Loading [Contrib]/a11y/accessibility-menu.js
Research article

Positive ground state solutions for asymptotically periodic generalized quasilinear Schrödinger equations

  • Received: 14 August 2021 Accepted: 08 October 2021 Published: 19 October 2021
  • MSC : 35J60, 35J50

  • In this paper, we study the existence of a positive ground state solution for a class of generalized quasilinear Schrödinger equations with asymptotically periodic potential. By the variational method, a positive ground state solution is obtained. Compared with the existing results, our results improve and generalize some existing related results.

    Citation: Shulin Zhang. Positive ground state solutions for asymptotically periodic generalized quasilinear Schrödinger equations[J]. AIMS Mathematics, 2022, 7(1): 1015-1034. doi: 10.3934/math.2022061

    Related Papers:

  • In this paper, we study the existence of a positive ground state solution for a class of generalized quasilinear Schrödinger equations with asymptotically periodic potential. By the variational method, a positive ground state solution is obtained. Compared with the existing results, our results improve and generalize some existing related results.



    In this paper, we are concerned with the following generalized quasilinear Schrödinger equations

    $ \begin{equation} -\text{div}(g^{2}(u)\nabla u)+ g(u)g'(u)|\nabla u|^{2}+ V(x)u = h(x, u), \; \; x\in \mathbb{R}^{N}, \end{equation} $ (1.1)

    where $ N\geq 3 $, $ g\in C^{1}(\mathbb{R}, \mathbb{R}^{+}) $ is an even function with $ g'(t)\geq 0 $ for all $ t\geq 0 $ and $ g(0) = 1 $, $ h\in C(\mathbb{R}^{N}\times \mathbb{R}, \mathbb{R}) $ and $ V(x) $ is 1-periodic or asymptotically periodic potential.

    The equations are related to the solitary wave solutions for quasilinear Schrödinger equations

    $ \begin{equation} iz_{t} = -\triangle z + W(x)z - h(x, z) - \triangle(l(|z|^{2}))l'(|z|^{2})z, \; \; x\in \mathbb{R}^{N}, \end{equation} $ (1.2)

    where $ W: \mathbb{R}^{N} \rightarrow \mathbb{R} $ is a given potential, $ z: \mathbb{R} \times \mathbb{R}^{N}\rightarrow \mathbb{C} $, $ h $ and $ l $ can be used to model physical phenomenon. The form of (1.2) has many applications in physics. For instance, the case $ l(s) = s $ models the time evolution of the condensate wave function in superfluid film [13,21]. In the case of $ l(s) = \sqrt{1+s} $, problem (1.2) models the self-channeling of a high-power ultrashort laser in matter [1,30]. For more physical background, we can refer to [2,5,20,29] and references therein.

    Putting $ z(t, x) = \exp(-iEt)u(x) $ in (1.2), where $ E\in \mathbb{R} $ and $ u $ is a real function, we obtain the following elliptic equation

    $ \begin{equation} -\triangle u + V(x)u -\triangle(l(u^{2}))l'(u^{2})u = h(x, u), \; \; x\in \mathbb{R}^{N}. \end{equation} $ (1.3)

    When $ l(s) = s $, Eq (1.3) is the superfluid film equation in plasma physics

    $ \begin{equation} -\triangle u + V(x)u -\triangle(u^{2})u = h(x, u), \; \; x\in \mathbb{R}^{N}. \end{equation} $ (1.4)

    Note that Eq (1.4) is a special case of (1.1) if we choose $ g^{2}(u) = 1+ 2u^{2} $. Equation (1.4) has been extensively studied, to see [3,23,27,33,34].

    When $ l(s) = \sqrt{1+s} $, (1.3) derives the following equation

    $ \begin{equation} -\triangle u + V(x)u -[\triangle(1+u^{2})^{\frac{1}{2}}]\frac{u}{2(1+u^{2})^{\frac{1}{2}}} = h(x, u), \; \; x\in \mathbb{R}^{N}, \end{equation} $ (1.5)

    which is used as a model of the self-channeling of a high-power ultrashort laser in matter. We obverse that Eq (1.5) is also a particular case of (1.1) if we take $ g^{2}(u) = 1+ \frac{u^{2}}{2(1+u^{2})} $.

    For (1.5), there were many papers studying the existence of solutions, to see [9,12,32,39] and references therein. In [19], when the nonlinear term is autonomous, the authors obtained a positive ground state solution by a perturbation approach. In [4], the authors proved that (1.5) has at least a positive solution by using a change of variable, monotonicity trick developed by Jeanjean and a priori estimate. In [16], the authors got the existence of infinitely many nontrivial solutions by using a revised Clark theorem and a priori estimate of the solution.

    Furthermore, if we take $ g^{2}(u) = 1+ \frac{[(l(u^{2}))']^{2}}{2} $, then (1.3) turns into (1.1).

    In recent years, many scholars have studied (1.1), for example [6,7,10,11,22,24,28,31,35]. Particularly, in [31], the authors obtained a positive solution with the autonomous nonlinear term. In [6], the authors acquired the existence of ground state solutions with periodic potential. In [10,11], the authors established the existence of positive solutions with the critical exponents, where critical exponents are $ 2^{*} $ and $ \alpha2^{*} $, respectively. With regard to generalized quasilinear Schrödinger-Maxwell systems and generalized quasilinear Schrödinger equation of Kirchhoff type, we can refer to [8,26,40] and references therein. To the best of our knowledge, there is no work concerning with the unified asymptotic process on $ V $ and $ h $ at infinity for general quasilinear Schrödinger equations.

    Motivated by above papers, we establish the existence of a positive ground state solution for Eq (1.1) and the corresponding periodic equation. We point out that Eq (1.1) is more general than (1.4) and (1.5). There are several difficulties in dealing with Eq (1.1). The first one is that the energy functional associated with Eq (1.1) is not well defined in the whole space $ H^{1}(\mathbb{R}^{N}) $ due to the presence of second order nonhomogeneous term. Another difficulty is the lack of compactness owing to the unboundedness of the domain. The final difficulty is that the functional loses the translation invariance because of the asymptotically periodic potential. To overcome the above difficulties, we firstly introduce a change of variable and we reformulate quasilinear elliptic Eq (1.1) into semilinear elliptic equation, whose associated functional is well defined in our working space. Secondly, we employ the energy comparison method to overcome the loss of translation invariance. Finally, to find a positive ground state solution, we use the Concentration-Compactness Principle, Nehari manifold method and the strong maximun principle. Our results are a complement and generalization of some results obtained by [18,34,37].

    Now, let us recall some basic facts. Set

    $ \begin{equation*} H^{1}(\mathbb{R}^{N}) = \{u\in L^{2}(\mathbb{R}^{N}): \nabla u\in L^{2}(\mathbb{R}^{N})\}, \end{equation*} $

    with the norm

    $ \begin{equation*} ||u||_{H} = \Big(\int_{\mathbb{R}^{N}}(|\nabla u|^{2} + |u|^{2})dx \Big)^{\frac{1}{2}}. \end{equation*} $

    It is well known that in the study of the elliptic equations, the potential function $ V $ plays an important role in dealing with compactness problem. Let us introduce the following working space

    $ \begin{equation*} X = \{u\in H^{1}(\mathbb{R}^{N}): \int_{\mathbb{R}^{N}}V(x)|u|^{2}dx < \infty \}, \end{equation*} $

    endowed with the norm

    $ \begin{equation*} ||u|| = \Big(\int_{\mathbb{R}^{N}}(|\nabla u|^{2} +V(x)|u|^{2})dx\Big)^{\frac{1}{2}}, u\in X. \end{equation*} $

    Then, the subsequent condition $ (V) $ implies that the norm $ ||\cdot || $ is equivalent to the norm $ ||\cdot ||_{H} $ (see [18]).

    We define the energy functional associated with (1.1) by

    $ \begin{equation} \Psi(u) = \frac{1}{2}\int_{\mathbb{R}^{N}}g^{2}(u)|\nabla u|^{2}dx + \frac{1}{2}\int_{\mathbb{R}^{N}}V(x)u^{2}dx - \int_{\mathbb{R}^{N}}H(x, u)dx, \end{equation} $ (1.6)

    where $ H(x, u) = \int_{0}^{u}h(x, \tau)d\tau $. However, $ \Psi $ is not well defined in the usual Sobolev space $ H^{1}(\mathbb{R}^{N}) $ because of the term $ \int_{\mathbb{R}^{N}}g^{2}(u)|\nabla u|^{2}dx $. To overcome this difficulty, we make use of the change of variable introduced by [35],

    $ \begin{equation*} v = G(u) = \int_{0}^{u}g(t)dt. \end{equation*} $

    After the change of variable, we can obtain the following functional

    $ \begin{equation} \Phi(v) = \frac{1}{2}\int_{\mathbb{R}^{N}}\Big[|\nabla v|^{2} + V(x)(G^{-1}(v))^{2}\Big]dx - \int_{\mathbb{R}^{N}}H(x, G^{-1}(v))dx, \end{equation} $ (1.7)

    where $ G^{-1}(v) $ is the inverse function of $ G(u) $. Since $ g $ is a nondecreasing positive function, we can get $ |G^{-1}(v)|\leq \frac{1}{g(0)}|v| $. From this and our hypotheses, it is easy to verify that $ \Phi $ is well defined on $ X $ and $ \Phi\in \mathcal{C}^{1} $.

    In order to obtain a critical point of (1.1), it suffices to study the following semilinear equations

    $ \begin{equation} -\Delta v + V(x)\frac{G^{-1}(v)}{g(G^{-1}(v))} - \frac{h(x, G^{-1}(v))}{g(G^{-1}(v))} = 0, \; x\in \mathbb{R}^{N}. \end{equation} $ (1.8)

    If $ v $ is said to be a weak solution for Eq (1.8), then it should satisfy

    $ \begin{equation} \langle \Phi'(v), \varphi \rangle = \int_{\mathbb{R}^{N}}\Big[\nabla v \nabla \varphi + V(x)\frac{G^{-1}(v)}{g(G^{-1}(v))}\varphi -\frac{h(x, G^{-1}(v))}{g(G^{-1}(v))}\varphi\Big]dx = 0, \end{equation} $ (1.9)

    for any $ \varphi \in C_{0}^{\infty}(\mathbb{R}^{N}) $. We note that if $ v $ is a critical point of the functional $ \Phi $, then $ u = G^{-1}(v) $ is a critical point of the functional $ \Psi $, i.e. $ u = G^{-1}(v) $ is a solution of Eq (1.1).

