\documentclass[twoside]{article} \usepackage{amssymb} % font used for R in Real numbers \pagestyle{myheadings} \markboth{\hfil Neumann and periodic boundary-value problems \hfil EJDE--2000/63} {EJDE--2000/63\hfil Petr Girg \hfil} \begin{document} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent {\sc Electronic Journal of Differential Equations}, Vol.~{\bf 2000}(2000), No.~63, pp.~1--28. \newline ISSN: 1072-6691. URL: http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu \quad ftp ejde.math.unt.edu (login: ftp)} \vspace{\bigskipamount} \\ % Neumann and periodic boundary-value problems for quasilinear ordinary differential equations with a nonlinearity in the derivative % \thanks{ {\em Mathematics Subject Classifications:} 34B15, 47H14. \hfil\break\indent {\em Key words:} p-Laplacian, Leray-Schauder degree, Landesmann-Lazer condition. \hfil\break\indent \copyright 2000 Southwest Texas State University. \hfil\break\indent Submitted July 18, 2000. Published October 16, 2000. \hfil\break\indent Supported by grants GA\v{C}R \# 201/97/0395, and M\v{S}MT\v{C}R \# VS 97156} } \date{} % \author{Petr Girg} \maketitle \begin{abstract} We present sufficient conditions for the existence of solutions to Neumann and periodic boundary-value problems for some class of quasilinear ordinary differential equations. We also show that this condition is necessary for certain nonlinearities. Our results involve the p-Laplacian, the mean-curvature operator and nonlinearities blowing up. \end{abstract} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{definition}{Definition}[section] \newtheorem{example}{Example}[section] \newtheorem{proposition}[theorem]{Proposition} \newtheorem{remark}{Remark}[section] \renewcommand{\theequation}{\thesection.\arabic{equation}} \catcode`@=11 \@addtoreset{equation}{section} \catcode`@=12 \section{Introduction} The semilinear boundary-value problems \begin{eqnarray} \label{eq1d} &u''(t)+g(u'(t))+h(u(t))=f(t) \quad t\in (0,T)\,, &\\ &u(0)=u(T)\,,\quad u'(0)=u'(T)\,,& \label{per_con} %\nonumber \end{eqnarray} and \begin{eqnarray} \label{eq1n} &u''(t)+g(u'(t))=f(t) \quad t\in (0,T)\,,& \\ \label{Neu_con} &u'(0)=0\,,\quad u'(T)=0\,,& %\nonumber \end{eqnarray} have been extensively studied by many authors (see e.g. \cite{CD,Danc,Maw}). In this paper we extend their results to the quasilinear boundary-value problems: \begin{equation} \label{fund_eq} \bigl(\varphi(u'(t))\bigr)'+g(u'(t))+h(u(t)) = f(t)\quad t\in (0,T)\,, \end{equation} subject to (\ref{per_con}) or to (\ref{Neu_con}). Overall, we will assume the following: \begin{description} \item[{\rm(P)}] $\varphi$ is an increasing homeomorphism of $I_1$ onto $I_2,$ where $I_1, I_2 \subset \mathbb R$ are open intervals containing zero and $\varphi(0)=0,$ \item[{\rm(G)}] $g$ is a continuous real function, \item[{\rm(H)}] $h$ is a continuous, bounded real function having limits in $\pm\infty$: $$h(-\infty):=\lim_{\xi\rightarrow -\infty}h(\xi)<\lim_{\xi\rightarrow +\infty}h(\xi)=:h(+\infty)\,.$$ \end{description} \noindent We also need to impose some of the following assumptions to prove several particular results. \begin{description} \item[{\rm(P')}] the inverse of $\varphi$ (denoted by $\varphi_{-1}$) is continuously differentiable, \item[{\rm(P'')}] $\varphi'_{-1}(0)>0$, \item[{\rm(PH)}] $\varphi$~is odd and there exist $c, \delta >0$ and $p > 1$ such that for all $z\in (-\delta,\delta)\cap\mathop{\rm Dom}\varphi:\,c|z|^{p-1}\leq|\varphi(z)|$, \item[{\rm(G')}] $g$ is a continuously differentiable real function. \end{description} \noindent The results presented involve blow-up-type nonlinearities such as $\varphi(z)=\tan(z)$, bounded nonlinearities of the type $\varphi(z)=\arctan(z)$, the $p$-Laplacian when $\varphi(z) =|z|^{p-2}z, 10\,. \end{equation} Then for any $\widetilde f\in\widetilde C_T$ ( $\widetilde f\in\widetilde C[0,T]$ ) satisfying \begin{equation}\label{M_con} \|\widetilde f\|_{L^2}<\sqrt{\frac{3}{T}}b-\sqrt{T}\sup_{\xi\in\mathbb R}|h(\xi)| \end{equation} the BVP (\ref{fund_eq})--(\ref{per_con}) ( (\ref{fund_eq})--(\ref{Neu_con}) ) has a solution if \begin{equation} \label{D_like} s(\widetilde f) + h(-\infty) < \overline f< s(\widetilde f) + h(+\infty)\,, \end{equation} where $s(\widetilde f)$ is given by Theorem \ref{Th1} ( Theorem \ref{Th3} ). Suppose, moreover, that $$h(-\infty) b\,, \end{equation} then the BVP (\ref{fund_eq})--(\ref{per_con}) or (\ref{fund_eq})--(\ref{Neu_con}), respectively, does not admit any solution with $f=\widetilde f+\overline f$. Compare also with results in \cite{KaSe}. } \end{remark} \begin{remark} \label{rem1} {\rm It is worth noting that if we consider the periodic BVP for quasilinear ODE: $$(\varphi(u'))' + \lambda u' + h(u) = f,\quad t\in (0,T)\,,$$ where $\lambda\in\mathbb R$, then the condition (\ref{D_like}) from Theorem \ref{Th4} is reduced to \begin{equation} \label{LanLaz} h(-\infty)<\overline f0$ and $g,h \in C(\mathbb R,\mathbb R)$, $h$ has finite limits $h(-\infty)0$. Then, for all $\widetilde f\in \widetilde C_T\,:\quad \|\widetilde f\|_{L^2} < \sqrt{\frac{3}{T}}\frac{2\sqrt{3}}{9}-\frac{\pi}{2}\sqrt{T}$, this problem has a solution if and only if $\overline f\in \mathbb R$ satisfy \begin{equation} \label{ex_eq} -\frac{\pi}{2}<\overline f<\frac{\pi}{2}\,. \end{equation} To prove this result, we can use Theorem \ref{Th4}. Indeed, $\varphi$ is odd homeomorphism of $(-\frac{1}{\sqrt{3}},\frac{1}{\sqrt{3}})$ onto $(-\frac{2\sqrt{3}}{9},\frac{2\sqrt{3}}{9})$, so that $b=-a=\frac{2\sqrt{3}}{9}$. Moreover, $$\lim_{z\pm\infty}\arctan\,z = \pm \frac{\pi}{2}\quad \mbox{and}\quad -\frac{\pi}{2}<\arctan\, z < \frac{\pi}{2},\\ \mbox{for all } z\in\mathbb R\,.$$ Hence the condition (\ref{ex_eq}) follows from (\ref{D_like}), where $s(\widetilde f)\equiv 0$, because $g(z)\equiv z$ (see Remark \ref{rem1}). This problem describes forced oscillations of voltage $u$ in an electrical circuit with {\sl ferro-resonance} (see e.g. \cite{KotKub}). Such nonlinear circuits are used in radiotechnics.