\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small 2005-Oujda International Conference on Nonlinear Analysis. \newline {\em Electronic Journal of Differential Equations}, Conference 14, 2006, pp. 173--180.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu (login: ftp)} \thanks{\copyright 2006 Texas State University - San Marcos.} \vspace{9mm}} \setcounter{page}{173} \begin{document} \title[\hfilneg EJDE/Conf/14 \hfil A generalization of Ekeland's variational principle] {A generalization of Ekeland's variational principle with applications} \author[A. R. El Amrouss, N. Tsouli \hfil EJDE/Conf/14 \hfilneg] {Abdel R. El Amrouss, Najib Tsouli} % in alphabetical order \address{Abdel R. El Amrouss \newline University Mohamed 1er, Faculty of sciences, Department of Mathematics, Oujda, Morocco} \email{amrouss@sciences.univ-oujda.ac.ma} \address{Najib Tsouli \newline University Mohamed 1er, Faculty of sciences, Department of Mathematics, Oujda, Morocco} \email{tsouli@sciences.univ-oujda.ac.ma} \date{} \thanks{Published September 20, 2006.} \subjclass[2000]{58E05, 35J65, 49B27} \keywords{Ekeland's principle variational; Palais-Smale condition; optimization} \begin{abstract} In this paper, we establish a variant of Ekeland's variational principle. This result suggest to introduce a generalization of the famous Palais-Smale condition. An example is provided showing how it is used to give the existence of minimizer for functions for which the Palais-Smale condition and the one introduced by Cerami are not satisfied. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{remark}[theorem]{Remark} \newtheorem{definition}[theorem]{Definition} \section{Introduction} Let $E$ be a complete metric space with metric $d$ and $\Phi : E\to \mathbb{R} \cup \{\infty\} $ a lower semicontinuous function which is bounded from below and not identically to $+\infty$. The Ekeland's variational principle, see \cite{E1}, allows for each $\varepsilon > 0$, each $\delta > 0$ and each $x \in E$ such as $$ \Phi(x) \leq \inf_E \Phi + \varepsilon, $$ to build an element $v \in E$ minimizing the functional $\Phi_v$ given by $$ \Phi_v(x) = \Phi(x) + \frac{\varepsilon}{\delta}d(x,v). $$ This principle has wide applications in optimization and nonlinear analysis \cite{E1, E2, C}. If $E$ is a Banach space and $\Phi: E \to \mathbb{R}$ is G\^ateaux differentiable, lower semi-continuous and bounded from below, then the Ekeland's variational principle provides the existence of a minimizing sequence $(u_n)$ such as $\Phi'(u_n) \to 0$, when $n \to \infty$. It is well known that if $\Phi$ satisfies the Palais-Smale condition then $\Phi$ reaches its minimum. But, it is possible to find a minimizing sequence $(u_n)$ such as $\Phi'(u_n) \to 0$, when $n \to \infty$, not having any convergent subsequence. Let us take the example of the function $\Phi(s) = \arctan(s)$. Ekeland \cite{E2} prove that if $\Phi$ is bounded below and satisfies the Cerami condition for every $c \in \mathbb{R}$, introduced by \cite{Ce}, then $\Phi$ has a minimal point. In this note, we prove a variant of Ekeland's