    Next, we give the following condition on potential $ V $:

    $ \begin{equation*} \begin{aligned} &(V)\; 0 < V_{0}\leq V(x)\leq V_{p}(x)\in L^{\infty}(\mathbb{R}^{N}) \; \text{and} \; V(x)-V_{p}(x)\in \mathcal{F}_{0}, \; \text{where}\\ &\qquad\qquad\qquad\mathcal{F}_{0}: = \Big\{k(x): \forall \varepsilon > 0, \; \lim\limits_{|y|\rightarrow \infty}\text{meas}\; \{x\in B_{1}(y):|k(x)|\geq \varepsilon\} = 0\Big\}, \\ &\qquad V_{0} \; \text{is a positive constant and}\; V_{p} \; \text{is} \; 1-\text{periodic in}\; x_{i}, 1\leq i\leq N. \end{aligned} \end{equation*} $

    On nonlinearity term $ h $, since we look for a positive ground state solution, we assume that $ h(x, t) = 0 $ for all $ (x, t)\in \mathbb{R}^{N}\times \mathbb{R}^{-} $. Moreover, the function $ h $ satisfies the following assumptions:

    $ \begin{equation*} \begin{aligned} &(H_{1})\; \lim\limits_{t\rightarrow 0^{+}} \frac{h(x, t)}{g(t)G(t)} = 0 \; \text{uniformly for}\; x\in \mathbb{R}^{N}.\\ &(H_{2})\; \lim\limits_{t\rightarrow +\infty} \frac{h(x, t)}{g(t)G(t)^{2^{*}-1}} = 0 \; \text{uniformly for}\; x\in \mathbb{R}^{N}, \text{where}\; 2^{*} = \frac{2N}{N-2}\; \text{for}\; N\geq3.\\ &(H_{3})\; t\mapsto \frac{h(x, t)}{g(t)G(t)^{3}}\; \text{is nondecreasing on}\; (0, +\infty).\\ &(H_{4})\; \text{There exists a periodic function}\; h_{p}\in C(\mathbb{R}^{N}\times \mathbb{R^{+}}, \mathbb{R})\; \text{and}\; h_{p}(x, t) = 0\; \text{for all}\; (x, t)\in \\ &\mathbb{R}^{N}\times \mathbb{R^{-}}, \; \text{which is}\; 1-\text{periodic in}\; x_{i}, 1\leq i\leq N, \; \text{such that}\\ &\qquad(1)\; h(x, t)\geq h_{p}(x, t)\; \text{for all}\; (x, t)\in \mathbb{R}^{N}\times \mathbb{R^{+}}\; \text{and}\; h(x, t)-h_{p}(x, t)\in \mathcal{F}, \; \text{where}\\ &\qquad\mathcal{F}: = \Big\{k(x, t):\forall \varepsilon > 0, \lim\limits_{|y|\rightarrow \infty}\text{meas}\{x\in B_{1}(y):|k(x, t)|\geq \varepsilon\} = 0 \; \text{uniformly for}\; |t|\; \text{bound} \Big\}.\\ &\qquad(2)\; t\mapsto \frac{h_{p}(x, t)}{g(t)G(t)^{3}}\; \text{is nondecreasing on}\; (0, +\infty).\\ &\qquad(3)\; \lim\limits_{t\rightarrow +\infty}\frac{H_{p}(x, t)}{G(t)^{2}} = +\infty\; \text{uniformly for}\; x\in \mathbb{R}^{N}, \text{where}\; H_{p}(x, t) = \int_{0}^{t}h_{p}(x, \tau)d\tau.\\ \end{aligned} \end{equation*} $

    We employ $ (H_{1}) $, $ (H_{2}) $ and $ (H_{4}) $ conditions to ensure that the energy functional $ \Phi $ has the mountain pass geometry structure. However, under these hypotheses, this functional does not satisfy the Cerami compactness condition, since the domain is all $ \mathbb{R}^{N} $. We observe that the condition $ (H_{3}) $ and (3) of $ (H_{4}) $ are used in the proof of the boundedness of the Cerami sequence of the functional associated with (1.1). We also observe that the asymptotic process of $ (V) $ and (1) of $ (H_{4}) $ is uniform at infinity due to [18], and then was used in [17,25,38].

    Now, we present the first result of this paper.

    Theorem 1.1. (Asymptotically periodic case). Assume that $ (V) $ and $ (H_{1})-(H_{4}) $ hold. Then, Eq (1.1) has a positive ground state solution.

    Remark 1.2. Compared with the known results in [6,18,24,34,37], our results are new and different due to the following some facts:

    (1) While [18,34,37], our model is more general and they are our special case. In our results, there is no need to assume $ h(x, t)\in C^{1}(\mathbb{R}^{N}, \mathbb{R}) $. The constrained manifold also need not be of class $ C^{1} $.

    (2) Contrasting with [34], we give a new asymptotic process of potential and nonlinearity term at infinity. To some extent, we extend and complement their results.

    (3) In [6,24], they verified the mountain pass geometry structure under the following condition

    $ \begin{equation*} (H^{\prime}_{2}) \; |h(x, t)|\leq C(1 + g(t)|G(t)|^{\alpha})\; \text{for some} \; C > 0\; \text{and}\; \alpha\in(2, 2^{*}). \end{equation*} $

    Our condition $ (H_{2}) $ is somewhat weaker than $ (H^{\prime}_{2}) $.

    (4) We choose condition $ (H_{3}) $ and (3) of $ (H_{4}) $ to be weaker than Ambrosetti-Rabinowitz type condition.

    In the special case: $ V = V_{p}, h = h_{p} $, Theorem 1.1 clearly gives us a solution for the periodic equation. Indeed, considering the following equation

    $ \begin{equation} -\text{div}(g^{2}(u)\nabla u)+ g(u)g'(u)|\nabla u|^{2}+ V_{p}(x)u = h_{p}(x, u), \; \; x\in \mathbb{R}^{N}. \end{equation} $ (1.10)

    under the hypothesis:

    $ (V_{1}) $ The function $ V_{p} $ is 1-periodic in $ x_{i}, 1\leq i\leq N $ and there exists a constant $ V_{0} > 0 $ such that

    $ \begin{equation*} 0 < V_{0}\leq V_{p}(x)\in L^{\infty}(\mathbb{R}^{N}), \; \; \text{for all}\; x\in \mathbb{R}^{N}. \end{equation*} $

    Corollary 1.3. (Periodic case). Assume that $ (V_{1}) $ holds, $ h_{p} $ satisfies $ (H_{1}), (H_{2}) $, (2) of $ (H_{4}) $ and (3) of $ (H_{4}) $. Then, Eq (1.10) has a positive ground state solution.

    Remark 1.4. To the best of our knowledge, even for the periodic case, our method is new. In [26], under condition $ (V_{1}) $, $ h_{p} $ satisfies $ (H_{1}), (H_{2}) $ and other conditions. They could get at least one nontrivial ground state solution by the Mountain Pass Theorem. While, we obtain the existence of a positive ground state solution with Eq (1.10) by Nehari manifold method.

    The rest of this article is organized as follows: In Section 2, we present some preliminary lemmas. In Section 3, we give the proof for our result.

    Notations

    $ \bullet $ $ L^{s}(\mathbb{R}^{N}) $ is the usual Lebesgue space endowed with the norm

    $ \begin{equation*} ||u||_{s} = \Big(\int_{\mathbb{R}^{N}}|u|^{s}dx \Big)^{\frac{1}{s}}, \forall s\in[1, +\infty). \end{equation*} $

    $ \bullet $ $ ||u||_{\infty} = \text{ess}\sup_{x\in \mathbb{R}^{N}}|u(x)| $ denotes the usual norm in $ L^{\infty}(\mathbb{R}^{N}) $.

    $ \bullet $ $ (X^{*}, || \cdot ||_{*}) $ is the dual space of $ (X, || \cdot ||) $.

    $ \bullet $ $ B_{R}(x_{0}): = \{x\in \mathbb{R}^{N}:|x-x_{0}| < R\} $.

    $ \bullet $ $ C $ represent different positive constants.

    Before we prove the existence of a positive ground state solution for Eq (1.1), we present some useful lemmas.

    Lemma 2.1. [24,40] The functions $ g, G, G^{-1} $ satisfy the following properties:

    $\begin{array}{l} (1)\; \text{the functions}\; G(\cdot)\; \text{and}\; G^{-1}(\cdot) \; \text{are strictly increasing and odd};\\ (2)\; G(s)\leq g(s)s\; \text{for all}\; s\geq 0; G(s)\geq g(s)s\; \text{for all}\; s\leq 0;\\ (3)\; g(G^{-1}(s))\geq g(0) = 1\; \text{for all}\; s\in \mathbb{R};\\ (4)\; |((G^{-1}(s))'| = |\frac{1}{g(G^{-1}(s))}| < 1;\\ (5)\; \frac{G^{-1}(s)}{s}\; \text{is decreasing on}\; (0, +\infty)\; \text{and increasing on}\; (-\infty, 0);\\ (6)\; |G^{-1}(s)|\leq \frac{1}{g(0)}|s| = |s|\; \text{for all}\; s\in \mathbb{R};\\ (7)\; \frac{|G^{-1}(s)|}{g(G^{-1}(s))}\leq \frac{1}{g^{2}(0)}|s| = |s|\; \text{for all}\; s\in \mathbb{R};\\ (8)\; \frac{G^{-1}(s)s}{g(G^{-1}(s))}\leq |G^{-1}(s)|^{2}\; \text{for all}\; s\in \mathbb{R};\\ (9)\; \frac{G^{-1}(s)}{s^{3}g(G^{-1}(s))}\; \text{is nonincreasing on}\; (0, \infty);\\ (10)\; \lim\limits_{|s|\rightarrow 0}\frac{G^{-1}(s)}{s} = \frac{1}{g(0)} = 1\; \text{and} \end{array} $
    $ \begin{align*} \begin{split} \lim\limits_{|s|\rightarrow \infty}\frac{G^{-1}(s)}{s} = \left \{ \begin{array}{ll} \frac{1}{g(\infty)}, &\text{if}\; g\; \text{is bounded}, \\ 0, & \text{if}\; g\; \text{is unbounded}. \end{array} \right. \end{split} \end{align*} $

    Lemma 2.2. Assume that $ (H_{1})-(H_{4}) $ hold, then for all $ (x, s)\in \mathbb{R}^{N}\times \mathbb{R} $, we have

    $ \begin{equation} \frac{h(x, G^{-1}(s))}{4g(G^{-1}(s))}- H(x, G^{-1}(s))\geq 0, \; \; \frac{h_{p}(x, G^{-1}(s))}{4g(G^{-1}(s))}- H_{p}(x, G^{-1}(s))\geq 0. \end{equation} $ (2.1)

    For any $ \delta > 0 $, there exist $ r_{\delta} > 0, C_{\delta} > 0 $ and $ \alpha\in(2, 2^{*}) $ such that

