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

Finite-dimensionality of attractors for wave equations with degenerate nonlocal damping

Zhijun Tang Senlin Yan Yao Xu  and Chengkui Zhong Corresponding author.
Abstract

In this paper we study the fractal dimension of global attractors for a class of wave equations with (single-point) degenerate nonlocal damping. Both the equation and its linearization degenerate into linear wave equations at the degenerate point and the usual approaches to bound the dimension of the entirety of attractors do not work directly. Instead, we develop a new process concerning the dimension near the degenerate point individually and show the finite dimensionality of the attractor.

Keywords: wave equations, degenerate nonlocal damping, global attractors, fractal dimension.

1 Introduction

In this paper, we consider the dimension problem of the global attractor for the initial-boundary value problem

{uttΔu+(up+utp)ut+f(u)=0in Ω×+,u(x,0)=u0(x),ut(x,0)=u1(x)in Ω,u(x,t)=0on Ω,\begin{cases}u_{tt}-\Delta u+(\|\nabla u\|^{p}+\|u_{t}\|^{p})u_{t}+f(u)=0&\text{in }\Omega\times\mathbb{R}^{+},\\ u(x,0)=u_{0}(x),u_{t}(x,0)=u_{1}(x)&\text{in }\Omega,\\ u(x,t)=0&\text{on }\partial\Omega,\end{cases} (1.1)

where p(1,2)p\in(1,2), Ω\Omega is a bounded domain in 3\mathbb{R}^{3} with smooth boundary Ω\partial\Omega and ff satisfies the elementary Assumption 1.1 below.

Assumption 1.1.

fC2()f\in C^{2}(\mathbb{R}) and satisfies the critical growth condition

|f′′(s)|C(1+|s|),|f^{\prime\prime}(s)|\leq C(1+|s|), (1.2)

and the dissipative condition

lim inf|s|f(s)μ>λ1,\liminf_{|s|\rightarrow\infty}f^{\prime}(s)\equiv\mu>-\lambda_{1}, (1.3)

where λ1\lambda_{1} is the first eigenvalue of AA, the negative Laplacian operator with homogeneous Dirichlet boundary condition.

There are massive works on asymptotic behaviours of solutions of nonlinear wave equations since 1970s. People are primarily concerned about the topics of well-posedness of the equations, and the existence, attracting speed as well as fractal dimension of the corresponding attractors. Models with various damping terms have been investigated, including weak damping kutku_{t}, strong damping kΔut-k\Delta u_{t}, fractional damping (Δ)αut(-\Delta)^{\alpha}u_{t} with α(0,1)\alpha\in(0,1) and nonlinear damping g(ut)g(u_{t}). For more information we refer readers to literature [2, 5, 16, 15, 19, 24, 25, 21, 18, 14] and references therein.

The dynamics of equations with nonlocal damping also attracted widespread attention. Authors in [17] have studied well-posedness of a class of extensible beam models with nonlocal energy damping (Δu2+ut2)qΔut\left(\|\Delta u\|^{2}+\|u_{t}\|^{2}\right)^{q}\Delta u_{t}, which was first proposed by Balakrishnan-Taylor[1]. Y. Sun and Z. Yang in [22] have investigated the existence of strong global and exponential attractors for the equation

utt+Δ2uκϕ(u2)ΔuM(Δu2+ut2)Δut+f(u)=h,u_{tt}+\Delta^{2}u-\kappa\phi(\|\nabla u\|^{2})\Delta u-M(\|\Delta u\|^{2}+\|u_{t}\|^{2})\Delta u_{t}+f(u)=h,

in which they assume that MM is positive on +\mathbb{R}^{+}. We also refer to [4, 7, 6] for more information about wave equations with damping term of the form M(uL2(Ω)2)(Δ)θutM(\|\nabla u\|^{2}_{L^{2}(\Omega)})(-\Delta)^{\theta}u_{t} and M(uL2(Ω)2)g(ut)M(\|\nabla u\|^{2}_{L^{2}(\Omega)})g(u_{t}), where the intensity of the damping involves only the potential part of the energy.

In this paper we are concerned about the dissipative wave equation (1.1) with nonlocal damping (up+utp)ut(\|\nabla u\|^{p}+\|u_{t}\|^{p})u_{t}, which is an extended Krasovskii model first studied by Chueshov[8]. It has weaker damping and the energy decays more slowly near the origin than the equations with usual weak damping. Particularly, degeneration happens at the origin, which from the geometrical perspective complicates the asymptotic behaviours intrinsically. It is therefore interesting to ask if the fractal dimension of the global attractor is still finite, since as will be mentioned below the classical methods to bound fractal dimensions fail in our setting and in most cases degeneration causes infinite-dimensionality. Our main result is as follows.

Theorem 1.2.

Suppose Assumption 1.1 holds and in addition

f(0)=0,f(0)>(1p4)λ1,|f(s)f(0)|=o(|s|p),f(0)=0,\quad f^{\prime}(0)>-\big{(}1-\frac{p}{4}\big{)}\lambda_{1},\quad|f^{\prime}(s)-f^{\prime}(0)|=o(|s|^{p}), (1.4)

Then the global attractor of Problem (1.1) in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega) has finite fractal dimension.

We should mention that condition (1.4) is only involved in the proof of Lemma 5.2 which declares the uniform Lipschitz stability of the semigroup near the origin. However, to prove Theorem 1.2 only the uniform Hölder property on the attractor near the origin is required. Therefore, condition (1.4) shall not be sharp. Especially, by calculating (5.10) more accurately, one can improve p4\frac{p}{4} in the condition f(0)>(1p4)λ1f^{\prime}(0)>-\big{(}1-\frac{p}{4}\big{)}\lambda_{1} with a slightly smaller number CpC_{p}.

Dimension problem is an essential subject in the area of infinite-dimensional dynamical systems, which may reflect the complexity of the system. If the fractal dimension of the global attractor is finite, in some sense one can reduce the dynamics on the infinite-dimensional phase space to a simpler one on a finite-dimensional phase space. There have been already some classical approaches to bound the (finite) fractal dimension specifically for negatively invariant sets, including global attractors. For instance, one can use the method of Lyapunov exponent in [23, 20] to study finite-dimensional behaviors if the quasi-derivative of S(T)S(T) is a compact perturbation of some uniform contractive operator for some fixed T>0T>0. Another alternative is the quasi-stability inequality method if the semigroup S(t)S(t), not necessarily differentiable, satisfies

S(T)v1S(T)v2ηv1v2+nZ(Kv1Kv2),\|S(T)v_{1}-S(T)v_{2}\|\leq\eta\|v_{1}-v_{2}\|+n_{Z}(Kv_{1}-Kv_{2}), (1.5)

where T>0,0<η<1T>0,0<\eta<1, KK is a Lipschitz mapping from the attractor 𝒜\mathscr{A} to a Banach space ZZ and nZn_{Z} is a compact seminorm on ZZ. All these methods require some kind of uniform contraction property on the entirety of the attractors. We refer readers to [8] for more details and to [28] for exponential attractors of semigroups on Banach spaces.

However, the existing methods all lose efficacy for our problem due to the degeneration occurring at the origin. In fact, the semigroup behaves just like the wave equation

uttΔu=0u_{tt}-\Delta u=0

near the origin if f(0)=0f(0)=0, in which situation the energy does not decay along the direction of each eigenfunction of AA. Therefore, we can neither write the derivative of S(T)S(T) into a compact perturbation of some uniform contractive operator, nor find a constant η<1\eta<1 to fulfill (1.5) uniformly (see Remark (5.3) for details).

It is worth pointing out that there are just few works dealing with the dimension problem of degenerate models, and almost all of them possess infinite-dimensional attractors. In [13] M. Efendiev and S. Zelik have proved the finite-dimensionality of global attractors for a class of porous medium equations when the nonlinearity g(u)g(u) satisfies g(0)>0g^{\prime}(0)>0. Moreover, they also have obtained the infinite-dimensionality in the case g(0)<0g^{\prime}(0)<0 by estimating Kolmogorov ϵ\epsilon-entropy. See also [11, 12] for the corresponding infinite-dimensionality results on parabolic equations with pp-Laplacian. Besides, using another method of the Z2Z_{2} index, C. Zhong and W. Niu have also shown in [27] that the fractal dimension of the global attractor is infinite for a class of pp-Laplacian equations.

We now explain the main difficulties and new ideas in the proof while describing the outline of this paper. In Section 2, we establish the well-posedness of strong solutions and generalized solutions to problem (1.1) by the monotone operator method under Assumption 1.1. In the same section, the existence of the global attractor 𝒜\mathscr{A} in H01(Ω)×L2(Ω)H^{1}_{0}(\Omega)\times L^{2}(\Omega) is proved through the dissipation and asymptotic smoothness. To verify the asymptotic smoothness, we make a decomposition u=v+wu=v+w, such that, vv decays uniformly to zero and ww possesses higher regularity for each t>0t>0 for bounded initial data. This special decomposition is also valid for establishing higher regularity of the global attractor under additional assumption 3.1 on ff in the next section, which infers directly, as a byproduct, the temporal Hölder continuity of S(t)S(t) uniformly on 𝒜\mathscr{A}. More detailedly, the regularity of 𝒜\mathscr{A} can be deduced from the higher regularity of the attracting set of ww and the latter conclusion can be achieved by a process of weighted energy estimate, inspired by the structure of the damping, together with an extended discrete-like Gronwall lemma. This is a novelty of the paper. Another essential estimate in Section 3 is Lemma 3.6 concerning the accurate decay rate of solutions near the origin on the phase space. Compared to the usual equations with nondegenerate weak damping whose solutions decay exponentially fast near the origin, in our problem the rates of both the norm and the energy are as slow as t1pt^{-\frac{1}{p}}. These inherent estimates obtained in Section 3 are useful to perform the dimension estimation in the last section.

Section 4 is devoted to an abstract framework of the dimension calculation for single-point degenerate equations. In consideration of problem (1.1), we employ the setting that the degenerate point 0 is a locally attracting point with a quantitative speed on the attractor and the semigroup is uniformly (in time and space) Hölder continuous near the origin. Intuitively, the attracting behaviour near 0 plays a role of contraction effect as the trajectories collapse to 0. Therefore, it is reasonable to believe the geometry of the attractor near 0 would not be worse than that of the region away from 0. Indeed, providing that the dimension of the complement on the attractor of neighbourhoods of 0 is finite, we are able to declare the finite dimensionality of neighbourhoods of 0, as stated in Theorems 4.3 and 4.4. In this process we decompose the entire attractor into the degenerate part (the neighbourhood of 0) and the nondegenerate part (the complement of the neighbourhood). The former set is positively invariant and the latter one is negatively invariant. Hence, it is possible to figure out the dimension of the latter one through the pioneering classical methods while the dimension problem of the former part has been reduced to that of the latter part. This is the central idea in Section 4.

In Section 5, we apply the previous theory to problem (1.1). To do this, we first show the uniform Lipschitz stability of S(t)S(t) near the origin in Lemma 5.2 by estimating the operator norm of the derivative DS(t)DS(t). To calculate the dimension of the nondegenerate part, we split DS(t)DS(t) into a uniform contractive linear operator and a compact operator, which gives rise to the finite-dimensionality of this negatively invariant set. As a result, Theorem 4.4 leads to Theorem 1.2 immediately.

Throughout the paper, we use the notations p=Lp(Ω)\|\cdot\|_{p}=\|\cdot\|_{L^{p}(\Omega)} (particularly, =2\|\cdot\|=\|\cdot\|_{2} for brevity) and, for a function uu

Iu(t)=u2+ut2,Iu,p(t)=up+utp.I_{u}(t)=\|\nabla u\|^{2}+\|u_{t}\|^{2},\quad I_{u,p}(t)=\|\nabla u\|^{p}+\|u_{t}\|^{p}. (1.6)

Without causing misunderstanding, we also write both the inner product in L2(Ω)L^{2}(\Omega) and the ordered pair of u,vu,v as (u,v)(u,v). Besides, we may use CC to denote any positive constant which may differ from each other, even in the same line.

2 well-posedness and global attractor

In this section we investigate the global well-posedness of Problem (1.1) as well as the existence of global attractors under Assumption 1.1.

Definition 2.1.

Suppose uC([0,T];H01(Ω))C1([0,T];L2(Ω))u\in C([0,T];H^{1}_{0}(\Omega))\cap C^{1}([0,T];L^{2}(\Omega)) with u(0)=u0u(0)=u_{0} and ut(0)=u1u_{t}(0)=u_{1}. Then we say

uu is a strong solution if

  • a)

    uW1,1(a,b;H01(Ω))u\in W^{1,1}(a,b;H_{0}^{1}(\Omega)) and utW1,1(a,b;L2(Ω)), 0<a<b<Tu_{t}\in W^{1,1}(a,b;L^{2}(\Omega)),\forall\,0<a<b<T;

  • b)

    Δu(t)L2(Ω),t[0,T]-\Delta u(t)\in L^{2}(\Omega),\forall\,t\in[0,T];

  • c)

    (1.1) holds in L2(Ω)L^{2}(\Omega) for almost every t[0,T]t\in[0,T].

uu is a generalized solution if there exists a sequence of strong solutions {u(j)}\{u^{(j)}\} with initial data (u0(j),u1(j))(u^{(j)}_{0},u^{(j)}_{1}) such that

(u(j),ut(j))(u,ut) in C([0,T];H01(Ω)×L2(Ω)).(u^{(j)},u_{t}^{(j)})\rightarrow(u,u_{t})\text{ in }C([0,T];H_{0}^{1}(\Omega)\times L^{2}(\Omega)).

uu is a weak solution if

Ωut(x)ψ(x)dx+0t[Ωu(τ,x)ψ(x)dx+Ωf(u(τ,x))ψ(x)dx+(u(τ)p+ut(τ)p)Ωut(τ,x)ψ(x)dx]dτ=Ωu1(x)ψ(x)dx\displaystyle\begin{split}&\int_{\Omega}u_{t}(x)\psi(x)dx+\int_{0}^{t}\Big{[}\int_{\Omega}\nabla u(\tau,x)\nabla\psi(x)dx+\int_{\Omega}f(u(\tau,x))\psi(x)dx\\ &\qquad+(\|\nabla u(\tau)\|^{p}+\|u_{t}(\tau)\|^{p})\int_{\Omega}u_{t}(\tau,x)\psi(x)dx\Big{]}d\tau=\int_{\Omega}u_{1}(x)\psi(x)dx\end{split} (2.1)

holds for all ψH01(Ω)\psi\in H_{0}^{1}(\Omega) and almost every t[0,T]t\in[0,T].

Theorem 2.2.

Under Assumption 1.1, we have the following conclusions:

If (u0,u1)(H2(Ω)H01(Ω))×H01(Ω)(u_{0},u_{1})\in(H^{2}(\Omega)\cap H_{0}^{1}(\Omega))\times H_{0}^{1}(\Omega) , then Problem (1.1) admits a unique strong solution uu, which satisfies that

(ut,utt)L(0,T;H01(Ω)×L2(Ω)),\displaystyle(u_{t},u_{tt})\in L^{\infty}(0,T;H_{0}^{1}(\Omega)\times L^{2}(\Omega)),
utCr([0,T),H01(Ω)),uttCr([0,T),L2(Ω)),\displaystyle u_{t}\in C_{r}([0,T),H_{0}^{1}(\Omega)),\quad u_{tt}\in C_{r}([0,T),L^{2}(\Omega)),
Δu+(u2+ut2)utCr([0,T),L2(Ω)),\displaystyle-\Delta u+(\|\nabla u\|^{2}+\|u_{t}\|^{2})u_{t}\in C_{r}([0,T),L^{2}(\Omega)),

where CrC_{r} is the space of right-continuous functions.

If (u0,u1)H01(Ω)×L2(Ω)(u_{0},u_{1})\in H_{0}^{1}(\Omega)\times L^{2}(\Omega), Problem (1.1) admits a unique generalized solution, which is also a weak solution. Furthermore, uu satisfies the energy equality

E(t)+0t(u(τ)p+ut(τ)p)ut(τ)2𝑑τ=E(0),E(t)+\int_{0}^{t}(\|\nabla u(\tau)\|^{p}+\|u_{t}(\tau)\|^{p})\|u_{t}(\tau)\|^{2}d\tau=E(0), (2.2)

where E(t)12(u(t)2+ut(t)2)+ΩF(u(t))𝑑xE(t)\triangleq\frac{1}{2}(\|\nabla u(t)\|^{2}+\|u_{t}(t)\|^{2})+\int_{\Omega}F(u(t))dx with F(s)0sf(τ)𝑑τF(s)\triangleq\int_{0}^{s}f(\tau)d\tau.

To prove Theorem 2.2, we use the monotone operator method as in [9].

Lemma 2.3.

Let A:D(A)HHA:D(A)\subset H\to H is a maximal monotone operator on a Hilbert space HH, i.e., (Ax1Ax2,x1x2)H0(Ax_{1}-Ax_{2},x_{1}-x_{2})_{H}\geq 0 for any x1,x2D(A)x_{1},x_{2}\in D(A) and Rg(I+A)=HRg(I+A)=H; besides, assume that 0A00\in A0. Let B:HHB:H\to H be locally Lipschitz. If u0D(A),fW1,1(0,t;H)u_{0}\in D(A),f\in W^{1,1}(0,t;H) for all t>0t>0, then there exists tmaxt_{max}\leq\infty such that the initial value problem

ut+Au+Buf and u=u0u_{t}+Au+Bu\ni f\text{ and }u=u_{0} (2.3)

has a unique strong solution uu on the interval [0,tmax)[0,t_{max}).

Whereas, if u0D(A)¯,fL1(0,t;H)u_{0}\in\overline{D(A)},f\in L^{1}(0,t;H) for all t>0t>0, then problem (2.3) has a unique generalized solution uC([0,tmax);H)u\in C([0,t_{max});H).

Moreover, in both case we have limttmaxu(t)H=\lim_{t\to t_{max}}\|u(t)\|_{H}=\infty provided tmax<t_{max}<\infty.

Proof of Theorem 2.2.

Step 1. We establish the existence and uniqueness of strong solutions and generalized solutions.

