\documentclass[reqno]{amsart} \usepackage{graphicx} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2004(2004), No. 72, pp. 1--24.\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 2004 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2004/72\hfil Exact multiplicity results] {Exact multiplicity results for a $p$-Laplacian positone problem with concave-convex-concave nonlinearities} \author[Idris Addou \& Shin-Hwa Wang\hfil EJDE-2004/72\hfilneg] {Idris Addou \& Shin-Hwa Wang} \address{Idris Addou \hfill\break D\'{e}partement de Math\'{e}matiques et Statistiques, Universit\'{e} de Montr\'{e}al\\ C.P. 6128, Succ. Centre-ville, Montreal, Quebec, Canada, H3C2J7} \email{addou@dms.umontreal.ca} \address{Shin-Hwa Wang \hfill\break Department of Mathematics, National Tsing Hua University\\ Hsinchu, Taiwan 300, Republic of China} \email{shwang@math.nthu.edu.tw} \date{} \thanks{Submitted March 8, 2004. Published May 20, 2004.} \subjclass[2000]{34B18, 34B15} \keywords{Exact multiplicity result, $p$-Laplacian, positone problem, bifurcation, \hfill\break\indent concave-convex-concave nonlinearity, positive solution, dead core solution, time map} \begin{abstract} We study the exact number of positive solutions of a two-point Dirichlet boundary-value problem involving the $p$-Laplacian operator. We consider the case $p=2$ and the case $p>1$, when the nonlinearity $f$ satisfies $f(0)>0$ (positone) and has three distinct simple positive zeros and such that $f''$ changes sign exactly \emph{twice} on $(0,\infty )$. Note that we may allow $f''$ to change sign more than twice on $(0,\infty )$. We also present some interesting examples. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{remark}[theorem]{Remark} \allowdisplaybreaks \section{Introduction} In this paper we present \emph{exact} multiplicity results of \emph{positive} \emph{solutions} for the nonlinear two-point Dirichlet boundary-value problem \begin{equation} \begin{gathered} -( \varphi _{p}(u'(x))) '=\lambda f(u(x)),\; -11$, $\varphi _{p}(y)=| y| ^{p-2}y$ and $( \varphi_{p}(u')) '$ is the one-dimensional $p$-Laplacian$, \lambda >0$ and $f$ is a concave-convex-concave nonlinearity. Precise conditions are listed below. This paper is intended as a second part of a previous paper by the present authors \cite{a3}. In fact, whereas the previous paper was a study of \eqref{1.1} with $f\in C^{2}[0,\infty )$ satisfying $f(0)=0$ and has \emph{two} distinct simple positive zeros $b0$ (positone) and has \emph{three} distinct simple positive zeros $a1$? None known results in the literature prove or disprove this feature. This paper provide an example which disproves this fact. Indeed, for some $p>1$, $p\neq 2$, \emph{we have obtained some positive solutions of \eqref{1.1} such that the interior zero set of their derivative is not a connected set.} Note that this situation is not known even for \eqref{1.1} when $f(0)=0$ in our previous paper Addou and Wang \cite{a3}. For $p=2$, $( \varphi _{p}(u')) '=u''$, and \eqref{1.1} reduces to \begin{equation} \begin{gathered} -u''(x)=\lambda f(u(x)),\text{ }-12b-a$ $( \Leftrightarrow \int_{a}^{c}f(u)du>0) $, and a certain condition; see also Wang \cite{w1}. One can note that, here, $f''(u)$ changes sign exactly once on $(0,\infty )$. Subsequently, Wang and Kazarinoff \cite{w3} and Wang \cite{w2} studied \eqref{1.2} when $f$ is a cubic-like nonlinearity. In particular, Wang and Kazarinoff proved the next theorem. Define $F(u)=\int_{0}^{u}f(t)dt$. \begin{theorem}[{\cite[Theorem 1 and Remark 2]{w3}}] \label{thm1.1} Suppose $f\in C^{2}[0,\infty )$ and there exist $00\quad \text{for }u\in (0,a), \\ f(u)>0\quad \text{for }u\in (b,c), \\ f(u)<0\quad \text{for }u\in (a,b)\cup (c,\infty ); \end{gathered} \label{1.4} \end{equation} \begin{equation} \int_{a}^{c}f(u)du>0; \label{1.5} \end{equation} there exists a unique $\beta \in (b,c)$ defined by $\int_{a}^{\beta }f(u)du=0$ and such that $2F(a)-\beta f(\beta )<0$; \\ there exists $r\in (0,c)$ such that $f''(u)>0$ for $00$ such that \begin{itemize} \item[(i)] for $0<\lambda <\lambda _{0}$, problem \eqref{1.2} has exactly one positive solution $u_{0}$ satisfying $0<\| u_{0}\| \lambda _{0}$, problem \eqref{1.2} has exactly three positive solutions $u_{0}0$) and $f''$ changes sign exactly \emph{once} on $(0,\infty )$ were proved by Korman and Shi \cite{k3} (resp. Korman \textit{et al.} \cite{k2}.) But for \eqref{1.1} with $p\neq 2$, little is known. In fact for the case where $0=a1$, has been recently studied in Addou and Wang \cite{a3} for the case where $f(0)=0$ and $f''$ changes sign exactly \emph{twice} on $(0,\infty )$. We note that the case where $f(0)>0$ and $f''$ changes sign exactly \emph{twice} has not been studied yet. The paper is organized as follows. Section 2 is devoted to the definitions of the sets which contain the solutions of \eqref{1.1} and stating the main tool used subsequently, namely, the quadrature method. Next, in Section 3, we state our main results. In Section 4, a weakened condition and two examples are given. Finally, in Section 5, we prove the main results. \section{Quadrature method} To state the main results, we first define the subsets of $C^{1}[-1,1]$ which contain the solutions of \eqref{1.1}. By a positive solution to \eqref{1.1} we mean a positive function $u\in C^{1}[-1,1]$ with $\varphi _{p}(u')\in C^{1}[-1,1]$ satisfying \eqref{1.1}. Recall that $% Z(u)=\{x\in [-1,1] :u'(x)=0\}$. We note that it is easy to show that, if $f\in C$ and $u$ is a positive solution of \eqref{1.1}, then $u\in C^{2}[-1,1] $ if $12$. For the proof we refer to Addou \cite[Lemma 6]{a1}. \begin{figure}[ht] \begin{center} \includegraphics[width=0.95\textwidth]{fig1.pdf} \end{center} \caption{ Typical graph: (a) of $u\in A_{0}^{+}$; (b) of $u\in A_{1}^{+}$; (c) of $u\in A_{00}^{+} $; (d) of $u\in A_{01}^{+}$; (e) of $u\in A_{10}^{+}$; (f) of $u\in A_{11}^{+}$.} \end{figure} \begin{figure}[ht] \begin{center} \includegraphics[width=0.95\textwidth]{fig2.pdf} \end{center} \caption{Typical graph: (a) of $u\in B_{0}^{+}$; (b) of $u\in B_{1}^{+}$; (c) of $u\in B_{00}^{+} $; (d) of $u\in B_{01}^{+}$; (e) of $u\in B_{10}^{+}$; (f) of $u\in B_{11}^{+}$.} \end{figure} Let $A^{+}$ (resp. $B^{+}$) be the subset of $C^{1}[-1,1]$ consisting of the functions $u$ satisfying \begin{itemize} \item[(i)] $u(x)>0$\ for all $x\in ( -1,1) $,\ $u(-1)=u(1)=0$ and $u'(-1)>0$ (resp. $u'(-1)=0$), \item[(ii)] $u$ is symmetrical with respect to $0$ (i.e.