_\Omega(1+|u|^{\rho-1})^{\frac{2n}{n+2}}\right]^{\frac{n+2}{2n}}\leqslant \tilde{c}\beta(\epsilon)(1+\|u\|^{\rho-1}_{L^{\frac{n(\rho-1)}{2}}(\Omega)})\\ &\leqslant c\beta(\epsilon)(1+\|u\|^{\rho-1}_{H^1_0(\Omega)}), \end{array}$$

and the result follows.

Corollary 24.

If(H4)and(C)hold, we have that there exists a constantC^>0$\begin{array}{} \displaystyle \hat C > 0 \end{array}$such that

suphϵGϵd2e(hϵe,fe)C^β(ϵ).$$\begin{array}{} \displaystyle \sup_{h_\epsilon\in \mathscr{G}_\epsilon}d^e_2(h^e_\epsilon,f^e)\leqslant \hat{C}\beta(\epsilon). \end{array}$$

Proposition 25.

Assume that(H4)and(C)hold true and consider the familyε }ε∈[01]given in Definition 10. Then we have that given sequencesn}n ∈ ℕ ⊂ (0,1] with εn → 0+ and Hn ∈ Σεn, for each n ∈ ℕ, there exists a convergent subsequence of (Hn}n ∈ ℕ in (𝒞*, d*), with its limit belonging to Σ0.

Proof. Since Hn ∈ , item (b) of Lemma 12 implies that there exists λn ∈ Γεn and hn𝒢εn such that

Hn = BXnhen - AXn, for each n e N.

Hn=BλnhneA˜λn, for each n.$$\begin{array}{} \displaystyle H_n = B_{\lambda_n}h_n^e-\tilde{A}_{\lambda_n}, \hbox{ for each } n\in \mathbb{N}. \end{array}$$

Since Γ is compact (recall Remark 7), there exists a subsequence nk} that converges to a function λ0 in (𝒞1,d1). Now, Corollary 24 shows that d2e(hne,fe)C^β(ϵn)0$\begin{array}{} \displaystyle d_2^e(h_n^e,{f^e}) \leqslant \hat C\beta ({\epsilon_n}) \to 0 \end{array}$ as n → ∞ and hence hne$\begin{array}{} \displaystyle h_n^e \end{array}$ converges to fe in (C2e,d2e)$\begin{array}{} \displaystyle ({\scr C}_2^e,d_2^e) \end{array}$. Therefore, we can easily see that Hnk converges to Bλ0feA˜λ0$\begin{array}{} \displaystyle B_{\lambda_0}f^e-\tilde{A}_{\lambda_0} \end{array}$ in (𝒞*, d*), which is in Σ0 by item (a) of Lemma 12.

With these preliminaries results, we are able to begin the proof of the upper semicontinuity of the family of the global attractors (åε}ε∈[01] of the skew-product semiflows, at ε = 0.

Lemma 26.

If {(uε,Hε)}ε ∈(0,1] is such that (uϵ,Hϵ)Xϵ$\begin{array}{} \displaystyle (u_\epsilon,H_\epsilon)\subset \mathbb{X}_\epsilon \end{array}$ and there exists (u0,H0)X0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{X}_0 \end{array}$ such that d[(uϵ,Hϵ),(u0,H0)]0$\begin{array}{} \displaystyle \mathfrak{d}[(u_\epsilon,H_\epsilon),(u_0,H_0)]\to 0 \end{array}$ as ε → 0+, we have

d[Πϵ(t)(uϵ,Hϵ),Π0(t)(u0,H0)]ϵ0+0,foreacht0.$$\begin{array}{} \displaystyle \mathfrak{d}[\Pi_\epsilon(t)(u_\epsilon,H_\epsilon),\Pi_0(t)(u_0,H_0)]\stackrel{\epsilon\to 0^+}{\longrightarrow} 0, \;{ for \; each }\; t \geqslant 0. \end{array}$$

Proof. We can write

Π(t)(uϵ,Hϵ)=(φϵ(t,Hϵ)uϵ,θtHϵ), for each ϵ[0,1].$$\begin{array}{} \displaystyle \Pi(t)(u_\epsilon,H_\epsilon)=(\varphi_\epsilon(t,H_\epsilon)u_\epsilon,\theta_tH_\epsilon), \hbox{ for each } \epsilon \in [0,1]. \end{array}$$

Since d*(Hε,H0) → 0 and {θ: t⩾0} is continuous in 𝒞* for each t ⩾ 0, we easily obtain that d*tHεtH0) → 0, as ε → 0+.

It remains to show that φϵ(t,Hϵ)uϵφ0(t,H0)u0H01(Ω)0$\begin{array}{} \|\varphi_\epsilon(t,H_\epsilon)u_\epsilon-\varphi_0(t,H_0)u_0\|_{H^1_0(\Omega)}\to 0 \end{array}$, as ε → 0+. Using Lemma 18, we can write

φϵ(t,Hϵ)uϵφ0(t,H0)u0=φλϵ(t)uϵφλ0(t)u0+0t[Bλϵ(S)hϵe(s,φϵ(s,Hϵ)uϵ)Bλ0(S)fe(φ0(s,H0)u0)]ds,$$\begin{array}{} \displaystyle \varphi_\epsilon&(t,H_\epsilon)u_\epsilon-\varphi_0(t,H_0)u_0 \\ &= \varphi_{\lambda_\epsilon}(t)u_\epsilon-\varphi_{\lambda_0}(t)u_0+\int_0^t [B_{\lambda_\epsilon}(s)h_\epsilon^e(s,\varphi_\epsilon(s,H_\epsilon)u_\epsilon)-B_{\lambda_0}(s)f^e(\varphi_0(s,H_0)u_0)]ds, \end{array}$$

where we assumed Hϵ=BλϵhϵeA˜λϵ$\begin{array}{} \displaystyle H_\epsilon=B_{\lambda_\epsilon}h_\epsilon^e-\tilde{A}_{\lambda_\epsilon} \end{array}$ and H0=Bλ0feA˜λ0$\begin{array}{} \displaystyle H_0=B_{\lambda_0}f^e-\tilde{A}_{\lambda_0} \end{array}$.