Let U=(u,v)TU=(u,v)^{T}, where v=utv=u_{t}. Define 𝒜:D(𝒜)H01(Ω)×L2(Ω)H01(Ω)×L2(Ω)\mathcal{A}:D(\mathcal{A})\subset H_{0}^{1}(\Omega)\times L^{2}(\Omega)\rightarrow H_{0}^{1}(\Omega)\times L^{2}(\Omega) and :H01(Ω)×L2(Ω)H01(Ω)×L2(Ω)\mathcal{B}:H_{0}^{1}(\Omega)\times L^{2}(\Omega)\rightarrow H_{0}^{1}(\Omega)\times L^{2}(\Omega) by

𝒜U=(vΔu),U=(0f(u)+(vp+up)v),\displaystyle\mathcal{A}U=\left(\begin{array}[]{c}-v\\ -\Delta u\end{array}\right),\quad\mathcal{B}U=\left(\begin{array}[]{c}0\\ f(u)+(\|v\|^{p}+\|\nabla u\|^{p})v\end{array}\right),

respectively, where D(𝒜)=(H2(Ω)H01(Ω))×H01(Ω)D(\mathcal{A})=(H^{2}(\Omega)\cap H_{0}^{1}(\Omega))\times H_{0}^{1}(\Omega). Then Problem (1.1) can be written into

{Ut+𝒜U+U=0,U(0)=U0,\begin{cases}U_{t}+\mathcal{A}U+\mathcal{B}U=0,\\ U(0)=U_{0},\end{cases}

where U0=(u0,u1)TU_{0}=(u_{0},u_{1})^{T}. It is easy to see that 𝒜\mathcal{A} is monotone. To prove 𝒜\mathcal{A} is maximal monotone, it is sufficient to show that

Rg(𝒜+I)=H01(Ω)×L2(Ω).Rg(\mathcal{A}+I)=H_{0}^{1}(\Omega)\times L^{2}(\Omega). (2.4)

Indeed, for any (f0,f1)H01(Ω)×L2(Ω)(f_{0},f_{1})\in H_{0}^{1}(\Omega)\times L^{2}(\Omega), we consider the following equation

(𝒜+I)U=(v+uΔu+v)=(f0f1).(\mathcal{A}+I)U=\left(\begin{aligned} -v+u\\ -\Delta u+v\end{aligned}\right)=\left(\begin{aligned} f_{0}\\ f_{1}\end{aligned}\right). (2.5)

Eliminating uu and vv respectively, we transform (2.5) into

{Δu+u=f0+f1L2(Ω),Δv+v=Δf0+f1H1(Ω).\begin{cases}-\Delta u+u=f_{0}+f_{1}\in L^{2}(\Omega),\\ -\Delta v+v=\Delta f_{0}+f_{1}\in H^{-1}(\Omega).\end{cases} (2.6)

By the method of variations, it is well known that (2.6) is well-posed, and so is (2.5). Therefore, we have proved (2.4) and 𝒜\mathcal{A} is maximal monotone. Denote θu=θu1+(1θ)u2,θv=θv1+(1θ)v2\theta_{u}=\theta u_{1}+(1-\theta)u_{2},\theta_{v}=\theta v_{1}+(1-\theta)v_{2}. It holds that

f(u1)f(u2)2\displaystyle\|f(u_{1})-f(u_{2})\|^{2} =Ω|01f(θu)(u1u2)𝑑θ|2𝑑x\displaystyle=\int_{\Omega}\left|\int_{0}^{1}f^{\prime}(\theta_{u})(u_{1}-u_{2})d\theta\right|^{2}dx
CΩ|u1u2|2(01(1+θu2)𝑑θ)2𝑑x\displaystyle\leq C\int_{\Omega}|u_{1}-u_{2}|^{2}\left(\int_{0}^{1}(1+\theta_{u}^{2})d\theta\right)^{2}dx
Cu1u262(Ω(1+|u1|2+|u2|2)3𝑑x)23\displaystyle\leq C\|u_{1}-u_{2}\|_{6}^{2}\left(\int_{\Omega}(1+|u_{1}|^{2}+|u_{2}|^{2})^{3}dx\right)^{\frac{2}{3}}
C(1+u1+u2)4u1u22,\displaystyle\leq C(1+\|\nabla u_{1}\|+\|\nabla u_{2}\|)^{4}\|\nabla u_{1}-\nabla u_{2}\|^{2},

i.e., f(u1)f(u2)C(1+u1+u2)2u1u2\|f(u_{1})-f(u_{2})\|\leq C(1+\|\nabla u_{1}\|+\|\nabla u_{2}\|)^{2}\|\nabla u_{1}-\nabla u_{2}\|. Besides, we have

v1pv1v2pv2\displaystyle\left\|\|v_{1}\|^{p}v_{1}-\|v_{2}\|^{p}v_{2}\right\| |v1pv2p|v1+v2pv1v2\displaystyle\leq\left|\|v_{1}\|^{p}-\|v_{2}\|^{p}\right|\|v_{1}\|+\|v_{2}\|^{p}\|v_{1}-v_{2}\|
=|p01θvp2(θv,v1v2)𝑑θ|v1+v2pv1v2\displaystyle=\left|p\int_{0}^{1}\|\theta_{v}\|^{p-2}(\theta_{v},v_{1}-v_{2})d\theta\right|\|v_{1}\|+\|v_{2}\|^{p}\|v_{1}-v_{2}\|
(Cp(v1+v2)p1v1+v2p)v1v2,\displaystyle\leq\left(C_{p}(\|v_{1}\|+\|v_{2}\|)^{p-1}\|v_{1}\|+\|v_{2}\|^{p}\right)\|v_{1}-v_{2}\|,

and

u1pv1u2pv2=(u1pu2p)v1+u2p(v1v2)\displaystyle\|\|\nabla u_{1}\|^{p}v_{1}-\|\nabla u_{2}\|^{p}v_{2}\|=\left\|(\|\nabla u_{1}\|^{p}-\|\nabla u_{2}\|^{p})v_{1}+\|\nabla u_{2}\|^{p}(v_{1}-v_{2})\right\|
\displaystyle\leq |p01θup2(θu,u1u2)𝑑θ|v1+u2pv1v2\displaystyle\left|p\int_{0}^{1}\|\nabla\theta_{u}\|^{p-2}(\nabla\theta_{u},\nabla u_{1}-\nabla u_{2})d\theta\right|\|v_{1}\|+\|\nabla u_{2}\|^{p}\|v_{1}-v_{2}\|
\displaystyle\leq Cpv1(u1+u2)p1u1u2+u2pv1v2.\displaystyle C_{p}\|v_{1}\|(\|\nabla u_{1}\|+\|\nabla u_{2}\|)^{p-1}\|\nabla u_{1}-\nabla u_{2}\|+\|\nabla u_{2}\|^{p}\|v_{1}-v_{2}\|.

Therefore, \mathcal{B} is locally Lipschitz on H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega). Hence, it follows from Lemma 2.3 that, for any (u0,u1)(H2(Ω)H01(Ω))×H01(Ω)(u_{0},u_{1})\in(H^{2}(\Omega)\cap H_{0}^{1}(\Omega))\times H_{0}^{1}(\Omega), there exists tmaxt_{max}\leq\infty such that Problem (1.1) has a unique strong solution on [0,tmax)[0,t_{max}). Noticing that D(𝒜)D(\mathcal{A}) is dense in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega), we also know that, for any (u0,u1)H01(Ω)×L2(Ω)(u_{0},u_{1})\in H_{0}^{1}(\Omega)\times L^{2}(\Omega), Problem (1.1) has a unique generalized solution. Besides, both strong solutions and generalized solutions belong to C([0,tmax);H01(Ω)×L2(Ω))C([0,t_{max});H_{0}^{1}(\Omega)\times L^{2}(\Omega)) and tmaxt_{max} is maximal in the sense that

if tmax<, then limttmax(u,ut)H01(Ω)×L2(Ω)=.\text{if }t_{max}<\infty,\text{ then }\lim_{t\rightarrow t_{max}}\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}=\infty.

To prove the global well-posedness, it suffices to show that (u,ut)H01(Ω)×L2(Ω)\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)} would not blow up in finite time. Taking μ0(μ,λ1)\mu_{0}\in(-\mu,\lambda_{1}), we know from (1.3) that there exists M>0M>0 such that

f(s)>μ0for |s|>M,f^{\prime}(s)>-\mu_{0}\quad\textrm{for }|s|>M, (2.7)

which implies that

{F(s)μ02s2C1,|s|>M,F(s)C1,|s|M.\begin{cases}F(s)\geq-\frac{\mu_{0}}{2}s^{2}-C_{1},&|s|>M,\\ F(s)\leq C_{1},&|s|\leq M.\end{cases}

Therefore, we have that

ΩF(u)𝑑x\displaystyle\int_{\Omega}F(u)dx ={|u|>M}F(u)𝑑x+{|u|M}F(u)𝑑xμ02u2C\displaystyle=\int_{\{|u|>M\}}F(u)dx+\int_{\{|u|\leq M\}}F(u)dx\geq-\frac{\mu_{0}}{2}\|u\|^{2}-C
μ02λ1u2C,\displaystyle\geq-\frac{\mu_{0}}{2\lambda_{1}}\|\nabla u\|^{2}-C,

and further

E(t)12(1μ0λ1)(u2+ut2)C.E(t)\geq\frac{1}{2}\Big{(}1-\frac{\mu_{0}}{\lambda_{1}}\Big{)}(\|\nabla u\|^{2}+\|u_{t}\|^{2})-C. (2.8)

Multiplying equation (1.1) by utu_{t} and integrating over Ω×(0,tmax)\Omega\times(0,t_{max}), we know that uu satisfies the energy equality (2.2), which implies E(t)E(t) is non-increasing with tt. This tells us that E(t)E(t) would not blow up in finite time, and so is (u,ut)H01(Ω)×L2(Ω)\|(u,u_{t})\|_{H^{1}_{0}(\Omega)\times L^{2}(\Omega)} by (2.8). Hence the strong solutions and generalized solutions are globally well-posed.

Step 2. We prove that generalized solutions are also weak solutions. Let (u,ut)(u,u_{t}) be a generalized solution with initial data (u0,u1)(u_{0},u_{1}). Then there exists a sequence of strong solutions (u(j),ut(j))(u^{(j)},u_{t}^{(j)}) with initial data (u0(j),u1(j))(u_{0}^{(j)},u_{1}^{(j)}) such that

(u(j),ut(j))(u,ut) in C([0,T];H01(Ω)×L2(Ω)) as j.(u^{(j)},u_{t}^{(j)})\rightarrow(u,u_{t})\text{ in }C([0,T];H_{0}^{1}(\Omega)\times L^{2}(\Omega))\text{ as }j\rightarrow\infty.

Obviously, (2.1) holds for each (u(j),ut(j))(u^{(j)},u_{t}^{(j)}), i.e.

Ωut(j)(x)ψ(x)dx+0t[Ωu(j)(τ,x)ψ(x)dx+Ωf(u(j)(τ,x))ψ(x)dx\displaystyle\int_{\Omega}u^{(j)}_{t}(x)\psi(x)dx+\int_{0}^{t}\left[\int_{\Omega}\nabla u^{(j)}(\tau,x)\nabla\psi(x)dx+\int_{\Omega}f(u^{(j)}(\tau,x))\psi(x)dx\right.
+(u(j)(τ)p+ut(j)(τ)p)Ωut(j)(τ,x)ψ(x)dx]dτ=Ωu(j)1(x)ψ(x)dx.\displaystyle+\left.(\|\nabla u^{(j)}(\tau)\|^{p}+\|u^{(j)}_{t}(\tau)\|^{p})\int_{\Omega}u^{(j)}_{t}(\tau,x)\psi(x)dx\right]d\tau=\int_{\Omega}u^{(j)}_{1}(x)\psi(x)dx.

It is easy to see that

0tΩu(j)ψ+f(u(j))ψdxdt0tΩuψ+f(u)ψdxdt.\displaystyle\int_{0}^{t}\int_{\Omega}\nabla u^{(j)}\nabla\psi+f(u^{(j)})\psi dxdt\rightarrow\int_{0}^{t}\int_{\Omega}\nabla u\nabla\psi+f(u)\psi dxdt.

Besides, since Ωut(j)ψ𝑑x\int_{\Omega}u_{t}^{(j)}\psi dx converges to Ωutψ𝑑x\int_{\Omega}u_{t}\psi dx for all t[0,T]t\in[0,T] and (u(j)p+ut(j)p)(ut(j),ψ)(\|\nabla u^{(j)}\|^{p}+\|u^{(j)}_{t}\|^{p})(u_{t}^{(j)},\psi) is bounded on [0,T][0,T], we infer from Lebesgue Dominated Convergence Theorem that

0t(u(j)p+ut(j)p)Ωut(j)ψ𝑑x𝑑t0t(up+utp)Ωutψ𝑑x𝑑t.\int_{0}^{t}(\|\nabla u^{(j)}\|^{p}+\|u^{(j)}_{t}\|^{p})\int_{\Omega}u_{t}^{(j)}\psi dxdt\rightarrow\int_{0}^{t}(\|\nabla u\|^{p}+\|u_{t}\|^{p})\int_{\Omega}u_{t}\psi dxdt.

Hence, equality (2.1) holds for uu, i.e., uu is also a weak solution. Following a similar argument, one can verify the energy equality (2.2). ∎

By Theorem 2.2, Problem (1.1) generates a semigroup on H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega), denoted by S(t)S(t). We will show S(t)S(t) possesses a global attractor. Let us start with the dissipation.

Theorem 2.4.

Under Assumption 1.1, the dynamical system (H01(Ω)×L2(Ω),S(t))(H_{0}^{1}(\Omega)\times L^{2}(\Omega),S(t)) generated by Problem (1.1) is dissipative. In other words, there exists a constant R>0R>0 such that for any bounded set BH01(Ω)×L2(Ω)B\subset H_{0}^{1}(\Omega)\times L^{2}(\Omega), there exists t0=t0(B)t_{0}=t_{0}(B) satisfying, for any tt0t\geq t_{0} and any (u,v)B(u,v)\in B,

S(t)(u,v)H01(Ω)×L2(Ω)R.\|S(t)(u,v)\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\leq R.
Proof.

Denote K=sup(u,v)B(u,v)H01(Ω)×L2(Ω)<K=\sup_{(u,v)\in B}\|(u,v)\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}<\infty. Multiplying equation (1.1) by ut+ϵuu_{t}+\epsilon u for some ϵ>0\epsilon>0 small and integrating over Ω\Omega, we have

ddtEϵ(t)+ϵEϵ(t)+(u(t)p+ut(t)p)ut(t)2+ϵ2u(t)2=3ϵ2ut(t)2+ϵ2(ut(t),u(t))+ϵΩF(u(t))𝑑xϵ(f(u(t)),u(t))ϵ(u(t)p+ut(t)p)(ut(t),u(t)),\displaystyle\begin{split}\frac{d}{dt}E_{\epsilon}(t)+\epsilon E_{\epsilon}(t)&+(\|\nabla u(t)\|^{p}+\|u_{t}(t)\|^{p})\|u_{t}(t)\|^{2}+\frac{\epsilon}{2}\|\nabla u(t)\|^{2}\\ =\frac{3\epsilon}{2}\|u_{t}(t)\|^{2}&+\epsilon^{2}(u_{t}(t),u(t))+\epsilon\int_{\Omega}F(u(t))dx-\epsilon(f(u(t)),u(t))\\ &-\epsilon(\|\nabla u(t)\|^{p}+\|u_{t}(t)\|^{p})(u_{t}(t),u(t)),\end{split} (2.9)

where Eϵ(t)=12(u(t)2+ut(t)2)+ΩF(u(t))𝑑x+ϵ(u(t),ut(t))E_{\epsilon}(t)=\frac{1}{2}(\|\nabla u(t)\|^{2}+\|u_{t}(t)\|^{2})+\int_{\Omega}F(u(t))dx+\epsilon(u(t),u_{t}(t)). By (2.8) we know that if ϵ<ϵ1\epsilon<\epsilon_{1} for some ϵ1\epsilon_{1} small

Eϵ(t)\displaystyle E_{\epsilon}(t) =E(t)+ϵ(u(t),ut(t))12(1μ0λ1)(u2+ut2)C+ϵ(u(t),ut(t))\displaystyle=E(t)+\epsilon(u(t),u_{t}(t))\geq\frac{1}{2}\left(1-\frac{\mu_{0}}{\lambda_{1}}\right)(\|\nabla u\|^{2}+\|u_{t}\|^{2})-C+\epsilon(u(t),u_{t}(t))
14(1μ0λ1)(u2+ut2)C.\displaystyle\geq\frac{1}{4}\left(1-\frac{\mu_{0}}{\lambda_{1}}\right)(\|\nabla u\|^{2}+\|u_{t}\|^{2})-C. (2.10)

Let us estimate the right hand side of (2.9) term by term. Firstly, it holds that

3ϵ2ut(t)2Cpϵp+2p+14utp+2,\frac{3\epsilon}{2}\|u_{t}(t)\|^{2}\leq C_{p}\epsilon^{\frac{p+2}{p}}+\frac{1}{4}\|u_{t}\|^{p+2}, (2.11)

and

ϵ2|(ut(t),u(t))|λ112ϵ2utuϵ22λ1(ut2up+u2p)ϵ22λ1(ut2up+u2+Cp)14ut2up+ϵ8(1μ0λ1)u2+Cpϵ,\displaystyle\begin{split}\epsilon^{2}|(u_{t}(t),u(t))|&\leq\lambda_{1}^{-\frac{1}{2}}\epsilon^{2}\|u_{t}\|\|\nabla u\|\leq\frac{\epsilon^{2}}{2\sqrt{\lambda_{1}}}(\|u_{t}\|^{2}\|\nabla u\|^{p}+\|\nabla u\|^{2-p})\\ &\leq\frac{\epsilon^{2}}{2\sqrt{\lambda_{1}}}(\|u_{t}\|^{2}\|\nabla u\|^{p}+\|\nabla u\|^{2}+C_{p})\\ &\leq\frac{1}{4}\|u_{t}\|^{2}\|\nabla u\|^{p}+\frac{\epsilon}{8}(1-\frac{\mu_{0}}{\lambda_{1}})\|\nabla u\|^{2}+C_{p}\epsilon,\end{split} (2.12)

if ϵ<ϵ2\epsilon<\epsilon_{2} for some ϵ2\epsilon_{2} small. For |s|M|s|\geq M, we can infer from (2.7) that

(12μ0s2+F(s))|Ms=Ms(f(τ)+μ0τ)𝑑τ(f(s)+μ0s)(sM),\displaystyle\left.\left(\frac{1}{2}\mu_{0}s^{2}+F(s)\right)\right|_{M}^{s}=\int_{M}^{s}(f(\tau)+\mu_{0}\tau)d\tau\leq(f(s)+\mu_{0}s)(s-M),

which implies

F(s)f(s)s+μ02s2+Cfor all |s|>M.F(s)\leq f(s)s+\frac{\mu_{0}}{2}s^{2}+C\quad\textrm{for all }|s|>M.

Hence,

ϵΩF(u)𝑑xϵ(f(u),u)\displaystyle\epsilon\int_{\Omega}F(u)dx-\epsilon(f(u),u) ϵ({|u|>M}(F(u)f(u)u)𝑑x+{|u|M}(F(u)f(u)u)𝑑x)\displaystyle\leq\epsilon\Big{(}\int_{\{|u|>M\}}(F(u)-f(u)u)dx+\int_{\{|u|\leq M\}}(F(u)-f(u)u)dx\Big{)}
ϵ{|u|>M}(μ02|u|2+C)+ϵΩCM\displaystyle\leq\epsilon\int_{\{|u|>M\}}\left(\frac{\mu_{0}}{2}|u|^{2}+C\right)+\epsilon\int_{\Omega}C_{M}
μ0ϵ2λ1u2+Cϵ.\displaystyle\leq\frac{\mu_{0}\epsilon}{2\lambda_{1}}\|\nabla u\|^{2}+C\epsilon. (2.13)

Noticing that E(t)E(t) is non-increasing, we know from (2.8) that

u(t)2+ut(t)2CK and Eϵ(t)CK,for all t0,\|\nabla u(t)\|^{2}+\|u_{t}(t)\|^{2}\leq C_{K}\textrm{ and }E_{\epsilon}(t)\leq C_{K},\quad\textrm{for all }t\geq 0,

where CK>0C_{K}>0 is a constant depending on KK. Setting ϵ<ϵ3=min{14λ112,λ1μ08CKp}\epsilon<\epsilon_{3}=\min\{\frac{1}{4}\lambda_{1}^{\frac{1}{2}},\frac{\lambda_{1}-\mu_{0}}{8C_{K}^{p}}\}, it follows from Young’s inequality that

ϵ(u(t)p+ut(t)p)(ut(t),u(t))ϵλ112up+1ut+ϵλ112uutp+1ϵ8(1μ0λ1)u2+2ϵλ1(1μ0λ1)1u2put2+ϵλ112(1puput2+p1putp+2)ϵ8(1μ0λ1)u2+12(up+utp)ut2.\displaystyle\begin{split}&\quad\epsilon(\|\nabla u(t)\|^{p}+\|u_{t}(t)\|^{p})(u_{t}(t),u(t))\leq\epsilon\lambda_{1}^{-\frac{1}{2}}\|\nabla u\|^{p+1}\|u_{t}\|+\epsilon\lambda_{1}^{-\frac{1}{2}}\|\nabla u\|\|u_{t}\|^{p+1}\\ &\leq\frac{\epsilon}{8}\left(1-\frac{\mu_{0}}{\lambda_{1}}\right)\|\nabla u\|^{2}+\frac{2\epsilon}{\lambda_{1}}\left(1-\frac{\mu_{0}}{\lambda_{1}}\right)^{-1}\|\nabla u\|^{2p}\|u_{t}\|^{2}\\ &\qquad\qquad+\epsilon\lambda_{1}^{-\frac{1}{2}}\left(\frac{1}{p}\|\nabla u\|^{p}\|u_{t}\|^{2}+\frac{p-1}{p}\|u_{t}\|^{p+2}\right)\\ &\leq\frac{\epsilon}{8}\left(1-\frac{\mu_{0}}{\lambda_{1}}\right)\|\nabla u\|^{2}+\frac{1}{2}(\|\nabla u\|^{p}+\|u_{t}\|^{p})\|u_{t}\|^{2}.\end{split} (2.14)

Inserting (2.11)-(2.14) into (2.9), we conclude that

ddtEϵ(t)+ϵEϵ(t)+14u(t)put(t)2+12utp+2+ϵ4(1μ0λ1)u2Cpϵ,\frac{d}{dt}E_{\epsilon}(t)+\epsilon E_{\epsilon}(t)+\frac{1}{4}\|\nabla u(t)\|^{p}\|u_{t}(t)\|^{2}+\frac{1}{2}\|u_{t}\|^{p+2}+\frac{\epsilon}{4}\left(1-\frac{\mu_{0}}{\lambda_{1}}\right)\|\nabla u\|^{2}\leq C_{p}\epsilon,

whenever ϵ<min{ϵ1,ϵ2,ϵ3}\epsilon<\min\{\epsilon_{1},\epsilon_{2},\epsilon_{3}\} is fixed. Hence, by the Gronwall lemma, we get

Eϵ(t)Eϵ(0)eϵt+C(1eϵt)CKeϵt+C,E_{\epsilon}(t)\leq E_{\epsilon}(0)e^{-\epsilon t}+C(1-e^{-\epsilon t})\leq C_{K}e^{-\epsilon t}+C,

which, by choosing t0=ϵ1ln(CK)t_{0}=\epsilon^{-1}\ln(C_{K}), implies

Eϵ(t)C+1for all tt0.E_{\epsilon}(t)\leq C+1\quad\textrm{for all }t\geq t_{0}.

