1 Introduction

In this paper, we study the existence of a global attractor for the following evolutionary partial differential equation:

$$\begin{aligned} \left\{ \begin{array}{rll} \displaystyle u_{t}-\Delta _{p}u + \mu |u|^{q-2}u + f(x,u) &{} =\;\; g(x), &{} x\in \Omega ,\ t>0,\\ \displaystyle u &{} =\;\; 0, &{} x\in \partial \Omega ,\; t>0,\\ \displaystyle u(0,x) &{} =\;\; u_0, &{} x\in \Omega , \end{array} \right. \end{aligned}$$
(1.1)

where \(\Omega \subseteq {\mathbb {R}}^N\) is an arbitrary bounded or unbounded smooth domain, \(\Delta _{p}u=\text{ div }(|\nabla u|^{p-2}\nabla u)\) is the so-called p-Laplacian, \(\mu >0, 1<p<N\), and \(q>2\). Moreover, the external force f(xu) is a Carathéodory function, i.e., \(f(\cdot , u)\) is measurable and \(f(x, \cdot )\) is continuous, whose properties will be specified later, and the external source \(g\in L^2(\Omega )\cap L^\infty (\Omega )\) is a given function independent of time.

The equation of type (1.1) is usually called a parabolic p-Laplace equation, which appears in many applications in mechanics, physics and biology (non-Newtonian fluids, gas flows in porous media, spread of biological populations, etc., see [6, 9, 10, 18] and the references therein). In the past decades, the well-posedness and the asymptotic behavior for the p-Laplace equation have been thoroughly investigated (see [3,4,5, 11,12,13, 16, 17, 22, 23, 29, 30, 33,34,37]). In particular, Geredeli [16] considered the existence of the global attractor for the equation

$$\begin{aligned} \begin{aligned} u_{t}-\Delta _{p}u + f(u)=g, \ x\in \Omega ,\ t>0, \end{aligned} \end{aligned}$$
(1.2)

where \(\Omega \subset {\mathbb {R}}^N\) is bounded domain and the nonlinearity f satisfies a very general condition, namely,

$$\begin{aligned} \begin{aligned} f\in \mathcal {C}^{1}({\mathbb {R}}),\ \ f(0)=0\ \text{ and } \; f^{\prime }(u)\ge -\ell , \end{aligned} \end{aligned}$$
(1.3)

for some \(\ell \ge 0.\) This condition implies that \(-\frac{\ell }{2} |u|^2 \le F(u)\le uf(u) + \frac{\ell }{2}|u|^2\) where \(F(u)=\int _0^u f(s)\,\text{d}s.\)

As in the previous results, for (weighted) p-Laplace equations, nonlinear reaction-diffusion equations or nonlinear wave equations, the existence of a global attractor is usually proved when the nonlinear external force satisfies a doubly bounded condition

$$\begin{aligned} \begin{aligned} \alpha _1|u|^{q}-\beta \le uf(u)\le \alpha _2|u|^{q}+\beta , \end{aligned} \end{aligned}$$
(1.4)

for some \(\alpha _1, \alpha _2>0\) and \(\beta \ge 0\). In particular, Yang-Sun and Zhong in [34] considered the existence of a global attractor for a p-Laplace equation in \({\mathbb {R}}^N\) under a condition similar to both (1.3) and (1.4). There are many papers referring to conditions either of the type (1.3) or (1.4), see e.g. [25, 29, 35]. As pointed out in [16], a main difficulty arises in the proof of the well-posedness for (1.2) when we simplify the condition of f only involving (1.3). The critical issue in the proof of well-posedness is to deal with the limiting procedure on f which is overcome by weak convergence techniques in Orlicz spaces. However, the boundedness of the domain \(\Omega \subset {\mathbb {R}}^N\) plays an essential role in the proof, since the author of [16] relies on the Poincaré inequality to get the well-posedness of (1.2) (global estimates of solutions) and an absorbing set for the semigroup. With a doubly bounded condition on f, the authors of [34] also obtain the global well-posedness and global attractor for the p-Laplace equation in unbounded domains (or \({\mathbb {R}}^N\)) without the help of the Poincaré inequality. Moreover, the authors in [4] also considered quasilinear parabolic equations with weighted p-Laplacian operators of the type \(-\text{ div }(\sigma (x)|\nabla u|^{p-2}\nabla u)\) on unbounded domains. In that situation, they find suitable conditions on the weight function \(\sigma (x)\) to obtain global existence of solutions. However, in our situation, we cannot use this property to get the desired result. Finally, we also point out that the condition \(f^{\prime }(u)\ge -\ell\) also helps to get uniqueness of solutions; actually, it is another form of the estimate

$$\begin{aligned} (f(u)-f(v))(u-v)\ge -\ell |u-v|^2. \end{aligned}$$
(1.5)

In this paper, we intend to generalize the conditions (1.3) and (1.5), and assume

$$\begin{aligned} f(x,u)=f_1(u)+m(x)f_2(u),\end{aligned}$$
(1.6)

where \(f_1, f_2\) and the weight function m(x) satisfy the following conditions:

(\(H_m\))::

\(m\in L^{\frac{N}{p}}(\Omega )\cap L^{\infty }(\Omega )\), \(m\ge 0\) and, if \(\Omega\) is unbounded, then additionally \(m\not \equiv 0.\)

(\(H^1_{f}\))::

There exists a constant \(a^*, \nu ,\nu _- \in (0,\infty )\) such that, with \(p^*=\frac{pN}{N-p}\) and \(2<q_-<q\),

$$\begin{aligned}&\liminf _{|u|\rightarrow 0}\frac{f_1(u)u}{|u|^2}\ge a^*,\quad \underset{u\ne 0}{{\text {ess}}\inf \,} \frac{f_1(u)u}{|u|^{q_-}}\ge -\nu _-, \\&\underset{u\ne 0}{{\text {ess}}\inf \,} \frac{f_1(u)u}{|u|^{q}}\ge -\nu \ \text{ and } \ \liminf _{|u|\rightarrow 0}\frac{f_2(u)u}{|u|^{p^*}}\ge -a^*. \end{aligned}$$
(1.7)
(\(H_{f}^2\))::

The functions \(f_1\) and \(f_2\) are continuous on \({\mathbb {R}}\), \(f_1(0)=f_2(0)=0\) and there holds for all \(u,v\in {\mathbb {R}}\)

$$\begin{aligned}&\left| f_1(u)-f_1(v)\right| \le \nu (|u-v|+ |u-v|^{q-1}) , \end{aligned}$$
(1.8)
$$\begin{aligned} 2< p<N:&\left| f_2(u)-f_2(v)\right| \le \lambda |u-v|^{p-1}, \end{aligned}$$
(1.9)
$$\begin{aligned} 1<p\le 2:&\left| f_2(u)-f_2(v)\right| \le \lambda \min \big \{|u-v|,|u-v|^{p-1}\big \}. \end{aligned}$$
(1.10)
(\(H_c\))::

The coefficients \(\nu ,\lambda\) in \(H_f^2\) satisfy the conditions

$$\begin{aligned} \lambda< \min (1,c_p) \lambda _1 \quad \text { and }\; \nu <\min (1,c_q')\mu , \end{aligned}$$

where \(\lambda _1=\lambda _1(\Omega )>0\) is the first eigenvalue of a modified eigenvalue problem of \(-\Delta _p\) on \(\Omega\). Moreover, \(c_p>0\) is the constant in the strong monotonicity estimate of \(-\Delta _p\), i.e., \((|\nabla u|^{p-2}\nabla u - |\nabla v|^{p-2}\nabla v, \nabla (u-v)) \ge c_p|\nabla u-\nabla v|^{\max (p,2)}\) and \(c_q'>0\) is an analogous constant in an estimate for the operator defined by \(|u|^{q-2}u\); for estimates of that type see [27, Lemma 5.1.19]. We note that the proof of an absorbing set in Sect. 4 requires a further smallness assumption on \(\lambda\), namely

$$\begin{aligned} \frac{\lambda }{\lambda _1} < \Big (\frac{p}{p+k_0(q-2)}\Big )^p (k_0(q-2)+1) \end{aligned}$$
(1.11)

where \(k_0\in {\mathbb {N}}\) is defined to satisfy the assumption \(k_0(q-2)+2>\frac{N(p+q-2)}{N-p}\), see (4.18) and (4.28) in the proof of Proposition 4.4.

The first (principal) eigenvalue \(\lambda _1\) of the eigenvalue problem

$$\begin{aligned} -\Delta _p u = \lambda m(x) |u|^{p-2}u\quad \text {in }\Omega ,\; u=0\;\;\text {on } \partial \Omega \end{aligned}$$
(1.12)

is positive, see [15] and, for more details, Proposition 2.7 below. For a bounded smooth domain \(\Omega\), thanks to the pioneering work [2, 8, 19], there is a largest constant \(\lambda _1=\lambda _1(\Omega )>0\) such that

$$\begin{aligned} \int _\Omega |u|^p\,\text{d}x \le \frac{1}{\lambda _1} \int _\Omega |\nabla u|^p\,\text{d}x \end{aligned}$$
(1.13)

for all functions u vanishing on \(\partial \Omega\). In other words, \(\lambda _1\) is the smallest eigenvalue of the eigenvalue problem \(-\Delta _p u = \lambda |u|^{p-2}u\) on \(\Omega\) with \(u=0\) on \(\partial \Omega\). However, as in the proof of Theorem 3.2 below, when \(\Omega\) is increasing to an unbounded domain, \(\lambda _1(\Omega )\) will tend to 0 as is well-known in the case \(p=2\). To this aim, \(\lambda _1\) is defined as the principal eigenvalue of (1.12) where the assumptions on m imply that the first eigenvalue of the operator \(-\Delta _p\) is positive. This allows for estimates of Poincaré type (1.13) on a sequence of bounded invading domains \(\Omega _R\subset \Omega\) which are independent of R.

Our main results can be summarized in the following theorem.

Theorem 1.1

Let \(1<p<N, q>2\) and let \(\Omega\) be an arbitrary (bounded or unbounded) smooth domain in \({\mathbb {R}}^{N}, N\ge 2\). Assume further that (\(H_m\)), (\(H^1_{f}\)) and (\(H_{f}^2\)) hold.

(1) For each \(g\in L^{2}(\Omega )\cap L^{\infty }(\Omega )\) and \(u_0\in L^{2}(\Omega )\) problem (1.1) has a unique global weak solution u such that

$$\begin{aligned} u\in L^{\infty }(0,T; L^{2}(\Omega ))\cap L^{p}(0,T; \mathcal {D}^{1,p}_0(\Omega ))\cap L^{q}(0,T; L^{q}(\Omega ))\ \ \text{ for } \text{ any }\ T>0. \end{aligned}$$
(1.14)

(2) Further assume that \(p>\frac{2N}{N+2}\) and \(\lambda \in (0, \lambda _1)\) is sufficiently small in (\(H_{f}^2\)). Then the nonlinear operator semigroup \((S(t))_{t\ge 0}\) on \(L^{2}(\Omega )\) defined by problem (1.1), i.e., \(S(t)u_0=u(t)\), possesses an invariant compact global attractor \(\mathcal {A}\) in \(L^{2}(\Omega )\cap L^{q}(\Omega )\).

(3) If \(\Omega\) is a bounded domain and \(p\ge 2\), then the global attractor has finite fractal dimension.

The main achievements of this paper are as follows:

(1) We get the existence of global-in-time solutions and of a global attractor for (1.1) when the nonlinear term f(u) does not satisfy the doubly bounded condition (1.4), even not in case of unbounded domains. Actually, our results in this paper improve the disadvantages of the above-mentioned results of [16, 34]. On one hand, we extend the result of [16] to unbounded domains. On the other hand, by comparing with the result of [34], our main progress is to get the global attractor for Eq. (1.1) in an unbounded domain with a generalized condition on f(xu) related to the critical exponent.

(2) In the case of unbounded domains, usually the Poincaré inequality or a geometric property of the domain is exploited to get a damping impact; see, for example, Abergel [1] and Marín-Rubio, Real [28] who studied the 2D Navier–Stokes equations on a strip in \({\mathbb {R}}^2\) and arbitrary domains of \({\mathbb {R}}^2\) satisfying the Poincaré inequality, respectively. Especially, the existence of a global attractor is obtained for equations with a damping term (see [14] for the whole space case). In this paper, we not only remove any property of the shape of the domain and the Poincaré inequality, but also weaken the globally acting damping to a local one; actually, only the term \(f_1(u)\) is assumed to have a local damping near the origin.

(3) Furthermore, based on the method of \(\ell\)-trajectories introduced by Málek and Pra\(\check{\text{ z }}\)ák [26], we also get estimates of the finite fractal dimension of the global attractor when \(\Omega\) is a bounded domain. From our proof, one can see that we also obtain an exponential attractor for (1.1).

We will divide the proof of Theorem 1.1 into several sections. For convenience of the reader, we recall the definition of continuous semigroups, the \(\omega\)-limit compactness of semigroups, describe an abstract result on the existence of a global attractor, and discuss the eigenvalue problem (1.12). The existence of global solutions of (1.1) will be proved in Sect. 3. With a further condition on \(\lambda\) and p, we get the existence of the global attractor by proving the \(\omega\)-limit compactness of the semigroup in Sect. 4. In the last section, using the method of \(\ell\)-trajectories, we estimate the finite fractal dimension of the global attractor when \(\Omega\) is a bounded domain.

2 Preliminary results

Let \(\Omega\) be a domain of \({\mathbb {R}}^N, N\ge 2\), and let \(\Omega _T\) be the space-time cylinder \(\Omega \times [0,T]\). The symbols \( L^q (\Omega ), L^r (0,T ;L^q (\Omega )),\) and so forth, denote the usual Lebesgue and Bochner spaces. The homogeneous Sobolev space \(\mathcal {D}^{1,p}_0 (\Omega )\) is defined as the closure of \(\mathcal {C}_0^\infty (\Omega )\) in the norm

$$\begin{aligned} \Vert u\Vert _{\mathcal {D}^{1,p}_0}=\left( \int _{\Omega }|\nabla u|^p\,\text{d}x\right) ^\frac{1}{p}. \end{aligned}$$

Let \(\mathcal {D}^{-1,p^\prime } (\Omega )\) be the dual space of \(\mathcal {D}_0^{1,p} (\Omega )\). Furthermore, we denote

$$\begin{aligned}&V=L^{p}(0,T;\mathcal {D}_0^{1,p} (\Omega ))\cap L^{2}(\Omega _T)\cap L^{q}(\Omega _T),\\&V^*=L^{p^\prime }(0,T;\mathcal {D}^{-1,p^\prime } (\Omega ))+ L^{2}(\Omega _T)+ L^{q^{\prime }}(\Omega _T), \end{aligned}$$

where \(p^\prime\) and \(q^{\prime }\) are the conjugates of p and q, respectively, i.e., \(1/p+1/p^{\prime }=1\) and \(1/q+1/q^{\prime }=1\). In order to simplify the notation, we will write \(L^q(\Omega )\), \(L^r(0,T;L^q(\Omega ))\) and \(\mathcal D^{1,p}_0(\Omega )\) as \(L^q\), \(L^r(0,T;L^q)\) and \(\mathcal D^{1,p}_0\), respectively, except for special circumstances. The letter C denotes a generic constant which may vary from line to line.

