\documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2005(2005), No. 19, pp. 1--21.\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 2005 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2005/19\hfil Existence and uniqueness of solutions] {Existence and uniqueness of solutions to a super-linear three-point boundary-value problem} \author[B. Calvert, C. P. Gupta\hfil EJDE-2005/19\hfilneg] {Bruce Calvert, Chaitan P. Gupta} % in alphabetical order \address{Bruce Calvert \hfill\break Department of Mathematics\\ University of Auckland\\ Auckland, New Zealand} \email{calvert@math.auckland.ac.nz} \address{Chaitan P. Gupta \hfill\break Department of Mathematics, 084\\ University of Nevada, Reno, NV 89557, USA} \email{gupta@unr.edu} \date{} \thanks{Submitted November 16, 2004. Published February 7, 2005.} \subjclass[2000]{34B10, 34B15} \keywords{Super-linear; three-point boundary-value problem} \begin{abstract} In previous papers, degree theory for nonlinear operators has been used to study a class of three-point boundary-value problems for second order ordinary differential equations having a super-linear term, and existence of a sequence of solutions has been shown. In this paper, we forgo the previous approach for the shooting method, which gives a drastically simpler existence theory, with less assumptions, and easy calculation of solutions. We even obtain uniqueness in the simplest case. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{definition}[theorem]{Definition} \section{Introduction} In the papers \cite{cap,cap2,hen1,Hthesis,hen2} the authors use degree theory to give existence of a sequence of solutions to a super-linear boundary value problem. More specifically, in \cite{Hthesis,hen2} they give existence of solutions to \begin{gather} x''+g(x) =p(t,x,x') \label{1} \\ x(0) =0\mbox,\quad x(\eta )=\beta x(1) \label{2} \end{gather} Here $\eta \in (0,1)$, making this a three point boundary value problem. The function $g$ is assumed to be super-linear, that is, it satisfies $g(x)/x\to \infty $ as $|x|\to \infty $, and $\beta=1$. In \cite{cal} the case $\beta \neq 1$ is argued along similar lines. In this paper, we obtain existence of solutions to (\ref{1}), (\ref{2}) for $\beta \neq 1$ via the intermediate value theorem, i.e. the shooting method, giving a drastically simpler existence theory, with less assumptions. Calculation of solutions numerically may be carried out by the shooting method. The shooting method is used theoretically in \cite{din,kol,kwo}, and elsewhere. Uniqueness is studied by Kwong in \cite{kwo}, which recovers results such as Moroney's theorem, giving uniqueness of a positive solution of a boundary-value problem involving a superlinear function. This builds on Kolodner's paper \cite{kol}, which gave the exact number of solutions of a rotating string problem, given the angular velocity. Similarly, in \cite{cos}, the boundary value problem \begin{gather*} x''+\lambda x^+-\alpha x^- = \sin(t) \\ x(0)=x(\pi )=0 \end{gather*} was shown to have exactly $2k$ solutions if $0<\alpha < 1$ and $k^2 < \lambda < (k+1)^2$, and Dinca and Sanchez \cite{din} pose the question of whether this uniqueness result can be obtained by elementary methods. Our uniqueness result, giving uniqueness of solutions to (\ref{1}) and (\ref{2}) in case $p=0$, is elementary and presumably new. Our approach does not readily lend itself to the case of nonzero $p$, and this gives an open question. There has been much recent work on 3-point boundary value problems, and much of it has concentrated on positive solutions, as in \cite{he,inf,kos,ma}. He and Ge \cite{he} give the existence of three positive solutions to the B.V.P. (\ref{1}), (\ref{2}), but the condition (\ref{3.02}) of our uniqueness theorem and their conditions (D2), (D3) cannot hold at the same time. Thus their work cannot be used to show that Theorem \ref{Theorem2} may not hold for all $k$. Similarly Infante and Webb \cite[Th 4.2]{inf} cannot be used because their conditions ($S_1$) and ($S_2$) are incompatible with (\ref{3.02}). Ma \cite{ma} shows that one can get existence of positive solutions to the B.V.P. (\ref{1}), (\ref{2}), assuming $g(x)/x \to 0$ as $x \to 0$, and $p=0$, which does show that one can obtain existence theorems like our Theorem \ref{Theorem1} for small $k$. Infante and Webb \cite{inf} show that one need not have positive coefficients in an m-point boundary value problem, and in this work we can indeed take $\beta$ to be negative. Capietto and Dambrosio \cite{cap} consider the case of asymmetric $g(x)$, superlinear for positive $x$, and give an extensive review of superlinear boundary value problems. \section{Assumptions and Preliminaries} A background on o.d.e.s involving functions satisfying Caratheodory's conditions is given in Chapter 18 of \cite{kur}. \noindent{\bf Assumption A}: - Assume that $g:{\mathbb{R}} \mapsto {\mathbb{R}}$ is a continuous super-linear function, that is, it satisfies $\frac{g(x)}{x}\to \infty $ as $|x|\to \infty $. Let $p:[0,1]\times \mathbb{R}^{2}\to \mathbb{R}$ be a function satisfying Caratheodory's conditions, i.e. for every $(x,y)\in \mathbb{R}^{2}$, $p(t,x,y)$ is Lebesgue measurable in $t$, and for a.e. $t\in [ 0,1]$, $p(t,x,y)$ is continuous in $(x,y)$. Suppose there exists an $M_{1}:[0,1]\times [ 0,\infty )\mapsto [ 0,\infty )$ such that (a) for each $s\in [ 0,\infty )$, $M_{1}(\cdot ,s)$ is integrable on $[0,1]$, (b) for each $t\in [ 0,1]$, $M_{1}(t,\cdot )$ is increasing on $[0,\infty )$ with $s^{-1}\int_{0}^{1}M_{1}(t,s)dt\to 0$ as $s\to \infty$, and (c) for all $t\in [ 0,1]$, and $(x,y)\in \mathbb{R}^{2}$, \[ |p(t,x,y)|\leq M_{1}(t,\max (|x|,|y|))\mbox{.} \] We need the next result, proved in \cite{cal} as Lemma 2. \begin{lemma} \label{Lemma1} Let $g$, $p$, and $M_{1}$ satisfy Assumption A. Suppose that $\frac{g(x)}{x}\geq 1$ for $x\neq 0$. Suppose that $(x(t),y(t))$ is an absolutely continuous solution for the initial value problem \begin{gather} x'(t) =y(t), \label{eq1-1} \\ y'(t)+g(x(t)) = p(t,x(t),y(t)),\quad \text{for a.e. }t\in [ 0,1] , \label{eq1-2} \\ x(0) =0\,, \label{eq1-3} \\ y(0) =\alpha \,. \label{eq1-4} \end{gather} For $x\in \mathbb{R}$, let $G(x)=\int_{0}^{x}g(s)ds$. Let $\varepsilon >0$ be given. Then for $\alpha >0$, large enough, we have \begin{gather} |y(t)| \leq \alpha (1+\varepsilon ) \label{eq2-1} \\ 2G(x(t)) \leq \alpha ^{2}(1+\varepsilon ), \label{eq2-2} \end{gather} for every $t\in [ 0,1]$. Moreover, \begin{equation} |\frac{d}{dt}(y^{2}(t)+2G(x(t)))|\leq 2|y(t)|M_{1}(t,\max (|x(t)|,|y(t)|)), \label{eq3} \end{equation} for $t\in [ 0,1]$ a.e. \end{lemma} We note that if we assume $g$ continuous and $g(x)/x \geq 1$ then the function $G(x)=\int_{0}^{x}g(s)ds$ is defined for $x\in \mathbb{R}$ and is such that $G$ is strictly increasing on $[0,\infty )$ and is strictly decreasing on $(-\infty ,0]$. Also, $G(x)>0$ for $x\in \mathbb{R}$, $x\neq 0$ and $G(0)=0$. \ We denote the inverse of the function $G$ restricted to $[0,\infty )$, $G\big|_{[0,\infty )}$, by $G_{+}^{-1}$ and the inverse of the function $G\big|_{(-\infty ,0]}$ by $G_{-}^{-1}$. We now need a new version of \cite[Lemma 3]{cal}, in which (\ref{eq7}) and (\ref{eq8}) replace (13) and (14) of \cite{cal}. \begin{lemma} \label{Lemma2} Let $\varepsilon >0$ be given and $g$, $p$, $M_{1}$ be as in Lemma \ref{Lemma1}. Then there exists an $A>0$ such that if $(x(t),y(t))$ is a solution for the initial value problem \eqref{eq1-1}, \eqref{eq1-2}, \eqref{eq1-3}, \eqref{eq1-4} and $t_{0}\in (0,1]$ is such that $x(t_{0})>0$, $y(t_{0})=0$; then \begin{equation} G_{+}^{-1}(\frac{\alpha ^{2}}{2(1+\varepsilon )})\leq x(t_{0})\leq G_{+}^{-1}(\frac{\alpha ^{2}}{2}(1+\varepsilon )) \label{eq7} \end{equation} if $|\alpha |>A$. \ Similarly, if $x(t_{0})<0$, $y(t_{0})=0$; then \begin{equation} G_{-}^{-1}(\frac{\alpha ^{2}}{2(1+\varepsilon )})\leq x(t_{0})\leq G_{-}^{-1}(\frac{\alpha ^{2}}{2}(1+\varepsilon )) \label{eq8} \end{equation} if $|\alpha |>A$. Also, \[ \min_{t\in [ 0,1]}\sqrt{x^{2}(t)+y^{2}(t)}\geq \frac{1}{2}\min \{G_{+}^{-1}(\frac{\alpha ^{2}}{8}),\frac{\alpha }{2}\}. \] \end{lemma} \begin{proof} We observe that the right inequality in (\ref{eq7}) follows immediately from (\ref{eq2-2}). Accordingly, it suffices to show that \begin{equation} \frac{\alpha ^{2}}{2}\leq (1+\varepsilon )G(x(t_{0})), \label{eq9} \end{equation} to prove that (\ref{eq7}) holds. Let us choose $A>0$, such that for $|\alpha |>A$, both (\ref{eq2-1}) and (\ref{eq2-2}) hold with $\alpha $ replaced by $|\alpha |$. With $h(t):=\sqrt{y^2(t)+2G(x(t)) }$, we get, by integrating (\ref{eq3}) from $0$ to $t$ and using (\ref{eq2-1}), \begin{equation} h^{2}(t)-\alpha ^{2}+2|\alpha |(1+\varepsilon )\int_{0}^{t}M_{1}(s,\max (|x(s)|,|y(s)|))ds\geq 0. \nonumber \label{eq10} \end{equation} We now take an $\varepsilon _{1}>0$ such that $2\varepsilon _{1}(1+\varepsilon )^{2}\leq \min \{\frac{\varepsilon }{1+\varepsilon }, \frac{1}{2}\}$. Next, we use the assumption $s^{-1}\int_{0}^{1}M_{1}(t,s)dt \to 0$ as $s\to \infty $, from Assumption A, to choose an $A>0$ so that for $\alpha >A$ the inequalities (\ref{eq2-1}), (\ref{eq2-2}) hold for $s\in [ 0,1]$. When for $s\in [ 0,1]$ \begin{equation} \max \{|x(s)|,|y(s)|\}\geq \frac{1}{2}\min \{G_{+}^{-1}(\frac{\alpha ^{2}}{8}),\frac{\alpha }{2}\} \label{eq11} \end{equation} we have, with $M(x):=\int_0^1M_1 (t,x)dt$, \begin{equation} M(\max \{|x(s)|,|y(s)|\})<\varepsilon _{1}\max \{|x(s)|,|y(s)|\}. \nonumber \label{eq12} \end{equation} For $\alpha >A$, we get on using the inequalities (\ref{eq2-1}), (\ref{eq2-2}), (\ref{eq12}) and the assumption $\frac{g(x)}{x}\geq 1$ for $x\neq 0$, that \begin{equation} M(\max \{|x(s)|,|y(s)|\})<\varepsilon _{1}\alpha (1+\varepsilon ), \label{eq13} \end{equation} and hence, using (\ref{eq10}), we get \begin{equation} h^{2}(t)\geq \alpha ^{2}(1-2\varepsilon _{1}(1+\varepsilon )^{2}), \label{eq14} \end{equation} provided (\ref{eq11}) holds for all $s\in [ 0,t]$. Since we chose $\varepsilon _{1}>0$ such that $2\varepsilon _{1}(1+\varepsilon )^{2}\leq \min \{\frac{\varepsilon }{1+\varepsilon },\frac{1}{2}\}$, we see that \[ y^{2}(t)+2G(x(t))\geq \frac{\alpha ^{2}}{2}, \] provided (\ref{eq11}) holds for all $s\in [ 0,t]$. Accordingly, either $y^{2}(t)\geq \frac{\alpha ^{2}}{4}$ or $|x(t)|\geq G_{+}^{-1}(\frac{\alpha ^{2}}{8})$ and hence \begin{equation} \max \{|x(t)|,|y(t)|\}\geq \min \{G_{+}^{-1}(\frac{\alpha ^{2}}{8}),\frac{ \alpha }{2}\}, \label{eq15} \end{equation} provided (\ref{eq11}) holds for all $s\in [ 0,t]$. We observe that ( \ref{eq11}) holds near $s=0$ since $y(0)=\alpha $. Let us next assume that (\ref{eq11}) holds for all $s\in [ 0,t]$, for some $t\in (0,1]$. If $0t$ such that (\ref{eq11}) holds for all $s\in [ 0,t_{1}]$. Accordingly, it follows that (\ref{eq11}) holds for all $s\in [ 0,1]$. Finally, if $y(t_{0})=0$, we see from (\ref{eq14}) and the assumption that $2\varepsilon _{1}(1+\varepsilon )^{2}\leq \min \{\frac{\varepsilon }{1+\varepsilon }, \frac{1}{2}\}$, that \[ \frac{\alpha ^{2}}{2}\leq (1+\varepsilon )G(x(t_{0})), \] and (\ref{eq9}) holds. This completes the proof that (\ref{eq7}) holds. A similar proof works to prove that (\ref{eq8}) holds. \end{proof} \begin{definition} \label{def}\rm For ${\bf u}(t)=(x(t),y(t))\in C^{1}([0,1],\mathbb{R}^{2}\backslash \{(0,0)\})$ we define \[ \varphi _{1}({\bf u})=\varphi _{1}(x,y)=-\int_{0}^{1}\frac{x(t)y'(t)-y(t)x'(t)}{x^{2}(t)+y^{2}(t)}dt, \] as the angle traversed clockwise from ${\bf u}(0)$ to ${\bf u}(1)$. \end{definition} We need a variant of \cite[Lemma 4.3]{hen1} and of \cite[Lemma 3]{hen2} to show that the angle $\varphi _{1}(x,y)\to \infty $, for solutions $(x,y)$ to (\ref{eq1-1})-(\ref{eq1-2}), when $\min_{t\in [ 0,1]}\|(x(t),y(t))\| \to \infty $. We use the following assumption: \smallskip \noindent{\bf Assumption B}: Let $g:\mathbb{R}\to \mathbb{R}$ be continuous and super-linear. Let $p:[0,1]\times \mathbb{R}{\bf \times }\mathbb{R} \to \mathbb{R}$ be a function satisfying Caratheodory's conditions. \ Suppose that there exists a $\mu \in (0,1]$, $\beta \geq 0$, $0\leq \gamma \in L^{1}[0,1]$, and $M_{2}:\mathbb{R}^{2}\to \mathbb{R}$, with $\frac{M_{2}(x,y)}{\|(x,y)\|}\to 0$ as $\|(x,y)\|\to \infty $, such that for a.e. $t\in [ 0,1]$, and $(x,y)\in \mathbb{R}^{2}$, \begin{equation} \mathop{\rm sign}(x)p(t,x,y)\leq (1-\mu ){\rm sign}(x)g(x)+\beta |y|+\gamma (t)M_{2}(x,y). \label{eq17} \end{equation} The inequality (\ref{eq17}) corresponds to inequality (4.3) in \cite{hen1}, i.e., \begin{equation} |p(t,x,y)|\leq (1-\mu )|g(x)|+\beta |y|+\gamma , \label{eq18} \end{equation} where $\gamma \in \mathbb{R}$. We shall provide the slight modifications needed in the proof of Lemma 4.3 of \cite{Hthesis}, and Lemma 3 of \cite{hen2} to cater for the difference between (\ref{eq17}) and (\ref{eq18}). \begin{lemma} \label{Lemma3} Suppose $g$ and $p$ satisfy Assumption B. Then for all $N\geq 0 $ there exists an $R\geq 0$ such that for all absolutely continuous solutions $(x(t),y(t))$ for the system \eqref{eq1-1}, \eqref{eq1-2} with $\min_{t\in [ 0,1]}\| (x(t),y(t))\| \geq R$, we have $\varphi _{1}(x,y)\geq N$. \end{lemma} \begin{proof} Since $\varphi _{1}({\bf u})=-\int_{0}^{1} \frac{x(t)y'(t)-y(t)x'(t)}{x^{2}(t)+y^{2}(t)}dt$, we see using (\ref{eq1-1}),(\ref{eq1-2}) that \[ -x(t)y'(t)+y(t)x'(t)=y^{2}(t)+x(t)g(x(t))-x(t)p(t,x(t),y(t)). \] Let us set \[ \theta (t)-\theta (0)=\int_{0}^{t}\frac{x(s)y'(s)-y(s)x'(s)}{x^{2}(s)+y^{2}(s)}ds, \] so that \begin{align*} -\theta '(t) &= -\frac{x(t)y'(t)-y(t)x'(t)}{ x^{2}(t)+y^{2}(t)} \\ &= \frac{y^{2}(t)+x(t)g(x(t))-x(t)p(t,x(t),y(t))}{x^{2}(t)+y^{2}(t)}. \end{align*} Let $N>0$ be given. Since $g$ is super-linear, we have for $K>0$ (to be chosen later) there is an $M=M(K)$ such that if $|x|\geq M$ then $\mu \frac{g(x)}{x}\geq K$. Hence \[ \mu \frac{g(x)}{x}+\frac{KM^{2}}{x^{2}}\geq K\text{ } \] for all $x\neq 0$, and $\mu xg(x)\geq Kx^{2}-KM^{2}$ for all $x\in \mathbb{R}$. Hence, \begin{align*} &y^{2}+xg(x)-x(t)p(t,x,y) \\ &\geq y^{2}+xg(x)-(1-\mu )xg(x)-\beta |x\|y|-\gamma (t)|x|M_{2}(x,y) \\ &\geq y^{2}+Kx^{2}-KM^{2}-\frac{\beta }{2}(\beta x^{2}+\frac{y^{2}}{\beta} )-\gamma (t)|x|M_{2}(x,y) \\ &\geq \frac{y^{2}}{2}+(K-\frac{\beta ^{2}}{2})x^{2}-KM^{2}-\gamma (t)|x|M_{2}(x,y) \\ &= \frac{y^{2}}{2}+\frac{k}{2}x^{2}-KM^{2}-\gamma (t)|x|M_{2}(x,y), \end{align*} where $k=2(K-\frac{\beta ^{2}}{2})$. Then, \begin{align*} -\theta '(t) &= \frac{y^{2}(t)+x(t)g(x(t))-x(t)p(t,x(t),y(t))}{ x^{2}(t)+y^{2}(t)} \\ &\geq \frac{\frac{y^{2}}{2}+\frac{k}{2}x^{2}-KM^{2}-\gamma (t)|x|M_{2}(x,y)}{x^{2}(t)+y^{2}(t)} \\ &\geq \frac{k}{2}(\frac{x^{2}+k^{-1}y^{2}}{x^{2}+y^{2}})-\frac{1}{\sqrt{k}}- \frac{\gamma (t)M_{2}(x,y)}{\|(x,y)\|}, \end{align*} assuming $\min_{t}\|(x(t),y(t))\|\geq M\sqrt{K}\sqrt[4]{k}$. We can then write \[ -\theta '(t)\geq \frac{k}{2}(\cos ^{2}\theta +k^{-1}\sin ^{2}\theta )-\frac{1}{\sqrt{k}}-\frac{\gamma (t)M_{2}(x,y)}{\|(x,y)\|}. \] Next we estimate \[ \int_{\theta (1)}^{\theta (0)}\frac{d\theta }{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta }, \] rather than estimating the integral $\int_{\theta (1)}^{\theta (0)}d\theta$. Since $\frac{1}{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta }\leq k$ we get \[ -\frac{\theta '(t)}{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta } \geq \frac{k}{2}-\sqrt{k}-\frac{\gamma (t)M_{2}(x,y)k}{\|(x,y)\|} \geq \frac{k}{2}-\sqrt{k}-\gamma (t), \] assuming $\frac{M_{2}(x,y)}{\|(x,y)\|}\leq \frac{1}{k}$, which holds if $\min_{t}\|(x(t),y(t))\|\geq \xi (k)$, say. Note \begin{align*} \int_{0}^{\frac{\pi }{2}}\frac{d\theta }{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta } &= \int_{0}^{\frac{\pi }{2}}\frac{k\sec ^{2}\theta d\theta }{ k+\tan ^{2}\theta } \\ &= \int_{0}^{\infty }\frac{kdu}{k+u^{2}} \\ &= \frac{k}{\sqrt{k}}\tan ^{-1}(\frac{u}{\sqrt{k}})|_{0}^{\infty } \\ &= \frac{\sqrt{k}\pi }{2}. \end{align*} Since $\cos ^{2}\theta +k^{-1}\sin ^{2}\theta $ has period $\pi $, given an interval $(a,b)$ we write $b-a=(n-f)\pi $, where $n$ is an integer and $f\in [ 0,1)$. \ Then \begin{align*} \int_{a}^{b}\frac{d\theta }{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta } &\leq \int_{0}^{n\pi }\frac{d\theta }{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta } \\ &\leq 2n\frac{\sqrt{k}\pi }{2} \\ &= \sqrt{k}\pi (\frac{b-a}{\pi }+f) \\ &\leq \sqrt{k}(b-a+\pi ). \end{align*} In particular, we get \[ -\int_{\theta (0)}^{\theta (1)}\frac{d\theta }{\cos ^{2}\theta +k^{-1}\sin ^{2}\theta }\leq \sqrt{k}(\theta (0)-\theta (1)+\pi ). \] Next we change the variable of integration from $\theta $ to $t$ to get \begin{align*} \sqrt{k}(\varphi _{1}(x,y)+\pi ) &\geq -\int_{0}^{1}\frac{\theta '(t)dt}{\cos ^{2}\theta (t)+k^{-1}\sin ^{2}\theta (t)} \\ &\geq \int_{0}^{1}(\frac{k}{2}-\sqrt{k}-\gamma (t))dt \\ &= \frac{k}{2}-\sqrt{k}-\int_{0}^{1}\gamma . \end{align*} This gives \begin{align*} \varphi _{1}(x,y) &\geq \frac{\sqrt{k}}{2}-1-\int_{0}^{1}\gamma -\pi, \quad \text{if }k\geq 1, \\ &= N,\quad \text{if }k=4(N+1+\int_{0}^{1}\gamma +\pi )^{2}\geq 1. \end{align*} Since \[ k=2(K-\frac{\beta ^{2}}{2}), \] we choose $K$ $=2(N+1+\int_{0}^{1}\gamma +\pi )^{2}+\frac{\beta ^{2}}{2}$. Finally choosing \[ R=\max (M\sqrt{K}\sqrt[4]{k},\xi (k)), \] we see that $\varphi _{1}(x,y)\geq N$ if $\min_{t\in [ 0,1]}\| (x(t),y(t))\| \geq R$. \end{proof} \section{Existence of Solutions} To fix ideas we study the case $\beta > 1$, as in \cite{cal}. \begin{theorem}\label{Theorem1} Let $\eta \in (0,1)$ and $\beta >1$ be given. Let $g$ and $p$ satisfy Assumptions A and B. Then, for each $k$ sufficiently large, there are (at least) two solutions ${\bf u}(t)=(x(t),y(t))$ of \begin{gather} x'(t) = y(t) \label{2.1} \\ y'(t) = -g(x(t))+p(t,x(t),y(t)) \label{2.2} \\ x(0) = 0 \label{2.3} \\ x(\eta ) = \beta x(1) \label{2.4} \end{gather} with $\varphi _{1}({\bf u})\in (\frac{\pi }{2}+k\pi ,\frac{\pi }{2}+(k+1)\pi)$, one with $x'(0)>0$ and the other with $x'(0)<0$. \end{theorem} We may, when thinking of calculating these solutions, say that we have one sequence $x_{n}$ of solutions to (\ref{1}), (\ref{2}) with $x_{n}'(0)\to \infty $ and another sequence $x_{n}$ of solutions to (\ref {1}), (\ref{2}) with $x_{n}'(0)\to -\infty $, with the angles they traverse as above. We break the proof of the theorem into two parts. We first prove existence of the solution $x$ when $p$ is smooth enough, and then we see that we can approximate $p$ by a sequence of smooth $p_{n}$, giving solutions $x_{n}$, and then then take limits to obtain existence when $p$ is not smooth. A sufficiently smooth $p$ will satisfy the following Caratheodory-Lipschitz condition. \begin{definition} \cite{kur} \rm Let $U$ be open in $\mathbb{R}^{n}$, and let $[a,b]$ be an interval of real numbers. Let $F:[a,b]\times U\to \mathbb{R}^{n}$ be given. We say $F$ satisfies a Caratheodory-Lipschitz condition if for all $x$, $t\mapsto F(t,x)$ is Lebesgue measurable, and for any $(t_{0},x_{0})\in D$, there are real valued integrable functions $m$ and $L$, such that \begin{gather} \| F(t,x)-F(t,y)\| \leq L(t)\| x-y\| \label{5} \\ \| F(t,x)\| \leq m(t) \label{6} \end{gather} for all $x$ and $y$ in some neighbourhood of $x_{0}$, and $t$ a.e. in some neighbourhood of $t_{0}$. \end{definition} We need the following definition. \begin{definition}\label{def1}\rm For $\alpha \in \mathbb{R}$ let $(x,y)$ be a solution of (\ref{eq1-1}), (\ref{eq1-2}), (\ref{eq1-3}) and (\ref{eq1-4}), and let $\eta \in (0,1)$ and $\beta >1$ be as in Theorem \ref{Theorem1}. We define a function $H:(0,\infty )\to \mathbb{R}$ by $H(\alpha )=\beta x(1)-x(\eta )$, for $\alpha \in (0,\infty )$. \end{definition} \begin{remark} \rm Since the function $g:\mathbb{R}\to \mathbb{R}$ in Theorem \ref{Theorem1} is assumed to be super-linear, we see that there exists an $M>0$ such that $\frac{g(x)}{x}\geq 1$ for $|x|\geq M$. Let us, now, define a function $\widetilde{g}:\mathbb{R}\to \mathbb{R}$ by \[ \widetilde{g}(x)=\begin{cases} g(x), & \text{ for }x\geq M \\ \frac{g(M)}{M}x\text, & \text{for }0\leq x\leq M \\ \frac{g(-M)}{-M}x, & \text{for }-M\leq x\leq 0 \\ g(x), & \text{for }x\leq -M. \end{cases} \] It then follows that $\frac{\widetilde{g}(x)}{x}\geq 1$ for all $x\neq 0$ and $g-\widetilde{g}$ is a bounded function on $\mathbb{R}$. Also, $p+ \widetilde{g}-g$ satisfies the same conditions as $p$ in Theorem \ref {Theorem1}. Accordingly, we shall assume in the following that the function $g$ in Theorem \ref{Theorem1} is such that $\frac{g(x)}{x}\geq 1$ for all $x\neq 0$, by replacing $g$ by $\widetilde{g}$ and $p$ by $p+\widetilde{g}-g$, if necessary. \end{remark} \begin{proof}[Proof of Theorem: Smooth case] Here we assume, in addition to Assumption A, that $p$ is Caratheodory-Lipschitz on $[0,1]\times \mathbb{R}^{2}$, $g$ is locally Lipschitz, and for all nonzero $x$, $g(x)/x \geq 1$. We first see from Lemma \ref{Lemma3} and the last claim in Lemma \ref{Lemma2}, that $\varphi _{1}(x,y)\to \infty $, as $y(0)\to \infty $. Accordingly, for every positive integer $k$, sufficiently large, there exists an $h_{k}\in \mathbb{R}$ such that if $(x,y)$ is a solution of \begin{equation} \begin{aligned} x' = y \\ y' = -g(x)+p(t,x,y) \\ x(0) = 0 \\ y(0) = h_{k}, \end{aligned}\label{h_k} \end{equation} then $\varphi _{1}(x,y)=\pi /2+k\pi $. There may be more than one value for $h_{k}$, so we let $h_{k}^{min}$ and $h_{k}^{max}$ be the smallest and largest such numbers. Then for the function $H$, defined in Definition \ref{def1}, we claim $H(h_{k})>0$, if $k$ is even. Since, now, $\varphi _{1}(x,y)=\pi /2+k\pi $ and $k$ is even, we see that $x(1)>0$ and $y(1)=0$, from the definition of $\varphi _{1}(x,y)$ (see Definition \ref{def}). Suppose, now, $x$ is maximised at $\eta ^{\ast }\in (0,1]$. We then get from (\ref{eq7}) of Lemma \ref{Lemma2} that $x(\eta )\leq x(\eta ^{\ast })\leq G_{+}^{-1}(\frac{ h_{k}^{2}}{2}(1+\epsilon ))$ and $\beta x(1)\geq \beta G_{+}^{-1}(\frac{ h_{k}^{2}}{2(1+\epsilon )})$, since $x(1)>0$ and $y(1)=0$. Now, \begin{equation} \begin{aligned} H(h_{k}) &= \beta x(1)-x(\eta ) \\ &\geq \beta G_{+}^{-1}(\frac{h_{k}^{2}}{2(1+\epsilon )})-G_{+}^{-1}(\frac{ h_{k}^{2}}{2}(1+\epsilon )) \\ &= (\beta -1)G_{+}^{-1}(\frac{h_{k}^{2}}{2(1+\epsilon )})+G_{+}^{-1}(\frac{ h_{k}^{2}}{2(1+\epsilon )})-G_{+}^{-1}(\frac{h_{k}^{2}}{2}(1+\epsilon )). \end{aligned}\label{eq16x} \end{equation} We may assume that $0<\epsilon <1$, and let us set \begin{equation} t=G_{+}^{-1}(\frac{h_{k}^{2}}{2(1+\epsilon )}) \label{eq16a} \end{equation} and \[ t+\delta =G_{+}^{-1}(\frac{h_{k}^{2}}{2}(1+\epsilon )). \] Then \[ G(t)=\frac{h_{k}^{2}}{2(1+\epsilon )} \quad\mbox{and}\quad G(t+\delta )=\frac{h_{k}^{2}}{2}(1+\epsilon ). \] Next, we see that \begin{equation} h_{k}^{2}\epsilon \geq \frac{h_{k}^{2}}{2}(1+\epsilon )-\frac{h_{k}^{2}}{2(1+\epsilon )} = G(t+\delta )-G(t) = \int_{t}^{t+\delta }g(s)ds \geq t\delta , \label{eq16b} \end{equation} in view of our assumption $\frac{g(x)}{x}\geq 1$ for all $x\neq 0$. It then follows from (\ref{eq16x}), (\ref{eq16a}), (\ref{eq16b}), the assumption $0<\epsilon <1$ and the fact that $G_{+}^{-1}$ is an increasing function that \[ H(h_{k}) = \beta x(1)-x(\eta ) \geq (\beta -1)G_{+}^{-1}(\frac{h_{k}^{2}}{4}) -\frac{h_{k}^{2}\epsilon}{G_{+}^{-1}(\frac{h_{k}^{2}}{4})} >0 \] if $\epsilon >0$ is chosen sufficiently small. Hence $H(h_{k})>0$. Similarly, $H(h_{k})<0$ when $k$ is odd. Now $H$ is continuous, indeed the map $(0,\alpha )\mapsto (x(1),x(\eta ))$ is locally Lipschitz by \cite{kur}. By the intermediate value theorem, there is an $\alpha \in (h_{2k}^{max},h_{2k+1}^{min})$ and a solution $(x,y)$ of (\ref{eq1-1}), (\ref {eq1-2}), (\ref{eq1-3}), (\ref{eq1-4}) such that $H(\alpha )=0$. We claim $\varphi _{1}(x,y)\in (\pi /2+2k\pi ,\pi /2+(2k+1)\pi )$. Suppose $\varphi _{1}(x,y)\leq \pi /2+2k\pi $. Then by the intermediate value theorem there is an $h_{2k}>\alpha $, contradicting $\alpha >h_{2k}$. This concludes the proof of Theorem \ref{Theorem1} in the smooth case. \end{proof} The next Lemma is needed in the proof for the non-smooth case. Writing $(x,y)={\bf u}$, we mollify the function $p(t,{\bf u})$ with respect to the second variable ${\bf