We have, using Lemma 17, we obtain

φλϵ(t)uϵφλ0(t)u0H01(Ω)φλϵ(t)uϵφλϵ(t)u0H01(Ω)+φλϵ(t)u0φλ0(t)u0H01(Ω)Kuϵu0H01(Ω)ekt+φλϵ(t)u0φλ0(t)u0H01(Ω).$$\begin{array}{} \displaystyle \|\varphi_{\lambda_\epsilon}(t)u_\epsilon&-\varphi_{\lambda_0}(t)u_0\|_{H^1_0(\Omega)} \leqslant \|\varphi_{\lambda_\epsilon}(t)u_\epsilon-\varphi_{\lambda_\epsilon}(t)u_0\|_{H^1_0(\Omega)}+\|\varphi_{\lambda_\epsilon}(t)u_0-\varphi_{\lambda_0}(t)u_0\|_{H^1_0(\Omega)}\\ &\leqslant K\|u_\epsilon-u_0\|_{H^1_0(\Omega)}e^{-kt}+\|\varphi_{\lambda_\epsilon}(t)u_0-\varphi_{\lambda_0}(t)u_0\|_{H^1_0(\Omega)}. \end{array}$$

Again, using Lemma 17, item 2 of Remark 7 and the Gronwall inequality, we obtain that

φλϵ(t)uϵφλ0(t)u0H01(Ω)=O(ϵ).$$\begin{array}{} \displaystyle \|\varphi_{\lambda_\epsilon}(t)u_\epsilon-\varphi_{\lambda_0}(t)u_0\|_{H^1_0(\Omega)} ={O}(\epsilon). \end{array}$$

For the second term, we have

Bλϵ(s)hϵe(s,φϵ(s,Hϵ)uϵ)Bλ0(s)fe(φ0(s,H0)u0)H01(Ω)Bλϵ(s)[hϵe(s,φϵ(s,Hϵ)uϵ)fe(φϵ(s,Hϵ)uϵ)]H01(Ω)+[Bλϵ(s)Bλ0(s)]fe(φ0(s,H0)u0)H01(Ω)$$\begin{equation} \begin{split} \|B_{\lambda_\epsilon}(s)h_\epsilon^e(s,\varphi_\epsilon(s,H_\epsilon)u_\epsilon)&-B_{\lambda_0}(s)f^e(\varphi_0(s,H_0)u_0)\|_{H^1_0(\Omega)}\\ & \leqslant \|B_{\lambda_\epsilon}(s)[h_\epsilon^e(s,\varphi_\epsilon(s,H_\epsilon)u_\epsilon) - f^e(\varphi_\epsilon(s,H_\epsilon)u_\epsilon)]\|_{H^1_0(\Omega)} \\&+\|[B_{\lambda_\epsilon}(s)-B_{\lambda_0}(s)]f^e(\varphi_0(s,H_0)u_0)\|_{H^1_0(\Omega)} \end{split} \end{equation}$$

and hence, using again item 2 of Remark 7 we obtain that

0tBλϵ(s)hϵe(s,φϵ(s,Hϵ)uϵ)Bλ0(s)fe(φ0(s,H0)u0)H01(Ω)O(ϵ)+0tφϵ(s,Hϵ)uϵφ0(s,H0)u0H01(Ω)ds.$$\begin{array}{} \displaystyle \int_0^t\|B_{\lambda_\epsilon}(s)h_\epsilon^e(s,\varphi_\epsilon(s,H_\epsilon)u_\epsilon)&-B_{\lambda_0}(s)f^e(\varphi_0(s,H_0)u_0)\|_{H^1_0(\Omega)}\\ & \leqslant {O}(\epsilon)+\int_0^t \|\varphi_\epsilon(s,H_\epsilon)u_\epsilon-\varphi_0(s,H_0)u_0\|_{H^1_0(\Omega)}ds. \end{array}$$

Finally, joining the estimates and applying again the Gronwall inequality, we obtain that

φϵ(t,Hϵ)uϵφ0(t,H0)u0H01(Ω)O(ϵ),$$\begin{array}{} \displaystyle \|\varphi_\epsilon(t,H_\epsilon)u_\epsilon-\varphi_0(t,H_0)u_0\|_{H^1_0(\Omega)}\leqslant {O}(\epsilon), \end{array}$$

and concludes the result.

We can prove the following:

Lemma 27.

If {(uε,Hε)}ε∈(0,1]is such that(uϵ,Hϵ)Aϵ$\begin{array}{} \displaystyle (u_\epsilon,H_\epsilon)\in \mathbb{A}_\epsilon \end{array}$and

limϵ0+d[(uϵ,Hϵ),(u0,H0)]=0$$\begin{array}{} \displaystyle \lim_{\epsilon\to 0^+}\mathfrak{d}[(u_\epsilon,H_\epsilon),(u_0,H_0)]= 0 \end{array}$$

for some(u0,H0)H01(Ω)×C$\begin{array}{} \displaystyle (u_0,H_0)\in H^1_0(\Omega)\times \mathscr{C}_\ast \end{array}$, then(u0,H0)A0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{A}_0 \end{array}$.

Proof. Using the characterization of global attractors in Subsection 2.1, we know that through each (uε, Hε) ∈ 𝔸ε we have a global bounded solution ξϵ:Xϵ$\begin{array}{} \displaystyle \xi_\epsilon\colon \mathbb{R}\to \mathbb{X}_\epsilon \end{array}$ of {∏ε(t) : t ⩾ 0}. To show that (u0,H0)A0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{A}_0 \end{array}$, it is sufficient to prove that through (u0, H0) there exists a global bounded solution ξ0:X0$\begin{array}{} \displaystyle \xi_0\colon \mathbb{R}\to \mathbb{X}_0 \end{array}$ of {∏ε(t) : t ⩾ 0}.

For any t ⩾ 0, we define ξ0(t) = ∏(t)(u0,H0). Since {Π0(t):t ⩾ 0} has a global attractor, the set ξ0([0,∞)) is bounded.

