1 Introduction

Let Ω be a bounded domain in \(\mathbb{R}^{n}\)\((n\geq 3)\) with smooth boundary, we consider the long-time behavior of the solutions for the following nonclassical reaction–diffusion equation:

$$ \textstyle\begin{cases} u_{t}-\varepsilon (t)\Delta u_{t}-\Delta u+f(u)=g(x) &\text{in } \varOmega \times (\tau,\infty ), \\ u=0 &\text{on } \partial \varOmega \times (\tau,\infty ), \\ u(x,\tau )=u_{\tau }, &x\in \varOmega, \end{cases} $$
(1.1)

where \(t>\tau \), \(\tau \in \mathbb{R}\) is the initial time, \(g(x)\in H^{-1}(\varOmega )\) is an external force term, \(\varepsilon (t)\in C^{1}(\mathbb{R})\) is a decreasing bounded function satisfying

$$\begin{aligned} \lim_{t\rightarrow +\infty }\varepsilon (t)=0, \end{aligned}$$
(1.2)

and there exists \(L>0\) such that

$$\begin{aligned} \sup_{t\in \mathbb{R}} \bigl( \bigl\vert \varepsilon (t) \bigr\vert + \bigl\vert \varepsilon '(t) \bigr\vert \bigr)\leq L. \end{aligned}$$
(1.3)

For the nonlinear term \(f\in C(\mathbb{R},\mathbb{R})\), similar to that in [3, 20, 24], we make the following classical assumptions:

$$ f'(u)\geq -l, \quad\forall u\in \mathbb{R,} $$
(1.4)

and

$$ -c_{0}+c_{1} \vert u \vert ^{p} \leq f(u)u\leq c_{0}+c_{2} \vert u \vert ^{p},\quad p\geq 2, $$
(1.5)

for some positive constants \(c_{0}, c_{1}, c_{2}\).

Let \(\mathcal{F}(u)=\int _{0}^{u}f(r)\,dr\), then there are constants \(\tilde{c}_{i}>0\)\((i=0,1,2)\) such that

$$\begin{aligned} -\tilde{c}_{0}+\tilde{c}_{1} \vert u \vert ^{p}\leq \mathcal{F}(u)\leq \tilde{c}_{0}+ \tilde{c}_{2} \vert u \vert ^{p},\quad \forall u\in \mathbb{R}. \end{aligned}$$
(1.6)

For Eq. (1.1), when \(\varepsilon (t)>0\) is a constant, the existence and long-time behavior of solutions have been extensively studied by several authors; see, e.g., [1, 4, 5, 23, 2527, 29, 30, 32]. In [4, 5, 29], the authors main considered the existence of solutions for this type of equations. In [1, 23, 2527, 30], the authors main considered the existence of the global attractors (see [23, 2527]) and the pullback (or the uniform) attractors (see [1, 23, 30]) in \(H_{0}^{1}(\varOmega )\) (or \(H^{1}(\mathbb{R}^{N})\)). In particular, in [32], we obtained the existence of the pullback attractors in \(C_{H_{0}^{1}(\varOmega )}\) (rather than in \(H_{0}^{1}(\varOmega )\)) for the nonclassical reaction–diffusion equations with delays.

When \(\varepsilon (t)=0\), Eq. (1.1) becomes the classical reaction–diffusion equation. The existence and the long-time behavior of solutions have also been extensively investigated by several authors; see, e.g., [2, 11, 12, 17, 21, 28, 31]. In [2, 11, 12, 28], the authors mainly considered the existence (or the blowup), uniqueness and the long-time decay of the solutions for the semilinear parabolic equation [11, 12], the nonlinear parabolic equation [2] and the coupled parabolic systems [28]. In [17, 21, 31], the authors have proved the existence of the global attractors in \(L^{p}(\varOmega )\), \(H_{0}^{1}(\varOmega )\), \(L^{2p-2}(\varOmega )\), \(H^{2}(\varOmega )\) (see [31]) and the existence of the pullback attractors in \(L^{p}(\varOmega )\) and \(H_{0}^{1}(\varOmega )\) (see [17] and [21], respectively).

When \(\varepsilon (t)\in C^{1}(\mathbb{R})\) satisfies (1.2)–(1.3), the long-time behavior of solutions for Eq. (1.1) has been considered by some researchers; see, e.g., [16, 18]. In [16], the authors have proved the existence of the time-dependent global attractors in \(\mathcal{H}_{t}\) with the nonlinearity f satisfying \(|f''(u)|\leq c(1+|u|)\) (see Theorem 3.4 in [16] for details). Furthermore, in [18], the authors have considered the case of the nonlinearity f satisfying the critical exponent growth and proved the existence of the time-dependent global attractors in \(\mathcal{H}_{t}\) (see Theorem 3.3 in [18] for details).

In this paper, we consider Eq. (1.1) with the nonlinearity f satisfying polynomial growth of arbitrary \(p-1\)\((p\geq 2)\) order, which makes that the Sobolev compact embedding is no longer valid and brings more difficulty for verifying the corresponding asymptotic compactness of the solutions process \(\{U(t,\tau )\}_{t\geq \tau }\). In order to overcome the difficulty mentioned above, we verify the existence of the time-dependent global attractors \(\hat{\mathcal{A}}\) in \(\mathcal{H}_{t}\) for the process \(\{U(t,\tau )\}_{t\geq \tau }\) by applying the contractive function methods as in [6, 13, 14, 19, 22, 27] (see Theorem 3.8).

2 Preliminaries

In this section, we firstly review briefly some notations, basic definitions and results about processes on time-dependent spaces (see [79, 19] for details).

2.1 Notations

Let \(\{X_{t}\}_{t\in \mathbb{R}}\) be a family of normed spaces, we introduce the R-ball of \(X_{t}\) as

$$ \mathbb{B}_{t}(R)=\bigl\{ z\in X_{t}: \Vert z \Vert _{X_{t}}\leq R\bigr\} . $$

For any given \(\epsilon >0\), the ϵ-neighborhood of a set \(B\subset X_{t}\) is defined as

$$ \mathcal{O}_{t}^{\epsilon }(B)=\bigcup _{x\in B}\bigl\{ y\in X_{t}: \Vert x-y \Vert _{X_{t}}< \epsilon \bigr\} =\bigcup_{x\in B}\bigl\{ x+\mathbb{B}_{t}(\epsilon )\bigr\} . $$

We denote the Hausdorff semidistance of two (nonempty) sets \(B,C\subset X_{t}\) by

$$ \delta _{t}(B,C)=\sup_{x\in B}\inf _{y\in C} \Vert x-y \Vert _{X_{t}}. $$

Moreover, we introduce the time-dependent space \(\mathcal{H}_{t}\) endowed with the norms

$$ \Vert u \Vert ^{2}_{\mathcal{H}_{t}}= \Vert u \Vert ^{2}_{2}+\varepsilon (t) \Vert \nabla u \Vert ^{2}_{2}, $$

where \(\|\cdot \|_{2}\) denotes the usual norm in \(L^{2}(\varOmega )\).

2.2 Some concepts

In this subsection, we give some concepts about the time-dependent global attractors.

Definition 2.1

Let \(\{X_{t}\}_{t\in \mathbb{R}}\) be a family of normed spaces. A process is a two-parameter family of mappings \(U(t,\tau ): X_{\tau }\rightarrow X_{t}\), \(t\geq \tau \), \(\tau \in \mathbb{R}\) with properties

  1. (i)

    \(U(\tau,\tau )=\mathrm{Id}\) is the identity operator on \(X_{\tau }\), \(\tau \in \mathbb{R}\);

  2. (ii)

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

Definition 2.2

A family \(\hat{C}=\{C_{t}\}_{t\in \mathbb{R}}\) of bounded sets \(C_{t}\subset X_{t}\) is called uniformly bounded if there exists a constant \(R>0\) such that \(C_{t}\subset \mathbb{B}_{t}(R)\) for all \(t\in \mathbb{R}\).

