\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2006(2006), No. 09, pp. 1--6.\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 2006 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2006/??\hfil Nonexistence of solutions] {Nonexistence of solutions to KPP-type equations of dimension greater than or equal to one} \author[J. Engl\"{a}nder, P. L. Simon\hfil EJDE-2006/??\hfilneg] {J\'{a}nos Engl\"{a}nder, P\'{e}ter L. Simon} % in alphabetical order \address{J\'{a}nos Engl\"{a}nder \hfill\break Department of Statistics and Applied Probability\\ University of California, Santa Barbara\\ CA 93106-3110, USA} \email{englander@pstat.ucsb.edu} \urladdr{http://www.pstat.ucsb.edu/faculty/englander} \address{P\'{e}ter L. Simon \hfill\break Department of Applied Analysis, E\"otv\"os Lor\'and University\\ P\'{a}zm\'{a}ny P\'{e}ter S\'{e}t\'{a}ny 1/C, H-1117 Budapest, Hungary} \email{simonp@cs.elte.hu} \urladdr{http://www.cs.elte.hu/$\sim$simonp} \date{} \thanks{Submitted September 19, 2005. Published January 24, 2006.} \subjclass[2000]{35J60, 35J65, 60J80} \keywords{KPP-equation; semilinear elliptic equations; \hfill\break\indent positive bounded solutions; branching Brownian-motion.} \begin{abstract} In this article, we consider a semilinear elliptic equations of the form $\Delta u+f(u)=0$, where $f$ is a concave function. We prove for arbitrary dimensions that there is no solution bounded in $(0,1)$. The significance of this result in probability theory is also discussed. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \section{Introduction and statement of main result} In this article, we study semilinear elliptic equations of the form $\Delta u+f(u)=0$. On the nonlinear term $f:[0,1] \to \mathbb{R}$ we assume that \begin{itemize} \item[(i)] $f$ is continuous, \item[(ii)] $f$ is positive on $(0,1)$, \item[(iii)] the mapping $z\mapsto f(z)/z$ is strictly decreasing. \end{itemize} Under these three conditions, we consider the Kolmogorov Petrovskii Piscunov-type (KPP-type) equation \begin{gather} \Delta u+f(u)=0 \label{eqn} \\ 00$. (In fact this particular nonlinearity is intimately related to the distribution of a \emph{branching Brownian motion}; see more on the subject in the next paragraph.) We will present a proof for our result that works basically for concave functions; in fact, (iii) of Assumption 1 is related to the concaveness of the function. The connection between the KPP equation and branching Brownian motion has already been discovered by H. P. McKean --- it first appeared in the classic work \cite{McK1975,McK1976}. Let $Z=(Z(t))_{t\geq 0}$ be the $d$-dimensional binary branching Brownian motion with a spatially and temporally constant branching rate $\beta>0$. The informal description of this process is as follows. A single particle starts at the origin, performs a Brownian motion on $\mathbb{R}^d$, after a mean--$1/\beta$ exponential time dies and produces two offspring, the two offspring perform independent Brownian motions from their birth location, die and produce two offspring after independent mean--$1/\beta$ exponential times, etc. Think of $Z(t)$ as the subset of $\mathbb{R}^d$ indicating the locations of the particles $z_1^t,\dots ,z^{N_t}_t$ alive at time $t$ (where $N_t$ denote the number of particles at $t$). Write $P_{x}$ to denote the law of $Z$ when the initial particle starts at $x$. The natural filtration is denoted by $\{\mathcal {F}_t,\; t\ge 0\}$. Then, as is well known (see e.g. \cite[Chapter 1]{D02}), the law of the process can be described via its Laplace functional as follows. If $f$ is a positive measurable function, then \begin{equation}\label{Laplace.func} E_x\exp \Big(-\sum_{i=1}^{N_t} f(z_i^t)\Big)=1-u(x,t), \end{equation} where $u$ solves the initial value problem \begin{equation} \label{par.eqn} \begin{gathered} \dot{u}=\frac{1}{2}\Delta u+f(u) \quad \text{in } \mathbb{R}^d\times\mathbb{R}_+ \\ u(\cdot,0)=1-e^{-f(\cdot)} \quad \text{in }\mathbb{R}^d \\ %\label{IC}\\ 0\le u\le 1 \quad \text{in } \mathbb{R}^d\times\mathbb{R}_+,%\label{same.cond} \end{gathered} \end{equation} with $f$ of the form \eqref{specialNL}. Equation \eqref{eqn}-\eqref{cond} appears when one studies certain `natural' martingales associated with branching Brownian motion (see e.g. \cite{EK}). To understand this, let $\mathcal{\widehat F}_t:=\sigma(\bigcup_{s\ge t}\mathcal{F}_s)$ and consider the tail $\sigma$-algebra $\mathcal{\widehat F}_{\infty}:=\bigcap_{t\ge 0}\mathcal{\widehat F}_s$. Choosing appropriate (sequences of) $f$'s one can then express the probabilities of various events $A_t\in \mathcal{\widehat F}_t$, for $t>0$, in terms of the function $u$ in \eqref{par.eqn}. Letting $t\to\infty$ then leads to the conclusion that if $A\in \mathcal{\widehat F}_{\infty}$ denotes a certain tail event (e.g. having strictly positive limit for a certain nonnegative `natural' martingale, or local/global extinction) then the