Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

Pullback attractors for the nonclassical diffusion equations with memory in time-dependent spaces

  • In this paper, we consider the asymptotic behavior of nonclassical diffusion equations with hereditary memory and time-dependent perturbed parameter on whole space Rn. Under a general assumption on the memory kernel k, the existence and regularity of time-dependent global attractors are proven using a new analytical technique. It is remarkable that the nonlinearity f has no restriction on the upper growth.

    Citation: Ke Li, Yongqin Xie, Yong Ren, Jun Li. Pullback attractors for the nonclassical diffusion equations with memory in time-dependent spaces[J]. AIMS Mathematics, 2023, 8(12): 30537-30561. doi: 10.3934/math.20231561

    Related Papers:

    [1] Xiaolei Dong, Yuming Qin . Finite fractal dimension of pullback attractors for a nonclassical diffusion equation. AIMS Mathematics, 2022, 7(5): 8064-8079. doi: 10.3934/math.2022449
    [2] Jing Wang, Qiaozhen Ma, Wenxue Zhou . Attractor of the nonclassical diffusion equation with memory on time- dependent space. AIMS Mathematics, 2023, 8(6): 14820-14841. doi: 10.3934/math.2023757
    [3] Tingting Liu, Tasneem Mustafa Hussain Sharfi, Qiaozhen Ma . Time-dependent asymptotic behavior of the solution for evolution equation with linear memory. AIMS Mathematics, 2023, 8(7): 16208-16227. doi: 10.3934/math.2023829
    [4] Jianbo Yuan, Shixuan Zhang, Yongqin Xie, Jiangwei Zhang . Exponential attractors for nonclassical diffusion equations with arbitrary polynomial growth nonlinearity. AIMS Mathematics, 2021, 6(11): 11778-11795. doi: 10.3934/math.2021684
    [5] Ruonan Liu, Tomás Caraballo . Random dynamics for a stochastic nonlocal reaction-diffusion equation with an energy functional. AIMS Mathematics, 2024, 9(4): 8020-8042. doi: 10.3934/math.2024390
    [6] Xiaobin Yao . Random attractors for non-autonomous stochastic plate equations with multiplicative noise and nonlinear damping. AIMS Mathematics, 2020, 5(3): 2577-2607. doi: 10.3934/math.2020169
    [7] Xin Liu . Stability of random attractors for non-autonomous stochastic p-Laplacian lattice equations with random viscosity. AIMS Mathematics, 2025, 10(3): 7396-7413. doi: 10.3934/math.2025339
    [8] Xiaoxia Wang, Jinping Jiang . The pullback attractor for the 2D g-Navier-Stokes equation with nonlinear damping and time delay. AIMS Mathematics, 2023, 8(11): 26650-26664. doi: 10.3934/math.20231363
    [9] Feng Zhou, Hongfang Li, Kaixuan Zhu, Xin Li . Dynamics of a damped quintic wave equation with time-dependent coefficients. AIMS Mathematics, 2024, 9(9): 24677-24698. doi: 10.3934/math.20241202
    [10] Li Yang . Pullback random attractors of stochastic strongly damped wave equations with variable delays on unbounded domains. AIMS Mathematics, 2021, 6(12): 13634-13664. doi: 10.3934/math.2021793
  • In this paper, we consider the asymptotic behavior of nonclassical diffusion equations with hereditary memory and time-dependent perturbed parameter on whole space Rn. Under a general assumption on the memory kernel k, the existence and regularity of time-dependent global attractors are proven using a new analytical technique. It is remarkable that the nonlinearity f has no restriction on the upper growth.



    This paper concentrates on the following nonclassical diffusion equation:

    utε(t)ΔutΔu0k(s)Δu(ts)ds+λu+f(x,u)=g, in Rn×Rτ, (1.1)

    and initial value

    u(x,t)=uτ(x,t), in Rn×(,τ], (1.2)

    where the forcing term g=g(x)L2(Rn)(n3) is known, λ>0 is a constant and Rτ=[τ,).

    To study problem (1.1) with (1.2), we assume that the time-dependent perturbed parameter ε(t), the nonlinearity f and the hereditary memory k(s) satisfy the following conditions, respectively:

    (H1) The time-dependent perturbed parameter ε(t)C1(R) is a decreasing bounded function satisfying

    limtε(t)=0, (1.3)

    and there exists a constant L>0, such that

    suptR(|ε(t)|+|ε(t)|)L. (1.4)

    (H2) The hereditary memory kernel k is a nonnegative summable function having the explicit form

    k(s)=sμ(σ)dσ,

    and μ(s)L1(R+) is a decreasing piecewise absolutely continuous in each interval [0,T] with T>0. In particular, μ(s) is allowed to exhibit (infinitely many) jumps. Moreover, we require that μ(0)< and

    k(s)θμ(s) (1.5)

    holds for some θ>0 and almost every s>0. As shown in [1,2], this is completely equivalent to the requirement that

    μ(s+r)Meδrμ(s)

    for some M1,δ>0, every r0, and almost every s>0, then we obtain

    μ()=limsμ(s)=0.

    For convenience, we also assume

    k(0)=0μ(s)ds=1.

    (H3) The nonlinearity fC1(Rn×R,R) satisfies

    f(x,s)sα|s|2φ(x),(x,s)Rn×R, (1.6)

    and

    sf(x,s)l,(x,s)Rn×R, (1.7)

    where φ(x)L1(Rn),f0(x)=f(x,0)L2(Rn) is given, α, l are two positive constants and α<λ.

    We denote by F the function F(x,s)=s0f(x,υ)dυ, then,

    F(x,s)f(x,s)s+l2s2.

    Let f1(x,s)=f(x,s)+αs and F1(x,s)=sτf1(x,s)ds, then,

    f1(x,s)+φ(x)0, (1.8)
    F1(x,s)f0(x)s+lα2s20, (1.9)

    and

    F1(x,s)f1(x,s)s+lα2s2. (1.10)

    Remark 1.1. The assumption of the nonlinearity comes from [3], but some constraints have been relaxed. For example, we will not require that φ(x) is a nonnegative function and l<λ. The class of nonlinearities studied in [4,5,6] have a restriction on the upper growth by which an exponential nonlinearity (e.g., f(u)=eu) does not hold ture.

    According to the Dafermos' idea in [7], we also add a new variable ηt to describe the past history of u, which is defined as follows:

    ηt=ηt(x,s):=s0u(x,tr)dr,sR+. (1.11)

    Let ηtt=tηt,ηts=sηt, then it gets

    ηtt=ηts+u, (1.12)

    where R+=[0,+). Historical variable uτ(,τs) of u satisfies

    0eσ(sτ) (1.13)

    where \Re > 0 and \sigma \leq \delta ( \delta is from (1.5)).

    Thus the system (1.1) with (1.2) can be rewritten as follows:

    \begin{eqnarray} \left\{ \begin{aligned}\ &u_t-\varepsilon(t)\Delta u_t-\Delta u-\int_{0}^{\infty}\mu(s)\Delta\eta^t(s)ds+ (\lambda-\alpha) u+f_1(x,u) = g, \\ &\eta_t^t = -\eta_s^t+u, \end{aligned} \right. \end{eqnarray} (1.14)

    and the initial data

    \begin{equation} u(x,\tau) = u_{\tau}(x,\tau),\,\eta^{\tau}(x,s) = \int_{0}^{s}u_{\tau}(x,\tau-r)dr. \end{equation} (1.15)

    From (1.13), it is easy to obtain the following estimate:

    \begin{equation*} \int_0^{\infty}\mu(s)\|\eta^{\tau}(s)\|_{H^1(\mathbb{R}^n)}^2ds\leq \Re. \end{equation*}

    For Eq (1.1), it is often used to describe some physical phenomena, for example, non-Newtonian flows, soil mechanics and heat conduction theory. Specifically, when we study heat conduction problems in fluid mechanics or solid mechanics, if the influence of viscosity is emphasized, the classic heat-conduction equation is often extended to the following form (see e.g., [8,9,10]):

    \begin{equation*} c\dot{u}-ca\Delta \dot{u}-k\Delta u = 0. \end{equation*}

    However, when we consider polymer and high viscous liquid, etc., some important factors such as the historical influence of u and the disturbance coefficient of viscosity must be included [11], that is, the following evolution equation:

    \begin{align} u_{t}-\varepsilon\Delta u_t-\nu\Delta u-\int_{0}^{\infty}k(s)\Delta u(t-s)ds+f(u) = g. \end{align} (1.16)

    The asymptotic behavior of solutions of Eq (1.16) has been studied by many scholars (see e.g., [3,12,13,14,15,16,17,18,19] and the references therein). However, the research focused on the nonclassical diffusion equation with constant coefficient and bounded smooth domain early on, see e.g., [20,21,22,23,24,25,26]. In [11], the diffusion equation with memory was proposed in the study of heat conduction and relaxation of high viscosity liquids. The convolution term represents the influence of past history on its future evolution and describes more accurately the diffusive process in certain materials, such as high viscosity liquids at low temperatures and polymers. Hence, it is necessary and scientifically significant to study the nonclassical diffusion equation with the time-dependent coefficient (i.e., variable coefficient) and memory. Furthermore, for Eq (1.16) , we focus on the nonclassical diffusion equation with memory and a time-dependent perturbed parameter \varepsilon(t) on a bounded domain. For example, for the case of k(s) = 0 in (1.1), in [27,28], the authors proved the existence and regularity of the time-dependent global attractors on time-dependent spaces when the nonlinearity satisfies |f''(u)|\leq c(1+|u|) and critically exponential growth, respectively. In addition, Wang and Ma studied the existence, regularity and asymptotic structure of the time-dependent global attractors in [29] for this equation when f meets polynomial growth of arbitrary order. In particular, Wang et al. [6] proved the existence of the time-dependent global attractors for the problem with the nonlinearity of critical exponential growth.

    More recently, the fourth author of the article and other co-authors considered the autonomous and nonautonomous nonclassical diffusion equation with memory or without memory on bounded domains, see e.g., [4,5,19,30,31,32]. In these papers, the operator decomposition and contractive functional methods are used to obtain the asymptotic regularity of the solutions and verify asymptotic compactness. It is worth mentioning that Xie et al. [4] have obtained the existence of the time-dependent global attractors on bounded domain for the nonclassical diffusion equation (1.16), lacking instantaneous damping with the nonlinearity that satisfies the polynomial growth of arbitrary order. However, we focus on the unbounded domain. Therefore, we can see that there is few relevant studies for the asymptotic behavior of solutions of Eq (1.1) in time-dependent whole spaces under assumptions (H1)(H3). This is because there are two major difficulties to obtain the existence of time-dependent global attractors.

    (i) First, because of the nonlinearity with no restriction on the upper growth, the higher asymptotic regularity of the solutions of Eq (1.1) can not be obtained using the method of [33,34].

    (ii) Second, due to the influence of the time-dependent perturbed parameter \varepsilon(t) and the lacking of compact embedding theorem on unbounded domains \mathbb{R}^n , it is impossible to directly construct the contractive function to prove the asymptotic compactness for the corresponding process \{U(t, \tau)\}_{t\geq \tau} of Eq (1.1) (see e.g., [29,30,35]).

    For solving these problems, a new analytical technique combined with the operator decomposition method is used to obtain contractive function, and then the pullback asymptotic compactness for the process \{U(t, \tau)\}_{t\geq \tau} of Eq (1.14) is proved. Furthermore, using this operator decomposition method, the asymptotic regularity of the solutions for Eq (1.14) is also proved. Then, the regularity of time-dependent global attractors for this equation on \mathcal{H}_t^1\times L_{\mu}^2(\mathbb{R}^{+}; H^1(\mathbb{R}^n)) ( \mathcal{H}_t^1 is defined later) are established.

    For conveniences, hereafter let |\cdot |_p be the norm of L^p(\mathbb{R}^n) \; (p\geq 1) . Let (\cdot, \cdot) , (\cdot, \cdot)+(\nabla\cdot, \nabla\cdot) = \langle\cdot, \cdot \rangle_{H^1(\mathbb{R}^n)} and (\cdot, \cdot)+(\nabla\cdot, \nabla\cdot)+(\Delta\cdot, \Delta\cdot) = \langle\cdot, \cdot \rangle_{H^2(\mathbb{R}^n)} be the inner product of L^2(\mathbb{R}^n) , H^1(\mathbb{R}^n) and H^2(\mathbb{R}^n) respectively.