Definition 2.3

A family \(\hat{B}=\{B_{t}\}_{t\in \mathbb{R}}\) is called pullback absorbing if it is uniformly bounded and for every \(R>0\), there exists a constant \(t_{0}=t_{0}(t,R)\leq t\) such that \(U(t,\tau )\mathbb{B}_{\tau }(R)\subset B_{t}\) for all \(\tau \leq t_{0}\).

The process \(\{U(t,\tau )\}_{t\geq \tau }\) is called dissipative whenever it admits a pullback absorbing family.

Definition 2.4

A time-dependent absorbing set for the process \(\{U(t,\tau )\}_{t\geq \tau }\) is a uniformly bounded family \(\hat{B}=\{B_{t}\}_{t\in \mathbb{R}}\) with the following property: for every \(R\geq 0\) there exists a \(t_{0}=t_{0}(R)\geq 0\) such that

$$ U(t,\tau )\mathbb{B}_{\tau }(R)\subset B_{t} \quad\text{for all } \tau \leq t-t_{0}. $$

Definition 2.5

The process \(\{U(t,\tau )\}_{t\geq \tau }\) is said to be pullback asymptotically compact if for any \(t\in \mathbb{R}\), any bounded sequence \(\{x_{n}\}_{n=1}^{\infty }\subset X_{\tau _{n}}\) and any sequence \(\{\tau _{n}\}_{n=1}^{\infty }\) with \(\tau _{n}\rightarrow -\infty \) as \(n\rightarrow \infty \), the sequence \(\{U(t,\tau _{n})x_{n}\}_{n=1}^{\infty }\) is precompact in \(\{X_{t}\}_{t\in \mathbb{R}}\).

Definition 2.6

The time-dependent global attractor for the process \(\{U(t,\tau )\}_{t\geq \tau }\) is the smallest family \(\hat{\mathcal{A}}=\{\mathcal{A}_{t}\}_{t\in \mathbb{R}}\) such that

  1. (i)

    \(\mathcal{A}_{t}\) is compact in \(X_{t}\);

  2. (ii)

    \(\hat{\mathcal{A}}\) is invariant, i.e., \(U(t,\tau )\mathcal{A}_{\tau }=\mathcal{A}_{t}, \forall t\geq \tau \);

  3. (iii)

    \(\hat{\mathcal{A}}\) is pullback attracting, i.e., it is uniformly bounded and the limit

    $$ \lim_{\tau \rightarrow -\infty }\delta _{t}\bigl(U(t,\tau )C_{\tau }, \mathcal{A}_{t}\bigr)=0 $$

    holds for every uniformly bounded family \(\hat{C}=\{C_{t}\}_{t\in \mathbb{R}}\) and every fixed \(t\in \mathbb{R}\).

Remark 2.7

The attracting property can be equivalently stated in terms of pullback absorbing: a (uniformly bounded) family \(\mathcal{K}=\{K_{t}\}_{t\in \mathbb{R}}\) is called pullback attracting if for any \(\epsilon >0\) the family \(\{\mathcal{O}_{t}^{\epsilon }(K_{t})\}_{t\in \mathbb{R}}\) is pullback absorbing.

Similarly to Theorem 4.2 in [8], we have the following theorem.

Theorem 2.8

The time-dependent global attractor\(\hat{\mathcal{A}}\)exists and it is unique if and only if the process\(\{U(t,\tau )\}_{t\geq \tau }\)is asymptotically compact, namely, the set

$$ \mathbb{K}=\bigl\{ \mathcal{K}=\{K_{t}\}_{t\in \mathbb{R}}: K_{t}\subset X_{t} \textit{ is compact }, \mathbb{K} \textit{ is pullback attracting}\bigr\} $$

is not empty.

2.3 Some results

In order to obtain the time-dependent global attractors of Eq. (1.1), we need the following definitions and conclusions, which are similar to those in [6, 13, 14, 19, 22, 27].

Definition 2.9

Let \(\{X_{t}\}_{t\in \mathbb{R}}\) be a family of Banach spaces and \(\hat{C}=\{C_{t}\}_{t\in \mathbb{R}}\) be a family of uniformly bounded subset of \(\{X_{t}\}_{t\in \mathbb{R}}\). We call a function \(\psi _{\tau }^{t}(\cdot,\cdot )\), defined on \(\{X_{t}\}_{t\in \mathbb{R}}\times \{X_{t}\}_{t\in \mathbb{R}}\), a contractive function on \(C_{\tau }\times C_{\tau }\) if for fixed \(t\in \mathbb{R}\) and any sequence \(\{x_{n}\}_{n=1}^{\infty }\subset C_{\tau }\), there is a subsequence \(\{x_{n_{k}}\}_{k=1}^{\infty }\subset \{x_{n}\}_{n=1}^{\infty }\) such that

$$\begin{aligned} \lim_{k\rightarrow \infty }\lim_{l\rightarrow \infty }\psi _{\tau }^{t}(x_{n_{k}},x_{n_{l}})=0\quad \text{for all } t\geq \tau. \end{aligned}$$

We denote the set of all contractive functions on \(C_{\tau }\times C_{\tau }\) by \(Contr(C_{\tau})\).

Theorem 2.10

Let\(\{U(t,\tau )\}_{t\geq \tau }\)be a process on Banach spaces\(\{X_{t}\}_{t\in \mathbb{R}}\)and have a pullback absorbing set\(\hat{B}=\{B_{t}\}_{t\in \mathbb{R}}\). Moreover, assume that, for any\(\epsilon >0\), there exist\(\tau _{0}=\tau _{0}(\epsilon )< t\)and\(\psi _{\tau _{0}}^{t}(\cdot,\cdot )\in \hat{C}(B_{\tau _{0}})\)such that

$$\begin{aligned} \bigl\Vert U(t,\tau _{0})x-U(t,\tau _{0})y \bigr\Vert _{X_{t}}\leq \epsilon +\psi _{ \tau _{0}}^{t}(x,y),\quad \forall x,y\in B_{\tau _{0}}, \end{aligned}$$

for any\(t\in \mathbb{R}\). Then\(\{U(t,\tau )\}_{t\geq \tau }\)is pullback asymptotically compact in\(\{X_{t}\}_{t\in \mathbb{R}}\).

Proof

We need to prove that, for any \(\{x_{n}\}_{n=1}^{\infty }\subset B_{\tau _{n}}\) and any \(\tau _{n}\rightarrow -\infty \) as \(n\rightarrow \infty \),

$$ \text{the sequence } \bigl\{ U(t,\tau _{n})x_{n}\bigr\} _{n=1}^{\infty } \text{ is precompact in } \{X_{t} \}_{t\in \mathbb{R}}. $$

In the following, we will show that \(\{U(t,\tau _{n})x_{n}\}_{n=1}^{\infty }\) has a convergent subsequence via diagonal methods.

Taking \(\epsilon _{m}>0\) with \(\epsilon _{m}\rightarrow 0\) as \(m\rightarrow \infty \).

Then, for \(\epsilon _{1}>0\), by the assumptions, there exist \(\tau _{0}=\tau _{0}(\epsilon _{1})< t\) and \(\psi ^{t}_{\tau _{0}}(\cdot,\cdot )\in \hat{C}(B_{\tau _{0}})\) such that

$$\begin{aligned} \bigl\Vert U(t,\tau _{0})x-U(t,\tau _{0})y \bigr\Vert _{X_{t}}\leq \epsilon _{1}+\psi ^{t}_{ \tau _{0}}(x,y), \quad\forall x,y\in B_{\tau _{0}}, \end{aligned}$$
(2.1)

for any \(t\in \mathbb{R}\), where \(\psi ^{t}_{\tau _{0}}\) depends on \(\tau _{0}\).