We now use an induction argument to define the solution for negative values of t. Consider the family {ξε(—1)}ε∈[0,1], which we can write as

ξϵ(1)=(uϵ(1),θ1Hϵ).$$\begin{array}{} \displaystyle \xi_\epsilon(-1)=(u_\epsilon(-1),\theta_{-1}H_\epsilon). \end{array}$$

Using (21), we have that the family {uε(—1)}ε∈[0,1] and hence there exists a sequence ε1,n → 0+ and a point u1H01(Ω)$\begin{array}{} \displaystyle {u_{ - 1}} \in H_0^1(\Omega ) \end{array}$ such that

uϵ1,n(1)u1, in H01(Ω).$$\begin{array}{} \displaystyle u_{\epsilon_{1,n}}(-1)\to u_{-1}, \hbox{ in } H^1_0(\Omega). \end{array}$$

We define then ξ0(—1) = (u—1, θ—1H0) and ξ(t) = ∏0(t + 1)(u—1—1H0), for —1t < 0. Clearly, using Lemma 26, we have

(uϵ1,n,Hϵ1,n)=Πϵ1,n(1)ξϵ1,n(1)Π0(1)ξ0(1),$$\begin{array}{} \displaystyle (u_{\epsilon_{1,n}},H_{\epsilon_{1,n}})=\Pi_{\epsilon_{1,n}}(1)\xi_{\epsilon_{1,n}}(-1)\to \Pi_0(1)\xi_0(-1), \end{array}$$

and thus (u0,H0) = ξ0(0) = ∏0(1)ξ0(—1).

Proceeding inductively, for each k ∈ ℕ, we obtain a subsequence {εk,n}n∈ℕ of {εk−1,n}n∈ℕ with εk,n → 0+ as n → ⊂ and a point ukH01(Ω)$\begin{array}{} \displaystyle u_{-k}\in H^1_0(\Omega) \end{array}$ such that if ξε(—k) = (uε(—k), θkHε) we have

uϵk,n(j)uj, for all j=1,,k.$$\begin{array}{} \displaystyle u_{\epsilon_{k,n}}(-j)\to u_{-j}, \hbox{ for all } j=1,\cdots, k. \end{array}$$

Defining ξ0(—k) = (uk, θkH0) and ξ0(t) = ∏0(t + k0(—k) for —kt < —k + 1, we have that

(uϵk,n,Hϵk,n)=Πϵk,n(1)ξϵk,n(j)Π0(j+1)ξ0(j+1), for each j=1,k,$$\begin{array}{} \displaystyle (u_{\epsilon_{k,n}},H_{\epsilon_{k,n}})=\Pi_{\epsilon_{k,n}}(1)\xi_{\epsilon_{k,n}}(-j)\to \Pi_0(-j+1)\xi_0(-j+1), \hbox{ for each } j=1\cdots,k, \end{array}$$

and thus (u0,H0) = ξ0(0) = ∏0(k0(—k).

Therefore, we obtain that ξ0:X0$\begin{array}{} \displaystyle \xi_0\colon \mathbb{R}\to \mathbb{X}_0 \end{array}$ is a bounded global solution of {∏0(t): t ⩾ 0} through (u0,H0), which implies that (u0,H0)A0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{A}_0 \end{array}$ and concludes the proof.

Now, we can easily prove the upper semicontinuity of the family of global attractors {𝔸ε}ε∈[0,1].

Theorem 28.

The family of global attractors {𝔸ε}ε∈[0,1]is upper semicontinuous at 0.

Proof. If {εn, Hn)}n∈ℕ ⊂ (0, 1] with εn → 0 and (un, Hn) ∈ 𝔸εn, it is clear that there exists a convergent subsequence of {(un, Hn)}n∈ℕ to a point (u0,H0)X0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{X}_0 \end{array}$ (using (21)). Hence, Lemma 27 shows that (u0,H0)A0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{A}_0 \end{array}$, which concludes the proof.

Upper semicontinuity for other attractors

As in immediate consequence of Theorem 28 we have (recall Theorem 21):

Corollary 29.

Assume that(H1)-(H4)and(C)hold true. Then we have that

the family of uniform attractors {𝒜ε }ε∈[0,1];

the family of cocycle attractors {A(Hε)}Hε∈Σεand

the family of pullback attractors {AHε(t)}t∈ℝ

are upper semicontinuous at 0 inH01(Ω)$\begin{array}{} \displaystyle H^1_0(\Omega) \end{array}$.

Topological structure of attractors

Following the results of [7, Section 3], we will study the structure of the global attractors 𝔸ε for the skew-product semiflows (∏ε(t): t ⩾ 0} and using this structure we obtain informations about the structure for the other attractors defined in Theorem 21.

Structure of 𝔸0

To study the structure of the global attractors 𝔸ε, we will study in more detail the structure of the global attractor 𝔸0 and we will make the following assumption:

There exists a finite number of isolated equilibia = {e1, … ,ep} of

{Δu=f(u),in Ωu=0,on Ω.$$\begin{array}{} \displaystyle \left\{ \begin{array}{*{20}{l}} { - {\rm{\Delta }}u = f(u),}&{{\rm{in\;\Omega }}}\\ {\quad \;\:{\rm{n\& }}\quad u = 0,}&{{\rm{on\;}}\partial {\rm{\Omega }}.} \end{array}\right. \end{array}$$

With this assumption, we define

Ei={ei}×Σ0X0, for i=1,,p.$$\begin{array}{} \displaystyle \mathbb{E}_i=\{e_i\}\times \Sigma_0\subset \mathbb{X}_0, \hbox{ for } i=1,\cdots,p. \end{array}$$

Lemma 30.

Each setEi$\begin{array}{} \displaystyle \mathbb{E}_i \end{array}$, i = 1, …, p, is invariant by the skew-product semiflow (∏0(t): t ⩾ 0} and EiA0$\begin{array}{} \displaystyle \mathbb{E}_i\subset \mathbb{A}_0 \end{array}$, for each i = 1, …, p.

Proof. Clearly, if we take H0 ∈ Σ0, the solution [0, ∞) ∋ ↦ φ(t,H0)ei of (10) is the constant solution φ0(t,H0)ei = ei, for all t ⩾ 0; hence

Π0(t)(ei,H0)=(ei,θtH0)Ei, for all t0.$$\begin{array}{} \displaystyle {\Pi _0}(t)({e_i},{H_0}) = ({e_i},{\theta _t}{H_0}) \in {\mathbb{E}_i},{\rm{\;for\;all\;}}t \geqslant 0. \end{array}$$

Conversely, if t ⩾ 0 and H ∈ Σ0 are given, set H0 = θtH. We have

Π0(t)(ei,H)=(ei,θtH)=(ei,H0),$$\begin{array}{} \displaystyle {\Pi _0}(t)({e_i},H) = ({e_i},{\theta _t}H) = ({e_i},{H_0}), \end{array}$$

therefore (ei,H0)Π0(t)Ei$\begin{array}{} \displaystyle ({e_i},{H_0}) \in {\Pi _0}(t){\mathbb{E}_i} \end{array}$. The last claim follows since Ei$\begin{array}{} \displaystyle {{\Bbb E}_i} \end{array}$ is abounded invariant subset of X0$\begin{array}{} \displaystyle {{\Bbb X}_0} \end{array}$.