variational principle. This result suggest to introduce a generalization of the classical Palais-Smale condition. An example is provided showing how it is used to give the existence of minimizer for functions for which the Palais-Smale condition or the Cerami condition are not satisfied. We also generalize some results cited in \cite{E1}, \cite{CS}, which the Palais-Smale condition or Cerami condition has failed. \section{Variants of Ekeland's variational principle} In this section we will prove the following variant of Ekeland's variational principle. We start with a definition. \begin{definition} \rm We say that $\alpha: [0,\infty[ \to ]0,\infty[$ is a comparison function of order $k$ if for every $q \geq k$ there exist $c,d \geq 0$ such that $$ \frac{\alpha((t+1)s)}{\alpha(t)} \leq cs^q + d, \forall t,s \in \mathbb{R}^+. $$ {\bf Examples:} \begin{enumerate} \item $\alpha(s) = (1+s)^k$ \item $\alpha(s) = (1+s)^k Log(2+s)$ \end{enumerate} \end{definition} Let $(E,d)$ be a complet space metric and $u \in E$. Denote by $\bar{B}(u,r) = \{x \in E \mid d(u,x) \leq r\}$ the closed boule and $B(u,r) = \{x \in E \mid d(u,x) < r\}$ the open boule. \begin{theorem} \label{thm2.1} Let $(E,d)$ be a complete space metric, $x_0 \in E$ fixed, $ \Phi: E \to \mathbb{R}$ a lower semi-continuous and bounded below. Let $\alpha: [0,\infty[ \to ]0,\infty[$ be a comparison function of order $k$ continuous nondecreasing. Thus for each $\varepsilon > 0$, each $\delta >0$ and each $u \in E$ such that $$\Phi(u) \leq \inf_E \Phi + \varepsilon$$ there exists a convergent sequence $(z_n)_{n\geq 1}$ of $E$ satisfies: \begin{itemize} \item[(i)] $z_1 = u, z_n \in \bar{B}(u, \gamma(u))$ with $\gamma(u)$ be a positive constant such that $u \mapsto \frac{\gamma(u)}{1+d(x_0,u)}$ is bounded in $E$ \item[(ii)] The sequence $(d(x_0,z_n))_{n\geq 1}$ is nondecreasing \item[(iii)] $\sum_{n=1}^{j} \frac{d(z_n,z_{n+1})}{\alpha(d(x_0,z_{n+1}))} < 2\delta$, for all $j \geq 1$ \item[(iv)] for $v =\lim_{n\to\infty} z_n, \Phi(v) \leq \Phi(u)$ \item[(v)] $d(u,v) \leq min\{\delta\alpha(d(x_0,v)), \gamma(u)\}$ \item[(vi)] for every $w \in \bar{B}(u, \gamma(u)) \setminus B(u, d(x_0,u))$, $$ \Phi(w) \geq \Phi(v) - \frac{\varepsilon}{\delta\alpha(d(x_0,w))}d(v,w). $$ \end{itemize} \end{theorem} \begin{proof} Let us define a partial order in $E$ by letting \begin{equation}\label{e0} \Phi(r) \leq \Phi(s) - \frac{\varepsilon}{\delta\alpha(d(x_0,r))}d(r,s) \end{equation} and \begin{equation}\label{e1} d(x_0,r) \geq d(x_0,s). \end{equation} This relation is easily seen to be reflexive, antisymmetry and transitive. Indeed, it is clear that $r \prec r$, for every $ r \in E$. The partial order $\prec$ is antisymmetry. Indeed, if $r \prec s$ and $s \prec r$ then $d(x_0,r) = d(x_0,s)$, $$ \Phi(r) \leq \Phi(s) - \frac{\varepsilon}{\delta\alpha(d(x_0,r))}d(r,s) \leq \Phi(r) - \frac{2\varepsilon}{\delta\alpha(d(x_0,r))}d(r,s). $$ However $d(r,s) = 0$ and so $r = s$. $\prec$ is transitive, because if $r \prec s$ and $s \prec t$ then $d(x_0,r) \geq d(x_0,s) \geq d(x_0,t)$ and \begin{equation}\label{e2} \Phi(r) \leq \Phi(s) - \frac{\varepsilon}{\delta\alpha(d(x_0,r))}d(r,s), \quad \Phi(s) \leq \Phi(t) - \frac{\varepsilon}{\delta\alpha(d(x_0,s))}d(t,s). \end{equation} From $d(t,s) \leq d(t,r) + d(r,s)$, (\ref{e2}) becomes $$ \Phi(r) \leq \Phi(s) - \frac{\varepsilon}{\delta\alpha(d(x_0,r))}[d(t,r) - d(t,s)], \quad \Phi(s) \leq \Phi(t) - \frac{\varepsilon}{\delta\alpha(d(x_0,s))}d(t,s). $$ This implies $$ \Phi(r) \leq \Phi(t) + \Big[\frac{\varepsilon}{\delta\alpha(d(x_0,r))} - \frac{\varepsilon}{\delta\alpha(d(x_0,s))}\Big]d(t,s) - \frac{\varepsilon}{\delta\alpha(d(x_0,r))}d(r,t). $$ Since $\alpha(.)$ is nondecreasing and $d(x_0,r) \geq d(x_0,s)$, we obtain \begin{align*} r \prec s \mbox{ and } s \prec t &\Rightarrow \left\{ \begin{array}{l} \Phi(r) \leq \Phi(t) - \frac{\varepsilon}{\delta\alpha(d(x_0,r))}d(r,t) \\ d(x_0,r) \geq d(x_0,t) \end{array}\right. \\ &\Rightarrow r \prec t. \end{align*} Moreover, if we denote $S = \{r \in E \mid r \prec s\}$, by lower semi-continuity of $E$, $S$ is closed. Let $\varepsilon$, $\delta$, $u$ and $\gamma(u)$ given by theorem. Now we define a sequence $S_n$ of subsets as follows. Start with $z_1 = u$ and define $$ S_1 = \{w \in E \mid w \prec z_1\} \cap \bar B(u, \gamma(u)), $$ and inductively $$ S_{n} = \{w \in E \mid w \prec z_{n} \} \cap \bar B(u, \gamma(u)), z_{n+1} \in S_n $$ such that \begin{equation}\label{e3} \Phi(z_{n+1}) \leq \inf_{S_n} \Phi + \frac{1}{(n+1)\alpha(d(x_0,z_n))}. \end{equation} Clearly by transitivity of $\prec$ the sequence $(S_n)_n$ is a decreasing sequence of non empty closed sets. Hence also $(d(x_0,z_n))_n$ is a bounded nondecreasing sequence and converges in $[d(x_0,u), d(x_0,u)+ \gamma(u)]$. Now we prove that the diameters of these sets go to zero: $diamS_n \to 0$. Indeed, on one hand $w \in S_{n+1}$ implies $$ \Phi(w) \leq \Phi(z_{n+1}) - \frac{\varepsilon}{\delta \alpha(d(x_0,w))} d(w,z_{n+1}) \quad\text{and}\quad d(x_0,w) \geq d(x_0,z_{n+1}). $$ From (\ref{e3}), it results $$ \Phi(w) \leq \inf_{S_n} \Phi + \frac{1}{(n+1)\alpha(d(x_0,z_n))} - \frac{\varepsilon}{\delta \alpha(d(x_0,w))} d(w,z_{n+1}). $$ This implies $$ d(w,z_{n+1}) \leq \frac{\delta}{\varepsilon(n+1)} \frac{\alpha(d(x_0,w))}{\alpha(d(x_0,z_n))}. \\ $$ On the other hand, we have that $w$ belongs to $\bar B(u, \gamma(u))$, we obtain \begin{equation}\label{e4} d(w,z_{n+1}) \leq \frac{\delta}{\varepsilon (n+1)} \frac{\alpha(\gamma(u)+d(x_0,u))}{\alpha(d(x_0,z_n))}. \end{equation} Since the function $u \mapsto \frac{\gamma(u)}{1+d(x_0,u)}$ is bounded, then there exists $M > 0$ such that \begin{equation}\label{e5} \gamma(u) \leq M(1 + d(x_0,u)). \end{equation} From (\ref{e4}), (\ref{e5}) and $\alpha(.)$ is a nondecreasing function, it results \begin{equation}\label{e6} d(w, z_{n+1}) \leq \frac{\delta}{ \varepsilon (n+1)} \frac{\alpha((M+1)(1+d(x_0,z_n)))}{\alpha(d(x_0,z_n))}. \end{equation} By (\ref{e6}) and $\alpha(.)