This, together with (2.10), yields the dissipation. ∎

As mentioned in the introduction, we can not deal with the damping term directly since it is neither compact nor contractive. To overcome this difficulty, we would make a decomposition and prove the asymptotic smoothness based on Propositions 2.5 and 2.6.

Proposition 2.5 ([8]).

Let S(t)S(t) be an evolution operator in a Banach space XX. Assume that for each t>0t>0 there exists a decomposition S(t)=S(1)(t)+S(2)(t)S(t)=S^{(1)}(t)+S^{(2)}(t), where S(2)(t)S^{(2)}(t) is a mapping in XX satisfying that for any bounded set BB

rB(t)=sup{S(2)(t)xX:xB}0 as t,r_{B}(t)=\sup\left\{\|S^{(2)}(t)x\|_{X}:x\in B\right\}\to 0\text{ as }t\to\infty, (2.15)

and S(1)(t)S^{(1)}(t) is compact in the sense that for each t>0t>0 the set S(1)(t)BS^{(1)}(t)B is a relatively compact set in XX for every t>0t>0 large enough and every bounded forward invariant set BB in XX. Then S(t)S(t) is asymptotically smooth.

Proposition 2.6 ([10]).

Let (X,S(t))(X,S(t)) be a dissipative dynamical system in a complete metric space XX. Then (X,S(t))(X,S(t)) possesses a compact global attractor if and only if (X,S(t))(X,S(t)) is asymptotically smooth.

Given λ>λ1\lambda>-\lambda_{1}, we decompose u=v+wu=v+w, where

{vttΔv+12utput12wtpwt+(up+12utp)vt+λv=0,v|Ω=0,v(0)=u0,vt(0)=u1,\begin{cases}v_{tt}-\Delta v+\frac{1}{2}\|u_{t}\|^{p}u_{t}-\frac{1}{2}\|w_{t}\|^{p}w_{t}+\left(\|\nabla u\|^{p}+\frac{1}{2}\|u_{t}\|^{p}\right)v_{t}+\lambda v=0,\\ v|_{\partial\Omega}=0,\\ v(0)=u_{0},v_{t}(0)=u_{1},\end{cases} (2.16)

and

{wttΔw+(12wtp+up+12utp)wt+f(u)λu+λw=0,w|Ω=0,w(0)=wt(0)=0.\begin{cases}w_{tt}-\Delta w+\left(\frac{1}{2}\|w_{t}\|^{p}+\|\nabla u\|^{p}+\frac{1}{2}\|u_{t}\|^{p}\right)w_{t}+f(u)-\lambda u+\lambda w=0,\\ w|_{\partial\Omega}=0,\\ w(0)=w_{t}(0)=0.\end{cases} (2.17)

We point out that the parameter λ\lambda is introduced in order to weaken the conditions on ff in Section 3. Lemmas 2.7 and 2.8 below are established under Assumption 1.1.

Lemma 2.7.

Let B0B_{0} be a bounded subset of H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega). Then any solution of (2.16) starting from B0B_{0} satisfies that

v2+vt20 as t,\|\nabla v\|^{2}+\|v_{t}\|^{2}\rightarrow 0\text{ as }t\to\infty,

In particular, the decay is uniform with respect to (u0,u1)B0(u_{0},u_{1})\in B_{0}.

Proof.

Multiplying equation (2.16) by vtv_{t} and integrating over [0,t]×Ω[0,t]\times\Omega, we get

I~v(t)I~v(0)+20t(g(vt),vt)=0,\tilde{I}_{v}(t)-\tilde{I}_{v}(0)+2\int_{0}^{t}(g(v_{t}),v_{t})=0,

where

I~v\displaystyle\tilde{I}_{v} =v2+vt2+λv2,\displaystyle=\|\nabla v\|^{2}+\|v_{t}\|^{2}+\lambda\|v\|^{2},
g(vt)\displaystyle g(v_{t}) =12utput12wtpwt+(up+12utp)vt.\displaystyle=\frac{1}{2}\|u_{t}\|^{p}u_{t}-\frac{1}{2}\|w_{t}\|^{p}w_{t}+\Big{(}\|\nabla u\|^{p}+\frac{1}{2}\|u_{t}\|^{p}\Big{)}v_{t}.

Note that c1IvI~vc2Ivc_{1}I_{v}\leq\tilde{I}_{v}\leq c_{2}I_{v} for some c2>c1>0c_{2}>c_{1}>0, as λ>λ1\lambda>-\lambda_{1}, where IvI_{v} is defined by (1.6). By the monotonicity inequality (Lemma 2.2 in [26]), it holds that

(utputwtpwt,vt)Cvtp+2,\displaystyle(\|u_{t}\|^{p}u_{t}-\|w_{t}\|^{p}w_{t},v_{t})\geq C\|v_{t}\|^{p+2},

and thereby

(g(vt),vt)Cvtp+2+(up+12utp)vt20,(g(v_{t}),v_{t})\geq C\|v_{t}\|^{p+2}+\Big{(}\|\nabla u\|^{p}+\frac{1}{2}\|u_{t}\|^{p}\Big{)}\|v_{t}\|^{2}\geq 0, (2.18)

which means that I~v(t)\tilde{I}_{v}(t) is monotonically decreasing. This, together with Theorem 2.4, tells us that u\|\nabla u\|, ut\|u_{t}\|, v\|\nabla v\| and vt\|v_{t}\| are uniformly bounded for t[0,)t\in[0,\infty).

To prove the lemma, we claim that, for fixed T>0T>0 and any M>0M>0, there exists a positive constant K(M)K(M) such that

I~v(T)I~v(0)K(M),\tilde{I}_{v}(T)-\tilde{I}_{v}(0)\leq-K(M),

whenever (u0,u1)B0(u_{0},u_{1})\in B_{0} and I~v(T)M\tilde{I}_{v}(T)\geq M. Otherwise, there exist M>0M>0 and a sequence (vn,vtn)(v^{n},v^{n}_{t}) associated with (u0n,u1n)B0(u_{0}^{n},u_{1}^{n})\in B_{0} such that

I~vn(T)Mandlimn0T(g(vn),vtn)𝑑t=0.\displaystyle\tilde{I}_{v^{n}}(T)\geq M~{}\mathrm{and}~{}\lim_{n\rightarrow\infty}\int_{0}^{T}(g(v^{n}),v^{n}_{t})dt=0. (2.19)

We infer from (2.18) and (2.19) that

vtn0 in Lp+2(0,T).\|v^{n}_{t}\|\to 0\text{\rm in }L^{p+2}(0,T). (2.20)

Besides, since {(vn,vtn)}\{(v^{n},v^{n}_{t})\} is bounded in L(0,T;H01(Ω)×L2(Ω))L^{\infty}(0,T;H_{0}^{1}(\Omega)\times L^{2}(\Omega)), we can find a function vv such that, up to a subsequence

vnv in L(0,T;H01(Ω)),\displaystyle v^{n}\stackrel{{\scriptstyle*}}{{\rightharpoonup}}v\text{\rm in }L^{\infty}(0,T;H_{0}^{1}(\Omega)),
vnv a.e. in [0,T]×Ω,\displaystyle v^{n}\to v\text{\rm a.e. in }[0,T]\times\Omega,
vnv in C([0,T];L2(Ω))\displaystyle v^{n}\to v\text{\rm in }C([0,T];L^{2}(\Omega)) (2.21)

and also

vtn0, a.e. in [0,T]×Ω,v^{n}_{t}\to 0,\text{\rm a.e. in }[0,T]\times\Omega,

where we have used Aubin-Lions Lemma in (2.21). Since

2g(vtn)\displaystyle 2\|g(v^{n}_{t})\| utnputnutnvtnp(utnvtn)+(2unp+utnp)vtn\displaystyle\leq\left\|\|u^{n}_{t}\|^{p}u^{n}_{t}-\|u^{n}_{t}-v^{n}_{t}\|^{p}(u^{n}_{t}-v^{n}_{t})\right\|+\big{\|}(2\|\nabla u^{n}\|^{p}+\|u^{n}_{t}\|^{p})v^{n}_{t}\big{\|}
|utnputnvtnp|utn+utnvtnpvtn+(2unp+utnp)vtn\displaystyle\leq\left|\|u^{n}_{t}\|^{p}-\|u^{n}_{t}-v^{n}_{t}\|^{p}\right|\|u^{n}_{t}\|+\|u^{n}_{t}-v^{n}_{t}\|^{p}\|v^{n}_{t}\|+(2\|\nabla u^{n}\|^{p}+\|u^{n}_{t}\|^{p})\|v^{n}_{t}\|
CB0vtn,\displaystyle\leq C_{B_{0}}\|v^{n}_{t}\|,

we also have g(vtn)0g(v^{n}_{t})\to 0 a.e. in [0,T]×Ω[0,T]\times\Omega. Therefore, taking nn to \infty in (2.16), one can see the limit function vv satisfies

{Δv+λv=0,inΩ×[0,T],v|Ω=0,\begin{cases}-\Delta v+\lambda v=0,&\mathrm{in}~{}\Omega\times[0,T],\\ v|_{\partial\Omega}=0,\end{cases}

which implies that v=0v=0 in H01(Ω)H_{0}^{1}(\Omega) for each t[0,T]t\in[0,T]. Furthermore, multiplying the equation of vnv^{n} by vnv^{n} and integrating over [0,T]×Ω[0,T]\times\Omega, we know

0Tvn2𝑑t=(vtn(0),vn(0))(vtn(T),vn(T))+0Tvtn2(g(vtn),vn)dtλvn2,\int_{0}^{T}\|\nabla v^{n}\|^{2}dt=(v^{n}_{t}(0),v^{n}(0))-(v^{n}_{t}(T),v^{n}(T))+\int_{0}^{T}\|v^{n}_{t}\|^{2}-(g(v^{n}_{t}),v^{n})dt-\lambda\|v^{n}\|^{2},

which, together with (2.20) and (2.21), infers that 0Tvn2𝑑t0\int_{0}^{T}\|\nabla v^{n}\|^{2}dt\to 0 as nn tends to \infty. Therefore, it follows from the monotonicity of I~v\tilde{I}_{v} that

Mlim supnI~vn(T)lim supn1T0TI~vn(t)𝑑t=0,M\leq\limsup_{n\to\infty}\tilde{I}_{v^{n}}(T)\leq\limsup_{n\rightarrow\infty}\frac{1}{T}\int_{0}^{T}\tilde{I}_{v^{n}}(t)dt=0,

which is a contraction.

As the conclusion, the claim is true and, for any ϵ>0\epsilon>0, there exists Nϵ+N_{\epsilon}\in\mathbb{N}^{+} such that

0I~v(t)ϵfor any tNϵT.0\leq\tilde{I}_{v}(t)\leq\epsilon\quad\textrm{for any }t\geq N_{\epsilon}T.

By the equivalence of I~v\tilde{I}_{v} and IvI_{v}, we complete the proof. ∎

Lemma 2.8.

For each t>0t>0, the solution (w(t),wt(t))(w(t),w_{t}(t)) of (2.17) at tt is bounded in Hβ+1(Ω)×Hβ(Ω)H^{\beta+1}(\Omega)\times H^{\beta}(\Omega) for some β>0\beta>0 small enough, where the bound is uniform with respect to (u0,u1)B0(u_{0},u_{1})\in B_{0}.

Proof.

Note that (u,ut)(u,u_{t}) is globally bounded in H01(Ω)×L2(Ω)H^{1}_{0}(\Omega)\times L^{2}(\Omega), where the bound depends on B0B_{0}. For any T>0T>0, by virtue of the strichartz estimate (Corollary 1.2 in [3]), we know

uL7/2(0,T;L14(Ω))C(u(T/2)H1(Ω)+ut(T/2)L2(Ω)+f(u)L1(0,T;L2(Ω))+(up+utp)utL1(0,T;L2(Ω)))C(B0)+CuL3(0,T;L6(Ω))3+C(T,B0)C(T,B0).\begin{split}\|u\|_{L^{7/2}(0,T;L^{14}(\Omega))}\leq&C\Big{(}\|u(T/2)\|_{H^{1}(\Omega)}+\|u_{t}(T/2)\|_{L^{2}(\Omega)}+\|f(u)\|_{L^{1}(0,T;L^{2}(\Omega))}\\ &\quad+\big{\|}(\|\nabla u\|^{p}+\|u_{t}\|^{p})u_{t}\big{\|}_{L^{1}(0,T;L^{2}(\Omega))}\Big{)}\\ \leq&C(B_{0})+C\|u\|_{L^{3}(0,T;L^{6}(\Omega))}^{3}+C(T,B_{0})\leq C(T,B_{0}).\end{split} (2.22)

Denote f(s)=f(s)λsf_{*}(s)=f(s)-\lambda s. It follows from the condition on ff that

f(u)W1,6/5(Ω)C(1+uH1(Ω)3),\|f_{*}(u)\|_{W^{1,6/5}(\Omega)}\leq C(1+\|u\|_{H^{1}(\Omega)}^{3}),

and

f(u)L7/6(0,T;L14/3(Ω))C(1+uL7/2(0,T;L14(Ω))3),\|f_{*}(u)\|_{L^{7/6}(0,T;L^{14/3}(\Omega))}\leq C(1+\|u\|_{L^{7/2}(0,T;L^{14}(\Omega))}^{3}),

which, together with the interpolation inequality gH6/13(Ω)gW1,6/5(Ω)6/13gL14/3(Ω)7/13\|g\|_{H^{6/13}(\Omega)}\leq\|g\|_{W^{1,6/5}(\Omega)}^{6/13}\|g\|_{L^{14/3}(\Omega)}^{7/13}, gives

f(u)L13/6(0,T;H6/13(Ω))f(u)L(0,T;W1,6/5(Ω))6/13f(u)L7/6(0,T;L14/3(Ω))7/13C(T,B0).\displaystyle\|f_{*}(u)\|_{L^{13/6}(0,T;H^{6/13}(\Omega))}\leq\|f_{*}(u)\|_{L^{\infty}(0,T;W^{1,6/5}(\Omega))}^{6/13}\|f_{*}(u)\|_{L^{7/6}(0,T;L^{14/3}(\Omega))}^{7/13}\leq C(T,B_{0}).

Since D(As)=H2s(Ω)D(A^{s})=H^{2s}(\Omega) for s<1/4s<1/4, we can denote w~=A313w\tilde{w}=A^{\frac{3}{13}}w and consider the equation

w~ttΔw~+λw~+(12wtp+up+12utp)w~t+A313f(u)=0\tilde{w}_{tt}-\Delta\tilde{w}+\lambda\tilde{w}+\Big{(}\frac{1}{2}\|w_{t}\|^{p}+\|\nabla u\|^{p}+\frac{1}{2}\|u_{t}\|^{p}\Big{)}\tilde{w}_{t}+A^{\frac{3}{13}}f_{*}(u)=0

with (w~(0),w~t(0))=0(\tilde{w}(0),\tilde{w}_{t}(0))=0 and Dirichlet boundary. Then we know (w~(t),w~t(t))(\tilde{w}(t),\tilde{w}_{t}(t)) is bounded in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega) by energy estimate, which implies the conclusion with β=613\beta=\frac{6}{13}. ∎

Define S(2)(t)(u0,u1)=(v(t),vt(t))S^{(2)}(t)(u_{0},u_{1})=(v(t),v_{t}(t)) and S(1)(t)(u0,u1)=(w(t),wt(t))S^{(1)}(t)(u_{0},u_{1})=(w(t),w_{t}(t)). We know that S(2)(t)S^{(2)}(t) satisfies (2.15) by Lemma 2.7 and S(1)(t)S^{(1)}(t) is compact for each t>0t>0 by Lemma 2.8. Therefore, by means of Proposition 2.5, S(t)S(t) is asymptotically smooth. Furthermore, since S(t)S(t) is bounded dissipative by Theorem 2.4, we obtain directly from Proposition 2.6 the existence of the global attractor.

Theorem 2.9.

Under Assumption 1.1, the dynamical system generated by Problem (1.1) has a global attractor 𝒜\mathscr{A} in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega).

3 Regularity and Decay

In this section, we establish some good properties, including the higher regularity of 𝒜\mathscr{A} and the polynomial decay rate of trajectories near the origin, which are useful to study the dimension of the global attractor. Besides Assumption 1.1, in this section we always assume the following condition:

Assumption 3.1.

f(0)=0f(0)=0, f(0)>λ1f^{\prime}(0)>-\lambda_{1}.

We start with a technical lemma, which derives the boundedness from the uniform local information.

Lemma 3.2.

Suppose that FF, ψ\psi and φ\varphi are nonnegative functions on [0,)[0,\infty) and there exist constants C1,C2C_{1},C_{2} such that

  • (i)

    ψL1(t,t+1)+φL1(t,t+1)C1\|\psi\|_{L^{1}(t,t+1)}+\|\varphi\|_{L^{1}(t,t+1)}\leq C_{1} for any t0t\geq 0;

  • (ii)

    sups[t,t+1]φ(s)C2infs[t,t+1]φ(s)\sup_{s\in[t,t+1]}\varphi(s)\leq C_{2}\inf_{s\in[t,t+1]}\varphi(s) for any t0t\geq 0;

  • (iii)

    FF is absolutely continuous and satisfies F+φFψφF^{\prime}+\varphi F\leq\psi\varphi.

Then there exists M>0M>0, depending only on F(0),C1,C2F(0),C_{1},C_{2}, such that F(t)MF(t)\leq M for all t[0,)t\in[0,\infty).

Proof.