Since \(\tau _{n}\rightarrow -\infty \), without loss of generality, we assume that \(\tau _{n}\leq \tau _{0}\) such that \(U(\tau _{0},\tau _{n})x_{n}\in B_{\tau _{0}}\) for each \(n\in \mathbb{N}\). Set \(y_{n}=U(\tau _{0},\tau _{n})x_{n}\), then from (2.1) we have

$$\begin{aligned} \bigl\Vert U(t,\tau _{n})x_{n}-U(t,\tau _{m})x_{m} \bigr\Vert _{X_{t}} &= \bigl\Vert U(t,\tau _{0})U( \tau _{0},\tau _{n})x_{n}-U(t, \tau _{0})U(\tau _{0},\tau _{m})x_{m} \bigr\Vert _{X_{t}} \\ &= \bigl\Vert U(t,\tau _{0})y_{n}-U(t,\tau _{0})y_{m} \bigr\Vert _{X_{t}} \\ &\leq \epsilon _{1}+\psi ^{t}_{\tau _{0}}(y_{n},y_{m}). \end{aligned}$$
(2.2)

By the definition of \(\hat{C}(B_{\tau _{0}})\) and \(\psi ^{t}_{\tau _{0}}\in \hat{C}(B_{\tau _{0}})\), we know that \(\{y_{n}\}_{n=1}^{\infty }\) have a subsequence \(\{y_{n_{k}}^{(1)}\}_{k=1}^{\infty }\) such that

$$\begin{aligned} \lim_{k\rightarrow \infty }\lim_{l\rightarrow \infty } \psi ^{t}_{ \tau _{0}}\bigl(y_{n_{k}}^{(1)},y_{n_{l}}^{(1)} \bigr)\leq \epsilon _{1}. \end{aligned}$$
(2.3)

Similarly to [13, 22, 27], we have

$$\begin{aligned} &\lim_{k\rightarrow \infty }\sup_{q\in \mathbb{N}} \bigl\Vert U\bigl(t, \tau _{n_{k+q}}^{(1)}\bigr)x_{n_{k+q}}^{(1)}-U\bigl(t, \tau _{n_{k}}^{(1)}\bigr)x_{n_{k}}^{(1)} \bigr\Vert _{X_{t}} \\ & \quad\leq \lim_{k\rightarrow \infty }\sup_{q\in \mathbb{N}} \limsup _{l\rightarrow \infty } \bigl\Vert U\bigl(t,\tau _{n_{k+q}}^{(1)} \bigr)x_{n_{k+q}}^{(1)}-U\bigl(t, \tau _{n_{l}}^{(1)} \bigr)x_{n_{l}}^{(1)} \bigr\Vert _{X_{t}} \\ &\qquad{} +\limsup_{k\rightarrow \infty }\limsup_{l\rightarrow \infty } \bigl\Vert U \bigl(t,\tau _{n_{k}}^{(1)}\bigr)x_{n_{k}}^{(1)}-U \bigl(t,\tau _{n_{l}}^{(1)}\bigr)x_{n_{l}}^{(1)} \bigr\Vert _{X_{t}} \\ & \quad\leq \epsilon _{1}+\lim_{k\rightarrow \infty }\sup _{q\in \mathbb{N}}\limsup_{l\rightarrow \infty } \psi ^{t}_{\tau _{0}} \bigl(y_{n_{k+q}}^{(1)},y_{n_{l}}^{(1)}\bigr) + \epsilon _{1}+\lim_{k\rightarrow \infty }\lim_{l\rightarrow \infty } \psi ^{t}_{\tau _{0}}\bigl(y_{n_{k}}^{(1)},y_{n_{l}}^{(1)} \bigr), \end{aligned}$$

which, combining with (2.2) and (2.3), implies that

$$\begin{aligned} \lim_{k\rightarrow \infty }\sup_{q\in \mathbb{N}} \bigl\Vert U\bigl(t, \tau _{n_{k+q}}^{(1)}\bigr)x_{n_{k+q}}^{(1)}-U\bigl(t, \tau _{n_{k}}^{(1)}\bigr)x_{n_{k}}^{(1)} \bigr\Vert _{X_{t}}\leq 4\epsilon _{1}. \end{aligned}$$

Therefore, there exists a \(K_{1}\in \mathbb{N}\) such that

$$\begin{aligned} \lim_{k\rightarrow \infty }\sup_{q\in \mathbb{N}} \bigl\Vert U\bigl(t, \tau _{n_{k}}^{(1)}\bigr)x_{n_{k}}^{(1)}-U\bigl(t, \tau _{n_{l}}^{(1)}\bigr)x_{n_{l}}^{(1)} \bigr\Vert _{X_{t}}\leq 5\epsilon _{1},\quad \text{for all } k,l\geq K_{1}. \end{aligned}$$

By induction, we can obtain that, for each \(m\geq 1\), there exists a subsequence \(\{U(t, \tau _{n_{k}}^{(m+1)})x_{n_{k}}^{(m+1)}\}_{k=1}^{\infty }\) of \(\{U(t,\tau _{n_{k}}^{(m)})x_{n_{k}}^{(m)}\}_{k=1}^{\infty }\) and certain \(K_{m+1}\) such that

$$\begin{aligned} \lim_{k\rightarrow \infty }\sup_{q\in \mathbb{N}} \bigl\Vert U\bigl(t, \tau _{n_{k}}^{(m+1)}\bigr)x_{n_{k}}^{(m+1)}-U\bigl(t, \tau _{n_{l}}^{(m+1)}\bigr)x_{n_{l}}^{(m+1)} \bigr\Vert _{X_{t}}\leq 5\epsilon _{m+1},\quad \text{for all } k,l\geq K_{m+1}. \end{aligned}$$

Now, we consider the diagonal subsequence \(\{U(t,\tau _{n_{k}}^{(k)})x_{n_{k}}^{(k)}\}_{k=1}^{\infty }\). Since for each \(m\in \mathbb{N}\), \(\{U(t,\tau _{n_{k}}^{(k)})x_{n_{k}}^{(k)}\}_{k=m}^{\infty }\) is a subsequence of \(\{U(t,\tau _{n_{k}}^{(k)})x_{n_{k}}^{(k)}\}_{k=1}^{\infty }\), then

$$\begin{aligned} \lim_{k\rightarrow \infty }\sup_{q\in \mathbb{N}} \bigl\Vert U\bigl(t, \tau _{n_{k}}^{(k)}\bigr)x_{n_{k}}^{(k)}-U\bigl(t, \tau _{n_{l}}^{(l)}\bigr)x_{n_{l}}^{(l)} \bigr\Vert _{X_{t}}\leq 6\epsilon _{m}, \quad\text{for all } k,l\geq \max \{m,K_{m}\}, \end{aligned}$$

which combining with \(\epsilon _{m}\rightarrow 0\) as \(m\rightarrow \infty \), implies that \(\{U(t,\tau _{n_{k}}^{(k)})x_{n_{k}}^{(k)}\}_{k=1}^{\infty }\) is a Cauchy sequence in \(\{X_{t}\}_{t\in \mathbb{R}}\). This shows that \(\{U(t,\tau _{n})x_{n}\}_{n=1}^{\infty }\) is precompact in \(\{X_{t}\}_{t\in \mathbb{R}}\). □

Similarly to Theorem 3.3 in [19], we have the following conclusion, which will be used to verify the existence of the time-dependent global attractor.

