\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2014 (2014), No. 139, pp. 1--11.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2014 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2014/139\hfil Eigenvalue problems] {Eigenvalue problems with $p$-Laplacian operators} \author[Y.-H. Cheng \hfil EJDE-2014/139\hfilneg] {Yan-Hsiou Cheng} % in alphabetical order \address{Yan-Hsiou Cheng\\ Department of Mathematics and Information Education\\ National Taipei University of Education\\ Taipei City 106, Taiwan} \email{yhcheng@tea.ntue.edu.tw} \thanks{Submitted November 7, 2013. Published June 16, 2014.} \subjclass[2000]{34A55, 34L15} \keywords{$p$-Laplacian; inverse spectral problem; instability interval} \begin{abstract} In this article, we study eigenvalue problems with the $p$-Laplacian operator: $$ -(|y'|^{p-2}y')'= (p-1)(\lambda\rho(x)-q(x))|y|^{p-2}y \quad \text{on } (0,\pi_{p}), $$ where $p>1$ and $\pi_{p}\equiv 2\pi/(p\sin(\pi/p))$. We show that if $\rho \equiv 1$ and $q$ is single-well with transition point $a=\pi_{p}/2$, then the second Neumann eigenvalue is greater than or equal to the first Dirichlet eigenvalue; the equality holds if and only if $q$ is constant. The same result also holds for $p$-Laplacian problem with single-barrier $\rho$ and $q \equiv 0$. Applying these results, we extend and improve a result by \cite{S07} by using finitely many eigenvalues and by generalizing the string equation to $p$-Laplacian problem. Moreover, our results also extend a result of Huang \cite{H97} on the estimate of the first instability interval for Hill equation to single-well function $q$. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{definition}[theorem]{Definition} \allowdisplaybreaks \section{Introduction} Recently there are many studies on the $p$-Laplacian operator: \begin{equation}-(|y'|^{p-2}y')'= (p-1)(\lambda\rho(x) -q(x))|y|^{p-2}y\quad \text{ on } (0,\pi_{p}), \label{eq1.1} \end{equation} where $p>1$ and $\pi_{p}\equiv 2\pi/(p\sin(\pi/p))$. An application for \eqref{eq1.1}, the most cited nowadays, is that of a highly viscid fluid flow (cf. Ladyzhenskaya \cite{la}, and Lions \cite{li}). This involves partial differential equations, but for symmetric flows, only the ordinary differential operator (perhaps in radial form) is involved (see, e.g., Binding and Dr\'{a}bek \cite{BD2003}, del Pino, Elgueta and Manasevich \cite{dem}, del Pino and Manasevich \cite{dm}, Rabinowitz \cite{r}, and Walter \cite{W98}). In 1979, Elbert \cite{E79} showed that the inverse function $S_{p} (x)\equiv w$ of the integral $$ x=\int_0^{w}\frac{dt}{(1-t^{p})^{1/p}}\qquad \text{for}\ 0\leq w\leq 1\,,$$ satisfies the initial valued problem \begin{equation*} \label{eq3.1} -(|u'|^{p-2}u')' = (p-1)|u|^{p-2}u ,\qquad u(0)=0 ,\ u'(0)=1. \end{equation*} The function $S_{p} (x)$ is called a generalized sine function and the value $\pi_{p} \equiv 2 \int_0^{1} (1-t^{p})^{-1/p} dt = 2\pi/(p \sin(\pi/p))$ is the first zero of $S_{p}(x)$. Continuing $S_{p}(x)$ symmetrically over $x\in [ \pi_{p}/2,\pi_{p}] $ and antisymmetrically outside $[0,\pi_{p}]$ by defining $$ S_{p}(x)=\begin{cases} S_{p}(\pi_{p}-x), & \text{if } \frac{\pi_{p}}{2}\leq x\leq \pi_{p}\,,\\ -S_{p}(x-\pi_{p}), & \text{if } \pi_{p}\leq x\leq 2\pi_{p}\,, \end{cases} $$ and $S_{p}(x)=S_{p}(x-2n\pi_{p})$ for $n=\pm 1, \pm 2, \dots$, he obtained a sine-like function defined on $\mathbb{R}$. Furthermore, he found the Pythagorean trigonometric identity for $p$-version: \begin{equation*} \label{eq2.3} |S_{p} (x)|^{p}+|S_{p}' (x)|^{p}=1\,. \end{equation*} Similarly, it