\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 150, pp. 1--16.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/150\hfil Homoclinic orbits] {Homoclinic orbits of second-order nonlinear difference equations} \author[H. Shi, X. Liu, Y. Zhang \hfil EJDE-2015/150\hfilneg] {Haiping Shi, Xia Liu, Yuanbiao Zhang} \address{Haiping Shi \newline Modern Business and Management Department, Guangdong Construction Vocational Technology Institute, Guangzhou 510440, China} \email{shp7971@163.com} \address{Xia Liu \newline Oriental Science and Technology College, Hunan Agricultural University, Changsha 410128, China\newline Science College, Hunan Agricultural University, Changsha 410128, China} \email{xia991002@163.com} \address{Yuanbiao Zhang \newline Packaging Engineering Institute, Jinan University, Zhuhai 519070, China} \email{abiaoa@163.com} \thanks{Submitted January 23, 2015. Published June 10, 2015.} \subjclass[2010]{34C37, 37J45, 39A12} \keywords{Homoclinic orbits; second order; nonlinear difference equations; \hfill\break\indent discrete variational methods} \begin{abstract} We establish existence criteria for homoclinic orbits of second-order nonlinear difference equations by using the critical point theory in combination with periodic approximations. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{example}[theorem]{Example} \allowdisplaybreaks \section{Introduction} Homoclinic orbits play an important role in analyzing the chaos of dynamical systems, and have been the subject of many investigations. If a system has the transversely intersected homoclinic orbits, then it must be chaotic. If it has the smoothly connected homoclinic orbits, then it cannot stand the perturbation, its perturbed system probably produce chaotic phenomenon. So homoclinic orbits have been extensively investigated since the time of Poincar\'e, see \cite{GuAWO,GuOA,GuOXA1,GuOXA2,GuOXA3,GuOXA4,Po,Ra2} and the references therein. Difference equations \cite{Ag,CuF} are closely related to differential equations in the sense that a differential equation model is often derived from a difference equation, and numerical solutions of a differential equation are obtained by discretizing the differential equation. Therefore, the study of homoclinic orbits \cite{ChT1,ChT2,ChT3,ChT4,ChW,DeC,FaZ,Lo1,Lo2,MaG1,MaG2,ZhYC} of difference equation is meaningful. Here $\mathbb{N}$, $\mathbb{Z}$ and $\mathbb{R}$ denote the sets of all natural numbers, integers and real numbers respectively. For any $a,b\in \mathbb{Z}$, define $\mathbb{Z}(a)=\{a,a+1,\dots\}$, $\mathbb{Z}(a,b)=\{a,a+1,\dots,b\}$ when $a\leq b$. The symbol $l^2$ denotes the space of real functions whose second powers are summable on $\mathbb{Z}$. Also, * denotes the transpose of a vector. This article considers the existence for homoclinic orbits of second-order nonlinear difference equation \begin{equation}\label{e1.1} Lu(t)=f(t,u(t+T),u(t),u(t-T)),\quad t\in \mathbb{Z} \end{equation} containing both advance and retardation. Here the operator $L$ is the Jacobi operator $$ Lu(t)=a(t)u(t+1)+a(t-1)u(t-1)+b(t)u(t), $$ where $a(t)$ and $b(t)$ are real valued for each $t\in \mathbb{Z}$, $T$ is a given nonnegative integer, $f\in C(\mathbb{Z}\times\mathbb{R}^3,\mathbb{R})$, $a(t)$, $b(t)$ and $f(t,v_1,v_2,v_3)$ are $M$-periodic in $t$ for a given positive integer $M$. Jacobi operators appear in a variety of applications \cite{Te}. We may think of \eqref{e1.1} as being a discrete analogue of the second-order nonlinear differential equation \begin{equation}\label{e1.2} Su(s)= f(s,u(s+T),u(s),u(s-T)),\quad s\in \mathbb{R}, \end{equation} where $S$ is the Sturm-Liouville differential expression, $f\in C(\mathbb{R}^4,\mathbb{R})$. Equations similar in structure to \eqref{e1.2} arise in the study of homoclinic orbits \cite{GuOA,GuOXA2,GuOXA3,GuOXA4} of functional differential equations. For the case $T=1$, Chen and Fang \cite{ChF} obtained the existence of periodic and subharmonic solutions of the second-order $p$-Laplacian difference equation $$ \Delta(\varphi_p(\Delta u(t-1)))+f(t,u(t+1),u(t),u(t-1))=0,\quad t\in \mathbb{Z}, $$ and Chen and Tang \cite{ChT1} obtained the existence of infinitely many homoclinic orbits of the fourth-order difference equation $$ \Delta^4 u(t-2)+q(t)u(t)=f(t,u(t+1),u(t),u(t-1)),\quad t\in \mathbb{Z} $$ containing both advance and retardation. It is well known that critical point theory is a powerful tool that deals with the problems of differential equations \cite{CaM,GuAWO,GuOA,GuOXA1,GuOXA2,GuOXA3,GuOXA4,MaW,Ra1}. Only since 2003, critical point theory has been employed to establish sufficient conditions on the existence of periodic solutions for second-order difference equations \cite{GuY1,GuY2}. Along this direction, Ma and Guo \cite{MaG1} (without periodicity assumption) and \cite{MaG2} (with periodicity assumption) applied variational methods to prove the existence of homoclinic orbits for the special form of \eqref{e1.1} (with $T=0$). The Ambrosetti-Rabinowitz condition plays a crucial role to ensure the boundedness of Palais-Smale sequences. This is very crucial in applying the critical point theory. Some