Next, we recall the basic concept of Kuratowski’s measure of non-compactness which will be used to establish the \(\omega\)-limit compactness of semigroup. The following definitions and propositions can be found in [24, 37].

Definition 2.1

Let M be a complete metric space. A one-parameter family \(\{S(t)\}_{t\ge 0}\) of operators \(S(t): M\rightarrow M, t\ge 0\), is called a \(C^0\) or continuous semigroup if

  1. 1.

    \( S(0) \text{ is } \text{ the } \text{ identity } \text{ map } \text{ on } M,\)

  2. 2.

    \( \ S(t+s)=S(t)S(s)\ \text{ for } \text{ all } t,s\ge 0,\)

  3. 3.

    \( \text{ for } \text{ each } x\in M \text{ the } \text{ function } S(t)x \text{ is } \text{ continuous } \text{ in } t\ge 0\).

Definition 2.2

Let M be a complete metric space and A be bounded subset of M. The measure of non-compactness \(\kappa (A)\) of A is defined by

$$\begin{aligned} \kappa (A)=\inf \{\delta >0: A \ \text{ admits } \text{ a } \text{ finite } \text{ cover } \text{ by } \text{ sets } \text{ of } \text{ diameter }\le \delta \}. \end{aligned}$$

Definition 2.3

A continuous semigroup \(\{S(t)\}_{t\ge 0}\) in a complete metric space M is called \(\omega\)-limit compact, if for any bounded set \(B\subset M\) and any \(\varepsilon >0\), there exists a time \(t^*\ge 0\) such that

$$\begin{aligned} \kappa \left( \bigcup _{t\ge t^*}S(t)B\right) \le \varepsilon . \end{aligned}$$

Proposition 2.4

Let M be an infinite dimensional Banach space, and \(B(\varepsilon ) \subset M\) a ball of radius \(\varepsilon >0\). Then \(\kappa (B(\varepsilon ))=2\varepsilon .\)

Proposition 2.5

Let \(\{S(t)\}_{t\ge 0}\) be a continuous semigroup in a complete metric space \((M,\rho )\). Then \(\{S(t)\}_{t\ge 0}\) has a global attractor in M if and only if

  1. 1.

    there is a bounded absorbing set \(B\subset M\), and

  2. 2.

    \(\{S(t)\}_{t\ge 0}\) is \(\omega\)-limit compact.

Moreover, we recall the Uniform Gronwall lemma:

Lemma 2.6

(Uniform Gronwall Lemma) Let gh and y be non-negative locally integrable functions on \([0,\infty [\) such that the differential inequality \( \frac{{\text {d}}y(t)}{{\text {d}}t}\le g(t)y(t)+h(t)\) holds for a.a. \(t\ge 0\) and

$$\begin{aligned} \int _{t}^{t+1}g(s)\,{\text {d}}s\le a_1,\quad \int _{t}^{t+1}h(s)\,{\text { d}}s\le a_2,\quad \int _{t}^{t+1}y(s)\,{\text {d}}s\le a_3, \end{aligned}$$

where \(a_1, a_2\) and \(a_3\) are positive constants. Then

$$\begin{aligned} y(t+1)\le \left( a_3 +a_2\right) e^{a_1} \quad \forall t \ge 0. \end{aligned}$$

For the following results we refer to J. Fleckinger-Pellé et al., see [15, Theorem 4.1, Proposition 5.2].

Proposition 2.7

(Principal Eigenvalue) Let \(\Omega \subset {\mathbb {R}}^N\) be an unbounded domain with smooth boundary and let m satisfy \((H_m)\). Then the equation

$$\begin{aligned} -\Delta _p u = \lambda m(x) |u|^{p-2}u \quad \text { in } \Omega ,\; u=0\;\;\text {on }\partial \Omega , \end{aligned}$$
(2.1)

has a smallest positive principal eigenvalue \(\lambda _1=\lambda _1(\Omega )>0\) and a corresponding eigenfunction \(u_1 \in W^{1,p}_m(\Omega ) = \overline{C_0^\infty (\Omega )}^{\Vert \cdot \Vert _{W^{1,p}_m}}\) where

$$\begin{aligned} \Vert u\Vert _{W^{1,p}_m} = \Vert \nabla u\Vert _{L^p} + \Bigg ( \int _\Omega m(x)|u|^p\,\mathrm{d}x\Bigg )^{1/p}.\end{aligned}$$

Moreover,

$$\begin{aligned} \lambda _1=\min \Bigg \{\int _\Omega |\nabla u|^p\,\mathrm{d}x: u\in W^{1,p}_m(\Omega ) ,\; \int _\Omega m(x)|u|^p\,\mathrm{d}x=1 \Bigg \}. \end{aligned}$$
(2.2)

As a consequence we mention that for any smooth bounded subdomain \(\Omega '\subset \Omega\) the positive principal eigenvalue \(\lambda _1(\Omega ')\) for the eigenvalue problem (2.1) in \(\Omega '\) satisfies

$$\begin{aligned} \lambda _1(\Omega ') \ge \lambda _1(\Omega ). \end{aligned}$$
(2.3)

Indeed, first we note that Proposition 2.7 also holds for the domain \(\Omega '\). Now, given any \(u\in C_0^\infty (\Omega ')\) with \(\Vert m^{1/p} u\Vert _{L^p(\Omega ')}=1\), there holds

$$\begin{aligned} \lambda _1(\Omega ') \int _{\Omega '} m|u|^p\,\mathrm{d}x\le \int _{\Omega '} |\nabla u|^p\,\mathrm{d}x= \int _{\Omega } |\nabla {\overline{u}}|^p\,\mathrm{d}x\end{aligned}$$

where \({\overline{u}}\) denotes the extension of u from \(\Omega '\) to \(\Omega\) by 0. Since \(\Vert m^{1/p} \,{\overline{u}}\Vert _{L^p(\Omega )}=1\) as well, we get (2.3).

3 Global well-posedness

In this section, we prove the global well-posedness of Eq. (1.1).

Definition 3.1

A function u(tx) is called a weak solution of (1.1) on [0, T] iff \(u\in V,\) \( u_t\in V^*\) and \(u|_{t=0}=u_0\) almost everywhere in \(\Omega\) such that

$$\begin{aligned}&\int _{\Omega _T} \left( \frac{\partial u}{\partial t}\xi + |\nabla u|^{p-2}\nabla u\nabla \xi + \mu |u|^{q-2}u\xi \right) \text{d }x\,\text{d}t + \int _{\Omega _T} f(x,u)\xi \,\text{d}x\,\text{d}t \\&\quad = \int _{\Omega _T} g(x)\xi \,\text{d}x\,\text{d}t \end{aligned}$$
(3.1)

for all test functions \(\xi \in V\).

Let us define the nonlinear operator

$$\begin{aligned} A: \mathcal {D}_0^{1,p}\rightarrow \mathcal {D}^{-1,p^\prime }, \quad A\varphi := - \text{ div }(|\nabla \varphi |^{p-2}\nabla \varphi ). \end{aligned}$$

It is well-known that the operator A is bounded, monotone and hemicontinuous. Next, we prove the existence of the weak solution to (1.1) by Galerkin’s method.

Theorem 3.2

Let the assumptions \((H^1_{f}),(H^2_{f})\) and \((H_m)\) hold. Then for any \(u_0\in L^2\) and any fixed \(T>0,\) there exists a unique weak solution to (1.1).

Proof

In the first part of the proof we assume that \(\Omega \subset {\mathbb {R}}^N\) is a bounded domain. For the final part of the proof concerning unbounded domains it will be crucial that estimates are independent of the domain \(\Omega\). Consider approximate solutions \(u_n(t)\) of the form

$$\begin{aligned} u_n(t)=\sum _{k=1}^{n}u_{nk}(t)e_k, \end{aligned}$$

where \(\{e_j\}_{j=1}^\infty\) is an orthonormal basis of \(L^2\) and has a dense linear hull in \(\mathcal {D}^{1,p}\cap L^{q}\cap L^{2}\). We get \(u_n\) from solving the nonlinear ODE problem

$$\begin{aligned} \Big \langle \frac{\,\text{d}u_n}{\,\text{d}t},e_k\Big \rangle =&- \langle Au_n,e_k\rangle -\langle \mu |u_n|^{q-2}u_n, e_k\rangle - \langle f_1(u_n)+m(x)f_2(u_n), e_k\rangle + \langle g, e_k\rangle , \end{aligned}$$
(3.2)
$$\begin{aligned} \langle u_n(0),e_k\rangle =&\,\langle u_0,e_k\rangle ,\quad k=1,2,\ldots ,n. \end{aligned}$$
(3.3)

By the boundedness, monotonicity and hemicontinuity of \(A: \mathcal {D}_0^{1,p}\rightarrow \mathcal {D}^{-1,p^\prime }\), it follows that the operator A is demicontinuous (see [31, Lemma 2.1 and Lemma 2.2, p. 38]). So, since \(\langle e_j, e_k\rangle )=\delta _{j,k}\) and \(f(x,\cdot )\) is a continuous function, the Peano existence theorem yields at least one local solution \(u_n\) to (3.2) in some interval \([0, T_n)\). Multiplying Eq. (3.2)\(_k\) by the function \(u_{nk}(t)\) for each k, and adding these relations for \(k= 1,\ldots ,n\), we have

$$\begin{aligned}&\frac{1}{2}\frac{\,\text{ d }}{\,\text{d}t}\Vert u_n\Vert _{L^{2}}^2+\Vert \nabla u_n\Vert _{L^{p}}^p+\mu \Vert u_n\Vert _{L^{q}}^q+\int _{\Omega }\left( f_1(u_n)+m(x)f_2(u_n)\right) u_n\,\text{d}x \\&\quad = \int _{\Omega }g(x)u_n\,\text{d}x. \end{aligned}$$
(3.4)

Next, we prove that we can extend the approximate solution \(u_n\) from \([0, T_n)\) to the interval [0, T], for every \(T > 0\). Firstly, by (1.7) and (1.9), one has

$$\begin{aligned} f_2(u)u&\ge - 2a^* |u|^{p^*}\\&\ge -2a^* \sigma _0^{p^*-p}|u|^{p}\ \ \text{ with } \ |u|<\sigma _0 \ \text{ for } \text{ some }\ \sigma _0>0 , \\ f_2(u)u&\ge -\lambda |u|^{p} \quad \text{ for } \ |u|\ge \sigma _0, \end{aligned}$$

so that

$$\begin{aligned} f_2(u)u \ge -\lambda _0 |u|^{p}\ \ \text{ for } \text{ all } u\in {\mathbb {R}}, \end{aligned}$$
(3.5)

where \(\lambda _0 := 2a^* \sigma _0^{p^*-p} + \lambda\). Since \(\lambda < \lambda _1(\Omega )\) we find \(\sigma _0>0\) small enough such that \(\lambda _0 < \lambda _1(\Omega )\). Then the property of the first eigenvalue \(\lambda _1(\Omega )\) of \(-\Delta _q\), see Proposition 2.7, and Hölder’s inequality imply that

$$\begin{aligned} \int _{{\Omega }} m(x)f_2(u_n)u_n\,\text{d}x \ge -\lambda _0\int _{{\Omega }}m(x)|u_n|^{p}\text{d}x \ge - \frac{\lambda _0}{\lambda _1} \int _{{\Omega }} |\nabla u_n|^{p}\,\text{d}x, \end{aligned}$$
(3.6)

where we note that \(m\ge 0\). Concerning \(f_1\), by (1.7) and (1.8), we have

$$\begin{aligned} f_1(u)u&\ge \frac{a^*}{2} |u|^2, \ \ \text{ with } \ |u|<\delta _0 \ \text{ for } \text{ some }\ \delta _0>0,\\ f_1(u)u&\ge -\nu |u|^{q}. \end{aligned}$$

These estimates can be summarized to the pointwise lower bound

$$\begin{aligned} f_1(u_n) u_n(x) \ge \frac{a^*}{2} |u_n|^2 - \frac{a^*}{2\delta _0^{q-2}} |u_n|^q - \nu |u_n|^q, \quad x\in \Omega , \end{aligned}$$
(3.7)

Since the second last term \(\frac{a^*}{2\delta _0^{q-2}} \Vert u_n\Vert _{L^q}^q\) will finally appear as a problematic term on the right-hand side of the a priori estimate and has to be absorbed, we also derive a more localized version of (3.7). To this aim, we fix a cut-off function \(a\in C^\infty _0({\mathbb {R}}^N)\) such that \(0\le a\le a^*\) and \(\Vert a\Vert _{L^\infty } = a^*\) and obtain that

$$\begin{aligned} f_1(u_n)u_n(x) \ge \frac{1}{2} a(x)|u_n|^2 - \frac{1}{2\delta _0^{q_- -2}} a(x)|u_n|^{q_-} - \nu |u_n|^q, \quad x\in \Omega . \end{aligned}$$
(3.8)

Integrating (3.7) and (3.8) over \(\Omega\) we get the estimates

$$\begin{aligned} \int _{{\Omega }} f_1(u_n)u_n\,\text{d}x&\ge \frac{a^*}{2} \int _{{\Omega }} |u_n|^2\,\text{d}x - \frac{a^*}{2\delta _0^{q-2}} \int _{{\Omega }} |u_n|^{q}\,\text{d}x - \nu \int _{{\Omega }} |u_n|^q\,\text{d}x, \end{aligned}$$
(3.9)
$$\begin{aligned} \int _{{\Omega }} f_1(u_n)u_n\,\text{d}x&\ge \frac{1}{2} \int _{{\Omega }} a(x)|u_n|^2\,\text{d}x - \frac{1}{2\delta _0^{q_- -2}} \int _{{\Omega }} a(x)|u_n|^{q_-}\,\text{d}x - \nu \int _{{\Omega }} |u_n|^q\,\text{d}x, \end{aligned}$$
(3.10)

respectively. For the second term on the right-hand side of (3.10), we use the compactness of the support of a and deduce from Young’s inequality that

$$\begin{aligned} \int _{\Omega } a(x)|u_n|^{q_-}\,\text{d}x \le \frac{\mu -\nu }{4} \int _{\Omega } |u_n|^{q}\,\text{d}x + C(a,q,q_-). \end{aligned}$$
(3.11)

Moreover, it is easy to see that

$$\begin{aligned} \int _{\Omega }g(x)u_n\,\text{d}x \le \frac{a^*}{4}\Vert u_n\Vert _{L^{2}}^2 + C(a^*)\Vert g\Vert _{L^{2}}^2. \end{aligned}$$
(3.12)