} \end{example} \begin{table} \begin{center} {\small \tabcolsep=1pt \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|} \hline & \multicolumn{4}{|c|}{Periodic, $g(s)=s^3$} & \multicolumn{3}{|c|}{Neumann, $g(s)=s^3$} & \multicolumn{2}{|c|}{both, $h(s)=\arctan s$} \\ \cline{2-10} $\varphi(z)$ & \multicolumn{1}{|c|}{Thm 1} & \multicolumn{3}{|c|}{Theorem 2} & \multicolumn{3}{|c|}{Theorem 3} & \multicolumn{2}{|c|}{Theorem 4} \\ \hline & (P)+(G) & +(P') & +(G') & +(P'') & (P)+(G) & +(P') & +(G') & \multicolumn{2}{|c|}{(P)+(G)+(PH)}\\ \hline $\begin{array}{c} |z|^{p-2}z, \\ 12 \end{array}$ & {\sf A} & \multicolumn{3}{|c|}{\sf NA} & {\sf A} & \multicolumn{2}{|c|}{\sf NA} & \multicolumn{2}{|c|}{\sf A} \\ \hline $\frac{z}{\sqrt{1-z^2}}$ & {\sf A} & \multicolumn{3}{|c|}{\sf A} & \multicolumn{3}{|c|}{\sf A} & \multicolumn{2}{|c|}{\sf A} \\ \hline $\frac{z}{\sqrt{1+z^2}}$ & {\sf A$^{*}$} & \multicolumn{3}{|c|}{\sf A$^{*}$} & \multicolumn{3}{|c|}{\sf A$^{*}$} & $\begin{array}{c} T<\frac{2\sqrt{3}}{\pi} \\ {\sf A}^{\dag} \end{array}$ & $\begin{array}{c} T>\frac{2\sqrt{3}}{\pi} \\ {\sf NA} \end{array}$ \\ \hline \multicolumn{1}{|c|}{$\exp\left(\frac{1}{|z|}\right)z$} & {\sf A} & \multicolumn{2}{|c|}{\sf A} & {\sf NA} & \multicolumn{3}{|c|}{\sf A} & \multicolumn{2}{|c|}{\sf NA} \\ \hline \end{tabular} } \end{center} \caption{\quad Legend: $^\#$ With periodic conditions only,\hfill\break $^{*}$ Provided that $\|\widetilde f\|_{L^2} < \sqrt{\frac{3}{T}}$, \quad $^{\dag}$ Assuming that $\|\widetilde f\|_{L^2}<\sqrt{\frac{3}{T}}-\frac{\pi}{2}\sqrt{T}$.} \label{tbl1} \end{table} Table \ref{tbl1} illustrates the applicability of Theorems \ref{Th1}--\ref{Th4} for some particular nonlinearities $\varphi$. The symbol {\bf\sf A} (applicable) indicates that the corresponding assumption is satisfied and thus using an appropriate Theorem one obtains the desired conclusion; on the other hand {\bf\sf NA} (not applicable) indicates that the corresponding assumption is not satisfied and that the selected Theorem is not applicable in that case). We also relate our results to the results already known. \newpage \section{Proofs of Theorems \ref{Th1} -- \ref{Th3}} We start this section with the following lemma: \begin{lemma} \label{lemma1} Let $f\in C[0,T]$, $q:J_1\rightarrow J_2$ be continuous, where $J_1,J_2$ are nonempty intervals and $J_1$ contains zero in its interior. Let $w\in C^1[0,T]$, $w(0)=w(T)$ and $w$ satisfies \begin{equation} \label{qeqa} \quad w'(t)+q(w(t))=f(t),\quad t\in(0,T) \end{equation} then \begin{equation} \label{ODH2} |w(t)-w(s)|\leq |t-s|^{\frac{1}{2}}\|\widetilde f\|_{L^2}\,. \end{equation} Moreover, if there exists $t_0\in [0,T]$ such that $w(t_0)=0$ then \begin{equation} \label{ODH1} \|w\|_{C} \leq \sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}\,. \end{equation} \end{lemma} \begin{remark} {\rm The previous lemma will be used to estimate solutions of (\ref{f_eq})--(\ref{f_con}) or (\ref{f_eq})--(\ref{N_con}), respectively. Indeed, the second of (\ref{f_con}) yields $w(0)=w(T)$ directly. On the other hand, the first of (\ref{f_con}), the continuity of $w$ and the fact, that $\varphi$ satisfies {\rm (P)} imply that there exists $t_0\in [0,T]$ such that $w(t_0)=0$. In the case of (\ref{N_con}) we have $w(0)=w(T)=0$. } \end{remark} \paragraph{Proof} Using the Cauchy--Schwarz inequality it is easy to show that \begin{equation} \label{wswth} |w(t)-w(s)|=\left|\int_s^t w'(\tau)d\tau\right|\leq |t-s|^{\frac{1}{2}}\|w'\|_{L^2}\,. \end{equation} Multiplying both sides of equation (\ref{f_eq}) by $w'$, integrating from $0$ to $T$, splitting $f$ as in (\ref{red_f}) and using $w(0)=w(T)$, we find that $\| w'\|^2_{L^2}=\left(\widetilde f, w'\right)_{L^2}$ (where $\left(\widetilde f, w'\right)_{L^2}:=\int_0^T\widetilde f w'$ is the scalar product in $L^2$). The Cauchy-Schwarz inequality yields $\|w'\|_{L^2}\leq\|\widetilde f\|_{L^2}$, which together with (\ref{wswth}) establishes (\ref{ODH2}). Now we are going to estimate $\|w\|_{C}.$ Decompose the function $w$ as $w = \widetilde w + \overline w$, where $\int_0^T\widetilde w(t)dt=0$ and $\overline w\in \mathbb R$. Since $\widetilde w\in C^1[0,T] \subset W^{1,2}[0,T]$ and $w(0)=w(T)$, the Sobolev inequality \cite[Proposition 1.3]{MawW} yields $\|\widetilde w\|_{C} \leq \sqrt{\frac{T}{12}}\| w'\|_{L^2}\leq\sqrt{\frac{T}{12}}\|\widetilde f\|_{L^2}.$ Since we assume that there exists a $t_0\in [0,T]$ such that $w(t_0)=0$, it shall be $|\overline w|\leq \|\widetilde w\|_{C}$. Hence $\|w\|_{C}\leq 2\sqrt{\frac{T}{12}}\|\widetilde f\|_{L^2}=\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2},$ which is the desired inequality (\ref{ODH1}). \hfill$\diamondsuit$\medskip Now we proceed to proofs of Theorems \ref{Th1}--\ref{Th3}. \paragraph{Proof of Theorem \ref{Th1}} We divide the proof into six steps: \noindent{\bf Step 1} {\it Let $q$ be a continuously differentiable and bounded real function. We prove that for all $\widetilde f\in \widetilde C_T$ there exists $\overline f\in \mathbb R$ such that the equation \begin{equation} \label{a_w} w'(t) + q\bigl(w(t)\bigr) = \widetilde f(t) + \overline f \end{equation} has a solution $w\in C^1[0,T]$ satisfying (\ref{f_con}).} Let $\widetilde f\in \widetilde C_T$ be given. Since $q$ is continuously differentiable and bounded function, the solution ( denoted by $w(t,\overline f,\alpha)$ ) to the initial-value problem \begin{eqnarray} \label{In_a} &w'(t) + q\bigl(w(t)\bigr) = \widetilde f(t) + \overline f,&\\ &w(0) = \alpha& \nonumber \end{eqnarray} exists on $(0,T)$, it is unique and is continuously differentiable with respect to parameters $\alpha$ and $\overline f$ (see e.g \cite{CodLew}). Let $M=\sup_{y\in\mathbb R}|q(y)|$. Taking $\overline f > M+\varepsilon$ and integrating (\ref{a_w}) from 0 to T, we find $w(T,\overline f, \alpha) > \alpha$. On the other hand, if $\overline f < -M-\varepsilon$ then we get $w(T,\overline f, \alpha)<\alpha.$ Since $w$ depends continuously on $\overline f$, for all $\alpha\in\mathbb R$, there exists $\overline f_{\alpha}\in\mathbb R$ such that $w(T,\overline f_{\alpha}, \alpha)=\alpha.$ Moreover, the partial derivative with respect to the parameter $\overline f$ , $w_{\overline f}(t,\overline f, \alpha)$, is a solution to the linear initial-value problem \begin{eqnarray} \label{inz} &z'(t) = 1-q'\bigl(w(t,\overline f, \alpha)\bigr )z(t)\,, &\\ &z(0) = 0\,.& \nonumber \end{eqnarray} The explicit formula of the solution of (\ref{inz}) yields $w_{\overline f}(T,\overline f,\alpha) > 0$, which means that $w(T,\cdot, \alpha)$ is increasing and $\overline f_{\alpha}$ is unique. Thus, we can define the mapping $\psi_1:\mathbb R\rightarrow [-M,M]$ by $\alpha\mapsto\overline f_{\alpha}$. As $w_{\overline f}(T,\overline f,\alpha) > 0$, by the abstract implicit function theorem (see \cite[Theorem 2.2.3]{Am_Pro}), we find that $\psi_1:\mathbb R\rightarrow [-M,M]$ is continuous. Rewrite (\ref{In_a}) as an integral equation: $$ w(t)=\int_0^t \left(\widetilde f(\tau)-q(w(\tau))\right)\,d\tau +\overline f t +\alpha\,. $$ Taking $\alpha > \int_0^T|\widetilde f(\tau)|d\tau+2TM$, we see that $w(t,\overline f,\alpha)>0$ for all $t\in[0,T]$ and $\overline f \in [-M,M]$. Consequently $\int_0^T\varphi_{-1}\bigl (w(\tau,\overline f,\alpha)\bigr )d\tau >0.