\textit{, }$u$ is even). \end{itemize} Note that the derivative of any function $u\in A^{+}$ (resp. $B^{+}$) satisfies $u'(0)=0$. Therefore $Z^{+}(u)$ contains at least $0$. Also, $Z^{+}(u)$ may be connected or is an union of many connected components. Furthermore, each connected component is either a single point or an interval $[ \tilde{a},\tilde{b}] ,\;\tilde{a}<\tilde{b}$. (Note that $u'$ is continuous). So, for each integer $k=1,2,\dots$., one can consider the subsets of $A^{+}$ (resp. $B^{+}$) which are composed by functions $u$ such that $Z^{+}(u)$ is an union of $k$ connected components exactly. These sets can be designed by $A_{a_{1}a_{2}\ldots a_{k}}^{+}$ (resp. $B_{b_{1}b_{2}\ldots b_{k}}^{+}$) where for all $j\in \{ 1,2,\ldots ,k\} $, $a_{j}=0$ (resp. $b_{j}=0$) if the $j^{th}$ connected component is a single point and $a_{j}=1$ (resp. $b_{j}=1$) if it is an interval (not reduced to a single point). For example, $A_{0}^{+}$ (resp. $B_{0}^{+}$) is the subset of $A^{+}$ (resp. $B^{+}$) consisting of the functions $u$ such that their derivative $u'$ vanishes once and only once (at $0$ necessarily). An example of a function in $A_{0}^{+}$ (resp. $B_{0}^{+}$) is given by Fig. 1(a) (resp. Fig. 2(a)). Also, $% A_{1}^{+} $ (resp. $B_{1}^{+}$) is the subset of $A^{+}$ (resp. $B^{+}$) such that $u\in A_{1}^{+}$ (resp. $u\in B_{1}^{+}$) if and only if $u\in A^{+}$ (resp. $u\in B^{+}$) and there exists $x_{0}\in ( 0,1) $ such that for all $x\in [0,1]$, $u'(x)=0$ if and only if $0\leq x\leq x_{0}$ (resp. $0\leq x\leq x_{0}$ or $x=1$). An example of a function in $A_{1}^{+}$ (resp. $B_{1}^{+}$) is given by Fig. 1(b) (resp. Fig. 2(b)). \noindent An example of a function in $A_{00}^{+}$ (resp. $B_{00}^{+}$) is given by Fig. 1(c) (resp. Fig. 2(c)). That is, there exists $x_{0}\in (0,1) $ such that, for all $0\leq x\leq 1$, \begin{equation*} u'(x)=0\text{ if and only if }x\in \{ 0,x_{0}\} \text{(resp. }x\in \{ 0,x_{0},1\} ). \end{equation*} An example of a function in $A_{01}^{+}$ (resp. $B_{01}^{+}$) is given by Fig. 1(d) (resp. Fig. 2(d)). That is, there exist $01$ and assume that $u$ is a positive solution of \eqref{1.1}, then $( | u'(x)|^{p}+p'\lambda F(u(x))) '=0$ for all $x\in [-1,1]$. \end{lemma} \begin{lemma} \label{Lemma2.2} Suppose $f$ satisfies conditions \eqref{1.3}--\eqref{1.5} and $u$ is a positive solution of problem \eqref{1.1}. Then $u\in A_{0}^{+}\cup A_{1}^{+}\cup A_{00}^{+}\cup A_{01}^{+}$. \end{lemma} \begin{proof} Suppose $f$ satisfies conditions \eqref{1.3}--\eqref{1.5}, $f(0)>0$ and $f$ changes sign exactly twice on $(0,\infty ) $. Suppose $u$ is a positive solution of \eqref{1.1}, then $u$ is symmetrical with respect to $0$. It can be easily proved that either $0<\|u\| _{\infty }\leq a$ or $\beta \leq \|u\| _{\infty }\leq c$ by applying Lemma \ref{Lemma2.1}; cf. Remark \ref{Remark1}(i). Thus \begin{equation*} u\in A_{0}^{+}\cup A_{1}^{+}\cup A_{00}^{+}\cup A_{01}^{+}\cup A_{10}^{+}\cup A_{11}^{+}\cup B_{0}^{+}\cup B_{1}^{+}\cup B_{00}^{+}\cup B_{01}^{+}\cup B_{10}^{+}\cup B_{11}^{+}. \end{equation*} The proof is easy but tedious, so we omit it. More precisely, (i) Since $f(0)>0$, if $u$ is a positive solution $u$ of \eqref{1.1} satisfying $\| u\| _{\infty }=u(0)=\eta \in ( 0,a] \cup [ \beta ,c] $, then $u'(-1)=( p^{\prime }\lambda F( \eta ) ) ^{1/p}>0$ by applying Lemma \ref {Lemma2.1}. Hence $u\notin B_{0}^{+}\cup B_{1}^{+}\cup B_{00}^{+}\cup B_{01}^{+}\cup B_{10}^{+}\cup B_{11}^{+}$. (ii) We then show that $u\notin A_{10}^{+}\cup A_{11}^{+}$. Suppose that $% u\in A_{10}^{+}\cup A_{11}^{+}$. Then either $\| u\| _{\infty }=c$ or $\| u\| _{\infty }=a$, and there exists $x_{1}\in ( 0,1) $ such that $u'(x_{1})=0$. If $\| u\| _{\infty }=c$, then by applying Lemma \ref{Lemma2.1}, $p'\lambda F(c)=p'\lambda F(x_{1})$, which contradicts the fact that $% F(c)>F(x_{1})$. So $\| u\| _{\infty }\neq c$. Similarly, $\| u\| _{\infty }\neq a$. We conclude that $u\notin A_{10}^{+}\cup A_{11}^{+}$. By above (i) and (ii), we obtain that $u\in A_{0}^{+}\cup A_{1}^{+}\cup A_{00}^{+}\cup A_{01}^{+}$. \end{proof} To study \eqref{1.1}, we make use of the quadrature method. Suppose $f\in C[0,\infty )$ satisfies conditions \eqref{1.3}--\eqref{1.5}. For any $E\geq 0$ and $s>0$, let $G(E,s):=E^{p}-p'\lambda F(s)$. It can be shown that, the function $G(E,\cdot )$ has at most four zeros in $( 0,\infty ) $. For any $E\geq 0$, define \begin{equation*} X_{1}(E)=\{ s>0:s\in \mathop{\rm dom}G(E,\cdot )\text{ and }G(E,u)>0\text{ for all }% u\in (0,s)\} \end{equation*} and \begin{equation*} r_{1}(E)=\begin{cases} 0 & \text{if }X_{1}(E)=\emptyset , \\ \sup ( X_{1}(E)) & \text{otherwise.} \end{cases} \end{equation*} Next for any $E\geq 0$, define \begin{equation*} X_{2}(E)=\{ s>r_{1}(E):s\in \mathop{\rm dom}G(E,\cdot )\text{ and }G(E,u)>0\text{ for all }u\in (r_{1}(E),s)\} \end{equation*} and \begin{equation*} r_{2}(E)=\begin{cases} \infty & \text{if }X_{2}(E)=\emptyset , \\ \sup ( X_{2}(E)) & \text{otherwise.} \end{cases} \end{equation*} Note that $X_{2}(E)$ and $r_{2}(E)$ are well defined even if $r_{1}(E)=\infty $. In fact, in this case, $X_{2}(E)=\emptyset $ and $r_{2}(E)=\infty $. Let \begin{align*} \tilde{D}_{1}=\big\{& E\geq 0: r_{1}(E)\in \mathop{\rm dom}G(E,\cdot ),\, G(E,r_{1}(E))=0, \\ &\ \text{and }\int_{0}^{r_{1}(E)}( E^{p}-p'\lambda F(t) ) ^{-1/p}dt<\infty \big\} , \end{align*} \begin{align*} \tilde{D}_{2}=\big\{& E\geq 0: r_{2}(E)\in \mathop{\rm dom}G(E,\cdot ),\, G(E,r_{2}(E))=0, \\ & \text{and }\int_{0}^{r_{2}(E)}( E^{p}-p'\lambda F(t) ) ^{-1/p}dt<\infty \big\} . \end{align*} Define the time maps \begin{gather*} T_{1}(E)=\int_{0}^{r_{1}(E)}( E^{p}-p'\lambda F( t) ) ^{-1/p}dt,\;E\in \tilde{D}_{1}, \\ T_{2}(E)=\int_{0}^{r_{2}(E)}( E^{p}-p'\lambda F( t) ) ^{-1/p}dt,\;E\in \tilde{D}_{2}, \end{gather*} whenever $\tilde{D}_{1}\neq \emptyset $ (resp. $\tilde{D}_{2}\neq \emptyset $). By Lemma \ref{Lemma2.1} and arguments in \cite{g1}, we have the following theorem. Note that in this paper, by Lemma \ref{Lemma2.2}, we restrict ourself on positive solutions $u\in A_{0}^{+}\cup A_{1}^{+}\cup A_{00}^{+}\cup A_{01}^{+}$. \begin{theorem}[Quadrature method]\label{Thm2.2} Consider \eqref{1.1}. Suppose $f\in C[0,\infty )$ satisfies conditions \eqref{1.3}--\eqref{1.5}. Let $E\geq 0$. Then $T_{1}$ (resp. $T_{2}$) is a continuous function of $E\in \tilde{D}_{1}$. (resp. $E\in \tilde{D}_{2}$). Moreover, \begin{itemize} \item[(i)] Problem \eqref{1.1} has a solution $u\in A_{0}^{+}$ satisfying $% u'(-1)=E>0$ if and only if $E\in \tilde{D}_{1}-\{ 0\} $, $f(r_{1}(E))\geq 0$ and $T_{1}(E)=1,$\ and in this case the solution is