数学杂志  2016, Vol. 36 Issue (6): 1173-1182   PDF    
扩展功能
加入收藏夹
复制引文信息
加入引用管理器
Email Alert
RSS
本文作者相关文章
MAI A-li
SUN Guo-wei
GROUND STATE SOLUTIONS FOR NONLINEAR DIFIERENCE EQUATIONS WITH PERIODIC COEFFICIENTS
MAI A-li, SUN Guo-wei    
Department of Applied Mathematics, Yuncheng University, Yuncheng 044000, China
Abstract: In this paper, we study the existence of ground state solutions for nonlinear second order diffierence equations with periodic coefficients. Using the critical point theory in combination with the Nehari manifold approach, the existence of ground state solutions is established. Under a more general super-quadratic condition than the classical Ambrosetti-Rabinowitz condition, the results considerably generalize some existing ones. Finally, an example is also presented to demonstrate our results.
Key words: nonlinear diffierence equations     Nehari manifold     ground state solutions     critical point theory    
一类周期非线性差分方程的基态解
买阿丽, 孙国伟    
运城学院应用数学系, 山西 运城 044000
摘要:本文研究了一类二阶周期非线性差分方程基态解的存在性问题.利用临界点理论结合Nehari流形方法, 获得了此类方程基态解的存在性.在比经典AR条件更一般的超二次条件下, 本文结论推广了已有的结果, 并举例说明此类方程解的存在性.
关键词非线性差分方程    Nehari流形    基态解    临界点理论    
1 Introduction

In this paper, we consider the following nonlinear second order difference equation

$ \begin{eqnarray}\label{eqn1.1} \left\{\begin{array}{ll} \Delta[a(k)\Delta u(k-1)]-b(k)u(k)+f(k, u(k))=0, \quad k\in\mathbb{Z}, \\ u(k)\rightarrow 0, \quad \quad \quad |k|\rightarrow\infty, \end{array} \right. \end{eqnarray} $ (1.1)

where $a(k)$, $b(k)$ and $f(k, u)$ are $T$-periodic in $k$ $f(k, u):\mathbb{Z}\times \mathbb{R}\rightarrow\mathbb{R}$ is a continuous function on $u$. The forward difference operator $\Delta$ is defined by

$ \Delta u(k)=u(k+1)-u(k)\quad {\rm ~~for~~all~}k\in\mathbb{Z}, $

where $\mathbb{Z}$ and $\mathbb{R}$ denote the sets of all integers and real numbers, respectively.

The solutions of (1.1) are referred to as homoclinic solutions of the equation

$ \begin{eqnarray}\label{eqn1.2} \Delta[a(k)\Delta u(k-1)]-b(k)u(k)+f(k, u(k))=0, \quad k\in\mathbb{Z}. \end{eqnarray} $ (1.2)

In the theory of differential equations, homoclinic orbits play an important role in analyzing the chaos of dynamical systems. If a system has the transversely intersected homoclinic orbits, then it must be chaotic. If a system has the smoothly connected homoclinic orbits, then it can not stand the perturbation, its perturbed system probably produces chaotic. Therefore, it is of practical importance and mathematical significance to study the existence of homoclinic solutions.

Difference equations represent the discrete counterpart of ordinary differential equations. The classical methods are used in difference equations, such as numerical analysis, fixed point methods, linear and nonlinear operator theory, see [1-4]. In recent years, the existence and multiplicity of homoclinic solutions for difference equations have been studied in many papers by variational methods, see [5-12] and the reference therein.

Assume the following conditions hold:

(A) $a(k)>0$ and $a(k+T)=a(k)$ for all $k\in \mathbb{Z}$.

(B) $b(k)>0$ and $b(k+T)=b(k)$ for all $k\in \mathbb{Z}$.

(H1) $f\in C(\mathbb{Z}\times \mathbb{R}, \mathbb{R})$, and there exist $C>0$ and $p\in(2, \infty)$such that

$ \begin{eqnarray} |f(k, u)|\leq C(1+|u|^{p-1})~~{\rm for ~all} ~k\in \mathbb{Z}, u\in \mathbb R.\nonumber \end{eqnarray} $

(H2) $\lim\limits_{|u|\to 0}{f(k, u)/u}=0$uniformly for $ k\in \mathbb{Z}$.

(H3) $\lim\limits_{|u|\to\infty}{F(k, u)/|u|^{2}}=+\infty$uniformly for $ k\in\mathbb{Z}$, where $F(k, u)$ is the primitive function of $f(k, u)$, i.e.,

$ F(k, u)=\int_{0}^{u}f(k, s) ds. $

(H4) $u\mapsto{f(k, u)/|u|}$is strictly increasing on $(-\infty, 0)$ and $(0, \infty)$.