$ Conversely, if $\alpha<-\int_0^T|\widetilde f(\tau)|d\tau-2TM$ then $w(t,\overline f, \alpha)<0$ for all $t\in [0,T]$ and $\overline f\in [-M,M]$. Hence $\int_0^T\varphi_{-1}\bigl(w(\tau,\overline f,\alpha)\bigr)d\tau<0.$ Since $\varphi_{-1}$ is a continuous real function, the mapping $\alpha \mapsto \int_0^T\varphi_{-1}\bigl(w(t,\overline f,\alpha)\bigr)dt$ is continuous and there exists at least one $\alpha_{\overline f}\in\mathbb R$ such that $\int_0^T\varphi_{-1}\bigl(w(t,\overline f,\alpha)\bigr)dt = 0.$ Moreover the partial derivative of $w(t,\overline f,\alpha)$ with respect to initial condition $\alpha$, $w_{\alpha}(t,\overline f,\alpha)$, is the solution of the linear initial-value problem \begin{eqnarray}\label{in_da} &v'(t) + q'\bigl(w(t,\overline f,\alpha)\bigr)v(t)=0\,, &\\ &v(0) = 1\,.&\nonumber \end{eqnarray} Taking into account explicit formula of the solution of (\ref{in_da}), it is easy to verify that $w_{\alpha}(t,\overline f, \alpha)>0$ for all $t\in [0,T]$ and $\overline f, \alpha\in\mathbb R$. Thus, for all $\overline f\in\mathbb R$ there exists unique $\alpha=\alpha_{\overline f}$ such that $\int_0^T\varphi_{-1}\bigl(w(t,\overline f,\alpha_{\overline f})\bigr)dt=0$. By the same reason as above, the mapping $\psi_2:\mathbb R\rightarrow\mathbb R$ defined by $\overline f\mapsto \alpha_{\overline f}$ is continuous. Let us consider the continuous mapping $\psi_1\circ\psi_2:[-M,M]\rightarrow [-M,M]$. Due to the Brouwer fixed point theorem there exists at least one $\overline f_0\in [-M,M]$ such that $\psi_1\bigl(\psi_2(\overline f_0)\bigr)=\overline f_0$. Hence the function $w(t,\overline f_0, \psi_2(\overline f_0))$ is a solution of the equation (\ref{a_w}) subject to (\ref{f_con}). \noindent {\bf Step 2} {\it We show that if {\rm (P)}, {\rm (G)} are satisfied then, for all $\widetilde f\in\widetilde C_T$, there exists at least one $\overline f\in\mathbb R$ such that the boundary-value problem for the equation \begin{equation} \label{wfg} w'(t)+g\bigl(\varphi_{-1}\bigl(w(t)\bigr)\bigr)=\widetilde f(t) +\overline f \end{equation} subject to (\ref{f_con}) has at least one solution.} Let $w$ be the solution of (\ref{wfg})--(\ref{f_con}). Integrating (\ref{wfg}) from $0$ to $T$ and then dividing by $T$, we obtain $$\frac{1}{T}\int_0^T g\bigl(\varphi_{-1}\penalty0{}(w(t))\bigr)d\,t=\overline f\,.$$ With respect to (\ref{ODH1}) (consider $q:=g\circ\varphi_{-1}$ in Lemma \ref{lemma1}) the following a priori bound on $w$ results: $\|w\|_{C} \leq\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}$. Hence we estimate $\overline f$ as follows: \begin{equation} \label{ODH3} |\overline f|\leq\sup\left\{|g(\xi)|: \varphi_{-1}\left(-\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}\right)\leq\xi\leq \varphi_{-1}\left(\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}\right)\right\}. \end{equation} As $\|w\|_{C}\leq\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}$, the restriction of $g\circ\varphi_{-1}$ on $I=\left[-\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}\right.,$ $\left.\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}\right]$ is essential in further considerations. Note that the assumption $\|\widetilde f\|_{L^2}\leq\sqrt{\frac{T}{3}}\mbox{min}\{|a|,b\}$ imply $I\subset (a,b)$ $(=\mathop{\rm Dom}\varphi_{-1})$; hence $g\circ \varphi_{-1}$ is well defined on $I$. Let us introduce the following sequence of functions $\gamma_n(x):=n\int_{I}\varrho\left (\frac{x-y}{n}\right )g\bigl(\varphi_{-1}(y)\bigr )dy,$ where $\varrho:\mathbb R\rightarrow\mathbb R$ is the regularization kernel given by \begin{equation} \varrho(x):=\left\{ \begin{array}{ll} c_0 e^{\frac{1}{|x|^2-1}} &\mbox{for } |x|<1\,, \\[1pt] 0 & \mbox{for } |x|\geq 1\,, \end{array} \right. \end{equation} where $c_0$ is a constant such that $\int_{-1}^1 \varrho(x)dx = 1.$ It is easy to verify that $\bigl\{\gamma_n\bigr\}_{n=1}^{\infty}$ converges to $g\circ\varphi_{-1}$ uniformly on $I,$ that $\gamma_n$ is continuously differentiable for any $n\in\mathbb N$ and that $\|\gamma_n\|_{C(I)}\leq\|g\circ\varphi_{-1}\|_{C(I)}$ ( $C(I)$ denotes the space of continuous real functions defined on $I$ ). From the first step we know that there exist $w_n$ and $\overline f_n$ such that $$ w'_{n}(t) + \gamma_n\bigl(w_n(t)\bigr)=\widetilde f(t) + \overline f_n $$ and $$\int_0^T\varphi_{-1}\bigl(w_n(t)\bigr)\,dt=0,\quad w_n(0)=w_n(T).$$ Utilizing (\ref{ODH1}) and $\|\gamma\|_{C(I)}\leq \|g\circ\varphi_{-1}\|_{C(I)}<+\infty$ we find that $\bigl\{\overline f_n\bigr\}_{n=1}^{\infty}$ is a bounded sequence. By (\ref{ODH1}), $w_n$ are equibounded and, by (\ref{ODH2}), $w_n$ are equicontinuous (we apply Lemma \ref{lemma1} with $q:=\gamma_n$ for each $n\in\mathbb N$, employing the fact that resulting inequalities do not contain $q$). Hence we can select subsequences $w_{n_k}, \overline f_{n_k}$ such that $w_{n_k}\rightarrow w$ in $C_T$ and $\overline f_{n_k}\rightarrow \overline f$. Since $$ w_{n_k}(t)=w_{n_k}(0)+\int_0^t\left(\widetilde f(\tau)+\overline f_{n_k}-\gamma_{n_k}(w_{n_k}(\tau))\right)d\tau\,,$$ passing to the limit we obtain $$w(t)=w(0)+\int_0^t\left(\widetilde f(\tau) + \overline f -g\bigl(\varphi_{-1}(w(\tau))\bigr) \right)d\tau\,.$$ As the integrand in the right-hand-side is continuous, $w\in C^{1}_T$ and satisfies (\ref{wfg}) in $(0,T)$. Thus Step 2 is over. \noindent{\bf Step 3} {\it We prove that if $\overline f_1, \overline f_2 \in\mathbb R$ and the equation (\ref{f_eq}) with $f = \widetilde f+\overline f_i, i=1,2$ has a solution satisfying (\ref{f_con}), then $\overline f_1 = \overline f_2$}. Conversely, assume that $\overline f_1 > \overline f_2$ and there exists $w_i\in C^1_T, i=1,2$ such that $w'_i(t) +g\left(\varphi_{-1}(w_i(t))\right)=\widetilde f(t)+\overline f_i,$ $i=1,2$ subject to (\ref{f_con}). Then from (\ref{per_con}) it follows that the function $\varphi_{-1}\bigl(w_1(t)\bigr)-\varphi_{-1}\bigl(w_2(t)\bigr)$ is a $T$-periodic function with mean value zero, so that there exist $t_0$ and $\delta_1>0$ such that $\varphi_{-1}\bigl(w_1(t_0)\bigr)-\varphi_{-1}\bigl(w_2(t_0)\bigr)=0$ and $\varphi_{-1}\bigl(w_1(t)\bigr)-\varphi_{-1}\bigl(w_2(t)\bigr)<0$ for all $t_0\overline f_2$ and since the functions $g\left(\varphi_{-1}(w_1(\cdot))\right)$ and $g\left(\varphi_{-1}(w_2(\cdot))\right)$ are continuous, there exists $\delta_2>0$ such that $\left|g\left(\varphi_{-1}(w_1(t))\right) -g\left(\varphi_{-1}(w_2(t))\right)\right|<(\overline f_1-\overline f_2)/2$ for any $t_00$$ for all $t_00$. The solution of the linear initial value problem (\ref{df_eq}) has the explicit form: \begin{equation} \label{exp_sol} \omega(t)=\kappa\frac{1}{A(t)} + \frac{1}{A(t)} \int_0^t\widetilde\phi(s)A(s)ds + \sigma \frac{1}{A(t)}\int_0^t A(s)ds\,. \end{equation} Substituting (\ref{exp_sol}) into (\ref{df_con}) we obtain system of two linear equations for $\sigma$ and $\kappa$. This system can be uniquely solved if corresponding determinant $\mbox{D}$ is different from zero. By a straightforward calculation we obtain: \begin{eqnarray}\label{deter} \mbox{D}&=& \frac{1}{A(T)}\Big [\int_0^T A(s)ds \int_0^T\frac{\varphi_{-1}'\left(w_0(t)\right)}{A(t)}dt \\ &&- \int_0^T\frac{\varphi_{-1}'(w_0(t))}{A(t)}\int_0^tA(s)\,ds\,dt\Big]+ \int_0^T\frac{\varphi_{-1}'(w_0(t))}{A(t)}\int_0^tA(s)\,ds\,dt \nonumber\\ &=& \frac{1}{A(T)}\int_0^T A(s)\int_0^s\frac{\varphi'_{-1}(w_0(t))}{A(t)}\,dt\,ds + \int_0^T\frac{\varphi'_{-1}\bigl(w_0(t)\bigr)}{A(t)}\int_0^t A(s)\,ds\,dt\,. \nonumber \end{eqnarray} Since $\varphi_{-1}:I_2\rightarrow I_1$ is an increasing homeomorphism, it follows that $\varphi_{-1}'(z)>0$ a.e. in $I_2$. By a contradiction, one can show that if $\widetilde f\not\equiv 0$ then $w_0$ satisfying (\ref{f_eq})--(\ref{f_con}) is not a constant function. Thus $\varphi_{-1}'\left(w_0(t)\right)>0$ on some subset of $[0,T]$ of positive measure, which taking into account (\ref{deter}) implies $\mbox{D}>0$. Now let us consider $\widetilde f\equiv 0$. We have to impose $\varphi'(0)>0$ to conclude $D>0$. The reason consists in the fact that $w_0\equiv 0$ is the unique solution of (\ref{f_eq})--(\ref{f_con}) with $\widetilde f \equiv 0$. We proved that $G_{(w,\overline f,\alpha)}(w_0,\overline f_0,\alpha_0):C^1_T\times\mathbb R\times\mathbb R\rightarrow C^1_T\times\mathbb R\times\mathbb R$ is bijective linear mapping. Then the inverse of $G_{(w,\overline f,\alpha)}(w_0,\overline f_0,\alpha_0)$ is continuous due to the Banach open mapping theorem (see \cite{Zeid}). Thus $G_{(w,\overline f,\alpha)}(w_0,\overline f_0,\alpha_0)$ is an isomorphism of $C^1_T\times\mathbb R\times\mathbb R$ onto itself and the assumptions of the implicit function theorem \cite[Theorem 2.2.3]{Am_Pro} are satisfied. This completes the proof of Theorem \ref{Th2}. \hfill$\diamondsuit$ \paragraph{Proof of Theorem \ref{Th3}} At first we perform the proof under the assumptions {\rm (P)}, {\rm (P')} and {\rm (G')}. Considering (\ref{ODH1}) we can define continuously differentiable and bounded function $q$ satisfying $q(y)=g\bigl(\varphi_{-1}(y)\bigr)$ for all $|y|\leq\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}.$ Since the related first order problem (\ref{f_eq})--(\ref{N_con}) is the same as that one considered in \cite{CD}, the existence of the solution and the differentiability of $w(\widetilde f)$ and $s(\widetilde f)$ follows from \cite[Theorem 3.4]{CD}. The assumption {\rm(P')} can be omitted and the assumption {\rm(G')} can be replaced by {\rm(G)} using the same argument as in the Step 2 of the proof of Theorem \ref{Th1}. From \cite[Theorem 3]{Maw} we obtain the uniqueness of $\overline f$ corresponding to a fixed $\widetilde f$. The continuity of $s(\widetilde f)$ can be proved in the same manner as in Step 5 of the proof of Theorem \ref{Th1}. A priori bound (\ref{odh_s}) is a consequence of (\ref{ODH1}) (cf. Step 4 in the proof of Theorem \ref{Th1}). \hfill$\diamondsuit$ \section{Proof of Theorem \ref{Th4}} To prove Theorem \ref{Th4}, we need the following comparison results: \begin{lemma}\label{lem2} Let $v'(t)+g\biggl(\varphi_{-1}\bigl(v(t)\bigr)\biggr)\leq\alpha+ \widetilde f(t)$ on $[0,T]$ and $v$ satisfies (\ref{f_con}). Then $\alpha\geq s(\widetilde f)$ (where $s(\widetilde f)$ comes from Theorem \ref{Th1}). \end{lemma} \paragraph{Proof} Assume conversely that $\alpha0$ such that $\varphi_{-1}\bigl(v(t_0)\bigr)- \varphi_{-1}\bigl(w(t_0)\bigr)=0$ and $\varphi_{-1}\bigl(v(t)\bigr)- \varphi_{-1}\bigl(w(t)\bigr) > 0$ for all $t_0w(t)$ for all $t_00$ such that $$( v'- w')(t)\leq\alpha-s(\widetilde f)-g\biggl(\varphi_{-1}\bigl(v(t)\bigr)\biggr) +g\biggl(\varphi_{-1}\bigl(w(t)\bigr)\biggr)<0$$ for all $t: |t-t_0|<\delta'$. Since $v(t_0)=w(t_0)$, we obtain $v(t)-w(t)<0$ for all $t_00$, such that $v(t)\frac{w(t)-w(t_0)}{t-t_0}\,,$$ i.e. $ v'(t_0)\geq w'(t_0)$. On the other hand, we have \begin{eqnarray*} w'(t_0)&=&-g\bigl(\varphi_{-1}\bigl(w(t_0)\bigr)\bigr)+\widetilde f(t_0)+s(\widetilde f) \\ &=&-g\bigl(\varphi_{-1}\bigl(v(t_0)\bigr)\bigr)+\widetilde f(t_0) +s(\widetilde f)\\ &>&-g\bigl(\varphi_{-1}\bigl(v(t_0)\bigr)\bigr)+\widetilde f(t_0) +\alpha\geq v'(t_0)\,, \end{eqnarray*} which is a contradiction.\hfill$\diamondsuit$ \begin{remark} {\rm It is possible to show that the assertions of Lemmas \ref{lem2} and \ref{lem3} hold true also with inverted inequalities. This is in agreement with the semilinear periodic problem studied in \cite{Danc}. These inequalities are used to prove the `dual' version of Theorem \ref{Th4} with $h(+\infty)0$ and $p>1$: $c|z|^{p-1}\leq|\varphi(z)| \leq C(|z|^{p-1}+1)$ for all $z\in\mathbb R$. Then, for any $y\in C_T$ ( $y\in C[0,T]$ ), there exists exactly one $u\in C^1[0,T]$ with $\varphi(u')\in C^1[0,T]$ and satisfying \begin{equation} \label{p_eq} \bigl(\varphi(u'(t))\bigr)'-\varphi(u(t))=y(t)\quad t\in (0,T) \end{equation} subject to (\ref{per_con}) ( (\ref{Neu_con}) ). \end{lemma} \paragraph{Proof} For the sake of brevity we present the proof only for the periodic conditions. In the case of the Neumann conditions, the proof is analogous. At first we prove that, for any given $y\in C_T$, there exists precisely one weak solution of (\ref{p_eq})--(\ref{per_con}), where the weak solution of (\ref{p_eq})--(\ref{per_con}) is any function $u\in W^{1,p}_T$ such that the following identity \begin{equation} \label{w_eq} \int_0^T\left\{\varphi(u')v'+\varphi(u)v\right\}=-\int_0^Tyv \end{equation} is satisfied for each $v\in W^{1,p}_T$. Then, using an regularity argument, we show that this function $u$ is smooth enough and satisfies (\ref{p_eq})--(\ref{per_con}) in the sense indicated in the assertion of the lemma. {\it Existence and uniqueness of the weak solution. } Let us define $\psi:W_T^{1,p}\rightarrow L^{p'}\quad$ (where $p':=p/(p-1)$) by $u\mapsto \psi(u)$ if and only if $\langle\psi(u),v\rangle=\int_0^T\varphi(u')v'\,,$ $u,v\in W^{1,p}_T$ for all $v\in W^{1,p}_T$. Since $|\varphi(z)|\leq C(|z|^{p-1}+1)$ for all $z\in\mathbb R$, it is easy to verify that $\psi$ is a continuous operator; this follows from the Nemitskii theorem (see \cite[Theorem 1.2.2]{Am_Pro}). From the fact that $\varphi$ is an increasing function we get strict monotonicity of $\psi$. Since any monotone continuous operator is also hemicontinuous (see \cite{Zeid}), we get that $\psi$ is a hemicontinuous one. The assumption $|\varphi(z)|\geq c|z|^{p-1}$ for all $z\in\mathbb R$ implies that $\psi$ is weakly coercive. Now the existence and uniqueness of the weak solution of (\ref{p_eq})--(\ref{per_con}) follows from \cite[Theorem 32.H]{Zeid}. {\it Regularity.