1 Introduction and main results

Let us consider the following (stationary) Dirac equation

$$\begin{aligned} -i\sum ^3_{k=1}\alpha _k\partial _ku + a\beta u + V(x)u = G_u(x,u) \end{aligned}$$
(1.1)

for \(x=(x_1,x_2,x_3)\in \mathbb {R}^3\), where \(\partial _k=\partial /\partial x_k, a>0\) is a constant, \(\alpha _1, \alpha _2, \alpha _3\) and \(\beta \) are \(4\times 4\) Pauli-Dirac matrices:

$$\begin{aligned} \beta = \left( \begin{array}{cc} I &{} 0\\ 0 &{} -I \end{array} \right) , \qquad \alpha _k=\left( \begin{array}{cc} 0 &{} \sigma _k\\ \sigma _k &{} 0 \end{array} \right) , \quad k=1,2,3 \end{aligned}$$

with

$$\begin{aligned} \sigma _1= \left( \begin{array}{cc} 0 &{} 1\\ 1 &{} 0 \end{array} \right) , \quad \sigma _2= \left( \begin{array}{cc} 0 &{} -i\\ i &{} 0 \end{array} \right) , \quad \sigma _3= \left( \begin{array}{cc} 1 &{} 0\\ 0 &{} -1 \end{array} \right) . \end{aligned}$$

This equation arises when one seeks for the standing wave solutions of the nonlinear Dirac equation (see [25])

$$\begin{aligned} -i\hbar \partial _t\psi =ic\hbar \sum ^3_{k=1}\alpha _k\partial _k \psi - mc^2\beta \psi - M(x)\psi + F_\psi (x,\psi ). \end{aligned}$$
(1.2)

Assuming that \(F(x,e^{i\theta } \psi ) = F(x,\psi )\) for all \(\theta \in [0,2\pi ]\), a standing wave solution of (1.2) is a solution of the form \(\psi (t,x) = e^{\frac{i\mu t}{\hbar }}u(x)\). It is clear that \(\psi (t,x)\) solves (1.2) if and only if u(x) solves (1.1) with \(a=mc/\hbar , V(x)=M(x)/c\hbar +\mu I_4/\hbar \) and \(G(x,u)=F(x,u)/c\hbar \).

For notational convenience, denoting

$$\begin{aligned} \alpha =(\alpha _1,\alpha _2,\alpha _3)\ \ \text {and}\ \ \alpha \cdot \nabla =\sum ^3_{k=1}\alpha _k\partial _k, \end{aligned}$$

we rewrite the Eq. (1.1) as

figure a

There are many papers studying the existence and multiplicity of standing wave of the equations under different assumptions on the potentials V and G, see, [3, 811, 1418, 21, 23] and their references. Recall that, mathematically, the conditions that the potential functions depend periodically on x is used for describing a class of self-interaction of quantum electrodynamics in, e.g. [1, 2, 4, 5, 19, 20, 24, 26] for Schrödinger equations and [3] for Dirac equations. Note that if the potentials are periodic in x one may also study the existence and multiplicity of periodic solutions. Naturally, a periodic solution of \((D_V)\) may be referred as a standing periodic wave of (1.2). In recently paper [12], we have investigated periodic solutions of \((D_V)\) in both cases that the nonlinearity \(G_u(x,u)\) is of superlinear and subcritical growth as \(|u|\rightarrow \infty \). The case of concave and convex has been researched in the paper [13].

In the present paper, we are interested in the case that G(xu) is asymptotically quadratic at 0 and \(\infty \) and obtain the existence and multiplicity results of periodic solutions.

We make the following periodicity hypothesis on V(x) and G(xu):

(V):

\(V\in C(\mathbb {R}^3,\mathbb {R})\), and V(x) is 1-periodic in \(x_k, k=1,2,3\).

\((G_0)\) :

\(G\in C^1(\mathbb {R}^3\times \mathbb {C}^4,[0,\infty ))\), and G(xu) is 1-periodic in \(x_k, k=1,2,3\).

We are looking for periodic solutions of \((D_V)\): \(u(x+z)=u(x)\) for any \(z\in \mathbb {Z}^3\).

Setting \(Q=[0,1]\times [0,1]\times [0,1]\), if u is a solution of \((D_V)\), its energy will be denoted by

$$\begin{aligned} \Phi (u)=\int _Q\left[ \frac{1}{2}(-i\alpha \cdot \nabla u + a\beta u + V(x) u)\cdot u-G(x,u)\right] dx, \end{aligned}$$
(1.3)

where (here and in the following) by \(v\cdot w\) we denote the scalar product in \(\mathbb {C}^4\) of v and w.

In order to state our results, let \(A_0=-i\alpha \cdot \nabla +a\beta \) and \(A_V=A_0+V\) denote the self-adjoint operators acting in \(L^2(Q,\mathbb {C}^4)\). Let \(\{\lambda _j\}_{j\in \mathbb {Z}}\) denote the sequence of all eigenvalues of \(A_V\) counted by multiplicity:

$$\begin{aligned} \ldots \le \lambda _{-2}\le \lambda _{-1}<\lambda _0=0<\lambda _1\le \lambda _2\le \ldots , \end{aligned}$$

and let \(\{e_j\}_{j\in \mathbb {Z}}\) be the associated sequence of eigenvectors of \(A_V\):

$$\begin{aligned} A_Ve_j=\lambda _je_j,\quad |e_j|_{L^2}=1,\quad j=\pm 1,\pm 2,\dots . \end{aligned}$$
(1.4)

Remark 1.1

We can find out all eigenvalues and the associated eigenfunctions of \(A_0\). Let

$$\begin{aligned} z=(k_1,k_2,k_3)\in \mathbb {N}^3,\quad x=(x_1,x_2,x_3)\in Q,\quad z x=k_1x_1+k_2x_2+k_3x_3, \end{aligned}$$

and \(| z|=\sqrt{k^2_1+k^2_2+k^2_3}\). Note that

$$\begin{aligned} A_0= \left( \begin{array}{ll} aI &{}\quad -i(\sigma _1\partial _1+\sigma _2\partial _2+\sigma _3\partial _3)\\ -i(\sigma _1\partial _1+\sigma _2\partial _2+\sigma _3\partial _3) &{}\quad -aI \end{array}\right) \end{aligned}$$

and

$$\begin{aligned}-i(\sigma _1\partial _1{\text{ e }}^{2\pi zxi} +\sigma _2\partial _2{\text{ e }}^{2\pi zxi}+\sigma _3\partial _3{\text{ e }}^{2\pi zxi}) =2\pi {\text{ e }}^{2\pi zxi}W,\end{aligned}$$

where \(W=\left( \begin{array}{ll} k_3 &{} \quad k_1-ik_2\\ k_1+ik_2 &{} \quad -k_3 \end{array}\right) .\) Setting \(D=\left( \begin{array}{ll} aI &{} 2\pi W\\ 2\pi W &{} -aI \end{array}\right) ,\) one can verify that if \(\lambda \ne 0\) is a eigenvalue of the matrix D and \(\mathbf v\) is a eigenvector corresponding to \(\lambda \), then \(\lambda \) must be a eigenvalue of \(A_0\) and \({\text{ e }}^{2\pi zxi}\mathbf v\) is a eigenfunction corresponding to \(\lambda \). By \(|\lambda I-D|=0\) we obtain

$$\begin{aligned} \begin{aligned}&\quad \ \left| \begin{array}{cc} (\lambda -a)I &{} -2\pi W \\ -2\pi W &{} (\lambda +a)I \end{array} \right| \\&\quad = \left| \begin{array}{llll} (\lambda -a) &{}\quad 0 &{} \quad -2\pi k_3 &{}\quad -2\pi (k_1-ik_2)\\ 0 &{}\quad (\lambda -a) &{}\quad -2\pi (k_1+ik_2) &{}\quad 2\pi k_3\\ -2\pi k_3 &{}\quad - 2\pi (k_1-ik_2) &{}\quad (\lambda +a) &{}\quad 0\\ -2\pi (k_1+ik_2) &{}\quad 2\pi k_3 &{} \quad 0 &{}\quad (\lambda +a) \end{array} \right| \\&\quad =(\lambda ^2-a^2-4\pi ^2|z|^2)^2=0, \end{aligned}\end{aligned}$$

and therefore

$$\begin{aligned} \lambda =\pm \sqrt{a^2+4\pi ^2|z|^2}. \end{aligned}$$

For \(\mathbf{v}=(c_1,c_2,c_3,c_4)\), in virtue of \(D\mathbf{v}^T=\lambda \mathbf{v}^T \) we get

$$\begin{aligned} \left\{ \begin{array}{l} 2\pi k_3c_3+2\pi (k_1-ik_2)c_4=(\lambda -a)c_1, \\ 2\pi (k_1+ik_2)c_3-2\pi k_3c_4=(\lambda -a)c_2, \end{array}\right. \end{aligned}$$

and so

$$\begin{aligned}\left\{ \begin{array}{l} \mathbf{v^{(1)}_\lambda }=(2\pi |z|^2,\ 0,\ (\lambda -a)k_3,\ (\lambda -a)(k_1+ik_2)),\\ \mathbf{v^{(2)}_\lambda }=(0,\ 2\pi |z|^2,\ (\lambda -a)(k_1-ik_2),\ (a-\lambda ) k_3). \end{array}\right. \end{aligned}$$

Put

$$\begin{aligned} {\bar{\mathbf{e}}_1}=(1,0,0,0), {\bar{\mathbf{e}}_2}=(0,1,0,0), {\bar{\mathbf{e}}_3}=(0,0,1,0), {\bar{\mathbf{e}}_4}=(0,0,0,1), \end{aligned}$$

then

$$\begin{aligned} \left. \begin{array}{l} \varphi ^{(1)}_\lambda (x):= {\text{ e }}^{2\pi i zx}[ 2\pi |z|^2{\bar{\mathbf{e}}_1}+ (\lambda -a) k_3{\bar{\mathbf{e}}_3}+(\lambda -a)(k_1+ik_2){\bar{\mathbf{e}}_4}],\\ \varphi ^{(2)}_\lambda (x):= {\text{ e }}^{2\pi i zx}[2\pi |z|^2 {\bar{\mathbf{e}}_2} +(\lambda -a)(k_1-ik_2) {\bar{\mathbf{e}}_3}-(\lambda -a) k_3 {\bar{\mathbf{e}}_4}] \end{array} \right. \end{aligned}$$
(1.5)

satisfy \(A_0\varphi ^{(j)}_\lambda =\lambda \varphi ^{(j)}_\lambda , j=1,2\).