We can now define a functional on X0$\begin{array}{} \displaystyle {{\Bbb X}_0} \end{array}$, which will help us understand the intern structure of 𝔸0.

Definition 12.

Define the functional V:X0$\begin{array}{} \displaystyle V:{\mathbb{X}_0} \to \mathbb{R} \end{array}$ by

V(v,H0)=12vH01(Ω)2ΩW(v),$$\begin{array}{} \displaystyle V(v,{H_0}) = \frac{1}{2}v_{H_0^1(\Omega )}^2 - {\smallint _\Omega }W(v), \end{array}$$

where W(r)=0rf(θ)dθ$\begin{array}{} \displaystyle W(r) = \smallint _0^rf(\theta )d\theta \end{array}$, for each (v,H0)X0$\begin{array}{} \displaystyle (v,{H_0}) \in {\mathbb{X}_0} \end{array}$.

Lemma 31.

Let (∏0(t): t ⩾ 0} be the skew-product semiflow defined in (17) for ε = 0. If(u0,H0)X0$\begin{array}{} \displaystyle ({u_0},{H_0}) \in {\mathbb{X}_0} \end{array}$and V is the functional in defined in (23), we have that

(a)the map [0, ∞) ∋ tV(∏0(t)(u0,H0)) is non-increasing, for each(u0,H0)X0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{X}_0 \end{array}$and it is constant inEi$\begin{array}{} \displaystyle {{\Bbb E}_i} \end{array}$, i = 1,… ,p.

(b)If the map [0, ∞) ∋ tV(∏0(t)(u0,H0)) is constant, then(u0,H0)Ei$\begin{array}{} \displaystyle ({u_0},{H_0}) \in {{\Bbb E}_i} \end{array}$, for some i = 1, …, p.

Proof. Since V(∏0(t)(u0, H0)) = V0(t)(u0, H0)), it is a straightforward computation to see, if H0 = Bλfe - Ãλ for some λ∈Г, that

ddtV(Π0(t)(u0,H0))=ddtφ(t,H0)u0L2(Ω)2λ(t)ddtφ(t,H0)u0H01(Ω)20,$$ \frac{d}{dt}V(\Pi_0(t)(u_0,H_0))=-\left\|\frac{d}{dt}\varphi(t,H_0)u_0\right\|^2_{L^2(\Omega)}-\lambda(t)\left\|\frac{d}{dt}\varphi(t,H_0)u_0\right\|^2_{H^1_0(\Omega)}\leqslant 0, $$

since 0 < γ ⩽ λ(t) for all t ∈ ℝ so the map [0, ∞) ∋ t ↦ V(∏0(t)(u0,Ho)) is non-increasing, and it is clearly constant in each 𝔼i.

Now, if map [0, ∞) ∋ t ↦ V(∏0(t)(u0,Ho)) is constant, we have that φ0(t,H0)u0 = u0, for all t ⩾ 0, and hence u0 is a equilibrium of —∆u = f(u) in H01(Ω)$\begin{array}{} \displaystyle H_0^1(\Omega ) \end{array}$(Ω), which implies that u0= ei, for some i = 1, … , p.

Definition 13.

Let (u0,H0)A0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{A}_0 \end{array}$. We define the ω-limit of (u0,H0) by

ω(u0,H0)={(u,H)A0: there exists a sequence tn  such that Π0(tn)(u0,H0)n(u,H)},$$\begin{array}{} \displaystyle \begin{array}{*{20}{r}} {\omega ({u_0},{H_0}) = \{ (u,H) \in {_0}:{\rm{\;there\;exists\;a\;sequence\;}}{t_n} \to \infty }\\ {\begin{array}{*{20}{l}} {{\rm{\;\;}}\qquad {\rm{such\;that\;}}{\Pi _0}({t_n})({u_0},{H_0})\mathop \to \limits^{n \to \infty } (u,H)\} ,} \end{array}} \end{array} \end{array}$$

and if ξ:A0$\begin{array}{} \displaystyle \xi:\mathbb{R}\to \mathbb{A}_0 \end{array}$ is a global solution of {∏0(t): ⩾ 0} through (u0,H0), we define the αξ -limit of (u0,H0) by

αξ(u0,H0)={(u,H)A0: there exists a sequence tn  such that ξ(tn)n(u,H)},$$\begin{array}{} \displaystyle \begin{array}{*{20}{l}} {{\alpha _\xi }({u_0},{H_0}) = \{ (u,H) \in \,{_0}:{\rm{\;there\;exists\;a\;sequence\;}}{t_n} \to \infty }\\ {{\rm{\;\;}}\qquad \qquad \qquad \qquad {\rm{such\;that\;}}\xi ( - {t_n})\mathop \to \limits^{n \to \infty } (u,H)\} ,} \end{array} \end{array}$$

It is a well known result, since {∏0(t): t ⩾ 0} has a global attractor A0$\begin{array}{} \displaystyle \mathbb{A}_0 \end{array}$, that both ω(u0, H0) and αξ (u0, H0) are non-empty, compact, invariant for {∏0(t): t ⩾ 0} and connected. With these, we can prove the following result.

Lemma 32.

For any (u0,H0)A0$\begin{array}{} \displaystyle (u_0,H_0)\in \mathbb{A}_0 \end{array}$and any global solution ξ through (u0, H0), there exists i, j = 1, … , p such that

ω(u0,H0)Eiandαξ(u0,H0)Ej.$$\begin{array}{} \displaystyle \omega(u_0,H_0) \subset \mathbb{E}_i \quad { and } \quad \alpha_\xi(u_0,H_0)\subset \mathbb{E}_j. \end{array}$$

Moreover, if i = j, then u0 = eu

Proof. Let (u,H) ∈ ω(u0,H0) and tn → ∞ such that ∏0(tn)(u0,H0) → (u,H). Since V is a continuous functional in X0$\begin{array}{} \displaystyle {{\Bbb X}_0} \end{array}$, we have that V(∏0(tn)(u0,H0)) → V(u,H), as n — ∞.