$ is a comparison function of order $k$, there exist $c, d > 0$ such that $$ d(w, z_{n+1}) \leq \frac{\delta}{ \varepsilon (n+1)}(c(M+1)^k+d), n\in \mathbb{N} $$ which gives $\mathop{\rm diam} S_{n+1}$ go to $0$, when $n \to \infty$. Now we claim that the unique point $v \in E$ in the intersection of the $S_n$'s satisfies conditions (iii)--(vi) of Theorem \ref{thm2.1}. Let then $\cap_n S_n = \{v\}$ and $z_n$ converges to $v$. Since $z_j \prec z_{j-1}\prec \dots \prec z_1$; and by (\ref{e0}), we have \begin{align*} \Phi(z_{j+1}) &\leq \Phi(z_j) - \frac{\varepsilon}{\delta \alpha(d(x_0,z_{j+1}))} d(z_j,z_{j+1}) \\ &\leq \Phi(z_1) - \sum_{n=1}^{j} \frac{\varepsilon d(z_j,z_{j+1})}{\delta\alpha(d(x_0,z_{j+1}))} \end{align*} or \begin{align*} \sum_{n=1}^{j} \frac{\varepsilon d(z_j,z_{j+1})}{\delta\alpha(d(x_0,z_{j+1}))} &\leq \Phi(u) - \Phi(z_{j+1}) \\ &\leq \inf_{E} \Phi + \varepsilon - \Phi(z_{j+1}) \leq \varepsilon. \end{align*} Thus assertion (iii) is shown. Since $v \in S_1$, (iv) is clear. It also results from it that $$ \frac{\varepsilon}{\delta\alpha(d(x_0,v))}d(v,u) \leq \Phi(u) - \Phi(v) \leq \inf_E \Phi + \varepsilon -\Phi(v) \leq \varepsilon. $$ The assertion (v) is shown. Finally, we prove (vi), let $w \in E$ such that $w \prec v$ and $w \in \bar B(u, \gamma(u))$, then we have $ w \prec z_n$ for every $n $. This gives $w \in \cap_n S_n$ and $w = v$, which means that $v$ be an minimal element in $\bar B(u, \gamma(u))$, i.e. $$ w \in \bar B(u, \gamma(u)) \quad\text{and}\quad w\prec v \Rightarrow w = v. $$ Consequently, $$ \Phi(w) > \Phi(v) - \frac{\varepsilon}{\delta\alpha(d(x_0,w))}d(v,w) $$ for every $ w \in \bar B(u, \gamma(u)) \setminus B(x_0, d(x_0,v))$. The proof is complete. \end{proof} \section{Applications} In this section $H$ denotes a Hilbert space, recall that a function $ \Phi: H \to \mathbb{R}$ is called G\^ateaux differerentiable if at every point $x_0$, there exists a continuous linear functional $f'(x_0)$ such that, for every $e \in X$, $$ \lim_{t \to 0}\frac{f(x_0+te) - f(x_0)}{t} = \langle f'(x_0), e\rangle . $$ We always assume that $\alpha: [0,\infty[ \to ]0,\infty[$ is a continuous nondecreasing comparison function of order $k$. For the rest of the text we will write $$ \Phi^c =\{u \in H : \Phi(u) \leq c\}, $$ for the sublevel sets as usual. \begin{definition} \rm We say that $\Phi$ satisfies $(C^{\alpha}_c)$ if: Every sequence $(u_n)_n \subset H$ such that $ \Phi(u_n) \to c$ and $\Phi'(u_n) \alpha(\|u_n\|) \to 0$ possesses a convergent subsequence. \end{definition} \begin{remark} \label{rmk3.1}\rm Note that if $\alpha(s) = cte$, the $(C^{\alpha}_c)$ condition is just the famous Palais-Smale condition and if $\alpha(s) = s + 1$, $(C^{\alpha}_c)$ is $(C)$ condition introduced by Cerami in \cite{Ce}. \end{remark} We can now state the following result. \begin{theorem} \label{thm3.1} Let $H$ be a Hilbert space, $ \Phi: H \to \mathbb{R}$ lower semi-continuous, bounded below and G\^ateaux differentiable. Let $\alpha: [0,\infty[ \to ]0,\infty[$ a continuous nondecreasing comparison function of order $k$. Assume that for every $\varepsilon > 0$, $$ \Phi^{a + \varepsilon} \cap K \neq \emptyset, $$ with $K$ is bounded in $H$ and $\Phi$ satisfies $(C^{\alpha}_a)$, with $a = \inf_{H}\Phi$, then $\Phi$ has a minimal