Since

(e0tφ(s)𝑑sF(t))ψ(t)φ(t)e0tφ(s)𝑑s,\left(e^{\int_{0}^{t}\varphi(s)ds}F(t)\right)^{\prime}\leq\psi(t)\varphi(t)e^{\int_{0}^{t}\varphi(s)ds},

integrating over [0,t][0,t], we can get

e0tφF(t)F(0)0tψ(s)φ(s)e0sφ𝑑sk=1tk1kψ(s)φ(s)e0sφ𝑑s,\displaystyle e^{\int_{0}^{t}\varphi}F(t)-F(0)\leq\int_{0}^{t}\psi(s)\varphi(s)e^{\int_{0}^{s}\varphi}ds\leq\sum_{k=1}^{\lceil t\rceil}\int_{k-1}^{k}\psi(s)\varphi(s)e^{\int_{0}^{s}\varphi}ds,

where t\lceil t\rceil denotes the smallest integer larger than or equal to tt. For each kk,

k1kψ(s)φ(s)e0sφ𝑑s\displaystyle\int_{k-1}^{k}\psi(s)\varphi(s)e^{\int_{0}^{s}\varphi}ds C2k1kψ(s)𝑑sφ(k)e0kφC1C2k1kφ(k)e0kφ𝑑s\displaystyle\leq C_{2}\int_{k-1}^{k}\psi(s)ds\cdot\varphi(k)e^{\int_{0}^{k}\varphi}\leq C_{1}C_{2}\int_{k-1}^{k}\varphi(k)e^{\int_{0}^{k}\varphi}ds
C1C22eC1k1kφ(s)e0sφ𝑑s,\displaystyle\leq C_{1}C_{2}^{2}e^{C_{1}}\int_{k-1}^{k}\varphi(s)e^{\int_{0}^{s}\varphi}ds,

which implies that

e0tφF(t)F(0)C0tφ(s)e0sφ𝑑s=C[e0tφ1].\displaystyle e^{\int_{0}^{t}\varphi}F(t)-F(0)\leq C\int_{0}^{\lceil t\rceil}\varphi(s)e^{\int_{0}^{s}\varphi}ds=C\big{[}e^{\int_{0}^{\lceil t\rceil}\varphi}-1\big{]}.

Therefore,

F(t)F(0)e0tφ+C[ettφe0tφ]M,\displaystyle F(t)\leq F(0)e^{-\int_{0}^{t}\varphi}+C\big{[}e^{\int_{t}^{\lceil t\rceil}\varphi}-e^{-\int_{0}^{t}\varphi}\big{]}\leq M,

where MM depends only on F(0),C1,C2F(0),C_{1},C_{2}. ∎

Lemma 3.3.

Let (u,ut)(u,u_{t}) start from a bounded set B0H01(Ω)×L2(Ω)B_{0}\subset H_{0}^{1}(\Omega)\times L^{2}(\Omega). Set λ=f(0)\lambda=f^{\prime}(0) in equation (2.17). Then the solution (w(t),wt(t))(w(t),w_{t}(t)) keeps (globally in time) bounded in H1+β(Ω)×Hβ(Ω)H^{1+\beta}(\Omega)\times H^{\beta}(\Omega) with β=27\beta=\frac{2}{7}.

Proof.

Similar to (2.22), for any tt and any T>0T>0, we derive that

uL72(t,t+T;L14(Ω))\displaystyle\|u\|_{L^{\frac{7}{2}}(t,t+T;L^{14}(\Omega))} C(u(t+T/2)+ut(t+T/2)+f(u)L1(t,t+T;L2(Ω))\displaystyle\leq C\Big{(}\|\nabla u(t+T/2)\|+\|u_{t}(t+T/2)\|+\|f(u)\|_{L^{1}(t,t+T;L^{2}(\Omega))}
+(up+utp)utL1(t,t+T;L2(Ω)))\displaystyle\qquad\qquad+\big{\|}(\|\nabla u\|^{p}+\|u_{t}\|^{p})u_{t}\big{\|}_{L^{1}(t,t+T;L^{2}(\Omega))}\Big{)}
C(T,B0).\displaystyle\leq C(T,B_{0}).

Denote still f(s)=f(s)λsf_{*}(s)=f(s)-\lambda s. Since f(0)=f(0)=0f_{*}(0)=f_{*}^{\prime}(0)=0 and fC2f_{*}\in C^{2}, it holds that |f(s)|C(|s|+|s|2)|f_{*}^{\prime}(s)|\leq C(|s|+|s|^{2}). Setting 1q=114+16+12\frac{1}{q}=\frac{1}{14}+\frac{1}{6}+\frac{1}{2}, we calculate

f(u)W1,q\displaystyle\|f_{*}(u)\|_{W^{1,q}} Cf(u)uLqC(|u|+|u|2)uLqC(u2+u14u2)\displaystyle\leq C\|f_{*}^{\prime}(u)\nabla u\|_{L^{q}}\leq C\|(|u|+|u|^{2})\nabla u\|_{L^{q}}\leq C(\|\nabla u\|^{2}+\|u\|_{14}\|\nabla u\|^{2})
C(1+u14)u2,\displaystyle\leq C\left(1+\|u\|_{14}\right)\|\nabla u\|^{2},

and further, for β=27\beta=\frac{2}{7}

f(u)D(Aβ2)Cf(u)HβCf(u)W1,qC(1+u14)u2.\|f_{*}(u)\|_{D(A^{\frac{\beta}{2}})}\leq C\|f_{*}(u)\|_{H^{\beta}}\leq C\|f_{*}(u)\|_{W^{1,q}}\leq C\left(1+\|u\|_{14}\right)\|\nabla u\|^{2}. (3.1)

Denote w~=Aβ2w\tilde{w}=A^{\frac{\beta}{2}}w, which satisfies

{w~ttΔw~+λw~+(12wtp+up+12utp)w~t+Aβ2f(u)=0,w~(0)=w~t(0)=0.\begin{cases}\tilde{w}_{tt}-\Delta\tilde{w}+\lambda\tilde{w}+\left(\frac{1}{2}\|w_{t}\|^{p}+\|\nabla u\|^{p}+\frac{1}{2}\|u_{t}\|^{p}\right)\tilde{w}_{t}+A^{\frac{\beta}{2}}f_{*}(u)=0,\\ \tilde{w}(0)=\tilde{w}_{t}(0)=0.\end{cases} (3.2)

Multiplying (3.2) by w~t\tilde{w}_{t} and integrating over Ω\Omega, we deduce

12ddt(w~t2+w~2+λw~2)+12Iu,pw~t2+(Aβ2f(u),w~t)0,\frac{1}{2}\frac{d}{dt}(\|\tilde{w}_{t}\|^{2}+\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2})+\frac{1}{2}I_{u,p}\|\tilde{w}_{t}\|^{2}+(A^{\frac{\beta}{2}}f_{*}(u),\tilde{w}_{t})\leq 0, (3.3)

where recall that Iu,p=up+utpI_{u,p}=\|\nabla u\|^{p}+\|u_{t}\|^{p}. Multiplying (3.2) by w~\tilde{w} and integrating over Ω\Omega, we have

ddt(w~t,w~)w~t2\displaystyle\frac{d}{dt}(\tilde{w}_{t},\tilde{w})-\|\tilde{w}_{t}\|^{2} +w~2+λw~2+(Aβ2f(u),w~)\displaystyle+\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2}+(A^{\frac{\beta}{2}}f_{*}(u),\tilde{w}) (3.4)
+(12wtp+12up+12Iu,p)(w~t,w~)=0.\displaystyle+\Big{(}\frac{1}{2}\|w_{t}\|^{p}+\frac{1}{2}\|\nabla u\|^{p}+\frac{1}{2}I_{u,p}\Big{)}(\tilde{w}_{t},\tilde{w})=0.

Then we multiply (3.4) by ϵIup2\epsilon I_{u}^{\frac{p}{2}} and sum it with (3.3) to obtain

12ddt(w~t2+w~2+λw~2)+ϵddt(Iup2(w~t,w~))+(12ϵ)Iup2w~t2\displaystyle\frac{1}{2}\frac{d}{dt}(\|\tilde{w}_{t}\|^{2}+\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2})+\epsilon\frac{d}{dt}\left(I_{u}^{\frac{p}{2}}(\tilde{w}_{t},\tilde{w})\right)+\Big{(}\frac{1}{2}-\epsilon\Big{)}I_{u}^{\frac{p}{2}}\|\tilde{w}_{t}\|^{2} (3.5)
+ϵIup2(w~2+λw~2)pϵ2|(w~t,w~)Iup21ddtIu|+|(Aβ2f(u),w~t)|\displaystyle\quad+\epsilon I_{u}^{\frac{p}{2}}(\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2})\leq\frac{p\epsilon}{2}\left|(\tilde{w}_{t},\tilde{w})I_{u}^{\frac{p}{2}-1}\frac{d}{dt}I_{u}\right|+|(A^{\frac{\beta}{2}}f_{*}(u),\tilde{w}_{t})|
+ϵIup2|(Aβ2f(u),w~)|+CpϵIup2(12wtp+12up+12Iu,p)|(w~t,w~)|,\displaystyle\qquad+\epsilon I_{u}^{\frac{p}{2}}|(A^{\frac{\beta}{2}}f_{*}(u),\tilde{w})|+C_{p}\epsilon I_{u}^{\frac{p}{2}}\Big{(}\frac{1}{2}\|w_{t}\|^{p}+\frac{1}{2}\|\nabla u\|^{p}+\frac{1}{2}I_{u,p}\Big{)}|(\tilde{w}_{t},\tilde{w})|,

where the fact Iup2Iu,pI_{u}^{\frac{p}{2}}\approx I_{u,p} is used. For the first term on the right hand side of the above inequality, it holds

pϵ2|(w~t,w~)Iup21ddtIu|\displaystyle\frac{p\epsilon}{2}\left|(\tilde{w}_{t},\tilde{w})I_{u}^{\frac{p}{2}-1}\frac{d}{dt}I_{u}\right| =pϵ2|(w~t,w~)Iup21[(up+utp)ut2+(f(u),ut)]|\displaystyle=\frac{p\epsilon}{2}\left|(\tilde{w}_{t},\tilde{w})I_{u}^{\frac{p}{2}-1}\left[(\|\nabla u\|^{p}+\|u_{t}\|^{p})\|u_{t}\|^{2}+(f(u),u_{t})\right]\right|
CϵIup2w~tw~\displaystyle\leq C\epsilon I_{u}^{\frac{p}{2}}\|\tilde{w}_{t}\|\|\nabla\tilde{w}\|
18Iup2w~t2+Cϵ2Iup2w~2,\displaystyle\leq\frac{1}{8}I_{u}^{\frac{p}{2}}\|\tilde{w}_{t}\|^{2}+C\epsilon^{2}I_{u}^{\frac{p}{2}}\|\nabla\tilde{w}\|^{2},

where we have used the fact |f(u)|C(|u|+|u|3)|f(u)|\leq C(|u|+|u|^{3}). By (3.1) and Hölder’s inequality, we have

|(Aβ2f(u),w~t)|Cf(u)Hβw~t\displaystyle|(A^{\frac{\beta}{2}}f_{*}(u),\tilde{w}_{t})|\leq C\|f_{*}(u)\|_{H^{\beta}}\|\tilde{w}_{t}\| 18u2w~t2+Cu2(1+u142)\displaystyle\leq\frac{1}{8}\|\nabla u\|^{2}\|\tilde{w}_{t}\|^{2}+C\|\nabla u\|^{2}\left(1+\|u\|_{14}^{2}\right)
18Iup2w~t2+CIup2(1+u142).\displaystyle\leq\frac{1}{8}I_{u}^{\frac{p}{2}}\|\tilde{w}_{t}\|^{2}+CI_{u}^{\frac{p}{2}}(1+\|u\|_{14}^{2}).

The last two terms can be estimated similarly since they all contain higher-power factors Iuq2I_{u}^{\frac{q}{2}} with qpq\geq p. As a result, for ϵ\epsilon small enough we conclude

ddt(12(w~t2+w~2+λw~2)+ϵIup2(w~t,w~))+2ϵ3Iup2(12(w~t2+w~2+λw~2)\displaystyle\frac{d}{dt}\Big{(}\frac{1}{2}(\|\tilde{w}_{t}\|^{2}+\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(\tilde{w}_{t},\tilde{w})\Big{)}+\frac{2\epsilon}{3}I_{u}^{\frac{p}{2}}\Big{(}\frac{1}{2}(\|\tilde{w}_{t}\|^{2}+\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2})
+ϵIup2(w~t,w~))CIup2(1+u142).\displaystyle\qquad+\epsilon I_{u}^{\frac{p}{2}}(\tilde{w}_{t},\tilde{w})\Big{)}\leq CI_{u}^{\frac{p}{2}}(1+\|u\|_{14}^{2}). (3.6)

Before applying Lemma 3.2 to (3.6), we still need to verify that

sups[t,t+1]Iu(s)C2infs[t,t+1]Iu(s)\displaystyle\sup_{s\in[t,t+1]}I_{u}(s)\leq C_{2}\inf_{s\in[t,t+1]}I_{u}(s) (3.7)

for any t0t\geq 0. To do this, note that

Iu(t)+2Iu,put2+2(f(u),ut)=0,\displaystyle I_{u}^{\prime}(t)+2I_{u,p}\|u_{t}\|^{2}+2(f(u),u_{t})=0,

which yields that

C(1+Iu)C(Iup2+|(f(u),ut)|Iu)IuIu\displaystyle-C(1+I_{u})\leq-C\Big{(}I_{u}^{\frac{p}{2}}+\frac{|(f(u),u_{t})|}{I_{u}}\Big{)}\leq\frac{I_{u}^{\prime}}{I_{u}} 2|(f(u),ut)|IuC(1+Iu).\displaystyle\leq\frac{2|(f(u),u_{t})|}{I_{u}}\leq C(1+I_{u}).

Thanks to the dissipation of uu, we know |IuIu|CB0\Big{|}\frac{I_{u}^{\prime}}{I_{u}}\Big{|}\leq C_{B_{0}}, i.e., |(lnIu)|CB0|(\ln I_{u})^{\prime}|\leq C_{B_{0}}. Therefore,

eC(ts)Iu(t)Iu(s)eC(ts)forany0st,\displaystyle e^{-C(t-s)}\leq\frac{I_{u}(t)}{I_{u}(s)}\leq e^{C(t-s)}\quad\mathrm{for~{}any~{}}0\leq s\leq t,

which implies exactly (3.7).

Now, by applying Lemma 3.2 to (3.6) and taking ϵ\epsilon small enough such that

12(w~t2+w~2+λw~2)+ϵIup2(w~t,w~)c(w~t2+w~2),\displaystyle\frac{1}{2}(\|\tilde{w}_{t}\|^{2}+\|\nabla\tilde{w}\|^{2}+\lambda\|\tilde{w}\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(\tilde{w}_{t},\tilde{w})\geq c(\|\tilde{w}_{t}\|^{2}+\|\nabla\tilde{w}\|^{2}),

we obtain the desired result. ∎

From Lemma (2.7) and Lemma (3.3), we know there exists R>0R>0 such that the set

{(u,v)H1+β(Ω)×Hβ(Ω)|uH1+β2+vHβ2R2}\big{\{}(u,v)\in H^{1+\beta}(\Omega)\times H^{\beta}(\Omega)\big{|}\|u\|_{H^{1+\beta}}^{2}+\|v\|_{H^{\beta}}^{2}\leq R^{2}\big{\}}

is attracting in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega). Therefore,

Theorem 3.4.

The global attractor 𝒜\mathscr{A} is bounded in Hβ+1(Ω)×Hβ(Ω)H^{\beta+1}(\Omega)\times H^{\beta}(\Omega) with β=27\beta=\frac{2}{7}.

Corollary 3.5.

The solutions in 𝒜\mathscr{A} are uniformly Hölder continuous in tt, i.e. there exist C>0C>0 and θ>0\theta>0, such that, for (u,ut)𝒜(u,u_{t})\in\mathscr{A}

(u(t1),ut(t1))(u(t2),ut(t2))H01(Ω)×L2(Ω)C|t1t2|θ,t1,t2,\|(u(t_{1}),u_{t}(t_{1}))-(u(t_{2}),u_{t}(t_{2}))\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\leq C|t_{1}-t_{2}|^{\theta},\quad\forall\,t_{1},t_{2}\in\mathbb{R},

where CC and θ\theta are independent of uu.

Proof.

By virtue of the interpolation inequality, we infer that

u(t1)u(t2)H01\displaystyle\|u(t_{1})-u(t_{2})\|_{H^{1}_{0}} Cu(t1)u(t2)θu(t1)u(t2)H1+β1θC𝒜t1t2ut(s)𝑑sθ\displaystyle\leq C\|u(t_{1})-u(t_{2})\|^{\theta}\|u(t_{1})-u(t_{2})\|^{1-\theta}_{H^{1+\beta}}\leq C_{\mathscr{A}}\Big{\|}\int_{t_{1}}^{t_{2}}u_{t}(s)ds\Big{\|}^{\theta}
C𝒜|t2t1|θ,\displaystyle\leq C_{\mathscr{A}}|t_{2}-t_{1}|^{\theta},

where θ=β1+β\theta=\frac{\beta}{1+\beta}, β=27\beta=\frac{2}{7}, and similarly

ut(t1)ut(t2)\displaystyle\|u_{t}(t_{1})-u_{t}(t_{2})\| Cut(t1)ut(t2)H1θut(t1)ut(t2)Hβ1θC𝒜t1t2utt(s)𝑑sH1θ\displaystyle\leq C\|u_{t}(t_{1})-u_{t}(t_{2})\|_{H^{-1}}^{\theta}\|u_{t}(t_{1})-u_{t}(t_{2})\|_{H^{\beta}}^{1-\theta}\leq C_{\mathscr{A}}\Big{\|}\int_{t_{1}}^{t_{2}}u_{tt}(s)ds\Big{\|}^{\theta}_{H^{-1}}
C𝒜|t2t1|θ.\displaystyle\leq C_{\mathscr{A}}|t_{2}-t_{1}|^{\theta}.

Lemma 3.6.

There exists r0>0r_{0}>0 such that any (u,ut)(u,u_{t}) with initial data (u0,u1)B(0,r0)H01(Ω)×L2(Ω)(u_{0},u_{1})\in B(0,r_{0})\subset H_{0}^{1}(\Omega)\times L^{2}(\Omega) converges to zero at a polynomial rate as follows:

C1(t+k1Iu(0)p2)1p(u,ut)H01(Ω)×L2(Ω)C2(t+k2Iu(0)p2)1p,C_{1}(t+k_{1}I_{u}(0)^{-\frac{p}{2}})^{-\frac{1}{p}}\leq\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\leq C_{2}(t+k_{2}I_{u}(0)^{-\frac{p}{2}})^{-\frac{1}{p}}, (3.8)

where Iu=u2+ut2I_{u}=\|\nabla u\|^{2}+\|u_{t}\|^{2} and Ci,ki>0C_{i},k_{i}>0 are independent of uu and tt.

Proof.

Part I. Firstly we prove the left inequality by the continuity argument. Since this process will be used several times in the paper, we present here the details.

Multiplying equation (1.1) by utu_{t} and integrating over Ω\Omega, we have

ddtE(t)+Iu,put2=0,\frac{d}{dt}E(t)+I_{u,p}\|u_{t}\|^{2}=0, (3.9)

where E(t)E(t) is defined in Lemma 2.2. Setting

F(s)F(s)12f(0)s2=0s0τ0ξf′′(ζ)𝑑ζ𝑑ξ𝑑τ,F_{*}(s)\triangleq F(s)-\frac{1}{2}f^{\prime}(0)s^{2}=\int_{0}^{s}\int_{0}^{\tau}\int_{0}^{\xi}f^{\prime\prime}(\zeta)d\zeta d\xi d\tau,

we know from the assumption on ff that |F(s)|C(|s|3+|s|4)|F_{*}(s)|\leq C(|s|^{3}+|s|^{4}) for ss\in\mathbb{R} and thereby

|ΩF(u)𝑑x|C(u3+u4).\left|\int_{\Omega}F_{*}(u)dx\right|\leq C(\|\nabla u\|^{3}+\|\nabla u\|^{4}).

