\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 134, pp. 1--15.\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-2015/134\hfil Existence of periodic solutions] {Existence of periodic solutions for sub-linear first-order Hamiltonian systems} \author[M. Timoumi \hfil EJDE-2015/134\hfilneg] {Mohsen Timoumi} \address{Mohsen Timoumi \newline Department of Mathematics, Faculty of Sciences, 5000 Monastir, Tunisia} \email{m\_timoumi@yahoo.com} \thanks{Submitted March 13, 2014. Published May 15, 2015.} \subjclass[2010]{34C25} \keywords{Hamiltonian systems; periodic solutions; saddle point theorem; \hfill\break\indent least action principle; sub-linear conditions} \begin{abstract} We prove the existence solutions for the sub-linear first-order Hamiltonian system $J\dot{u}(t)+Au(t)+\nabla H(t,u(t))=h(t)$ by using the least action principle and a version of the Saddle Point Theorem. \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} In this article, we consider the first-order Hamiltonian system \begin{equation} J\dot{u}(t)+Au(t)+\nabla H(t,u(t))=h(t) \label{e1.1} \end{equation} where $A$ is a $(2N\times 2N)$ symmetric matrix, $H\in C^{1}(\mathbb{R}\times\mathbb{R}^{2N},\mathbb{R})$ is $T$-periodic in the first variable $(T>0)$ and $h\in C(\mathbb{R},\mathbb{R}^{2N})$ is $T$-periodic. When $A=0$ and $h=0$, it has been proved that system $\eqref{e1.1}$ has at least one $T$-periodic solution by the use of critical point theory and minimax methods \cite{a1,c1,e1,f1,g1,l1,l2, t1,t3,x1}. Many solvability conditions are given, such as the convex condition (see [3,5]), the super-quadratic condition (see \cite{a1,f1,l1,l2,l4,r2,t1,x1}), the sub-linear condition (see \cite{c1,t3}). When $A$ is not identically null, the existence of periodic solutions for \eqref{e1.1} has been studied in \cite{l2,t2}. In all these last papers, the Hamiltonian is assumed to be super-quadratic. As far as the general case ($A$ not identically null) is concerned, to our best knowledge, there is no research about the existence of periodic solutions for \eqref{e1.1} when $H$ is sub-linear. In \cite{c1}, the authors considered the special case $A=0$ and $h=0$ and obtain the existence of subharmonic solutions for \eqref{e1.1} under the following assumptions: \smallskip \noindent(A1) There exist constants $a,b,c>0$, $\alpha\in [0,1[$, functions $p\in L^{\frac{2}{1-\alpha}}(0,T;\mathbb{R}^{+})$, $q\in L^2(0,T;\mathbb{R}^{+})$ and a nondecreasing function $\gamma\in C(\mathbb{R}^{+},\mathbb{R}^{+})$ with the following properties: \begin{itemize} \item[(i)] $\gamma(s+t)\leq c(\gamma(s)+\gamma(t))$ for all $s,t\in\mathbb{R}^{+}$, \item[(ii)] $\gamma(t)\leq at^{\alpha}+b$ for all $t\in\mathbb{R}^{+}$, \item[(iii)] $\gamma(t)\to +\infty\ as\ t\to+\infty$, such that \begin{gather*} |\nabla H(t,x)|\leq p(t)\gamma(|x|)+q(t),\quad \forall x\in\mathbb{R}^{2N}, \text{ a.e. } t\in[0,T];\\ \lim_{|x|\to\infty}\frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt=\pm\infty. \end{gather*} \end{itemize} Similarly, in \cite{t3} the author considered the case $A=0$ and $h=0$ and obtained the existence os subharmonic solutions for \eqref{e1.1} under the following assumptions: \smallskip \noindent(A2) There exist a positive constant $a$, $g\in L^2(0,T;\mathbb{R})$ and a non-increasing function $\omega\in C(\mathbb{R}^{+},\mathbb{R}^{+})$ with the properties: \begin{gather*} \liminf_{s\to\infty}\frac{\omega(s)}{\omega(\sqrt{s})} > 0,\\ \omega(s)\to 0,\quad \omega(s)s\to\infty\quad\text{as }s\to\infty, \end{gather*} such that \begin{gather*} |\nabla H(t,x)|\leq a\omega(|x|)|x|+g(t),\quad \forall x\in\mathbb{R}^{2N},\text{ a.e. } t\in [0,T];\\ \frac{1}{[\omega(|x|)|x|]^2}\int^T_0H(t,x)dt \to+\infty \quad \text{as }|x|\to\infty. \end{gather*} In Sections 4,5, we will use the Least Action Principle and a version of the Saddle Point Theorem to study the existence of periodic solutions for \eqref{e1.1}, when $A$ and $h$ are not necessary null and $H$ satisfies some more general variants conditions replacing conditions (A1), (A2). \section{Preliminaries} Let $T>0$ and $A$ be a $(2N\times 2N)$ symmetric matrix. Consider the Hilbert space $H^{1/2}(S^{1},\mathbb{R}^{2N})$ where $S^{1}=\mathbb{R}/(T\mathbb{Z})$ and the continuous quadratic form $Q$ defined on $E$ by $$ Q(u)=\frac{1}{2}\int^T_0(J\dot{u}(t)\cdot u(t) +Au(t)\cdot u(t))dt $$ where $x\cdot y$ is the inner product of $x,y\in\mathbb{R}^{2N}$. Let us denote by $E^{0}$, $E^{-}$, $E^{+}$ respectively the subspaces of $E$ on which $Q$ is null, negative definite and positive definite. It is well known that these subspaces are mutually orthogonal in $L^2(S^{1},\mathbb{R}^{2N})$ and in $E$ with respect to the bilinear form $$ B(u,v)=\frac{1}{2}\int^T_0(J\dot{u}(t)\cdot v(t) +Au(t)\cdot v(t))dt,\ u,v\in E $$ associated with $Q$. If $u\in E^{+}$ and $v\in E^{-}$, then $B(u,v)=0$ and $Q(u+v)=Q(u)+Q(v)$. For $u=u^{-}+u^{0}+u^{+}\in E$, the expression $\|u\|=[Q(u^{+})-Q(u^{-})+|u^{0}|^2]^{1/2}$ is an equivalent norm in $E$. It is well known that the space $E$ is compactly embedded in $L^{s}(S^{1},\mathbb{R}^{2N})$ for all $s\in [1,\infty[$. In particular, for all $s\in [1,\infty[$, there exists $\lambda_{s}>0$ such that