    Let \|\cdot\|_0, \|\cdot\|_1 be the norm of H^1(\mathbb{R}^n) and H^2(\mathbb{R}^n) respectively, then,

    \begin{align*} \|x\|_0^2 & = |x|_2^2+|\nabla x|_2^2, \; &\forall x\in H^1(\mathbb{R}^n); \\ \|x\|_1^2& = |x|_2^2+|\nabla x|_2^2+|\Delta x|_2^2, \; &\forall x\in H^2(\mathbb{R}^n). \end{align*}

    The time-dependent space \mathcal{H}^1_t = H^1(\mathbb{R}^n) and \mathcal{H}_t^2 = H^2(\mathbb{R}^n) are equipped with the norms:

    \begin{equation*} \|\cdot\|^2_{\mathcal{H}^1_t} = |\cdot|^2_2+\varepsilon(t)|\nabla\cdot|_2^2\; \; \text{ and }\; \; \|\cdot\|^2_{\mathcal{H}_t^2} = \|\cdot\|_0^2+\varepsilon(t)|\triangle\cdot|^2_2. \end{equation*}

    It is necessary to point out here that \|\cdot\|^2_{\mathcal{H}^1_t} and |\cdot|^2_2+\varepsilon(t)\|\cdot\|_0^2 are equivalent. In fact the following inequality is obvious:

    \begin{equation*} \|\cdot\|^2_{\mathcal{H}^1_t}\leq |\cdot|^2_2+\varepsilon(t)\|\cdot\|_0^2, \end{equation*}

    and

    \begin{align*} \|\cdot\|^2_{\mathcal{H}^1_t}& = \frac{1}{2} |\cdot|^2_2+\frac{1}{2} |\cdot|^2_2+\varepsilon(t)|\nabla\cdot|_2^2 \\ & \geq \frac{1}{2} |\cdot|^2_2+\frac{1}{2L} \varepsilon(t)|\cdot|^2_2+\varepsilon(t)|\nabla\cdot|_2^2\\ &\geq\frac{1}{2(1+L)}\left(|\cdot|^2_2+\varepsilon(t)\|\cdot\|_0^2\right). \end{align*}

    Denote the weight spaces \mathcal{V}_0 = L_\mu ^2(\mathbb{R}^+; L^2(\mathbb{R}^n)) , \mathcal{V}_1 = L_\mu ^2(\mathbb{R}^+; H^1(\mathbb{R}^n)) and \mathcal{V}_2 = L_\mu ^2(\mathbb{R}^+; H^2(\mathbb{R}^n)) , and their inner products and norms are defined as follows:

    \langle\psi,\eta \rangle_{\mathcal{V}_0} = \int_0^\infty \mu(s) \left(\psi, \eta \right) ds,\ \ |\eta^t|_{\mu,{2}}^2 = \int_0^\infty {\mu (s)} {|{\eta^t(s) }|_2^2}ds;
    \langle\psi,\eta \rangle_{\mathcal{V}_1} = \int_0^\infty \mu(s) \left < \psi, \eta \right > _{H^1(\mathbb{R}^n)} ds,\ \ \left\| \eta^t \right\|_{\mu,{0}}^2 = \int_0^\infty {\mu \left( s \right)} {\|{\eta^t(s) } \|_0^2}ds;

    and

    \langle\psi,\eta \rangle_{\mathcal{V}_2} = \int_0^\infty \mu(s) \left < \psi,\eta \right > _{H^2(\mathbb{R}^n)}ds,\ \ \left\| \eta^t \right\|_{\mu,{1}}^2 = \int_0^\infty {\mu \left( s \right)} {\| {\eta^t }(s) \|^2_{1}}ds.

    With the above notation, the phase spaces of Eq (1.14) can be denoted as

    \mathcal{M}_t: = \mathcal{M}_t(\mathbb{R}^n) = \mathcal{H}_t^1 \times {\mathcal{V}_1}\,\text{ and }\, \mathcal{M}_t^1: = \mathcal{M}^1_t(\mathbb{R}^n) = \mathcal{H}_t^2 \times {\mathcal{V}_2}

    equipped the following norms:

    \left\| \cdot \right\|^2_{\mathcal{M}_t} = \left\| \cdot \right\|^2_{\mathcal{H}^1_t} + \left\| \cdot \right\|^2_{\mu,0} \,\text{ and }\,\left\| \cdot \right\|^2_{\mathcal{M}_t^{1}} = \left\| \cdot \right\|^2_{\mathcal{H}_t^{2}} + \left\| \cdot \right\|^2_{\mu, {1}}

    respectively.

    Particularly, we use \{\mathcal{M}_t\}_{t\in\mathbb{R}} to denote a family of normed time-dependent spaces. Moreover, we introduce some common notations based on processes of time-dependent space(see e.g., [35,36,37,38]).

    Let \{\mathcal{M}_t\}_{t\in\mathbb{R}} be a family of normed time-dependent space. Note that the ball with radius of R in \mathcal{M}_t is

    \mathscr{B}_t(R) = \{w\in \mathcal{M}_t:\|w\|_{\mathcal{M}_t}\leq R\}.

    For any given \varepsilon > 0 , we define the \varepsilon neighborhood of set B\subset \mathcal{M}_t as follows:

    \mathcal{O}_t^{\varepsilon}(B) = \bigcup\limits_{x\in B}\{y\in \mathcal{M}_t:\|x-y\|_{\mathcal{M}_t} < \varepsilon\} = \bigcup\limits_{x\in B}\{x+\mathscr{B}_t(\varepsilon)\}.

    Hausdorff semidistance of between two nonempty sets A, B\subset X_t is defined as

    {dist}_{\mathcal{M}_t}(A,B) = \sup\limits_{x\in A}\inf\limits_{y\in B}\|x-y\|_{\mathcal{M}_t}.

    The plan of this paper is as follows. In Section 2, we recall some basic concepts as the time-dependent global attractors and useful results that will be used later. In Section 3, we first prove pullback asymptotic compactness of the process corresponding to problem (1.14) with (1.15) by constructing contractive function, and then we obtain the existence and the regularity of time-dependent global attractors to problem (1.14) with (1.15) in whole space \mathbb{R}^n .

    In this section, we will recall some basic concepts of time-dependent global attractors and theories of the existence of time-dependent global attractors (see e.g., [35,36,37]).

    Definition 2.1. Let \{\mathcal{M}_t\}_{t\in\mathbb{R}} be a family of normed spaces. A two-parameter family of operators U(t, \tau): \mathcal{M}_{\tau}\rightarrow \mathcal{M}_t is called a process if it satisfies the following properties:

    (i) U(\tau, \tau) = Id, \tau\in \mathbb{R} (identity operator);

    (ii) U(t, s)U(s, \tau) = U(t, \tau), \; \forall t\geq s\geq \tau\in \mathbb{R} .

    Definition 2.2. A family of sets \tilde{C} = \{C_t:C_t\subset \mathcal{M}_t \, \, is \, \, bounded \}_{t\in \mathbb{R}} is called uniformly bounded if there exists a constant R > 0 , such that C_t\subset \mathscr{B}_t(R) for any t\in \mathbb{R} .

    Definition 2.3. A family of sets \tilde{B} = \{B_t\}_{t\in \mathbb{R}} is called pullback absorbing if \tilde{B} = \{B_t\}_{t\in \mathbb{R}} is uniformly bounded and for all R > 0 , there exists a constant t_0 = t_0(t, R)\leq t such that U(t, \tau)\mathscr{B}_{\tau}(R)\subset B_t for any \tau\leq t_0 .

    The process \{U(t, \tau)\}_{t\geq \tau} is called dissipative whenever it enters a pullback absorbing family \tilde{B}_0 = \{B^0_t\}_{t\in \mathbb{R}} .

    Definition 2.4. A time-dependent absorbing set for the process U(t, \tau) is a uniformly bounded family \tilde{B} = \{B_t\}_{t\in \mathbb{R}} with the following characteristics: For any R > 0 , there exists t_0 = t_0(t, R)\geq 0 , such that

    \begin{equation*} U(t,\tau)\mathscr{B}_{\tau}(R)\subset B_t, \quad for\, \, all \, \, \tau\leq t-t_0. \end{equation*}

    Definition 2.5. The process U(t, \tau) is called pullback asymptotic compact if for any t\in \mathbb{R} , any bounded sequence \{z_n\}_{n = 1}^{\infty}\subset \mathcal{M}_{\tau_n} and \{\tau_n\}_{n = 1}^{\infty}\subset (-\infty, t] with \tau_n\rightarrow {-\infty} as n\rightarrow \infty , the sequence \{U(t, \tau_n)z_n\}_{n = 1}^{\infty} has a convergent subsequence in \mathcal{M}_t .

    Definition 2.6. A time-dependent global attractor of the process U(t, \tau) is the smallest family \tilde{\mathscr{A}} = \{\mathscr{A}_t\}_{t\in \mathbb{R}} such that

    (i) for every t\in\mathbb{R} , \mathscr{A}_t is compact in \mathcal{M}_t ;

    (ii) \tilde{\mathscr{A}} is pullback attracting, namely, \tilde{\mathscr{A}} is uniformly bounded and

    \begin{equation*} \lim\limits_{\tau\rightarrow {-\infty}}{dist}_{\mathcal{M}_t}(U(t,\tau)C_{\tau},\mathscr{A}_t) = 0 \end{equation*}

    holds for any uniformly bounded family \tilde{C} = \{C_{\tau}\}_{\tau\in \mathbb{R}} and every fixed t\in \mathbb{R} and \tau\leq t .

    Remark 2.1. The pullback attracting nature can be equivalently described in the light of pullback absorbing: A (uniformly bounded) family \mathcal{K} = \{K_t\}_{t\in\mathbb{R}} is said to be pullback attracting if for any \varepsilon > 0 the family \{\mathcal{O}^{\varepsilon}_t(K_t)\}_{t\in\mathbb{R}} is pullback absorbing.

    Theorem 2.1. A time-dependent global attractor {\tilde{\mathscr{A}}} exists and it is unique if and only if the process U(t, \tau) is asymptotic compact, i.e., the set

    \mathbb{K} = \{\mathcal{K} = \{K_t\}_{t\in\mathbb{R}}:K_t\subset \mathcal{M}_t \;is\; compact, \mathbb{K}\; is\; pullback\; attracting\}

    is non-empty.

    It can be seen from Definition 2.6 that the time-dependent global attractor is not necessarily invariant. This is mainly because that the process is not required to meet some continuity. If the process U(t, \tau) satisfies the appropriate continuity, then the invariance of time-dependent global attractor {\mathscr{tilde{A}}} can be obtained.

    Definition 2.7. We say that \tilde{\mathscr{A}} = \{\mathscr{A}_t\}_{t\in \mathbb{R}} is invariant if

    \begin{equation*} U(t, \tau)\mathscr{A}_{\tau} = \mathscr{A}_{t}, \quad t\geq \tau\in\mathbb{R}. \end{equation*}

    Lemma 2.1. If the time-dependent global attractor \tilde{\mathscr{A}} exists and the process U(t, \tau) is a strongly continuous process, then \tilde{\mathscr{A}} is invariant.

    Next, we will state the definitions of contractive function and contractive process, which will be used to obtain asymptotic compactness of a family of process \{U(t, \tau)\}_{t\geq \tau} (see e.g., [23,35,39,40,41,42,43]).

    Definition 2.8. Let \{\mathcal{M}_t\}_{t\in\mathbb{R}} be a family of Banach spaces and \tilde{B} = \{B_t\subset \mathcal{M}_t\}_{t\in \mathbb{R}} be a family of uniformly bounded subset. We call function \varphi(\cdot, \cdot) , defined on \mathcal{M}_{\tau}\times \mathcal{M}_{\tau} , to be a contractive function on B_{\tau}\times {B}_{\tau} if for any sequence \{z_n\}_{n = 1}^{\infty}\subset B_{\tau} , there exists a subsequence \{z_{n_k}\}_{k = 1}^{\infty}\subset \{z_n\} _ {n = 1} ^ {\infty} such that

    \lim\limits_{k\to\infty} \lim\limits_{l\to\infty}\varphi^t_{\tau}(z_{n_k}, z_{n_l}) = 0,\quad \forall\,t\geq\tau\in\mathbb{R}.

    We use \mathfrak{E}(B_{\tau}) to denote the set all contractive function on B_{\tau}\times B_{\tau} .

    Definition 2.9. Let U(t, \tau) be a process on \{\mathcal{M}_t\}_{t\in\mathbb{R}} and have a pullback bounded absorbing set \tilde{B} = \{B_t\}_{t\in \mathbb{R}} . U(t, \tau) is called \mathcal{M}_t -contractive process if for any given \varepsilon > 0 , there exist T = T(\varepsilon) and \varphi^t_{T}(\cdot, \cdot)\in \mathfrak{E}(B_T) such that

    \|U(t,T)z_1-U(t,T)z_2\|_{\mathcal{M}_t}\le\varepsilon+ \varphi^t_{T}(z_1, z_2), \quad \forall\,z_i\in B_{T}\; (i = 1,2),

    where \varphi^t_{T} depends on T .