Theorem 2.11

Let\(\{U(t,\tau )\}_{t\geq \tau }\)be a process on Banach space\(\{X_{t}\}_{t\in \mathbb{R}}\), then\(\{U(t,\tau )\}_{t\geq \tau }\)has a time-dependent global attractor in\(\{X_{t}\}_{t\in \mathbb{R}}\)if the following conditions hold:

  1. (i)

    \(\{U(t,\tau )\}_{t\geq \tau }\)has a pullback absorbing set\(\hat{B}=\{B_{t}\}_{t\in \mathbb{R}}\)in\(\{X_{t}\}_{t\in \mathbb{R}}\);

  2. (ii)

    \(\{U(t,\tau )\}_{t\geq \tau }\)is pullback asymptotically compact in\(\hat{B}=\{B_{t}\}_{t\in \mathbb{R}}\).

3 Time-dependent global attractors

In this section, we will establish the existence of the time-dependent global attractors.

3.1 Existence and uniqueness of solutions

In this subsection, we consider the well-posedness of the solutions for Eq. (1.1) with (1.4)–(1.5). At first, we define the weak solutions as follows.

Definition 3.1

A weak solution of Eq. (1.1) is a function \(u\in C([\tau, T]; \mathcal{H}_{t})\cap L^{2}(\tau, T; H_{0}^{1}( \varOmega ))\cap L^{p}(\tau, T; L^{p}(\varOmega ))\) for all \(T>\tau \), with \(u(\tau )=u_{\tau }\) and such that, for all \(\varphi \in H_{0}^{1}(\varOmega )\), it satisfies

$$\begin{aligned} &\frac{d}{dt}\bigl[\bigl(u(t),\varphi \bigr)+\varepsilon (t) \bigl(\nabla u(t),\nabla \varphi \bigr)\bigr] +\bigl(1-\varepsilon '(t)\bigr) \bigl(\nabla u(t),\nabla \varphi \bigr)+\bigl(f\bigl(u(t)\bigr), \varphi \bigr) \\ &\quad =\bigl(g(x),\varphi \bigr),\quad \text{in } \mathcal{D}'(\tau,+\infty ). \end{aligned}$$

Remark 3.2

We notice that, if \(u(t)\) is a weak solution of Eq. (1.1), then it satisfies the energy equality

$$\begin{aligned} & \bigl\Vert u(t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla u(t) \bigr\Vert _{2}^{2}+ \int _{s}^{t}\bigl(2- \varepsilon '(r) \bigr) \bigl\Vert \nabla u(r) \bigr\Vert _{2}^{2}\,dr +2 \int _{s}^{t}\bigl(f\bigl(u(r)\bigr),u(r)\bigr)\,dr \\ &\quad= \bigl\Vert u(s) \bigr\Vert _{2}^{2}+\varepsilon (s) \bigl\Vert \nabla u(s) \bigr\Vert _{2}^{2} +2 \int _{s}^{t}\bigl(g(r),u(r)\bigr)\,dr \quad\text{for all } \tau \leq s\leq t. \end{aligned}$$

The following theorem gives the existence of the weak solutions, which is similar to that in [10] and can be obtained by the Faedo–Galerkin methods.

Theorem 3.3

Letfsatisfy (1.4)(1.5), \(g\in H^{-1}(\varOmega )\)and\(u_{\tau }\in \mathcal{H}_{\tau }\). Then, for any\(\tau \in \mathbb{R}\)and\(t>\tau \), there exists a weak solution\(u(t)\)to Eq. (1.1), which satisfies\(u\in C([\tau,t]; \mathcal{H}_{t})\cap L^{2}(\tau,t;H_{0}^{1}( \varOmega )) \cap L^{p}(\tau,t; L^{p}(\varOmega ))\), \(u_{t}\in L^{2}(\tau,t;\mathcal{H}_{t})\).

Proof

Let \(\{w_{j}\}_{j\geq 1}\subset H_{0}^{1}(\varOmega )\cap L^{p}(\varOmega )\) be a Hilbert basis of \(L^{2}(\varOmega )\) such that \(\operatorname{span} \{w_{j}\}_{j\geq 1}\) is dense in \(H_{0}^{1}(\varOmega )\cap L^{p}(\varOmega )\). In order to establish the existence of the weak solutions, we need the approximate system for any \(m\geq n\) seeking \(\tilde{u}^{m}(t, x)=\varSigma _{j=1}^{m}\gamma _{mj}(t)\omega _{j}(x)\) that satisfies

$$\begin{aligned} \textstyle\begin{cases} \frac{d}{dt}[(\tilde{u}^{m}(t), \omega _{j})+\varepsilon (t)(\nabla \tilde{u}^{m}(t),\nabla \omega _{j})] +(1-\varepsilon '(t))(\nabla \tilde{u}^{m}(t), \nabla \omega _{j}) +(f(\tilde{u}^{m}(t)),\omega _{j})\\ \quad=(g(x), \omega _{j}), \\ \tilde{u}_{\tau }^{m}=u_{\tau }, \end{cases}\displaystyle \end{aligned}$$

for a.e. \(t>\tau, 1\leq j\leq m\).

We will provide a priori estimates that show that these solutions are well-defined in the interval \([\tau,t]\) for any \(t>\tau \).

Step 1: First a priori estimates. Multiplying each equation in the above system by \(\gamma _{mj}(t)\), respectively, and summing from \(j=1\) to m, we obtain

$$\begin{aligned} &\frac{1}{2}\frac{d}{dt} \bigl( \bigl\Vert \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} + \varepsilon (t) \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} \bigr) + \bigl(1- \varepsilon '(t)\bigr) \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} \\ &\quad{} +\bigl(f\bigl(\tilde{u}^{m}(t)\bigr),\tilde{u}^{m}(t) \bigr)=\bigl(g(x),\tilde{u}^{m}(t)\bigr) \leq \frac{1}{2} \Vert g \Vert _{H^{-1}}^{2} +\frac{1}{2} \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2},\quad \text{a.e. }t> \tau, \end{aligned}$$

where we have used the Hölder and Young inequalities.

Furthermore, by (1.5), we know that

$$\begin{aligned} &\frac{d}{dt} \bigl( \bigl\Vert \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} +\varepsilon (t) \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} \bigr)+ \bigl(1-2\varepsilon '(t)\bigr) \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2}+2c_{1} \bigl\Vert \tilde{u}^{m}(t) \bigr\Vert _{p}^{p} \\ &\quad\leq 2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2},\quad \text{a.e. } t>\tau. \end{aligned}$$

Integrating it in \([\tau,t]\), we have

$$\begin{aligned} & \bigl\Vert \tilde{u}^{m}(t) \bigr\Vert _{2}^{2}+ \varepsilon (t) \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} + \int _{\tau }^{t}\bigl(1-2\varepsilon '(s) \bigr) \bigl\Vert \nabla \tilde{u}^{m}(s) \bigr\Vert _{2}^{2}\,ds +2c_{1} \int _{\tau }^{t} \bigl\Vert \tilde{u}^{m}(s) \bigr\Vert _{p}^{p}\,ds \\ & \quad\leq \bigl\Vert \tilde{u}^{m}(\tau ) \bigr\Vert _{2}^{2}+\varepsilon (\tau ) \bigl\Vert \nabla \tilde{u}^{m}(\tau ) \bigr\Vert _{2}^{2} + \bigl(2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}\bigr) (t- \tau )\quad \text{for all } t\geq \tau. \end{aligned}$$

Hence,

$$\begin{aligned} & \bigl\Vert \tilde{u}^{m}(t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} + \int _{\tau }^{t} \bigl\Vert \nabla \tilde{u}^{m}(s) \bigr\Vert _{2}^{2}\,ds +2c_{1} \int _{\tau }^{t} \bigl\Vert \tilde{u}^{m}(s) \bigr\Vert _{p}^{p}\,ds \\ & \quad\leq \bigl\Vert \tilde{u}^{m}(\tau ) \bigr\Vert _{2}^{2}+\varepsilon (\tau ) \bigl\Vert \nabla \tilde{u}^{m}(\tau ) \bigr\Vert _{2}^{2} + \bigl(2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}\bigr) (t- \tau )\quad \text{for all } t\geq \tau. \end{aligned}$$
(3.1)

So, from (3.1), we can get

$$\begin{aligned} \bigl\{ \tilde{u}^{m}\bigr\} _{m\geq n}\quad \text{is bounded in } L^{\infty }(\tau,t; \mathcal{H}_{t})\cap L^{2}\bigl(\tau,t;H_{0}^{1}(\varOmega )\bigr)\cap L^{p}\bigl( \tau,t;L^{p}(\varOmega )\bigr) \end{aligned}$$
(3.2)

for all \(t>\tau \).