Furthermore, since f(0)>λ1f^{\prime}(0)>-\lambda_{1} and

E=12ut2+12u2+f(0)2u2+ΩF(u)𝑑x,E=\frac{1}{2}\|u_{t}\|^{2}+\frac{1}{2}\|\nabla u\|^{2}+\frac{f^{\prime}(0)}{2}\|u\|^{2}+\int_{\Omega}F_{*}(u)dx,

there exists ϵ1>0\epsilon_{1}>0 such that if uϵ1\|\nabla u\|\leq\epsilon_{1}, we can find uniform constants l2>l1>0l_{2}>l_{1}>0 satisfying

l1IuEl2Iu.l_{1}I_{u}\leq E\leq l_{2}I_{u}. (3.10)

Let r0=l1l2ϵ12r_{0}=\sqrt{\frac{l_{1}}{l_{2}}}\frac{\epsilon_{1}}{2} and (u0,u1)B(0,r0)H01(Ω)×L2(Ω)(u_{0},u_{1})\in B(0,r_{0})\subset H^{1}_{0}(\Omega)\times L^{2}(\Omega), i.e. Iu(0)<r02=l1ϵ124l2I_{u}(0)<r_{0}^{2}=\frac{l_{1}\epsilon_{1}^{2}}{4l_{2}}. We claim that u(t)\|\nabla u(t)\| keeps smaller than ϵ1\epsilon_{1} for all t0t\geq 0. In fact, since uC([0,);H01(Ω))u\in C([0,\infty);H^{1}_{0}(\Omega)) and u0<r0<ϵ12\|\nabla u_{0}\|<r_{0}<\frac{\epsilon_{1}}{2}, there exists a maximal interval [0,T)[0,T) on which u<ϵ1\|\nabla u\|<\epsilon_{1} and (3.10) holds. Since E(t)E(t) is monotonically decreasing in tt, we have for s[0,T)s\in[0,T) E(s)E(0)l2Iu(0)=l1ϵ124E(s)\leq E(0)\leq l_{2}I_{u}(0)=\frac{l_{1}\epsilon_{1}^{2}}{4} and further Iu(s)l11E(s)ϵ124I_{u}(s)\leq l_{1}^{-1}E(s)\leq\frac{\epsilon_{1}^{2}}{4}. Therefore,

u(s)Iu(s)12ϵ12fors[0,T).\|\nabla u(s)\|\leq I_{u}(s)^{\frac{1}{2}}\leq\frac{\epsilon_{1}}{2}\quad\mathrm{for}~{}s\in[0,T).

By the continuity argument, we conclude that T=T=\infty, i.e., the claim is true, and (3.10) holds for all t0t\geq 0.

As a result, we infer from (3.9) and (3.10) that

0ddtE(t)+Iu,p(t)Iu(t)ddtE(t)+CpIu(t)p2+1ddtE(t)+Cpl1p21Ep2+1(t),0\leq\frac{d}{dt}E(t)+I_{u,p}(t)I_{u}(t)\leq\frac{d}{dt}E(t)+C_{p}I_{u}(t)^{\frac{p}{2}+1}\leq\frac{d}{dt}E(t)+C_{p}l_{1}^{-\frac{p}{2}-1}E^{\frac{p}{2}+1}(t),

and thereby, for t0t\geq 0

u(t)2+ut(t)2l21E(t)l21(E(0)p2+Ct)2pl21([l1Iu(0)]p2+Ct)2p.\|\nabla u(t)\|^{2}+\|u_{t}(t)\|^{2}\geq l_{2}^{-1}E(t)\geq l_{2}^{-1}\left(E(0)^{-\frac{p}{2}}+Ct\right)^{-\frac{2}{p}}\geq l_{2}^{-1}\Big{(}[l_{1}I_{u}(0)]^{-\frac{p}{2}}+Ct\Big{)}^{-\frac{2}{p}}.

This yields the first inequality by choosing C1,k1C_{1},k_{1} properly.

Part II. Now we prove the right one. Multiplying equation (1.1) by uu and integrating over Ω\Omega, we have

ddt(ut,u)ut2+u2+(f(u),u)+(up+utp)(ut,u)=0.\frac{d}{dt}(u_{t},u)-\|u_{t}\|^{2}+\|\nabla u\|^{2}+(f(u),u)+(\|\nabla u\|^{p}+\|u_{t}\|^{p})(u_{t},u)=0.

Note that

f(s)s=f(0)s2+0s0τf′′(ζ)𝑑ζ𝑑τs.\displaystyle f(s)s=f^{\prime}(0)s^{2}+\int_{0}^{s}\int_{0}^{\tau}f^{\prime\prime}(\zeta)d\zeta d\tau\cdot s.

Using the conditions on ff and the fact that IuI_{u} keeps always small by the continuity argument in Part I, one can take r0r_{0} small enough to obtain

ddt(ut,u)+Cu22ut2.\frac{d}{dt}(u_{t},u)+C\|\nabla u\|^{2}\leq 2\|u_{t}\|^{2}. (3.11)

Combining (3.9) and (3.10), we also have

ddtE(t)+CpE(t)p2ut20.\frac{d}{dt}E(t)+C_{p}E(t)^{\frac{p}{2}}\|u_{t}\|^{2}\leq 0. (3.12)

Now choose ϵ(0,Cp4)\epsilon\in(0,\frac{C_{p}}{4}). Multiplying (3.11) by ϵE(t)p2\epsilon E(t)^{\frac{p}{2}} and summing it with (3.12), we get

ddtE(t)+CpE(t)p2ut2+ϵE(t)p2ddt(ut,u)+ϵCE(t)p2u2\displaystyle\frac{d}{dt}E(t)+C_{p}E(t)^{\frac{p}{2}}\|u_{t}\|^{2}+\epsilon E(t)^{\frac{p}{2}}\frac{d}{dt}(u_{t},u)+\epsilon CE(t)^{\frac{p}{2}}\|\nabla u\|^{2} 2ϵE(t)p2ut2,\displaystyle\leq 2\epsilon E(t)^{\frac{p}{2}}\|u_{t}\|^{2},
(1ϵp2E(t)p21(ut,u))ddtE(t)+ϵddt[E(t)p2(ut,u)]+Cp2\displaystyle\left(1-\frac{\epsilon p}{2}E(t)^{\frac{p}{2}-1}(u_{t},u)\right)\frac{d}{dt}E(t)+\epsilon\frac{d}{dt}\left[E(t)^{\frac{p}{2}}(u_{t},u)\right]+\frac{C_{p}}{2} E(t)p2ut2\displaystyle E(t)^{\frac{p}{2}}\|u_{t}\|^{2}
+ϵC\displaystyle+\epsilon C E(t)p2u20.\displaystyle E(t)^{\frac{p}{2}}\|\nabla u\|^{2}\leq 0.

Since E(t)E(t) is decreasing, we deduce that for r0r_{0} small enough

(1ϵp2E(t)p21(ut,u))ddtE(t)2ddtE(t),\displaystyle\left(1-\frac{\epsilon p}{2}E(t)^{\frac{p}{2}-1}(u_{t},u)\right)\frac{d}{dt}E(t)\geq 2\frac{d}{dt}E(t),

and therefore

2ddt[E(t)+ϵ2E(t)p2(ut,u)]+ϵCE(t)p2(u2+ut2)0.2\frac{d}{dt}\left[E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u)\right]+\epsilon CE(t)^{\frac{p}{2}}(\|\nabla u\|^{2}+\|u_{t}\|^{2})\leq 0.

Noticing that the quantities IuI_{u}, E(t)E(t) and E(t)+ϵ2E(t)p2(ut,u)E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u) are equivalent, we end up with

ddt[E(t)+ϵ2E(t)p2(ut,u)]+C[E(t)+ϵ2E(t)p2(ut,u)]p2+10.\frac{d}{dt}\left[E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u)\right]+C\left[E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u)\right]^{\frac{p}{2}+1}\leq 0.

Therefore, by choosing properly C2C_{2} and k2k_{2}, the desired inequality follows. ∎

Remark 3.7.

Lemma 3.6 can be partially strengthened as follows:

For each r>0r>0, there exist C1,k1C_{1},k_{1}, depending on rr, such that any (u,ut)(u,u_{t}) with initial data (u0,u1)B(0,r)H01(Ω)×L2(Ω)(u_{0},u_{1})\in B(0,r)\subset H_{0}^{1}(\Omega)\times L^{2}(\Omega) satisfies

(u,ut)H01(Ω)×L2(Ω)C1(t+k1Iu(0)p2)1p.\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\geq C_{1}(t+k_{1}I_{u}(0)^{-\frac{p}{2}})^{-\frac{1}{p}}.

To do this, it is sufficient to choose C1C_{1} smaller such that C1k11pC_{1}\leq k_{1}^{\frac{1}{p}}. In fact, if (u0,u1)H01(Ω)×L2(Ω)>r0\|(u_{0},u_{1})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}>r_{0} and (u,ut)H01(Ω)×L2(Ω)\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)} reaches r0r_{0} for the first time at t0t_{0}, we know from Lemma 3.6 that for this smaller C1C_{1} and all tt0t\geq t_{0}

(u,ut)H01(Ω)×L2(Ω)\displaystyle\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)} C1(tt0+k1r0p)1pC1(t+k1r0prpIu(0)p2)1p.\displaystyle\geq C_{1}(t-t_{0}+k_{1}r_{0}^{-p})^{-\frac{1}{p}}\geq C_{1}(t+k_{1}r_{0}^{-p}r^{p}I_{u}(0)^{-\frac{p}{2}})^{-\frac{1}{p}}.

On the other hand, since (u,ut)H01(Ω)×L2(Ω)r0\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\geq r_{0} for t[0,t0]t\in[0,t_{0}], it also holds

(u,ut)H01(Ω)×L2(Ω)r0C1(k1r0p)1pC1(t+k1r0prpIu(0)p2)1p,\displaystyle\|(u,u_{t})\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\geq r_{0}\geq C_{1}(k_{1}r_{0}^{-p})^{-\frac{1}{p}}\geq C_{1}(t+k_{1}r_{0}^{-p}r^{p}I_{u}(0)^{-\frac{p}{2}})^{-\frac{1}{p}},

where C1k11pC_{1}\leq k_{1}^{\frac{1}{p}} is used. This gives the desired estimate by replacing k1k_{1} with k1r0prpk_{1}r_{0}^{-p}r^{p}.

4 Dimension estimates

Definition 4.1.

Let XX be a metric space and KXK\subset X. Let N(K,ϵ)N(K,\epsilon) denote the minimum number of closed balls of radius ϵ\epsilon with centres in KK required to cover KK. The fractal dimension (or upper box-counting dimension) of KK is defined by

dB(K)=lim supϵ0lnN(K,ϵ)lnϵ.d_{B}(K)=\limsup_{\epsilon\to 0}\frac{\ln N(K,\epsilon)}{-\ln\epsilon}.

Usually it is simpler to calculate the fractal dimension by taking the (superior) limit through a discrete sequence {ϵk}\{\epsilon_{k}\}, rather than a continuous one, as in the following lemma.

Lemma 4.2 ([20]).

If {ϵk}\{\epsilon_{k}\} is a decreasing sequence tending to zero with ϵk+1αϵk\epsilon_{k+1}\geq\alpha\epsilon_{k} for some α(0,1)\alpha\in(0,1), then

dB(K)=lim supklnN(K,ϵk)lnϵk.d_{B}(K)=\limsup_{k\to\infty}\frac{\ln N(K,\epsilon_{k})}{-\ln\epsilon_{k}}.

Consider the following degenerate evolutionary problem

ut+Au=0u_{t}+Au=0 (4.1)

on a Hilbert space XX, where AA is a nonlinear unbounded operator and is degenerate at the origin. Suppose that this abstract problem possesses a global attractor 𝒜\mathscr{A}. We would like to figure out the fractal dimension of 𝒜\mathscr{A}. However, it is not able to apply directly the method of [20] in this setting because the derivative of S(t)S(t) may not be a compact perturbation of a contractive operator near the origin where the degeneration occurs. By the similar reason, the quasi-stability method does not work either, since the quasi-stability inequality is violated near the origin.

Observing that AA is non-degenerate outside of a neighborhood of the origin, we embrace tentatively the belief that the properties of 𝒜\mathscr{A} away from the origin are good and would like to deal with the degenerate region primarily.

Theorem 4.3.

Let 𝒜\mathscr{A} be the global attractor of the dynamic system S(t)S(t) on a Hilbert space (X,)(X,\|\cdot\|). Assume that 0𝒜0\in\mathscr{A} and the fractal dimension of 𝒜\B(0,ϵ)\mathscr{A}\backslash B(0,\epsilon) is finite for each ϵ>0\epsilon>0. Let {ϵm}m=0\{\epsilon_{m}\}_{m=0}^{\infty} be a decreasing sequence and {tm}m=0\{t_{m}\}_{m=0}^{\infty} an increasing sequence with t0=0t_{0}=0, such that

ϵm0,ϵm+1αϵmfor some α(0,1),\displaystyle\epsilon_{m}\rightarrow 0,~{}~{}\epsilon_{m+1}\geq\alpha\epsilon_{m}\quad\textrm{for some }\alpha\in(0,1), (4.2)
supu𝒜B(0,ϵ0)¯S(t)uϵmfor ttm\displaystyle\sup_{u\in\mathscr{A}\cap\overline{B(0,\epsilon_{0})}}\|S(t)u\|\leq\epsilon_{m}\quad\textrm{for }t\geq t_{m} (4.3)

and

d0=Δlim supmlntmlnϵm<.d_{0}\stackrel{{\scriptstyle\Delta}}{{=}}\limsup_{m\to\infty}\frac{\ln t_{m}}{-\ln\epsilon_{m}}<\infty. (4.4)

Suppose further that S(t)u0S(t)u_{0} is uniformly Hölder continuous on (0,)×[𝒜B(0,ϵ0)](0,\infty)\times[\mathscr{A}\cap B(0,\epsilon_{0})], i.e., there exist θ(0,1]\theta\in(0,1] and L>0L>0 such that for t1,t2>0t_{1},t_{2}>0 and u1,u2𝒜B(0,ϵ0)u_{1},u_{2}\in\mathscr{A}\cap B(0,\epsilon_{0})

S(t1)u1S(t2)u2L(|t1t2|+u1u2)θ.\|S(t_{1})u_{1}-S(t_{2})u_{2}\|\leq L(|t_{1}-t_{2}|+\|u_{1}-u_{2}\|)^{\theta}. (4.5)

Then dB(𝒜)<d_{B}(\mathscr{A})<\infty.

Proof.

Denote 𝒜ϵ0=𝒜B(0,ϵ0)\mathscr{A}_{\epsilon_{0}}=\mathscr{A}\cap B(0,\epsilon_{0}) and 𝒜ϵmc=𝒜[B(0,ϵ0)B(0,ϵm)¯]\mathscr{A}_{\epsilon_{m}}^{c}=\mathscr{A}\cap[B(0,\epsilon_{0})\setminus\overline{B(0,\epsilon_{m})}]. Without loss of generality, we suppose that ϵ1<ϵ0\epsilon_{1}<\epsilon_{0}. It is easy to verify that s=0tS(s)𝒜ϵ1c𝒜ϵmc\cup_{s=0}^{t}S(s)\mathscr{A}_{\epsilon_{1}}^{c}\supset\mathscr{A}_{\epsilon_{m}}^{c} for ttmt\geq t_{m} by (4.3). For each mm, we have

N(𝒜ϵ1c×[0,tm],2ϵm)N(𝒜ϵ1c,ϵm)×ϵm1tm,\displaystyle N(\mathscr{A}_{\epsilon_{1}}^{c}\times[0,t_{m}],\sqrt{2}\epsilon_{m})\leq N(\mathscr{A}_{\epsilon_{1}}^{c},\epsilon_{m})\times\lceil\epsilon_{m}^{-1}t_{m}\rceil,

where s\lceil s\rceil denotes the smallest integer larger than or equal to ss. This means 𝒜ϵ1c×[0,tm]\mathscr{A}_{\epsilon_{1}}^{c}\times[0,t_{m}] has a covering of no more than N(𝒜ϵ1c,ϵm)×ϵm1tmN(\mathscr{A}_{\epsilon_{1}}^{c},\epsilon_{m})\times\lceil\epsilon_{m}^{-1}t_{m}\rceil balls of radius 2ϵm\sqrt{2}\epsilon_{m}. The image of this covering under S()S(\cdot) provides a covering of s=0tmS(s)𝒜ϵ1c\cup_{s=0}^{t_{m}}S(s)\mathscr{A}_{\epsilon_{1}}^{c} by sets of diameter no larger than L(22ϵm)θL(2\sqrt{2}\epsilon_{m})^{\theta}. Each of these sets is certainly contained in a closed ball of radius 2L(22ϵm)θ2L(2\sqrt{2}\epsilon_{m})^{\theta}, i.e.,

N(s=0tmS(s)𝒜ϵ1c,2L(22ϵm)θ)N(𝒜ϵ1c,ϵm)×ϵm1tm.N(\cup_{s=0}^{t_{m}}S(s)\mathscr{A}_{\epsilon_{1}}^{c},2L(2\sqrt{2}\epsilon_{m})^{\theta})\leq N(\mathscr{A}_{\epsilon_{1}}^{c},\epsilon_{m})\times\lceil\epsilon_{m}^{-1}t_{m}\rceil.

Note that 2L(22ϵm)θϵm2L(2\sqrt{2}\epsilon_{m})^{\theta}\geq\epsilon_{m} for ϵm\epsilon_{m} small enough. In this situation, it follows that

N(𝒜ϵ0,2L(22ϵm)θ)\displaystyle N(\mathscr{A}_{\epsilon_{0}},2L(2\sqrt{2}\epsilon_{m})^{\theta}) N(𝒜ϵmc,2L(22ϵm)θ)+1N(s=0tmS(s)𝒜ϵ1c,2L(22ϵm)θ)+1\displaystyle\leq N(\mathscr{A}_{\epsilon_{m}}^{c},2L(2\sqrt{2}\epsilon_{m})^{\theta})+1\leq N(\cup_{s=0}^{t_{m}}S(s)\mathscr{A}_{\epsilon_{1}}^{c},2L(2\sqrt{2}\epsilon_{m})^{\theta})+1
N(𝒜ϵ1c,ϵm)×ϵm1tm+1,\displaystyle\leq N(\mathscr{A}_{\epsilon_{1}}^{c},\epsilon_{m})\times\lceil\epsilon_{m}^{-1}t_{m}\rceil+1,

and thereby

dB(𝒜)\displaystyle d_{B}(\mathscr{A}) dB(𝒜𝒜ϵ0)+lim supmlnN(𝒜ϵ0,2L(22ϵm)θ)ln[2L(22ϵm)θ]\displaystyle\leq d_{B}(\mathscr{A}\setminus\mathscr{A}_{\epsilon_{0}})+\limsup_{m\to\infty}\frac{\ln N(\mathscr{A}_{\epsilon_{0}},2L(2\sqrt{2}\epsilon_{m})^{\theta})}{-\ln[2L(2\sqrt{2}\epsilon_{m})^{\theta}]}
dB(𝒜𝒜ϵ0)+lim supmln[N(𝒜ϵ1c,ϵm)×ϵm1tm+1]θlnϵmln(21+3θ/2L)\displaystyle\leq d_{B}(\mathscr{A}\setminus\mathscr{A}_{\epsilon_{0}})+\limsup_{m\to\infty}\frac{\ln[N(\mathscr{A}_{\epsilon_{1}}^{c},\epsilon_{m})\times\lceil\epsilon_{m}^{-1}t_{m}\rceil+1]}{-\theta\ln\epsilon_{m}-\ln(2^{1+3\theta/2}L)}
=dB(𝒜𝒜ϵ0)+lim supmlnN(𝒜ϵ1c,ϵm)+lntmlnϵmθlnϵm\displaystyle=d_{B}(\mathscr{A}\setminus\mathscr{A}_{\epsilon_{0}})+\limsup_{m\to\infty}\frac{\ln N(\mathscr{A}_{\epsilon_{1}}^{c},\epsilon_{m})+\ln t_{m}-\ln\epsilon_{m}}{-\theta\ln\epsilon_{m}}
=dB(𝒜𝒜ϵ0)+θ1dB(𝒜ϵ1c)+θ1d0+θ1\displaystyle=d_{B}(\mathscr{A}\setminus\mathscr{A}_{\epsilon_{0}})+\theta^{-1}d_{B}(\mathscr{A}_{\epsilon_{1}}^{c})+\theta^{-1}d_{0}+\theta^{-1}
<.\displaystyle<\infty.

