\documentclass[reqno]{amsart} \usepackage{amssymb} \usepackage{hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2005(2005), No. 54, pp. 1--22.\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-2005/54\hfil Local stability of spike steady states] {Local stability of spike steady states in a simplified Gierer-Meinhardt system} \author[G. E. Karadzhov, D. E. Edmunds, P.P.N. de Groen\hfil EJDE-2005/54\hfilneg] {Georgi E. Karadzhov, David Edmunds, Pieter de Groen} \address{Georgi E. Karadzhov \hfill\break Institute of Mathematics and Informatics \\ Bulgarian Academy of Sciences\\ 1113 Sofia, Bulgaria} \email{geremika@math.bas.bg} \address{David E. Edmunds \hfill\break Department of Mathematics, University of Sussex\\ Brighton BN1 9RF, U. K.} \email{d.e.edmunds@sussex.ac.uk} \address{Pieter P.N. de Groen \hfill\break Department of Mathematics, Vrije Universiteit Brussel \\ Pleinlaan 2, B-1050 Brussels, Belgium} \email{pdegroen@vub.ac.be} \date{} \thanks{Submitted March 3, 2005. Published May 23, 2005.} \subjclass[2000]{35B25, 35K60} \keywords{Spike solution; singular perturbations; reaction-diffusion equations; \hfill\break\indent Gierer-Meinhardt equations} \begin{abstract} In this paper we study the stability of the single internal spike solution of a simplified Gierer-Meinhardt' system of equations in one space dimension. The linearization around this spike consists of a selfadjoint differential operator plus a non-local term, which is a non-selfadjoint compact integral operator. We find the asymptotic behaviour of the small eigenvalues and we prove stability of the steady state for the parameter $(p,q,r,\mu)$ in a four-dimensional region (the same as for the shadow equation, \cite{gk1}) and for any finite $D$ if $\varepsilon$ is sufficiently small. Moreover, there exists an exponentially large $D(\varepsilon)$ such that the stability is still valid for $D0,\\ \tau H_t= D \Delta H-\mu H+U^r H^{-s}\quad x\in\Omega,\; t>0\,,\\ \partial_nU=0=\partial_nH \quad x\in\partial\Omega,\; t\ge 0, \end{gathered} \end{equation} where $U$ and $H$ represent activator and inhibitor concentrations, $\varepsilon$ and $D$ their diffusivities, and where $\tau$ and $\mu$ are the reaction time rate and the decay rate of the inhibitor; $D$ is assumed to be positive and $\varepsilon$ and $\tau$ small (positive). $\Omega$ is a bounded domain; we shall restrict our analysis to one space dimension and choose $\Omega:=[-1,1]$. The exponents $\{p>1\,, ~q>0\,,~r>1\}$ satisfy the inequality \begin{equation} \label{eq1s1} \gamma_r:=\frac{qr}{p-1}>1\,. \end{equation} Iron, Ward \& Wei \cite{ward2} analyze by formal asymptotic expansions the stability of approximate $N-$spike solutions for the simplified system, namely, that obtained by taking $\tau=0$. Rigorous results are obtained in \cite{gk1} for the case of the so-called shadow equation $$ U_t= \varepsilon^2 U_{xx}-U+2^q U^{p}\Bigl(\int_{-1}^1 U^r dx\Bigr)^{-q} \,,\quad U_x(-1,t)=U_x(1,t), $$ derived from the system in the limit $D\to \infty$ and $\tau\to 0$. In this paper we propose to study rigourously the simplified system when $s=0$. After rescaling $U \to \varepsilon^{-\nu_1} U$, $H \to \varepsilon^{-\nu_2} H$, $\nu_1:=\frac{q}{1-p+qr}$, $\nu_2:=\frac{p-1}{1-p+qr}$, we get the system \begin{equation} \label{simplgm} \begin{gathered} U_t= \varepsilon^2\Delta U-U+U^p H^{-q} \quad x\in\Omega,\; t>0,\\ 0= D\Delta H-\mu H+\varepsilon^{-1}U^r \quad x\in\Omega,\; t>0\\ \partial_nU=0=\partial_nH \quad x\in\partial\Omega,\; t\ge 0 \end{gathered} \end{equation} Our goal is a rigorous study of stability of the single internal spike solution via the spectrum of the linearized operator. In \cite{takagi} it is shown by the implicit function theorem, that such a spike solution exists for \begin{quote} $p>1$, $r>1$, $q>0$, $qr\neq (p-1)(s+1)$\\ and $D$ exponentially large w.r.t. $\varepsilon>0$, \end{quote} which is close to the shadow spike corresponding to $D=\infty$. A different approach, based on geometric singular perturbation theory, is applied in \cite{doelman1} for the same problem on the whole line. In \cite{ww} a rigorous treatment of the stability of multiple-peaked spike solutions is given, based on the Liapunov-Schmidt reduction method. See also \cite{ward3}, \cite{ward4}, where stability and Hopf bifurcation of the one-spike solution is studied. In this paper we construct a single spike solution (on a bounded interval) by fix-point iteration and we establish stability by a rigorous analysis of the spectrum of the first variation around this spike. In section \ref{spike} we construct a positive (stationary) solution with a single internal spike for $p>1$, $r>1$, $q>0$, $qr\neq (p-1)(s+1)$ and for any fixed $D$, using contraction around another shadow spike that exists for all $D>0$. The existence of such a solution is proved in \cite{wei} in a larger domain $\sqrt D \gg \varepsilon$. In section \ref{reference} we study the spectrum of the differential operator $L_\varepsilon$. The eigenvalues are estimated using Rayleigh's quotient. In section \ref{perturb} we make a detailed study of the influence of the nonlocal term on the eigenvalues as a function of the parameters $p$, $q$, $r$ and $D$ using perturbational methods. We construct an asymptotic approximation of the small eigenvalue $\lambda_\varepsilon$ of the perturbed non-selfadjoint operator and show that $Re\lambda_\varepsilon >0$ for any finite $D$ and for sufficiently small $\varepsilon$. We cover not only the usual known cases $r=p+1$, or $r=2$, $10$ for $D0$ and all $\varepsilon$, $0<\varepsilon<\varepsilon_0(\delta)$ if $p>1 ,r>1, q>0, qr\neq p-1$. By definition, the (single) spike solution (or spike) is such a steady state for which $S(x)=O(1)$ as $\varepsilon\to 0$ in a neighbourhood of the origin and $S(x)$ is exponentially small outside. Let $h$ be the solution of the linear equation $$ h''-\delta^2 h =-f,\; h'(\pm1)=0\,, $$ then $$ h(x)=\int_{-1}^1 {\widetilde G}_\delta (x,y)f(y)dy, $$ where Green's function ${\widetilde G}_\delta$ is given by \begin{equation} \label{green3} {\widetilde G}_\delta (x,y)=\frac{1}{\delta \sinh 2\delta}\cosh \delta(1+x) \cosh\delta(1-y)\quad \mbox{if }x1$ this equation happens to have the closed form solution, cf. \cite{gk1}, \begin{equation} \label{eq5s1} w_p(\xi):={\textstyle\left(\frac{p+1}{2}\right)}^{1\over p-1} \left(\cosh({\textstyle{\frac{p-1}{2}}}\,\xi)\right)^{-{2\over p-1}}\,, \end{equation} which for large $|\xi|$ has the asymptotic behaviour \begin{equation} \label{eq6s1} w_p(\xi)=\alpha\,e^{-|\xi|}\,(1+O(e^{-(p-1)|\xi|}))\,, \quad \alpha:=(2p+2)^{1\over p-1}\,. \end{equation} Now we want to solve the equation \eqref{spike1} for all $\delta>0$. To this end we introduce an extra