Remark1.1  (H2) implies that $u(k)\equiv 0$ is a trivial solution of (1.1).

Our main result is following.

Theorem1.1  Suppose that conditions (A), (B) and (H1)-(H4) are satisfied. Then equation (1.1) has at least a nontrivial ground state solution, i.e., solution corresponding to the least energy of the action functional of (1.1).

Remark1.2  In [6], the authors also considered (1.1) and assumed that (H2) and the following classical Ambrosetti-Rabinowitz superlinear condition (see [16, 17]): There exists a constant $\mu>2$ such that

$ \begin{eqnarray}\label {eqnAR} 0<\mu F(k, u)\leq f(k, u)u, u\neq 0, \end{eqnarray} $ (1.3)

(1.3) implies that for each $k\in\mathbb{Z}$, there exists a constant $C>0$ such that

$ \begin{eqnarray}\label {eqnAR1} F(k, u)\geq C|u|^{\mu}~ {\rm~for~} |u|\geq 1. \end{eqnarray} $ (1.4)

This implies (H3) holds. There exists a superlinear function, such as

$ f(k, u)=u\ln(1+|u|) $

does not satisfy (1.3). However, it satisfies the conditions (H1)-(H4). So our conditions are weaker than conditions presented in [6]. And in our paper, we do not need periodic approximation technique to obtain homoclinic solutions. Furthermore, the existence of ground state solutions can be obtained.

Remark1.3  In [13], the authors considered the following difference equation

$ \begin{eqnarray}\label{eqnshi} \Delta(\varphi_{p}(\Delta u(n-1))-q(n)\varphi_{p}(u_{n})+f(n, u_{n+M}, u_{n}, u_{n-M})=0, \quad n\in\mathbb{Z}. \end{eqnarray} $ (1.5)

Let $p=0$ and $M=0$, (1.2) is the special case of (1.5).

The following hypotheses with $p=0$ and $M=0$ are satisfied in [13]:

(F1) $F\in C(\mathbb{Z}\times \mathbb{R}, \mathbb{R})$ with $F(n+T, u)=F(n, u)$ and it satisfies $\frac{\partial F}{\partial u}=f(n, u)$;

(F2) there exist positive constants $\varrho$ and $a<\frac{\underline{q}}{4}(\frac{\kappa_{1}}{\kappa_{2}})^{2}$ such that $|F(n, u)|\leq a|u|^{2}$, for all $n\in \mathbb{Z}$ and $|u|\leq\varrho$;

(F3) there exist constants $\rho, $ $c>\frac{1}{4}(\frac{\kappa_{1}}{\kappa_{2}})^{2}(4+\overline{q})$and $b$ such that $F(n, u)\geq c|u|^{2}+b$, for all $n\in \mathbb{Z}$ and $|u|\geq \rho$;

(F4) $fu-2F>0, $ for all $n\in \mathbb{Z}$ and $|u|\neq 0$;

(F5) $fu-2F\rightarrow\infty, $ as $|u|\rightarrow\infty$.

Note that (H2)-(H4) imply that (F2)-(F4). A nontrivial homoclinic orbit of (1.5) is obtained by Mountain Pass Lemma in combination with periodic approximations in [13]. However, in our paper, we employ the Nehari manifold approach instead of periodic approximation technique to obtain the ground state solutions. Furthermore, we show that the functional is coercive on Nehari manifold (Lemma 3.2), which is weaker than P.S. condition in [13].

The rest of this paper is organized as follows. In Section (1.1), we establish the variational framework associated with (1.1), and transfer the existence of solutions of boundary value problem (1.1) into the existence of critical points of the corresponding functional. By employing the critical point theory, we give proofs of the main results in Section 3. Finally, we give a simple example to demonstrate our results.

2 Variational Framework

In this section, we firstly establish the corresponding variational setting associated with (1.1). Let

$ X=\{u=\{u(k)\}:u(k)\in\mathbb{R}, k\in\mathbb{Z}\} $

be the set of all real sequences

$ u=\{u(k)\}_{k\in\mathbb{Z}}=\left(\cdots, u(-k), u(-k+1), \cdots, u(-1), u(0), u(1), \cdots, u(k), \cdots\right). $

Then $X$ is a vector space with $au+bv=\{au(k)+bu(k)\}$ for $u, v\in X$, $a, b\in\mathbb{R}$.