We will use the following hypotheses:

\((G_1)\) :

there is \(b_0\ge 0\) such that and \(G_u(x,u)-b_0u=o(|u|)\) as \(u\rightarrow 0\) uniformly in \(x\in Q\);

\((G_2)\) :

there is \(b_\infty >0\) satisfying \(G_u(x,u)-b_\infty u=o(|u|)\) as \(|u|\rightarrow \infty \) uniformly in \(x\in Q\);

\((G_3)\) :

either (i) \(b_\infty \notin \sigma (A_V)\) or (ii) \(G_u(x,u)-b_\infty u\) is bounded and \(G(x,u)-\frac{1}{2}b_\infty |u|^2\rightarrow \infty \) as \(|u|\rightarrow \infty \) uniformly in \(x\in Q\).

Set

$$\begin{aligned} G^0(x,u):=G(x,u)-\frac{1}{2}b_0|u|^2,\ G^\infty (x,u):=G(x,u)-\frac{1}{2}b_\infty |u|^2,\end{aligned}$$

and define

$$\begin{aligned} b_0^+:=\min [\sigma (A_V)\cap (b_0,\infty )],\ \ b_\infty ^-:=\max [\sigma (A_V)\cap (b_0,b_\infty )].\end{aligned}$$

The first result reads as follows.

Theorem 1.2

Let \((V), (G_0)\) and \((G_1)-(G_3)\) be satisfied and \(b_\infty >b^+_0\). Then

  1. (a)

    if \(G^0(x,u)\ge 0\), then \((D_V)\) has at least one nontrivial periodic solution in \(H^1(Q,\mathbb {C}^4)\);

  2. (b)

    if G is even in u, then \((D_V)\) has at least \(d(b_0,b_\infty )\) pairs of periodic solutions, where \(d(b_0,b_\infty )\) denotes the dimensionality of the eigenspace associated to \(\sigma (A_V)\cap (b_0,b_\infty )\).

If \(b_0\equiv 0\), then \(b^+_0=\lambda _1\), we have

Corollary 1.3

Assume that \((V), (G_0)\) and \((G_1)-(G_3)\) hold with \(b_0=0\). If \(b_\infty >\lambda _1\), then \((D_V)\) has at least one nontrivial periodic solution in \(H^1(Q,\mathbb {C}^4)\). If G is in addition even in u, then \((D_V)\) has at least \(d(0, b_\infty )\) pairs of periodic solutions.

If \(V(x)\equiv 0\), that is, \(A_V=A_0\), then the equation \((D_V)\) becomes the following

figure b

We write \(\{\mu _j\}\) the sequence of all eigenvalues of \(A_0\) according to the size of order, not by multiplicity:

$$\begin{aligned} ...<\mu _{-2}<\mu _{-1}<\mu _0=0<\mu _1<\mu _2<.... \end{aligned}$$

Let \(\sharp _{\mu _k}\) define the multiplicity of \(\mu _k\), and \(\lambda ^{(\mu _k)}_{j}\) the eigenvalues such that \(\lambda ^{(\mu _k)}_{j}=\mu _k,\ j=1,\dots ,\ \sharp _{\mu _k}\).

Let N[j] denote the number of \(z\in {\mathbb {N}}^3\) corresponding to \(|z|^2=j\). For \(0\le |z|^2\le 10\), we have:

$$\begin{aligned} N[0]=N[3]=1;\ N[j]=3,\ j=1,2,4,6,8; \end{aligned}$$
$$\begin{aligned} N[k]=6,\ k=5,9,10;\ \ N[7]=0, \end{aligned}$$

then by Remark 1.1,

$$\begin{aligned} \mu _j=\sqrt{a^2+4(j-1)\pi ^2},\ 1\le j\le 7;\ \mu _k=\sqrt{a^2+4k\pi ^2},\ k=8,9,10,\end{aligned}$$

and

$$\begin{aligned}\sharp _{\mu _1}=\sharp _{\mu _4}=1;\ \sharp _{\mu _j}=3,j=2,3,5,7,8;\ \sharp _{\mu _k}=6,k=6,9,10.\end{aligned}$$

Accordingly, we see

$$\begin{aligned}\begin{aligned}&\lambda ^{(\mu _1)}_1=\mu _1=a,\ \lambda ^{(\mu _2)}_1=\lambda ^{(\mu _2)}_2=\lambda ^{(\mu _2)}_3 =\sqrt{a^2+4\pi ^2},\\&\lambda ^{(\mu _3)}_1=\lambda ^{(\mu _3)}_2=\lambda ^{(\mu _3)}_3 =\sqrt{a^2+8\pi ^2},\ \lambda ^{(\mu _4)}_8=\sqrt{a^2+12\pi ^2},\\&\lambda ^{(\mu _5)}_1=\lambda ^{(\mu _5)}_2=\lambda ^{(\mu _5)}_3=\sqrt{a^2+16\pi ^2},\ \lambda ^{(\mu _6)}_1=\cdots =\lambda ^{(\mu _6)}_6=\sqrt{a^2+20\pi ^2},\\&\lambda ^{(\mu _7)}_1=\lambda ^{(\mu _7)}_2=\lambda ^{(\mu _7)}_3=\sqrt{a^2+24\pi ^2},\ \lambda ^{(\mu _8)}_1=\lambda ^{(\mu _8)}_2=\lambda ^{(\mu _8)}_3=\sqrt{a^2+32\pi ^2},\\&\lambda ^{(\mu _9)}_1=\cdots =\lambda ^{(\mu _9)}_6=\sqrt{a^2+36\pi ^2},\ \lambda ^{(\mu _{10})}_1=\cdots =\lambda ^{(\mu _{10})}_6=\sqrt{a^2+40\pi ^2}. \end{aligned}\end{aligned}$$

By (1.5), we can list the first 10 eigenvalues \(\lambda _j\) and eigenfunctions \(e_j\) corresponding to \(\lambda _j\) as follows:

$$\begin{aligned}\begin{aligned} \lambda _1&=\lambda _2=\mu _1=a \ \ \text {with}\ \ z=(0,\ 0,\ 0),\\ e_1&=(1,\ 0,\ 0,\ 0),\ \ e_2=(0,\ 1,\ 0,\ 0);\\ \lambda _3&=\lambda _4=\mu _2=\sqrt{a^2+4\pi ^2} \ \ \text {with}\ \ z=(1,0,0),\\ e_3&=\Delta _1{\text{ e }}^{2\pi x_1i}(2\pi ,\ 0,\ 0,\ \mu _2-a),\\ e_4&=\Delta _1{\text{ e }}^{2\pi x_1i}(0,\ 2\pi ,\ \mu _2-a,\ 0);\\ \lambda _5&=\lambda _6=\mu _2 \ \ \text {with}\ \ z=(0,1,0),\\ e_5&=\Delta _1{\text{ e }}^{2\pi x_2i}(2\pi ,\ 0,\ 0,\ (\mu _2-a)i),\\ e_6&=\Delta _1{\text{ e }}^{2\pi x_2i}(0,\ 2\pi ,\ (a-\mu _2)i,\ 0);\\ \lambda _7&=\lambda _8=\mu _2 \ \ \text {with}\ \ z=(0,\ 0,\ 1),\\ e_7&=\Delta _1{\text{ e }}^{2\pi x_3i}(2\pi ,\ 0,\ \mu _2-a,0),\\ e_8&=\Delta _1{\text{ e }}^{2\pi x_3i}(0,\ 2\pi ,\ 0,\ a-\mu _2);\\ \lambda _9&=\lambda _{10}=\mu _3=\sqrt{a^2+8\pi ^2} \ \ \text {with}\ \ z=(1,\ 1,\ 0),\\ e_9&=\Delta _2{\text{ e }}^{2\pi (x_1+x_2)i}(4\pi ,\ 0,\ (\mu _3-a)(1+i),\ 0),\\ e_{10}&=\Delta _2{\text{ e }}^{2\pi (x_1+x_2)i}(0,\ 4\pi ,\ 0,\ (\mu _3-a)(1-i)), \end{aligned} \end{aligned}$$

where \(\Delta _1=\frac{1}{\sqrt{4\pi ^2+(\mu _2-a)^2}},\ \Delta _2=\frac{1}{\sqrt{16\pi ^2+2(\mu _3 -a)^2} }.\)

Now we have a special consequence corresponding to the equation \((D_0)\).

Corollary 1.4

Let \((G_0)\) and \((G_1)-(G_3)\) be satisfied with \(b_0=0\). Then \((D_0)\) has at least one nontrivial periodic solution in \(H^1(Q,\mathbb {C}^4)\), provided \(b_\infty >a\). If moreover G is in even in u and \(b_\infty ^-=\mu _k\) for some positive integer k, then \((D_0)\) has at least \(l:=2(\sharp _{\mu _1}+\dots +\sharp _{\mu _k})\) pairs of periodic solutions.

A more general result can be obtained if \((G_1)\) is replaced by

\((G'_1)\) :

there is \(b_0\in C(Q,[0,\infty ))\) such that \(b_0(x)\) is 1-period with \(b_0(x)\ge 0\) and \(G_u(x,u)-b_0(x)u=o(|u|)\) as \(|u|\rightarrow \infty \) uniformly in \(x\in Q\),

\((G_2)\) is replaced by

\((G'_2)\) :

there is \(b_\infty \in C(Q,(0,\infty ))\) such that \(b_\infty (x)\) is 1-period and \(G_u(x,u)-b_\infty (x)u=o(|u|)\) as \(|u|\rightarrow \infty \) uniformly in \(x\in Q\),

and \((G_3)\) is replaced by

\((G'_3)\) :

either (i) \(0\notin \sigma (A_V-b_\infty )\) or (ii) \({\hat{G}}(x,u):= \frac{1}{2}\hat{G}_u(x,u)u- G(x,u)\ge 0\) and \({\hat{G}}(x,u)\rightarrow \infty \) as \(|u|\rightarrow \infty \) uniformly in \(x\in Q\).