On the one hand, by combining (3.4), (3.6), (3.9), and (3.12), we infer that

$$\begin{aligned} \frac{\,\text{ d }}{\,\text{d}t}\Vert u_n\Vert _{L^{2}}^2 + \widetilde{\lambda }\Vert \nabla u_n\Vert _{L^{p}}^p+\widetilde{\mu }\Vert u_n\Vert _{L^{q}}^q + a^*\Vert u_n\Vert _{L^{2}}^2&\le \frac{a^*}{\delta _0^{q-2}} \Vert u_n\Vert _{L^{p}}^p + C, \end{aligned}$$
(3.13)

where \(\widetilde{\lambda } = \frac{2}{\lambda _1} (\lambda _1-\lambda _0)\) and \(\widetilde{\mu }:=\mu -\nu >0.\) On the other hand, by (3.4), (3.6), (3.10), (3.11), (3.12), and even omitting the integral \(\int _{{\Omega }} |a(x)||u_n|^2\,\text{d}x\), we get the modified estimate

$$\begin{aligned} \frac{\,\text{d}}{\,\text{d}t}\Vert u_n\Vert _{L^{2}}^2 + \widetilde{\lambda }\Vert \nabla u_n\Vert _{L^{p}}^p+\widetilde{\mu }\Vert u_n\Vert _{L^{q}}^q \le C. \end{aligned}$$
(3.14)

Next we choose \(\mu _0>0\) small enough such that \(\frac{\mu _0a^*}{\delta _0^{q-2}}\le \frac{\widetilde{\mu }}{2}\), multiply (3.13) by \(\mu _0\) and add (3.14) to see that

$$\begin{aligned} \frac{\,\text{d}}{\,\text{d}t} \Vert u_n\Vert _{L^{2}}^2 + \widetilde{\lambda } \Vert \nabla u_n\Vert _{L^{p}}^p + \frac{\widetilde{\mu }(1+\mu _0/2)}{1+\mu _0} \Vert u_n\Vert _{L^{q}}^q + \frac{a^*\mu _0 }{\mu _0+1} \Vert u_n\Vert _{L^{2}}^2&\le C. \end{aligned}$$
(3.15)

Then by Gronwall’s inequality, we obtain

$$\begin{aligned} \Vert u_n\Vert _{L^{\infty }(0, t, L^2)}^2 \le C + \Vert u_n(0)\Vert _{L^{2}}^2 \exp \Big (-\frac{a^*\mu _0 t }{\mu _0+1} \Big ), \end{aligned}$$
(3.16)

where \(\Vert u_n(0)\Vert _{L^{2}}\le \Vert u(0)\Vert _{L^{2}}\). Therefore, we can extend the approximate solution to the interval [0, T],  for every \(T > 0.\) On the other hand, by (3.15), we also have

$$\begin{aligned} \Vert u_n(t)\Vert _{L^{2}}^2 + \widetilde{\lambda } \int _0^t \Vert \nabla u_n\Vert _{L^{p}}^p\,\text{d}\tau + \frac{\widetilde{\mu }(1+\mu _0/2)}{1+\mu _0} \int _0^t \Vert u_n\Vert _{L^{q}}^q\,\text{d}\tau \le \Vert u_n(0)\Vert _{L^{2}}^2+Ct, \end{aligned}$$
(3.17)

for any \(t\in [0,T]\). We note that the generic constant C so far depends only on the constants in the assumptions \((H_m), (H_f^1),(H_f^2)\), but neither on the solution, its initial value, the time interval nor the domain \(\Omega\).

Now, multiplying Eq. (3.2)\(_k\) by the function \(u^\prime _{nk}(t)\), adding these relations for \(k= 1,\ldots ,n\), integrating over (sT) and taking into account (3.3), we have

$$\begin{aligned}&\int _s^T \Vert \partial _t u_{n}(t)\Vert _{L^{2}}^2\,\text{d}t + \frac{1}{p}\Vert \nabla u_n(T)\Vert _{L^{p}}^p + \frac{\mu }{q} \Vert u_n(T)\Vert _{L^{q}}^q \\&\qquad + \int _{\Omega } F(x,u_{n}(T,x))\,\text{d}x -\int _{\Omega } g(x)u_{n}(T,x)\,\text{d}x \\&\quad \le \frac{1}{p}\Vert \nabla u_n(s)\Vert _{L^{p}}^p + \frac{\mu }{q} \Vert u_n(s)\Vert _{L^{q}}^q + \int _{\Omega } F(x,u_{n}(s,x))\,\text{d}x - \int _{\Omega } g(x)u_{n}(s,x)\,\text{d}x, \end{aligned}$$
(3.18)

where \(F(x,u)=\int _0^uf(x,\xi )\,\text{d}\xi\). By conditions (1.8) and (1.9), we see that

$$\begin{aligned} |F(x,u)|&\le \int _0^u \big (|f_1(\xi )|+|m(x)|f_2(\xi )|\big )\,\text{d}\xi \le C\big (|u|^2+|u|^{q}+|u|^{p}\big ). \end{aligned}$$
(3.19)

By using the Sobolev embedding \(\mathcal {D}_0^{1,p}\hookrightarrow L^{r}\) with \(\min (2,p)\le r \le p^*\) and then integrating (3.18) over (0, T) with respect to the variable s, we have from (3.6)–(3.19)

$$\begin{aligned} \Vert t^{1/2} \partial _t u_{n}\Vert _{L^{2}(0, T; L^{2})}\le C(T)=C(T,a,g,C_f,||u_0||_{L^{2}}). \end{aligned}$$
(3.20)

Note that so far all constants are independent of the bounded domain \(\Omega\).

In the second step of the proof we show that the sequence \((u_n)\) converges to a weak solution of (1.1). The boundedness of A, (3.17) and (3.20) imply the existence of functions \(u\in L^{\infty }(0,T;L^{2})\) and \(\chi \in \ L^{p^\prime }(0,T; \mathcal {D}^{-1,p^\prime })\) such that—ignoring the notion of subsequences

$$\begin{aligned} \left\{ \begin{array}{cll} \displaystyle \; u_n\rightarrow u &{} \text{ weakly-* } \text{ in } &{} L^{\infty }(0,T;L^{2}),\\ \partial _t u_{n}\rightarrow u_t\;\; &{} \text{ weakly } \text{ in } &{} L^{2}(\epsilon ,T;L^{2})\ \ \text{ for } \, \epsilon \in (0,T),\\ \;u_n\rightarrow u &{} \text{ weakly } \text{ in } &{} L^{p}(0,T;\mathcal {D}_0^{1,p}) \text{ and } \text{ in } L^q(0,T;L^q)\\ \! Au_n\rightarrow \chi &{} \text{ weakly } \text{ in } &{} L^{p^\prime }(0,T; \mathcal {D}^{-1,p^\prime }), \end{array} \right. \end{aligned}$$
(3.21)

as \(n\rightarrow \infty\) where \(\chi\) will later shown to satisfy \(\chi =Au.\) From \((3.21)_{1,2}\) we deduce that

$$\begin{aligned} u \in {\mathcal {C}}((0, T]; L^{2}). \end{aligned}$$
(3.22)

Since \(\Omega\) is bounded, an Aubin-type compact embedding theorem ( [32, Corollary 4]), yields the precompactness of the set \(\{u_n: n\in {\mathbb {N}}\}\) in \(L^{p}(\epsilon , T; L^{r})\) for every \(\epsilon \in (0,T)\) and \(\min (2,p)\le r<p^*\). Hence, given any sequence \(\{\epsilon _k\}_k\) with \(\epsilon _k \searrow 0\), there exist subsequences \(\{u_{n_m}^{(k)}\}_{m=1}^\infty \subset \{u_{n_m}^{(k-1)}\}_{m=1}^\infty \subset \cdots \subset \{u_{n_m}\}_{m=1}^\infty\) such that for each \(k\in {\mathbb {N}}\)

$$\begin{aligned} u_{n_m}^{(k)}&\rightarrow u\; \text{ in } \;L^{p}(\epsilon _k, T; L^{r}),\\ u_{n_m}^{(k)}(t)&\rightarrow u(t)\;\; t\text{-a.e. } \text{ in } \;\;L^{r},\\ u_{n_m}^{(k)}&\rightarrow u\; \text{ a.e. } \text{ on } \; (\epsilon _k, T)\times \Omega \; \text{ as } \; m\rightarrow \infty . \end{aligned}$$

Now, by Cantor’s diagonal argument, we obtain (up to a subsequence \(\{u_{n_k}^{(k)}\}\)) for \(\min (2,p)\le r<p^*\)

$$\begin{aligned} u_n(t)\rightarrow u(t)\;\; t \text {-a.e. in }\; L^{r} \;\text { and }\; u_n\rightarrow u\; \text{ a.e. } \text{ in } \ \Omega _T=\Omega \times [0,T]. \end{aligned}$$
(3.23)

Hence \(f(x,u_n)\rightarrow f(x,u)\) a.e. in \(\Omega _T\). Moreover, the boundedness of \(\{u_n\}\) in \( L^{p}(0,T; L^{p^{*}})\cap L^{q}(0,T; L^{q})\) and assumption (1.8) and (1.9) imply that \(\{f(x,u_n)\}\) is bounded in \(L^{\rho _1}(0,T; L^{\rho _2})\), where

$$\begin{aligned} \rho _1=\min \Big \{\frac{p}{p-1},\frac{q}{q-1}\Big \}, \quad \rho _2 = \min \Big \{2,\frac{p^*}{p-1},\frac{q}{q-1}\Big \}, \end{aligned}$$

Consequently, \(f(x,u_n)\rightarrow \psi\) weakly in \(L^{\rho _1}(0,T; L^{\rho _2})\). Thus \(\psi =f(x,u)\) thanks to [21, Lemma 1.3]. By the same process, we see that \(|u_n|^{q-2}u_n\rightarrow |u|^{q-2}u\) weakly in \(L^{q^\prime }(\Omega _T)\). Therefore, we have

$$\begin{aligned} u_t=-\chi -\mu |u|^{q-2}u-f(x,u)+g \ \ \text{ in }\ V^*+L^{\rho _1}(0,T; L^{\rho _2}), \end{aligned}$$
(3.24)

Note that \(V^* + L^{\rho _1}(0,T; L^{\rho _2}) \subseteq L^{1}\big (0,T; \mathcal {D}^{-1,p^\prime } + L^{q^{\prime }}+L^{\rho _2}\big )\), so that we get \(u\in \mathcal {C}([0,T];\mathcal {D}^{-1,p^\prime }+L^{q^{\prime }}+L^{\rho _2})\). Because of \(u\in L^{\infty }(0,T; L^{2})\) and \(L^2 \hookrightarrow L^{q'}\), [20, Lemma 8.1, p. 275] implies that

$$\begin{aligned} u\in \mathcal {C}_{w}([0,T];L^{2}), \end{aligned}$$
(3.25)

i.e., u is continuous in time with respect to the weak topology of \(L^2\). From (3.21) and (3.24), expressing \(u_n(0)\) and u(0) as integrals with respect to time of \(u_n,\partial _tu_n\) and \(u,\partial _tu\), respectively, we have

$$\begin{aligned} (u_{n}(0), e_k)\rightarrow (u(0), e_k),\; k=1,\ldots ,n, \end{aligned}$$

and then, by (3.3), \(u(0)=u_0\). Hence, taking into account (3.21) and passing to the limit in (3.17) when \(n\rightarrow \infty\), we get at first that \(\Vert u(t)\Vert _{L^{2}}^2\le \Vert u_0\Vert _{L^{2}}^2 +Ct\) and hence

$$\begin{aligned} \limsup _{t\rightarrow 0}\Vert u(t)\Vert _{L^{2}}^2\le \Vert u_0\Vert _{L^{2}}^2. \end{aligned}$$

Moreover, by (3.25), \(\Vert u_0\Vert _{L^{2}} \le \liminf _{t\rightarrow 0} \Vert u(t)\Vert _{L^{2}}\) so that with the above inequality u(t) is continuous at \(t=0\) in \(L^2\). By (3.22) we conclude that

$$\begin{aligned} u\in \mathcal {C}([0,T];L^{2}). \end{aligned}$$
(3.26)

Next, we prove \(\chi =Au\). It suffices to prove that \(\limsup _{n\rightarrow \infty }\int _0^T\langle Au_n, u_n\rangle \,\text{d}t\le \int _0^T \langle \chi ,u\rangle \,\text{d}t\) (see [31, Lemma 2.1, p. 38]. From (3.4) we have

$$\begin{aligned} \int _0^T \langle Au_n, u_n\rangle \,\text{d}t&= \int _0^T \int _{\Omega } \left( g-f(x,u_n) -\mu |u_n|^{q-2} u_n\right) u_n\,\text{d}x\,\text{d}t \\&\quad + \Vert u_n(0)\Vert _{L^{2}}^2-\Vert u_n(T)\Vert _{L^{2}}^2. \end{aligned}$$
(3.27)

Note that due to (3.3) \(u_{n}(0)\rightarrow u_0\) in \(L^{2}\), whereas, by weak convergence properties,

$$\begin{aligned} \begin{aligned} \Vert u\Vert _{L^q(0,T;L^{q})}&\le \liminf _{n\rightarrow \infty } \Vert u_n\Vert _{L^q(0,T;L^{q})},\\ \int _0^T \int _{\Omega } gu\,\text{d}x\,\text{d}t&= \lim _{n\rightarrow \infty } \int _0^T \int _{\Omega } g u_n\,\text{d}x\,\text{d}t, \\ \int _0^T \int _{\Omega } f(x,u)u\,\text{d}x\,\text{d}t&\le \liminf _{n\rightarrow \infty } \int _0^T \int _{\Omega } f(x,u_n)u_n\,\text{d}x\,\text{d}t. \end{aligned} \end{aligned}$$
(3.28)

For the proof of \((3.28)_3\) we use \((H_f^1)\), \((H_f^2)\) to get on the one hand that \(f_1(x,v)v + \nu _-|v|^{q_-}\ge 0\) and \(f_2(v)v + \lambda |v|^p\ge 0\) for all \(v\in {\mathbb {R}}\). On the other hand, by \((H_f^1)\), \((H_f^2)\), the non-negative sequence \(\{f_1(x,u_n)u_n + \nu _-|u_n|^{q_-} + m(x)(f_2(u_n)u_n + \lambda |u_n|^p)\}_n\) is uniformly bounded in \(L^1((0,T)\times \Omega )\). Hence Fatou’s Lemma and the pointwise convergence properties in (3.23) imply that

$$\begin{aligned}&\int _0^T \int _\Omega \big \{f_1(x,u)u + \nu _-|u|^{q_-} + m(x)f_2(x,u)u + \lambda m(x)|u|^p\big \}\,\text{d}x\,\text{d}t \\&\qquad \le \liminf _{n\rightarrow \infty } \int _0^T \int _\Omega \big \{ f_1(x,u_n)u_n + \nu _-|u_n|^{q_-} + m(x)f_2(u_n)u_n + \lambda m(x)|u_n|^p\big \}\,\text{d}x\,\text{d}t. \end{aligned}$$
(3.29)