Moreover, combining with (1.5) and (3.2), we obtain

$$\begin{aligned} \bigl\{ f\bigl(\tilde{u}^{m}\bigr)\bigr\} _{m\geq n} \quad\text{is bounded in } L^{q}\bigl(\tau,t;L^{q}( \varOmega )\bigr) \text{ for all } t>\tau, \end{aligned}$$

where \(q=p/(p-1)\).

Then there exist functions \(\tilde{u}\in L^{\infty }(\tau,t; \mathcal{H}_{t}) \cap L^{2}(\tau,t;H_{0}^{1}( \varOmega ))\cap L^{p}(\tau,t;L^{p}(\varOmega ))\) and \(\tilde{\chi }\in L^{q}(\tau, t; L^{q}(\varOmega ))\) for all \(t>\tau \), and a subsequence such that

$$\begin{aligned} \textstyle\begin{cases} \tilde{u}^{m}\rightarrow \tilde{u} &\text{weakly-star in } L^{ \infty }(\tau,t;\mathcal{H}_{t}), \\ \tilde{u}^{m}\rightarrow \tilde{u} &\text{weakly in } L^{2} (\tau, t; H_{0}^{1}( \varOmega )), \\ \tilde{u}^{m}\rightarrow \tilde{u} &\text{weakly in } L^{p}(\tau, t; L^{p}( \varOmega )), \\ f(\tilde{u}^{m})\rightarrow \tilde{\chi } &\text{weakly in } L^{q}( \tau, t; L^{q}(\varOmega )). \end{cases}\displaystyle \end{aligned}$$
(3.3)

Step 2: Uniform estimate for the time derivatives. Multiplying each equation of the approximate system by \(\gamma '_{mj}(t)\) and summing from \(j=1\) to m, we arrive at

$$\begin{aligned} &\bigl\Vert \bigl(\tilde{u}^{m}\bigr)'(t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \bigl(\nabla \tilde{u}^{m}\bigr)'(t) \bigr\Vert _{2}^{2} +\frac{1}{2}\frac{d}{dt} \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} \\ &\quad{}+\bigl(f\bigl(\tilde{u}^{m}\bigr),\bigl(\tilde{u}^{m} \bigr)'(t)\bigr)=\bigl(g(x),\bigl(\tilde{u}^{m} \bigr)'(t)\bigr),\quad \text{a.e. } t>\tau. \end{aligned}$$

By the Hölder and Young inequalities, we have

$$\begin{aligned} &\bigl\Vert \bigl(\tilde{u}^{m}\bigr)'(t) \bigr\Vert _{2}^{2}+2\varepsilon (t) \bigl\Vert \bigl(\nabla \tilde{u}^{m}\bigr)'(t) \bigr\Vert _{2}^{2} +\frac{d}{dt} \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} \\ &\quad +2\frac{d}{dt} \int _{\varOmega }\mathcal{F}\bigl(\tilde{u}^{m}(t,x)\bigr) \,dx \leq \Vert g \Vert _{2}^{2},\quad \text{a.e. }t>\tau. \end{aligned}$$

Integrating it from τ to t, and from (1.6) we can get

$$\begin{aligned} & \bigl\Vert \nabla \tilde{u}^{m}(t) \bigr\Vert _{2}^{2} +2\tilde{c}_{1} \bigl\Vert \tilde{u}^{m}(t) \bigr\Vert _{p}^{p}+ \int _{\tau }^{t} \bigl( \bigl\Vert \bigl( \tilde{u}^{m}\bigr)'(s) \bigr\Vert _{2}^{2} + \varepsilon (t) \bigl\Vert \bigl(\nabla \tilde{u}^{m} \bigr)'(s) \bigr\Vert _{2}^{2} \bigr)\,ds \\ &\quad \leq 4\tilde{c}_{0} \vert \varOmega \vert + \bigl\Vert \nabla \tilde{u}^{m}(\tau ) \bigr\Vert _{2}^{2} +2 \tilde{c}_{2} \bigl\Vert \tilde{u}^{m}(\tau ) \bigr\Vert _{p}^{p}+ \Vert g \Vert _{2}^{2}(t- \tau ) \end{aligned}$$
(3.4)

for all \(t\geq \tau \) and any \(m\geq n\).

Since \(\tilde{u}_{\tau }^{m}=u_{\tau }\) for all \(m\geq n\) and \(\tilde{u}^{m}_{\tau }\in H_{0}^{1}(\varOmega )\cap L^{p}(\varOmega )\), by (3.4), we obtain

$$\begin{aligned} \bigl\{ \tilde{u}^{m}(t)\bigr\} _{m\geq n} \quad\text{is bounded in } L^{\infty }\bigl(\tau, t; H_{0}^{1}( \varOmega )\cap L^{p}(\varOmega )\bigr) \end{aligned}$$
(3.5)

and

$$\begin{aligned} \bigl\{ \bigl(\tilde{u}^{m}\bigr)'(t)\bigr\} _{m\geq n}\quad \text{is bounded in } L^{2}(\tau, t; \mathcal{H}_{t}) \end{aligned}$$
(3.6)

for all \(t>\tau \). Then there exist functions \(\tilde{u}\in L^{\infty }(\tau, t; H_{0}^{1}(\varOmega ) \cap L^{p}( \varOmega ))\) and \(\tilde{u}_{t}\in L^{2}(\tau, t; \mathcal{H}_{t})\) for all \(t>\tau \), which improve the regularity of ũ obtained in Step 1.

For any fixed \(t>\tau \), since

$$\begin{aligned} \bigl\Vert \tilde{u}^{m}(t_{2})- \tilde{u}^{m}(t_{1}) \bigr\Vert _{\mathcal{H}_{t}}^{2} &= \bigl\Vert \tilde{u}^{m}(t_{2})-\tilde{u}^{m}(t_{1}) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla \tilde{u}^{m}(t_{2})-\nabla \tilde{u}^{m}(t_{1}) \bigr\Vert _{2}^{2} \\ &= \biggl\Vert \int _{t_{1}}^{t_{2}}\bigl(\tilde{u}^{m} \bigr)'(s)\,ds \biggr\Vert _{2}^{2} + \varepsilon (t) \biggl\Vert \int _{t_{1}}^{t_{2}}\bigl(\nabla \tilde{u}^{m} \bigr)'(s)\,ds \biggr\Vert _{2}^{2} \\ &\leq \bigl( \bigl\Vert \bigl(\tilde{u}^{m}\bigr)' \bigr\Vert _{L^{2}(\tau,t;L^{2}(\varOmega ))}^{2} + \varepsilon (t) \bigl\Vert \bigl( \nabla \tilde{u}^{m}\bigr)' \bigr\Vert _{L^{2}(\tau,t;L^{2}( \varOmega ))}^{2}\bigr) \vert t_{2}-t_{1} \vert \\ &= \bigl\Vert \bigl(\tilde{u}^{m}\bigr)' \bigr\Vert _{L^{2}(\tau,t;\mathcal{H}_{t})}^{2} \vert t_{2}-t_{1} \vert , \end{aligned}$$
(3.7)

for all \(t_{1},t_{2}\in [\tau,t]\), from (3.5), (3.6) and (3.7), by the Ascoli–Arzelà Theorem, and taking into account the initial data for all the sequence, we deduce that there is a subsequence such that

$$\begin{aligned} \tilde{u}^{m}\rightarrow \tilde{u} \quad\text{in } C\bigl([ \tau,t]; \mathcal{H}_{t}\bigr) \end{aligned}$$
(3.8)

for all \(t>\tau \) and a.e. in \(\varOmega \times (\tau,\infty )\).