special cases of \eqref{e1.1} have been studied by many researchers via variational methods, see \cite{GuY1,GuY2,MaG1,MaG2}. However, to our best knowledge, the results on homoclinic orbits of \eqref{e1.1} are scarce in the literature. Since \eqref{e1.1} contains both advance and retardation, there are very few manuscripts dealing with this subject, the traditional ways of establishing the functional in \cite{DeC,GuY1,GuY2,Lo1,MaG1,MaG2,MoR} are inapplicable to our case. The main purpose of this article is to give some sufficient conditions for the existence of a nontrivial homoclinic orbit for \eqref{e1.1} without the classical Ambrosetti-Rabinowitz condition. In particular, our results generalize and improve the existing results; see Remarks \ref{rmk1.3} and \ref{rmk1.4}. The motivation for the present work stems from the recent papers \cite{ChF,ChT4,GuOXA4}. Let $$ \underline{\lambda}=\min_{t\in {\mathbb{Z}}(1,M)}(b(t)-|a(t-1)|-|a(t)|),\quad \bar{\lambda}=\max_{t\in {\mathbb{Z}}(1,M)}(b(t)+|a(t-1)|+|a(t)|). $$ In this article we use the following hypotheses: \begin{itemize} \item[(H1)] $b(t)-|a(t-1)|-|a(t)|>0$, for all $t\in \mathbb{Z}$; \item[(H2)] there exists a functional $F(t,v_1,v_2)\in C^1(\mathbb{Z}\times \mathbb{R}^2,\mathbb{R})$ with $F(t+M,v_1,v_2)=F(t,v_1,v_2)$ and it satisfies $$ \frac{\partial F(t-T,v_2,v_3)}{\partial v_2} +\frac{\partial F(t,v_1,v_2)}{\partial v_2} =f(t,v_1,v_2,v_3); $$ \item[(H3)] there exist positive constants $\delta_1$ and $a_1<\underline{\lambda}/4$ such that \[ |F(t,v_1,v_2)|\leq a_1\left(v_1^2+v_2^2\right) \] for all $t\in \mathbb{Z}$ and $\sqrt{v_1^2+v_2^2}\leq\delta_1$; \item[(H4)] there exist constants $\rho_1, c_1>\bar{\lambda}/4$ and $b_1$ such that \[ F(t,v_1,v_2)\geq c_1\left(v_1^2+v_2^2\right)+b_1 \] for all $t\in \mathbb{Z}$ and $\sqrt{v_1^2+v_2^2}\geq\rho_1$; \item[(H5)] \[ \frac{\partial F(t,v_1,v_2)}{\partial v_1}v_1+ \frac{\partial F(t,v_1,v_2)}{\partial v_2}v_2-2F(t,v_1,v_2)>0, \] for all $(t,v_1,v_2)\in \mathbb{Z}\times \mathbb{R}^2\setminus\{(0,0)\}$; \item[(H6)] \[ \frac{\partial F(t,v_1,v_2)}{\partial v_1}v_1+ \frac{\partial F(t,v_1,v_2)}{\partial v_2}v_2-2F(t,v_1,v_2)\to+\infty \] as $\sqrt{v_1^2+v_2^2}\to+\infty$. \end{itemize} Our main results are the following theorem. \begin{theorem} \label{thm1.1} Suppose that {\rm (H1)--(H6)} are satisfied. Then \eqref{e1.1} has a nontrivial homoclinic orbit. \end{theorem} \begin{remark} \label{rmk1.2} \rm By (H4), it is easy to see that there exists a constant $\zeta_1>0$ such that \begin{itemize} \item[(H4')] $F(t,v_1,v_2)\geq c_1\left(v_1^2+v_2^2\right)+b_1-\zeta_1$, for all $(t,v_1,v_2)\in \mathbb{Z}\times \mathbb{R}^2$. \end{itemize} As a matter of fact, letting \[ \zeta_1=\max\big\{|F(n,v_1,v_2)-c_1\left(v_1^2+v_2^2\right)-b_1|: n\in \mathbb{Z}, \sqrt{v_1^2+v_2^2}\leq \rho_1\big\}, \] we can easily get the desired result. \end{remark} \begin{remark} \label{rmk1.3} \rm As a special case of Theorem \ref{thm1.1} with $T=0$ and $a(t)<0$, we obtain \cite[Theorem 1.1]{MaG2}. \end{remark} \begin{remark} \label{rmk1.4} \rm In many studies (see e.g. \cite{GuY1,GuY2,MaG1,MaG2}) of second-order difference equations, the following classical Ambrosetti-Rabinowitz condition is assumed. \begin{itemize} \item[(AR)] There exists a constant $\beta>2$ such that $0<\beta F(t,u)\leq uf(t,u)$ for all $t\in \mathbb{Z}$ and $u\in \mathbb{R}\setminus\{0\}$. \end{itemize} Note that (H4)--(H6) are much weaker than (AR). Thus our result improves that the existing results. \end{remark} For the next theorem, we use the hypotheses: \begin{itemize} \item[(H7)] there exist positive constants $\delta_2$ and $a_2>\frac{\bar{\lambda}}{4}$ such that \[ |F(t,v_1,v_2)|\geq a_2\left(v_1^2+v_2^2\right) \] for all $t\in \mathbb{Z}$ and $\sqrt{v_1^2+v_2^2}\leq\delta_2$; \item[(H8)] there exists a constant $1<\mu<2$ such that \[ 0<\frac{\partial F(t,v_1,v_2)}{\partial v_1}v_1+ \frac{\partial F(t,v_1,v_2)}{\partial v_2}v_2\leq \mu F(t,v_1,v_2), \] for all $(t,v_1,v_2)\in \mathbb{Z}\times \mathbb{R}^2\setminus\{(0,0)\}$. \end{itemize} \begin{theorem} \label{thm1.5} Suppose that {\rm (H1), (H2), (H7), (H8)} are satisfied. Then \eqref{e1.1} has a nontrivial homoclinic orbit. \end{theorem} \begin{remark} \label{rmk1.6} \rm By (H8), there exist constants $a_3>0$ and $b_2$ such that \[ F(t,v_1,v_2)\leq a_3\left(v_1^2+v_2^2\right)^{\mu/2}+b_2 \quad \text{for all }t\in \mathbb{Z}, \] which implies that there exist constants $\rho_2>0$ and $c_2<\frac{\underline{\lambda}}{4}$ such that \begin{itemize} \item[(H9)] $F(t,v_1,v_2)\leq c_2\left(v_1^2+v_2^2\right)+b_2$ for all $t\in \mathbb{Z}$ and $\sqrt{v_1^2+v_2^2}\geq\rho_2$. \end{itemize} \end{remark} By (H9), it is easy to see that there exists a constant $\zeta_2>0$ such that \begin{itemize} \item[(H9')] $F(t,v_1,v_2)\leq c_2\left(v_1^2+v_2^2\right)+b_2+\zeta_2$, for all $(t,v_1,v_2)\in \mathbb{Z}\times \mathbb{R}^2$. \end{itemize} The remainder of this paper is organized as follows. In Section 2, we shall establish the variational framework associated with \eqref{e1.1} and transfer the problem of the existence of homoclinic orbits of \eqref{e1.1} into that