Furthermore, the sequences \(\{|u_n|^{q_-}\}_n\) and \(\{m(x)|u_n|^{p}\}_n\) are converging pointwise a.e. to \(|u|^{q_-}\) and \(m(x)|u|^p\), respectively, and are uniformly bounded in the space \(L^{r}((0,T)\times \Omega )\) for some \(1<r<\infty\), since \(q_-<q\) and \(p<p^*\). This, by the weak convergence result of [21, Lemma 1.3] and using the admissible test function \(1\in L^{r'}\),

$$\begin{aligned} \Vert u_n\Vert _{L^{q_-}}^{q_-} = \int _0^T\int _\Omega |u_n|^{q_-}\cdot 1\,\text{d}x\,\text{d}x \;\rightarrow \; \int _0^T\int _\Omega |u|^{q_-}\cdot 1\,\text{d}x\,\text{d}t = \Vert u\Vert _{L^{q_-}}^{q_-} \end{aligned}$$

as \(n\rightarrow \infty\). Similarly, we obtain that the space-time integral of \(\{m(x)|u_n|^{p}\}\) converges to the corresponding integral of \(m(x)|u|^p\). Therefore, we are allowed to remove the artificial terms \(\nu _-|u_n|^{q_-}\), \(\lambda m(x)|u_n|^p\), and their limits, in (3.29) to get \((3.28)_3\). Finally, writing \(u_n(T)-u_n(t)\) and \(u(T)-u(t)\) as integrals on [tT] using time derivatives, \((3.21)_2\) and (3.23) imply that

$$\begin{aligned} \Vert u(T)\Vert _{L^{2}} \le \lim _{n\rightarrow \infty } \Vert u_n(T)\Vert _{L^{2}}. \end{aligned}$$
(3.30)

In view of (3.24), we have

$$\begin{aligned} \int _0^T \int _{\Omega } \left( g-f(x,u) - \mu |u|^{q-2}\right) u\,\text{d}x\,\text{d}t + \Vert u(0)\Vert _{L^{2}}^2 - \Vert u(T)\Vert _{L^{2}}^2 =\int _0^T \langle \chi ,u\rangle \,\text{d}t. \end{aligned}$$
(3.31)

Hence (3.27)–(3.31) imply that

$$\begin{aligned} \limsup _{n\rightarrow \infty }\int _0^T\langle Au_n, u_n\rangle \,\text{d}t\le \int _0^T\langle \chi ,u\rangle \,\text{d}t. \end{aligned}$$
(3.32)

Finally, we prove the uniqueness of the solution. Let u and v be the weak solutions of (1.1), with initial data \(u_0\) and \(v_0\), respectively. Denoting \(\omega =u-v\), we have

$$\begin{aligned} \begin{aligned} \omega _{t}&= \left( \text{div }(|\nabla u|^{p-2}\nabla u)-\text{ div }(|\nabla v|^{p-2}\nabla v)\right) \\&\quad -\mu \left( |u|^{q-2}u-|v|^{q-2}v\right) - (f(x,u)-f(x,v)),\\ \omega (0)&= u_0-v_0. \end{aligned}\end{aligned}$$
(3.33)

Multiplying (3.33) by \(\omega\), using (1.8) and the strong monotonicity of the p-Laplacian, we get

$$\begin{aligned} \frac{\,\text{d}}{\,\text{d}t}\Vert \omega \Vert _{L^{2}}^2 + c_p\Vert \nabla \omega \Vert _{L^{\max (p,2)}}^{\max (p,2)} + c_q'\mu \Vert \omega \Vert _{L^{q}}^q \le \nu \Vert \omega \Vert _{L^{q}}^q + \lambda \int _{\Omega } m(x)|\omega |^{p}\,\text{d}x, \end{aligned}$$

so that the assumptions \((H_c)\) on \(\nu ,\lambda\) imply that

$$\begin{aligned} \frac{\,\text{d}}{\,\text{d}t}\Vert \omega \Vert _{L^{2}}^2 \le 0. \end{aligned}$$
(3.34)

Now it is obvious to obtain \(\omega =0\) if \(u_0=v_0\). Therefore, the solution is unique.

In the final part of the proof let \(\Omega \subseteq {\mathbb {R}}^N\) be an unbounded domain. We will use the method of invading domains to get the desired result following the idea of Babin-Vishik (see [7] for details). For the convenience of the reader, we give the details below. Let \(({\tilde{u}}_{R}),\, R\rightarrow +\infty ,\) be a sequence of solutions of problem (1.1) on the bounded open sets \({\Omega }_R=\{x\in \Omega : |x|<R\}\) with initial data \(u_{0,R}=\psi _R(|x|)u_0(x)\) and boundary value \(\tilde{u}_R=0\) on \(\partial \Omega _R\), where \(\psi _R\in C^\infty ({\mathbb {R}};[0,1])\) is a cut-off function satisfying \(\psi _R(\xi )=1\) for \(|\xi |\le R-1\) and \(\psi _R(\xi )=0\) for \(|\xi |\ge R\); moreover, let \(\Vert \psi _R^{(\ell )}\Vert _ {L^\infty } \le c_\ell <\infty\) for \(\ell \in {\mathbb {N}}\) and all \(R\ge 2\). Note that

$$\begin{aligned} \Vert u_0-u_{0,R}\Vert _{L^{2}}^2 = \int _{\Omega } (1-\psi _R^2(|x|)) |u_0|^2\,\text{d}x \rightarrow 0\;\; \text{ as }\; R\rightarrow \infty . \end{aligned}$$
(3.35)

Since \(\Vert u_{0,R}\Vert _{L^2(\Omega _R)} \le \Vert u_0\Vert _{L^2(\Omega )}\), the estimates (3.4)–(3.20) hold uniformly with respect to R for \({\tilde{u}}_R\). To extend the function \({\tilde{u}}_R\) from \(\Omega _R\) to \({\mathbb {R}}^N\) we multiply \({\tilde{u}}_R\) by \(\psi _R(|x|)\) and extend the product by zero for \(|x|\ge R\), to get the function by \(u_R\). To show that \(u_R\) is a solution of an equation of type (1.1) on \(\Omega\) we note that by Poincaré’s inequality

$$\begin{aligned} \Vert \nabla u_R\Vert _{L^p(\Omega )} \le \Vert \nabla {\tilde{u}}_R\Vert _{L^p(\Omega _R)} + c_1\Vert {\tilde{u}}_R\Vert _{L^p(\Omega _R\setminus \Omega _{R-1})} \le c\Vert \nabla {\tilde{u}}_R\Vert _{L^p(\Omega _R)} \end{aligned}$$

due to the special structure of the set \(\Omega _R\setminus \Omega _{R-1}\) of “thickness” 1. From (3.4) to (3.20) it follows that there exists a subsequence \(\{u_j\}_j=\{u_{R_j}\}\) of \(\{u_R\}_R\), which is weakly convergent in \(L^p([0,T]; \mathcal {D}_0^{1,p}({\Omega }))\) and weakly\(-*\) convergent in \(L^{\infty }(0,T; L^{2}({\Omega }))\). Denote the limit of \(\{u_j\}\) by \(u_\infty =u_{\infty }(t,x)\).

For fixed \(R=R_k\), let \(u_{kj}:=\psi _k(x)u_j\) for \(j>k\). Via arguments as above for \(\{u_R\}\), the sequence \(\{\psi _ku_j\}_j\), for fixed \(k\in {\mathbb {N}}\), satisfies the estimates (3.4)–(3.20) uniformly with respect to j (\(j>k)\) and with norms evaluated on the domain \({\Omega }_{R_k}\). Exploiting Cantor’s method there exists a subsequence \(\{u_{j'}\}_{j'}\) of \(\{u_{j}\}_{j}\) such that for each \(k\in {\mathbb {N}}\) the sequence \(\{\psi _k u_{j'}\}_{j'}\) is weakly convergent in the above mentioned spaces on \(\Omega _k\). Obviously, due to pointwise- a.e. convergence, see (3.23), the limit \(u_{k\infty }\) satisfies the identity

$$\begin{aligned} u_{k\infty } = \psi _ku_\infty ,\quad k\in {\mathbb {N}}.\end{aligned}$$

As in the case of a bounded domain, we prove that \(\psi _ku_\infty\) is a weak solution of (1.1) in \([0,T]\times \Omega _{R_k}\) so that for any \(v\in \mathcal {C}_0^\infty ([0,T]\times {\Omega }_{ R_k-1})\)

$$\begin{aligned} \int _{0}^T&\int _{\Omega } |\nabla u_{\infty }|^{p-2}\nabla u_{\infty }\nabla v(t,x))\,\text{d}x\,\text{d}t \\&= - \int _{0}^T \int _{\Omega } \left( \frac{\partial u_\infty }{\partial t} + \mu |u_\infty |^{q-2}u_\infty + f(x,u_\infty )-g\right) v(t,x)\,\text{d}x\,\text{d}t. \end{aligned}$$
(3.36)

Since \(R=R_k\) is arbitrary, we deduce that (3.36) is fulfilled for arbitrary \(v\in \mathcal {C}_0^\infty ([0,T]\times \Omega )\), and \(u_\infty\) is a solution of (1.1). By (3.35), the solution satisfies (1.1)\(_2\). In the estimates (3.4)–(3.20) we pass to the limit as \(R\rightarrow \infty\) and we find that these estimates hold for \(u=u_\infty\).

The uniqueness of the solution \(u_\infty\) is proved as in the bounded domain case. Now the proof of this theorem is finished. \(\square\)

Thus, by Theorem 3.2, the solution operator \(S(t)u_0 = u(t)\) of problem (1.1) generates a continuous semigroup in \(L^{2}\).

4 Existence of the global attractor

We begin this section with the existence of an absorbing set for the semigroup \({S(t)}_{t\ge 0}\).

Proposition 4.1

Under the assumption of Theorem 3.2, the semigroup \({S(t)}_{t\ge 0}\) associated with problem (1.1) admits an absorbing set in \(\mathcal {D}_0^{1,p}\cap L^{2}\cap L^{q}\); i.e., there is a bounded set \(B_0\subset \mathcal {D}_0^{1,p}\cap L^{2}\cap L^{q}\) such that, for any bounded set B in \(L^{2}\), there exists a \(T_1>0\), depending only on B, such that

$$\begin{aligned} S(t)B\subset B_0\ \text{ for } \text{ any }\ t\ge T_1.\end{aligned}$$

Proof

Firstly, we multiply (1.1) by u and get that

$$\begin{aligned} \frac{1}{2}\frac{\,\text{ d }}{\,\text{d}t}\Vert u\Vert _{L^{2}}^2+\Vert \nabla u\Vert _{L^{p}}^p+\mu \Vert u\Vert _{L^{q}}^q=-\int _{\Omega }f(x,u)u\,\text{d}x+\int _{\Omega }g(x)u\,\text{d}x. \end{aligned}$$
(4.1)

Considering the a priori estimate (3.16) on any interval [0, t) and passing to the limit \(n\rightarrow \infty\), we get an absorbing set in \(L^{2}({\mathbb {R}}^2)\); namely, for any bounded set \(B\subset L^2\) there exists a \(t_1=t_1(B)>0\) such that

$$\begin{aligned} \Vert u(t)\Vert _{L^{2}}^2\le C\ \ \text{ for } \text{ any }\ t\ge t_1 \ \text{ and } u_0\in B \end{aligned}$$
(4.2)

where \(C>0\) is a constant independent of the initial data \(u_0\in B\). Moreover, by (3.14), we see that

$$\begin{aligned} \int _{0}^{t}\left (\Vert \nabla u\Vert _{L^{p}}^p + \Vert u\Vert _{L^{q}}^q\right)\,\text{d}\tau \le \Vert u_0\Vert _{L^2}^2+Ct. \end{aligned}$$
(4.3)

Let

$$\begin{aligned} E(u(t)):=\frac{1}{p}\int _\Omega |\nabla u|^p\,\text{d}x+\frac{\mu }{q}\int _\Omega |u|^q\,\text{d}x+\int _{\Omega }F(x,u)\,\text{d}x- \int _{\Omega }g(x)u\,\text{d}x, \end{aligned}$$

where \(F(x,u)=\int _0^u f(x,\xi )\,\text{d}\xi\). By condition (1.8)–(1.10), (3.19) and (4.3), we see that

$$\begin{aligned}&\left| \int _{\Omega } F(x,u)\,\text{d}x\right| + \left| \int _{\Omega } g(x)u\,\text{d}x\right| \\&\quad \le \int _{\Omega } \left( |u|^q+m(x)|u|^p + (1+m(x))|u|^2\right) \text{d}x + \Vert g\Vert _{L^2}\Vert u\Vert _{L^2} \\&\quad \le C \left( \Vert u(t)\Vert _{L^q}^{q}+\Vert m\Vert _{L^{\frac{N}{p}}}\Vert u(t)\Vert _{L^{p^*}}^{p} + \Vert u(t)\Vert _{L^2} + \Vert u(t)\Vert _{L^2}^{2}\right) \ \ \text{ for } \text{ any }\ t\ge t_1. \end{aligned}$$
(4.4)

Multiplying (1.1) by \(u_t\) we find

$$\begin{aligned} \frac{\,\text{ d }}{\,\text{d}t}E(u(t)) = - \int _\Omega |u_t|^2\,\text{d}x\le 0, \end{aligned}$$
(4.5)

so that for any \(t\ge s\ge 0\)

$$\begin{aligned} E(u(t)) + \int _s^t\int _\Omega |u_t|^2\,\text{d}x\,\text{d}\tau \le E(u(s)). \end{aligned}$$

A further integration of the previous inequality over (0, t) with respect to the variable s, and (4.2)–(4.4) imply that

$$\begin{aligned} t E(u(t))&\le C\int _0^t\big (\Vert \nabla u\Vert _{L^{p}}^p + \mu \Vert u\Vert _{L^{q}}^q\big )\,\text{d}s + C\int _0^t \Vert u\Vert _{L^{p^*}}^{p}\,\text{d}s + Ct \\&\le C\Vert u_0\Vert _{L^{2}}^2+C t\ \ \text{ for } \text{ any }\ t\ge t_1, \end{aligned}$$
(4.6)

where the estimate \(\int _0^t \Vert u\Vert _{L^{p^*}}^{p} \,\text{d}s \le c \int _0^t\Vert \nabla u\Vert _{L^{p}}^{p}\) was used. To get a lower bound for the left-hand side of (4.6), the estimate \(\int _\Omega gu\,\text{d}x \ge -\Vert g\Vert _{L^2}\Vert u\Vert _{L^2} \ge -C\) is obvious. Concerning the estimate for the integral of F consider the term involving \(f_1\). By (1.7) and for \(u>0\)

$$\begin{aligned} \int _0^{u} f_1(u)\,\text{d}u = \int _0^{u} \frac{f_1(u)u}{u}\,\text{d}u \ge \int _0^{u} -\nu _-|u|^{q_- -1}\,\text{d}u = - \frac{\nu _-}{q_-} |u|^{q_-}.\end{aligned}$$