    $ \begin{align} &0\leq \frac{h_{p}(x, G^{-1}(s))}{g(G^{-1}(s))}\leq\frac{h(x, G^{-1}(s))}{g(G^{-1}(s))}\leq \delta |s|, \; \forall (x, s)\in \mathbb{R}^{N} \times [-r_{\delta}, r_{\delta}], \end{align} $ (2.2)
    $ \begin{align} &0\leq\frac{h_{p}(x, G^{-1}(s))}{g(G^{-1}(s))}\leq \frac{h(x, G^{-1}(s))}{g(G^{-1}(s))}\leq \delta|s|+ C_{\delta}|s|^{2^{*}-1}, \; \forall (x, s)\in \mathbb{R}^{N}\times \mathbb{R}, \end{align} $ (2.3)
    $ \begin{align} &0\leq\frac{h_{p}(x, G^{-1}(s))}{g(G^{-1}(s))}\leq \frac{h(x, G^{-1}(s))}{g(G^{-1}(s))}\leq C_{\delta}|s|+ \delta|s|^{2^{*}-1}, \; \forall (x, s)\in \mathbb{R}^{N}\times \mathbb{R}, \end{align} $ (2.4)
    $ \begin{align} &0\leq\frac{h_{p}(x, G^{-1}(s))}{g(G^{-1}(s))}\leq \frac{h(x, G^{-1}(s))}{g(G^{-1}(s))}\leq \delta(|s|+ |s|^{2^{*}-1})+ C_{\delta}|s|^{\alpha-1}, \; \forall (x, s)\in \mathbb{R}^{N}\times\mathbb{ R}, \end{align} $ (2.5)
    $ \begin{align} &0\leq H_{p}(x, G^{-1}(s))\leq H(x, G^{-1}(s))\leq \frac{\delta}{2}|s|^{2}+ \frac{C_{\delta}}{2^{*}}|s|^{2^{*}}, \; \forall (x, s)\in \mathbb{R}^{N}\times \mathbb{R}, \end{align} $ (2.6)
    $ \begin{align} &0\leq H_{p}(x, G^{-1}(s))\leq H(x, G^{-1}(s))\leq \frac{\delta}{2}|s|^{2}+ \frac{\delta}{2^{*}}|s|^{2^{*}}+ \frac{C_{\delta}}{\alpha}|s|^{\alpha}, \; \forall (x, s)\in \mathbb{R}^{N}\times \mathbb{R}. \end{align} $ (2.7)

    Proof. We give the proof the first inequality of (2.1). From $ (H_{3}) $, we have

    $ \begin{align*} H(x, t)& = \int^{t}_{0}h(x, \tau)d\tau = \int^{t}_{0}\frac{h(x, \tau)}{g(\tau)G(\tau)^{3}}g(\tau)G(\tau)^{3}d\tau \\ &\leq \frac{h(x, t)}{g(t)G(t)^{3}}\int^{t}_{0}g(\tau)G(\tau)^{3}d\tau = \frac{h(x, t)}{g(t)G(t)^{3}}\int^{t}_{0}G(\tau)^{3}dG(\tau) = \frac{h(x, t)G(t)}{4g(t)}, \end{align*} $

    that is

    $ \begin{equation*} H(x, t)\leq \frac{h(x, t)G(t)}{4g(t)}, \end{equation*} $

    then, taking $ t = G^{-1}(s) $, (2.1) is proved. According to $ (H_{1})-(H_{4}) $ conditions, it is easy to deduce the inequalities of (2.2)–(2.7).

    Lemma 2.3. [18] Assume that condition $ (V) $ holds. Then, there are two positive constants $ d_{1} $ and $ d_{2} $ such that $ d_{1}||u||_{H}\leq ||u||\leq d_{2}||u||_{H} $ for all $ u\in X $.

    Define the Nehari manifold

    $ \begin{equation*} \mathcal{N}: = \{v\in X\setminus \{0\}: \langle\Phi'(v), v\rangle = 0\}, \end{equation*} $

    and set $ c: = \inf_{v\in \mathcal{N}}\Phi(v) $.

    Lemma 2.4. Assume that $ (V) $ and $ (H_{1})-(H_{4}) $ are satisfied, then for any $ v\in X, v\neq 0 $, there exists a unique $ t_{v} > 0 $ such that $ t_{v}v\in \mathcal{N} $. Moreover, the maximun of $ \Phi(tv) $ for $ t\geq 0 $ is arrived at $ t_{v} $.

    Proof. Let $ v\in X\setminus\{0\} $ and define a function $ f(t): = \Phi(tv) $ on $ [0, \infty) $. It follows from Lemma 2.1-(10), $ (H_{1}) $ and the Lebesgue dominated convergence theorem that

    $ \begin{align*} \frac{f(t)}{t^{2}}& = \frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla v|^{2}dx + \frac{1}{2}\int_{\mathbb{R}^{N}}V(x)\frac{(G^{-1}(tv))^{2}}{t^{2}}dx - \int_{\mathbb{R}^{N}}\frac{H(x, G^{-1}(tv))}{t^{2}}dx\\ & = \frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla v|^{2}dx + \frac{1}{2}\int_{\mathbb{R}^{N}}V(x)\frac{(G^{-1}(tv))^{2}}{t^{2}v^{2}}v^{2}dx - \int_{\mathbb{R}^{N}}\frac{H(x, G^{-1}(tv))}{t^{2}v^{2}}v^{2}dx\\ &\rightarrow \frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla v|^{2}dx + \frac{1}{2}\int_{\mathbb{R}^{N}}V(x)v^{2}dx, \end{align*} $

    when $ t\rightarrow 0 $, which shows that $ f(t) > 0 $ for $ t > 0 $ small enough. Set $ \Omega = \{x\in \mathbb{R}^{N}: v(x) > 0\} $. Using the fact that $ (1), (3) $ of $ (H_{4}) $ and the Fatou lemma, We have

    $ \begin{equation} \underset{t\rightarrow +\infty}{\lim\inf}\int_{\Omega}\frac{H(x, G^{-1}(tv))}{t^{2}v^{2}}v^{2}dx\geq \underset{t\rightarrow +\infty}{\lim\inf}\int_{\Omega}\frac{H_{p}(x, G^{-1}(tv))}{t^{2}v^{2}}v^{2}dx = +\infty. \end{equation} $ (2.8)

    Hence, from Lemma 2.1-(6) and (2.8), we can get

    $ \begin{equation*} \underset{t\rightarrow +\infty}{\lim\sup}\frac{f(t)}{t^{2}} \leq\frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla v|^{2}dx + \frac{1}{2}\int_{\mathbb{R}^{N}}V(x)v^{2}dx - \underset{t\rightarrow +\infty}{ \lim\inf}\int_{\Omega}\frac{H_{p}(x, G^{-1}(tv))}{t^{2}v^{2}}v^{2}dx, \end{equation*} $

    which deduces $ f(t)\rightarrow -\infty $ as $ t\rightarrow +\infty $. So there exists $ t_{v} > 0 $ such that $ f(t_{v}) = \max_{t > 0}f(t) $ and $ f'(t_{v}) = 0 $, i.e., $ \Phi(t_{v}v) = \max_{t > 0}\Phi(tv) $ and $ t_{v}v\in \mathcal{N} $.

    The condition $ f'(t) = 0 $ is equivalent to

    $ \begin{equation} \frac{1}{t^{2}}||\nabla v||^{2}_{2} = \int_{v\neq 0}\Big[\frac{h(x, G^{-1}(tv))}{g(G^{-1}(tv))t^{3}v^{3}} - V(x)\frac{G^{-1}(tv)}{g(G^{-1}(tv))t^{3}v^{3}}\Big]v^{4}dx. \end{equation} $ (2.9)

    By $ (H_{3}) $ and Lemma 2.1-(9), the right side of (2.9) is increasing for $ t > 0 $. So, there is a unique $ t_{v} > 0 $ such that $ f'(t_{v}) = 0 $.

    From Lemma 2.4, we can obtain the following lemma easily.

    Lemma 2.5. Assume that $ (V) $ and $ (H_{1})-(H_{4}) $ are satisfied. Then, the functional $ \Phi $ satisfies the following mountain pass geometry structure:

    (i) there exist positive constants $ \rho $ and $ b $ such that $ \Phi(v)\geq b $ for $ ||v|| = \rho $;

    (ii) there exists a function $ v_{0}\in X $ such that $ ||v_{0}|| > \rho $ and $ \Phi(v_{0}) < 0 $.

    Lemma 2.6. Assume that $ (V) $ and $ (H_{1})-(H_{4}) $ hold, then there exists a bounded Cerami sequence $ \{v_{n}\}\subset X $ for $ \Phi $.

    Proof. From Lemma 2.5, we know that $ \Phi $ satisfies the mountain pass geometry structure. Thus, the mountain pass theorem deduces that there exists a Cerami sequence $ \{v_{n}\}\subset X $ such that

    $ \begin{equation} \Phi(v_{n}) \rightarrow c \; \text{and}\; (1 + ||v_{n}||)||\Phi'(v_{n})||_{*}\rightarrow 0, \end{equation} $ (2.10)

    where $ c: = \inf_{v\in \mathcal{N}}\Phi(v) $.

    We claim that a Cerami sequence $ \{v_{n}\}\subset X $ satisfies

    $ \begin{equation} \int_{\mathbb{R}^{N}}|\nabla v_{n}|^{2}dx + \int_{\mathbb{R}^{N}}V(x)(G^{-1}(v_{n}))^{2}dx \leq M, \end{equation} $ (2.11)

    for some $ M > 0 $.

    From (2.10), we have

    $ \begin{equation} \Phi(v_{n}) = \frac{1}{2}\int_{\mathbb{R}^{N}}\Big[|\nabla v_{n}|^{2} + V(x)(G^{-1}(v_{n}))^{2}\Big]dx - \int_{\mathbb{R}^{N}}H(x, G^{-1}(v_{n}))dx \rightarrow c, \end{equation} $ (2.12)

    and

    $ \begin{equation} \langle \Phi'(v_{n}), v_{n} \rangle = \int_{\mathbb{R}^{N}}\Big[|\nabla v_{n}|^{2} + V(x)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}v_{n} -\frac{h(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}v_{n}\Big]dx = o_{n}(1)||v_{n}||. \end{equation} $ (2.13)

    From (2.1), (2.12), (2.13) and Lemma 2.1-(8), we obtain

    $ \begin{align*} c +1& \geq\Phi(v_{n}) - \frac{1}{4}\langle \Phi'(v_{n}), v_{n} \rangle\\ & = \frac{1}{4}\int_{\mathbb{R}^{N}}|\nabla v_{n}|^{2}dx + \frac{1}{4}\int_{\mathbb{R}^{N}}V(x)(G^{-1}(v_{n}))^{2}dx +\frac{1}{4}\int_{\mathbb{R}^{N}}V(x)\Big[(G^{-1}(v_{n}))^{2}- \frac{G^{-1}(v_{n})v_{n}}{g(G^{-1}(v_{n})}\Big] \\ &\quad + \int_{\mathbb{R}^{N}}\Big[\frac{h(x, G^{-1}(v_{n}))v_{n}}{4g(G^{-1}(v_{n}))} - H(x, G^{-1}(v_{n}))\Big]dx\\ &\geq \frac{1}{4}\int_{\mathbb{R}^{N}}|\nabla v_{n}|^{2}dx + \frac{1}{4}\int_{\mathbb{R}^{N}}V(x)(G^{-1}(v_{n}))^{2}dx, \end{align*} $

    that is

    $ \begin{equation} \int_{\mathbb{R}^{N}}|\nabla v_{n}|^{2} + \int_{\mathbb{R}^{N}}V(x)(G^{-1}(v_{n}))^{2}dx \leq 4(c+1) : = M. \end{equation} $ (2.14)

    Thus, (2.11) holds.