Define the space

$ \begin{eqnarray*} E:=\{~u\in X: \sum_{k\in\mathbb{Z}}\left[a(k)(\Delta u(k-1))^{2}+b(k)(u(k))^{2}\right]<+\infty\}. \end{eqnarray*} $

Then $E$ is a Banach space equipped with the corresponding norm

$ \begin{eqnarray*} \|u\|^{2}=\sum_{k\in\mathbb{Z}}\left[a(k)(\Delta u(k-1))^{2}+b(k) (u(k))^{2}\right], \quad \forall u\in E. \end{eqnarray*} $

For $1\leq p<+\infty$, denote

$ l^{p}=\{u=\{u(k)\}\in X:\sum\limits_{k\in\mathbb{Z}}|u(k)|^{p}<\infty\} $

equipped with the norm

$ \|u{{\|}_{{{l}^{p}}}}={{\left( \sum\limits_{k\in \mathbb{Z}}{|}u(k){{|}^{p}} \right)}^{\frac{1}{p}}}, $

$|\cdot|$ is the usual absolute value in $\mathbb{R}$. Then the following embedding between $l^{p}$ spaces holds

$ \begin{eqnarray}\label{eqn2.1} l^{q}\subset l^{p}, \|u\|_{l^{p}}\leq\|u\|_{l^{q}}, 1\leq q\leq p\leq \infty. \end{eqnarray} $ (2.1)

Now we consider the variational functional $J$ defined on $E$ by

$ \begin{eqnarray*} J(u)&=&\sum_{k\in\mathbb{Z}}\left[\frac{1}{2}a(k)(\Delta u(k-1))^{2}+\frac{1}{2}b(k) (u(k))^{2}-F(k, u(k))\right]\\ &=&\frac{1}{2}\|u\|^{2}-\sum_{k\in\mathbb{Z}}F(k, u(k)). \end{eqnarray*} $

Then $J\in C^{1}(E, \mathbb{R})$, and for all $u, v\in E$, we have

$ \begin{eqnarray}\label{eqn2.4} (J'(u), v)=\sum_{k\in\mathbb{Z}}\left[a(k)\Delta u(k-1)\Delta v(k-1)+b(k)u(k)v(k)-f(k, u(k))v(k)\right]. \end{eqnarray} $ (2.2)

Then, we easily get the variational formulation for (1.1):

Lemma2.1  Every critical point $u\in E$ of $J$ is a solution of (1.1).

Proof  We assume that $u\in E$ is a critical point of $J$, then $J'(u)=0$. According to (2.2), this is equivalent to

$ \begin{eqnarray}\label{eqn2.3} \sum_{k\in\mathbb{Z}}\left[a(k)\Delta u(k-1)\Delta v(k-1)+b(k)u(k)v(k)-f(k, u(k))v(k)\right]=0, ~~~\forall v\in E. \end{eqnarray} $ (2.3)

For any $h\in \mathbb{Z}$, we define $e_h\in E$ by putting $e_h(k)=\delta_{hk}$ for all $k\in \mathbb{Z}$, where $\delta_{hk}=1$ if $h=k$; $\delta_{hk}=0$ if $h\neq k$. If we apply (2.3) with $v=e_h$, then

$ -\Delta[a(k)\Delta u(k-1)]+b(k)u(k)-f(k, u(k))=0, $

i.e., $u$ is a solution of (1.1). The proof is completed.

Therefore, $u\neq 0$ is a critical point of $J$ on $E$ if and only if $u$ is a solution of (1.1) with $u(\pm\infty)=0$, that is to say, $u$ is a homoclinic solution emanating from $0$. Thus, we have reduced the problem of finding homoclinic solutions of (1.1) to that of seeking critical points of the functional $J$ on $E$.

3 Proof of Main Results

Now, we consider the Nehari manifold $ \mathcal{N}=\{u\in E\setminus\{0\}:J'(u)u=0\}, $ and let $c=_{u\in \mathcal{N}}^{\ \ \inf }J(u).$ By the definition of $\mathcal{N}$, we know $\mathcal{N}$ contains all nontrivial critical points of $J$.

Lemma3.1  Assume that (A), (B) and (H2)-(H4) are satisfied, then $\mathcal{N}$ is homeomorphic to the unit sphere $S$ in $E$, where $S=\{u\in E: \|u\|=1\}$.

Proof  Let $\varphi(u)=\sum\limits_{k\in\mathbb{Z}}F(k, u(k))$. By (H2), we have

$ \begin{eqnarray}\label{eqn3.01} \varphi'(u)=o(\|u\|)~{\rm~as~}~u\rightarrow 0. \end{eqnarray} $ (3.1)

Let $U\subset E\setminus\{0\}$ be a weakly compact subset, we know that

$ \begin{eqnarray}\label{eqn3.03} \varphi(su)/s^{2}\rightarrow\infty {\rm~uniformly ~for~} u {\rm~on~} U, {\rm~as~} s\rightarrow\infty. \end{eqnarray} $ (3.2)

In fact, let $\{u_n\}\subset U$. It needs to show that

$ \begin{eqnarray*} {\rm if~} s_n\rightarrow\infty, \quad \varphi(s_nu_n)/(s_n)^{2}\rightarrow\infty \end{eqnarray*} $

as $n\rightarrow\infty$. Passing to a subsequence if necessary, $u_n\rightharpoonup u\in E\setminus\{0\}$ and $u_{n}(k)\rightarrow u(k)$ for every $k$, as $n\rightarrow\infty$.