Theorem 1.5

Suppose that \((V), (G_0), (G'_1)-(G'_3)\) are satisfied and \(q_\infty >q_0^+\), where \(q_\infty :=\min \limits _{x\in Q}b_\infty (x),\ q_0^+:=\min [\sigma (A_V)\cap (q_0,\infty )]\) and \(q_0:= \max \limits _{x\in Q}b_0(x) \). Then

  1. (a)

    if \(G(x,u)-\frac{1}{2} q_0|u|^2\ge 0\), then \((D_V)\) has at least one nontrivial periodic solution in \(H^1(Q,\mathbb {C}^4)\);

  2. (b)

    if G is even in u, then \((D_V)\) has at least \(d(q_0,q_\infty )\) pairs of periodic solutions.

This paper is organized as follows. In Sect. 2, we state the variational setting and establish a deformation theorem and abstract critical point theorems under the Cerami condition (\((C)_c\)-condition). The proofs of the main results are given in Sect. 3.

2 Variational setting and abstract critical point theorems

To prove our main results, some preliminaries are first in order.

In what follows by \(|\cdot |_q\) we denote the usual \(L^q\)-norm, and \((\cdot ,\cdot )_2\) the usual \(L^2\)-inner product. Let

$$\begin{aligned} L_T^q(Q):=\{u\in L_{loc}^q(\mathbb {R}^3,\mathbb {C}^4): u(x+{\hat{e}}_i)=u(x)\ a.e.\ ,\quad i=1,2,3\}, \end{aligned}$$

where \({\hat{e}}_1=(1,0,0), {\hat{e}}_2=(0,1,0), \hat{e}_3=(0,0,1)\). Let \(A_0=-i\alpha \cdot \nabla +a\beta ,\, A_V=A_0+V\) denote the self-adjoint operators on \(L^2(Q,{\mathbb {C}}^4)\) with domain

$$\begin{aligned}&\mathcal {D}(A_V)=\mathcal {D}(A_0)=H_T^1(Q)\\&\quad :=\{u\in H^1_{loc}(\mathbb {R}^3,\mathbb {C}^4):u(x+{\hat{e}}_i)=u(x)\ a.e.\ ,i=1,2,3\}. \end{aligned}$$

Set \(E:=\mathcal {D}(|A_V|^{\frac{1}{2}})\) which is a Hilbert space with the inner product and norm, for \(u=\sum _{j\in \mathbb {Z}} a_je_j\) and \(v=\sum _{j\in \mathbb {Z}} b_je_j\in E\),

$$\begin{aligned} (u, v)=\sum _{j\not =0}|\lambda _j|a_j\cdot b_j+(u^0,v^0)_2 \quad \text {and}\quad \Vert u\Vert ^2=\sum _{j\not =0}|\lambda _j||a_j|^2+|u^0|^2_2, \end{aligned}$$
(2.1)

here \(\{e_j\}_{j\in \mathbb {Z}}\) are the eigenvectors of \(A_V\).

Then we have an orthogonal decomposition \(E=E^-\oplus E^0\oplus E^+\) with \(E^-:=\hbox {span}\{e_j: j<0\}, E^+:=\hbox {span}\{e_j: j>0\}\), and \(E^0:=\ker (A_V)\). Note that if \(0\not \in \sigma (A_V)\) then \(E^0=\{0\}\).

The functional \(\Phi \) defined by (1.3) can be rewritten by

$$\begin{aligned} \Phi (u)=\frac{1}{2}\big (\Vert u^+\Vert ^2-\Vert u^-\Vert ^2\big )-\int _{Q}G(x,u) \end{aligned}$$

for \(u=u^-+u^0+u^+\in E\). Then \(\Phi \in C^1(E,\mathbb {R})\) and critical points of \(\Phi \) are solutions of \((D_V)\).

First we have the following (see [8, 11])

Lemma 2.1

\(E=H^{1/2}(Q,\mathbb {C}^4)\) with equivalent norms, hence E embeds compactly into \(L_T^s (Q)\) for all \(s\in [1,3)\). In particular there is a constant \(a_s>0\) such that

$$\begin{aligned} |u|_s\le a_s\Vert u\Vert \quad \text { for all }\;\; u\in E. \end{aligned}$$
(2.2)

We also use the following result, the proof is similar to that of Proposition B.10 in [22].

Lemma 2.2

Assume that

  1. (i)

    \(G\in C^1(Q\times \mathbb {C}^4,\mathbb {R})\), and

  2. (ii)

    there are \(k_1, k_2>0\) such that

    $$\begin{aligned} |G_u(x,u)|\le k_1+k_2|u|^s, \quad \forall (x,u)\in Q\times \mathbb {C}^4, \end{aligned}$$

    where \(0\le s<3\).

Then

$$\begin{aligned} \psi (u):=\int _QG(x,u) \end{aligned}$$
(2.3)

is weakly continuous and \(\psi '\in C(E, \mathbb {R})\) is compact.

Recall that a sequence \(\{u_j\}\) in E is said to be a \((C)_c\)-sequence of \(\Phi \), if \(\Phi (u_j)\rightarrow c\) and \((1+\Vert u_j\Vert )\Phi '(u_j)\rightarrow 0\) as \(j\rightarrow \infty \). We say that \(\Phi \) satisfies the \((C)_c\)-condition if any \((C)_c\)-sequence possesses a convergent subsequence ([6]).

Let X be a Banach space, and

$$\begin{aligned} \Phi ^b_a:=\Phi _a\cap \Phi ^b,\ \Phi _a:=\{u\in X: \Phi (u)\ge a \},\, \Phi ^b:=\{u\in X: \Phi (u)\le b \}. \end{aligned}$$

We first establish a deformation theorem which plays an important role in the multiplicity for \((D_V)\).

Theorem 2.3

Let \(\Phi \in C^1(X,\mathbb {R})\) and satisfy the \((C)_c\)-condition, \( K_c=\{u\in X: \Phi (u)=c\) and \(\Phi '(u)=0\}\). If \({\bar{\varepsilon }}>0\) and \(\mathcal {O}\) is any neighborhood of \(K_c\), then there exists an \(\varepsilon \in (0,{\bar{\varepsilon }})\) and a deformation \(\eta \in C([0,1]\times X,X)\) such that

\(1^\circ \, \eta (0,u)=u\) for all \(u\in X\).

\( 2^\circ \,\eta (t,u)=u\) for all \(t\in [0,1]\) if \(u\notin \Phi ^{c+\varepsilon }_{c-\varepsilon }\).

\( 3^\circ \,\eta (t,\cdot ): X\rightarrow X\) is homeomorphism for \(t\in [0,1]\).

\( 4^\circ \,\Phi (\eta (\cdot ,u))\) is nonincreasing on [0, 1] for \(u\in E\).

\( 5^\circ \,\eta (1,\Phi ^{c+\varepsilon }\setminus \mathcal {O})\subset \Phi ^{c-\varepsilon }\).

\( 6^\circ \) If \(K_c=\emptyset , \eta (1,\Phi ^{c+\varepsilon })\subset \Phi ^{c-\varepsilon }\).

\( 7^\circ \) If \(\Phi (u)\) is even in \(u, \eta (t,u)\) is odd in u.

Proof

By the \((C)_c\)-condition, \(K_c\) is compact. Set \(U_\delta =\{u\in X:\ d(u,K_c)<\delta \}\). Choosing \(\delta \) suitably small (\(\delta <1\)), \(U_\delta \subset \mathcal {O}\). Therefore it suffices to prove \(5^\circ \) with \(\mathcal {O}\) replaced by \(U_\delta \). Note that \(U_\delta =\emptyset \) when \(K_c=\emptyset \), and so we get \(6^\circ \) instead.

Let \(M>0\) such that \(\Vert u\Vert \le M\) for all \(u\in U_\delta \).

One can easy to verify that there are \({\hat{\varepsilon }}>0\) and \(\alpha >0\) such that

$$\begin{aligned} (1+\Vert u\Vert )\Vert \Phi '(u)\Vert \ge \alpha ,\ \ \text {for all}\ \ u\in \Phi ^{c+{\hat{\varepsilon }}}_{c-{\hat{\varepsilon }}}\setminus U_{\delta /2}. \end{aligned}$$
(2.4)

We may assume that

$$\begin{aligned} 0<{\hat{\varepsilon }}< \frac{3\delta }{8(1+M)}\min \left\{ {\bar{\varepsilon }},\alpha ^2,\frac{1}{4}\right\} . \end{aligned}$$
(2.5)

Let \(\tilde{X}:=\{u\in X\, | \, \Phi '(u)\not =0\}\) and \(V: \tilde{X}\rightarrow X\) be a pseudo gradient such that V is odd if \(\Phi \) is even (see [22]). Choosing any \(\varepsilon \in (0,{\hat{\varepsilon }})\), define

$$\begin{aligned} h(s)=\left\{ \begin{array}{ll} 1,\, &{}\quad {\text {if}}\, 0\le s\le 1,\\ \frac{1}{s},\, &{}\quad {\text {if}}\, s>1, \end{array}\right. \end{aligned}$$
$$\begin{aligned} f(u)=\frac{d(u,X\setminus \Phi ^{c+{\hat{\varepsilon }}}_{c-{\hat{\varepsilon }}})}{d(u,X\setminus \Phi ^{c+{\hat{\varepsilon }}}_{c-{\hat{\varepsilon }}})+d(u,\Phi ^{c+\varepsilon }_{c-\varepsilon })},\ g(u)=\frac{d(u, U_{\delta /8})}{d(u, U_{\delta /8})+d(u,X\setminus U_{\delta /4})}. \end{aligned}$$

Then

$$\begin{aligned}f |_{\Phi ^{c+\varepsilon }_{c-\varepsilon }}=g |_{X\setminus U_{\delta /4}}=1,\, \, f |_{X\setminus \Phi ^{c+{\hat{\varepsilon }}}_{c-{\hat{\varepsilon }}}}=g |_{U_{\delta /8}}=0.\end{aligned}$$

Let

$$\begin{aligned} W(u)=\left\{ \begin{array}{ll} -f(u)g(u)h((1+\Vert u\Vert )\Vert V(u)\Vert )(1+\Vert u\Vert )^2V(u),\, &{}\quad u\in {\tilde{X}},\\ 0,\, &{}\quad \text {otherwise}. \end{array}\right. \end{aligned}$$

It is easy to verify that

$$\begin{aligned} \Vert W(u)\Vert \le 1+\Vert u\Vert \qquad \text {for all }u. \end{aligned}$$
(2.6)

Then by construction, W is locally Lipschitz continuous on X and W is odd if \(\Phi \) is even.