unique. \item[(ii)] Problem \eqref{1.1} has a solution $u\in A_{1}^{+}$ satisfying $% u'(-1)=E>0$ if and only if $E\in \tilde{D}_{1}-\{ 0\} $, $f(r_{1}(E))=0$ and $T_{1}(E)<1$, and in this case the solution is unique. \item[(iii)] Problem \eqref{1.1} has a solution $u\in A_{00}^{+}$ satisfying $u'(-1)=E>0$ if and only if $E\in \tilde{D}_{2}-\{ 0\} $, $f(r_{1}(E))\geq 0,\;f(r_{2}(E))\geq 0$, and $T_{2}(E)=1$, and in this case the solution is unique. \item[(iv)] Problem \eqref{1.1} has a solution $u\in A_{01}^{+}$ satisfying $u'(-1)=E>0$ if and only if $E\in \tilde{D}_{2}-\{ 0\} $, $f(r_{1}(E))=0$, $f(r_{2}(E))\geq 0$, and $T_{2}(E)<1$, and in this case the solution is unique. \end{itemize} \end{theorem} \begin{remark} \label{Remark4} \rm In practice, we first study the variations of the real-valued function $G(E,\cdot )$, then compute $X_{1}(E)$ and deduce $r_{1}(E)$ (resp. compute $X_{2}(E)$ and deduce $r_{2}(E)$). Next, we compute $\tilde{D}_{1}$ (resp. $\tilde{D}_{2}$). For this, we first compute the set \[ D_{1}=\{ E>0:r_{1}(E)\in \mathop{\rm dom} G(E,\cdot ), \; G(E,r_{1}(E))=0,\; f(r_{1}(E))>0 \}, \] (resp. \[ D_{2}=\{ E>0:r_{2}(E)\in \mathop{\rm dom}G(E,\cdot ),\; G(E,r_{2}(E))=0,\; f(r_{2}(E))>0\}), \] and then we deduce $\tilde{D}_{1}$ (resp. $\tilde{D}_{2}$) by observing that $D_{1}\subset \tilde{D}_{1}-\{ 0\} \subset \overline{D}_{1}$ (resp. $D_{2}\subset \tilde{D}_{2}-\{ 0\} \subset \overline{D}_{2} $) ; we omit the proof. (Note that $\overline{D}_{1}$ is the closure of $D_{1}$ (resp. $\overline{D}_{2}$ is the closure of $D_{2}$).) After that, we define the time map $T_{1}$ on $\tilde{D}_{1}$ and then compute its limits at the boundary points of $\tilde{D}_{1}$. We next study the variations of $T_{1}$ on $\tilde{D}_{1}$. For $T_{2}$, we shall show that its definition domain $\tilde{D}_{2}$ is restricted to a single point; there is no variation to study for $T_{2}$. We achieve our study by discussing the number of solutions to \begin{itemize} \item[(i)] Equation $T_{1}(E)=1$ and $f(r_{1}(E))\geq 0$ for $E\in \tilde{D}_{1}-\{ 0\} $ in case of looking for solutions $u$ in $A_{0}^{+}$. \item[(ii)] Inequality $T_{1}(E)<1$ and $f(r_{1}(E))=0$ for $E\in \tilde{D}_{1}-\{ 0\} $ in case of looking for solutions $u$ in $A_{1}^{+}$. \item[(iii)] Equation $T_{2}(E)=1$ and $f(r_{1}(E))\geq 0$, $f(r_{2}(E))\geq 0$, for $E\in \tilde{D}_{2}-\{ 0\} $ in case of looking for solutions $u$ in $A_{00}^{+}$. \item[(iv)] Inequality $T_{2}(E)<1$ and $f(r_{1}(E))=0$, $f(r_{2}(E))\geq 0, $ for $E\in \tilde{D}_{2}-\{ 0\} $ in case of looking for solutions $u$ in $A_{01}^{+}$. \end{itemize} \end{remark} \section{Main results} We determine the exact multiplicity of positive solutions of \eqref{1.1} for $\lambda >0$ under hypotheses (H1)-(H5) stated below. In particular, we assume that $f$ satisfies the $``$\emph{convexity}$"$ condition (H4) which implies for the particular case $p=2$, that $f''$ changes sign exactly \emph{twice} on $(0,\infty )$, i.e., $f$ is concave-convex-concave on $(0,\infty )$. Note that if $f$ satisfies (H1)-(H3) then it satisfies \eqref{1.3}--\eqref{1.5}. Also note that we may allow that $f''$ changes sign more than twice, i.e., we may allow that $f$ is concave-convex-concave-convex; see Section 4. For $f$, recalling that $F(u)=\int_{0}^{u}f(t)dt$, we let \begin{gather} \theta _{p}(u):=pF(u)-uf(u), \nonumber\\ \Psi _{p}(u):=u\theta _{p}'(u)-\theta _{p}(u)=puf(u)-u^{2}f^{\prime }(u)-pF(u), \nonumber\\ \nu _{p}:=\big\{ \int_{0}^{c}( F(c)-F(u)) ^{-1/p}du\big\} ^{p}/p'\in (0,\infty ], \nonumber\\ \alpha _{p}:=\big\{ \int_{0}^{a}( F(a)-F(u)) ^{-1/p}du\big\} ^{p}/p'\in (0,\infty ], \label{3.3} \\ \lambda _{p}:=\big\{ \int_{0}^{\beta }( F(\beta )-F(u)) ^{-1/p}\big\} ^{p}/p'\in (0,\infty ], \label{3.4} \\ \mu _{p}:=\inf_{\beta \leq \xi \leq c}\big\{ \int\nolimits_{0}^{\xi }( F(\xi )-F(u)) ^{-1/p}du\big\} ^{p}/p', \nonumber \end{gather} where $01$. For all $\lambda >0$, we denote $S_{\lambda }$ the positive solution set of \eqref{1.1}. For fixed $p>1$, suppose $f\in C^{2}[0,\infty )$ and there exist $00$ \item[(H2)] $f(u)>0$ for $00$ for $bc$ \item[(H3)] $\int_{a}^{c}f(u)du>0$, and there exists $\beta ^{*}\in ( 0,\beta] $ such that $\theta _{p}(\beta ^{*})=pF(\beta ^{*})-\beta^{*}f(\beta ^{*})<0$, where $\beta \in (b,c)$ is defined by $\int_{a}^{\beta}f(u)du=0$, \item[(H4)] There exist $00\quad \text{for }00\quad \text{for }s_{p}\mu _{p}$, there exist $u_{\lambda }$, $v_{\lambda }$ and $w_{\lambda }\in A_{0}^{+}$ such that $u_{\lambda}0$ and $1\theta _{2}( \sigma _{2}) \approx -0.8789$. \protect\linebreak (c) Graph of $\Psi _{2}(u)$. $0.8792\approx \Psi _{2}(\sigma _{2}) >\Psi _{2}(r_{2})\approx 0.5127$.} \end{figure} Next, we give two interesting examples of quartic polynomials of Theorem \ref{Thm3.5} with $p=2$, of which one satisfies $\theta _{2}(\beta )>0$ and the other satisfies $\theta _{2}(\beta )<0$. \subsection*{Two examples of Theorem \ref{Thm3.5}} (i) (See Fig. 4, $\theta _{2}(\beta )>0$) Let $p=2$. The function $% f=f_{1}(u)=-(u+1/4)(u-1)(u-2)(u-3)-0.13$ satisfies all conditions (H1)-(H5) in Theorem \ref{Thm3.5} with $a\approx 0.9497$, $b\approx 2.0565$, $c\approx 2.9792$, $\int_{a}^{c}f_{1}(s)ds\approx 0.004695>0$, $r_{2}\approx 0.7425$, $% s_{2}\approx 2.1325$, $2.5787\approx \sigma _{2}<\beta \approx 2.9387$, $% 1.4751\approx \theta _{2}(\beta )>\theta _{2}(\sigma )\approx -0.8789$. Note that, in (H4)-(H5), $0.8792\approx \Psi _{2}(\sigma _{2})>\Psi _{2}(r_{2})\approx 0.5127$. \noindent (ii) ($\theta _{2}(\beta )<0$) Let $p=2$. Let $f=f_{2}(u)=-(u-d)(u-a)(u-b)(u-c)$ with \begin{equation*} d=-\frac{1}{6}<0c_{1}\Leftrightarrow \int_{a}^{c}f_{2}(u)du=\frac{1}{360}% (c-1)^{2}(18c^{3}-49c^{2}-26c+25)>0\text{ on }( 2,\infty ) . \end{equation*} So for $c>c_{1}\approx 2.8380$, there exists a unique $\beta =\beta (c)\in (b,c)=( 2,c) $ satisfying \begin{equation*} \int_{a}^{\beta }f_{2}(u)du=0. \end{equation*} or, \begin{equation*} \frac{1}{360}(\beta -1)^{2}[ -72\beta^{3}+(90c+111)\beta ^{2}+(-160c+114)\beta -140c+57] =0. \end{equation*} Note that $\beta (c)$ can be expressed explicitly by Cartan's formulas; see e.g. \cite{j1}. We compute that \begin{equation*} \theta _{2}(u)=\frac{3}{5}u^{5}-\frac{6c+17}{12}u^{4}+\frac{17c+9}{18}u^{3}+% \frac{c}{3}u \end{equation*} and \begin{equation*} \theta _{2}(\beta (c))<0\text{ for }c>c_{2}\approx 2.9056, \end{equation*} where $c_{2}$ is the unique zero of $\theta _{2}(\beta (c))$ on $( c_{1},\infty ) $. Thus $f_{2}$ satisfies (H3) for $c>c_{2}\approx 2.9056$. Finally, we check (H5) for $c>c_{2}$. Let $\sigma _{2}=\sigma _{2}(c)$ be the unique zero of \begin{equation*} \theta _{2}'(u)=3u^{4}-\frac{6c+17}{3}u^{3}+\frac{17c+9}{6}u^{2}+% \frac{c}{3} \end{equation*} in $(s_{2},c)$. Note that $\sigma _{2}(c)$ can be expressed explicitly; see e.g. \cite{j1}. We compute that \begin{gather*} \Psi _{2}(u)=\frac{1}{180}u^{3}[ 432u^{2}-(270c+765)u+340c+180]\,,\\ \Psi _{2}(\sigma _{2})>\Psi _{2}(s_{2})\text{ for }c>c_{2}\approx 2.9056. \end{gather*} We summarize above results and conclude that $f_{2}$ satisfies (H1)-(H5) for $c>c_{2}\approx 2.9056$. \subsection{The case $p>2$} By Lemma \ref{Lemma2.2}, $S_{\lambda }\subset A_{0}^{+}\cup A_{1}^{+}\cup A_{00}^{+}\cup A_{01}^{+}$. Hence \begin{equation*} S_{\lambda }=( S_{\lambda }\cap A_{0}^{+}) \cup ( S_{\lambda }\cap A_{1}^{+}) \cup ( S_{\lambda }\cap A_{00}^{+}) \cup ( S_{\lambda }\cap A_{01}^{+}) ; \end{equation*} Theorem \ref{Thm3.6} (resp. \ref{Thm3.7}, \ref{Thm3.8}, \ref{Thm3.9}) gives complete description of the set $S_{\lambda }\cap A_{0}^{+}$ (resp. $% S_{\lambda }\cap A_{1}^{+},\;S_{\lambda }\cap A_{00}^{+},\;S_{\lambda }\cap A_{01}^{+}$); see Fig. 5. \begin{figure}[ht] \begin{center} \includegraphics[width=0.95\textwidth] {fig5.pdf} \end{center} \caption{Bifurcation diagrams for Eq. \eqref{1.1} with $p>2$ and $0<\alpha _{p}<\mu _{p}$. (a) $\mu _{p}<\nu _{p}<\lambda_{p}$; (b) $\mu _{p}<\nu _{p}=\lambda _{p};$ (c) $\mu _{p}<\lambda _{p}<\nu_{p}$.} \end{figure} \begin{remark}\label{Remark8}\rm Fig. 5 describes the solution set of \eqref{1.1} when $p>2$ and $0<\alpha _{p}<\mu _{p}$. There are two connected branches. The lower branch bifurcates at the origin and represents solutions in $A_{0}^{+}$ until $\lambda =\alpha _{p};$ for all $\lambda >\alpha _{p}$, the branch is horizontal and represents solutions in $A_{1}^{+}$. The upper branch is formed by two parts: The upper horizontal curve represents solutions in $% A_{1}^{+}$ with norm equal to $c$, and the $\subset $-shaped curve represents solutions in $A_{0}^{+}\,$until $\lambda =\lambda _{p}$ where there is a solution in $A_{00}^{+}$ with norm equal to $\beta $ and on its right the lower horizontal curve which represents solutions in $A_{01}^{+}$ with norm equal to $\beta $. \end{remark} We shall show that, for $p>2$, $0<\mu _{p}<\lambda _{p}<\infty $ and $0<\mu_{p}<\nu _{p}<\infty $; see Lemmas \ref{Lemma4.9}-\ref{Lemma4.10}. We also show that, for $p>2$, $0<\alpha _{p}<\lambda _{p}<\infty $; see Lemma \ref{Lemma4.11}. For $p>1$, let \begin{gather*} B(0,a):= \{ u\in C^{1}[-1,1]:\| u\| _{\infty }2$, see Fig. 5] \label{Thm3.6} Assume that $p>2$ and $f\in C^{2}[0,\infty )$ satisfies (H1)-(H5). Then $0<\mu _{p}<\lambda _{p}<\infty $, $0<\mu _{p}<\nu _{p}<\infty $, and $0<\alpha _{p}<\lambda _{p}<\infty $. $S_{\lambda }\cap A_{0}^{+}\cap ( B(0,\beta )-\overline{B(0,a)}) =\emptyset $. Also, \begin{itemize} \item[(i)] For $0<\lambda \leq \alpha _{p}$, there exists $u_{\lambda }\in A_{0}^{+}\cap \overline{B(0,a)}$ such that $S_{\lambda }\cap A_{0}^{+}\cap \overline{B(0,a)}=\{ u_{\lambda }\} $. Moreover, $0<\| u_{\lambda }\| _{\infty }\leq a$, and $\| u_{\lambda }\| _{\infty }=a$ if and only if $\lambda =\alpha _{p}$. \item[(ii)] For $\lambda >\alpha _{p}$, $S_{\lambda }\cap A_{0}^{+}\cap \overline{B(0,a)}=\emptyset $. \end{itemize} Moreover, \textbf{(a) If }$\mu _{p}<\nu _{p}<\lambda _{p}$, \textbf{then:} \begin{itemize} \item[(iii)] For $0<\lambda <\mu _{p}$, $S_{\lambda }\cap A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) =\emptyset $. \item[(iv)] For $\lambda =\mu _{p}$, there exists $u_{\lambda }\in A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) $ such that $% \newline S_{\lambda }\cap A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) =\{ u_{\lambda }\} $ and $\beta <\| u_{\lambda }\| _{\infty }\lambda _{p}=\nu _{p}$, $S_{\lambda }\cap A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) =\emptyset $. \end{itemize} \noindent\textbf{(c) If }$\mu _{p}<\lambda _{p}<\nu _{p}$, \textbf{then:} \begin{itemize} \item[(xiii)] For $0<\lambda <\mu _{p}$, $S_{\lambda }\cap A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) =\emptyset $. \item[(xiv)] For $\lambda =\mu _{p}$, there exists $u_{\lambda }\in A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) $ such that \newline $S_{\lambda }\cap A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) =\{ u_{\lambda }\} $. Moreover, $\beta <\| u_{\lambda}\| _{\infty }\nu _{p}$, $S_{\lambda }\cap A_{0}^{+}\cap ( \overline{B(0,c)}-B(0,\beta )) =\emptyset $. \end{itemize} \end{theorem} \begin{theorem}[$S_{\lambda }\cap A_{1}^{+}$ for $p>2$, see Fig. 5]\label{Thm3.7} Assume that $p>2$ and $f\in C^{2}[0,\infty )$\ satisfies (H1)-(H5). Then each solution $u_{\lambda }$ of \eqref{1.1} in $A_{1}^{+}$ satisfies $\| u_{\lambda }\| _{\infty }=a$ or $\| u_{\lambda }\| _{\infty }=c$. Moreover: \begin{itemize} \item[(i)] For $0<\lambda \leq \alpha _{p}$, $S_{\lambda }\cap A_{1}^{+}\cap \partial \overline{B(0,a)}=\emptyset $. \item[(ii)] For $\lambda >\alpha _{p}$, there exists $u_{\lambda }\in A_{1}^{+}\cap \partial \overline{B(0,a)}$ such that $S_{\lambda }\cap A_{1}^{+}\cap \partial \overline{B(0,a)}=\{ u_{\lambda }\} $ and $% \| u_{\lambda }\| _{\infty }=a$. \item[(iii)] For $0<\lambda \leq \nu _{p}$, $S_{\lambda }\cap A_{1}^{+}\cap \partial \overline{B(0,c)}=\emptyset $. \item[(iv)] For $\lambda >\nu _{p}$, there exists $u_{\lambda }\in A_{1}^{+}\cap \partial \overline{B(0,c)}$ such that $S_{\lambda }\cap A_{1}^{+}\cap \partial \overline{B(0,c)}=\{ u_{\lambda }\} $ and $% \| u_{\lambda }\| _{\infty }=c$. \end{itemize} \end{theorem} \begin{theorem}[$S_{\lambda }\cap A_{00}^{+}$ for $p>2$, see Fig. 5]\label{Thm3.8} Assume that $p>2$ and $f\in C^{2}[0,\infty )$\ satisfies (H1)-(H5). Then each solution $u_{\lambda }$ of \eqref{1.1} in $A_{00}^{+}$ satisfies $\| u_{\lambda }\| _{\infty }=\beta $. Moreover, \begin{itemize} \item[(i)] For $\lambda \neq \lambda _{p}\;$and\ $\lambda >0,\;S_{\lambda }\cap A_{00}^{+}=\emptyset $. \item[(ii)] For $\lambda =\lambda _{p}$, there exists $u_{\lambda }\in A_{00}^{+}$ such that $S_{\lambda }\cap A_{00}^{+}=\{ u_{\lambda }\} $ and $\| u_{\lambda }\| _{\infty }=\beta $. \end{itemize} \end{theorem} \begin{theorem}[$S_{\lambda }\cap A_{01}^{+}$ for $p>2$, see Fig. 5]\label{Thm3.9} Assume that $p>2$ and $f\in C^{2}[0,\infty )$\ satisfies (H1)-(H5). Then each solution $u_{\lambda }$ of \eqref{1.1} in $A_{01}^{+}$ satisfies $\|u_{\lambda }\| _{\infty }=\beta $. Moreover, \begin{itemize} \item[(i)] For $0<\lambda \leq \lambda _{p},\;S_{\lambda }\cap A_{01}^{+}=\emptyset $. \item[(ii)] For $\lambda >\lambda _{p}$, there exists $u_{\lambda }\in A_{01}^{+}$ such that $S_{\lambda }\cap A_{01}^{+}=\{ u_{\lambda }\} $ and $\| u_{\lambda }\| _{\infty }=\beta $. \end{itemize} \end{theorem} \section{A weakened condition and two examples} We point out that the \emph{convexity} condition of $\theta _{p}(u)=pF(u)-uf(u)$ on $(0,c)$ in (H4) in Theorems \ref{Thm3.5}-\ref{Thm3.9} can actually be weakened; cf. Remark 9 in Addou and Wang \cite{a3}, which holds true in the case $f(0)=0$ as well as in the positone case $f(0)>0$. More precisely, condition (H4) can be weakened as\smallskip \begin{itemize} \item[(H4$'$)] There exist $0\leq r_{p}0\text{ for\ }00\text{ for\ }s_{p}1$, $\Psi _{p}(\sigma _{p})>0=\Psi _{p}(r_{p})$. We also note that in (H4$'$) the condition \begin{equation*} (p-2)f'(u)-uf''(u)>0\quad \text{for }s_{p}\theta _{p}(t_{p}) \text{ for all }\xi \in (u,c]\} & \text{otherwise,} \end{cases} \label{3.9} \end{equation} where $t_{p}$ is the unique zero of $\theta _{p}'(u)$ on $(r_{p},s_{p}) $. Thus we summarize that (H4$'$) can be weakened as follows: \begin{itemize} \item[(H4$''$)] There exist $0\leq r_{p}0\text{ for\ }00$, $p=2$ and $r_{2}=0$, Theorem \ref{Theorem4.1} generalizes \cite[Theorem 1]{w3}. We give two examples of classes of nonlinearities of Theorem \ref{Theorem4.1}. \begin{proposition} \label{Theorem4.2} Let $p=2$. $f=f_{3}(u)=-(u-a)(u-b)(u-c)$ with $a=1$, $b=3$, and \begin{equation*} c>2b-a=5\text{ }( \Leftrightarrow \int_{a}^{c}f_{3}(u)du>0) . \end{equation*} Then $f_{3}$ satisfies all conditions (H1)-(H3), (H4$'$) and (H5) in Theorem \ref{Theorem4.1}. \end{proposition} \begin{proof} It is easy to see that $f_{3}(u)=-(u-1)(u-3)(u-c)$ satisfies (H1), (H2), (H5) and (H4$'$) with $r_{2}=0$ for $c>5$. Finally, we check (H3) for $c>5$. We compute that \begin{gather*} \theta _{2}(u)=\frac{1}{2}u^{4}-\frac{1}{3}(c+4)u^{3}+3cu,\\ \beta =\beta (c)=\frac{1}{3}( 2c+5-2\sqrt{c^{2}-7c+10}) \in ( 3,c) , \end{gather*} and $\theta _{2}(\beta )<0$ for $c>\tilde{c}\approx 5.1193$, where $\tilde{c} $ is the unique zero of $\theta _{2}(\beta )$ on $( 5,\infty ) $. Although for $55$. \end{proof} \begin{figure}[ht] \begin{center} \includegraphics[width=0.95\textwidth]{fig6.pdf} \end{center} \caption{(a) Graph of $f_{4}(u)=-\sin u+\varepsilon $ on $( 0,c) $ for $\varepsilon =0.2$, $a\approx 0.2014$, $b\approx 2.9402$, $c\approx 6.4845$, $\beta \approx 4.7770$, $r_{2}=0$, $s_{2}=\pi \approx 3.1416$, $\sigma _{2}\approx 4.4473$. (b) Graph of $\theta _{2}(u)$ on $( 0,c) $. $t_{2}\approx 0.8650$, $d\approx 6.1084$.} \end{figure} \begin{proposition}[See Fig. 6 for $\varepsilon =0.2$]\label{Theorem4.3} Let $p=2$. For $0<\varepsilon <1$, let \begin{equation*} 0c. \end{cases} \end{equation*} Then $f_{4}$\ satisfies all conditions (H1)-(H3), (H4$''$) and (H5) in Theorem \ref{Theorem4.1} for $\varepsilon >0$ small enough. \end{proposition} \begin{proof} For $f=f_{4}(u)$, we find that \begin{gather*} \theta _{2}(u)=u\sin u+2\cos u+\varepsilon u-2,\\ \theta _{2}'(u)=u\cos u-\sin u+\varepsilon . \end{gather*} Let $\sigma _{2}=\sigma _{2}(\varepsilon )$ be the unique zero of $\theta _{2}'(u)$ on $( \pi ,2\pi ) \subset ( b,c) $ and $\beta =\beta (\varepsilon )$ be the unique zero of \begin{equation*} \int_{a}^{u}f_{4}(s)ds=\varepsilon u+\cos u-\varepsilon \sin ^{-1}\varepsilon -\sqrt{1-\varepsilon ^{2}} \end{equation*} on $( b,c) $. It can be checked easily that \begin{itemize} \item[(i)] $f_{4}$ satisfies (H1) and (H2) for $0<\varepsilon <1$. \item[(ii)] $f_{4}$ satisfies the condition $\int_{a}^{c}f_{4}(u)du>0$ in (H3) for $0<\varepsilon <1$. Also, for $\varepsilon >0$ small enough, by continuity, $0<\sigma _{2}(\varepsilon )<\beta (\varepsilon )$ since $\sigma _{2}(0)\approx 4.4934<\beta (0)=2\pi $, and \begin{equation} \theta _{2}(\sigma _{2}(\varepsilon ))<0 \label{4.4} \end{equation} since $\theta _{2}(\sigma _{2}(0))\approx -6.8206<0$. So $f_{4}$ satisfies (H3) for $\varepsilon >0$ small enough. \item[(iii)] We check that $f_{4}$ satisfies (H4$''$). First \begin{equation} \theta _{2}'(0)=f_{4}(0)=\varepsilon >0, \label{4.5} \end{equation} \begin{equation} \theta _{2}''(u)=-uf_{4}''(u)=-u\sin u \begin{cases} <0 & \text{for }0=r_{2}0 & \text{for }s_{2}=\pi 0$ small enough, let $t_{2}=t_{2}(\varepsilon )$ be the unique zero of $\theta _{2}'(u)=u\cos u-\sin u+\varepsilon $ on $( 0,\pi ) $. It can be proved that $\lim_{\varepsilon \to 0^{+}}t_{2}(\varepsilon )=0$. More precisely, we compute that \begin{equation*} t_{2}(\varepsilon )\sim (2\varepsilon )^{1/3}\text{ as }\varepsilon \to 0^{+} \end{equation*} and hence \begin{equation*} \theta _{2}(t_{2}(\varepsilon ))\sim 2^{1/3}\varepsilon ^{4/3}\text{ as }% \varepsilon \to 0^{+}. \end{equation*} Thus \begin{equation} \theta _{2}(2\pi )=2\varepsilon \pi >\theta _{2}(t_{2}(\varepsilon ))\text{ for }\varepsilon >0\text{ small enough.} \label{4.7} \end{equation} We also find that \begin{gather} \theta _{2}'(2\pi )=2\pi +\varepsilon >0, \label{4.8}\\ \theta _{2}'(c)=c\cos c=( 2\pi +\sin ^{-1}\varepsilon ) \sqrt{1-\varepsilon ^{2}}>0. \label{4.9} \end{gather} So by (\ref{4.4})-(\ref{4.9}), it can be proved that there exists $d\in ( \sigma _{2},2\pi ] $ such that $f_{4}$ satisfies \begin{gather*} \theta _{2}''(u)=-uf_{4}''(u)=-u\sin u<0 \quad \text{for } 0=r_{2}0 & \text{for }d\leq u\leq c=2\pi +\sin ^{-1}\varepsilon , \end{cases} \\ \theta _{2}''(u)=-uf_{4}''(u)=-u\sin u>0\quad \text{for }\sigma _{2}\leq u0 $ small enough. \end{proof} \section{Proofs of main results} First, we have the next lemma which holds for nonlinearities $f\in C[0,\infty )$ satisfying (H1), (H2) and the condition $\int_{a}^{c}f(s)ds>0$ in (H3). We omit the proof. \begin{lemma} \label{Lemma4.7} Assume that $f\in C[0,\infty )$ satisfies (H1), (H2) and the condition $\int_{a}^{c}f(s)ds>0$ in (H3). Consider the function defined by \begin{equation*} s\longmapsto G(\lambda ,E,s):=E^{p}-p'\lambda F(s), \end{equation*} where $p>1$, $E\geq 0$ and $\lambda >0$ are real parameters. Then \begin{itemize} \item[(i)] If $E>E_{c}:=( p'\lambda F(c)) ^{1/p}>0$, then the function $G(\lambda ,E,\cdot )$ is strictly positive on $(0,\infty ) $. \item[(ii)] If $E=E_{c}$, then the function $G(\lambda ,E,\cdot )$ is strictly positive on $( 0,c) $ and vanishes at $c$. \item[(iii)] If $E_{a}:=( p'\lambda F(a)) ^{1/p}0\;\text{for all }E\in ( E_{a},E_{c}) . \end{equation*} \item[(b)] $\lim_{E\to E_{a}{}^{+}}s_{1}(\lambda ,E)=\beta $ and $\lim_{E\to E_{c}^{-}}s_{1}(\lambda ,E)=c$. \end{itemize} \item[(iv)] If $E=E_{a}$, then \[ G(\lambda ,E)\begin{cases} >0 & \text{for }00 & \text{for }a0\text{ for all }E\in ( 0,E_{a}) . \end{equation*} \item[(b)] $\lim_{E\to 0^{+}}s_{2}(\lambda ,E)=0$ and $% \lim_{E\to E_{a}^{-}}s_{2}(\lambda ,E)=a$. \end{itemize} \item[(vi)] If $E=0$, then $G(\lambda ,0,s) \begin{cases} <0 & \text{for }01$, $\lambda >0$ and $E\geq 0$, we let \begin{equation*} X_{1}(\lambda ,E):=\{ s\in \mathop{\rm dom}G(\lambda ,E,\cdot )=( 0,\infty ) :G(\lambda ,E,u)>0\quad\text{for all }u\in (0,s)\} . \end{equation*} In view of Lemma \ref{Lemma4.7}, it follows that \begin{equation*} X_{1}(\lambda ,E)= \begin{cases} ( 0,\infty ) & \text{if }E>E_{c}, \\ ( 0,c] & \text{if }E=E_{c}, \\ ( 0,s_{1}(\lambda ,E)] & \text{if }E_{a}E_{c}, \\ c & \text{if }E=E_{c}, \\ s_{1}(\lambda ,E) & \text{if }E_{a}r_{1}(\lambda ,E): s\in \mathop{\rm dom} G(\lambda ,E,\cdot )=( 0,\infty ) , \\ &\quad G(\lambda ,E,u)>0\text{ for all }u\in (r_{1}(\lambda ,E),s) \big\} \end{align*} In view of Lemma \ref{Lemma4.7}, \begin{equation*} X_{2}(\lambda ,E)=\begin{cases} \emptyset & \text{if }E>E_{c}, \\ ( c,\infty ) & \text{if }E=E_{c}, \\ \emptyset & \text{if }E_{a}0: r_{1}(\lambda ,E)\in \mathop{\rm dom}G(\lambda ,E,\cdot )=( 0,\infty ),\\ &\quad G(\lambda ,E,r_{1}(\lambda ,E))=0, \text{ and }f(r_{1}(\lambda ,E))>0 \big\}\\ &=( 0,E_{a}) \cup (E_{a},E_{c}) \end{align*} and \begin{align*} D_{2}(p,\lambda )&:=\big\{ E>0: r_{2}(\lambda ,E)\in \mathop{\rm dom}G(\lambda ,E,\cdot )=( 0,\infty ),\\ &\quad G(\lambda ,E,r_{2}(\lambda ,E))=0,\text{ and }f(r_{2}(\lambda ,E))>0 \\ &=\{ E_{a}\} . \end{align*} Note that the definition domains of the time maps $T_{1}$ and $T_{2}$ are \begin{align*} \tilde{D}_{1}(p,\lambda ):=\big\{&E\geq 0: r_{1}(\lambda ,E)\in \mathop{\rm dom}G(\lambda ,E,\cdot )=( 0,\infty ) , G(\lambda ,E,r_{1}(\lambda ,E))=0,\\ &\text{ and }\int\nolimits_{0}^{r_{1}(\lambda ,E)}( G( \lambda ,E,u) ) ^{-1/p}du <\infty \big\}, \end{align*} \begin{align*} \tilde{D}_{2}(p,\lambda ):=\big\{ &E\geq 0: r_{2}(\lambda ,E)\in \mathop{\rm dom}G(\lambda ,E,\cdot )=( 0,\infty ) , G(\lambda ,E,r_{2}(\lambda ,E))=0, \\ &\text{and }\int\nolimits_{0}^{r_{2}(\lambda ,E)}( G( \lambda ,E,u) ) ^{-1/p}du <\infty \big\} \end{align*} In the present case, $( 0,E_{a}) \cup ( E_{a},E_{c}) \subset \tilde{D}_{1}(p,\lambda )\subset [ 0,E_{a}] \cup [ E_{a},E_{c}] $, and $\tilde{D}_{2}(p,\lambda )\subset \{ E_{a}\} $. We define, for $E\in \tilde{D}_{1}(p,\lambda )$, the time map \begin{align*} T_{1}(\lambda ,E)&:= \int\nolimits_{0}^{r_{1}(\lambda ,E)}( G(\lambda ,E,u)) ^{-1/p}du \\ &=\int\nolimits_{0}^{r_{1}(\lambda ,E)}( E^{p}-p'\lambda F(u)) ^{-1/p}du \\ &=(p'\lambda )^{-1/p}\int_{0}^{r_{1}(\lambda ,E)}( F(r_{1}(\lambda ,E))-F(u)) ^{-1/p}du, \end{align*} since $G(\lambda ,E,r_{1}(\lambda ,E))=E^{p}-p'\lambda F(r_{1}(\lambda ,E))=0$. For all $\lambda >0$, $r_{1}(\lambda ,\cdot )$ is an increasing $C^{1}$-diffeomorphism from $( 0,E_{a}] $ onto $( 0,a] $ and from $( E_{a},E_{c}] $ onto $( \beta ,c] $. Thus $T_{1}$ may be written as \begin{equation*} T_{1}(p,\lambda ,E)=(p'\lambda )^{-1/p}S(p,r_{1}(\lambda ,E)) \text{ for } E\in \overline{D}_{1}(p,\lambda ), \end{equation*} where for all $p>1$, $S(p,\cdot )$ is defined by \begin{equation} S(p,\alpha ):=\int\nolimits_{0}^{\alpha }( F(\alpha )-F(u)) ^{-1/p}du\text{\ for\ all\ }\alpha \in ( 0,a] \cup ( \beta ,c] . \label{4.12} \end{equation} Note that $S(p,\cdot )$ takes its values in $[ 0,\infty ] $. We define, for $E\in \tilde{D}_{2}(p,\lambda )$, the time map \begin{equation*} T_{2}(p,\lambda ,E):=\int\nolimits_{0}^{r_{2}(\lambda ,E)}( G( \lambda ,E,u) ) ^{-1/p}du=(p'\lambda )^{-1/p}S(p,r_{2}(\lambda ,E)). \end{equation*} Note that, if $\tilde{D}_{2}(p,\lambda )\neq \emptyset $ then $\tilde{D}% _{2}(p,\lambda )=\{ E_{a}\} $ and $r_{2}(\lambda ,E)=\beta $. That is why we extend the definition domain of $S(p,\cdot )$ by including the eventual range of $r_{2}(\lambda ,\cdot );$ that is, we define $% S(p,\cdot )$ on $( 0,a] \cup [ \beta ,c] $. On the other hand, continuity arguments imply that if $\tilde{D}_{2}(p,\lambda )=\{ E_{a}\} $ then \begin{align*} T_{2}(p,\lambda ,E_{a}) &=(p'\lambda )^{-1/p}S(p,r_{2}(\lambda ,E))=(p'\lambda )^{-1/p}S(p,\beta )\\ &=\lim_{\alpha \to \beta^{+}}(p'\lambda )^{-1/p}S(p,\alpha ) =\lim_{E\to E_{a}^{+}}(p'\lambda)^{-1/p}S(p,r_{1}(\lambda ,E))\\ &=\lim_{E\to E_{a}^{+}}T_{1}(p,\lambda ,E). \end{align*} So we simply study the function $\alpha \mapsto S(p,\alpha )$ for $\alpha \in ( 0,a] \cup [ \beta ,c] $, and if $S(p,\alpha)<\infty $, we intend that \begin{equation*} S(p,\alpha )= \begin{cases} (p'\lambda )^{1/p}T_{1}(p,\lambda ,E_{\alpha }:=r_{1}^{-1}(\lambda ,\alpha )) & \text{if }\alpha \in ( 0,a] \cup ( \beta ,c] , \\ (p'\lambda )^{1/p}T_{2}(p,\lambda ,E_{a}) & \text{if }\alpha =\beta . \end{cases} \end{equation*} \begin{lemma} \label{Lemma4.8} $f'(a)<0$ and $f'(c)<0$. \end{lemma} The proof of Lemma \ref{Lemma4.8} is easy but tedious; we omit it. \begin{lemma} \label{Lemma4.9} \begin{itemize} \item[(i)] $S(p,0)=0$ if $p>1$. \item[(ii)] $S(p,a)=\infty $ if and only if $11$ and $0<\alpha 1$, $S(p,a)=\int_{0}^{a}( F(a)-F(u))^{-1/p}du$. Note that $F(a)-F(u)=-\frac{1}{2}f'(a)(a-u)^{2}+o((u-a)^{2})$ near $a^{-}$ and by Lemma \ref{Lemma4.8}, $f'(a)<0$. Therefore, \begin{equation*} ( F(a)-F(u)) ^{-1/p}\approx ( -f'(a)/2)^{-1/p}( a-u) ^{-2/p}\quad \text{near}\ a^{-}. \end{equation*} Then easy computation shows that $S(p,a)=\infty $\ if and only if $10$, $( F(\beta )-F(u)) ^{-1/p}\approx ( f(\beta )) ^{-1/p}(\beta -u)^{-1/p}$ near $\beta ^{-}$. Then easy computation shows that $\int_{\beta -\varepsilon }^{\beta }( F(\beta )-F(u))^{-1/p}du<\infty $ for $p>1$ and $\varepsilon >0$ sufficiently small. \noindent(\textbf{Eventual singularity at }$a^{-}$) Since $F(\beta )=F(a)$ by (H3), the same arguments as those used in the proof of part (ii) above imply that $\int_{0}^{a}( F(\beta )-F(u)) ^{-1/p}du=\infty $ if and only if $10$ sufficiently small. In above analysis, $S(p,\beta )=\infty $ if and only if $11$, $S(p,c)=\int_{0}^{c}( F(c)-F(u))^{-1/p}du$. Note that $F(c)-F(u)=-\frac{1}{2}f'(c)(c-u)^{2}+o((u-c)^{2})$ near $c^{-}$ and by Lemma \ref{Lemma4.8}, $f'(c)<0$. Therefore, \begin{equation*} ( F(c)-F(u)) ^{-1/p}\approx ( -f'(c)/2) ^{-1/p}( c-u) ^{-2/p}\quad \text{near }c^{-}. \end{equation*} Then an easy computation shows that $S(p,c)=\infty$ if and only if $11$, $S'(p,\alpha )$ is given by \begin{equation} S'(p,\alpha )=\frac{1}{p\alpha }\int_{0}^{\alpha }\frac{\theta _{p}(\alpha )-\theta _{p}(u)}{( F(\alpha )-F(u)) ^{1/p}}du\quad \text{for }\alpha \in ( 0,a) \cup (\beta ,c), \label{4.13} \end{equation} where $\theta _{p}(u)=pF(u)-uf(u)$. This implies \begin{gather*} \theta _{p}'(u)=( p-1) f(u)-uf'(u), \\ \theta _{p}''(u)=( p-2) f'(u)-uf'' (u). \end{gather*} Thus by (H1) and (H4), \begin{equation} \label{4.14} \begin{gathered} \theta _{p}(0)=0, \\ \theta _{p}'(0)=( p-1) f(0)>0, \\ \theta _{p}''(u) \begin{cases} >0 & \text{for }00 & \text{for }s_{p}0, \\ \theta _{p}'(c)=(p-1)f(c)-cf'(c)=-cf'(c)>0. \end{gather*} Hence there exist $t_{p}\in ( r_{p},s_{p}) $ and $\sigma _{p}\in ( s_{p},c) $ such that \begin{gather} \theta _{p}\quad \text{is strictly increasing on }( 0,t_{p}) , \label{4.15}\\ \theta _{p}\quad \text{is strictly decreasing on }( t_{p},\sigma _{p}), \label{4.16}\\ \theta _{p}\quad \text{is strictly increasing on }( \sigma _{p},c) .