for all $u\in E$, \begin{equation} \|u\|_{L^{s}}\leq\lambda_{s}\|u\|. \label{e2.1} \end{equation} Next, we have a version of the Saddle Point Theorem \cite{r1}. \begin{lemma} \label{lem2.1} Let $E=E^{1}\oplus E^2$ be a real Hilbert space with $E^2=(E^{1})^{\bot}$. Suppose that $f\in C^{1}(E,\mathbb{R})$ satisfies \begin{itemize} \item[(a)] $f(u)=\frac{1}{2}\langle Lu,u\rangle +g(u)$ and $Lu=L_1P_1u+L_2P_2u$ with $L_{i}:E^{i}\to E^{i}$ bounded and self-adjoint, $i=1,2$; \item[(b)] $g'$ is compact; \item[(c)] There exists $\beta\in\mathbb{R}$ such that $f(u)\leq\beta$ for all $u\in E^{1}$; \item[(d)] There exists $\gamma\in\mathbb{R}$ such that $f(u)\geq\gamma$ for all $u\in E^2$. \end{itemize} Furthermore, if $f$ satisfies the Palais-Smale condition $(PS)_{c}$ for all $c\geq\gamma$, then $f$ possesses a critical value $c\in[\gamma,\beta]$. \end{lemma} \section{Linear Hamiltonian systems} Let $A$ be a $(2N\times 2N)$ symmetric matrix, we consider the linear Hamiltonian system \begin{equation} \dot{x}=JAx. \label{e3.1} \end{equation} Let $\lambda_1,\dots ,\lambda_{s}$ be all the distinct eigenvalues of $B=JA$ and $F_1,\dots ,F_{s}$ be the corresponding root subspaces. The dimension of the root subspace $F_{\sigma}$ is equal to the multiplicity $m_{\sigma}$ of the corresponding root $\lambda_{\sigma}$ of the characteristic equation $det(B-\lambda I_{2N})=0$ $(m_1+\dots +m_{s}=2N)$. The space $\mathbb{R}^{2N}$ splits into a direct sum of the $B$-invariant subspaces $F_{\sigma}$: \begin{equation} \mathbb{R}^{2N}=F_1\oplus\dots \oplus F_{s}. \label{e3.2} \end{equation} Each subspace $F_{\sigma}$ possesses a basis $(a^{\sigma}_1,\dots ,a^{\sigma}_{m_{\sigma}})$ satisfying $$ Ba^{\sigma}_1=\lambda_{\sigma}a^{\sigma}_1,\ Ba^{\sigma}_2 =\lambda_{\sigma}a^{\sigma}_2+a^{\sigma}_1,\dots , Ba^{\sigma}_{m_{\sigma}}=\lambda_{\sigma}a^{\sigma}_{m_{\sigma}} +a^{\sigma}_{m_{\sigma}-1}. $$ The $(m_{\sigma}\times m_{\sigma})$ matrix \[ Q_{\sigma}(\lambda_{\sigma})= \begin{pmatrix} \lambda_{\sigma} & 1 & 0 & 0 & \dots & 0 \\ 0 & \lambda_{\sigma} & 1 & 0 & \dots & 0 \\ . & . & . & . & \dots & . \\ 0 & 0 & 0 & \dots & \lambda_{\sigma} & 1 \\ 0 & 0 & 0 &\dots & 0 & \lambda_{\sigma} \\ \end{pmatrix} \] is called an elementary Jordan matrix. We have $B=SQS^{-1}$ where $Q$ is a direct sum of elementary Jordan matrices \[ Q= \begin{pmatrix} Q_1(\lambda_1) & 0 & 0 & \dots & 0 \\ 0 & Q_2(\lambda_2) & 0 & \dots & 0 \\ . & . & . & \dots & . \\ . & 0 & 0 & \dots & Q_{s}(\lambda_{s}) \\ \end{pmatrix} = Q_1(\lambda_1)\oplus\dots \oplus Q_{s}(\lambda_{s}) \] the columns of the matrix $S$, $$ a^{1}_1,\dots ,a^{1}_{m_1};\ a^2_1,\dots ,a^2_{m_2}; \dots ;\ a^{s}_1,\dots ,a^{s}_{m_{s}}$$ form a basis for $\mathbb{R}^{2N}$ and so $det(S)\neq 0$. The matrizant of equation \eqref{e3.1} is given by $$ R(t)=e^{tB}=S[\exp(tQ_1(\lambda_1))\oplus \dots \oplus \exp(tQ_{s}(\lambda_{s}))]S^{-1}=S e^{tQ}S^{-1}. $$ then the solution of equation \eqref{e3.1} with initial condition $x(0)$ is $$ x(t)=e^{tB}x(0). $$ Therefore to each eigenvalue $\lambda_{\sigma}$ corresponds a group of $m_{\sigma}$-linearly independent solutions: \begin{equation} \begin{gathered} x^{\sigma}_1(t)=e^{\lambda_{\sigma}t}a^{\sigma}_1\\ x^{\sigma}_2(t)=e^{\lambda_{\sigma}t}(ta^{\sigma}_1+a^{\sigma}_2)\\ \dots \\ x^{\sigma}_{m_{\sigma}}(t)=e^{\lambda_{\sigma}t} (\frac{1}{(m_{\sigma}-1)!}t^{m_{\sigma}-1}a^{\sigma}_1 +\dots +a^{\sigma}_{m_{\sigma}}). \end{gathered} \label{e3.3} \end{equation} Moreover, combining the solutions of all the groups \eqref{e3.3} (there are obviously $2N$ in all, since $m_1+\dots +m_{s}=2N$), we obtain a complete system of linearly independent solutions of \eqref{e3.1}. Now, assume that $\lambda_1=0$ is an eigenvalue of $B=JA$ and let $1\leq m\leq m_1$ be the dimension of the corresponding eigenspace $E_1$. We can replace the basis $(a^{1}_1,\dots ,a^{1}_{m_1})$ of the root subspace $F_1$ by the basis $(b^{1}_1,\dots ,b^{1}_{m_1})$ where $(b^{1}_1,\dots ,b^{1}_{m})$ is a basis of $E_1$, $b^{1}_{j}=a^{1}_{j}$ for $m+1\leq j \leq m_1$ and such that $b^{1}_{m+1}=B b^{1}_{m}$. To this basis corresponds the group of $2N$ linearly independent solutions: \begin{equation} \begin{gathered} u^{1}_1(t)=b^{1}_1\\ \dots \\ u^{1}_{m}(t)=b^{1}_{m}\\ u^{1}_{m+1}(t)=b^{1}_{m}t+b^{1}_{m+1}\\ \dots \\ u^{1}_{m_1}(t)=\frac{1}{(m_1-m)!}b^{1}_{m}t^{m_1-m}+\dots +b^{1}_{m_1}\\ u^{\sigma}_{k}(t)=x^{\sigma}_{k}(t),\quad 2\leq\sigma\leq s,\; 1\leq k\leq m_{\sigma}. \end{gathered} \label{e3.4} \end{equation} A solution $u$ of \eqref{e3.1} may be written in the form $$ u(t)=\sum^{s}_{\sigma=1}\sum^{m_{\sigma}}_{j=1} \alpha^{\sigma}_{j}u^{\sigma}_{j}(t). $$ Let $T>0$ be such that $\lambda_{\sigma}T\notin 2i\pi\mathbb{Z}$ for all $1\leq\sigma\leq s$. If $u$ is $T$-periodic, then for any $1\leq \sigma\leq s$, we have $$ \sum^{m_{\sigma}}_{j=1}\alpha^{\sigma}_{j}u^{\sigma}_{j}(kT) =\sum^{m_{\sigma}}_{j=1}\alpha^{\sigma}_{j}u^{\sigma}_{j}(0),\quad \forall k\in\mathbb{Z}. $$ It is easy to see that $\alpha^{1}_{j}=0$ for $m+1\leq j\leq m_1$ and $\alpha^{\sigma}_{j}=0$ for $2\leq\sigma\leq s$ and $1\leq j\leq m_{m_{\sigma}}$. Therefore, $u(t)=\sum^{m}_{j=1}\alpha^{1}_{j}b^{1}_{j}$. Hence the set of $T$-periodic solutions of \eqref{e3.1} is equal to $N(A)$. \begin{example} \label{examp3.1}\rm Let \[ A= \begin{pmatrix} -12 & 6 & 5 & 1 \\ -2 & 1 & 0 & 1 \\ 2 & -1 & 0 & -1\\ 2 & -1 & 0 & -1\end{pmatrix} \] The characteristic equation corresponding to $B=JA$ is $\det(JA-XI_4)=X^{3}(X-5)=0$. To the eigenvalue $\lambda_1=0$ corresponds the eigenspace $$E_1=\operatorname{span}\{e_1,e_2\} $$ and the root subspace $$ F_1=\operatorname{span}\{e_1,e_2,e_3\} $$ where $e_1=(1,2,0,0)$, $e_2=(1,1,1,1)$, $e_3=(0,0,0,1)$ with $Be_3=e_2$. To the eigenvalue $\lambda_2=5$ corresponds the root subspace $$ E_2=F_2=\operatorname{span}\{e_4\}, $$ where $e_4=(0,0,1,0)$. Then we have $JA=SQS^{-1}$ with \[ S= \begin{pmatrix} 1 & 1 & 0 & 0 \\ 2 & 1 & 0 & 1 \\ 0 & 1 & 0 & 1\\ 0 & 1 & 1 & 0\end{pmatrix}, \quad Q= \begin{pmatrix} 0 & 0 & 0 & 0 \\ 0 & 0 & 1 & 0 \\ 0 & 0 & 0 & 0\\ 0 & 0 & 0 & 5\end{pmatrix} \] The matrizant of the corresponding equation \eqref{e3.1} is then \[ R(t)=SQS^{-1}= S\begin{pmatrix} 1 & 0 & 0 & 0 \\ 0 & 1 & t & 0 \\ 0 & 0 & 1 & 0\\ 0 & 0 & 0 & e^{5t}\end{pmatrix} S^{-1}. \] To the basis $(e_1,e_2,e_3,e_4)$ corresponds the group of $4$-linearly independent solutions \begin{equation} \begin{gathered} u_1(t)=e_1\\ u_2(t)=e_2\\ u_3(t)=te_2+e_3\\ u_4(t)=e^{5t}e_4. \end{gathered} \label{e3.5} \end{equation} A solution of equation \eqref{e3.1} takes the form $$ u(t)=\alpha_1u_1(t)+\alpha_2u_2(t)+\alpha_3u_3(t) +\alpha_4u_4(t) $$ and it is easy to verify that $u$ is $T$-periodic for $T>0$ if and only if $\alpha_3=\alpha_4=0$, i.e. $u\in N(A)$. \end{example} \section{First class of sub-linear Hamiltonian systems} Consider the first-order Hamiltonian system \begin{equation} J\dot{u}(t)+Au(t)+\nabla H(t,u(t))=h(t) \label{eHcal} \end{equation} where $A$ is a $(2N\times 2N)$ symmetric matrix, $H:\mathbb{R}\times\mathbb{R}^{2N}\to\mathbb{R}$ is a continuous function, $T$-periodic in the first variable $(T>0)$ and differentiable with respect to the second variable with continuous derivative $\nabla H(t,x)=\frac{\partial H}{\partial x}(t,x)$, $h\in C(\mathbb{R},\mathbb{R}^{2N})$ is $T$-periodic and $J$ is the standard symplectic matrix $J=\left(\begin{smallmatrix} 0 & -I\\ I & 0 \end{smallmatrix}\right)$. Let $\gamma:\mathbb{R}^{+}\to\mathbb{R}^{+}$ be a nondecreasing continuous function satisfying the properties: \begin{itemize} \item[(i)] $\gamma(s+t)\leq c(\gamma(s)+\gamma(t))$ for all $s,t\in\mathbb{R}^{+}$, \item[(ii)] $\gamma(t)\leq at^{\alpha}+b$ for all $t\in\mathbb{R}^{+}$, \item[(iii)] $\gamma(t)\to +\infty$ as $t\to+\infty$, \end{itemize} where $a,b,c$ are positive constants and $\alpha\in[0,1[$. Consider the following assumptions \begin{itemize} \item[(C1)] $\dim(N(A))=m\geq 1$ and $A$ has no eigenvalue of the form $ki\frac{2\pi}{T}$ $(k\in \mathbb{N}^{*})$; \item[(H1)] There exist two functions $p\in L^{\frac{2}{1-\alpha}}(0,T;\mathbb{R}^{+})$ and $q\in L^2(0,T;\mathbb{R}^{+})$ such that $$ |\nabla H(t,x)|\leq p(t)\gamma(|x|)+q(t),\quad \forall x\in\mathbb{R}^{2N},\text{ a.e. } t\in[0,T]. $$ \end{itemize} Our main results in this section are the following theorems. \begin{theorem} \label{thm4.1} Assume {\rm (C1)} and {\rm (H1)} hold and \begin{itemize} \item[(H2)] $H$ satisfies \[ \limsup_{|x|\to\infty, x\in N(A)}\frac{|x|}{\gamma^2(|x|)}<+\infty,\quad \lim_{|x|\to\infty,x\in N(A)}\frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt =+\infty. \] \end{itemize} Then \eqref{eHcal} possesses at least one $T$-periodic solution. \end{theorem} \begin{example} \label{examp4.1} \rm Let $A$ be the matrix defined in Example \ref{examp3.1} and let $$ H(t,x)=(\frac{3}{4}T-t)|x|^{8/5},\quad \forall x\in\mathbb{R}^{2N},\; \forall t\in [0,T]. $$ Then $$ |\nabla H(t,x)|=\frac{8}{5}|\frac{3}{4}T-t||x|^{3/5}. $$ Let $\gamma(t)=t^{3/5}, t\geq 0$. It is clear that properties (i), (ii), (iii) are satisfied. Moreover, we have \begin{gather*} \limsup_{|x|\to\infty, x\in N(A)}\frac{|x|}{\gamma^2(|x|)} = \limsup_{|x|\to\infty, x\in N(A)}\frac{|x|}{|x|^{\frac{6}{5}}}=0<+\infty,\\ \lim_{|x|\to\infty, x\in N(A)} \frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt =\lim_{|x|\to\infty, x\in N(A)}\frac{\frac{1}{4}T^2|x|^{8/5}}{|x|^{\frac{6}{5}}} =+\infty \end{gather*} Hence, by Theorem \ref{thm4.1}, the corresponding system \eqref{eHcal} possesses at least one $T$-periodic solution. \end{example} \begin{theorem} \label{thm4.2} Assume {\rm (C1)} and {\rm (H1)} hold and \begin{itemize} \item[(H3)] $H$ satisfies \[ \limsup_{|x|\to\infty, x\in N(A)}\frac{\gamma^2(|x|)}{|x|}<\infty,\quad \lim_{|x|\to\infty}\frac{1}{|x|}\int^T_0H(t,x)dt=+\infty. \] \end{itemize} Then \eqref{eHcal} possesses at least one $T$-periodic solution. \end{theorem} \begin{theorem} \label{thm4.3} Assume {\rm (C1)} and {\rm (H1)} hold and \begin{itemize} \item[(H4)] $H$ satisfies \[ \limsup_{|x|\to\infty, x\in N(A)}\frac{\gamma^2(|x|)}{|x|}=0,\quad \lim_{|x|\to\infty}\frac{1}{|x|}\int^T_0H(t,x)dt>\int^T_0|h(t)|dt. \] \end{itemize} Then \eqref{eHcal} possesses at least one $T$-periodic solution. \end{theorem} \begin{example} \label{examp4.2} \rm Let $A$ be the matrix defined in Example \ref{examp3.1} and let $$ H(t,x)=(\frac{1}{2}T-t)ln^{\frac{3}{2}}(1+|x|^2)+\frac{l(t)|x|^{3}}{1+|x|^2}, \quad \forall x\in\mathbb{R}^{2N},\; \forall t\in [0,T], $$ where $l\in C([0,T],\mathbb{R}^{+})$ with $\int^T_0l(t)dt>\int^T_0|h(t)|dt$. Then \begin{align*} |\nabla H(t,x)| &\leq\frac{3}{2}|\frac{1}{2}T-t|\big(\ln(1+|x|^2)\big)^{1/2} \frac{|x|}{1+|x|^2}+\frac{l(t)(5|x|^{4})+3|x|^2)}{1+2|x|^2+|x|^{4}}\\ &\leq \frac{3}{2}|\frac{1}{2}T-t|\big(\ln(1+|x|^2)\big)^{1/2} \frac{|x|}{1+|x|^2}+c_1 \end{align*} where $c_1$ is a positive constant. Let $\gamma(t)=\big(\ln(1+|t|^2)\big)^{1/2}$, $t\geq 0$. It is clear that conditions (i), (ii), (iii) are satisfied. Moreover, \begin{gather*} \limsup_{|x|\to\infty, x\in N(A)}\frac{\gamma^2(|x|)}{|x|} = \limsup_{|x|\to\infty, x\in N(A)}\frac{ln(1+|x|^2)}{|x|}=0<+\infty, \\ \lim_{|x|\to\infty, x\in N(A)}\frac{1}{|x|}\int^T_0H(t,x)dt =\int^T_0l(t)dt>\int^T_0|h(t)|dt. \end{gather*} Hence, by Theorem \ref{thm4.3}, the corresponding system \eqref{eHcal} possesses at least one $T$-periodic solution. \end{example} \begin{theorem} \label{thm4.4} Assume {\rm (C1)} and {\rm (H1)} hold and \begin{itemize} \item[(H5)] $H$ satisfies \[ \int^T_0h(t)dt\bot N(A),\ \lim_{|x|\to\infty, x\in N(A)} \frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt=+\infty. \] \end{itemize} Then \eqref{eHcal} possesses at least one $T$-periodic solution. \end{theorem} Theorem \ref{thm4.4} generalizes the result concerning the existence of periodic solutions for \eqref{eHcal} in \cite[Theorem 3.1]{c1}. \begin{example} \label{examp4.3}\rm Let $A$ be the matrix defined in Example \ref{examp3.1} and let $$ H(t,x)=(\frac{3}{4}T-t)ln^{\frac{3}{2}}(1+|x|^2)+l(t)\big(\ln(1+|x|^2)\big)^{1/2}, \quad x\in\mathbb{R}^{2N},\; t\in [0,T], $$ where $l\in C([0,T],\mathbb{R}^{+})$ and $h(t)=c(t)v_1+d(t)v_2$, with $v_1=(2,-1,0,-1), v_2=(0,0,1,-1) \in(N(A))^{\bot}$, $c,d\in C(\mathbb{R},\mathbb{R})$. Then $\int^T_0h(t)dt\bot N(A)$ and $$ |\nabla H(t,x)|\leq\frac{3}{2}|\frac{3}{4}T-t|\big(\ln(1+|x|^2)\big)^{1/2}+ l(t). $$ Let $\gamma(t)=\big(\ln(1+|x|^2)\big)^{1/2}$, $t\geq 0$. It is easy to verify that $\gamma$ satisfies conditions (i), (ii), (iii). Moreover, $$ \lim_{|x|\to\infty, x\in N(A)}\frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt= \lim_{|x|\to\infty, x\in N(A)}\frac{T^2}{4}\big(\ln(1+|x|^2)\big)^{1/2}=+\infty $$ Hence, by Theorem \ref{thm4.4}, the corresponding system \eqref{eHcal} possesses at least one $T$-periodic solution. \end{example} \begin{remark} \label{rmk4.1} \rm Let $u(t)$ be a periodic solution of \eqref{eHcal}, then by replacing $t$ by $-t$ in \eqref{eHcal}, we obtain $$ \dot{u}(-t) = JH'(-t,u(-t)). $$ So it is clear that the function $v(t)=u(-t)$ is a periodic solution of the system $$ \dot{v}(t) = -JH'(-t,v(t)). $$ Moreover, $-H(-t,x)$ satisfies (H2)--(H5) whenever $H(t,x)$ satisfies the following assumptions \begin{itemize} \item[(H2')] \[ \limsup_{|x|\to\infty, x\in N(A)}\frac{|x|}{\gamma^2(|x|)}<+\infty,\quad \lim_{|x|\to\infty,x\in N(A}\frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt=-\infty; \] \item[(H3')] \[ \limsup_{|x|\to\infty, x\in N(A)}\frac{\gamma^2(|x|)}{|x|}<\infty,\quad \lim_{|x|\to\infty}\frac{1}{|x|}\int^T_0H(t,x)dt=-\infty; \] \item[(H4')] \[ \limsup_{|x|\to\infty, x\in N(A)}\frac{\gamma^2(|x|)}{|x|}=0,\quad \lim_{|x|\to\infty}\frac{1}{|x|}\int^T_0H(t,x)dt<-\int^T_0|h(t)|dt; \] \item[(H5')] \[ \int^T_0h(t)dt\bot N(A),\ \lim_{|x|\to\infty, x\in N(A)} \frac{1}{\gamma^2(|x|)}\int^T_0H(t,x)dt=-\infty. \] \end{itemize} Consequently, the previous Theorems remains true if we replace (H2)--(H5) by (H2')--(H5'). \end{remark} \subsection*{Proofs of Theorems} Consider the functional $$ \varphi(u)=\frac{1}{2}\int^T_0(J\dot{u}(t)\cdot u(t)+Au(t)\cdot u(t))dt +\int^T_0H(t,u(t))]\,dt -\int^T_0h(t)\cdot u(t))\,dt $$ Let $E$ be the space introduced in Section 2. By assumption (H1) and the property (ii) of $\gamma$, \cite[Proposition B37]{r1} implies that $\varphi\in C^{1}(E,\mathbb{R})$ and the critical points of $\varphi$ on $E$ correspond to the $T$-periodic solutions of \eqref{eHcal}, moreover $$ \varphi'(u)v=\int^T_0[J\dot{u}(t)+Au(t)+\nabla H(t,u(t))]\cdot v(t)\,dt -\int^T_0h(t)\cdot v(t)\,dt. $$ \begin{lemma} \label{lem4.1} Assume {\rm (H1)} holds. Then for any $(PS)$ sequence $(u_{n})\subset E$ of the functional $\varphi$, there exists a constant $c_0>0$ such that \begin{equation} \|\tilde{u}_{n}\|\leq c_0(\gamma(\|u^{0}_{n}\|)+1),\quad \forall n\in\mathbb{N}\label{e4.1} \end{equation} where $\tilde{u}_{n}=u^{+}_{n}+u^{-}_{n}=u_{n}-u^{0}_{n}$, with $u^{0}_{n}\in E^{0}$, $u^{-}_{n}\in E^{-}$, $u^{+}_{n}\in E^{+}$. \end{lemma} \begin{proof} Let $(u_{n})_{n\in\mathbb{N}}$ be a $(PS)$ sequence, i.e. $\varphi(u_{n})$ is bounded and $\varphi'(u_{n})\to 0$ as $n\to\infty$. We have $$ \varphi'(u_{n})(u^{+}_{n}-u^{-}_{n})=2\|\tilde{u}_{n}\|^2 +\int^T_0\nabla H(t,u_{n})\cdot (u^{+}_{n}-u^{-}_{n})dt -\int^T_0h(t)\cdot(u^{+}_{n}-u^{-}_{n})dt. $$ Since $\varphi'(u_{n})\to 0$ as $n\to\infty$, there exists a constant $c_2>0$ such that \[ \big|\varphi'(u_{n})(u^{+}_{n}-u^{-}_{n})\big| \\ \leq c_2\|\tilde{u}_{n}\|,\ \forall n\in\mathbb{N}. \] By H\"older's inequality and (H1), we have \begin{equation} \begin{aligned} \big|\int^T_0\nabla H(t,u_{n})\cdot(u^{+}_{n}-u^{-}_{n})dt\big| &\leq \|\tilde{u}_{n}\|_{L^2}(\int^T_0|\nabla H(t,u_{n})|^2dt)^{1/2}\\ &\leq\|\tilde{u}_{n}\|_{L^2}(\int^T_0[p(t)\gamma(|u_{n}|)+q(t)]dt)^{1/2} \\ &\leq\|\tilde{u}_{n}\|_{L^2}\big[(\int^T_0p^2(t)\gamma^2(|u_{n}|)dt)^{1/2} +\|q\|_{L^2}\big]. \end{aligned}\label{e4.2} \end{equation} Now, by nondecreasing condition and the properties (i) and (ii) of $\gamma$, we have \begin{align*} \Big(\int^T_0p^2(t)\gamma^2(|u_{n}|)dt\Big)^{1/2} &\leq\Big(\int^T_0p^2(t)\gamma^2(|\tilde{u}_{n}|+|u^{0}_{n}|)dt\Big)^{1/2}\\ &\leq c\Big(\int^T_0 [p^2(t)[\gamma(|\tilde{u}_{n}|) +\gamma(|u^{0}_{n}|)]^2dt\Big)^{1/2}\\ &\leq c\Big[\Big(\int^T_0p^2(t)\gamma^2(|\tilde{u}_{n}|)dt\Big)^{1/2} +\|p\|_{L^2}\gamma(|u^{0}_{n}|\Big)\Big]\\ &\leq c\Big[\Big(\int^T_0p^2(t)(a|\tilde{u}_{n}|^{\alpha}+b)^2dt\Big)^{1/2} + \|p\|_{L^2} \gamma(|u^{0}_{n}|)\Big]\\ &\leq c\Big[a\Big(\int^T_0p^2(t)|\tilde{u}_{n}|^{2\alpha}dt\Big)^{1/2} + b\|p\|_{L^2}+ \|p\|_{L^2} \gamma(|u^{0}_{n}|)\Big]\\ &\leq c\big[a\|p\|_{L^{\frac{2}{1-\alpha}}}\|\tilde{u}_{n}\|^{\alpha}_{L^2} + b\|p\|_{L^2}+\|p\|_{L^2}\gamma(|u^{0}_{n}|)\big]. \end{align*} On the other hand, by \eqref{e2.1} we have \begin{align*} 2\|\tilde{u}_{n}\|^2 &\leq|\varphi'(u_{n})(u^{+}_{n}-u^{-}_{n})| +|\int^T_0\nabla H(t,u_{n})\cdot(u^{+}_{n}-u^{-}_{n})dt|\\ &\quad +|\int^T_0h(t)\cdot(u^{+}_{n}-u^{-}_{n})| \leq c_2\|\tilde{u}_{n}\|+\|\tilde{u}_{n}\|_{L^2} c\big[a\|p\|_{L^{\frac{2}{1-\alpha}}}\|\tilde{u}_{n}\|^{\alpha}_{L^2}\\ &\quad + b\|p\|_{L^2} +\|q\|_{L^2}+\|p\|_{L^2}\gamma(|u^{0}_{n}|)\big]+\|h\|_{L^2}\|\tilde{u}_{n}\|_{L^2}\\ &\leq ac\|p\|_{L^{\frac{2}{1-\alpha}}}\lambda^{\alpha+1}_2 \|\tilde{u}_{n}\|^{\alpha+1}+ [c_1+cb\|p\|_{L^2}\\ &\quad +\|q\|_{L^2}+\|h\|_{L^2}]\lambda_2\|\tilde{u}_{n}\| +c\lambda_2\|p\|_{L^2}\gamma(|u^{0}_{n}|)\|\tilde{u}_{n}\|. \end{align*} Since $0\leq\alpha<1$, we deduce that there exists a constant $c_0>0$ satisfying \eqref{e4.1}. \end{proof} We will apply Lemma \ref{lem2.1} to the functional $\varphi$ to obtain critical points. \begin{lemma} \label{lem4.2} If {\rm (H1)} holds and $H$ satisfies one of the assumptions {\rm (H2)--(H5)}, then $\varphi$ satisfies the $(PS)_{c}$ condition for all $c\in\mathbb{R}$. \end{lemma} \begin{proof} Let $(u_{n})_{n\in\mathbb{N}}$ be a $(PS)_{c}$ sequence, that is $\varphi(u_{n})\to c$ and $\varphi'(u_{n}) \to 0$ as $n\to\infty$. Then there exists a positive constant $c_3$ such that $$ |\varphi(u_{n})|\leq c_3,\quad \|\varphi'(u_{n})\|\leq c_3,\quad \forall n\in\mathbb{N}. $$ By the Mean Value Theorem and H\"older's inequality, we have \begin{equation} \begin{aligned} &\big|\int^T_0(H(t,u_{n})-H(t,u^{0}_{n}))dt\big|\\ &=\big|\int^T_0\int^{1}_0\nabla H(t,u^{0}_{n}+s\tilde{u}_{n})\cdot \tilde{u}_{n}\,ds\,dt\big|\\ &\leq\|\tilde{u}_{n}\|_{L^2}\int^{1}_0\Big(\int^T_0| \nabla H(t,u^{0}_{n} +s\tilde{u}_{n})|^2dt\Big)^{1/2}ds. \end{aligned} \label{e4.3} \end{equation} As in the proof of Lemma \ref{lem4.1}, we have \begin{equation} \begin{aligned} &\Big(\int^T_0| \nabla H(t,u^{0}_{n}+s\tilde{u}_{n})|^2dt\Big)^{1/2}\\ &\leq ac\|p\|_{L^{\frac{2}{1-\alpha}}}\|\tilde{u}_{n}\|^{\alpha}_{L^2} + cb\|p\|_{L^2}+\|q\|_{L^2}+c\|p\|_{L^2}\gamma(|u^{0}_{n}|)\|q\|_{L^2}\big]. \end{aligned} \label{e4.4} \end{equation} Therefore, by properties \eqref{e2.1}, \eqref{e4.1}, \eqref{e4.3}, \eqref{e4.4} and since $0\leq\alpha<1$, there exists a positive constant $c_4$ such that \begin{equation} \begin{aligned} \big|\int^T_0(H(t,u_{n})-H(t,u^{0}_{n}))dt\big| &\leq c_0(\gamma(|u^{0}_{n}|)+1)[ac\|p\|_{L^{\frac{2}{1-\alpha}}} c^{\alpha}_0(\gamma(|u^{0}_{n}|)+1)^{\alpha}\\ &\quad +c\|p\|_{L^2}\gamma(|u^{0}_{n}|)+cb\|p\|_{L^2}+\|q\|_{L^2}]\\ &\leq c_4(\gamma^2(|u^{0}_{n}|)+1). \end{aligned} \label{e4.5} \end{equation} Combining \eqref{e2.1}, \eqref{e4.1}, \eqref{e4.2} and \eqref{e4.5} yields \begin{equation} \begin{aligned} c_3&\geq \varphi(u_{n})\\ &\geq -\|\tilde{u}_{n}\|^2+\int^T_0(H(t,u_{n})-H(t,u^{0}_{n}))dt +\int^T_0H(t,u^{0}_{n})dt \\ &\quad -\int^T_0h(t)(\tilde{u}_{n}+u^{0}_{n})dt\\ &\geq -c^2_0(\gamma(|u^{0}_{n})|+1)^2-c_4(\gamma^2(|u^{0}_{n})|)+1) -c_0\|h\|_{L^2}(\gamma(|u^{0}_{n})|+1) \\ &\quad -\|h\|_{L^{1}}|u^{0}_{n}|+ \int^T_0H(t,u^{0}_{n})dt\\ &\geq -c_5(\gamma^2(|u^{0}_{n})|)+1)-\|h\|_{L^{1}}|u^{0}_{n}| +\int^T_0H(t,u^{0}_{n})dt, \end{aligned} \label{e4.6} \end{equation} where $c_5$ is a positive constant. \smallskip \noindent\textbf{Case 1:} $H$ satisfies (H2). By \eqref{e4.6}, we have $$ c_3\geq \gamma^2(|u^{0}_{n})|)[-c_5-\|h\|_{L^{1}} \frac{|u^{0}_{n}|}{\gamma^2(|u^{0}_{n})|)} +\frac{1}{\gamma^2(|u^{0}_{n})|)}\int^T_0H(t,u^{0}_{n})dt]-c_5. $$ It follows from (H2) that $(u^{0}_{n})$ is bounded. \smallskip \noindent\textbf{Case 2:} $H$ satisfies (H3) or (H4). Note that by \eqref{e4.6} $$ c_3\geq |u^{0}_{n}|[-c_5\frac{\gamma^2(|u^{0}_{n})|)}{|u^{0}_{n})|} -\|h\|_{L^{1}}+\frac{1}{|u^{0}_{n}|}\int^T_0H(t,u^{0}_{n})dt]-c_5. $$ Hence (H3) or (H4) implies that $(u^{0}_{n})$ is bounded. \smallskip \noindent\textbf{Case 2:} $H$ satisfies (H5). Since $\int^T_0h(t)dt\bot N(A)$, we get as in \eqref{e4.6} \begin{equation} \begin{aligned} c_3&\geq \varphi(u_{n})\\ &\geq -\|\tilde{u}_{n}\|^2 +\int^T_0(H(t,u_{n})-H(t,u^{0}_{n}))dt +\int^T_0H(t,u^{0}_{n})dt\\ &\quad -\int^T_0h(t)\cdot\tilde{u}_{n}dt\\ &\geq -c_5(\gamma^2(|u^{0}_{n})|)+1)+\int^T_0H(t,u^{0}_{n})dt, \\ &\geq \gamma^2(|u^{0}_{n})|)[-c_5 +\frac{1}{\gamma^2(|u^{0}_{n})|)}\int^T_0H(t,u^{0}_{n})dt]-c_5. \end{aligned} \label{e4.6b} \end{equation} Hence (H5) implies that $(u^{0}_{n})$ is bounded. In all the above cases, $(u^{0}_{n})$ is bounded. We deduce from Lemma \ref{lem4.1} that $(u_{n})$ is also bounded in $E$. By a standard argument, we conclude that $(u_{n})$ possesses a convergent subsequence. The proof of Lemma \ref{lem4.2} is complete. \end{proof} Now, decompose $E=E^{-}\oplus (E^{0}\oplus E^{+})$ and let $E^{1}=E^{-}$ and $E^2=E^{0}\oplus E^{+}$. Remark that by Section 3, we have $E^{0}=N(A)$. We will verify that $\varphi$ satisfies condition c) of Lemma \ref{lem2.1}. For $u\in E^{1}$, we have $$ \varphi(u)= -\|u\|^2+\int^T_0(H(t,u)-H(t,0))dt +\int^T_0H(t,0)dt -\int^T_0h(t)\cdot u dt. $$ As in the proof of Lemma \ref{lem4.2}, $$ \big|\int^T_0(H(t,u)-H(t,0))dt\big| \leq [ac\|p\|_{L^{\frac{2}{1-\alpha}}}\|u\|^{\alpha}_{L^2} +cb\|p\|_{L^2}+\|q\|_{L^2}]\|u\|_{L^2}. $$ Hence by \eqref{e2.1}, we deduce \begin{equation} \begin{aligned} \varphi(u) &\leq -\|u\|^2 +ac\|p\|_{L^{\frac{2}{1-\alpha}}}\lambda^{\alpha+1}_2 \|u\|^{\alpha+1} +(cb\|p\|_{L^2}\\ &\quad +\|q\|_{L^2}+\|h\|_{L^2})\lambda_2\|u\|+\int^T_0H(t,0)dt. \end{aligned}\label{e4.7} \end{equation} Since $0\leq\alpha< 1$, \eqref{e4.7} implies that $\varphi(u)\to-\infty$ as $\|u\|\to\infty$. Hence there exists $\beta\in\mathbb{R}$ such that $f(u)\leq\beta$ for all $u\in E^{1}$. Condition (c) of Lemma \ref{lem2.1} is then proved. Let us verify that $\varphi$ satisfies condition (d) of Lemma \ref{lem2.1}. In fact, for $u\in E^2=E^{0}\oplus E^{+}$, as in the proof of Lemma \ref{lem4.2}, we have \begin{equation} \begin{aligned} &\big|\int^T_0(H(t,u)-H(t,u^{0}))dt\big|\\ &\leq\big[ac\|p\|_{L^{\frac{2}{1-\alpha}}}\|u^{+}\|^{\alpha}_{L^2} + cb\|p\|_{L^2}+c\|p\|_{L^2}\gamma(|u^{0}|)+\|q\|_{L^2}\big]\|u^{+}\|^{\alpha}_{L^2}. \end{aligned} \label{e4.8} \end{equation} From \eqref{e2.1} and \eqref{e4.8}, we deduce that \begin{equation} \begin{aligned} \varphi(u) &\geq\|u^{+}\|^2-ac\|p\|_{L^{\frac{2}{1-\alpha}}}\lambda^{\alpha+1}_2\|u^{+} \|^{\alpha+1}-c\|p\|_{L^2}\lambda_2\|u^{+}\|\gamma(|u^{0}|)\\ &\quad -(cb\|p\|_{L^2}+\|q\|_{L^2}+\|h\|_{L^2})\lambda_2\|u^{+}\| -\int^T_0|h|dt|u^{0}|+\int^T_0H(t,u^{0})dt. \end{aligned} \label{e4.9} \end{equation} For $\epsilon>0$, there exists a constant $C(\epsilon)$ such that $$ c\|p\|_{L^2}\lambda_2\|u^{+}\|\gamma(|u^{0}|) \leq\epsilon\|u^{+}\|^2+C(\epsilon)\gamma^2(|u^{0}|). $$ Taking $\epsilon=1/2$, it follows from \eqref{e4.9} that \begin{equation} \begin{aligned} \varphi(u) &\geq\frac{1}{2}\|u^{+}\|^2-ac\|p\|_{L^{\frac{2}{1-\alpha}}} \lambda^{\alpha+1}_2\|u^{+}\|^{\alpha+1} -\lambda_2(cb\|p\|_{L^2}+\|q\|_{L^2}\\ &\quad +\|h\|_{L^2})\|u^{+}\| -C(\frac{1}{2})\gamma^2(|u^{0}|) -\int^T_0|h|dt|u^{0}|+\int^T_0H(t,u^{0})dt. \end{aligned} \label{e4.10} \end{equation} Since $0\leq\alpha<1$, the term $$ \frac{1}{2}\|u^{+}\|^2-ac\|p\|_{L^{\frac{2}{1-\alpha}}}\lambda^{\alpha+1}_2 \|u^{+}\|^{\alpha+1} -\lambda_2(cb\|p\|_{L^2}+\|q\|_{L^2}+\|h\|_{L^2})\|u^{+}\| $$ approaches $+\infty$ as $\|u^{+}\|\to\infty$. It remains to study the following member of \eqref{e4.10} $$ \psi(u^{0})=-C(\frac{1}{2})\gamma^2(|u^{0}|) -\int^T_0|h|dt|u^{0}|+\int^T_0H(t,u^{0})dt. $$ \smallskip \noindent\textbf{Case 1:} (H2) holds. We have $$ \psi(u^{0})\geq\gamma^2(|u^{0}|)\big(-C(\frac{1}{2}) -\int^T_0|h|dt\frac{|u^{0}|}{\gamma^2(|u^{0}|)} + \frac{1}{\gamma^2(|u^{0}|)}\int^T_0H(t,u^{0})dt\big). $$ It follows from (H2) that $\psi(u^{0})\to +\infty$ as $|u^{0}|\to \infty$. \smallskip \noindent\textbf{Case 2:} (H3) or (H4) holds. We have $$ \psi(u^{0})\geq|u^{0}|\big(-C(\frac{1}{2})\frac{\gamma^2(|u^{0}|)}{|u^{0}|} -\int^T_0|h|dt+\frac{1}{|u^{0}|}\int^T_0H(t,u^{0})dt\big). $$ It follows from (H3) or (H4) that $\psi(u^{0})\to +\infty$ as $|u^{0}|\to \infty$. \smallskip \noindent\textbf{Case 3:} (H5) holds. Since $\int^T_0h(t)dt\bot N(A)$, we have $$ \psi(u^{0})\geq\gamma^2(|u^{0}|)\big(-C(\frac{1}{2}) + \frac{1}{\gamma^2(|u^{0}|)}\int^T_0H(t,u^{0})dt\big). $$ It follows from (H5) that $\psi(u^{0})\to +\infty$ as $|u^{0}|\to \infty$. Therefore, if one of assumptions (H2)--(H5) is satisfied, then $\varphi(u)\to +\infty$ as $\|u\|\to \infty$. So there exists a constant $\rho$ such that $\varphi(u)\geq\rho$ for all $u\in E^2$. Condition d) of Lemma \ref{lem2.1} is satisfied. Moreover, it is well known that the derivative of the functional $d(u)=\int^T_0H(t,u)dt-\int^T_0hu dt$ is compact. All the conditions of Lemma \ref{lem2.1} are satisfied, so $\varphi$ possesses a critical point $u$ which is a $T$-periodic solution of system \eqref{eHcal} \section{Second class of Hamiltonian systems} For $A, H$ and $h$ be defined as in Section 4, we have the following result. \begin{theorem} \label{thm5.1} Let $\omega\in C(\mathbb{R}^{+},\mathbb{R}^{+})$ be a non-increasing function with the following properties: \begin{itemize} \item[(a)] $\liminf_{s\to\infty}\frac{\omega(s)}{\omega(\sqrt{s})}>0$, \item[(b)] $\omega(s)\to 0$ and $\omega(s)s\to+\infty$ as $ s\to\infty$. \end{itemize} Assume that $A$ satisfies {\rm (C1)}, and $H$ satisfies \begin{itemize} \item[(H6)] There exist a positive constant $a$ and a function $g\in L^2(0,T;\mathbb{R})$ such that $$ |\nabla H(t,x)|\leq a\omega(|x|)+g(t),\quad \forall x\in\mathbb{R}^{2N}, \text{ a.e. } t\in[0,T]; $$ \item[(H7)] $$ \lim_{|x|\to\infty,x\in N(A)}\frac{1}{(\omega(|x|)|x|)^2}\int^T_0H(t,x)dt =+\infty; $$ \item[(H8)] There exists $f\in L^{1}(0,T;\mathbb{R})$ such that $$ H(t,x)\geq f(t),\quad \forall x\in\mathbb{R}^{2N},\text{ a.e. } t\in [0,T]. $$ \end{itemize} Then system \eqref{eHcal} possesses at least one $T$-periodic solution. \end{theorem} The above theorem generalizes \cite[Theorem 1.1]{t3}. \begin{example} \label{examp5.1}\rm Take $\omega(s)=\frac{1}{\ln(2+s^2)},\ s\geq 0,$ $$ H(t,x)=(\frac{1}{2}+\cos(\frac{2\pi}{T}t))\frac{|x|^2}{\ln(2+|x|^2)},\quad \forall t\in[0,T],\; \forall x\in\mathbb{R}^{2N} $$ and let $A$ be the matrix defined in Section 3, $h\in C([0,T],\mathbb{R})$. Then $A, H, h$ satisfy assumptions of Theorem \ref{thm5.1}. \end{example} \subsection*{Proof of Theorem \ref{thm5.1}} As in Section 4, we will apply Lemma \ref{lem2.1} to the functional $\varphi$ defined on the space $E$ introduced in section 2. \begin{lemma}[\cite{t3}] \label{lem5.1} Assume {\rm (H6)} and {\rm (H7)} hold, then there exists a non-increasing function $\theta\in C(]0,+\infty[,\mathbb{R}^{+})$ and a positive constant $c_0$ such that \begin{itemize} \item[(i)] $\theta(s)\to 0$ and $\theta(s)s\to\infty$ as $s\to\infty$, \item[(ii)] $\|\nabla H(t,u)\|_{L^2}\leq c_0(\theta(\|u\|)\|u\|+1)$ for all $u\in E$ \item[(iii)] \[ \frac{1}{(\theta(\|u^{0}\|)\|u^{0}\|)^2}\int^T_0H(t,u^{0})dt\to+\infty \quad\text{as } \|u^{0}\|\to\infty. \] \end{itemize} \end{lemma} \begin{lemma} \label{lem5.2} Assume {\rm (H6)} holds. Then for any $(PS)$ sequence of the functional $\varphi$, there exists a constant $c_1>0$ such that \begin{equation} \|\tilde{u}_{n}\|\leq c_1\big(\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\|+1\big). \label{e5.1} \end{equation} \end{lemma} \begin{proof} Let $(u_{n})$ be a Palais-Smale sequence, that is $(\varphi(u_{n}))$ is bonded and $\varphi'(u_{n})\to 0, as\ n\to\infty$. We have $$ \varphi'(u_{n})(u^{+}_{n}-u^{-}_{n})=2\|\tilde{u}_{n}\|^2 +\int^T_0\nabla H(t,u_{n}(t))\cdot (u^{+}_{n}-u^{-}_{n})dt -\int^T_0h(t)\cdot(u^{+}_{n}-u^{-}_{n})dt. $$ Since $\theta$ is non-increasing and $\|u\|\geq max(\|\tilde{u}\|,\|u^{0}\|)$, we have \begin{equation} \theta(\|u\|)\leq \min(\theta(\|\tilde{u}\|),\theta(\|u^{0}\|)). \label{e5.2} \end{equation} By H\"older's inequality, inequalities \eqref{e2.1}, \eqref{e5.1}, \eqref{e5.2} and Lemma \ref{lem5.1}, we have \begin{align*} &\big|\int^T_0\nabla H(t,u_{n}(t))\cdot(u^{+}_{n}-u^{-}_{n})dt\big| \\ &\leq\|u^{+}_{n}-u^{-}_{n}\|_{L^2}(\int^T_0|\nabla H(t,u_{n})|^2dt)^{1/2}\\ &\leq c_2\|\tilde{u}_{n}\|(\theta(\|u_{n}\|)\|u_{n}\|+1) \\ &\leq c_2\|\tilde{u}_{n}\|\big(\theta(\|\tilde{u}_{n}\|) \|\tilde{u}_{n}\|+\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\|+1\big). \end{align*} Thus there exists positive constants $c_3, c_4$ such that \begin{align*} c_3\|\tilde{u}_{n}\| &\geq\varphi'(u_{n})(u^{+}_{n}-u^{-}_{n})\\ &\geq 2\|\tilde{u}_{n}\|^2-c_2\|\tilde{u}_{n}\| \big(\theta(\|\tilde{u}_{n}\|)\|\tilde{u}_{n}\| +\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\|+1\big)-c_4\|\tilde{u}_{n}\|. \end{align*} Hence $$ c_2\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\| \geq\|\tilde{u}_{n}\|[2-c_2\|\tilde{u}_{n}\|]-c_3-c_4. $$ Since $\theta(s)\to 0$ as $s\to\infty$, this implies the existence of a constant $c_1$ satisfying \eqref{e5.1}. \end{proof} \begin{lemma} \label{lem5.3} $\varphi$ satisfies the $(PS)_{c}$ condition for all real $c$. \end{lemma} \begin{proof} Let $(u_{n})$ be a $(PS)_{c}$-sequence. Assume that $(u^{0}_{n})$ is unbounded. Going to a subsequence if necessary, we can assume that $\|u^{0}_{n}\|\to\infty$ as $n\to\infty$. By the Mean Value Theorem, H\"older's inequality, inequality \eqref{e2.1} and Lemma \ref{lem5.1} (ii), there exists a positive constant $c_5$ such that \begin{equation} \begin{aligned} &\big|\int^T_0(H(t,u_{n})-H(t,u^{0}_{n}))dt\big|\\ &=\big|\int^T_0\int^{1}_0\nabla H(t,u^{0}_{n} +s\tilde{u}_{n})\cdot\tilde{u}_{n}\,ds\,dt\big|\\ &\leq \|\tilde{u}_{n}\|_{L^2}\int^{1}_0 \Big(\int^T_0|\nabla H(t,u^{0}_{n}+s\tilde{u}_{n})dt |\Big)^{1/2}ds\\ &\leq c_5\|\tilde{u}_{n}\|\big[\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\| + \theta(\|u^{0}_{n}\|)\|\tilde{u}_{n}\|+1\big]. \end{aligned} \label{e5.3} \end{equation} Hence by Lemma \ref{lem5.2}, there exists a positive constant $c_{6}$ such that \begin{equation} \big|\int^T_0(H(t,u_{n})-H(t,u^{0}_{n}))dt\big| \leq c_{6}\big([\theta(\|u^{0}_{n}\|)\|\tilde{u}^{0}_{n}\|]^2+1\big). \label{e5.4} \end{equation} Combining \eqref{e2.1}, \eqref{e5.1} and \eqref{e5.4} yields $$ \varphi(u_{n})\geq-c_{7}\big([\theta(\|u^{0}_{n}\|)\|\tilde{u}^{0}_{n}\|]^2+1\big) -\frac{1}{T}\int^T_0|h(t)|dt\|u^{0}_{n}\|+ \int^T_0H(t,u^{0}_{n})dt $$ where $c_{7}$ is a positive constant. On the other hand, it is easy to see that $\lim inf_{s\to\infty}\frac{\theta(s)}{\theta(\sqrt{s})}>0$. So there exists a positive constant $c_{8}$ such that for $s$ large