function $u(x):=P_x(A)$ is either constant ($=0$ or $=1$), or it must solve \eqref{eqn}-\eqref{cond}. Hence, it immediately follows from our main theorem that \emph{the tail $\sigma$-algebra is trivial}, that is, all those events $A$ satisfy $P_{\cdot}(A)\equiv 0$ or $P_{\cdot}(A)\equiv1$. Note that if $\beta>0$ is replaced by a smooth nonnegative function $\beta(\cdot)$ that does not vanish everywhere, then this corresponds to having \emph{spatially dependent} branching rate for the branching Brownian motion. It would be desirable therefore to investigate whether our main theorem can be generalized for such $\beta$'s. \section{Proof of the theorem} The proof is based on two ideas: The application of the semilinear elliptic maximum principle, which is generalized here fore concave functions, and a comparison between the semilinear and the linear problems. Using these two ideas we will show that the \emph{minimal positive solution} of \eqref{eqn} is $u_{\min}\equiv 1$, hence \eqref{eqn} has no solution satisfying \eqref{cond}. First we state and prove a semilinear maximum principle. The results in this form is a generalization of \cite[Proposition 7.1]{EP99} for the particular case when the elliptic operator is $L=\Delta$. \begin{lemma}[Semilinear elliptic maximum principle]\label{emp} Let $f:[0,\infty ) \to \mathbb{R}$ be a continuous function, for which the mapping $z\mapsto f(z)/z$ is strictly decreasing. Let $D\subset \mathbb{R}^{d}$ be a bounded domain with smooth boundary. If $v_{i}\in C^2(D)\cap C(\bar D)$\ satisfy $v_{i}>0$ in $D$, $\Delta v_{i}+f( v_{i})=0$, in $D$ for $i=1,2$, and $v_{1}\ge v_{2}$ on $\partial D$, then $v_{1}\ge v_{2}$ in $\bar D$. \end{lemma} \begin{proof} Note that the function $w:=v_1-v_2$ satisfies \begin{equation} \Delta w + f(v_1)-f(v_2)=0 . \label{e3} \end{equation} We show that $w\ge 0$ in $D$. Suppose to the contrary that there exists a point $y\in D$ where $w$ is negative. Let $\Omega_0:=\{x\in D\mid w(x)<0\}$. Let $\Omega$ be the connected component of $\Omega_0$ containing $y$. Since $w \geq 0$ on $\partial D$, one has $\Omega \subset \subset D$ and \begin{equation} w<0 \quad \mbox{in } \Omega, \quad w=0 \quad \mbox{on } \partial \Omega . \label{e4} \end{equation} Let us multiply the equation $\Delta v_{1}+f( v_{1})=0$ by $w$ and equation \eqref{e3} by $v_1$, then subtract the second equation from the first, and integrate on $\Omega$. Using that $w=v_1-v_2$ one obtains \begin{equation} I+II:=\int_{\Omega} (w \Delta v_1 - v_1 \Delta w) + \int_{\Omega} (v_1 f(v_2) -v_2 f(v_1)) =0. \label{5} \end{equation} Using Green's second identity and that $w=0$ on $\partial \Omega$, along with the fact that $\partial_{\nu} w \geq 0$ on $\partial \Omega$, we obtain $$ I= -\int_{\partial \Omega} v_1 \partial_{\nu} w \leq 0, $$ where $\nu$ denotes the unit outward normal to $\partial \Omega$. Furthermore, since $v_10$ (this is automatically satisfied under assumption that the mapping $z\mapsto f(z)/z$ is strictly decreasing). Then for any $y\in \mathbb{R}^d$ and $p\in (0,1)$ there exists a ball $\Omega:= B_R(y)$ (with some $R>0$) and a radially symmetric $C^2$ function $v:\Omega \to \mathbb{R}$ such that \begin{gather*} \Delta v + f(v)= 0 \\ v>0 \quad \mbox{in } \Omega\\ v=0 \quad \mbox{on } \partial\Omega\,\quad v(y)=p \,. \end{gather*} \end{lemma} \begin{proof} We show the existence of a radially symmetric solution of the form $v(x)=V(|x-y|)$. Let $V\in C^2([0,\infty))$ be the solution of the initial value problem \begin{equation} \begin{gathered} (r^{d-1}V'(r))'+r^{d-1}f(V(r))=0 \\ V(0)=p, \ V'(0)=0 . \end{gathered}\label{e6} \end{equation} Writing $\Delta$ in polar coordinates, one sees that it is sufficient to prove that there exists an $R>0$ such that $V(R)=0$ and $V(r)>0$ for all $r\in [0,R)$. To this end, consider the \emph{linear} initial value problem \begin{equation} \label{e7} \begin{gathered} (r^{d-1}W'(r))'+r^{d-1}m W(r)=0 \\ W(0)=p, \ W'(0)=0 , \end{gathered} \end{equation} where $m>0$ is chosen so that $f(u)>mu$ holds for all $u\in (0,p)$. (Our assumptions on $f$ guarantee the existence of such an $m$.) It is known that $W$ has a first root, which we denote by $\rho$. Note that in this case $-m$ is the first eigenvalue of the Laplacian on the ball $B_{\rho}$. We now show that $V$ has a root in $(0,\rho]$. In order to do so let us multiply \eqref{e7} by $V$ and \eqref{e6} by $W$, then subtract one equation from the other, and finally, integrate on $[0,\rho]$. We obtain \begin{equation} \begin{aligned} I+II&:=\int_0^{\rho} [(r^{d-1}W'(r))'V(r) - (r^{d-1}V'(r))'W(r)]\, \mbox{d}r \\ &\quad +\int_0^{\rho} r^{d-1}[mW(r)V(r) - W(r)f(V(r))]\, \mbox{d}r = 0\,. \end{aligned}\label{e8} \end{equation} Suppose now that $V$ has no root in $(0,\rho]$. Then, integrating by parts, $ I = \rho^{d-1}W'(\rho)V(\rho)<0. $ Next, observe that by integrating \eqref{e6}, one gets $V'(r)<0$ (i.e. $V$ is decreasing). Hence $V(r)