parameter $\nu\leq \varepsilon$ in the non-linear part of \eqref{spike1} defining \begin{equation} \label{modspike0} Q_\nu[\varphi](\xi) :=|\varphi(\xi)|^p \Bigl( \int_{-1/\varepsilon}^{1/\varepsilon} G_\delta (\nu\xi,\nu\eta) |\varphi(\eta)|^rd\eta\Bigr)^{-q} \end{equation} and rewrite \eqref{spike1} in the form \begin{equation} \label{modspike1} \varphi''(\xi)-\varphi(\xi)+Q_\varepsilon[\varphi]=0,\;\;\; \varphi'(\pm 1/\varepsilon)=0. \end{equation} Setting the parameter $\nu$ to zero, we get a simplified equation, that we shall call the corresponding {\it shadow equation} (and which differs from Takagi's \cite{takagi} by a multiplicative constant): $$ \widetilde\varphi''(\xi)-\widetilde\varphi(\xi)+Q_0[\widetilde\varphi]=0,\; \widetilde\varphi'(\pm 1/\varepsilon)=0. $$ The solution is given by \begin{equation} \label{shad2} \widetilde\varphi:=\Big(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \psi_\varepsilon^r(\eta)d\eta\Big)^{-\alpha_r}\psi_\varepsilon, \end{equation} where \[ \psi_\varepsilon''-\psi_\varepsilon+\psi_\varepsilon^p=0,\quad \psi_\varepsilon'(\pm1/\varepsilon)=0. \] This equation has a unique solution with a single spike located in the interior of the domain; its properties are well known \cite{gk1} Section 2.1 and \cite{takagi}. Thus we have constructed a shadow spike solution for any fixed $\delta>0$, which coincides with the shadow solution from \cite{gk1}, \cite{takagi} if $\delta=0$. The main idea is to find a solution of the problem \eqref{modspike1} in a small neighbourhood of our shadow spike solution $\widetilde\varphi$. If $\varphi=\widetilde\varphi+u$ we get an equation for $u$: \begin{equation} \label{int1a} \begin{gathered} u''-u+\{Q_0[{\widetilde \varphi}+u]- Q_0[\widetilde\varphi]\}+\{Q_\varepsilon[\widetilde\varphi+u]- Q_0[\widetilde\varphi+u]\}=0\,,\\ u'(\pm 1/\varepsilon)=0\,. \end{gathered} \end{equation} Using the Taylor formula we can write \begin{equation} \label{int2} \begin{gathered} Q_0[\widetilde\varphi+u]-Q_0[\widetilde\varphi]=Q'_0[\widetilde\varphi]u+f(u)\,,\\ f(u):=\int_0^1\{\partial_\sigma Q_0[\widetilde\varphi+\sigma u]- \partial_\sigma Q_0[\widetilde\varphi+\sigma u]_{\sigma=0}\} d\sigma, \end{gathered} \end{equation} and \begin{equation} \label{int2a}g_\varepsilon(u):= Q_\varepsilon[\widetilde\varphi+u]-Q_0[\widetilde\varphi+u]= \int_0^\varepsilon \partial_\nu Q_\nu[\widetilde\varphi+ u] d\nu, \end{equation} where $Q'_0[\widetilde\varphi]$ is the (non-local) linear operator \begin{align*} Q'_0[\widetilde\varphi] u &:=p{\widetilde\varphi}^{p-1} u \Big(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta {\widetilde\varphi}^r(\eta)d\eta\Big)^{-q}\\ &\quad -rq{\widetilde\varphi}^p \Big(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta {\widetilde\varphi}^r(\eta)d\eta\Big)^{-q-1} \int_{-1/\varepsilon}^{1/\varepsilon} g_\delta {\widetilde\varphi}^{r-1}(\eta)u(\eta)d\eta. \end{align*} The linear part of the operator in equation \eqref{int1a} is given by $\widetilde A $, \begin{align*} \widetilde A u&:=-u''+u-Q_0'[\widetilde\varphi]u\\ &=-u''+u-p\widetilde\varphi^{p-1} \Bigl(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta\widetilde\varphi^r(\eta)d\eta\Bigr)^{-q}u\\ &\quad + rq\widetilde\varphi^p \Bigl(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta\widetilde\varphi^r(\eta)d\eta\Bigr)^{-q-1} \int_{-1/\varepsilon}^{1/\varepsilon} g_\delta\widetilde\varphi^{r-1}(\eta)u(\eta)d\eta, \end{align*} or \begin{equation} \label{shad1a} \widetilde A u=-u''+u-p\psi^{p-1}u+\frac{rq \psi_\varepsilon^p \langle u,\psi_\varepsilon^{r-1}\rangle}{\langle 1,\psi_\varepsilon^{r}\rangle}. \end{equation} It is equal to the operator associated with the shadow equation as in \cite[eqs. (1.3), (2.1), (2.20), (2.21)]{gk1}. In \cite{gk1} it is shown that $\widetilde L u:=-u''+u-p\psi^{p-1}u $ restricted to even functions is invertible in $L^2$. Now we remark that $\widetilde A$ is also invertible if $qr\neq p{-}1$. Indeed, it is sufficient to show that zero is not an eigenvalue of $\widetilde A$. Suppose, on the contrary, that $\widetilde A u=0$. Then, since $\widetilde L \psi_\varepsilon =(1-p)\psi_\varepsilon^p$, we see that $u=c\psi_\varepsilon$ and $\widetilde A \psi_\varepsilon =(1-p+qr)\psi_\varepsilon^p\neq 0$. Thus $\widetilde A^{-1}$, restricted to even functions, is a bounded operator in $L^2$, uniformly w.r.t. $\varepsilon$ (cf. \cite{gk1}). Here $L^2$ is the space of quadratically integrable functions on the interval $(-1/\varepsilon,1/\varepsilon)$. Let $H^2$ be the associated Sobolev space, equipped with the usual norm $\|u\|_2:=\|u''\|+\|u\|$. Since $\|u\|_2 \asymp \|(A+c)u\|$ for some large constant $c>0$, we conclude that $\widetilde A^{-1}$, restricted to even functions, is a bounded operator from $L^2$ to $H^2$, uniformly w.r.t. $\varepsilon$. In this way we reduce the problem \eqref{int1a} to the integral equation \begin{equation} \label{int1} u = M u,\quad \mbox{where}\quad Mu:= {\widetilde A}^{-1}[f(u)+g_\varepsilon(u)]\,, \end{equation} with $f$ and $g$ as defined in \eqref{int2}, \eqref{int2a}. We are going to apply the contraction method in the ball \begin{align*} X_\varepsilon:=\{&u\in H^2(-1/\varepsilon,1/\varepsilon) : \mbox{$u$ is even, }, u'(\pm1/\varepsilon)=0\,,\\ &\|u\|_\omega:=\|u\|_2+\max |u(\xi)|/\omega(\xi)\leq \varrho\}\,, \end{align*} where $0<\varepsilon<\varepsilon_0(\delta)$ and where $\omega(\xi)$ is the weight function \begin{equation} \label{shad2aa} \omega(\xi):=\begin{cases} e^{-(p-1)|\xi|}& \mbox{if } 12\,. \end{cases} \end{equation} Since by \cite[eq. (2.11)]{gk1}, \begin{equation} \label{shad2a} |\psi_\varepsilon(\xi)-w_p(\xi)|\leq c\,e^{-1/\varepsilon},\quad |\xi|\leq 1/\varepsilon, \end{equation} we can find a constant $\xi_\varepsilon=\log(C/\varepsilon)$ (where $C$ is a generic positive constant in the sequel) so that $\psi_\varepsilon>\varepsilon^\kappa$ on $[-\xi_\varepsilon,\xi_\varepsilon]$, where $\kappa$ satisfies $\max(1/2,1/r)<\kappa<1$. Then $$ \widetilde\varphi(\xi)>C\,g_\delta^{-\alpha_r} \varepsilon^\kappa,\quad\mbox{if } |\xi|\le\xi_\varepsilon\,. $$ Therefore, choosing \begin{equation} \label{choice1} \varrho:=C\,g_\delta^{-\alpha_r} \varepsilon^\kappa, \end{equation} we get $$\widetilde\varphi+\sigma u>0 \quad \mbox{on }[-\xi_\varepsilon,\xi_\varepsilon] \mbox{ for any } u\in X_\varepsilon \mbox{ and } 0<\sigma<1. $$ Hence \begin{align*} V&:= \int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\nu\xi,\nu\eta)|\widetilde \varphi(\eta)+\sigma u(\eta)|^r d\eta>\\ &> \int_{-\xi_\varepsilon}^{\xi_\varepsilon} G_\delta(\nu\xi,\nu\eta)(\widetilde \varphi(\eta) +\sigma u(\eta))^r d\eta- \int_{|\xi|>\xi_\varepsilon} G_\delta(\nu\xi,\nu\eta)|\widetilde \varphi(\eta) +\sigma u(\eta)|^r d\eta. \end{align*} Since $\widetilde \varphi +\sigma u>C g_\delta^{-\alpha_r}$ on $[-1,1]$, since $$ \frac{g_\delta}{\cosh^2 \delta} =\frac{\delta}{\mu\sinh 2\delta}\leq G_\delta(x,y) \leq 2g_\delta,\quad \mbox{if } -1\leq x,y\leq 1\,, $$ and since $\xi_\varepsilon>1$ for $\varepsilon<\varepsilon_0(\delta)$, we get $$ V>C g_\delta^{1-r\alpha_r} [1-C_\delta \varepsilon^{\kappa r-1}]\cosh^{-2} \delta >C g_\delta^{1-r\alpha_r} \cosh^{-2} \delta $$ if $\kappa r-1>0$ and if $0<\varepsilon<\varepsilon_0(\delta)$. Therefore, \begin{equation} \label{positiv1} V>C_\delta\quad \mbox{if } 0<\varepsilon<\varepsilon_0(\delta), \end{equation} uniformly w.r.t. $\varepsilon$, where the positive quantities $C_\delta$ and $\varepsilon_0(\delta)$ are equivalent to $1$ w.r.t. $\delta$ on any compact interval $[0,\delta_0]$. All other estimates below will be uniform in the same sense. We shall first prove the estimates for $ u\in X_\varepsilon$, \begin{equation} \label{contr1} \|f(u)\|\leq C_\delta \varrho^\gamma \|u\|_2^2 \end{equation} and \begin{equation} \label{contr2} \|g_\varepsilon(u)\|\leq C_\delta \varepsilon , \end{equation} where here and below $\gamma$ is a generic positive number that depends on $p$ and $r$. To prove \eqref{contr1} we use the definition of $f(u)$ in \eqref{int2} and write it as a sum of five terms $f(u)=\sum _{j=1}^5 f_j(u)\,$, where $$ \begin{aligned} f_1(u):=&\, p u\int_{0}^{1}\langle |\widetilde\varphi+\sigma u|^{r},g_\delta\rangle^{-q} \left[|\widetilde\varphi +\sigma u|^{p-1} \mathop{\rm sign} (\widetilde\varphi+\sigma u)-\widetilde\varphi^{p-1}\right] \, d\sigma\,,\\ f_2(u):=&\, p \widetilde\varphi^{p-1} u \int_{0}^{1}\left[\langle |\widetilde\varphi+ \sigma u|^{r},g_\delta\rangle^{-q}- \langle \widetilde\varphi^r,g_\delta\rangle^{-q}\right]\, d\sigma\,,\\ f_3(u):=& {}-qr\int_{0}^{1} \langle |\widetilde\varphi+\sigma u|^r,g_\delta\rangle^{-q-1} |\widetilde\varphi +\sigma u|^{p}\\ &\quad \quad \quad \quad \quad \times \langle |\widetilde\varphi+\sigma u|^{r-1} \mathop{\rm sign} (\widetilde\varphi+\sigma u)-\widetilde\varphi^{r-1},g_\delta u\rangle \, d\sigma\,, \\ f_4(u):=& {} -qr\int_{0}^{1}\langle |\widetilde\varphi+\sigma u|^r,g_\delta\rangle^{-q-1} [|\widetilde\varphi +\sigma u|^{p}-\widetilde\varphi^p] \langle \widetilde\varphi^{r-1},g_\delta u\rangle \, d\sigma\,,\\ f_5(u):=& {}-qr\int_{0}^{1}\left[\langle |\widetilde\varphi+\sigma u|^r, g_\delta\rangle^{-q-1} -\langle \widetilde\varphi^r,g_\delta\rangle^{-q-1}\right] \langle \widetilde\varphi^{r-1},g_\delta u\rangle \,\widetilde \varphi^p \, d\sigma\,. \end{aligned} $$ Denote the second factor in the integrand of $f_1$ by $$ f_0(u):=|\widetilde\varphi+\sigma u|^{p-1} \mathop{\rm sign}(\widetilde\varphi+ \sigma u) -\widetilde\varphi^{p-1}. $$ For all $\varepsilon>0$, $0\le\sigma\le 1$ and all functions $u$ it satisfies \begin{equation} \label{est0} |f_0(u)|\leq \begin{cases} 2\min\{\widetilde\varphi^{p{-}2}| u|,| u|^{p-1}\}& \text{ if }1 2 \end{cases} \end{equation} This is a consequence of the following inequalities:\\ If $a>0$ and $00$ and $b>1$ then \begin{enumerate} \item $0 \leq (a+x)^b - a^b \leq \begin{cases} 2^bxa^{b-1}&\text{if } 0\le x\le a\,,\\ 2^bx^b&\text{if } x\ge a\,, \end{cases}$ \item $0 \leq a^b-(a-y)^b \leq by a^{b-1}$ if $0\leq y \leq a$ \item $0 \leq a^b + t^b \leq (a+t)^b$ if $t \geq 0$ \end{enumerate} Substituting $y=-x$ and $t=-a{-}x$, $b=p{-}1$, $a=\widetilde\varphi$ and $x=\sigma u$ this proves \eqref{est0}. Restricting this inequality to functions $u\in X_\varepsilon$, which are uniformly bounded by $\varrho\omega$ $\bigl($with $\varrho<1$, cf. \eqref{choice1}$\bigr)$, we find the estimate \begin{equation} \label{est0a} |f_0(u)|\leq C|u|^{\sigma_p},\quad \sigma_p:=\min(1,p-1),\; u\in X_\varepsilon\,, \end{equation} for some $C>0$ not depending on $\delta$ or $\varepsilon$, cf. \eqref{positiv1}. Essentially, the restriction $| u|\le\varrho\omega$ in this inequality is necessary only if $p>2$. Using \eqref{est0a}, \eqref{choice1} and the definitions of $f_j(u)$ we find the following uniform estimates if $|u|\leq \varrho\omega$: \setlength{\arraycolsep}{.2em} \def\vsep#1{{\vrule height #1 depth 0pt width 0pt}} \begin{eqnarray} \label{est1}|f_1(u)|&\leq&\displaystyle C\,|u|^{1+\sigma_p}\leq C_\delta \rho^{1+\sigma_p}\omega^{1+\sigma_p} \leq C\, \rho^{1+\sigma_p}\widetilde\omega, \\ &&\displaystyle\vsep{1.6em} \mbox{where }\widetilde\omega(\xi):=\omega(\xi)\ \mbox{if}\ 1 1, \; \alpha:=(2p+2)^{\frac{1}{p-1}}. $$ Hence it is sufficient to prove the estimate \begin{equation} \label{est5} |\varphi(1/\varepsilon)-\widetilde\varphi(1/\varepsilon)|\leq C_\delta \varepsilon^\gamma e^{-1/\varepsilon},\quad 0<\varepsilon<\varepsilon_0(\delta). \end{equation} We can estimate this difference using the integral equation \eqref{contr9}, where $Q_\nu[\varphi]$ is defined by \eqref{modspike0}. We have (using Taylor's formula, \eqref{positiv1}, \eqref{est2a}), \begin{equation} \label{contr10} Q_\nu[\varphi](\eta)= Q_0[\varphi](\eta)+ (1+|\eta|)\varphi^p(\eta)\,O(\nu),\quad |\eta\|\leq 1/\varepsilon, \end{equation} where $$ Q_0[\varphi](\eta)=\varphi^p(\eta) \Bigl( \int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi^r(\xi)d\xi\Bigr)^{-q}. $$ Using also the estimate $|\varphi(\xi)-\widetilde\varphi(\xi)|\leq \varrho\omega(\xi)$, $0<\varepsilon<\varepsilon_0(\delta)$, we find $$ Q_0[\varphi](\eta)= Q_0[\widetilde\varphi](\eta)+ |\varphi^p(\eta)-\widetilde\varphi^p(\eta)|\,O(1) + \varphi^p(\eta)\varrho\,. $$ Hence \begin{equation} \label{contr11} \varphi(\xi)=\widetilde\varphi(\xi){ +}\frac{1}{\varepsilon} \int_{-1/\varepsilon}^{1/\varepsilon}{\widetilde G}_{1/\varepsilon} (\varepsilon\xi, \varepsilon\eta) \Bigl[|\varphi^{p}(\eta){-}\widetilde\varphi^{p}(\eta)|\,O(1) +(1+|\eta|)\varphi^p(\eta)\,O(\varrho)\Bigr]d\eta\,. \end{equation} To estimate this integral, we need a better estimate of $\varphi$, namely \begin{equation} \label{est8} \varphi(\xi)=O_\delta(e^{-|\xi|}),\quad |\xi|\leq 1/\varepsilon,\; 0<\varepsilon<\varepsilon_0(\delta). \end{equation} Indeed $\varphi$ satisfies the equation $$ \varphi''=q \varphi,\; q=1-\varphi^{p-1} \Bigl(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta)\varphi^r(\eta)d\eta\Bigr)^{-q}. $$ Since $\varphi\leq \widetilde\varphi +\varrho\omega$, it follows that $$q(\xi)=1-O(e^{-\gamma|\xi|}),$$ hence applying the classical asymptotic theory we get \eqref{est8}. On the other hand, to estimate the difference $\varphi^p -\widetilde\varphi^p$, we use the estimates $|\varphi-\widetilde\varphi|\leq \varrho\omega$ and $|\varphi|,\;|\widetilde\varphi|\leq C_\delta e^{-|\eta|}$. Thus $$ |\varphi^p(\eta)-\widetilde\varphi^p(\eta)|\leq C_\delta |\varphi(\eta)-\widetilde\varphi(\eta)| e^{-(p-1)\eta} \leq C_\delta |\varphi(\eta)-\widetilde\varphi(\eta)|^b e^{-(p-b)\eta} $$ if $03/2$.