\label{4.17} \end{gather} In addition, there exist $\delta _{p}\in ( t_{p},\sigma _{p}) $ and $\gamma _{p}\in ( \sigma _{p},c) $ such that \begin{equation} \theta _{p}(\delta _{p})=\theta _{p}(\gamma _{p})=0. \label{4.18} \end{equation} The typical graph of $\theta _{p}(u)$ on $[0,c]$ is depicted in Fig. 7. \begin{lemma} \label{Lemma4.10} For $p>1$, the following statements hold \begin{itemize} \item[(i)] $S(p,\alpha )$ is strictly increasing on $( 0,a) $. \item[(ii)] $S(p,\alpha )$ has exactly one critical point, a minimum, on $(\beta ,c)$. More precisely, there exists a unique $m_{p}\in (\beta ,c)$ such that $S(p,\alpha )$ is strictly decreasing on $( \beta,m_{p}) $ and is strictly increasing on $( m_{p},c) $. \end{itemize} \end{lemma} \begin{proof} Part (i). By (\ref{4.13})-(\ref{4.15}), it suffices to show that \begin{equation} 01. \label{Cond} \end{equation} Note that \begin{equation*} \theta _{p}(a)=pF(a)-af(a)=pF(a)>0, \end{equation*} and by Lemma \ref{Lemma4.8}, \begin{equation*} \theta _{p}'(a)=(p-1)f(a)-af'(a)=-af'(a)>0, \end{equation*} then $a\in ( 0,t_{p}) \cup ( \gamma _{p},c) $ by \eqref{4.14}--\eqref{4.18}. If $a\in ( \gamma _{p},c) $ ($\subset ( \sigma _{p},c) $), then (\ref{4.17}) implies that $\theta _{p}$% \ is strictly increasing on $( a,c) $. Hence \begin{equation} \theta _{p}'(u)>0\quad \text{for }u\in ( a,c) . \label{additive} \end{equation} However, (H2) implies that there exists $\eta _{p}\in ( a,b) $ ($% \subset ( a,c) $) such that $f(\eta _{p})<0$\ and\ $f^{\prime }(\eta _{p})=0$. Therefore, \begin{equation*} \theta _{p}'(\eta _{p})=(p-1)f(\eta _{p})-\eta _{p}f' (\eta _{p})=(p-1)f(\eta _{p})<0\quad \text{ for }p>1, \end{equation*} which leads to a contradiction with (\ref{additive}). Therefore, (\ref{Cond}) holds and hence part (i) follows. Part (ii) follows by exactly the same arguments used to prove \cite[Lemma 4.7]{a3}. To this end, it suffices to prove the following two lemmas. \end{proof} \begin{lemma} \label{Lemma4.4a} For $p>1$, $S''(p,\alpha )+(p/\alpha)S'(p,\alpha )>0$ for all $\alpha \in ( \max \{ \sigma_{p},\beta \} ,c) $. \end{lemma} The proof of Lemma \ref{Lemma4.4a} is the same as that of \cite[Lemma 4.6]{a3}; we omit it. \begin{lemma} \label{Lemma4.5a} Assume that $p>1$. \begin{itemize} \item[(i)] If $p>2$ then $S'( p,c) =\infty $. \item[(ii)] If $\beta <\sigma _{p}$\ then $S'(p,\alpha )<0$ for all $\alpha \in (\beta ,\sigma _{p}]$. \item[(iii)] If $\beta =\sigma _{p}$\ then $S(p,\beta )<\infty $ then $-\infty \leq S'(p,\beta )<0$. \item[(iv)] If $\beta >\sigma _{p}\;$and $S(p,\beta )<\infty $ then $S'(p,\beta )=-\infty $. \end{itemize} \end{lemma} \begin{proof} The proofs of parts (i) and (iii) follow exactly as those of parts (i) and (iii) of \cite[Lemma 4.5]{a3}; we omit them. For part (ii) we point out that by (H3) ($\theta _{p}(\beta^{*})<0$ where $\beta ^{*}\leq \beta $) it follows that $\delta_{p}<\beta $. Then the argument used to prove Lemma 4.5(ii) of \cite{a3} can apply to prove $S'(p,\alpha )<0$ for all $\alpha \in (\beta ,\sigma_{p}]$. So part (ii) holds. Proof of part (iv). Since $F(\beta )=F(a)>0$, it follows that the integral representing $S'(p,\beta )$, has two singularities; one at $a$ and the other at $\beta $. So we write \begin{equation*} S'(p,\beta )=(p\beta) ^{-1}(I_{a^{-}}+I_{a^{+}}+I_{\beta }) , \end{equation*} where \begin{gather*} I_{a^{-}}:= \int_{0}^{a}\frac{\theta _{p}(\beta )-\theta _{p}(u)}{( F(\beta )-F(u)) ^{(p+1)/p}}du, \\ I_{a^{+}}:= \int_{a}^{(a+\beta )/2}\frac{\theta _{p}(\beta )-\theta _{p}(u)}{( F(\beta )-F(u)) ^{(p+1)/p}}du, \\ I_{\beta }:= \int_{(a+\beta )/2}^{\beta }\frac{\theta _{p}(\beta ) -\theta _{p}(u)}{( F(\beta )-F(u)) ^{(p+1)/p}}\,du. \end{gather*} We next show $I_{\beta }<\infty$ and $I_{a^{\pm }}=-\infty$. First we show $I_{\beta }<\infty $. Since $(c>)$ $\beta >\sigma _{p}$, it follows that $\theta _{p}'(\beta )>0$. Therefore, \begin{equation*} \frac{\theta _{p}(\beta )-\theta _{p}(u)}{( F(\beta )-F(u)) ^{(p+1)/p}}\approx \frac{\theta _{p}'(\beta )}{( f(\beta )) ^{(p+1)/p}}\frac{1}{(\beta -u)^{1/p}}\quad \text{near }\beta ^{-}. \end{equation*} Since $1/p<1$ and $\theta _{p}'(\beta )( f(\beta )) ^{-(p+1)/p}>0$, easy computations show that $I_{\beta }<\infty $. We then show $I_{a^{\pm }}=-\infty $. Note that \begin{equation*} \frac{\theta _{p}(\beta )-\theta _{p}(u)}{( F(\beta )-F(u)) ^{(p+1)/p}}\approx \frac{\theta _{p}(\beta )-\theta _{p}(a)}{( -f'(a)/2) ^{(p+1)/p}}\frac{1}{( a-u) ^{2(p+1)/p}}\quad \text{near }a. \end{equation*} Since $2(p+1)/p>1$ and \begin{equation*} \frac{\theta _{p}(\beta )-\theta _{p}(a)}{( -f'(a)/2)^{(p+1)/p}}= \frac{-\beta f(\beta)}{( -f'(a)/2)^{(p+1)/p}}<0, \end{equation*} easy computations show that $I_{a^{\pm }}=-\infty $. This completes the proof of Lemma \ref{Lemma4.5a}. Therefore, the proof Lemma \ref{Lemma4.10} is also complete. \end{proof} \begin{lemma} \label{Lemma4.11} For $p>2$, $0<\alpha _{p}<\lambda _{p}<\infty $. \end{lemma} \begin{proof} By (\ref{3.4}), (\ref{4.12}) and Lemma \ref{Lemma4.9}(iii), for $p>2$, \begin{equation*} \lambda _{p}=\big\{ \int_{0}^{\beta }( F(\beta )-F(u)) ^{-1/p}du\big\} ^{p}/p'=\big\{ \lim_{\alpha \to \beta ^{+}}S(p,\alpha )\big\} ^{p}/p'<\infty . \end{equation*} We then find that \begin{align*} \lambda _{p} &=\{ \int_{0}^{\beta }( F(\beta )-F(u))^{-1/p}du\} ^{p}/p' \\ &=\{ \int_{0}^{a}( F(\beta )-F(u))^{-1/p}du+\int_{a}^{\beta }( F(\beta )-F(u)) ^{-1/p}du\} ^{p}/p' \\ &>\{ \int_{0}^{a}( F(\beta )-F(u)) ^{-1/p}du\}^{p}/p' \\ &=\{ \int_{0}^{a}( F(a)-F(u)) ^{-1/p}du\}^{p}/p' \quad\text{(since }F(\beta )=F(a)\text{)} \\ &=\alpha _{p}>0\quad \text{(by (\ref{3.3})).} \end{align*} This completes the proof. \end{proof} Let $u$ be a positive solution of \eqref{1.1}, then $0<\left\Vert u\right\Vert _{\infty }\leq a$ or $\beta \leq \left\Vert u\right\Vert _{\infty }\leq c$. In addition, $u\in A_{0}^{+}\cup A_{1}^{+}\cup A_{00}^{+}\cup A_{01}^{+}$ by Lemma \ref{Lemma2.2}. By Lemma \ref{Lemma4.9}(ii)-(iv), for $p>2$, $S(p,a)<\infty$, $S(p,\beta )<\infty$, $S(p,c)<\infty$. In this case we have the following three statements: \noindent\textbf{(i)} Suppose for $\lambda =\alpha _{p}=( S(p,a)) ^{p}/p'$, $u_{\alpha _{p}}$ is the corresponding solution of (\ref {1.1}) satisfying $\| u_{\alpha _{p}}\| _{\infty }=u_{\alpha _{p}}(0)=a$. Then \begin{equation*} u_{\alpha _{p}}'(x)=\{ p'\alpha _{p}[F(a)-F(u(x))] \} ^{1/p}>0\quad \text{ for }-1\leq x<0 \end{equation*} by Lemma \ref{Lemma2.1}. So $u_{\alpha _{p}}\in A_{0}^{+}$. Then for each $\lambda >\alpha _{p}$, \begin{equation*} u_{\lambda }(x):=\begin{cases} u_{\alpha _{p}}\big( ( \frac{\lambda }{\alpha _{p}}) ^{1/p}( | x| -1+( \frac{\alpha _{p}}{\lambda }) ^{1/p}) \big) & \text{if }1-( \frac{\alpha _{p}}{\lambda }) ^{1/p}<| x| \leq 1, \\ a & \text{if }| x| \leq 1-( \frac{\alpha _{p}}{\lambda }) ^{1/p} \end{cases} \end{equation*} is a $C^{1}$ dead core solution of \eqref{1.1} satisfying $\|u_{\lambda }\| _{\infty }=a$, \begin{equation*} u_{\lambda }'(-1)=( p'\lambda F(a))^{1/p}>( p'\alpha _{p}F(a)) ^{1/p}=u_{\alpha _{p}}'(-1)>0, \end{equation*} and $u_{\lambda }\in A_{1}^{+}$. \noindent\textbf{(ii)} Suppose for $\lambda =\nu _{p}=( S_{p}(c)) ^{p}/p'$, $u_{\nu _{p}}$ is the corresponding solution of \eqref{1.1} satisfying $\| u_{\nu _{p}}\| _{\infty }=u_{\nu _{p}}(0)=c$. Then \begin{equation*} u_{\nu _{p}}'(x)=\{ p'\nu _{p}[ F(c)-F(u(x))] \} ^{1/p}>0\text{ for }-1\leq x<0 \end{equation*} by Lemma \ref{Lemma2.1}. So $u_{\nu _{p}}\in A_{0}^{+}$. Then for each $\lambda >\nu _{p}$, \begin{equation*} u_{\lambda }(x):=\begin{cases} u_{\nu _{p}}\big( ( \frac{\lambda }{\nu _{p}}) ^{1/p}( | x| -1+( \frac{\nu _{p}}{\lambda }) ^{1/p}) \big) & \text{if }1-( \frac{\nu _{p}}{\lambda }) ^{1/p}<| x| \leq 1, \\ c & \text{if }| x| \leq 1-( \frac{\nu _{p}}{\lambda }) ^{1/p} \end{cases} \end{equation*} is a $C^{1}$ dead core solution of \eqref{1.1} satisfying $\|u_{\lambda }\| _{\infty }=c$, \begin{equation*} u_{\lambda }'(-1)=( p'\lambda F(c)) ^{1/p}>( p'\nu _{p}F(c)) ^{1/p}=u_{\nu _{p}}^{\prime }(-1)>0, \end{equation*} and $u_{\lambda }\in A_{1}^{+}$. \noindent\textbf{(iii)} Suppose for $\lambda =\lambda _{p}=(S(p,\beta ))^{p}/p'$, $u_{\lambda _{p}}\,$is the corresponding solution of \eqref{1.1} satisfying $\| u_{\lambda _{p}}\| _{\infty }=u_{\lambda_{p}}(0)=\beta $. Then, by Lemma \ref{Lemma2.1}, there exists a unique negative number $-x_{0}\in (-1,0)$ such that \begin{equation*} u_{\lambda _{p}}(-x_{0})=a\text{ and }u_{\lambda _{p}}'(-x_{0})=0 \end{equation*} and \begin{equation*} u_{\lambda _{p}}'(x)=\{ p'\lambda _{p}[ F(\beta)-F(u(x))] \} ^{1/p}>0\quad \text{ for }x\in [ -1,0) -\{ -x_{0}\} . \end{equation*} So $u_{\lambda _{p}}\in A_{00}^{+}$. Then for each $\lambda >\lambda _{p}$, \begin{equation*} u_{\lambda }(x):=\begin{cases} u_{\lambda _{p}}\big( ( \frac{\lambda }{\lambda _{p}})^{1/p}| x| \big) & \text{if }| x| \leq ( \frac{K_{0a}}{\lambda }) ^{1/p}, \\ a & \text{if }( \frac{K_{0a}}{\lambda }) ^{1/p}\leq |x| \leq 1-( \frac{K_{a\beta }}{\lambda }) ^{1/p}, \\ u_{\lambda _{p}}\Big( \frac{| x| -1 +( \frac{K_{a\beta }}{\lambda }) ^{1/p} +( \frac{K_{a\beta }}{\lambda _{p}}) ^{1/p}}{( \frac{K_{a\beta }}{\lambda }) ^{1/p} +( \frac{K_{a\beta }}{\lambda _{p}}) ^{1/p}}\Big) & \text{if }1-( \frac{K_{a\beta } }{\lambda }) ^{1/p}\leq | x| \leq 1 \end{cases} \end{equation*} is a positive solution of \eqref{1.1} satisfying $\| u_{\lambda}\| _{\infty }=\beta $, \begin{equation*} u_{\lambda }'(-1)=( p'\lambda F(\beta )) ^{1/p}>( p'\lambda _{p}F(\beta )) ^{1/p}=u_{\lambda _{p}}'(-1)>0, \end{equation*} and $u_{\lambda }\in A_{01}^{+}$, where \begin{eqnarray*} &&K_{0a}:= ( \int_{0}^{a}( F(\beta )-F(u)) ^{-1/p}du) ^{p}/p', \\ &&K_{a\beta }:= ( \int_{a}^{\beta }( F(\beta )-F(u)) ^{-1/p}du) ^{p}/p'. \end{eqnarray*} Hence Theorems \ref{Thm3.5}-\ref{Thm3.9} follow immediately by Theorem \ref{Thm2.2} and Lemmas \ref{Lemma4.9}-\ref{Lemma4.11}. \subsection*{Acknowledgments} The authors thank the anonymous referee for his/her valuable remarks. Much of the computation in this paper has been checked using the symbolic manipulator Mathcad 7 Professional for Addou and Mathematica 4.0 for Wang. \begin{thebibliography}{99} \bibitem{a1} I. Addou,\ Multiplicity of solutions for quasilinear elliptic boundary-value problems,\ \textit{Electron. J. Differential Equations }% \textbf{1999} (1999), No. 21, 1-27. \bibitem{a2} I. Addou, Exact multiplicity results for quasilinear boundary-value problems with cubic-like nonlinearities, \textit{Electron. J. Differential Equations} \textbf{2000} (2000), No. 01, 1-26, (Addendum, 27-29). \bibitem{a3} I. Addou, and S.-H. Wang, Exact multiplicity results for some $p $-Laplacian nonpositone problems with concave-convex-concave nonlinearities. \textit{Nonlinear Analysis} \textbf{53} (2003), 111-137. \bibitem{g1} M. Guedda and L. Veron,\ Bifurcation phenomena associated to the $p$-Laplace operator,\ \textit{Trans. Amer. Math. Soc.} \textbf{310} (1988), 419-431. \bibitem{j1} N. Jacobson, \textit{Basic Algebra}, Freeman, New York, 1989. \bibitem{k1} P. Korman, Y. Li and T. Ouyang,\ Exact multiplicity results for boundary problems with nonlinearities generalizing cubic, \textit{Proc. Roy. Soc. Edinburgh} \textbf{126A} (1996), 599-616. \bibitem{k2} P. Korman, Y. Li and T. Ouyang,\ Perturbation of global solution curves for semilinear problems, \textit{Adv. Nonlinear Stud.} \textbf{3} (2003), 289--299. \bibitem{k3} P. Korman and J. Shi,\ Instability and exact multiplicity of solutions of semilinear equations, \textit{Electron. J. Diff. Equ. Conf.,} \textbf{5} (2000), 311-322. \bibitem{s1} J. Smoller and A. Wasserman,\ Global bifurcation of steady-state solutions, \textit{J. Differential Equations} \textbf{39} (1981), 269-290. \bibitem{w1} S.-H. Wang, A correction for a paper by J. Smoller and A. Wasserman, \textit{J. Differential Equations} \textbf{77} (1989), 199-202. \bibitem{w2} S.-H. Wang, On the time map of a nonlinear two-point boundary value problem, \textit{Differential Integral Equations} \textbf{7} (1994), 49-55. \bibitem{w3} S.-H. Wang and N. D. Kazarinoff, Bifurcation of steady-state solutions of a scalar reaction-diffusion equation in one space variable, \textit{J. Austral. Math. Soc.} (Series A) \textbf{52} (1992), 343-355. \end{thebibliography} \end{document}