    Next, we claim that $ \int_{\mathbb{R}^{N}}v_{n}^{2}dx $ is bounded. In fact, by Lemma 2.1-(2) and (2.14), we have

    $ \begin{equation} \begin{aligned} \int_{|G^{-1}(v_{n})|\leq1}v_{n}^{2}dx &\leq g^{2}(1)\int_{|G^{-1}(v_{n})|\leq1}(G^{-1}(v_{n}))^{2}dx \\ &\leq \frac{C^{2}}{V_{0}}\int_{|G^{-1}(v_{n})|\leq1}V(x)(G^{-1}(v_{n}))^{2}dx \leq \frac{C^{2}}{V_{0}}M. \end{aligned} \end{equation} $ (2.15)

    Moreover, by the Sobolev inequality and (2.14), we deduce that

    $ \begin{equation} \begin{aligned} \int_{|G^{-1}(v_{n})| > 1}v_{n}^{2}dx &\leq \int_{|G^{-1}(v_{n})| > 1}v_{n}^{2^{*}}dx \leq C\Big(\int_{|G^{-1}(v_{n})| > 1}|\nabla v_{n}|^{2}dx\Big)^{\frac{2^{*}}{2}}\\ &\leq C\Big(\int_{\mathbb{R}^{N}}|\nabla v_{n}|^{2}dx\Big)^{\frac{2^{*}}{2}} \leq C M^{\frac{2^{*}}{2}}. \end{aligned} \end{equation} $ (2.16)

    Obviously, there is a constant $ C > 0 $ such that

    $ \begin{equation} \int_{\mathbb{R}^{N}}v_{n}^{2}dx = \int_{|G^{-1}(v_{n})|\leq1}v_{n}^{2}dx + \int_{|G^{-1}(v_{n})| > 1}v_{n}^{2}dx \leq C. \end{equation} $ (2.17)

    The claim is proved, then combining (2.14) with (2.17), $ \{v_{n}\}\subset X $ is bounded.

    In this section, before proving Theorem 1.1, we give three important lemmas to help us complete the proof of Theorem 1.1.

    Lemma 3.1. [18] Assume that $ (V) $ and $ (H_{1})-(H_{4}) $ are satisfied. If $ v\in \mathcal{N} $ and $ \Phi(v) = c $, then $ v $ is a solution of problem (1.1).

    Since $ V(x) $ is asymptotically periodic. In this case, the functional $ \Phi $ loses the translation invariance. The following two lemmas give careful estimates among $ V $ and $ V_{p} $, $ h $ and $ h_{p} $, $ H $ and $ H_{p} $, which is inspired by [18,37].

    Lemma 3.2. Assume that $ (V), (H_{1}), (H_{2}) $ and (1) of $ (H_{4}) $ hold. Suppose that $ \{v_{n}\} $ is bounded in $ X $ and $ v_{n}\rightarrow 0 $ in $ L_{loc}^{\gamma}(\mathbb{R}^{N}) $, for any $ \gamma\in [2, 2^{*}) $, then up to a subsequence, one has

    $ \begin{align} &( {\rm{i}}) \int_{\mathbb{R}^{N}}\Big(V(x) - V_{p}(x)\Big)(G^{-1}(v_{n}))^{2}dx = o_{n}(1). \end{align} $ (3.1)
    $ \begin{align} &( {\rm{ii}}) \int_{\mathbb{R}^{N}}\Big[H(x, G^{-1}(v_{n})) - H_{p}(x, G^{-1}(v_{n}))\Big]dx = o_{n}(1). \end{align} $ (3.2)

    Proof. (i) The proof of (3.1). Firstly, we assert that for any given $ k(x)\in \mathcal{F}_{0} $ and $ \varepsilon > 0 $, there exists $ R_{\varepsilon} > 0 $ such that

    $ \begin{equation} \int_{\{x: |k(x)|\geq \varepsilon\}}|v|^{2}dx \leq C\int_{B_{R_{\varepsilon}+1}(0)}|v|^{2}dx + C\varepsilon^{\frac{2}{N}}||v||_{H}^{2}, \; \; \forall v\in X, \end{equation} $ (3.3)

    where $ C $ is a constant and independent on $ \varepsilon $. (3.3) has already been proved in [18]. For the convenience of readers, we give a brief proof as following. According to the definition of $ \mathcal{F}_{0} $, for any $ \varepsilon > 0 $, there exists $ R_{\varepsilon} > 0 $ such that

    $ \begin{equation*} \text{meas}\{x\in B_{1}(y): |k(x)|\geq\varepsilon\} < \varepsilon, \; \; \forall |y|\geq R_{\varepsilon}. \end{equation*} $

    Covering $ \mathbb{R}^{N} $ by balls $ B_{1}(y_{i}), i\in \mathbb{N} $, where each point of $ \mathbb{R}^{N} $ is contained in at most $ N+1 $ balls. Without loss of generality, we assume that $ |y_{i}| < R_{\varepsilon} $, $ i = 1, 2, \cdot \cdot \cdot, n_{\varepsilon} $ and $ |y_{i}|\geq R_{\varepsilon} $, $ i = n_{\varepsilon}+1, n_{\varepsilon}+2, \cdot \cdot \cdot, +\infty $. By the H$ \ddot{o} $lder and Sobolev inequalities, we have

    $ \begin{equation*} \begin{aligned} \int_{\{x: |k(x)|\geq \varepsilon\}}|v|^{2}dx &\leq \sum^{+\infty}_{i = 1}\int_{\Omega_{i}}|v|^{2}dx\\ & = \sum^{n_{\varepsilon}}_{i = 1}\int_{\Omega_{i}}|v|^{2}dx + \sum^{+\infty}_{i = n_{\varepsilon}+1}\int_{\Omega_{i}}|v|^{2}dx\\ &\leq (N+1)\int_{\{x\in B_{R_{\varepsilon}+1}(0): k(x)\geq \varepsilon\}} |v|^{2}dx + \sum^{+\infty}_{i = n_{\varepsilon}+1} (\text{meas}\Omega_{i})^{\frac{2}{N}}\Big(\int_{\Omega_{i}}|v|^{2^{*}}dx\Big)^{\frac{N-2}{N}}\\ &\leq (N+1)\int_{B_{R_{\varepsilon}+1}(0)} |v|^{2}dx + C\varepsilon^{\frac{2}{N}}\sum^{+\infty}_{i = n_{\varepsilon}+1}\int_{\Omega_{i}}(|\nabla v|^{2}+ v^{2})dx\\ &\leq C\int_{B_{R_{\varepsilon}+1}(0)} |v|^{2}dx + C\varepsilon^{\frac{2}{N}}||v||^{2}_{H}, \end{aligned} \end{equation*} $

    where $ \Omega_{i} = \{ x\in B_{1}(y_{i}): |k(x)|\geq \varepsilon\} $.

    Set $ k(x): = V(x) - V_{p}(x)\in \mathcal{F}_{0} $, by Lemma 2.1-(6), we have

    $ \begin{align*} &\Big|\int_{\mathbb{R}^{N}}\Big(V(x) - V_{p}(x)\Big)(G^{-1}(v_{n}))^{2}dx\Big| \leq \int_{\mathbb{R}^{N}}|k(x)||G^{-1}(v_{n})|^{2}dx\\ &\leq \int_{\mathbb{R}^{N}}|k(x)v_{n}^{2}|dx = \int_{\{x: |k(x)|\geq \varepsilon\}}|k(x)v_{n}^{2}|dx + \int_{\{x: |k(x)| < \varepsilon\}}|k(x)v_{n}^{2}|dx\\ &\leq 2||V_{p}||_{\infty}\Big[C\int_{B_{R_{\varepsilon}+1}(0)}v_{n}^{2}dx + C\varepsilon^{\frac{2}{N}}||v_{n}||_{H}^{2}\Big] + \varepsilon \int_{\mathbb{R}^{N}}|v_{n}|^{2}dx\\ & = C\varepsilon^{\frac{2}{N}} + C\varepsilon + o_{n}(1). \end{align*} $

    Let $ \varepsilon\rightarrow 0 $, this completes the proof of (3.1).

    (ii) The proof of (3.2). Let $ \overline{h}(x, s): = h(x, s)- h_{p}(x, s) \in \mathcal{F} $. By the definition of $ \mathcal{F} $, for any $ \varepsilon > 0 $, there exists $ R_{\varepsilon} > 0 $ such that

    $ \begin{equation*} \text{meas}\; \{x\in B_{1}(y): |\bar{h}(x, s)|\geq \varepsilon \} < \varepsilon, \; \; \forall\; |y|\geq R_{\varepsilon}, \; |s|\leq \frac{1}{\varepsilon}, \end{equation*} $

    covering $ \mathbb{R}^{N} $ by balls $ B_{1}(y_{i}), i\in \mathbb{N} $, where each point of $ \mathbb{R}^{N} $ is contained in at most $ N+1 $ balls. Without loss of generality, we assume that $ |y_{i}| < R_{\varepsilon} $, $ i = 1, 2, \cdot \cdot \cdot, n_{\varepsilon} $ and $ |y_{i}|\geq R_{\varepsilon} $, $ i = n_{\varepsilon}+1, n_{\varepsilon}+2, \cdot \cdot \cdot, +\infty $. Using the mean value theorem and Lemma 2.1-(4), there exists $ t_{n}\in (0, 1) $ such that

    $ \begin{align*} H(x, G^{-1}(v_{n}))- H_{p}(x, G^{-1}(v_{n}))& = \Big[h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))\Big](G^{-1}(t_{n}v_{n}))'v_{n}\\ & = \frac{\Big[h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))\Big]v_{n}}{g(G^{-1}(t_{n}v_{n}))}. \end{align*} $