may be defined an analogue of the hyperbolic sine function (see \cite{lind}) $Sh_{p} (x)\equiv v$ by the inverse function of the integral $x=\int_0^{v} (1+|t|^{p})^{-1/p} dt$. It is clearly that $Sh_{p} (x)=(-1)^{-1/p}S_{p} ((-1)^{1/p}x)$ and $Sh_{p}' (x)=S_{p}' ((-1)^{1/p}x)$ where $(-1)^{1/p}=e^{\pi i/p}$. Furthermore, we have $Sh_{p}''(x)=\frac{|Sh_{p}(x)|^{p-2}Sh_{p}(x)}{Sh_{p}'^{p-2}(x)}$ and \[ %\label{eq2.5} Sh_{p}'^{p}(x)-|Sh_{p}(x)|^{p}=1. \] Denote by $\sigma_{2k}$ ($\sigma_{2k-1}$) the set of periodic (anti-periodic) eigenvalues of \eqref{eq1.1} which admit the corresponding eigenfunctions with exactly $2k$ zeros in $[0,\pi_{p})$. In 2001, Zhang \cite{Z01} used a rotation number function to show the existence of the minimal eigenvalue $\underline{\lambda}_{n}=\min \sigma_{n}$ and the maximal eigenvalue $\overline{\lambda}_{n}=\max \sigma_{n}$, respectively. In particular, Binding and Rynne in a series of papers \cite{BR07,BR08,BR09} showed that \eqref{eq1.1} has an infinite sequence of variational and non-variational periodic eigenvalues and the multiplicity of the periodic eigenvalue can be arbitrary. They also showed that the Dirichlet eigenvalues $\{\mu_{n}\}_{n\geq 1}$ and Neumann eigenvalues $\{\nu_{n}\}_{n\geq 0}$ for \eqref{eq1.1} acting on $(0,\pi_{p})$ satisfy \begin{gather*} \dots \leq \overline{\lambda}_{2n-2}< \underline{\lambda}_{2n-1} \leq \mu_{2n-1},\\ \nu_{2n-1} \leq \overline{\lambda}_{2n-1} <\underline{\lambda}_{2n} \leq \mu_{2n}, \nu_{2n} \leq \overline{\lambda}_{2n}<\underline{\lambda}_{2n+1}\leq \dots. \end{gather*} Note that, for $q\equiv 0$ and $\rho \equiv 1$, we find $\nu_0=0$ and $\mu_{n}=\nu_{n}=n^{p}$ for $n\geq 1$. Recently, the eigenvalue gap/ratio are concerned. We say a function $f$ is single-well with transition point $a$ if $f$ is decreasing on $(0,a)$ and increasing on $(a,\pi_{p})$; $f$ is single-barrier if $-f$ is single-well. In 2010, Bogn\'ar and Dosly \cite{BD10} used the Pr\"ufer transformation derived by generalized sine function to show that the Dirichlet eigenvalues for \eqref{eq1.1} with $\rho \equiv 1$ and nonnegative single-well $q(x)$ satisfy $\mu_{n}/\mu_{m} \leq n^p/m^p$. Furthermore, Chen, Law, Lian and Wang \cite{CLLW13} also used the generalized Pr\"ufer transformation to show that $\mu_{n}/\mu_1\leq n^p$ for \eqref{eq1.1} with $\rho \equiv 1$ and nonnegative continuous $q(x)$. On the other hand, the authors in \cite{CLW} studied the first two Dirichlet eigenvalues for \eqref{eq1.1} and showed that (i) $\mu_2-\mu_1 \geq 2^{p}-1$ if $\rho \equiv 1$ and $q(x)$ is single-well with transition point at $\pi_{p}/2$; (ii) $\mu_2/\mu_1 \geq 2^{p}$ if $q(x) \equiv 0$ and $\rho(x)$ is single-barrier with transition point at $\pi_{p}/2$. In this article, we study the gap between the Dirichlet eigenvalues and Neumann eigenvalues. In \cite[Theorem 2.5]{S07}, Shen considered the spectra $\sigma_{D} = \{ \mu_1,\mu_2, \dots \}$, $\sigma_{DN} =\{ \tau_1,\tau_2, \dots \}$, $\sigma_{ND}= \{ \gamma_1,\gamma_2, \dots \}$, and $\sigma_{N} =\{\nu_0, \nu_1,\nu_2, \dots \}$ for the following string equations \begin{gather*} %\label{eq1.5} y''(x) +\mu \rho(x)y(x)=0,\quad y(0)=y(\pi)=0,\\ %\label{eq1.6} u''(x) +\tau \rho(x)u(x)=0,\quad u(0)=u'(\pi)=0,\\ %\label{eq1.7} z''(x) +\gamma \rho(x)z(x)=0,\quad z'(0)=z(\pi)=0,\\ %\label{eq1.8} v''(x) +\nu \rho(x)v(x)=0,\quad v'(0)=v'(\pi)=0, \end{gather*} respectively, where $\rho$ is a positive piecewisely continuous function defined on $[0, \pi]$. He showed that