\end{quote} Then \begin{equation} \label{est6} |\varphi^p(\eta)-\widetilde\varphi^p(\eta)|\leq C_\delta \varrho^b \omega^b e^{-(p-b)\eta}. \end{equation} Returning now to \eqref{contr11} we estimate the integral by \eqref{est6}: \begin{align*} \frac{1}{\varepsilon}\int_{-1/\varepsilon}^{1/\varepsilon}&{\widetilde G}_{1/\varepsilon} (1, \varepsilon\eta)|\varphi^p(\eta)-\widetilde\varphi^p(\eta)|d\eta\leq \\ &\leq \frac{C_\delta\varrho^b}{\sinh 2/\varepsilon}\int_{-1/\varepsilon}^{1/\varepsilon} \cosh(1/\varepsilon+\eta) e^{-|\eta|(p-b+b(p-1))} d\eta \leq C_\delta \varrho^b e^{-1/\varepsilon} \end{align*} if $10$\,: $\lambda_{0}(\varepsilon)<\lambda_1(\varepsilon)<\lambda_2(\varepsilon)<\dots$ with corresponding eigenfunctions $\psi_{0}(\cdot,\varepsilon)$, $\psi_1(\cdot,\varepsilon)$, $\psi_2(\cdot,\varepsilon)$, \dots. Its spectrum converges for $\varepsilon\to 0$ (and for all selfadjoint boundary conditions) to the spectrum of $L_{0}$, see e.g.~\cite[ch.~9]{codd}. We shall calculate the rate of convergence. The ``limiting'' operator $L_{0}$ (on the whole real axis) has the continuous spectrum $[ 1,\infty)$ and may have discrete eigenvalues below this interval (see \cite[p.~140]{henry}). Simple calculations show: \begin{equation} \label{eq1s2} \begin{array}{lll} \psi_{\rm o}:=w_p^{p+1\over 2}&L_{\rm o}\,\psi_{\rm o}= -\frac14(p-1)(p+3)\,\psi_{\rm o}\,,\hspace{2em}&p>1\,,\\ \psi_1:=\dot w_p&L_{\rm o}\,\psi_1=0\,&p>1\,,\vsep{1.7em}\\ \psi_2:=w_p^{3-p\over 2}-\frac12{\textstyle{p+3\over p+1}}\,w_p^{p+1\over 2}\hspace{1em}& L_{\rm o}\,\psi_2 =\frac14 (p-1)(5-p)\,\psi_2\,,~~&13$, we substitute $\psi(\xi)=\vartheta({p-1\over 2}\xi) $ in the eigenvalue equation $L_{0}\psi=\lambda\psi$ using the explicit form of $w_p$ from \eqref{eq5s1}. This yields the equation $$ M_p \vartheta:=-\ddot\vartheta-2p(p+1)(p-1)^{-2}\cosh^{-2}(\eta)\vartheta =({\textstyle{2\over p-1}})^2\,(\lambda-1)\vartheta=\mu \vartheta\,. $$ Since the ``potential'' in $M_p$ is an increasing function of $p$, its eigenvalues are increasing functions of $p$ by the minimax theorem \eqref{minmax}. Since $\lambda_2\to1$ if $p\to 3$ from below, the second eigenvalue of $M_p$ tends to zero for $p\nearrow3$ and gets absorbed into the continuous spectrum if $p\ge 3$. So $L_{0}$ has only two eigenvalues below 1 if $p\ge 3$. In order to compute the rate of convergence of the smallest eigenvalues $\lambda_{0}(\varepsilon)$ and $\lambda_1(\varepsilon)$ (and $\lambda_2(\varepsilon)$ if $p<3$) of $L_\varepsilon$, we can use the technique of \cite{pdg} and \cite{gk}. We compute (formally) approximate eigenfunctions and project them onto the true eigenfunctions; the residuals yields estimates for the eigenvalues. Let $\widetilde L_\varepsilon, \widetilde B_\varepsilon$ be the corresponding operators resulting in linearization around the shadow spike solution $\widetilde\varphi_\varepsilon$. (We use the notation $\widetilde\varphi_\varepsilon$ for $\widetilde\varphi$.) More precisely (see \eqref{shad1a}) we have, \[ \widetilde L_\varepsilon\,u:=-\ddot u+u-p\,\widetilde\varphi_\varepsilon^{p-1}\, \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(0,0) \widetilde\varphi_\varepsilon^r(\eta)d\eta\Big)^{-q}u = -\ddot u+u-\psi_\varepsilon^{p-1} u, \] \begin{align*} \widetilde B_\varepsilon v &:=qr \widetilde\varphi_\varepsilon^p \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(0,0) \widetilde\varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1} \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(0,0) \widetilde\varphi_\varepsilon^{r-1}(\eta)v(\eta)d\eta\Big)\\ &= qr\psi_\varepsilon^p \frac{\langle u,\psi_\varepsilon^{r-1}\rangle} {\langle \psi_\varepsilon^r,1\rangle}. \end{align*} Thus the operator $\widetilde A_\varepsilon =\widetilde L_\varepsilon +\widetilde B_\varepsilon$ does not depend on $\delta$ and coincides with the shadow operator from \cite{gk1}. Since $$ L_\varepsilon =\widetilde L_\varepsilon +O\bigl(|\varphi_\varepsilon^{p-1}-\widetilde\varphi_\varepsilon^{p-1}|+ \bigl[\varepsilon(1+|\xi|)+|\langle \varphi_\varepsilon^{r}- \widetilde\varphi_\varepsilon^r,g_\delta\rangle|\bigr] \widetilde \varphi_\varepsilon^{p-1}\bigr)$$ it follows the uniform estimate \begin{equation} \label{oper3} \|L_\varepsilon-\widetilde L_\varepsilon\| =O(\varepsilon^\gamma )\,,\quad 0<\varepsilon<\varepsilon_0(\delta). \end{equation} Here and later on the positive quantity $O(1)$ depends on $\delta$ and is equivalent to $1$ on any compact interval $[0,\delta_0]$. All estimates will be uniform in the same sense. Analogously, \begin{equation} \label{oper3a} \|B_\varepsilon-\widetilde B_\varepsilon\| =O(\varepsilon^\gamma)\,,\quad 0<\varepsilon<\varepsilon_0(\delta). \end{equation} In particular, using the asymptotic behaviour of the eigenvalues of $\widetilde L_\varepsilon$ \cite{gk1}, we find $$ \lambda_0(\varepsilon)=\lambda_0 +O(\varepsilon^\gamma) \,,\quad \lambda_2(\varepsilon)=\mu_0 +O(\varepsilon^\gamma )\,, $$ where $\mu_0:=\lambda_2$ if $p<3$ and $\mu_0:=1$ if $p\geq 3$. To find the asymptotic behaviour of the small eigenvalue $\lambda_1(\varepsilon)$ we shall use $\dot\varphi_\varepsilon$ as an approximate eigenfunction. Differentiating \eqref{spike1}, we get $$ L_\varepsilon \dot\varphi_\varepsilon=-q\varepsilon\varphi_\varepsilon^p \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1} \int_{-1/\varepsilon}^{1/\varepsilon}\partial_x G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta. $$ We evaluate this expression as follows. We have $$ \partial_x G_\delta(\varepsilon\xi,\varepsilon\eta)=\pm \delta^2/2\mu +\delta^2 \varepsilon(|\xi|+|\eta|)\,O(1),\quad \xi\neq \eta, $$ where ``+" corresponds to the region $\xi<\eta$ and "--" corresponds to $\xi>\eta$; \begin{equation} \label{est10} \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1}= \Big(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1}+O(\delta^2 \varepsilon (1+|\xi|))\,. \end{equation} Therefore, $$ L_\varepsilon \dot\varphi_\varepsilon=-\frac{q\,\varepsilon\,\delta^2\varphi_\varepsilon^p }{2\mu}\, \Big(\int_\xi^{1/\varepsilon} \varphi_\varepsilon^r(\eta)d\eta- \int_{-1/\varepsilon}^\xi \varphi_\varepsilon^r(\eta)d\eta\Big) +O(\delta^2 \varepsilon^2(1+|\xi|) \varphi_\varepsilon^p) \,. $$ In particular, $$ \langle L_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon \rangle= -\frac{q\,\varepsilon\,\delta^2}{2\mu\,(p+1)} \int_{-1/\varepsilon}^{1/\varepsilon} \varphi_\varepsilon^{p+r+1}(\eta)\,d\eta\, \Big(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1}+O(\delta^2 \varepsilon^2 ). $$ The asymptotic expansion of $\lambda_1(\varepsilon)$ will be calculated using the same technique as in \cite{pdg} and \cite{gk}. We compute an approximate