It is easy to see that the same estimate holds when \(u<0\). Since \(2<q_-<q\), by Hölder’s and Young’s inequality, the term \(\frac{\nu _-}{q_-}\Vert u\Vert _{L^{q_-}}^{q_-}\) can be estimated by \(\frac{\mu }{2q}\Vert u\Vert _{L^q}^q +C\). For the integral with \(f_2\) the assumption (1.7) implies for \(u>0\) that

$$\begin{aligned} \int _\Omega \int _0^{u} mf_2(u)\,\text{d}u\,\text{d}x&\ge \int _\Omega \int _0^{u} -\lambda m(x)|u|^{p-1}\,\text{d}u\,\text{d}x = -\frac{\lambda }{p} \int _\Omega m(x)|u|^p\,\text{d}x\\&\ge -\frac{\lambda }{\lambda _1 p} \Vert \nabla u\Vert _{L^p}^p. \end{aligned}$$

The same estimate holds for negative u. Summarizing the last two estimates we conclude that

$$\begin{aligned} E(u(t))&\ge \frac{1}{p}\Vert \nabla u\Vert _{L^p}^p + \frac{\mu }{q}\Vert u\Vert _{L^q}^q - \frac{\nu _-}{q_-} \Vert u\Vert _{L^{q_-}}^{q_-} -\frac{\lambda }{\lambda _1 p} \Vert \nabla u\Vert _{L^p}^p - C \\&\ge \frac{1}{p}\big (1-\frac{\lambda }{\lambda _1}\big ) \Vert \nabla u\Vert _{L^p}^p + \frac{\mu }{2q}\Vert u\Vert _{L^q}^q - C. \end{aligned}$$

where \(1-\frac{\lambda }{\lambda _1}>0\). Together with (4.6) we obtain the final pointwise estimate

$$\begin{aligned} \Vert \nabla u(t)\Vert _{L^{p}}^p + \Vert u(t)\Vert _{L^{q}}^q\le \frac{C}{t}\Vert u_0\Vert _{L^{2}}^2 +C\ \ \text{ for } \text{ any }\ t\ge t_1. \end{aligned}$$
(4.7)

In view of (4.7) there exists an absorbing set in \(\mathcal {D}_0^{1,p}\cap L^{q}\) for any \(t\ge T_1\) for some \(T_1\ge t_1\) large enough. This completes the proof. \(\square\)

Proposition 4.2

Under the assumption of Theorem 3.2, the semigroup \({S(t)}_{t\ge 0}\) associated with problem (1.1) admits an absorbing set in \( L^{2q-2}\) if \(\lambda >0\) in \((H_f^2)\) is sufficiently small; i.e., there is a bounded set \(B_0\subset L^{2q-2}\) such that, for any bounded set \(B \subset L^{2}\), there exists a \(T_2>0\) depending only on B such that

$$\begin{aligned} S(t)B\subset B_0\ \text{ for } \text{ any }\ t\ge T_2. \end{aligned}$$

Proof

Similarly to Proposition 4.1, we multiply (1.1) by \(|u|^{q-2}u\) to get

$$\begin{aligned}&\frac{1}{q} \frac{\,\text{d}}{\,\text{d}t}\Vert u\Vert _{L^{q}}^q + (q\!-\!1)\!\int _{\Omega } |\nabla u|^p|u|^{q-2}\,\text{d}x + \mu \int _{\Omega } |u|^{2q-2}\,\text{v}x \\&\quad + \int _{\Omega } f(x,u)|u|^{q-2}u\,\text{d}x = \int _{\Omega } g|u|^{q-2}u\,\text{d}x. \end{aligned}$$
(4.8)

We note that

$$\begin{aligned} (q-1)\int _{\Omega }|\nabla u|^p|u|^{q-2}\,\text{d}x=\frac{p^{p}(q-1)}{(p+q-2)^p}\int _{\Omega }\left| \nabla |u|^{\frac{p+q-2}{p}}\right| ^p\,\text{d}x \end{aligned}$$

and rewrite (4.8) in the form

$$\begin{aligned}&\frac{1}{q}\frac{\,\text{d}}{\,\text{d}t}\Vert u\Vert _{L^{q}}^q + \frac{p^{p}(q-1)}{(p+q-2)^p} \int _{\Omega } \left| \nabla |u|^{\frac{p+q-2}{p}}\right| ^p\,\text{d}x + \mu \int _{\Omega } |u|^{2q-2}\,\text{d}x \\&\quad = -\int _{\Omega } f(x,u)|u|^{q-2}u\,\text{d}x + \int _{\Omega } g|u|^{q-2}u\,\text{d}x. \end{aligned}$$
(4.9)

With (1.8), (1.9), and (3.5), (3.6), and by applying (2.2) to \(|u|^{\frac{q+p-2}{p}}\) we deduce that

$$\begin{aligned} \int _{\Omega } f(x,u)|u|^{q-2}u\,\text{d}x&= \int _{\Omega } \left( f_1(u)+m(x)f_2(u)\right) |u|^{q-2}u\,\text{d}x \\&\ge -\nu \int _{\Omega } |u|^{2q-2}\,\text{d}x - \lambda \int _{{\Omega }} m(x)|u|^{q+p-2}\,\text{d}x \\&\ge -\nu \int _{\Omega } |u|^{2q-2}\,\text{d}x - \frac{\lambda }{\lambda _1} \int _{\Omega } \left| \nabla |u|^{\frac{p+q-2}{p}}\right| ^p\,\text{d}x \end{aligned}$$
(4.10)

We choose \(\lambda\) small enough such that

$$\begin{aligned} \frac{\lambda }{\lambda _1} < \Big (\frac{p}{p+q-2}\Big )^p(q-1). \end{aligned}$$
(4.11)

Moreover, note that

$$\begin{aligned} \int _{\Omega } g(x)|u|^{q-2}u\,\text{d}x \le \Vert g\Vert _{L^2} \Vert u\Vert _{L^{2q-2}}^{q-1} \le \frac{\mu -\nu }{2} \Vert u\Vert _{L^{2q-2}}^{2q-2} + C\Vert g\Vert _{L^2}^2. \end{aligned}$$

Therefore, with \({\widetilde{\mu }}=\mu -\nu\), we obtain from (4.9), (4.10) and (4.11) that

$$\begin{aligned} \frac{\,\text{ d }}{\,\text{d}t}\Vert u\Vert _{L^{q}}^q + \widetilde{\mu } \Vert u\Vert _{L^{2q-2}}^{2q-2} \le C. \end{aligned}$$
(4.12)

Integrating (4.12) on \([t, t+1]\) with \(t\ge T_1\), Proposition 4.1 implies that

$$\begin{aligned} \int _t^{t+1}\int _{\Omega }|u|^{2q-2}\,\text{d}x\,\text{d}s\le C+\Vert u(t)\Vert _{L^{q}}\le C\ \ \ \text{ for } \text{ any }\ t\ge T_1. \end{aligned}$$
(4.13)

Now, we multiply (1.1) by \(|u|^{2(q-2)}u\), and by a procedure similar as for (4.8)–(4.10) we arrive at the estimate

$$\begin{aligned}&\frac{1}{2q-2} \frac{\,\text{ d }}{\,\text{d}t}\Vert u\Vert _{L^{2q-2}}^{2q-2} + (2q-3) \int _{\Omega } |\nabla u|^p|u|^{2(q-2)}\,\text{d}x + {\widetilde{\mu }} \int _{\Omega }|u|^{3q-4}\,\text{d}x \\&\quad \le C\int _{\Omega } g(x)|u|^{2q-3}\,\text{d}x + \lambda \int _\Omega m(x)|u|^{p+2(q-2)}\,\text{d}x. \end{aligned}$$
(4.14)

For the first integral on the right-hand side we use Hölder’s inequality to get that

$$\begin{aligned} C\int _{\Omega } g(x)|u|^{2q-3}\,\text{d}x \le C\Vert g\Vert _{L^{\frac{3q-4}{q-1}}} \Vert u\Vert _{L^{3q-4}}^{2q-3} \le \frac{{\widetilde{\mu }}}{2} \Vert u\Vert _{L^{3q-4}}^{3q-4} + C, \end{aligned}$$

where the term \(\frac{{\widetilde{\mu }}}{2} \Vert u\Vert _{L^{3q-4}}^{3q-4}\) can be absorbed. The second integral on the right-hand side is rewritten in the form \(\lambda \int _\Omega m(x)\big (|u|^{\frac{p+2(q-2)}{p}}\big )^p\,\text{d}x\) and estimated with the help of (2.2) and finally absorbed by the integral \((2q-3) \int _{\Omega } |\nabla u|^p|u|^{2(q-2)}\,\text{d}x\) in (4.14). For this final step we need the smallness assumption \( \frac{\lambda }{\lambda _1} < \big (\frac{p}{p+2(q-2)}\big )^p (2q-3).\) This gives, cf. (4.12),

$$\begin{aligned} \frac{\,\text{ d }}{\,\text{d}t} \Vert u\Vert _{L^{2q-2}}^{2q-2} + \widetilde{\mu } \int _{\Omega } |u|^{3q-4}\,\text{d}x \le C. \end{aligned}$$

Applying the Uniform Gronwall Lemma 2.6, (4.13) and (4.14), we finally obtain

$$\begin{aligned} \Vert u(t)\Vert _{L^{2q-2}}^{2q-2}\le C \quad \text{ for } \text{ any }\ t\ge T_1+1, \end{aligned}$$
(4.15)

and we let \(T_2=T_1+1\) to finish the proof of this proposition. \(\square\)

Remark 4.3

By using of (4.14) and (4.15), one also has

$$\begin{aligned} \int _t^{t+1}\int _{\Omega }|u|^{3q-4}\,\text{d}x\,\text{d}s\le C\ \ \ \text{ for } \text{ any }\ t\ge T_1+1. \end{aligned}$$
(4.16)

Multiplying (1.1) by \(|u|^{3(q-2)}u\), we have

$$\begin{aligned} \frac{1}{3q-4}\frac{\,\text{ d }}{\,\text{d}t}\Vert u\Vert _{L^{3q-4}}^{3q-4} + {\widetilde{\mu }} \int _{\Omega } |u|^{4q-6}\,\text{d}x&\le C\int _{\Omega } g(x)|u|^{3q-5}\,\text{d}x + C \\&\le C\Vert g\Vert _{L^{\frac{4q-6}{q-1}}} \Vert u\Vert _{L^{4q-6}}^{3q-5} \le \frac{{\widetilde{\mu }}}{2} \Vert u\Vert _{L^{4q-6}}^{4q-6} + C. \end{aligned}$$

A further argument in this step is the absorption of the integral \(\int _\Omega m(x) |u|^{p+3(q-2)}\,\text{d}x\) with the help of (2.2) which requires the smallness \(\frac{\lambda }{\lambda _1} < \Big (\frac{p}{p+3(q-2)}\Big )^p (3(q-2)+1).\) With the same arguments as before, we conclude that

$$\begin{aligned} \int _t^{t+1}\int _{\Omega }|u|^{4q-6}\,\text{d}x\,\text{d}s\le C\ \ \ \text{ for } \text{ any }\ t\ge T_1+1,\\ \Vert u\Vert _{L^{3q-4}}^{3q-4}\le C \ \ \ \text{ for } \text{ any }\ t\ge T_1+2. \end{aligned}$$

Finally, adopting the smallness assumption on \(\lambda\) step by step, we obtain for a finite number \(k\in {\mathbb {N}}\) the uniform bound

$$\begin{aligned} \Vert u\Vert _{L^{k(q-2)+2}}^{k(q-2)+2}\le C \ \ \ \text{ for } \text{ any }\ t\ge T_1+k-1 \end{aligned}$$
(4.17)

provided

$$\begin{aligned} \frac{\lambda }{\lambda _1} < \Big (\frac{p}{p+k(q-2)}\Big )^p (k(q-2)+1). \end{aligned}$$
(4.18)

Proposition 4.4

Under the assumptions of Theorem 3.2additionally suppose that \(\frac{2N}{N+2}<p<N\). Then for any \(\varepsilon >0\) and any bounded subset \(B\subset L^{2}\), there exist constants \(T_3=T_3(B,\varepsilon )>0\) and \(R=R(\varepsilon )>0\) such that

$$\begin{aligned} \int _{\{x\in \Omega :\, |x|\ge R\}} |u|^2\,\text{d}x + \int _{\{x\in \Omega :\, |x|\ge R\}}|u|^q\,\text{d}x \le C \varepsilon \quad \text{ for } \text{ any } \ t\ge {T}_3\ \text{ and } \ u_0\in B, \end{aligned}$$
(4.19)

where the constant C is independent of \(\varepsilon\) and B.

Proof

Fix any \(\vartheta \in \mathcal {C}^{\infty }({\mathbb {R}})\) such that \(0\le \vartheta \le 1\),

$$\begin{aligned} \vartheta (s)=0\ \text{ for }\ 0\le |s|\le 1,\ \ \vartheta (s)=1\ \text{ for }\ \ |s|\ge 2, \end{aligned}$$

and define \(\vartheta _{R}(x)=\vartheta (\frac{|x|^2}{R^2})\). Now, multiplying (1.1) by \(\vartheta _{R}^{\beta }|u|^{\alpha -1}u\) with \(\alpha =1\) and \(\alpha =q-1\), \(\beta >p\) and integrating in \(\Omega\), one has

$$\begin{aligned} \frac{1}{\alpha +1}&\frac{\,\text{ d }}{\,\text{d}t} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x + \mu \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x \\&= \big \langle \text{ div }(|\nabla u|^{p-2}\nabla u), \vartheta _{R}^{\beta }|u|^{\alpha -1}u\big \rangle - \int _{\Omega } \vartheta _{R}^{\beta }f(x,u)|u|^{\alpha -1}u\,\text{d}x + \int _{\Omega } \vartheta _{R}^{\beta }g|u|^{\alpha -1}u\,\text{d}x. \end{aligned}$$
(4.20)

Concerning \(f(x,u) = f_1(u)+mf_2(u)\) where \(0\le m=m(x)\) satisfies \((H_m)\), one gets, similar to the estimates of (3.6), (3.7), (3.8) and (3.9), (3.10),

$$\begin{aligned} \int _{{\Omega }}\vartheta _{R}^{\beta }m(x)f_2(u)u|u|^{\alpha -1}\,\text{d}x&\ge -\lambda \int _{{\Omega }}\vartheta _{R}^{\beta } m|u|^{p+\alpha -1}\,\text{d}x \\&\ge - C\left( \int _{{\Omega }}\vartheta _{R}^{\beta }m^{\frac{N}{p}}\,\text{d}x \right) ^{\frac{p}{N}} \left( \int _{{\Omega }}\vartheta _{R}^{\beta }|u|^{\frac{N(p+\alpha -1)}{N-p}} \,\text{d}x\right) ^{\frac{N-p}{N}}. \end{aligned}$$
(4.21)