This completes the proof. ∎

Intuitively, conditions (4.2)-(4.4) can be replaced by the following assumption of Theorem 4.4 in the setting of problem (4.1).

Theorem 4.4.

Let 𝒜\mathscr{A} be the global attractor of the dynamic system S(t)S(t) generated by equation (4.1) on Hilbert space (X,)(X,\|\cdot\|). Assume that 0𝒜0\in\mathscr{A} and the fractal dimension of 𝒜\B(0,ϵ)\mathscr{A}\backslash B(0,\epsilon) is finite for each ϵ>0\epsilon>0. Suppose that there exist ϵ1>0\epsilon_{1}>0, C>0C>0, α>0\alpha>0, l211>0l_{2}\geq 1_{1}>0 and a functional E(u)E(u) defined on 𝒜B(0,ϵ1)\mathscr{A}\cap B(0,\epsilon_{1}) such that

l1u2E(u)l2u2for u𝒜B(0,ϵ1),\displaystyle l_{1}\|u\|^{2}\leq E(u)\leq l_{2}\|u\|^{2}\quad\textrm{for }u\in\mathscr{A}\cap B(0,\epsilon_{1}), (4.6)

and, for u(t)=S(t)u0u(t)=S(t)u_{0}

ddtE(u(t))+CE(u(t))1+α0,if u(t)𝒜B(0,ϵ1).\frac{d}{dt}E(u(t))+CE(u(t))^{1+\alpha}\leq 0,\quad\textrm{if }u(t)\in\mathscr{A}\cap B(0,\epsilon_{1}). (4.7)

Suppose further that S(t)u0S(t)u_{0} is uniformly θ\theta-Hölder continuous on (0,)×[𝒜B(0,ϵ1)](0,\infty)\times[\mathscr{A}\cap B(0,\epsilon_{1})]. Then dB(𝒜)<d_{B}(\mathscr{A})<\infty.

Proof.

Set ϵ0=l21/2l11/2ϵ1\epsilon_{0}=l_{2}^{-1/2}l_{1}^{1/2}\epsilon_{1}. It is easy to verify that u(t)u(t) stays in 𝒜B(0,ϵ1)\mathscr{A}\cap B(0,\epsilon_{1}) for all t0t\geq 0 if u0𝒜B(0,ϵ0)u_{0}\in\mathscr{A}\cap B(0,\epsilon_{0}). Therefore, it is sufficient to show there exist a decreasing sequence {ϵm}m=0\{\epsilon_{m}\}_{m=0}^{\infty} with ϵ0\epsilon_{0} given above and an increasing sequence {tm}m=0\{t_{m}\}_{m=0}^{\infty} with t0=0t_{0}=0 satisfying (4.2)-(4.4). Multiplying (4.7) by E(u)1αE(u)^{-1-\alpha} and integrating from 0 to tt, we obtain

E(u(t))α\displaystyle E(u(t))^{-\alpha} αCt+E(u(0))α,\displaystyle\geq\alpha Ct+E(u(0))^{-\alpha},
E(u(t))\displaystyle E(u(t)) (αCt+E(u(0))α)1α,\displaystyle\leq(\alpha Ct+E(u(0))^{-\alpha})^{-\frac{1}{\alpha}},

which implies that

u(t)2\displaystyle\|u(t)\|^{2} (αl1αCt+l1αl2αu(0)2α)1α.\displaystyle\leq(\alpha l_{1}^{\alpha}Ct+l_{1}^{\alpha}l_{2}^{-\alpha}\|u(0)\|^{-2\alpha})^{-\frac{1}{\alpha}}.

For m1m\geq 1, set ϵm=2mϵ0\epsilon_{m}=2^{-m}\epsilon_{0} and tm=22mαl1αl2ααl1αCϵ02αt_{m}=\frac{2^{2m\alpha}-l_{1}^{\alpha}l_{2}^{-\alpha}}{\alpha l_{1}^{\alpha}C\epsilon_{0}^{2\alpha}}. One can easily verify that (4.2) and (4.3) hold. Besides, we infer that

lim supmlntmlnϵm=lim supmln(22mαl1αl2α)ln(αl1αCϵ02α)mln2lnϵ0=2α,\limsup_{m\to\infty}\frac{\ln t_{m}}{-\ln\epsilon_{m}}=\limsup_{m\to\infty}\frac{\ln(2^{2m\alpha}-l_{1}^{\alpha}l_{2}^{-\alpha})-\ln(\alpha l_{1}^{\alpha}C\epsilon_{0}^{2\alpha})}{m\ln 2-\ln\epsilon_{0}}=2\alpha,

which implies (4.4). ∎

5 Finite dimensionality

In this section, we use Theorem 4.4 to prove the finite dimensionality of the global attractor for problem (1.1).

The formal linearization of (1.1) is given by

{UttΔU+(up+utp)Ut+f(u)U+p[up2(u,U)+utp2(ut,Ut)]ut=0,U(0)=ξ,Ut(0)=ζ.\begin{cases}U_{tt}-\Delta U+(\|\nabla u\|^{p}+\|u_{t}\|^{p})U_{t}+f^{\prime}(u)U\\ \qquad\qquad\qquad+p\left[\|\nabla u\|^{p-2}(\nabla u,\nabla U)+\|u_{t}\|^{p-2}(u_{t},U_{t})\right]u_{t}=0,\\ U(0)=\xi,U_{t}(0)=\zeta.\end{cases} (5.1)
Lemma 5.1.

Suppose Assumption 1.1 holds. For t>0t>0, S(t)S(t) is Fréchet differentiable on H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega). Its derivative at ω0=(u0,u1)\omega_{0}=(u_{0},u_{1}) is the linear operator

L(t;ω0):(ξ,ζ)(U(t),Ut(t)),L(t;\omega_{0}):(\xi,\zeta)\mapsto(U(t),U_{t}(t)),

where UU is the solution of (5.1). Furthermore, for each t>0t>0, L(t;ω0)L(t;\omega_{0}) is continuous in ω0\omega_{0}.

Proof.

Similarly to the proof of Theorem 2.2, one may show the existence and uniqueness of (5.1) in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega) by the monotone operator theory. We omit the details and prove only DS(t)ω0=L(t;ω0)DS(t)\omega_{0}=L(t;\omega_{0}).

Denote X=H01(Ω)×L2(Ω)X=H_{0}^{1}(\Omega)\times L^{2}(\Omega). Let ω0=(u0,u1)X\omega_{0}=(u_{0},u_{1})\in X and ω~0=ω0+(ξ,ζ)X\tilde{\omega}_{0}=\omega_{0}+(\xi,\zeta)\in X. For t0t\geq 0, denote ω(t)=S(t)ω0=(u(t),ut(t))\omega(t)=S(t)\omega_{0}=(u(t),u_{t}(t)) and ω~(t)=S(t)ω~0=(u~(t),u~t(t))\tilde{\omega}(t)=S(t)\tilde{\omega}_{0}=(\tilde{u}(t),\tilde{u}_{t}(t)), both of which are uniformly bounded in XX due to the dissipation, with the bound depending on ω0\|\omega_{0}\| and ω~0\|\tilde{\omega}_{0}\|.

We first show S(t)S(t) is locally Lipschitz in XX for each t0t\geq 0. Indeed, the difference ψ=u~u\psi=\tilde{u}-u satisfies

{ψttΔψ+(u~p+u~tp)u~t(up+utp)ut+f(u~)f(u)=0,ψ(0)=ξ,ψt(0)=ζ.\displaystyle\begin{cases}\psi_{tt}-\Delta\psi+(\|\nabla\tilde{u}\|^{p}+\|\tilde{u}_{t}\|^{p})\tilde{u}_{t}-(\|\nabla u\|^{p}+\|u_{t}\|^{p})u_{t}+f(\tilde{u})-f(u)=0,\\ \psi(0)=\xi,\psi_{t}(0)=\zeta.\end{cases} (5.2)

Multiplying (5.2) by ψt\psi_{t} and integrating over Ω\Omega, we have

12ddtψt2+12ddtψ2+(u~tpu~tutput,ψt)+u~pψt2\displaystyle\quad\frac{1}{2}\frac{d}{dt}\|\psi_{t}\|^{2}+\frac{1}{2}\frac{d}{dt}\|\nabla\psi\|^{2}+(\|\tilde{u}_{t}\|^{p}\tilde{u}_{t}-\|u_{t}\|^{p}u_{t},\psi_{t})+\|\nabla\tilde{u}\|^{p}\|\psi_{t}\|^{2}
=(upu~p)(ut,ψt)+(f(u)f(u~),ψt)\displaystyle=(\|\nabla u\|^{p}-\|\nabla\tilde{u}\|^{p})(u_{t},\psi_{t})+(f(u)-f(\tilde{u}),\psi_{t})
=01puθp2(uθ,ψ)𝑑θ(ut,ψt)Ω01f(uθ)ψ𝑑θψt𝑑x\displaystyle=-\int_{0}^{1}p\|\nabla u_{\theta}\|^{p-2}(\nabla u_{\theta},\nabla\psi)d\theta\cdot(u_{t},\psi_{t})-\int_{\Omega}\int_{0}^{1}f^{\prime}(u_{\theta})\psi d\theta\cdot\psi_{t}dx
C(ψ2+ψt2),\displaystyle\leq C(\|\nabla\psi\|^{2}+\|\psi_{t}\|^{2}),

where uθ=θu+(1θ)u~u_{\theta}=\theta u+(1-\theta)\tilde{u}. Noticing that (u~tpu~tutput,ψt)0(\|\tilde{u}_{t}\|^{p}\tilde{u}_{t}-\|u_{t}\|^{p}u_{t},\psi_{t})\geq 0, by means of the Gronwall lemma we have

ω~(t)ω(t)X2e2Ct(ξ,ζ)X2.\|\tilde{\omega}(t)-\omega(t)\|_{X}^{2}\leq e^{2Ct}\|(\xi,\zeta)\|_{X}^{2}. (5.3)

To continue, we introduce the operators from C([0,T];H01(Ω))C1([0,T];L2(Ω))C([0,T];H_{0}^{1}(\Omega))\cap C^{1}([0,T];L^{2}(\Omega)) to C([0,T];L2(Ω))C([0,T];L^{2}(\Omega)) by

𝒢(u)=uput,(u)=utput.\mathcal{G}(u)=\|\nabla u\|^{p}u_{t},\quad\mathcal{H}(u)=\|u_{t}\|^{p}u_{t}.

Their Fréchet derivatives are given by

𝒢(u)v\displaystyle\mathcal{G}^{\prime}(u)v =upvt+pup2(u,v)ut,\displaystyle=\|\nabla u\|^{p}v_{t}+p\|\nabla u\|^{p-2}(\nabla u,\nabla v)u_{t},
(u)v\displaystyle\mathcal{H}^{\prime}(u)v =utpvt+putp2(ut,vt)ut.\displaystyle=\|u_{t}\|^{p}v_{t}+p\|u_{t}\|^{p-2}(u_{t},v_{t})u_{t}.

For each tt, we set

𝔊\displaystyle\mathfrak{G} =Δ𝒢(u~)𝒢(u)𝒢(u)(u~u)=𝔊1+𝔊2,\displaystyle\stackrel{{\scriptstyle\Delta}}{{=}}\mathcal{G}(\tilde{u})-\mathcal{G}(u)-\mathcal{G}^{\prime}(u)(\tilde{u}-u)=\mathfrak{G}_{1}+\mathfrak{G}_{2},
\displaystyle\mathfrak{H} =Δ(u~)(u)(u)(u~u),\displaystyle\stackrel{{\scriptstyle\Delta}}{{=}}\mathcal{H}(\tilde{u})-\mathcal{H}(u)-\mathcal{H}^{\prime}(u)(\tilde{u}-u),

where

𝔊1\displaystyle\mathfrak{G}_{1} =(u~pup)utpup2(u,u~u)ut,\displaystyle=(\|\nabla\tilde{u}\|^{p}-\|\nabla u\|^{p})u_{t}-p\|\nabla u\|^{p-2}(\nabla u,\nabla\tilde{u}-\nabla u)u_{t},
𝔊2\displaystyle\mathfrak{G}_{2} =(u~pup)(u~tut).\displaystyle=(\|\nabla\tilde{u}\|^{p}-\|\nabla u\|^{p})(\tilde{u}_{t}-u_{t}).

For any ϵ>0\epsilon>0, since S(t)S(t) is Lipschitz, we have, if (ξ,ζ)(\xi,\zeta) is close to zero

𝔊1\displaystyle\|\mathfrak{G}_{1}\| =p01uθp2(uθ,u~u)𝑑θutp(up2u,u~u)ut\displaystyle=\Big{\|}p\int_{0}^{1}\|\nabla u_{\theta}\|^{p-2}(\nabla u_{\theta},\nabla\tilde{u}-\nabla u)d\theta\cdot u_{t}-p(\|\nabla u\|^{p-2}\nabla u,\nabla\tilde{u}-\nabla u)u_{t}\Big{\|}
p01uθp2uθup2u𝑑θu~uut\displaystyle\leq p\int_{0}^{1}\big{\|}\|\nabla u_{\theta}\|^{p-2}\nabla u_{\theta}-\|\nabla u\|^{p-2}\nabla u\big{\|}d\theta\cdot\|\nabla\tilde{u}-\nabla u\|\|u_{t}\|
ϵ2u~u\displaystyle\leq\frac{\epsilon}{2}\|\nabla\tilde{u}-\nabla u\|

and

𝔊2ϵ2u~tut.\|\mathfrak{G}_{2}\|\leq\frac{\epsilon}{2}\|\tilde{u}_{t}-u_{t}\|.

Therefore, we infer that

𝔊ϵω~ωX,\|\mathfrak{G}\|\leq\epsilon\|\tilde{\omega}-\omega\|_{X}, (5.4)

and following a similar argument,

ϵω~ωX.\|\mathfrak{H}\|\leq\epsilon\|\tilde{\omega}-\omega\|_{X}. (5.5)

Now, denote Ψ=u~uU\Psi=\tilde{u}-u-U with UU the solution of (5.1). By calculation we know that Ψ\Psi satisfies

ΨttΔΨ\displaystyle\Psi_{tt}-\Delta\Psi +𝒢(u~)𝒢(u)𝒢(u)U+(u~)(u)(u)U\displaystyle+\mathcal{G}(\tilde{u})-\mathcal{G}(u)-\mathcal{G}^{\prime}(u)U+\mathcal{H}(\tilde{u})-\mathcal{H}(u)-\mathcal{H}^{\prime}(u)U
+f(u~)f(u)f(u)U=0.\displaystyle+f(\tilde{u})-f(u)-f^{\prime}(u)U=0.

Denoting 𝔉=f(u~)f(u)f(u)(u~u)\mathfrak{F}=f(\tilde{u})-f(u)-f^{\prime}(u)(\tilde{u}-u), we can rewrite the above equation into

ΨttΔΨ+𝒢(u)Ψ+(u)Ψ+f(u)Ψ+𝔊++𝔉=0.\displaystyle\Psi_{tt}-\Delta\Psi+\mathcal{G}^{\prime}(u)\Psi+\mathcal{H}^{\prime}(u)\Psi+f^{\prime}(u)\Psi+\mathfrak{G}+\mathfrak{H}+\mathfrak{F}=0. (5.6)

Noticing that

𝔉(t)=01[f(θu~+(1θ)u)f(u)](u~u)𝑑θ=0101f′′(u+τθ(u~u))θψ2𝑑τ𝑑θ,\mathfrak{F}(t)=\int_{0}^{1}[f^{\prime}(\theta\tilde{u}+(1-\theta)u)-f^{\prime}(u)](\tilde{u}-u)d\theta=\int_{0}^{1}\int_{0}^{1}f^{\prime\prime}(u+\tau\theta(\tilde{u}-u))\theta\psi^{2}d\tau d\theta,

it holds

|(𝔉(t),Ψt)|C(1+u+u~)6ψ23ΨtCψ2Ψt.|(\mathfrak{F}(t),\Psi_{t})|\leq C\|(1+u+\tilde{u})\|_{6}\|\psi^{2}\|_{3}\|\Psi_{t}\|\leq C\|\nabla\psi\|^{2}\|\Psi_{t}\|.

Multiplying (5.6) by Ψt\Psi_{t} and integrating over Ω\Omega, we obtain

12ddt(Ψt2+Ψ2)\displaystyle\frac{1}{2}\frac{d}{dt}(\|\Psi_{t}\|^{2}+\|\nabla\Psi\|^{2}) =([𝒢(u)+(u)+f(u)]Ψ,Ψt)(𝔊++𝔉,Ψt)\displaystyle=-([\mathcal{G}^{\prime}(u)+\mathcal{H}^{\prime}(u)+f^{\prime}(u)]\Psi,\Psi_{t})-(\mathfrak{G}+\mathfrak{H}+\mathfrak{F},\Psi_{t})
C(Ψt2+Ψ2)+(𝔊+)Ψt+Cψ2Ψt.\displaystyle\leq C(\|\Psi_{t}\|^{2}+\|\nabla\Psi\|^{2})+(\|\mathfrak{G}\|+\|\mathfrak{H}\|)\|\Psi_{t}\|+C\|\nabla\psi\|^{2}\|\Psi_{t}\|.

Let (ξ,ζ)(\xi,\zeta) be close enough to zero. Then, by (5.3)–(5.5), we have

12ddt(Ψt2+Ψ2)\displaystyle\frac{1}{2}\frac{d}{dt}(\|\Psi_{t}\|^{2}+\|\nabla\Psi\|^{2}) C(Ψt2+Ψ2)+(2ϵω~ωX+Cψ2)Ψt\displaystyle\leq C(\|\Psi_{t}\|^{2}+\|\nabla\Psi\|^{2})+(2\epsilon\|\tilde{\omega}-\omega\|_{X}+C\|\nabla\psi\|^{2})\|\Psi_{t}\|
C(Ψt2+Ψ2)+3ϵω~ωXΨt\displaystyle\leq C(\|\Psi_{t}\|^{2}+\|\nabla\Psi\|^{2})+3\epsilon\|\tilde{\omega}-\omega\|_{X}\|\Psi_{t}\|
(C+1)(Ψt2+Ψ2)+94ϵ2ω~ωX2.\displaystyle\leq(C+1)(\|\Psi_{t}\|^{2}+\|\nabla\Psi\|^{2})+\frac{9}{4}\epsilon^{2}\|\tilde{\omega}-\omega\|_{X}^{2}.

Therefore, it follows from the Gronwall lemma, as well as (5.3), that

ΨX294ϵ2eCt(ξ,ζ)X2,\|\Psi\|_{X}^{2}\leq\frac{9}{4}\epsilon^{2}e^{Ct}\|(\xi,\zeta)\|_{X}^{2},

since Ψ(0)=0\Psi(0)=0. In other words,

ω~(t)ω(t)(U(t),Ut(t))X2(ξ,ζ)X294ϵ2eCtfor (ξ,ζ) close enough to 0 in X.\frac{\|\tilde{\omega}(t)-\omega(t)-(U(t),U_{t}(t))\|_{X}^{2}}{\|(\xi,\zeta)\|_{X}^{2}}\leq\frac{9}{4}\epsilon^{2}e^{Ct}\quad\textrm{for }(\xi,\zeta)\textrm{ close enough to }0\textrm{ in }X.