Since V(∏0(‧)(u0, H0)) is non-increasing and has a convergent subsequence, we obtain that

V(Π0(t)(u0,H0))V(u,H), as t.$$\begin{array}{} \displaystyle V({\Pi _0}(t)({u_0},{H_0})) \to V(u,H),{\rm{\;as\;}}t \to \infty . \end{array}$$

V(no(t)(uo,Ho)) → V(u,H), as t → o.

Hence, if (u1, H1) is any point in ω(u0, H0), we have that V(u1, H1) = V(u,H). Since ω(u0,H0) is invariant for {∏0(t): t ⩾ 0} we have that V(∏0(t)(u,H)) = V(u,H), and then [0,∞) ∋ t → V(∏0(t)(u,H)) is constant, which implies that (u,H)Ei$\begin{array}{} \displaystyle (u,H)\in \mathbb{E}_i \end{array}$, for some i = 1, … , p. The connectedness of ω (u0, H0) shows us that ω(u0,H0)Ei$\begin{array}{} \displaystyle \omega(u_0,H_0)\subset \mathbb{E}_i \end{array}$.

The proof for αξ (u0, H0) is analogous and the last assertion is straightforward

Proposition 33.

The family E={E1,,Ep}$\begin{array}{} \displaystyle \mathfrak{E}=\{\mathbb{E}_1,\cdots, \mathbb{E}_p\} \end{array}$ is a disjoint family of isolated invariants for {∏0(t): t ⩾ 0}; that is, 𝔼i ⊂ 𝔸0, Π(0)(t)𝔼i = 𝔼i for all t ⩾ 0, there exists δ > 0 such that 𝔼iis the maximal invariant set for {∏0(t): t ⩾ 0} in

Oδ(Ei)={(v,H)X0:veiH01(Ω)<δ},$$\begin{array}{} \displaystyle \mathscr{O}_{\delta}(\mathbb{E}_i)=\{(v,H)\in \mathbb{X}_0\colon \|v-e_i\|_{H^1_0(\Omega)}<\delta\}, \end{array}$$

for each i = 1, … , p, and also 𝔼i ∩ 𝔼j = ∅, if 1 ⩽ i ≠ jp.

Proof. It only remains to prove that there exists δ > 0 such that 𝔼i is the maximal invariant set in Oδ(Ei)$\begin{array}{} \displaystyle {{\scr O}_\delta }({{\Bbb E}_i}) \end{array}$. Let δ=12min1ijpeiejH01(Ω)$\delta = \frac{1}{2}{\min _{1 \leqslant i \ne j \leqslant p}}{e_i} - {e_j}{_{H_0^1(\Omega )}}$.

If 𝔼i is not the maximal invariant in Oδ(Ei)$\begin{array}{} \displaystyle {{\scr O}_\delta }({_i}) \end{array}$, there exists a global solution ξ of {∏0(t): t ⩾ 0} such that ξ(ℝ) ⊂ Oδ(𝔼i), with ξ(ℝ)\𝔼i ∩ 𝔼j = ∅. But then the previous lemma shows that ω(ξ(0)) ⊂ 𝔼i and αξ(ξ(0))⊂𝔼i, which implies that ξ(0) = (ei,H0), and therefore ξ(ℝ) ⊂ 𝔼i, so we reached a contradiction.

All these results combined show us that the semigroup {∏0(t): t ⩾ 0} is, in fact, a generalized gradient semigroup (see [7,15,26] for more details) with disjoint family of isolated invariants E={E1,,Ep}$\begin{array}{} \displaystyle \mathfrak{E}=\{\mathbb{E}_1,\cdots,\mathbb{E}_p\} \end{array}$ and as a consequence, we can write the global attractor 𝔸0 as

A0=i=1pWu(Ei),$$ \mathbb{A}_0 = \bigcup_{i=1}^p \mathbb{W}^u(\mathbb{E}_i), $$

where

Wu(Ei)={(u,H)A0: there exists a global solution ξ of {Π0(t):t0} with ξ(0)=(u,H) such that d(ξ(t),Ei)0, as t}.$$\begin{equation} \begin{split} \mathbb{W}^u(\mathbb{E}_i)= \{(u,H)&\in \mathbb{A}_0\colon \hbox{ there exists a global solution } \xi \hbox{ of }\{\Pi_0(t)\colon t\geqslant 0\} \\ &\hbox{ with } \xi(0)=(u,H) \hbox{ such that }\mathfrak{d}(\xi(t),\mathbb{E}_i)\to 0, \hbox{ as } t\to -\infty\}. \end{split} \end{equation}$$

If H0 ∈ Σ0, there exists λ ∈ Γ such that H0 = Bλfe — Ãλ. Thus we can define

Wu(ei,H0)={u&H01(Ω): there exists a global solution η (10)for  H0with η(0)=u such that η(t)eiH01(Ω)0, as t}.$$\begin{array}{} \displaystyle \begin{array}{*{20}{l}} {{W^u}({e_i},{H_0}) = \{ u \in \,{\rm{\& }}H_0^1({\rm{\Omega }}):{\rm{\;there\;exists\;a\;global\;solution\;}}\eta {\rm{\;(10)for \;}}{H_0}}\\ {\qquad \qquad \qquad \qquad \quad {\rm{with\;}}\eta (0) = u{\rm{\;such\;that\;}}\eta (t) - {e_i}{_{H_0^1(\Omega )}} \to 0,{\rm{\;as\;}}t \to - \infty \} .} \end{array} \end{array}$$

It is simple to see that

Wu(Ei)=Wu(ei,H0)×{H0}, for each i=1,,p,$$\begin{array}{} \displaystyle {\mathbb{W}^u}({\mathbb{E}_i}) = {W^u}({e_i},{H_0}) \times \left\{ {{H_0}} \right\},{\rm{\;for\;each\;}}i = 1, \cdots ,p, \end{array}$$

and thus we have that

A0=i=1pH0Σ0Wu(ei,H0)×{H0},$$ \mathbb{A}_0=\bigcup_{i=1}^p\bigcup_{H_0\in \Sigma_0}W^u(e_i,H_0)\times \{H_0\}, $$

and Theorem 21 shows us that the uniform attractor 𝒜0of the non-autonomous dynamical system (φ0,θ)(H01(Ω),Σ0)$\begin{array}{} \displaystyle {({\varphi _0},\theta )_{(H_0^1(\Omega ),{\Sigma _0})}} \end{array}$ is given by