    Set

    $ \begin{equation*} \begin{aligned} \Omega_{1}&: = \{x\in B_{1}(y_{i}): |\overline{h}(x, G^{-1}(t_{n}v_{n}))| < \varepsilon\}, \\ \Omega_{2}&: = \{x\in B_{1}(y_{i}): |G^{-1}(t_{n}v_{n})|\leq \frac{1}{\varepsilon}, |\overline{h}(x, G^{-1}(t_{n}v_{n}))|\geq \varepsilon\}, \\ \Omega_{3}&: = \{x\in B_{1}(y_{i}): |G^{-1}(t_{n}v_{n})| > \frac{1}{\varepsilon}, |\overline{h}(x, G^{-1}(t_{n}v_{n}))|\geq \varepsilon\}. \end{aligned} \end{equation*} $

    Then, we obtain

    $ \begin{align*} &\Big|\int_{\mathbb{R}^{N}}\Big[H(x, G^{-1}(v_{n})) - [H_{p}(x, G^{-1}(v_{n}))\Big]dx\Big|\\ &\leq \sum^{+\infty}_{i = 1}\int_{B_{1}(y_{i})}\frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx\\ & = \sum^{n_{\varepsilon}}_{i = 1}\int_{B_{1}(y_{i})}\frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx \\ &\quad + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{B_{1}(y_{i})} \frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx\\ & = \sum^{n_{\varepsilon}}_{i = 1}\int_{B_{1}(y_{i})}\frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx \\ &\quad + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{1}}\frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx\\ &\quad + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{2}} \frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx \\ &\quad +\sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{3}} \frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx\\ &: = \Phi_{1} + \Phi_{2}+\Phi_{3}+\Phi_{4}. \end{align*} $

    From (2.4), we have

    $ \begin{equation*} \begin{aligned} \Phi_{1}& \leq (N+1)\int_{B_{R_{\varepsilon}+1}(0)}\frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n})}dx\\ &\leq (N+1)\int_{B_{R_{\varepsilon}+1}(0)}2(C_{\delta}|t_{n}v_{n}| + \delta |t_{n}v_{n}|^{2^{*}-1})|v_{n}|dx\\ & \leq 2(N+1)C_{\delta}\int_{B_{R_{\varepsilon}+1}(0)}|v_{n}|^{2}dx + 2(N+1)\delta \int_{B_{R_{\varepsilon}+1}(0)}|v_{n}|^{2^{*}}dx\\ & \leq C\delta + o_{n}(1). \end{aligned} \end{equation*} $

    Set

    $ \begin{equation*} \begin{aligned} \Omega_{11}&: = \{x\in B_{1}(y_{i}): |\overline{h}(x, G^{-1}(t_{n}v_{n}))| < \varepsilon, |G^{-1}(t_{n}v_{n})| \leq r_{\delta}\}, \\ \Omega_{12}&: = \{x\in B_{1}(y_{i}): |\overline{h}(x, G^{-1}(t_{n}v_{n}))| < \varepsilon, |G^{-1}(t_{n}v_{n})| > r_{\delta}\}.\\ \end{aligned} \end{equation*} $

    By (2.2), we obtain

    $ \begin{equation*} \begin{aligned} \Phi_{2}& = \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{11}} \frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx \\ &\quad + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{12}} \frac{|h(x, G^{-1}(t_{n}v_{n}))- h_{p}(x, G^{-1}(t_{n}v_{n}))||v_{n}|}{g(G^{-1}(t_{n}v_{n}))}dx\\ &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{11}} 2\delta|t_{n}v_{n}||v_{n}|dx + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{12}}\frac{\varepsilon}{r_{\delta}}|t_{n}v_{n}||v_{n}|dx\\ &\leq 2\delta \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{11}} |v_{n}|^{2}dx + \frac{\varepsilon}{r_{\delta}} \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{12}}|v_{n}|^{2}dx \\ &\leq 2(N+1)\delta \int_{\mathbb{R}^{N}} |v_{n}|^{2}dx + \frac{(N+1)\varepsilon}{r_{\delta}} \int_{\mathbb{R}^{N}} |v_{n}|^{2}dx\\ &\leq C\delta + C\varepsilon. \end{aligned} \end{equation*} $

    From (2.4), H$ \ddot{o} $lder and Sobolev inequalities, we get

    $ \begin{equation*} \begin{aligned} \Phi_{3}& \leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{2}}2\Big[(C_{\delta}|t_{n}v_{n}| + \delta |t_{n}v_{n}|^{2^{*}-1})v_{n}\Big]dx\\ &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \Big[2C_{\delta} \int_{\Omega_{2}}|v_{n}|^{2} dx + 2\delta \int_{\Omega_{2}}|v_{n}|^{2^{*}}dx\Big]\\ &\leq 2C_{\delta} \sum^{+\infty}_{i = n_{\varepsilon}+1}(\text{meas}\; \Omega_{2})^{\frac{2}{N}} \Big(\int_{\Omega_{2}} |v_{n}|^{2^{*}}dx\Big)^{\frac{N-2}{N}} + 2(N+1)\delta\int_{\mathbb{R}^{N}}|v_{n}|^{2^{*}}dx\\ &\leq 2C_{\delta}(N+1)\varepsilon^{\frac{2}{N}}\int_{\mathbb{R}^{N}}(|\nabla v_{n}|^{2}+ |v_{n}|^{2})dx + C\delta \\ & \leq C\varepsilon^{\frac{2}{N}} + C\delta. \end{aligned} \end{equation*} $

    Due to (2.5), one has

    $ \begin{equation*} \begin{aligned} \Phi_{4} &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{3}}2\Big[(\delta |t_{n}v_{n}|+ \delta |t_{n}v_{n}|^{2^{*}-1} + C_{\delta}|t_{n}v_{n}|^{\alpha-1})v_{n}\Big]dx\\ &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{3}}2\Big(\delta |v_{n}|^{2}+ \delta |v_{n}|^{2^{*}} + C_{\delta}|v_{n}|^{\alpha}\Big)dx\\ &\leq 2\delta(N+1) \int_{\mathbb{R}^{N}}(|v_{n}|^{2}+|v_{n}|^{2^{*}})dx + 2C_{\delta} \varepsilon^{2^{*}-\alpha} \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{3}}|v_{n}|^{2^{*}}dx\\ &\leq C \delta + C \varepsilon^{2^{*}-\alpha}. \end{aligned} \end{equation*} $

    Now, let $ \varepsilon\rightarrow 0 $ and then $ \delta\rightarrow 0 $, we have $ \Phi_{i}\rightarrow 0, i = 1, 2, 3, 4 $, the proof of (3.2) is completed.

    Lemma 3.3. Assume that $ (V), (H_{1}), (H_{2}) $ and (1) of $ (H_{4}) $ are satisfied. If $ \{v_{n}\} $ is bounded in $ X $ and $ |y_{n}|\rightarrow +\infty $. Then, for any $ \varphi\in C_{0}^{\infty}(\mathbb{R}^{N}) $, one has

    $ \begin{align} &( {\rm{i}}) \int_{\mathbb{R}^{N}}\Big(V(x) - V_{p}(x)\Big)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi(x-y_{n}) dx = o_{n}(1). \end{align} $ (3.4)
    $ \begin{align} &( {\rm{ii}}) \int_{\mathbb{R}^{N}}\frac{h(x, G^{-1}(v_{n})) - h_{p}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})dx = o_{n}(1). \end{align} $ (3.5)

    Proof. (i) The proof of (3.4). By Lemma 2.1-(3) and (6), we have

    $ \begin{equation} \Big|\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi\Big|\leq |G^{-1}(v_{n})\varphi|\leq |v_{n}\varphi|. \end{equation} $ (3.6)

    Due to $ \varphi\in C_{0}^{\infty}(\mathbb{R}^{N}) $, one has

    $ \begin{equation} \int_{B_{R_{\varepsilon}+1}(0)}|\varphi(x-y_{n})|^{2}dx = o_{n}(1). \end{equation} $ (3.7)

    Set $ k(x): = V(x) - V_{p}(x)\in \mathcal{F} $, in view of (3.3), (3.6), (3.7) and the H$ \ddot{o} $lder inequality, we have

    $ \begin{align*} &\Big|\int_{\mathbb{R}^{N}}\Big(V(x) - V_{p}(x)\Big)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi(x-y_{n}) dx \Big|\leq \int_{\mathbb{R}^{N}}\Big|\Big(V(x) - V_{p}(x)\Big)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big| dx\\ & = \int_{\{x:|k(x)|\geq\varepsilon\}}\Big|k(x)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx + \int_{\{x:|k(x)| < \varepsilon\}}\Big|k(x)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx\\ &\leq2||V_{p}||_{\infty}\int_{\{x:|k(x)|\geq\varepsilon\}}|v_{n}\varphi(x-y_{n})|dx + \varepsilon\int_{\{x:|k(x)| < \varepsilon\}}|v_{n}\varphi(x-y_{n})|dx\\ &\leq 2||V_{p}||_{\infty}||v_{n}||_{2}\Big(\int_{\{x:|k(x)|\geq \varepsilon\}}|\varphi(x- y_{n})|^{2}dx \Big)^{\frac{1}{2}} + \varepsilon ||v_{n}||_{2} ||\varphi||_{2}\\ &\leq 2||V_{p}||_{\infty}||v_{n}||_{2} \Big(C\int_{B_{R_{\varepsilon}+1}(0)}|\varphi(x- y_{n})|^{2}dx + C\varepsilon^{\frac{2}{N}}||\varphi||^{2}_{H} \Big)^{\frac{1}{2}}+ \varepsilon ||v_{n}||_{2} ||\varphi||_{2}\\ &\leq C\varepsilon^{\frac{1}{N}}+ C\varepsilon +o_{n}(1). \end{align*} $

    Let $ \varepsilon\rightarrow 0 $, (3.4) is proved.