u}$. \begin{lemma} Suppose $p:[0,1]\times \mathbb{R}^{2}\to \mathbb{R}$ satisfies (a) the Caratheodory-Lipschitz conditions, and $M_{1}:[0,1]\times [ 0,\infty )\to [ 0,\infty )$ satisfies: \begin{itemize} \item[(b)] for all $t\in [ 0,1]$, $M_{1}(t,\cdot )$ is increasing on $[0,\infty )$, \item[(c)] for all $s\in [ 0,\infty )$, $M_{1}(\cdot ,s)$ is integrable on $[0,1]$, and \item[(d)] $s^{-1}\int_{0}^{1}M_{1}(t,s)dt\to 0$ as $s\to \infty $. \end{itemize} Suppose that for all $t$ and ${\bf u}$, \begin{itemize} \item[(e)] $|p(t,{\bf u})|\leq M_{1}(t,\| {\bf u}\| _{\infty })$. \end{itemize} Let $\varphi \in C_{0}^{\infty }(\mathbb{R}^{2})$ have support in $\{{\bf u}\in \mathbb{R}^{2}:\| {\bf u}\| _{\infty }\leq 1\}$, $\varphi \geq 0$, $\int \varphi =1$. Let $\epsilon >0$ be given. Let $p^{\epsilon }(t,{\bf u})=\int p(t,{\bf u}-\epsilon {\bf v})\varphi ({\bf v})d{\bf v}$, and let $M_{1}^{\epsilon }(t,s)=M_{1}(t,s+\epsilon )$. Then the pair of functions $p^{\epsilon }$ and ${M_{1}}^{\epsilon }$ satisfy conditions (a) through (e). \end{lemma} \begin{proof} To show $p^{\epsilon }$ satisfies (a), we first let ${\bf u}$ be given, and claim $t\mapsto p^{\epsilon }(t,{\bf u})$ is Lebesgue measurable. That is, \[ t\mapsto \epsilon ^{-2}\int p(t,{\bf x})\varphi ({\frac{{\bf u}-{\bf x}}{ \epsilon }})dx_{1}dx_{2} \] is measurable. For a.e. $t\in [ 0,1]$ , $p(t,x)$ is continuous in $x$, and so the integral is a Riemann integral. Accordingly, \[ \epsilon ^{-2}\int p(t,{\bf x})\varphi ({\frac{{\bf u}-{\bf x}}{\epsilon }} )dx_{1}dx_{2}=\lim_{n\to \infty }\sum_{{\bf x}\in P(n)}p(t,{\bf x} )\varphi ({\frac{{\bf u}-{\bf x}}{\epsilon }}), \] where $\{P(n)\}$ is a sequence of partitions of $[0,1]\times [ 0,1]$. Each of these sums is a measurable function of $t$ since $p$ satisfies the Caratheodory conditions, Lebesgue measure is complete, and the Lebesgue measurable functions form a vector space. The limit of a sequence of measurable functions is measurable, and so we have proved the claim. To show $p^{\epsilon }$ satisfies (a), let $(t_{0},{\bf u}_{0})\in D$ be given. We claim there are real valued integrable functions $m$ and $L$, such that \begin{gather} \| p^{\epsilon }(t,{\bf u})-p^{\epsilon }(t,{\bf w})\| \leq L(t)\| {\bf u}-{\bf w}\| \label{15} \\ \| p^{\epsilon }(t,{\bf u})\| \leq m(t) \label{16} \end{gather} for all ${\bf u}$ and ${\bf w}$ in some neighbourhood of ${\bf u}_{0}$, and for a.e. $t$ in some neighbourhood of $t_{0}$, i.e. (\ref{5}) and (\ref{6}) hold. Now \[ \| p^{\epsilon }(t,{\bf u})-p^{\epsilon }(t,{\bf w})\| \leq \epsilon ^{-2}\int_{N(\epsilon )}p(t,{\bf x})|\varphi ({\frac{{\bf u}-{\bf x}}{ \epsilon }})-\varphi ({\frac{{\bf w}-{\bf x}}{\epsilon }})|dx_{1}dx_{2} \] where $N(\epsilon )$ stands for $\{{\bf x}:\| {\bf x}-{\bf u}\| \leq \epsilon \}\cup \{x:\| {\bf x}-{\bf w}\| \leq \epsilon \}$. Note $N(\epsilon )\subset \{{\bf x}:\| {\bf x}\| \leq \max (\| {\bf u} \| ,\| {\bf v}\| )+\epsilon )$. By (b), with $K(\varphi )$ the Lipschitz constant of $\varphi $, \begin{align*} RHS &\leq \epsilon ^{-3}\int_{N(\epsilon )}M_{1}(t,\max (\| {\bf u}\| ,\| {\bf v}\| )+\epsilon )K(\varphi )\| {\bf u}-{\bf w}\| dx_{1}dx_{2} \\ &\leq 2\epsilon ^{-1}M_{1}(t,\max (\| {\bf u}\| ,\| {\bf v}\| )+\epsilon )K(\varphi )\| {\bf u}-{\bf w}\| . \end{align*} Since $M_{1}$ is increasing, (b) shows that (\ref{15}) holds for all ${\bf u} $ and ${\bf w}$ in any given bounded set, and all $t\in [ 0,1]$. For (\ref{16}), we check $\| p^{\epsilon }(t,0)\| $ is integrable, and this and (\ref{15}) gives (\ref{16}). \begin{align*} |p^{\epsilon }(t,0)| &= \| \int p(t,-\epsilon {\bf v})\varphi ({\bf v})d {\bf v}\| \\ &\leq \int |p(t,-\epsilon {\bf v})\varphi ({\bf v})|d{\bf v} \\ &\leq \int_{\| v\| \leq 1}M_{1}(t,\epsilon )\varphi ({\bf v})dv, \quad \text{by (e),} \\ &= M_{1}(t,\epsilon ). \end{align*} To show (b) for ${M_{1}}^{\epsilon }$ we note that $s\mapsto M_{1}(t,s+\epsilon )$ is increasing on $[0,\infty )$. To show (c) for ${M_{1}}^{\epsilon }$ we note that for all $s$, $t\mapsto M_{1}(t,s+\epsilon )$ is integrable on $[0,1]$. To show (d) for ${M_{1}}^{\epsilon }$ we note that $s^{-1} \int_{0}^{1}M_{1}(t,s+\epsilon )dt\to 0$ as $s\to \infty $. To show (e) for $p^{\epsilon }(t,0)$ and ${M_{1}}^{\epsilon }$ we note that \begin{align*} |p^{\epsilon }(t,{\bf u})| &\leq \int |p(t,{\bf u}-\epsilon {\bf v})\varphi ({\bf v})d{\bf v} \\ &\leq \int M_{1}(t,\| {\bf u}\| +\epsilon )\varphi ({\bf v})d{\bf v} \\ &\leq \int M_{1}^{\epsilon }(t,\| {\bf u}\| )\varphi ({\bf v})d{\bf v} \\ &= M_{1}^{\epsilon }(t,\| {\bf u}\| ). \end{align*} \end{proof} \begin{proof}[Proof of Theorem: Non-smooth case] Given $g$, we will, by adding a term to $g$ and subtracting it from $p$, assume that $g(x)/x\geq 1$ for all $x\neq 0$. For $\epsilon >0$, we take $g^{\epsilon }$ which is locally Lipschitz and such that $g^{\epsilon }(x)\to g(x)$ uniformly on bounded sets, and $g^{\epsilon }(x)/x\geq 1$ for all $x\neq 0$. For each large integer $k$, and $\epsilon >0$, we let $(x_{\epsilon },y_{\epsilon })$ be a solution of (\ref{2.1}) to (\ref{2.4}) with $g^{\epsilon }$ and $p^{\epsilon }$ replacing $g$ and $p$, satisfying $\varphi _{1}(x_{\epsilon },y_{\epsilon })\in (\frac{\pi }{2}+k\pi ,\frac{\pi }{2}+(k+1)\pi )$, with $y_{\epsilon }(0)>0$. Now we can check the $(x_{\epsilon },y_{\epsilon })$ are uniformly bounded, and we can check they are equi-continuous, since their derivatives are uniformly bounded. By the Arzela-Ascoli Theorem, there is a sequence $\epsilon (n)\to 0$ with $(x_{\epsilon (n)},y_{\epsilon (n)})$ converging to $(x,y)$, say, in $C([0,1]; \mathbb{R}^{2})$. Now for all $t\in [ 0,1]$, \begin{equation} \begin{pmatrix} x_{\epsilon } \\ y_{\epsilon } \end{pmatrix}(t) =\begin{pmatrix} 0 \\ y_{\epsilon }(0) \end{pmatrix} +\int_{0}^{t}\begin{pmatrix} y_{\epsilon }(s) \\ -g^{\epsilon }(x_{\epsilon }(s))+p^{\epsilon }(s,x_{\epsilon }(s),y_{\epsilon }(s)) \end{pmatrix} ds \label{17} \end{equation} We use the dominated convergence theorem to let $n\to \infty $ with $\epsilon =\epsilon (n)$.\\ (a) We claim $p^{\epsilon }(s,x_{\epsilon}(s),y_{\epsilon }(s))$ converges to $p(s,x(s),y(s))$ for $s$ a.e. in $[0,1]$. Take $s$ so that $p(s,{\bf u})$ is continuous in ${\bf u}$. Then we note \[ \int_{\mathbb{R}^{2}}p(s,x_{\epsilon (n)}(s)-{\epsilon }(n)v_{1},y_{\epsilon (n)}(s)-{\epsilon }(n)v_{2})\varphi (v_{1},v_{2})dv_{1}dv_{2}\to p(s,x(s),y(s)), \] proving the claim. \\ (b) We note $g^{\epsilon (n)}(s,x_{\epsilon (n)}(s))$ converges to $g(x(s))$ for any $s$. \\ (c) We claim $p^{\epsilon (n)}(s,x_{\epsilon (n)}(s),y_{\epsilon (n)}(s))\leq M_{1}(s,K)$ for some $K>0$. Just take $K\geq \sup_{s\in [ 0,1]}\sup_{n}\max (|x_{\epsilon (n)}(s)|,|y_{\epsilon (n)}(s)|)+\max_{n}\epsilon (n)$. \\ (d) We note there is $K$ such that for all $n$ and $x$, $|g^{\epsilon (n)}(x_{\epsilon (n)}(s))|\leq K$. \\ The dominated convergence theorem is applicable by (a) -- (d). Hence \begin{equation} \begin{pmatrix} x \\ y \end{pmatrix}(t)=\begin{pmatrix} 0 \\ y(0) \end{pmatrix} +\int_{0}^{t}\begin{pmatrix} y(s) \\ -g(x(s))+p(s,x(s),y(s)) \end{pmatrix} ds \label{18} \end{equation} Hence the o.d.e. (\ref{2.1}) and (\ref{2.2}) holds for $(x,y)$. The boundary conditions (\ref{2.3}) and (\ref{2.4}) hold for $(x,y)$, since they held for the approximations $(x_{\epsilon },y_{\epsilon })$. We note that $\varphi _{1}(x_{\epsilon (n)},y_{\epsilon (n)})\to \varphi _{1}(x,y)$, noting that for all $n$, $(x_{\epsilon (n)},y_{\epsilon (n)})$ are outside some neighbourhood of $(0,0)$. Hence $\varphi _{1}(x,y)\in [ \frac{\pi }{2}+k\pi ,\frac{\pi }{2}+(k+1)\pi ]$, since $\varphi_{1}(x_{\epsilon (n)},y_{\epsilon (n)})\in (\frac{\pi }{2}+k\pi ,\frac{\pi}{2}+(k+1)\pi )$. Because of the boundary condition (\ref{2.4}), $\varphi_{1}(x,y)\neq \frac{\pi }{2}+k\pi $ for all large $k$. This ends the proof of the theorem. \end{proof} \section{Uniqueness} We proved in Theorem \ref{Theorem1} that the equations (\ref{2.1}), (\ref{2.2}), (\ref{2.3}), (\ref{2.4}) have at least one solution $(x,x')$ with $\varphi_{1}(x,x')\in (\frac{\pi }{2}+k\pi ,\frac{\pi }{2}+(k+1)\pi )$, and $x'(0)>0$. In this section we shall show that the equations (\ref{2.1}), (\ref{2.2}), (\ref{2.3}), (\ref{2.4}) have exactly one solution $(x,x')$ with $\varphi _{1}(x,x')\in (\frac{\pi }{2}+k\pi ,\frac{\pi }{2} +(k+1)\pi )$ and $x'(0)>0$, when $p\equiv 0$, $g$ is like the function $x\mapsto |x|^{s}sgn(x)$, for some $s>1$, and $\beta $, $\eta $ satisfy a suitable inequality. The arguments can be easily modified to prove that the equations (\ref{2.1}), (\ref{2.2}), (\ref{2.3}), (\ref{2.4}) have exactly one solution $(x,x')$ with $\varphi _{1}(x,x')\in (\frac{\pi }{2} +k\pi ,\frac{\pi }{2}+(k+1)\pi )$, and $x'(0)<0$. In Remark \ref{rem14} we give a result for $\beta < 1$. \begin{theorem}\label{Theorem2} Let $g:\mathbb{R}\to \mathbb{R}$ be a continuously differentiable function. Suppose that there exist $p_{0}>0$, $p_{1}>0$, and an $s>1$ such that for all $x\in \mathbb{R}$, \begin{equation} p_{0}|x|^{s}\leq g(x)\mathop{\rm sgn}(x)\leq p_{1}|x|^{s}, \label{3.01} \end{equation} and there exists a $h>0$ such that \begin{equation} \frac{g(x)}{x^{1+h}}\text{ is increasing on }(0,\infty ) \text{ and } (-\infty , 0). \label{3.02} \end{equation} Let $\beta >1$ and $\eta \in (0,1)$ be such that \begin{equation} \beta ^{2}>\sqrt{1+\frac{\eta ^{4}}{4}}+\frac{\eta ^{2}}{2}. \label{3.1} \end{equation} Then, for $k$ (an integer) sufficiently large, the solution of the system of equations \begin{gather} x'(t)=y(t), \label{3-03} \\ y'(t)=-g(x(t)),\label{3-04} \\ x(0) = 0, \label{3-05} \\ x(\eta ) = \beta x(1), \label{3-06} \end{gather} with $\varphi _{1}(x,x')\in (\frac{\pi }{2}+k\pi ,\frac{\pi }{2} +(k+1)\pi )$, and $x'(0)>0$, is unique. \end{theorem} \begin{remark} {\rm The existence of a solution for the system of equations (\ref{3-03}), (\ref {3-04}), (\ref{3-05}), (\ref{3-06}) is obtained using Theorem \ref{Theorem1}.} \end{remark} \begin{proof}[Proof of Theorem \ref{Theorem2}] Let $(x_{\alpha }(t),y_{\alpha }(t))$ be the solution of the equations (\ref{3-03}), (\ref{3-04}) with $x(0)=0$, $y(0)=\alpha $. Recalling $G(x)=\int_{0}^{x}g(t)dt$, we see, from (\ref{3-03}), (\ref{3-04}) with $(x,y)=(x_{\alpha }(t),y_{\alpha }(t))$, that \begin{equation} G(x)+\frac{y^{2}}{2}=\frac{\alpha ^{2}}{2}=G(\gamma )=G(-\gamma ^{\ast }), \nonumber \label{3.3} \end{equation} where $\gamma =G_+^{-1}(\alpha^2/2)$ and $-\gamma ^{\ast }=G_-^{-1}(\alpha^2/2)$. Note that both $\gamma $ and $\gamma ^{\ast }$ are positive. In the following, we shall use $\gamma $ to parametrise the solution, giving $(x,y)=(x_{\alpha }(t),y_{\alpha }(t)):=(x(t,\gamma ),y(t,\gamma ))$. We define $\gamma _{k}$ by setting $\gamma =\gamma _{k}$ when $\varphi _{1}(x,y)=\frac{\pi }{2}+k\pi $. This corresponds to $\alpha =h_{k}$ (see equation (\ref{h_k})). We next consider $\beta x(1,\gamma )$ and $x(\eta ,\gamma )$ for $\gamma \in (\gamma _{k},\gamma _{k+1})$. Now, from Theorem \ref{Theorem1} we see for $k$ sufficiently large that there exists a $\gamma _{0}\in (\gamma _{k},\gamma _{k+1})$ such that $\beta x(1,\gamma _{0})=x(\eta ,\gamma _{0})$. To show uniqueness of $\gamma _{0}$ it suffices to show that \begin{equation} |\beta \frac{\partial x}{\partial \gamma }(1,\gamma _{0})|>| \frac{\partial x}{\partial \gamma }(\eta ,\gamma _{0})|. \label{3.2} \end{equation} Let us define $\widetilde{\varphi }(t,\gamma )$ by setting \begin{equation} \widetilde{\varphi }(t,\gamma ) =\int_{0}^{t}\frac{x'(s)y(s)-y'(s)x(s)}{x^{2}(s)+y^{2}(s)}ds, \nonumber \label{3.4} \end{equation} where $(x,y)=(x(t,\gamma ), y(t,\gamma ))$. Now, we define a function $\widetilde{t}(\varphi ,\gamma )$ by \begin{equation} t=\widetilde{t}(\varphi ,\gamma )\text{ }\Longleftrightarrow \varphi = \widetilde{\varphi }(t,\gamma ). \label{3.5} \end{equation} We note that $\widetilde{t}(\varphi ,\gamma )$ is the time taken for the solution $(x(t,\gamma ),y(t,\gamma ))$ to traverse the angle $\varphi $. For $t=1$, we then have \[ 1=\widetilde{t}(\widetilde{\varphi }(1,\gamma ),\gamma ), \] from (\ref{3.5}). Next we use the implicit function theorem to get \begin{equation} \frac{\partial }{\partial \gamma }\widetilde{\varphi }(1,\gamma ) = -\frac{\frac{\partial \widetilde{t}}{\partial \gamma }(1,\gamma )}{\frac{\partial \widetilde{t}}{\partial \varphi }(1,\gamma )} = -\frac{\partial \widetilde{t}}{\partial \gamma }(\widetilde{\varphi } (1,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial t} (1,\gamma ). \label{3.7} \end{equation} Let us define $\widetilde{x}(\varphi ,\gamma )$ as follows: since traversing the angle $\varphi $ clockwise along the curve $G(x)+\frac{y^{2}}{2} =G(\gamma )$ from $(0,\alpha )$ brings us to a point $(x,y)$, we define \begin{equation} \widetilde{x}(\varphi ,\gamma )=x. \nonumber \label{3.8} \end{equation} Note \[ \widetilde{x}(\widetilde{\varphi }(1,\gamma ),\gamma )=x(1,\gamma ). \] >From the chain rule, \begin{equation} \frac{\partial x}{\partial \gamma }(1,\gamma )=\frac{\partial \widetilde{x}}{ \partial \gamma }(\widetilde{\varphi }(1,\gamma ),\gamma )+\frac{\partial \widetilde{x}}{\partial \varphi }(\widetilde{\varphi }(1,\gamma ),\gamma ) \frac{\partial \widetilde{\varphi }}{\partial \gamma }(1,\gamma ). \label{3.9} \end{equation} Similarly, we get \begin{equation} \frac{\partial x}{\partial \gamma }(\eta ,\gamma )=\frac{\partial \widetilde{ x}}{\partial \gamma }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )+\frac{ \partial \widetilde{x}}{\partial \varphi }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial \gamma }(\eta ,\gamma ). \label{3.91} \end{equation} We express $\widetilde{t}(\varphi ,\gamma )$ in terms of $n$, the number of times the solution $(x,y)$ goes around the origin, and the time taken to traverse the angle $2\pi $. Let \begin{equation} \widetilde{t}(\varphi ,\gamma )=n(\varphi ,\gamma )t_{P}(\gamma )+t_{Q}(\varphi ,\gamma ), \label{3.10} \end{equation} where $t_{P}$ is the time for a full revolution or the time taken to traverse the angle $2\pi $, and $t_{Q}\in [ 0,t_{P})$. Note \begin{equation} \frac{n(\widetilde{\varphi }(\eta ,\gamma ),\gamma )}{n(\widetilde{\varphi }(1,\gamma ),\gamma )}\to \eta \text{ as }\gamma \to \infty . \nonumber% \label{3.11} \end{equation} We differentiate $\widetilde{t}(\varphi ,\gamma )$ with respect to $\gamma $ in (\ref{3.10}) when $\widetilde{\varphi }(1,\gamma )$ is not a multiple of ${2\pi }$ to get \begin{equation} \frac{\partial \widetilde{t}}{\partial \gamma }(\varphi ,\gamma )=n\frac{ \partial t_{P}}{\partial \gamma }+\frac{\partial t_{Q}}{\partial \gamma } . \label{3.110} \end{equation} Next, define \begin{equation} \varphi _{\eta }:=\widetilde{\varphi }(\eta ,\gamma _{0})\text{ and }\varphi _{1}:=\widetilde{\varphi }(1,\gamma _{0}). \nonumber \label{3.111} \end{equation} We evaluate (\ref{3.110}) at two points. \begin{gather} \frac{\partial \widetilde{t}}{\partial \gamma }(\varphi _{1},\gamma _{0})=n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma } (\varphi _{1},\gamma _{0})+\frac{\partial t_{Q}}{\partial \gamma }(\varphi _{1},\gamma _{0}). \label{3.12} \\ \frac{\partial \widetilde{t}}{\partial \gamma }(\varphi _{\eta },\gamma _{0})=n(\varphi _{\eta },\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\varphi _{\eta },\gamma _{0})+\frac{\partial t_{Q}}{\partial \gamma } (\varphi _{\eta },\gamma _{0}). \label{3.121} \end{gather} Now, by \cite[Lemma 4]{cal} we see that \begin{equation} \frac{\partial }{\partial \gamma }\frac{\partial \widetilde{t}}{\partial \varphi }(\varphi ,\gamma )\leq 0\text{, for all }(\varphi ,\gamma ). \nonumber \label{3.122} \end{equation} Also, notice that $\frac{\partial \widetilde{t}}{\partial \gamma }(0,\gamma )=0$ for all $\gamma$. So we get \[ 0\geq \frac{\partial t_{Q}}{\partial \gamma }(\varphi ,\gamma _{0})\geq \frac{\partial t_{P}}{\partial \gamma }(\gamma _{0}). \] Now, to prove (\ref{3.2}) we compute the ratio of (\ref{3.91}) to (\ref{3.9}). This gives \[ \frac{\frac{\partial x}{\partial \gamma }(\eta ,\gamma )}{\frac{\partial x}{ \partial \gamma }(1,\gamma )}=\frac{\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )+\frac{\partial \widetilde{x}}{\partial \varphi }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial \gamma }(\eta ,\gamma )}{ \frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{\varphi } (1,\gamma ),\gamma )+\frac{\partial \widetilde{x}}{\partial \varphi }( \widetilde{\varphi }(1,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial \gamma }(1,\gamma )}. \] Using (\ref{3.7}) we then get \begin{equation} \frac{\frac{\partial x}{\partial \gamma }(\eta ,\gamma )}{\frac{\partial x}{ \partial \gamma }(1,\gamma )}=\frac{\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )-\frac{\partial \widetilde{x}}{\partial \varphi }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )\frac{\partial \widetilde{t}}{\partial \gamma }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial t}(\eta ,\gamma )}{\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{ \varphi }(1,\gamma ),\gamma )-\frac{\partial \widetilde{x}}{\partial \varphi }(\widetilde{\varphi }(1,\gamma ),\gamma )\frac{\partial \widetilde{t}}{ \partial \gamma }(\widetilde{\varphi }(1,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial t}(1,\gamma )}. \label{3.130} \end{equation} We note that \begin{equation} \frac{\partial x}{\partial t}(t,\gamma )=\frac{\partial \widetilde{x}}{ \partial \varphi }(\widetilde{\varphi }(t,\gamma ),\gamma )\frac{\partial \widetilde{\varphi }}{\partial t}(t,\gamma )\text{ for all }(t,\gamma ). \nonumber \label{3.13} \end{equation} Hence, (\ref{3.130}) becomes \[ \frac{\frac{\partial x}{\partial \gamma }(\eta ,\gamma )}{\frac{\partial x}{ \partial \gamma }(1,\gamma )}=\frac{\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{\varphi }(\eta ,\gamma ),\gamma )-\frac{\partial \widetilde{t}}{\partial \gamma }(\widetilde{\varphi }(\eta ,\gamma ),\gamma ) \frac{\partial x}{\partial t}(\eta ,\gamma )}{\frac{\partial \widetilde{x}}{ \partial \gamma }(\widetilde{\varphi }(1,\gamma ),\gamma )-\frac{\partial \widetilde{t}}{\partial \gamma }(\widetilde{\varphi }(1,\gamma ),\gamma ) \frac{\partial x}{\partial t}(1,\gamma )}. \] Now, by (\ref{3.12}) and (\ref{3.121}) we get \begin{equation} \frac{\frac{\partial x}{\partial \gamma }(\eta ,\gamma _{0})}{\frac{\partial x}{\partial \gamma }(1,\gamma _{0})}=\frac{NUM}{DENOM}, \label{3.14} \end{equation} where \begin{equation} NUM=\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{\varphi } (\eta ,\gamma _{0}),\gamma _{0})-\frac{\partial x}{\partial t}(\eta ,\gamma _{0})[n(\varphi _{\eta },\gamma _{0})\frac{\partial t_{P}}{\partial \gamma } (\gamma _{0})+\frac{\partial t_{Q}}{\partial \gamma }(\varphi _{\eta },\gamma _{0})], \label{3.15} \end{equation} and \begin{equation} DENOM=\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{\varphi } (1,\gamma _{0}),\gamma _{0})-\frac{\partial x}{\partial t}(1,\gamma _{0})[n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma } (\gamma _{0})+\frac{\partial t_{Q}}{\partial \gamma }(\varphi _{1},\gamma _{0})]. \label{3.16} \end{equation} Next, we need to show that the first term in (\ref{3.16}) is small compared to the second term. To do this we need the following lemma. \end{proof} \begin{lemma}\label{Lemma5} Suppose that the assumptions (\ref{3.01}) and (\ref{3.02}) of Theorem \ref{Theorem2} hold. Then \begin{equation} \frac{|\frac{\partial \widetilde{x}}{\partial \gamma }(\varphi _{1},\gamma _{0})|}{n(\varphi _{1},\gamma _{0})|\frac{\partial t_{P}}{\partial \gamma } (\gamma _{0})|\cdot |\frac{\partial x}{\partial t}(1,\gamma _{0})|} \to 0\text, \label{4.0} \end{equation} as $\gamma _{0}\to \infty $, (note that $\gamma _{0}$ depends on $k$ and as $k\to \infty $, $\gamma _{0}\to \infty $). \end{lemma} \begin{proof} For better readability we break the proof into a sequence of items.\\ \textbf{Item 1:-} We show that \begin{gather} \frac{\partial \widetilde{x}}{\partial \gamma }(\varphi ,\gamma )=\frac{ xg(\gamma )}{xg(x)+2(G(\gamma )-G(x))}=\frac{xg(\gamma )}{xg(x)+y^{2}}, \label{4.03} \\ \frac{\partial \widetilde{y}}{\partial \gamma }(\varphi ,\gamma )=\frac{ yg(\gamma )}{xg(x)+2(G(\gamma )-G(x))}=\frac{yg(\gamma )}{xg(x)+y^{2}}. \label{4.04} \end{gather} We first note that as we hold $\varphi $ constant, we hold $y/x$ constant i.e. \[ \frac{\sqrt{2(G(\gamma )-G(x))}}{x} \] is constant, which in turn implies that \[ \frac{\partial }{\partial \gamma }\Big( \frac{G(\gamma )-G(x)}{x^{2}}\Big) =0. \] Hence, \[ (g(\gamma )-g(x)\frac{\partial \widetilde{x}}{\partial \gamma })x^{2}-2x \frac{\partial \widetilde{x}}{\partial \gamma }(G(\gamma )-G(x))=0, \] i.e. \[ \frac{\partial \widetilde{x}}{\partial \gamma }[g(x)x^{2}+2x(G(\gamma )-G(x))]=x^{2}g(\gamma ), \] and so (\ref{4.03}) holds. Also, since we hold $\varphi $ constant, $y/x$ is constant and we get \[ \frac{\partial \widetilde{y}}{\partial \gamma }=\frac{y}{x}\frac{\partial \widetilde{x}}{\partial \gamma }, \] giving (\ref{4.04}). \\ \textbf{Item 2:-} There exists a constant $K_{0}=K_{0}(s,p_{0},p_{1})$ such that \begin{equation} |\frac{\partial \widetilde{x}}{\partial \gamma }|\leq K_{0}(s,p_{0},p_{1}) , \nonumber \label{4.1} \end{equation} for all $(\varphi ,$ $\gamma )$. We first claim that for all $x\in \mathbb{R}$, \begin{equation} xg(x)-2G(x)\geq 0. \nonumber \label{4.05} \end{equation} Indeed, both for $x>0$ and for $x<0$, \[ G(x)=\int_{0}^{x}g(t)dt\leq \int_{0}^{x}\frac{tg(x)}{x}dt=\frac{xg(x)}{2}, \] since $\frac{g(x)}{x\text{ }}$ is increasing. Hence the claim. Next, we claim that for all $x$, \begin{equation} \frac{p_{1}|x|^{s+1}}{s+1}\geq G(x)\geq \frac{p_{0}|x|^{s+1}}{s+1}. \label{4.06} \end{equation} We see from (\ref{3.01}) that for $x>0$, \[ G(x)\geq \int_{0}^{x}p_{0}t^{s}dt=\frac{p_{0}x^{s+1}}{s+1}. \] Similarly, we get for $x<0$, \[ G(x)\geq \frac{p_{0}|x|^{s+1}}{s+1}. \] The proof of the left half of the inequality in (\ref{4.06}) is similar. >From (\ref{4.03}) we have \begin{equation} \begin{aligned} \frac{\partial \widetilde{x}}{\partial \gamma }(\varphi ,\gamma ) &= \frac{xg(\gamma )}{xg(x)+2(G(\gamma )-G(x))} \\ &\leq \frac{|x|g(\gamma )}{2G(\gamma )}\text{, by (\ref{4.05}),} \\ &\leq \max (\gamma ,\gamma ^{\ast })\frac{p_{1}\gamma ^{s}(s+1)}{ 2p_{0}\gamma ^{s+1}}. \end{aligned} \label{4.061} \end{equation} Next, we see from (\ref{4.06}) that there exists a constant $K$, depending on $p_{0}$, $p_{1}$, and $s$ such that \begin{equation} \gamma ^{\ast }\leq K\gamma \quad\text{and}\quad \gamma \leq K\gamma ^{\ast }. \label{4.07} \end{equation} Using (\ref{4.07}) in (\ref{4.061}) the proof of Item 2 is immediate. \\ \textbf{Item 3:-} There exists a constant $K_{1}=K_{1}(s,p_{0},p_{1})>0$ such that \begin{equation} |\frac{\partial x}{\partial t}(1,\gamma _{0})|\geq K_{1}\gamma _{0}^{\frac{ s+1}{2}}, \label{4.4} \end{equation} for all $\gamma _{0}$. Since $g(x)/x$ is increasing, we have $G(x)/x$ is increasing by \cite[Lemma 3]{cal}. So \[ G(\frac{\gamma }{\beta })\leq \frac{G(\gamma )}{\beta }. \] Hence, assuming $x(1,\gamma _{0})\geq 0$, so that $x(1,\gamma _{0})\leq \frac{\gamma }{\beta }$, we have \begin{align*} \frac{1}{2}(\frac{\partial x}{\partial t}(1,\gamma _{0}))^{2} &= G(\gamma _{0})-G(x(1,\gamma _{0})) \\ &\geq G(\gamma _{0})-G(\frac{\gamma _{0}}{\beta }) \\ &\geq (1-\frac{1}{\beta })G(\gamma _{0}) \\ &\geq (1-\frac{1}{\beta })\frac{p_{0}\gamma ^{s+1}}{s+1}, \end{align*} by (\ref{4.06}). Taking square roots, we get (\ref{4.4}), proving Item 3. \\ \textbf{Item 4:-} Let $t_{R}$ be the time taken to go from $(0,\alpha )$ to $(\gamma ,0)$. Also let $t_{L}$ be the time taken to go from $(0,-\alpha)$ to $(-\gamma ^{\ast },0)$. We shall show that \begin{equation} \frac{dt_{R}}{d\gamma }=\int_{0}^{\frac{\pi }{2}}\frac{xg(\gamma )(g(x)-xg'(x))}{(xg(x)+y^{2})^{3}}(x^{2}+y^{2})d\varphi . \label{4.5} \end{equation} We note that since \begin{gather*} x' = y, \\ y' = -g(x), \end{gather*} we have \[ \frac{\partial \widetilde{\varphi }}{\partial t}(t,\gamma )=-\frac{ xy'-yx'}{x^{2}+y^{2}}=\frac{xg(x)+y^{2}}{x^{2}+y^{2}}. \] Accordingly, \[ t_{R}(\gamma )=\int_{0}^{\frac{\pi }{2}}\frac{x^{2}+y^{2}}{xg(x)+y^{2}} d\varphi . \] Using Leibnitz's rule, (\ref{4.03}) and (\ref{4.04}), we