Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-24T07:52:30.197Z Has data issue: false hasContentIssue false

Groundstates of the planar Schrödinger–Poisson system with potential well and lack of symmetry

Published online by Cambridge University Press:  22 May 2023

Zhisu Liu
Affiliation:
Center for Mathematical Sciences/School of Mathematics and Physics, China University of Geosciences, Wuhan, Hubei 430074, People's Republic of China liuzhisu@cug.edu.cn
Vicenţiu D. Rădulescu
Affiliation:
Faculty of Applied Mathematics, AGH University of Science and Technology, 30-059 Kraków, Poland Faculty of Electrical Engineering and Communication, Brno University of Technology, Technická 3058/10, Brno 61600, Czech Republic Department of Mathematics, University of Craiova, Street A.I. Cuza 13, 200585 Craiova, Romania Simion Stoilow Institute of Mathematics of the Romanian Academy, P.O. Box 1-764, 014700 Bucharest, Romania School of Mathematics, Zhejiang Normal University, Jinhua 321004, Zhejiang, People's Republic of China radulescu@inf.ucv.ro
Jianjun Zhang
Affiliation:
College of Mathematics and Statistics, Chongqing Jiaotong University, Chongqing 400074, People's Republic of China zhangjianjun09@tsinghua.org.cn
Rights & Permissions [Opens in a new window]

Abstract

The Schrödinger–Poisson system describes standing waves for the nonlinear Schrödinger equation interacting with the electrostatic field. In this paper, we are concerned with the existence of positive ground states to the planar Schrödinger–Poisson system with a nonlinearity having either a subcritical or a critical exponential growth in the sense of Trudinger–Moser. A feature of this paper is that neither the finite steep potential nor the reaction satisfies any symmetry or periodicity hypotheses. The analysis developed in this paper seems to be the first attempt in the study of planar Schrödinger–Poisson systems with lack of symmetry.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s), 2023. Published by Cambridge University Press on behalf of The Royal Society of Edinburgh

1. Introduction

This paper deals with the qualitative analysis of solutions to Schrödinger–Poisson systems of the type

(1.1)\begin{equation} \left\{\begin{array}{@{}ll} i\psi_t -\Delta\psi+E(x)\psi +\nu\phi \psi=0, & (x,t)\in\mathbb{R}^2\times\mathbb{R}, \\ \Delta \phi=|\psi|^2, & (x,t)\in\mathbb{R}^2\times\mathbb{R}, \end{array}\right. \end{equation}

where $\psi :\mathbb {R}^2\times \mathbb {R}\rightarrow {\mathbb {C}}$ is the wave function, $E:\mathbb {R}^2\rightarrow \mathbb {R}$ is an external potential and $\nu$ is a real parameter. The function $\phi$ represents an internal potential for a nonlocal self-interaction of the wave function $\psi$.

By the standing wave ansatz $\psi (x,t)={\rm e}^{-i\lambda t}u(x)$ (with $\lambda \in \mathbb {R}$), problem (1.1) reduces to the following stationary planar Schrödinger–Poisson system:

(1.2)\begin{equation} \left\{\begin{array}{@{}ll} -\Delta u+V(x)u +\nu\phi u=0, & x\in\mathbb{R}^2, \\ \Delta \phi=u^2, & x\in\mathbb{R}^2, \end{array}\right. \end{equation}

where $V(x)=E(x)+\lambda$. In some recent works, local nonlinear terms of the form $f(u)$ have been added to the right-hand side of the first equation in (1.2). In this case, the nonlinear term $f(u)$ describes the interaction effect among particles; see Benci and Fortunato [Reference Benci and Fortunato9]. We shall be concerned in this paper with the case where $f$ is a continuous function with exponential critical or subcritical growth in the Trudinger–Moser sense.

The analysis developed in this paper is performed in the case where $V(x)$ is a finite steep potential well. This is a variation on the infinite potential well, in which a particle is trapped in a ‘box’ with limited potential ‘walls’. Unlike the infinite potential well, there is a probability that the particle will be detected outside the box. The quantum mechanical interpretation contrasts from the classical interpretation in that the particle cannot be detected outside the box if its total energy is less than the potential energy barrier of the walls. Even when the particle's energy is less than the potential energy barrier of the walls, there is a non-zero probability of the particle surviving outside the box according to the quantum interpretation.

During the last few decades, quantum modelling of semiconductors has become a very active area of research. The (local or nonlocal) Schrödinger–Poisson system explains the thermodynamical and electrostatic equilibrium of electrons trapped in tiny quantum wells. In reality, the interaction of a charge particle with an electromagnetic field can be characterized by coupling the nonlinear Schrödinger's and Poisson's equations, according to a classical model. Physicists proposed the Schrödinger–Poisson system to quantify the precise energy levels of electrons in semiconductor heterostructures; see Nier [Reference Nier39].

The first equation in system (1.2) is referred as the Schrödinger equation while the second equation in (1.2) is known as the Poisson equation. The Schrödinger equation plays the role of Newton's laws and conservation of energy in classical mechanics, that is, it predicts the future behaviour of a dynamic system. The linear Schrödinger equation is a central tool of quantum mechanics, which provides a thorough description of a particle in a non-relativistic setting.

In recent decades, a considerable amount of research has been conducted on the nature and behaviour of solutions to the Schrödinger–Poisson system. This is due in part to the fact that this class of nonlinear problems contains a large number of fundamental physical models including, for instance the interaction of a charge particle with the electrostatic field in quantum mechanics. In this case, the unknowns $u$ and $\phi$ represent the wave functions associated with the particle and the electric potential, respectively. Related applications include the study of obstacle problem, the seepage surface problem or Elenbaas's equation. In astrophysics, the Schrödinger–Poisson system has been suggested to model certain theoretical concepts; see Schunck and Mielke [Reference Schunck and Mielke45]. Self-gravitating boson stars, for example, may be a source of exotic laser interferometer gravitational-wave observatory detections in addition to the predicted gravitational wave merger signals of black hole and neutron star binary systems; see Sennett et al. [Reference Sennett, Hinderer, Steinhoff, Buonanno and Ossokine46].

The structure of the nonlinear Schrödinger equation is much more complicated. This equation is a prototypical dispersive nonlinear partial differential equation that has been central for almost four decades now to a variety of areas in mathematical physics. The relevant fields of application vary from Bose–Einstein condensates and nonlinear optics (see Byeon and Wang [Reference Byeon and Wang10]), propagation of the electric field in optical fibers (see Malomed [Reference Malomed36]) to the self-focusing and collapse of Langmuir waves in plasma physics (see Zakharov [Reference Zakharov, Galeev and Sudan56]) and the behaviour of deep water waves and freak waves (the so-called rogue waves) in the ocean (see Onorato et al. [Reference Onorato, Osborne, Serio and Bertone41]). The nonlinear Schrödinger equation also describes various phenomena arising in the theory of Heisenberg ferromagnets and magnons, self-channelling of a high-power ultra-short laser in matter, condensed matter theory, dissipative quantum mechanics, electromagnetic fields (see Avron et al. [Reference Avron, Herbst and Simon6]), plasma physics (e.g., the Kurihara superfluid film equation). We refer to Sulem and Sulem [Reference Sulem and Sulem49] for a modern overview, including applications.

Schrödinger also established the classical derivation of his equation, based upon the analogy between mechanics and optics, and closer to de Broglie's ideas. Schrödinger developed a perturbation method, inspired by the work of Lord Rayleigh in acoustics, and he proved the equivalence between his wave mechanics and Heisenberg's matrix. The importance of Schrödinger's perturbation method was pointed out by Einstein [Reference Einstein27], who wrote: ‘The Schrödinger method, which has in a certain sense the character of a field theory, does indeed deduce the existence of only discrete states, in surprising agreement with empirical facts. It does so on the basis of differential equations applying a kind of resonance argument.’

In the literature, in order to overcome the lack of compactness, such problems have been investigated under periodicity assumptions or various symmetry hypotheses, namely, either in radially symmetric spaces or in axially symmetry spaces. For instance, Chen and Tang [Reference Chen and Tang22] proposed a new approach to recover the compactness for Cerami sequences, while Tang [Reference Tang51] developed a direct method (the non-Nehari manifold method) to find a minimizing Cerami sequence for the energy functional outside the Nehari–Pankov manifold by using the diagonal method. We also refer to Albuquerque et al. [Reference Albuquerque, Carvalho, Figueiredo and Medeiros3], Alves et al. [Reference Alves, Cassani, Tarsi and Yang4], Chen and Tang [Reference Chen and Tang20, Reference Chen and Tang21], Sun and Ma [Reference Sun and Ma50], Wen et al. [Reference Wen, Chen and Rădulescu55], etc. In a nonlocal setting, Tang and Cheng [Reference Tang and Cheng52] introduced an original method to recover the compactness of Palais–Smale sequences. A key of the present paper is that in our framework, the compactness is recovered by virtue of the Trudinger–Moser inequality, see Trudinger [Reference Trudinger53] and Moser [Reference Moser38].

The main novelty in the present paper is that we establish the existence of groundstates under general hypotheses and without assuming any symmetry or periodicity restrictions neither for the steep potential nor for the reaction.

1.1 Overview and historical comments

In the present paper, we are concerned with the following planar Schrödinger–Poisson system

(1.3)\begin{equation} \left\{\begin{array}{@{}ll} -\Delta u+V(x)u+\nu\phi u=f(u), & x\in\mathbb{R}^2, \\ \Delta \phi=u^2, & x\in\mathbb{R}^2, \end{array}\right. \end{equation}

where $\nu \in \mathbb {R}$ and $f\in C(\mathbb {R},\mathbb {R})$. We assume that $V\in C(\mathbb {R}^2,\mathbb {R})$ satisfies the following hypotheses:

  1. (V1) $V$ is weakly differentiable and satisfies $(\nabla V(x), x)\in L^{\infty }(\mathbb {R}^3)\cup L^{\kappa }(\mathbb {R}^3)$ for $\kappa >1$ and

    \[ V(x)+\frac{1}{2}(\nabla V(x), x)\geq0,\quad\text{for all}\ x\in\mathbb{R}^2, \]
    where $(\cdot,\cdot )$ is the usual inner product in $\mathbb {R}^2$;
  2. (V2) for all $x\in \mathbb {R}^2$, $V(x)\leq \lim \limits _{|y|\rightarrow +\infty }V(y)=V_\infty <+\infty$ and the inequality is strict in a subset of positive Lebesgue measure;

  3. (V3) $\inf \sigma (-\Delta +V(\cdot ))>0$, where $\sigma (-\Delta +V(\cdot ))$ denotes the spectrum of the self-adjoint operator $-\Delta +V(\cdot ):H^1(\mathbb {R}^2)\rightarrow L^2(\mathbb {R}^2)$, that is,

    \[ \inf\sigma(-\Delta+V({\cdot}))=\inf\limits_{u\in H^1(\mathbb{R}^2){\setminus}\{0\} }\frac{\int_{\mathbb{R}^2}(|\nabla u|^2+V(x)u^2)\,{\rm d}\kern 0.06em x}{\int_{\mathbb{R}^2}u^2\,{\rm d}\kern 0.06em x}>0. \]

Condition (V2) expresses that $V$ is a finite steep potential well. Assumptions of this type have been used in many recent papers for various types of elliptic problems; we refer only to Rabinowitz [Reference Rabinowitz43] for the study of a nonlinear Schrödinger equation with the nonlinear subcritical growth. The existence of a potential well rather than simply a local minimum has advantages in several situations. It is, for example, a crucial requirement when using a Lyapunov function to determine the stability of a stationary solution of an infinite dimensional dynamical system; see Ball and Marsden [Reference Ball and Marsden8, § 4] and Marsden and Hughes [Reference Marsden and Hughes37, § 6.6] for a discussion of this issue in the context of nonlinear elasticity.

As one of the typical examples of nonlinear Schrödinger equations with nonlocal nonlinearities, there has been a large amount of literature to Schrödinger–Poisson systems in dimension three, see [Reference Cerami and Vaira18, Reference He and Zou28, Reference Ianni and Ruiz29, Reference Liu, Zhang and Huang33, Reference Ruiz44, Reference Sun and Ma50, Reference Wang and Zhou54] and the references therein. This kind of systems arises in the physical literature as an approximation of the Hartree–Fock model of a quantum many-body system of electrons under the presence of the external potential $V(x)$; see [Reference Benci and Fortunato9, Reference Lieb and Simon30, Reference Lions32] and the references therein.

In the last decade, planar Schrödinger–Poisson systems have attracted a lot of attention after Stubbe [Reference Stubbe48] introduced an analytic framework to a system of this type. Indeed, the second equation in system (1.3) is called the Poisson equation, which can be solved by

\[ \phi(x) = \Gamma(x) \ast u^2(x)=\int_{\mathbb{R}^2}\Gamma(x-y)u^2(y)\,{\rm d}y, \]

where $\Gamma$ is the Newtonian kernel in dimension $2$ and is expressed by

\[ \Gamma(x)=\frac{1}{2\pi}\ln|x|,\quad x\in\mathbb{R}^2{\setminus}\{0\}. \]

And so, formally, problem (1.3) has a variational structure with the associated energy functional

\begin{align*} I(u)& =\frac{1}{2} \int_{\mathbb{R}^{2}}\left(|\nabla u|^{2}+V(x) u^{2}\right) {\rm d}\kern 0.06em x\\ & \quad+\frac{\nu}{4} \int_{\mathbb{R}^{4}} \Gamma(|x-y|) u^{2}(x) u^{2}(y)\,{\rm d}\kern 0.06em x\,{\rm d} y-\int_{\mathbb{R}^{2}}F(u)\,{\rm d}\kern 0.06em x, \end{align*}

where $F(t)=\int _{0}^{t}f(\tau )\,{\rm d}\tau$. Note that the approaches dealing with higher-dimensional cases seem difficult to be adapted to the planar case, since $\Gamma (x)=\frac {1}{2\pi }\ln |x|$ is sign-changing and presents singularities both at zero and infinity, and the corresponding energy functional is not well-defined on $H^1(\mathbb {R}^2)$. Precisely, the energy functional $I$ involves a convolution term

\[ \int_{\mathbb{R}^{2}} \int_{\mathbb{R}^{2}} \ln(|x-y|) u^{2}(x) u^{2}(y)\,{\rm d}\kern 0.06em x\,{\rm d}y \]

which is not well defined for all $u\in H^1(\mathbb {R}^2)$. So the rigorous study of planar Schrödinger–Newton systems had remained open for a long time. This is why much less is known in the planar case.