    (ii) The proof of (3.5). Let $ \overline{h}(x, s): = h(x, s)- h_{p}(x, s)\in \mathcal{F} $. As the proof of Lemma 3.2, covering $ \mathbb{R}^{N} $ by balls $ B_{1}(y_{i}) $. Set

    $ \begin{equation*} \begin{aligned} \Omega_{4}&: = \{x\in B_{1}(y_{i}): |\overline{h}(x, G^{-1}(v_{n}))| < \varepsilon\}, \\ \Omega_{5}&: = \{x\in B_{1}(y_{i}): |G^{-1}(v_{n})|\leq \frac{1}{\varepsilon}, |\overline{h}(x, G^{-1}(v_{n}))|\geq \varepsilon\}, \\ \Omega_{6}&: = \{x\in B_{1}(y_{i}): |G^{-1}(v_{n})| > \frac{1}{\varepsilon}, |\overline{h}(x, G^{-1}(v_{n}))|\geq \varepsilon\}. \end{aligned} \end{equation*} $

    We obtain

    $ \begin{equation*} \begin{aligned} &\int_{\mathbb{R}^{N}} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx \\ &\leq \sum^{n_{\varepsilon}}_{i = 1}\int_{B_{1}(y_{i})} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{B_{1}(y_{i})} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx \\ & = \sum^{n_{\varepsilon}}_{i = 1}\int_{B_{1}(y_{i})} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{4}} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx\\ &\quad + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{5}} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{6}} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx\\ &: = \Phi_{5} + \Phi_{6} + \Phi_{7} + \Phi_{8}. \end{aligned} \end{equation*} $

    From (2.4) and (3.7), we have

    $ \begin{equation*} \begin{aligned} \Phi_{5} &\leq (N+1)\int_{B_{R_{\varepsilon}+1}(0)} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx\\ &\leq (N+1) \int_{B_{R_{\varepsilon}+1}(0)} 2(C_{\delta}|v_{n}| + \delta |v_{n}|^{2^{*}-1}) |\varphi(x-y_{n})|dx\\ &\leq 2(N+1)C_{\delta} \int_{B_{R_{\varepsilon}+1}(0)} |v_{n}||\varphi(x- y_{n})| dx + 2(N+1)\delta\int_{B_{R_{\varepsilon}+1}(0)} |v_{n}|^{2^{*}-1}|\varphi(x- y_{n})| dx\\ &\leq 2C_{\delta} ||v_{n}||_{2} \Big(\int_{B_{R_{\varepsilon}+1}(0)} |\varphi(x- y_{n})|^{2}\Big)^{\frac{1}{2}} + 2(N+1)\delta ||v_{n}||^{2^{*}-1}_{2^{*}}||\varphi||_{2^{*}}\\ &\leq C\delta +o_{n}(1). \end{aligned} \end{equation*} $

    On account of Lemma 2.1-(3), we obtain

    $ \begin{equation*} \begin{aligned} \Phi_{6} & = \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{4}} \Big|\frac{\overline{h}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi(x-y_{n})\Big|dx\\ &\leq \varepsilon \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{4}}|\varphi(x- y_{n})|dx\\ &\leq (N+1)\varepsilon \int_{\mathbb{R}^{N}}|\varphi(x- y_{n})|dx\\ &\leq C\varepsilon. \end{aligned} \end{equation*} $

    From (2.4), the H$ \ddot{o} $lder, Yong and Sobolev inequalities, one has

    $ \begin{equation*} \begin{aligned} \Phi_{7} &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{5}}2(C_{\delta}|v_{n}| + \delta|v_{n}|^{2^{*}-1})|\varphi(x-y_{n})|dx\\ &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{5}}2C_{\delta}|v_{n}\varphi(x- y_{n})|dx + \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{5}}2\delta|v_{n}|^{2^{*}-1} |\varphi(x- y_{n}|dx\\ &\leq 2 C_{\delta}\sum^{+\infty}_{i = n_{\varepsilon}+1}(\text{meas}\; \Omega_{5})^{\frac{2}{N}}\times \Big(\int_{\Omega_{5}} (|v_{n}\varphi(x- y_{n})| )^{{\frac{N}{N-2}} } dx \Big)^{ \frac{N-2}{N}} + 2(N+1)\delta ||v_{n}||^{2^{*}-1}_{2^{*}} ||\varphi||_{2^{*}}\\ &\leq 2 C_{\delta}\varepsilon^{\frac{2}{N}}\sum^{+\infty}_{i = n_{\varepsilon}+1}\Big(\int_{\Omega_{5}}(\frac{|v_{n}|^{2^{*}}}{2} + \frac{|\varphi(x- y_{n})|^{2^{*}}}{2} )dx \Big)^{\frac{N-2}{N}} + C\delta\\ &\leq 2 C_{\delta}\varepsilon^{\frac{2}{N}}\sum^{+\infty}_{i = n_{\varepsilon}+1}\Big[\Big(\frac{1}{2}\int_{\Omega_{5}}|v_{n}|^{2^{*}}dx \Big)^{\frac{N-2}{N}} + \Big(\frac{1}{2}\int_{\Omega_{5}}|\varphi(x- y_{n})|^{2^{*}}dx \Big)^{\frac{N-2}{N}}\Big] + C\delta \\ &\leq 2 C_{\delta}\varepsilon^{\frac{2}{N}}(N+1)(\frac{1}{2})^{\frac{N-2}{N}}C\Big[\int_{\mathbb{R}^{N}}(|\nabla v_{n}|^{2} +|v_{n}|^{2})dx + \int_{\mathbb{R}^{N}}(|\nabla \varphi(x- y_{n})|^{2} +|\varphi(x- y_{n})|^{2})dx \Big] + C\delta \\ & = C\varepsilon^{\frac{2}{N}}+ C\delta.\\ \end{aligned} \end{equation*} $

    By using (2.5) and H$ \ddot{o} $lder inequality, we have

    $ \begin{equation*} \begin{aligned} \Phi_{8} &\leq \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{6}}2\Big(\delta|v_{n}| + \delta|v_{n}|^{2^{*}-1} + C_{\delta}|v_{n}|^{\alpha-1}\Big) |\varphi(x- y_{n})|dx\\ &\leq 2\delta(N+1)\int_{\mathbb{R}^{N}}|v_{n}\varphi(x- y_{n})| )| dx + 2\delta(N+1)\int_{\mathbb{R}^{N}}|v_{n}|^{2^{*}-1} |\varphi(x- y_{n})|dx \\ & \quad + 2C_{\delta}\varepsilon^{2^{*}-\alpha} \sum^{+\infty}_{i = n_{\varepsilon}+1} \int_{\Omega_{6}} |v_{n}|^{2^{*}-1}|\varphi(x- y_{n})|dx\\ &\leq 2\delta (N+1)(||v_{n}||_{2}||\varphi||_{2} + ||v_{n}||^{2^{*}-1}_{2^{*}}||\varphi||_{2^{*}}) + 2(N+1)C_{\delta}\varepsilon^{2^{*}-\alpha}||v_{n}||^{2^{*}-1}_{2^{*}}||\varphi||_{2^{*}}\\ &\leq C\delta + C_{\delta}\varepsilon^{2^{*}-\alpha}. \end{aligned} \end{equation*} $

    As before, let $ \varepsilon\rightarrow 0 $ and then $ \delta\rightarrow 0 $, we get $ \Phi_{i}\rightarrow 0, i = 5, 6, 7, 8 $, we complete the proof of (3.5).

    Now we are ready to prove Theorem 1.1.

    Proof of Theorem 1.1. We can follow the energy comparison method in [15] to give the proof of Theorem 1.1. The functional $ \Phi $ is rewritten as

    $ \begin{equation*} \Phi(v) = \frac{1}{2}\int_{\mathbb{R}^{N}}\Big[|\nabla v|^{2} + V(x)(G^{-1}(v))^{2}\Big]dx - \int_{\mathbb{R}^{N}}H(x, G^{-1}(v))dx. \end{equation*} $

    We also need to consider the corresponding periodic functional $ \Phi_{p}: X\rightarrow \mathbb{R} $,

    $ \begin{equation*} \Phi_{p}(v) = \frac{1}{2}\int_{\mathbb{R}^{N}}\Big[|\nabla v|^{2} + V_{p}(x)(G^{-1}(v))^{2}\Big]dx - \int_{\mathbb{R}^{N}}H_{p}(x, G^{-1}(v))dx. \end{equation*} $

    By Lemma 2.5, there exists a Cerami sequence $ \{v_{n}\}\subset X $, such that

    $ \begin{equation*} \Phi(v_{n})\rightarrow c\; \text{and}\; (1 + ||v_{n}||)||\Phi'(v_{n})||_{*}\rightarrow 0. \end{equation*} $

    According to Lemma 2.6, the Cerami sequence $ \{v_{n}\} $ is bounded in $ X $. Thus, there exists $ v\in X $ such that $ v_{n}\rightharpoonup v $ in $ X $ and then $ \Phi'(v) = 0 $, that is $ v $ is a weak solution of problem (1.8). The proof of this result is standard, so we omit here.

    (i) The case $ v = 0 $. Then $ v_{n}\rightharpoonup 0 $ in $ X $, $ v_{n}\rightarrow 0 $ in $ L_{loc}^{\gamma}(\mathbb{R}^{N}), 2\leq \gamma < 2^{*} $ and $ v_{n}(x)\rightarrow 0 $ a.e. on $ \mathbb{R}^{N} $. From Lemma 3.2 and Lemma 3.3, we can deduce that

    $ \begin{equation} \begin{aligned} \Big|\Phi(v_{n}) - \Phi_{p}(v_{n})\Big|&\leq\frac{1}{2}\int_{\mathbb{R}^{N}}\Big|\Big(V(x)- V_{p}(x)\Big)(G^{-1}(v_{n}))^{2}\Big|dx \\ &\quad + \int_{\mathbb{R}^{N}}\Big|H(x, G^{-1}(v_{n})) - H_{p}(x, G^{-1}(v_{n}))\Big|dx = o_{n}(1), \end{aligned} \end{equation} $ (3.8)

    and taking $ \varphi \in X $ with $ ||\varphi|| = 1 $, we obtain that

    $ \begin{equation} \begin{aligned} \Big|\Big|\langle\Phi'(v_{n}) - \Phi'_{p}(v_{n}) \Big|\Big|_{*} &\leq \sup\limits_{\varphi \in X, ||\varphi|| = 1}\Big[\int_{\mathbb{R}^{N}}\Big|\Big(V(x)- V_{p}(x)\Big)\frac{G^{-1}(v_{n})}{g(G^{-1}(v_{n}))}\varphi \Big|dx\\ &\quad + \int_{\mathbb{R}^{N}}\Big|\frac{h(x, G^{-1}(v_{n}))- h_{p}(x, G^{-1}(v_{n}))}{g(G^{-1}(v_{n}))}\varphi \Big|dx\Big]\\ & = o_{n}(1). \end{aligned} \end{equation} $ (3.9)

    From (3.8) and (3.9), we can get that $ \{v_{n}\} $ is also a Cerami sequence for $ \Phi_{p} $. Namely,

    $ \begin{equation} \Phi_{p}(v_{n})\rightarrow c\; \text{and}\; (1 + ||v_{n}||)||\Phi'_{p}(v_{n})||_{*}\rightarrow 0. \end{equation} $ (3.10)

    Define

    $ \begin{equation*} \beta: = \underset{n\rightarrow \infty}{\lim\sup}\sup\limits_{y\in \mathbb{R}^{N}}\int_{B_{1}(y)}v_{n}^{2}dx, \end{equation*} $

    if $ \beta = 0 $, the Lions lemma [14], we have $ v_{n}\rightarrow 0 $ in $ L^{\alpha}(\mathbb{R}^{N}) $ for all $ \alpha\in (2, 2^{*}) $.