A0=i=1pH0Σ0Wu(ei,H0).$$ \mathscr{A}_0 = \bigcup_{i=1}^p \bigcup_{H_0\in \Sigma_0} W^u(e_i,H_0). $$

Moreover, the cocycle attractor {A(H0)}H0∈Σ0 of the non-autonomous dynamical system (φ0,θ)(H01(Ω),Σ0)$\begin{array}{} \displaystyle {({\varphi _0},\theta )_{(H_0^1(\Omega ),{\Sigma _0})}} \end{array}$ is given by

A(H0)=i=1pWu(ei,H0), for each H0Σ0.$$ A(H_0) = \bigcup_{i=1}^p W^u(e_i,H_0), \hbox{ for each } H_0\in \Sigma_0. $$

And finally, for each H0 ∈ Σ0, the pullback attractor {AH0(t)}t∈ℝ of the evolution process TH0(t,s) = φ0(t — s, θsH0) is given by

AH0(t)=i=1pWu(ei,θtH0), for all tR and each H0Σ0.$$ A_{H_0}(t) = \bigcup_{i=1}^p W^u(e_i,\theta_tH_0), \hbox{ for all } t\in \mathbb{R} \hbox{ and each } H_0\in \Sigma_0. $$

Remark 11.

If H0 = Bγfe — Ãγ, defining

A(t)=AH0(t), for each t,$$ \mathscr{A}(t)=A_{H_0}(t), \hbox{ for each }t\in \mathbb{R}, $$

we obtain the pullback attractor for (5); that is, for each t ∈ ℝ, we have

A(t)={η(t):η is a bounded global solution of (5)}.$$\begin{array}{} \displaystyle \mathscr{A}(t) = \{ \eta (t):\eta {\rm{\;is\;a\;bounded\;global\;solution of (}}5)\} . \end{array}$$

If γ(·) is a constant function, equation (5) is autonomous, Σ0 = Bγfe — Ãγ} is a singleton and we obtain, as a particular case, that the autonomous system generated by (5) is a gradient semigroup with a finite collection of equilibria ℰ= {e0, …, ep} and that 𝔸0 = 𝒜 × Σ0, where 𝒜0 = Wu(ei, H0) and H0 = Bγfe —Ãγ.

Structure of 𝔸e

In this subsection we prove that, under suitable conditions, the attractors 𝔸ε inherit the same generalized gradient structure from 𝔸0. To this end, we first need to take the following fact in account: we have, following the ideas of the proof of Lemma 26, if Hε → H0 in (𝒞*,d*), that

limϵ0+d[Πϵ(t)(u,Hϵ),Π0(t)(u,H0)]=0,$$\begin{array}{} \displaystyle \lim_{\epsilon \to 0^+}\mathfrak{d}[\Pi_\epsilon(t)(u,H_\epsilon),\Pi_0(t)(u,H_0)]=0, \end{array}$$

uniformly for t in bounded subsets of ℝ, u in bounded subsets of H01(Ω)$\begin{array}{} \displaystyle H^1_0(\Omega) \end{array}$.

With all these considerations and previous results, we are able to state the following structural result.

Theorem 34

Assume that hypotheses (H1)-(H4), (C) and (F) hold true. Assume also that

(a) for each ε ∈ (0,1] there exists a disjoint family of isolated invariants Eϵ={E1,ϵ,,Ep,ϵ}$\begin{array}{} \displaystyle \mathfrak{E}_\epsilon=\{\mathbb{E}_{1,\epsilon},\cdots,\mathbb{E}_{p,\epsilon}\} \end{array}$ for {∏ε(t): t ⩾ 0} such that

dH[Ei,ϵ,Ei]+dH[Ei,Ei,ϵ]0, as ϵ0+;$$\begin{array}{} \displaystyle \mathfrak{d}_H[\mathbb{E}_{i,\epsilon},\mathbb{E}_{i}]+\mathfrak{d}_H[\mathbb{E}_{i},\mathbb{E}_{i,\epsilon}]\to 0, \hbox{ as } \epsilon \to 0^+; \end{array}$$

(b) there exists ε0 > 0 and neighborhoods 𝕌i of 𝔼i such that 𝔼i,εis the maximal invariant set of {∏ε(t): t ⩾ 0} in 𝕌i, for each i = 1, …, p and 0 < ε ⩽ ε0.

Then there exists ε0 > 0 such that {∏ε(t): t ⩾ 0} is a generalized gradient semigroup with a disjoint family of isolated invariantsℱε, for each 0 < εε1. Moreover for 0 ⩽ ε ⩽ ε1, we have

Aϵ=i=1pWu(Ei,ϵ).$$\mathbb{A}_\epsilon=\bigcup_{i=1}^p\mathbb{W}^u(\mathbb{E}_{i,\epsilon}).$$

Proof. The proof of this theorem is analogous to the proof of [15, Theorem 1.5], with the aid of Proposition 25.

Global hyperbolic solutions for (10)

Let ξε be a global solution of {∏ε(t): t ⩾ 0} in 𝔸ε. Thus we have that there exists Hε ∈ ∑ε such that ξε(t) = (ηε(t), θtHε), for all t ∈ ℝ, where ηε is a global bounded solution of (10).

Writing Hϵ=BλhϵeA˜λ$\begin{array}{} \displaystyle H_\epsilon=B_\lambda h_\epsilon^e-\tilde{A}_\lambda \end{array}$, for some λ ∈ Γ and hε ∈ 𝒞2, we can consider the variational problem for the solution ηε, given by

utλ(t)ΔutΔu=Dshϵe(t,ηϵ(t))u, in Ωu=0, on Ω,$$\begin{equation} \left\{\begin{split} &u_t-\lambda(t)\Delta u_t-\Delta u=D_sh^e_\epsilon(t,\eta_\epsilon(t))u, \hbox{ in } \Omega\\ &u=0, \hbox{ on } \partial \Omega, \end{split}\right. \end{equation}$$

where [Dshϵe(t,ηϵ(t))u](x)=shϵ(t,ηϵ(t)(x))u(x)$\begin{array}{} \displaystyle [D_sh_\epsilon^e(t,\eta_\epsilon(t))u](x)=\partial_sh_\epsilon(t,\eta_\epsilon(t)(x))u(x) \end{array}$,a.e x ∈ Ω, for uH01(Ω)$\begin{array}{} \displaystyle u\in H^1_0(\Omega) \end{array}$.