    Next, we will give the method to prove the existence of time-dependent global attractors for evolution equations, which will be used in the later discussion.

    Theorem 2.2. [35] Let \{\mathcal{M}_t\}_{t\in\mathbb{R}} be a family of Banach spaces, then U(t, \tau) has a time-dependent global attractor in \{\mathcal{M}_t\}_{t\in\mathbb{R}} , if the following conditions hold:

    (i) U(t, \tau) has a pullback absorbing set \tilde{B} = \{B_t\}_{t\in \mathbb{R}} in \{\mathcal{M}_t\}_{t\in\mathbb{R}} ;

    (ii) U(t, \tau) is a \mathcal{M}_t -contractive process.

    Lemma 2.2. [44] Let X\subset\subset H\subset Y be Banach spaces, with X reflexive. Suppose that \{u_n\}_{n = 0}^\infty is a sequence, uniformly bounded in L^2(\tau, T; X) and du_n/ dt is uniformly bounded in L^p(\tau, T; Y) , for some p > 1 . Then, there is a subsequence of \{u_n\}_{n = 0}^\infty that converges strongly in L^2(\tau, T; H) .

    In this section, we shall consider the existence of time-dependent global attractors in \{\mathcal{M}_t\}_{t\in\mathbb{R}} . For this purpose, we have to first discuss the well-posedness for Eq (1.14) with (1.15).

    The well-posedness for Eq (1.14) with (1.15) can be obtained by using Faedo-Galerkin method (see e.g., [38,45]). To this end, we first give the definition of weak solution.

    Definition 3.1. For any R > 0 and T > \tau , let I = [\tau, T] , then the function z = (u, \eta^t) defined on \mathbb{R}^n\times I is called a weak solution of the problem (1.14) with initial value z_\tau = (u_\tau, \eta^\tau)\in \mathscr{B}_\tau(R)\subset \mathcal{M}_\tau if

    u \in C(I;\mathcal{H}_t^1),\, u_t \in L^2(I;\mathcal{H}_t^1), \,\eta^t \in C(I; L_{\mu}^2(\mathbb{R}^+; H^1(\mathbb{R}^n))),\\ \eta^t_t+\eta^t_s\in L^{\infty}(I; L_{\mu}^2(\mathbb{R}^+; H^1(\mathbb{R}^n)))\cap L^{2}(I; L_{\mu}^2(\mathbb{R}^+ ;H^1(\mathbb{R}^n))).

    Furthermore, the following identity hold:

    \begin{eqnarray} \left\{ \begin{aligned} &( u_t, \omega)+\varepsilon(t)\langle u_t, \omega\rangle_{H^1}+\langle u, \omega\rangle_{H^1}+\langle\eta^t, \omega\rangle_{\mathcal{V}_1} +\langle f(u), \omega\rangle = ( g, \omega), \\ &\langle\eta_t^t+\eta_s^t, \varphi\rangle_{\mathcal{V}_1} = \langle u, \varphi\rangle_{\mathcal{V}_1}, \end{aligned} \right. \end{eqnarray} (3.1)

    holds for any (\omega, \varphi)\in C^{\infty}(I, \mathcal{H}_t^1)\times C^{\infty} (I, L_{\mu}^{2}(\mathbb{R}^+; H^1(\mathbb{R}^n))) and a.e. \, t\in I .

    Lemma 3.1. For any R > 0 , T > \tau and z_{\tau} = (u_{\tau}, \eta^{\tau})\in \mathscr{B}_\tau(R)\subset\mathcal{M}_{\tau} , then the problem (1.14) with (1.15) has unique weak solution

    \begin{equation*} z(t) = (u(x, t), \eta^{t})\in C([\tau,T];\mathcal{H}_t^1\times\mathcal{V}), \end{equation*}

    which continuously depends on the initial data in \mathcal{M}_\tau , i.e., there exists a constant \kappa > 0 not related to t such that the process U(t, \tau) is Lipschitz continuous

    \begin{equation} \|U(t,\tau)z_{\tau}^1-U(t,\tau)z_{\tau}^2\|_{\mathcal{M}_t} \leq Ce^{\kappa(T-\tau)}\|z_{\tau}^1-z_{\tau}^2\|_{\mathcal{M}_{\tau}},\quad \forall t\in [\tau,T]. \end{equation} (3.2)

    By Lemma 3.1, we may define the process of solutions on time-dependent space \{\mathcal{M}_t\}_{t\in\mathbb{R}} :

    \begin{equation} U(t,\tau):\mathcal{M}_{\tau}\rightarrow \mathcal{M}_t,\; U(t,\tau)z_{\tau} = z(t),\quad \forall t\geq \tau. \end{equation} (3.3)

    In addition, it's easy to obtain that the process U(t, \tau) is a strongly continuous process on the time-dependent phase space \{\mathcal{M}_t\}_{t\in\mathbb{R}} .

    In the following discussion, let C mean any positive constant and Q(\cdot) be a monotonically increasing function on [0, \, \infty) which may be different from line to line even in the same line.

    We always assume that: The assumptions {(\bf H1)}–{(\bf H3)} are true. Furthermore, z(t) = (u(t), \eta^t) is a sufficiently regular solution of (1.14) with (1.15).

    Lemma 3.2. For any R > 0 and z_\tau = (u_\tau, \eta^\tau)\in \mathscr{B}_\tau(R)\subset \mathcal{M}_\tau . Then, there exist positive two constants \delta_1 and k_1 , such that

    \begin{equation*} |u|^2_2+\varepsilon(t)|\nabla u|^2_2+\|\eta^t\|^2_{\mu, 0}\leq Q(R)e^{-\delta_1(t-\tau)}+k_1 \end{equation*}

    holds for any \tau\leq t .

    Proof. Multiplying the first equation of (1.14) by u in L^2(\mathbb{R}^n) and the second one by \eta^t in \mathcal{V}_0 , then we obtain

    \begin{align} \frac{1}{2}&\frac{d}{dt}\left(|u|_2^2+\varepsilon(t)|\nabla u|_2^2+|\nabla\eta^t|_{\mu,2}^2\right)-\frac{1}{2}\varepsilon'(t)|\nabla u|_2^2-\frac{1}{2}\int_0^\infty\mu'(s)|\nabla\eta^t(s)|_2^2ds\\ +&|\nabla u|_2^2+\int_{\mathbb{R}^n}f_1(x,u)u+\frac{\lambda-\alpha}{2}|u|_2^2\leq\frac{1}{2(\lambda-\alpha)}|g|_2^2, \end{align} (3.4)

    and

    \begin{align} \frac{1}{2} \frac{d}{dt}|\eta^t|_{\mu,2}^2-\frac{1}{2}\int_0^{\infty}\mu'(s)|\eta^t(s)|_2^2ds\leq 2|u|_2^2+\frac{1}{8}|\eta^t(s)|^2_{\mu,2}. \end{align} (3.5)

    Furthermore, we also get

    \begin{align} \frac{1}{2}\frac{d}{dt}\int_0^{\infty}k(s)|\eta^t(s)|_2^2ds+\frac{3}{8}|\eta^t|_{\mu,2}^2\leq 2\theta^2|u|_2^2, \end{align} (3.6)

    and

    \begin{align} \frac{1}{2}\frac{d}{dt}\int_0^{\infty}k(s)(\nabla\eta^t(s),\nabla\eta^t(s))ds+\frac{1}{4}|\nabla\eta^t|_{\mu,2}^2\leq \theta^2|\nabla u|_2^2. \end{align} (3.7)

    From (3.5) and (3.6), one gets

    \begin{align} \frac{1}{2}\frac{d}{dt}\left(|\eta^t|_{\mu,2}^2+|\eta^t|_{k,2}^2\right)+\frac{1}{4}|\eta^t|_{\mu,2}^2\leq 2(1+\theta^2)|u|_2^2. \end{align} (3.8)

    Taking \sigma = \frac{\lambda-\alpha}{8(1+\theta^2)} and let

    \begin{align*} E(t) = &\frac{1}{2}\left[|u|_2^2+\varepsilon(t)|\nabla u|_2^2+|\nabla\eta^t|_{\mu,2}^2+\sigma\left(|\eta^t|_{\mu,2}^2+|\eta^t|_{k,2}^2\right)+\frac{1}{2\theta^2}|\nabla\eta^t|_{k,2}^2\right];\\ H(t) = &-\frac{1}{2}\varepsilon'(t)|\nabla u|_2^2-\frac{1}{2}\int_0^\infty\mu'(s)|\nabla\eta^t(s)|_2^2ds +\frac{1}{2}|\nabla u|_2^2+\frac{1}{8\theta^2}|\nabla\eta^t|_{\mu,2}^2\\&+\int_{\mathbb{R}^n}f_1(x,u)u+\frac{\lambda-\alpha}{4}|u|_2^2+\frac{\sigma}{4}|\eta^t|_{\mu,2}^2. \end{align*}

    Then combining (3.4) and (3.5), it follows that

    \begin{equation} \frac{d}{dt}E(t)+H(t)\leq \frac{1}{2(\lambda-\alpha)}|g|_2^2. \end{equation} (3.9)

    Furthermore, let

    \begin{equation*} a_0 = \max\{1+\frac{1}{2\theta}, \sigma(1+\theta)\}, \,\text{ and }\, \widetilde{a}_0 = \min\{1, \sigma\}. \end{equation*}

    Then we get

    \begin{equation} E(t) \leq a_0\left(|u|_2^2+\varepsilon(t)|\nabla u|_2^2+\|\eta^t\|_{\mu,0}^2\right), \end{equation} (3.10)
    \begin{equation} E(t)\geq \widetilde{a}_0\left(|u|_2^2+\varepsilon(t)|\nabla u|_2^2+\|\eta^t\|_{\mu,0}^2\right), \end{equation} (3.11)

    and

    \begin{align} H(t) \geq& \frac{1}{2}|\nabla u|_2^2+\frac{1}{8\theta^2}|\nabla\eta^t|_{\mu,2}^2+\int_{\mathbb{R}^n}f_1(x,u)u+\frac{\lambda-\alpha}{4}|u|_2^2+\frac{\sigma}{4}|\eta^t|_{\mu,2}^2\\ \geq & \frac{\lambda-\alpha}{4}|u|_2^2+\frac{\varepsilon(t)}{2L}|\nabla u|_2^2+\frac{1}{8\theta^2}|\nabla\eta^t|_{\mu,2}^2+\frac{\sigma}{4}|\eta^t|_{\mu,2}^2+\int_{\mathbb{R}^n}\left(f_1(x,u)u+\varphi(x)\right)-|\varphi|_1\\ \geq & \delta_1a_0\left(|u|_2^2+\varepsilon(t)|\nabla u|_2^2+\|\eta^t\|_{\mu,0}^2\right)+\int_{\mathbb{R}^n}\left(f_1(x,u)u+\varphi(x)\right)-|\varphi|_1\\ \geq & \delta_1E(t)-|\varphi|_1, \end{align} (3.12)

    where \delta_1 = \frac{1}{2a_0}\min\{\frac{1}{L}, \frac{1}{4\theta^2}, \frac{\sigma}{2}, \frac{\lambda-\alpha}{2}\} .

    Then from (3.9) and (3.12) we get

    \begin{equation*} \frac{d}{dt}E(t)+\delta_1 E(t)\leq \frac{1}{2(\lambda-\alpha)}|g|_2^2+|\varphi|_1. \end{equation*}

    Applying Grönwall lemma, we have

    \begin{equation*} E(t)\leq E(\tau)e^{-\delta_1(t-\tau)}+\frac{1}{2\delta_1(\lambda-\alpha)}|g|_2^2+\frac{1}{\delta_1}|\varphi|_1. \end{equation*}

    By (3.11), it follows that

    \begin{equation*} |u|^2_2+\varepsilon(t)|\nabla u|^2_2+\|\eta^t\|^2_{\mu, 0}\leq Q(R)e^{-\delta_1(t-\tau)}+\frac{1}{\widetilde{a}_0}\left(\frac{1}{2\delta_1(\lambda-\alpha)}|g|_2^2+\frac{1}{\delta_1}|\varphi|_1\right). \end{equation*}

    Let

    \begin{equation*} k_1 = \frac{1}{\widetilde{a}_0}\left(\frac{1}{2\delta_1(\lambda-\alpha)}|g|_2^2+\frac{1}{\delta_1}|\varphi|_1\right). \end{equation*}

    Then the proof is complete.