eigenfunction $w$, $\|w\|=1$ of the operator $L_\varepsilon$ and we show that \begin{equation} \label{ev9} \langle L_\varepsilon w,w\rangle =\nu_\varepsilon(1+O(\widetilde R_\varepsilon))\quad \mbox{and}\quad \|L_\varepsilon w\|^2=O(\widetilde R_\varepsilon\,R_\varepsilon)\,, \end{equation} where $R_\varepsilon=o(1)$ and $\widetilde R_\varepsilon=o(1)$ for $\varepsilon\to 0$. The generalized Fourier expansion of $w$ in the true eigenfunctions $\{\psi_k: k=0,1,\dots\}$ of $L_\varepsilon$ is $$ w=\sum_{k=0}^\infty c_k \psi_k\quad \mbox{with}\quad \sum_{k=0}^\infty| c_k|^2=\|w\|^2=1\,. $$ Since all eigenvalues of $L_\varepsilon$ except $\lambda_1(\varepsilon)$ are uniformly bounded away from $\lambda_1(0)=0$ by a distance $d>0$, we find from \eqref{ev9} $$ 1-|c_1|^2=\sum_{k=0\,,~k\ne1}^\infty |c_k|^2~\le~ d^{-2}\, \sum_{k=0\,,~k\ne1}^\infty \lambda_k^2\,|c_k|^2~\le~ d^{-2}\,\|L_\varepsilon w\|^2=O(\widetilde R_\varepsilon\,R_\varepsilon)\,, $$ implying that $|c_1|^2=1+O(\widetilde R_\varepsilon\,R_\varepsilon)$. The estimate for the inner product in \eqref{ev9} now implies that $$ \langle L_\varepsilon w,w\rangle -\nu_\varepsilon=|c_1|^2\lambda_1(\varepsilon)- \nu_\varepsilon+\sum_{k=0,\; k\ne1}^\infty \lambda_k |c_k|^2 =O\bigl(\widetilde R_\varepsilon(\nu_\varepsilon + R_\varepsilon)\bigr) $$ and hence that $$ \lambda_1(\varepsilon)=\nu_\varepsilon+O\bigl(\widetilde R_\varepsilon(\nu_\varepsilon+ R_\varepsilon )\bigr). $$ Let $\psi_1(\cdot,\varepsilon)$ be the true eigenfunction of $L_\varepsilon$ corresponding to $\lambda_1(\varepsilon)$. We look for an approximate eigenfunction of the form $\psi_1(\cdot,\varepsilon)\approx\dot\varphi_\varepsilon+$ boundary layer corrections. Within the interval $[-1/\varepsilon,1/\varepsilon]$ the tails of $\dot\varphi_\varepsilon$ are exponentially small by \eqref{asympt1} and \eqref{spike1}, \begin{equation} \label{eq4s2} \dot\varphi_\varepsilon({\pm 1\over\varepsilon})=0 \quad \mbox{and}\quad \ddot\varphi_\varepsilon(\pm {1\over\varepsilon})= a e^{-{1/\varepsilon}}(1+O(\varepsilon^\gamma)),\quad 0<\varepsilon<\varepsilon_0(\delta),\;\gamma>0, \end{equation} where $a:=2\alpha(\int_{-\infty}^\infty g_\delta w_p^r(\eta)d\eta)^{-\alpha_r}$. \vrule height 0pt depth 10pt width 0pt. We construct boundary layer terms at both endpoints by standard matched asymptotic expansions. Suitable boundary layer corrections at the right and left endpoints are \begin{equation} \label{eq5as2} \begin{gathered} h(\xi):={}-\ddot\varphi_\varepsilon({\textstyle{1\over\varepsilon}}) \varrho(\varepsilon\xi) \,\exp\big(\xi-{1\over\varepsilon})\big)\,, \\ k(\xi):= \ddot\varphi_\varepsilon({-\textstyle{1\over\varepsilon}})\,\varrho(-\varepsilon\xi) \,\exp\big(-\xi-{1\over\varepsilon}) \big)\,, \\ \widetilde\psi_1 :=\dot\varphi_\varepsilon+h+k \end{gathered} \end{equation} where $\varrho$ is a monotonic $\mathcal{C}^\infty$ cut-off function satisfying $\varrho(x)=1$ if $x\ge 3/4$ and $\varrho(x)=0$ if $x\le 1/2$. From the definition it is clear that $\widetilde\psi_1$ satisfies the boundary conditions at $\xi={\pm 1/\varepsilon}$ and $$ L_\varepsilon h={}-p \,\varphi_\varepsilon^{p-1}h \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta (\varepsilon\xi,\varepsilon\eta)\varphi_\varepsilon^r (\eta)d\eta\Big)^{-q}+\ddot\varphi_\varepsilon({\textstyle{1\over\varepsilon}}) \left(\varepsilon^2\varrho''+2\varepsilon\varrho'\right)\exp(\xi- \textstyle{1\over\varepsilon})\,. $$ For $p>1$ we have \begin{equation} \label{eq6s2} \|L_\varepsilon \widetilde\psi_1\|^2=\delta^4 \varepsilon^2+ O(e^{-(2+\gamma)/\varepsilon})\,, \quad 0<\varepsilon<\varepsilon_0(\delta). \end{equation} Further, $$ \langle L_\varepsilon \widetilde\psi_1,\widetilde\psi_1\rangle = \langle L_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon \rangle - \langle L_\varepsilon (h+k), h+k \rangle - [(\dot h +\dot k)\widetilde\psi_1]_{-1/\varepsilon}^{1/\varepsilon}. $$ We can calculate the last two terms, hence \begin{equation} \label{eq6bs2} \langle L_\varepsilon \widetilde\psi_1,\widetilde\psi_1\rangle = \langle L_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon \rangle - 2[\ddot\varphi_\varepsilon(1/\varepsilon)]^2+e^{-(2+\gamma)/\varepsilon} O(1). \end{equation} On the other hand, $$ \|\dot\varphi_\varepsilon\|^2=\Big(\int_{-\infty}^\infty g_\delta w_p^r(\eta)d\eta\Big)^{-2\alpha_r}\int_{-\infty}^\infty (\dot w_p)^2d\eta\,(1 +\varepsilon^\gamma\, O(1)). $$ Therefore, the above estimates show that \begin{equation} \label{eq7as2} \lambda_1(\varepsilon)=-a(\delta) \varepsilon -a_1 e^{-2/\varepsilon} + (\delta^2 \varepsilon^{1+\gamma} + e^{-(2+\gamma)/\varepsilon})\,O(1),\quad 0<\varepsilon<\varepsilon_0(\delta), \end{equation} where $a(\delta)>0,\; a(\delta)=\delta^2\, O(1)$ and \begin{equation} \label{est12} a_1=8\alpha^2 \Big(\int_{-\infty}^\infty (\dot w_p)^2d\eta\Big)^{-1}. \end{equation} In particular, for any fixed $\delta>0$ we have the asymptotic $$ \lambda_1(\varepsilon)=-a(\delta) \varepsilon + \varepsilon^{1+\gamma} \,O(1),\quad 0<\varepsilon<\varepsilon_0(\delta). $$ Thus the small eigenvalue $\lambda_1(\varepsilon)$ of the differential operator $L_\varepsilon$ is always negative. In contrast, in the next section we shall prove that the small eigenvalue $\lambda_\varepsilon$ of the perturbed operator $A_\varepsilon$ is positive for any fixed $\delta>0$ if $0<\varepsilon<\varepsilon_0(\delta)$. If we allow dependence of $\delta$ on $\varepsilon$, then $\lambda_\varepsilon$ is positive for all $\delta>\delta(\varepsilon)$, where $\delta(\varepsilon)$ is exponentially small w.r.t. $\varepsilon\in (0,\varepsilon_0)$. To prove these facts, we need two type of estimates: for any fixed $\delta>0$ or for all small $\delta$. \section{Perturbation of the spectrum by the non-local term \label{perturb}} In this section we consider how the nonlocal operator $B_\varepsilon$ perturbs the eigenvalues of $L_\varepsilon$. Since $\|B_\varepsilon\|=O(1)$ it follows that the spectrum of $A_\varepsilon$ lies in a strip around the real axis. Hence this is an operator with compact resolvent and according to Kato, p. 237 \cite{kato}, its spectrum consists of eigenvalues with finite multiplicity. Our goal is to find conditions on the parameters $p,q,r$ so that the spectrum of $A_\varepsilon$ lies in the right half-plane. We shall prove that this is true under the same conditions on the parameters $p,q,r$ as in the shadow case, cf. \cite{gk1}. \subsection{ \label{perturb1} Perturbation of the small eigenvalue by the non-local term} In this subsection we consider how the non-local operator $B_\varepsilon$ perturbs the small eigenvalue $\lambda_1 (\varepsilon)$ of $L_\varepsilon$. Both the operators $L_\varepsilon$ and $B_\varepsilon$ are invariant under