Since ϵ\epsilon could be arbitrarily small, we obtain the differentiability of S(t)S(t). ∎

Lemma 5.2.

Suppose that ff satisfies Assumption 1.1 and (1.4). Then there exists r0>0r_{0}>0 such that S(t)ω0S(t)\omega_{0} is uniformly (with respect to tt) Lipschitz continuous on B(0,r0)H01(Ω)×L2(Ω)B(0,r_{0})\subset H^{1}_{0}(\Omega)\times L^{2}(\Omega).

Proof.

Let (ξ,ζ)X1\|(\xi,\zeta)\|_{X}\leq 1 and denote f(s)=f(s)f(0)sf_{*}(s)=f(s)-f^{\prime}(0)s. We multiply equation (5.1) by UtU_{t} and integrate over Ω\Omega to get

12ddt[Ut2+U2+f(0)U2]+(up+utp)Ut2+pup2(u,U)(ut,Ut)+putp2|(ut,Ut)|2+(f(u)U,Ut)=0.\displaystyle\begin{split}&\quad\frac{1}{2}\frac{d}{dt}[\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2}]+(\|\nabla u\|^{p}+\|u_{t}\|^{p})\|U_{t}\|^{2}\\ &+p\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U_{t})+p\|u_{t}\|^{p-2}|(u_{t},U_{t})|^{2}+(f_{*}^{\prime}(u)U,U_{t})=0.\end{split} (5.7)

And similarly, by the test function UU, it holds

ddt(Ut,U)Ut2+U2+f(0)U2+(up+utp)(Ut,U)+pup2(u,U)(ut,U)+putp2(ut,Ut)(ut,U)+(f(u)U,U)=0.\displaystyle\begin{split}&\quad\frac{d}{dt}(U_{t},U)-\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2}+(\|\nabla u\|^{p}+\|u_{t}\|^{p})(U_{t},U)\\ &+p\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U)+p\|u_{t}\|^{p-2}(u_{t},U_{t})(u_{t},U)+(f_{*}^{\prime}(u)U,U)=0.\end{split} (5.8)

Now, taking ϵ>0\epsilon>0, we multiply (5.8) with ϵIup2(t)\epsilon I_{u}^{\frac{p}{2}}(t) and sum it with (5.7) to obtain

12ddt(Ut2+U2+f(0)U2)+ϵIup2ddt(Ut,U)+putp2|(ut,Ut)|2+Iu,pUt2+ϵIup2(U2+f(0)U2)=pup2(u,U)(ut,Ut)(f(u)U,Ut)+ϵIup2Ut2ϵIup2Iu,p(Ut,U)ϵpIup2up2(u,U)(ut,U)ϵpIup2utp2(ut,Ut)(ut,U)ϵIup2(f(u)U,U),\displaystyle\begin{split}&\quad\frac{1}{2}\frac{d}{dt}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon I_{u}^{\frac{p}{2}}\frac{d}{dt}(U_{t},U)+p\|u_{t}\|^{p-2}|(u_{t},U_{t})|^{2}\\ &+I_{u,p}\|U_{t}\|^{2}+\epsilon I_{u}^{\frac{p}{2}}(\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})=-p\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U_{t})\\ &-(f_{*}^{\prime}(u)U,U_{t})+\epsilon I_{u}^{\frac{p}{2}}\|U_{t}\|^{2}-\epsilon I_{u}^{\frac{p}{2}}I_{u,p}(U_{t},U)-\epsilon pI_{u}^{\frac{p}{2}}\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U)\\ &-\epsilon pI_{u}^{\frac{p}{2}}\|u_{t}\|^{p-2}(u_{t},U_{t})(u_{t},U)-\epsilon I_{u}^{\frac{p}{2}}(f_{*}^{\prime}(u)U,U),\end{split} (5.9)

where Iu,Iu,pI_{u},I_{u,p} have been given in (1.6). Note that

ϵIup2ddt(Ut,U)\displaystyle\epsilon I_{u}^{\frac{p}{2}}\frac{d}{dt}(U_{t},U) =ϵddt[Iup2(Ut,U)]ϵp2(Ut,U)Iup21ddt(u2+ut2)\displaystyle=\epsilon\frac{d}{dt}[I_{u}^{\frac{p}{2}}(U_{t},U)]-\frac{\epsilon p}{2}(U_{t},U)I_{u}^{\frac{p}{2}-1}\frac{d}{dt}(\|\nabla u\|^{2}+\|u_{t}\|^{2})
=ϵddt[Iup2(Ut,U)]ϵp2(Ut,U)Iup21[Iu,put2(f(u),ut)]\displaystyle=\epsilon\frac{d}{dt}[I_{u}^{\frac{p}{2}}(U_{t},U)]-\frac{\epsilon p}{2}(U_{t},U)I_{u}^{\frac{p}{2}-1}\left[-I_{u,p}\|u_{t}\|^{2}-(f(u),u_{t})\right]
ϵddt[Iup2(Ut,U)]ϵC(Ut2+U2)Iup21[Iu,put2+|(f(u),ut)|],\displaystyle\geq\epsilon\frac{d}{dt}[I_{u}^{\frac{p}{2}}(U_{t},U)]-\epsilon C(\|U_{t}\|^{2}+\|\nabla U\|^{2})I_{u}^{\frac{p}{2}-1}\left[I_{u,p}\|u_{t}\|^{2}+|(f(u),u_{t})|\right],

and

|pup2(u,U)(ut,Ut)|pup1U|(ut,Ut)|p4u2p2U2ut2p+putp2|(ut,Ut)|2p4(u2+ut2)p/2U2+putp2|(ut,Ut)|2.\displaystyle\begin{split}&\quad\big{|}p\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U_{t})\big{|}\leq p\|\nabla u\|^{p-1}\|\nabla U\|\cdot|(u_{t},U_{t})|\\ &\leq\frac{p}{4}\|\nabla u\|^{2p-2}\|\nabla U\|^{2}\|u_{t}\|^{2-p}+p\|u_{t}\|^{p-2}|(u_{t},U_{t})|^{2}\\ &\leq\frac{p}{4}(\|\nabla u\|^{2}+\|u_{t}\|^{2})^{p/2}\|\nabla U\|^{2}+p\|u_{t}\|^{p-2}|(u_{t},U_{t})|^{2}.\end{split} (5.10)

We derive from (5.9) that

12ddt(Ut2+U2+f(0)U2)+ϵddt(Iup2(Ut,U))+(1ϵ)Iup2Ut2+(ϵp4)Iup2U2+ϵIup2f(0)U2(f(u)U,Ut)ϵIup2Iu,p(Ut,U)ϵpIup2up2(u,U)(ut,U)ϵpIup2utp2(ut,Ut)(ut,U)ϵIup2(f(u)U,U)+ϵC(Ut2+U2)Iup21[Iu,put2+|(f(u),ut)|],\displaystyle\begin{split}&\frac{1}{2}\frac{d}{dt}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon\frac{d}{dt}\Big{(}I_{u}^{\frac{p}{2}}(U_{t},U)\Big{)}+(1-\epsilon)I_{u}^{\frac{p}{2}}\|U_{t}\|^{2}\\ &+\big{(}\epsilon-\frac{p}{4}\big{)}I_{u}^{\frac{p}{2}}\|\nabla U\|^{2}+\epsilon I_{u}^{\frac{p}{2}}f^{\prime}(0)\|U\|^{2}\leq-(f_{*}^{\prime}(u)U,U_{t})-\epsilon I_{u}^{\frac{p}{2}}I_{u,p}(U_{t},U)\\ &-\epsilon pI_{u}^{\frac{p}{2}}\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U)-\epsilon pI_{u}^{\frac{p}{2}}\|u_{t}\|^{p-2}(u_{t},U_{t})(u_{t},U)-\epsilon I_{u}^{\frac{p}{2}}(f_{*}^{\prime}(u)U,U)\\ &\quad+\epsilon C(\|U_{t}\|^{2}+\|\nabla U\|^{2})I_{u}^{\frac{p}{2}-1}\left[I_{u,p}\|u_{t}\|^{2}+|(f(u),u_{t})|\right],\end{split} (5.11)

where we have also used the fact Iup2Iu,pI_{u}^{\frac{p}{2}}\leq I_{u,p}. By the condition f(0)>(1p4)λ1f^{\prime}(0)>-(1-\frac{p}{4})\lambda_{1}, we can choose μ(0,1p4)\mu\in(0,1-\frac{p}{4}) such that f(0)>μλ1>(1p4)λ1f^{\prime}(0)>-\mu\lambda_{1}>-(1-\frac{p}{4})\lambda_{1} and deduce from Poincaré’s inequality

(ϵp4)Iup2U2+ϵIup2f(0)U2>(ϵp4μϵ)Iup2U2.\displaystyle\big{(}\epsilon-\frac{p}{4}\big{)}I_{u}^{\frac{p}{2}}\|\nabla U\|^{2}+\epsilon I_{u}^{\frac{p}{2}}f^{\prime}(0)\|U\|^{2}>\big{(}\epsilon-\frac{p}{4}-\mu\epsilon\big{)}I_{u}^{\frac{p}{2}}\|\nabla U\|^{2}.

This allows us to find ϵ(0,1)\epsilon\in(0,1) and c>0c>0 such that

(1ϵ)Iup2Ut2+(ϵp4)Iup2U2+ϵIup2f(0)U23cIup2(Ut2+U2).\displaystyle(1-\epsilon)I_{u}^{\frac{p}{2}}\|U_{t}\|^{2}+\big{(}\epsilon-\frac{p}{4}\big{)}I_{u}^{\frac{p}{2}}\|\nabla U\|^{2}+\epsilon I_{u}^{\frac{p}{2}}f^{\prime}(0)\|U\|^{2}\geq 3cI_{u}^{\frac{p}{2}}(\|U_{t}\|^{2}+\|\nabla U\|^{2}).

Therefore, (5.11) gives

12ddt(Ut2+U2+f(0)U2)+ϵddt(Iup2(Ut,U))+3cIup2(Ut2+U2)(f(u)U,Ut)ϵIup2Iu,p(Ut,U)ϵIup2up2(u,U)(ut,U)ϵpIup2utp2(ut,Ut)(ut,U)ϵIup2(f(u)U,U)+ϵC(Ut2+U2)[Iu,put2+|(f(u),ut)|].\displaystyle\begin{split}&\frac{1}{2}\frac{d}{dt}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon\frac{d}{dt}\left(I_{u}^{\frac{p}{2}}(U_{t},U)\right)+3cI_{u}^{\frac{p}{2}}(\|U_{t}\|^{2}+\|\nabla U\|^{2})\\ &\leq-(f_{*}^{\prime}(u)U,U_{t})-\epsilon I_{u}^{\frac{p}{2}}I_{u,p}(U_{t},U)-\epsilon I_{u}^{\frac{p}{2}}\|\nabla u\|^{p-2}(\nabla u,\nabla U)(u_{t},U)\\ &\qquad-\epsilon pI_{u}^{\frac{p}{2}}\|u_{t}\|^{p-2}(u_{t},U_{t})(u_{t},U)-\epsilon I_{u}^{\frac{p}{2}}(f_{*}^{\prime}(u)U,U)\\ &\qquad\qquad+\epsilon C(\|U_{t}\|^{2}+\|\nabla U\|^{2})\left[I_{u,p}\|u_{t}\|^{2}+|(f(u),u_{t})|\right].\end{split} (5.12)

It remains to estimate the right hand side of the above inequality. From (1.4) we know for any η>0\eta>0 there exists δ>0\delta>0 such that

|f(s)|{η|s|p,|s|<δ,C(|s|+|s|2)Cδ|s|2,|s|δ.|f_{*}^{\prime}(s)|\leq\begin{cases}\eta|s|^{p},&|s|<\delta,\\ C(|s|+|s|^{2})\leq C_{\delta}|s|^{2},&|s|\geq\delta.\end{cases}

Note that u(t)\|\nabla u(t)\| and ut(t)\|u_{t}(t)\| keep small for all t0t\geq 0 if we take r0r_{0} small enough by the continuity argument as in Lemma 3.6. In this way, we can find C1>0C_{1}>0 such that if r0<C1r_{0}<C_{1}, we have

f(u)3\displaystyle\|f_{*}^{\prime}(u)\|_{3} ({|u|<δ}|f(u)|3𝑑x+{|u|δ}|f(u)|3𝑑x)13ηuL3p(Ω)p+CuL6(Ω)2\displaystyle\leq\Big{(}\int_{\{|u|<\delta\}}|f_{*}^{\prime}(u)|^{3}dx+\int_{\{|u|\geq\delta\}}|f_{*}^{\prime}(u)|^{3}dx\Big{)}^{\frac{1}{3}}\leq\eta\|u\|_{L^{3p}(\Omega)}^{p}+C\|u\|_{L^{6}(\Omega)}^{2}
Cηup+Cu2cup,\displaystyle\leq C\eta\|\nabla u\|^{p}+C\|\nabla u\|^{2}\leq c\|\nabla u\|^{p},

if we choose η\eta small enough in advance, and thereby

|(f(u)U,Ut)|f(u)3UtU6cIup2(U2+Ut2).|(f_{*}^{\prime}(u)U,U_{t})|\leq\|f_{*}^{\prime}(u)\|_{3}\|U_{t}\|\|U\|_{6}\leq cI_{u}^{\frac{p}{2}}(\|\nabla U\|^{2}+\|U_{t}\|^{2}).

For the rest terms, one can estimate them more easily since they all contain higher-power factors Iuq2I_{u}^{\frac{q}{2}} with q>pq>p. In detail, there exists C2>0C_{2}>0 such that if r0<C2r_{0}<C_{2}, the rest of the right hand side of inequality (5.12) will be smaller than cIup2(U2+Ut2)cI_{u}^{\frac{p}{2}}(\|\nabla U\|^{2}+\|U_{t}\|^{2}). Hence, we end up with

ddt(12(Ut2+U2+f(0)U2)+ϵIup2(Ut,U))+cIup2(Ut2+U2)0.\frac{d}{dt}\left(\frac{1}{2}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(U_{t},U)\right)+cI_{u}^{\frac{p}{2}}(\|U_{t}\|^{2}+\|\nabla U\|^{2})\leq 0. (5.13)

Note that 12(Ut2+U2+f(0)U2)+ϵIup2(Ut,U)\frac{1}{2}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(U_{t},U) is equivalent to Ut2+U2\|U_{t}\|^{2}+\|\nabla U\|^{2} if Iup2I_{u}^{\frac{p}{2}} is small enough. Therefore, following the continuity argument as in Lemma 3.6 again, we can find a constant C3>0C_{3}>0 such that if r0<C3r_{0}<C_{3}, then

Ut(t)2+U(t)2\displaystyle\|U_{t}(t)\|^{2}+\|\nabla U(t)\|^{2} C[12(Ut2+U2+f(0)U2)+ϵIup2(Ut,U)]|t\displaystyle\leq C\Big{[}\frac{1}{2}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(U_{t},U)\Big{]}\Big{|}_{t}
C[12(Ut2+U2+f(0)U2)+ϵIup2(Ut,U)]|t=0\displaystyle\leq C\Big{[}\frac{1}{2}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(U_{t},U)\Big{]}\Big{|}_{t=0}
C.\displaystyle\leq C.

This means that L(t;)L(t;\cdot) is uniformly bounded, i.e., L(t;)XXC\|L(t;\cdot)\|_{X\to X}\leq C for all t0t\geq 0.

As the conclusion, for every ω1,ω2B(0,r0)\omega_{1},\omega_{2}\in B(0,r_{0}) and t0t\geq 0, we obtain

S(t)ω1S(t)ω2X01L(t;θω1+(1θ)ω2)(ω1ω2)𝑑θXCω1ω2X.\|S(t)\omega_{1}-S(t)\omega_{2}\|_{X}\leq\Big{\|}\int_{0}^{1}L(t;\theta\omega_{1}+(1-\theta)\omega_{2})(\omega_{1}-\omega_{2})d\theta\Big{\|}_{X}\leq C\|\omega_{1}-\omega_{2}\|_{X}.

This completes the proof. ∎

Remark 5.3.

Inequality (5.13) implies that

ddtΛU(t)+cIup2(t)ΛU(t)0,\frac{d}{dt}\Lambda_{U}(t)+cI_{u}^{\frac{p}{2}}(t)\Lambda_{U}(t)\leq 0,

where ΛU=12(Ut2+U2+f(0)U2)+ϵIup2(Ut,U)\Lambda_{U}=\frac{1}{2}(\|U_{t}\|^{2}+\|\nabla U\|^{2}+f^{\prime}(0)\|U\|^{2})+\epsilon I_{u}^{\frac{p}{2}}(U_{t},U). By the Gronwall lemma, we get

ΛU(t)ΛU(0)exp(0tcIup2(s)ds).\Lambda_{U}(t)\leq\Lambda_{U}(0)\exp\Big{(}\int_{0}^{t}-cI_{u}^{\frac{p}{2}}(s)ds\Big{)}.

Then it follows from Lemma 3.6 that

ΛU(t)\displaystyle\Lambda_{U}(t) ΛU(0)exp(C1pln(k1(Iu(0))p/2)C1pln(t+k1(Iu(0))p/2))\displaystyle\leq\Lambda_{U}(0)\exp\left(C_{1}^{p}\ln(k_{1}(I_{u}(0))^{-p/2})-C_{1}^{p}\ln(t+k_{1}(I_{u}(0))^{-p/2})\right)
=ΛU(0)(k1(Iu(0))p/2t+k1(Iu(0))p/2)C1p,\displaystyle=\Lambda_{U}(0)\left(\frac{k_{1}(I_{u}(0))^{-p/2}}{t+k_{1}(I_{u}(0))^{-p/2}}\right)^{C_{1}^{p}},

where C1,k1C_{1},k_{1} are the constants given in Lemma 3.6. Note that ΛU1/2\Lambda_{U}^{1/2} can be regarded as an equivalent quantity of (U,Ut)H01(Ω)×L2(Ω)\|(U,U_{t})\|_{H^{1}_{0}(\Omega)\times L^{2}(\Omega)} when IuI_{u} is small enough. This tells us that, for each (u0,u1)B(0,r0)(u_{0},u_{1})\in B(0,r_{0}), the derivative L(t;u0,u1)L(t;u_{0},u_{1}) would have norm less than 1, in fact tending to zero, as long as tt is large enough. However, this is not the case if we consider the supremum of the operator norm of L(t;u0,u1)L(t;u_{0},u_{1}) over (u0,u1)B(0,r0)(u_{0},u_{1})\in B(0,r_{0}), since for fixed T>0T>0

k1(Iu(0))p/2T+k1(Iu(0))p/21asIu(0)0.\frac{k_{1}(I_{u}(0))^{-p/2}}{T+k_{1}(I_{u}(0))^{-p/2}}\to 1\quad\mathrm{as}~{}I_{u}(0)\to 0.

In fact, if (u0,u1)=(0,0)(u_{0},u_{1})=(0,0), the linearized equation (5.1) turns exactly into the wave equation

UttΔU+f(0)U=0,\displaystyle U_{tt}-\Delta U+f^{\prime}(0)U=0,

which conserves the quantity ΛU\Lambda_{U}. Hence, the linearized operator has no any uniform contraction property and the known methods to estimate the fractal dimension do not work near the origin.