Note that from (H2) and (H4), it is easy to get that

$ \begin{eqnarray}\label{eqn3.1} F(k, u)>0, {\rm ~~for ~all~} u\neq 0. \end{eqnarray} $ (3.3)

Since $|s_n u_{n}(k)|\rightarrow\infty$ and $u_n\neq 0$, by (H3) and (3.3), we have

$ \frac{\varphi ({{s}_{n}}{{u}_{n}})}{{{({{s}_{n}})}^{2}}}=\sum\limits_{k\in \mathbb{Z}}{\frac{F(k,{{s}_{n}}{{u}_{n}}(k))}{{{({{s}_{n}}{{u}_{n}}(k))}^{2}}}}{{({{u}_{n}}(k))}^{2}}\to \infty ~~\rm{as}~\mathit{n}\to \infty . $

Thus, we obtain (3.2) holds.

From (H4), for all $u\neq 0$ and $s>0$, we have

$ \begin{eqnarray}\label{eqn3.02} s\mapsto \varphi'(su)u/s {\rm~~is ~strictly~increasing}. \end{eqnarray} $ (3.4)

Let $h(s):=J(sw), s>0$. Then

$ h'(s)=J'(sw)w=s(\|w\|^{2}-s^{-1}\varphi(sw)w) $

from (3.1)-(3.4), then there exists a unique $s_{w}$, such that, when $0<s<s_{w}$, $h'(s)>0 $; and when $s>s_{w}$, $h'(s)<0 $. Therefore, $h'(s_{w})=J'(s_{w}w)w=0$, and $s_{w}w\in \mathcal{N}$.

Therefore, $s_{w}$ is a unique maximum of $h(s)$, and we can define the mapping $\hat{m}:E\setminus\{0\}\rightarrow \mathcal{N}$ by setting $\hat{m}(w):=s_{w}w$. Then the mapping $\hat{m}$ is continuous. Indeed, suppose $w_{n}\rightarrow w\neq 0$. Since $\hat{m}(tu)=\hat{m}(u)$ for each $t>0$, we may assume $w_{n}\in S$ for all $n$. Write $\hat{m}(w_{n})=s_{w_n}w_{n}$. Then $\{s_{w_n}\}$ is bounded. If not, $s_{w_n}\rightarrow\infty$ as $n\rightarrow\infty$.

Note that by (H4), for all $u\neq 0$,

$ \begin{eqnarray*} \frac{1}{2}f(k, u)u-F(k, u)&=&\frac{1}{2} f(k, u)u-\int_{0}^{u} f(k, s)ds\\ &>&\frac{1}{2} f(k, u)u-\frac{f(k, u)}{u}\int_0^u sds=0. \end{eqnarray*} $

So, for all $u\in \mathcal{N}$, we have

$ \begin{eqnarray}\label{eqn4.1} J(u)=J(u)-\frac{1}{2}J'(u)u=\sum_{k\in \mathbb{Z}}\left(\frac{1}{2}f(k, u(k))u(k)-F(k, u(k))\right)>0. \end{eqnarray} $ (3.5)

By (H3), we have

$ 0<\frac{J({{s}_{wn}}w)}{{{({{s}_{wn}})}^{2}}}=\frac{1}{2}\|w{{\|}^{2}}-\sum\limits_{k\in \mathbb{Z}}{\frac{F(k,{{s}_{wn}}w(k))}{{{({{s}_{wn}}w(k))}^{2}}}}{{(w(k))}^{2}}\to -\infty ,~~~\rm{as}~\ \mathit{n}\to \infty , $

which is a contradiction. Therefore, $s_{w_n}\rightarrow s>0$ after passing to a subsequence if needed. Since $\mathcal{N}$ is closed and $\hat{m}(w_{n})=s_{w_n}w_{n}\rightarrow sw, sw\in \mathcal{N}$. Hence $sw=s_{w}w=\hat{m}(w)$ by the uniqueness of $s_{w}$.

Therefore, we define a mapping $m:S\rightarrow \mathcal{N}$ by setting $m:=\hat{m}|_{S}$, then $m$ is a homeomorphism between $S$ and $\mathcal{N}$.