of the existence of critical points of the corresponding functional. Some related fundamental results will also be recalled. In Section 3, we shall complete the proof of the results by using the critical point method. Finally, in Section 4, we shall give two examples to illustrate the results. \section{Preliminaries} To apply the critical point theory, we shall establish the corresponding variational framework for \eqref{e1.1} and give some lemmas which will be of fundamental importance in proving our results. We start by giving the basic notation. Let $S$ be the set of sequences \[ u=\{u(t)\}_{t\in \mathbb{Z}} =(\dots,u(-t), \dots,u(-1),u(0),u(1),\dots,u(t),\dots); \] that is, $$ S=\{\{u(t)\}: u(t)\in \mathbb{R},\; t\in \mathbb{Z}\}. $$ For any $u,v\in S$, $a,b\in \mathbb{R}$, $au+bv$ is defined by $$ au+bv=\{au(t)+bv(t)\}_{t=-\infty}^{+\infty}. $$ Then $S$ is a vector space. For any given positive integers $M$ and $m$, we define $$ E_{m}=\{u\in S|u(t+2mM)=u(t),\; \forall t\in \mathbb{Z}\}. $$ Clearly, $E_m$ is isomorphic to $\mathbb{R}^{2mM}$. $E_m$ can be equipped with the inner product \begin{equation}\label{e2.1} \left(u,v\right)=\sum^{mM-1}_{t=-mM}u(t)\cdot v(t),\quad \forall u,v\in E_m, \end{equation} by which the norm $\|\cdot\|$ can be induced by \begin{equation}\label{e2.2} \|u\|=\Big(\sum^{mM-1}_{t=-mM}u^2(t)\Big)^{1/2},\quad \forall u\in E_m. \end{equation} It is obvious that $E_m$ with the inner product \eqref{e2.1} is a finite dimensional Hilbert space and linearly homeomorphic to $\mathbb{R}^{2mM}$. In what follows, we define a norm in $E_m$ by $$ \|u\|_{\infty}=\max_{t\in { \mathbb{Z}}(-mM,mM-1)}|u(t)|,\quad \forall u\in E_m. $$ For $u\in E_m$, we define the functional $J_m$ by \begin{equation}\label{e2.3} J_m(u)=\frac{1}{2}\sum_{t=-mM}^{mM-1}Lu(t)\cdot u(t) -\sum_{t=-mM}^{mM-1}F(t,u(t+T),u(t)). \end{equation} Clearly, $J_m\in C^1(E_m,\mathbb{R})$ and for any $u=\{u(t)\}_{t\in { \mathbb{Z}}}\in E_m$, by the periodicity of $\{u(t)\}_{t\in { \mathbb{Z}}}$, we can compute the partial derivative as \begin{equation}\label{e2.4} \frac{\partial J_m(u)}{\partial u(t)}= Lu(t)-f(t,u(t+T),u(t),u(t-T)),\ \forall t\in \mathbb{Z}(-mM,mM-1). \end{equation} Thus, $u$ is a critical point of $J_m$ on $E_m$ if and only if $$ Lu(t)=f(t,u(t+T),u(t),u(t-T)),\ \forall t\in \mathbb{Z}(-mM,mM-1). $$ Due to the periodicity of $u=\{ u(t)\}_{t\in { \mathbb{Z}}}\in E_m$ and $f(t,v_1,v_2,v_3)$ in the first variable $t$, we reduce the existence of periodic solutions of \eqref{e1.1} to the existence of critical points of $J_m$ on $E_m$. That is, the functional $J_m$ is just the variational framework of \eqref{e1.1}. Let $E$ be a real Banach space, $J\in C^1(E,\mathbb{R})$, i.e., $J$ is a continuously Fr\'echet-differentiable functional defined on $E$. $J$ is said to satisfy the Palais-Smale condition (PS condition for short) if any sequence $\left\{u(t)\right\}\subset E$ for which $\left\{J\left(u(t)\right)\right\}$ is bounded and $J^\prime \left(u(t)\right)\to 0$ $(t\to \infty)$ possesses a convergent subsequence in $E$. Let $B_\rho$ denote the open ball in $E$ about 0 of radius $\rho$ and let $\partial B_\rho$ denote its boundary. \begin{lemma}[Mountain Pass Lemma \cite{Ra1}] \label{lem2.1} Let $E$ be a real Banach space and $J\in C^1(E,\mathbb{R})$ satisfy the PS condition. If $J(0)=0$ and \begin{itemize} \item[(J1)] there exist constants $\rho, \alpha>0$ such that $J|_{\partial B_\rho}\geq \alpha$, and \item[(J2)] there exists $e\in E\setminus B_\rho$ such that $J(e)\leq 0$. \end{itemize} Then $J$ possesses a critical value $c\geq \alpha$ given by \begin{equation}\label{e2.5} c=\inf_{g\in \Gamma}\max_{s\in[0,1]} J(g(s)), \end{equation} where \begin{equation}\label{e2.6} \Gamma =\{g\in C([0,1],E)|g(0)=0,\ g(1)=e\}. \end{equation} \end{lemma} \begin{lemma} \label{lem2.2} Assume that {\rm (H1)} holds. Then there exist constants $\underline{\lambda}$ and $\bar{\lambda}$ independent of $m$, such that \begin{equation}\label{e2.7} \underline{\lambda}\|u\|^2\leq\sum_{t=-mM}^{mM-1}Lu(t)\cdot u(t) \leq\bar{\lambda}\|u\|^2. \end{equation} \end{lemma} \begin{proof} Let $$ \sum_{t=-mM}^{mM-1}Lu(t)\cdot u(t)=(P_mu,u), $$ where $u=(u(-mM),\dots,u(-1),u(0),u(1),\dots,u(mM-1))^\ast$ and \[ P_m= \left(\begin{smallmatrix} b(-mM)& a(-mM)& 0& \dots & 0& a(-mM-1) \\ a(-mM)& b(-mM+1)& a(-mM+1)& \dots & 0& 0 \\ \dots &\dots &\dots &\dots &\dots &\dots \\ 0& 0& 0& \dots & b(mM-2)& a(mM-2) \\ a(mM-1)& 0& 0& \dots & a(mM-2)& b(mM-1) \end{smallmatrix} \right) \] which is a $2mM\times 2mM$ matrix. By (H1), $P_m$ is positive definite. Let $\lambda_{-mM}, \lambda_{-mM+1}, \dots, \lambda_{-1}, \lambda_0, \lambda_1, \dots, \lambda_{mM-2},\lambda_{mM-1} $ be the eigenvalues of $P_m$. Applying matrix theory, we see that $\underline{\lambda}\leq \lambda_i \leq \overline{\lambda}$, $i\in \mathbb{Z}(-mM,mM-1)$. From the definition of the norm $\|\cdot\|$, \eqref{e2.7} is obviously true. \end{proof} \section{Proof of main results} In this section, we shall prove the results stated in Section 1 by using the critical point theory. \subsection{Proof of Theorem \ref{thm1.1}} \begin{lemma} \label{lem3.1} Suppose