Since \(f\in C(\mathbb{R}, \mathbb{R})\), we conclude that \(f(\tilde{u}^{m})\rightarrow f(\tilde{u}) a.e\). in \(\varOmega \times (\tau, \infty )\). So, combining with (3.3) and [15] (Lemma 1.3, p. 12) we obtain \(\tilde{\chi }=f(\tilde{u})\).

Thus, together with (3.3) and (3.8), by taking the limit in the equations satisfied by \(\{\tilde{u}^{m}\}\) and, thanks to the fact that \(\operatorname{span} \{\omega _{j}\}_{j\geq 1}\) is dense in \(H_{0}^{1}(\varOmega )\cap L^{p}(\varOmega )\), we conclude that ũ is a weak solution of Eq. (1.1).

Step 3: Proof of the general statement by density. For each \(n\in \mathbb{N}\), we define \(u^{n}_{\tau }=\varSigma _{j=1}^{n}(u_{\tau }, \omega _{j})\omega _{j}\). (Due to the fact that \(\{\omega _{j}\}_{j\geq 1}\) is a Hilbert basis of \(L^{2}(\varOmega )\), it is easy to check that \(u^{n}_{\tau }\rightarrow u_{\tau }\) in \(\mathcal{H}_{\tau }\).)

Let also consider a sequence \(\{g^{n}\}_{n=1}^{\infty }\subset L^{2}(\varOmega )\) converging to \(g\in H^{-1}(\varOmega )\).

Denote by \(u^{n}\) the corresponding solution to Eq. (1.1) with g replaced by \(g^{n}\) and initial data \(u_{\tau }^{n}\).

Then, by the energy equality for each \(u^{n}\), we have

$$\begin{aligned} & \bigl\Vert u^{n}(t) \bigr\Vert _{2}^{2}+ \varepsilon (t) \bigl\Vert \nabla u^{n}(t) \bigr\Vert _{2}^{2} +2 \int _{\tau }^{t} \bigl\Vert \nabla u^{n}(s) \bigr\Vert _{2}^{2}\,ds+2 \int _{\tau }^{t}\bigl(f\bigl(u^{n}(s)\bigr), u^{n}(s)\bigr)\,ds \\ &\quad = \bigl\Vert u^{n}(\tau ) \bigr\Vert _{2}^{2}+ \varepsilon (\tau ) \bigl\Vert \nabla u^{n}(\tau ) \bigr\Vert _{2}^{2} +2 \int _{\tau }^{t}\bigl(g^{n}(x), u^{n}(s)\bigr)\,ds,\quad \forall t\geq \tau. \end{aligned}$$

Similar to the reasoning process in Step 1, we get

$$\begin{aligned} \bigl\{ u^{n}\bigr\} \quad \text{is bounded in } L^{\infty }(\tau,t;\mathcal{H}_{t}) \cap L^{2}\bigl( \tau,t;H_{0}^{1}(\varOmega )\bigr) \cap L^{p}\bigl( \tau,t;L^{p}( \varOmega )\bigr) \end{aligned}$$
(3.9)

for all \(t>\tau \).

Now, combining with (1.5) and (3.9), we see that \(\{f(u^{n})\}\) is bounded in \(L^{q}(\tau,t;L^{q}(\varOmega ))\) for all \(t>\tau \).

Therefore, there exist functions \(u\in L^{\infty }(\tau,t;\mathcal{H}_{t})\cap L^{2}(\tau,t;H_{0}^{1}( \varOmega )) \cap L^{p}(\tau,t;L^{p}(\varOmega ))\) and \(\chi \in L^{q}(\tau,t;L^{q}(\varOmega ))\) for all \(t>\tau \), and a subsequence such that

$$\begin{aligned} \textstyle\begin{cases} u^{n}\rightarrow u &\text{weakly-star in } L^{\infty }( \tau, t; \mathcal{H}_{t}), \\ u^{n}\rightarrow u &\text{weakly in } L^{2}(\tau, t; H_{0}^{1}(\varOmega )), \\ u^{n}\rightarrow u &\text{weakly in } L^{p}(\tau, t; L^{p}(\varOmega )), \\ f(u^{n})\rightarrow \chi &\text{weakly in } L^{q}(\tau, t; L^{q}( \varOmega )), \end{cases}\displaystyle \end{aligned}$$
(3.10)

for all \(t> \tau \).

Moreover, we may improve some of the above convergence. Taking into account the energy equality for \(u^{n}-u^{m}\), we have

$$\begin{aligned} &\bigl\Vert u^{n}(t)-u^{m}(t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla u^{n}(t)- \nabla u^{m}(t) \bigr\Vert _{2}^{2} + \int _{\tau }^{t} \bigl\Vert \nabla u^{n}(s)-\nabla u^{m}(s) \bigr\Vert _{2}^{2} \,ds \\ &\quad \leq \bigl\Vert u^{n}(\tau )-u^{m}(\tau ) \bigr\Vert _{2}^{2} +\varepsilon (\tau ) \bigl\Vert \nabla u^{n}(\tau )-\nabla u^{m}(\tau ) \bigr\Vert _{2}^{2} +2l \int _{\tau }^{t} \bigl\Vert u^{n}(s)-u^{m}(s) \bigr\Vert _{2}^{2}\,ds \\ &\qquad{}+ \bigl\Vert g^{n}-g^{m} \bigr\Vert _{H^{-1}}^{2}(t-\tau ),\quad \forall t\geq \tau. \end{aligned}$$
(3.11)

By (3.11), we know that

$$\begin{aligned} \bigl\{ u^{n}\bigr\} \quad \text{is a Cauchy sequence in } C\bigl([\tau,t]; \mathcal{H}_{t}\bigr) \cap L^{2}\bigl(\tau,t; H_{0}^{1}(\varOmega )\bigr) \text{ for all } t>\tau. \end{aligned}$$

Thus, we have \(u^{n}\rightarrow u\) a.e. in \(\varOmega \times (\tau, \infty )\).

Therefore, as before, combining with (3.10) and [15] (Lemma 1.3, p. 12) we obtain \(\chi =f(u)\); and from (3.10) we may take the limit in the equations satisfied by \(u^{n}\) and conclude that u is a weak solution of Eq. (1.1). □

For the solutions of Eq. (1.1), the following theorem shows the uniqueness and continuity with respect to initial data.