Now we consider the Cauchy problem:

$$\begin{aligned} \frac{{\text {d}}\eta }{{\text {d}}t}=W(\eta ), \, \, \eta (0,u)=u. \end{aligned}$$
(2.7)

By virtue of the locally Lipschitz continuity of W and (2.6), the basic existence uniqueness theorem for ordinary differentia equations implies that for each \(u\in X\), (2.7) has a unique solution \(\eta (t, u)\) defined for \(t\in [0, \infty )\), and \(\eta \in C([0,1]\times X, X)\). (2.7) implies that \(1^\circ \) holds. Since \(f(u)=0\) on \(X\setminus \Phi ^{c+{\hat{\varepsilon }}}_{c-{\hat{\varepsilon }}}\), so \(2^\circ \) is true. The semigroup property for solutions of (2.7) gives \(3^\circ \). The oddness of W when \(\Phi \) is even yields \(7^\circ \).

If \(W(u)\not =0,\ u\in {\tilde{X}}\) so V(u) is defined as is \(V(\eta (t,u)) \) and

$$\begin{aligned} \begin{aligned} \frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}&= \, \ (\Phi '(\eta (t,u)), W(\eta (t,u))) \\&= \, -f(\eta )g(\eta )h((1+\Vert \eta \Vert )\Vert V(\eta )\Vert )(1+\Vert \eta \Vert )^2(\Phi '(\eta ),V(\eta ))\\&\le \, -f(\eta )g(\eta )h((1+\Vert \eta \Vert )\Vert V(\eta )\Vert )(1+\Vert \eta \Vert )^2\Vert \Phi '(\eta )\Vert ^2\le 0. \end{aligned} \end{aligned}$$
(2.8)

It follows that \(4^\circ \) holds.

Finally, we verify \(\eta (1,\Phi ^{c+\varepsilon }\setminus U_\delta )\subset \Phi ^{c-\varepsilon }\). Let \(u\in \Phi ^{c+\varepsilon }\setminus U_\delta \), then \(\Phi (\eta (t,u))\le c+\varepsilon \) by \( 4^\circ \) and \( 1^\circ \). We need only prove that there exists \(t_0\in [0,1]\) such that \(\Phi (\eta (t_0,u))\le c-\varepsilon \), then \(4^\circ \) gives \(\Phi (\eta (1,u))\le c-\varepsilon \).

If otherwise, then \(\Phi (\eta (t,u))>c-\varepsilon \) for all \(t\in [0,1]\), and thus \(\eta (t,u)\in \Phi ^{c+\varepsilon }_{c-\varepsilon }\), which implies

$$\begin{aligned} \Phi (\eta (0,u))-\Phi (\eta (t,u))\le 2\varepsilon <2{\hat{\varepsilon }},\ \forall t\in [0,1]. \end{aligned}$$
(2.9)

If \(\eta (t,u)\in X\setminus U_{\delta /2}\) for all \(t\in [0,1]\), we see \( \eta (t,u)\in \Phi ^{c+\varepsilon }_{c-\varepsilon }\setminus U_{\delta /2}\). This shows \(f(\eta (t,u))=g(\eta (t,u))=1\) and by (2.4),

$$\begin{aligned} (1+\Vert \eta (t,u)\Vert )\Vert \Phi '(\eta (t,u))\Vert \ge \alpha ,\ \forall t\in [0,1]. \end{aligned}$$
(2.10)

This yields

$$\begin{aligned} \begin{aligned} \frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}&= -h((1+\Vert \eta \Vert )\Vert V(\eta )\Vert )(1+\Vert \eta \Vert )^2(\Phi '(\eta ),V(\eta ))\\&\le -h((1+\Vert \eta \Vert )\Vert V(\eta )\Vert )(1+\Vert \eta \Vert )^2\Vert \Phi '(\eta )\Vert ^2,\ \ \forall t\in [0,1]. \end{aligned} \end{aligned}$$
(2.11)

If \((1+\Vert \eta \Vert )\Vert V(\eta )\Vert \le 1\), then \(h((1+\Vert \eta \Vert )\Vert V(\eta )\Vert )=1\). It follows from (2.10) and (2.11) that

$$\begin{aligned} \frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}\le -\alpha ^2. \end{aligned}$$
(2.12)

If \((1+\Vert \eta \Vert )\Vert V(\eta )\Vert >1\), then

$$\begin{aligned} h((1+\Vert \eta \Vert )\Vert V(\eta )\Vert )=[(1+\Vert \eta \Vert )\Vert V(\eta )\Vert ]^{-1}, \end{aligned}$$

so (2.11) and the property of \(V(\cdot )\) imply

$$\begin{aligned} \frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}\le -(1+\Vert \eta \Vert )\Vert V(\eta )\Vert \left[ \frac{\Vert \Phi '(\eta )\Vert }{\Vert V(\eta )\Vert }\right] ^2 \le -\frac{1}{4}. \end{aligned}$$
(2.13)

Consequently, by (2.12) and (2.13) we have

$$\begin{aligned} \frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}\le -\min \left\{ \alpha ^2,\frac{1}{4}\right\} \ \ \text {for all}\ t\in [0,1]. \end{aligned}$$
(2.14)

Integrating (2.14) and combing the result with (2.9) gives

$$\begin{aligned} \begin{aligned}&2{\hat{\varepsilon }}\ge \Phi (\eta (0,u))-\Phi (\eta (1,u))\\&\quad = \int _0^1-\frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}\ge \min \left\{ \alpha ^2,\frac{1}{4}\right\} , \end{aligned} \end{aligned}$$
(2.15)

this is contrary to (2.5). Consequently, we infer that there is \({\bar{t}}\in [0,1]\) such that \(\eta ({\bar{t}},u)\in U_{\delta /2} \). Obviously, \({\bar{t}}>0\) since \(\eta (0,u)=u \not \in U_\delta \). The continuity of \(\eta (t,u)\) guarantees that there are \(s_1,s_2\in [0,1]\) with \(s_1\not =s_2\) such that \(\eta (s_1,u)\in \partial U_{\delta /4},\ \eta (s_1,u)\in \partial U_\delta \) and \(\eta (t,u)\in U_\delta \setminus \overline{U}_{\delta /4}\) for all \(t\in (s_1,s_2)\) or \(t\in (s_2,s_1)\), where \(\overline{B}\) denotes the closure of B. This yields

$$\begin{aligned} \Vert \eta (s_1,u)-\eta (s_2,u)\Vert \ge 3\delta /4.\end{aligned}$$
(2.16)

By (2.6) we see \(\Vert W(u)\Vert \le 1+ M\) for all \(u\in U_\delta \), and so

$$\begin{aligned} \Vert \eta (s_2,u)-\eta (s_1,u)\Vert \le (1+M)|s_2-s_1| \end{aligned}$$

which together with (2.16) shows

$$\begin{aligned} |s_2-s_1|\ge \frac{3\delta }{4(1+M)}. \end{aligned}$$

We may assume that \(s_1<s_2\).

On the other hand, similarly to (2.15) we get that

$$\begin{aligned}\begin{aligned} 2{\hat{\varepsilon }}&\ge \Phi (\eta (s_1,u))-\Phi (\eta (s_2,u))\\&= \int _{s_1}^{s_2}-\frac{{\text {d}}\Phi (\eta (t,u))}{{\text {d}}t}\\&\ge \min \left\{ \alpha ^2,\frac{1}{4}\right\} (s_2-s_1)\\&\ge \frac{3\delta }{4(1+M)}\min \left\{ \alpha ^2,\frac{1}{4}\right\} . \end{aligned} \end{aligned}$$

This, however, leads to a contradiction. The proof is complete. \(\square \)

Remark 2.4

In paper [12] (or [13]), we established a deformation theorem under the \((C)_c\)-condition. However, it is difficult to use for the multiplicity. Therefore, Theorem 2.3 improves the corresponding result in [12].

In order to study the functional \(\Phi \), we need certain abstract critical point theorems. In the following, we suppose that E is a real Hilbert space with \(E=X\oplus Y.\)

Theorem 2.5

Let \(e\in X\setminus \{0\}\) and \(\Omega =\{u=s e+v:\Vert u\Vert < R, s>0,v\in Y\}\). Suppose that

\((\Phi _1)\) :

\(\Phi \in C^1(E,\mathbb {R})\), satisfies the \((C)_c\)-condition for any \(c\in \mathbb {R}\);

\((\Phi _2)\) :

there is a \(r\in (0,R)\) such that \(\rho :=\inf \Phi (X\cap \partial B_r)>\omega :=\sup \Phi (\partial \Omega )\), where \(\partial \Omega \) refers to the boundary of \(\Omega \) relative to \(\hbox {span}\{e\}\oplus Y\), and \(B_r=\{u\in E:\ \Vert u\Vert <r\}\).

Then \(\Phi \) has a critical value \(c\ge \rho \), with

$$\begin{aligned} c=\inf _{h\in \Gamma }\sup _{u\in \Omega }\,\Phi (h(u)), \end{aligned}$$

here

$$\begin{aligned} \begin{aligned} \Gamma =\{h\in C(E, E): h|_{\partial \Omega }=\text {id},\ \Phi (h(u))\le \Phi (u)\ \text {for}\ u\in \overline{\Omega }\}. \end{aligned} \end{aligned}$$
(2.17)

Proof

Put \(S=X\cap \partial B_r\). We first show that for any \(h\in \Gamma , h(\Omega )\cap S\ne \emptyset .\) We may assume \(\Vert e\Vert =1\). Chose \({\hat{e}}\in Y\) with \(\Vert {\hat{e}}\Vert =1\), and write \(F:=\hbox {span}\{e,{\hat{e}}\}, \Omega _F:=F\cap \Omega .\) Let \(\overline{\Omega }_F, \partial \Omega _F \) denote the closure and bound of \(\Omega \) in F, respectively, P the project of E onto Y. For \(u\in \overline{\Omega }_F, t\in [0,1],\) define

$$\begin{aligned} H(t,u)=t[\Vert (\text {id}-P)h(u)\Vert e+Ph(u)]+(1-t)u. \end{aligned}$$

Then \(H:[0,1]\times \overline{\Omega }_F\rightarrow E\) is continuous. Obviously H is a compact operator. Since \(h|_{\partial \Omega }=\text {id},\) if \(u\in \partial \Omega _F\),

$$\begin{aligned} H(t,u)=t[\Vert u-Pu\Vert e+Pu]+(1-t)u=u, \end{aligned}$$

i.e., \(H(t,\cdot )|_{\partial \Omega _F}=\text {id}\) for \(t\in [0,1]\). In particular \(H(t,u)\ne re\) for \(t\in [0,1], u\in \partial \Omega _F\). By the property of Brouwer degree, we have

$$\begin{aligned} \deg (H(1,\cdot ),\Omega _F, re)=\deg (H(0,\cdot ),\Omega _F, re)=\deg (\text {id},\Omega _F, re)=1 \end{aligned}$$

which implies that there exists \(u\in \Omega _F\) such that \(H(1,u)=re\in S\). We find \(Ph(u)=0, \Vert h(u)\Vert =r\), i.e. \(h(u)\in S\), and therefore \(c\ge \rho \).