if $\sigma_{DN}= \sigma_{ND}$ and $\sigma_{N}= \sigma_{D}\cup \{0\}$, then $\rho (x)$ is a constant function at its points of continuity. Consider \eqref{eq1.1} and assume $q$ and $\rho$ satisfy (i) $\rho\equiv 1$ and $q$ is single-well with transition point $a=\pi_{p}/2$, or (ii) $q\equiv 0$ and $\rho$ is single-barrier with transition point $a=\pi_{p}/2$. In this paper, we show that $\mu_1=\nu_1$ if and only if (i) $q$ is constant, or (ii) $\rho$ is constant, respectively. Our results extend and improve the result of Shen \cite[Theorem 2.5]{S07} by using finitely many eigenvalues and by generalizing the string equation to $p$-Laplacian eigenvalue problem. \begin{theorem} \label{thm1.1} Consider \eqref{eq1.1} with $q(x)\in L^{1}(0,\pi_{p})$ and $\rho\equiv 1$. If $q(x)$ is single-well on $(0,\pi_{p})$ with transition point $a=\pi_{p}/2$, then $\nu_1\geq \mu_1$. Equality holds if and only if $q$ is constant. If $a\neq \pi_{p}/2$, then there exist some functions $q$ giving $\nu_1< \mu_1$. \end{theorem} \begin{theorem} \label{thm1.2} Consider \eqref{eq1.1} with a positive piecewisely continuous function $\rho$ and $q\equiv 0$. If $\rho(x)$ is single-barrier on $(0,\pi_{p})$ with transition point $a=\pi_{p}/2$, then $\nu_1\geq \mu_1$. Equality holds if and only if $\rho$ is constant. If $a\neq \pi_{p}/2$, then there exist some functions $\rho$ giving $\nu_1< \mu_1$. \end{theorem} The proof of Theorem \ref{thm1.1} follows the method developed by Horv\'ath \cite{H02}. We first perturb the extremal function $q$ and study the identity for $\frac{d}{dt}(\nu_1(t)-\mu_1(t))$ where $t$ is a parameter. We will show that the optimal function $q$ is a step function with a jump at $\pi_{p}/2$ and then compel it to be constant. Furthermore, by the principle of duality, the same method also works for \eqref{eq1.1} with $q\equiv 0$ and single-barrier $\rho$. We shall remark that Theorems \ref{thm1.1} and \ref{thm1.2} can be used to solve inverse problems of the instability interval for $p=2$: \begin{equation} -y''=(\lambda\rho(x) -q(x))y\,.\label{eq1.10} \end{equation} Denote by $\{\lambda_{n}\}_{n\geq 0}$ and $\{\lambda'_{n}\}_{n\geq 1}$ the eigenvalues of \eqref{eq1.10} with $q(x)=q(x+\pi), \rho(x)=\rho(x+\pi)$ under the periodic ($y(0)=y(\pi)$, $y'(0)=y'(\pi)$), and anti-periodic ($y(0)=-y(\pi)$, $y'(0)=-y'(\pi)$) boundary conditions, respectively. It is known \cite{CL55} (see also \cite{B46,MW66}) that $\nu_0\leq \lambda_0$ and \begin{equation} \begin{gathered} \dots\leq\lambda_{2n-2} <\lambda'_{2n-1} \leq \nu_{2n-1}, \\ \mu_{2n-1} \leq \lambda'_{2n}< \lambda_{2n-1} \leq \nu_{2n}, \\ \mu_{2n} \leq \lambda_{2n}<\lambda'_{2n+1}\leq \dots\\ \end{gathered}\label{eq1.12} \end{equation} The intervals $(\lambda'_{2n-1},\lambda'_{2n})$ and $(\lambda_{2n-1},\lambda_{2n})$ are called the $(2n-1)$-th and $2n$-th instability intervals. The interval $(-\infty,\lambda_0)$ is called the zero-th instability interval. In 1946, Borg \cite{B46} studied an inverse problem for Hill's equation. He showed that the potential $q(x)$ is constant if and only if all instability intervals, except the zero-th, are absent. Later, Hochstadt \cite{H84} generalized Borg's result and showed that if $q$ is $C^{1}$, then $q$ has period $1/n$ if and only if all those finite instability intervals whose index is not a multiple of $n$ vanish. In 1997, Huang \cite{H97} proved that if $q$ is symmetric single-well (or symmetric single-barrier), then $q$ is constant if and only if the first instability interval is