Theorem 3.4

Letfsatisfy (1.4)(1.5), \(g\in H^{-1}(\varOmega )\)and\(u_{\tau }\in \mathcal{H}_{\tau }\), then the weak solution of Eq. (1.1) is unique. Moreover, for every two solutions\(u^{1}(t)\)and\(u^{2}(t)\) (with different initial data), the following Lipschitz continuity holds:

$$\begin{aligned} \bigl\Vert \omega (t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla \omega (t) \bigr\Vert _{2}^{2} \leq \bigl( \Vert \omega _{\tau } \Vert _{2}^{2}+ \varepsilon (\tau ) \Vert \nabla \omega _{ \tau } \Vert _{2}^{2}\bigr)e^{2l(t-\tau )}, \quad\forall t\geq \tau, \end{aligned}$$

where\(\omega (t)=u^{1}(t)-u^{2}(t)\).

Proof

Let \(\omega (t)=u^{1}(t)-u^{2}(t)\), then \(\omega (t)\) satisfies the following equation:

$$\begin{aligned} \textstyle\begin{cases} \omega _{t}-\varepsilon (t)\Delta \omega _{t}-\Delta \omega =f(u^{1})-f(u^{2}) &\text{in } \varOmega \times (\tau,\infty ), \\ \omega (x, t)=0 &\text{on } \partial \varOmega \times (\tau,\infty ), \\ \omega (x, \tau )=u^{1}_{\tau }-u^{2}_{\tau }, &x\in \varOmega. \end{cases}\displaystyle \end{aligned}$$
(3.12)

Taking the \(L^{2}\)-inner product between (3.12) and ω, and using (1.4), we have

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert \omega \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla \omega \Vert _{2}^{2}\bigr) + \bigl(2-\varepsilon '(t)\bigr) \Vert \nabla \omega \Vert _{2}^{2} \leq 2l \Vert \omega \Vert _{2}^{2}. \end{aligned}$$

Then

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert \omega \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla \omega \Vert _{2}^{2}\bigr) \leq 2l\bigl( \Vert \omega \Vert _{2}^{2}+\varepsilon (t) \Vert \nabla \omega \Vert _{2}^{2}\bigr). \end{aligned}$$

By the Gronwall lemma, it yields

$$\begin{aligned} \bigl\Vert \omega (t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla \omega (t) \bigr\Vert _{2}^{2} \leq \bigl( \Vert \omega _{\tau } \Vert _{2}^{2}+ \varepsilon (\tau ) \Vert \nabla \omega _{ \tau } \Vert _{2}^{2}\bigr)e^{2l(t-\tau )}, \end{aligned}$$

and the uniqueness holds. □

Thus, we define the solution processes \(\{U(t,\tau )\}_{t\geq \tau }\) in the spaces \(\mathcal{H}_{t}\) as:

$$\begin{aligned} U(t,\tau ): \mathcal{H}_{\tau }\rightarrow \mathcal{H}_{t},\qquad U(t, \tau )u_{\tau }=u(t),\quad \forall t\geq \tau. \end{aligned}$$
(3.13)

Moreover, Theorem 3.4 shows that the process \(\{U(t,\tau )\}_{t\geq \tau }\) is Lipschitz in \(\mathcal{H}_{t}\):

$$ \bigl\Vert U(t,\tau )u_{\tau }^{1}-U(t,\tau )u_{\tau }^{2} \bigr\Vert _{\mathcal{H}_{t}} \leq \bigl\Vert u_{\tau }^{1}-u_{\tau }^{2} \bigr\Vert _{\mathcal{H}_{\tau }}e^{2l(t- \tau )}, \quad\forall t\geq \tau. $$

3.2 Time-dependent global attractors

In this subsection, we will verify the existence of the time-dependent global attractors in \(\mathcal{H}_{t}\) for the process \(\{U(t, \tau )\}_{t\geq \tau }\) defined by (3.13).

3.2.1 Time-dependent absorbing sets

In the following, we will obtain the time-dependent global absorbing sets.

Lemma 3.5

Letfsatisfy (1.4)(1.5), \(g\in H^{-1}(\varOmega )\)and\(u_{\tau }\in \mathbb{B}_{\tau }(R)\subset \mathcal{H}_{\tau }\). Then there exists a\(R_{0}>0\)such that the family\(\hat{B}=\{B_{t}(R_{0})\}_{t\in \mathbb{R}}\)is a time-dependent absorbing set for the process\(\{U(t, \tau )\}_{t\geq \tau }\).

Proof

Multiplying (1.1) by \(u(t)\) and integrating over \(x\in \varOmega \), we arrive at

$$\begin{aligned} \frac{1}{2}\frac{d}{dt}\bigl( \Vert u \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla u \Vert _{2}^{2}\bigr) + \biggl(1-\frac{1}{2}\varepsilon '(t)\biggr) \Vert \nabla u \Vert _{2}^{2}+\bigl(f(u), u\bigr)= \bigl\langle g(x), u \bigr\rangle . \end{aligned}$$

Thanks to (1.5) and the Hölder inequality, we have

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert u \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla u \Vert _{2}^{2}\bigr) + \bigl(1- \varepsilon '(t)\bigr) \Vert \nabla u \Vert _{2}^{2}+2c_{1} \Vert u \Vert _{p}^{p}\leq 2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}. \end{aligned}$$

Furthermore, by (1.3), we can get

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert u \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla u \Vert _{2}^{2}\bigr) + \frac{1}{1+L}\bigl(\lambda _{1} \Vert u \Vert _{2}^{2}+\varepsilon (t) \Vert \nabla u \Vert _{2}^{2}\bigr) \leq 2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}. \end{aligned}$$

Setting \(\lambda =\min \{\lambda _{1},1\}\) and \(\beta =\frac{\lambda }{1+L}\), we deduce that

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert u \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla u \Vert _{2}^{2}\bigr) + \beta \bigl( \Vert u \Vert _{2}^{2}+\varepsilon (t) \Vert \nabla u \Vert _{2}^{2}\bigr)\leq 2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}. \end{aligned}$$
(3.14)

Multiplying (3.14) by \(e^{\beta t}\) and integrating it in \([\tau, t]\), we obtain

$$\begin{aligned} &\bigl( \bigl\Vert u(t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla u(t) \bigr\Vert _{2}^{2} \bigr)e^{\beta t} \\ & \quad\leq \bigl( \bigl\Vert u(\tau ) \bigr\Vert _{2}^{2}+ \varepsilon (\tau ) \bigl\Vert \nabla u(\tau ) \bigr\Vert _{2}^{2} \bigr)e^{\beta \tau } +\bigl(2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}\bigr) \int _{ \tau }^{t}e^{\beta s}\,ds,\quad \forall t\geq \tau. \end{aligned}$$

Therefore,

$$\begin{aligned} \bigl( \bigl\Vert u(t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla u(t) \bigr\Vert _{2}^{2}\bigr) \leq {}&\bigl( \bigl\Vert u( \tau ) \bigr\Vert _{2}^{2}+ \varepsilon (\tau ) \bigl\Vert \nabla u(\tau ) \bigr\Vert _{2}^{2} \bigr)e^{- \beta (t-\tau )} +\frac{1}{\beta }\bigl(2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}\bigr) \\ \leq {}&1+\frac{1}{\beta }\bigl(2c_{0} \vert \varOmega \vert + \Vert g \Vert _{H^{-1}}^{2}\bigr)=R_{0}, \end{aligned}$$

provided that \(t-\tau \geq t_{0}\) with \(t_{0}=\frac{1}{\beta }\ln (\|u_{\tau }\|_{2}^{2}+\varepsilon (\tau ) \|\nabla u_{\tau }\|_{2}^{2})\), from which we obtain the existence of the time-dependent absorbing set. □

3.2.2 Time-dependent global attractors

At first, we have the following lemma, which is similar to that in [15].