that {\rm (H1), (H2)--(H6)} are satisfied. Then $J_m$ satisfies the {\rm PS} condition. \end{lemma} \begin{proof} Assume that $\{u_j\}_{j\in\mathbb{N}}$ in $E_m$ is a sequence such that $\{J_m(u_j)\}_{j\in\mathbb{N}}$ is bounded. Then there exists a constant $K_1>0$ such that $-K_1\leq J_m(u_j)$. By \eqref{e2.7} and (H4'), it is easy to see that \begin{align*} -K_1\leq J_m(u_j) &\leq\frac{\bar{\lambda}}{2}\|u_j\|^2 -\sum_{t=-mM}^{mM-1}\{c_1[u_j^2(t+T)+u_j^2(t)]+b_1-\zeta_1\}\\ &=\frac{\bar{\lambda}}{2}\|u_j\|^2-2c_1\|u_j\|^2 +2mM\left(\zeta_1-b_1\right),\ \forall j\in \mathbb{N}. \end{align*} Therefore, \begin{equation}\label{e3.1} \big(2c_1-\frac{\bar{\lambda}}{2}\big)\|u_j\|^2 \leq 2mM(\zeta_1-b_1)+K_1. \end{equation} Since $c_1>\bar{\lambda}/4$, \eqref{e3.1} implies that $\{u_j\}_{j\in\mathbb{N}}$ is bounded in $E_m$. Thus, $\{u_j\}_{j\in\mathbb{N}}$ possesses a convergence subsequence in $E_m$. The desired result follows. \end{proof} \begin{lemma} \label{lem3.2} Suppose that {\rm (H1)--(H6)} are satisfied. Then for any given positive integer $m$, \eqref{e1.1} possesses a $2mM$-periodic solution $u_m\in E_m$. \end{lemma} \begin{proof} In our case, it is clear that $J_m(0)=0$. By Lemma \ref{lem3.1}, $J_m$ satisfies the PS condition. By (H3), we have \begin{align*} J_m(u) &\geq \frac{\underline{\lambda}}{2}\|u\|^2-a_1\sum^{mM-1}_{t=-mM} [u^2(t)+u^2(t+T)]\\ &\geq \frac{\underline{\lambda}}{2}\|u\|^2-2a_1\|u\|^2 \\ &=\big(\frac{\underline{\lambda}}{2}-2a_1\big)\|u\|^2. \end{align*} Taking $\alpha_1=(\frac{\underline{\lambda}}{2}-2a_1)\delta_1^2>0$, we obtain $$ J_m(u)|_{\partial B_{\delta_1}}\geq\alpha_1>0, $$ which implies that $J_m$ satisfies the condition (J1) of the Mountain Pass Lemma. Next, we shall verify the condition (J2) of the Mountain Pass Lemma.. There exists a sufficiently large number $\rho>\max\{\rho_1,\delta_1\}$ such that \begin{equation}\label{e3.2} \big(2c_1-\frac{\bar{\lambda}}{2}\big)\rho^2\geq |b_1|. \end{equation} Let $e_m^{(1)}\in E_m$ and \begin{gather*} e_m^{(1)}(t)= \begin{cases} \rho, & \text{if }t=0,\\ 0, & \text{if } t\in\{j\in\mathbb{Z}:-mM\leq j\leq mM-1\text{ and } j\neq0\}, \end{cases} \\ e_m^{(1)}(t+T)= \begin{cases} \rho, & \text{if } t=0,\\ 0, & \text{if } t\in\{j\in\mathbb{Z}:-mM\leq j\leq mM-1\text{ and } j\neq0\}. \end{cases} \end{gather*} Then \begin{align*} &F\big(t,e_m^{(1)}(t+T),e_m^{(1)}(t)\big)\\ &= \begin{cases} F(0,\rho,\rho), & \text{if } t=0,\\ 0, & \text{if } t\in\{j\in\mathbb{Z}:-mM\leq j\leq mM-1\text{ and } j\neq0\}. \end{cases} \end{align*} With \eqref{e3.2} and (H4), we have \begin{equation}\label{e3.3} \begin{aligned} J_m\big(e_m^{(1)}\big) &=\frac{1}{2}\sum_{t=-mM}^{mM-1}L\big(e_m^{(1)}(t)\big)\cdot\big(e_m^{(1)}(t)\big)\\ &\quad -\sum_{t=-mM}^{mM-1}F\left(t,\big(e_m^{(1)}(t+T)\big),\big(e_m^{(1)}(t)\big)\right)\\ &\leq\frac{\bar{\lambda}}{2}\|e_m^{(1)}\|^2-2c_1\rho^2-b_1\\ & =-\big(2c_1-\frac{\bar{\lambda}}{2}\big)\rho^2-b_1\leq0. \end{aligned} \end{equation} All the assumptions of the Mountain Pass Lemma have been verified. Consequently, $J_m$ possesses a critical value $c_m$ given by \eqref{e2.5} and \eqref{e2.6} with $E=E_m$ and $\Gamma=\Gamma_m$, where \[ \Gamma_m=\big\{g_m\in C([0,1], E_m)|g_m(0)=0,\ g_m(1)=e_m^{(1)}, \ e_m^{(1)}\in E_m\backslash B_{\rho}\big\}. \] Let $u_m$ denote the corresponding critical point of $J_m$ on $E_m$. Note that $\|u_m\|\neq0$ since $c_m>0$. \end{proof} \begin{lemma} \label{lem3.3} Suppose that {\rm (H1)--(H6)} are satisfied. Then there exist positive constants $\delta_1$ and $\eta_1$ independent of $m$ such that \begin{equation}\label{e3.4} \delta_1\leq \|u_m\|_\infty\leq \eta_1. \end{equation} \end{lemma} \begin{proof} The continuity of $F(0,v_1,v_2)$ with respect to the second and third variables implies that there exists a constant $\tau_1>0$ such that $|F(0,v_1,v_2)|\leq \tau_1$ for $\sqrt{v_1^2+v_2^2}\leq\delta_1$. It is clear that \begin{align*} J_m(u_m) &\leq\max_{0\leq s\leq1}\Big\{\big|\frac{1}{2}\sum_{t=-mM}^{mM-1}L \big(se_m^{(1)}(t)\big)\cdot\big(se_m^{(1)}(t)\big)\big|\\ &\quad -\sum_{t=-mM}^{mM-1}F\left(t,se_m^{(1)}(t+T),se_m^{(1)}(t)\right)\Big\}\\ &\leq\frac{\bar{\lambda}}{2}\|e_m^{(1)}\|^2+\tau_1\\ &=\frac{\bar{\lambda}}{2}\rho^2+\tau_1. \end{align*} Let $\xi_1=\frac{\bar{\lambda}}{2}\rho^2+\tau_1$. Then $J_m(u_m)\leq\xi_1$, which is a bound independent of $m$. From \eqref{e2.3} and \eqref{e2.4}, we have \begin{align*} J_m(u_m) &=\frac{1}{2}\sum^{mM-1}_{t=-mM}\Big[\frac{\partial F(t-T,u_m(t),u_m(t-T))} {\partial v_2}u_m(t)\\ &\quad +\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}u_m(t)\Big] -\sum^{mM-1}_{t=-mM}F(t,u_m(t+T),u_m(t)) \\ &=\frac{1}{2}\sum^{mM-1}_{t=-mM}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}u_m(t+T)\\ &\quad +\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}u_m(t)\Big] -\sum^{mM-1}_{t=-mM}F(t,u_m(t+T),u_m(t)) \\ &\leq \xi_1. \end{align*} By (H5) and (H6), there exists a constant $\eta_1>0$ such that \[ \frac{1}{2}\Big(\frac{\partial F(t,v_1,v_2)}{\partial v_1}v_1+ \frac{\partial F(t,v_1,v_2)}{\partial v_2}v_2\Big)-F(t,v_1,v_2)>\xi_1, \] for all $t\in \mathbb{Z}$ and $\sqrt{v_1^2+v_2^2}\geq \eta_1$, which implies that $|u_m(t)|\leq \eta_1$ for all $t\in \mathbb{Z}$; that is, $\|u_m\|_\infty\leq \eta_1$. From the definition of $J_m$, we have \begin{align*} 0&=\left(J'_m(u_m),u_m\right) \\ &\geq\underline{\lambda}\|u_m\|^2 -\sum^{mM-1}_{t=-mM}\Big[\frac{\partial F(t-T,u_m(t),u_m(t-T))} {\partial v_2}u_m(t)\\ &\quad +\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}u_m(t)\Big]. \end{align*} This inequality and (H3) yield \begin{align*} \underline{\lambda}\|u_m\|^2 &\leq \sum^{mM-1}_{t=-mM}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}u_m(t+T)\\ &\quad +\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}u_m(t)\Big] \\ &\leq \Big\{\sum^{mM-1}_{t=-mM}\big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}\big]^2\Big\}^{1/2}\|u_m\|\\ &\quad + \Big\{\sum^{mM-1}_{t=-mM} \big[\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}\big]^2\Big\}^{1/2}\|u_m\|. \end{align*} That is, \begin{align*} \underline{\lambda}\|u_m\| &\leq \Big\{\sum^{mM-1}_{t=-mM}\big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}\big]^2\Big\}^{1/2}\\ &\quad + \Big\{\sum^{mM-1}_{t=-mM}\big[\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}\big]^2\Big\}^{1/2}. \end{align*} Thus, \begin{equation}\label{e3.5} \begin{aligned} \underline{\lambda}^2\|u_m\|^2 &\leq 2\sum^{mM-1}_{t=-mM}\big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}\big]^2\\ &\quad +2\sum^{mM-1}_{t=-mM}\big[\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}\big]^2. \end{aligned} \end{equation} From this inequality and (H3), we obtain \[ \underline{\lambda}^2\|u_m\|^2 \leq2\sum^{mM-1}_{t=-mM}\left[2a_1u_m(t+T)\right]^2 +2\sum^{mM-1}_{t=-mM}\left[2a_1u_m(t)\right]^2=16a_1^2\|u_m\|^2. \] Thus, we have $u_m=0$. This contradicts $\|u_m\|\neq0$, which shows that $$ \|u_m\|_\infty\geq\delta_1, $$ and the proof is complete. \end{proof} \begin{proof}[Proof of Theorem \ref{thm1.1}] Now we shall give the existence of a nontrivial homoclinic orbit. Consider the sequence $\{u_m(t)\}_{t\in \mathbb{Z}}$ of $2mM$-periodic solutions found in Section 3.1. First, by \eqref{e3.4}, for any $m\in \mathbb{N}$, there exists a constant $t_m\in\mathbb{Z}$ independent of $m$ such that \begin{equation}\label{e3.6} |u_m(t_m)|\geq \delta_1. \end{equation} Since $a(t)$, $b(t)$ and $f(t,v_1,v_2,v_3)$ are $M$-periodic in $t$, $\{u_m(t+jM)\}$ is also $2mM$-periodic solution of \eqref{e1.1} (for all $j\in \mathbb{N}$). Hence, making such shifts, we can assume that $t_m\in \mathbb{Z}(0,M-1)$ in \eqref{e3.6}. Moreover, passing to a subsequence of $m$s, we can even assume that $t_m=t_0$ is independent of $m$. Next, we extract a subsequence, still denote by $u_m$, such that $$ u_m(t)\to u(t),\text{as } m\to\infty,\; \forall t\in \mathbb{Z}. $$ Inequality \eqref{e3.6} implies that $|u(t_0)|\geq \xi$ and, hence, $u=\{u(t)\}$ is a nonzero sequence. Moreover, \begin{align*} &Lu(t)-f(t,u(t+T),u(t),u(t-T))\\ &=\lim_{m\to \infty}[Lu_m(t)-f(t,u_m(t+T),u_m(t),u_m(t-T))]=0. \end{align*} So $u=\{u(t)\}$ is a solution of \eqref{e1.1}. Finally, we show that $u\in l^2$. For $u_m\in E_m$, let \begin{gather*} P_m=\big\{t\in \mathbb{Z}: |u_m(t)|<\frac{\sqrt{2}}{2}\delta_1, -mM\leq t\leq mM-1\big\}, \\ Q_m=\big\{t\in \mathbb{Z}: |u_m(t)| \geq\frac{\sqrt{2}}{2}\delta_1, -mM\leq t\leq mM-1\big\}. \end{gather*} Since $F(t,v_1,v_2)\in C^1(\mathbb{Z}\times \mathbb{R}^2,\mathbb{R})$, there exist constants $\bar{\xi}>0$, $\underline{\xi}>0$ such that \begin{gather*} \max\Big\{\Big[\frac{\partial F(t,v_1,v_2)}{\partial v_1}\Big]^2 + \Big[\frac{\partial F(t,v_1,v_2)}{\partial v_2}\Big]^2: \delta_1\leq\sqrt{v_1^2+v_2^2}\leq\eta_1,\; t\in \mathbb{Z}\Big\} \leq \bar{\xi}, \\ \begin{aligned} \min\Big\{&\frac{1}{2}\Big[\frac{\partial F(t,v_1,v_2)}{\partial v_1}v_1+ \frac{\partial F(t,v_1,v_2)}{\partial v_2}v_2\Big] -F(t,v_1,v_2):\\ & \delta_1\leq\sqrt{v_1^2+v_2^2}\leq\eta_1,\; t\in \mathbb{Z}\Big\} \geq \underline{\xi}. \end{aligned} \end{gather*} For $t\in Q_m$, \begin{align*} &\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}\Big]^2 +\Big[\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}\Big]^2 \\ &\leq\frac{\bar{\xi}}{\underline{\xi}} \Big\{\frac{1}{2}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}u_m(t+T) +\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}u_m(t)\Big]\\ &\quad -F(t,u_m(t+T),u_m(t))\Big\}. \end{align*} By \eqref{e3.5}, we have \begin{align*} &\underline{\lambda}^2\|u_m\|^2\\ &\leq 2\sum_{t\in P_m}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}\Big]^2 +2\sum_{t\in P_m}\big[\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}\big]^2\\ &\quad +2\sum_{t\in Q_m}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}\Big]^2 +2\sum_{t\in Q_m}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}\Big]^2 \\ &\leq 2\sum_{t\in P_m}[2a_1u_m(t+T)]^2 +2\sum_{t\in P_m}[2a_1u_m(t)]^2\\ &\quad +\frac{\bar{\xi}}{\underline{\xi}}\sum_{t\in Q_m} \Big\{\frac{1}{2}\Big[\frac{\partial F(t,u_m(t+T),u_m(t))} {\partial v_1}u_m(t+T)\\ &\quad +\frac{\partial F(t,u_m(t+T),u_m(t))}{\partial v_2}u_m(t)\Big] -F(t,u_m(t+T),u_m(t))\Big\} \\ &\leq 16a_1^2\|u_m\|^2+\frac{\bar{\xi}\xi_1}{\underline{\xi}}. \end{align*} Thus, $$ \|u_m\|^2\leq \frac{\bar{\xi}\xi_1}{\underline{\xi} \left(\underline{\lambda}^2-16a_1^2\right)}. $$ For any fixed $D\in \mathbb{Z}$ and $m$ large enough, we have $$ \sum^D_{t=-D}u_m^2(t)\leq\|u_m\|^2 \leq\frac{\bar{\xi}\xi_1}{\underline{\xi}\left(\underline{\lambda}^2-16a_1^2\right)}. $$ Since $\bar{\xi}, \underline{\xi}, \xi_1, \underline{\lambda}$ and $a_1$ are constants independent of $m$, passing to the limit, we have $$ \sum^D_{t=-D}u^2(t) \leq \frac{\bar{\xi}\xi_1}{\underline{\xi}\left(\underline{\lambda}^2-16a_1^2\right)}. $$ By the arbitrariness of $D$, $u\in l^2$. Therefore, $u$ satisfies $u(t)\to 0$ as $|t|\to \infty$. The existence of a nontrivial homoclinic orbit is obtained. \end{proof} \subsection{Proof Theorem \ref{thm1.5}} Let \begin{equation}\label{e3.7} J_m^\ast(u)=-\frac{1}{2}\sum_{t=-mM}^{mM-1}Lu(t)\cdot u(t)+ \sum_{t=-mM}^{mM-1}F(t,u(t+T),u(t)). \end{equation} Then \begin{equation}\label{e3.8} \frac{\partial J_m^\ast(u)}{\partial u(t)}= -Lu(t)+f(t,u(t+T),u(t),u(t-T)), \end{equation} for all $t\in \mathbb{Z}(-mM,mM-1)$. \begin{lemma} \label{lem3.4} Suppose that {\rm (H1), (H2), (H7), (H8)} are satisfied. Then $J_m^\ast$ satisfies the PS condition. \end{lemma}. \begin{proof} Assume that $\{u_j\}_{j\in\mathbb{N}}$ in $E_m$ is a sequence such that $\{J_m^\ast(u_j)\}_{j\in\mathbb{N}}$ is bounded. Then there exists a constant $K_2>0$ such that $-K_2\leq J_m^\ast(u_j)$. By \eqref{e2.7} and (H9'), it is easy to see that \[ -K_2\leq J^\ast_m(u_j)\leq -\frac{\underline{\lambda}}{2}\|u_j\|^2+2c_2\|u_j\|^2+2mM\left(\zeta_2+b_2\right), \quad \forall j\in \mathbb{N}. \] Therefore, \begin{equation}\label{e3.9} -\big(2c_2-\frac{\underline{\lambda}}{2}\big)\|u_j\|^2 \leq 2mM\left(\zeta_2+b_2\right)+K_2. \end{equation} Since $c_2<\underline{\lambda}/4$, \eqref{e3.9} implies that $\{u_j\}_{j\in\mathbb{N}}$ is bounded in $E_m$. Thus, $\{u_j\}_{j\in\mathbb{N}}$ possesses a convergence subsequence in $E_m$. The desired result follows. \end{proof} \begin{lemma} \label{lem3.5} Suppose that {\rm (H1), (H2), (H7), (H8)} are satisfied. Then for any given positive integer $m$, \eqref{e1.1} possesses a $2mM$-periodic solution $u_m^\ast\in E_m$. \end{lemma} \begin{proof} In our case, it is clear that $J_m^\ast(0)=0$. By Lemma \ref{lem3.4}, $J_m^\ast$ satisfies the PS condition. By (H7), we have \begin{align*} J_m^\ast(u) &\geq -\frac{\bar{\lambda}}{2}\|u\|^2 +a_2\sum^{mM-1}_{t=-mM}[u^2(t)+u^2(t+T)]\\ &\geq -\frac{\bar{\lambda}}{2}\|u\|^2+2a_2\|u\|^2\\ &=-\big(\frac{\bar{\lambda}}{2}-2a_2\big)\|u\|^2. \end{align*} Taking $\alpha_2=-(\frac{\bar{\lambda}}{2}-2a_2)\delta_2^2>0$, we obtain $$ J_m^\ast(u)|_{\partial B_{\delta_2}}\geq\alpha_2>0, $$ which implies that $J_m^\ast$ satisfies the condition (J1) of the Mountain Pass Lemma. Next, we shall verify the condition (J2) of the Mountain Pass Lemma. There exists a sufficiently large number $\eta>\max\{\rho_2,\delta_2\}$ such that \begin{equation}\label{e3.10} \big(2c_2-\frac{\bar{\lambda}}{2}\big)\eta^2\geq|b_2|. \end{equation} Let $e_m^{(2)}\in E_m$ and \begin{gather*} e_m^{(2)}(t)=\begin{cases} \eta, & \text{if } t=0,\\ 0, & \text{if } t\in\{j\in\mathbb{Z}:-mM\leq j\leq mM-1\text{ and } j\neq0\}, \end{cases} \\ e_m^{(2)}(t+T)= \begin{cases} \eta, & \text{if } t=0,\\ 0, & \text{if } t\in\{j\in\mathbb{Z}:-mM\leq j\leq mM-1\text{ and }j\neq0\}. \end{cases} \end{gather*} Then \begin{align*} &F\big(t,e_m^{(2)}(t+T),e_m^{(2)}(t)\big)\\ &= \begin{cases} F(0,\eta,\eta), & \text{if } t=0,\\ 0, & \text{if } t\in\{j\in\mathbb{Z}:-mM\leq j\leq mM-1\text{ and } j\neq0\}, \end{cases} \end{align*} With \eqref{e3.10} and (H9), we have \begin{equation}\label{e3.11} \begin{aligned} J_m^\ast\big(e_m^{(2)}\big) &=-\frac{1}{2}\sum_{t=-mM}^{mM-1}L\big(e_m^{(2)}(t)\big) \cdot\big(e_m^{(2)}(t)\big)\\ &\quad +\sum_{t=-mM}^{mM-1}F\Big(t,e_m^{(2)}(t+T),e_m^{(2)}(t)\Big)\\ &\leq-\frac{\underline{\lambda}}{2}\|e_m^{(2)}\|^2+2c_2\eta^2+b_2\\ & =-\big(\frac{\underline{\lambda}}{2}-2c_2\big)\eta^2+b_2\leq0. \end{aligned} \end{equation} All the assumptions of the Mountain Pass Lemma have been verified. Consequently, $J_m^\ast$ possesses a critical value $c_m^\ast$ given by \eqref{e2.5} and \eqref{e2.6} with $E=E_m$ and $\Gamma=\Gamma_m$, where \[ \Gamma_m=\big\{g_m\in C([0,1], E_m)|g_m(0)=0,\; g_m(1)=e_m^{(2)},\; e_m^{(2)}\in E_m\backslash B_{\eta}\big\}. \] Let $u_m^\ast$ denote the corresponding critical point of $J_m^\ast$ on $E_m$. Note that $\|u_m^\ast\|\neq0$ since $c_m^\ast>0$. \end{proof} \begin{lemma} \label{lem3.6} Suppose that {\rm (H1), (H2), (H7), (H8)} are satisfied. Then there exist positive constants $\delta_2$ and $\eta_2$ independent of $m$ such that \begin{equation}\label{e3.12} \delta_2\leq \|u_m^\ast\|_\infty\leq \eta_2. \end{equation} \end{lemma} \begin{proof} The continuity of $F(0,v_1,v_2)$ with respect to the second and third variables implies that there exists a constant $\tau_2>0$ such that $|F(0,v_1,v_2)|\leq \tau_2$ for $\sqrt{v_1^2+v_2^2}\leq\delta_2$. It is clear that \begin{equation}\label{e3.13} \begin{aligned} |J_m^\ast\left(u_m^\ast\right)| &\leq\max_{0\leq s\leq1}\Big\{\big|-\frac{1}{2}\sum_{t=-mM}^{mM-1} L\big(se_m^{(2)}(t)\big)\cdot\big(se_m^{(2)}(t)\big)\big|\\ &\quad +\sum_{t=-mM}^{mM-1}F\left(t,se_m^{(2)}(t+T),se_m^{(2)}(t)\right)\Big\}\\ &\leq\frac{\bar{\lambda}}{2}\|e_m^{(2)}\|^2+\tau_2 \\ &=\frac{\bar{\lambda}}{2}\eta^2+\tau_2. \end{aligned} \end{equation} Let $\xi_2=\frac{\bar{\lambda}}{2}\eta^2+\tau_2$, we have that $|J_m^\ast\left(u_m^\ast\right)|\leq\xi_2$, which is a bound independent of $m$. Then by \eqref{e3.7} and \eqref{e3.8}, we have \begin{align*} \xi_2 & \geq J_m^\ast(u_m)\\ &=-\frac{1}{2}\sum^{mM-1}_{t=-mM} \Big[\frac{\partial F(t-T,u_m^\ast(t),u_m^\ast(t-T))} {\partial v_2}u_m^\ast(t)\\ &\quad +\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))}{\partial v_2}u_m^\ast(t)\Big] +\sum^{mM-1}_{t=-mM}F(t,u_m^\ast(t+T),u_m^\ast(t)) \\ &=-\frac{1}{2}\sum^{mM-1}_{t=-mM} \Big[\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))} {\partial v_1}u_m^\ast(t+T)\\ &\quad +\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))}{\partial v_2}u_m^\ast(t)\Big] +\sum^{mM-1}_{t=-mM}F(t,u_m^\ast(t+T),u_m^\ast(t)) \\ &\geq\big(\frac{2-\mu}{2}\big)\sum^{mM-1}_{t=-mM}F(t,u_m^\ast(t+T),u_m^\ast(t))\,. \end{align*} Then \begin{equation}\label{e3.14} \sum^{mM-1}_{t=-mM}F(t,u_m^\ast(t+T),u_m^\ast(t))\leq \frac{2\xi_2}{2-\mu}. \end{equation} Since \begin{align*} J_m^\ast(u_m^\ast) &=-\frac{1}{2}\sum^{mM-1}_{t=-mM} \Big[\frac{\partial F(t-T,u_m^\ast(t),u_m^\ast(t-T))} {\partial v_2}u_m^\ast(t)\\ &\quad +\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))}{\partial v_2}u_m^\ast(t)\Big] +\sum^{mM-1}_{t=-mM}F(t,u_m^\ast(t+T),u_m^\ast(t))\\ &\geq-\xi_2. \end{align*} This inequality combined with \eqref{e3.14} gives us \begin{equation}\label{e3.15} \begin{aligned} \frac{1}{2}\underline{\lambda}\|u_m^\ast\| &\leq\frac{1}{2}\sum^{mM-1}_{t=-mM} \Big[\frac{\partial F(t-T,u_m^\ast(t),u_m^\ast(t-T))} {\partial v_2}u_m^\ast(t)\\ &\quad +\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))}{\partial v_2}u_m^\ast(t)\Big]\\ &\leq\sum^{mM-1}_{t=-mM}F(t,u_m^\ast(t+T),u_m^\ast(t))+\xi_2\\ &\leq \frac{(4-\mu)\xi_2}{2-\mu}, \end{aligned} \end{equation} \begin{equation}\label{e3.16} \|u_m^\ast\|\leq \frac{2(4-\mu)\xi_2}{(2-\mu)\underline{\lambda}}, \end{equation} whose right-hand side is independent of $m$. Then $\|u_m^\ast\|\leq \eta_2$, which implies $$ \|u_m^\ast\|_\infty\leq \eta_2. $$ From the definition of $J_m^\ast$, we have \[ 0=\big(J_m^{\ast'}(u_m^\ast),u_m^\ast\big) \geq-\bar{\lambda}\|u_m^\ast\|^2 +\sum^{mM-1}_{t=-mM}f(t,u_m^\ast(t+T),u_m^\ast(t),u_m^\ast(t-T))u_m^\ast(t). \] This inequality combined with (H7) yields \begin{align*} \bar{\lambda}\|u_m^\ast\|^2 &\geq \sum^{mM-1}_{t=-mM}\Big[\frac{\partial F(t-T,u_m^\ast(t),u_m^\ast(t-T))} {\partial v_2}u_m^\ast(t) \\ &\quad +\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))}{\partial v_2}u_m^\ast(t)\Big] \\ &=\sum^{mM-1}_{t=-mM}\Big[\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))} {\partial v_1}u_m^\ast(t+T)\\ &\quad +\frac{\partial F(t,u_m^\ast(t+T),u_m^\ast(t))}{\partial v_2}u_m^\ast(t)\Big]\\ &\geq 2a_2\sum^{mM-1}_{t=-mM}\big[\left(u_m^\ast(t+T)\right)^2 +\left(u_m^\ast(t)\right)^2\big] \\ &=4a_2\|u_m^\ast\|^2. \end{align*} Thus, we have $u_m^\ast=0$. This contradicts $\|u_m^\ast\|\neq0$, which shows that $\|u_m^\ast\|_\infty\geq\delta_2$, and the proof is complete. \end{proof} The proof of Theorem \ref{thm1.5} is done similarly to the proof of Theorem \ref{thm1.1}. We omit it here for simplicity. \section{Examples} As an application of Theorems \ref{thm1.1} and \ref{thm1.5}, we give two examples that illustrate our main results. \begin{example} \label{examp4.1} \rm Let \begin{gather*} f(t,v_1,v_2,v_3) =\gamma v_2\Big(\frac{v_1^2+v_2^2}{v_1^2+v_2^2+1} +\frac{v_2^2+v_3^2}{v_2^2+v_3^2+1}\Big), \\ F(t,v_1,v_2)=\frac{\gamma}{2}[v_1^2+v_2^2-\ln(v_1^2+v_2^2+1)], \end{gather*} where $\gamma>\bar{\lambda}$. If (H1) is satisfied, then it is easy to verify that all the assumptions of Theorem \ref{thm1.1} are satisfied. Consequently, there exists a nontrivial homoclinic orbit. \end{example} \begin{example} \label{examp4.2} \rm Let $$ f(t,v_1,v_2,v_3)= \begin{cases} v_2[(v_1^2+v_2^2)^{\frac{\mu}{2}-1}+(v_2^2+v_3^2)^{\frac{\mu}{2}-1}],\\ \quad \text{if } (v_1,v_2)\neq(0,0)\text{ and } (v_2,v_3)\neq(0,0),\\[4pt] 0, \quad\text{if } (v_1,v_2)=(0,0)\text{ or } (v_2,v_3)=(0,0), \end{cases} $$ and $$ F(t,v_1,v_2)=\frac{1}{\mu}(v_1^2+v_2^2)^{\mu/2}, $$ where $1<\mu<2$. If (H1) is satisfied, then it is easy to verify all the assumptions of Theorem \ref{thm1.5} are satisfied. Consequently, there exists a nontrivial homoclinic orbit. \end{example} \subsection*{Acknowledgments} This project is supported by the National Natural Science Foundation of China (No. 11401121), by the Natural Science Foundation of Guangdong Province (No. S2013010014460), and by the Hunan Provincial Natural Science Foundation of China (No. 2015JJ2075). \begin{thebibliography}{99} \bibitem{Ag} R. P. Agarwal; \emph{Difference Equations and Inequalities: Theory, Methods and Applications}, Marcel Dekker: New York, 2000. \bibitem{CaM} P. Candito, G. Molica Bisci; \emph{Existence of solutions for nonlinear algebraic systems with a parameter}, Appl. Math. Comput., 218(23) (2012), pp. 