absent, i.e. $\lambda'_1=\lambda'_2$. Thus, for all instability intervals, the first instability gives the most information about the potential $q$. Using Theorems \ref{thm1.1} and \ref{thm1.2}, and \eqref{eq1.12}, we may eliminate the assumption on the symmetric of $q$ and obtain the following results immediately. \begin{corollary} \label{coro1.3} Consider \eqref{eq1.10} with $\pi$-periodic functions $\rho$ and $q$. Then the first instability interval is absent if and only if one of the following conditions holds: \begin{itemize} \item[(i)] $\rho\equiv 1$ and $q$ is single-well with transition point $a=\pi/2$. \item[(ii)] $q\equiv 0$ and $\rho$ is single-barrier with transition point $a= \pi/2$. \end{itemize} \end{corollary} The paper is organized as follows. In section 2, we use a modified Pr\"ufer substitution and comparison theorem to derive properties of eigenfunctions. In section 3, we study two generalized trigonometric equations. The Dirichlet and Neumann eigenvalues are corresponding to the roots of two generalized trigonometric equations, respectively. Finally, in section 4, we give proofs of our main theorems \ref{thm1.1} and \ref{thm1.2}. \section{Preliminaries} At the beginning of this section, we give two formulas of generalized trigonometric functions. The proof is similar to the classical trigonometric functions, so we omit it here. \begin{lemma} \label{lem2.1} Define the generalized tangent function by $T_{p}(x)\equiv S_{p}(x)/ S_{p}'(x)$ for $x\neq (k+1/2)\pi_{p}$ and the generalized reciprocal tangent function by $RT_{p}(x)\equiv S_{p}'(x)/ S_{p}(x)$ for $x\neq k\pi_{p}$. Then we have \begin{itemize} \item[(i)] $T_{p}'(x)=1+|T_{p}(x)|^{p} .$ \item[(ii)] $RT_{p}'(x)= -(RT_{p}(x))^{2} (1+|T_{p}(x)|^{p}).$ \end{itemize} \end{lemma} Denote by $(\mu_i, \phi_i)_{i\geq 1}$ the normalized Dirichlet eigenpair and $(\nu_i, \psi_i)_{i\geq 0}$ the normalized Neumann eigenpair of \eqref{eq1.1} with $\phi_i(x)>0$, $\psi_i(x)>0$ for $x$ near $0^{+}$. The normalized condition means $\int_0^{\pi_{p}}\rho(x)|\phi_i(x)|^{p}dx =\int_0^{\pi_{p}}\rho(x)|\psi_i(x)|^{p}dx =1 $ for all $i$. \begin{definition} {def1} \rm Let $f$ and $g$ be continuous functions and $g(x)\neq 0$. Define $h(x)\equiv f(x)/g(x)$. We say $\alpha_0$ is a crossing point of $f$ and $g$ if $h(\alpha_0)=1$ and $h$ satisfies one of the following conditions \begin{itemize} \item[(i)] $h(\alpha_0^{+})>1$ and $h(\alpha_0^{-})<1$. \item[(ii)] $h(\alpha_0^{+})<1$ and $h(\alpha_0^{-})>1$. \end{itemize} \end{definition} \begin{lemma} \label{lem2.2} There are exactly two crossing points of $|\phi_1(x)|$ and $|\psi_1(x)|$ in $(0,\pi_{p})$. \end{lemma} \begin{proof} First, we introduce a generalized Pr\"ufer substitution derived by $S_{p}$ and $S_{p}'$: \begin{gather*} \phi_1(x)=r(x) S_{p}(\theta_{D}(x))\,,\quad \phi_1'(x) = r(x)S_{p}'(\theta_{D}(x))\,, \\ \psi_1(x)=R(x) S_{p}(\theta_{N}(x))\,,\quad \psi_1'(x) = R(x)S_{p}'(\theta_{N}(x))\,, \end{gather*} where $\theta_{D}(0)=0$ and $\theta_{N}(0)=\pi_{p}/2$. Here, $\theta_{D}(x)$ and $\theta_{N}(x)$ are called the Pr\"ufer angles of $\phi_1(x)$ and $\psi_1(x)$, respectively. By direct calculation, we find that \begin{gather} \theta_{D}'(x)=|S_{p}'(\theta_{D}(x))|^{p} +(\mu_1\rho(x)-q(x))|S_{p}(\theta_{D}(x))|^{p},\label{eq2.0.1}\\ \theta_{N}'(x)=|S_{p}'(\theta_{N}(x))|^{p} +(\nu_1\rho(x)-q(x))|S_{p}(\theta_{N}(x))|^{p}.