Moreover, for the integral involving \(f_1\),

$$\begin{aligned}&\int _{\Omega }\vartheta _{R}^{\beta }f_1(u)u|u|^{\alpha -1}\,\text{d}x \ge \frac{a^* }{2}\int _{{\Omega }} \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x \\&\quad - \frac{a^* }{2\delta _0^{q-2}} \int _{{\Omega }} \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x -\nu \int _{{\Omega }} \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x, \end{aligned}$$
(4.22)

as well as with the cut-off function \(a\in C^\infty _0({\mathbb {R}}^N)\) satisfying \(0\le a\le a^*=\Vert a\Vert _{L^\infty }\),

$$\begin{aligned}\int _{\Omega }&\vartheta _{R}^{\beta } f_1(u)u|u|^{\alpha -1}\,\text{d}x \\&\ge \frac{1}{2}\int _{\Omega } \vartheta _{R}^{\beta } a|u|^{\alpha +1}\,\text{d}x -\frac{1}{2\delta _0^{q_- -2}} \int _{\Omega } \vartheta _{R}^{\beta } a |u|^{q_-+\alpha -1}\,\text{d}x - \nu \int _{{\Omega }} \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x \\&\ge \frac{1}{2}\int _{\Omega }\vartheta _{R}^{\beta } a |u|^{\alpha +1}\,\text{d}x -\left( \frac{\mu -\nu }{2}+\nu \right) \int _{{\Omega }}\vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x \\&\quad - C\int _{{\Omega }} \vartheta _{R}^{\beta }|a(x)|^{\frac{q+\alpha -1}{q-q_-}}\,\text{d}x. \end{aligned}$$
(4.23)

Finally, the first term on the right hand side of (4.20) is given by

$$\begin{aligned} \big\langle \text{div }&(|\nabla u|^{p-2}\nabla u),\, \vartheta _{R}^{\beta }|u|^{\alpha -1}u\big \rangle \\&= -\alpha \int _{\Omega }\vartheta _{R}^{\beta }|\nabla u|^{p}\,|u|^{\alpha -1}\,\text{d}x - \int _{\Omega }\frac{2\beta x}{R^2}\vartheta _R^{\prime }\vartheta _{R}^{\beta -1}|\nabla u|^{p-2}\nabla u\, |u|^{\alpha -1}u\,\text{d}x \\&= -\alpha \int _{\Omega } \vartheta _{R}^{\beta }|\nabla u|^{p}| u|^{\alpha -1}\,\text{d}x - \int _{A_R}\frac{2\beta x}{R^2}\vartheta _R^{\prime }\vartheta _{R}^{\beta -1}|\nabla u|^{p-2}\nabla u\, |u|^{\alpha -1}u\,\text{d}x, \end{aligned}$$
(4.24)

where \(A_{R}\) denotes the “annulus” \(A_{R}=\{x\in \Omega : R\le |x|\le 2R\}\).

Thus, by (4.21), (4.22) and (4.21), (4.23) respectively, we obtain two different estimates as follows:

$$\begin{aligned}&\frac{1}{\alpha +1} \frac{\,\text{ d }}{\,\text{d}t}\int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x + \alpha \int _{\Omega } \vartheta _{R}^{\beta } |\nabla u|^{p}|u|^{\alpha -1}\,\text{d}x+\frac{\widetilde{\mu }}{2} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x \\&\qquad + \frac{a^*}{2}\int _{\Omega } \vartheta _{R}^{\beta }|u_n|^{\alpha +1}\,\text{d}x \\&\quad \le \; \frac{a^*}{2\delta _0^{q-2}} \int _{\Omega } \vartheta _{R}^{\beta }|u_n|^{q+\alpha -1}\,\text{d}x + \frac{C\beta }{R} \int _{A_R} \vartheta _{R}^{\beta -1} |\nabla u|^{p-1} | u|^{\alpha }\,\text{d}x \\&\qquad + C\left( \int _{\Omega } \vartheta _{R}^{\beta }\, m^{\frac{N}{p}}\,\text{d}x \right) ^{\frac{p}{N}} \left( \int _{\Omega } \vartheta _{R}^{\beta } |u|^{\frac{N(p+\alpha -1)}{N-p}}\,\text{d}x \right) ^{\frac{N-p}{N}} + \int _{\Omega } \vartheta _{R}^{\beta }\, g|u|^{\alpha -1}u\,\text{d}x, \end{aligned}$$
(4.25)

as well as

$$\begin{aligned}&\frac{1}{\alpha +1} \frac{\,\text{ d }}{\,\text{d}t} \int _{\Omega } \vartheta _{R}^{\beta } |u|^{\alpha +1}\,\text{d}x + \alpha \int _{\Omega } \vartheta _{R}^{\beta } |\nabla u|^{p} |u|^{\alpha -1}\,\text{d}x + \frac{\widetilde{\mu }}{2} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x + \frac{1}{2} \int _{\Omega } \vartheta _{R}^{\beta } \,a |u|^{\alpha +1}\,\text{d}x \\&\quad \le \frac{C\beta }{R}\int _{A_R} \vartheta _{R}^{\beta -1}|\nabla u|^{p-1}|u|^{\alpha }\,\text{d}x + C\int _{\Omega } \vartheta _{R}^{\beta }\, a^{\frac{q+\alpha -1}{q-q_-}}\,\text{d}x \\&\qquad + C\left( \int _{\Omega } \vartheta _{R}^\beta \, m^{\frac{N}{p}}\,\text{d}x \right) ^{\frac{p}{N}} \left( \int _{{\Omega }} \vartheta _{R}^{\beta }|u|^{\frac{N(p+\alpha -1)}{N-p}}\,\text{d}x \right) ^{\frac{N-p}{N}} + \int _{\Omega } \vartheta _{R}^{\beta }\,g|u|^{\alpha -1}u\,\text{d}x. \end{aligned}$$
(4.26)

Since \(a(x)\in C_0^\infty ({\mathbb {R}}^N)\) and \(m(x)\in L^{\frac{N}{p}}\cap L^{\infty }\), for any \(\varepsilon >0\), there exists \(R_0\) sufficiently large such that for any \(R\ge R_0\), one has \(\int _{{\Omega }} \vartheta _{R}^{\beta } a^{\frac{q+\alpha -1}{q-q_-}}\,\text{d}x = 0\) and

$$\begin{aligned} \left( \int _{{\Omega }} \vartheta _{R}^{\beta }|m(x)|^{\frac{N}{p}}\,\text{d}x \right) ^{\frac{p}{N}} \le \varepsilon . \end{aligned}$$

We also note that

$$\begin{aligned}&\frac{C\beta }{R} \int _{A_R} \vartheta _{R}^{\beta -1}|\nabla u|^{p-1}| u|^{\alpha }\,\text{d}x \\&\le \frac{C\beta }{R} \left( \int _{A_R} \vartheta _{R}^{\beta }|\nabla u|^{p}| u|^{\alpha -1}\,\text{d}x\right) ^{\frac{p-1}{p}} \left( \int _{A_R} \vartheta _{R}^{\beta -p}| u|^{\alpha +p-1}\,\text{d}x\right) ^{\frac{1}{p}} \\&\le \frac{\alpha }{2} \int _{\Omega }\vartheta _{R}^{\beta } |\nabla u|^{p} |u|^{\alpha -1}\,\text{d}x + \frac{C}{R^p} \int _{A_R} \vartheta _{R}^{\beta -p}| u|^{\alpha +p-1}\,\text{d}x. \end{aligned}$$
(4.27)

Since \(p>\frac{2N}{N+2}\), we find \(k_0\in {\mathbb {N}}\) such that

$$\begin{aligned} \frac{N(p+\alpha -1)}{N-p}<k_0(q-2)+2\quad \text {for }\; \alpha =1\text { and } \alpha =q-1. \end{aligned}$$
(4.28)

Then by an interpolation inequality and uniform estimates of \(\Vert u(t)\Vert _{L^{2}}\) and \(\Vert u(t)\Vert _{L^{k_0(q-2)+2}}\) with respect to time t, see (4.7), (4.17), we obtain in view of Remark 4.3 with an adequate \(\theta \in (0,1)\) the estimate

$$\begin{aligned}&\left( \int _{\Omega } \vartheta _{R}^{\beta }|m(x)|^{\frac{N}{p}} \,\text{d}x \right) ^{\frac{p}{N}} \left( \int _{{\Omega }}\vartheta _{R}^{\beta }|u|^{\frac{N(p+\alpha -1)}{N-p}}\,\text{d}x \right) ^{\frac{N-p}{N}} \le \varepsilon \Vert u\Vert _{L^{\frac{N(p+\alpha -1)}{N-p}}}^{p+\alpha -1} \\&\le \varepsilon \Vert u\Vert _{L^{2}}^{\theta (p+\alpha -1)} \Vert u\Vert _{L^{k_0(q-2)+2}}^{(1-\theta )(p+\alpha -1)} \\&\le C\varepsilon \ \ \ \text{ for } \text{ any }\ t\ge T_1+k_0-1 \text{ and } R\ge R_0. \end{aligned}$$
(4.29)

Therefore, we deduce from (4.25)–(4.29) for \(t\ge T_1+k_0-1\) and \(R\ge R_0\) that

$$\begin{aligned}&\frac{1}{\alpha +1} \frac{\,\text{ d }}{\,\text{d}t}\int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x +\frac{\alpha }{2}\int _{\Omega } \vartheta _{R}^{\beta }|\nabla u|^{p}| u|^{\alpha -1}\,\text{d}x \\&\qquad + \frac{\widetilde{\mu }}{2} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x + \frac{a^*}{2} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x \\&\quad \le \frac{a^*}{2\delta _0^{q-2}} \int _{A_R} \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x \\&\qquad + \frac{C}{R^p}\int _{\Omega } \vartheta _{R}^{\beta -p}|u|^{\alpha +p-1}\,\text{d}x + \int _{\Omega } \vartheta _{R}^{\beta }\, g|u|^{\alpha -1}u\,\text{d}x + C\varepsilon ,\end{aligned}$$
(4.30)
$$\begin{aligned}&\frac{1}{\alpha +1} \frac{\,\text{ d }}{\,\text{d}t}\int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x +\frac{\alpha }{2}\int _{\Omega }\vartheta _{R}^{\beta }|\nabla u|^{p}| u|^{\alpha -1}\,\text{d}x \\&\qquad + \frac{\widetilde{\mu }}{2} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x + \frac{1}{2}\int _{\Omega }\vartheta _{R}^{\beta }\, a |u|^{\alpha +1}\,\text{d}x \\&\le \frac{C}{R^p}\int _{A_R} \vartheta _{R}^{\beta -p}| u|^{\alpha +p-1}\,\text{d}x \\&\qquad + \int _{\Omega } \vartheta _{R}^{\beta }\, g|u|^{\alpha -1}u\,\text{d}x + C\varepsilon . \end{aligned}$$
(4.31)

Similar to the proof of (3.13)–(3.15), we obtain, choosing \(\mu _0>0\) sufficiently small, that

$$\begin{aligned}&\frac{\mu _0+1}{\alpha +1} \frac{\,\text{ d }}{\,\text{d}t} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x + \frac{\alpha (\mu _0+1)}{2} \int _{\Omega } \vartheta _{R}^{\beta }|\nabla u|^{p}| u|^{\alpha -1}\,\text{d}x \\&\ \ \ +\frac{\widetilde{\mu }(2\mu _0+1)}{4} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q+\alpha -1}\,\text{d}x + \frac{\mu _0 a^*}{2} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{\alpha +1}\,\text{d}x \\&\le \frac{C}{R^p} \int _{A_R} \vartheta _{R}^{\beta -p}| u|^{\alpha +p-1}\,\text{d}x + C\int _{\Omega } \vartheta _{R}^{\beta }\, g|u|^{\alpha -1}u\,\text{d}x + C\varepsilon . \end{aligned}$$
(4.32)

Firstly, we take \(\alpha =1\), apply (4.17) in Remark 4.3 with \(k_0\) as in (4.28) so that \(k_0(q-2)+2> p\), and Hölder’s inequality to get with a suitable number \(\theta \in (0,1)\) that

$$\begin{aligned} \int _{A_R}\vartheta _{R}^{\beta -p}|u|^{p}\,\text{d}x \le \left\{ \begin{array}{l} \displaystyle \int _{\Omega }| u|^{p}\,\text{d}x = \Vert u\Vert _{L^{p}}^p \le \Vert u\Vert _{L^{2}}^{p\theta }\Vert u\Vert _{L^{k_0(q-2)+2}}^{(1-\theta )p} \ \ \text{ if } p\ge 2,\\ \displaystyle \int _{A_R} | u|^{p}\,\text{d}x \le C\Vert u\Vert _{L^{2}}^{p}R^{\frac{N(2-p)}{2}}\qquad \qquad \quad \; \text{ if } \frac{2N}{2+N}<p< 2. \end{array} \right. \end{aligned}$$
(4.33)

By using Proposition 4.1 and (4.17) and choosing \(R_0\) sufficiently large, it follows that

$$\begin{aligned} \frac{C}{R^p} \int _{A_R} \vartheta _{R}^{\beta -p}|u|^{p}\,\text{d}x \le C\varepsilon \ \ \ \text{ for } \text{ any }\ t\ge T_1+k_0-1 \text{ and } R\ge R_0. \end{aligned}$$

Moreover, we have for any \( t\ge T_1\) and \(R\ge R_0\)

$$\begin{aligned} C\int _{\Omega } \vartheta _{R}^{\beta }\, gu\,\text{d}x\le C \left( \int _{{\Omega }} \vartheta _{R}^{\beta }\, |g|^{2}\,\text{d}x \right) ^{\frac{1}{2}}\Vert u\Vert _{L^{2}} \le C\varepsilon . \end{aligned}$$

This implies that for any \(t\ge T_1+k_0-1 \text{ and } R\ge R_0\)

$$\begin{aligned}&\frac{\,\text{ d }}{\,\text{d}t} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{2}\,\text{d}x + \int _{\Omega }\vartheta _{R}^{\beta }|\nabla u|^{p}\,\text{d}x + \frac{{\widetilde{\mu }}(2\mu _0+1)}{2(\mu _0+1)} \int _{\Omega }\vartheta _{R}^{\beta }|u|^{q}\,\text{d}x \\&\quad + \frac{\mu _0 a^*}{\mu _0+1} \int _{\Omega }\vartheta _{R}^{\beta }|u|^{2}\,\text{d}x \le C\varepsilon . \end{aligned}$$
(4.34)

Applying Gronwall’s Lemma, we have

$$\begin{aligned} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{2}\,\text{d}x\le e^{-\frac{\mu _0 a^*}{\mu _0+1} t} \Vert u_0\Vert _{L^{2}} + C\varepsilon \le C\varepsilon \ \ \text{ for } \text{ any }\ t\ge \widehat{T}_{3},\end{aligned}$$
(4.35)

for some \(\widehat{T}_3>T_1+k_0-1\). Next, we insert (4.35) into (4.34) and integrate on \([t, t+1]\) with \(t\ge \widehat{T}_3\) to get the estimate

$$\begin{aligned} \int _t^{t+1} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{q}\,\text{d}x\,\text{d}t \le C\varepsilon +\int _{\Omega } \vartheta _{R}^{\beta }|u(t)|^{2}\,\text{d}x \le C\varepsilon \quad \text{ for } \text{ any }\ t\ge \widehat{T}_3. \end{aligned}$$
(4.36)