the change of sign $\xi\mapsto{-}\xi$ and hence leave the subspaces of even and odd functions invariant. Hence in this subsection we can consider the operator $A_\varepsilon=L_\varepsilon +B_\varepsilon$ on the subspace of odd functions only. Then $A_\varepsilon$ is a small perturbation of the selfadjoint operator $L_\varepsilon$. Indeed, since $$ B_\varepsilon = B_{0\varepsilon} +\delta^2 \varepsilon (1+|\xi|)\varphi_\varepsilon^p O(1), $$ where $$ B_{0\varepsilon} v =q\,r\,\varphi_\varepsilon^p \Big(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1} \int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi_\varepsilon^{r-1}(\eta)v(\eta)d\eta\,, $$ and $ B_{0\varepsilon}=0$ on odd functions, it follows $$ \|B_\varepsilon\|=\delta^2 \varepsilon O(1),\quad 0<\varepsilon<\varepsilon_0(\delta). $$ Hence by Kato \cite[p. 364]{kato} the spectrum of $A_\varepsilon$ (on odd functions) consists of one simple small eigenvalue $\lambda_\varepsilon$ (see \eqref{eq7as2}), \begin{equation} \label{est14a} \lambda_\varepsilon =(\delta^2 \varepsilon+e^{-2/\varepsilon})\,O(1), \end{equation} and eigenvalues close to the real axis and in the half plane $Re\lambda>1/2$. Thus the problem is reduced to determine the sign of $Re\lambda_\varepsilon$. To this end we shall find its asymptotic behaviour. This will be done in two steps. In the first step we use the a priori estimate \eqref{est14a} and derive a better estimate for $\lambda_\varepsilon$ (see \eqref{eigenv11a} below). To this end we use the same technique as for the selfadjoint operator $L_\varepsilon$, exploiting the fact that the non-selfadjoint operator $A_\varepsilon$ is a sufficiently small perturbation of $L_\varepsilon$. Let $$ A_\varepsilon \psi_\varepsilon =\lambda_\varepsilon \psi_\varepsilon,\quad \|\psi_\varepsilon\|=1 $$ (the eigenfunction being odd one). As an approximate eigenfunction we use the same function $\widetilde\psi$ as before: $\widetilde\psi=\dot\varphi_\varepsilon+h+k$. Note that this is also an odd function. Let $$ \widetilde\psi=c\psi_\varepsilon +dg,\quad \|g\|=1, \mbox{ with $g$ orthogonal to }\psi_\varepsilon . $$ Then \begin{equation} \label{eig1} \|L_\varepsilon \widetilde\psi\|^2=|c|^2 \|L_\varepsilon \psi_\varepsilon\|^2 +|d|^2 \|L_\varepsilon g\|^2 \end{equation} and $L_\varepsilon \psi_\varepsilon =\lambda_\varepsilon \psi_\varepsilon +\delta^2 \varepsilon \,O(1)$, hence $$\|L_\varepsilon \psi_\varepsilon\| =(\delta^2 \varepsilon+e^{-2/\varepsilon})\,O(1). $$ On the other hand we already know that (see \eqref{eq6s2}), \begin{equation} \label{eig2} \|L_\varepsilon \widetilde\psi\| =(\delta^2 \varepsilon+e^{-(1+\gamma)/\varepsilon})\,O(1),\quad p>1. \end{equation} Now we need the uniform estimate $\|L_\varepsilon g\|\geq C$. Suppose on the contrary that $\|L_\varepsilon g\|=o(1)$ as $\varepsilon \to 0$. Let $L_\varepsilon \omega_1= \lambda_1 \omega_1$, $\|\omega_1\|=1$. If $g=c_1\omega_1+d_1h_1$, $|c_1|^2+|d_1|^2=1$ is the orthogonal decomposition, we find that $g=\omega_1+o(1)$. On the other hand, if $\psi_\varepsilon=c_2\omega_1+d_2h_2$, $|c_2|^2+|d_2|^2=1$ is the orthogonal decomposition of $\psi_\varepsilon$, then since $\|L_\varepsilon \psi_\varepsilon\|=O(\lambda_1)$ we find that $\psi_\varepsilon=\omega_1 +O(\lambda_1)$. Then $\langle g, \psi_\varepsilon \rangle=1+o(1)$, what contradicts orthogonality of $g$ and $\psi_\varepsilon$. Hence \begin{equation} \label{est14} |d| =(\delta^2 \varepsilon+e^{-(1+\gamma)/\varepsilon})\,O(1). \end{equation} Further, $$ \langle A_\varepsilon \widetilde\psi,\widetilde\psi\rangle = |c|^2\lambda_\varepsilon +d{\bar c} \langle A_\varepsilon g, \psi_\varepsilon \rangle + |d|^2 \langle A_\varepsilon g, g \rangle . $$ Since $$ \langle A_\varepsilon g, \psi_\varepsilon \rangle = \langle L_\varepsilon g, \psi_\varepsilon \rangle +\delta^2 \varepsilon \,O(1) $$ it follows $$ \langle A_\varepsilon g, \psi_\varepsilon \rangle = (\delta^2 \varepsilon +e^{-2/\varepsilon})\, O(1). $$ On the other hand, \eqref{eig1}, \eqref{eig2} imply $$ |d|\|A_\varepsilon g\| = (\delta^2 \varepsilon+e^{-(1+\gamma)/\varepsilon})O(1). $$ Therefore, $$ \langle A_\varepsilon \widetilde\psi,\widetilde\psi\rangle = \lambda_\varepsilon \|\widetilde\psi\|^2 +(\delta^4 \varepsilon^2 +e^{-(2+\gamma)/\varepsilon})\,O(1). $$ To evaluate the quadratic form we write $$ \langle A_\varepsilon \widetilde\psi,\widetilde\psi\rangle = \langle A_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon\rangle -[(\dot h+\dot k)\widetilde\psi] _{-1/\varepsilon}^{1/\varepsilon} - \langle L_\varepsilon (h+k), h+k\rangle + \langle B_\varepsilon \dot\varphi_\varepsilon, h+k\rangle . $$ Since $$ \langle B_\varepsilon \dot\varphi_\varepsilon, h+k\rangle =\delta^2 e^{-2/\varepsilon}\, O(1), \quad p>1, $$ we get from above estimates \begin{equation} \label{rel2} \lambda_\varepsilon =\frac{\langle A_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon\rangle} {\|\dot\varphi_\varepsilon\|^2} -\frac{2[\ddot\varphi_\varepsilon(1/\varepsilon)]^2} {\|\dot\varphi_\varepsilon\|^2} + (\delta^4 \varepsilon^2+e^{-(2+\gamma)/\varepsilon}+\delta^2 e^{-2/\varepsilon})\,O(1), \end{equation} where $0<\varepsilon<\varepsilon_0(\delta),\; p>1$, and it remains to evaluate $\langle A_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon\rangle$. We have $$ L_\varepsilon \dot\varphi_\varepsilon={}-q\,\varepsilon\,\varphi_\varepsilon^p\, \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1} \int_{-1/\varepsilon}^{1/\varepsilon}\partial_x G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta $$ and $$ B_\varepsilon \dot\varphi_\varepsilon=q\,\varphi_\varepsilon^p\, \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1} \int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) d\varphi_\varepsilon^r(\eta). $$ Since \begin{align*} &\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) d\varphi_\varepsilon^r(\eta)\,=\\ &\hspace{4em}=\,\big[ G_\delta(\varepsilon\xi,1)- G_\delta(\varepsilon\xi,-1)\big]\varphi_\varepsilon^r(1/\varepsilon) -\varepsilon\int_{-1/\varepsilon}^{1/\varepsilon}\partial_y G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta, \\ &\partial_y G_\delta(\varepsilon\xi,\varepsilon\eta)+ \partial_x G_\delta(\varepsilon\xi,\varepsilon\eta)= \delta c_\delta \sinh\delta\varepsilon(\xi+\eta)\,,\quad \text{ where } c_\delta:=\frac{\delta}{\mu\sinh 2\delta} \\ &G_\delta(\varepsilon\xi,1)-G_\delta(\varepsilon\xi,-1)=\delta^2 \,O(1) \end{align*} we find \begin{align*} A_\varepsilon \dot\varphi_\varepsilon&={}-q\delta c_\delta \varepsilon \varphi_\varepsilon^p \int_{-1/\varepsilon}^{1/\varepsilon} \sinh \delta \varepsilon(\xi+\eta) \varphi_\varepsilon^r(\eta) d\eta \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1} \\ &\quad{} +O\Bigl(\delta^2 \varphi_\varepsilon^r(1/\varepsilon) \varphi_\varepsilon^p\, \Big(\int_{-1/\varepsilon}^{1/\varepsilon} G_\delta(\varepsilon\xi,\varepsilon\eta) \varphi_\varepsilon^r(\eta)d\eta\Big)^{-q-1}\Bigr). \end{align*} We simplify this expression as follows. Since $$ \sinh \delta\varepsilon(\xi+\eta)=\delta\varepsilon(\xi+\eta)+O(\delta^3\varepsilon^3(|\xi|^3+ |\eta|^3)), $$ it follows $$ \int_{-1/\varepsilon}^{1/\varepsilon} \sinh \delta \varepsilon(\xi+\eta) \varphi_\varepsilon^r(\eta) d\eta = \delta\varepsilon\xi\int_{-1/\varepsilon}^{1/\varepsilon} \varphi_\varepsilon^r(\eta) d\eta +O(\delta^3\varepsilon^3(1+|\xi|^3)). $$ Using also \eqref{est10} we get \begin{align*} &\langle A_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon\rangle=\\ &= -q\delta^2 \varepsilon^2 c_\delta\int_{-1/\varepsilon}^{1/\varepsilon}\varphi_\varepsilon^r(\eta)d\eta \int_{-1/\varepsilon}^{1/\varepsilon}\xi\varphi_\varepsilon^p(\eta)d\varphi_\varepsilon \Bigl(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi_\varepsilon^r(\eta)d\eta\Bigr)^{-q-1}+O(\delta^2\varepsilon^3)\\ &=\frac{q}{p+1}\delta^2 \varepsilon^2 c_\delta\int_{-1/\varepsilon}^{1/\varepsilon} \varphi_\varepsilon^r(\eta)d\eta \int_{-1/\varepsilon}^{1/\varepsilon}\varphi_\varepsilon^{p+1}(\eta)d\eta \Bigl(\int_{-1/\varepsilon}^{1/\varepsilon} g_\delta \varphi_\varepsilon^r(\eta)d\eta\Bigr)^{-q-1}+O(\delta^2\varepsilon^3 ). \end{align*} In this expression we can replace as before $\varphi_\varepsilon$ by $\widetilde\varphi_\varepsilon$ and then by $w_p$. As a result we get \begin{equation} \label{eigenv10} \frac{\langle A_\varepsilon \dot\varphi_\varepsilon,\dot\varphi_\varepsilon\rangle} {\|\dot\varphi_\varepsilon\|^2}= \frac{ q\,\delta^2 \varepsilon^2\int_{-\infty}^\infty w_p^{p+1}(\eta)d\eta} {(p+1)\cosh^2 \delta\int_{-\infty}^\infty (\dot w_p)^2d\eta}+ O(\delta^2\varepsilon^{5/2} ). \end{equation} Finally, \eqref{rel2}, \eqref{eigenv10} and \eqref{eq4s2} give \begin{equation} \label{eigenv11} \frac{ q\,\delta^2 \varepsilon^2\int_{-\infty}^\infty w_p^{p+1}(\eta)d\eta} {(p+1)\cosh^2 \delta\int_{-\infty}^\infty (\dot w_p)^2 d\eta}- \frac{8\alpha^2 e^{-2/\varepsilon}} {\int_{-\infty}^\infty (\dot w_p)^2 d\eta}+ (\delta^2\varepsilon^{5/2}+\delta^4\varepsilon^2 +O(e^{-(2+\gamma)/\varepsilon})) \end{equation} if $0<\varepsilon<\varepsilon_0(\delta)$. In particular, \begin{equation} \label{eigenv11a} \lambda_\varepsilon=O(\delta^2 \varepsilon^2 +e^{-2/\varepsilon})\,,\hspace{2em} 0<\varepsilon<\varepsilon_0(\delta). \end{equation} To find the asymptotic behaviour of $\lambda_\varepsilon$, we notice that using \eqref{eigenv11a} we can improve the bound for $d$ (cf. \eqref{est14}): \begin{equation} \label{est14ab} |d|=O(\delta^2\varepsilon^2+e^{-(1+\gamma)/\varepsilon})\,. \end{equation} Indeed, since $A_\varepsilon \widetilde\psi =c\lambda_\varepsilon \psi_\varepsilon+dA_\varepsilon g$ and $A_\varepsilon g=L_\varepsilon g+\delta^2\varepsilon\, O(1)$, hence $\|A_\varepsilon g\|\geq C>0$, it follows $|d|=(\|A_\varepsilon \widetilde\psi\|+\delta^2\varepsilon^2 +e^{-2/\varepsilon})$. Using $\|A_\varepsilon \dot\varphi_\varepsilon\| =\delta^2\varepsilon^2 \,O(1)$ and $\|A_\varepsilon h\|=e^{-(1+\gamma)/\varepsilon} \,O(1)$ we get \eqref{est14ab}. Now, having the better estimate \eqref{est14ab} we can repeat the above arguments and show that instead of \eqref{eigenv11} we have for $ 0<\varepsilon<\varepsilon_0(\delta)$\,: \begin{equation} \label{eigenv11b} \lambda_\varepsilon= \frac{q\delta^2 \varepsilon^2\int_{-\infty}^\infty w_p^{p+1}(\eta)d\eta} {(p+1)\cosh^2 \delta\int_{-\infty}^\infty (\dot w_p)^2 d\eta}- \frac{8\alpha^2 e^{-2/\varepsilon}}{\int_{-\infty}^\infty (\dot w_p)^2 d\eta}+ (\delta^2\varepsilon^{5/2}+e^{-(2+\gamma)/\varepsilon})\,O(1)\,. \end{equation} In particular, for any fixed $\delta>0$ we have the asymptotic $$ \lambda_\varepsilon= \frac{q\delta^2 \varepsilon^2\int_{-\infty}^\infty w_p^{p+1}(\eta)d\eta} {(p+1)\cosh^2 \delta\int_{-\infty}^\infty (\dot w_p)^2 d\eta} +\varepsilon^{5/2} \,O(1),\; 0<\varepsilon<\varepsilon_0(\delta). $$ If $\delta$ is not fixed and we allow $\delta\to 0$ as $\varepsilon\to 0$, then $Re \lambda_\varepsilon$ changes sign around the point $\delta(\varepsilon)$ given as a solution to the equation $$ \frac{q}{p+1}\,\frac{\mu}{8\alpha^2}\, \varepsilon^2 \delta^2 e^{2/\varepsilon}\int_{-\infty} ^\infty w_p^{p+1}(\eta)d\eta=1. $$ Note that the same expression is obtained in \cite{ward2} using formal asymptotic methods. \subsection{ \label{perturb2}Perturbation of the non-small eigenvalues and uniform estimates of the resolvent} According to \eqref{oper3}, \eqref{oper3a} and Kato \cite[p.\,364]{kato}, the eigenvalues of $A_\varepsilon$ lie in a $O(\varepsilon^\gamma)$ neighbourhood of the eigenvalues of the shadow operator $\widetilde A_\varepsilon$. Hence, under the same conditions on the parameters $p,q,r$ as in \cite{gk1}, the spectrum of $A_\varepsilon$ lies in the right-half plane for all $0<\varepsilon<\varepsilon(D)$. In particular, there exists an angle $\chi_D\in (0,\pi/2)$, such that the resolvent set of $A_\varepsilon$ contains the sector $$ \Lambda:=\{\lambda\in \mathbb{C} : \chi_D \leq |\arg (\lambda-\mu_\varepsilon) |\leq \pi \}, $$ where $\mu_\varepsilon=\frac12 \mathop{\rm Re} \lambda_\varepsilon $. Moreover, in this sector the resolvent satisfies for some constant $M_{\varepsilon,D}$, for all $0<\varepsilon<\varepsilon(D)$, the estimate \begin{equation} \label{resest1} \|(A_\varepsilon-\lambda)^{-1}\|\leq \frac{M_{\varepsilon,D}}{|\lambda-\mu_\varepsilon|}\quad \mbox{for all } \lambda\in\Lambda. \end{equation} To prove this estimate, we use the formula \begin{gather*} (A_\varepsilon-\lambda)^{-1}=(1+(L_\varepsilon-\lambda)^{-1}B_\varepsilon)^{-1} (L_\varepsilon-\lambda)^{-1},\\ \|(L_\varepsilon-\lambda)^{-1}\|\leq 1/\mathop{\rm dist}(\lambda,\sigma(L_\varepsilon)). \end{gather*} Since $$ \|(L_\varepsilon-\lambda)^{-1}\|\leq \frac{C}{|\lambda|},\quad \|(L_\varepsilon-\lambda)^{-1}B_\varepsilon\|\leq \frac{1}{2}, $$ uniformly for all $|\lambda|>N_D$, $\lambda\in \Lambda$, $0<\varepsilon<\varepsilon(D)$ for some $N_D$ large enough, it follows $$ \|(A_\varepsilon-\lambda)^{-1}\|\leq \frac{C}{|\lambda-\mu_\varepsilon|}, $$ uniformly for all $|\lambda|>N_D$, $\lambda\in \Lambda$, and $ 0<\varepsilon<\varepsilon(D)$. If $|\lambda|\leq N_D$, $\lambda\in \Lambda$, then $$ \|(\widetilde A_\varepsilon-\lambda)^{-1}\|\leq C_{\varepsilon,D}. $$ \section{Contraction around the steady state $S(x,\varepsilon)$ \label{contrac}} In this section we study stability of the spike solution $S$ of \eqref{simplgm} as given in \eqref{spike0}. We assume that the parameters $(p,q,r,\mu,\varepsilon)$ are such that all eigenvalues of $A_\varepsilon$ are located in the right half plane. %{\sc Norms.