\label{eq2.0.2} \end{gather} Let $x_0$ be the unique zero of $\psi_1(x)$ in $(0,\pi_{p})$. Since $\phi_1(x)>0$ on $(0,x_0)$, $0=\phi_1(0)<\psi_1(0)$ and $\phi_1(x_0)>\psi_1(x_0)=0$, we find the number of the crossing points of $\phi_1(x)$ and $\psi_1(x)$ in $(0,x_0)$ must be odd. Assume $00$. Let $t_1$ be the first root of $f(t)=-f(t-m)$ and $t_2$ be the second root of $g(t)=-g(t-m)$. Then $t_2> t_1$. \end{lemma} \begin{proof} First, note that $t_1\in (1,\min\{1+m, 2^{p}\})$ and $t_2\in (1, 3^{p})$ for $m>0$. (i) Assume $t\geq 0$. Then, by Lemma \ref{lem2.1}, we find \begin{align*} g'(t) &=\frac{1}{p}t^{\frac{1-p}{p}}RT_{p}(t^{1/p}\frac{\pi_{p}}{2} +\frac{\pi_{p}}{2})-t^{1/p}\big(1+|T_{p}(t^{1/p}\frac{\pi_{p}}{2} +\frac{\pi_{p}}{2})|^{p}\big)\\ &\quad\times RT_{p}^{2}(t^{1/p}\frac{\pi_{p}}{2} +\frac{\pi_{p}}{2})\cdot\frac{1}{p}t^{\frac{1-p}{p}}\frac{\pi_{p}}{2}\\ &=\frac{t^{\frac{1-p}{p}}}{2p|S_{p}(t^{1/p}\frac{\pi_{p}}{2} +\frac{\pi_{p}}{2})|^{2}}\Big\{2S_{p}(t^{1/p}\frac{\pi_{p}}{2} +\frac{\pi_{p}}{2})S_{p}'(t^{1/p}\frac{\pi_{p}}{2} +\frac{\pi_{p}}{2})\\ &\quad -t^{1/p}\pi_{p}|S_{p}'(t^{1/p} \frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})|^{2-p}\Big\}\\ &\equiv \frac{t^{\frac{1-p}{p}}}{2p|S_{p}(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})|^{2}}\tilde{g}(t). \end{align*} If $S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) >0$, in this case $t^{1/p}\in (2+4n,4+4n)$ for $n\geq 0$, then \begin{align*} \tilde{g}(t) &= S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) [2S_{p}(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})-t^{1/p}\pi_{p}|S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})|^{1-p}]\ \\ &\leq S_{p}'(t^{1/p}\frac{\pi_{p}}{2}) [2S_{p}(t^{1/p}\frac{\pi_{p}}{2})-t^{1/p}\pi_{p}]\\ &\equiv S_{p}'(t^{1/p}\frac{\pi_{p}}{2}) h(t) . \end{align*} Since $h((2+4n)^{p})<0$ for $n\geq 0$, and $h'(t)=\frac{t^{\frac{1-p}{p}}\pi_{p}}{p}(S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})-1)<0$ for $t^{1/p}\in (2+4n,4+4n)$ and $n\geq 1$, we have $h(t)<0$ for $t^{1/p}\in (2+4n,4+4n)$ and $n\geq 0$. Hence $g'(t)<0$ for $t^{1/p}\in (2+4n,3+4n)\cup (3+4n,4+4n)$ and $n\geq 0$. Similarly, if $S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) <0$, in this case $t^{1/p}\in (4n,4n+2)$ for $n\geq 0$, then \begin{align*} \tilde{g}(t) &= S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) [2S_{p}(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})+t^{1/p}\pi_{p}|S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})|^{1-p}] \\ &\leq S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) [2S_{p}(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})+t^{1/p}\pi_{p}] \\ &\equiv S_{p}'(t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) \tilde{h}(t)\,. \end{align*} Since $\tilde{h}((4n)^{p})>0$ for $n\geq 0$ and $\tilde{h}'(t)=\frac{t^{\frac{1-p}{p}}\pi_{p}}{p}(S_{p}' (t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2})+1)>0$ for $t^{1/p}\in (4n,4n+2)$ and $n\geq 0$, we have $\tilde{h}(t)>0$ for $t^{1/p}\in (4n,4n+2)$ and $n\geq 0$ and hence $g'(t)<0$ for $t^{1/p}\in (4n,4n+1)\cup(4n+1,4n+2)$ and $n\geq 0$. \smallskip (ii) Assume $t< 0$. Let $\hat{t}=-t$ and $\tilde{t}=\hat{t}^{1/p}\frac{\pi_{p}}{2} +(-1)^{-1/p}\frac{\pi_{p}}{2}$. Since $$ g(t)=t^{1/p}RT_{p} (t^{1/p}\frac{\pi_{p}}{2}+\frac{\pi_{p}}{2}) =(-1)^{1/p}\hat{t}^{1/p}\frac{S_{p}'((-1)^{1/p}\tilde{t})}{S_{p}((-1)^{1/p}\tilde{t})} =\hat{t}^{1/p}\frac{Sh_{p}'(\tilde{t})}{Sh_{p}(\tilde{t})}\,, $$ we have \begin{align*} g'(t) &= -\frac{1}{p}\hat{t}^{\frac{1-p}{p}}\frac{Sh_{p}'(\tilde{t})}{Sh_{p}(\tilde{t})} +\hat{t}^{1/p}(-\frac{1}{p}\hat{t}^{\frac{1-p}{p}}) \frac{\pi_{p}}{2}\Big(\frac{Sh_{p}''(\tilde{t})}{Sh_{p}(\tilde{t})} -\frac{Sh_{p}'^{2}(\tilde{t})}{Sh_{p}^{2}(\tilde{t})}\Big)\\ &= \frac{-\frac{1}{p}\hat{t}^{\frac{1-p}{p}}}{Sh_{p}^{2}(\tilde{t})} \Big[Sh_{p}'(\tilde{t})Sh_{p}(\tilde{t}) +\frac{\pi_{p}}{2}\hat{t}^{1/p}\Big(\frac{|Sh_{p}(\tilde{t})|^{p}}{Sh_{p}'^{p-2} (\tilde{t})}-Sh_{p}'^{2}(\tilde{t})\Big) \Big]\\ &= \frac{-\frac{1}{p}\hat{t}^{\frac{1-p}{p}}}{Sh_{p}^{2}(\tilde{t})} [ Sh_{p}'(\tilde{t})Sh_{p}(\tilde{t})-\frac{\pi_{p}}{2} \hat{t}^{1/p}Sh_{p}'^{2-p}(\tilde{t})]\\ &\equiv \frac{-\frac{1}{p}\hat{t}^{\frac{1-p}{p}}}{Sh_{p}^{2}(\tilde{t})} \hat{g}(t)\,. \end{align*} Using similar argument as step (i), we can show $\hat{g}(t)>0$ and hence $g'(t)<0$ for all $t<0$. \smallskip (iii) Let $t=t(m)$. If $g(t)=-g(t-m)$, then $g'(t)\frac{dt}{dm}=-g'(t-m)(\frac{dt}{dm}-1)$ and hence $$0<\frac{dt}{dm}=\frac{g'(t-m)}{g'(t)+g'(t-m)}<1\,.$$ This implies $t_2(m)$ is strictly increasing for $m>0$. On the other hand, when $m=2^{p}$, we have $t_2=2^{p}$ and $$t_2-t_1>0\quad \text{for}\ m\geq 2^{p}\,.$$ Theqrefore, we only need to consider $00$. \end{proof} \begin{lemma}\label{lem3.2} Let $m>1$. Let $s_1$ be the first root of $f(s)=-f(sm)$ and $s_2$ be the second root of $g(s)=-g(sm)$. Then $s_2> s_1$. \end{lemma} \begin{proof} Note that $s_1, s_2\in (\frac{1}{m}, \min\{1, \frac{2^{p}}{m}\} )$. If $s_1\geq s_2$ for some $m>1$, then $\frac{1}{m}\leq s_2-f(s_2)\,. $$ This implies $s_2=1$. Hence $s_1\leq s_2$ for $m>1$. \end{proof} \section{Proof of Main Theorems} \begin{proof}[Proof of Theorem \ref{thm1.1}] For $M>0$, denote $$ A_{M}=\big\{ 0\leq q(x)\leq M : q \text{ is single-well with transition point } \frac{\pi_{p}}{2}\big\}. $$ Then $A_{M}$ is closed and $E(q)\equiv (\nu_1-\mu_1)(q)$ is bounded on $A_{M}$. Hence there exists an optimal function $q_0$ giving the minimal eigenvalue gap $\nu_1-\mu_1$. Recall the definitions of $x_{-}$ and $x_{+}$ in \eqref{eq2.1}. We shall define $q(x,t)=(1-t)q_0(x)+tq_1(x)$ for $t\in [0,\pi_{p}]$ for some appropriated function $q_1$. First, assume $x_{-}\leq \pi_{p}/2 \leq x_{+}$. Let $$ q_1(x)=\begin{cases} q_0(x_{-})&\text{on } (0,\frac{\pi_{p}}{2}),\\ q_0(x_{+})&\text{on } (\frac{\pi_{p}}{2},\pi_{p}). \end{cases} $$ By the optimality of $q_0$ and Lemma \ref{lem2.3}, we have \begin{align*} 0&\leq\ frac{d}{dt}(\nu_1(t)-\mu_1(t))|_{t=0}\\ &= \int_0^{\pi_{p}}(q_1(x)-q_0(x))(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dt\,, \end{align*} which is nonpositive. Hence, $q_0(x)=q_1(x)$ a.e. on $[0,\pi_{p}]$. If $x_{-}>\pi_{p}/2$, we let $$ q_1(x)=\{\begin{cases} 0& \text{on } (0,x_{-}),\\ M& \text{on } (x_{-},\pi_{p}). \end{cases} $$ By the normality of $\phi_1$ and $\psi_1$, we have $$ \int_0^{x_{-}}(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dx > 0> \int_{x_{-}}^{\pi_{p}}(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dx . $$ Hence, by the optimality of $q_0$, we have \begin{align*} 0&\leq \frac{d}{dt}(\nu_1(t)-\mu_1(t))|_{t=0}\\ &= \int_0^{\pi_{p}}(q_1(x)-q_0(x))(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dx\\ &= \int_0^{x_{-}}(-q_0(x))(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dx\\ &\quad + \int_{x_{-}}^{\pi_{p}}(M-q_0(x))(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dx\\ &\leq -q_0(\frac{\pi_{p}}{2})\int_0^{x_{-}}(|\psi_1(x,0)|^{p} -|\phi_1(x,0)|^{p})dt\\ &\quad +(M-q_0(x_{+}))\int_{x_{-}}^{\pi_{p}}(|\psi_1(x,0)|^{p} -|\phi_1(x,0)|^{p})dt\,, \end{align*} which is non-positive. This implies that $q_0=0$ on $(0,x_{-})$ and $=M$ on $(x_{-},\pi_{p})$. But this makes a contradiction to Lemma \ref{lem2.4}. Hence this case is refuted. The