    Corollary 3.1. For any given R\in\mathbb{R}^+ , let z_{\tau} = (u_{\tau}, \eta^{\tau})\in\mathscr{B}_{\tau}(R) \subset\mathcal{M}_{\tau} , then the process U(t, \tau) corresponding to Eq (1.14) with (1.15) possesses a time-dependent bounded absorbing set \tilde{B_0} = \{B^0_t = \mathscr{B}_{t}(\rho_0)\}_{t\in \mathbb{R}} , that is there exists t_0 = t_0(t, R) < t , such that

    \begin{equation*} U(t,\tau)\mathscr{B}_{\tau}(R)\subset B^0_t, \quad \forall \, \tau\leq t_0. \end{equation*}

    In fact, let t_0 = t_0(t, R) = t- \frac{1}{\delta_1}\ln\frac{Q(R)}{k_1} < t and \rho_0 = 2k_1 , then the conclusion can be directly obtained from Lemma 3.2.

    For brevity, later in this article, let \tilde{B}^0 be the bounded uniformly absorbing set obtained in Lemma 3.2, i.e.,

    \begin{equation} \tilde{B}^0 = \{B_t: z = (u,\eta^t)\in B_t,\, \|z\|^2_{\mathcal{M}_t} = |u|_2^2+\varepsilon(t)|\nabla u|^2_{2}+\|\eta^t\|^2_{\mu, 0}\leq \rho_0\}_{t\in \mathbb{R}}. \end{equation} (3.13)

    From Lemma 3.2, it follows that:

    Corollary 3.2. For any R > 0 and z_{\tau} = (u_{\tau}, \eta^{\tau})\in\mathscr{B}_{\tau}(R) \subset\mathcal{M}_{\tau} , then there exists \mathcal{K}_0 = \mathcal{K}_0(R) such that

    \begin{equation*} |u(t)|_2^2 + \varepsilon(t)|\nabla u(t)|_2^2+\|\eta^t \|_{\mu ,0}^2\leq \mathcal{K}_0 \end{equation*}

    holds for all t-\tau \geq 0 .

    Corollary 3.3. For any R > 0 and z_{\tau} = (u_{\tau}, \eta^{\tau})\in\mathscr{B}_{\tau}(R) \subset\mathcal{M}_{\tau} , then there exists \rho_1 = \rho_1(k_1) > 0 , such that the following estimate:

    \begin{eqnarray*} \int_t^{t+1}\left(|u(s)|^2_{2}+ |\nabla u(s)|^2_{2}+ \int_{\mathbb{R}^n}f_1(x,u(s))u(s)\right)ds \leq Q(R)e^{-\delta_1(t-\tau)}+\rho_1 \end{eqnarray*}

    holds for any \tau\leq t .

    Proof. By (3.4), it's easy to obtain

    \begin{align} &\frac{1}{2}\frac{d}{dt}\left(|u|_2^2+\varepsilon(t)|\nabla u|_2^2+|\nabla\eta^t|_{\mu,2}^2\right)+\frac{\lambda-\alpha}{2}|u|_2^2+|\nabla u|_2^2\\ +&\int_{\mathbb{R}^n}\left(f_1(x,u)u+\varphi(x)\right)\leq\frac{1}{2(\lambda-\alpha)}|g|_2^2+|\varphi|_1. \end{align} (3.14)

    Let b_1 = \min\{1, \frac{\lambda-\alpha}{2}\} and integrating (3.14) on [t, t+1] , we get

    \begin{align*} &\int_t^{t+1}\left(|u(s)|_2^2+|\nabla u(s)|^2_{2}+\int_{\mathbb{R}^n}\big(f_1(x,u(s))u(s)+\varphi(x)\big)\right)ds\\ \leq &\frac{1}{b_1} \left(\frac{1}{2(\lambda-\beta_1)})|g|^2_2+|\varphi|_1+\frac{1}{2}\left(|u(t)|_2^2+\varepsilon(t)|\nabla u(t)|_2^2+|\nabla\eta^t|_{\mu,2}^2\right)\right)\\ \leq&Q(R)e^{-\delta_1(t-\tau)}+\frac{1}{b_1} \left(\frac{1}{2(\lambda-\beta_1)})|g|^2_2+|\varphi|_1\right)+\frac{1}{2b_1} k_1. \end{align*}

    Let \rho_1 = \frac{1}{b_1} \left(\frac{1}{2(\lambda-\beta_1)})|g|^2_2+|\varphi|_1\right)+\frac{1}{2b_1} k_1+|\varphi|_1 and \tau\leq t , then the proof is complete.

    Lemma 3.3. For any R > 0 and z_\tau = (u_\tau, \eta^\tau)\in \mathscr{B}_\tau(R)\subset \mathcal{M}_\tau . Then, there exist positive two constants \delta_2 and k_2 , such that

    \begin{equation*} |u(t)|^2_2+|\nabla u(t)|^2_2+\int_{\mathbb{R}^n}F_1(x,u)\leq Q(R)e^{-\delta_2(t-\tau)}+k_2 \end{equation*}

    holds for any \tau < t .

    Proof. We now multiply the first equation of (1.14) by u_t in L^2(\mathbb{R}^n) , then we get

    \begin{align} &\frac{1}{2}\frac{d}{dt}\left[(\lambda-\alpha)|u|_2^2+|\nabla u|_2^2+ 2\int_{\mathbb{R}^n} F_1(x,u)+2\int_0^{\infty}\mu(s)\left(\nabla u,\nabla\eta^t(s)\right)ds\right] +\frac{1}{2}|u_t|^2_2+\frac{\varepsilon(t)}{2}|\nabla u_t|^2_2 \\ \leq &2|\nabla u|_2^2-\frac{\mu^2(0)}{4}\int_0^{\infty}\mu'(s)|\nabla\eta^t(s)|_2^2ds+\frac{1}{2}|g|_2^2. \end{align} (3.15)

    Setting

    \begin{align*} \mathcal{F}(u) = &\int_{\mathbb{R}^n}\left(F_1(x,u)-f_0(x)u+\frac{l-\alpha}{2}|u|^2\right)\geq 0,\\ N(t) = &\frac{1}{2}\left[(\lambda-\alpha)|u|_2^2+|\nabla u|_2^2+ 2\int_{\mathbb{R}^n} F_1(x,u)+2\int_0^{\infty}\mu(s)\left(\nabla u,\nabla\eta^t(s)\right)ds\right]. \end{align*}

    From Lemma 3.2 and Corollary 3.2, we get that for any t > \tau , there exists t^{\star}\in(\tau, t] such that

    \begin{equation*} |u(t^{\star})|_2^2+|\nabla u(t^{\star})|_2^2+|\nabla\eta^{t^{\star}}|_{\mu,2}^2+\int_{\mathbb{R}^n}\left(f_1(x,u(t^{\star}))u(t^{\star})+\varphi(x)\right) \leq Q(R). \end{equation*}

    Furthermore, associating with (1.10), it is easy to obtain the following inequality:

    \begin{equation} |u(t^{\star})|_2^2+|\nabla u(t^{\star})|_2^2+|\nabla\eta^{t^{\star}}|_{\mu,2}^2+ \mathcal{F}(u(t^{\star}))\leq Q(R) \end{equation} (3.16)

    holds for some t^{\star}\in (\tau, t] . Let

    \begin{equation*} B_{\beta}(t) = \frac{1}{2}\left(|u(t)|_2^2+\varepsilon(t)|\nabla u(t)|_2^2+|\nabla\eta^t|_{\mu,2}^2\right)+ \beta N(t)+ \frac{\beta}{2\theta^2}\int_0^{\infty}k(s)|\nabla\eta^t(s)|_{2}^2ds, \end{equation*}

    where

    \begin{equation*} \beta = \frac{1}{2}\min\{\frac{1}{1+|\lambda-l|},\frac{1}{4}, \frac{2}{\mu^2(0)},\sqrt{4|\alpha-l|^2+2(\lambda-\alpha)}-2|\alpha-l| \} < \frac{1}{4}. \end{equation*}

    Then, 1-(1+|\lambda-l|)\beta > \frac{1}{2}, 1-2\beta > \frac{1}{2} . Let

    \begin{align*} a_1 & = \frac{1}{2}\min\{1-(1+|\lambda-l|)\beta,2\beta, 1-2\beta \},\\ \tilde{a}_1 & = \frac{1}{2}\max\{1+(1+|\lambda-l|)\beta,2\beta, L+2\beta, 1+\beta+\frac{\beta}{\theta}\}. \end{align*}

    Therefore, we get

    \begin{align} B_{\beta}(t)& \geq \frac{1}{2}\left(1-(1+|\lambda-l|\beta)\right)|u(t)|_2^2+\frac{\beta}{4}|\nabla u(t)|_2^2 +\frac{1}{2}(1-2\beta)|\nabla\eta^t|_{\mu,2}^2+ \beta\mathcal{F}(u)-\frac{\beta}{2}|f_0|_2^2\\ & \geq a_1\left(|u(t)|_2^2+|\nabla u(t)|_2^2+|\nabla\eta^t|_{\mu,2}^2+\mathcal{F}(u)\right)-\frac{\beta}{2}|f_0|_2^2 \end{align} (3.17)

    and

    \begin{align} B_{\beta}(t) \leq& \frac{1}{2}\left(1+(1+|\lambda-l|\beta)\right)|u(t)|_2^2+\frac{1}{2}(L+2\beta)|\nabla u(t)|_2^2 + \frac{1}{2}(1+\beta+\frac{\beta}{\theta})|\nabla\eta^t|_{\mu,2}^2+ \beta\mathcal{F}(u)+\frac{\beta}{2}|f_0|_2^2\\ \leq &\tilde{a}_1\left( |u(t)|_2^2+|\nabla u(t)|_2^2+|\nabla\eta^t|_{\mu,2}^2+\mathcal{F}(u)\right)+\frac{1}{4}|f_0|_2^2. \end{align} (3.18)

    Combining with (3.4) and (3.15), it follows that

    \begin{align} &\frac{d}{dt}B_{\beta}(t)+\frac{1}{2}|u_t|^2_2+\frac{\varepsilon(t)}{2}|\nabla u_t|^2_2-\frac{1}{2}\varepsilon'(t)|\nabla u|_2^2-\frac{1}{2}\left(1-\frac{\mu^2(0)\beta}{2}\right)\int^{\infty}_0\mu'(s)|\nabla\eta^t(s)|_2^2ds\\ +&(1-3\beta)|\nabla u|_2^2+\frac{\lambda-\alpha}{2}|u|_2^2+\frac{\beta}{4\theta^2}|\nabla\eta^t|_{\mu,2}^2+\int_{\mathbb{R}^n}f_1(x,u)u\leq C|g|^2_2. \end{align} (3.19)

    Let

    \begin{align*} \tilde{B}_{\beta}(t) = &\frac{1}{2}|u_t|^2_2+\frac{\varepsilon(t)}{2}|\nabla u_t|^2_2-\frac{1}{2}\varepsilon'(t)|\nabla u|_2^2-\frac{1}{2}\left(1-\frac{\mu^2(0)\beta}{2}\right)\int^{\infty}_0\mu'(s)|\nabla\eta^t(s)|_2^2ds\\ &+(1-3\beta)|\nabla u|_2^2+\frac{\lambda-\alpha}{2}|u|_2^2+\frac{\beta}{4\theta^2}|\nabla\eta^t|_{\mu,2}^2+\int_{\mathbb{R}^n}f_1(x,u)u. \end{align*}

    Then,

    \begin{align*} \tilde{B}_{\beta}(t)\geq& (1-3\beta)|\nabla u|_2^2+\frac{\lambda-\alpha}{2}|u|_2^2+\frac{\beta}{4\theta^2}|\nabla\eta^t|_{\mu,2}^2+\int_{\mathbb{R}^n}f_1(x,u)u\\ \geq&(1-3\beta)|\nabla u|_2^2+\frac{\lambda-\alpha}{2}|u|_2^2+\frac{\beta}{4\theta^2}|\nabla\eta^t|_{\mu,2}^2+\beta\int_{\mathbb{R}^n}\left(f_1(x,u)u+\varphi(x)\right)-|\varphi|_1\\ \geq&(1-3\beta)|\nabla u|_2^2+\frac{\lambda-\alpha}{2}|u|_2^2+\frac{\beta}{4\theta^2}|\nabla\eta^t|_{\mu,2}^2 + \beta\mathcal{F}(u)+\beta\int_{\mathbb{R}^n}\left(f_0(x)u+(\alpha-l)u^2+\varphi(x)\right)-|\varphi|_1\\ \geq&(1-3\beta)|\nabla u|_2^2+\left(\frac{\lambda-\alpha}{2}-|\alpha-l|\beta-\frac{\beta^2}{4}\right)|u|_2^2 + \frac{\beta}{4\theta^2}|\nabla\eta^t|_{\mu,2}^2+\beta\mathcal{F}(u)-C(|f_0|_2^2+|\varphi|_1). \end{align*}

    Then, there is constant \delta_2 > 0 , such that

    \begin{equation} \tilde{B}_{\beta}(t)\geq\delta_2B_{\beta}(t)-C(|f_0|_2^2+|\varphi|_1). \end{equation} (3.20)

    Thus, we can rewrite (3.17) as follows:

    \begin{equation*} \frac{d}{dt}B_{\beta}(t)+\delta_2 B_{\beta}(t)\leq C(|g|_2^2+|f_0|_2^2+|\varphi|_1). \end{equation*}

    Applying Grönwall lemma, we have

    \begin{align} & B_{\beta}(t)\leq B_{\beta}(t^{\star})e^{-\delta_2(t-t^{\star})}+C(|\varphi|_1+|g|^2_2+|f_0|_2^2) \\ \leq &\tilde{a}_1\left( |u(t^{\star})|_2^2+|\nabla u(t^{\star})|_2^2+|\nabla\eta^{t^{\star}}|_{\mu,2}^2+\mathcal{F}(u(t^{\star}))\right)e^{-\delta_2(t-\tau)} + C(|\varphi|_1+|g|^2_2+|f_0|_2^2)\\ \leq& Q(R)e^{-\delta_2(t-\tau)}+C(|\varphi|_1+|g|^2_2+|f_0|_2^2) \end{align} (3.21)

    holds for any \tau < t . Then the proof is complete.