} Besides the standard $L^2$-norm for functions on the interval $[-1,1]$ denoted by $\|\cdot\|\,$, we use in this section the ``energy norm'' $\|\cdot\|_1$, which is associated naturally to a problem with a small parameter like \eqref{eq11s1} and is defined by: $ \|u\|_1^{2}:=\|u\|^2+\|\varepsilon\,u'\|^2$. For fixed positive $a$ large enough and uniformly for all $\varepsilon\in(0,\varepsilon_{0}]$ this norm satisfies the equivalences $$ \langle (A+a)u,u\rangle^{1/2} \asymp \|u\|_1. $$ We study perturbations around the steady state spike solution $S$, using the contraction method as in \cite{gk1} The perturbation satisfies equation \eqref{eq13s1}, which reads: $$ v_t + A v =f[v]\,,\quad v(x,0)=v_{0}(x), $$ where the quadratic term $ f$ is given by \eqref{eq14s1} and the (linear) operator $A$ is defined by \eqref{eq15s1}. Obviously, this operator $A$ has the same spectral properties as its (stretched) cousin $A_\varepsilon$ has in sections \ref{reference} and \ref{perturb}. Hence, under the positivity condition, stated above, $A$ is a sectorial operator, see \cite{henry}, it satisfies the estimate \eqref{resest1}. Associated to $A$ is the semigroup $$ e^{-At} :=\frac{1}{2\pi i} \int_{\Gamma} (A-\lambda)^{-1}e^{-\lambda t} d\lambda\,,\quad t>0\,, $$ where $\Gamma$ is a suitable contour in the resolvent set $\Lambda$. As in \cite{gk1} we prove the following statement. \begin{lemma} \label{l2bounds} For all $t>0$, all $\varepsilon\in(0,\varepsilon(D)]$ and for some constant $C_{\varepsilon,D}$, not depending on $t$, this semigroup satisfies: \begin{gather} \label{eq3s4} \|e^{-At}\| \leq C_{\varepsilon,D} e^{-\mu_\varepsilon t}\,,\\ \label{eq4s4} \|e^{-At}\|_1 \leq C_{\varepsilon,D} e^{-\mu_\varepsilon t}\, \begin{cases} \|u\|_1^{}\,,\\ (1+t^{-1/2})\|u\| \end{cases} \end{gather} \end{lemma} Hence we can apply the contraction method as in \cite{henry}, \cite{gk1}, and prove the local stability of the single internal spike solution $S(x,\varepsilon)$ in the Sobolev norm $\|\cdot\|_1$. \begin{theorem}[Local stability of the single internal spike for $0<\varepsilon<\varepsilon(D)$] There exist positive constants $C(D), C_\varepsilon(D)$ and $ \varepsilon(D)$, depending also on $p$, $q$, $r$ and $\mu$, and small $\varrho_\varepsilon(D)$ such that the solution $(U,H)$ of the system \eqref{simplgm} exists for all times $t>0$ and satisfies \begin{gather*} \|U(\cdot,t)-S\|_1 \leq \varrho e^{-\mu_\varepsilon t},\\ \|H(\cdot,t)-H\|_1 \leq C(D)\varrho \varepsilon^{-1}e^{-\mu_\varepsilon t}, \end{gather*} for all $\varepsilon$ and $\varrho$ satisfying \[ 0<\varepsilon<\varepsilon(D),\quad 0<\varrho<\varrho_\varepsilon(D), \] for all initial conditions $U_{0}\in H^1 (-1,1)$ in the vicinity of $S$, that satisfy the compatibility conditions $U_{0}' (-1)=U_{0}' (1)=0$ and satisfy the bound $$ \|U_{0}-S\|_1 \leq C_\varepsilon(D) \varrho. $$ \end{theorem} \subsection*{Acknowledgement} We are grateful to the referee for a number of useful suggestions. The first author (G. E. Karadzhov) was partially supported by the Research Community ``Advanced Numerical Methods for Mathematical Modelling'' of the Fund for Scientific Research - Flanders, and by the University of Sussex at Brighton. \begin{thebibliography}{66} \bibitem{codd} E. A.\ Coddington \& N.\ Levinson, {\it Theory of ordinary differential equations}, McGraw-Hill, New York, 1955. \bibitem{doelman} A.\ Doelman, R. A.\ Gardner \& T. J.\ Kaper, {\it Large stable pulse solutions in reaction diffusion equations}, Indiana Univ. Math. J., 50(1), pp. 443--507, 2001. \bibitem{doelman1} A.\ Doelman, T. J.\ Kaper \& H.Van der Ploeg, {\it Spatially periodic and aperiodic multi-pulse solutions in the one-dimensional Gierer-Meinhardt equation}, Methods and Appl. of Anal., 8, pp. 1--28, 2001. \bibitem{gierer} A.\ Gierer \& H.\ Meinhardt, {\it A theory of biological pattern formation}, Kybernetik, {\bf 12}, pp.\ 30--39, 1972. \bibitem{golub} G. H. Golub \& C. F. Van~Loan, {\it Matrix Computations}, $2^{\rm nd}$~ed., The Johns Hopkins University Press, Baltimore, 1989. \bibitem{pdg} P. P. N. de Groen, {\it The nature of resonance in a singular perturbation problem of turning point type}, SIAM J. Math. Anal., {\bf 11}, pp.\ 1--22, 1980. \bibitem{gk} P. de Groen \& G.E. Karadzhov, {\it Exponentially Slow Travelling Waves on a Finite Interval for Burgers' Type Equation}, Electronic Journal of Differential Equations, Vol. 1998(1998), No. 30, pp. 1--38. \bibitem{gk1} P. de Groen \& G. E. Karadzhov, {\it Metastability in the shadow system for Gierer-Meinhardt's equations}, Electronic Journal of Differential Equations, vol. 2002(2002). \bibitem{henry} D. Henry, {\it Geometric theory of semilinear parabolic equations}, Lecture Notes in Mathematics {\bf 840}, Springer-Verlag, Berlin, 1981. \bibitem{kato} T.\ Kato, {\it Perturbation theory for linear operators}, Springer verlag, New York, 1966. \bibitem{lin} C.-S. Lin, W.-M. Ni \& I. Takagi, {\it Large amplitude stationary solutions to a chemotaxis system}, J. Diff. Eq., {\bf 72} pp. 1--27 (1988). \bibitem{ni} Wei-Ming Ni, {\it Diffusion, cross-diffusion, and their spike layer steady states}, Notices of the AMS, {\bf 45} pp. 9--18 (1998). \bibitem{reed} M. Reed \& B. Simon, {\it Methods of modern mathematical physics \#4, Analysis of operators}, Academic Press, London, 1978. - 396 p. - ISBN 0125850042. \bibitem{takagi} I. Takagi, {\it Point-condensation for a reaction-diffusion system}, J. of Diff. Eq., {\bf 61} pp. 208--249 (1986). \bibitem{turing} A.\ Turing, {\it The chemical basis of morphogenesis}, Phil. Trans. Roy. Soc. London, Series B, {\bf 237} pp.\ 37--72, 1952. \bibitem{ward1} D.\ Iron \& M. J.\ Ward, {\it A metastable spike solution for a non-local reaction-diffusion model}, SIAM J. Applied Mathematics, {\bf 60}, pp.\ 778--802, (2000). \bibitem{ward2} D.\ Iron, M. J.\ Ward \& J. Wei, {\it The stability of spike solutions to the one-dimensional Gierer-Meinhardt model}, Phys. D, 150 (2001), pp. 2562. \bibitem{ward3} M.J.\ Ward \& J. Wei, {\it Hopf bifurcation and oscillatory instabilities of spike solutions for the one-dimensional Gierer-Meinhardt model}, J. Nonlin. Sci. {\bf 14}, pp. 671--711 (2003). \bibitem{ward4} M. J. Ward \& J. Wei, {\it Hopf bifurcation and oscillation for the shadow Gierer-Meinhardt system}, European J. of Applied Math. {\bf 14} (2003). \bibitem{wei} J. Wei, {\it On single interior spike solutions of Gierer-Meinhardt system: uniqueness and spectrum estimates}, Europeam J. of Applied Math. {\bf 10}, pp 353--378 (1999). \bibitem{ww} J. Wei \& M. Winter, {\it Existence and stability analysis of multiple-peaked solutions}, preprint: www.math.cuhk.edu.hk/$\sim$wei \end{thebibliography} \end{document}