point. \end{theorem} For the proof of this theorem we will use the following lemmas. \begin{lemma} \label{lem3.1} Under the conditions of Theorem \ref{thm3.1}, for every $\varepsilon > 0$, every $u \in H$ such that $\Phi(u) \leq \inf_H \Phi + \varepsilon$ and every $\delta > 0$ such that $$ \delta \leq \frac{\|u\|+1}{2\alpha(3(1+\|u\|))} $$ there exists $v \in H$ that satisfies \begin{enumerate} \item $\Phi(v) \leq \Phi(u)$ \item $\frac{\|v-u\|}{\alpha(\|v\|)} \leq \delta$ \item for every $h \in H, t \in \mathbb{R}$ such that$ \|h\| = 1, |t| \leq 1$ and $t \geq 0$ we have $$ \Phi(v + th) \geq \Phi(v) - \frac{\varepsilon}{\delta\alpha(\|v + th\|)}|t|. $$ \end{enumerate} \end{lemma} \begin{proof} Let in Theorem \ref{thm2.1}, $x_0 = 0, \gamma(u) = 2(\|u\| + 1)$ and $d(x,y) =\|x - y\|$ for every $x, y \in H$. Then, by $iv)$ and $v)$ of Theorem \ref{thm2.1}, there exists $v \in H$ ( $v = \lim_{n \to \infty} z_n, (z_n)$ the sequence built in theorem \ref{thm2.1}) such that \begin{equation}\label{e7} \Phi(v) \leq \Phi(u) ~~{\rm and}~ \|v-u\| \leq \delta\alpha(\|v\|). \end{equation} Thus assertions 1. and 2. follow. Now we prove the assertion 3. Let $h \in H$ such that $\|h\| = 1$ and $|t| \leq 1$ we have $v \in \bar B(u, \|u\|+1)$. Indeed, if not $\|u\|+1 < \|v - u \|$. Since $\alpha(.)$ is nondecreasing, $ \delta \leq \frac{\|u\|+1}{2\alpha(3(1+\|u\|))} $ and by (\ref{e7}), it results $$ \|u\|+1 < \|v- u\| \leq \delta \alpha(\|v\|) \leq \delta \alpha(3(1+\|u\|))) \leq \frac{\|u\|+1}{2}. $$ This is a contradiction. Furthermore, we have \begin{align*} \|v + th\| & \leq \|v\| + |t|\|h\| = \|v\| + |t| \\ &\leq 2\|u\| + 1 + 1 = \gamma(u). \end{align*} On the other hand, it is clear , since $t\langle v,h\rangle \geq 0$, that $$ \|v +th\| = [\|v\|^2+\|th\|^2+2t]^{1/2} \geq \|v\|. $$ Thus, by (iv), (v), (vi) of Theorem \ref{thm2.1}, assertions 1, 2, 3 of the lemma follow. \end{proof} \begin{lemma} \label{lem3.2} Under the conditions of Theorem \ref{thm3.1}, we have \begin{equation}\label{e8} |\langle \Phi'(v),h\rangle | \leq \frac{\varepsilon}{\delta\alpha(\|v\|)}, \quad \forall h \in H, \|h\| = 1. \end{equation} \end{lemma} \begin{proof} Let $h \in H$ such that $\|h\| = 1$ and consider two cases:\\ Case 1. If $\langle v,h\rangle \geq 0$ and $t > 0$, from 3. of Lemma \ref{lem3.1} and $\Phi$ being G\^ateaux differentiable, letting $t$ approach $0$, we obtain $$ \langle \Phi'(v),h\rangle \geq - \frac{\varepsilon}{\delta\alpha(\|v\|)}. $$ Case 2. In the similar way, if $ \langle v,h \rangle \leq 0$ and $t < 0$ goes to $0$, we have $$ \langle \Phi'(v),h\rangle \leq \frac{\varepsilon}{\delta\alpha(\|v\|)}, \quad \forall h, \|h\| = 1. $$ Thus the Lemma \ref{lem3.2} follows. \end{proof} \begin{proof}[Proof of Theorem \ref{thm3.1}] For $\varepsilon = \frac{1}{n}$, with $n \geq 1$, there exists a sequence $(u_n) \subset K$ such that $$ \Phi(u_n) \leq a + \frac{1}{n} $$ and, since $(u_n)$ is bounded, there exists $\delta > 0$ such that $$ \delta \leq \frac{\|u_n\|+1}{\alpha(3(1+\|u_n\|))}, \forall n \geq 1. $$ Consequently, by Lemma \ref{lem3.1} and Lemma \ref{lem3.2}, there