Secondly, we take \(\alpha =q-1\) in (4.32), and have

$$\begin{aligned}&\frac{\mu _0+1}{q} \frac{\,\text{ d }}{\,\text{d}t} \int _{\Omega }\vartheta _{R}^{\beta }|u|^{q}\,\text{d}x+ \frac{(q-1)(\mu _0+1)}{2}\int _{\Omega } \vartheta _{R}^{\beta }|\nabla u|^{p}| u|^{q-2}\,\text{d}x \\&\ \ \ \ +\frac{{\widetilde{\mu }}(2\mu _0+1)}{4} \int _{\Omega } \vartheta _{R}^{\beta }|u|^{2q-2}\,\text{d}x + \frac{\mu _0 a^*}{4} \int _{{\Omega }}\vartheta _{R}^{\beta }|u|^{q}\,\text{d}x \\&\le \frac{C}{R^p} \int _{A_R} \vartheta _{R}^{\beta -p}| u|^{q+p-2}\,\text{d}x + C\int _{\Omega } \vartheta _{R}^{\beta }\, g|u|^{q-2}u\,\text{d}x+C \varepsilon . \end{aligned}$$
(4.37)

Applying (4.17) in Remark 4.3 with \(k_0\) as above, i.e. \((k_0+1)(q-2)+2> p+q-2\), we get with some \(\theta \in (0,1)\) that

$$\begin{aligned} \int _{A_R} \vartheta _{R}^{\beta -p}|u|^{p+q-2}\,\text{d}x \le \left\{ \begin{array}{l} \displaystyle \Vert u\Vert _{L^{p+q-2}}^{p+q-2} \le \Vert u\Vert _{L^{2}}^{(p+q-2)\theta } \Vert u\Vert _{L^{(k_0+1)(q-2)+2}}^{(1-\theta )(p+q-2)} \ \;\; \text{ if } p\ge 2, \\ \displaystyle \int _{A_R} |u|^{p+q-2}\,\text{d}x\le C\Vert u\Vert _{L^{q}}^{p+q-2}R^{\frac{N(2-p)}{q}}\ \qquad \text{ if } \frac{2N}{2+N}<p< 2. \end{array} \right. \end{aligned}$$

Using Proposition 4.1 and (4.37) and choosing \(R_1>R_0\) sufficiently large, it follows that

$$\begin{aligned} \frac{C}{R^p} \int _{A_R} \vartheta _{R}^{\beta -p}|u|^{q+p-2}\,\text{d}x \le C\varepsilon \quad \text{ for } \text{ any }\ t\ge T_1+k_0-1 \text{ and } R\ge R_1. \end{aligned}$$

Moreover, by (4.15), we also get for any \( t\ge T_1+1\) and \(R\ge R_1\)

$$\begin{aligned} \int _{\Omega }\vartheta _{R}^{\beta }\, g |u|^{q-2}u\,\text{d}x \le \left( \int _{\Omega } \vartheta _{R}^{2\beta }\, |g|^{2}\,\text{d}x \right) ^{\frac{1}{2}} \left( \int _{\Omega } |u|^{2q-2}\,\text{d}x \right) ^{\frac{1}{2}} \le C\varepsilon . \end{aligned}$$

Therefore,

$$\begin{aligned} \frac{\,\text{ d }}{\,\text{d}t}\int _{\Omega } \vartheta _{R}^{\beta }|u|^{q}\,\text{d}x + \frac{q{\widetilde{\mu }}(2\mu _0+1)}{4(\mu _0+1)}\int _{\Omega } \vartheta _{R}^{\beta }|u|^{2q-2}\,\text{d}x \le C\varepsilon \ \ \ \text{ for } \text{ any }\ t\ge T_1+k_0 \text{ and } R\ge R_1. \end{aligned}$$

Now the Uniform Gronwall Lemma 2.6, (4.36) and (4.38) yield the inequality

$$\begin{aligned} \int _{\Omega } \vartheta _{R}^{\beta }|u(t)|^{q}\,\text{d}x \le C\varepsilon \ \ \ \text{ for } \text{ any }\ t\ge \max \{T_1+k_0,\widehat{T}_3\}, \end{aligned}$$
(4.38)

which finishes the proof of this proposition with \(T_3:=\max \{T_1+k_0,\widehat{T}_3\}\). \(\square\)

Theorem 4.5

Under the assumptions of Theorem 1.1(1), (2) and for \(\frac{2N}{N+2}<p<N\) the semigroup \({S(t)}_{t\ge 0}\) generated by (1.1) with initial data \(u_0\in L^2\) has an \((L^2,L^2\cap L^q)\)-global attractor \(\mathcal {A}\), that is, \(\mathcal {A}\) is compact, invariant in \(L^2\cap L^q\) and attracts every bounded subset of \(L^2\) in the topology of \(L^2\cap L^q\).

Proof

Let \(B\subset L^2(\Omega )\) be bounded. It suffices to verify that the semigroup \({S(t)}_{t\ge 0}\) is \(\omega\)–limit compact. For any fixed R, by Proposition 4.1 and Proposition 4.2, there exists \(T_2=T_2(B,R)\) such that

$$\begin{aligned} \bigcup \nolimits _{t\ge T_2} \bigcup \nolimits _{u_0\in B} (1-\vartheta _R)S(t)u_0 \quad \text{ is } \text{ bounded } \text{ in } \;\; W^{1,p}(\Omega _{2R})\, \text{ and } \; L^{2q-2}(\Omega _{2R}).\end{aligned}$$

Then by the compact embedding \( W^{1,p}(\Omega _{2R})\hookrightarrow L^{2}(\Omega _{2R})\) and the embedding \(L^2\cap L^{2q-2} \hookrightarrow L^q\) we obtain that \(\bigcup \nolimits _{t\ge T_2}\bigcup \nolimits _{u_0\in B}(1-\vartheta _R)S(t)u_0\) is precompact in \(L^{2}(\Omega _{2R})\cap L^{q}(\Omega _{2R})\); hence

$$\begin{aligned} \kappa \left( \bigcup \nolimits _{t\ge T_2}\bigcup \nolimits _{u_0\in B} (1-\vartheta _R)S(t)u_0\right) _{L^{2}\cap L^{q}}=0\ \ \text{ for } \text{ any }\ R>0. \end{aligned}$$

On the other hand, due to Proposition 4.4, there exists positive constants \(T_3=T_3(B,\varepsilon )\) and \(R=R(\varepsilon )\) such that

$$\begin{aligned} \Big \Vert \bigcup \nolimits _{t\ge T_3} \bigcup \nolimits _{u_0\in B} \vartheta _RS(t)u_0\Big \Vert _{L^{2}\bigcap L^{q}}\le C\varepsilon , \end{aligned}$$

so that in view of Proposition 2.4

$$\begin{aligned} \kappa \left( \bigcup \nolimits _{t\ge T_3} \bigcup \nolimits _{u_0\in B}\vartheta _RS(t)u_0\right) _{L^{2}\cap L^{q}}<C\varepsilon . \end{aligned}$$

Choosing \(T_3>T_2\), we see that

$$\begin{aligned}&\kappa \left( \bigcup \nolimits _{t\ge T_3} \bigcup \nolimits _{u_0\in B}S(t)u_0\right) _{L^{2}\cap L^{q}} \\&\quad \le \kappa \left( \bigcup \nolimits _{t\ge T_3}\bigcup \nolimits _{u_0\in B}(1-\vartheta _R)S(t)u_0\right) _{L^{2}\cap L^{q}}+\kappa \left( \bigcup \nolimits _{t\ge T_3}\bigcup \nolimits _{u_0\in B}\vartheta _RS(t)u_0\right) _{L^{2}\cap L^{q}} \\&\quad <C\varepsilon , \end{aligned}$$
(4.39)

from which we conclude that the semigroup \(\{S(t)\}_{t\ge 0}\) is \(\omega\)–limit compact. Therefore, the existence of a global attractor attracting every bounded subset \(B\subset L^2\) in the topology of \(L^2\cap L^q\) is proved. \(\square\)

5 Dimension of the global attractor

In this section, we use the method of \(\ell\)-trajectories (see [26]) to estimate the finite dimension of the global attractor in case of a bounded domain \(\Omega \subset {\mathbb {R}}^N\) and \(p\ge 2\). We first recall the definition of the fractal dimension of a compact subset \(\mathscr {C}\) of a metric space \(\mathscr {X}\), denoted by

$$\begin{aligned} d^{\mathscr {X}}_{f}(\mathscr {C}) := \limsup _{\varepsilon \rightarrow 0} \frac{\log N_{\varepsilon }^{\mathscr {X}}(\mathscr {C})}{\log (1/\varepsilon )}, \end{aligned}$$

where \(N_{\varepsilon }^{\mathscr {X}}(\mathscr {C})\) is the minimal number of \(\varepsilon\)-ball (with respect to the metric of \(\mathscr {X}\)) needed to cover \(\mathscr {C}\).

Lemma 5.1

[26, Lemma 1.3] Let \(\mathscr {X},\mathscr {Y}\) by normed spaces such that \(\mathscr {Y}\hookrightarrow \hookrightarrow \mathscr {X}\) and \(\mathscr {C}\subset \mathscr {X}\) is bounded. Assume that there exists a mapping \(\mathscr {L}: \mathscr {X} \mapsto \mathscr {Y}\) such that \(\mathscr {L}\mathscr {C}=\mathscr {C}\) and \(\mathscr {L}\) is Lipschitz continuous on \(\mathscr {C}\). Then \(d^{\mathscr {X}}_{f}(\mathscr {C})\) is finite.

Remark 5.2

In view of the proof in [26], the same result as Lemma 5.1 holds if \(\mathscr {L}:\mathscr {X}\mapsto \mathscr {Y}\) is Hölder continuous on \(\mathscr {C}\).

For convenience, we introduce the space of trajectories

$$\begin{aligned}\mathcal {X}_{\ell } = \{\omega : (0, \ell )\rightarrow L^{2}: \ \omega \ \text{ is } \text{ a } \text{ solution } \text{ of } (1.1)\ \text{ on }\ (0,\ell )\}, \end{aligned}$$

where \(\ell >0\) is a constant to be fixed, which we endow with the topology of \(\mathcal {X}_{\ell }:=L^{2}(0,\ell ; L^{2})\). Then we define the semigroup \(L_t\) on \(\mathcal {X}_\ell\) by

$$\begin{aligned} \{L_t\}(\omega )(\tau ):=u(t+\tau ), \ \ \tau \in [0,\ell ], \end{aligned}$$

where u is the unique solution on \([0,\ell +t]\) such that \(u|_{[0,\ell ]}=\omega\). We take

$$\begin{aligned} \mathcal {B}_0 = \overline{\bigcup \nolimits _{t\ge T_1} S(t)B_0}^{L^2_w}, \end{aligned}$$

where \(B_0\) is bounded absorbing set in \(L^{2}\) and \(T_1\) is given in Proposition 4.1 such that \(S(t)B_0\subset B_0\) for all \(t\ge T_1\); the closure is taken in the weak topology of \(L^{2}\). Furthermore, let

$$\begin{aligned} \mathscr {B}_{\ell }=\{\omega \in \mathcal {X}_{\ell }: \ \omega (0)\in \mathcal {B}_0\}. \end{aligned}$$

Theorem 5.3

The dynamical system \((L_t, \mathcal {X}_{\ell })\) possesses a global attractor \(\mathscr {A}_{\ell }\).

Proof

From Proposition 4.1, \(\mathscr {B}_{\ell }\) is positively invariant with respect to \(L_t\). In view of the proof in Theorem 3.2, the solutions to (1.1) are compact in the following sense:

$$\begin{aligned} \text {(CP)} \quad {\left\{ \begin{array}{ll} \text {Let }\{u_n(t)\}\text { be a sequence of solutions on }[0,T]\text { in the sense of }(3.1)\text { with initial data}\\ \text {given by a sequence }\{u_{0n}\}\text { which is bounded in }L^{2}\text { and converges in }L^2\text { to }u_0.\\ \text {Then there exists a subsequence of the solutions }\{u_n(t)\}\text { converging weakly in }V\\ \text {to a solution }u\text { of }(3.1)\text { on }[0,T]\text { with the initial data }u_{0}. \end{array}\right. } \end{aligned}$$

Applying (CP), we have

$$\begin{aligned} \overline{\mathscr {B}_{\ell }}^{\mathcal {X}_{\ell }}\subset \mathscr {B}_{\ell }. \end{aligned}$$
(5.1)

In fact, let \(\{\omega _n\}\subset \mathscr {B}_{\ell }\) be a sequence of trajectories such that \(\omega _n\rightarrow \omega\) in \(\mathcal {X}_{\ell }\). Then, by (CP), \(\omega\) is also a trajectory, i.e. \(\omega \in X_\ell\). It remains to show that \(\omega (0)\in \mathcal {B}_0\). Firstly, by Proposition 4.1, \(\omega _{n}(t)\in \mathcal {B}_0\) for all t. Secondly, at least for a subsequence, by (3.23), \(\omega _n(t)\rightarrow \omega (t)\) in \(L^{2}\) for almost all t, and since \(\mathcal {B}_0\) is closed, \(\omega (t)\in \mathcal {B}_0\) for almost all \(t\in (0,\ell )\). In particular, \(\omega (t)\in \mathcal {B}_0\) for points t arbitrarily close to 0, which by the continuity of \(\omega :[0,\ell ]\mapsto L^{2}\) and the closedness of \(\mathcal {B}_0\) implies that \(\omega (0)\in \mathcal {B}_0\). Therefore, Theorem 3.2, Proposition 4.1 and (5.1) guarantee that all the assumptions of [26, Lemma 1.1] are satisfied with \(\Sigma _t=L_t, \mathscr {X}=\mathcal {X}_\ell\) and \(\mathscr {B}^{1}=\overline{L_{t}\mathscr {B}_{\ell }}^{\mathcal {X}_\ell }\), and we finish the proof of this theorem. \(\square\)

Lemma 5.4

Let \(u_1\) and \(u_2\) be two solutions of (1.1) starting from \(\mathcal {B}_0\). Then, there exists a constant \(\ell >0\) such that

$$\begin{aligned} \int _\ell ^{2\ell } \Vert \nabla (u_1-u_2)\Vert _{L^{p}}^p\,\text{d}s + \int _\ell ^{2\ell } \Vert (u_1-u_2)\Vert _{L^{q}}^q\,\text{d}s \le C\int _0^{\ell } \Vert u_1-u_2\Vert _{L^{2}}^2\,\text{d}s, \end{aligned}$$
(5.2)

where the constant \(C>0\) is independent of \(u_1(0)\) and \(u_2(0)\).