Lemma 3.6

Letfsatisfy (1.4)(1.5), \(g\in H^{-1}(\varOmega )\), \(u_{\tau }\in \mathcal{H}_{\tau }\)and\(\{u^{n}(t)\}_{n=1}^{\infty }\)be a sequence of solutions for Eq. (1.1) with initial data\(u^{n}_{\tau }\in \mathcal{H}_{\tau }\)\((n=1,2,\ldots )\), then there exists a subsequence of\(\{u^{n}(t)\}_{n=1}^{\infty }\)that converges strongly in\(L^{2}(\tau,t; L^{2}(\varOmega ))\).

Proof

By (1.5) and Theorem 3.3, we know that there exists a sequence \(\{u^{n}(t)\}_{n=1}^{\infty }\subset L^{2}(\tau,T;H_{0}^{1}(\varOmega ))\), \(\{f(u^{n}(t))\}_{n=1}^{\infty }\subset L^{q}(\tau,T;L^{q}(\varOmega ))\). Then, from Eq. (1.1), we obtain \(\partial _{t}u^{n}-\varepsilon (t)\partial _{t}\Delta u^{n}=\Delta u^{n}-f(u^{n})+g(x) \in L^{2}(\tau,T;H^{-1}(\varOmega ))+L^{q}(\tau,T;L^{q}(\varOmega )) \subset L^{2}(\tau,T;H^{-2}(\varOmega ))\). By the regularization theory for elliptic equations, we know that \(\partial _{t}u^{n}\in L^{2}(\tau,T;L^{2}(\varOmega ))\). As in [15], there exists a subsequence of \(\{u^{n}(t)\}_{n=1}^{\infty }\) (still denoted by \(\{u^{n}(t)\}_{n=1}^{\infty }\)) that converges strongly in \(L^{2}(\tau,T; L^{2}(\varOmega ))\). □

Then we have the following theorem, which will obtain the pullback asymptotic compactness for the process \(\{U(t,\tau )\}_{t\geq \tau }\) defined by (3.13).

Theorem 3.7

Letfsatisfy (1.4)(1.5), \(g\in H^{-1}(\varOmega )\)and\(u_{\tau }\in \mathbb{B}_{\tau }(R)\subset \mathcal{H}_{\tau }\), then\(\{U(t, \tau )\}_{t\geq \tau }\)is pullback asymptotically compact in\(\mathcal{H}_{t}\).

Proof

Let \(u^{i}(t)\) (\(i=1,2\)) be the solutions corresponding to initial data \(u^{i}_{\tau}\in\mathbb{B}_{\tau}(R)\subset\mathcal{H}_{\tau}\), that is, \(u^{i}(t)\) satisfies the following equation:

$$\begin{aligned} u_{t}-\varepsilon (t)\Delta u_{t}-\Delta u+f(u)=g(x),\quad \text{in } \varOmega \times (\tau,\infty ), \end{aligned}$$

with initial data

$$\begin{aligned} u^{i}(x,\tau )=u^{i}_{\tau },\quad x\in \varOmega. \end{aligned}$$

Denoting \(\omega (t)=u^{1}(t)-u^{2}(t)\), then \(\omega (t)\) satisfies the following equation:

$$\begin{aligned} \omega _{t}-\varepsilon (t)\Delta \omega _{t}-\Delta \omega +f\bigl(u^{1}\bigr)-f \bigl(u^{2}\bigr)=0,\quad \text{in } \varOmega \times (\tau,\infty ), \end{aligned}$$
(3.15)

with initial data

$$\begin{aligned} \omega (x,\tau )=u^{1}_{\tau }-u^{2}_{\tau },\quad x\in \varOmega. \end{aligned}$$

Multiplying (3.15) by \(\omega (t)\) and integrating it in Ω, then, by (1.4), we obtain

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert \omega \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla \omega \Vert _{2}^{2}\bigr) + \bigl(2-\varepsilon '(t)\bigr) \Vert \nabla \omega \Vert _{2}^{2}\leq 2l \Vert \omega \Vert _{2}^{2}. \end{aligned}$$

By the Poincaré inequality, we have

$$\begin{aligned} \frac{d}{dt}\bigl( \Vert \omega \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla \omega \Vert _{2}^{2}\bigr) +\beta _{1}\bigl( \Vert \omega \Vert _{2}^{2}+ \varepsilon (t) \Vert \nabla \omega \Vert _{2}^{2}\bigr) \leq 2l \Vert \omega \Vert _{2}^{2}, \end{aligned}$$

where \(\beta _{1}=2\beta \), β is given by (3.14).

Thanks to the Gronwall lemma, we get

$$\begin{aligned} &\bigl\Vert \omega (t) \bigr\Vert _{2}^{2}+\varepsilon (t) \bigl\Vert \nabla \omega (t) \bigr\Vert _{2}^{2}\\ &\quad \leq \bigl( \bigl\Vert \omega (\tau ) \bigr\Vert _{2}^{2}+ \varepsilon (\tau ) \bigl\Vert \nabla \omega ( \tau ) \bigr\Vert _{2}^{2}\bigr) e^{-\beta _{1}(t-\tau )}+2l \int _{\tau }^{t} \bigl\Vert \omega (s) \bigr\Vert _{2}^{2}\,ds,\quad \forall t\geq \tau. \end{aligned}$$

Setting

$$\begin{aligned} \psi_{\tau}^{t}(u_{\tau}^{1},u_{\tau}^{2})=&2l \int _{\tau }^{t} \bigl\Vert \omega (s) \bigr\Vert _{2}^{2}\,ds, \end{aligned}$$

combining with Definition 2.9 and Lemma 3.6, we know that \(\psi_{\tau}^{t}(\cdot,\cdot)\) is a contractive function. Then, for any \(\epsilon >0\) and any fixed \(t\in \mathbb{R}\), let \(\tau _{0}=t-\frac{1}{\beta _{1}}\ln \frac{\|w_{\tau }\|_{2}^{2} +\varepsilon (\tau )\|\nabla w_{\tau }\|_{2}^{2}}{\epsilon }\), we easily see that \(\{U(t,\tau )\}_{t\geq \tau }\) is pullback asymptotically compact in \(\mathcal{H}_{t}\) by Theorem 2.10. □

Combining with Lemma 3.5 and Theorem 3.7, we have the main result of this paper.

Theorem 3.8

Letfsatisfy (1.4)(1.5), \(g\in H^{-1}(\varOmega )\)and\(u_{\tau }\in \mathbb{B}_{\tau }(R)\subset \mathcal{H}_{\tau }\), then\(\{U(t,\tau )\}_{t\geq \tau }\)possesses a time-dependent global attractor\(\hat{\mathcal{A}}=\{\mathcal{A}_{t}\}_{t\in \mathbb{R}}\)in\(\mathcal{H}_{t}\); that is, \(\mathcal{A}_{t}\)is compact, \(\hat{\mathcal{A}}\)is nonempty, invariant in\(\mathcal{H}_{t}\)and pullback attracts every bounded subset of\(\mathcal{H}_{t}\)with respect to the\(\mathcal{H}_{t}\)-norm.

Remark 3.9

In Theorem 3.8, we have obtained the time-dependent global attractor \(\hat{\mathcal{A}}=\{\mathcal{A}_{t}\}_{t\in \mathbb{R}}\) in \(\mathcal{H}_{t}\). From (1.2) we know that \(\varepsilon (t)\rightarrow 0\) as \(t\rightarrow +\infty \), then Eq. (1.1) becomes the classical reaction–diffusion equation \(u_{t}-\Delta u+f(u)=g(x)\). An interesting question is about the limitation of \(\mathcal{A}_{t}\) as \(t\rightarrow +\infty \), that is, how to describe \(\lim_{t\rightarrow +\infty }\mathcal{A}_{t}\)? We will consider this problem in our next work.