We also consider the functional $\hat{\Psi}:E\setminus \{0\}\rightarrow \mathbb{R}$ and $\Psi:S\rightarrow \mathbb{R}$ by

$ \begin{eqnarray*} \hat{\Psi}(w):=J(\hat{m}(w))~~{\rm~and}~~\Psi(w):=\hat{\Psi}|_{S}. \end{eqnarray*} $

Then we have $\hat{\Psi}\in C^{1}(E\setminus\{0\}, \mathbb{R})$ and

$ \begin{eqnarray*} \hat{\Psi}'(w)z=\frac{\|\hat{m}(w)\|}{\|w\|}J'(\hat{m}(w))z ~~{\rm~for~all}~w, z\in E, w\neq 0. \end{eqnarray*} $

In fact, let $w\in E\setminus\{0\}$ and $z\in E$. By Lemma 3.1 and the mean value theorem, we obtain

$ \begin{eqnarray*} \hat{\Psi}(w+tz)-\hat{\Psi}(w)&=&J(s_{w+tz}(w+tz))-J(s_{w}w)\\ &\leq&J(s_{w+tz}(w+tz))-J(s_{w+tz}(w))\\ &=& J'(s_{w+tz}(w+\tau_{t}tz))s_{w+tz}tz, \end{eqnarray*} $

where $|t|$ is small enough and $\tau_{t}\in(0, 1)$. Similarly,

$ \begin{eqnarray*} \hat{\Psi}(w+tz)-\hat{\Psi}(w)&=&J(s_{w+tz}(w+tz))-J(s_{w}w)\\ &\geq&J(s_{w}(w+tz))-J(s_{w}(w))\\ &=& J'(s_{w}(w+\eta_{t}tz))s_{w}tz, \end{eqnarray*} $

where $\eta_{t}\in(0, 1)$. Combining these two inequalities and the continuity of function $w\mapsto s_{w}$, we have

$ \begin{eqnarray*} \lim_{t\rightarrow 0}\frac{\hat{\Psi}(w+tz)-\hat{\Psi}(w)}{t}=s_{w}J'(s_{w}w)z=\frac{\|\hat{m}(w)\|}{\|w\|}J'(\hat{m}(w))z. \end{eqnarray*} $

Hence the G$\hat{a}$teaux derivative of $\hat{\Psi}$ is bounded linear in $z$ and continuous in $w$. It follows that $\hat{\Psi}$ is a class of $C^{1}$ (see [15], Proposition1.3).

Since $w\in S$, $m(w)=\hat{m}(w)$, so we have $\Psi\in C^{1}(S, \mathbb{R})$ and

$ \ {\Psi }'(w)z=\|m(w)\|{J}'(m(w))z~~~\rm{for}~\rm{all}~z\in {{T}_{w}}(S)=\{v\in E:(w,v)=0\}. $

Lemma3.2  Assume that (A), (B) and (H1)-(H3) are satisfied, for $u\in \mathcal{N}$ then $J(u)\rightarrow \infty$ as $\|u\|\rightarrow \infty$.

Proof  By way of contradiction, we assume that there exists a sequence $\{u_n\}\subset\mathcal{N}$ such that $J(u_n)\leq d$, as $\|{u_n}\|\rightarrow \infty$. Set $v_n=\frac{u_n}{\|u_n\|}$, and then there exists a subsequence, still denoted by $v_n$, such that $v_n\rightharpoonup v$, and therefore $v_{n}(k)\rightarrow v(k)$ for every $k$, as $n\rightarrow \infty$.

First we can prove that there exist $\delta>0$ and $k_j\in \mathbb{Z}$ such that

$ \begin{eqnarray}\label{eqnv} |v_n(k_j)|\geq \delta. \end{eqnarray} $ (3.6)

In fact, if not, then $v_n\rightarrow 0$ in $l^{\infty}$ as $n\rightarrow \infty$. For $q>2$,

$ \begin{eqnarray*} \|v_n\|^{q}_{l^{q}}\leq \|v_n\|^{q-2}_{l^{\infty}}\|v_n\|^{2}_{l^{2}} \end{eqnarray*} $

so we have $v_n\rightarrow 0$ in all $l^{q}, ~q>2$.