In [Reference Stubbe48], Stubbe introduced a variational framework for (1.3) with $V(x)\equiv 1$ by setting a weighted Sobolev space

\[ X:= \left\{u\in H^1(\mathbb{R}^2):\int_{\mathbb{R}^2}\ln(1+|x|)|u(x)|^{2}\,{\rm d}\kern 0.06em x<{+}\infty\right\}, \]

endowed with the norm

\[ \|u\|_{X}^{2}=\int_{\mathbb{R}^{2}}\left(|\nabla u|^{2}+V(x)|u|^{2}\right) {\rm d}\kern 0.06em x+\int_{\mathbb{R}^{2}} \ln (1+|x|)|u(x)|^{2}\,{\rm d}\kern 0.06em x, \]

which yields that the associated energy functional is well-defined and continuously differentiable on the space $X$. Cingolani and Weth [Reference Cingolani and Weth23] further developed the above variational framework to (1.3) with $f(x,u)=|u|^{p-2}u,\ p\geq 4$ and gave a variational characterization of ground state solutions when $V(x)$ is positive and $1$-periodic. Later, Du and Weth [Reference Du and Weth26] extended the above results to the case $p\in (2,4)$. Chen and Tang [Reference Chen and Tang20] proved that there exists at least a ground state solution to (1.3) in an axially symmetric functions space, when $f$ satisfies some subcritical polynomial growth conditions, see also [Reference Azzollini7, Reference Cao, Dai and Zhang14, Reference Chen and Tang21, Reference Wen, Chen and Rădulescu55] and so on. Some results on the existence and multiplicity of nontrivial solutions are obtained in [Reference Battaglia and Van Schaftingen12, Reference Chen and Tang21, Reference Liu, Rădulescu, Tang and Zhang35] for the subcritical exponential growth case. In particular, the authors in [Reference Liu, Rădulescu, Tang and Zhang35] developed an asymptotic approximation procedure to set the problem (1.3) in the standard Sobolev space $H^1(\mathbb {R}^2)$. It is worthy in [Reference Bucur, Cassani and Tarsi11, Reference Cassani and Tarsi17] that a different approach has been also developed by establishing new weighted versions of the Trudinger–Moser inequality, for which the problems are well defined in a log-weighted Sobolev space where variational methods can be applied up to cover the maximal possible nonlinear growth. For the critical exponential growth case, we also refer to [Reference Alves and Figueiredo5, Reference Chen and Tang19, Reference Chen and Tang22], which is introduced later.

In all of these works, we should point out that, what has been considered on system (1.3) is the potential $V$ is either autonomous or periodic, see [Reference Alves and Figueiredo5, Reference Cao, Dai and Zhang14, Reference Stubbe48], or axially symmetric, see [Reference Chen and Tang19, Reference Chen and Tang20, Reference Chen and Tang22, Reference Wen, Chen and Rădulescu55]. So it is quite natural to ask if there exist nontrivial solutions for planar Schrödinger–Poisson systems without any symmetry or periodicity assumption on $V$. The main focus of the present paper is at the existence of positive solutions to system (1.3) with $V$ satisfying some suitable finite potential well condition.

2. Main results

As is well known, the classical Sobolev embedding theorem asserts that

(2.1)\begin{equation} W^{1,p}_0(\Omega)\subset L^q(\Omega)\ \text{for}\ 1\leq q\leq p^* \ \text{and}\ p< N, \end{equation}

where $\Omega \subset \mathbb {R}^N$ is a bounded domain and $p^* = Np/(N-p)$ is the critical Sobolev exponent. In the limiting case $p = N$, the critical Sobolev exponent becomes infinite and $W^{1,N}(\Omega )\subset L^q(\Omega )$ for $1 \leq q < \infty$. However, one cannot take the limit as $q\nearrow N$ in (2.1), that is, the embedding $W^{1,N}_0(\Omega )\subset L^\infty (\Omega )$ is no longer valid. To fill this gap, Trudinger [Reference Trudinger53] discovered a borderline embedding result; see also Pohozaev [Reference Pohozaev42]. Roughly speaking, this is an exponential-type inequality which asserts that

\[ u\in W^{1,N}_0(\Omega) \Rightarrow \int_\Omega\,{\rm e}^{u^2}\,{\rm d}\kern 0.06em x<\infty. \]

This inequality was subsequently sharpened by Moser [Reference Moser38] as follows:

\[ \sup_{\|\nabla u\|_{L^N(\Omega)}\leq 1}\int_\Omega\,{\rm e}^{\mu |u|^{N'}}\,{\rm d}\kern 0.06em x\ \left\{ \begin{array}{@{}ll} \leq C\,|\Omega| & \text{if}\ \mu\leq \mu_N:=N\omega_N^{1/(N-1)}\\ ={+}\infty & \text{if}\ \mu>\mu_N, \end{array}\right. \]

where $N':=N/(N-1)$ and $\mu _{N-1}$ is the measure of the unit sphere in $\mathbb {R}^N$.

Since we consider planar Schrödinger–Poisson systems involving nonlinearities of exponential growth in the sense of Trudinger–Moser, we first recall the Trudinger–Moser inequality, in the sense established in [Reference Cao13]; see also [Reference Adachi and Tanaka1, Reference Cassani, Sani and Tarsi15]. This inequality plays a crucial role in estimating subcritical or critical nonlinearities of Trudinger–Moser type.

Lemma 2.1 [Reference Cao13] If $\alpha >0$ and $u\in H^1(\mathbb {R}^2)$, then

\[ \int_{\mathbb{R}^2}\left({\rm e}^{\alpha u^2}-1\right){\rm d}\kern 0.06em x<\infty. \]

Moreover, if $u\in H^1(\mathbb {R}^2),\|\nabla u\|_2^2\leq 1, \|u\|_2^2<\theta <\infty$ and $\alpha <4\pi$, then there exists a constant $C_{\theta,\alpha }$ which depends only on $\theta,\alpha$ such that

\[ \int_{\mathbb{R}^2}\left({\rm e}^{\alpha u^2}-1\right){\rm d}\kern 0.06em x\leq C_{\theta,\alpha}. \]

We also recall a notion of criticality which is totally different from the Sobolev type.

  1. (f 0) There exists $\alpha _0>0$ such that

    \[ \lim\limits_{|t|\rightarrow\infty}\frac{f(t)}{{\rm e}^{\alpha t^2}}=0,\ \forall \alpha>\alpha_0,\quad \lim\limits_{|t|\rightarrow\infty}\frac{f(t)}{{\rm e}^{\alpha t^2}}={+}\infty,\ \forall \alpha<\alpha_0, \]

which was introduced by Adimurthi and Yadava [Reference Adimurthi and Yadava2], see also de Figueiredo et al. [Reference de Figueiredo, Miyagaki and Ruf24] to the planar nonlinear elliptic problems. At present, there have been a large number of works in the literature to nonlinear elliptic problems involving critical growth of Trudinger–Moser type. We refer the readers to [Reference Alves, Cassani, Tarsi and Yang4, Reference Cassani, Tavares and Zhang16, Reference do Ó and Ruf40] and the references therein.

2.1 The subcritical case

For the subcritical exponential growth case, we make the following assumptions on the nonlinearity $f$.

  1. (f 1) For every $\theta >0$, there exists $C_\theta >0$ such that $|f(s)|\leq C_\theta \min \{1,|s|\}\,{\rm e}^{\theta |s|^2}, \forall s>0$.

  2. (f 2) $f\in C(\mathbb {R},\mathbb {R})$ and $f(s)=o(s)$ as $s\rightarrow 0$, $f(s)\equiv 0$ for $s\leq 0$ and $f(s)>0$ for $s>0$.

  3. (f 3) The function $\frac {f(t)t-F(t)}{t^3}: t\mapsto \mathbb {R}$ is nondecreasing in $(0,+\infty )$.

  4. (f 4) There exist $M_0>0$ and $t_0>0$ such that

    \[ F(t)\leq M_0|f(t)|,\quad \forall |t|\geq t_0. \]

Now we state our result about the existence of positive ground state solutions to problem (1.3).

Theorem 2.2 Assume (V1)(V3) and $(f_1)$$(f_4)$ hold. Then, for any $\nu >0$, problem (1.3) has at least a positive ground state solution $u\in X$ satisfying

(2.2)\begin{equation} \left|\int_{\mathbb{R}^2}\int_{\mathbb{R}^2}\ln|x-y|u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y\right|<{+}\infty. \end{equation}

Remark 2.3 In the present paper, we can also endow $X$ with the norm (see [Reference Chen and Tang19, Reference Chen and Tang20])

\[ \|u\|_{X}^{2}=\int_{\mathbb{R}^{2}}\left(|\nabla u|^{2}+V(x)|u|^{2}\right) {\rm d}\kern 0.06em x+\int_{\mathbb{R}^{2}} \ln (2+|x|)|u(x)|^{2}\,{\rm d}\kern 0.06em x, \]

from which we can easily see that

\[ \int_{\mathbb{R}^{2}} \ln (2+|x|)|u(x)|^{2}\,{\rm d}\kern 0.06em x>\int_{\mathbb{R}^{2}} \ln 2|u(x)|^{2}\,{\rm d}\kern 0.06em x. \]

As a clear estimate on bound from below for the term $\int _{\mathbb {R}^{2}} \ln (2+|x|)|u(x)|^{2}\,{\rm d}\kern 0.06em x$, it is good for us to develop an energy estimate inequality in analysis. Thus, it seems possible to improve condition (V3) by relaxing the lower bound of $V$.

Remark 2.4 In order to obtain the compactness directly, the authors in [Reference Chen and Tang19Reference Chen and Tang22, Reference Wen, Chen and Rădulescu55] studied system (1.3) in an axially symmetric space $E:=X\cap H_{as}^1$ with

\[ H_{as}^1=\left\{u\in H^1(\mathbb{R}^2):\, u(x):=u(x_1,x_2)=u(|x_1|,|x_2|),\ \forall x\in\mathbb{R}^2\right\}. \]

This is a natural constraint set, since critical points of the functional $I$ restricted to $E$ are also critical points of the functional $I$ in $X$. More importantly, for any axially symmetric function $u\in E$, by decomposing the convolution term in functional $I$, they can obtain an estimate

\[ \int_{\mathbb{R}^4}\ln(1+|x-y|)u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y\geq \frac{1}{4}\|u\|_2^2\int_{\mathbb{R}^2}\ln(1+|x|)u^2\,{\rm d}\kern 0.06em x,\quad u\in E, \]

which is crucial in proving that $I$ satisfies the Cerami condition at arbitrary energy level in $E$. Moreover, compared with [Reference Chen and Tang21] where the authors studied the case of subcritical polynomial growth, and used a monotonicity condition on $V$:

\begin{align*} & V\in C^1(\mathbb{R}^2,\mathbb{R}),\, t\mapsto t^2[2V(tx)\\ & \quad - \nabla V(tx)\cdot(tx)]\ \text{is nondecreasing in}\ (0,\infty)\ \text{for all}\ x\in\mathbb{R}^2 \end{align*}

to prove the existence of ground state solutions of (1.3), this condition is removed in theorem 2.2. However, we do not provide a minimax characterization of this ground state energy.

In addition to the difficulties that the quadratic part of $I$ is not coercive on $X$ and the norm of $X$ is not translation invariant, the boundedness and compactness of any Cerami sequence $\{u_n\}$ associated with functional $I$ are also the main obstacles that we need to overcome in our arguments. In the autonomous or periodic case, see [Reference Alves and Figueiredo5], some strong compactness lemma can be established with the help of the translation invariance of $I$. However, their approaches do not work any more for the non-autonomous equation (1.3), since it seems difficult to construct a Cerami sequence satisfying asymptotically some Pohozaev identity due to the lack of the translation invariance of $I$.

In contrast to the symmetry or axial symmetry case (see [Reference Chen and Tang21]), where one can establish a new inequality on $V_1(u)$ to prove the boundedness of any Cerami sequence $\{u_n\}$ for functional $I$, as mentioned in remark 2.4, it becomes tougher to prove the boundedness of $\{u_n\}$ in $X$ for our case without any symmetry assumption on $V$. To bypass this obstacle, a new nonlocal perturbation approach is introduced to prove firstly that any Cerami sequence $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$. And then by virtue of Lions’ vanishing lemma combined with estimates on the convolution term, we can rule out the vanishing case for sequence $\{u_n\}$. Finally, a delicate analysis on the mountain-pass values corresponding to the functional $I$ and the associated limit functional $I_\infty$ is given to exclude energy bubbling of sequence $\{u_n\}$ at infinity.

2.2 The critical case

We now turn our attention to the critical case. In the literature, there have been a few results on planar Schrödinger–Poisson systems with critical growth of Trudinger–Moser type. Recently, Alves and Figueiredo [Reference Alves and Figueiredo5] investigated the existence of positive ground state solutions for (1.3) when $V(x)\equiv 1$ and $f$ satisfies ($f_0$),($f_2$) and the following conditions:

  1. (f'3) $\frac {f(t)}{t^3}$ is increasing in $(0,\infty )$;

  2. (f 5) there exists $\theta >2$ such that $0<\theta F(t)\leq f(t)t$ for all $t>0$;

  3. (f 6) there exist constants $p>4$ and

    \[ \lambda_0>\max\left\{1,\left[\frac{2(p-2)\alpha_0c_p}{\pi(p-4)}\right]^{(p-2)/2}\right\} \]
    such that $f(t)\geq \lambda _0t^{p-1}$ for $t\geq 0$, where $c_p=\inf _{\mathcal {N}_p} I_p$ with
    \[ \mathcal{N}_p:=\{u\in X{\setminus}\{0\}:\,I'_p(u)u=0\} \]
    and
    \begin{align*} I_p(u)& =\frac{1}{2}\int_{\mathbb{R}^2}(|\nabla u|^2+u^2)\,{\rm d}\kern 0.06em x\\ & \quad +\frac{\nu}{8\pi}\int_{\mathbb{R}^4}\ln(|x-y|)u^2(x)u^2(y)\,{\rm d}\kern 0.06em x{\rm d}y-\frac{1}{p}\int_{\mathbb{R}^2}|u|^p\,{\rm d}\kern 0.06em x. \end{align*}

Observe from [Reference Alves and Figueiredo5] that the monotonicity condition ($f'_3$) is often used to guarantee the boundedness of the Palais–Smale sequence $\{u_n\}$ associated with $I$. With the aid of ($f_6$), the authors established directly an estimate on the norm of sequence $\{u_n\}$ in $H^1(\mathbb {R}^2)$. Then thanks to the Trudinger–Moser inequality, the compactness was obtained. However, as a global condition, ($f_6$) requires $f(t)$ to be super-cubic for all $t\geq 0$, which seems a little bit strict especially for $t>0$ small.