    In subsection, we will prove the existence of time-dependent global attractors in \{\mathcal{M}_t\}_{t\in\mathbb{R}} through the process U(t, \tau) defined by (3.3). In order to prove Theorem 3.1, we first give the following lemmas.

    Lemma 3.4. For any R > 0 and z_{\tau} = (u_{\tau}, \eta^{\tau})\in\mathscr{B}_{\tau}(R) \subset\mathcal{M}_{\tau} , then there exists a positive constant \mathcal{K}_1 = \mathcal{K}_1(R) , such that

    \begin{eqnarray*} \int^{t+1}_{t}\big(|u_{t}(s)|_2^2+\varepsilon(s)|\nabla u_{t}(s)|_2^2\big)ds \leq \mathcal{K}_1 \end{eqnarray*}

    holds for any t\geq \tau .

    Proof. From (3.19) and the value of \beta , one gets

    \begin{equation} \frac{d}{dt}B_{\beta}(t)+\frac{1}{2}|u_t|^2_2+\frac{\varepsilon(t)}{2}|\nabla u_t|^2_2\leq C(|\varphi|_1+|g|^2_2+|f_0|_2^2). \end{equation} (3.22)

    Integrating t from t to t+1 on both sides of (3.22), and organizing it to obtain

    \begin{align*} &\int_t^{t+1}\left(|u_t(s)|^2_2+\varepsilon(s)|\nabla u_t(s)|^2_2\right)ds\leq B_{\beta}(t)+C(|\varphi|_1+|g|^2_2+|f_0|_2^2)\\ \leq& Q(R)e^{-\delta_2(t-\tau)}+C(|\varphi|_1+|g|^2_2+|f_0|_2^2)+a_1\beta|f_0|_2^2\\ \leq& Q(R)+C(|\varphi|_1+|g|^2_2+|f_0|_2^2)+a_1\beta|f_0|_2^2 \end{align*}

    holds for any \tau < t . Let \mathcal{K}_1 = Q(R)+C(|\varphi|_1+|g|^2_2+|f_0|_2^2)+a_1\beta|f_0|_2^2 , then the proof is complete.

    Lemma 3.5. Let B be any bounded subset \mathcal{M}_{\tau} and z_{\tau}\in B . Then for any \varepsilon > 0 , there exist positive constants K_1 large enough and t_1\leq t , such that

    \begin{equation*} \int_{{B}_{k}^c}\left(|u|^2+\varepsilon(t)|\nabla u|^2\right)+\int_0^{\infty}\mu(s)\int_{B_{k}^c}|\nabla \eta^t(s)|^2ds\leq C\varepsilon \end{equation*}

    holds for every \tau\leq t_1 \leq t and k\geq K_1 , where B_{k}^c = \{x\in\mathbb{R}^n: |x|\geq k\} .

    Proof. Let \theta(\cdot)\in C^{\infty}(\mathbb{R}^+) satisfies

    \begin{equation*} 0\leq \theta(s)\leq 1, \ \ \, 0\leq \theta'(s)\leq b, \; \forall s\in \mathbb{R}^+, \end{equation*}

    and

    \begin{equation*} \theta(s) = 0, \; \forall s\in [0, 1), \ \ \theta(s) = 1,\; \forall s\in [2, \infty), \end{equation*}

    where b > 0 is a constant. Setting \theta_k = \theta(\frac{2|x|^2}{k^2}) and multiplying the first equation of (1.14) by \theta_k^2u in L^2(\mathbb{R}^n) , then we obtain

    \begin{align} &\frac{d}{dt}E_k(t)-\frac{\varepsilon'(t)}{2}|\theta_k\nabla u|_2^2-\frac{1}{2}\int_0^\infty\mu'(s)|\theta_k\nabla\eta^t(s)|_2^2ds+|\theta_k\nabla u|_2^2+\frac{\lambda-\alpha}{2}|\theta_ku|_2^2\\ \leq&-\frac{8}{k^2}\int_0^{\infty}\mu(s)\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla \eta^t(s))ds-\frac{8\varepsilon(t)}{k^2}\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla u_t)\\ &-\frac{8}{k^2}\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla u)+\frac{1}{2(\lambda-\alpha)}|\theta_kg|_2^2-\int_{\mathbb{R}^n}\theta^2_k f_1(x,u)u, \end{align} (3.23)

    where E_k(t) = \frac{1}{2}\left[|\theta_ku|_2^2+\varepsilon(t)|\theta_k\nabla u|_2^2+|\theta_k\nabla\eta^t|_{\mu, 2}^2\right] . Similar to (3.7), we define the function N_k(t) as follows:

    \begin{equation} N_k(t) = \int_0^{\infty}k(s)\left(\theta_k\nabla\eta^t,\theta_k\nabla\eta^t\right)ds. \end{equation} (3.24)

    Then it is follows that

    \begin{align} &\frac{\mathrm{d}}{\mathrm{d}t}N_{k}(t)+\frac{1}{2}|\theta_k\nabla \eta^t|_2^{2}\leq 2\theta^2|\theta_k\nabla u|_2^2. \end{align} (3.25)

    We can define a functional with a undetermined coefficient \kappa as follows:

    \begin{equation} H_k(t) = E_k(t)+\frac{\kappa}{2} N_k(t). \end{equation} (3.26)

    Let 0 < \kappa < \frac{1}{2\theta^2} be sufficiently small, then it follows that

    \begin{equation} H_k(t)\geq \frac{1}{2}\left(|\theta_ku|_2^2+\varepsilon(t)|\theta_k\nabla u|_2^2+|\theta_k\nabla\eta^t|_{\mu,2}^2\right), \end{equation} (3.27)

    and

    \begin{equation} H_k(t)\leq \frac{1+\theta\kappa}{2}\left(|\theta_ku|_2^2+\varepsilon(t)|\theta_k\nabla u|_2^2+|\theta_k\nabla\eta^t|_{\mu,2}^2\right). \end{equation} (3.28)

    Combining with (3.23), (3.25) and (3.26), we obtain

    \begin{align} &\frac{d}{dt}H_k(t)+\frac{\lambda-\alpha}{2}|\theta_ku|_2^2+\frac{1}{2}(1-\theta^2\kappa)|\theta_k\nabla u|_2^2+\frac{\kappa}{4}|\theta_k\nabla\eta^t|^2_{\mu,2}\\ \leq&-\frac{8}{k^2}\int_0^{\infty}\mu(s)\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla \eta^t(s))ds-\frac{8\varepsilon(t)}{k^2}\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla u_t)\\ &+\frac{64}{k^4}\int_{\mathbb{R}^n}(\theta'_k)^2|u|^2|x|^2+\frac{1}{2(\lambda-\alpha)}|\theta_kg|_2^2-\int_{\mathbb{R}^n}\theta_k^2f_1(x,u)u. \end{align} (3.29)

    From (1.9), there is

    \begin{align*} -\int_{\mathbb{R}^n}\theta_k^2f_1(x,u)u & = -\int_{\mathbb{R}^n}\theta_k^2\left(f_1(x,u)u+\varphi(x)\right)+\int_{\mathbb{R}^n}\theta_k^2\varphi(x)\\ \leq&-\int_{\mathbb{R}^n}\theta_k^2\left(f_1(x,u)u+\varphi(x)\right)+|\theta_k\varphi|_1\\ \leq &|\theta_k\varphi|_1. \end{align*}

    Let \delta_3 = \min\{\lambda-\alpha, \frac{1-\theta^2\kappa}{L}, \frac{\kappa}{2}\} , then we have that

    \begin{align} &\frac{d}{dt}H_{k}(t)+\frac{\delta_3}{2}\left(|\theta_ku|^2_2+\varepsilon(t)|\theta_k\nabla u|^2_2+|\theta_k\nabla\eta^t|^2_{\mu, 2}\right)\\ \leq&-\frac{8}{k^2}\int_0^{\infty}\mu(s)\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla \eta^t(s))ds-\frac{8\varepsilon(t)}{k^2}\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla u_t)\\ &-\frac{8}{k^2}\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla u)+\frac{1}{2(\lambda-\alpha)}|\theta_kg|_2^2+|\theta_k\varphi|_1. \end{align} (3.30)

    Next, we will estimate each item on the right side of (3.30). According to the definition of \theta , it can be seen that

    \begin{equation*} \theta_k' = 0, \quad |x| < \frac{k}{2}\text{ or } |x| > k. \end{equation*}

    Therefore,

    \begin{align*} \frac{64}{k^4} \int_{\mathbb{R}^n}(\theta'_k)^2|u|^2|x|^2 = & \frac{64}{k^4}\int_{\frac{k}{2}\leq|x|\leq k}(\theta'_k)2|u|^2|x|^2 \\ \leq & \frac{64}{k^2}\int_{\frac{k}{2}\leq|x|\leq k}b^2|u|^2\\ \leq &\frac{C}{k^2}|u|_2^2. \end{align*}

    Similarly, we also can obtain the following estimates:

    \begin{align*} -\frac{8}{k^2}\int_0^{\infty}\mu(s)\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla \eta^t(s))ds & \leq \frac{C}{k}\left(|u|_2^2+|\nabla\eta^t|^2_{\mu,2}\right); \\ -\frac{8\varepsilon(t)}{k^2}\int_{\mathbb{R}^n}\theta_k\theta'_ku(x\cdot\nabla u_t(s))& \leq \frac{C}{k}\left(|u|_2^2+\varepsilon(t)|\nabla u_t|^2_{2}\right). \end{align*}

    Combining with (3.28), then (3.30) can be rewritten as follows:

    \begin{align} &\frac{d}{dt}H_{k}(t)+\delta_2H_k(t)\\ \leq &\frac{C}{k}\left(|u|_2^2+|\nabla\eta^t|^2_{\mu,2}+\varepsilon(t)|\nabla u_t|^2_{2}\right) + C\left(|\theta_kg|^2_2+|\theta_k\varphi|_1\right). \end{align} (3.31)

    By Grönwall lemma, it follows that

    \begin{align*} & H_{k}(t) \\ \leq& H_{k}(\tau)e^{-\delta_3(t-\tau)}+C\left(|\theta_kg|^2_2+|\theta_kf_0|^2_2+|\theta_k\varphi|_1\right) + \frac{C}{k}e^{-\delta_3t} \int^{t}_{\tau}e^{\delta_2s}\left(|u(s)|_2^2+|\nabla\eta^s|^2_{\mu,2}+\varepsilon(s)|\nabla u_t(s)|^2_{2}\right)ds. \end{align*}