exists a sequence $(v_n)$ satisfying \begin{itemize} \item[(i)] $\Phi(v_n) \leq \Phi(u_n)$ \item[(ii)] $\|\Phi'(v_n)\|\alpha(\|v_n\|) \to 0$, as $n \to \infty$. \end{itemize} The $(C^\alpha_a)$ condition implies that $(v_n)$ has a subsequence $(v_{n_k})$ convergent to some point $u$. Since $\Phi$ is lower semi-continuous, we get $$ \inf_{H}\Phi \leq \Phi(u) \leq \liminf_{k\to \infty}\Phi(v_{n_k}) \leq \inf_{H}\Phi. $$ Therefore, $\Phi(u) = \inf_{H}\Phi$. \end{proof} Now, we illustrate Theorem \ref{thm3.1} by an example where the function $\Phi$ checks the conditions of Theorem \ref{thm3.1}, but the Palais-Smale condition and Cerami condition do not hold. \noindent{\bf Example.} Consider $$ f(s) = \begin{cases} \arctan(s)& \text{if } s \leq 0 \\ \sin(s) & \text{if } 0 \leq s \leq 2\pi \\ \arctan(s-2\pi) & \text{if } s \geq 2\pi. \end{cases} $$ and $\Phi(u) = f(2\pi + Log(\|u\|^2+1) - (\|u\|^2 + 1)^\frac{1}{2})$ for $ u \in H$. It is clear that $\Phi$ is $C^1$ functional and $a = \inf_H\Phi = -1$. Take $$ K = \{ u\in H : \log(\|u\|^2+1) - (\|u\|^2 + 1)^\frac{1}{2} \in [-2\pi, 0]\}, $$ it is easy to see that $\Phi^{-1+\varepsilon} \cap K \neq \emptyset$ for every $\varepsilon > 0$. On the other hand, $\Phi$ satisfies $(C^{\alpha}_c)$, with $\alpha(s) = s^2+1$, and by Theorem \ref{thm3.1}, $\Phi$ has a minimal point $u_0$ which $\Phi'(u_0) = 0$. \begin{theorem} \label{thm3.2} Let $\Phi : H \to \mathbb{R}$ be G\^ateaux differentiable and bounded below, says $a \inf_{H}\Phi$. Assume that $\alpha: [0,\infty[ \to ]0,\infty[$ be a continuous nondecreasing function such that $\int_{1}^{\infty} \frac{1}{\alpha(s)} \,ds = +\infty$. If $ \Phi$ satisfies $(C^\alpha_a)$ then the set $\Phi^{a+\beta}$ is bounded, for some $\beta > 0$. \end{theorem} The main point to prove Theorem \ref{thm3.2} is the following. \begin{lemma} \label{lem3.3} Under the conditions of Theorem \ref{thm3.2}, for every $\varepsilon > 0$, every $u \in H$ such that $\Phi(u) \leq \inf_H \Phi + \varepsilon$ and every $\delta > 0$ there exists $v \in H$ satisfies \begin{enumerate} \item $\Phi(v) \leq \Phi(u)$ \item $\frac{\|v-u\|}{\alpha(\|v\|)} \leq \delta$ \item for every $h \in H$ such that $ \|h\| = 1$, we have $$ |\langle \Phi'(v),h\rangle | \leq \frac{\varepsilon}{\delta\alpha(\|v\|)}. $$ \end{enumerate} \end{lemma} \begin{proof} Let in Theorem \ref{thm2.1}, $x_0 = 0$ and $d(x,y) = \|x - y\|$ for every $x, y \in H$. From theorem \ref{thm2.1} there exists a sequence $(z_n)_{n\geq 1}$ satisfying $(\|z_n\|)$ is nondecreasing and \begin{equation}\label{e9} \sum_{n=1}^{j} \frac{\|z_n-z_{n+1}\|}{\alpha(\|z_{n+1}\|)} < 2\delta, \quad \forall j \geq 1. \end{equation} However, since $\int_{1}^{\infty} \frac{1}{\alpha(s)} \,ds = +\infty$ there exists $\gamma > 0$ such that \begin{equation}\label{e10} \delta \leq \frac{1}{2}\int_{\|u\|}^{\|u\|+\gamma} \frac{1}{\alpha(s)}\,ds. \end{equation} Put $v = \lim_{n\to\infty} z_n$ and $\gamma(u) = 2\|u\| + \gamma +1$ in Theorem \ref{thm2.1}. Thus, by (iv)-(v) of Theorem \ref{thm2.1}, we obtain $$ \Phi(v) \leq \Phi(u) \quad\text{and}\quad \|v-u\| \leq \delta\alpha(\|v\|). $$ For the proof of assertion 3, it is enough to verify that $h \in H$ such that $\|h\| = 1$ we have $v + th \in \bar B(u,\gamma(u))$ for every $t$ sufficiently small. Now we prove that \begin{equation}\label{e11} \|z_n\| \leq \|u\|+\gamma, \quad \forall n \geq 1. \end{equation} If not, there exists $j \geq 1$ such that $\|z_{j+1}\| > \|u\|+\gamma$. However, by (\ref{e10}) and $\alpha$ is nondecreasing, we obtain \begin{align*} 2\delta &\leq \int_{\|z_1\|}^{\|z_{j+1}\|} \frac{1}{\alpha(s)} \,ds \\ &\leq \sum_{n=1}^{j} \int_{\|z_n\|}^{\|z_{n+1}\|} \frac{1}{\alpha(s)} \,ds \\ &\leq \sum_{n=1}^{j} \frac{\|z_{n+1}\| - \|z_n\|}{\alpha(\|z_{n+1}\|)} \\ &\leq \sum_{n=1}^{j} \frac{\|z_n - z_{n+1}\|}{\alpha(\|z_{n+1}\|)}. \end{align*} This contradicts (\ref{e9}). Using (\ref{e11}), we have \begin{equation}\label{e12} \|v - u\| \leq 2\|u\| + \gamma. \end{equation} Thus, for $|t| \leq 1$ and $h \in H$ such that $ \|h\| = 1$ and by (\ref{e12}), it results $$\|v + th - u\| \leq 2\|u\| + \gamma +1 = \gamma(u).$$ Finally, the Lemma \ref{lem3.2} allows to conclude. The proof is complete. \end{proof} \begin{proof}[Proof of theorem \ref{thm3.2}] Suppose, by contradiction, that $\Phi^{a + \beta}$ is unbounded for all $\beta > 0$. Then, there exists $(u_n)$ such that $\|u_n\| \geq n$ and $$ \Phi(u_n) \leq a + \frac{1}{n}. $$ and Lemma \ref{lem3.3} with $\varepsilon = (\frac{1}{n})^2, \delta = \frac{1}{n}$ implies the existence of $(v_n)$ satisfying \begin{itemize} \item[(i)] $\Phi(v_n) \leq \Phi(u_n)$ \item[(ii)]$ \|v_n - u_n\| \leq \frac{1}{n} \alpha(\|v_n\|)$ \item [(iii)] $\|\Phi'(v_n)\|\alpha(\|v_n\|) \to 0$, as $n \to \infty$. \end{itemize} We reach a contradiction with $(C^\alpha_a)$, since (i)-(iii) give respectively \begin{enumerate} \item $\Phi(v_n) \to a$, as $n \to \infty$, \item $ \|v_n\| \to \infty$, as $n \to \infty$, \item $\|\Phi'(v_n)\|\alpha(\|v_n\|) \to 0$, as $n \to \infty$. \end{enumerate} \end{proof} As an immediate consequence of the above results we have the following result. \begin{corollary} Let $H$ be a Hilbert space, $ \Phi: H \to \mathbb{R}$ lower semi-continuous, bounded below and G\^ateaux differentiable. Assume that $\alpha: [0,\infty[ \to ]0,\infty[$ be a continuous nondecreasing function such that $\int_{1}^{\infty} \frac{1}{\alpha(s)} \,ds = +\infty$. If $ \Phi$ satisfies $(C^\alpha_a)$, with $a = \inf_{H}\Phi$, then $\Phi$ has a minimal point. \end{corollary} \begin{thebibliography}{00} \bibitem{E1} Ekeland I., \emph{On the variational principle}, J. Math. Anal. Applic., 47 (1974) 324-357. \bibitem{E2} Ekeland I., \emph{Convexity methods in Hamiltonian mechanics}, Springer, Berlin, (1990). \bibitem{Ce} Cerami G., \emph{Un criterio de esistenza per i punti critici su variet\'a ilimitate}, Rc. Ist. Lomb. Sci. Lett. 121,(1978) 332-336. \bibitem{C} Clark, F. H. \emph{Optimisation and Nonsmooth Analysis}, Wiley, NewYork1983. \bibitem{CS} Costa, D. G., Elves De B. eSilva; \emph{The Palais-Smale condition versus coercivity}, Nonlinear Analysis, T. M. A. 16 (1991) 371-381. \end{thebibliography} \end{document}