    From (2.5) and (2.7), we have

    $ \begin{equation} \lim\limits_{n\rightarrow \infty}\int_{\mathbb{R}^{N}}H(x, G^{-1}(v_{n}))dx = \lim\limits_{n\rightarrow \infty}\int_{\mathbb{R}^{N}}\frac{h(x, G^{-1}(v_{n}))v_{n}}{g(G^{-1}(v_{n})}dx = 0. \end{equation} $ (3.11)

    It is implied by Lemma 2.1-(8) that

    $ \begin{equation} 0\leq |G^{-1}(s)|^{2} - \frac{G^{-1}(s)s}{g(G^{-1}(s))}\rightarrow 0 (s\rightarrow 0). \end{equation} $ (3.12)

    Then, combining with (3.11) and (3.12), we obtain

    $ \begin{align*} c& = \Phi(v_{n}) - \frac{1}{2}\langle \Phi'(v_{n}), v_{n} \rangle + o_{n}(1)\\ & = \frac{1}{2}\int_{\mathbb{R}^{N}}V(x)\Big[(G^{-1}(v_{n}))^{2}- \frac{G^{-1}(v_{n})v_{n}}{g(G^{-1}(v_{n})}\Big]dx + \int_{\mathbb{R}^{N}}\Big[\frac{h(x, G^{-1}(v_{n}))v_{n}}{2g(G^{-1}(v_{n}))} - H(x, G^{-1}(v_{n}))\Big]dx + o_{n}(1)\\ &\rightarrow 0. \end{align*} $

    This contradiction shows $ \beta > 0 $. Up to a subsequence, there exist a sequence $ \{y_{n}\}\subset \mathbb{Z}^{N} $ and $ r > 0 $ such that $ |y_{n}|\rightarrow \infty $,

    $ \begin{equation} \int_{B_{r(0)}}w_{n}^{2}dx = \int_{B_{r(y_{n})}}v_{n}^{2}dx\geq \frac{\beta}{2} > 0, \end{equation} $ (3.13)

    where $ w_{n}(x): = v_{n}(x + y_{n}) $. Since $ ||w_{n}|| = ||v_{n}|| $, we see that $ \{w_{n}\} $ is bounded. Going if necessary to a subsequence, we have $ w_{n}\rightharpoonup w $ in $ X $, $ w_{n}\rightarrow w $ in $ L_{loc}^{\gamma}(\mathbb{R}^{N}), 2\leq \gamma < 2^{*} $ and $ w_{n}(x)\rightarrow w(x) $ a.e. on $ \mathbb{R}^{N} $. Thus, (3.13) implies that $ w\neq 0 $. For any $ \varphi\in C^{\infty}_{0}(\mathbb{R}^{N}) $, we have

    $ \begin{equation*} \langle \Phi'_{p}(w), \varphi\rangle = \lim\limits_{n\rightarrow \infty}\langle \Phi'_{p}(w_{n}), \varphi\rangle = 0. \end{equation*} $

    Hence $ \Phi'_{p}(w) = 0 $. Next, our task is to verify that $ \Phi_{p}(w)\leq c $. Since

    $ \begin{equation} ||w_{n}|| = ||v_{n}||, \; \; \Phi_{p}(v_{n}) = \Phi_{p}(w_{n}), \; \; \Phi'_{p}(v_{n}) = \Phi'_{p}(w_{n}). \end{equation} $ (3.14)

    From (2.1), (3.10), (3.14), Lemma 2.1-(8), (1) of $ (H_{4}) $ and the Fatou lemma, we obtain

    $ \begin{equation} \begin{aligned} c& = \underset{n\rightarrow \infty}{\lim\inf}\Big[\Phi_{p}(w_{n}) - \frac{1}{4}\langle \Phi'_{p}(w_{n}), w_{n} \rangle\Big]\\ & = \underset{n\rightarrow \infty}{\lim\inf}\frac{1}{4}\int_{\mathbb{R}^{N}}\Big[|\nabla w_{n}|^{2} + V_{p}(x)(G^{-1}(w_{n}))^{2}\Big]dx +\underset{n\rightarrow \infty}{\lim\inf}\frac{1}{4}\int_{\mathbb{R}^{N}}V_{p}(x)\Big[(G^{-1}(w_{n}))^{2}- \frac{G^{-1}(w_{n})w_{n}}{g(G^{-1}(w_{n})}\Big]dx \\ &\quad + \underset{n\rightarrow \infty}{\lim\inf}\int_{\mathbb{R}^{N}}\Big[\frac{h_{p}(x, G^{-1}(w_{n}))w_{n}}{4g(G^{-1}(w_{n}))} - H_{p}(x, G^{-1}(w_{n}))\Big]dx\\ &\geq\frac{1}{4}\int_{\mathbb{R}^{N}}\Big[|\nabla w|^{2} + V_{p}(x)(G^{-1}(w))^{2}\Big]dx + \frac{1}{4}\int_{\mathbb{R}^{N}}V_{p}(x)\Big[(G^{-1}(w))^{2}- \frac{G^{-1}(w)w}{g(G^{-1}(w)}\Big]dx \\ &\quad + \int_{\mathbb{R}^{N}}\Big[\frac{h_{p}(x, G^{-1}(w))w_{n}}{4g(G^{-1}(w))} - H_{p}(x, G^{-1}(w))\Big]dx\\ & = \Phi_{p}(w) - \frac{1}{4}\langle\Phi'_{p}(w), w\rangle = \Phi_{p}(w). \end{aligned} \end{equation} $ (3.15)

    By (3.15), (1) of $ (H_{4}) $ and the definition of $ c $

    $ \begin{equation*} c\leq \Phi(t_{w}w)\leq \Phi_{p}(t_{w}w)\leq\Phi_{p}(w)\leq c. \end{equation*} $

    This implies that $ \Phi(t_{w}w) = c $. Let $ w_{0} = t_{w}w $, then $ w_{0}\in \mathcal{N} $ and $ \Phi(w_{0}) = c $. In view of Lemma 3.1, $ \Phi'(w_{0}) = 0 $. This shows that $ w_{0} $ is a ground state solution of problem (1.8).

    (ii) The case $ v\neq 0 $. Since $ v $ is a weak solution of problem (1.8), $ \Phi(v)\geq c $. It follows from (2.1), Lemma 2.1-(8) and Fatou's lemma that

    $ \begin{align*} c& = \underset{n\rightarrow \infty}{\lim\inf}\Big[\Phi(v_{n}) - \frac{1}{4}\langle \Phi'(v_{n}), v_{n} \rangle\Big]\\ & = \underset{n\rightarrow \infty}{\lim\inf}\frac{1}{4}\int_{\mathbb{R}^{N}}\Big[|\nabla v_{n}|^{2} + V(x)(G^{-1}(v_{n}))^{2}\Big]dx +\underset{n\rightarrow \infty}{\lim\inf}\frac{1}{4}\int_{\mathbb{R}^{N}}V(x)\Big[(G^{-1}(v_{n}))^{2}- \frac{G^{-1}(v_{n})v_{n}}{g(G^{-1}(v_{n})}\Big]dx \\ &\quad + \underset{n\rightarrow \infty}{\lim\inf}\int_{\mathbb{R}^{N}}\Big[\frac{h(x, G^{-1}(v_{n}))v_{n}}{4g(G^{-1}(v_{n}))} - H(x, G^{-1}(v_{n}))\Big]dx\\ &\geq\frac{1}{4}\int_{\mathbb{R}^{N}}\Big[|\nabla v|^{2} + V(x)(G^{-1}(v))^{2}\Big]dx + \frac{1}{4}\int_{\mathbb{R}^{N}}V(x)\Big[(G^{-1}(v))^{2}- \frac{G^{-1}(v)v}{g(G^{-1}(v)}\Big]dx \\ &\quad + \int_{\mathbb{R}^{N}}\Big[\frac{h(x, G^{-1}(v))v}{4g(G^{-1}(v))} - H(x, G^{-1}(v))\Big]dx\\ & = \Phi(v) - \frac{1}{4}\langle\Phi'(v), v\rangle = \Phi(v). \end{align*} $

    This shows that $ \Phi(v) = c $ and $ \Phi'(v) = 0 $, which implies that $ v $ is a ground state solution of problem (1.8).

    From (i) and (ii), we can obtain that Eq (1.8) has a ground state solution $ v\in X $. By using the strong maximum principle [36], we can get that $ v $ is a positive ground state solution of (1.8). Namely, Eq (1.1) possesses a positive ground state solution $ u = G^{-1}(v) $ and the proof is completed.

    In this paper, we have established the existence of a positive ground state solution for Eqs (1.1) and (1.10) by the variational method. In comparison with previous works, this paper has several new features. Firstly, we consider a more general model than the literature [18,34,37]. Secondly, compared to the literature [34], we give a new asymptotic process of potential and nonlinearity term. Finally, we choose the more general nonlinear term than Ambrosetti-Rabinowitz condition. Therefore, to some extent, we have improved and extended the results of the existing literature.

    The author expresses his appreciation to the reviewers and the handling editor whose careful reading of the manuscript and valuable comments greatly improved the original manuscript.

    The author declares that he has no competing interests concerning the publication of this article.