} Suppose that $u$ is a weak solution of (\ref{p_eq})--(\ref{per_con}). We show that $u\in C^1_T$, $\varphi(u')\in C^1[0,T]$ and that (\ref{p_eq}) holds pointwise in $(0,T)$. Integrating by parts we can rewrite the equation (\ref{w_eq}) into the form: \begin{eqnarray}\label{i_eq} \int_0^T \left(\varphi(u'(t))-\int_0^t\bigl[\varphi \bigl(u(\tau)\bigr)+y(\tau)\bigr]d\tau\right) v'(t)dt\, + && \\ \left[\int_0^{t}\bigl(\varphi\bigl(u(\tau)\bigr)+y(\tau)\bigr)d\tau\, v(t) \right]^T_0 &=& 0\,. \nonumber \end{eqnarray} Let us define a function $M:[0,T]\rightarrow \mathbb R$, $$t\mapsto\varphi(u')-\int_0^t \bigl[\varphi\bigl(u(\tau)\bigr)+y(\tau)\bigr]d\tau.$$ It is easy to see that $M\in L^{p'}$ and from (\ref{i_eq}) we get $$\int_0^TM(t)v'(t)dt=0\mbox{ for all } v\in C^{\infty}_0(0,T)\,.$$ Hence $$\int^T_0\frac{\delta M}{\delta t}v=0\mbox{ for all } v\in C^{\infty}_0(0,T),$$ where $\frac{\delta M}{\delta t}$ denotes the distributional derivative of $M$. Since $M\in L^{p'}\hookrightarrow L^1$ and $\frac{\delta M}{\delta t}=0$, we obtain that $M(t)=k$ a.e. in $[0,T]$ and $k\in \mathbb R$. Thus \begin{equation} \label{veinte} \varphi(u'(t))-\int_0^t\bigl[\varphi \bigl(u(\tau)\bigr)+y(\tau)\bigr]d\tau - k = 0\quad \mbox{a.e. in}\quad [0,T]\,. \end{equation} Since $\varphi$ is an increasing homeomorphism of $\mathbb R$ onto itself, we can rewrite the previous equation into the following form \begin{equation}\label{phi_eq} u'(t)-\varphi_{-1}\left(\int_0^t\bigl[\varphi \bigl(u(\tau)\bigr)+y(\tau)\bigr]d\tau - k\right) = 0\,. \end{equation} Now let us define a function $F:\mathbb R\times [0,T]\rightarrow\mathbb R,$ $$(z,t)\mapsto z-\varphi_{-1}\left(\int_0^t\bigl[\varphi \bigl(u(\tau)\bigr)+y(\tau)\bigr]d\tau - k\right)\,.$$ It is possible to show that $F$ is continuous on $\mathbb R\times [0,T]$. Moreover, $F(\cdot, t_0)$ is an increasing function for all $t_0\in [0,T]$, and $$\lim_{z\rightarrow -\infty}F(z,t_0)=-\infty, \quad \lim_{z\rightarrow +\infty}F(z,t_0)=+\infty\,.$$ Hence for each $t\in [0,T]$ there exists exactly one $z(t)\in\mathbb R$, such that $$F(z(t),t)=0\,.$$ Since $\frac{\partial F}{\partial z}$ is continuous and $\frac{\partial F}{\partial z} = 1 \not = 0$, we can apply the implicit function theorem to show that $z(\cdot)\in C[0,T]$. From (\ref{phi_eq}) we get $$F(u'(t),t)=0\quad\mbox{a.e. in}\quad [0,T]\,.$$ Thus we arrive at $$z(t)=u'(t)\quad\mbox{a.e. in}\quad [0,T]\,.$$ Since $u\in W_T^{1,p}$ is absolutely continuous, this identity holds true for all $t\in [0,T]$ and thus $u\in C^1[0,T]\cap W^{1,p}_T.$ Now it remains to prove that $\varphi(u')\in C^1[0,T].$ Let us define $G:\mathbb R\times [0,T]\rightarrow\mathbb R$, $$(z,t)\mapsto z -\int_0^t\biggl(\varphi(u(\tau)) - y(\tau)\biggr)d\tau - k\,.$$ The function $G$ is continuous on $\mathbb R\times [0,T]$. Moreover, for all $t_0\in [0,T]$, $G(\cdot, t_0)$ is an increasing function and $\lim_{z\rightarrow \pm\infty}G(z,t_0)=\pm\infty$. Hence for each $t\in[0,T]$ there exists exactly one $z(t)$ such that $$G\bigl(z(t),t\bigr)=0\,.$$ Since $u\in C^1[0,T]$ and $y\in C[0,T]$, the partial derivatives $\frac{\partial G}{\partial z}$ and $\frac{\partial G}{\partial t}$ are continuous; moreover $\frac{\partial G}{\partial z}=1\not = 0$. Then the implicit function theorem yields $z(t)\in C^1[0,T]$. Taking into account (\ref{veinte}) and the fact that $u\in C^1[0,T]$, we see that $z(t)=\varphi(u')$ for all $t\in [0,T]$; thus $\varphi(u')\in C^1[0,T]$. Now, since $u\in C^1[0,T]\cap W^{1,p}_T$ and $\varphi(u')\in C^1[0,T]$, integrating (\ref{w_eq}) by parts we show that $u$ satisfies (\ref{p_eq}) in $(0,T)$ and that $u'(0)=u'(T)$. This concludes the proof. \hfill$\diamondsuit$\medskip Now we can define a solution operator $K: C_T\rightarrow C^1_T$, \begin{equation} \label{define_K} y \mapsto K(y)\,, \end{equation} where $K(y)$ is the solution of (\ref{p_eq})--(\ref{per_con}). Analogously we can define a solution operator $K': C[0,T]\rightarrow C^1[0,T]$ corresponding to the Neumann problem (\ref{p_eq})--(\ref{Neu_con}). \begin{lemma}\label{lem5} Let $K$ and $K'$ be defined as above. Then $K$ is compact as a mapping between $C_T$ and $C^1_T$ and $K'$ is compact as a mapping between $C[0,T]$ and $C^1[0,T]$. \end{lemma} \paragraph{Proof} We prove the compactness of $K$. The proof of the compactness of $K'$ is analogous. Let us consider the following sequence of equations: \begin{equation}\label{s_eq} \left(\varphi(u'_n(t))\right)'-\varphi(u_n(t)) = y_n(t)\,, \end{equation} subject to (\ref{per_con}), where $\{y_n\}_{n=1}^{\infty}\subset C[0,T]$ is bounded. We are going to show that one can select a convergent subsequence from $\{u_n\}_{n=1}^{\infty}\subset C^1_T$. Multiplying the equation (\ref{s_eq}) by $u_n$, integrating from $0$ to $T$, integrating the first term in the left-hand-side by parts and using the periodic conditions (\ref{per_con}), we obtain \begin{equation}\label{s_int} -\int_0^T\bigl\{\varphi(u'_n(t))u'_n(t)+ \varphi(u_n(t))u_n(t)\bigr\}dt=\int_0^T y_n(t)u_n(t)dt\,. \end{equation} Since we assume that $|\varphi(z)|\geq c|z|^{p-1}$ for all $z\in\mathbb R$, we find that $$ \|u_n\|^p_{W^{1,p}_T}\leq \frac{1}{c}\int_0^T\bigl\{\varphi(u'_n(t)) u'_n(t)+ \varphi(u_n(t))u_n(t)\bigr\}dt\,. $$ Using this we estimate the terms on the left-hand side of (\ref{s_int}). The right hand-side of (\ref{s_int}) is estimated by the H\"older inequality. Therefore we find that $$\|u_n\|^p_{W^{1,p}_T}\leq\frac{1}{c}\|y_n\|_{L^{p'}}\|u_n\|_{L^p}\leq \frac{1}{c}\|y_n\|_{L^{p'}}\|u_n\|_{W^{1,p}_T}\,,$$ which implies \begin{equation}\label{w_est} \|u_n\|_{W^{1,p}_T}\leq\frac{1}{c}\left(\|y_n\|_{L^{p'}}\right)^{\frac{1}{p-1}}\,. \end{equation} Thus $\{u_n\}^{\infty}_{n=1}$ is bounded in $W^{1,p}_T$. Since $W^{1,p}_T$ is compactly imbedded in $C_T$, we can select $\{u_{n_k}\}_{k=1}^{\infty}$ such that $u_{n_k}\rightarrow w$ in $C_T$. Let $h_k(t)=y_{n_k}(t)-\varphi\bigl(u_{n_k}(t)\bigr)$. One can see that there exists $\beta>0$ such that $\|h_k\|_{C}\leq\beta$. Hence $\|\left(\varphi\left(u'_{n_k}\right)\right)'\|_{C}\leq\beta$. It follows from (\ref{per_con}) that there exists $t^k_0\in [0,T]$, such that $u'_{n_k}(t^k_0)=0$. From (\ref{s_eq}) we obtain $\varphi(u'_{n_k}(t))=\int_{t_0^k}^t h_k(\tau)d\tau$ and consequently $\|\varphi(u_{n_k}'(t))\|_{C}\leq T\beta$. Thus $\varphi(u'_{n_k}(t))$ is bounded in $C^1_T$ norm. Due to the compact imbedding of $C^1_T$ into $C_T$ we can select $\varphi(u_{n_{k_j}}')\rightarrow v$ in $C_T$. Since $u_{n_{k_j}}'=\varphi_{-1}\bigl(\varphi(u'_{n_{k_j}})\bigr)$ and $\varphi_{-1}$ is continuous, $u'_{n_{k_j}}\rightarrow \varphi_{-1}(v)$ in $C_T.