Later, under weaker assumptions than ($f'_3$) and ($f_6$), Chen and Tang [Reference Chen and Tang22] studied the existence of nontrivial solutions to (1.3) when $f(u)$ is replaced by $f(x,u)$. Motivated by [Reference de Figueiredo, Miyagaki and Ruf24], $f(x,u)$ is required to satisfy the following conditions:

  1. (F1) $f(x,t)t>0$ for all $(x,t)\in \mathbb {R}^2\times (\mathbb {R}{\setminus} \{0\})$ and there exist $M_0>0$ and $t_0>0$ such that

    \[ F(x,t)\leq M_0|f(x,t)|,\quad \forall x\in\mathbb{R}^2,\ |t|\geq t_0; \]
  2. (F2) $\liminf _{t\rightarrow \infty }\frac {t^2F(x,t)}{{\rm e}^{\alpha _0 t^2}}\geq \kappa >\frac {2}{\alpha _0^2\rho ^2}$ where $\rho \in (0,1/2)$ such that $\rho ^2\max _{|x|\leq \rho } V(x) \leq 1$.

Conditions (F1) and (F2) are devoted to giving a sharp estimate on the minimax level to guarantee that any Cerami sequence or any minimizing sequence $\{u_n\}$ of the associated functional does not vanish. Moreover, (F1) together with a weaker monotonicity condition

  1. (f 7) for all $x\in \mathbb {R}^2$, the mapping $(0,\infty )\ni t\mapsto \displaystyle \frac {f(t)-V(x)t}{t^3}$ is non-decreasing,

than ($f'_3$) can be used to verify that the weak limit of any Cerami sequence $\{u_n\}$ is a nontrivial solution of (1.3) in [Reference Chen and Tang22]. It is worth pointing out that the authors in [Reference Chen and Tang22] need to introduce some sort of axially symmetric assumptions on $V$ and $f$.

Another feature of the present paper is that we study the existence of nontrivial solutions to system (1.3) without hypotheses (F1) and (F2). Let $\mathcal {S}_2$ be the best constant of Sobolev embedding $H_0^1(B_{1/4}(0))\hookrightarrow L^2(B_{1/4}(0))$, that is,

(2.3)\begin{equation} \mathcal{S}_2\left(\int_{B_{1/4}(0)}u^2\,{\rm d}\kern 0.06em x\right)^{1/2}\leq\left(\int_{B_{1/4}(0)}|\nabla u|^2+V(x)u^2\,{\rm d}\kern 0.06em x\right)^{1/2}, \end{equation}

which has been one well-known fact. Moreover, the compactness of the embedding guarantees the achievement of $\mathcal {S}_2$, since $B_{1/4}(0)$ is a specific bounded domain. Recently, one fine bound of constant $\mathcal {S}_2$ has been also obtained in Du [Reference Du25], which makes possible to give one specific estimate of $\nu$ from below in the following result.

Theorem 2.5 Assume that conditions (V2)–(V3) and (f$_0$), (f$_2$), (f$_7$) hold. Then, for

\[ \nu>\frac{\alpha_0 \mathcal{S}_2^4}{4\pi\ln 2}, \]

equation (1.3) has at least a positive ground state solution $u\in X$ satisfying

(2.4)\begin{equation} \left|\int_{\mathbb{R}^2}\int_{\mathbb{R}^2}\ln|x-y|u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y\right|<{+}\infty. \end{equation}

Essentially, we adopt some ideas used in the subcritical exponential growth case to prove theorem 2.5. However, instead of the nonlocal perturbation approach used in theorem 2.2, the boundedness of any Cerami sequence $\{u_n\}$ in $H^1(\mathbb {R}^2)$ is obtained by a contradiction argument combining with Lions’ vanishing lemma.

In order to obtain the existence of weak solutions to the limiting problem and further to exclude energy bubbling of sequence $\{u_n\}$ at infinity, a refined estimate on the mountain pass value is given under an additional restriction on $\nu$.

Remark 2.6 Different from [Reference Chen and Tang19, Reference Chen and Tang22], the potential $V$ and the nonlinearity $f$ are only required to satisfy some weaker assumptions as in theorem 2.5, and are neither axially symmetric nor satisfy condition (F1) or (F2). Moreover, the conditions of theorem 2.5 do not involve ($f'_3$), ($f_5$) and ($f_6$) in [Reference Alves and Figueiredo5] where the authors studied system (1.3) in the radially symmetric setting. Indeed, ($f_7$) is weaker than ($f'_3$). Since nonlinearity $f$ in theorem 2.5 is very general, the restriction on $\nu$ has to be stated to ensure that the associated mountain pass value is less than $\frac {\pi }{\alpha _0}$, so that the compactness is recovered by virtue of the Trudinger–Moser inequality.

Throughout the paper, we need the following notations. Denote by $L^s(\mathbb {R}^2)$, $s\in [1,\infty ]$ the usual Lebesgue space with the norm $\|\cdot \|_s$. For any $r>0$ and any $z\in \mathbb {R}^2$, $B_r(z)$ stands for the ball of radius $r$ centred at $z$. $X^*$ denotes the dual space of $X$. At last, $C,C_1,C_2,\ldots$ denote various positive generic constants.

3. Preliminary results

Consider the Hilbert space $H^1(\mathbb {R}^2)$ with the norm

\[ \|u\|:=\left(\int_{\mathbb{R}^2}|\nabla u|^2+V(x)u^2\,{\rm d}\kern 0.06em x\right)^{\frac{1}{2}}, \]

which is equivalent to the standard norm in $H^1(\mathbb {R}^2)$ with (V2)–(V3). For any $u\in X$, we also define

\[ \|u\|_\star:=\left(\int_{\mathbb{R}^2}\ln(1+|x|)u^2\,{\rm d}\kern 0.06em x\right)^{\frac{1}{2}} \]

and

\[ \|u\|_X=\left(\|u\|^2+\|u\|_\star^2\right)^{\frac{1}{2}}. \]

In what follows, we recall a few basic properties about the Newton kernel to problem (1.3). Define the symmetric bilinear forms

\begin{align*} & (u,v)\mapsto B_1(u,v)=\frac{1}{2\pi}\int_{\mathbb{R}^4}\ln(1+|x-y|)u(x)u(y)\,{\rm d}\kern 0.06em x\,{\rm d}y,\\ & (u,v)\mapsto B_2(u,v)=\frac{1}{2\pi}\int_{\mathbb{R}^4}\ln\left(1+\frac{1}{|x-y|}\right)u(x)u(y)\,{\rm d}\kern 0.06em x\,{\rm d}y,\\ & (u,v)\mapsto B_0(u,v)=B_1(u,v)-B_2(u,v)=\frac{1}{2\pi}\int_{\mathbb{R}^4}\ln(|x-y|)u(x)u(y)\,{\rm d}\kern 0.06em x\,{\rm d}y, \end{align*}

where $u,v:\mathbb {R}^2\rightarrow \mathbb {R}$ are measurable functions in the Lebesgue sense. We define on $X$ the associated functionals

\begin{align*} & V_1(u)=B_1(u^2,u^2)=\frac{1}{2\pi}\int_{\mathbb{R}^4}\ln(1+|x-y|)u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y,\\ & V_2(u)= B_2(u^2,u^2)=\frac{1}{2\pi}\int_{\mathbb{R}^4}\ln\left(1+\frac{1}{|x-y|}\right)u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y,\\ & u\mapsto V_0(u)=V_1(u)-V_2(u)=\frac{1}{2\pi}\int_{\mathbb{R}^4}\ln(|x-y|)u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y. \end{align*}

Observe that

\[ \ln(1+|x-y|)\leq \ln(1+|x|+|y|)\leq \ln(1+|x|)+\ln(1+|y|)\quad \text{for}\ x,y\in\mathbb{R}^2, \]

then we have the estimate for $u,v,w,z\in X$

(3.1)\begin{align} B_1(uv,wz)& \leq \frac{1}{2\pi}\int_{\mathbb{R}^4}[\ln(1+|x|)+\ln(1+|y|)]|u(x)v(x)||w(y)z(y)|\,{\rm d}\kern 0.06em x\,{\rm d}y\nonumber\\ & \leq\frac{1}{2\pi}(\|u\|_\star\|v\|_\star\|w\|_2\|z\|_2+\|u\|_2\|v\|_2\|w\|_\star\|z\|_\star). \end{align}

Due to the Hardy–Littlewood–Sobolev inequality, we deduce that

(3.2)\begin{equation} |B_2(u,v)|\leq \frac{1}{2\pi}\int_{\mathbb{R}^4}\frac{|u(x)u(y)|}{|x-y|}\,{\rm d}\kern 0.06em x\,{\rm d}y\leq C\|u\|_{4/3}\|v\|_{4/3},\quad u,v\in X, \end{equation}

which implies that

(3.3)\begin{equation} |V_2(u)|\leq C\|u\|^4_{8/3},\quad u\in X. \end{equation}

Lemma 3.1 [Reference Cingolani and Weth23] Let $\{u_n\}$ be a sequence in $L^2(\mathbb {R}^2)$ such that $u_n\rightarrow u\in L^2(\mathbb {R}^2){\setminus} \{ 0\}$ pointwise almost everywhere on $\mathbb {R}^2$. Moreover, let $\{v_n\}$ be a bounded sequence in $L^2(\mathbb {R}^2)$ such that $\sup _{n\in \mathbb {N}}B_1(u_n^2,v_n^2)<\infty$. Then there exist $n_0\in \mathbb {N}$ and $C>0$ such that $\|v_n\|_\star < C$ for $n\geq n_0$. If, moreover,

\[ B_1(u_n^2,v_n^2)\rightarrow0\quad \text{and}\quad \|v_n\|\rightarrow0\quad \text{as}\ n\rightarrow\infty, \]

then $\|v_n\|_\star \rightarrow 0$ as $n\rightarrow \infty$.

Lemma 3.2 [Reference Cingolani and Weth23] Let $\{u_n\},\{v_n\},\{w_n\}$ be bounded sequences in $X$ such that $u_n\rightharpoonup u$ weakly in $X$. Then, for every $z \in X$, we have $B_1(v_n w_n,z(u_n-u))\rightarrow 0$ as $n\rightarrow \infty$.

Lemma 3.3 [Reference Cingolani and Weth23] The following conclusions hold true.

  1. (i) The space $X$ is compactly embedded into $L^{s}(\mathbb {R}^2)$ for all $s\in [2,+\infty )$.

  2. (ii) The functionals $V_0,V_1,V_2$ are of $C^1$ class on $X$. Moreover, $V'_i(u)v=4B_i(u^2,uv)$ for $u,v\in X$ and $i=0,1,2$.

  3. (iii) $V_2$ is continuously differentiable on $L^{\frac {8}{3}}(\mathbb {R}^2)$.

  4. (iv) $V_1$ is weakly lower semicontinuous on $H^1(\mathbb {R}^2)$.

We now recall a version of the Mountain Pass Theorem which plays a crucial role in proving the existence of nontrivial solutions.

Theorem 3.4 [Reference Silva and Vieira47] Let $X$ be a real Banach space and let $I\in C^1(X,\mathbb {R})$. Let $S$ be a closed subset of $X$ which disconnects (arcwise) $X$ in distinct connected components $X_1$ and $X_2$. Suppose further that $I(0) = 0$ and

  1. (i) $0\in X_1$ and there is $\alpha >0$ such that $I_S\geq \alpha$,

  2. (ii) there is $e\in X_2$ such that $I(e)\leq 0$.

Then $I$ possesses a sequence $\{u_n\}\subset X$(called as $(Ce)_c$ sequence) satisfying

\[ I(u_n)\rightarrow c,\ \|I'(u_n)\|_{X^*}(1+\|u_n\|_X)\rightarrow0 \]

with $c\geq \alpha > 0$ given by $c=\inf \limits _{\gamma \in \Gamma }\max \limits _{t\in [0,1]}I(\gamma (t))$, where

\[ \Gamma=\{\gamma\in C([0,1],X):\,\gamma(0)=0,\,\gamma(1)=e\}. \]

4. The subcritical case

Without loss of generality, we assume $\nu =1$ in this section. We now give more details to describe the nonlocal perturbation approach mentioned in §2. Since we do not impose the well-known Ambrosetti–Rabinowitz condition, the boundedness of the Palais–Smale sequence cannot be obtained easily. In order to overcome this difficulty, we introduce a perturbation technique developed in [Reference Liu, Zhang and Huang33, Reference Liu, Lou and Zhang34] to equation (1.3). Set

\[ \lambda\in(0,1],\quad r\in\left(\max\{p,4\},+\infty\right). \]

Consider the following modified problem:

(4.1)\begin{equation} \left\{\begin{array}{@{}ll} -\Delta u+V(x)u+\phi u+\lambda\left(\int_{\mathbb{R}^2}u^2\,{\rm d}\kern 0.06em x\right)^{\dfrac14}u =f(u)+\lambda|u|^{r-2}u, & x\in\mathbb{R}^2, \\ \Delta \phi=u^2, & x\in\mathbb{R}^2, \end{array}\right. \end{equation}

whose associated functional is given by

\begin{align*} I_{\lambda}(u)& =\frac{1}{2}\|u\|^2+\frac{1}{8\pi}\int_{\mathbb{R}^4}\ln(|x-y|)u^2(x)u^2(y)\,{\rm d}\kern 0.06em x\,{\rm d}y+\frac{2\lambda}{5} \|u\|_2^{\frac52}\\ & \quad -\int_{\mathbb{R}^2}F(u)\,{\rm d}\kern 0.06em x-\frac{\lambda}{r}\|u\|_r^r. \end{align*}

Now we provide a Pohozaev type identity for the modified equation (4.1).

Lemma 4.1 Suppose that $u\in X$ is a weak solution of (4.1). Then we have the following identity:

\begin{align*} P_\lambda(u)& :=\int_{\mathbb{R}^2}V(x)u^2\,{\rm d}\kern 0.06em x+\frac{1}{2}\int_{\mathbb{R}^2}(\nabla V(x),x)u^2\,{\rm d}\kern 0.06em x+V_0(u)+\lambda\|u\|_2^{5/2} +\frac{1}{4}\|u\|_2^4\nonumber\\ & \quad-2\int_{\mathbb{R}^2}F(u)\,{\rm d}\kern 0.06em x-\frac{2\lambda}{r}\|u\|_r^r=0. \end{align*}

In the following, we verify the geometry assumption of theorem 3.4 so that we can get the associated $(Ce)_{c_\lambda }$ sequence $\{u_{n,\lambda }\}$ (still denoted by $\{u_{n}\}$) with $c_\lambda \geq \alpha > 0$ and

(4.2)\begin{equation} c_\lambda:=\inf\limits_{\gamma\in\Gamma}\max\limits_{t\in[0,1]}I_\lambda(\gamma(t)). \end{equation}

From the following lemma, we can observe that $c_\lambda$ has an uniform bound independently of $\lambda$. That is, there exist $a,b>0$ such that $c_\lambda \in [a,b]$, where $a,b$ do not depend on $\lambda$.

Lemma 4.2 Assume (V1)–(V3) and $(f_1)$$(f_3)$ hold, then assumption (i) and (ii) of theorem 3.4 hold true.