case $x_{+}\leq \pi_{p}/2$ is similar. After simplification, the optimal function $q_0$ is a $1$-step function. Without loss of generality, let $$ q_0(x)=\begin{cases} 0&\text{on } (0,\frac{\pi_{p}}{2}),\\ m&\text{on } (\frac{\pi_{p}}{2},\pi_{p}). \end{cases} $$ By equating the corresponding ratio by $y'/y$ at $\pi_{p}/2$, $\nu_1$ is the second root of the functional equation $\lambda^{1/p}RT_{p}(\frac{\pi_{p}}{2}\lambda^{1/p}+\frac{\pi_{p}}{2}) =-(\lambda-m)^{1/p}RT_{p}(\frac{\pi_{p}}{2}(\lambda-m)^{1/p}+\frac{\pi_{p}}{2})$, and, similarly, $\mu_1$ is the first root of $\lambda^{1/p}RT_{p}(\frac{\pi_{p}}{2}\lambda^{1/p})=-(\lambda -m)^{1/p}RT_{p}(\frac{\pi_{p}}{2}(\lambda-m)^{1/p})$. Using Lemma \ref{lem3.1}, we obtain $\nu_1-\mu_1> 0$. Finally, if the transition point $a$ is not $\pi_{p}/2$, we let $$ q(x,t)=\begin{cases} t&\text{on } [0,a],\\ 0&\text{on } [a,\pi_{p}]. \end{cases} $$ Since $\phi_1(x,0)=(p/\pi_{p})^{1/p}S_{p} (x)$, $\psi_1(x,0)=(p/\pi_{p})^{1/p}S_{p}(x+\pi_{p}/2)$, and $$ \int_0^{\pi_{p}/2}(|\psi_1(x,0)|^{p}-|\phi_1(x,0)|^{p})dx=0, $$ we have $$ \frac{d}{dt}(\nu_1(t)-\mu_1(t))|_{t=0}=\int_0^{a}(|\psi_1(x,0)|^{p} -|\phi_1(x,0)|^{p})dx < 0\,, $$ when $00$, $\nu_1(t)-\mu_1(t)<0$ when $01$, denote $$ A_{M}=\big\{ \frac{1}{M}\leq \rho(x)\leq M : \rho\ \text{is single-barrier with transition point}\ \frac{\pi_{p}}{2}\big\}. $$ Then there exists an optimal function $\rho_0$ giving the minimal eigenvalue ratio $\nu_1/\mu_1$. Similar to the proof of Theorem \ref{thm1.1} and by Lemma \ref{lem2.4}, the cases $x_{+}<\pi_{p}/2$ and $x_{-}>\pi_{p}/2$ are refuted by using suitable $\rho_0$'s. Hence we have $x_{-}\leq \pi_{p}/2\leq x_{+}$ and $$ \rho_0(x)=\begin{cases} \rho_0(x_{-})&\text{on }(0,\frac{\pi_{p}}{2}),\\ \rho_0(x_{+})&\text{on }(\frac{\pi_{p}}{2},\pi_{p}). \end{cases} $$ That is the optimal function $\rho_0$ is a $1$-step function. Without loss of generality, let $$ \rho_0(x)=\begin{cases} 1&\text{on } (0,\frac{\pi_{p}}{2}),\\ m&\text{on } (\frac{\pi_{p}}{2},\pi_{p}), \end{cases} $$ for some $m>1$. Then $\nu_1$ is the second root of $$ RT_{p}(\frac{\pi_{p}}{2}\lambda^{1/p}+\frac{\pi_{p}}{2}) =-m^{1/p}RT_{p}(\frac{\pi_{p}}{2}(m\lambda)^{1/p}+\frac{\pi_{p}}{2}), $$ and $\mu_1$ is the first root of $$ RT_{p}(\frac{\pi_{p}}{2}\lambda^{1/p}) =-m^{1/p}RT_{p}(\frac{\pi_{p}}{2}(m\lambda)^{1/p}). $$ Hence, by Lemma \ref{lem3.2}, $\nu_1/\mu_1>1$. Finally, we let $$ \rho(x,t)=\begin{cases} t&\text{on } [0,a],\\ 1&\text{on } [a,\pi_{p}]. \end{cases} $$ Then it can be shown that, if $0< \pi_{p}/2 -a<< 1$, the function $\rho(x,t)$ gives $\nu_1/\mu_1<1$ for small $t>1$. \end{proof} \subsection*{Acknowledgments} The author is partially supported by Ministry of Science and Technology, Taiwan, Republic of China, under contract nos. NSC 102-2115-M-152-002. \begin{thebibliography}{10} \bibitem{BD2003} P. A. Binding and P. Dr\'{a}bek; Sturm-Liouville theory for the $p$-Laplacian, \emph{Studia Scientiarum Mathematicarum Hungarica} \textbf{40} (2003), 373-396. \bibitem{BR07} P. A. Binding, B. P. Rynne; The spectrum of the periodic $p$-Laplacian; \emph{J. Differential Equations} \textbf{235} (2007), 199-218. \bibitem{BR08} P. A. Binding and B. P. Rynne; Variational and non-variational eigenvalues of the $p$-Laplacian, \emph{J. Differential Equations} \textbf{244} (2008), 24-39. \bibitem{BR09} P. A. Binding, B. P. Rynne; Oscillation and interlacing for various spectra of the $p$-Laplacian, \emph{Nonlinear Analysis} \textbf{71} (2009), 2780-2791. \bibitem{br89} G. Birkhoff, G. C. Rota; \emph{Ordinary Differential Equations 4th ed.