$ On the other hand, $u_{n_{k_j}}$ can be written in the form $$u_{n_{k_j}}(t)=u_{n_{k_j}}(0)+\int_0^t\varphi_{-1}\left(\varphi(u'_{n_{k_j}}(\tau))\right)d\tau\,. $$ Since $u_{n_{k_j}}\rightarrow w$ as $n_{k_j}\rightarrow$ in $C_T$, we find that $w(t)=w(0)+\int_0^t\varphi_{-1}\left(v(\tau)\right)d\tau$. As the integrand is continuous, differentiating the former equation we get $w'=\varphi_{-1}(v)$. Thus $u_{n_{k_j}} \rightarrow w$ in $C^1_T$. This ends the proof. \hfill$\diamondsuit$ \begin{remark} {\rm The proof of the previous lemma is based on the ideas from \cite{DelP}.} \end{remark} Let $g,h,\widetilde f, T$ be as in Theorem \ref{Th4} and let $z_0:= 2\sup_{\xi\in\mathbb R}|g(\xi)| + |h(+\infty)| + |h(-\infty)|+2\sup_{\xi\in\mathbb R}|h(\xi)|+\|\widetilde f\|_{C}$. We define a function $l:\mathbb R\rightarrow\mathbb R$ by \begin{equation} \label{def_l} l(z):=\left\{ \begin{array}{ll} z & \mbox{for}\ 0\leq z\leq z_0\,, \\ z_0 & \mbox{for}\ z>z_0\,, \end{array}\right. \end{equation} for $z\geq 0$ and $l(z)=-l(-z)$ for $z<0$. We also define: \begin{equation}\label{dfm} \Gamma(p,c, F) := \Big|\sqrt{\frac{T}{3}}\frac{F}{c}\Big|^{p-1} \end{equation} for any $p>1$, $c>0$, $F\geq 0$. \begin{lemma} \label{lem6} Let $\varphi$ be an increasing homeomorphism of $\mathbb R$ onto $\mathbb R$ and there exist $c,C>0$ and $p>1$: $c|z|^{p-1}\leq |\varphi(z)|\leq C\left(|z|^{p-1}+1\right)$. Let {\rm (G)} be satisfied and $g$ be bounded. Then, for any $\lambda\in [0,1]$ and $|\overline f|\leq\sup_{\xi\in\mathbb R}|h(\xi)|+\sup_{\xi\in\mathbb R}|g(\xi)|$, all solutions of \begin{equation} \label{lam_eq} \bigl(\varphi(u')\bigr)'+\lambda g(u')+l(u)=\lambda (\widetilde f+\overline f) \end{equation} subject to (\ref{per_con}) ( (\ref{Neu_con}) ) are a priori bounded by \newline $\|u\|_{C^1}\leq (1+2T)\,\Gamma\left(p,c, \|\widetilde f \|_{L^2}+\sqrt{T}z_0\right)+z_0$. \end{lemma} \paragraph{Proof} We rewrite the equation (\ref{lam_eq}) as $$\bigl(\varphi(u')\bigr)'+ \lambda g(u')=\lambda (\widetilde f+\overline f)-l(u)\,.$$ Then it follows from (\ref{ODH1}) (consider $q:=\lambda g$ in Lemma \ref{lemma1}) that any solution $u$ of (\ref{lam_eq})--(\ref{per_con}) satisfy $$ \|\varphi(u')\|_{C}\leq\sqrt{\frac{T}{3}}\left\|\lambda \widetilde f-l(u)\right\|_{L^2}\,. $$ Taking into account the assumption $|\varphi(z)|\geq c|z|^{p-1}$, the following inequality $$c |u'|^{p-1}\leq |\varphi(u')|\leq\sqrt{\frac{T}{3}}\left\|\lambda\widetilde f-l(u)\right\|_{L^2}$$ is satisfied for any $t\in [0,T]$. This implies $$ \|u'\|_{C}\leq\left(\frac{1}{c}\sqrt{\frac{T}{3}}\left\|\lambda \widetilde f-l(u)\right\|_{L^2}\right)^{\frac{1}{p-1}}\,. $$ Since $\|\lambda\widetilde f-l(u)\|_{L^2}\leq\|\widetilde f\|_{L^2}+\sqrt{T}z_0$ (recall that $\|\lambda \widetilde f\|_{L^2}=|\lambda|\|\widetilde f\|_{L^2}$, where $\lambda\in [0,1]$ and $\|l(u)\|_{L^2}\leq\sqrt{T}\sup_{\xi\in\mathbb R}|l(\xi)|=\sqrt{T}z_0$), we get \begin{equation} \label{iuy} \|u'\|_{C}\leq\Gamma\left(p,c,\|\widetilde f\|_{L^2}+\sqrt{T}z_0\right)\,. \end{equation} Let us split the function $u$ as $u=\widetilde u+\overline u$, where $\overline u = \frac{1}{T}\int_0^Tu(t)dt$. As $\left\|\widetilde u\right\|_{C}\leq\int_0^T\left|u'\right|\leq T\left\| u'\right\|_{C}$, from (\ref{iuy}) we have \begin{equation} \label{iduy} \|\widetilde u\|_{C}\leq T\,\Gamma\left(p,c,\|\widetilde f\|_{L^2}+\sqrt{T}z_0\right)\,. \end{equation} Now rewrite (\ref{lam_eq}) as $$\bigl(\varphi(u')\bigr)'=\lambda \left(\widetilde f+\overline f - g(u')\right)-l(u)$$ and suppose that $\overline u>z_0+ T\,\Gamma\left(p,c,\|\widetilde f\|_{L^2}+\sqrt{T}z_0\right)$. Then (\ref{iduy}) implies $u(t)>z_0$ for all $t\in [0,T]$ and, by the definition of $l$, we have that $l(u)=z_0$ for all $t\in [0,T]$. Hence we can rewrite (\ref{lam_eq}) as $$ \left(\varphi(u')\right)'=\lambda\left(\widetilde f+\overline f-g(u')\right)-z_0\,. $$ Integrating from $0$ to $T$, using $g(u')<\sup_{\xi\in\mathbb R}|g(\xi)|$ and $\widetilde f<\|\widetilde f\|_{C}$ we get that $$\varphi(u'(T))<\left(\lambda(\|\widetilde f\|_{C}+\overline f+\sup_{\xi\in\mathbb R}|g(\xi)|)-z_0\right)T + \varphi(u'(0))\,.$$ Since $\lambda\in [0,1]$ and $z_0>\sup_{\xi\in\mathbb R}|g(\xi)|+\|\widetilde f\|_{C}+\overline f$, we find that $\varphi(u'(T))<\varphi(u'(0))$, which contradicts the periodic conditions. Analogously as above, the possibility $\overline u<-z_0-T\,\Gamma\left(p,c,\|\widetilde f\|_{L^2}+\sqrt{T}z_0\right)$ leads to a contradiction. Hence $|\overline u|0$ and $p>1$: $c|z|^{p-1}\leq|\varphi(z)|\leq C(|z|^{p-1}+1)$ for all $z\in\mathbb R$, $g$ satisfies {\rm(G)} and is bounded. Let us define $V : C^1_T\rightarrow C^1_T$ (~$C^1[0,T]\rightarrow C^1[0,T]$~) by $$u\mapsto u-K\bigl(-\varphi(u)-g(u')-l(u)+f\bigl)\,, $$ where the operator $K$ is defined by (\ref{define_K}) for the periodic ( the Neumann ) BVP. Then there exists $R_0>0$ such that, if $V(u)=0$ then $\|u\|_{C^1}\leq R_0$. Moreover the Leray--Schauder degree $\deg (V, B(0,R), 0)$ is well defined and non-zero if $B(0,R)=\{u\in C^1_T \ \mbox{( }C^1[0,T]\mbox{ )}$: $\|u\|_{C^1} R_0$. \end{lemma} \paragraph{Proof} Let us define $U :C^1_T\times [0,1]\rightarrow C^1_T$ by $$(u,\lambda) \rightarrow u-K\bigl(-\varphi(u)-\lambda g(u')-l(u)+\lambda f\bigl)\,.