get \begin{align*} \frac{dt_{R}}{d\gamma } &= \int_{0}^{\frac{\pi }{2}}\frac{1}{ (xg(x)+y^{2})^{2}}\{(2x\frac{xg(\gamma )}{xg(x)+y^{2}} +2y\frac{yg(\gamma)}{xg(x)+y^{2}})(xg(x)+y^{2}) \\ &\quad -(x^{2}+y^{2})[(xg'(x)+g(x))\frac{xg(\gamma )}{xg(x)+y^{2}}+2y \frac{yg(\gamma )}{xg(x)+y^{2}}\}d\varphi \\ &= \int_{0}^{\frac{\pi }{2}}\frac{2x^{2}g(\gamma )+2y^{2}g(\gamma )-\frac{ x^{2}+y^{2}}{xg(x)+y^{2}}[(xg'(x)+g(x))xg(\gamma )+2y^{2}g(\gamma)] }{(xg(x)+y^{2})^{2}}d\varphi \\ &= \int_{0}^{\frac{\pi }{2}}\frac{2g(\gamma )(xg(x)+y^{2}) -(xg'(x)+g(x))xg(\gamma )-2y^{2}g(\gamma )}{(xg(x)+y^{2})^{3}} (x^{2}+y^{2})d\varphi , \end{align*} which gives (\ref{4.5}). \\ \textbf{Item 5:-} We change variables from $\varphi $ to $u$ to calculate\ (\ref{4.5}). We parametrise the curve \[ \frac{y^{2}}{2}+G(x)=G(\gamma ), \] in the first quadrant, by setting \begin{gather*} y = \sqrt{2G(\gamma )}\sin u \\ x = G_{+}^{-1}(G(\gamma )\cos ^{2}u). \end{gather*} So, with $\gamma $ kept fixed, \begin{align*} \frac{dy}{du} &= \sqrt{2G(\gamma )}\cos u \\ g(x)\frac{dx}{du} &= -2G(\gamma )\cos u\sin u, \end{align*} i.e. \[ \frac{dx}{du}=-\frac{2G(\gamma )\cos u\sin u}{g(x)}. \] Now, \begin{equation} (x^{2}+y^{2})\frac{d\varphi }{du}=y\frac{dx}{du}-x\frac{dy}{du}. \nonumber %\label{4.6} \end{equation} Hence, \[ (x^{2}+y^{2})\frac{d\varphi }{du}=-\frac{(2G(\gamma ))^{\frac{3}{2}}\cos ^{2}u\sin u}{g(x)}-x\sqrt{2G(\gamma )}\cos u. \] Therefore, \begin{equation} \frac{dt_{R}}{d\gamma }=\int_{0}^{\frac{\pi }{2}}\frac{xg(\gamma )(xg'(x)-g(x))}{(xg(x)+y^{2})^{3}}\Big[ \frac{(2G(\gamma ))^{\frac{3}{2}}\cos ^{2}u\sin u}{g(x)}+x\sqrt{2G(\gamma )}\cos u\Big] du, \label{4.7} \end{equation} noting that as $\varphi $ varies from $0$ to $\pi/2$, $u$ also varies from $0$ to $\pi/2$. In the next items we estimate the various terms in (\ref{4.7}). \\ \textbf{Item 6:-} There exist constants $K_{3}>0$ and $K_{4}>0$ depending on $p_{0}$, $p_{1}$, and $s$ (but not $u$, $\gamma $) such that \begin{gather} x \leq K_{3}\gamma (\cos u)^{\frac{2}{s+1}}, \label{4.8} \\ x \geq K_{4}\gamma (\cos u)^{\frac{2}{s+1}}. \label{4.9} \end{gather} Indeed, \[ \frac{p_{1}\gamma ^{s+1}}{s+1}\cos ^{2}u\geq G(\gamma )\cos ^{2}u=G(x)\geq \frac{p_{1}x^{s+1}}{s+1}. \] The inequality given by the two outside terms above gives (\ref{4.8}). We obtain (\ref{4.9}) similarly. \\ \textbf{Item 7:-} There exists a constant $K_{5}>0$ depending on $p_{0}$, $p_{1}$, and $s$ (but not $\gamma $) such that \begin{equation} xg(x)+y^{2}\leq K_{5}\gamma ^{s+1}. \label{4.10} \end{equation} Indeed, \begin{gather*} xg(x) \leq p_{1}x^{s+1}\leq \frac{p_{1}}{p_{0}}(s+1)G(x)\leq \frac{p_{1}}{p_{0}}(s+1)G(\gamma )\leq \frac{p_{1}^{2}}{p_{0}}\gamma ^{s+1}, \\ y^{2} = 2G(\gamma )\sin ^{2}u\leq 2p_{1}\frac{\gamma ^{s+1}}{s+1}\sin ^{2}u, \end{gather*} and (\ref{4.10}) follows. \\ \textbf{Item 8:-} There exists a constant $K_{6}>0$ depending on $p_{0}$, $p_{1}$, and $s$ (but not $\gamma $) such that \begin{equation} xg'(x)-g(x)\geq K_{6}x^{s}. \label{4.11} \end{equation} We see from (\ref{3.02}) that $\frac{g(x)}{x^{1+h}}$ is increasing, which implies $\log \frac{g(x)}{x^{1+h}}$ is increasing, which implies $\frac{d}{dx}(\log \frac{g(x)}{x^{1+h}})\geq 0$, which implies $\frac{g'(x) }{g(x)}-\frac{1+h}{x}\geq 0$, which implies $xg'(x)-g(x)\geq hg(x)\geq kp_{0}x^{s}$, thereby proving (\ref{4.11}). \\ \textbf{Item 9:-} We find a lower bound for $\frac{(2G(\gamma))^{\frac{3}{2}}\cos ^{2}u\sin u}{g(x)}+x\sqrt{2G(\gamma )}\cos u$, using (\ref{4.06}), (\ref{4.8}) and (\ref{4.9}). Indeed, \begin{align*} &\frac{(2G(\gamma ))^{\frac{3}{2}}\cos ^{2}u\sin u}{g(x)}+x\sqrt{2G(\gamma )}\cos u \\ &\geq 2\sqrt{2}\frac{p_{0}^{\frac{3}{2}}\gamma ^{\frac{3(s+1)}{2}}}{(s+1)^{ \frac{3}{2}}p_{1}x^{s}}\cos ^{2}u\sin u +K_{4}\gamma (\cos u)^{\frac{2}{s+1}} \sqrt{2\frac{p_{0}}{s+1}}\gamma ^{\frac{s+1}{2}}\cos u \\ &\geq 2\sqrt{2}\frac{p_{0}^{\frac{3}{2}}\gamma ^{\frac{3(s+1)}{2}}}{(s+1)^{ \frac{3}{2}}p_{1}K_{3}^{s}\gamma ^{s}(\cos u)^{\frac{2s}{s+1}}}\cos ^{2}u\sin u+K_{4}\gamma (\cos u)^{\frac{2}{s+1}}\sqrt{2\frac{p_{0}}{s+1}} \gamma ^{\frac{s+1}{2}}\cos u \nonumber \\ &\geq K_{7}\gamma ^{\frac{s+3}{2}}[(\cos u)^{\frac{2}{s+1}}\sin u +(\cos u)^{\frac{s+3}{s+1}}]. \end{align*} %\label{4.12} \textbf{Item 10:-} Now we use our estimates to find a lower bound for (\ref{4.7}). \begin{align*} &\frac{dt_{R}}{d\gamma }\\ &= \int_{0}^{\frac{\pi }{2}}\frac{xg(\gamma )(xg'(x)-g(x))}{(xg(x)+y^{2})^{3}}\Big[ \frac{(2G(\gamma ))^{\frac{ 3}{2}}\cos ^{2}u\sin u}{g(x)}+x\sqrt{2G(\gamma )}\cos u\Big] du \\ &\geq \int_{0}^{\frac{\pi }{2}}\frac{K_{4}\gamma (\cos u)^{\frac{2}{s+1} }g(\gamma )(K_{6}x^{s})}{(K_{5}\gamma ^{s+1})^{3}}K_{7}\gamma ^{\frac{s+3}{2}}[(\cos u)^{\frac{2}{s+1}}\sin u+(\cos u)^{\frac{s+3}{s+1}}]du \\ &\geq \int_{0}^{\frac{\pi }{2}}\frac{K_{4}\gamma (\cos u)^{\frac{2}{s+1} }p_{0}\gamma ^{s}K_{6}(K_{4}\gamma (\cos u)^{\frac{2}{s+1}})^{s}}{(K_{5} \gamma ^{s+1})^{3}}\\ &\quad\times K_{7}\gamma ^{\frac{s+3}{2}}\big[(\cos u)^{\frac{2}{s+1} }\sin u+(\cos u)^{\frac{s+3}{s+1}}\big]du \\ &\geq K_{8}\gamma ^{-\frac{s+1}{2}}. %\label{4.13} \end{align*} We note in a similar way that \[ \frac{dt_{L}}{d\gamma }\geq K_{9}\gamma ^{-\frac{s+1}{2}}\text, \] and hence \[ \frac{dt_{P}}{d\gamma }\geq K_{10}\gamma ^{-\frac{s+1}{2}}. \] Next, to complete the proof of the lemma we use the various estimates obtained above, in (\ref{4.0}). \[ \frac{|\frac{\partial \widetilde{x}}{\partial \gamma }(\varphi _{1},\gamma _{0})|}{n(\varphi _{1},\gamma _{0})|\frac{\partial t_{P}}{\partial \gamma } (\gamma _{0})|\cdot |\frac{\partial x}{\partial t}(1,\gamma _{0})|}\leq \frac{K_{11}}{n(\varphi _{1},\gamma _{0})\gamma _{0}^{\frac{s+1}{2}}\gamma _{0}^{-\frac{s+1}{2}}}=\frac{K_{11}}{n(\varphi _{1},\gamma _{0})} \to 0, \] as $\gamma _{0}\to \infty $. \end{proof} Next, we find an upper bound for $\frac{|y(\eta ,\gamma _{0})|}{|y(1,\gamma _{0})|}$ $\ $, using (\ref{3.1}). \begin{lemma}\label{Lemma6} Suppose $g\ $is continuous, super-linear and $\frac{g(x)}{x}$ is increasing on $(0,\infty )$ and decreasing on $(-\infty ,0)$. Suppose (\ref{3.1}) holds. \ Then there exists a $\beta _{0}<\beta $ such that \begin{equation} \frac{|y(\eta ,\gamma _{0})|}{|y(1,\gamma _{0})|}\leq \frac{\beta _{0}}{ \eta }. \label{5.1} \end{equation} \end{lemma} \begin{proof} We have $x(1,\gamma _{0})=\frac{x(\eta ,\gamma _{0})}{\beta }$ giving \[ -\frac{\gamma _{0}^{\ast }}{\beta }\leq x(1,\gamma _{0})\leq \frac{\gamma _{0}}{\beta }, \] where $\gamma _{0}^{\ast }>0$ is defined by $G(-\gamma _{0}^{\ast })=G(\gamma _{0})$. Suppose $x(\eta ,\gamma _{0})\geq \frac{\gamma _{0}}{\beta }$. Then $x(\eta ,\gamma _{0})\geq x(1,\gamma _{0})$, giving $|y(\eta ,\gamma _{0})|\leq |y(1,\gamma _{0})|$, and (\ref{5.1}) holds for any $\beta _{0}\in [ \eta ,\beta )$. Similarly, if $x(\eta ,\gamma _{0})\leq -\frac{ \gamma _{0}^{\ast }}{\beta }$, then (\ref{5.1}) holds. Hence we assume that \[ -\frac{\gamma _{0}^{\ast }}{\beta }\leq x(\eta ,\gamma _{0})\leq \frac{ \gamma _{0}}{\beta }, \] giving \[ -\frac{\gamma _{0}^{\ast }}{\beta ^{2}}\leq x(1,\gamma _{0})\leq \frac{ \gamma _{0}}{\beta ^{2}}. \] This implies if $x(1,\gamma _{0})\geq 0$, that \[ |y(1,\gamma _{0})|\geq \sqrt{2(G(\gamma _{0})-G(\frac{\gamma _{0}}{\beta ^{2} }))}. \] Similarly, if $x(1,\gamma _{0})\leq 0$, then \[ |y(1,\gamma _{0})|\geq \sqrt{2(G(\gamma _{0})-G(-\frac{\gamma _{0}^{\ast }}{ \beta ^{2}}))}. \] Since $|y(\eta ,\gamma _{0})|\leq \sqrt{2G(\gamma _{0})}$, (\ref{5.1}) holds if \[ \frac{\beta _{0}^{2}}{\eta ^{2}}\geq \frac{G(\gamma _{0})}{G(\gamma _{0})-G( \frac{\gamma _{0}}{\beta ^{2}})}, \] i.e. \begin{equation} (\beta _{0}^{2}-\eta ^{2})G(\gamma _{0})\geq \beta _{0}^{2}G(\frac{\gamma _{0}}{\beta ^{2}}), \label{5.2} \end{equation} and also \begin{equation} (\beta _{0}^{2}-\eta ^{2})G(\gamma _{0})\geq \beta _{0}^{2}G(-\frac{\gamma _{0}^{\ast }}{\beta ^{2}}). \label{5.3} \end{equation} We claim that for all $z\in \mathbb{R}$ and $\delta \geq 1$, \begin{equation} G(\delta z)\geq \delta ^{2}G(z). \label{5.4} \end{equation} Suppose $s>0$. Then $\frac{g(\delta s)}{\delta s}\geq \frac{g(s)}{s}$. For $z>0$, \[ G(\delta z)=\int_{0}^{\delta z}g(t)dt=\int_{0}^{z}g(\delta s)\delta ds\geq \delta ^{2}\int_{0}^{z}g(s)ds, \] proving the claim. Similarly the claim can be proved for $z<0$. In (\ref{5.4}) put $z=\frac{\gamma _{0}}{\beta ^{2}}$, and $\delta =\beta ^{2}$, giving $G(\gamma _{0})\geq \beta ^{4}G(\frac{\gamma _{0}}{\beta ^{2}} )$ and $G(\gamma _{0})\geq \beta ^{4}G(-\frac{\gamma _{0}^{\ast }}{\beta ^{2}})$. Hence, (\ref{5.2}) and (\ref{5.3}) hold if \begin{equation} (1-\frac{\eta ^{2}}{\beta _{0}^{2}})\beta ^{4}\geq 1. \label{5.5} \end{equation} Taking $\beta _{0}=(1-\varepsilon )\beta $, (\ref{5.5}) holds if \begin{equation} \beta ^{2}\geq \frac{(\frac{\eta }{1-\varepsilon })^{2}}{2}+\sqrt{\frac{( \frac{\eta }{1-\varepsilon })^{4}}{4}+1}, \label{5.6} \end{equation} and (\ref{5.6}) holds for some $\varepsilon >0$ by (\ref{3.1}). \end{proof} \begin{proof}[Proof of Theorem \ref{Theorem2}, continued] We divide $NUM$ from (\ref{3.15}) and $DENOM$ from (\ref{3.16}) by $\frac{\partial x}{\partial t}(1,\gamma _{0})n(\varphi _{1},\gamma _{0}) \frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})$ to define \begin{gather*} NEWNUM = \frac{NUM}{\frac{\partial x}{\partial t}(1,\gamma _{0})n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})}\quad \text{and} \\ NEWDENOM = \frac{DENOM}{\frac{\partial x}{\partial t}(1,\gamma _{0})n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma } (\gamma _{0})}. \end{gather*} Now, $NEWDENOM\to -1$ as $\gamma _{0}\to \infty $ since the first summand converges to zero by Lemma \ref{Lemma5} and the second summand is \[ -(1+\frac{\frac{\partial t_{Q}}{\partial \gamma }(\varphi _{1},\gamma _{0})}{ n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})}), \] which converges to $-1$ since \[ \frac{\frac{\partial t_{Q}}{\partial \gamma} (\varphi _{1},\gamma _{0})}{\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})}\leq 1 \] and $n(\varphi _{1},\gamma _{0})\to \infty $. Now, \begin{align*} NEWNUM &= \frac{\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{ \varphi }(\eta ,\gamma _{0}),\gamma _{0})}{\frac{\partial x}{\partial t} (1,\gamma _{0})n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})} \\ &\quad -\frac{\frac{\partial x}{\partial t}(\eta ,\gamma _{0})[n(\varphi _{\eta },\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})+\frac{ \partial t_{Q}}{\partial \gamma }(\varphi _{\eta },\gamma _{0})]}{\frac{ \partial x}{\partial t}(1,\gamma _{0})n(\varphi _{1},\gamma _{0})\frac{ \partial t_{P}}{\partial \gamma }(\gamma _{0})}, \end{align*} and we set \begin{gather*} NEWNUM1 = \frac{\frac{\partial \widetilde{x}}{\partial \gamma }(\widetilde{ \varphi }(\eta ,\gamma _{0}),\gamma _{0})}{\frac{\partial x}{\partial t} (1,\gamma _{0})n(\varphi _{1},\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})}, \\ NEWNUM2 = \frac{\frac{\partial x}{\partial t}(\eta ,\gamma _{0})[n(\varphi _{\eta },\gamma _{0})\frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})+ \frac{\partial t_{Q}}{\partial \gamma }(\varphi _{\eta },\gamma _{0})]}{ \frac{\partial x}{\partial t}(1,\gamma _{0})n(\varphi _{1},\gamma _{0})\frac{ \partial t_{P}}{\partial \gamma }(\gamma _{0})}. \end{gather*} Item 2 of Lemma \ref{Lemma5} shows that $|\frac{\partial \widetilde{x}}{ \partial \gamma }(\widetilde{\varphi }(\eta ,\gamma _{0}),\gamma _{0})|\leq K_{0}$, and hence Lemma \ref{Lemma5} applies to show that $NEWNUM1$ converges to zero as $\gamma _{0}\to \infty $. We rewrite $NEWNUM2$ as \[ \frac{\frac{\partial x}{\partial t}(\eta ,\gamma _{0})}{\frac{\partial x}{ \partial t}(1,\gamma _{0})}(\frac{n(\varphi _{\eta },\gamma _{0})}{n(\varphi _{1},\gamma _{0})}+\frac{\frac{\partial t_{Q}}{\partial \gamma }(\varphi _{\eta },\gamma _{0})}{n(\varphi _{1},\gamma _{0})\ \frac{\partial t_{P}}{ \partial \gamma }(\gamma _{0})}). \] Hence, we get from Lemma \ref{Lemma5} that \[ |NEWNUM2|\leq \frac{\beta _{0}}{\eta }\Big(\frac{n(\varphi _{\eta },\gamma _{0}) }{n(\varphi _{1},\gamma _{0})}+\big|\frac{\frac{\partial t_{Q}}{\partial \gamma} (\varphi _{\eta },\gamma _{0})}{n(\varphi _{1},\gamma _{0}) \frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})}\big|\Big). \] Since $\frac{n(\varphi _{\eta },\gamma _{0})}{n(\varphi_{1},\gamma _{0})}\to \eta $ and \[ \Big|\frac{\frac{\partial t_{Q}}{\partial \gamma }(\varphi _{\eta }, \gamma _{0})}{n(\varphi _{1},\gamma_{0}) \frac{\partial t_{P}}{\partial \gamma }(\gamma _{0})}\Big|\to 0 \] as $\gamma _{0}\to \infty $, we have $|\frac{NEWNUM2}{\beta }|<1$ for $\gamma _{0}$ sufficiently large. Hence, (\ref{3.2}) holds and the proof of the theorem is complete. \end{proof} \begin{remark}\label{rem14} \rm Given $\eta \in (0,1)$, and $\beta \in (0,1)$, satisfying \begin{equation}\label{93} \beta < \frac{\eta}{\sqrt{0.5+\sqrt{ 0.25+\eta^4} } }, \end{equation} the solution of (\ref{3-03}) - (\ref{3-06}) satisfying $\varphi_{\eta}(x,x') \in (\frac{\pi}{2} +k\pi ,\frac{\pi}{2} +(k+1)\pi)$ exists and is unique, if $k$ is large. Note we have replaced 1 by $\eta$ in Definition \ref{def}, to give \[ \varphi _{\eta}(x,y)=-\int_{0}^{\eta}\frac{x(t)y'(t)-y(t)x'(t)}{x^{2}(t) +y^{2}(t)}dt\,. \] The inequality (\ref{93}) follows from (\ref{3.1}) by replacing $\eta$ and $\beta$ by their inverses. The change of variable $\tau = \eta^{-1}t$ leads to this, using the fact that Theorem \ref{Theorem2} holds for $\eta > 1$ too. \end{remark} \begin{thebibliography}{00} \bibitem{cal} B. Calvert and C. Gupta; {\em Multiple solutions for a super-linear three-point boundary value problem}, Nonlinear Analysis {\bf 50} (2002), 115--128. \bibitem{cap} A. Capietto and W. Dambrosio; {\em Multiplicity results for some two-point superlinear asymmetric boundary value problem}, Nonlinear Analysis {\bf 38}(7) (1999), 869--896. \bibitem{cap2} A. Capietto, M. Henrard, J. Mawhin, and F. Zanolin; {\em Continuation approach to superlinear two points boundary value problem}, Topol. Methods Nonlinear Analysis {\bf 3} (1994), 81--100. \bibitem{cos} D. Costa, D. G. de Figuerido, and P. Srikanth; {\em The exact number of solutions for a class of ordinary differential equations through Morse index computation}, J. Diff. Eq. {\bf 96} (1992), 185--199. \bibitem{din} G. Dinka and L. Sanchez; {\em Multiple solutions of boundary value problems: an elementary approach via the shooting method}, Nonlinear Differential Equations and Applications {\bf 1} (1994), 163--178. \bibitem{he} X. He and W. Ge; {\em Triple solutions for second-order three-point boundary value problems}, J. Math. Anal. Appl. {\bf 268} (2002) 256--265. \bibitem{hen1} M. Henrard; {\em Degre topologique et existence d'une infinite de solutions d'un probleme aux limites pour une equation singuliere}, Portugal. Math., {\bf 52} (1995), 153--165. \bibitem{Hthesis} M. Henrard; {\em Topological degree in boundary value problems: Existence and multiplicity results for second order differential equations}, Thesis, Universit\'{e} Catholique de Louvain, Louvain-la-Neuve, Belgium, 123pp (1995). \bibitem{hen2} M. Henrard; {\em Infinitely many solutions of weakly coupled super-linear systems}, Adv. Differential Equations {\bf 2} (1997), 753--778. \bibitem{inf} G. Infante and J. R. L. Webb; {\em Positive solutions of some nonlocal boundary value problems}, Abstract and Applied Analysis {bf 2003} 18 (2003), 1047--1060. \bibitem{kol} I. I. Kolodner; {\em Heavy rotating string - a nonlinear eigenvalue problem}, Comm Pure Appl Math {\bf 8} (1955), 395--408. \bibitem{kos} N. Kosmatov; {\em Semipositone m-point boundary value problems}, Electron. J. Diff. Eqns. {\bf 2004} No 119 (2004), 1--7. \bibitem{kwo} M. K. Kwong; {\em On the Kolodner-Coffman method for the uniqueness problem of Emden-Fowler BVP}, J. of Applied Math and Physics(ZAMP), {\bf 41} (1990), 79--103. \bibitem{kur} J. Kurzweil; Ordinary Differential Equations, Elsevier, Amsterdam, 1986. \bibitem{ma} R. Ma; {\em Positive solutions of a nonlinear 3-point boundary value problem}, Electronic J. Diff. Eqns. {\bf 34} (1999), 1--8. \end{thebibliography} \end{document}