This equation generates a linear evolution process {Lξϵ(t,s):ts}L(H01(Ω))$\begin{array}{} \displaystyle \{L_{\xi_\epsilon}(t,s)\colon t \geqslant s\}\subset \mathscr{L}(H^1_0(\Omega)) \end{array}$, where u(t) = Lξε(t, s)u0 is the solution at time t of (24) with u(s) = u0.

Definition 14.

We say that Lξε(t, s): t ⩾ s} has an exponential dichotomy if there exists a family of projections {Q(t)}t∉ℝ in L(H01(Ω))$\begin{array}{} \displaystyle \mathscr{L}(H^1_0(\Omega)) \end{array}$satisfying:

(a) Q(t)Lξε(t, s) = (t,s)Q(s),forall t ⩾ s.

(b) The restriction Lξε(t,s)|R(Q(s)), t ⩾ s, is an isomorphism from R(Q(s)) into R(Q(t)); and we denote its inverse by Lξε(s, t): R(Q(t)) → R(Q(s)).

(c) There are constants ω > 0 and M ⩾ 1 such that

Lξϵ(t,s)(IQ(s))L(H01(Ω))Meω(ts), for ts;Lξϵ(s,t)Q(t)L(H01(Ω))Meω(st), for s<t.$$\begin{array}{} \displaystyle & \|L_{\xi_\epsilon}(t,s)(I-Q(s))\|_{\mathscr{L}(H^1_0(\Omega))}\leqslant Me^{-\omega(t-s)}, \hbox{ for } t \geqslant s;\\ & \|L_{\xi_\epsilon}(s,t)Q(t)\|_{\mathscr{L}(H^1_0(\Omega))}\leqslant Me^{\omega(s-t)}, \hbox{ for } s<t. \end{array}$$

In this case, we say that ξε is a hyperbolic global solution of {Πε (t): t ⩾ 0}.

Now we make the following assumption:

 Each global solution ξi,H0(t)=(ei,θtH0),tR and H0Σ0, is hyperbolic.\begin{equation*} \begin{split} \hbox{ Each global } &\hbox{solution } \xi_{i,H_0}^\ast(t)=(e_i,\theta_tH_0), \ t\in \mathbb{R} \hbox{ and } \\&H_0\in \Sigma_0, \hbox{ is hyperbolic}. \end{split} \end{equation*}

Remark 12.

Since Σ0={BλfeA˜λ}λΓ$\begin{array}{} \displaystyle \Sigma_0=\{B_\lambda f^e-\tilde{A}_\lambda\}_{\lambda\in \Gamma} \end{array}$, we have a family {ξ*i,λ}λ∊T of hyperbolic global solutions of {Π0(t): t ⩾ 0}, where ξi,λ=ξi,H0$\begin{array}{} \displaystyle \{\xi_{i,\lambda}^\ast\}_{\lambda\in \Gamma} \end{array}$ for H0=BλfeA˜λ$\begin{array}{} \displaystyle H_0=B_\lambda f^e-\tilde{A}_\lambda \end{array}$.

Now, proceeding as in [16, Section 7.2] and [26], we have the following result:

Proposition 35.

Assume that hypotheses(H1)-(H5), (C), (F)and(Hy)hold true. Thus, there exist ε0 > 0 andδ > 0 such that for each i = 1, ··· , p, λ ∊ Γ and ε ∊ (0, ε0], there exist a uniquehϵGϵ$\begin{array}{} \displaystyle h_\epsilon^\ast\in \mathscr{G}_\epsilon \end{array}$and a unique bounded global solutionηi,λ,ϵ:Aϵ$\begin{array}{} \displaystyle \eta_{i,\lambda,\epsilon}^\ast \colon \mathbb{R}\to \mathscr{A}_\epsilon \end{array}$of (10), withHϵ=Bλhϵ,eA˜λ$\begin{array}{} \displaystyle H_\epsilon=B_\lambda h_\epsilon^{\ast,e}-\tilde{A}_\lambda \end{array}$, which we denote byHϵλ$\begin{array}{} \displaystyle H_\epsilon^\lambda \end{array}$, such that:

(i)the functiontξi,λ,ϵ(t)=(ηi,λ,ϵ(t),θtHϵλ)Aϵ$\begin{array}{} \displaystyle \mathbb{R}\ni t\mapsto \xi_{i,\lambda,\epsilon}^\ast(t)=(\eta_{i,\lambda,\epsilon}^\ast(t),\theta_tH_\epsilon^\lambda)\in \mathbb{A}_\epsilon \end{array}$is a hyperbolic global solution of {Πε (t): t ⩾ 0};

(ii)suptd[ξi,λ,ϵ(t),(ei,θtH0λ)]<δ$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}\mathfrak{d}[\xi_{i,\lambda,\epsilon}^\ast(t),(e_i,\theta_tH_0^\lambda)]<\delta \end{array}$; whereH0λ=BλfeA˜λ$\begin{array}{} \displaystyle H_0^\lambda=B_\lambda f^e-\tilde{A}_\lambda \end{array}$, and

(iii)suptd[ξi,λ,ϵ(t),(ei,θtH0λ)]0$\begin{array}{} \displaystyle \sup_{t\in \mathbb{R}}\mathfrak{d}[\xi_{i,\lambda,\epsilon}^\ast(t),(e_i,\theta_tH_0^\lambda)]\to 0 \end{array}$, asε → 0+.