$$ Since, by Lemma \ref{lem5}, $K: C_T\rightarrow C^1_T$ is compact and $$-\varphi(u)-\lambda g(u')-l(u)+\lambda f$$ is an continuous operator from $C^1_T$ to $C_T\,$, $U(\cdot,\lambda)$ is a compact perturbation of the identity of $C^1_T$ onto itself for all $\lambda\in [0,1]$. Furthermore, it follows that $V=U(\cdot,1)$ from the definition of $U$. It is easy to see that the operator equation $U(\cdot,\lambda)=0$ is equivalent to the equation (\ref{lam_eq}): $$ \bigl(\varphi(u')\bigr)'+\lambda g(u')+l(u)=\lambda f $$ subject to (\ref{per_con}). Due to Lemma \ref{lem6}, for any $\lambda\in [0,1]$, all solutions of (\ref{lam_eq})--(\ref{per_con}) satisfy $\|u\|_{C^1}\leq (1+2T)\Gamma\left(p,c, \|\widetilde f\|_{L^2}+\sqrt{T}z_0\right)+z_0=:R_0$. Hence the degree is well defined for every ball $B(0,R)$ with radius $R>R_0$. Since $\varphi$ and $l$ are odd functions, $U(\cdot,0)$ is odd mapping; consequently $$\deg (M(\cdot,0),B(0,R),0)\not=0\,.$$ Finally, the homotopy invariance property of the degree implies that $$\deg (U(\cdot,\lambda),B(0,R),0)\not=0\mbox{ for all }\lambda\in[0,1].$$ \hfill$\diamondsuit$\medskip Now, we prove Theorem \ref{Th4}. The method follows the idea of the proof of Theorem 2 in \cite{Danc}. For the sake of brevity we will present a proof for the periodic conditions. The argument for the Neumann conditions is analogous. \paragraph{Proof of Theorem \ref{Th4}} The proof is divided into three steps: \noindent{\bf Step 1} {\it First, we will prove the theorem under additional assumptions: $\mathop{\rm Dom}\varphi = \mathbb R$, there exists $c,C>0$ and $p>1$ such that $c|z|^{p-1}\leq |\varphi(z)|\leq C(|z|^{p-1}+1)$ for all $z\in\mathbb R$ and $g$ is bounded.} We use the degree argument. Define the operator $N:C^1_T\times\mathbb R\rightarrow C^1_T$ by $$(u,\lambda)\mapsto u-K\left(-\varphi(u)- g(u')-\lambda h(u)-(1-\lambda)l(u) + \widetilde f+\overline f\right),$$ where $K:C_T\rightarrow C^1_T$ is introduced after Lemma \ref{lem4} and $l:\mathbb R\rightarrow\mathbb R$ is defined by (\ref{def_l}). Now we can rewrite the boundary-value problem (\ref{fund_eq})--(\ref{per_con}) into an equivalent operator equation $N(u,1)=0$. From the definition of $N$ it follows that $N(\cdot,\lambda)$ is a compact perturbation of the identity of $C^1_T$ onto itself for all $\lambda\in [0,1]$. We start our homotopy argument with $\lambda=0$. Then $N(\cdot,0)=V$, where $V$ is defined in Lemma \ref{lem7}. Hence, by Lemma \ref{lem7}, there exists $R_0>0$ such that if $u$ is a solution of $N(u,0)=0$ then $\|u\|_{C^1}\leq R_0$; moreover, $\deg (N(\cdot,0), B(0,R), 0)\not = 0$ provided that $R>R_0$. We show later that there exists $R_0'> R_0>0$ such that, for all $\lambda\in [0,1]$, every solution $u$ of the operator equation $N(u,\lambda)=0$ satisfies $\|u\|_{C^1}0$, independent of $\lambda$, such that $\|u'\|_{C}\leq A$. Therefore $\|\widetilde u\|_{C}\leq AT$ and it remains to prove that there exists a constant $j>0$, independent of $\lambda$ such that $|\overline u|\leq j$. Let $\gamma := \frac{1}{2}\left(h(+\infty)+s(\widetilde f)-\overline f\right)$; by (\ref{D_like}) $\gamma>0$. Hence we can find $m>0$ (independent of $\lambda$) such that \begin{equation}\label{p_int} (1-\lambda)h(y)+\lambda l(y)>h(+\infty)-\gamma \end{equation} for all $y\geq m$ and $0\leq \lambda\leq 1$. Assuming that $\overline u>m+AT$, we find $$\left(\varphi(u')\right)' + g(u') = \overline f + \widetilde f - (1-\lambda) h(u)- \lambda l(u) <\overline f + \widetilde f - h(+\infty) + \gamma = s(\widetilde f) - \gamma + \widetilde f.$$ Taking $\alpha=s(\widetilde f)-\gamma$ we see from Lemma \ref{lem2} that $s(\widetilde f) - \gamma\geq s(\widetilde f)$, a contrary to $\gamma>0$. Thus $\overline u\leq m+AT$. Similarly, we exclude the possibility $\overline u\geq -m-AT$. Hence $\|u\|_{C}\leq AT+m =: j$ and $\|u\|_{C^1}\leq A(1+T)+m$, where $A$ and $m$ do not depend on $\lambda$. So that the desired radius is any $R_0'> \max\{A(1+T)+m, R_0\}$. \noindent{\bf Step 2} {\it Now we remove the additional assumptions on $\varphi$ and $g$.} Let us take fixed $\widetilde f\in C_T$ and define $$ M:=\varphi_{-1}\Big(\sqrt{\frac{T}{3}}\big(\|\widetilde f\|_{L^2}+ \sqrt{T}\sup_{\xi\in\mathbb R}|h(\xi)|\big)\Big)>0\,. $$ Note that (\ref{M_con}) imply that $0<\varphi(M)\varphi_{-1}\left(\sqrt{\frac{T}{3}}\|\widetilde f\|_{L^2}\right)$ yield $q(\varphi_{-1}(v))=g(\varphi_{-1}(v))$. So that any solution to (\ref{adjkf}) satisfies also \begin{eqnarray}\label{lj6rl} &v'+g(\varphi_{-1}(v))=\widetilde f + s(\widetilde f)\,,& \\ &\int_0^T\varphi_{-1}(v(t))dt=0\,,\quad v(0)=v(T)&\nonumber \end{eqnarray} and vice versa. Integrating equations in (\ref{adjkf}) and (\ref{lj6rl}),respectively, from $0$ to $T$ and taking into account corresponding boundary conditions, we arrive at $$s(\widetilde f)=\frac{1}{T}\int_0^T q\left(\psi_{-1}(v(t))\right)\,dt=\frac{1}{T} \int_0^T q\left(\varphi_{-1}(v(t))\right)\,dt=r(\widetilde f)\,.$$ Thus (\ref{dajkf}) becomes (\ref{D_like}): $$s(\widetilde f)+h(-\infty)<\overline f\eta\quad \mbox{ implies }\quad F(\cdot, \xi, \cdot)>F(\cdot, \eta,\cdot)\,, \end{equation} then there exists at most one solution of the BVP: \begin{eqnarray}\label{equ_incp} &\left(\varphi(u')\right)'=F(t,u,u')\,,\quad t\in(0,T)& \\ &u(0)=u(2\pi),\quad u'(0)=u'(2\pi)\,. &\nonumber \end{eqnarray} \end{proposition} \paragraph{Proof} If $u_1,\,\,\, u_2$ are two distinct solutions of (\ref{equ_incp}) then \begin{equation} \label{substr_eq} \left(\varphi(u_1')\right)'-\left(\varphi(u_2')\right)' = F(t,u_1,u_1')-F(t,u_2,u_2')\,. \end{equation} Set $z=u_1-u_2$; obviously $z\in C[0,T]$ and hence $z$ attains its maximum value, say at $t_M$ on [0,T]. Without lost of generality, we can suppose that $u_1(t_M)>u_2(t_M)$. This implies $\max_{t\in [0,T]}z(t)>0$. Since $z(0)=z(T)$, we may assume $t_M\in [0,T)$. Taking into account that $z\in C^1[0,T]$ and that $z$ satisfies (\ref{per_con}), it follows that $z'(t_M)=0$, which is equivalent to $u_1'(t_M)=u'_2(t_M)$. Upon integrating from $t_M$ to $t$, the equation (\ref{substr_eq}) becomes $$\varphi\left(\rule{0pt}{3mm}u_1'(t)\right)-\varphi\left(\rule{0pt}{3mm}u_2'(t)\right) = \int_{t_M}^{t} \left(\rule{0pt}{5mm} F\left(\rule{0pt}{3mm}\tau,u_1\left(\tau\right), u_1'\left(\tau\right)\right)-F\left(\rule{0pt}{3mm}\tau,u_2\left(\tau\right),u_2'\left(\tau\right) \right)\right)d\tau\,.$$ Considering $u_1(t_M)>u_2(t_M)$, $u_1'(t_M)=u_2'(t_M)$, the continuity of \newline $F\left(\rule{0pt}{3mm}t,u_1(t),u_1'(t)\right)-F\left(\rule{0pt}{3mm}t,u_2(t),u_2'(t)\right)$ and (\ref{con_inc}), we find that there exists $\delta>0$ such that $$\varphi(u_1'(t))-\varphi(u_2'(t))>0 \mbox{ for all } t\in(t_M,t_M+\delta)\,.