Note that from (H1) and (H2), we have for any $\varepsilon>0$, there exists $c_\varepsilon>0$ such that

$ \begin{eqnarray}\label{eqnf} |f(k, u)|\leq \varepsilon|u|+c_{\varepsilon}|u|^{p-1} ~~{\rm~and~}~~ |F(k, u)|\leq \varepsilon|u|^{2}+c_{\varepsilon}|u|^{p}. \end{eqnarray} $ (3.7)

Then for each $s>0$, we have

$ \sum\limits_{k\in \mathbb{Z}}{F}(k, s{{v}_{n}}(k))\le \varepsilon {{s}^{2}}\|{{v}_{n}}\|_{{{l}^{2}}}^{2}+{{c}_{\varepsilon }}{{s}^{p}}\|{{v}_{n}}\|_{{{l}^{p}}}^{p}, $

which implies that $\sum\limits_{k\in\mathbb{Z}}F(k, sv_{n}(k))\rightarrow 0 {\rm~as~} n\rightarrow\infty$.So

$ \begin{eqnarray}\label{eqnd} d\geq J(u_n)\geq J(sv_n)=\frac{s^{2}}{2}\|v^{(k)}\|^{2}-\sum_{k\in\mathbb{Z}}F(k, sv_{n}(k))\rightarrow \frac{s^{2}}{2}, \end{eqnarray} $ (3.8)

as $n\rightarrow \infty$. This is a contradiction if $s>\sqrt{2d}$.

By periodicity of coefficients, we know $J$ and $\mathcal{N}$ are both invariant under $ \mathit{T}$-translation. Making such shifts, we can assume that $1\leq k_j\leq T-1$ in (3.6). Moreover, passing to a subsequence, we can assume that $k_{j}=k_{0}$ is independent of $j$.

Next we can extract a subsequence, still denoted by $\{v_n\}$, such that $v_{n}(k)\rightarrow v(k)$ for all $k\in\mathbb{Z}$. Specially, for $k=k_{0}$, inequality (3.6) shows that $|v(k_0)|\geq \delta$, so $v\neq 0$. Since $|u_{n}(k)|\rightarrow \infty$ as $n\rightarrow \infty$, it follows again from (H3) that

$ 0\le \frac{J({{u}_{n}})}{\|{{u}_{n}}{{\|}^{2}}}=\frac{1}{2}-\sum\limits_{k\in \mathbb{Z}}{\frac{F(k, {{u}_{n}}(k))}{{{({{u}_{n}}(k))}^{2}}}}{{({{v}_{n}}(k))}^{2}}\to -\infty ~~~~~~~as~n\to \infty, $

a contradiction again.

From above, we have the following lemma, which is important in this paper.

Lemma3.3   $\{w_{n}\}$ is a Palais-Smale sequence for $\Psi$ if and only if $\{m(w_{n})\}$ is a Palais-Smale sequence for $J$.

Proof  Let $\{w_{n}\}$ be a Palais-Smale sequence for $\Psi$, and let $u_{n}=m(w_{n})\in\mathcal{ N}$. Since for every $w_{n}\in S$ we have an orthogonal splitting $E=T_{w_{n}}S\oplus\mathbb{R}w_{n}$, we have

$ \begin{eqnarray*} \|\Psi'(w_{n})\|=\sup_{\substack {z\in T_{w_{n}}S \\ \|z\|=1}}\Psi'(w_{n})z=\|m(w_{n})\|\sup_{\substack {z\in T_{w_{n}}S \\ \|z\|=1}}J'(m(w_{n}))z=\|u_{n}\|\sup_{\substack {z\in T_{w_{n}}S \\ \|z\|=1}}J'(u_{n})z. \end{eqnarray*} $

Then

$ \begin{eqnarray*} \|\Psi'(w_{n})\|&\leq& \|u_{n}\|\|J'(u_{n})\|=\|u_{n}\|\sup_{\substack {z\in T_{w_{n}}S, ~t\in\mathbb{R} \\ z+tw\neq 0}}\frac{J'(u_{n})(z+tw)}{\|z+tw\|}\\ &\leq&\|u_{n}\|\sup_{z\in T_{w_{n}}S\backslash\{0\}}\frac{J'(u_{n})(z)}{\|z\|}=\|\Psi'(w_{n})\|. \\ \end{eqnarray*} $

Therefore

$ \begin{eqnarray}\label{eqn3.7} \|\Psi'(w_{n})\|=\|u_{n}\|\|J'(u_{n})\|. \end{eqnarray} $ (3.9)

By (3.5), for $u_{n}\in\mathcal{ N}$, $J(u_{n})>0$, so there exists a constant $c_0>0$ such that $J(u_{n})>c_0$. And since $c_0\leq J(u_{n})=\frac{1}{2}~\|u_{n}\|^{2}-I(u_{n})\leq \frac{1}{2}~\|u_{n}\|^{2}$, $\|u_{n}\|\geq \sqrt{2c_0}$. Together with Lemma3.2, $\sqrt{2c_0}\leq \|u_{n}\|\leq\sup_{n}\|u_{n}\|<\infty$. Hence $\{w_{n}\}$ is a Palais-Smale sequence for $\Psi$ if and only if $\{u_{n}\}$ is a Palais-Smale sequence for $J$.