Proof

We set \(v(t)=u_1(t)-u_2(t)\). We have

$$\begin{aligned} v_t-(\Delta _{p}u_1-\Delta _{p}u_2)+\mu (|u_1| ^{q-2}u_1-|u_2|^{q-2}u_2)=(f(x,u_1)-f(x,u_2)). \end{aligned}$$
(5.3)

Similarly to (3.34), we obtain the inequality

$$\begin{aligned} \frac{1}{2}\frac{\,\text{ d }}{\,\text{d}t}\Vert v\Vert _{L^{2}}^2+\widetilde{\lambda }\Vert \nabla v\Vert _{L^{p}}^p+\widetilde{\mu } \Vert v\Vert _{L^{q}}^q\le 0. \end{aligned}$$
(5.4)

For \(\ell >0\), one has

$$\begin{aligned} \Vert v(\tau )\Vert _{L^{2}}^2\le \Vert v(s)\Vert _{L^{2}}^2 \ \text{ for } 0\le \tau -s\le 2\ell . \end{aligned}$$
(5.5)

We fix \(s\in (0,\ell )\) and integrate (5.4) over \(\tau \in [s,2\ell ]\) to obtain

$$\begin{aligned} \Vert v(2\ell )\Vert _{L^{2}}^2 + 2\widetilde{\lambda } \int _s^{2\ell } \Vert \nabla v(\tau )\Vert _{L^{p}}^p\,\text{d}\tau + 2\widetilde{\mu } \int _s^{2\ell } \Vert v(\tau )\Vert _{L^{q}} ^q\,\text{d}\tau \le \Vert v(s)\Vert _{L^{2}}^2. \end{aligned}$$
(5.6)

From above inequality, we have

$$\begin{aligned} \int _\ell ^{2\ell } \Vert \nabla v(\tau )\Vert _{L^{p}}^p\,\text{d}\tau + \int _\ell ^{2\ell } \Vert v(\tau )\Vert _{L^{q}}^q\,\text{d}\tau \le C\Vert v(s)\Vert _{L^{2}}^2, \end{aligned}$$
(5.7)

where C depends on \(\widetilde{\lambda }\) and \(\widetilde{\mu }\). Finally, we finish the proof by integrating (5.7) over \(s\in (0,\ell ).\) \(\square\)

Lemma 5.5

Let \(p\ge 2\) and let \(u_1, u_2\) be as in Lemma 5.4. Then \(v=u_1-u_2\) satisfies the estimate

$$\begin{aligned} \left\| \frac{\partial v}{\partial t}\right\| _{L^1(\ell ,2\ell ; W^*)} \le C\left( \Vert v\Vert _{L^2(0,\ell ; L^2)}^{\frac{2}{p}} + \Vert v\Vert _{L^2(0,\ell ; L^2)}^{\frac{2}{q}} + \Vert v\Vert _{L^2(0,\ell ; L^2)}^{\frac{2(p-1)}{p}} + \Vert v\Vert _{L^2(0,\ell ; L^2)}^{\frac{2(q-1)}{q}} \right) , \end{aligned}$$

where \(W^*=\mathcal {D}^{-1,p^\prime }+ L^{q^\prime }\) is the dual of the space \(W=\mathcal {D}_0^{1,p}\cap L^{q}\).

Proof

We begin with (5.3) written in the form

$$\begin{aligned} v_t&= (\Delta _{p}u_1 -\Delta _{p}u_2)-\mu (|u_1|^{q-2}u_1-|u_2|^{q-2}u_2)+(f(x,u_1)-f(x,u_2)) \\&=: I+II+III. \end{aligned}$$
(5.8)

Let us estimate the norm of I, II, and III considered as functionals on W. Concerning I we have with \(c_p=(p-1)\max (1,2^{p-1})\)

$$\begin{aligned} \left| \int _\ell ^{2\ell } \left\langle (\Delta _pu_1-\Delta _pu_2),\varphi \right\rangle \,\text{d}t\right|&\le \int _\ell ^{2\ell } \int _{\Omega } \left| |\nabla u_1|^{p-2}\nabla u_1 -|\nabla u_2|^{p-2}\nabla u_2\right| |\nabla \varphi |\,\text{d}x\,\text{d}t\\&\le \int _\ell ^{2\ell } \int _{\Omega } \left( |\nabla v| \nabla u_1 + |\nabla u_2|^{p-2}\nabla v\right) |\nabla \varphi |\,\text{d}x\,\text{d}t\;\\&\le c_p\big (\Vert u_1\Vert ^{p-2}_{L^{p}(\ell ,2\ell ; \mathcal {D}^{1,p}_0)} \\&\quad + \Vert u_2\Vert ^{p-2}_{L^{p}(\ell ,2\ell ; \mathcal {D}^{1,p}_0)}\big )\Vert v\Vert _{L^{p}(\ell ,2\ell ; \mathcal {D}^{1,p}_0)}\Vert \varphi \Vert _{L^{p}(\ell ,2\ell ; \mathcal {D}^{1,p}_0)}. \end{aligned}$$

Since \(u_1\) and \(u_2\) are solutions to (1.1), from (5.2) we conclude that

$$\begin{aligned} \Vert I\Vert _{L^{p^\prime }(\ell ,2\ell ; \mathcal {D}^{-1,p^\prime })} \le C\Vert v\Vert _{L^2(0,\ell ;L^2)}^{\frac{2}{p}}. \end{aligned}$$
(5.9)

For II, since

$$\begin{aligned}&\Bigg | \int _\ell ^{2\ell } \big \langle (|u_1|^{q-2}u_1- |u_2|^{q-2}u_2),\varphi \big \rangle \,\text{d}t \Bigg | \le c_q\int _\ell ^{2\ell } \int _\Omega \big (|u_1|^{q-2}+|u_2|^{q-2}\big ) |v||\varphi |\,\text{d}x\,\text{d}t\\&\quad \le c_q\big (\Vert u_1\Vert ^{q-2}_{L^q(\ell ,2\ell ; L^q)} + \Vert u_2\Vert ^{q-2}_{L^q(\ell ,2\ell ; L^q)}\big ) \Vert v\Vert _{L^q(\ell ,2\ell ; L^q)} \Vert \varphi \Vert _{L^q(\ell ,2\ell ;L^q)}, \end{aligned}$$

by (5.2) we get that

$$\begin{aligned} \Vert II\Vert _{L^{q^\prime }(\ell ,2\ell ; L^{q^\prime })} \le C\Vert v\Vert _{L^{2}(0,\ell ;L^{2})}^{\frac{2}{q}} . \end{aligned}$$
(5.10)

Finally, by (1.8) and (1.9), we deduce

$$\begin{aligned}&\Bigg | \int _\ell ^{2\ell } \big \langle (f(x,u_1)-f(x,u_2)),\varphi \big \rangle \,\text{d}t\Bigg | \\&\quad \le \int _\ell ^{2\ell } \int _\Omega \Big (\nu |u_1-u_2|^{q-1} + m(x)\lambda |u_1-u_2|^{p-1}\Big ) |\varphi |\,\text{d}x\,\text{d}t\\&\quad \le \nu \int _\ell ^{2\ell } \int _\Omega |v|^{q-1} |\varphi |\,\text{d}x\,\text{d}t + \lambda \int _\ell ^{2\ell } \int _\Omega m(x) |v|^{p-1}| \varphi |\,\text{d}x\,\text{d}t \\&\quad \le \nu \int _\ell ^{2\ell } \int _\Omega |v|^{q-1} |\varphi |\,\text{d}x\,\text{d}t\\&\qquad + \lambda \Bigg (\int _\ell ^{2\ell } \int _\Omega m(x) |v|^p| \,\text{d}x\,\text{d}t\Bigg )^{\frac{p-1}{p}} \Bigg (\int _\ell ^{2\ell } \int _\Omega m(x) |\varphi |^p\,\text{d}x\,\text{d}t \Bigg )^{1/p}\\&\quad \le C\Vert v\Vert ^{q-1}_{L^q(\ell ,2\ell ;L^q)} \Vert \varphi \Vert _{L^q(\ell ,2\ell ;L^q)} + C\Vert v\Vert ^{p-1}_{L^p(\ell ,2\ell ;{\mathcal {D}}^p)} \Vert \varphi \Vert _{L^p(\ell ,2\ell ; {\mathcal {D}}^p)} . \end{aligned}$$

Now it follows from (5.2) that

$$\begin{aligned} \Vert III\Vert _{L^{p^\prime }(\ell ,2\ell ; \mathcal {D}^{-1,p^\prime })+L^{q^\prime }(\ell ,2\ell ; L^{q^\prime })}&\le C\Big (\Vert u_1-u_2\Vert _{L^{2}(0,\ell ; L^{2})}^{\frac{2(q-1)}{q}}+\Vert u_1-u_2\Vert ^{\frac{2(p-1)}{p}}_{L^{2}(\ell ,2\ell ; L^2)}\Big ). \end{aligned}$$
(5.11)

Since \(q>2\) and \(\Omega\) is bounded, we see that \(\mathcal {D}_0^{1,p}\cap L^{2}\cap L^{q}=\mathcal {D}_0^{1,p}\cap L^{q}\), and hence \(W^*+L^{2}=W^*\). Besides, it is obvious that

$$\begin{aligned}L^{p^\prime }(\ell ,2\ell ; \mathcal {D}^{-1,p^\prime })+L^{q^\prime }(\ell ,2\ell ; L^{q^\prime })+L^{2}(\ell ,2\ell ; L^{2})\subset L^{1}(\ell ,2\ell ;W^*+L^{2})=L^{1}(\ell ,2\ell ;W^*).\end{aligned}$$

By (5.3), (5.8), (5.10), (5.11), we get the result of the lemma. \(\square\)

Theorem 5.6

The fractal dimension of the global attractor \(\mathscr {A}_\ell \subset \mathcal {X}_\ell\) introduced in Theorem 5.3is finite, i.e., \(d_{f}^{\mathcal {X}_\ell }(\mathscr {A}_\ell )\) is finite.

Proof

We set

$$\begin{aligned} \mathcal {W}_{\ell }:&= \{u\in L^{2}(0,\ell ; \mathcal {D}_{0}^{1,p}\cap L^{q}); \ u^{\prime }\in L^{1}(0,\ell ; W^*)\}, \end{aligned}$$
(5.12)

and note the compact embedding \(\mathcal {W}_\ell \hookrightarrow \hookrightarrow \mathcal {X}_{\ell }\), since \(\mathcal {D}_{0}^{1,p}\cap L^{q}\hookrightarrow \hookrightarrow L^{2}\). Then by Lemmas 5.4 and 5.5, we obtain with \(\mathcal {L}:=L_{\ell }\) for all \(\omega _1,\omega _2\in \mathscr {B}_{\ell }\) that

$$\begin{aligned} \Vert \mathcal {L}\omega _1-\mathcal {L}\omega _2\Vert _{L^{2}(0,\ell ; \mathcal {D}_{0}^{1,p}\cap L^{q})}&\le C\Big (\Vert \omega _1-\omega _2\Vert _{L^{2}(0,\ell ; L^{2})}^{\frac{2}{p}}+\Vert \omega _1-\omega _2\Vert _{L^{2}(0,\ell ; L^{2})}^{\frac{2}{q}}\Big ), \end{aligned}$$

and

$$\begin{aligned}&\big \Vert \partial _t\big (\mathcal {L}\omega _1-\mathcal {L}\omega _2\big )\big \Vert _{L^{1}(0,\ell ; W^*)}\\&\quad \le C\Big ( \Vert \omega _1-\omega _2\Vert _{L^{2}(0,\ell ;L^{2})}^{\frac{2(p-1)}{p}} + \Vert \omega _1-\omega _2\Vert _{L^{2}(0,\ell ; L^{2})}^{\frac{2}{p}}\\&\qquad + \Vert \omega _1-\omega _2\Vert _{L^{2}(0,\ell ; L^{2})}^{\frac{2(q-1)}{q}} + \Vert \omega _1-\omega _2\Vert _{L^{2}(0,\ell ;L^{2})}^{\frac{2}{q}} \Big ). \end{aligned}$$

Therefore, \(\mathcal {L}:\mathcal {X}_{\ell }\mapsto \mathcal {X}_{\ell }\) is Hölder continuous on \(\mathscr {A}_{\ell }\). We apply Lemma 5.1 with \(\mathscr {X}=\mathcal {X}_{\ell }, \mathscr {Y}=\mathcal {W}_{\ell }, \mathscr {L}=\mathcal {L}\) and \(\mathscr {C}=\mathscr {A}_\ell\). Since \(\mathscr {A}_\ell \subset \mathcal {X}_\ell\), \(L_{t}\mathscr {A}_\ell =\mathscr {A}_\ell\), we finish the proof using Remark 5.2. \(\square\)

Finally, for all \(\tau >0\), we apply (5.5) to get that \(S_t: L^{2}\mapsto L^{2}\) is uniformly (with respect to \(t\in [0,\tau ]\)) Lipschitz continuous on \(L^{2}\). Then, by [26, Lemma 2.1], the operator \(L_t: \mathcal {X}_{\ell }\mapsto \mathcal {X}_{\ell }\) is uniformly Lipschitz continuous in \(t\in [0,\tau ]\) on \(\mathscr {B}_{\ell }\). Hence, also the operator \(e :\mathcal {X}_{\ell }\rightarrow L^{2}\) defined by \(e(v) = v(l)\) i.e., e maps an l-trajectory onto its endpoint, is Lipschitz continuous on \(\mathscr {B}_\ell\). Therefore, by [26, Theorem 2.4], the global attractor \(\mathscr {A}_\ell\) is finite dimensional.

Besides, by (4.5), we see that the unique solution u satisfies \(u^{\prime }\in L^{2}(0,\ell ; L^{2})\). Then, by [26, Lemma 2.2], the mapping \(t\mapsto L_t\omega\) has the following property: there exists \(c>0\) such that for all \(\omega \in \mathscr {B}_{\ell }\) and \(t_1, t_2\in [0,\tau ]\) it holds

$$\begin{aligned} \Vert L_{t_1}\omega -L_{t_2}\omega \Vert _{X_{\ell }}\le c|t_1-t_2|^{1/2}. \end{aligned}$$

Thus, it can be proved (see [26] for details of the proof) that the dynamical system \((L_t, \mathcal {X}_{\ell })\) possesses an exponential attractor \(\mathscr {E}_{\ell }\). Since e is Lipschitz, it follows that \(\mathscr {E}=\mathscr {E}_{\ell }\) is an exponential attractor for S(t) on \(L^{2}\). Since an exponential attractor has, by definition, finite fractal dimension and always contains the global attractor, in view of Theorem 5.6, we get the following theorem.

Theorem 5.7

If \(\Omega\) is a bounded domain of \({\mathbb {R}}^N\) and \(p\ge 2\), then the global attractor associated with (1.1) has finite fractal dimension.