Next we prove there is a sequence \(\{u_j\}\) in \(\Omega \) such that

$$\begin{aligned} (1+\Vert u_j\Vert )\Vert \Phi '(u_j)\Vert \rightarrow 0\ \ \text {for}\ j\rightarrow \infty . \end{aligned}$$
(2.18)

Indeed otherwise there exist \(\alpha _0>0\) and \(\varepsilon _0>0\) such that

$$\begin{aligned} (1+\Vert u\Vert )\Vert \Phi '(u)\Vert \ge \alpha _0\ \ \ \text {for all}\ u\in \Omega \cap \Phi ^{c+\varepsilon _0}_{c-\varepsilon _0}. \end{aligned}$$

Set \({\bar{\varepsilon }}=\min \{\frac{1}{2}(\rho -\omega ),\varepsilon _0\}\). There is an \(\varepsilon \in (0,{\bar{\varepsilon }})\) and \(\eta \in C([0,1]\times E, E)\) given by Theorem 2.3 such that \(1^\circ -4^\circ \) and \(6^\circ \) are satisfied. Chose \(h\in \Gamma \) such that \(\sup \Phi (h(\Omega ))\le c+\varepsilon .\) Consequently

$$\begin{aligned} h(\Omega )\subset \Phi ^{c+\varepsilon }. \end{aligned}$$
(2.19)

Let \(g(u):=\eta (1,h(u))\), then \(g\in C(E,E)\). It follows from \(3^\circ \) and \(1^\circ \) that

$$\begin{aligned} \Phi (g(u))=\Phi (\eta (1,h(u)))\le \Phi (\eta (0,h(u)))=\Phi (h(u))\le \Phi (u) \end{aligned}$$

for all \(u\in \overline{\Omega }\). For \(u\in \partial \Omega , (\Phi _2)\) shows

$$\begin{aligned} \Phi (u)\le \omega <\rho -{\bar{\varepsilon }} \le c-{\bar{\varepsilon }}\le c-\varepsilon \end{aligned}$$

which, by \(2^\circ \), implies \(\eta (1,u)=u\), and so

$$\begin{aligned} g(u)=\eta (1,h(u))=\eta (1,u)=u. \end{aligned}$$

Thus \(g\in \Gamma \). (2.19) and \(6^\circ \) yield \(g(\Omega )=\eta (1,h(\Omega ))\subset \Phi ^{c-\varepsilon }\) which leads to the contradiction \(c\le \sup \Phi (g(\Omega ))\le c-\varepsilon \).

Now we find that there is a sequence \(\{u_j\}\) in \(\Omega \) satisfying (2.18). Since \(\Phi \) satisfies \((\Phi _1)\) (the \((C)_c\)-condition), there exists a convergent subsequence \(\{u_{j_k}\}\) of \(\{u_j\}\) such that \(u_{j_k}\rightarrow {\bar{u}}\). The conclusion follows by \(\Phi \in C^1(E,E)\). \(\square \)

Remark 2.6

In [22], Theorem 5.3], under the conditions that Y is finite dimensional and \(\Phi \) satisfies the (PS)-condition, the same result was proved. Clearly, the conditions of Theorem 2.5 are weaker than that of Theorem 5.3.

Next, we consider a kind of pseudo-index (see [7]). Let \(\Sigma \) denote the class of closed subsets of E symmetric with respect to the origin, and \(\gamma : \Sigma \rightarrow \mathbb {N}\cup \{\infty \}\) the \(\mathbb {Z}_2\) genus map (see [22]). Let \(\Phi \in C(E,{\mathbb {R}}),\ J=(\sigma ,\infty )\),

$$\begin{aligned} {\mathcal {H}}=\{h\in C(E,E): \ h\ \text {is a homeomorphism and is odd}\}, \end{aligned}$$
$$\begin{aligned} {\mathcal {M}}_J=\{g\in {\mathcal {H}} : \ g|_{\Phi ^{-1}(\mathbb {R}\setminus J)}=\text {id} \ \text {and}\ \Phi (g(u))\le \Phi (u)\ \text {for}\ u\in E\}, \end{aligned}$$

and \(\Lambda _*=\{h\in {\mathcal {M}}_J : \ h(B_1Y)\subset \Phi ^{-1}(J)\cup B_rY\}\).

Now we define the pseudo-index \((\Sigma , i^*)\) relative to \({\mathcal {M}}_J\) for the genus \(\gamma \) as follows

$$\begin{aligned} i^*(A)=\inf \limits _{h\in \Lambda _*}\gamma (A\cap h(S_1Y)). \end{aligned}$$

One can verify the following

Lemma 2.7

Let \(\Sigma ^*=\Sigma \), then \((\Sigma ^*, i^*)\) satisfies all properties for pseudo-index ([7]):

(P1):

\(\Sigma ^*\subset \Sigma ,\ \overline{A\setminus B}\in \Sigma ^*\) and \( \overline{g(A)}\in \Sigma ^*\) for all \(A\in \Sigma ^*,\ B\in \Sigma \) and \(g\in {\mathcal {M}}_J\);

(P2):

\(A\subset B\) implies \(i^*(A)\le i^*(B) \) for all \(A, B\in \Sigma ^*\);

(P3):

\(i^*(\overline{A\setminus B})\ge i^*(A)-\gamma (B)\) for all \(A\in \Sigma ^*\) and \(B\in \Sigma \);

(P4):

\(i^*(\overline{g(A)})\ge i^*(A)\) for all \(A\in \Sigma ^*\) and \(g\in {\mathcal {M}}_J\).

Now, we give a abstract critical point theorem as follows.

Theorem 2.8

Assume that \(\Phi \) is even and satisfies \( (\Phi _1)\). If

\( (\Phi _3)\) :

there exists \(r>0\) with \(\rho :=\inf \Phi (S_rY)>\Phi (0)=0\), where \(S_r:=\partial B_r,\ AB:=A\cap B\);

\( (\Phi _4)\) :

there exists a finite dimensional subspace \(Y_0\subset Y\) and \(R>r\) such that for \(E_*:=X\oplus Y_0, M_*=\sup \Phi (E_*)<+\infty \) and \(\sigma :=\sup \Phi (E_*\setminus B_R)<\rho \),

then \(\Phi \) possesses at least m distinct pairs of critical points, where \(m=\dim Y_0\).

Proof

Let

$$\begin{aligned} \Sigma _k=\{A\in \Sigma : i^*(A)\ge k\}, \quad k=1,2,\ldots , m.\end{aligned}$$

Define

$$\begin{aligned} c_k =\inf \limits _{A\in \Sigma _k }\sup \limits _{u\in A}\Phi (u), \quad k=1,2,\ldots ,m. \end{aligned}$$
(2.20)

We first show \(\Sigma _k\not =\emptyset \). Set \({\tilde{A}}:=B_RE_*\). \((\Phi _4)\) implies \(\Phi ^{-1}(J)\subset (E\setminus E_*)\cup B_R\), and hence

$$\begin{aligned} {\tilde{A}}\supset Y_0\cap (\Phi ^{-1}(J)\cup B_RY)\supset Y_0\cap h(B_1Y) \end{aligned}$$

for each \(h\in \Lambda _*\), which yields

$$\begin{aligned} {\tilde{A}}\cap h(S_1Y)\supset Y_0\cap h(S_1Y)\supset \partial (Y_0\cap h(B_1Y)), \end{aligned}$$

and we get

$$\begin{aligned} \gamma ({\tilde{A}}\cap h(S_1Y))\ge \gamma (\partial (Y_0\cap h(B_1Y)))\ge m. \end{aligned}$$

Consequently, \(\Sigma _k\not =\emptyset \), and \(c_m\le M_*\) by \((\Phi _4)\). For any \(A\in \Sigma _k\), by \(h:=r\text {id}\in \Lambda _*\) one has

$$\begin{aligned} \gamma (A\cap S_rY)=\gamma (A\cap h(S_1Y)\ge i^*(A)\ge k \end{aligned}$$

which yields \(c_k\ge \rho \) by \((\Phi _3)\). Noting that \(\Sigma _1\supset \Sigma _2\supset \cdots \supset \Sigma _m\), we have

$$\begin{aligned} \sigma <\rho \le c_1\le c_2\le \cdots \le c_m\le M_*. \end{aligned}$$

It is obvious that \(K_c:=\{u\in X: \Phi (u)=c\) and \(\Phi '(u)=0\}\in \Sigma \), and \(K_c\) is compact by the \((C)_c \)-condition.

Finally, we claim:

\((P^*)\) :

If \( 1\le j,\ j+l\le m\), and \(c_j=\cdots =c_{j+l}\equiv c\), then \(\gamma (K_c)\ge l+1\).