    Combining with Lemma 3.3, we have

    \begin{equation*} H_{k}(t)\leq Q(R)e^{-\delta_3(t-\tau)}+C\left(|\theta_kg|^2_2+|\theta_kf_0|^2_2+|\theta_k\varphi|_1\right) +\frac{C}{k} \mathcal{K}_0, \end{equation*}

    and for any \varepsilon > 0 , there exists K_1\geq 0 large enough, such that for every k\geq K_1 ,

    \begin{align*} &\left(|\theta_kg|^2_2+|\theta_kf_0|^2_2+|\theta_k\varphi|_1\right) = \int_{B_k^c}\left(|\theta_kg|^2+|\theta_kf_0|^2+|\theta_k\varphi|\right) < C\varepsilon;\\ &\frac{C}{k}\mathcal{K}_0 < C\varepsilon, \end{align*}

    and let t_1 = t -\frac{1}{\delta_3}\ln \frac{Q(R)}{\varepsilon} , for every \tau\leq t_1\leq t , it follows that

    \begin{equation*} Q(R)e^{-\delta_3(t-\tau)} < C\varepsilon. \end{equation*}

    By (3.27), then,

    \begin{equation*} |\theta_ku|_2^2+\varepsilon(t)|\theta_k\nabla u|_2^2+|\theta_k\nabla\eta^t|_{\mu,2}^2+\varepsilon(t)\mathcal{F}_1(\theta_ku)\leq C\varepsilon \end{equation*}

    holds for every \tau\leq t_1\leq t and k\geq K_1 . So we have that

    \begin{equation*} \int_{{B}_{k}^c}\left(|u|^2+\varepsilon(t)|\nabla u|^2\right)+\int_0^{\infty}\mu(s)\int_{B_{k}^c}|\nabla \eta^t(s)|^2ds\leq C\varepsilon. \end{equation*}

    In order to obtain the asymptotic regularity estimates later, we decompose the solution U(t, \tau)z_{\tau} = (u(t), \eta^t) into the following sum:

    \begin{equation} U(t, \tau)z_{\tau} = U_1(t,\tau)z_{\tau}+K(t,\tau)z_{\tau}, \end{equation} (3.32)

    where U_1(t, \tau)z_{\tau} = (v(t), \xi^t) and K(t, \tau)z_{\tau} = (\omega(t), \zeta^t) solve the following equations respectively:

    \begin{eqnarray} \left\{ \begin{aligned}\ &v_{t}-\varepsilon(t)\Delta v_t-\Delta v-\int_{0}^{\infty}\mu(s)\Delta\xi^t(s)ds+f(x,u)-f(x,\omega)+(\lambda+l) v = 0, \\ &\xi_t^t = -\xi_s^t+v, \end{aligned} \right. \end{eqnarray} (3.33)

    where l is a constant from (1.5). It has an initial data

    \begin{equation} v(x,\tau) = u_{\tau}(x),\,\xi^{\tau}(x,s) = \int_{{0}}^{s}u_{\tau}(x,{\tau}-\theta)d\theta, \quad (x,s)\in\mathbb{R}^n\times \mathbb{R}^+. \end{equation} (3.34)

    And

    \begin{eqnarray} \left\{ \begin{aligned}\ &\omega_{t}-\varepsilon(t)\Delta \omega_t-\Delta \omega-\int_{0}^{\infty}\mu(s)\Delta \zeta^t(s)ds+f(x,\omega)+(\lambda+l)\omega = lu+g, \\ &\zeta_t^t = -\zeta_s^t+\omega, \end{aligned} \right. \end{eqnarray} (3.35)

    with the initial value conditions

    \begin{equation} \omega(x,\tau) = 0,\,\zeta^{\tau}(x,s) = 0, \quad (x,s)\in\mathbb{R}^n\times \mathbb{R}^+. \end{equation} (3.36)

    Lemma 3.6. Assume that K(t, \tau)z_{\tau} = (\omega(t), \zeta^t) is the solution of Eq (3.35) with (3.36). Then there exist positive constants k_i(i = 3, 4) , such that

    \begin{equation*} |\omega(t)|_{2}^2 +|\nabla\omega(t)|^2_{2}+\|\zeta^t\|^2_{\mu,0}+\mathcal{F}(\omega(t))\leq k_3 \end{equation*}

    and

    \begin{equation*} \int^{t+1}_t\left(|\omega_t(s)|_{2}^2 +\varepsilon(s)|\nabla\omega_t(s)|^2_{2}\right)ds \leq k_4 \end{equation*}

    hold for any t-\tau > 0 .

    Proof. This proof can imitate the proof of Lemma 3.2, Corollary 3.3 and Lemma 3.4 word by word, it should be noted that (3.17), (3.21) and \|(\omega(\tau), \zeta^{\tau})\|_{\mathcal{M}_{\tau}} = 0 . Then hence is omitted.

    Lemma 3.7. Assume that U_1(t, \tau)z_{\tau} = (v(t), \xi^t) is the solution of Eq (3.33) with (3.34). Then,

    \begin{equation*} \lim\limits_{\tau\rightarrow -\infty}\left( \|U_1(t,\tau)z_{\tau}\|^2_{\mathcal{M}_t}\right) = 0 \end{equation*}

    holds for every t\in \mathbb{R}^\tau fixed.

    Proof. Multiplying the first equation of (3.33) by v(t) and integrating in L^2(\mathbb{R}^n) and the second one by \xi^t in \mathcal{V} , we have

    \begin{align} \begin{split} \frac{1}{2}\frac{d}{dt} \left(|v|^2_2+\varepsilon(t)|\nabla v|^2_2+|\nabla\xi^t|^2_{\mu,2}\right)+\frac{1}{2}\int^\infty_0-\mu'(s)|\nabla\xi^t(s)|^2_2ds -\varepsilon{'}(t)|\nabla v|^2_2+|\nabla v|^2_2+\lambda|v|^2_2\leq 0, \end{split} \end{align} (3.37)

    and

    \begin{equation} \frac{1}{2}\frac{d}{dt}\int^\infty_0\mu(s)|\xi^t(s)|^2_2ds-\frac{1}{2}\int^\infty_0\mu'(s)|\xi^t(s)|^2_2ds\leq |v|_2|\xi^t|_{\mu,2}. \end{equation} (3.38)

    Furthermore,

    \begin{equation} \frac{1}{2}\frac{d}{dt}\int^\infty_0k(s)|\xi^t(s)|^2_2ds+\frac{1}{4}|\xi^t(s)|^2_{\mu,2}\leq \theta^2|v|_2^2. \end{equation} (3.39)

    Similar to (3.7), we introduce the functional

    \begin{equation*} N_1(t) = \int_0^{\infty}k(s)(\nabla \xi^t(s),\nabla\xi^t(s))ds. \end{equation*}

    Then,

    \begin{align} \frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}N_{1}(t)+\frac{1}{4}|\nabla\xi(t)|_{\mu,2}^{2}&\leq\theta^2|\nabla v(t)|^2_{2}. \end{align} (3.40)

    Setting up

    \begin{equation} \gamma = \frac{1}{2(2+\theta^2)}. \end{equation} (3.41)

    Furthermore, we define the energy-like functional

    \begin{equation} D_{\gamma}(t,\tau) = \frac{1}{2}\left(|v|^2_2+\varepsilon(t)|\nabla v|^2_2+|\nabla\xi^{t}|^2_{\mu,2}+\gamma|\xi^{t}|^2_{\mu,2}+\gamma\int_0^{\infty}k(s)|\xi^{t}(s)|_2^2ds+\gamma N_1(t)\right). \end{equation} (3.42)

    Then we get

    \begin{equation} D_{\gamma}(t,\tau)\leq a_3\left(|v|^2_2+\varepsilon(t)|\nabla v|^2_2+\|\xi^t\|^2_{\mu,0}\right) = a_3\|U_1(t,\tau)z_{\tau}\|^2_{\mathcal{M}_t}, \end{equation} (3.43)

    and

    \begin{equation} D_{\gamma}(t,\tau)\geq \tilde{a}_3\left(|v|^2_2+\varepsilon(t)|\nabla v|^2_2+\|\xi^t\|^2_{\mu,0}\right) = \tilde{a}_3\|U_1(t,\tau)z_{\tau}\|^2_{\mathcal{M}_t}, \end{equation} (3.44)

    where a_3 = \frac{1+\gamma\theta}{2} and \tilde{a}_3 = \frac{\gamma}{2(1+\gamma)} > 0 .

    Therefore, it follows that

    \begin{align} \begin{split} &\frac{d}{dt}D_{\gamma}(t,\tau)-\frac{1}{2}\int^{\infty}_0\mu'(s)|\nabla\xi^t(s)|_2^2ds-\varepsilon'(t)|\nabla v|_2^2 +\frac{\gamma}{4}|\nabla\xi^t|_{\mu,2}^2+\frac{\gamma}{8}|\xi^t|^2_{\mu,2}\\&+(\lambda-2\gamma-\gamma\theta^2)|v|^2_2+\frac{1-\gamma\theta^2}{L}\varepsilon(t)|\nabla v|^2_2\leq 0. \end{split} \end{align} (3.45)

    By (3.41), we have

    \begin{equation*} \lambda-(2+\theta^2)\gamma = \frac{1}{2},\,1-\gamma\theta^2 > \frac{1}{2}. \end{equation*}

    Let \delta_4 = \frac{1}{4a_3}\min\{\frac{\gamma}{8}, \lambda-2\gamma-\gamma\theta^2, \frac{1-\gamma\theta^2}{L}\} , then (3.45) can be rewritten as

    \begin{align} \frac{d}{dt}D_{\gamma}(t,\tau)+\delta_4D_{\gamma}(t,\tau)\leq 0. \end{align} (3.46)

    Applying Grönwall lemma, we get that

    \begin{equation*} D_{\gamma}(t,\tau)\leq D_{\gamma}(\tau,\tau)e^{-\delta_4(t-\tau)}. \end{equation*}

    Then, combining with (3.43) and (3.44), we have

    \begin{equation*} \lim\limits_{\tau\rightarrow -\infty}\|U_1(t,\tau)z_{\tau}\|^2_{\mathcal{M}_t} = 0 \end{equation*}

    is true for any t\in \mathbb{R}^\tau .

    Lemma 3.8. There exists a positive constant \mathcal{K}_3 = \mathcal{K}_3(R) which depends on t and \tau , such that

    \begin{equation} |\nabla\omega(t)|_{2}^2 +\varepsilon(t)|\triangle\omega(t)|^2_{2}+\|\zeta^t\|^2_{\mu,1}\leq \mathcal{K}_3 \end{equation} (3.47)

    holds for any \tau \leq t\in\mathbb{R} .

    Proof. Acting on the first equation of (3.35) by -\Delta\omega(t) on L^2(\mathbb{R}^n) , we have

    \begin{align} \begin{split} &\frac{1}{2}\frac{d}{dt}\left(|\nabla\omega|^2_2+\varepsilon(t)|\Delta\omega|^2_2+|\Delta\zeta^t|^2_{\mu,2}\right)-\frac{\varepsilon'(t)}{2}|\Delta \omega|_2^2-\frac{1}{2}\int_0^{\infty}\mu'(s)|\Delta\zeta^t(s)|_2^2ds\\ &+\lambda|\nabla\omega|_2^2+\frac{1}{2}|\Delta\omega|_2^2\leq l^2|u|^2_2+|g|^2_2. \end{split} \end{align} (3.48)

    Similar to (3.7), we introduce the functional

    \begin{equation*} N_2(t) = \int_0^{\infty}k(s)(\Delta \zeta^t(s),\Delta\zeta^t(s))ds. \end{equation*}

    Then,

    \begin{align} \frac{1}{2}\frac{\mathrm{d}}{\mathrm{d}t}N_{2}(t)+\frac{1}{4}|\Delta\zeta^t|_{\mu,2}^{2}&\leq\theta^2|\Delta\omega|^2_{2}. \end{align} (3.49)

    Therefore, from (3.48) and (3.49), it is that

    \begin{align} \begin{split} &\frac{1}{2}\frac{d}{dt}\left(|\nabla\omega|^2_2+\varepsilon(t)|\Delta\omega|^2_2+|\Delta\zeta^t|^2_{\mu,2}+\frac{1}{4\theta^2}N_2(t)\right)-\frac{\varepsilon'(t)}{2}|\Delta \omega|_2^2-\frac{1}{2}\int_0^{\infty}\mu'(s)|\Delta\zeta^t(s)|_2^2ds\\ & +\frac{1}{16\theta^2}|\Delta\zeta^t|_{\mu,2}^{2}+\lambda|\nabla\omega|_2^2+\frac{1}{4}|\Delta\omega|_2^2 \leq l^2|u|^2_2+|g|^2_2. \end{split} \end{align} (3.50)

    By Corollary 3.2 and the Grönwall lemma, we have

    \begin{equation*} |\nabla\omega(t)|_{2}^2 +\varepsilon(t)|\Delta\omega(t)|^2_{2}+|\Delta\zeta^t|^2_{\mu,2}\leq C\left(\mathcal{K}_0+|g|^2_2\right) \end{equation*}

    holds for any \tau\leq t \in \mathbb{R} . Let

    \begin{equation*} \mathcal{K}_3 = C(\mathcal{K}_0+|g|^2_2), \end{equation*}

    then it follows that

    \begin{equation*} |\nabla\omega(t)|_{2}^2 +\varepsilon(t)|\Delta\omega(s)|^2_{2}+|\Delta\zeta^t|^2_{\mu,2}\leq \mathcal{K}_3 \end{equation*}

    holds for any \tau\leq t\in\mathbb{R} .