}, Wiley (New York), 1989. \bibitem{BD10} G. Bogn\'ar, O. Dosly; The ratio of eigenvalues of the Dirichlet eigenvalue problem for equations with one-dimensional $p$-Laplacian, \emph{Abstract and Applied Analyis} \textbf{2010} (2010), doi:10.1155/2010/123975. \bibitem{B46} G. Borg; Eine Umkehrung der Sturm-Liouvilleschen Eigenwertaufgabe. Bestimmung der Differentialgleichung durch die Eigenwerte, \emph{Acta Math} \textbf{78} (1946), 1-96. \bibitem{CLLW13} C. Z. Chen, C. K. Law, W. C. Lian, W. C. Wang; Optimal upper bounds for the eigenvalue ratios of one-dimensional $p$-Laplacian, \emph{Proc. Amer. Math. Soc.} \textbf{141} (2013), 883-893. \bibitem{CLW} Y. H. Cheng, W. C. Lian, W. C. Wang; The dual eigenvalue problems for $p$-Laplacian, \emph{Acta Math Hungar.}, to appear. \bibitem{CL55} E. A. Coddington, N. Levinson; \emph{Theory of Ordinary Differential Equations}, New York: McGraw-Hill, 1955. \bibitem{E79} \'{A}. Elbert; A half-linear second order differential equation, \emph{Qualitative theory of differential equations Vol. I, II} (1979), 153-180; \emph{Colloq. Math. Soc. J\'{a}nos Bolyai} \textbf{30} (1981), North-Holland Amsterdam-New York. \bibitem{H84} H. Hochstadt; A generalization of Borg's inverse theorem for Hill's equation \emph{J. Math. Anal. Appl.} \textbf{102} (1984) 599-605. \bibitem{H02} M. Horv\'ath; On the first two eigenvalues of Sturm-Liouville operators, \emph{Proc. Amer. Math. Soc.} \textbf{131} (2002), 1215-1224. \bibitem{H97} M. J. Huang; The first instability interval for Hill equations with symmetric single well potentials, \emph{Proc. Amer. Math. Soc.} \textbf{125} (1997), 775-778. \bibitem{K76} J. B. Keller; The minimum ratio of two eigenvalues, \emph{SIAM J. Appl. Math.} \textbf{31} (1976), 485-491. \bibitem{la} O. A. Ladyzhenskaya; \emph{The Mathematical Theory of Viscous Incompressible Flow}, 2nd ed., Gordon and Breach (New York), 1969. \bibitem{L94} R. Lavine; The eigenvalue gap for one-dimensional convex potentials, \emph{Proc. Amer. Math. Soc.} \textbf{121} (1994), 815-821. \bibitem{lind} P. Lindqvist; Some remarkable sine and cosine functions, \emph{Ric. Mat.} \textbf{44} (1995), 269-290. \bibitem{li} J. L. Lions; \emph{Quelques methodes de resolution des problemes aux limites non lineaires}, Dunod (Paris), 1969. \bibitem{MW66} W. Magnus, S. Winkler; \emph{Hill's Equation}, New York : Interscience Publishers, 1966. \bibitem{dem} M. del Pino, M. Elgueta, R. Manasevich; A homotopic deformation along p of a Leray-Schauder degree result and existence for $(|u'|^{p-2}u')'+f(t,u)=0$, $u(0)=u(1)=0$, $p>1$, \emph{J. Differential Equations} \textbf{80} (1989), 1-13. \bibitem{dm} M. del Pino, R. Manasevich; Global bifurcation from the eigenvalues of the $p$-Laplacian, \emph{J. Differential Equations} \textbf{92} (1991), 226-251. \bibitem{r} P. H. Rabinowitz; Some global results for nonlinear eigenvalue problems, \emph{J. Functional Analysis} \textbf{7} (1971), 487-513. \bibitem{S07} C. L. Shen; On some inverse spectral problems related to the Ambarzumyan problem and the dual string of the string equation, \emph{Inverse Problems} \textbf{23} (2007), 2417-2436. \bibitem{W98} W. Walter; Sturm-Liouville theory for the radial $\triangle_{p}$-operator, \emph{Math. Z.} \textbf{227} (1998), 175-185. \bibitem{Z01} M. Zhang; The rotation number approach to eigenvalues of the one-dimensional $p$-Laplacian with periodic potentials, \emph{J. London Math. Soc.} \textbf{64} (2001), 125-143. \end{thebibliography} \end{document}