If \(\gamma (K_c)\le l\), then there is a \(\delta >0\) such that \(\gamma (U_\delta (K_c))=\gamma (K_c)\le l\). Invoking Theorem 2.3 with \(\mathcal {O}=U_\delta (K_c)\) and \({\bar{\varepsilon }}=\frac{\rho -\sigma }{2}\), there are \(\varepsilon \in (0,{\bar{\varepsilon }})\) and \(\eta \in C([0,1]\times E, E)\) such that \(\eta (1,\cdot )\) satisfies the properties \(1^\circ -7^\circ \) and

$$\begin{aligned} \eta (1,\Phi ^{c+\varepsilon }\setminus \mathcal {O})\subset \Phi ^{c-\varepsilon }. \end{aligned}$$
(2.21)

Choose \({\hat{A}}\in \Sigma _{j+l}\) such that \(\sup \Phi ({\hat{A}})\le c+\varepsilon ,\) and hence

$$\begin{aligned} {\hat{A}}\subset \Phi ^{c+\varepsilon }. \end{aligned}$$
(2.22)

By (P3) one has

$$\begin{aligned} i^*(\overline{{\hat{A}}\setminus \mathcal {O}})\ge i^*(\hat{A})-\gamma (\mathcal {O})\ge j+l-l=j. \end{aligned}$$
(2.23)

Using \(3^\circ \) and \(7^\circ \) we get \(\eta (1,\cdot )\in {\mathcal {H}}\). \(4^\circ \) gives \(\Phi (\eta (1,u))\le \Phi (u)\) for all \(u\in E\). Since \(\sigma <c-\varepsilon \), we have \(\Phi ^{-1}(\mathbb {R}\setminus J)\subset E\setminus \Phi ^{c+\varepsilon }_{c-\varepsilon }\), and \(2^\circ \) implies \(\eta (1,\cdot )|_{\Phi ^{-1}(\mathbb {R}\setminus J)}=\text {id}\). Therefore \(\eta (1,\cdot )\in {\mathcal {M}}_J\). Set \(A_*:=\eta (1,\overline{\hat{A}\setminus \mathcal {O}})\in \Sigma \). It follows from (P4) and (2.23) that

$$\begin{aligned} i^*(A_*)=i^*\left( \overline{\eta (1,\overline{\hat{A}\setminus \mathcal {O}})}\right) \ge i^*\left( \overline{{\hat{A}}\setminus \mathcal {O}}\right) \ge j, \end{aligned}$$

and thus \(A_*\in \Sigma _j\). Combing with (2.21), (2.22) and (2.20) we see

$$\begin{aligned} c\le \sup \Phi (A_*)\le c-\varepsilon <c, \end{aligned}$$

a contradiction. Therefore, the conclusion \((P^*)\) is valid and the proof is complete. \(\square \)

3 The proof of the main results

Throughout this section, we suppose that (V) and \((G_0)\) are satisfied.

Observe that, \((G_2)\) implies that for any \(\varepsilon >0\) there is \(R_\varepsilon >0\) such that

$$\begin{aligned} |G_u(x,u)-b_\infty u|\le \varepsilon |u| \hbox { whenever } |u|\ge R_\varepsilon , \end{aligned}$$
(3.1)

hence

$$\begin{aligned} |G_u(x,u)u-b_\infty |u|^2|\le |G_u(x,u)-b_\infty u||u|\le \varepsilon |u|^2 \end{aligned}$$

or

$$\begin{aligned} (b_\infty -\varepsilon )|u|^2\le G_u(x,u)u\le (b_\infty +\varepsilon )|u|^2 \quad \text {for all }|u|\ge R_\varepsilon . \end{aligned}$$

Fixed \(s_0\in (0,1)\), in virtue of \(G(x,u)\ge 0\) we get

$$\begin{aligned} \begin{aligned} G(x,u)&=\,G(x,s_0u)+\int ^1_{s_0}G_u(x,su)\cdot uds\\&\ge \,\int ^1_{s_0}\frac{1}{s}G_u(x,su)suds\\&\ge \,\frac{1}{2}(b_\infty -\varepsilon )(1-s^2_0)|u|^2 \end{aligned} \end{aligned}$$

for all \(|u|\ge \frac{1}{s_0}R_\varepsilon \), and so

$$\begin{aligned} G(x,u)\ge \frac{1}{2}(b_\infty -\varepsilon )(1-s^2_0)|u|^2-C_{s_0}\quad \text {for all }(x,u). \end{aligned}$$
(3.2)

First, we have the following lemma.

Lemma 3.1

Suppose that \((G_1)\) and \((G_2)\) hold and \(\{u_j\}\) is a bounded \((C)_c\)-sequence of \(\Phi \). Then there exists a critical point u of \(\Phi \) such that \(\Phi (u)=c\) and after passing to a subsequence, \(u_j\rightarrow u\) strongly in E.

Proof

By Lemma 2.1, without loss of generality, we may assume that

$$\begin{aligned} u_n\rightharpoonup u\ \text {in}\ E\ \text {and}\ u_u\rightarrow u\ \text {in}\ L_T^s(Q)\ \text {for}\ s\in [1,3). \end{aligned}$$
(3.3)

Plainly, u is a critical point of \(\Phi \). \((G_1)\) and \((G_2)\) yield that

$$\begin{aligned} |G_u(x,u)|\le C_1|u| \ \text {for}\ \text {all} \ (x,u) \end{aligned}$$
(3.4)

which shows that \(\psi '\) is continuous and compact by Lemma 2.2, where \(\psi \) is defined by (2.3). It follows from the representation of \(\Phi '\), together with (3.3), the facts \(\Phi '(u)=0\) and \(\Phi '(u_n)\rightarrow 0\), and the compactness of \(\psi '\), that

$$\begin{aligned} \ \ \Vert u_n^+-u^+\Vert ^2= & {} (\Phi '(u_n)-\Phi '(u),u_n^+-u^+)\\&+(\psi '(u_n)-\psi '(u),u_n^+-u^+ )\rightarrow 0 \ \ \text {as}\ \ n\rightarrow \infty . \end{aligned}$$

Similarly, \(\Vert u^-_n-u^-\Vert \rightarrow 0\) as \(n\rightarrow \infty \). It is clear that \(\{u_j^0\}\) has a convergent subsequence since \(E^0\) is finite dimensional. We have thus proved the lemma. \(\square \)

Lemma 3.2

If \(b_\infty >\lambda _1\) and \((G_3)\) holds, then any \((C)_c\)-sequence of \(\Phi \) is bounded.

Proof

Let \(\{u_j\}\subset E\) be such that \(\Phi (u_j)\rightarrow c\) and \((1+\Vert u_j\Vert )\Phi '(u_j)\rightarrow 0\).

Defining

$$\begin{aligned} {\tilde{E}}^+:= & {} \left\{ u\in E:\ u=\sum _{\lambda _j>b_\infty }a_je_j\right\} ,\\ {\tilde{E}}^0:= & {} \left\{ u\in E:\ u=\sum _{\lambda _j=b_\infty }a_je_j\right\} ,\\ {\tilde{E}}^-:= & {} \left\{ u\in E:\ u=\sum _{\lambda _j<b_\infty ,\lambda _j\ne 0}a_je_j+u^0,u^0\in E^0\right\} , \end{aligned}$$

we have \(E={\tilde{E}}^+\oplus {\tilde{E}}^0\oplus {\tilde{E}}^-\) and write \(u=\tilde{u}^++{\tilde{u}}^0+{\tilde{u}}^-\) for \(u\in E\) corresponding to this decomposition. Clearly, \({\tilde{E}}^0=\{0\}\) if \(\hbox {b}_\infty \notin \sigma (A_V)\).

Let \(P^{\pm }: E\rightarrow E^{\pm }\) be the orthogonal projections. One can see

$$\begin{aligned} (\Phi '(u),{\tilde{u}}^+)= & {} \Vert \tilde{u}^+\Vert ^2-b_\infty |{\tilde{u}}^+|^2_2 -\displaystyle \int _QG^\infty _u(x,u){\tilde{u}}^+,\nonumber \\ (\Phi '(u),{\tilde{u}}^-)= & {} (P^+u,{\tilde{u}}^-)-(P^-u,{\tilde{u}}^-)\nonumber \\&-b_\infty |{\tilde{u}}^-|^2_2-\displaystyle \int _QG^\infty _u(x,u){\tilde{u}}^-. \end{aligned}$$
(3.5)

For \(u=\sum _{j\in \mathbb {Z},j\ne 0} a_je_j+u^0\in E\ (u^0\in E^0)\), we have

$$\begin{aligned} {\tilde{u}}^+=\sum _{\lambda _j>b_\infty }a_je_j,\ \ \ \ {\tilde{u}}^-=\sum _{\lambda _j<b_\infty ,\lambda _j\ne 0}a_je_j+u^0. \end{aligned}$$

By (2.1) one finds

$$\begin{aligned} \begin{aligned} \Vert {\tilde{u}}^+\Vert ^2-b_\infty |{\tilde{u}}^+|^2_2&= \sum _{\lambda _j>b_\infty }\lambda _j|a_j|^2-b_\infty \sum _{\lambda _j>b_\infty }|a_j|^2\\&\ge \, \left( 1-\frac{b_\infty }{\lambda '}\right) \Vert {\tilde{u}}^+\Vert ^2, \end{aligned} \end{aligned}$$
(3.6)

where \(\lambda ':=\min (\sigma (A_V)\cap (b_\infty ,\infty ))\). Since \(b_\infty >\lambda _1, \sigma (A_V)\cap (0,b_\infty )\ne \emptyset \). Setting \(\lambda '':=\max (\sigma (A_V)\cap (0,b_\infty ))\), we obtain

$$\begin{aligned} \begin{aligned}&\,(P^+u,{\tilde{u}}^-)-(P^-u,{\tilde{u}}^-) -b_\infty |{\tilde{u}}^-|^2_2\\&\quad =\sum _{0<\lambda _j<b_\infty }\lambda _j|a_j|^2-\sum _{\lambda _j<0}|\lambda _j||a_j|^2 -b_\infty \sum _{\lambda _j<b_\infty ,\lambda _j\ne 0}|a_j|^2-b_\infty |u^0|^2_2\\&\quad \le \Vert {\tilde{u}}^-\Vert ^2 -2\sum _{\lambda _j<0}|\lambda _j| |a_j|^2 -\frac{b_\infty }{\lambda ''}\sum _{0<\lambda _j<b_\infty }\lambda _j|a_j|^2-(1+b_\infty )|u^0|^2_2, \end{aligned} \end{aligned}$$

and therefore

$$\begin{aligned} -(P^+u,{\tilde{u}}^-)+(P^-u,{\tilde{u}}^-)+b_\infty |\tilde{u}^-|^2_2\ge (w-1)\Vert {\tilde{u}}^-\Vert ^2, \end{aligned}$$
(3.7)

here \(w:=\min \{1+b_\infty ,2,\frac{b_\infty }{\lambda ''}\}\). For \(\delta >0\) small, it follows from (3.1) that

$$\begin{aligned} |G^\infty _u(x,u)|<\delta |u|+C_\delta ,\ \text {for all } (x,u). \end{aligned}$$
(3.8)