11700-11707. \bibitem{ChF} P. Chen, H. Fang; \emph{Existence of periodic and subharmonic solutions for second-order $p$-Laplacian difference equations}, Adv. Difference Equ., 2007 (2007), pp. 1-9. \bibitem{ChT1} P. Chen, X. H. Tang; \emph{Existence of infinitely many homoclinic orbits for fourth-order difference systems containing both advance and retardation}, Appl. Math. Comput., 217(9) (2011), pp. 4408-4415. \bibitem{ChT2} P. Chen, X. H. Tang; \emph{Existence and multiplicity of homoclinic orbits for 2$n$th-order nonlinear difference equations containing both many advances and retardations}, J. Math. Anal. Appl., 381(2) (2011), pp. 485-505. \bibitem{ChT3} P. Chen, X. H. Tang; \emph{Infinitely many homoclinic solutions for the second-order discrete $p$-Laplacian systems}, Bull. Belg. Math. Soc., 20(2) (2013), pp. 193-212. \bibitem{ChT4} P. Chen, X. H. Tang; \emph{Existence of homoclinic solutions for some second-order discrete Hamiltonian systems}, J. Difference Equ. Appl., 19(4) (2013), pp. 633-648. \bibitem{ChW} P. Chen, Z. M. Wang; \emph{Infinitely many homoclinic solutions for a class of nonlinear difference equations}, Electron. J. Qual. Theory Differ. Equ., (47) (2012), pp. 1-18. \bibitem{CuF} P. Cull, M. Flahive, R. Robson; \emph{Difference Equations: From Rabbits to Chaos}, Springer: New York, 2005. \bibitem{DeC} X.Q. Deng, G. Cheng; \emph{Homoclinic orbits for second order discrete Hamiltonian systems with potential changing sign}, Acta Appl. Math., 103(3) (2008), pp. 301-314. \bibitem{FaZ} H. Fang, D. P. Zhao; \emph{Existence of nontrivial homoclinic orbits for fourth-order difference equations}, Appl. Math. Comput., 214(1) (2009), pp. 163-170. \bibitem{GuAWO} C. J. Guo, R. P. Agarwal, C. J. Wang, D. O'Regan; \emph{The existence of homoclinic orbits for a class of first order superquadratic Hamiltonian systems}, Mem. Differential Equations Math. Phys., 61 (2014), pp. 83-102. \bibitem{GuOA} C. J. Guo, D. O'Regan, R. P. Agarwal; \emph{Existence of homoclinic solutions for a class of the second-order neutral differential equations with multiple deviating arguments}, Adv. Dyn. Syst. Appl., 5(1) (2010), pp. 75-85. \bibitem{GuOXA1} C. J. Guo, D. O'Regan, Y. T. Xu, R. P. Agarwal; \emph{Existence of subharmonic solutions and homoclinic orbits for a class of high-order differential equations}, Appl. Anal., 90(7) (2011), pp. 1169-1183. \bibitem{GuOXA2} C. J. Guo, D. O'Regan, Y. T. Xu, R. P. Agarwal; \emph{Homoclinic orbits for a singular second-order neutral differential equation}, J. Math. Anal. Appl., 366(2) (2010), pp. 550-560. \bibitem{GuOXA3} C. J. Guo, D. O'Regan, Y. T. Xu, R. P. Agarwal; \emph{Existence and multiplicity of homoclinic orbits of a second-order differential difference equation via variational methods}, Appl. Math. Inform. Mech., 4(1) (2012), pp. 1-15. \bibitem{GuOXA4} C. J. Guo, D. O'Regan, Y. T. Xu, R. P. Agarwal; \emph{Existence of homoclinic orbits of a class of second order differential difference equations}, Dyn. Contin. Discrete Impuls. Syst. Ser. B Appl. Algorithms, 20(6) (2013), pp. 675-690. \bibitem{GuY1} Z. M. Guo, J. S. Yu; \emph{Existence of periodic and subharmonic solutions for second-order superlinear difference equations}, Sci. China Math., 46(4) (2003), pp. 506-515. \bibitem{GuY2} Z. M. Guo, J. S. Yu; \emph{The existence of periodic and subharmonic solutions of subquadratic second order difference equations}, J. London Math. Soc., 68(2) (2003), pp. 419-430. \bibitem{Lo1} Y. H. Long; \emph{Homoclinic solutions of some second-order non-periodic discrete systems}, Adv. Difference Equ., 2011 (2011), pp. 1-12. \bibitem{Lo2} Y. H. Long; \emph{Homoclinic orbits for a class of noncoercive discrete Hamiltonian systems}, J. Appl. Math., 2012 (2012), pp. 1-21. \bibitem{MaG1} M.J. Ma, Z.M. Guo; \emph{Homoclinic orbits for second order self-adjoint difference equations}, J. Math. Anal. Appl., 323(1) (2006), pp. 513-521. \bibitem{MaG2} M.J. Ma, Z.M. Guo; \emph{Homoclinic orbits and subharmonics for nonlinear second order difference equations}, Nonlinear Anal., 67(6) (2007), pp. 1737-1745. \bibitem{MaW} J. Mawhin, M. Willem; \emph{Critical Point Theory and Hamiltonian Systems}, Springer: New York, 1989. \bibitem{MoR} G. Molica Bisci, D. Repov$\check{s}$; \emph{Existence of solutions for $p$-Laplacian discrete equations}, Appl. Math. Comput., 242(1) (2014), pp. 454-461. \bibitem{Po} H. Poincar\'e; \emph{Les m\'ethodes nouvelles de la m\'ecanique c\'eleste}, Gauthier-Villars: Paris, 1899. \bibitem{Ra1} P. H. Rabinowitz; \emph{Minimax Methods in Critical Point Theory with Applications to Differential Equations}, Providence, RI: New York, 1986. \bibitem{Ra2} P. H. Rabinowitz; \emph{Homoclinic orbits for a class of Hamiltonian systems}, Proc. Roy. Soc. Edinburgh Sect. A, 114(1-2) (1990), pp. 33-38. \bibitem{Te} G. Teschl; \emph{Jacobi Operators and Completely Integrable Nonlinear Lattices}, Providence, RI: New York, 2000. \bibitem{ZhYC} Z. Zhou, J.S. Yu, Y.M. Chen; \emph{Homoclinic solutions in periodic difference equations with saturable nonlinearity}, Sci. China Math, 54(1) (2011), pp. 83-93. \end{thebibliography} \end{document}