Proof. Choosing $\theta \in (0,4\pi )$ and $p>2$, it follows from ($f_1$)–($f_2$) that, for any $\varepsilon >0$, there exists $C_{\varepsilon } > 0$ such that

(4.3)\begin{equation} |F(s)|\leq \varepsilon ({\rm e}^{\theta|s|^2}-1)+C_{\varepsilon}|s|^p,\quad s\in\mathbb{R}. \end{equation}

By Moser–Trudinger's inequality (lemma 2.1), we claim that there exists $C>0$ (independent of $u$ and $\varepsilon$) such that, for any $u\in H^1(\mathbb {R}^2)$ with $\|u\|^2<1$, there holds that

(4.4)\begin{equation} \int_{\mathbb{R}^2}F(u)\,{\rm d}\kern 0.06em x\leq \varepsilon C\|u\|^2+C_{\varepsilon}\|u\|^{p}. \end{equation}

In fact, due to $e^x-1-x\leq x(e^x-1)$ for any $x\ge 0$,

\begin{align*} \int_{\mathbb{R}^2}({\rm e}^{\theta|u|^2}-1){\rm d}\kern 0.06em x& =\theta\int_{\mathbb{R}^2}u^2\,{\rm d}\kern 0.06em x+\int_{\mathbb{R}^2}({\rm e}^{\theta|u|^2}-1-\theta|u|^2)\,{\rm d}\kern 0.06em x\\ & \leq\theta\int_{\mathbb{R}^2}u^2\,{\rm d}\kern 0.06em x+\theta\int_{\mathbb{R}^2}u^2({\rm e}^{\theta|u|^2}-1)\,{\rm d}\kern 0.06em x. \end{align*}

By lemma 2.1 and Hölder's inequality, for some $c>0$ (independent of $u$), one has

\[ \int_{\mathbb{R}^2}u^2({\rm e}^{\theta|u|^2}-1)\,{\rm d}\kern 0.06em x\leq c\|u\|^2. \]

So (4.4) follows from (4.3) and Sobolev's embedding.

Then, we have by (3.3)

(4.5)\begin{align} I_\lambda(u)& =\frac{1}{2}\|u\|^2+\frac{1}{4}[V_1(u)-V_2(u)]+\frac{2\lambda}{5} \|u\|_2^{\frac52}-\int_{\mathbb{R}^2}F(u)\,{\rm d}\kern 0.06em x-\frac{\lambda}{r}\|u\|_r^r\nonumber\\ & \geq\frac{1-2\varepsilon C}{2}\|u\|^2-C_1\|u\|^4-C_\varepsilon\|u\|^p-C_3\|u\|^r, \end{align}

which implies that there exist $\alpha >0$ and $\rho >0$ small such that

(4.6)\begin{equation} I_\lambda(u)\geq \alpha,\quad \forall u\in S=\{u\in X: \|u\|=\rho\}. \end{equation}

On the other hand, take $e\in C_0^\infty (\mathbb {R}^2)$ such that $e(x)\equiv 0$ for $x\in \mathbb {R}^2{\setminus} {B_{\frac {1}{4}}(0)}$, $e(x)\equiv 1$ for $x\in B_{\frac {1}{8}}(0)$, and $|\nabla e(x)|\leq C$. Then we have the following estimate:

(4.7)\begin{align} I_\lambda(se) & =\frac{s^2}{2}\|e\|^2+\frac{s^4}{4}[V_1(e)-V_2(e)]+\frac{2\lambda s^{5/2}}{5} \|e\|_2^{\frac52}-\int_{\mathbb{R}^2}F(se)\,{\rm d}\kern 0.06em x-\frac{\lambda s^r}{r}\|e\|_r^r\nonumber\\ & \leq\frac{s^2}{2}\|e\|^2 -\frac{s^4}{8\pi}\int_{|x|\leq\frac{1}{4}}\int_{|y|\leq\frac{1}{4}}\ln\frac{1}{|x-y|}\,{\rm e}^2(y)\,{\rm e}^2(x)\,{\rm d}y\,{\rm d}\kern 0.06em x+\frac{2 s^{5/2}}{5} \|e\|_2^{\frac52}\nonumber\\ & \leq\frac{s^2}{2}\|e\|^2 -\frac{s^4\ln2}{8\pi}\left(\int_{\mathbb{R}^2}\,{\rm e}^2(x)\,{\rm d}\kern 0.06em x\right)^2+\frac{2 s^{5/2}}{5} \|e\|_2^{\frac52}. \end{align}

Hence, we can choose $t_0>0$ large enough such that $I_{\lambda }(t_0e)<0$.

Lemma 4.3 Assume (V1)–(V3) and $(f_1)$$(f_4)$ hold, then any $(Ce)_{c_\lambda }$ sequence $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$.

Proof. Assume $\{u_n\}$ is a $(Ce)_{c_\lambda }$ sequence. Then the following holds$:$

(4.8)\begin{align} C& \geq I_{\lambda}(u_n)-\frac{1}{4}I'_{\lambda}(u_n)u_n\nonumber\\ & = \frac{1}{4}\|u_n\|^2 +\frac{3\lambda}{20}\|u_n\|_2^{\frac{5}{2}}+\int_{A_n\cup\{\mathbb{R}^2{\setminus}{A_n}\}}\left(\frac{1}{4}f(u_n)u_n-F(u_n)\right){\rm d}\kern 0.06em x\nonumber\\ & \quad +\frac{r-4}{4r}\lambda\int_{\mathbb{R}^2}|u_n|^{r}\,{\rm d}\kern 0.06em x\nonumber\\ & \geq \frac{1}{4}\|u_n\|^2 +\frac{3\lambda}{20}\|u_n\|_2^{\frac{5}{2}}-\int_{A_n}(F(u_n)-\frac{1}{4}f(u_n)u_n)\,{\rm d}\kern 0.06em x +\frac{r-4}{4r}\lambda\int_{\mathbb{R}^2}|u_n|^{r}\,{\rm d}\kern 0.06em x, \end{align}

where $A_n:=\{x|\,\frac {1}{4}f(u_n)u_n-F(u_n)\leq 0\}$. Using ($f_4$), the definition of $A_n$ implies that there exists $T>0$ such that $|u_n|\leq T$ for $x\in A_n$. So, by ($f_2$) there exists $C_T>0$ such that

(4.9)\begin{equation} \int_{A_n}(F(u_n)-\frac{1}{4}f(u_n)u_n)\,{\rm d}\kern 0.06em x\leq C_{T}\int_{A_n}u_n^2\,{\rm d}\kern 0.06em x\leq C_{T}\|u_n\|_2^2. \end{equation}

Observe that for any large $B_1>0$, there exists $B_2>0$ such that $\frac {3}{20}\|u_n\|_2^{\frac {5}{2}}\geq B_1\|u_n\|_2^2-B_2$. So combining (4.8) and (4.9) we have

(4.10)\begin{equation} C+\lambda B_2\geq \frac{1}{8}\|u_n\|^2+\int_{\mathbb{R}^2}\left[(\lambda B_1-C_{T})|u_n|^2 +\frac{r-4}{4r}\lambda|u_n|^{r}\right]{\rm d}\kern 0.06em x. \end{equation}

Let $B_1$ be large enough, then for fixed $\lambda$, the following holds:

\[ (\lambda B_1-C_{T})t^2 +\frac{r-4}{4r}\lambda t^{r}\ge0 \]

for $t\geq 0$. Thus, it follows from (4.10) that $\|u_n\|\leq C$ for some $C$ (independent of $n$).

Lemma 4.4 Assume (V1)–(V3) and $(f_1)$, $(f_2)$ and $(f_4)$ hold true, if $I'_{\lambda,\infty }(u)=0$ and $u\in X{\setminus} \{0\}$, then $I_{\lambda,\infty }(u)=\max \limits _{t\in (0,+\infty )}I_{\lambda,\infty }(u_t)$, where $u_t:=t^2u(tx)$. Here,

\begin{align*} I_{\lambda,\infty}(u)& {:=}\frac{1}{2}\int_{\mathbb{R}^2}|\nabla u|^2\,{\rm d}\kern 0.06em x+\frac{1}{2}\int_{\mathbb{R}^2}V_\infty u^2\,{\rm d}\kern 0.06em x+\frac{1}{4}V_0(u) +\frac{2\lambda}{5}\|u\|_2^{5/2}\nonumber\\ & \quad -\int_{\mathbb{R}^2}F(u)\,{\rm d}\kern 0.06em x-\frac{\lambda}{r}\|u\|_r^r. \end{align*}

Proof. By $I'_{\lambda,\infty }(u)=0$, we have

(4.11)\begin{align} 2\|\nabla u\|_2^2& +\int_{\mathbb{R}^2} V_\infty u^2\,{\rm d}\kern 0.06em x+\lambda\|u\|_2^{5/2} +V_0(u)\nonumber\\ & \quad-\frac{1}{4}\|u\|_2^4-2\int_{\mathbb{R}^2}(f(u)u-F(u))\,{\rm d}\kern 0.06em x-\frac{2(r-1)\lambda}{r}\|u\|_r^r=0, \end{align}

which comes from $2I'_{\lambda,\infty }(u)u-\mathcal {P}_{\lambda,\infty }(u)=0$. Here,

\[ \mathcal{P}_{\lambda,\infty}(u)=\int_{\mathbb{R}^2}V_\infty u^2\,{\rm d}\kern 0.06em x+\lambda\|u\|_2^{5/2}+V_0(u)+\frac{1}{4}\|u\|_2^4 -2\int_{\mathbb{R}^2} F(u)\,{\rm d}\kern 0.06em x-\frac{2\lambda}{r}\|u\|_r^r. \]

Let us define a function $\chi (t):=I_{\lambda,\infty }(u_t)$ on $[0,+\infty )$. Obviously, $\chi (0)=0$ and $\chi (t)>0$ for $t>0$ small and $\chi (t)<0$ for $t$ sufficiently large. Thus, $\max _{t\in (0,+\infty )}I_{\lambda,\infty }(u_t)$ is achieved at some $t_u>0$. So $\chi '(t_u)=0$ and $u_{t_u}$ satisfies (4.11). Observe that

\begin{align*} \chi'(t)& =2t^3\|\nabla u\|_2^2+t\int_{\mathbb{R}^2} V_\infty u^2\,{\rm d}\kern 0.06em x+\lambda t^{3/2}\|u\|_2^{5/2}+ t^3V_0(u)-\int_{\mathbb{R}^2}\left(\frac{F(t^2u)}{t^2}\right)'_t\,{\rm d}\kern 0.06em x\\ & \quad-{t}^{3}\left(\ln t +\frac{1}{4}\right)\|u\|_{2}^{4} -\frac{2(r-1)\lambda t^{2r-3}}{r}\|u\|_r^r\\ & \quad = t^3\left[2\|\nabla u\|_2^2+\frac{1}{t}\int_{\mathbb{R}^2} V_\infty u^2\,{\rm d}\kern 0.06em x+\frac{\lambda}{t^{3/2}}\|u\|_2^{5/2}+V_0(u)-\left(\ln t +\frac{1}{4}\right)\|u\|_{2}^{4}\right.\\ & \quad\left.-2\int_{\mathbb{R}^2}\frac{(f(t^2u)t^2u-F(t^2u))}{t^6}\,{\rm d}\kern 0.06em x-\frac{2(r-1)\lambda t^{2r-6}}{r}\|u\|_r^r\right]. \end{align*}

From (4.11) we infer that $\chi '(1)=0$ and $\chi '(t)>0$ for $t<1$ and $\chi '(t)<0$ for $t>1$. Thus, $t_u=1$. The proof is complete.

Motivated by the strategy of proposition 3.1 in [Reference Cingolani and Weth23], we have

Lemma 4.5 Assume (V1)–(V3) and $(f_1)$$(f_4)$ hold, for $\lambda \in (0,1]$ fixed, there exists $u_0\in X{\setminus} \{0\}$ such that $I'_\lambda (u_0)=0$ with $I_\lambda (u_0)=c_\lambda$.

Proof. By using lemmas 4.2 and 4.3, we observe that there exists a $(Ce)_{c_\lambda }$ sequence $\{u_n\}\subset X$ with $\|u_n\|\leq C$ uniformly for $n$. The remaining proof will be divided into three steps.

Step 1. We claim that

(4.12)\begin{equation} \liminf\limits_{n\rightarrow\infty}\sup\limits_{y\in\mathbb{R}^2}\int_{B_2(y)}u_n^2(x)\,{\rm d}\kern 0.06em x>0. \end{equation}

If (4.12) does not occur, then by Lions’ vanishing lemma (see [Reference Lions31]), we have $u_n\rightarrow 0$ in $L^s(\mathbb {R}^2)$ for all $s>2$. From ($f_1$)–($f_2$), take $\theta$ small enough, we deduce that, for any $\varepsilon >0$, there exists $C_{\varepsilon } > 0$ such that for $p>2$

(4.13)\begin{equation} |f(u_n)|\leq \varepsilon ({\rm e}^{\theta|u_n|^2}+1)|u_n|+C_{\varepsilon}|u_n|^p, \end{equation}

from which we have $\int _{\mathbb {R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x=o(1)$ for large $n$. Thus, it follows from (4.4) and (3.3) that

\begin{align*} & \|u_n\|^2+V_1(u_n)+\lambda \|u_n\|_2^{5/2}=I_\lambda'(u_n)u_n+V_2(u_n)\\ & \quad +\int_{\mathbb{R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x+\lambda\|u_n\|_r^r+o(1)=o(1), \end{align*}

which yields that $u_n\rightarrow 0$ in $H^1(\mathbb {R}^2)$ and $V_1(u_n)\rightarrow 0$ as $n\rightarrow \infty$. Thus, by (4.4) one has

\[ I_\lambda(u_n)\!=\!\frac{1}{2}\|u_n\|^2\!+\!\frac{2\lambda}{5} \|u_n\|_2^{5/2}\!+\!\frac{1}{4}(V_1(u_n)\!-V_2(u_n))-\int_{\mathbb{R}^2}F(u_n)\,{\rm d}\kern 0.06em x-\frac{\lambda}{r}\|u_n\|_r^r\rightarrow0 \]

as $n\rightarrow \infty$, which contradicts $c_\lambda >a$. So the claim is true. Going if necessary to a subsequence, there exists a sequence $\{y_n\}\subset \mathbb {R}^2$ such that $v_n=u_n(\cdot +y_n)$ is still bounded in $H^1(\mathbb {R}^2)$ and $v_n\rightharpoonup v_0$ for some non-zero function $v_0\in H^1(\mathbb {R}^2)$, and $v_n\rightarrow v_0$ a.e. in $\mathbb {R}^2$.

Step 2. We claim that $\{y_n\}$ is bounded. Assume by contradiction that $|y_n|\rightarrow +\infty$. Since $\{u_n\}$ is a $(Ce)_{c_\lambda }$ sequence for $I_\lambda$, the following holds:

\[ V_1(v_n)=V_1(u_n)=o(1)+V_2(u_n)+\int_{\mathbb{R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x+\lambda\|u_n\|_r^r-\|u_n\|^2-\lambda\|u_n\|_2^{\frac{5}{2}}, \]

which yields that $V_1(v_n)$ is bounded uniformly for $n$ due to the boundedness of $\{u_n\}$ in $H^1(\mathbb {R}^2)$. Recalling lemma 3.1, we obtain that $\|v_n\|_{\star }$ are also bounded uniformly for $n$, and thus $\{v_n\}$ is bounded in $X$. Up to subsequence, we may assume that $v_n\rightharpoonup v_0$ in $X$. So, $v_0\in X$. It then follows by lemma 3.3 (i) that $v_n\rightarrow v_0$ in $L^s(\mathbb {R}^2)$ for $s\geq 2$ as $n\rightarrow \infty$. Observe that for small $r>0$, due to $y_n\rightarrow +\infty$, one has

\begin{align*} \|u_n\|^2_{{\star}}& =\int_{\mathbb{R}^2}\ln(1+|x+y_n|)v_n^2(x)\,{\rm d}\kern 0.06em x\geq \int_{B_r(0)}\ln(1+|x+y_n|)v_n^2(x)\,{\rm d}\kern 0.06em x\\ & \geq \frac{1}{2}\int_{B_r(0)}\ln(1+|y_n|)v_n^2(x)\,{\rm d}\kern 0.06em x\geq C_2\ln(1+|y_n|) \end{align*}

with some $C_2>0$, and

\begin{align*} \|v_0({\cdot}{-}y_n)\|^2_{{\star}}& =\int_{\mathbb{R}^2}\ln(1+|x+y_n|)v_0^2(x)\,{\rm d}\kern 0.06em x\\ & \leq \int_{B_r(0)}[\ln(1+|x|)+\ln(1+|y_n|)]v_0^2(x)\,{\rm d}\kern 0.06em x\\ & \leq C_3\ln(1+|y_n|) \end{align*}

for some $C_3>0$. In view of the above inequalities, we deduce from the Fatou lemma that there exists $C>0$ such that

(4.14)\begin{align} \|v_0({\cdot}{-}y_n)\|_X^2& \leq \|v_0\|^2+\|v_0({\cdot}{-}y_n)\|_\star^2\nonumber\\ & \leq C(\|u_n\|^2+\|v_0({\cdot}{-}y_n)\|_\star^2)\nonumber\\ & \leq C\|u_n\|_X^2. \end{align}

Define

\begin{align*} & \tilde{I}_{\lambda,n}(v_n):=\frac{1}{2}\int_{\mathbb{R}^2}|\nabla v_n|^2\,{\rm d}\kern 0.06em x+\frac{1}{2}\int_{\mathbb{R}^2}V(x+y_n)v_n^2\,{\rm d}\kern 0.06em x\\ & \quad +\frac{1}{4}V_0(v_n) +\frac{2\lambda}{5}\|v_n\|_2^{5/2}-\int_{\mathbb{R}^2}F(v_n)\,{\rm d}\kern 0.06em x-\frac{\lambda}{r}\|v_n\|_r^r. \end{align*}

Therefore, we have for every $n$,

(4.15)\begin{align} \left|\tilde{I}'_{\lambda,n}(v_n)(v_n-v_0)\right|& =\left|I'_\lambda(u_n)(u_n-v_0({\cdot}{-}y_n))\right|\nonumber\\ & \quad \leq \|I'_\lambda(u_n)\|_{X^*}(\|u_n\|_X+\|v_0({\cdot}{-}y_n))\|_X), \end{align}

which, together with (4.14), implies that

(4.16)\begin{equation} |\tilde{I}'_{\lambda,n}(v_n)(v_n-v_0)|\rightarrow0,\quad\text{as}\ n\rightarrow\infty. \end{equation}

Based on the fact that $v_n\rightarrow v_0$ in $L^s(\mathbb {R}^2)$ for $s\geq 2$ as $n\rightarrow \infty$, by assumption (V2) one has

(4.17)\begin{equation} \int_{\mathbb{R}^2}V(x+y_n)v_n(v_n-v_0)\,{\rm d}\kern 0.06em x\rightarrow0, \end{equation}

as $n\rightarrow \infty$, and by (4.13) and (3.2) one has

(4.18)\begin{equation} \begin{aligned} & \int_{\mathbb{R}^2}f(v_n)(v_n-v_0)\,{\rm d}\kern 0.06em x\rightarrow0,\quad \|v_n\|_2^{1/2}\int_{\mathbb{R}^2}v_n(v_n-v_0)\,{\rm d}\kern 0.06em x\rightarrow0,\\ & |\frac{1}{4}V'_2(v_n)(v_n-v_0)|\leq |B_2(v_n^2,v_n(v_n-v_0))|\leq \|v_n\|_{8/3}^3\|v_n-v_0\|_{8/3}\rightarrow0, \end{aligned} \end{equation}

as $n\rightarrow \infty$. By the definition of $\tilde {I}_{\lambda,n}(v_n)$, we have

(4.19)\begin{align} & B_1(v_n^2,v_n(v_n-v_0))\nonumber\\ & = \tilde{I}'_{\lambda,n}(v_n)(v_n-v_0)-\int_{\mathbb{R}^2}V(x+y_n)v_n(v_n-v_0)\,{\rm d}\kern 0.06em x-\|\nabla (v_n-v_0)\|_2^2\nonumber\\ & \quad+B_2(v_n^2,v_n(v_n-v_0))-\lambda\|v_n\|_2^{1/2}\int_{\mathbb{R}^2}v_n(v_n-v_0)\,{\rm d}\kern 0.06em x+\int_{\mathbb{R}^2}f(v_n)(v_n-v_0)\,{\rm d}\kern 0.06em x\nonumber\\ & \quad+\lambda\int_{\mathbb{R}^2}|v_n|^{r-2}v_n(v_n-v_0)\,{\rm d}\kern 0.06em x+o(1). \end{align}

Combining (4.16)–(4.19), we infer that $\|\nabla (v_n-v_0)\|_2^2\rightarrow 0$ and $B_1(v_n^2, v_n(v_n-v_0))\rightarrow 0$ as $n\rightarrow \infty$, and then $v_n\rightarrow v_0$ in $H^1(\mathbb {R}^2)$. Recalling lemmas 3.1 and 3.2, we have $\|v_n-v_0\|_\star \rightarrow 0$ as $n\rightarrow \infty$. We therefore deduce that $v_n\rightarrow v_0$ in $X$.

Note that, for any $\varphi \in C_0^\infty (\mathbb {R}^2)$, we have, after passing to a subsequence,

\[ \int_{\mathbb{R}^2}V(x+y_n)v_n\varphi\,{\rm d}\kern 0.06em x\rightarrow\int_{\mathbb{R}^2}V_\infty v_0\varphi\,{\rm d}\kern 0.06em x\quad \text{as}\ n\rightarrow\infty. \]

Thus, from (4.14) and (4.15) we deduce that for any $\varphi \in C_0^\infty (\mathbb {R}^2)$

\begin{align*} |I'_{\lambda,\infty}(v_0)\varphi|& =|\tilde{I}'_{\lambda,n}(v_n)\varphi|+o(1)=|I'_\lambda(u_n)\varphi({\cdot}{-}y_n)|+o(1)\\ & \leq \|I'_\lambda(u_n)\|_{X^*}\|\varphi({\cdot}{-}y_n)\|_{X}+o(1)\\ & \leq C\|I'_\lambda(u_n)\|_{X^*}\|\varphi\|_{X}+o(1), \end{align*}

which implies that $I'_{\lambda,\infty }(v_0)=0$. That is, $v_0$ is a nontrivial critical point of functional $I_{\lambda,\infty }$ with $I_{\lambda,\infty }(v_0)=c_\lambda$. Recalling the definition of $c_\lambda$ and lemma 4.4, we have

\[ c_\lambda\leq \max_{t\in(0,+\infty)}I_\lambda(v_{0t})<\max_{t\in(0,+\infty)}I_{\lambda,\infty}(v_{0t})=I_{\lambda,\infty}(v_{0})=c_\lambda, \quad v_{0t}=t^2v_0(tx), \]

which is a contradiction. Therefore, $\{y_n\}$ is a bounded sequence.

Step 3. We show that $u_n\rightarrow u$ in $X$. Since we have known from step 2 that $\{y_n\}$ is a bounded sequence, there exists $u_0\in H^1(\mathbb {R}^2){\setminus} \{0\}$ such that $u_n\rightharpoonup u_0$ in $H^1(\mathbb {R}^2)$ and $u_n\rightarrow u_0$ a.e. in $\mathbb {R}^2$. Arguing as in step 2, we deduce that $u_n\rightarrow u_0$ in $X$. We conclude that $u_0$ is critical point of $I_\lambda$ with $I_\lambda (u_\lambda )=c_\lambda$.

4.1 Proof of theorem 2.2

In view of lemma 4.5, for fixed $\lambda \in (0,1]$, we have $I'_{\lambda }(u_\lambda )=0$ for some $u_\lambda \in X{\setminus} \{0\}$. Choosing a sequence $\{\lambda _n\}\subset (0,1]$ satisfying $\lambda _n\rightarrow 0^+$, there exists a sequence of nontrivial critical points $\{u_{\lambda _n}\}$(still denoted by $\{u_n\}$) of $I_{\lambda _n}$ with $I_{\lambda _n}(u_n)=c_{\lambda _n}$. We now show that $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$. In fact, according to the definition of $I'_{\lambda _n}(u_n)u_n=0$, we have

(4.20)\begin{equation} V_0(u_n)={-}(\|\nabla u_n\|_2^2+\int_{\mathbb{R}^2}V(x)u_n^2\,{\rm d}\kern 0.06em x) -\lambda_n\|u_n\|_2^{5/2} +\lambda_n\|u_n\|_r^r+\int_{\mathbb{R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x. \end{equation}

Substituting $V_0(u_n)$ into $P_\lambda (u_n)=0$ gives

(4.21)\begin{align} & -\|\nabla u_n\|_2^2 +\frac{1}{2}\int_{\mathbb{R}^2}(\nabla V(x),x)u_n^2+\frac{1}{4}\|u_n\|_2^4\nonumber\\ & \quad +\int_{\mathbb{R}^2}f(u_n)u_n-2F(u_n)\,{\rm d}\kern 0.06em x +\frac{r-2}{r}\lambda_n\|u_n\|_r^r=0. \end{align}

We use the same fashion to get

(4.22)\begin{align} I_{\lambda_n}(u_n)& =\frac{1}{4}\|\nabla u_n\|_2^2+\frac{1}{4}\int_{\mathbb{R}^2}V(x)u_n^2\,{\rm d}\kern 0.06em x\nonumber\\ & \quad+\int_{\mathbb{R}^2}\frac{1}{4}f(u_n)u_n-F(u_n)\,{\rm d}\kern 0.06em x+\frac{3\lambda_n}{20}\|u_n\|_2^{5/2} +\frac{\lambda_n(r-4)}{4r}\|u_n\|_r^r. \end{align}

Putting (4.22) into (4.21), we get

(4.23)\begin{equation} \begin{aligned} & -[4I_{\lambda_n}(u_n)-\int_{\mathbb{R}^2}V(x)u_n^2\,{\rm d}\kern 0.06em x+\left(\frac{4}{r}-1\right)\lambda_n\|u_n\|_r^r\\ & \quad +\int_{\mathbb{R}^2}4F(u_n)-f(u_n)u_n\,{\rm d}\kern 0.06em x-\frac{3\lambda_n}{5}\|u_n\|_2^{5/2}] \\ & +\frac{1}{2}\int_{\mathbb{R}^2}(\nabla V(x),x)u_n^2 +\frac{1}{4}\|u_n\|_2^4 +\int_{\mathbb{R}^2}f(u_n)u_n-2F(u_n)\,{\rm d}\kern 0.06em x +\frac{r-2}{r}\lambda_n\|u_n\|_r^r=0, \end{aligned} \end{equation}

which can be rewritten as

(4.24)\begin{align} & \frac{1}{4}\|u_n\|_2^4+\int_{\mathbb{R}^2}V(x)u_n^2\,{\rm d}\kern 0.06em x+\left(2-\frac{6}{r}\right)\lambda_n\|u_n\|_r^r \nonumber\\ & \quad + \frac{1}{2}\int_{\mathbb{R}^2}(\nabla V(x),x)u_n^2\,{\rm d}\kern 0.06em x+\frac{3\lambda_n}{5}\|u_n\|_2^{5/2}\nonumber\\ & \quad +\int_{\mathbb{R}^2}2f(u_n)u_n-6F(u_n){\rm d}\kern 0.06em x=4I_{\lambda_n}(u_n). \end{align}

From condition ($f_3$), we can conclude that $f(u_n)u_n\geq 3F(u_n)$, see lemma 4.2 in [Reference Chen and Tang21]. Thus, (4.24) implies that $\{u_n\}$ is bounded in $L^2(\mathbb {R}^2)$ uniformly for $n$. Moreover, observe from (4.24) that

(4.25)\begin{equation} \begin{aligned} & \int_{\mathbb{R}^2}V(x)u_n^2\,{\rm d}\kern 0.06em x+\frac{1}{2}\int_{\mathbb{R}^2}\nabla V(x)\cdot xu_n^2\,{\rm d}\kern 0.06em x\leq C,\\ & \left(2-\frac{6}{r}\right)\lambda_n\|u_n\|_r^r\leq C,\quad \frac{3\lambda_n}{5}\|u_n\|_2^{5/2}\leq C,\\ & \int_{\mathbb{R}^2}2f(u_n)u_n-6F(u_n)\,{\rm d}\kern 0.06em x\leq C. \end{aligned} \end{equation}

In view of lemma 4.1, we have

(4.26)\begin{align} & 2I'_{\lambda_n}(u_n)u_n-P_{\lambda_n}(u_n)= 2\|\nabla u_n\|_2^2\nonumber\\ & \qquad +\int_{\mathbb{R}^2} (V(x)-\frac{1}{2}\nabla V(x)\cdot x) u_n^2\,{\rm d}\kern 0.06em x+\lambda_n\|u_n\|_2^{5/2} +V_0(u)\nonumber\\ & \qquad-\frac{1}{4}\|u_n\|_2^4-2\int_{\mathbb{R}^2}(f(u_n)u_n-F(u_n))\,{\rm d}\kern 0.06em x-\frac{2(r-1)\lambda_n}{r}\|u_n\|_r^r=0. \end{align}

For $t>0$, from (4.26) we deduce that

(4.27)\begin{align} & I_{\lambda_n}(u_n)-I_{\lambda_n}(u_{nt})= \frac{1-t^4}{2}\|\nabla u_n\|_2^2 +\frac{1}{2}\int_{\mathbb{R}^2} [V(x)-t^2V(t^{{-}1}x)] u_n^2\,{\rm d}\kern 0.06em x+\lambda_n\frac{2(1-t^{5/2})}{5}\|u_n\|_2^{5/2}\nonumber\\ & \quad+\frac{1-t^4}{4}V_0(u_n) +\frac{t^4\ln t}{4}\|u_n\|_2^4+\int_{\mathbb{R}^2}\left[\frac{F(t^2u_n)}{t^2}-F(u_n)\right]{\rm d}\kern 0.06em x-\frac{(1-t^{2(r-1)})\lambda_n}{r}\|u_n\|_r^r\nonumber\\ & \quad= \frac{1-t^4}{4}[2I'_{\lambda_n}(u_n)u_n-P_{\lambda_n}(u_n)]+\frac{1}{t^2}F(t^2u_n)\bigg]{\rm d}\kern 0.06em x\nonumber\\ & \quad+\frac{1}{4}\int_{\mathbb{R}^2}\left[(1+t^4)V(x)-2t^2V(t^{{-}1}x) +\frac{1-t^4}{2}\nabla V(x)\cdot x\right]u_n^2\,{\rm d}\kern 0.06em x\nonumber\\ & \quad+\frac{1-t^4+4t^4\ln t}{16}\|u_n\|_2^4+\int_{\mathbb{R}^2}\left[\frac{1-t^4}{2}f(u_n)u_n +\frac{t^4-3}{2}F(u_n)\right.\nonumber\\ & \quad+\left(\frac{3}{20}+\frac{t^4}{4}-\frac{2t^{5/2}}{5}\right)\lambda_n\|u_n\|_2^{5/2}+\left[\frac{(r-1)(1-t^4)}{2r}-\frac{(1-t^{2(r-1)})} {r}\right]\lambda_n\|u_n\|_r^r. \end{align}

where $u_{nt}(x)=t^2u_{n}(tx)$. Now we show that $\{\|\nabla u_n\|_2\}$ is bounded uniformly for $n$. Suppose by contradiction that $\|\nabla u_n\|_2\rightarrow \infty$. Take $t_n=(\sqrt {M}/\|\nabla u_n\|_2)^{1/2}$ for some $M>0$ large, then $t_n\rightarrow 0$. Obviously, $t_n^4\ln t_n\rightarrow 0$. Letting $t=t_n$ in (4.27), since $\{u_n\}$ is bounded in $L^2(\mathbb {R}^2)$ uniformly for $n$, by (V2) and ($f_1$) and (4.4), we have for large $t_n$

(4.28)\begin{align} & I_{\lambda_n}(u_n)-I_{\lambda_n}(t_n^2u_{nt_{n}})\nonumber\\ & = \frac{1}{4}\int_{\mathbb{R}^2}\left[V(x) +\frac{1}{2}\nabla V(x)\cdot x\right]u_n^2\,{\rm d}\kern 0.06em x\nonumber\\ & \quad+\frac{1}{16}\|u_n\|_2^4+\int_{\mathbb{R}^2}\left[\frac{1}{2}f(u_n)u_n-\frac{3}{2}F(u_n)\right]{\rm d}\kern 0.06em x,\nonumber\\ & \quad+ \frac{3}{20}\lambda_n\|u_n\|_2^{5/2}+\frac{r-3}{2r}\lambda_n\|u_n\|_r^r+o(1). \end{align}

Therefore, it follows from (3.3), (4.4), (4.25), (4.28) and (V1), ($f_1$), the Gagliardo–Nirenberg inequality that

(4.29)\begin{align} c_{\lambda_n}& =I_{\lambda_n}(u_n)\geq I_{\lambda_n}(t_n^2u_{nt_{n}})+o(1)\nonumber\\ & = \frac{t_n^4}{2}\|\nabla u_n\|_2^2+\frac{t_n^4}{4}V_0(u_n)-\frac{t_n^4\ln t_n}{4}\|u_n\|_2^4 -\frac{1}{t_n^2}\int_{\mathbb{R}^2}F(t_n^2u_n)\,{\rm d}\kern 0.06em x\nonumber\\ & \quad+\frac{1}{2}\int_{\mathbb{R}^2} t_n^2V(t_n^{{-}1}x) u_n^2\,{\rm d}\kern 0.06em x+\lambda_n\frac{2t_n^{5/2}}{5}\|u_n\|_2^{5/2}-\frac{t_n^{2r-2}}{r}\lambda_n\|u_n\|_r^{r}+o(1)\nonumber\\ & \geq\frac{t_n^4}{2}\|\nabla u_n\|_2^2-\frac{t_n^4}{4}V_2(u_n)-t_n^{2(p-1)}C\|u_n\|_p^p+o(1)\nonumber\\ & \geq\frac{M}{2}-\frac{t_n^4C}{4}\|u_n\|_2^3\|\nabla u_n\|_2-t_n^{2(p-1)}C\|u_n\|_2^2\|\nabla u_n\|_2^{p-2}+o(1)\nonumber\\ & \geq\frac{M}{2}-\frac{CM}{4\|\nabla u_n\|_2}\|u_n\|_2^3-\frac{CM^{(p-1)/2}}{\|\nabla u_n\|_2}\|u_n\|_2^2+o(1), \end{align}

which, together with the fact that $\|\nabla u_n\|_2\to +\infty$, implies a contradiction by letting $M>0$ large enough. Hence, $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$. Arguing similarly as in lemma 4.5, we obtain that $u_n\rightarrow u_0$ in $X$. Moreover, by letting $n\rightarrow +\infty$, one has $c_{\lambda _n}\rightarrow c^*>a>0$ and $I_{\lambda _n}(u_{n})=c^*+o(1)$. Here, $a$ is positive number given in (4.2) below. For any $\varphi \in C_0^\infty (\mathbb {R}^2)$, we have

\[ I'_{\lambda_n}(u_{n})\varphi= I'(u_{n})\varphi+\lambda_n\|u_n\|_2^\frac{1}{2}\int_{\mathbb{R}^2}u_{n}\varphi\,{\rm d}\kern 0.06em x +\lambda_n\int_{\mathbb{R}^2}|u_n|^{r-2}u_n\varphi\,{\rm d}\kern 0.06em x=o(1)\|\varphi\|_X. \]

Thus, $\{u_n\}$ is a Palais–Smale sequence of $I$ with level $c^*$. Therefore, arguing as that in lemma 4.5, there exists a nontrivial $u_0\in X$ such that $I'(u_0)=0$ and $I(u_0)=c^*$. Now let us define the set of solutions

\[ \mathcal{S}:=\{u\in X{\setminus}\{0\}:\,I'(u)=0\}. \]

It is clear that $\mathcal {S}\not =\emptyset$ and $\mathcal {S}$ is bounded away from zero. To be precise, a short estimate yields that for any $u\in \mathcal {S}$, the following holds:

(4.30)\begin{equation} \|u\|\geq C\quad\text{for}\; \text{some}\ C>0. \end{equation}

We claim that

\[ c_*:=\inf\limits_{u\in\mathcal{S}}I(u)>0. \]

For a contradiction, we assume $c_*=0$. Then there exists $\{u_n\}\subset \mathcal {S}$ such that $I(u_n)\rightarrow 0$ as $n\rightarrow \infty$. In view of (4.24), $I(u)\geq \frac {1}{16}\|u\|_2^4$ for all $u\in \mathcal {S}$. So, $\|u_n\|_2^4\rightarrow 0$. Arguing as above, we have $u_n$ is bounded in $H^1(\mathbb {R}^2)$ uniformly for $n$. Using the Gagliardo–Nirenberg inequality:

\[ \|u_n\|_p^p\leq C_p\|u_n\|_2^2\|\nabla u_n\|_2^{p-2} \]

we infer that $u_n\rightarrow 0$ in $L^p(\mathbb {R}^2)$ for $p\in [2,+\infty )$. By (4.13), we have ${\int _{\mathbb {R}^2}f(u_n)u_n{\rm d}\kern 0.06em x=o(1)}$ for large $n$, and furthermore $u_n\rightarrow 0$ in $H^1(\mathbb {R}^2)$ which contradicts (4.30). The claim is true. Take finally a minimizing sequence $\{u_n\}\subset \mathcal {S}$ so that $I(u_n)\rightarrow c_*$. Similarly to lemma 4.5, there exists $u_{*}\in X$ so that $u_n\rightarrow u_{*}$ in $X$ and $I'(u_{*})=0$. It follows that $u_{*}$ is a positive ground state solution of problem (1.3).

5. The critical case

In this section, we are devoted to the proof of theorem 2.5. Differently from the subcritical case, we prove the existence of critical point of functional $I$ directly by using the Mountain-Pass Theorem (see theorem 3.4). It is not hard to check the mountain pass geometry of $I$ as similar arguments to the subcritical case. Therefore, recalling theorem 3.4, we can also get the associated $(Ce)_{c_{mp}}$ sequence $\{u_{n}\}$ with $c_{mp}\geq \alpha > 0$ and $c_{mp}:=\inf \limits _{\gamma \in \Gamma }\max \limits _{t\in [0,1]}I(\gamma (t)),$ where $\Gamma =\{\gamma \in C([0,1],X):\,\gamma (0)=0,\,\gamma (1)=e\}$. Now let us verify the $(Ce)_{c_{mp}}$ sequence $\{u_{n}\}$ contains a bounded subsequence in $H^1(\mathbb {R}^2)$.

Lemma 5.1 Assume the conditions of theorem 2.5 hold, then if $\nu >\frac {\alpha _0 \mathcal {S}_2^4}{4\pi \ln 2}$, we have

(5.1)\begin{equation} c_{mp}<\frac{\pi}{\alpha_0}, \end{equation}

where $\alpha _0$ has been given by assumption ($f_0$).

Proof. Recalling (2.3), there exists $\tilde {e}\in H_0^1(B_{\frac {1}{4}}(0))$ such that

(5.2)\begin{equation} \|\tilde{e}\|=1\quad\text{and}\quad \|\tilde{e}\|_2=\mathcal{S}_2^{{-}1}. \end{equation}

Then, similarly to lemma 4.2, for any $s>0$ we have

(5.3)\begin{align} I(s\tilde{e}) & =\frac{s^2}{2}\|\tilde{e}\|^2+\frac{s^4\nu}{4}[V_1(\tilde{e})-V_2(\tilde{e})]-\int_{\mathbb{R}^2}F(s\tilde{e}){\rm d}\kern 0.06em x\nonumber\\ & \leq\frac{s^2}{2}\|\tilde{e}\|^2 -\frac{s^4\nu}{8\pi}\int_{|x|\leq\frac{1}{4}}\int_{|y|\leq\frac{1}{4}}\ln\frac{1}{|x-y|}\tilde{e}^2(y)\tilde{e}^2(x)\,{\rm d}y\,{\rm d}\kern 0.06em x\nonumber\\ & \leq\frac{s^2}{2} -\frac{s^4\nu\ln2}{8\pi}\mathcal{S}_2^{{-}4}. \end{align}

And so,

\[ c_{mp}<\max_{s\in(0,+\infty)}\left\{\frac{s^2}{2} -\frac{s^4\nu\ln2}{8\pi}\mathcal{S}_2^{{-}4}\right\}. \]

A direct computation shows that

\[ \max_{s\in(0,+\infty)}\left\{\frac{s^2}{2} -\frac{s^4\nu\ln2}{8\pi}\mathcal{S}_2^{{-}4}\right\}=\frac{\pi\mathcal{S}_2^4}{2\ln 2 \cdot\nu}. \]

Thus, the conclusion follows from $\nu >\frac {\alpha _0 \mathcal {S}_2^4}{2\ln 2}$.

Lemma 5.2 Assume the conditions of theorem 2.5 hold, then any $(Ce)_{c_{mp}}$ sequence $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$.

Proof. From ($f_{7}$) we can deduce that (see also lemma 2.3 in [Reference Chen and Tang20])

(5.4)\begin{align} & \frac{1-s^4}{4}f(t)t+F(st)-F(t)+\frac{(1-s^2)^2}{4}V(x)t^2\nonumber\\ & = \int_{1}^{s}\left[\frac{f(\tau t)-V(x)\tau t}{(\tau t)^3}-\frac{f(t)-V(x)t}{t^3}\right]\tau^3t^4\,{\rm d}\tau\geq 0,\quad \forall t\not=0,s\geq0. \end{align}

Then by the definition of $I$, we have

(5.5)\begin{align} C& \geq I(u_n)-\frac{1}{4}I'(u_n)u_n\nonumber\\ & = \frac{1}{4}\|\nabla u_n\|_2^2 +\int_{\mathbb{R}^2}\left(\frac{1}{4}f(u_n)u_n-F(u_n)+\frac{1}{4}V(x)u_n^{2}\right){\rm d}\kern 0.06em x, \end{align}

which implies that $\{\|\nabla u_n\|_2\}$ is bounded uniformly for $n$. In view of lemma 5.1, there exists $\varepsilon _0>0$ small such that

(5.6)\begin{equation} c_{mp}<\frac{\pi}{\alpha_0}(1-5\varepsilon_0)=:\tilde{c}<\frac{\pi}{\alpha_0}. \end{equation}

We now prove the boundedness of $\{u_n\}$ in $H^1(\mathbb {R}^2)$. Suppose by contradiction that $\|u_n\|\rightarrow \infty$. Set $v_n=\sqrt {4\tilde {c}} u_n/\|u_n\|$, then $\|v_n\|^2=4\tilde {c}$ and $\|\nabla v_n\|=o(1)$. So, we have $v_n\rightharpoonup v$ in $H^1(\mathbb {R}^2)$ and $v_n\rightarrow v$ a.e. in $\mathbb {R}^2$ after passing to a subsequence. Furthermore, we have either $\{v_n\}$ is vanishing, i.e.,

(5.7)\begin{equation} \lim\limits_{n\rightarrow\infty}\sup\limits_{y\in\mathbb{R}^2}\int_{B_2(y)}v_n^2(x)\,{\rm d}\kern 0.06em x=0 \end{equation}

or non-vanishing, i.e., there exist $\delta >0$ and a sequence $\{y_n\}\subset \mathbb {R}^2$ such that

(5.8)\begin{equation} \lim\limits_{n\rightarrow\infty}\int_{B_2(y_n)}v_n^2(x)\,{\rm d}\kern 0.06em x>\delta. \end{equation}

If (5.7) occurs, then it follows from Lions’ vanishing lemma (see [Reference Lions31]) that $v_n\rightarrow 0$ in $L^s(\mathbb {R}^2)$ for all $s>2$.

Moreover, a straightforward computation shows by assumption ($f_{7}$) and (5.4) that

(5.9)\begin{equation} I(u)\geq I(tu)+\frac{1-t^4}{4}I'(u)u,\quad \forall u\in X,t\geq 0. \end{equation}

It then follows from (3.3) and ($f_{7}$) that

(5.10)\begin{align} I(u_n)& \geq I(v_n)+\frac{1-16\tilde{c}^{2}\|u_n\|^{{-}4}}{4}I'(u_n)u_n\nonumber\\ & \geq \frac{1}{2}\|v_n\|^2+\frac{\nu}{4}V_0(v_n)-\int_{\mathbb{R}^2}F(v_n)\,{\rm d}\kern 0.06em x+o(1)\nonumber\\ & \geq \frac{1}{2}\|v_n\|^2-\int_{\mathbb{R}^2}\frac{1}{4}\left(f(v_n)v_n+V(x)v^2_n\right){\rm d}\kern 0.06em x+o(1). \end{align}

Since $\|\nabla v_n\|=o(1)$, by $(f_0)$ and Trudinger–Moser's inequality, for any $\varepsilon >0$ there exists $C_\varepsilon >0$ such that

\begin{align*} \int_{\mathbb{R}^2}f(v_n)v_n\,{\rm d}\kern 0.06em x& \leq \varepsilon\int_{\mathbb{R}^2}u_n^2\,{\rm d}\kern 0.06em x+C_\varepsilon\int_{\mathbb{R}^2}({\rm e}^{\alpha_0v_n^2}-1)v_n\,{\rm d}\kern 0.06em x\\ & \leq \varepsilon C\|v_n\|^2+C_\varepsilon\left[\int_{\mathbb{R}^2}({\rm e}^{\frac{4}{3}\alpha_0v_n^2}-1)\,{\rm d}\kern 0.06em x\right]^{\frac{3}{4}}\|v_n\|_4\\ & =\varepsilon C\|v_n\|^2+C_\varepsilon o(1), \end{align*}

which implies by the arbitrariness of $\varepsilon$ that

\[ \lim\limits_{n\rightarrow\infty}\int_{\mathbb{R}^2}f(v_n)v_n\,{\rm d}\kern 0.06em x=0, \]

which, together with (5.10), implies a contradiction due to (5.6) because if $\|\nabla v_n\|_2^2\to 0$, then $\int _{\mathbb {R}^2} V(x)v_n^2\,{\rm d}\kern 0.06em x\to 4\tilde c$. Thus non-vanishing must hold, namely relation (5.8) holds true. Since $\|\nabla v_n\|=o(1)$ and $\|v_n\|^2=4\tilde {c}$, in view of (5.10), the Trudinger–Moser inequality implies that there exists $C>0$ such that

(5.11)\begin{equation} V_1(v_n)\leq C. \end{equation}

Therefore, from ($f_1$), ($f_7$), (3.3), (5.8), (5.10) and (5.11) we infer that for $n$ large

(5.12)\begin{align} o(1)& =\frac{I'(u_n)u_n}{\|u_n\|^4}=\nu V_1(v_n)-\nu V_2(v_n)-\int_{\mathbb{R}^2}\frac{f(u_n)u_n}{\|u_n\|^4}\,{\rm d}\kern 0.06em x+o(1)\nonumber\\ & \leq C-\int_{\mathbb{R}^2}\frac{f(\sqrt{4\tilde{c}}\|u_n\|v_n)v^4_n}{\sqrt{4\tilde{c}}\|u_n\|^3v^3_n}\,{\rm d}\kern 0.06em x+o(1)\nonumber\\ & \leq C-\int_{B_2(y_n)}\frac{f(\sqrt{4\tilde{c}}\|u_n\|v_n)v^4_n}{\sqrt{4\tilde{c}}\|u_n\|^3v^3_n}\,{\rm d}\kern 0.06em x+o(1)={-}\infty. \end{align}

This is a contradiction and the conclusion follows.

Lemma 5.3 Assume the conditions of theorem 2.5 hold, there exists $u_0\in X{\setminus} \{0\}$ such that $I'(u_0)=0$ with $I(u_0)=c_{mp}$.

Proof. Assume that $\{u_{n}\}$ is a $(Ce)_{c_{mp}}$ sequence of functional $I$, then we have from lemma 5.2 that there exists $M>0$ such that $\|u_n\|\leq M$ uniformly for $n$. Similarly to lemma 4.5, the proof of this lemma will be also divided into three steps.

Step 1. We show that

(5.13)\begin{equation} \liminf\limits_{n\rightarrow\infty}\sup\limits_{y\in\mathbb{R}^2}\int_{B_2(y)}u_n^2(x)\,{\rm d}\kern 0.06em x>0. \end{equation}

Otherwise, it follows from Lions’ vanishing lemma (see [Reference Lions31]) that $u_n\rightarrow 0$ in $L^s(\mathbb {R}^2)$ for all $s>2$. By recalling (5.5) and lemma 5.1, there exists $\varepsilon _0>0$ small such that

(5.14)\begin{equation} \|\nabla u_n\|_2^2\leq 4c_{mp}<\frac{4\pi}{\alpha_0}(1-5\varepsilon_0). \end{equation}

In view of ($f_0$), for $s\in (1,2)$ and some $M_1>0$, we have

(5.15)\begin{equation} |f(u)|^s\leq C[{\rm e}^{\alpha_0(1+\varepsilon_0)su^2}-1],\quad |u|\geq M_1. \end{equation}

By choosing $s\in (1,2)$ such that

\[ (1+\varepsilon_0)(1-5\varepsilon_0)s<1, \]

we infer from lemma 2.1, (5.14) and (5.15), Hölder's inequality that

(5.16)\begin{align} \int_{|u_n|\geq M_1}f(u_n)u_n\,{\rm d}\kern 0.06em x & \leq \left(\int_{|u_n|\geq M_1}|f(u_n)|^s\,{\rm d}\kern 0.06em x\right)^{1/s}\|u_n\|_{s'}\nonumber\\ & \leq C\left(\int_{\mathbb{R}^2}[{\rm e}^{\alpha_0(1+\varepsilon_0)su_n^2}-1]\,{\rm d}\kern 0.06em x\right)^{1/s}\|u_n\|_{s'}\nonumber\\ & \leq C\|u_n\|_{s'}=o(1). \end{align}

Here, $s'=\frac {s}{s-1}\in (2,+\infty )$. So, for any $\varepsilon >0$, by ($f_2$), there exist $\bar {M}_\varepsilon >0$ small and $C_\varepsilon >0$ such that

(5.17)\begin{align} \int_{\mathbb{R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x& \leq\int_{|u_n|\leq\bar{M}_\varepsilon}\varepsilon u_n^2\,{\rm d}\kern 0.06em x +\int_{ \bar{M}_\varepsilon\leq|u_n|\leq M_1}f(u_n)u_n\,{\rm d}\kern 0.06em x \nonumber\\ & +\int_{|u_n|\geq M_1}f(u_n)u_n\,{\rm d}\kern 0.06em x \leq\varepsilon C+C_\varepsilon\int_{\tilde{M}_\varepsilon\leq|u_n|\leq M_\varepsilon}|u_n|^p\,{\rm d}\kern 0.06em x+\varepsilon\nonumber\\ & \leq\varepsilon C+C_\varepsilon o(1)+\varepsilon,\quad p\in(2,\infty). \end{align}

Thus,

(5.18)\begin{equation} \lim\limits_{n\rightarrow\infty}\int_{\mathbb{R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x=0. \end{equation}

Thus, it follows from (5.18) and (3.3) that

(5.19)\begin{equation} \|u_n\|^2+\nu V_1(u_n)=I_\lambda'(u_n)u_n+\nu V_2(u_n)+\int_{\mathbb{R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x+o(1)=o(1), \end{equation}

which implies that $u_n\rightarrow 0$ in $H^1(\mathbb {R}^2)$, $V_1(u_n)\rightarrow 0$. Moreover,

(5.20)\begin{equation} \int_{\mathbb{R}^2}F(u_n)\,{\rm d}\kern 0.06em x\leq \frac{1}{4}\int_{\mathbb{R}^2}\left(f(u_n)u_n+V(x)u^2_n\right){\rm d}\kern 0.06em x=o(1), \end{equation}

and then

\[ \lim_{n\rightarrow\infty}I(u_n)=0=c_{mp} \]

which contradicts $c_{mp}>0$. So (5.13) holds true. After passing to a subsequence, there exists a sequence $\{y_n\}\subset \mathbb {R}^2$ such that $v_n=u_n(\cdot +y_n)$ is still bounded in $H^1(\mathbb {R}^2)$ and $v_n\rightharpoonup v_0\in H^1(\mathbb {R}^2){\setminus} \{0\}$ and $v_n\rightarrow v_0$ a.e. in $\mathbb {R}^2$.

Step 2. We claim that $\{y_n\}$ is bounded. Assume by contradiction that $y_n\rightarrow +\infty$. In view of (5.17), we can easily see that $\int _{\mathbb {R}^2}f(u_n)u_n\,{\rm d}\kern 0.06em x\leq C$ uniformly for $n$. So, by the definition of $I$, we have

\begin{align*} \frac{\nu}{4}V_1(v_n)& =\frac{\nu}{4}V_1(u_n)\\ & =I(u_n)+\frac{\nu}{4}V_2(u_n) +\int_{\mathbb{R}^2}F(u_n)\,{\rm d}\kern 0.06em x-\frac{1}{2}\|u_n\|^2\\ & \leq I(u_n)+\frac{\nu}{4}V_2(u_n) +\int_{\mathbb{R}^2}\frac{1}{4}\left(f(u_n)u_n+V(x)u_n^2\right){\rm d}\kern 0.06em x-\frac{1}{2}\|u_n\|^2, \end{align*}

which implies that $V_1(v_n)$ is bounded uniformly for $n$, due to the boundedness of $\{u_n\}$ in $H^1(\mathbb {R}^2)$. It follows from lemma 3.1 that $\|v_n\|_{\star }$ is also bounded in $n$, and so $\{v_n\}$ is bounded in $X$. Up to subsequence, there exists $v_0\in X$ such that $v_n\rightharpoonup v_0$ in $X$ and $v_n\rightarrow v_0$ in $L^s(\mathbb {R}^2)$ for $s\geq 2$ as $n\rightarrow \infty$. Arguing as in lemma 4.5, we obtain that

(5.21)\begin{equation} |\tilde{I}'(v_n)(v_n-v_0)|\rightarrow0,\quad\text{as}\ n\rightarrow\infty. \end{equation}

where

\[ \tilde{I}(v_n):=\frac{1}{2}\int_{\mathbb{R}^2}|\nabla v_n|^2\,{\rm d}\kern 0.06em x+\frac{1}{2}\int_{\mathbb{R}^2}V(x+y_n)v_n^2\,{\rm d}\kern 0.06em x+\frac{\nu}{4}V_0(v_n) -\int_{\mathbb{R}^2}F(v_n)\,{\rm d}\kern 0.06em x. \]

Based on the fact that $v_n\rightarrow v_0$ in $L^s(\mathbb {R}^2)$ for $s\geq 2$ as $n\rightarrow \infty$, by assumption (V2) and (3.2) one has

(5.22)\begin{equation} \int_{\mathbb{R}^2}V(x+y_n)v_n(v_n-v_0)\,{\rm d}\kern 0.06em x\rightarrow0,\quad \left|\frac{1}{4}V'_2(v_n)(v_n-v_0)\right|\rightarrow0, \quad\text{as}\ n\rightarrow\infty . \end{equation}

Arguing similarly as (5.16) and (5.17), we have

(5.23)\begin{equation} \int_{\mathbb{R}^2}f(v_n)(v_n-v_0)\,{\rm d}\kern 0.06em x\rightarrow0,\quad \text{as}\ n\rightarrow\infty. \end{equation}

By the definition of $\tilde {I}(v_n)$, we have

(5.24)\begin{align} & B_1(v_n^2,v_n(v_n-v_0))+\|\nabla (v_n-v_0)\|_2^2\nonumber\\ & = \tilde{I}'(v_n)(v_n-v_0)-\int_{\mathbb{R}^2}V(x+y_n)v_n(v_n-v_0)\,{\rm d}\kern 0.06em x\nonumber\\ & \quad+B_2(v_n^2,v_n(v_n-v_0))+\int_{\mathbb{R}^2}f(v_n)(v_n-v_0)\,{\rm d}\kern 0.06em x +o(1). \end{align}

Combining (5.21)–(5.24), we infer that $\|\nabla (v_n-v_0)\|_2^2\rightarrow 0$ and $B_1(v_n^2,v_n(v_n-v_0))\rightarrow 0$ as $n\rightarrow \infty$, and then $v_n\rightarrow v_0$ in $H^1(\mathbb {R}^2)$. Recalling lemma 3.1, we have $\|v_n-v_0\|_\star \rightarrow 0$ as $n\rightarrow \infty$. We therefore deduce that $v_n\rightarrow v_0$ in $X$.

Furthermore, similarly to the proof of lemma 4.5, we obtain that $v_0$ is a nontrivial critical point of functional $I_\infty$ and

\[ I_\infty(v_0)=\lim\limits_{n\rightarrow\infty}I(v_n)=c_{mp}, \]

where

\[ I_\infty(v_0):=\frac{1}{2}\int_{\mathbb{R}^2}|\nabla v_0|^2\,{\rm d}\kern 0.06em x+\frac{1}{2}\int_{\mathbb{R}^2}V_\infty v_0^2\,{\rm d}\kern 0.06em x+\frac{\nu}{4}V_0(v_0) -\int_{\mathbb{R}^2}F(v_0)\,{\rm d}\kern 0.06em x. \]

Using the conditions of theorem 2.5, like (5.9) we can also estimate

\[ I_\infty(u)\geq I_\infty(tu)+\frac{1-t^4}{4}I'_\infty(u)u, \quad \forall u\in X, t\geq0, \]

which yields

(5.25)\begin{equation} I_\infty(v_0)\geq \max_{t\geq0}I_\infty(tv_0). \end{equation}

Recalling the definition of $c_{mp}$, we have

\[ c_{mp}\leq \max_{t\in(0,+\infty)}I(t v_{0})<\max_{t\in(0,+\infty)}I_{\infty}(tv_{0})\leq I_{\infty}(v_{0})=c_{mp}, \]

which is a contradiction. Therefore, $\{y_n\}$ is a bounded sequence.

Step 3. We show that $u_n\rightarrow u$ in $X$ and then $u$ is critical point of $I$ with $I(u)=c_{mp}$. The proof is the same as that in lemma 4.5, and we omit it.

5.1 Proof of theorem 2.5

Similarly to theorem 2.2, let us define the set of solutions

\[ \mathcal{\tilde{S}}:=\{u\in X{\setminus}\{0\}:\,I'(u)=0\}. \]

By lemma 5.3, $\mathcal {\tilde {S}}\not =\emptyset$. For $u\in \mathcal {\tilde {S}}$ by ($f_0$), for any $\varepsilon >0$ there exists $C_\varepsilon >0$ such that

\begin{align*} \|u\|^2& \leq \|u\|^2+\nu V_1(u)\\ & \leq\varepsilon\int_{\mathbb{R}^2}u^2\,{\rm d}\kern 0.06em x+C_\varepsilon\int_{\mathbb{R}^2}({\rm e}^{\alpha_0 u^2}-1)u^3\,{\rm d}\kern 0.06em x+\nu V_2(u)\\ & \leq \varepsilon\|u\|^2+C_\varepsilon\left(\int_{\mathbb{R}^2}\,{\rm e}^{2\alpha_0 u^2}-1)\right)^{\frac{1}{2}}\|u\|_6^3+\nu C\|u\|^4\\ & \leq\varepsilon\|u\|^2+C_\varepsilon C\|u\|^3+\nu C\|u\|^4, \end{align*}

which implies that there exists $C>0$ such that $\|u\|\geq C$ for any $u\in \mathcal {\tilde {S}}$.

We claim that

(5.26)\begin{equation} \tilde{c}_*:=\inf\limits_{u\in\mathcal{\tilde{S}}}I(u)>0. \end{equation}

Assume by contradiction that $\tilde {c}_*=0$ and $\{u_n\}\subset \mathcal {\tilde {S}}$ satisfies $I(u_n)\rightarrow 0$ as $n\rightarrow \infty$. Recalling lemma 5.2, we deduce that $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$ uniformly for $n$, and then by (5.5), $I(u_n)\geq \frac {1}{4}\|\nabla u_n\|_2^2$. Obviously, $\|\nabla u_n\|_2^2\rightarrow 0$. By the Gagliardo–Nirenberg inequality, we have $u_n\rightarrow 0$ in $L^p(\mathbb {R}^2)$ for $p>2$. The Trudinger–Moser inequality implies $\int _{\mathbb {R}^2}f(u_n)u_n{\rm d}\kern 0.06em x=o(1)$ as $n\rightarrow \infty$. And so using $I'(u_n)=0$ and (3.3), we have $V_1(u_n)\rightarrow 0$ and $u_n\rightarrow 0$ in $H^1(\mathbb {R}^2)$ which is impossible. (5.26) holds true. It is easy to obtain from lemma 5.1 that $\tilde {c}_*<\frac {\pi }{\alpha _0}$.

Finally, let $\{u_n\}\subset \mathcal {\tilde {S}}$ be a minimizing sequence, hence $I(u_n)\rightarrow \tilde {c}_*\in \left (0,\frac {\pi }{\alpha _0}\right )$. By lemma 5.2, we know that $\{u_n\}$ is bounded in $H^1(\mathbb {R}^2)$. Similar to the proof of lemma 5.3, there exists $\tilde {u}_{*}\in X$ such that $u_n\rightarrow \tilde {u}_{*}$ in $X$ and $I'(\tilde {u}_{*})=0$ and $I(\tilde {u}_{*})=\tilde {c}_*$. Thus, $\tilde {u}_{*}$ is a positive ground state solution of problem (1.3). The proof is now complete.

Acknowledgements

Zhisu Liu is supported by the National Natural Science Foundation of China (No. 12226331; 12226325), and the Fundamental Research Funds for the Central Universities, China University of Geosciences (Wuhan, Grants CUG2106211 and CUGST2). The research of Vicenţiu D. Rădulescu was supported by a grant of the Romanian Ministry of Research, Innovation and Digitization, CNCS/CCCDI–UEFISCDI, project number PCE 137/2021, within PNCDI III. Jianjun Zhang was partially supported by the National Natural Science Foundation of China (No. 11871123) and Team Building Project for Graduate Tutors in Chongqing (JDDSTD201802).

References

Adachi, S. and Tanaka, K.. Trudinger type inequalities in $\mathbb {R}^N$ and their best exponents. Proc. Am. Math. Soc. 128 (2000), 20512057.CrossRefGoogle Scholar
Adimurthi, and Yadava, S.. Multiplicity results for semilinear elliptic equations in bounded domain of $\mathbb {R}^2$ involving critical exponent. Ann. Sc. Norm. Super. Pisa 17 (1990), 481504.Google Scholar
Albuquerque, F., Carvalho, J., Figueiredo, G. and Medeiros, E.. On a planar non-autonomous Schrödinger-Poisson system involving exponential critical growth. Calc. Var. PDE 60 (2021), 30.CrossRefGoogle Scholar
Alves, C., Cassani, D., Tarsi, C. and Yang, M.. Existence and concentration of ground state solutions for a critical nonlocal Schrödinger equation in $\mathbb {R}^2$. J. Differ. Equ. 261 (2016), 19331972.CrossRefGoogle Scholar
Alves, C. and Figueiredo, G.. Existence of positive solution for a planar Schrödinger-Poisson system with exponential growth. J. Math. Phys. 60 (2019), 011503.CrossRefGoogle Scholar
Avron, J., Herbst, I. and Simon, B.. Schrödinger operators with electromagnetic fields. III. Atoms in homogeneous magnetic field. Commun. Math. Phys. 79 (1981), 529572.CrossRefGoogle Scholar
Azzollini, A.. The planar Schrödinger-Poisson system with a positive potential. Nonlinearity 34 (2021), 57995820.CrossRefGoogle Scholar
Ball, J. and Marsden, J.. Quasiconvexity at the boundary, positivity of the second variation and elastic stability. Arch. Ration. Mech. Anal. 86 (1984), 251277.CrossRefGoogle Scholar
Benci, V. and Fortunato, D.. An eigenvalue problem for the Schrödinger-Maxwell equations. Topol. Methods Nonlinear Anal. 11 (1998), 283293.CrossRefGoogle Scholar
Byeon, J. and Wang, Z. Q.. Standing waves with a critical frequency for nonlinear Schrödinger equations. Arch. Ration. Mech. Anal. 165 (2002), 295316.CrossRefGoogle Scholar
Bucur, C., Cassani, D. and Tarsi, C.. Quasilinear logarithmic Choquard equations with exponential growth in $R^N$. J. Differ. Equ. 328 (2022), 261294.CrossRefGoogle Scholar
Battaglia, L. and Van Schaftingen, J.. Ground states for the Chquard equations with a sign-changing self-interaction potential. Z. Angew. Math. Phys. 69 (2018), 16.CrossRefGoogle Scholar
Cao, D.. Nontrivial solution of semilinear elliptic equation with critical exponent in $\mathbb {R}^2$. Commun. Partial Differ. Equ. 17 (1992), 407435.CrossRefGoogle Scholar
Cao, D., Dai, W. and Zhang, Y.. Existence and symmetry of solutions to $2$-D Schrödinger-Newton equations. Dyn. Partial Differ. Equ. 18 (2021), 113156.CrossRefGoogle Scholar
Cassani, D., Sani, F. and Tarsi, C.. Equivalent Moser type inequalities in $\mathbb {R}^2$ and the zero mass case. J. Funct. Anal. 267 (2014), 42364263.CrossRefGoogle Scholar
Cassani, D., Tavares, H. and Zhang, J.. Bose fluids and positive solutions to weakly coupled systems with critical growth in dimension two. J. Differ. Equ. 269 (2020), 23282385.CrossRefGoogle Scholar
Cassani, D. and Tarsi, C.. Schrödinger-Newton equations in dimension two via a Pohozaev-Trudinger log-weighted inequality. Calc. Var. Partial Differ. Equ. 60 (2021), 31.CrossRefGoogle Scholar
Cerami, G. and Vaira, G.. Positive solutions for some non-autonomous Schrödinger-Poisson systems. J. Differ. Equ. 248 (2010), 521543.CrossRefGoogle Scholar
Chen, S. and Tang, X.. On the planar Schrödinger-Poisson system with zero mass and critical exponential growth. Adv. Differ. Equ. 25 (2020), 687708.Google Scholar
Chen, S. and Tang, X.. Existence of ground state solutions for the planar axially symmetric Schrödinger-Poisson system. Discrete Contin. Dyn. Syst., Ser. B 24 (2019), 46854702.Google Scholar
Chen, S. and Tang, X.. On the planar Schrödinger-Poisson system with the axially symmetric potential. J. Differ. Equ. 268 (2020), 945976.CrossRefGoogle Scholar
Chen, S. and Tang, X.. Axially symmetric solutions for the planar Schrödinger-Poisson system with critical exponential growth. J. Differ. Equ. 269 (2020), 91449174.CrossRefGoogle Scholar
Cingolani, S. and Weth, T.. On the planar Schrödinger-Poisson systems. Ann. Inst. Henri Poincaré, Anal. Non Linéaire 33 (2016), 169197.CrossRefGoogle Scholar
de Figueiredo, D., Miyagaki, O. and Ruf, B.. Elliptic equations in $\mathbb {R}^2$ with nonlinearities in the critical growth range. Calc. Var. Partial Differ. Equ. 3 (1995), 139153.CrossRefGoogle Scholar
Du, L.. Bounds for subcritical best Sobolev constants in $W^1,p$. Commun. Pure Appl. Anal. 20 (2021), 38713886.CrossRefGoogle Scholar
Du, M. and Weth, T.. Ground states and high energy solutions of the planar Schrödinger-Poisson system. Nonlinearity 30 (2017), 34923515.CrossRefGoogle Scholar
Einstein, A.. Ideas and opinions (New York: Crown Trade Paperbacks, 1954).Google Scholar
He, X. and Zou, W.. Existence and concentration of ground states for Schrödinger-Poisson equations with critical growth. J. Math. Phys. 53 (2012), 023702.CrossRefGoogle Scholar
Ianni, I. and Ruiz, D.. Ground and bound states for a static Schrödinger-Poisson-Slater problem. Commun. Contemp. Math. 14 (2012), 1250003.CrossRefGoogle Scholar
Lieb, E. and Simon, B.. The Thomas-Fermi theory of atoms, molecules and solids. Adv. Math. 23 (1977), 22116.CrossRefGoogle Scholar
Lions, P.-L.. The concentration compactness principle in the calculus of variations: the locally compact case. Parts 2. Ann. Inst. H. Poincaré, Analyse Non Linéaire 2 (1984), 223283. Ann. Inst. H. Poincaré, Analyse Non Linéaire 2 (1984), 223–283.CrossRefGoogle Scholar
Lions, P.-L.. Solutions of Hartree-Fock equations for Coulomb systems. Commun. Math. Phys. 109 (1987), 3397.CrossRefGoogle Scholar
Liu, Z., Zhang, Z. and Huang, S.. Existence and nonexistence of positive solutions for a static Schrödinger-Poisson-Slater equation. J. Differ. Equ. 266 (2019), 59125941.CrossRefGoogle Scholar
Liu, Z., Lou, Y. and Zhang, J.. A perturbation approach to studying sign-changing solutions of Kirchhoff equations with a general nonlinearity. Ann. Mat. Pura Appl. 201 (2022), 12291255.CrossRefGoogle Scholar
Liu, Z., Rădulescu, V. D., Tang, C. and Zhang, J.. Another look at planar Schrödinger-Newton system. J. Differ. Equ. 328 (2022), 65104.CrossRefGoogle Scholar
Malomed, B.. Variational methods in nonlinear fiber optics and related fields. Progr. Opt. 43 (2002), 69191.Google Scholar
Marsden, J. and Hughes, T.. Mathematical foundations of elasticity (New York: Dover Edition, 1994).Google Scholar
Moser, J.. A sharp form of an inequality by N. Trudinger. Indiana Univ. Math. J. 20 (1970/71), 10771092.CrossRefGoogle Scholar
Nier, F.. A variational formulation of Schrödinger-Poisson systems in dimension $d\leq 3$. Commun. Partial Differ. Equ. 18 (1993), 11251147.CrossRefGoogle Scholar
do Ó, J. M. and Ruf, B.. On a Schrödinger equation with periodic potential and critical growth in $\mathbb {R}^2$. Nonlinear Differ. Equ. Appl. 13 (2006), 167192.Google Scholar
Onorato, M., Osborne, A. R., Serio, M. and Bertone, S.. Freak waves in random oceanic sea states. Phys. Rev. Lett. 86 (2001), 58315834.CrossRefGoogle ScholarPubMed
Pohozaev, S.. The Sobolev embedding in the special case $p = n$. In Proceedings of the Technical Scientific Conference on Advances of Scientific Research, Mathematics Sections, Moscow (1965), 158–170.Google Scholar
Rabinowitz, P.. On a class of nonlinear Schrödinger equations. Z. Angew. Math. Phys. 43 (1992), 270291.CrossRefGoogle Scholar
Ruiz, D.. The Schrödinger-Poisson equation under the effect of a nonlinear local term. J. Funct. Anal. 237 (2006), 655674.CrossRefGoogle Scholar
Schunck, F. and Mielke, E.. General relativistic boson stars. Classical Quantum Gravity 20 (2003), R301R356.CrossRefGoogle Scholar
Sennett, B., Hinderer, T., Steinhoff, J., Buonanno, A. and Ossokine, S.. Distinguishing boson stars from black holes and neutron stars from tidal interactions in inspiraling binary systems. Phys. Rev. D 96 (2017), 024002.CrossRefGoogle Scholar
Silva, E. and Vieira, G.. Quasilinear asymptotically periodic Schrödinger equations with critical growth. Calc. Var. Partial Differ. Equ. 39 (2010), 133.CrossRefGoogle Scholar
Stubbe, J.. Bound states of two-dimensional Schrödinger-Newton equations, e-print arXiv:0807.4059.Google Scholar
Sulem, C. and Sulem, P.-L.. The nonlinear Schrödinger equation. Self-focusing and wave collapse. Applied Mathematical Sciences, Vol. 139 (New York: Springer-Verlag, 1999).Google Scholar
Sun, J. and Ma, S.. Ground state solutions for some Schrödinger-Poisson systems with periodic potentials. J. Differ. Equ. 260 (2016), 21192149.CrossRefGoogle Scholar
Tang, X.. Non-Nehari manifold method for asymptotically periodic Schrödinger equations. Sci. China Math. 58 (2015), 715728.CrossRefGoogle Scholar
Tang, X. and Cheng, B.. Ground state sign-changing solutions for Kirchhoff type problems in bounded domains. J. Differ. Equ. 261 (2016), 23842402.CrossRefGoogle Scholar
Trudinger, N. S.. On imbeddings into Orlicz spaces and some applications. J. Math. Mech. 17 (1967), 473483.Google Scholar
Wang, Z. and Zhou, H.. Positive solution for a nonlinear stationary Schrödinger-Poisson system in $\mathbb {R}^3$. Discrete Contin. Dyn. Syst. 18 (2007), 809816.CrossRefGoogle Scholar
Wen, L., Chen, S. and Rădulescu, V. D.. Axially symmetric solutions of the Schrödinger-Poisson system with zero mass potential in $\mathbb {R}^2$. Appl. Math. Lett. 104 (2020), 106244.CrossRefGoogle Scholar
Zakharov, V. E.. Collapse and self-focusing of Langmuir waves. In Handbook of Plasma Physics, Vol.2, Basic Plasma Physics (eds. Galeev, A. A and Sudan, R. N.), pp. 81121 (Elsevier North-Holland, 1984).Google Scholar