Putting \(u_j={\tilde{u}}^+_j+{\tilde{u}}^-_j+{\tilde{u}}^0_j\), by (3.5) we know

$$\begin{aligned} \left. \begin{array}{ll} &{}\ \Vert {\tilde{u}}^+_j\Vert ^2-b_\infty |\tilde{u}^+_j|^2_2 =(\Phi '(u_j),{\tilde{u}}^+_j) +\displaystyle \int _QG^\infty _u(x,u_j){\tilde{u}}^+_j,\\ &{}\,-(P^+u_j,{\tilde{u}}^-_j)+(P^-u_j,{\tilde{u}}^-_j)+b_\infty |{\tilde{u}}^-_j|_2^2\\ &{}\qquad \qquad \qquad =-(\Phi '(u_j),{\tilde{u}}^-_j)- \displaystyle \int _QG^\infty _u(x,u_j){\tilde{u}}^-_j. \end{array}\right. \end{aligned}$$
(3.9)

(3.6)–(3.9) and (2.2) yield

$$\begin{aligned} \begin{aligned} \xi \Vert {\tilde{u}}^+_j+{\tilde{u}}^-_j\Vert ^2&\le \Vert \Phi '(u_j)\Vert \Vert {\tilde{u}}^+_j+{\tilde{u}}^-_j\Vert \\&\quad \,\, + \delta C'\Vert u_j\Vert \Vert {\tilde{u}}^+_j+{\tilde{u}}^-_j\Vert +C'_\delta \Vert {\tilde{u}}^+_j+{\tilde{u}}^-_j\Vert \end{aligned} \end{aligned}$$
(3.10)

with \(\xi =\min \{1-\frac{b_\infty }{\lambda '},w-1\}\).

If (i) of \((G_3)\) holds, then \(u_j={\tilde{u}}^+_j+{\tilde{u}}^-_j\). (3.10) implies that

$$\begin{aligned} \xi \Vert u_j\Vert \le \Vert \Phi '(u_j)\Vert + \delta C'\Vert u_j\Vert +C'_\delta , \end{aligned}$$

and so \(\{u_j\}\) is bounded.

Next let (ii) of \((G_3)\) be satisfied. (3.6), (3.7) and (3.9) yield that \(\{{\tilde{u}}^+_j+{\tilde{u}}^-_j\}\) is bounded. We claim that \(\{{\tilde{u}}^0_j\}\) is bounded.

Assume by contradiction that \(\Vert {\tilde{u}}^0_j\Vert \rightarrow \infty \) as \(j\rightarrow \infty \). Since \({\tilde{E}}^0\) is finite dimensional, we have: along a subsequence, there exists \(Q_0\subset Q\) satisfying \(|Q_0|>0\) such that \(|{\tilde{u}}^0_j(x)|\rightarrow \infty \) as \(j\rightarrow \infty \) uniformly in \(x\in Q_0\). Here, we write |W| for the Lebesgue measure of \(W\subset {\mathbb {R}}^3\). It follows from the hypotheses that \(G^\infty (x,\tilde{u}^0_j)\rightarrow \infty \) as \(j\rightarrow \infty \) uniformly in \(x\in Q_0\), and thus

$$\begin{aligned} \begin{aligned} G^\infty (x,u_j)&=G^\infty (x,{\tilde{u}}^0_j) +\int _0^1G^\infty _u(x,s(u_j-{\tilde{u}}^0_j))({\tilde{u}}^+_j+{\tilde{u}}^-_j)ds\\&\ge G^\infty (x,{\tilde{u}}^0_j)-K_1\Vert {\tilde{u}}^+_j+{\tilde{u}}^-_j\Vert \rightarrow \infty \end{aligned} \end{aligned}$$
(3.11)

as \(j\rightarrow \infty \) uniformly in \(x\in Q_0\).

By (3.2) and \(G^\infty (x,u)\rightarrow \infty \) as \(|u|\rightarrow \infty \) we obtain that there exists \(m_0>0\) such that

$$\begin{aligned} G^\infty (x,u)\ge -m_0\ \ \text {for all }\ (x,u). \end{aligned}$$
(3.12)

Noting that

$$\begin{aligned} \begin{aligned}&\, \Vert u^+\Vert ^2-\Vert u^-\Vert ^2-b_\infty |u|^2_2\\&\quad =\,\Vert {\tilde{u}}^+\Vert ^2+\sum _{0<\lambda _j<b_\infty }\lambda _j|a_j|^2 -\sum _{\lambda _j<0}|\lambda _j||a_j|^2 -b_\infty |{\tilde{u}}^++{\tilde{u}}^-|^2_2, \end{aligned} \end{aligned}$$

we get by (2.2)

$$\begin{aligned} \left| \Vert u^+\Vert ^2-\Vert u^-\Vert ^2-b_\infty |u|^2_2\right| \le (1+a_2b_\infty )(\Vert {\tilde{u}}^+\Vert ^2+\Vert {\tilde{u}}^-\Vert ^2). \end{aligned}$$
(3.13)

On account of (3.13), (3.12) and (3.11) we see that

$$\begin{aligned} \begin{aligned} |\Phi (u_j)|&=\, \left| \frac{1}{2}(\Vert u_j^+\Vert ^2-\Vert u_j^-\Vert ^2-b_\infty |u_j|^2_2) -\int _QG^\infty (x,u_j)\right| \\&\ge \, \left| \int _QG^\infty (x,u_j)\right| -\frac{1}{2}(1+a_2b_\infty )(\Vert {\tilde{u}}^+_j\Vert ^2+\Vert {\tilde{u}}^-_j\Vert ^2)\\&\ge \, \int _{Q_0}G^\infty (x,u_j)-m_0-\frac{1}{2}(1+a_2b_\infty )(\Vert \tilde{u}^+_j\Vert ^2+\Vert {\tilde{u}}^-_j\Vert ^2) \rightarrow \infty \end{aligned} \end{aligned}$$

as \(j\rightarrow \infty \), a contradiction. Consequently \(\{x^0_j\}\) is bounded and the proof is complete. \(\square \)

We need to introduce another orthogonal decomposition: \(E=\hat{E}^+\oplus {\hat{E}}^0\oplus {\hat{E}}^-\), where

$$\begin{aligned} \left\{ \begin{array}{l} {\hat{E}}^+:=\left\{ u\in E:\ u=\displaystyle \sum _{\lambda _j>b_0}a_je_j\right\} ,\\ {\hat{E}}^0:=\left\{ u\in E:\ u=\displaystyle \sum _{\lambda _j=b_0}a_je_j\right\} , \\ {\hat{E}}^-:=\left\{ u\in E:\ u=\displaystyle \sum _{\lambda _j<b_0,\lambda _j\ne 0}a_je_j+u^0,u^0\in E^0\right\} ,\end{array}\right. \end{aligned}$$
(3.14)

One can verify that there is \(\xi _0\in (0,1)\) such that

$$\begin{aligned} \begin{aligned}&\, \Vert {\hat{u}}^+\Vert ^2-b_0|{\hat{u}}^+|^2_2 \ge \xi _0 \Vert {\hat{u}}^+\Vert ^2,\\&\ (P^+u,{\hat{u}}^-)-(P^-u,{\hat{u}}^-)-b_0|{\hat{u}}^-|^2_2\le -\xi _0 \Vert {\hat{u}}^-\Vert ^2 \end{aligned} \end{aligned}$$
(3.15)

for any \(u={\hat{u}}^++{\hat{u}}^0+{\hat{u}}^-\in E\), the proof is similar to that of (3.6) and (3.7).

Lemma 3.3

Suppose that \((G_1)\) and \((G_2)\) hold, then there exist \(r>0\) and \(\rho >0\) such that \(\inf \Phi ({\hat{E}}^+\cap B_r)\ge 0\) and \(\inf \Phi (\hat{E}^+\cap \partial B_r)\ge \rho \).

Proof

Choosing \(q\in (2,3)\), we have that, for any \(\varepsilon >0\), there is \(C_\varepsilon >0\) such that

$$\begin{aligned} G^0(x,u)\le \varepsilon |u|^2+C_\varepsilon |u|^q,\, \hbox { for all }\, (x,u). \end{aligned}$$

This implies

$$\begin{aligned}\begin{aligned} \Phi ({\hat{u}}^+)&=\, \frac{1}{2}(\Vert {\hat{u}}^+\Vert ^2-b_0|{\hat{u}}^+|^2_2)-\int _QG^0(x,{\hat{u}}^+)\\&\ge \, \frac{1}{2}\xi _0\Vert {\hat{u}}^+\Vert ^2-\varepsilon C_1'\Vert {\hat{u}}^+\Vert ^2-C'_2C_\varepsilon \Vert {\hat{u}}^+\Vert ^p \end{aligned} \end{aligned}$$

via (3.15) for \({\hat{u}}\in {\hat{E}}^+\), which follows that the conclusion is valid. \(\square \)

Lemma 3.4

Let \((G_2)\) be satisfied. If \(b_\infty >b_0^+\), then for any \(n\in {\mathbb {N}}\) with \(b^-_\infty =\lambda _n \), there exists \(R_n>r\) such that \(\sup \Phi ( E_n\setminus B_{R_n})<0\) and \(\sup \Phi (E_n)<\infty \), where r is as in Lemma 3.3, \(E_n:=E^-\oplus E^0\oplus \hbox {span}\{e_1,...,e_n\}\).