    Next, we will verify the existence and regularity of the pullback global attractors \tilde{\mathscr{A}} for Eq (1.1).

    Theorem 3.1. The family of process U(t, \tau) for Eq (1.14) with initial conditions (1.15) is \mathcal{M}_t -contractive process on B_T\in\tilde{B}^0 (from (3.13)).

    Proof. Let z_i(t) = (u_i(t), \xi_i^t) = U(t, \tau)z_{\tau}^i \; (i = 1, 2) be the solutions to Eq (1.14) with the parameter \varepsilon(t) and initial data z^i_{\tau}\in B_{\tau}\in\tilde{B}^0\; (i = 1, 2) ( \tilde{B}^0 is from (3.13)) respectively.

    By (3.32), there is

    \begin{equation*} z_i(t) = U(t,\tau)z_{\tau}^i = U_1(t,\tau)z_{\tau}^i+K(t)z_{\tau}^i = (v_i(t),\xi^t_i)+(\omega_i(t),\zeta^t_i). \end{equation*}

    It gets

    \begin{align} \begin{split} &\|U(t,\tau)z_{\tau}^1-U(t,\tau)z_{\tau}^2\|_{\mathcal{M}_t}^2\\ \leq& 2\|U_1(t,\tau)z_{\tau}^1-U_1(t,\tau)z_{\tau}^2\|_{\mathcal{M}_t}^2+2\left(|\omega_1(t)-\omega_2(t)|_2^2+\varepsilon(t)|\nabla\omega_1(t)-\nabla\omega_2(t)|_2^2+\|\zeta^t_1-\zeta^t_2\|_{\mu,0}^2\right), \end{split} \end{align} (3.51)

    and

    \begin{equation*} \lim\limits_{\tau\rightarrow -\infty}\|U_1(t,\tau)z_{\tau}^1-U_1(t,\tau)z_{\tau}^2\|_{\mathcal{M}_t}^2\leq 2\lim\limits_{\tau\rightarrow -\infty}\left(\|U_1(t,\tau)z_{\tau}^1\|_{\mathcal{M}_t}^2+\|U_1(t,\tau)z_{\tau}^2\|_{\mathcal{M}_t}^2\right) = 0. \end{equation*}

    Then for any \varepsilon > 0 , there is T = T(\varepsilon)\leq t such that

    \begin{equation} 2\|U(t,T)z_{T}^1-U(t,T)z_{T}^2\|_{\mathcal{M}_t}^2 < \varepsilon \end{equation} (3.52)

    holds for any fixed t\geq T .

    Let (\varpi(t), \varsigma^t) = (\omega_1(t)-\omega_2(t), \zeta^t_1-\zeta^t_2) be the solution of the following system:

    \begin{eqnarray*} \left\{ \begin{aligned}\ &\varpi_{t}-\varepsilon(t)\Delta \varpi_t-\Delta \varpi-\int_{0}^{\infty}\mu(s)\Delta \varsigma^t(s)ds+f(x,\omega_1)-f(x,\omega_2)+(\lambda+l) \varpi = l (u_1-u_2), \\ &\varsigma_t^t = -\varsigma_s^t+\varpi. \end{aligned} \right. \end{eqnarray*}

    It has an initial value conditions

    \begin{align*} \varpi(x,\tau) = 0,\,\varsigma^{\tau}(x,s) = 0, \quad (x,s)\in\mathbb{R}^n\times \mathbb{R}^+. \end{align*}

    Next, we also define the functional

    \begin{equation*} \Gamma(t) = \frac{1}{2}\left(|\varpi(t)|_2^2+\varepsilon(t)|\nabla\varpi(t)|_2^2+|\nabla\varsigma^t|_{\mu,2}^2+\frac{1}{2\theta^2}|\nabla\varsigma^t|_{k,2}^2+\frac{\lambda}{2(1+\theta^2)}\left(|\varsigma^t|_{\mu,2}^2+|\varsigma^t|_{k,2}^2\right)\right). \end{equation*}

    Then,

    \begin{align*} \Gamma(t) & \leq a_4\left(|\varpi(t)|_2^2+\varepsilon(t)|\nabla\varpi(t)|_2^2+\|\varsigma^t\|_{\mu,0}^2\right), \\ \Gamma(t) & \geq \tilde{a}_4\left(|\varpi(t)|_2^2+\varepsilon(t)|\nabla\varpi(t)|_2^2+\|\varsigma^t\|_{\mu,0}^2\right), \end{align*}

    where a_4 = \frac{1}{2}\max\{1+\frac{1}{2\theta}, \frac{\lambda(1+\theta)}{2(1+\theta^2)}\} and \tilde{a}_4 = \frac{1}{2}\min\{1, \frac{\lambda}{2(1+\theta^2)}\} .

    Similar to (3.5)–(3.7), we get

    \begin{align} \frac{1}{2} \frac{d}{dt}\left(|\varsigma^t|_{\mu,2}^2+|\varsigma^t|_{k,2}^2\right)+\frac{1}{4}|\varsigma^t|_{\mu,2}^2\leq 2(1+\theta^2)|\varpi|_2^2, \end{align} (3.53)

    and

    \begin{align*} \frac{1}{2}\frac{d}{dt}\int_0^{\infty}k(s)(\nabla\varsigma^t(s),\nabla\varsigma^t(s))+\frac{1}{4}|\nabla\varsigma^t|_{\mu,2}^2\leq \theta^2|\nabla \varpi|_2^2. \end{align*}

    Then it yields

    \begin{align} &\frac{d}{dt}\Gamma(t) -\frac{1}{2}\int^{\infty}_0\mu'(s)|\nabla\varsigma^t(s)|_{2}^2ds-\frac{\varepsilon'(t)}{2}|\nabla\varpi|_2^2+\frac{1}{2}|\nabla\varpi(t)|_2^2\\ &+\frac{\lambda}{4}|\varpi(t)|_2^2+\frac{1}{8\theta^2}|\nabla\varsigma^t|_{\mu,2}^2+\frac{\lambda}{16(1+\theta^2)}|\varsigma^t|_{\mu,2}^2\\ \leq&\frac{l^2}{\lambda}\left(|u_1(t)-u_2(t)|_0^2\right). \end{align} (3.54)

    Let \delta_5 = \frac{1}{2 a_4}\min\{\frac{1}{L}, \frac{1}{4\theta^2}, \frac{\lambda}{8(1+\theta^2)}\} , then it follows that

    \begin{align*} \frac{d}{dt}\Gamma(t)+\delta_5\Gamma(t)\leq\frac{l^2}{\lambda}|u_1(t)-u_2(t)|_0^2. \end{align*}

    So, we get

    \begin{align} &|\varpi(t)|_2^2+\varepsilon(t)|\nabla\varpi(t)|_2^2+\|\varsigma^t\|_{\mu,0}^2 \leq \frac{l^2}{\tilde{a}_4\lambda}e^{-\delta_5t}\int^t_{T}e^{\delta_5s}|u_1(s)-u_2(s)|_0^2ds. \end{align} (3.55)

    And

    \begin{align*} &e^{-\delta_5t}\int^t_{T}e^{\delta_5s}|u_1(s)-u_2(s)|_0^2ds \\ \leq&\int^t_{T}\int_{B_{k}}|u_1(s)-u_2(s)|^2ds+e^{-\delta_5t}\int^t_{T}e^{\delta_5s}\int_{B_{k}^c}|u_1(s)-u_2(s)|^2ds\\ \leq&\int^t_{T}\int_{B_{k}}|u_1(s)-u_2(s)|^2ds+C\varepsilon. \end{align*}

    Then setting

    \begin{align} \varphi_T^{t}(z_1, z_2) = C\int^t_T\int_{B_{k}}\left(|u_{1}(s)-u_{2}(s)|_2^2\right)ds. \end{align} (3.56)

    Combining with Lemmas 3.3 and 3.4. Applying Lemma 2.2, then the sequences \{u_{n}(s)\}_{n = 1}^{\infty} is relatively compact in L^2(T, t;L^2(B_{k})) . In other words, for any sequences \{z_{n}(T) = (u_{n}(T), \eta^{T}_n)\}\subset B_T \in\tilde{B}^0 , \{z_{n}(t) = (u_{n}(t), \eta^t_n)\} is the solution of Eq (1.14) with the initial data \{z_{n}(T) = (u_{n}(T), \eta^{T}_n)\} respectively. Then there exist subsequence \{z_{n_k}\}\subset \{z_n\} satisfying

    \begin{equation*} \lim\limits_{k\to\infty} \lim\limits_{l\to\infty}\varphi_T^{t}(z_{n_k}, z_{n_l}) = 0. \end{equation*}

    It follows that \varphi^t_T\in \mathfrak{E}(B_T) . Substituting (3.56) and (3.52) into (3.51), one gets

    \begin{equation*} \|U(t,T)x-U(t,T)y\|^2_{\mathcal{M}_t}\leq C\varepsilon+\varphi^t_{T}(x,y). \end{equation*}

    By Definitions 2.8 and 2.9, then \varphi_{T}^{t} is contractive function in B_T . Therefore, it's easy to obtain that the process U(t, \tau) is \mathcal{M}_t -contractive process on B_T\in\tilde{B}^0 (from (3.13)).

    As the end of this article, we will deduce the main conclusion as the following theorem:

    Theorem 3.2. The process U(t, \tau) defined by (3.3) possesses a time-dependent global attractor \mathscr{\tilde{A}} in \{\mathcal{M}_t\}_{t\in\mathbb{R}} , and {\mathscr{\tilde{A}}} is non-empty, compact, invariant in \{\mathcal{M}_t\}_{t\in\mathbb{R}} and pullback attracting in \{\mathcal{M}_t\}_{t\in\mathbb{R}} . Furthermore,

    \begin{equation*} {\mathscr{\tilde{A}}}\subset \{\mathcal{M}_t^1\}_{t\in\mathbb{R}}. \end{equation*}

    Proof. Thanks to Lemma 3.2 and Theorem 3.1, it's easy to get the existence of time-dependent global attractor \mathscr{\tilde{A}} for the process U(t, \tau) defined by (3.3) in time-dependent spaces \{\mathcal{M}_t\}_{t\in\mathbb{R}} . According to Lemmas 3.7 and 3.8, the pullback asymptotic regularity of the solutions of Eq (1.1) is proved, and the regularity of the time-dependent global attractor \mathscr{\tilde{A}} is obtained. By Lemma 2.1 and (3.2), it follows that the invariance of time-dependent global attractor \mathscr{\tilde{A}} .

    We conclude the existence, uniqueness and regularity of time-dependent global attractors on whole space. The findings of this study can be considered as a supplement to our previous works, such as [4,5]. We overcome some essential difficulties for studying this kind of problem, including that the compact embedding is no longer valid under the case of unbounded domain and the nonlinear term fulfills the supcritical growth as well as the memory kernel satisfying more general assumptions. However, our results show that the method of operator decomposition that was proposed in [46] is available for dealing with the case of unbounded domain like (1.1).

    Unfortunately, we fail to consider the existence of time-dependent global attractors for Eq (1.1) which lacks instantaneous damping on whole space, and further study the upper-semicontinuity of time-dependent global attractors between two kinds of equations. Future studies shall consider such issues using the ideas of the paper and [4,5], i.e., the asymptotic behavior of solutions for nonautonomous and autonomous equations (1.1) lacking instantaneous damping on whole space.

    The authors declare that they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The research was financially supported by National Natural Science Foundation of China (Nos. 11101053, 71471020) and Scientific Research Program Funds of NUDT (No. 22-ZZCX-016).

    The authors declare no conflicts of interest.