$$ Hence $u'_1(t)>u_2'(t)$ for all $t\in(t_M, t_M+\delta)$, which implies $z'(t)>0$ for all $t\in (t_M, t_M+\delta)$, a contrary to $z(t_M)\geq z(t)$ for all $t\in [0,T]$. \hfill$\diamondsuit$ \begin{proposition} \label{prop_un_Neu} Suppose that $F: [0,T]\times\mathbb R^2\rightarrow\mathbb R$ is continuous and satisfies (\ref{con_inc}). Then the BVP \begin{eqnarray}\label{equ_incn} &\left(\varphi(u')\right)'=F(t,u,u')\,,\quad t\in(0,T)& \\ &u'(0)=0,\quad u'(T)=0 & \nonumber \end{eqnarray} admits at most one solution. \end{proposition} \paragraph{Proof} The only difference to the proof of Proposition \ref{prop_un} is that we have to exclude the possibility $t_M=T$. Suppose that $z=u_1-u_2$ attains its maximum at $T$. Integrating (\ref{equ_incn}) we find that $$\varphi(u_1'(t))-\varphi(u_2'(t)) = - \int_{t}^{T} \left(\rule{0pt}{5mm} F\left(\rule{0pt}{3mm}\tau,u_1\left(\tau\right), u_1'\left(\tau\right)\right)-F\left(\rule{0pt}{3mm}\tau,u_2\left(\tau\right),u_2'\left(\tau\right) \right)\right)d\tau\,.$$ Taking into account $u_1(T)>u_2(T)\,,\quad u_1'(T)=u'_2(T)=0\,,$ the continuity of $F(t,u_1(t),u_1'(t))-F(t,u_2(t),u_2'(t))$ and (\ref{con_inc}), we find that there exists $\delta>0$ such that $$\varphi(u_1'(t))-\varphi(u_2'(t))<0 \mbox{ for all } t\in(T-\delta,T)\,.$$ Hence $u'_1(t)-u'_2(t)<0$ for all $t\in (T-\delta,T)$, which implies $z'(t)<0$ for all $t\in (T-\delta,T)$, a contrary to $z(T)\geq z(t)$ for all $t\in [0,T]$. \hfill$\diamondsuit$ \medskip Taking a particular function $F(t, u, v) = -h(u) - g(v) + f$, where $h\in C(\mathbb R, \mathbb R)$, is bounded and strictly decreasing, $g\in C(\mathbb R, \mathbb R)$ and $f\in C[0,T]$, we obtain uniqueness for the BVPs (\ref{fund_eq})--(\ref{per_con}) and (\ref{fund_eq})--(\ref{Neu_con}) studied previously: \begin{theorem} \label{uniquenessTH} Suppose {\rm (P)}, {\rm (G)}, $f\in C_T$ ( $f\in C[0,T]$ ), $h\in C(\mathbb R,\mathbb R)$ and moreover impose that $h$ is strictly decreasing. Then the BVP (\ref{fund_eq})--(\ref{per_con}) ( (\ref{fund_eq})--(\ref{Neu_con}) ) admits at most one solution. \end{theorem} We would like to point out that the monotonicity assumption, $h$ being decreasing, was essential to prove the uniqueness. The following simple example shows that there are no analogous statements to the previous theorem assuming $h$ increasing, even not for semilinear BVP: \begin{example} \label{un_cex} {\rm Consider BVP: \begin{equation} \begin{array}{cr} \label{un_cex_eq} u''(t)+h(u(t))=\cos 2t\,,&\quad t\in(0,T) \\ u(0)=u(2\pi),\quad u'(0)=u'(2\pi) & \end{array} \end{equation} with $$ h(s)\left\{ \begin{array}{lc} s \quad & \mbox{for } |s|\leq 1 \\ (2-\frac{1}{|s|}){\rm sgn}(s) \quad & \mbox{for } |s|> 1 \end{array} \right. $$ then $u=-\frac{1}{3}\cos(t) + c\sin(t+\phi)$, with parameters $|c|\leq\frac{2}{3}$ and $\phi\in [0,2\pi)$, is a two dimensional subset of the set of all solutions to (\ref{un_cex_eq}).} \end{example} For some particular results in this direction (the semilinear problem) we refer to \cite{Danc}. \paragraph{Acknowledgment.} The author wish to thank to Prof. Dr\'abek for his advice while writing this paper and to Prof. Man\'asevich for stimulating discussion during the conference EDPCHILE '99. The author would also like to thank to referees for their suggestions and for including references \cite{HMZ1}, and \cite{HMZ2}. \begin{thebibliography}{99} \bibitem{Am_Pro} Ambrosetti, A., and Prodi, G.,\quad {\it A Primer of Nonlinear Analysis}, Cambridge University Press, Cambridge, 1993. \bibitem{ArcOrs} Arcoya, D., and Orsina, L.,\quad Landesman--Lazer conditions and quasilinear elliptic equations, Nonlinear Anal., {\bf 28} (1997), 1623-1632. \bibitem{BoDrKu} Boccardo, L., Dr\'abek, P., and Ku\v cera, M.,\quad Landasman-Lazer conditions for strongly nonlinear boundary-value problems, Comment. Math. Univ. Carolin., {\bf 30} (1989), 411-427. \bibitem{CD} Ca\~nada, A., and Dr\'abek, P.,\quad On semilinear problems with nonlinearities depending only on derivatives, SIAM J. Math. Anal., {\bf 27} (1996), 543-557. \bibitem{CodLew} Coddington E.~A., Levinson, N.,\quad {\it Theory of ordinary differential equations}, McGraw-Hill, New York, 1955. \bibitem{Danc} Dancer, E.~N.,\quad On the ranges of certain damped nonlinear differential equation, Ann. Mat. Pura Appl., {\bf 119} (1979), 281-295. \bibitem{DelP} del Pino, M.,\quad {\it Acerca de un problema cuasilineal de secundo orden}, Diploma Thesis, Universidad de Chile, Santiago de Chile, 1987. \bibitem{Fuc} Fu\v c\'\i k, S., \quad {\it Solvability of Nonlinear Equation and Boundary Value Problems}, D. Reidel Publishing Company, Dordrecht, 1980. \bibitem{HMZ1} Garc\'{\i}a-Huidobro, M., Man\'asevich, R., and Zanolin, F.,\quad A Fredholm-like Result for Strongly Nonlinear Second Order ODE's, {\it J. Differential Equations}, {\bf 114} (1994), 132-167. \bibitem{HMZ2} Garc\'{\i}a-Huidobro, M., Man\'asevich, R., and Zanolin, F.,\quad Strongly Nonlinear Second Order ODE's with Unilateral Conditions, {\it Differential and Integral Equations}, {\bf 6} (1993), 1057-1078. \bibitem{KaSe} Kannan, R., and Seikkala, S., \quad Existence of solutions to a periodic boundary-value problem, in {\it `Function Spaces, Differential Operators and Nonlinear Analysis'} (Vesa Mustonen, Ji\v r\'{\i} R\'akosn\'\i k, editors), Mathematical Institute AS \v CR, Prague, 1999, 149-156. \bibitem{KotKub} Kotek, Z., Kub\'\i{k}, S.,\quad {\it Nonlinear circuits} (in Czech), SNTL, Prague, 1962. \bibitem{LanLaz} Landesman, E.~M., and Lazer, A.~C.,\quad Nonlinear perturbations of linear elliptic boundary-value problems at resonance, J. Math. Mech., {\bf 19} (1970), 609-623. \bibitem{ManMaw} Man\'asevich, R., and Mawhin, J.,\quad Periodic solutions for nonlinear systems with p-Laplacian-like operators, J. Differential Equations, {\bf 145} (1998), 367-393. \bibitem{MawW} Mawhin, J., Willem, M.,\quad {\it Critical point theory and Hamiltonian systems}, Springer Verlag, 1989. \bibitem{Maw} Mawhin, J.,\quad Some remarks on semilinear problems at resonance where the nonlinearity depends only on the derivatives, Acta Math. Inform. Univ. Ostraviensis, {\bf 2} (1994), 61-69. \bibitem{WFA} Ames, W. F., \quad {\it Nonlinear partial differential equations in engineering}, Academic press, New York, 1965. \bibitem{Zeid} Zeidler, E.,\quad {\it Nonlinear Functional Analysis and its Applications I}, Springer-Verlag, Berlin, 1986. \end{thebibliography} \noindent{\sc Petr Girg} \\ Department of Mathematics \\ University of West Bohemia \\ P. O. Box 314 \\ 306 14 Plze\v n, Czech Republic \\ e-mail: pgirg@kma.zcu.cz \end{document}