Proof

It will suffice to show that for \(u\in E_n\)

$$\begin{aligned} \Phi (u)\rightarrow -\infty \ \ \ \text {as} \ \ \Vert u\Vert \rightarrow \infty . \end{aligned}$$
(3.16)

Choose \(s_0\in \left( 0,\sqrt{1-\frac{b^-_\infty }{b_\infty }}\right) \) in (3.2). Noting that \( u^+=\sum \nolimits _{j=1}^n s_{j}e_j\) for \(u\in E_n\), by (3.2), for \(\varepsilon =\frac{1}{2} (b_\infty -\frac{b^-_\infty }{1-s^2_0})\), we find

$$\begin{aligned} \begin{aligned} 2\Phi (u)&= \Vert u^+\Vert ^2-\Vert u^-\Vert ^2-2\int _QG(x,u)\\&\le \Vert u^+\Vert ^2-\Vert u^-\Vert ^2- \alpha _0(|u^+|_2^2+|u^0|_2^2+|u^-|^2_2)+2C_{s_0}, \end{aligned} \end{aligned}$$
(3.17)

where \(\alpha _0:=(b_\infty -\varepsilon )(1-s^2_0)>b^-_\infty \). Since

$$\begin{aligned}\begin{aligned} \alpha _0|u^+|_2^2-\Vert u^+\Vert ^2 \ge (\alpha _0-\lambda _n)\sum \limits _{j=1}^n|s_{j}|^2 \ge \frac{\alpha _0-b^-_\infty }{\lambda _1}\Vert u^+\Vert ^2_2, \end{aligned} \end{aligned}$$

by (3.17) we get

$$\begin{aligned} \begin{aligned} 2\Phi (u)&\le \,-\frac{\alpha _0-b^-_\infty }{\lambda _1}\Vert u^+\Vert ^2 -C_2(\Vert u^-\Vert ^2+\Vert u^0\Vert ^2)+2C_{s_0}\\&\le -{\hat{C}}\Vert u\Vert ^2+2C_{s_0} \end{aligned} \end{aligned}$$

which implies that (3.16) is valid and \(\sup \Phi (E_n)<\infty \). \(\square \)

As a consequence, we have

Lemma 3.5

Under the conditions of Lemma 3.4, if \(G^0(x,u)\ge 0\), then there is \(R_0>r\) such that \(\sup \Phi (\partial \Omega )\le 0\), where

$$\begin{aligned} \Omega :=\{u={\hat{u}}^-+{\hat{u}}^0+se_m: {\hat{u}}^-+{\hat{u}}^0\in {\hat{E}}^-\oplus {\hat{E}}^0, s>0, \Vert u\Vert <R_0\} \end{aligned}$$

with \(A_Ve_m=b^+_0e_m\), and \(\partial \Omega \) refers to the boundary of \(\Omega \) relative to \(\hbox {span}\{e_m\}\oplus {\hat{E}}^-\oplus {\hat{E}}^0\).

Proof

Since \({\hat{E}}^-\oplus {\hat{E}}^0\oplus {\mathbb {R}}^+e_m\subset E_m\) and \(\lambda _m=b_0^+\le b^-_\infty \), by Lemma 3.4 we find that \(\Phi (u)<0 \) for \(u={\hat{u}}^-+{\hat{u}}^0+se_m\) with \(\Vert u\Vert =R_0\) and \(s>0\) when \(R_0>r\) large.

Let \(u={\hat{u}}^-+{\hat{u}}^0\) with \(\Vert u\Vert \le R_0\). By \(G^0(x,u)\ge 0\) and (3.15) one has

$$\begin{aligned} \begin{aligned} 2\Phi (u)&= (P^+u,{\hat{u}}^-)-(P^-u,{\hat{u}}^-)-b_0|{\hat{u}}^-|^2_2- 2\int _QG^0(x,u)\\&\le -\xi _0\Vert {\hat{u}}^-\Vert ^2-2\int _QG^0(x,u) \le 0 \end{aligned} \end{aligned}$$

which yields that the result is valid. \(\square \)

Now, with the above arguments, we are ready to prove Theorem 1.2.

Proof of Theorem 1.2

(Existence) Let us verify the conditions of Theorem 2.5. Let \(X=\hat{E}^+,\ Y={\hat{E}}^-\oplus {\hat{E}}^0,\ r>0\) be from Lemma 3.3. Lemma 3.1 and 3.2 imply that \((\Phi _1)\) is true. Lemma 3.3 yields \(\inf \Phi (X\cap \partial B_r)\ge \rho \), and Lemma 3.5 gives \(\Phi |_{\partial \Omega }<\sigma _0\) for \(\sigma _0\in (0,\rho )\). Therefore \((\Phi _2)\) holds. It follows from Theorem 2.5 that \(\Phi \) possesses a critical value \(c\ge \rho \), with

$$\begin{aligned} c=\inf _{h\in \Gamma }\sup _{u\in \Omega }\,\Phi (h(1,u)), \end{aligned}$$

where \(\Gamma \) is defined as (2.17).

Next, we proceed to prove the multiplicity. Since G is even in \(u, \Phi \) is even. Using Lemma 3.3 we know that the condition \((\Phi _3)\) holds with \(X={\hat{E}}^-\oplus Y^0\) and \(Y=\hat{E}^+\). Let \(\hbox {span}\{e_m,\dots ,e_n\}\) be the eigenspace associated to \(\sigma (A_V)\cap (b_0,b_\infty )\), and \(\lambda _j\) the eigenvalue corresponding to \(e_j\) (i.e., \(A_Ve_j=\lambda _je_j),\ j=m,\dots ,n\), then \(b^+_0=\lambda _m,\ b^-_\infty =\lambda _n\) and \(d(b_0,b_\infty )=n-m\). It follow from Lemma 3.4 that \(\Phi \) satisfies \((\Phi _4)\) with \(Y_0=\hbox {span}\{e_m,\dots , e_n\}, R=R_n,\ M_*=M_n\) and \(\sigma \in (0,\rho )\). Therefore, \(\Phi \) has at least \(n-m\) pairs of nontrivial critical points by Theorem 2.8. \(\square \)

We are now in a position to give the proof of Theorem 1.5.

Proof

The main difference to the proof of Theorem 1.2 lies in the boundedness of the (C)c-sequences.

Claim 1. Any (C)c-sequence is bounded.

Let \(\{u_j\}\subset E\) be such that

$$\begin{aligned} \Phi (u_j)\rightarrow c,\ (1+\Vert u_j\Vert )\Phi '(u_j)\rightarrow 0 \ \text {as}\ \ j\rightarrow \infty . \end{aligned}$$

We then have

$$\begin{aligned} \int _Q{\hat{G}}(x,u_j)=\Phi (u_j)-\frac{1}{2}\Phi '(u_j)\cdot u_j\le C_0. \end{aligned}$$
(3.18)

Assume by contradiction that \(\Vert u_j\Vert \rightarrow \infty \). Then the normalized sequence \(v_j=u_j/\Vert u_j\Vert \) satisfies (up to a subsequence) \(v_j\rightharpoonup v\) in E. Lemma 2.1 guarantees \(v_j\rightarrow v\) in \(L_T^s(Q)\) for \(s\in [1,3)\) and \(|v_j|_s\le a_s \) for all \(s\in [1,3]\). We write \(\tilde{u}_j=u^-_j+u^+_j, \tilde{v}_j=v^-_j+v^+_j\). Then

$$\begin{aligned}\Phi '(u_j)(u^+_j-u^-_j) =\Vert u_j\Vert ^2\left( \Vert \tilde{v}_j\Vert ^2-\int _Q \frac{G_u(x,u_j)(v^+_j-v^-_j)|v_j|}{|u_j|}\right) , \end{aligned}$$

and therefore

$$\begin{aligned} o(1) =\Vert \tilde{v}_j\Vert ^2-\int _Q \frac{G_u(x,u_j)(v^+_j-v^-_j)|v_j|}{|u_j|}. \end{aligned}$$
(3.19)

We distinguish the two cases: \(v=0\) or \(v\ne 0\).

let \(v=0\). \((G'_1)\) and \((G'_2)\) yield that (3.4) is true, this implies

$$\begin{aligned}\int _{Q}\frac{|G_u(x,u_j)|}{|u_j|}|v^+_j-v^-_j|\,|v_j| \le C_1 |v_j|^2_2 \end{aligned}$$

which jointly with (3.19) shows \(\Vert \tilde{v}_j\Vert ^2\rightarrow 0\), and so \(|\tilde{v}_j|_2\rightarrow 0\). \(|v_j|_2\rightarrow 0\) yields \(|v^0_j|_2\rightarrow 0\). We obtain \(1=\Vert v_j\Vert =\Vert \tilde{v}_j\Vert +|v^0_j|_2\rightarrow 0\), a contradiction.

Assume \(v\ne 0\). First let (i) of \((G_3)\) hold. Since \(|u_j(x)|=|v_j(x)|\ \Vert u_j\Vert \rightarrow \infty \), by (3.4) and Lebesgue dominated convergence theorem we obtain

$$\begin{aligned} \int _{Q}\frac{G_u(x,u_j)v_j\varphi }{|u_j|}\rightarrow \int _{Q}b_\infty (x)v\varphi \end{aligned}$$

for any \(\varphi \in C^\infty [Q,{\mathbb {C}}^4]\), hence \(A_V v= b_\infty v\), which contradicts \(0\notin \sigma (A_V-b_\infty )\).

Suppose that (ii) of \((G_3)\) is satisfied. \(v_j\rightarrow v\) in \(L_T^s(Q)\) guarantees (up to a subsequence) \(v_j(x)\rightarrow v(x) \) a.e. on Q. Since \(v\ne 0\), there exists \(Q_0\subset Q\) with \(|Q_0|>0\) such that

$$\begin{aligned} v_j(x)\rightarrow v(x)\ \text { as }\ j\rightarrow \infty \ \ \text {uniformly on }\ Q_0 \end{aligned}$$

and \(|v_j(x)|\ge \varepsilon _0>0\) for large j. Observe that \(|u_j(x)|=\Vert u_j\Vert |v_(x)|\ge \varepsilon _0\Vert u_j\Vert \rightarrow \infty \) for \(x\in Q_0\). By (ii) of \((G_3)\) we have

$$\begin{aligned} \int _{Q_0}\hat{G}(x,u_j)\rightarrow \infty , \end{aligned}$$

which contradicts (3.18).

Next we have

Claim 2. The conclusions of Lemmas 3.33.5 are true where \((G_1)\) and \((G_2)\) are replaced by \((G'_1)\) and \((G'_2)\) respectively, and \(b_0\) is replaced by \(q_0\) in (3.14).

Since (3.2) and (3.4) are satisfied, where \(b_\infty \) is replaced by \(q_\infty \), one can prove as before.

Finally, repeating the arguments of the proof of Theorem 1.2, we obtain the desired results. \(\square \)