Now, we give the detailed proof of Theorem 1.1.

Proof  From(3.9), $\Psi'(w)=0$ if and only if $J'(m(w))=0$. So $w$ is a critical point of $\Psi$ if and only if $m(w)$ is a nontrivial critical point of $J$. Moreover, the corresponding values of $\Psi$ and $J$ coincide and $\inf_{S}\Psi=\inf_{\mathcal{N}}J$.

Let $u_{0}\in \mathcal{N}$ such that $J(u_{0})=c$, then $m^{-1}(u_{0})\in S$ is a minimizer of $\Psi$ and therefore a critical point of $\Psi$, so $u_{0}$ is a critical point of $J$. It needs to show that there exists a minimizer $u\in \mathcal{N}$ of $J|_\mathcal{N}$.

Let $\{w_n\}\subset S$ be a minimizing sequence for $\Psi$. By Ekeland's variational principle we may assume $\Psi(w_n)\rightarrow c, ~\Psi'(w_n)\rightarrow 0$ as $n\rightarrow \infty$, hence $J(u_n)\rightarrow c, J'(u_n)\rightarrow 0$ as $n\rightarrow \infty$, where $u_n:=m(w_n)\in \mathcal{N}$.

It follows from Lemma 3.2 that $\{u_n\}$ is bounded in $\mathcal{N}$, then there exists a subsequence, still denoted by the same notation, such that $u_n$ weakly converges to some $u \in E$. We know that there exist $\delta>0$ and $k_j\in \mathbb{Z}$ such that

$ \begin{eqnarray}\label{eqnu} |u_n(k_j)|\geq \delta. \end{eqnarray} $ (3.10)

If not, then $u_n\rightarrow 0$ in $l^{\infty}$ as $n\rightarrow \infty$. Note that, for $q>2$,

$ \|{{u}_{n}}\|_{{{l}^{q}}}^{q}\le \|{{u}_{n}}\|_{{{l}^{\infty }}}^{q-2}\|{{u}_{n}}\|_{{{l}^{2}}}^{2}, $

then $u_n\rightarrow 0$ in all $l^{q}, ~q>2$. By (3.7), we have

$ \begin{eqnarray*} \sum_{k\in\mathbb{Z}}f(k, u_{n}(k))u_{n}(k) &\leq& \varepsilon\sum_{k\in \mathbb{Z}}|u_{n}(k)|\cdot|u_{n}(k)|+c_{\varepsilon}\sum_{k\in\mathbb{Z}} |u_{n}(k)|^{p-1}\cdot|u_{n}(k)|\\ &\leq& \varepsilon \|u_n\|_{l^{2}}\cdot\|u_n\|_{l^{2}}+c_{\varepsilon}\|u_n\|^{p-1}_{l^{p}}\cdot\|u_n\|_{l^{p}}\\ &\leq& \varepsilon \|u_n\|_{l^{2}}\cdot\|u_n\|+c_{\varepsilon}\|u_n\|^{p-1}_{l^{p}}\cdot\|u_n\|, \\ \end{eqnarray*} $

which implies that $\sum\limits_{k\in\mathbb{Z}}f(k, u_{n}(k))u_{n}(k)=o(\|u_n\|){\rm~as~} n\rightarrow\infty$. Therefore

$ \begin{eqnarray*} o(\|u_n\|)=\left(J'(u_n), u_n\right) = \|u_n\|^{2}-\sum_{k\in \mathbb{Z}}f(k, u_{n}(k))u_{n}(k)=\|u_n\|^{2}-o(\|u_n\|). \end{eqnarray*} $

So $\|u_n\|^{2}\rightarrow 0$, as $n\rightarrow\infty$, which contradicts with $u_n\in\mathcal{N}$.

Since $J$ and $J'$ are both invariant under $T$-translation. Making such shifts, we assume that $1\leq k_j\leq T-1$ in (3.10). Moreover passing to a subsequence, we assume that $k_j=k_{0}$ is independent of $j$. Extracting a subsequence, still denoted by $\{u_n\}$, we have $u_n\rightharpoonup u$ and $u_{n}(k)\rightarrow u(k)$ for all $k\in\mathbb{Z}$. Specially, for $k=k_{0}$, inequality (3.10) shows that $|u(k_0)|\geq \delta$, so $u\neq 0$. Hence $u\in \mathcal{N}$.