    [1] S. Gatti, A. Miranville, V. Pata, S. Zelik, Attractors for semi-linear equations of viscoelasticity with very low dissipation, Rocky Mountain J. Math., 38 (2008), 1117–1138. https://doi.org/10.1216/RMJ-2008-38-4-1117 doi: 10.1216/RMJ-2008-38-4-1117
    [2] T. T. Le, D. T. Nguyen, The nonclassical diffusion equations with time-dependent memory kernels and a new class of nonlinearities, Glasg. Math. J., 64 (2022), 716–733. https://doi.org/10.1002/mma.6791 doi: 10.1002/mma.6791
    [3] T. T. Le, D. T. Nguyen, Uniform attractors of nonclassical diffusion equations on \mathbb{R}^N with memory and singularly oscillating external forces, Math. Methods Appl. Sci., 44 (2021), 820–852. https://doi.org/10.1002/mma.6791 doi: 10.1002/mma.6791
    [4] Z. Xie, J. W. Zhang, Y. Q. Xie, Asymptotic behavior of quasi-linear evolution equations on time-dependent product spaces, Discrete Contin. Dyn. Syst. B, 28 (2023), 2316–2334. https://doi.org/10.3934/dcdsb.2022171 doi: 10.3934/dcdsb.2022171
    [5] Y. Q. Xie, D. Liu, J. W. Zhang, X. M. Liu, Uniform attractors for nonclassical diffusion equations with perturbed parameter and memory, J. Math. Phys., 64 (2023), 022701. https://doi.org/10.1063/5.0068029 doi: 10.1063/5.0068029
    [6] J. Wang, Q. Z. Ma, W. X. Zhou, Attractor of the nonclassical diffusion equation with memory on time-dependent space, AIMS Math., 8 (2023), 14820–14841. https://doi.org/10.3934/math.2023757 doi: 10.3934/math.2023757
    [7] C. M. Dafermos, Asymptotic stability in viscoelasticity, Arch. Rational Mech. Anal., 37 (1970), 297–308. https://doi.org/10.1007/BF00251609 doi: 10.1007/BF00251609
    [8] P. J. Chen, M. E. Gurtin, On a theory of heat conduction involving two temperatures, Z. Angew. Math. Phys., 19 (1968), 614–627. https://doi.org/10.1007/BF01594969 doi: 10.1007/BF01594969
    [9] G. I. Barenblatt, I. P. Zheltov, I. N. Kochina, Basic concepts in the theory of seepage of homogeneous liquids in fissured rocks, J. Appl. Math. Mech., 24 (1960), 1286–1303. https://doi.org/10.1016/0021-8928(60)90107-6 doi: 10.1016/0021-8928(60)90107-6
    [10] E. C. Aifantis, On the problem of diffusion in solids, Acta Mech., 37 (1980), 265–296. https://doi.org/10.1007/BF01202949 doi: 10.1007/BF01202949
    [11] J. Jäckle, Heat conduction and relaxation in liquids of high viscosity, Phys. A, 162 (1990), 377–404. https://doi.org/10.1016/0378-4371(90)90424-Q doi: 10.1016/0378-4371(90)90424-Q
    [12] C. T. Anh, N. D. Toan, Nonclassical diffusion equations on \mathbb{R}^N with singularly oscillating external forces, Appl. Math. Lett., 38 (2014), 20–26. https://doi.org/10.1016/j.aml.2014.06.008 doi: 10.1016/j.aml.2014.06.008
    [13] M. Conti, E. M. Marchini, A remark on nonclassical diffusion equations with memory, Appl. Math. Optim., 73 (2016), 1–21. https://doi.org/10.1007/s00245-015-9290-8 doi: 10.1007/s00245-015-9290-8
    [14] M. Conti, E. M. Marchini, V. Pata, Nonclassical diffusion with memory, Math. Methods Appl. Sci., 38 (2015), 948–958. https://doi.org/10.1002/mma.3120 doi: 10.1002/mma.3120
    [15] J. W. Zhang, Y. Q. Xie, Asymptotic behavior for a class of viscoelastic equations with memory lacking instantaneous damping, AIMS Math., 6 (2021), 9491–9509. https://doi.org/10.3934/math.2021552 doi: 10.3934/math.2021552
    [16] V. Pata, A. Zucchi, Attractors for a damped hyperbolic equation with linear memory, Adv. Math. Sci. Appl., 11 (2001), 505–529.
    [17] J. W. Zhang, Y. Q. Xie, Q. Q. Luo, Z. P. Tang, Asymptotic behavior for the semi-linear reaction-diffusion equations with memory, Adv. Differ. Equ., 2019 (2019), 510. https://doi.org/10.1186/s13662-019-2399-3 doi: 10.1186/s13662-019-2399-3
    [18] V. V. Chepyzhov, A. Miranville, On trajectory and global attractors for semilinear heat equations with fading memory, Indiana Univ. Math. J., 55 (2006), 119–168. https://doi.org/10.1512/iumj.2006.55.2597 doi: 10.1512/iumj.2006.55.2597
    [19] J. B. Yuan, S. X. Zhang, Y. Q. Xie, J. W. Zhang, Exponential attractors for nonclassical diffusion equations with arbitrary polynomial growth nonlinearity, AIMS Math., 6 (2021), 11778–11795. https://doi.org/10.3934/math.2021684 doi: 10.3934/math.2021684
    [20] C. Y. Sun, M. H. Yang, Dynamics of the nonclassical diffusion equation, Asymptot. Anal., 59 (2008), 51–81. https://doi.org/10.3233/ASY-2008-0886 doi: 10.3233/ASY-2008-0886
    [21] Y. L. Xiao, Attractors for a nonclassical diffusion equation, Acta Math. Appl. Sinca, 18 (2002), 273–276. https://doi.org/10.1007/s102550200026 doi: 10.1007/s102550200026
    [22] J. W. Zhang, Z. M. Liu, J. H. Huang, Upper semicontinuity of optimal attractors for viscoelastic equations lacking strong damping, Appl. Anal., 102 (2023), 3609–3628. https://doi.org/10.1080/00036811.2022.2088532 doi: 10.1080/00036811.2022.2088532
    [23] Y. Q. Xie, Q. S. Li, K. X. Zhu, Attractors for nonclassical diffusion equations with arbitrary polynomial growth nonlinearity, Nonlinear Anal. Real World Appl., 31 (2016), 23–37. https://doi.org/10.1016/j.nonrwa.2016.01.004 doi: 10.1016/j.nonrwa.2016.01.004
    [24] J. W. Zhang, Z. M. Liu, J. H. Huang, Weak mean random attractors for nonautonomous stochastic parabolic equation with variable exponents, Stoch. Dyn., 23 (2023), 2350019. https://doi.org/10.1142/S0219493723500193 doi: 10.1142/S0219493723500193
    [25] J. W. Zhang, Z. M. Liu, J. H. Huang, Upper semicontinuity of pullback \mathscr{D}-attractors for nonlinear parabolic equation with nonstandard growth condition, Math. Nachr., 2023, 1–24. https://doi.org/10.1002/mana.202100527
    [26] E. S. Baranovskii, Strong solutions of the incompressible Navier-Stokes-Voigt model, Mathematics, 8 (2020), 181. https://doi.org/10.3390/math8020181 doi: 10.3390/math8020181
    [27] T. Ding, Y. F. Liu, Time-dependent global attractor for the nonclassical diffusion equation, Appl. Anal., 94 (2015), 1439–1449. https://doi.org/10.1080/00036811.2014.933475 doi: 10.1080/00036811.2014.933475
    [28] Q. Z. Ma, X. P. Wang, L. Xu, Existence and regularity of time-dependent global attractors for the nonclassical reaction-diffusion equations with lower forcing term, Bound. Value Probl., 2016 (2016), 1–11. https://doi.org/10.1186/s13661-015-0513-3 doi: 10.1186/s13661-015-0513-3
    [29] J. Wang, Q. Z. Ma, Asymptotic dynamic of the nonclassical diffusion equation with time-dependent coefficient, J. Appl. Anal. Comput., 11 (2021), 445–463. https://doi.org/10.11948/20200055 doi: 10.11948/20200055
    [30] J. B. Yuan, S. X. Zhang, Y. Q. Xie, J. W. Zhang, Attractors for a class of perturbed nonclassical diffusion equations with memory, Discrete Contin. Dyn. Syst. B, 27 (2022), 4995–5007. https://doi.org/10.3934/dcdsb.2021261 doi: 10.3934/dcdsb.2021261
    [31] Y. Q. Xie, J. Li, K. X. Zhu, Upper semicontinuity of attractors for nonclassical diffusion equations with arbitrary polynomial growth, Adv. Differ. Equ., 2021 (2021), 1–17. https://doi.org/10.1186/s13662-020-03146-2 doi: 10.1186/s13662-020-03146-2
    [32] K. X. Zhu, Y. Q. Xie, F. Zhou, Attractors for the nonclassical reaction-diffusion equations on time-dependent spaces, Bound. Value Probl., 2020 (2020), 1–14. https://doi.org/10.1186/s13661-020-01392-7 doi: 10.1186/s13661-020-01392-7
    [33] M. Conti, F. Dell'Oro, V. Pata, Nonclassical diffusion with memory lacking instantaneous damping, Commun. Pure Appl. Anal., 19 (2020), 2035–2050. https://doi.org/10.3934/cpaa.2020090 doi: 10.3934/cpaa.2020090
    [34] N. D. Toan, Uniform attractors of nonclassical diffusion equations lacking instantaneous damping on \mathbb {R}^{N} with memory, Acta Appl. Math., 170 (2020), 789–822. https://doi.org/10.1007/s10440-020-00359-1 doi: 10.1007/s10440-020-00359-1
    [35] F. J. Meng, M. H. Yang, C. K. Zhong, Attractors for wave equation with nonlinear damping on time-dependent space, Discrete Contin. Dyn. Syst. B, 21 (2016), 205–225. https://doi.org/10.3934/dcdsb.2016.21.205 doi: 10.3934/dcdsb.2016.21.205
    [36] M. Conti, V. Pata, Asymptotic structure of the attractor for processes on time-dependent spaces, Nonlinear Anal. Real World Appl., 19 (2014), 1–10. https://doi.org/10.1016/j.nonrwa.2014.02.002 doi: 10.1016/j.nonrwa.2014.02.002
    [37] M. Conti, V. Pata, R. Temam, Attractors for process on time-dependent spaces. Applications to wave equations, J. Differ. Equ., 255 (2013), 1254–1277. https://doi.org/10.1016/j.jde.2013.05.013 doi: 10.1016/j.jde.2013.05.013
    [38] A. N. Carvalho, J. A. Langa, J. C. Robinson, Attractors for infinite-dimensional non-autonomous dynamical systems, New York: Springer, 2013. https://doi.org/10.1007/978-1-4614-4581-4
    [39] P. E. Kloeden, T. Lorenz, Pullback incremental attraction, Nonauton. Dyn. Syst., 1 (2014), 53-60. https://doi.org/10.2478/msds-2013-0004 doi: 10.2478/msds-2013-0004
    [40] C. Y. Sun, L. Yang, J. Q. Duan, Asymptotic behavior for a semilinear second order evolution equation, Trans. Amer. Math. Soc., 363 (2011), 6085–6109.
    [41] C. Y. Sun, D. M. Cao, J. Q. Duan, Non-autonomous dynamics of wave equations with nonlinear damping and critical nonlinearity, Nonlinearity, 19 (2006), 2645. https://doi.org/10.1088/0951-7715/19/11/0086 doi: 10.1088/0951-7715/19/11/0086
    [42] Y. Q. Xie, Y. Li, Y. Zeng, Uniform attractors for nonclassical diffusion equations with memory, J. Funct. Spaces, 2016 (2016), 1–11. https://doi.org/10.1155/2016/5340489 doi: 10.1155/2016/5340489
    [43] C. M. Dafermos, M. Slemrod, Asymptotic behavior of nonlinear contraction semigroups, J. Funct. Anal., 13 (1973), 97–106. https://doi.org/10.1016/0022-1236(73)90069-4 doi: 10.1016/0022-1236(73)90069-4
    [44] J. C. Robinson, Infinite-dimensional dynamical systems, Cambridge: Cambridge University Press, 2001.
    [45] Z. Tang, J. Zhang, D. Liu, Well-posedness of time-dependent nonclassical diffusion equation with memory, Math. Theor. Appl., 41 (2021), 102–111.
    [46] Y. Q. Xie, J. W. Zhang, C. X. Huang, Attractors for reaction-diffusion equation with memory, Acta Math. Sinica (Chin. Ser.), 64 (2021), 979–990. https://doi.org/10.12386/A2021sxxb0081 doi: 10.12386/A2021sxxb0081
  • This article has been cited by:

    1. Yadan Shi, Yongqin Xie, Ke Li, Zhipiao Tang, Attractors for the nonclassical diffusion equations with the driving delay term in time-dependent spaces, 2024, 32, 2688-1594, 6847, 10.3934/era.2024320
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1236) PDF downloads(66) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog