1 Introduction and main results

Consider the following Schödinger–Poisson equations:

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u+\lambda V(x)u+K(x)\phi u=f(x,u)\ \ \ \ \ &{} \ \text{ in } \mathbb {R}^3,\\ -\Delta \phi =K(x)u^2\ \ \ \ \ \ &{} \ \text{ in } \mathbb {R}^3, \end{array}\right. \qquad \qquad \qquad (SP)_{\lambda } \end{aligned}$$

where \(\lambda \ge 1\) is a parameter, \(V\in C(\mathbb {R}^3,\mathbb {R})\) and \(f\in C(\mathbb {R}^3\times \mathbb {R},\mathbb {R})\).

Problem \((SP)_\lambda \) (also called Schrödinger–Maxwell equation) arises in applications from mathematical physics, such as in quantum electrodynamics, to describe the interaction of a charged particle with the electromagnetic field, and also in semiconductor theory, in nonlinear optics and in plasma physics. For more details in physical aspects, we refer to [9, 12].

There has been a vast literature on the study of existence and multiplicity of solutions of system \((SP)_\lambda \) under various hypotheses on the potential \(V(x)\) and the nonlinearity \(f(x,u)\), see [13, 5, 914, 18, 19, 21, 22, 2428, 31, 3437] and the references therein. Most of them dealt with the situation where \(V(x)\) is a positive constant or being radially symmetric and \(f(x,u)=|u|^{p-1}u\), \(1<p<5\). In [25] the case \(p=5/3\) was studied. The authors applied a minimization procedure in an appropriate manifold to find a positive solution (possibly non-radial) for system \((SP)_1\) (i.e. \((SP)_\lambda \) with \(\lambda =1\)). In [11, 12], a radial positive solution of \((SP)_1\) was obtained for \(3\le p<5\), by taking advantage of the mountain pass theorem due to Ambrosetti and Rabinowitz [4]. In [13], a related Pohozǎev identity was found, and with this in hand, the authors proved that problem \((SP)_1\) has no nontrivial solutions for \(p\le 1\) or \(p>5\). This result was completed in [24], where Ruiz showed that if \(p\le 2\), problem \((SP)_1\) does not admit any nontrivial solution, and if \(2<p<5\), there exists a positive radial solution of \((SP)_1\). Ambrosetti and Ruiz [2] and Ambrosetti [3] considered problem \((SP)_1\) with a parameter, i.e.,

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u+u+\lambda \phi u=|u|^{p-1}u\ \ \ \ \ \ &{} \text{ in } \mathbb {R}^3,\\ -\Delta \phi =u^2\ \ \ \ \ \ \ &{} \text{ in } \mathbb {R}^3. \end{array}\right. \ \ \ \ \ \ \ \qquad \qquad \qquad (A)_{\lambda } \end{aligned}$$

Using variational methods, they constructed the existence of infinitely many pairs of radial solutions of problem \((A)_\lambda \), where \(2<p<5\), for all \(\lambda >0\), and also multiple solutions (but not infinitely many) of \((A)_\lambda \), where \(1<p\le 2\), for \(\lambda >0\) small sufficiently. In addition, the existence of infinitely many non-radial solutions of system \((SP)_1\) was constructed in d’Avenia et al. [14], when \(1<p<5\) and \(K(x)\) is a positive radial function decaying at infinity. See also [5, 19, 34, 37] for the critical case.

The case of positive and non-radial potential \(V\) has been discussed in [10, 22, 26, 28, 31, 35]. In particular, supposing that \(V(x)\) satifies:

\((V_1)\) :

\(V\in C(\mathbb {R}^3, \mathbb {R})\) and \(\inf _{x\in \mathbb {R}^3}V(x)\ge a>0\), where \(a\) is a positive constant;

\((V_2)\) :

For any \(b>0\), meas\(\left\{ x\in \mathbb {R}^3:V(x)\le b\right\} <+\infty \), where meas denotes the Lebesgue measure in \(\mathbb {R}^3\);

[10, 22, 31] established the existence of infinitely many high-energy solutions of problem \((SP)_1\), where \(f\) is 4-superlinear at infinity, while the existence of infinitely many small-energy solutions was proved in Sun [26] with sublinear nonlinearity. The proofs in [10, 22, 31] were based on the (variant) fountain theorem. It is worth mentioning that conditions \((V_1)\)\((V_2)\) were first introduced by Bartsch and Wang [8] to guarantee the compact embedding of the functional space (see [8, Remark 3.5]). If replacing \((V_2)\) by a more general assumption:

\((V_3)\) :

There is \(b>0\) such that meas\(\left\{ x\in \mathbb {R}^3:V(x)\le b\right\} <+\infty \),

the compactness of the embedding fails and this situation becomes more delicated.

Recently, [32, 35] considered this case. Yang et al. [32] investigated the semiclassical solutions of the Schrödinger–Poisson equations

$$\begin{aligned} \left\{ \begin{array}{ll} -\varepsilon ^2\Delta u+ V(x)u+\phi u=f(x,u)\ \ \ \ \ &{} \ \text{ in }\mathbb {R}^3,\\ -\Delta \phi =4\pi u^2\ \ \ \ \ \ &{} \ \text{ in } \mathbb {R}^3. \end{array}\right. \ \ \ \ \ \qquad \qquad \qquad (B)_{\varepsilon } \end{aligned}$$

They assumed that \((V_3)\) holds, \(V(0)=\) min\(V=0\) and \(f(x,u)\) satisfies:

\((g_1)\) :

\(f(x,u)=o(u)\) as \(u\rightarrow 0\) uniformly in \(x\);

\((g_2)\) :

There are \(c_0>0\) and \(q<6\) such that \(|f(x,u)|\le c_0(1+|u|^{q-1})\) for all \((x,u)\);

\((g_3)\) :

There are \(a_0>0\), \(p>4\) and \(\mu >4\) such that \(F(x,u)\ge a_0|u|^p\) and \(\mu F(x,u)\le f(x,u)u\) for all \((x,u)\), where \(F(x,u):=\int _0^u f(x,s)ds\).

They showed that for any \(\sigma >0\) there exists \(\varepsilon _\sigma >0\) such that \((B)_\varepsilon \) has at least one solution when \(\varepsilon \le \varepsilon _\sigma \); and if additionally \(f(x,u)\) is odd in \(u\), then given any \(\varepsilon >0\) small enough \((B)_\varepsilon \) has at least \(m\) pairs of solutions. Zhao et al. [35] studied the existence of nontrivial solution and concentration results (as \(\lambda \rightarrow +\infty \)) of \((SP)_\lambda \), provided that \(V\) satisfies \((V_3)\) and

\((V_4)\) :

\(V\in C(\mathbb {R}^3,\mathbb {R})\) and \(V\) is bounded below,

\((V_5)\) :

\(\Omega =int V^{-1}(0)\) is nonempty and has smooth boundary and \(\overline{\Omega }=V^{-1}(0)\),

and \(f(x,u)=|u|^{p-2}u\) \((2<p<6)\).

We also note that if \(K\equiv 0\), \((SP)_\lambda \) reduces to the Schödinger equation

$$\begin{aligned} -\Delta u+\lambda V(x)u=f(x,u),\ \ \ \ \ \ \ \ x\in \mathbb {R}^N,\ \ \ \qquad \qquad \qquad \qquad (C)_{\lambda } \end{aligned}$$

which has been the object of interest for many authors, see e.g. [15, 16, 29] and their references. In [16], Ding and Szulkin studied the existence and the number of decaying solutions of problem \((C)_\lambda \) when \(V\) may change sign, satisfies \((V_4)\) and

\((V_6)\) :

There exists \(b>0\) such that the set \(\left\{ x\in \mathbb {R}^N:V(x)<b\right\} \) is nonempty and has finite measure;

and \(f\) is either asymptotically linear or superlinear (but subcritical) in \(u\) as \(|u|\rightarrow \infty \). Wang and Zhou [29] dealt with the ground states of problem \((C)_\lambda \), where \(V(x)\) changes sign and may vanish at infinity, \(f(x,u)=K_1(x)g(u)\) and \(g\) is either of the form \(g(u)=|u|^{p-1}u\) with \(1<p<\frac{N+2}{N-2}\) or asymptotically linear.

Motivated by the works mentioned above, in the present paper, we are mostly interested in sign-changing potentials though in a few cases we need to have \(V\ge 0\). Under \((V_3)\)\((V_4)\) and some more generic 4-superlinear conditions on \(f(x,u)\), we prove the existence and multiplicity of solutions of problem \((SP)_\lambda \) when \(\lambda >0\) large, using variational method. Furthermore, we investigate the situation where the nonlinearity \(f(x,u)\) is sublinear with mild assumptions different from those studied previously. Infinitely many small-energy solutions are obtained for problem \((SP)_1\) by applying a new version of symmetric mountain pass lemma developed by Kajikiya. The main results are the following theorems.

First, we handle the 4-superlinear case, and hence make the following assumptions:

\((f_1)\) :

\(F(x,u)\ge 0\) for all \((x,u)\) and \(f(x,u)=o(u)\) uniformly in \(x\) as \(u\rightarrow 0\).

\((f_2)\) :

\(F(x,u)/u^4\rightarrow +\infty \) as \(|u|\rightarrow \infty \) uniformly in \(x\).

\((f_3)\) :

\( \mathcal {F}(x,u):=\frac{1}{4}f(x,u)u-F(x,u)\ge 0\) for all \((x,u)\in \mathbb {R}^3\times \mathbb {R}\).

\((f_4)\) :

There exist \(a_1\), \(L_1>0\) and \(\tau \in (3/2,2)\) such that

$$\begin{aligned} |f(x,u)|^\tau \le a_1\mathcal {F}(x,u)|u|^\tau ,\ \ \ \ \ \ \forall x\in \mathbb {R}^3,\ \ |u|\ge L_1. \end{aligned}$$
\((K)\) :

\(K\in L^2(\mathbb {R}^3)\cup L^\infty (\mathbb {R}^3)\) and \(K(x)\ge 0\) for all \(x\in \mathbb {R}^3\).

Remark 1.1

It follows from \((f_2)\) and \((f_4)\) that \(|f(x,u)|^\tau \le \frac{a_1}{4}|f(x,u)||u|^{\tau +1}\) for large \(u\). Thus, by (\(f_1\)), for any \(\varepsilon >0\), there is \(C_\varepsilon >0\) such that

$$\begin{aligned} |f(x,u)|\le \varepsilon |u|+C_\varepsilon |u|^{p-1},\ \ \ \ \ \ \ \forall (x,u)\in \mathbb {R}^3\times \mathbb {R} \end{aligned}$$
(1.1)

and

$$\begin{aligned} |F(x,u)|\le \varepsilon u^2+C_\varepsilon |u|^p,\ \ \ \ \ \ \forall (x,u)\in \mathbb {R}^3\times \mathbb {R}, \end{aligned}$$

where \(p=2\tau /(\tau -1)\in (4,2^*)\), \(2^*=6\) is the critical exponent for the Sobolev embedding in dimension 3.

Theorem 1.1

(Superlinear) Assume that \((V_3)\)\((V_4)\), \((K)\) and \((f_1)\)\((f_4)\) are satisfied.

  1. (i)

    If \(V(x)<0\) for some \(x\in \mathbb {R}^3\), then for each \(k\in \mathbb {N}\), there exist \(\lambda _k>k\) and \(b_k>0\) such that problem \((SP)_\lambda \) has a nontrivial solution \((u_\lambda ,\phi _\lambda )\in H^1(\mathbb {R}^3)\times \mathcal {D}^{1,2}(\mathbb {R}^3)\) for every \(\lambda =\lambda _k\) and \(|K|_2<b_k\) (or \(|K|_\infty <b_k\)).

  2. (ii)

    If \(V^{-1}(0)\) has nonempty interior, then there exist \(\Lambda >0\) and \(b_\lambda >0\) such that problem \((SP)_\lambda \) has a nontrivial solution \((u_\lambda ,\phi _\lambda )\in H^1(\mathbb {R}^3)\times \mathcal {D}^{1,2}(\mathbb {R}^3)\) for every \(\lambda >\Lambda \) and \(|K|_2<b_\lambda \) (or \(|K|_\infty <b_\lambda \)).

Remark 1.2

Theorem 1.1 (ii) generalizes [35, Theorem 1.1], which is the special case of Theorem 1.1 (ii) corresponding to \(f(x,u)=|u|^{p-2}u\) \((4<p<6)\).

If \(V\ge 0\), the restriction on the norm of \(K\) can be removed and we have the following theorem.

Theorem 1.2

(Superlinear) Assume that \(V\ge 0\), \((V_3)\)\((V_4)\), \((K)\) and \((f_1)\)\((f_4)\) are satisfied, and \(V^{-1}(0)\) has nonempty interior \(\Omega \). Then there exist \(\Lambda _*>0\) such that problem \((SP)_\lambda \) has at least one nontrivial solution \((u_\lambda ,\phi _\lambda )\in H^1(\mathbb {R}^3)\times \mathcal {D}^{1,2}(\mathbb {R}^3)\) whenever \(\lambda >\Lambda _*\). Moreover, if \(f\) is odd in \(t\), then for each \(k\ge 1\) there exists \(\Lambda _k>0\) such that problem \((SP)_\lambda \) has at least \(k\) pairs of nontrivial solutions whenever \(\lambda >\Lambda _k\).

Remark 1.3

Theorem 1.2 can be viewed as an improvement of the results in Yang et al. [32] and Zhao et al. [35]. Comparing with [32, Theorems 1.1 and 1.2], our hypotheses on \(f\) are much weaker. Indeed, assumption \((g_3)\) implies

$$\begin{aligned} 0<\mu F(x,u)\le f(x,u)u \ \ \ \ \text{ for } \text{ some } \mu >4 \text{ and } \text{ all } (x,u) \text{ with } u\ne 0. \end{aligned}$$

So, if \(f\) satisfies \((g_1)\) and \((g_3)\), it is easy to see that \((f_2)\)\((f_3)\) hold, and it will be showed as in the proof of [16, Lemma 2.2 \((i)\)] that so does \((f_4)\). As for [35], we consider a larger class of nonlinearities and discuss the multiplicity result.

Remark 1.4

There are functions \(f\) which match conditions \((f_1)\)\((f_4)\) but not satisfying the results in [32, 35]. For example, let

$$\begin{aligned} f(x,t)=h(x)t^3\left( 2\ln (1+t^2)+\frac{t^2}{1+t^2}\right) ,\ \ \ \ \ \forall (x,t)\in \mathbb {R}^3\times \mathbb {R}, \end{aligned}$$

where \(h\) is a continuous bounded function with \(\inf _{x\in \mathbb {R}^3}h(x)>0\).

Next, we treat the sublinear case. Assume that:

\((f_5)\) :

There exist constants \(\sigma \), \(\gamma \in (1,2)\) and functions \(m\in L^{2/(2-\sigma )}(\mathbb {R}^3,\mathbb {R}^+)\), \(h\in L^{2/(2-\gamma )}(\mathbb {R}^3,\mathbb {R}^+)\) such that

$$\begin{aligned} |f(x,u)|\le m(x)|u|^{\sigma -1}+h(x)|u|^{\gamma -1},\ \ \ \ \ \ \forall (x,u)\in \mathbb {R}^3\times \mathbb {R}. \end{aligned}$$
\((f_6)\) :

There exist \(x_0\in \mathbb {R}^3\), two sequences \(\left\{ \varepsilon _n\right\} \), \(\left\{ M_n\right\} \) and constants \(a_2\), \(\varepsilon \), \(\delta >0\) such that \(\varepsilon _n>0\), \(M_n>0\) and

$$\begin{aligned}&\lim _{n\rightarrow \infty }\varepsilon _n=0,\ \ \ \ \ \lim _{n\rightarrow \infty }M_n=+\infty ,\nonumber \\&\varepsilon _n^{-2}F(x,u)\ge M_n\ \ \ \ \ \ \ \text{ for } |x-x_0|\le \delta \text{ and } |u|=\varepsilon _n, \nonumber \\&F(x,u)\ge -a_2u^2\ \ \ \ \ \ \ \text{ for } |x-x_0|\le \delta \text{ and } |u|\le \varepsilon . \end{aligned}$$
(1.2)

Theorem 1.3

(Sublinear) Assume that \(V\ge 0\), \((V_3)\), \((K)\) and \((f_5)\)\((f_6)\) are satisfied and that \(f(x,u)\) is odd in \(u\). Then problem \((SP)_1\) possesses infinitely many nontrivial solutions \(\left\{ (u_k,\phi _k)\right\} \) such that

$$\begin{aligned} \frac{1}{2}\int \limits _{\mathbb {R}^3}(|\nabla u_k|^2+V(x)u_k^2)dx+\frac{1}{4}\int \limits _{\mathbb {R}^3}K(x)\phi _k u_k^2dx-\int \limits _{\mathbb {R}^3}F(x,u_k)dx\rightarrow 0^-\ \ \ \text{ as } \ \ k\rightarrow \infty . \end{aligned}$$

Remark 1.5

In Sun [26], the existence of infinitely many small-energy solutions was obtained for \((SP)_1\), where \(K\equiv 1\), under assumptions \((V_1)\)\((V_2)\) and:

\((f')\) :

\(f(x,u)=b(x)|u|^{\sigma -1}\), where \(b:\mathbb {R}^3\rightarrow \mathbb {R}^+\) is a positive continuous function such that \(b\in L^{2/(2-\sigma )}(\mathbb {R}^3,\mathbb {R})\) and \(1<\sigma <2\) is a constant.

Observing \((f')\) implies that there is an open set \(J\subset \mathbb {R}^3\) such that

$$\begin{aligned} F(x,t)/t^2\rightarrow +\infty \ \ \ \ \ \ \ \text{ as } t\rightarrow 0\ \ \text{ uniformly } \text{ for } x\in J, \end{aligned}$$

it is stronger than \((f_5)\)\((f_6)\). Hence Theorem 1.3 improves [26, Theorem 1.1] by weakening hypotheses on \(V\), \(K\) and \(f\). There are functions \(V\), \(K\) and \(f\) which match our setting but not satisfying the results in [21, 26]. For example, let

$$\begin{aligned} V\equiv c(>0),\ \ \ \ \ \ \ \ K(x)=|x|^{-4}, \end{aligned}$$

and

$$\begin{aligned} f(x,u)=\left\{ \begin{array}{ll} |x|e^{-|x|^2}\left[ \sigma |u|^{\sigma -2}u\sin ^2\left( \frac{1}{|u|^\varrho }\right) -\varrho |u|^{\sigma -\varrho -2}\sin \left( \frac{2}{|u|^\varrho }\right) \right] , \ \ \ \ &{} t\ne 0,\\ 0,\ \ \ \ &{} t=0, \end{array} \right. \end{aligned}$$

where \(\varrho >0\) small enough and \(\sigma \in (1+\varrho ,2)\). Simple calculation shows that

$$\begin{aligned} F(x,u)=\left\{ \begin{array}{ll} |x|e^{-|x|^2}|u|^\sigma \sin ^2\left( \frac{1}{|u|^\varrho }\right) , \ \ \ \ &{} t\ne 0,\\ 0,\ \ \ \ &{} t=0. \end{array} \right. \end{aligned}$$

It is easy to check that \((V_3)\)\((V_4)\), \((K)\) and \((f_5)\)\((f_6)\) are satisfied with \(\varepsilon _n=\left( \frac{2}{(2n+1)\pi }\right) ^{1/\varrho }\). However, in this case, \((V_2)\) and \((f')\) fail.

The paper is organized as follows. In Sect. 2 we introduce the variational setting and recall some related preliminaries. Section 3 is concerned with the 4-superlinear case and Sect. 4 with the sublinear case. In Sect. 5, concentration of solutions to problem \((SP)_\lambda \) on the set \(V^{-1}(0)\) as \(\lambda \rightarrow +\infty \) is discussed.

Notation

  • \(H^1(\mathbb {R}^3)\) is the usual Sobolev space endowed with the standard scalar and norm

    $$\begin{aligned} ( u,v)_{H^1}=\int \limits _{\mathbb {R}^3}(\nabla u\cdot \nabla v+uv)dx;\ \ \ \ \ \ \Vert u\Vert _{H^1}=(u,u)_{H^1}^{1/2}. \end{aligned}$$
  • \(\mathcal {D}^{1,2}(\mathbb {R}^3)\) is the completion of \(C_0^\infty (\mathbb {R}^3)\) with respect to the norm \(\Vert u\Vert _{\mathcal {D}^{1,2}}^2:=\int _{\mathbb {R}^3}|\nabla u|^2dx.\)

  • \(L^s(\Omega )\), \(1\le s\le +\infty \), \(\Omega \subset \mathbb {R}^3\), denotes a Lebesgue space; the norm in \(L^s(\Omega )\) is denoted by \(|u|_{s,\Omega }\), where \(\Omega \) is a proper subset of \(\mathbb {R}^3\), by \(|\cdot |_s\) when \(\Omega =\mathbb {R}^3\).

  • \(\bar{S}\) is the best Sobolev constant for the Sobolev embedding \(\mathcal {D}^{1,2}(\mathbb {R}^3)\hookrightarrow L^6(\mathbb {R}^3)\), i.e.,

    $$\begin{aligned} \bar{S}=\inf _{u\in H^1(\mathbb {R}^3)\backslash \left\{ 0\right\} }\frac{\Vert u\Vert _{\mathcal {D}^{1,2}}}{|u|_6}. \end{aligned}$$
  • For any \(r>0\) and \(z\in \mathbb {R}^3\), \(B_r(z)\) denotes the ball of radius \(r\) centered at \(z\).

  • The letter \(c\) will be used to denote various positive constants which may vary from line to line and are not essential to the problem.

2 Variational setting and preliminaries

Let

$$\begin{aligned} E:=\left\{ u\in H^1(\mathbb {R}^3):\int \limits _{\mathbb {R}^3}V^+(x)u^2dx<+\infty \right\} , \end{aligned}$$

where \(V^\pm (x)=\max \left\{ \pm V(x),0\right\} \). Then \(E\) is a Hilbert space with the inner product and norm

$$\begin{aligned} (u,v)=\int \limits _{\mathbb {R}^3}(\nabla u\cdot \nabla v+V^+(x)uv)dx,\ \ \ \ \ \ \Vert u\Vert =(u,u)^{1/2}. \end{aligned}$$

We also need the following inner product

$$\begin{aligned} (u,v)_\lambda =\int \limits _{\mathbb {R}^3}(\nabla u\cdot \nabla v+\lambda V^+(x)uv)dx, \end{aligned}$$

and the corresponding norm is denoted by \(\Vert u\Vert _\lambda =( u,u)_\lambda ^{1/2}\) (so \(\Vert \cdot \Vert =\Vert \cdot \Vert _1\)). Set \(E_\lambda =(E,\Vert \cdot \Vert _\lambda )\). It follows from \((V_3)\), \((V_4)\) and the Poincaré inequality that the embedding \(E_\lambda \hookrightarrow H^1(\mathbb {R}^3)\) is continuous, and hence, for \(s\in [2,2^*]\), there exists \(\nu _s>0\) (independent of \(\lambda \)) such that

$$\begin{aligned} |u|_s\le \nu _s\Vert u\Vert _\lambda ,\ \ \ \ \ \ \ \forall u\in E_\lambda . \end{aligned}$$
(2.1)

Let

$$\begin{aligned} F_\lambda :=\left\{ u\in E_\lambda :\text{ supp }u\subset V^{-1}([0,+\infty ))\right\} , \end{aligned}$$

and \(F_\lambda ^\bot \) denote the orthogonal complement of \(F_\lambda \) in \(E_\lambda \). Clearly, \(F_\lambda =E_\lambda \) if \(V\ge 0\), otherwise \(F_\lambda ^\bot \ne \left\{ 0\right\} \). Define

$$\begin{aligned} A_\lambda :=-\Delta +\lambda V, \end{aligned}$$

then \(A_\lambda \) is formally self-adjoint in \(L^2(\mathbb {R}^3)\) and the associated bilinear form

$$\begin{aligned} a_\lambda (u,v):=\int \limits _{\mathbb {R}^3}(\nabla u\cdot \nabla v+\lambda V(x)uv)dx \end{aligned}$$

is continuous in \(E_\lambda \). As in [16], we consider the eigenvalue problem

$$\begin{aligned} -\Delta u+\lambda V^+(x)u=\mu \lambda V^-(x)u,\ \ \ \ \ \ \ u\in F_\lambda ^\bot . \end{aligned}$$
(2.2)

In view of \((V_3)\)\((V_4)\), the functional \(I(u)=\int _{\mathbb {R}^3}V^-(x)u^2dx\) for \(u\in F_\lambda ^\bot \) is weakly continuous. Hence, as a result of [30, Theorems 4.45 and 4.46], we deduce the following proposition, which is the spectral theorem for compact self-adjoint operators jointly with the Courant-Fischer minimax characterization of eigenvalues.

Proposition 2.1

Assume that \((V_3)\)\((V_4)\) hold, then for any fixed \(\lambda >0\), problem (2.2) has a sequence of positive eigenvalues \(\left\{ \mu _j(\lambda )\right\} _{j=1}^\infty \), which may be characterized by

$$\begin{aligned} \mu _j(\lambda )=\inf _{\text{ dim }M\ge j,M\subset F_\lambda ^\bot }\sup \left\{ \Vert u\Vert _\lambda ^2:u\in M,\int \limits _{\mathbb {R}^3}\lambda V^-(x)u^2dx=1\right\} ,\ \ \ \ \ j=1,2,\dots . \end{aligned}$$

Furthermore, \(\mu _1(\lambda )\le \mu _2(\lambda )\le \dots \le \mu _j(\lambda )\mathop {\longrightarrow }\limits ^{j} +\infty \) and the corresponding eigenfunctions \(\left\{ e_j(\lambda )\right\} _{j=1}^\infty \), which may be chosen so that \((e_i(\lambda ),e_j(\lambda ))_\lambda =\delta _{ij}\), are a basis of \(F_\lambda ^\bot \).

For the eigenvalues \(\left\{ \mu _j(\lambda )\right\} \) defined above, we have the following properties.

Proposition 2.2

(see Lemma 2.1 in [16]) Assume that \((V_3)\)\((V_4)\) hold and \(V^-\not \equiv 0\). Then, for each fixed \(j\in \mathbb {N}\),

  1. (i)

    \(\mu _j(\lambda )\rightarrow 0\) as \(\lambda \rightarrow +\infty \).

  2. (ii)

    \(\mu _j(\lambda )\) is a non-increasing continuous function of \(\lambda \).

Remark 2.1

By Proposition 2.2 \((i)\), there exists \(\Lambda _0>0\) such that \(\mu _1(\lambda )\le 1\) for all \(\lambda > \Lambda _0\).

Take

$$\begin{aligned} E_\lambda ^-:=\text{ span }\left\{ e_j(\lambda ):\mu _j(\lambda )\le 1\right\} \ \ \ \ \ \text{ and }\ \ \ \ \ E_\lambda ^+:=\text{ span }\left\{ e_j(\lambda ):\mu _j(\lambda )>1\right\} . \end{aligned}$$

Then we have the following orthogonal decomposition:

$$\begin{aligned} E_\lambda =E_\lambda ^-\bigoplus E_\lambda ^+\bigoplus F_\lambda . \end{aligned}$$

From Remark 2.1, we have that dim\(E_\lambda ^-\ge 1\) when \(\lambda >\Lambda _0\). Moreover, dim\(E_\lambda ^-<+\infty \) for every fixed \(\lambda >0\) since \(\mu _j(\lambda )\mathop {\longrightarrow }\limits ^{j} +\infty \).

It is well known that problem \((SP)_\lambda \) can be transformed into a Schrödinger equation with a nonlocal term (see e.g. [24]). Indeed, the Lax-Milgram theorem implies that for all \(u\in E_\lambda \), there exists a unique \(\phi _u\in \mathcal {D}^{1,2}(\mathbb {R}^3)\), which can be expressed as \(\phi _u(x)=\frac{1}{4\pi }\int _{\mathbb {R}^3}\frac{K(y)u^2(y)}{|x-y|}dy\), satisfying

$$\begin{aligned} -\Delta \phi _u=K(x)u^2. \end{aligned}$$
(2.3)

If \(K\in L^\infty (\mathbb {R}^3)\), by Hölder and Sobolev inequality, we get

$$\begin{aligned} \Vert \phi _u\Vert _{\mathcal {D}^{1,2}}^2=\int \limits _{\mathbb {R}^3}K(x)\phi _u u^2dx\le \bar{S}^{-2}\nu _{12/5}^4|K|_\infty ^2\Vert u\Vert _\lambda ^4. \end{aligned}$$

Similarly, if \(K\in L^2(\mathbb {R}^3)\),

$$\begin{aligned} \Vert \phi _u\Vert _{\mathcal {D}^{1,2}}^2=\int \limits _{\mathbb {R}^3}K(x)\phi _u u^2dx\le \bar{S}^{-2}\nu _6^4|K|_2^2\Vert u\Vert _\lambda ^4. \end{aligned}$$

Thus, there exists \(C_0>0\) such that

$$\begin{aligned} \Vert \phi _u\Vert _{\mathcal {D}^{1,2}}^2=\int \limits _{\mathbb {R}^3}K(x)\phi _u u^2dx\le C_0 \Vert u\Vert _\lambda ^4,\ \ \ \ \ \ \ \ \ \forall K\in L^2(\mathbb {R}^3)\cup L^\infty (\mathbb {R}^3). \end{aligned}$$
(2.4)

Take

$$\begin{aligned} N(u)=\int \limits _{\mathbb {R}^3}K(x)\phi _u u^2dx=\frac{1}{4\pi }\int \int \limits _{\mathbb {R}^3\times \mathbb {R}^3}\frac{K(x)K(y)u^2(x)u^2(y)}{|x-y|}dxdy \end{aligned}$$

We recall some important properties of the functional \(N\).

Lemma 2.1

Let \(K\in L^\infty (\mathbb {R}^3)\cup L^2(\mathbb {R}^3)\). If \(u_n\rightharpoonup u\) in \(H^1(\mathbb {R}^3)\) and \(u_n(x)\rightarrow u(x)\) a.e. \(x\in \mathbb {R}^3\), then

  1. (i)

    \(\phi _{u_n}\rightharpoonup \phi _u\) in \(\mathcal {D}^{1,2}(\mathbb {R}^3)\), and \(N(u)\le \liminf _{n\rightarrow \infty }N(u_n)\);

  2. (ii)

    \(N(u_n-u)=N(u_n)-N(u)+o(1)\);

  3. (iii)

    \(N'(u_n-u)=N'(u_n)-N'(u)+o(1)\) in \(H^{-1}(\mathbb {R}^3)\).

Proof

A straightforward adaption of [37, Lemma 2.1] shows that (i) holds. If \(K\equiv 1\), the proofs of (ii) and (iii) have been given in [36], and it is easy to see that the conclusions remain valid if \(K\in L^\infty (\mathbb {R}^3)\). Hence we only consider the case \(K\in L^2(\mathbb {R}^3)\).

We claim that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(K(x)\phi _{u_n}u_n^2-K(x)\phi _uu^2)dx\mathop {\longrightarrow }\limits ^{n}0 \end{aligned}$$
(2.5)

and

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(K(x)\phi _{u_n}u_n\psi -K(x)\phi _uu\psi )dx\mathop {\longrightarrow }\limits ^{n}0 \end{aligned}$$
(2.6)

uniformly for \(\psi \in H^1(\mathbb {R}^3)\) with \(\Vert \psi \Vert _{H^1}\le 1\). It follows from (i) and Hölder’s inequality that

$$\begin{aligned}&\lim _{n\rightarrow \infty }\int \limits _{\mathbb {R}^3}(K(x)\phi _{u_n}u_n^2-K(x)\phi _uu^2)dx\nonumber \\&\quad \le \lim _{n\rightarrow \infty }\int \limits _{\mathbb {R}^3}\left[ K(x)\phi _{u_n}(u_n^2-u^2)+K(x)(\phi _{u_n}-\phi _u)u^2\right] dx\nonumber \\&\quad \le \lim _{n\rightarrow \infty }|\phi _{u_n}|_6|u_n+u|_6|K(x)(u_n-u)|_{3/2}\nonumber \\&\qquad +\lim _{n\rightarrow \infty }\int _{\mathbb {R}^3}K(x)u^2(\phi _{u_n}-\phi _u)dx. \end{aligned}$$
(2.7)

The first limit on the right is 0 by the fact \(K^{3/2}\in L^{4/3}(\mathbb {R}^3)\) and \((u_n-u)^{3/2}\rightharpoonup 0\) in \(L^4(\mathbb {R}^3)\), and so is the second limit because \((\phi _{u_n}-\phi _u)\rightharpoonup 0\) in \(L^6(\mathbb {R}^3)\) and \(K(x)u^2\in L^{6/5}(\mathbb {R}^3)\). Thus (2.5) holds. Moreover, observing that \(|K(x)u|^{6/5}\in L^{5/4}(\mathbb {R}^3)\) and \((\phi _{u_n}-\phi _u)^{6/5}\rightharpoonup 0\) in \(L^5(\mathbb {R}^3)\), we obtain

$$\begin{aligned}&\int \limits _{\mathbb {R}^3}(K(x)\phi _{u_n}u_n\psi -K(x)\phi _uu\psi )dx\nonumber \\&\quad \le \int \limits _{\mathbb {R}^3}\left[ K(x)\phi _{u_n}(u_n-u)\psi +K(x)(\phi _{u_n}-\phi _u)u\psi \right] dx\\&\quad \le |\phi _{u_n}|_6|\psi |_6|K(x)(u_n-u)|_{3/2}+|\psi |_6|K(x)u(\phi _{u_n}-\phi _u)|_{6/5}\\&\quad \le c|K(x)(u_n-u)|_{3/2}+c|K(x)u(\phi _{u_n}-\phi _u)|_{6/5}\\&\quad \rightarrow 0 \end{aligned}$$

uniformly with respect to \(\psi \), i.e., (2.6) is satisfied. Now (ii) and (iii) follow from (2.5) and (2.6), respectively. \(\square \)

By (1.1) and the above lemma, the functional \(\varphi _\lambda : E_\lambda \rightarrow \mathbb {R}\),

$$\begin{aligned} \varphi _\lambda (u)=\frac{1}{2}\int \limits _{\mathbb {R}^3}(|\nabla u|^2+\lambda V(x)u^2)dx+\frac{1}{4}\int \limits _{\mathbb {R}^3}K(x)\phi _uu^2dx-\int \limits _{\mathbb {R}^3}F(x,u)dx, \end{aligned}$$

is of class \(C^1\) with derivative

$$\begin{aligned} \langle \varphi _\lambda '(u),v\rangle =\int \limits _{\mathbb {R}^3}(\nabla u\cdot \nabla v+\lambda V(x)uv+K(x)\phi _uuv-f(x,u)v)dx \end{aligned}$$

for all \(u\), \(v\in E_\lambda \). It can be proved that the pair \((u,\phi )\in E_\lambda \times \mathcal {D}^{1,2}(\mathbb {R}^3)\) is a solution of problem \((SP)_\lambda \) if and only if \(u\in E_\lambda \) is a critical point of \(\varphi _\lambda \) and \(\phi =\phi _u\) (see [9]).

To conclude this section, we state the following propositions, which will be applied to prove Theorems 1.1–1.3. Recall that a \(C^1\) functional \(I\) satisfies Cerami condition at level \(c\) (\((C)_c\) condition for short) if any sequence \((u_n)\subset E\) such that \(I(u_n)\rightarrow c\) and \((1+\Vert u_n\Vert )\Vert I'(u_n)\Vert \rightarrow 0\) has a converging subsequence; such a sequence is then called a \((C)_c\) sequence.

Proposition 2.3

(see [17]) Let \(E\) be a real Banach space and \(I\in C^1(E,\mathbb {R})\) satisfying

$$\begin{aligned} \max \left\{ I(0),I(e)\right\} \le a<b\le \inf _{\Vert u\Vert =\rho }I(u) \end{aligned}$$

for some \(a<b\), \(\rho >0\) and \(e\in E\) with \(\Vert e\Vert >\rho \). Let \(c\ge b\) be characterized by

$$\begin{aligned} c=\inf _{\gamma \in \Gamma }\max _{t\in [0,1]}I(\gamma (t)), \end{aligned}$$

where \(\Gamma =\left\{ \gamma \in C([0,1],E):\gamma (0)=0,\gamma (1)=e\right\} \) is the set of continuous paths jointing 0 and \(e\), then \(I\) possesses a \((C)_c\) sequence.

If \(V(x)\) is sign-changing, we need the following linking theorem.

Proposition 2.4

(see [23]) Let \(E=X\bigoplus Y\) be a Banach space with \(\dim Y<+\infty \), \(I\in C^1(E,\mathbb {R})\). If there exist \(R>\rho >0\), \(\alpha >0\) and \(e_0\in X\) such that

$$\begin{aligned} \alpha :=\inf I(X\cap S_\rho )>\sup I(\partial Q) \end{aligned}$$

where \(S_\rho =\left\{ u\in E:\Vert u\Vert =\rho \right\} \), \(Q=\left\{ u=v+te_0:v\in Y, t\ge 0, \Vert u\Vert \le R\right\} \). Then \(I\) has a \((C)_c\) sequence with \(c\in [\alpha ,\sup I(Q)]\).

Proposition 2.5

(see [6]) Suppose that \(I\in C^1(E,\mathbb {R})\) is even, \(I(0)=0\) and there exist closed subspaces \(E_1\), \(E_2\) such that codim\(E_1<+\infty \), \(\inf I(E_1\cap S_\rho )\ge \alpha \) for some \(\rho \), \(\alpha >0\) and \(\sup I(E_2)<+\infty \). If \(I\) satisfies the \((C)_c\)-condition for all \(c\in [\alpha ,\sup I(E_2)]\), then \(I\) has at least \(\dim E_2-\)codim\(E_1\) pairs of critical points with corresponding critical values in \([\alpha ,\sup I(E_2)]\).

To establish the existence of infinitely many solutions in the sublinear case, we require the new version of symmetric mountain pass lemma of Kajikiya (see [20]). Let \(E\) be a Banach space and

$$\begin{aligned} \Gamma :=\left\{ A\subset E\backslash \left\{ 0\right\} : A \text{ is } \text{ closed } \text{ and } \text{ symmetric } \text{ with } \text{ respect } \text{ to } \text{ the } \text{ origin }\right\} . \end{aligned}$$

We define

$$\begin{aligned} \Gamma _k:=\left\{ A\in \Gamma :\gamma (A)\ge k\right\} , \end{aligned}$$

where \(\gamma (A):=\inf \left\{ m\in \mathbb {N}:\exists h\in C(A,\mathbb {R}^m\backslash \left\{ 0\right\} ), -h(x)=h(-x)\right\} \). If there is no such mapping \(h\) for any \(m\in \mathbb {N}\), we set \(\gamma (A)=+\infty \).

Proposition 2.6

(Symmetric mountain pass lemma) Let \(E\) be an infinite dimensional Banach space and \(I\in C^1(E,\mathbb {R})\) be even, \(I(0)=0\) and satisfies the following conditions:

  1. (i)

    \(I\) is bounded from below and satisfies the Palais-Smale condition (PS), i.e., \((u_n)\subset E\) has a converging subsequence whenever \(\left\{ I(u_n)\right\} \) is bounded and \(I'(u_n)\rightarrow 0\) as \(n\rightarrow \infty \).

  2. (ii)

    For each \(k\in \mathbb {N}\), there exists an \(A_k\in \Gamma _k\) such that \(\sup _{u\in A_k}I(u)<0\).

Then either (1) or (2) holds.

  1. (1)

    There exists a sequence \(\left\{ u_k\right\} \) such that \(I'(u_k)=0\), \(I(u_k)<0\) and \(\left\{ u_k\right\} \) converges to zero.

  2. (2)

    There exist two sequence \(\left\{ u_k\right\} \) and \(\left\{ v_k\right\} \) such that \(I'(u_k)=0\), \(I(u_k)=0\), \(u_k\ne 0\), \(\lim _{k\rightarrow \infty }u_k=0\), \(I'(v_k)=0\), \(I(v_k)<0\), \(\lim _{k\rightarrow \infty }I(v_k)=0\) and \(\left\{ v_k\right\} \) converges to a non-zero limit.

Remark 2.2

From Proposition 2.6, we deduce a sequence \(\left\{ u_k\right\} \) of critical points such that \(I(u_k)\le 0\), \(u_k\ne 0\) and \(\lim _{k\rightarrow \infty }u_k=0\).

3 Proofs of Theorems 1.1–1.2

We first discuss the \((C)_c\) sequence. We only consider the case \(K\in L^2(\mathbb {R}^3)\), the other case \(K\in L^\infty (\mathbb {R}^3)\) is similar.

Lemma 3.1

Let \((V_3)\)\((V_4)\), \((K)\), \((f_1)\)\((f_4)\) be satisfied. Then each \((C)_c\)-sequence (\(c\in \mathbb {R}\)) of \(\varphi _\lambda \) is bounded in \(E_\lambda \).

Proof

Let \((u_n)\subset E_\lambda \) be a \((C)_c\) sequence of \(\varphi _\lambda \). Arguing indirectly, we can assume that

$$\begin{aligned} \varphi _\lambda (u_n)\rightarrow c,\ \ \ \ \ \Vert \varphi _\lambda '(u_n)\Vert (1+\Vert u_n\Vert _\lambda )\rightarrow 0,\ \ \ \ \ \ \ \Vert u_n\Vert _\lambda \rightarrow \infty \end{aligned}$$
(3.1)

as \(n\rightarrow \infty \) after passing to a subsequence. Take \(w_n:=u_n/\Vert u_n\Vert _\lambda \). Then \(\Vert w_n\Vert _\lambda =1\), \(w_n\rightharpoonup w\) in \(E_\lambda \) and \(w_n(x)\rightarrow w(x)\) a.e. \(x\in \mathbb {R}^3\) after passing to a subsequence.

We first consider the case \(w=0\). Combining this with (3.1), \((f_3)\) and the fact \(w_n\rightarrow 0\) in \(L^2(\left\{ x\in \mathbb {R}^3:V(x)<0\right\} )\), we obtain

$$\begin{aligned} o(1)&= \frac{1}{\Vert u_n\Vert _\lambda ^2}\left( \varphi _\lambda (u_n)-\frac{1}{4}\langle \varphi _\lambda '(u_n),u_n\rangle \right) \\&\ge \frac{1}{4}\Vert w_n\Vert _\lambda ^2-\frac{\lambda }{4}\int \limits _{\mathbb {R}^3}V^-(x)w_n^2dx +\frac{1}{\Vert u_n\Vert _\lambda ^2}\int \limits _{\mathbb {R}^3}\mathcal {F}(x,u)dx\\&\ge \frac{1}{4}-\frac{\lambda }{4}|V^-|_\infty \int \limits _{supp V^-}w_n^2dx\\&= \frac{1}{4}+o(1), \end{aligned}$$

a contradiction.

If \(w\ne 0\), then the set \(\Omega _1=\left\{ x\in \mathbb {R}^3:w(x)\ne 0\right\} \) has positive Lebesgue measure. For \(x\in \Omega _1\), one has \(|u_n(x)|\rightarrow \infty \) as \(n\rightarrow \infty \), and then, by \((f_2)\),

$$\begin{aligned} \frac{F(x,u_n(x))}{u_n^4(x)}w_n^4(x)\rightarrow +\infty \ \ \ \ \text{ as } n\rightarrow \infty , \end{aligned}$$

which, jointly with Fatou’s lemma (see [33]), shows that

$$\begin{aligned} \int \limits _{\Omega _1}\frac{F(x,u_n)}{u_n^4}w_n^4dx\rightarrow +\infty \ \ \ \ \text{ as } n\rightarrow \infty . \end{aligned}$$
(3.2)

We see from \((f_1)\), (2.4), (3.2) and the first limit of (3.1) that

$$\begin{aligned} \frac{C_0}{4}&\ge \limsup _{n\rightarrow \infty }\int \limits _{\mathbb {R}^3}\frac{F(x,u_n)}{\Vert u_n\Vert _\lambda ^4}dx \ge \limsup _{n\rightarrow \infty }\int \limits _{\Omega _1}\frac{F(x,u_n)}{u_n^4}w_n^4dx=+\infty . \end{aligned}$$

This is impossible.

In any case, we deduce a contradiction. Hence \((u_n)\) is bounded in \(E_\lambda \). \(\square \)

Lemma 3.2

Suppose that \((V_3)\)\((V_4)\), \((K)\) and (1.1) are satisfied. If \(u_n\rightharpoonup u\) in \(E_\lambda \), \(u_n(x)\rightarrow u(x)\) a.e. in \(\mathbb {R}^3\), and we denote \(w_n:=u_n-u\), then

$$\begin{aligned} \varphi _\lambda (u_n)=\varphi _\lambda (w_n)+\varphi _\lambda (u)+o(1) \end{aligned}$$
(3.3)

and

$$\begin{aligned} \varphi _\lambda '(u_n)=\varphi _\lambda '(w_n)+\varphi _\lambda '(u)+o(1) \end{aligned}$$
(3.4)

as \(n\rightarrow \infty \). In particular, if \(\varphi _\lambda (u_n)\rightarrow d\) \((\in \mathbb {R})\) and \(\varphi _\lambda '(u_n)\rightarrow 0\) in \(E_\lambda ^*\) (the dual space of \(E_\lambda \)), then \(\varphi _\lambda '(u)=0\), and

$$\begin{aligned} \varphi _\lambda (w_n)\rightarrow d-\varphi _\lambda (u),\ \ \ \ \ \ \ \varphi _\lambda '(w_n)\rightarrow 0 \end{aligned}$$
(3.5)

after passing to a subsequence.

Proof

Since \(u_n\rightharpoonup u\) in \(E_\lambda \), one has \((u_n-u,u)_\lambda \rightarrow 0\) as \(n\rightarrow \infty \), which implies that

$$\begin{aligned} \Vert u_n\Vert _\lambda ^2=(w_n+u,w_n+u)_\lambda =\Vert w_n\Vert _\lambda ^2+\Vert u\Vert _\lambda ^2+o(1). \end{aligned}$$
(3.6)

Recall \((V_3)\) and \(w_n\rightharpoonup 0\), we have

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}V^-(x)w_nudx\right| = \left| \int \limits _{supp V^-}V^-(x)w_nudx\right| \le |V^-|_\infty \left( \int \limits _{supp V^-}w_n^2dx\right) ^{1/2}|u|_2\mathop {\longrightarrow }\limits ^{n} 0 \end{aligned}$$

by the Hölder inequality. Thus

$$\begin{aligned} \int \limits _{\mathbb {R}^3}V^-(x)u_n^2dx=\int \limits _{\mathbb {R}^3}V^-(x)w_n^2dx +\int \limits _{\mathbb {R}^3}V^-(x)u^2dx+o(1). \end{aligned}$$

Combining this with (3.6) and Lemma 2.1 (ii), we obtain

$$\begin{aligned} \frac{1}{2}a_\lambda (u_n,u_n)+\frac{1}{4}N(u_n)&= \left( \frac{1}{2}a_\lambda (w_n,w_n) +\frac{1}{4}N(w_n)\right) +\left( \frac{1}{2}a_\lambda (u,u)+\frac{1}{4}N(u)\right) +o(1). \end{aligned}$$

Similarly, by Lemma 2.1 (iii),

$$\begin{aligned} a_\lambda (u_n,h)+\int \limits _{\mathbb {R}^3}K(x)\phi _{u_n}u_nhdx&= \left( a_\lambda (w_n,h)+\int \limits _{\mathbb {R}^3}K(x)\phi _{w_n}w_nhdx\right) \\&+\left( a_\lambda (u,h)+\int \limits _{\mathbb {R}^3}K(x)\phi _uuhdx\right) +o(1),\ \ \ \ \ \ \forall h\in E_\lambda . \end{aligned}$$

Therefore, to obtain (3.3) and (3.4), it suffices to check that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(F(x,u_n)-F(x,w_n)-F(x,u))dx=o(1) \end{aligned}$$
(3.7)

and

$$\begin{aligned} \sup _{\Vert h\Vert _\lambda =1}\int \limits _{\mathbb {R}^3}(f(x,u_n)-f(x,w_n)-f(x,u))h dx=o(1). \end{aligned}$$
(3.8)

Here, we only prove (3.8), the verification of (3.7) is similar. Inspired by [1], we take \({\lim _{n\rightarrow \infty }\sup _{\Vert h\Vert _\lambda =1}}\big |\int \limits _{\mathbb {R}^3}(f(x,u_n)-f(x,w_n)-f(x,u))h dx\big |=A\). If \(A>0\), then, there is \(h_0\in E_\lambda \) with \(\Vert h_0\Vert _\lambda =1\) such that

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}(f(x,u_n)-f(x,w_n)-f(x,u))h_0dx\right| \ge \frac{A}{2} \end{aligned}$$
(3.9)

for \(n\) large enough. It follows form (1.1) and the Young inequality that

$$\begin{aligned} |(f(x,u_n)-f(x,w_n))h_0|&\le \varepsilon (|w_n+u|+|w_n|)|h_0|+C_\varepsilon (|w_n+u|^{p-1}+|w_n|^{p-1})|h_0|\\&\le c(\varepsilon |w_n||h_0|+\varepsilon |u||h_0|+C_\varepsilon |w_n|^{p-1}|h_0|+C_\varepsilon |u|^{p-1}|h_0|)\\&\le c(\varepsilon w_n^2+\varepsilon h_0^2+\varepsilon u^2+\varepsilon |w_n|^p+C_{\varepsilon ,1}|u|^p+C_{\varepsilon ,2}|h_0|^p) \end{aligned}$$

for all \(n\). Hence

$$\begin{aligned} |(f(x,u_n)\!-\!f(x,w_n)\!-\!f(x,u))h_0|\!\le \! c(\varepsilon w_n^2\!+\!\varepsilon h_0^2\!+\!\varepsilon u^2\!+\!\varepsilon |w_n|^p\!+\!C_{\varepsilon ,1}|u|^p\!+\!C_{\varepsilon ,2}|h_0|^p). \end{aligned}$$

Letting

$$\begin{aligned} g_n(x):=\max \left\{ |(f(x,u_n)-f(x,w_n)-f(x,u))h_0|-c\varepsilon ( w_n^2+|w_n|^p),0\right\} , \end{aligned}$$

we have

$$\begin{aligned} 0\le g_n(x)\le c(\varepsilon h_0^2+\varepsilon u^2+C_{\varepsilon ,1}|u|^p+C_{\varepsilon ,2}|h_0|^p)\in L^1(\mathbb {R}^3), \end{aligned}$$

which implies that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}g_n(x)dx\rightarrow 0\ \ \ \ \ \text{ as } \ \ n\rightarrow \infty \end{aligned}$$
(3.10)

because of the Lebesgue dominated convergence theorem and the fact \(w_n\rightarrow 0\) a.e. in \(\mathbb {R}^3\). The definition of \(g_n(x)\) implies that

$$\begin{aligned} |(f(x,u_n)-f(x,w_n)-f(x,u))h_0|\le g_n(x)+c\varepsilon ( w_n^2+|w_n|^p), \end{aligned}$$

which, together with (3.10) and (2.1), shows that

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}(f(x,u_n)-f(x,w_n)-f(x,u))h_0 dx\right| \le c\varepsilon \end{aligned}$$

for \(n\) sufficiently large. This contradicts (3.9). Hence \(A=0\) and (3.8) holds.

If moreover \(\varphi _\lambda '(u_n)\rightarrow 0\) as \(n\rightarrow \infty \), then \(\varphi _\lambda '(u)=0\). Indeed, for each \(\psi \in C_0^\infty (\mathbb {R}^3)\), we have

$$\begin{aligned} (u_n-u,\psi )_\lambda \mathop {\longrightarrow }\limits ^{n} 0, \end{aligned}$$
(3.11)

and

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}V^-(x)(u_n-u)\psi dx\right| \le |V^-|_\infty \left( \int \limits _{supp \psi }(u_n-u)^2dx\right) ^{1/2}|\psi |_2 \mathop {\longrightarrow }\limits ^{n} 0, \end{aligned}$$
(3.12)

since \(u_n\rightarrow u\) in \(L_{loc}^2(\mathbb {R}^3)\). By Lemma 2.1 (i), \(u_n\rightharpoonup u\) in \(E_\lambda \) yields \(\phi _{u_n}\rightharpoonup \phi _u\) in \(\mathcal {D}^{1,2}(\mathbb {R}^3)\). So

$$\begin{aligned} \phi _{u_n}\rightharpoonup \phi _u\ \ \ \ \ \text{ in } L^6(\mathbb {R}^3), \end{aligned}$$

and hence

$$\begin{aligned} \int \limits _{\mathbb {R}^3}K(x)(\phi _{u_n}-\phi _u)u\psi dx\rightarrow 0 \end{aligned}$$

since \(K(x)u\psi \in L^{6/5}(\mathbb {R}^3)\). Combining this with Hölder’s inequality, we obtain

$$\begin{aligned}&\left| \int \limits _{\mathbb {R}^3}(K(x)\phi _{u_n}u_n\psi -K(x)\phi _uu\psi )dx\right| \nonumber \\&\quad \le \int \limits _{\mathbb {R}^3}|K(x)\phi _{u_n}(u_n-u)\psi |dx+\int \limits _{\mathbb {R}^3}\left| K(x)(\phi _{u_n}-\phi _u)u\psi \right| dx\nonumber \\&\quad \le |\psi |_\infty |K|_2|\phi _{u_n}|_6|u_n-u|_{3,supp \psi }+\int _{\mathbb {R}^3}|K(x)(\phi _{u_n}-\phi _u)u\psi |dx\nonumber \\&\quad =o(1). \end{aligned}$$
(3.13)

Furthermore, it follows from (1.1) and the dominated convergence theorem that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(f(x,u_n)-f(x,u))\psi dx=\int \limits _{supp \psi }(f(x,u_n)-f(x,u))\psi dx=o(1). \end{aligned}$$

This, jointly with (3.13), (3.12) and (3.11), shows that

$$\begin{aligned} \langle \varphi _\lambda '(u),\psi \rangle =\lim _{n\rightarrow \infty }\langle \varphi _\lambda '(u_n),\psi \rangle =0,\ \ \ \ \ \ \ \ \forall \psi \in C_0^\infty (\mathbb {R}^3). \end{aligned}$$

Consequently, \(\varphi _\lambda '(u)=0\) and (3.5) follows from (3.3)–(3.4). The proof is complete. \(\square \)

Lemma 3.3

Let \(V\ge 0\), \((V_3)\)\((V_4)\), \((K)\), \((f_1)\)\((f_4)\) be satisfied. Then, for any \(M>0\), there exists \(\Lambda =\Lambda (M)>0\) such that \(\varphi _\lambda \) satisfies \((C)_c\) condition for all \(c<M\) and \(\lambda >\Lambda \).

Proof

Let \((u_n)\subset E_\lambda \) be a \((C)_c\) sequence with \(c< M\). According to Lemma 3.1, \((u_n)\) is bounded. Hence we may assume that

$$\begin{aligned} u_n\!\rightharpoonup \!u\ \ \text{ in } E_\lambda ,\ \ u_n\!\rightarrow \!u \text{ in } L_{loc}^s(\mathbb {R}^3)\ \ (2\le s<2^*),\ \ u_n(x)\!\rightarrow \!u(x) \text{ a.e. } x\in \mathbb {R}^3 \end{aligned}$$
(3.14)

after passing to a subsequence. Denote \(w_n:=u_n-u\), we claim that \(w_n\rightarrow 0\) in \(E_\lambda \) for \(\lambda >0\) large. In fact, Lemma 3.2 yields that \(\varphi _\lambda '(u)=0\), and

$$\begin{aligned} \varphi _\lambda (w_n)\rightarrow c-\varphi _\lambda (u),\ \ \ \ \varphi _\lambda '(w_n)\rightarrow 0\ \ \ \ \ \ \ \text{ as } n\rightarrow \infty . \end{aligned}$$
(3.15)

Noting \(V\ge 0\) and using \((f_3)\), we get

$$\begin{aligned} \varphi _\lambda (u)=\varphi _\lambda (u)-\frac{1}{4}\langle \varphi _\lambda '(u),u\rangle =\frac{1}{4}\Vert u\Vert _\lambda ^2+\int \limits _{\mathbb {R}^3}\mathcal {F}(x,u_n)dx\ge 0, \end{aligned}$$

and then, by (3.15),

$$\begin{aligned} \int \limits _{\mathbb {R}^3}\mathcal {F}(x,w_n)dx&\le \varphi _\lambda (w_n)-\frac{1}{4}\langle \varphi _{\lambda }'(w_n),w_n\rangle =c-\varphi _{\lambda }(u)+o(1)\le M+o(1).\qquad \qquad \end{aligned}$$
(3.16)

Since \(V(x)<b\) on a set of finite measure and \(w_n\rightharpoonup 0\),

$$\begin{aligned} |w_n|_2^2\le \frac{1}{\lambda b}\int \limits _{V\ge b}\lambda V^+(x)w_n^2dx+\int \limits _{V<b}w_n^2dx\le \frac{1}{\lambda b}\Vert w_n\Vert ^2_\lambda +o(1). \end{aligned}$$
(3.17)

For \(2<s<2^*\), by (3.17) and the Hölder and Sobolev inequality, we obtain

$$\begin{aligned} |w_n|_s^s&\le \left( \int \limits _{\mathbb {R}^3}w_n^2dx\right) ^{\frac{2^*-s}{2^*-2}} \left( \int \limits _{\mathbb {R}^3}w_n^{2^*}dx\right) ^\frac{s-2}{2^*-2}\nonumber \\&\le \left( \frac{1}{\lambda b}\Vert w_n\Vert _\lambda ^2\right) ^\frac{2^*-s}{2^*-2}\bar{S}^{-\frac{2^*(s-2)}{2^*-2}}\left( \int \limits _{\mathbb {R}^3}|\nabla w_n|^2dx\right) ^\frac{2^*(s-2)}{2(2^*-2)}+o(1)\nonumber \\&\le \bar{S}^{-\frac{2^*(s-2)}{2^*-2}}\left( \frac{1}{\lambda b}\right) ^\frac{2^*-s}{2^*-2}\Vert w_n\Vert _\lambda ^s+o(1). \end{aligned}$$
(3.18)

By \((f_1)\), for any \(\varepsilon >0\), there exists \(\delta =\delta (\varepsilon )>0\) such that \(|f(x,t)|\le \varepsilon |t|\) for all \(x\in \mathbb {R}^3\) and \(|t|\le \delta \), and \((f_4)\) is satisfied for \(|t|\ge \delta \) (with the same \(\tau \) but possibly larger \(a_1\)). Hence we obtain

$$\begin{aligned} \int \limits _{|w_n|\le \delta }f(x,w_n)w_ndx\le \varepsilon \int \limits _{|w_n|\le \delta }w_n^2dx\le \frac{\varepsilon }{\lambda b}\Vert w_n\Vert _\lambda ^2+o(1), \end{aligned}$$
(3.19)

and

$$\begin{aligned} \int \limits _{|w_n|\ge \delta }f(x,w_n)w_ndx&\le \left( \int _{|w_n|\ge \delta }\left| \frac{f(x,w_n)}{w_n}\right| ^\tau dx\right) ^{1/\tau }|w_n|_s^2\nonumber \\&\le \left( \int _{|w_n|\ge \delta }a_1\mathcal {F}(x,w_n)dx\right) ^{1/\tau }|w_n|_s^2\nonumber \\&\le (a_1M)^{1/\tau }\bar{S}^{-\frac{2^*(2s-4)}{s(2^*-2)}}\left( \frac{1}{\lambda b}\right) ^\theta \Vert w_n\Vert _\lambda ^2+o(1) \end{aligned}$$
(3.20)

by \((f_4)\), (3.16), (3.18) with \(s=2\tau /(\tau -1)\) and the Hölder inequality, where \(\theta =\frac{2(2^*-s)}{s(2^*-2)}>0\). Therefore, using (3.20), (3.19) and the second limit of (3.15),

$$\begin{aligned} o(1)&= \langle \varphi _\lambda '(w_n),w_n\rangle \nonumber \\&\ge \Vert w_n\Vert _\lambda ^2-\int \limits _{\mathbb {R}^3}f(x,w_n)w_ndx\nonumber \\&\ge \left[ 1-\frac{\varepsilon }{\lambda b}-(a_1M)^{1/\tau }\bar{S}^{-\frac{2^*(2s-4)}{s(2^*-2)}}\left( \frac{1}{\lambda b}\right) ^\theta \right] \Vert w_n\Vert _\lambda ^2+o(1). \end{aligned}$$
(3.21)

So, there exists \(\Lambda =\Lambda (M)>0\) such that \(w_n\rightarrow 0\) in \(E_\lambda \) when \(\lambda >\Lambda \). Since \(w_n=u_n-u\), it follows that \(u_n\rightarrow u\) in \(E_\lambda \). \(\square \)

Lemma 3.4

Suppose that \((V_3)\)\((V_4)\), \((K)\), \((f_1)\)\((f_4)\) are satisfied, and \((u_n)\subset E_\lambda \) be a \((C)_c\) (\(c>0\)) sequence of \(\varphi _\lambda \) satisfying \(u_n\rightharpoonup u\) as \(n\rightarrow \infty \). Then, for any \(M>0\), there exists \(\Lambda =\Lambda (M)>0\) such that, \(u\) is a nontrivial critical point of \(\varphi _\lambda \) and \(\varphi _\lambda (u)\le c\) for all \(c<M\) and \(\lambda >\Lambda \).

Proof

By Lemma 3.2, we have \(\varphi _\lambda '(u)=0\) and

$$\begin{aligned} \varphi _\lambda (w_n)\rightarrow c-\varphi _\lambda (u),\ \ \ \ \ \ \varphi _\lambda '(w_n)\rightarrow 0\ \ \ \ \ \ \text{ as } n\rightarrow \infty . \end{aligned}$$
(3.22)

Since \(V\) is allowed to be sign-changing, from

$$\begin{aligned} \varphi _\lambda (u)=\varphi _\lambda (u)-\frac{1}{4}\langle \varphi _\lambda '(u),u\rangle =\frac{1}{4}\Vert u\Vert _\lambda ^2- \frac{\lambda }{4}\int \limits _{\mathbb {R}^3}V^-(x)u^2dx+\int \limits _{\mathbb {R}^3}\mathcal {F}(x,u)dx, \end{aligned}$$

it cannot deduce \(\varphi _\lambda (u)\ge 0\). We consider two possibilities:

  1. (i)

    \(\varphi _\lambda (u)<0\),

  2. (ii)

    \(\varphi _\lambda (u)\ge 0\).

If \(\varphi _\lambda (u)<0\), then \(u\ne 0\) and the proof is done. If \(\varphi _\lambda (u)\ge 0\), following the same lines as the proof of Lemma 3.3, we can deduce \(u_n\rightarrow u\) in \(E_\lambda \). Indeed, using \((V_2)\) and the fact \(w_n\rightarrow 0\) in \(L^2(\left\{ x\in \mathbb {R}^3:V(x)<b\right\} )\), we have

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}V^-(x)w_n^2dx\right| \le |V^-|_\infty \int \limits _{supp V^-}w_n^2 dx=o(1). \end{aligned}$$

Combining this with (3.22), we obtain

$$\begin{aligned} \int \limits _{\mathbb {R}^3}\mathcal {F}(x,w_n)dx&= \varphi _\lambda (w_n)-\frac{1}{4}\langle \varphi _\lambda '(w_n),w_n\rangle +\frac{1}{4}\int \limits _{\mathbb {R}^3}\lambda V^-(x)w_n^2dx-\frac{1}{4}\Vert w_n\Vert _\lambda ^2\\&\le c-\varphi _\lambda (u)+o(1)\\&\le M+o(1). \end{aligned}$$

It follows that (3.20) and (3.21) remain valid. Hence \(u_n\rightarrow u\) in \(E_\lambda \) and \(\varphi _\lambda (u)=c\) \((>0)\). This completes the proof. \(\square \)

Next, we give some preliminary results, i.e., Lemmas 3.5 to 3.8, which ensure that the functional \(\varphi _\lambda \) has the linking structure.

Lemma 3.5

Suppose that \((V_3)\)\((V_4)\), \((K)\) and (1.1) with \(p\in (4,2^*)\) are satisfied. Then, for each \(\lambda >\Lambda _0\) (\(\Lambda _0\) is the constant given in Remark 2.1), there exist \(\alpha _\lambda \), \(\rho _\lambda >0\) such that

$$\begin{aligned} \varphi _\lambda (u)\ge \alpha _\lambda \ \ \ \ \ \ \text{ for } \text{ all } u\in E_\lambda ^+\bigoplus F_\lambda \text{ with } \Vert u\Vert _\lambda =\rho _\lambda . \end{aligned}$$
(3.23)

Furthermore, if \(V\ge 0\), we can choose \(\alpha \), \(\rho >0\) independent of \(\lambda \).

Proof

For any \(u\in E_\lambda ^+\bigoplus F_\lambda \), writing \(u=u_1+u_2\) with \(u_1\in E_\lambda ^+\) and \(u_2\in F_\lambda \). Clearly, \((u_1,u_2)_\lambda =0\), and

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(|\nabla u|^2+\lambda V(x)u^2)dx=\int \limits _{\mathbb {R}^3}(|\nabla u|_1^2+\lambda V(x)u_1^2)dx+\Vert u_2\Vert ^2_\lambda . \end{aligned}$$
(3.24)

For each fixed \(\lambda >\Lambda _0\), noticing \(\mu _j(\lambda )\mathop {\longrightarrow }\limits ^{j} +\infty \), there exists a positive integer \(n_\lambda \) such that \(\mu _j(\lambda )\le 1\) for \(j\le n_\lambda \) and \(\mu _j(\lambda )>1\) for \(j\ge n_\lambda +1\). For \(u_1\in E_\lambda ^+\), we set \(u_1=\sum _{j=n_\lambda +1}^\infty a_j(\lambda )e_j(\lambda )\). Thus

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(|\nabla u_1|^2+\lambda V(x)u_1^2)dx=\Vert u_1\Vert ^2_\lambda -\int \limits _{\mathbb {R}^3}\lambda V^-(x)u_1^2dx \ge \left( 1-\frac{1}{\mu _{n_\lambda +1}(\lambda )}\right) \Vert u_1\Vert _\lambda ^2\nonumber \\ \end{aligned}$$
(3.25)

Now, using (3.24), (3.25) and (2.1), we obtain

$$\begin{aligned} \varphi _\lambda (u)&\ge \frac{1}{2}\left( 1-\frac{1}{\mu _{n_\lambda +1}(\lambda )}\right) \Vert u\Vert _\lambda ^2-\varepsilon |u|_2^2-C_\varepsilon |u|_p^p\\&\ge \left[ \frac{1}{2}\left( 1-\frac{1}{\mu _{n_\lambda +1}(\lambda )}\right) -\varepsilon \nu _2^2\right] \Vert u\Vert ^2_\lambda -C_\varepsilon \nu _p^p\Vert u\Vert ^p_\lambda , \end{aligned}$$

consequently the conclusion follows because \(p>2\) and \(\varepsilon \) has been chosen arbitrarily.

If \(V\ge 0\), since \(E_\lambda =F_\lambda \), and

$$\begin{aligned} \int \limits _{\mathbb {R}^3}(|\nabla u|^2+\lambda V(x)u^2)dx=\Vert u\Vert ^2_\lambda , \end{aligned}$$

we can choose \(\alpha \), \(\rho >0\) (independent of \(\lambda \)) such that (3.23) holds. \(\square \)

Lemma 3.6

Let \((V_3)\)\((V_4)\), \((K)\), \((f_1)\) and \((f_2)\) be satisfied. Then, for any finite dimensional subspace \(\widetilde{E}_\lambda \subset E_\lambda \), there holds

$$\begin{aligned} \varphi _\lambda (u)\rightarrow -\infty \ \ \ \ \ \ \ \ \text{ as }\ \ \ \ \ \Vert u\Vert _\lambda \rightarrow \infty ,\ \ u\in \widetilde{E}_\lambda . \end{aligned}$$

Proof

Assuming the contrary, there is a sequence \((u_n)\subset \widetilde{E}_\lambda \) with \(\Vert u_n\Vert _\lambda \rightarrow \infty \) such that

$$\begin{aligned} -\infty <\inf _{n}\varphi _\lambda (u_n). \end{aligned}$$
(3.26)

Take \(v_n:=u_n/\Vert u_n\Vert _\lambda \). Since dim\(\widetilde{E}_\lambda <+\infty \), there exists \(v\in \widetilde{E}_\lambda \backslash \left\{ 0\right\} \) such that

$$\begin{aligned} v_n\rightarrow v \text{ in } \widetilde{E}_\lambda ,\ \ \ \ \ \ \ \ \ \ \ \ \ v_n(x)\rightarrow v(x) \text{ a.e. } x\in \mathbb {R}^3 \end{aligned}$$

after passing to a subsequence. If \(v(x)\ne 0\), then \(|u_n(x)|\mathop {\rightarrow }\limits ^{n} +\infty \), and hence by \((f_2)\),

$$\begin{aligned} \frac{F(x,u_n(x))}{u_n^4(x)}v_n^4(x)\rightarrow +\infty \ \ \ \ \ \ \ \text{ as } n\rightarrow \infty . \end{aligned}$$

Combining this with \((f_1)\), (2.4) and Fatou’s lemma, we obtain

$$\begin{aligned} \frac{\varphi _\lambda (u_n)}{\Vert u_n\Vert _\lambda ^4}&\le \frac{1}{2\Vert u_n\Vert _\lambda ^2}+\frac{C_0}{4}-\int \limits _{\mathbb {R}^3}\frac{F(x,u_n)}{\Vert u_n\Vert _\lambda ^4}dx\\&= \frac{1}{2\Vert u_n\Vert _\lambda ^2}+\frac{C_0}{4}-\left( \int \limits _{v=0}+\int \limits _{v\ne 0}\right) \frac{F(x,u_n)}{u_n^4}v_n^4dx\\&\le \frac{1}{2\Vert u_n\Vert _\lambda ^2}+\frac{C_0}{4}-\int \limits _{v\ne 0}\frac{F(x,u_n)}{u_n^4}v_n^4dx\\&\rightarrow -\infty , \end{aligned}$$

a contradiction with (3.26). \(\square \)

Lemma 3.7

Suppose that \((V_3)\)\((V_4)\), \((K)\), \((f_1)\) and \((f_2)\) are satisfied. If \(V(x)<0\) for some \(x\), then, for each \(k\in \mathbb {N}\), there exist \(\lambda _k>k\), \(w_k\in E_{\lambda _k}^+\bigoplus F_{\lambda _k}\), \(R_{\lambda _k}>\rho _{\lambda _k}\) (\(\rho _{\lambda _k}\) is the constant given in Lemma 3.5) and \(b_k>0\) such that, for \(|K|_2<b_k\) (or \(|K|_\infty <b_k\)),

  1. (a)

    \(\sup \varphi _{\lambda _k}(\partial Q_k)\le 0\),

  2. (b)

    \(\sup \varphi _{\lambda _k}(Q_k)\) is bounded above by a constant independent of \(\lambda _k\),

where \(Q_k:=\left\{ u=v+tw_k:v\in E_{\lambda _k}^-, t\ge 0, \Vert u\Vert \le R_{\lambda _k}\right\} \).

Proof

We adapt an argument in Ding and Szulkin [16]. For each \(k\in \mathbb {N}\), since \(\mu _j(k)\rightarrow +\infty \) as \(j\rightarrow \infty \), there is \(j_k\in \mathbb {N}\) such that \(\mu _{j_k}(k)>1\). By Proposition 2.2, there is \(\lambda _k>k\) such that

$$\begin{aligned} 1<\mu _{j_k}(\lambda _k)<1+\frac{1}{\lambda _k}. \end{aligned}$$

Taking \(w_k:=e_{j_k}(\lambda _k)\) be an eigenvalue of \(\mu _{j_k}(\lambda _k)\), then \(w_k\in E_{\lambda _k}^+\) as \(\mu _{j_k}(\lambda _k)>1\). Since dim\(E_{\lambda _k}^- \bigoplus \mathbb {R}w_k<+\infty \), it follows directly from Lemma 3.6 that \((a)\) holds with \(R_{\lambda _k}>0\) large.

By \((f_2)\), for each \(\eta >|V^-|_\infty \), there is \(r_\eta >0\) such that \(F(x,t)\ge \frac{1}{2}\eta t^2\) if \(|t|\ge r_\eta \). For \(u=v+w\in E_{\lambda _k}^-\bigoplus \mathbb {R}w_k\), we get

$$\begin{aligned} \int \limits _{\mathbb {R}^3}V^-(x)u^2dx=\int \limits _{\mathbb {R}^3}V^-(x)v^2dx+\int \limits _{\mathbb {R}^3}V^-(x)w^2dx \end{aligned}$$

by the orthogonality of \(E_{\lambda _k}^-\) and \(\mathbb {R}w_k\). Hence we obtain

$$\begin{aligned} \varphi _{\lambda _k}(u)&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}(|\nabla w|^2+\lambda _k V(x)w^2)dx+\frac{1}{4}\int \limits _{\mathbb {R}^3}K(x)\phi _uu^2dx-\int \limits _{supp V^-}F(x,u)dx\\&\le \frac{1}{2}\left( \mu _{j_k}(\lambda _k)-1\right) \lambda _k\int \limits _{\mathbb {R}^3}V^-(x)w^2dx-\int \limits _{supp V^-}\frac{1}{2}\eta u^2dx+\frac{1}{4}\bar{S}^{-2}\nu _6^4|K|_2^2\Vert u\Vert _{\lambda _k}^4\\&-\int \limits _{supp V^-,|u|\le r_\eta }\left( F(x,u)-\frac{1}{2}\eta u^2\right) dx\\&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}V^-(x)w^2dx-\frac{\eta }{2|V^-|_\infty }\int \limits _{\mathbb {R}^3}V^-(x)w^2dx +C_\eta +\frac{1}{4}\bar{S}^{-2}\nu _6^4|K|_2^2R_{\lambda _k}^4\\&\le C_\eta +1 \end{aligned}$$

for \(u=v+w\in E_{\lambda _k}^-\bigoplus \mathbb {R}w_k\) with \(\Vert u\Vert \le R_{\lambda _k}\) and \(|K|_2<b_k:=2\bar{S}(\nu _6R_{\lambda _k})^{-2}\), where \(C_\eta \) depends on \(\eta \) but not \(\lambda \). \(\square \)

Lemma 3.8

Suppose that \((V_3)\)\((V_4)\), \((K)\), \((f_1)\) and \((f_2)\) are satisfied. If \(\Omega :=\text{ int }V^{-1}(0)\) is nonempty, then, for each \(\lambda >\Lambda _0\), there exist \(w\in E_\lambda ^+\bigoplus F_\lambda \), \(R_\lambda >0\) and \(b_\lambda >0\) such that for \(|K|_{2} < b_{\lambda }\) (or\(|K|_{\infty } < b_{\lambda }\)),

  1. (a)

    \(\sup \varphi _\lambda (\partial Q)\le 0\),

  2. (b)

    \(\sup \varphi _\lambda (Q)\) is bounded above by a constant independent of \(\lambda \),

where \(Q=\left\{ u=v+tw:v\in E_\lambda ^-,t\ge 0, \Vert u\Vert \le R_\lambda \right\} \).

Proof

Choose \(e_0\in C_0^\infty (\Omega )\backslash \left\{ 0\right\} \), then \(e_0\in F_\lambda \). By Lemma 3.6, there is \(R_\lambda >0\) large such that \(\varphi _{\lambda }(u)\le 0\) whenever \(u\in E^-_{\lambda }\bigoplus \mathbb {R}e_0\) and \(\Vert u\Vert _\lambda \ge R_\lambda \).

For \(u=v+w\in E_\lambda ^-\bigoplus \mathbb {R}e_0\), we obtain

$$\begin{aligned} \varphi _\lambda (u)&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}|\nabla w|^2dx+\frac{1}{4} \int \limits _{\mathbb {R}^3}K(x)\phi _uu^2dx-\int _{\Omega }F(x,u)dx\nonumber \\&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}|\nabla w|^2dx\!-\!\frac{\eta }{2}\int \limits _{\Omega }u^2dx\!-\!\int \limits _{\Omega , |u|\!\le \! r_\eta }\left( F(x,u)\!-\!\frac{\eta }{2}u^2\right) dx \!+\!\frac{1}{4}\bar{S}^{-2}\nu _6^4|K|_2^2\Vert u\Vert _{\lambda _k}^4\nonumber \\&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}|\nabla w|^2dx-\frac{\eta }{2}\int \limits _{\Omega }u^2dx+C_\eta +\frac{1}{4}\bar{S}^{-2}\nu _6^4|K|_2^2\Vert u\Vert _{\lambda _k}^4. \end{aligned}$$
(3.27)

Observing \(w\in C_0^\infty (\Omega )\), one has

$$\begin{aligned} \int \limits _{\mathbb {R}^3}|\nabla w|^2dx=\int \limits _{\Omega }(-\Delta w)udx\le |\Delta w|_2|u|_{2,\Omega }\le d_0|\nabla w|_2|u|_{2,\Omega }\le \frac{d_0^2}{2\eta }|\nabla w|_2^2+\frac{\eta }{2}|u|_{2,\Omega }^2,\nonumber \\ \end{aligned}$$
(3.28)

where \(d_0\) is a constant depending on \(e_0\). Choosing \(\eta \ge d_0^2\), we have \(|\nabla w|_2^2\le \eta |u|_{2,\Omega }^2\), and it follows from (3.27) that

$$\begin{aligned} \varphi _\lambda (u)\le C_\eta +\frac{1}{4}\bar{S}^{-2}\nu _6^4|K|_2^2\Vert u\Vert _{\lambda _k}^4\le C_\eta +1 \end{aligned}$$

for all \(u\in E_\lambda ^-\bigoplus \mathbb {R}e_0\) with \(\Vert u\Vert \le R_\lambda \) and \(|K|_2<b_\lambda :=2\bar{S}(\nu _6R_\lambda )^{-2}\). \(\square \)

Now we are in a position to prove Theorems 1.1 and 1.2.

Proof of Theorem 1.1

Case (i). It follows from Lemmas 3.5, 3.7 and Proposition 2.4 that, for \(\lambda =\lambda _k\) and \(|K|_2\in (0,b_k)\), \(\varphi _{\lambda _k}\) has a \((C)_c\) sequence with \(c\in [\alpha _{\lambda _k},\sup \varphi _{\lambda _k}(Q_k)]\). Setting \(M:=\sup \varphi _{\lambda _k}(Q_k)\), then \(\varphi _{\lambda _k}\) has a nontrivial critical point according to Lemmas 3.1 and 3.4.

Case (ii). The conclusion follows from Lemmas 3.1, 3.4, 3.5, 3.8 and Proposition 2.4. \(\square \)

Proof of Theorem 1.2

(Existence) Suppose \(V\ge 0\). By Lemma 3.5, there exist constants \(\alpha \), \(\rho >0\) (independent of \(\lambda \)) such that

$$\begin{aligned} \varphi _\lambda (u)\ge \alpha \ \ \ \ \ \text{ for } u\in E_\lambda \text{ with } \Vert u\Vert _\lambda =\rho . \end{aligned}$$
(3.29)

Take \(e_0\in C_0^\infty (\Omega )\backslash \left\{ 0\right\} \). Then, by \((f_1)\), \((f_2)\) and Fatou’s lemma,

$$\begin{aligned} \frac{\varphi _\lambda (te_0)}{t^4} \le \frac{1}{2t^2}\int \limits _{\Omega }|\nabla e_0|^2dx+\frac{1}{4}N(e_0)-\int \limits _{\left\{ x\in \Omega :e_0(x)\ne 0\right\} }\frac{F(x,te_0)}{(te_0)^4}e_0^4dx \rightarrow -\infty \end{aligned}$$

as \(t\rightarrow +\infty \), which yields that \(\varphi _\lambda (te_0)<0\) for \(t>0\) large. Clearly, there is \(C_1>0\) (independent of \(\lambda \)) such that

$$\begin{aligned} c_\lambda :=\inf _{h\in \Gamma }\max _{t\in [0,1]}\varphi _\lambda (h(t))\le \sup _{t\ge 0}\varphi _\lambda (te_0)\le C_1, \end{aligned}$$
(3.30)

where \(\Gamma =\left\{ h\in C([0,1],E_\lambda ):h(0)=0, \Vert h(1)\Vert _\lambda \ge \rho ,\varphi _\lambda (h(1))<0\right\} \). By Proposition 2.3 and Lemma 3.3, we obtain a nontrivial critical point \(u_\lambda \) of \(\varphi _\lambda \) with \(\varphi _\lambda (u_\lambda )\in [\alpha , C_1]\) for \(\lambda \) large.

(Multiplicity) For each \(k\in \mathbb {N}\), we choose \(k\) functions \(e_i\in C_0^\infty (\Omega )\) such that supp\(e_i\cap \)supp\(e_j=\emptyset \) if \(i\ne j\). Let

$$\begin{aligned} W_k=\text{ span }\left\{ e_1,e_2,\dots ,e_k\right\} . \end{aligned}$$

According to (3.29), Lemma 3.3 and Proposition 2.5, it suffices to show that \(\sup \varphi _\lambda (W_k)\) is bounded above by a constant independent of \(\lambda \).

For \(u\in W_k\) and \(\eta >0\), we have [cf. (3.28)]

$$\begin{aligned} \int \limits _{\mathbb {R}^3}|\nabla u|^2dx\le \frac{d_k^2}{2\eta }|\nabla u|_2^2+\frac{\eta }{2}|u|^2_{2,\Omega } \end{aligned}$$

(\(d_k\) is a constant depending on \(W_k\)). It follows that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}|\nabla u|^2dx\le \eta |u|_{2,\Omega }^2, \ \ \ \ \ \ \ \ \ \ \ \ \text{ if } \eta \ge d_k^2. \end{aligned}$$
(3.31)

Combining this with (2.4) and the Hölder inequality, we obtain

$$\begin{aligned} N(u)\le C_0\Vert u\Vert _\lambda ^4=C_0\left( \int \limits _{\Omega }|\nabla u|^2dx\right) ^2 \le C_0\eta ^2\left( \int \limits _{\Omega }u^2dx\right) ^2\le C_0\eta ^2|\Omega |\int \limits _{\Omega }u^4dx \hbox { for all } u\in W_{k}.\nonumber \\ \end{aligned}$$
(3.32)

By \((f_2)\), for each \(\eta >d_k^2\), there is \(r_\eta >0\) such that

$$\begin{aligned} F(x,t)\ge \frac{1}{2}\eta t^2+\frac{1}{4}C_0\eta ^2|\Omega |t^4,\ \ \ \ \ \ \ \ \forall x\in \mathbb {R}^3,\ \ |t|\ge r_\eta . \end{aligned}$$
(3.33)

Hence we obtain, using (3.31)–(3.33),

$$\begin{aligned} \varphi _\lambda (u)&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}|\nabla u|^2dx+\frac{1}{4}N(u)-\int \limits _{\Omega }F(x,u)dx\\&\le \frac{1}{2}\int \limits _{\mathbb {R}^3}|\nabla u|^2dx+\frac{1}{4}N(u)-\frac{\eta }{2}\int \limits _\Omega u^2dx-\frac{1}{4}C_0\eta ^2|\Omega |\int \limits _{\Omega }u^4dx\\&-\int \limits _{\Omega ,|u|\le r_\eta }\left( F(x,u)-\frac{\eta }{2}u^2 -\frac{1}{4}C_0\eta ^2|\Omega |u^4\right) dx\\&\le C_\eta \end{aligned}$$

for all \(u\in W_k\), where \(C_\eta \) is independent of \(\lambda \). \(\square \)

4 Proof of Theorem 1.3

In this section, we are concerned with problem \((SP)_1\) with sublinear nonlinearity. We consider the functional \(\varphi _1\) (denoted by \(\varphi \) for simplicity) on \((E, \Vert \cdot \Vert )\):

$$\begin{aligned} \varphi (u)&= \frac{1}{2}\int \limits _{\mathbb {R}^3}(|\nabla u|^2+V(x)u^2)dx+\frac{1}{4}\int \limits _{\mathbb {R}^3}K(x)\phi _u u^2dx-\psi (u), \end{aligned}$$

where \(\psi (u)=\int _{\mathbb {R}^3}F(x,u)dx\). Since the constant \(\nu _s\) given in (2.1) is independent of \(\lambda \), it still holds

$$\begin{aligned} |u|_s\le \nu _s\Vert u\Vert ,\ \ \ \ \ \ \ \ \forall u\in E. \end{aligned}$$
(4.1)

It follows from \((f_5)\) that

$$\begin{aligned} |F(x,u)|\le m(x)|u|^\sigma +h(x)|u|^\gamma ,\ \ \ \ \ \ \forall (x,u)\in \mathbb {R}^3\times \mathbb {R}, \end{aligned}$$
(4.2)

which, jointly with (4.1) and Hölder’s inequality, shows that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}F(x,u)dx&\le \int \limits _{\mathbb {R}^3}(m(x)|u|^\sigma +h(x)|u|^\gamma )dx\nonumber \\&\le |m|_{\frac{2}{2-\sigma }}|u|_2^\sigma +|h|_{\frac{2}{2-\gamma }}|u|_2^\gamma \nonumber \\&\le |m|_{\frac{2}{2-\sigma }}\nu _2^\sigma \Vert u\Vert ^\sigma +|h|_{\frac{2}{2-\gamma }}\nu _2^\gamma \Vert u\Vert ^\gamma \nonumber \\&< +\infty . \end{aligned}$$
(4.3)

Hence, \(\psi \) and \(\varphi \) are well defined. In addition, we have the following lemmas.

Lemma 4.1

Assume that \((V_3)\), \((V_4)\) and \((f_5)\) hold and \(u_n\rightharpoonup u\) in \(E\), then

$$\begin{aligned} f(x,u_n)\rightarrow f(x,u)\ \ \ \ \ \ \text{ in } L^2(\mathbb {R}^3). \end{aligned}$$
(4.4)

Proof

Since \(u_n\rightharpoonup u\) in \(E\), there is a constant \(M>0\) such that

$$\begin{aligned} \Vert u_n\Vert \le M\ \ \ \ \ \ \text{ and }\ \ \ \ \ \Vert u\Vert \le M,\ \ \ \ \forall n\in \mathbb {N}. \end{aligned}$$
(4.5)

Up to a subsequence, we can assume that

$$\begin{aligned}&u_n\rightarrow u\ \ \ \ \ \text{ in } L_{loc}^2(\mathbb {R}^3),\nonumber \\&u_n(x)\rightarrow u(x)\ \ \ \text{ a.e. } x\in \mathbb {R}^3. \end{aligned}$$
(4.6)

By the properties of the functions \(m\) and \(h\), we have, for every \(\varepsilon >0\), there exists \(T_\varepsilon >0\) such that

$$\begin{aligned} \left( \int \limits _{|x|\ge T_\varepsilon }|m(x)|^\frac{2}{2-\sigma }dx\right) ^{\frac{2-\sigma }{2}}< \sqrt{\varepsilon }\ \ \ \ \ \text{ and }\ \ \ \ \left( \int \limits _{|x|\ge T_\varepsilon }|h(x)|^\frac{2}{2-\gamma }dx\right) ^\frac{2-\gamma }{2}<\sqrt{\varepsilon }. \end{aligned}$$
(4.7)

By (4.6), passing to a subsequence if necessary, we can assume that \(\sum _{n=1}^\infty \int _{|x|\!\le \! T_\varepsilon }|u_n-u|^2dx<+\!\infty \). Taking \({w(x)=\sum _{n=1}^\infty |u_n(x)-u(x)|}\) for \(|x|\le T_\varepsilon \), then \(\int _{|x|\le T_\varepsilon }w^2dx<+\infty \). It follows from \((f_5)\) that, for all \(n\in \mathbb {N}\) and \(|x|\le T_\varepsilon \),

$$\begin{aligned} |f(x,u_n)-f(x,u)|^2&\le [m(x)(|u_n|^{\sigma -1}+|u|^{\sigma -1})+h(x)(|u_n|^{\gamma -1}+|u|^{\gamma -1})]^2\\&\le 4m^2(x)(|u_n|^{2\sigma -2}+|u|^{2\sigma -2})+4h^2(x)(|u_n|^{2\gamma -2}+|u|^{2\gamma -2})\\&\le 2^{2\sigma +1}m^2(x)(|u_n-u|^{2\sigma -2}+|u|^{2\sigma -2})\\&+2^{2\gamma +1}h^2(x)(|u_n-u|^{2\gamma -2}+|u|^{2\gamma -2})\\&\le 2^{2\sigma +1}m^2(x)(|w|^{2\sigma -2}+|u|^{2\sigma -2})\\&+2^{2\gamma +1}h^2(x)(|w|^{2\gamma -2}+|u|^{2\gamma -2}), \end{aligned}$$

and, using Hölder’s inequality,

$$\begin{aligned}&\int \limits _{|x|\le T_\varepsilon }\left[ 2^{2\sigma +1}m^2(x)(|w|^{2\sigma -2}+|u|^{2\sigma -2}) +2^{2\gamma +1}h^2(x)(|w|^{2\gamma -2}+|u|^{2\gamma -2})\right] dx\\&\quad \le 2^{2\sigma +1}|m|_{\frac{2}{2-\sigma }}^2\left[ \left( \int \limits _{|x|\le T_\varepsilon }w^2dx\right) ^{\sigma -1}+\left( \int \limits _{|x|\le T_\varepsilon }u^2dx\right) ^{\sigma -1}\right] \\&\qquad +2^{2\gamma +1}|h|_{\frac{2}{2-\gamma }}^2\left[ \left( \int \limits _{|x|\le T_\varepsilon }w^2dx\right) ^{\gamma -1}+\left( \int \limits _{|x|\le T_\varepsilon }u^2dx\right) ^{\gamma -1}\right] \\&\quad <+\infty . \end{aligned}$$

Hence, by Lebesgue dominated convergence theorem, we obtain

$$\begin{aligned} \int \limits _{|x|\le T_\varepsilon }|f(x,u_n)-f(x,u)|^2dx\rightarrow 0\ \ \ \ \ \text{ as }\ \ n\rightarrow \infty . \end{aligned}$$
(4.8)

On the other hand, using (\(f_5\)), (4.7), (4.5), (4.1) and the Hölder inequality, we have

$$\begin{aligned}&\int \limits _{|x|\ge T_\varepsilon }|f(x,u_n)-f(x,u)|^2dx\\&\quad \le \int \limits _{|x|\ge T_\varepsilon }[m(x)(|u_n|^{\sigma -1}+|u|^{\sigma -1})+h(x)(|u_n|^{\gamma -1}+|u|^{\gamma -1})]^2dx\\&\quad \le 4\int \limits _{|x|\ge T_\varepsilon }m^2(x)(|u_n|^{2\sigma -2}+|u|^{2\sigma -2})dx\\&\qquad +4\int \limits _{|x|\ge T_\varepsilon }h^2(x)(|u_n|^{2\gamma -2}+|u|^{2\gamma -2})dx\\&\quad \le 4\left( \int \limits _{|x|\ge T_\varepsilon }|m|^\frac{2}{2-\sigma }dx\right) ^{2-\sigma }(|u_n|_2^{2\sigma -2}+|u|_2^{2\sigma -2})\\&\qquad +4\left( \int _{|x|\ge T_\varepsilon }|h|^\frac{2}{2-\gamma }dx\right) ^{2-\gamma }(|u_n|_2^{2\gamma -2}+|u|_2^{2\gamma -2})\\&\quad \le 8 \varepsilon \left( \nu _2^{2\sigma -2}M^{2\sigma -2}+\nu _2^{2\gamma -2}M^{2\gamma -2}\right) . \end{aligned}$$

This, together with (4.8), shows that (4.4) holds. This completes the proof. \(\square \)

Lemma 4.2

Assume that \(V\ge 0\), \((V_3)\), \((K)\) and \((f_5)\) hold. Then \(\psi \in C^1(E,\mathbb {R})\) and \(\psi ':E\rightarrow E^*\) (the dual space of \(E\)) is compact, and hence \(\varphi \in C^1(E,\mathbb {R})\),

$$\begin{aligned}&\langle \psi '(u),v\rangle =\int \limits _{\mathbb {R}^3}f(x,u)vdx,\\&\langle \varphi '(u),v \rangle =\int \limits _{\mathbb {R}^3}\left( \nabla u\cdot \nabla v+V(x)uv+K(x)\phi _u uv-f(x,u)v\right) dx\nonumber \end{aligned}$$
(4.9)

for all \(u,v\in E\). If \(u\) is a critical point of \(\varphi \), then the pair \((u,\phi _u)\) is a solution of problem \((SP)_1\).

Proof

In view of Lemma 4.1 and (4.1), the proof is standard and we refer to [23]. \(\square \)

Proof of Theorem 1.3

In view of Lemma 4.2 and the oddness of \(f\), we know that \(\varphi \in C^1(E,\mathbb {R})\) and \(\varphi (-u)=\varphi (u)\). It remains to verify the conditions (i) and (ii) of Proposition 2.6. We follow an argument in [20].

Verification of \((i)\). Since \(V\ge 0\), we get \(F_\lambda =E_\lambda \). It follows from (4.3) that

$$\begin{aligned} \varphi (u)&\ge \frac{1}{2}\Vert u\Vert ^2-|m|_{\frac{2}{2-\sigma }}\nu _2^\sigma \Vert u\Vert ^\sigma -|h|_{\frac{2}{2-\gamma }}\nu _2^\gamma \Vert u\Vert ^\gamma ,\ \ \ \ \ \ \ \forall u\in E. \end{aligned}$$

Noting that \(\sigma \), \(\gamma \in (1,2)\), we have

$$\begin{aligned} \varphi (u)\rightarrow +\infty \ \ \ \ \ \text{ as }\ \ \Vert u\Vert \rightarrow \infty . \end{aligned}$$
(4.10)

Thus \(\varphi \) is bounded from below.

Let \((u_n)\subset E\) be a (PS)-sequence of \(\varphi \), i.e., \(\left\{ \varphi (u_n)\right\} \) is bounded and \(\varphi '(u_n)\rightarrow 0\) as \(n\rightarrow \infty \). By (4.10), \((u_n)\) is bounded, and then \(u_n\rightharpoonup u\) in \(E\) for some \(u\in E\). Recall that

$$\begin{aligned} (xy)^{1/2}(x+y)\le x^2+y^2,\ \ \ \ \ \ \ \forall x,y\ge 0. \end{aligned}$$

Hence we obtain, by (2.3) and Hölder’s inequality,

$$\begin{aligned}&\int \limits _{\mathbb {R}^3}K(x)(\phi _{u_n}u_nu+\phi _uu_n u)dx\\&\quad \le \left( \int \limits _{\mathbb {R}^3}K(x)\phi _{u_n}u_n^2dx\right) ^{1/2}\left( \int \limits _{\mathbb {R}^3}K(x)\phi _{u_n}u^2dx\right) ^{1/2}\\&\qquad +\left( \int \limits _{\mathbb {R}^3}K(x)\phi _uu_n^2dx\right) ^{1/2}\left( \int \limits _{\mathbb {R}^3}K(x)\phi _uu^2dx\right) ^{1/2}\\&\quad =\left( \int \limits _{\mathbb {R}^3}\nabla \phi _{u_n}\cdot \nabla \phi _udx\right) ^{1/2}\left( \Vert \phi _{u_n}\Vert _{\mathcal {D}^{1,2}}+\Vert \phi _u\Vert _{\mathcal {D}^{1,2}}\right) \\&\quad \le \left( \int \limits _{\mathbb {R}^3}|\nabla \phi _{u_n}|^2dx\right) ^{1/4}\left( \int \limits _{\mathbb {R}^3}|\nabla \phi _{u}|^2dx\right) ^{1/4}\left( \Vert \phi _{u_n}\Vert _{\mathcal {D}^{1,2}}+\Vert \phi _u\Vert _{\mathcal {D}^{1,2}}\right) \\&\quad =\Vert \phi _{u_n}\Vert _{\mathcal {D}^{1,2}}^{1/2}\Vert \phi _u\Vert _{\mathcal {D}^{1,2}}^{1/2} \left( \Vert \phi _{u_n}\Vert _{\mathcal {D}^{1,2}}+\Vert \phi _u\Vert _{\mathcal {D}^{1,2}}\right) \\&\quad \le \Vert \phi _{u_n}\Vert _{\mathcal {D}^{1,2}}^2+\Vert \phi _u\Vert _{\mathcal {D}^{1,2}}^2\\&\quad =\int \limits _{\mathbb {R}^3}K(x)(\phi _{u_n}u_n^2+\phi _uu^2)dx, \end{aligned}$$

which implies that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}K(x)(\phi _{u_n}u_n-\phi _uu)(u_n-u)dx\ge 0. \end{aligned}$$

Combining this with Lemma 4.1, we obtain

$$\begin{aligned}&\Vert u_n-u\Vert ^2=\langle \varphi '(u_n)-\varphi '(u),u_n-u\rangle -\int \limits _{\mathbb {R}^3}K(x)(\phi _{u_n}u_n-\phi _u u)(u_n-u)dx\\&\qquad \qquad \qquad +\int \limits _{\mathbb {R}^3}(f(x,u_n)-f(x,u))(u_n-u)dx\\&\le \Vert \varphi '(u_n)\Vert _{E^*}\Vert u_n\!-\!u\Vert \!-\!\langle \varphi '(u),u_n\!-\!u\rangle \!+\!\left( \int \limits _{\mathbb {R}^3}|f(x,u_n)\!-\!f(x,u)|^2dx\right) ^{1/2}\cdot |u_n\!-\!u|_2\\&\quad \rightarrow 0, \end{aligned}$$

that is, \(u_n\rightarrow u\) \((n\rightarrow \infty )\). Hence the (PS) condition holds.

Verification of \((ii)\). For simplicity, we assume that \(x_0=0\) in \((f_6)\). For \(r>0\), let \(D(r)\) denotes the cube

$$\begin{aligned} D(r)=\left\{ (x_1,x_2,x_3):0\le x_i\le r, i=1,2,3\right\} . \end{aligned}$$

Fix \(r>0\) small enough such that \(D(r)\subset B(0,\delta )\), where \(\delta \) is the constant given in \((f_6)\). For arbitrary \(k\in \mathbb {N}\), we shall construct an \(A_k\in \Gamma _k\) satisfying \(\sup _{u\in A_k}\varphi (u)<0\).

Let \(m\in \mathbb {N}\) be the smallest integer such that \(m^3\ge k\). We divide \(D(r)\) equally into \(m^3\) small cubes by planes parallel to each face of \(D(r)\) and denote them by \(D_i\) with \(1\le i\le m^3\). We only use \(D_i\) with \(1\le i\le k\). Set \(a=r/m\). Then the edge of \(D_i\) has length \(a\). We consider a cube \(E_i\subset D_i\) \((i=1,2,\dots ,k)\) such that \(E_i\) has the same center as that of \(D_i\), the faces of \(E_i\) and \(D_i\) are parallel and the edge of \(E_i\) has length \(a/2\). Define \(\zeta \in C_0^\infty (\mathbb {R},[0,1])\) such that \(\zeta (t)=1\) for \(t\in [a/4,3a/4]\), \(\zeta (t)=0\) for \(t\in (-\infty ,0]\bigcup [a,+\infty )\). Define

$$\begin{aligned} \xi (x)=\zeta (x_1)\zeta (x_2)\zeta (x_3),\ \ \ \ \ \ (x_1,x_2,x_3)\in \mathbb {R}^3. \end{aligned}$$

Then supp\(\xi \subset [0,a]^3\). Now for each \(1\le i\le k\), we can choose a suitable \(y_i\in \mathbb {R}^3\) and define

$$\begin{aligned} \xi _i(x)=\xi (x-y_i),\ \ \ \ \ \ \forall x\in \mathbb {R}^3 \end{aligned}$$

such that

$$\begin{aligned} \text{ supp }\xi _i\subset D_i,\ \ \ \ \ \ \ \text{ supp }\xi _i\bigcap \text{ supp }\xi _j=\emptyset \ \ (i\ne j), \end{aligned}$$
(4.11)

and

$$\begin{aligned} \xi _i(x)=1\ \ (x\in E_i),\ \ \ \ \ \ \ 0\le \xi _i(x)\le 1\ \ (x\in \mathbb {R}^3). \end{aligned}$$

Set

$$\begin{aligned} V_k=\left\{ (t_1,t_2,\dots ,t_k)\in \mathbb {R}^k:\max _{1\le i\le k}|t_i|=1\right\} \end{aligned}$$
(4.12)

and

$$\begin{aligned} W_k=\left\{ \sum _{i=1}^kt_i\xi _i(x):(t_1,t_2,\dots ,t_k)\in V_k\right\} . \end{aligned}$$

Observing \(V_k\) is homeomorphic to the unit sphere in \(\mathbb {R}^k\) by an odd mapping, we get \(\gamma (V_k)=k\). Furthermore, \(\gamma (W_k)=\gamma (V_k)=k\) because the mapping \((t_1,\dots ,t_k)\longmapsto \sum _{i=1}^kt_i\xi _i(x)\) is odd and homeomorphic. Since \(W_k\) is compact, there exists \(C_k>0\) such that

$$\begin{aligned} \Vert u\Vert \le C_k,\ \ \ \ \ \ \ \ \forall u\in W_k. \end{aligned}$$
(4.13)

For \(0<s<\varepsilon \) (\(\varepsilon \) is the constant given in \((f_6)\)) and \(u=\sum _{i=1}^kt_i\xi _i(x)\in W_k\), we obtain

$$\begin{aligned} \varphi (su)&\le \frac{s^2}{2}\Vert u\Vert ^2+\frac{s^4}{4}\int \limits _{\mathbb {R}^3}K(x)\phi _uu^2dx -\int \limits _{\mathbb {R}_3}F\left( x,s\sum _{i=1}^kt_i\xi _i(x)\right) dx\nonumber \\&\le \frac{s^2}{2}C_k^2+\frac{s^4}{4}C_0C_k^4-\sum _{i=1}^k\int \limits _{D_i}F(x,st_i\xi _i(x))dx \end{aligned}$$
(4.14)

by (4.13), (4.11) and Lemma 3.5 \((i)\). Observing (4.12), there exists an integer \(i_0\in [1,k]\) such that \(|t_{i_0}|=1\). Then it follows that

$$\begin{aligned} \sum _{i=1}^k\int \limits _{D_i}F(x,st_i\xi _i(x))dx&= \int \limits _{E_{i_0}}F(x,st_{i_0}\xi _{i_0}(x))dx+\int \limits _{D_{i_0}\backslash E_{i_0}}F(x,st_{i_0}\xi _{i_0}(x))dx\nonumber \\&+\sum _{i\ne i_0}\int \limits _{D_i}F(x,st_i\xi _i(x))dx. \end{aligned}$$
(4.15)

Noting that \(|t_{i_0}|=1\), \(\xi _{i_0}\equiv 1\) on \(E_{i_0}\) and \(F(x,u)\) is even in \(u\), we get

$$\begin{aligned} \int \limits _{E_{i_0}}F(x,st_{i_0}\xi _{i_0}(x))dx=\int \limits _{E_{i_0}}F(x,s)dx. \end{aligned}$$
(4.16)

By (\(f_6\)),

$$\begin{aligned} \int \limits _{D_{i_0}\backslash E_{i_0}}F(x,st_{i_0}\xi _{i_0}(x))dx+\sum _{i\ne i_0}\int \limits _{D_i}F(x,st_i\xi _i(x))dx\ge -a_2\text{ vol }(D(r))s^2,\quad \end{aligned}$$
(4.17)

where vol\((D(r))\) denotes the volume of \(D(r)\), i.e. \(r^3\). Combining (4.14)–(4.17), one has

$$\begin{aligned} \varphi (su)\le \frac{s^2}{2}C_k^2+\frac{s^4}{4} C_0C_k^4+a_2r^3s^2-\int \limits _{E_{i_0}}F(x,s)dx. \end{aligned}$$

Substituting \(s=\varepsilon _n\) and using (1.2), we obtain

$$\begin{aligned} \varphi (\varepsilon _n u)\le \varepsilon _n^2\left[ \frac{C_k^2}{2}+\frac{\varepsilon _n^2}{4} C_0C_k^4+a_2r^3-\left( \frac{a}{2}\right) ^3M_n\right] . \end{aligned}$$

Since \(\varepsilon _n\rightarrow 0^+\) and \(M_n\rightarrow +\infty \) as \(n\rightarrow \infty \), we choose \(n_0\) large enough such that the right side of the last inequality is negative. Take

$$\begin{aligned} A_k=\varepsilon _{n_0}W_k. \end{aligned}$$

Then we have

$$\begin{aligned} \gamma (A_k)=\gamma (W_k)=k\ \ \ \ \ \ \ \text{ and }\ \ \ \ \ \sup _{u\in A_k}\varphi (u)<0. \end{aligned}$$

Consequently, Theorem 1.3 follows from Proposition 2.6. This completes the proof. \(\square \)

5 Concentration of solutions

In this section, we deal with problem \((SP)_\lambda \) with \(\lambda =\lambda _k\rightarrow +\infty \).

Theorem 5.1

Suppose that \((V_3)\)\((V_4)\) and \((K)\) are satisfied, \(V^{-1}(0)\) has nonempty interior \(\Omega \) and there exist \(a_3>0\), \(p\in (2,2^*)\) such that

$$\begin{aligned} |f(x,t)|\le a_3(|t|+|t|^{p-1}),\ \ \ \ \ \ \ \ \ \forall (x,t)\in \mathbb {R}^3\times \mathbb {R}. \end{aligned}$$
(5.1)

Let \((u_k)\subset E\) be a solution of \((SP)_\lambda \) with \(\lambda =\lambda _k\). If \(\lambda _k\rightarrow +\infty \) and \(\Vert u_k\Vert _{\lambda _k}\le C\) for some \(C>0\) and all \(k\), then, passing to a subsequence, \(u_k\rightarrow \bar{u}\) in \(L^s(\mathbb {R}^3)\) for \(s\in (2,2^*)\), \(\bar{u}\) is a weak solution of

$$\begin{aligned} \left\{ \begin{array}{ll} -\Delta u+\frac{1}{4\pi }\left( (K(x)u^2)*\frac{1}{|x|}\right) K(x)u=f(x,u)\ \ \ \ &{} \ \text{ in } \Omega ,\\ u=0\ \ \ \ &{} \text{ on } \partial \Omega ,\\ \end{array} \right. \end{aligned}$$
(5.2)

and \(\bar{u}=0\) a.e. in \(\mathbb {R}^3\backslash V^{-1}(0)\). If moreover \(V\ge 0\) and \((f_1)\) is satisfied, then \(u_k\rightarrow \bar{u}\) in \(E\).

We note that \(\bar{u}\in H_0^1(\Omega )\) if \(V^{-1}(0)=\overline{\Omega }\) and \(\partial \Omega \) is locally Lipschitz continuous (see [7]). Before proving the above theorem we point out some of its consequences.

Corollary 5.1

Let \((u_\lambda ,\phi _\lambda )\) be the solution obtained in Theorem 1.2 (existence result). Then \(u_\lambda \rightarrow \bar{u}\) in \(E\), \(\phi _\lambda \rightarrow \phi _{\bar{u}}\) in \(\mathcal {D}^{1,2}(\mathbb {R}^3)\) as \(\lambda \rightarrow +\infty \), and \(\bar{u}\) is a nontrivial solution of (5.2).

Proof

For \(\lambda _k\rightarrow +\infty \), set \(u_k:=u_{\lambda _k}\) be the critical point of \(\varphi _{\lambda _k}\) obtained in Theorem 1.2. It follows from (3.30) that

$$\begin{aligned} C_1\ge c_{\lambda _k}=\varphi _{\lambda _k}(u_k)-\frac{1}{4}\langle \varphi _{\lambda _k}'(u_k),u_k \rangle =\frac{1}{4}\Vert u_k\Vert _{\lambda _k}^2+\int \limits _{\mathbb {R}^3}\mathcal {F}(x,u_k)dx\ge \frac{1}{4}\Vert u_k\Vert _{\lambda _k}^2. \end{aligned}$$

Hence \(\left\{ \Vert u_k\Vert _{\lambda _k}\right\} \) is bounded. So the conclusion of Theorem 5.1 holds.

We show that \(\bar{u}\ne 0\). Since \(V\ge 0\) and \(\langle \varphi _{\lambda _k}'(u_k),u_k\rangle =0\), we have

$$\begin{aligned} \Vert u_k\Vert _{\lambda _k}^2+N(u_k)=\int \limits _{\mathbb {R}^3}f(x,u_k)u_kdx\le \varepsilon |u_k|_2^2+C_\varepsilon |u_k|_p^p. \end{aligned}$$

If \(\bar{u}=0\), then \(u_k\rightarrow 0\) in \(L^p(\mathbb {R}^3)\), and therefore

$$\begin{aligned} \Vert u_k\Vert _{\lambda _k}\rightarrow 0,\ \ \ \ \ \ \ N(u_k)\rightarrow 0\ \ \ \ \ \ \text{ as } k\rightarrow \infty \end{aligned}$$

(note \(|u_{\lambda _k}|_2\) is bounded and \(\varepsilon \) is arbitrary). Now it follows easily that \(\varphi _{\lambda _k}(u_k)\rightarrow 0\), a contradiction with the fact \(\varphi _{\lambda _k}(u_k)=c_{\lambda _k}\ge \alpha \). \(\square \)

Proof of Theorem 5.1

We adapt an argument in [7]. We divide the proof into three steps.

(1) Since \(\Vert u_k\Vert \le \Vert u_k\Vert _{\lambda _k}\le C\), one has

$$\begin{aligned} u_k\rightharpoonup \bar{u}\ \ \text{ in } E,\ \ \ \ \ \ u_k\rightarrow \bar{u} \ \ \text{ in } L_{loc}^s(\mathbb {R}^3)\ \ (2\le s<2^*),\ \ \ \ \ \ u_k(x)\rightarrow \bar{u}(x)\ \ \text{ a.e. } x\in \mathbb {R}^3. \end{aligned}$$

For any \(\psi \in C_0^\infty (\mathbb {R}^3)\), it follows from the fact \(\langle \varphi _{\lambda _k}'(u_k),\psi \rangle =0\) that

$$\begin{aligned}&\left| \int \limits _{\mathbb {R}^3}V(x)u_k\psi dx\right| \\&\quad \le \frac{1}{\lambda _k}\left( \int \limits _{\mathbb {R}^3}|f(x,u_k)\psi |dx+\int \limits _{\mathbb {R}^3}|K(x)\phi _{u_k}u_k\psi |dx +\int \limits _{\mathbb {R}^3}|\nabla u_k\cdot \nabla \psi |dx\right) \\&\quad \le \frac{1}{\lambda _k}\left[ a_3(|u_k|_2|\psi |_2+|u_k|_p^{p-1}|\psi |_p)+|K|_2 |\phi _{u_k}|_6|\psi |_\infty |u_k|_3+|\nabla u_k|_2|\nabla \psi |_2\right] \\&\quad \le \frac{c}{\lambda _k}\longrightarrow 0\ \ \ \ \ \ \text{ as } k\rightarrow \infty , \end{aligned}$$

and hence

$$\begin{aligned} \int \limits _{\mathbb {R}^3}V(x)\bar{u}\psi dx=0,\ \ \ \ \ \ \ \forall \psi \in C_0^\infty (\mathbb {R}^3), \end{aligned}$$

which implies that \(\bar{u}=0\) a.e. in \(\mathbb {R}^3\backslash V^{-1}(0)\). Now for each \(\psi \in C_0^\infty (\Omega )\), since \( \langle \varphi _{\lambda _k}'(u_k),\psi \rangle =0\), it follows that

$$\begin{aligned} \int \limits _{\mathbb {R}^3}\nabla \bar{u}\cdot \nabla \psi dx+\int \limits _{\mathbb {R}^3}K(x)\phi _{\bar{u}}\bar{u}\psi dx=\int \limits _{\mathbb {R}^3}f(x,\bar{u})\psi dx, \end{aligned}$$

i.e., \(\bar{u}\) is a weak solution of (5.2) by the density of \(C_0^\infty (\Omega )\) in \(H_0^1(\Omega )\).

(2) \(u_k\rightarrow \bar{u}\) in \(L^s(\mathbb {R}^3)\) for \(2<s<2^*\). Arguing indirectly, by Lion’s vanishing lemma, there exist \(\delta \), \(\rho >0\) and \((x_k)\subset \mathbb {R}^3\) such that

$$\begin{aligned} \int \limits _{B_\rho (x_k)}(u_k-\bar{u})^2dx\ge \delta . \end{aligned}$$

It is easy to see that \(|x_k|\mathop {\longrightarrow }\limits ^{k} \infty \). So meas\(\left( B_\rho (x_k)\cap \left\{ x\in \mathbb {R}^3:V(x)<b\right\} \right) \rightarrow 0\), and

$$\begin{aligned} \int \limits _{B_\rho (x_k)\cap \left\{ V<b\right\} }(u_k-\bar{u})^2dx\le |u_k-\bar{u}|_3^2\left( \text{ meas }(B_\rho (x_k)\cap \left\{ V<b\right\} )\right) ^{1/3}\mathop {\longrightarrow }\limits ^{k}0. \end{aligned}$$

Thus,

$$\begin{aligned} \Vert u_k\Vert _{\lambda _k}^2&\ge \lambda _k b\int \limits _{B_\rho (x_k)\cap \left\{ V\ge b\right\} }u_k^2dx\\&= \lambda _k b \int \limits _{B_\rho (x_k)\cap \left\{ V\ge b\right\} }(u_k-\bar{u})^2dx\\&= \lambda _k b \left( \int \limits _{B_\rho (x_k)}(u_k-\bar{u})^2dx-\int \limits _{B_\rho (x_k)\cap \left\{ V<b\right\} }(u_k-\bar{u})^2dx\right) \\&\rightarrow +\infty , \end{aligned}$$

a contradiction with the boundedness of \(\left\{ \Vert u_k\Vert _{\lambda _k}\right\} _k\).

(3) Suppose that \(V\ge 0\) and \((f_1)\) holds. We show that \(u_k\rightarrow \bar{u}\) in \(E\). Since \(\langle \varphi _{\lambda _k}'(u_k),u_k\rangle =0\) and \(\langle \varphi _{\lambda _k}'(u_k),\bar{u}\rangle =0\), we have

$$\begin{aligned} \Vert u_k\Vert ^2_{\lambda _k}=\int \limits _{\mathbb {R}^3}f(x,u_k)u_kdx-\int \limits _{\mathbb {R}^3}K(x)\phi _{u_k}u_k^2dx \end{aligned}$$
(5.3)

and

$$\begin{aligned} (u_k,\bar{u})_{\lambda _k}=\int \limits _{\mathbb {R}^3}f(x,u_k)\bar{u}dx-\int \limits _{\mathbb {R}^3}K(x)\phi _{u_k}u_k \bar{u}dx. \end{aligned}$$
(5.4)

From (5.1) and \((f_1)\), for any \(\varepsilon >0\), there exists \(C_\varepsilon >0\) such that

$$\begin{aligned} |f(x,t)|\le \varepsilon |t|+C_\varepsilon |t|^{p-1},\ \ \ \ \ \ \forall (x,t)\in \mathbb {R}^3\times \mathbb {R}. \end{aligned}$$

Hence we obtain

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}f(x,u_k)(u_k-\bar{u})dx\right|&\le \varepsilon \int \limits _{\mathbb {R}^3}|u_k||u_k-\bar{u}|dx+C_\varepsilon \int \limits _{\mathbb {R}^3}|u_k|^{p-1}|u_k-\bar{u}|dx\nonumber \\&\le \varepsilon |u_k|_2|u_k-\bar{u}|_2+C_\varepsilon |u_k|_p^{p-1}|u_k-\bar{u}|_p\nonumber \\&= o(1) \end{aligned}$$
(5.5)

since \(u_k\rightarrow \bar{u}\) in \(L^p(\mathbb {R}^3)\) \((2<p<6)\), \((u_k)\subset E\) is bounded and \(\varepsilon \) has been chosen arbitrarily. Similar to (2.7), we have

$$\begin{aligned} \left| \int \limits _{\mathbb {R}^3}K(x)\phi _{u_k}u_k(u_k-\bar{u})dx\right| \le |\phi _{u_k}|_6|u_k|_6\left( \int \limits _{\mathbb {R}^3}K(x)(u_k-\bar{u})^{3/2}dx\right) ^{2/3}\rightarrow 0. \end{aligned}$$
(5.6)

Using (5.3)-(5.6) and recalling \(\bar{u}(x)=0\) if \(V(x)>0\), we obtain

$$\begin{aligned} \Vert u_k\Vert ^2\le \Vert u_k\Vert _{\lambda _k}^2=(u_k,\bar{u})_{\lambda _k}+o(1)=\int \limits _{\mathbb {R}^3}\nabla u_k\cdot \nabla \bar{u} dx+o(1)=\Vert \bar{u}\Vert ^2+o(1). \end{aligned}$$
(5.7)

It follows from the weak lower semicontinuity that

$$\begin{aligned} \Vert \bar{u}\Vert ^2\le \liminf _{k\rightarrow \infty }\Vert u_k\Vert ^2, \end{aligned}$$

which, jointly with (5.7), shows that \(u_k\rightarrow \bar{u}\) in \(E\). The proof is complete. \(\square \)