To show finite dimensionality of the non-degenerate part of the attractor, we introduce the decomposition U=V+WU=V+W, where

{VttΔV+(up+utp)Vt=0,(V(0),Vt(0))=(ξ,ζ),\begin{cases}V_{tt}-\Delta V+(\|\nabla u\|^{p}+\|u_{t}\|^{p})V_{t}=0,\\ (V(0),V_{t}(0))=(\xi,\zeta),\end{cases} (5.14)

and

{WttΔW+(up+utp)Wt+p[up2(u,U)+utp2(ut,Ut)]ut+f(u)U=0,W(0)=Wt(0)=0.\begin{cases}W_{tt}-\Delta W+(\|\nabla u\|^{p}+\|u_{t}\|^{p})W_{t}+p[\|\nabla u\|^{p-2}(\nabla u,\nabla U)\\ \qquad\qquad\qquad\qquad\qquad\qquad\qquad\,\,+\|u_{t}\|^{p-2}(u_{t},U_{t})]u_{t}+f^{\prime}(u)U=0,\\ W(0)=W_{t}(0)=0.\end{cases} (5.15)
Lemma 5.4.

Suppose Assumptions 1.1 and 3.1 hold. Then there exist ϵ0>0,T0>0\epsilon_{0}>0,T_{0}>0 and 0<q<10<q<1 such that if (u,ut)𝒜(u,u_{t})\in\mathscr{A} with u(0)2+ut(0)2ϵ02\|\nabla u(0)\|^{2}+\|u_{t}(0)\|^{2}\geq\epsilon_{0}^{2}, then the solution of (5.14) satisfies

IV(T)qIV(0),TT0,I_{V}(T)\leq qI_{V}(0),\quad\forall T\geq T_{0},

where recall IV=V2+Vt2I_{V}=\|\nabla V\|^{2}+\|V_{t}\|^{2}.

Proof.

Let (u,ut)𝒜(u,u_{t})\in\mathscr{A} and Iu(0)ϵ02I_{u}(0)\geq\epsilon_{0}^{2} with ϵ0\epsilon_{0} to be determined. We multiply (5.14) by Vt+ϵVV_{t}+\epsilon V with ϵ>0\epsilon>0 and integrate over Ω\Omega to obtain

ddtIV,ϵ(t)+(u(t)p\displaystyle\frac{d}{dt}I_{V,\epsilon}(t)+(\|\nabla u(t)\|^{p} +ut(t)pϵ)Vt(t)2+ϵV(t)2\displaystyle+\|u_{t}(t)\|^{p}-\epsilon)\|V_{t}(t)\|^{2}+\epsilon\|\nabla V(t)\|^{2}
+ϵ(u(t)p+ut(t)p)(V(t),Vt(t))=0,\displaystyle+\epsilon(\|\nabla u(t)\|^{p}+\|u_{t}(t)\|^{p})(V(t),V_{t}(t))=0,

where IV,ϵ(t)=12Vt(t)2+12V(t)2+ϵ(V(t),Vt(t))I_{V,{\epsilon}}(t)=\frac{1}{2}\|V_{t}(t)\|^{2}+\frac{1}{2}\|\nabla V(t)\|^{2}+\epsilon(V(t),V_{t}(t)). Note that

|ϵ(up+utp)(V,Vt)|ϵC(V,Vt)ϵ2V2+ϵCVt2.\left|\epsilon(\|\nabla u\|^{p}+\|u_{t}\|^{p})(V,V_{t})\right|\leq\epsilon C(V,V_{t})\leq\frac{\epsilon}{2}\|\nabla V\|^{2}+\epsilon C\|V_{t}\|^{2}.

In terms of Remark 3.7 and (3.10) we have that

ddtIV,ϵ(t)+(C1t+k1ϵ0pC0ϵ)Vt(t)2+ϵ2V(t)20.\frac{d}{dt}I_{V,{\epsilon}}(t)+\Big{(}\frac{C_{1}}{t+k_{1}\epsilon_{0}^{-p}}-C_{0}\epsilon\Big{)}\|V_{t}(t)\|^{2}+\frac{\epsilon}{2}\|\nabla V(t)\|^{2}\leq 0.

Set T0=k1ϵ0pT_{0}=k_{1}\epsilon_{0}^{-p}. Then, for TT0T\geq T_{0}, it holds for t[0,T]t\in[0,T]

ddtIV,ϵ(t)+(C1T+k1ϵ0pC0ϵ)Vt(t)2+ϵ2V(t)20,\displaystyle\frac{d}{dt}I_{V,{\epsilon}}(t)+\Big{(}\frac{C_{1}}{T+k_{1}\epsilon_{0}^{-p}}-C_{0}\epsilon\Big{)}\|V_{t}(t)\|^{2}+\frac{\epsilon}{2}\|\nabla V(t)\|^{2}\leq 0,
ddtIV,ϵ(t)+C14C0(T+k1ϵ0p)(Vt(t)2+V(t)2)0,\displaystyle\frac{d}{dt}I_{V,{\epsilon}}(t)+\frac{C_{1}}{4C_{0}(T+k_{1}\epsilon_{0}^{-p})}(\|V_{t}(t)\|^{2}+\|\nabla V(t)\|^{2})\leq 0,

by setting ϵ=C12C0(T+k1ϵ0p)\epsilon=\frac{C_{1}}{2C_{0}(T+k_{1}\epsilon_{0}^{-p})}. Since

1ϵ/λ12IV(t)IV,ϵ(t)1+ϵ/λ12IV(t)for ϵ<λ1,\frac{1-\epsilon/\sqrt{\lambda_{1}}}{2}I_{V}(t)\leq I_{V,{\epsilon}}(t)\leq\frac{1+\epsilon/\sqrt{\lambda_{1}}}{2}I_{V}(t)\quad\textrm{for }\epsilon<\sqrt{\lambda_{1}},

we have for t[0,T]t\in[0,T]

ddtIV,ϵ(t)+C(1+ϵ/λ1)TIV,ϵ(t)0.\frac{d}{dt}I_{V,{\epsilon}}(t)+\frac{C}{(1+\epsilon/\sqrt{\lambda_{1}})T}I_{V,{\epsilon}}(t)\leq 0.

Hence, it follows that

IV(T)\displaystyle I_{V}(T) 21ϵ/λ1IV,ϵ(T)21ϵ/λ1IV,ϵ(0)eC(1+ϵ/λ1)T×T\displaystyle\leq\frac{2}{1-\epsilon/\sqrt{\lambda_{1}}}I_{V,{\epsilon}}(T)\leq\frac{2}{1-\epsilon/\sqrt{\lambda_{1}}}I_{V,{\epsilon}}(0)e^{-\frac{C}{(1+\epsilon/\sqrt{\lambda_{1}})T}\times T}
1+ϵ/λ11ϵ/λ1eC2IV(0).\displaystyle\leq\frac{1+\epsilon/\sqrt{\lambda_{1}}}{1-\epsilon/\sqrt{\lambda_{1}}}e^{-\frac{C}{2}}I_{V}(0).

Observe that

1+ϵ/λ11ϵ/λ1eC24λ1C0k1ϵ0p+C14λ1C0k1ϵ0pC1eC2=Δq\frac{1+\epsilon/\sqrt{\lambda_{1}}}{1-\epsilon/\sqrt{\lambda_{1}}}e^{-\frac{C}{2}}\leq\frac{4\sqrt{\lambda_{1}}C_{0}k_{1}\epsilon_{0}^{-p}+C_{1}}{4\sqrt{\lambda_{1}}C_{0}k_{1}\epsilon_{0}^{-p}-C_{1}}e^{-\frac{C}{2}}\stackrel{{\scriptstyle\Delta}}{{=}}q

and limε00+q=eC2<1\lim_{\varepsilon_{0}\rightarrow 0^{+}}q=e^{-\frac{C}{2}}<1. Therefore, we can take ϵ0\epsilon_{0} small enough, such that,

ϵ0<r0,ϵ<λ1 and q<1,\epsilon_{0}<r_{0},\epsilon<\sqrt{\lambda_{1}}\textrm{ and }q<1,

in which situation the desired result holds. ∎

Lemma 5.5.

Suppose Assumptions 1.1 and 3.1 hold. Let (u,ut)𝒜(u,u_{t})\in\mathscr{A} and (ξ,ζ)H01×L21\|(\xi,\zeta)\|_{H^{1}_{0}\times L^{2}}\leq 1. Then, for each T>0T>0, (W(T),Wt(T))Hβ+1(Ω)×Hβ(Ω)(\nabla W(T),W_{t}(T))\in H^{\beta+1}(\Omega)\times H^{\beta}(\Omega) for some β>0\beta>0.

Proof.

As in the proof of Lemma 2.8, it suffices to show the higher regularity of f(u)Uf^{\prime}(u)U and [up2(u,U)+utp2(ut,Ut)]ut[\|\nabla u\|^{p-2}(\nabla u,\nabla U)+\|u_{t}\|^{p-2}(u_{t},U_{t})]u_{t}. Note that (u,ut)(u,u_{t}) is uniformly bounded in H1+27(Ω)×H27(Ω)H^{1+\frac{2}{7}}(\Omega)\times H^{\frac{2}{7}}(\Omega) by Theorem 3.4. For β1=47\beta_{1}=\frac{4}{7} and q=149q=\frac{14}{9}

f(u)UHβ1Cf(u)UW1,qf′′(u)uULq+f(u)ULqCU,\displaystyle\|f^{\prime}(u)U\|_{H^{\beta_{1}}}\leq C\|f^{\prime}(u)U\|_{W^{1,q}}\leq\|f^{\prime\prime}(u)\nabla uU\|_{L^{q}}+\|f^{\prime}(u)\nabla U\|_{L^{q}}\leq C\|\nabla U\|,

and for β2=27\beta_{2}=\frac{2}{7}

[up2(u,U)+utp2(ut,Ut)]utHβ2C(U+Ut).\displaystyle\|[\|\nabla u\|^{p-2}(\nabla u,\nabla U)+\|u_{t}\|^{p-2}(u_{t},U_{t})]u_{t}\|_{H^{\beta_{2}}}\leq C(\|\nabla U\|+\|U_{t}\|).

Therefore, for β=27\beta=\frac{2}{7} it holds uniformly in tt that

f(u)UHβ+[up2(u,U)+utp2(ut,Ut)]utHβC(ξ+ζ)C,\displaystyle\|f^{\prime}(u)U\|_{H^{\beta}}+\|[\|\nabla u\|^{p-2}(\nabla u,\nabla U)+\|u_{t}\|^{p-2}(u_{t},U_{t})]u_{t}\|_{H^{\beta}}\leq C(\|\nabla\xi\|+\|\zeta\|)\leq C,

if (ξ,ζ)H01(Ω)×L2(Ω)1\|(\xi,\zeta)\|_{H_{0}^{1}(\Omega)\times L^{2}(\Omega)}\leq 1. Following the similar lines as in the proof of Lemma 2.8, we complete the proof. ∎

Now we are in position to show the finite dimensionality of the global attractor.

Proposition 5.6 ([8]).

Let MM be a compact set in a Banach space XX and V:MXV:M\mapsto X be uniformly quasi-differentiable on MM. Assume that the quasi-derivative L(u)L(u) can be split into two linear parts

L(u)=L1(u)+L2(u),uM,L(u)=L^{1}(u)+L^{2}(u),\quad u\in M,

where

supuML1(u)q<1\sup_{u\in M}\|L^{1}(u)\|\equiv q<1

and L2(u)L^{2}(u) is a compact operator on XX for each uMu\in M. We also assume that the function uL2(u)u\mapsto L^{2}(u) is continuous in the operator norm. If MVMM\subset VM, then dB(M)d_{B}(M) is finite.

Theorem 5.7.

Suppose Assumptions 1.1 and 3.1 hold. There exists ϵ0>0\epsilon_{0}>0 such that for any ϵ(0,ϵ0)\epsilon\in(0,\epsilon_{0}), 𝒜\B(0,ϵ)\mathscr{A}\backslash B(0,\epsilon) has finite fractal dimension in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega).

Proof.

We have already known that S(t)ω0S(t)\omega_{0} decays uniformly near 0 by Lemma 3.6, which also indicates that S(t)S(t) is negatively invariant on 𝒜\B(0,ϵ)\mathscr{A}\backslash B(0,\epsilon) for small ϵ\epsilon and large tt (depending on ϵ\epsilon). Due to the analysis above, we have the decomposition

L(t;ω0)(ξ,ζ)=(U(t),Ut(t))\displaystyle L(t;\omega_{0})(\xi,\zeta)=(U(t),U_{t}(t)) =(V(t),Vt(t))+(W(t),Wt(t))\displaystyle=(V(t),V_{t}(t))+(W(t),W_{t}(t))
=L1(t;ω0)(ξ,ζ)+L2(t;ω0)(ξ,ζ),\displaystyle=L_{1}(t;\omega_{0})(\xi,\zeta)+L_{2}(t;\omega_{0})(\xi,\zeta),

where L1(t;ω0)<1\|L_{1}(t;\omega_{0})\|<1 for t>T0t>T_{0} by Lemma 5.4, and L2(t;ω0)L_{2}(t;\omega_{0}) is compact in H01(Ω)×L2(Ω)H_{0}^{1}(\Omega)\times L^{2}(\Omega) for each t>0t>0 by Lemma 5.5. Moreover, it is easy to verify that L2(t;ω0)L^{2}(t;\omega_{0}) is continuous in ω0\omega_{0}. By virtue of Proposition 5.6 we complete the proof by choosing t>T0t>T_{0} large enough. ∎

Proof of Theorem 1.2.

Note that 0𝒜0\in\mathscr{A}. According to the proof of Lemma 3.6, it holds in 𝒜B(0,r0)\mathscr{A}\cap B(0,r_{0}) that

ddt[E(t)+ϵ2E(t)p2(ut,u)]+C[E(t)+ϵ2E(t)p2(ut,u)]p2+10,\frac{d}{dt}\left[E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u)\right]+C\left[E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u)\right]^{\frac{p}{2}+1}\leq 0,

and E(t)+ϵ2E(t)p2(ut,u)E(t)+\frac{\epsilon}{2}E(t)^{\frac{p}{2}}(u_{t},u) is equivalent to IuI_{u} in 𝒜B(0,r0)\mathscr{A}\cap B(0,r_{0}). Combining these facts with Corollary 3.5, Lemma 5.2 as well as Theorem 5.7, we complete the proof by applying Theorem 4.4. ∎

Acknowledgements

This work is supported by the National Natural Science Foundation of China (11731005, 12201604, 12371106) and the Fundamental Research Funds for the Central Universities (E3E40102X2).

Data Availability Statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

References

  • [1] AV Balakrishnan and LW Taylor. Distributed parameter nonlinear damping models for flight structures. In Proceedings Damping, volume 89, 1989.
  • [2] J. M. Ball. Global attractors for damped semilinear wave equations. volume 10, pages 31–52. 2004. Partial differential equations and applications.
  • [3] M. D. Blair, H. F. Smith, and C. D. Sogge. Strichartz estimates for the wave equation on manifolds with boundary. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 26(5):1817–1829, 2009.
  • [4] M. M. Cavalcanti, V. N. Domingos Cavalcanti, M. A. Jorge Silva, and C. M. Webler. Exponential stability for the wave equation with degenerate nonlocal weak damping. Israel J. Math., 219(1):189–213, 2017.
  • [5] V. V. Chepyzhov and M. I. Vishik. Attractors for equations of mathematical physics, volume 49 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2002.
  • [6] I. Chueshov. Global attractors for a class of Kirchhoff wave models with a structural nonlinear damping. J. Abstr. Differ. Equ. Appl., 1(1):86–106, 2010.
  • [7] I. Chueshov. Long-time dynamics of Kirchhoff wave models with strong nonlinear damping. J. Differential Equations, 252(2):1229–1262, 2012.
  • [8] I. Chueshov. Dynamics of quasi-stable dissipative systems. Universitext. Springer, Cham, 2015.
  • [9] I. Chueshov, M. Eller, and I. Lasiecka. On the attractor for a semilinear wave equation with critical exponent and nonlinear boundary dissipation. Comm. Partial Differential Equations, 27(9-10):1901–1951, 2002.
  • [10] I. Chueshov and I. Lasiecka. Long-time behavior of second order evolution equations with nonlinear damping. Mem. Amer. Math. Soc., 195(912):viii+183, 2008.
  • [11] M. Efendiev and M. Ôtani. Infinite-dimensional attractors for evolution equations with pp-Laplacian and their Kolmogorov entropy. Differential Integral Equations, 20(11):1201–1209, 2007.
  • [12] M. Efendiev and M. Ôtani. Infinite-dimensional attractors for parabolic equations with pp-Laplacian in heterogeneous medium. Ann. Inst. H. Poincaré C Anal. Non Linéaire, 28(4):565–582, 2011.
  • [13] M. Efendiev and S. Zelik. Finite- and infinite-dimensional attractors for porous media equations. Proc. Lond. Math. Soc. (3), 96(1):51–77, 2008.
  • [14] E. Feireisl and E. Zuazua. Global attractors for semilinear wave equations with locally distributed nonlinear damping and critical exponent. Comm. Partial Differential Equations, 18(9-10):1539–1555, 1993.
  • [15] J. M. Ghidaglia and A. Marzocchi. Longtime behaviour of strongly damped wave equations, global attractors and their dimension. SIAM J. Math. Anal., 22(4):879–895, 1991.
  • [16] M. Grasselli and V. Pata. On the damped semilinear wave equation with critical exponent. Number suppl., pages 351–358. 2003. Dynamical systems and differential equations (Wilmington, NC, 2002).
  • [17] M. A. Jorge Silva, V. Narciso, and A. Vicente. On a beam model related to flight structures with nonlocal energy damping. Discrete Contin. Dyn. Syst. Ser. B, 24(7):3281–3298, 2019.
  • [18] A. Kh. Khanmamedov. Remark on the regularity of the global attractor for the wave equation with nonlinear damping. Nonlinear Anal., 72(3-4):1993–1999, 2010.
  • [19] V. Pata and M. Squassina. On the strongly damped wave equation. Comm. Math. Phys., 253(3):511–533, 2005.
  • [20] J. C. Robinson. Dimensions, embeddings, and attractors, volume 186 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2011.
  • [21] C. Sun, M. Yang, and C. Zhong. Global attractors for the wave equation with nonlinear damping. J. Differential Equations, 227(2):427–443, 2006.
  • [22] Y. Sun and Z. Yang. Strong attractors and their robustness for an extensible beam model with energy damping. Discrete Contin. Dyn. Syst. Ser. B, 27(6):3101–3129, 2022.
  • [23] R. Temam. Infinite-dimensional dynamical systems in mechanics and physics, volume 68 of Applied Mathematical Sciences. Springer-Verlag, New York, second edition, 1997.
  • [24] Z. Yang, P. Ding, and L. Li. Longtime dynamics of the Kirchhoff equations with fractional damping and supercritical nonlinearity. J. Math. Anal. Appl., 442(2):485–510, 2016.
  • [25] Z. Yang, Z. Liu, and P. Niu. Exponential attractor for the wave equation with structural damping and supercritical exponent. Commun. Contemp. Math., 18(6):1550055, 13, 2016.
  • [26] C. Zhao, C. Zhao, and C. Zhong. The global attractor for a class of extensible beams with nonlocal weak damping. Discrete Contin. Dyn. Syst. Ser. B, 25(3):935–955, 2020.
  • [27] C. Zhong and W. Niu. On the Z2Z_{2} index of the global attractor for a class of pp-Laplacian equations. Nonlinear Anal., 73(12):3698–3704, 2010.
  • [28] Y. Zhong and C. Zhong. Exponential attractors for semigroups in Banach spaces. Nonlinear Anal., 75(4):1799–1809, 2012.

Zhijun Tang
Department of Mathematics, Nanjing University, Nanjing, 210093, China
E-mail: [email protected]

Senlin Yan
Department of Mathematics, Nanjing University, Nanjing, 210093, China
E-mail: [email protected]

Yao Xu
School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing, 100049, China
E-mail: [email protected], [email protected]

Chengkui Zhong
Department of Mathematics, Nanjing University, Nanjing, 210093, China
E-mail: [email protected]