Now, we prove that $J(u)=c$. By Fatou's lemma,

$ \begin{eqnarray*} c&=&\lim_{n\rightarrow\infty}\left(J(u_n)-\frac{1}{2}J'(u_n)u_n\right)=\lim_{n\rightarrow\infty}\sum_{k\in \mathbb{Z}}\left(\frac{1}{2}f(k, u_{n}(k))u_{n}(k)-F(k, u_{n}(k))\right)\\ &\geq& \sum_{k\in \mathbb{Z}} \left(\frac{1}{2}f(k, u(k))u(k)-F(k, u(k))\right)=J(u)-\frac{1}{2}J'(u)u=J(u)\geq c. \end{eqnarray*} $

So $J(u)=c$. The proof of Theorem 1.1 is completed.

Example1  Consider the difference equation

$ \begin{eqnarray} &&\Delta[|\sin k|\Delta u(k-1)]-|\cos k|u(k)\nonumber\label{eqnex} \\ &&+\left[c(\alpha+2)u(k)|u(k)|^{\alpha}+d(\beta+2)u(k)|u(k)|^{\beta}\right](\phi(k)+M)=0, \end{eqnarray} $ (3.11)

where $c>0, d>0, \alpha\geq\beta>0, M>0, \phi(k)$ is a bounded continuous $\pi-$periodic function and $|\phi(k)|<M$, $k\in\mathbb{Z}$. Let $a(k)=|\sin k|$, $b(k)=|\cos k|$,

$ f(k, u)=\left[c(\alpha+2)u|u|^{\alpha}+d(\beta+2)u|u|^{\beta}\right](\phi(k)+M) $

and

$ F(k, u)=\int_{0}^{u}f(k, s) ds=\left[c|u|^{\alpha+2}+d|u|^{\beta+2}\right](\phi(k)+M). $

It is easy to show that all the assumptions of Theorem 1.1 are satisfied. Therefore, equation (3.11) has at least one homoclinic solution.

References
[1] Zhang K, Xu J. Existence of solutions for a second order difference boundary value problem[J]. J. Math., 2014, 34: 856–862.
[2] Agarwal R P. Difference equations and inequalities: theory, methods and applications[M]. New York: Marcel Dekker, 1992.
[3] Kelly W G, Peterson A C. Difference equations: an introduction with applications[M]. New York: Academic Press, 1991.
[4] Lakshmikantham V, Trigiante D. Theory of difference equations: numerical methods and applica-tions[M]. New York: Academic Press, 1988.
[5] Ma Manjun, Guo Zhiming. Homoclinic orbits for second order self-adjoint difference equations[J]. J. Math. Anal. Appl., 2005, 323: 513–521.
[6] Ma Manjun, Guo Zhiming. Homoclinic orbits for nonliear second order difference equations[J]. Nonl. Anal., 2007, 67: 1737–1745. DOI:10.1016/j.na.2006.08.014
[7] Ma Defang, Zhou Zhan. Existence and multiplicity results of homoclinic solutions for the DNLS equations with unbounded potentials[J]. Abstr. Appl. Anal., 2012, 2012: 1–15.
[8] Mai Ali, Zhou Zhan. Ground state solutions for the periodic discrete nonlinear SchrÄodinger equations with superlinear nonlinearities[J]. Abstr. Appl. Anal., 2013, 2013: 1–11.
[9] Mai Ali, Zhou Zhan. Discrete solitons for periodic discrete nonlinear SchrÄodinger equations[J]. Appl. Math. Comput., 2013, 222: 34–41.
[10] Mai Ali, Zhou Zhan. Homoclinic solutions for a class of nonlinear difference equations[J]. J. Appl. Math., 2014, 2014: 1–8.
[11] Sun Guowei. On standing wave solutions for discrete nonlinear SchrÄodinger equations[J]. Abstr. Appl. Anal., 2013, 2013: 1–6.
[12] Shi Haiping, Zhang Hongqiang. Existence of gap solitons in periodic discrete nonlinear SchrÄodinger equations[J]. J. Math. Anal. Appl., 2010, 361(2): 411–419. DOI:10.1016/j.jmaa.2009.07.026
[13] Haiping Shi, Xia Liu, Yuanbiao Zhang. Homoclinic orbits for second order p-Laplacian difference equations containing both advance and retardation[J]. RACSAM, 2016, 110(1): 65–78. DOI:10.1007/s13398-015-0221-y
[14] Szulkin A, Weth T. The method of Nehari manifold[A]. Gao D Y, Motreanu D, eds. Handbook of nonconvex analysis and applications[C]. Boston: International Press, 2010.
[15] Willem M. Minimax theorems[M]. Boston: BirkhÄauser, 1996.
[16] Ambrosetti A, Rabinowitz P H. Dual variational methods in critical point theory and applications[J]. J. Funct. Anal., 1973, 14: 349–381. DOI:10.1016/0022-1236(73)90051-7
[17] Rabinowitz P H. Minimax methods in critical point theory with applications to differential equa-tions[C]. Providence RI: AMS, 1986.