enough $\theta(s)\geq c_{8}\theta(\sqrt{s})$. Hence for $n$ large enough $$ \frac{\|u^{0}_{n}\|}{[\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\|]^2} \geq\frac{1}{c^2_{8}[\theta(\|u^{0}_{n}\|^{1/2})\|u^{0}_{n}\|^{1/2}]^2} \to 0\quad \text{as } n\to\infty\,. $$ Therefore, \begin{align*} \varphi(u_{n}) &\geq [\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\|]^2[-c_{7} -\frac{1}{T}\int^T_0|h(t)|dt\frac{\|u^{0}_{n}\|}{[\theta(\|u^{0}_{n}\|) \|u^{0}_{n}\|]^2} \\ &\quad +\frac{1}{[\theta(\|u^{0}_{n}\|)\|u^{0}_{n}\|]^2} \int^T_0H(t,u^{0}_{n})dt]-c_{7} \to+\infty \end{align*} as $n\to\infty$, which contradicts the boundedness of $(\varphi(u_{n}))$. Hence $(\|u^{0}_{n}\|)$ is bounded, and by Lemma \ref{lem5.2}, $(u_{n})$ is also bounded. By a standard argument, we conclude that $(u_{n})$ possesses a convergent subsequence. The proof is complete. \end{proof} Now, for $u=u^{0}+u^{+}\in E^2=E^{0}\oplus E^{+}$, we have as in \eqref{e5.3}, $$ |\int^T_0(H(t,u)-H(t,u^{0}))dt|\leq c_5\|u^{+}\|\big[\theta(\|u^{0}\|)\|u^{0}\| + \theta(\|u^{0}\|)\|u^{+}\|+1\big]. $$ Since $c_5\theta(\|u^{0}\|)\|u^{0}\|\|u^{+}\| \leq\frac{1}{2}\|u^{+}\|^2+2c^2_5[\theta(\|u^{0}\|)\|u^{0}\|]^2$, we obtain \begin{align*} \varphi(u) &\geq(\frac{1}{2}-c_5\theta(\|u^{0}\|))\|u^{+}\|^2-c_5\|u^{+}\|\\ &\quad +\big[\theta(\|u^{0}\|)\|u^{0}\|\big]^2 \Big(-2c^2_5 -\frac{1}{T}\int^T_0|h(t)|dt \frac{\|u^{0}\|}{[\theta(\|u^{0}\|)\|u^{0}\|]^2} \\ &\quad +\frac{1}{[\theta(\|u^{0}\|)\|u^{0}\|]^2}\int^T_0H(t,u^{0})dt\Big). \end{align*} Since $\theta(s)\to 0$ as $s\to\infty$, there exists $r>0$ such that $c_5\theta(s)\leq\frac{1}{4}$ for $s\geq r$. Then, if $\|u^{0}\|\geq r$, we have \begin{align*} \varphi(u) &\geq\frac{1}{4}\|u^{+}\|^2-c_5\|u^{+}\| +[\theta(\|u^{0}\|)\|u^{0}\|]^2(-2c^2_5 \\ &\quad -\frac{1}{T}\int^T_0|h(t)|dt \frac{\|u^{0}\|}{[\theta(\|u^{0}\|)\|u^{0}\|]^2} +\frac{1}{[\theta(\|u^{0}\|)\|u^{0}\|]^2}\int^T_0H(t,u^{0})dt). \end{align*} then $\varphi(u)\to+\infty$ as $\|u^{0}+u^{+}\|\to\infty$, $\|u^{0}\|\geq r$. If $\|u^{0}\|\leq r$, we have by (H8) and \eqref{e2.1} $$ \varphi(u)\geq\|u^{+}\|^2+\int^T_0f(t)dt -\frac{r}{T}\int^T_0|h(t)|dt-\lambda_2\|h\|_{L^2}\|u^{+}\| $$ then $\varphi(u)\to+\infty$ as $\|u^{0}+u^{+}\|\to\infty$, $\|u^{0}\|\leq r$. Therefore $\varphi(u)\to+\infty$ as $\|u\|\to\infty$, $u\in E^2$. In $E^{1}$, as in \cite{t3}, we obtain $\varphi(u)\to-\infty$ as $\|u\|\to\infty$. Hence, by Lemma \ref{lem2.1}, $\varphi$ possesses at least a critical point $u$ which is a $T$-periodic solution of \eqref{eHcal}. \begin{thebibliography}{00} \bibitem{a1} T. An; \emph{Periodic solutions of superlinear autonomous Hamiltonian systems with prescribed period}, J. Math. Anal. Appl. 323 (2006), pp. 854-863. \bibitem{c1} A. R. Chouikha, M. Timoumi; \emph{Subharmonic solutions for nonautonomous sublinear first order Hamiltonian systems}, Arxiv, 1302-4309 V1 (math.DS) 18, Feb 2013. \bibitem{e1} I. Ekeland; \emph{Periodic solutions of Hamiltonian equations and a theorem of P. Rabinowitz}, J. Diff. Eq. (1979), pp. 523-534. \bibitem{f1} P-L. Felmer; \emph{Periodic solutions of superquadratic Hamiltonian systems}, J. Diff. Eq. 102 (1993), pp. 188-307. \bibitem{g1} M. Gilardi, M. Matzeu; \emph{Periodic solutions of convex autonomous Hamiltonian systems with a quadratic growth at the origin and superquadratic at infinity}, Annali di Matematica Pura and Applicata-(IV), vol. CXL VII, pp 21-72. \bibitem{l1} C. Li, Z-Q. Ou, C-L. Tang; \emph{Periodic and subharmonic solutions for a class of non-autonomous Hamiltonian systems}, Nonlinear Analysis 75 (2012), pp 2262-2272. \bibitem{l2} S. Li, A. Szulkin; \emph{Periodic solutions for a class of non-autonomous Hamiltonian systems}, J. Diff. Eq., (1994), pp. 226-238. \bibitem{l3} S. Li, M. Willem; \emph{Applications of local linking to critical point theory}, J. Math. Anal. Appl. 189, (1995), pp. 6-32. \bibitem{l4} S. Luan, A. Mao; \emph{Periodic solutions for a class of non-autonomous Hamiltonian systems}, Nonlinear Analysis 61, (2005), pp. 1413-1426. \bibitem{m1} J. Mawhin, M. Willem; \emph{Critical point theory and Hamiltonian systems}, Applied Mathematical Sciences, 74, Springer, New York, 1989. \bibitem{r1} P. H. Rabinowitz; \emph{Minimax methods in critical point theory with applications to differential equations}, CBMS. Reg. Conf. Ser. Math., vol 65, Amer. Math. Soc., Providence, RI (1986). \bibitem{r2} P. Rabinowitz; \emph{Periodic solutions of Hamiltonian systems}, Comm. Pure Appl. Math. 31 (1978), pp. 157-184. \bibitem{t1} M. Timoumi; \emph{Periodic and subharmonic solutions for a class of non coercive superqadratic Hamiltonian systems}, Nonl. Dyn. and Syst. Theory, 11(3) (2011), pp. 319-336. \bibitem{t2} M. Timoumi; \emph{Periodic solutions for non coercive super-quadratic Hamiltonian systems}, Dem. Mathematica, Vol. XI, No 2, (2007), pp. 331-346. \bibitem{t3} M. Timoumi; \emph{Subharmonic solutions for first-order Hamiltonian systems}, Elect. J. Diff. Eq., Vol. 2013 (2013), No. 197, pp. 1-14. \bibitem{x1} X. Xu; \emph{Periodic solutions for non-autonomous Hamiltonian systems possessing super-quadratic potentials}, Nonlinear Analysis 51, (2002), pp. 941-955. \end{thebibliography} \end{document}