Moreover, ifξε is a global solution of {Πε (t): t ⩾ 0} such thatd[ξϵ(t),ξi,λ,ϵ(t)]<δ$\begin{array}{} \displaystyle \mathfrak{d}[\xi_\epsilon(t),\xi_{i,\lambda,\epsilon}^\ast(t)]<\delta \end{array}$for all t ⩾ 0 (t ⩽ 0) then

d[ξϵ(t),ξi,λ,ϵ(t)]0 as t(t)$$\begin{array}{} \displaystyle \mathfrak{d}[\xi_\epsilon(t),\xi_{i,\lambda,\epsilon}^\ast(t)]\to 0 \hbox{ as } t\to \infty \ (t\to -\infty) \end{array}$$

With this proposition, we can construct a family Eϵ={E1,ϵ,,Ep,ϵ}$\begin{array}{} \displaystyle \mathfrak{E}_\epsilon=\{\mathbb{E}_{1,\epsilon},\cdots,\mathbb{E}_{p,\epsilon}\} \end{array}$, satifying the conditions of Theorem 34. In fact, for each λ ∊ Γ, let

Di,ϵ(λ)=tξi,λ,ϵ(t)=t(ηi,λ,ϵ(t),θtHϵλ), for each i=1,,p,$$\begin{array}{} \displaystyle \mathbb{D}_{i,\epsilon}(\lambda)= \bigcup_{t\in \mathbb{R}}\xi_{i,\lambda,\epsilon}^\ast(t)= \bigcup_{t\in \mathbb{R}}(\eta_{i,\lambda,\epsilon}^\ast(t), \theta_tH_\epsilon^\lambda), \hbox{ for each } i=1,\cdots,p, \end{array}$$

and define

Ei,ϵ=λΓDi,ϵ(λ), for each i=1,,p.$$\begin{array}{} \displaystyle \mathbb{E}_{i,\epsilon}=\bigcup_{\lambda\in \Gamma}\mathbb{D}_{i,\epsilon}(\lambda), \hbox{ for each } i=1,\cdots,p. \end{array}$$

Now, as an immediate consequence of Theorem 34 and Proposition 35, we have:

Theorem 36.

Assume that hypotheses(H1)-(H5), (C), (F)and(Hy)hold true. Consider the disjoint family of isolated invariantsEϵ={E1,ϵ,,Ep,ϵ}$\begin{array}{} \displaystyle \mathfrak{E}_\epsilon=\{\mathbb{E}_{1,\epsilon},\cdots,\mathbb{E}_{p,\epsilon}\} \end{array}$forε(t): t ⩾ 0} given by (25).

Then there exists ε1 > 0 such thatε (t): t ⩾ 0} is a generalized gradient semigroup with a disjoint family of isolated invariantsε,for each 0 ⩽ ε ⩽ ε1. Moreover, for 0 ⩽ εε1, we have

Aϵ=i=1pWu(Ei,ϵ),$$\begin{array}{} \displaystyle \mathbb{A}_\epsilon=\bigcup_{i=1}^p\mathbb{W}^u(\mathbb{E}_{i,\epsilon}), \end{array}$$

and

Wu(Ei,ϵ)=λΓWu(Di,ϵ(λ)).$$\begin{array}{} \displaystyle \mathbb{W}^u(\mathbb{E}_{i,\epsilon})=\bigcup_{\lambda\in \Gamma}\mathbb{W}^u(\mathbb{D}_{i,\epsilon}(\lambda)). \end{array}$$

With this result, we can derive structures for the other types of attractors, as we did for 𝔸0; namely, we have that the uniform attractor of (φϵ,θ)(H01(Ω),Σϵ)$\begin{array}{} \displaystyle (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma_\epsilon)} \end{array}$ is given by

Aϵ=πH01(Ω)Aϵ=i=1pWu(Ei,ϵ),$$\begin{array}{} \displaystyle \mathscr{A}_\epsilon = \pi_{H^1_0(\Omega)}\mathbb{A}_\epsilon = \bigcup_{i=1}^p W^u(\mathscr{E}_{i,\epsilon}), \end{array}$$

where Ei,ϵ=λΓtηi,λ,ϵ(t)$\begin{array}{} \displaystyle \mathscr{E}_{i,\epsilon}=\bigcup_{\lambda\in \Gamma}\bigcup_{t\in \mathbb{R}}\eta_{i,\lambda,\epsilon}^\ast(t) \end{array}$. Also, the cocycle attractor for (φϵ,θ)(H01(Ω),Σϵ)$\begin{array}{} \displaystyle (\varphi_\epsilon,\theta)_{(H^1_0(\Omega),\Sigma_\epsilon)} \end{array}$ is given by

A(Hϵ)=i=1pWu(Ei,ϵ(λ)),$$\begin{array}{} \displaystyle A(H_\epsilon)= \bigcup_{i=1}^p W^u(\mathscr{E}_{i,\epsilon}(\lambda)), \end{array}$$

where Ei,ϵ(λ)=tηi,λ,ϵ(t)$\begin{array}{} \displaystyle \mathscr{E}_{i,\epsilon}(\lambda)=\bigcup_{t\in \mathbb{R}}\eta_{i,\lambda,\epsilon}^\ast(t) \end{array}$ and Hϵ=BλhϵeA˜λ$\begin{array}{} \displaystyle H_\epsilon=B_\lambda h_{\epsilon}^e-\tilde{A}_\lambda \end{array}$.

Lastly, if Hε ∊ Σε, the pullback attractor {AHε (t)}t∊ℝ of the evolution process THε (t, s) = φε (t - s, θsHε) is given by

AHϵ(t)=i=1pWu(ηi,λ,ϵ)(t), for all t,$$\begin{array}{} \displaystyle A_{H_\epsilon}(t) = \bigcup_{i=1}^p W^u(\eta_{i,\lambda,\epsilon}^\ast)(t), \hbox{ for all } t\in \mathbb{R}, \end{array}$$

where

Wu(ξi,λ,ϵ)(t)={uH01(Ω): there exists a global solution ηϵ of (10) for Hϵwith ηϵ(t)=u such that ηϵ(s)ηi,λ,ϵ(s)H01(Ω)0, as s}.\begin{equation*} \begin{split} W^u(\xi_{i,\lambda,\epsilon}^\ast)(t) &=\{u\in \ H^1_0(\Omega) \colon \hbox{ there exists a global solution } \eta_\epsilon \hbox{ of }(10) \hbox{ for } H_\epsilon \\& \hbox{with }\eta_\epsilon(t)=u \hbox{ such that }\|\eta_\epsilon(s)-\eta_{i,\lambda,\epsilon}^\ast(s)\|_{H^1_0(\Omega)}\to 0, \hbox{ as } s\to -\infty\}. \end{split} \end{equation*}

Further remarks

As discussed extensively in [5] and [7], we could do the analysis presented in this work, removing condition (C) by assuming only that we have a semigroup of translations {θt: t ⩾ 0} in 𝒢ε with a global attractor Ξε𝒢ε. Thus, despite some complications with notations, the results would remain unchanged, replacing 𝒢ε by Ξε. This, in turn, allows us to give a more precise description of the attractors when gε is, for instance, asymptotically autonomous; since in this case, we notice that Ξε is a singleton, which is significantly smaller than 𝒢ε.

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