    [1] A. V. Borovskii, A. L. Galkin, Dynamical modulation of an ultrashort high-intensity laser pulse in matter, J. Exp. Theor. Phys., 77 (1993), 562–573. doi: 1063-7761/93/10056-12fanxiexian_myfh10.00.
    [2] A. D. Bouard, N. Hayashi, J. Saut, Global existence of small solutions to a relativistic nonlinear Schrödinger equation, Commun. Math. Phys., 189 (1997), 73–105. doi: 10.1007/s002200050191. doi: 10.1007/s002200050191
    [3] M. Colin, L. Jeanjean, Solutions for a quasilinear Schrödinger equation: A dual approach, Nonlinear Anal., 56 (2004), 213–226. doi: 10.1016/j.na.2003.09.008. doi: 10.1016/j.na.2003.09.008
    [4] C. M. Chu, H. D. Liu, Existence of positive solutions for a quasilinear Schrödinger equation, Nonlinear Anal-Real., 44 (2018), 118–127. doi: 10.1016/j.nonrwa.2018.04.007. doi: 10.1016/j.nonrwa.2018.04.007
    [5] X. L. Chen, R. N. Sudan, Necessary and sufficient conditions for self-focusing of short ultraintense laser pulse in underdense plasma, Phys. Rev. Lett., 70 (1993), 2082–2085. doi: 10.1103/PhysRevLett.70.2082. doi: 10.1103/PhysRevLett.70.2082
    [6] J. H. Chen, X. H. Tang, B. T. Cheng, Non-Nehari manifold for a class of generalized quasilinear Schrödinger equations, Appl. Math. Lett., 74 (2017), 20–26. doi: 10.1016/j.aml.2017.04.032. doi: 10.1016/j.aml.2017.04.032
    [7] J. H. Chen, X. H. Tang, B. T. Cheng, Ground states for a class of generalized quasilinear Schrödinger equations in $\mathbb{R}^{N}$, Mediterr J. Math., 14 (2017), 190. doi: 10.1007/s00009-017-0990-y. doi: 10.1007/s00009-017-0990-y
    [8] J. H. Chen, X. H. Tang, B. Cheng, Existence of ground state sign-changing solutions for a class of generalized quasilinear Schrödinger-Maxwell system in $\mathbb{R}^{3}$, Comput. Math. Appl., 74 (2017), 466–481. doi: 10.1016/j.camwa.2017.04.028. doi: 10.1016/j.camwa.2017.04.028
    [9] Y. B. Deng, W. T. Huang, Positive ground state solutions for a quasilinear elliptic equation with critical exponent, Discrete Cont. Dyn-A., 37 (2017), 4213–4230. doi: 10.3934/dcds.2017179. doi: 10.3934/dcds.2017179
    [10] Y. Deng, S. Peng, S. Yan, Positive solition solutions for generalized quasilinear Schrödinger equations with critical growth, J. Differ. Equations, 258 (2015), 115–147. doi: 10.1016/j.jde.2014.09.006. doi: 10.1016/j.jde.2014.09.006
    [11] Y. Deng, S. Peng, S. Yan, Critical exponents and solitary wave solutions for generalized quasilinear Schrödinger equations, J. Differ. Equations, 260 (2016), 1228–1262. doi: 10.1016/j.jde.2015.09.021. doi: 10.1016/j.jde.2015.09.021
    [12] W. T. Huang, J. L. Xiang, Solition solutions for a quasilinear Schrödinger equation with critical exponent, Commun. Pur. Appl. Anal., 15 (2016), 1309–1333. doi: 10.3934/cpaa.2016.15.1309. doi: 10.3934/cpaa.2016.15.1309
    [13] S. Kurihura, Large-amplitude quasi-solitions in suerfluid films, J. Phys. Soc. Japan, 50 (1981), 3262–3267. doi: 10.1143/JPSJ.50.3262. doi: 10.1143/JPSJ.50.3262
    [14] P. L. Lions, The concentration-compactness principle in the calculus of variations. The locally compact case part 1, Ann. I. H. Poincare-Anal., 1 (1984), 109–145. doi: 10.1016/S0294-1449(16)30428-0. doi: 10.1016/S0294-1449(16)30428-0
    [15] S. B. Liu, On ground states of superlinear p-Laplacian euqations in $\mathbb{R}^{N}$, J. Math. Anal. Appl., 361 (2010), 48–58. doi: 10.1016/j.jmaa.2009.09.016. doi: 10.1016/j.jmaa.2009.09.016
    [16] Z. P. Liang, J. F. Gao, A. R. Li, Infinitely many solutions to a quasilinear Schrödinger equation with a local sublinear term, Appl. Math. Lett., 89 (2019), 22–27. doi: 10.1016/j.aml.2018.09.015. doi: 10.1016/j.aml.2018.09.015
    [17] J. Liu, J. F. Liao, C. L. Tang, A positive ground state solution for a class of asymptotically periodic Schrödinger equations, Comput. Math. Appl., 71 (2016), 965–976. doi: 10.1016/j.camwa.2016.01.004. doi: 10.1016/j.camwa.2016.01.004
    [18] J. J. Liu, J. F. Liao, C. L. Tang, A positive ground state solution for a class of asymptotically periodic Schrödinger equations with critical exponent, Comput. Math. Appl., 72 (2016), 1851–1864. doi: 10.1016/j.camwa.2016.08.010. doi: 10.1016/j.camwa.2016.08.010
    [19] X. Q. Liu, J. Q. Liu, Z. Q. Wang, Quasilinear elliptic equations via perturbation method, P. Am. Math. Soc., 141 (2013), 253–263. doi: 10.1090/S0002-9939-2012-11293-6. doi: 10.1090/S0002-9939-2012-11293-6
    [20] H. Lange, M. Poppenberg, H. Teismann, Nash-Moser methods for the solution of quasilinear Schrödinger equations, Commun. Part. Diff. Eq., 24 (1999), 1399–1418. doi: 10.1080/03605309908821469. doi: 10.1080/03605309908821469
    [21] E. Laedke, K. Spatschek, L. Stenflo, Evolution theorem for a class of perturbed envelope soliton solutions, J. Math. Phys., 24 (1983), 2764–2769. doi: 10.1063/1.525675. doi: 10.1063/1.525675
    [22] Q. Li, K. Teng, X. Wu, Ground state solutions and geometrically distinct solutions for generalized quasilinear Schrödinger equation, Math. Methods Appl. Sci., 40 (2017), 2165–2176. doi: 10.1002/mma.4131. doi: 10.1002/mma.4131
    [23] J. Q. Liu, Y. Q. Wang, Z. Q. Wang, Soliton solutions for quasilinear Schrödinger equations, Ⅱ, J. Differ. Equations, 187 (2003), 473–493. doi: 10.1016/S0022-0396(02)00064-5. doi: 10.1016/S0022-0396(02)00064-5
    [24] Q. Q. Li, X. Wu, Multiple solutions for generalized quasilinear Schrödinger equations, Math. Method. Appl. Sci., 40 (2017), 1359–1366. doi: 10.1002/mma.4050. doi: 10.1002/mma.4050
    [25] Y. Y. Li, Y. F. Xue, C. L. Tang, Ground state solutions for asymptotically periodic modified Schrödinger-Poisson system involving critical exponent, Commun. Pur. Appl. Anal., 18 (2019), 2299–2324. doi: 10.3934/cpaa.2019104. doi: 10.3934/cpaa.2019104
    [26] F. Li, X. Zhu, Z. Liang, Multiple solutions to a class of generalized quasilinear Schrödinger equations with a Kirchhoff-type perturbation, J. Math. Anal. Appl., 443 (2016), 11–38. doi: 10.1016/j.jmaa.2016.05.005. doi: 10.1016/j.jmaa.2016.05.005
    [27] A. Moameni, Existence of soliton solutions for a quasilinear Schrödinger equation involving critical exponent in $\mathbb{R}^{N}$, J. Differ. Equations, 229 (2006), 570–587. doi: 10.1016/j.jde.2006.07.001. doi: 10.1016/j.jde.2006.07.001
    [28] J. C. Oliveira Junior, S. I. Moreira, Generalized quasilinear equations with sign changing unbounded potential, Appl. Anal., 2020, 1–18. doi: 10.1080/00036811.2020.1836356. doi: 10.1080/00036811.2020.1836356
    [29] M. Poppenberg, K. Schmitt, Z. Q. Wang, On the existence of soliton solutions to quasilinear Schrödinger equations, Calc. Var. Partial Dif., 14 (2002), 329–344. doi: 10.1007/s005260100105. doi: 10.1007/s005260100105
    [30] B. Ritchie, Relativistic self-focusing and channel formation in laser-plasma interactions, Phys. Rev. E, 50 (1994), 687–689. doi: 10.1103/PhysRevE.50.R687. doi: 10.1103/PhysRevE.50.R687
    [31] H. X. Shi, H. B. Chen, Positive solutions for generalized quasilinear Schrödinger equations with general vanishing at infinity, Appl. Math. Lett., 61 (2016), 137–142. doi: 10.1016/j.aml.2016.06.004. doi: 10.1016/j.aml.2016.06.004
    [32] T. T. Shang, R. X. Liang, Ground state solutions for a quasilinear elliptic equation with general critical nonlinearity, Complex Var. Elliptic, 66 (2021), 586–613. doi: 10.1080/17476933.2020.1731736. doi: 10.1080/17476933.2020.1731736
    [33] E. A. B. Silva, G. F. Vieira, Quasilinear asymptotically periodic Schrödinger equations with critical growth, Calc. Var. Partial Dif., 39 (2010), 1–33. doi: 10.1007/s00526-009-0299-1. doi: 10.1007/s00526-009-0299-1
    [34] E. A. B. Silva, G. F. Vieira, Quasilinear asymptotically periodic Schrödinger equations with subcritical growth, Nonlinear Anal., 72 (2010), 2935–2949. doi: 10.1016/j.na.2009.11.037. doi: 10.1016/j.na.2009.11.037
    [35] Y. Shen, Y. Wang, Soliton solutions for generalized quasilinear Schrödinger equations, Nonlinear Anal-Theor., 80 (2013), 194–201. doi: 10.1016/j.na.2012.10.005. doi: 10.1016/j.na.2012.10.005
    [36] J. L. Vazquez, A strong maximum principle for some quasilinear elliptic equations, Appl. Math. Opt., 12 (1984), 191–202. doi: 10.1007/BF01449041. doi: 10.1007/BF01449041
    [37] Y. F. Xue, J. Liu, C. L. Tang, A ground state solution for an asymptotically periodiic quasilinear Schrödinger equation, Comput. Math. Appl., 74 (2017), 1143–1157. doi: 10.1016/j.camwa.2017.05.033. doi: 10.1016/j.camwa.2017.05.033
    [38] Y. F. Xue, C. L. Tang, Ground state solutions for asymptotically periodic quasilinear Schrödinger equations with critical growth, Commun. Pur. Appl. Anal., 17 (2018), 1121–1145. doi: 10.3934/cpaa.2018054. doi: 10.3934/cpaa.2018054
    [39] J. Yang, Y. Wang, A. A. Abdelgadir, Soliton solutions for quasilinear Schrödinger equation, J. Math. Phys., 54 (2013), 071502. doi: 10.1063/1.4811394. doi: 10.1063/1.4811394
    [40] X. Zhu, F. Li, Z. Liang, Existence of ground state solutions to a generalized quasilinear Schrödinger-Maxwell system, J. Math. Phys., 57 (2016), 101505. doi: 10.1063/1.4965442. doi: 10.1063/1.4965442
  • This article has been cited by:

    1. Shulin Zhang, Existence of nontrivial positive solutions for generalized quasilinear elliptic equations with critical exponent, 2022, 7, 2473-6988, 9748, 10.3934/math.2022543
    2. Joshua Warigue Ndiaye, Iterative reconstruction algorithms for solving the Schrödinger equations on conical spaces, 2022, 12, 1664-2368, 10.1007/s13324-021-00645-7
    3. Jiameng Li, Huiwen Chen, Zhimin He, Zigen Ouyang, Infinitely many solutions for quasilinear Schrödinger equation with general superlinear nonlinearity, 2023, 2023, 1687-2770, 10.1186/s13661-023-01755-w
    4. Liang Xue, Jiafa Xu, Donal O'Regan, Positive solutions for a critical quasilinear Schrödinger equation, 2023, 8, 2473-6988, 19566, 10.3934/math.2023998
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(2425) PDF downloads(109) Cited by(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog