\documentclass[twoside]{article} \usepackage{amssymb} % used for R in Real numbers \pagestyle{myheadings} \markboth{ A singular boundary-value problem }% { Robert M. Houck \& Stephen B. Robinson } \begin{document} \setcounter{page}{75} \title{\vspace{-1in}\parbox{\linewidth}{\footnotesize\noindent Fourth Mississippi State Conference on Differential Equations and\newline Computational Simulations, Electronic Journal of Differential Equations, \newline Conference 03, 1999, pp 75--90. \newline http://ejde.math.swt.edu or http://ejde.math.unt.edu \newline ftp ejde.math.swt.edu or ejde.math.unt.edu (login: ftp)} \vspace{\bigskipamount} \\ % A singular nonlinear boundary-value problem \thanks{ {\em Mathematics Subject Classifications:} 34B15. \hfil\break\indent {\em Key words:} singular nonlinear boundary value problems, existence and uniqueness, \hfil\break\indent asymptotic estimates. \hfil\break\indent \copyright 2000 Southwest Texas State University and University of North Texas. \hfil\break\indent Published July 10, 2000. } } \date{} \author{ Robert M. Houck \& Stephen B. Robinson } \maketitle \begin{abstract} In this paper we prove an existence and uniqueness theorem for the singular nonlinear boundary-value problem $$\displaylines{ (|y'(t)|^py'(t))'+\frac{\phi}{y^{\lambda}(t)}=0 \mbox{ in } (0,1),\cr y(0)=0=y(1), }$$ where $p\geq 0$, $\lambda>0$ and $\phi$ is a positive function in $L^1_{{\rm loc}}(0,1)$. Moreover, we derive asymptotic estimates describing the behavior of the solution and its derivative at the boundary. \end{abstract} \newtheorem{lemma}{Lemma} \newtheorem{theorem}{Theorem} \newtheorem{corollary}{Corollary} \section{Introduction} In this paper we study the singular nonlinear boundary value problem \begin{equation}\begin{array}{c} \displaystyle (|y'(t)|^py'(t))'+\frac{\phi}{y^{\lambda}}=0 \quad \mbox{in } (0,1), \\[7pt] \displaystyle y(0)=y(1)=0\,, \end{array} \label{bvp} \end{equation} where we assume throughout that $p\geq 0$, $\lambda>0$ and $\phi$ is a positive function in $L^1_{{\rm loc}}(0,1)$. Boundary value problems such as (\ref{bvp}) occur in a wide variety of applications. For example, see \cite{CF}, \cite{CN} and \cite{S} for applications to fluid dynamics. The primary motivation for our work comes from \cite{T}, in which Taliaferro studied (\ref{bvp}) for the case where $p=0$ and $\phi\in C(0,1)$. Taliaferro showed that (\ref{bvp}) has a smooth positive solution iff \[ \int_0^1 t(1-t)\phi(t)\, dt<\infty,\] and that the given solution is unique. Taliaferro also showed that $ \lim_{t\to 0^+}y'(t)$ exists and is finite iff \[ \int_0^{1/2}\frac{\phi(t)}{t^{\lambda}}\, dt<\infty, \] where a similar condition applies at $t=1$. Moreover, Taliaferro derived asymptotic formulae describing the behavior of the solution near the boundary for both the finite and infinite slope cases. Generalizations of Taliaferro's paper include Gatica, Oliker, and Waltman \cite{GOW}, Gatica, Hernandez, and Waltman \cite{GHW}, Baxley \cite{B}, and Baxley and Martin \cite{BM}. The papers \cite{GOW}, \cite{GHW} and \cite{B} study the case $p=0$, and the work in \cite{BM} is related to the case $-2\leq p<0$. In this paper we generalize Taliaferro's work to the case $p\geq 0$. This generalization allows us to study fluids with velocity dependent diffusion, and it forces us to confront the inevitable difficulties that follow when a well understood linear differential operator is replaced by a nonlinear differential operator. One example of the differences between the cases $p=0$ and $p>0$ is that in the latter case the problem has an additional singularity where $y'(t)=0$. Moreover, this singularity cannot be avoided, because a positive solution of (\ref{bvp}) must be concave down and must be $0$ at the boundaries, and therefore must have a unique critical point where it achieves a maximum. To get a different point of view on this singularity assume that $y$ is twice differentiable and rewrite (\ref{bvp}) as $$\displaylines{ y''(t)+\frac{\phi}{(p+1)|y'|^p y^{\lambda}}=0 \quad \mbox{ in } (0,1), \cr y(0)=0=y(1), }$$ which is similar in form to the problem studied in \cite{BM}. From this representation of the problem it is clear that a solution cannot be twice differentiable at its critical point. In Section 2 we prove that (\ref{bvp}) has a positive solution iff $\phi\in L^1_{{\rm loc}}(0,1)$ such that \[ \int_0^1 \big(\int_{t}^{1/2}\phi \, ds \big)^{\frac{1}{p+1}}\,dt<\infty, \] and that this solution is unique. (This is equivalent to Taliaferro's condition when $p=0 $.) The argument begins by examining initial value problems that start in the interior of the interval and then ``shoot" towards the boundary. Of particular interest are the solutions whose initial slopes are $0$. Thus we shoot from the singularity mentioned above, and it is important to check that suitable existence, uniqueness, comparison, and continuous dependence theorems are still available. We prove these lemmas and then show that for each $T\in (0,1)$ there exist left and right {\it half solutions} $y_0$ and $y_1$ defined on $[0,T]$ and $[T,1]$, respectively, such that $y_0(0)=0=y_1(1)$ and $y_0'(T)=0=y_1'(T)$. Finally, we show that the parameter $T$ can be adjusted until $y_0(T)=y_1(T)$, thus bringing the half solutions together to create the unique solution to (\ref{bvp}). In Section 3 we examine the boundary behavior of the solution. We prove that the solution has finite slope at $t=0$ iff \[ \int_0^{1/2}\frac{\phi(t)}{t^{\lambda}}\, dt<\infty, \] just as in \cite{T}, where a similar condition applies at $t=1$. For the finite slope case we quickly obtain the asymptotic formulae \[ y'(t) = \left ( \frac{1}{A}\right )^{\frac{\lambda}{p+1}}\Big( \int _t^T\frac{\phi}{s^{\lambda}}(1+o(1))\, ds \Big)^{\frac{1}{p+1}}, \] and \[ y(t)=\left ( \frac{1}{A}\right )^{\frac{\lambda}{p+1}}\int _0^t \Big( \int _t^T\frac{\phi}{s^{\lambda}}(1+o(1))\, ds \Big)^{\frac{1}{p+1}}, \] which are similar to those in \cite{T}. For the remainder of Section 3 we concentrate on the infinite slope case. Using comparison arguments we prove that if $\psi$ is positive and locally integrable such that $ \lim_{t\to 0^+}\frac{\psi (t)}{\phi (t)}=1$ and if $z(t)$ is any solution of \begin{equation} \begin{array}{c} \displaystyle (|z'(t)|^pz'(t))'+\frac{\psi (t)}{z^{\lambda}(t)}=0 \mbox{ in } (0,\delta), \\[7pt] z(0)=0, \end{array} \label{compbvp} \end{equation} for some $\delta>0$, then $ \lim_{t\to 0^+}\frac{z'(t)}{y'(t)}=\lim_{t\to 0^+}\frac{z(t)}{y(t)}=1$. We apply this general result to the special case where $\phi$ behaves like a power of $t$. In this case we see that if $\phi$ is asymptotically comparable to $ct^r$, then $y(t)$ is asymptotically comparable to $\gamma t^{\rho}$ for appropriate $\gamma$ and $\rho$ . These results complement those in \cite{T} and \cite{BM}. \section{Existence and Uniqueness} In this section we prove \begin{theorem} The boundary-value problem (\ref{bvp}) has a positive solution iff \[ \int_0^1 \Big(\int_{t}^{1/2}\phi \, ds \Big)^{\frac{1}{p+1}}\, dt<\infty\,. \] Moreover, this solution is unique. \label{bvpex} \end{theorem} Let $T\in (0,1)$ and consider the initial value problem \begin{equation} \begin{array}{c} \displaystyle (|y'(t)|^py'(t))'+\frac{\phi}{y^{\lambda}}=0 , \\[7pt] y(T)=h,\, y'(T)=k, \end{array} \label{ivp} \end{equation} where $h>0$ and $k\geq 0$. We begin this section by investigating the positive solutions of (\ref{ivp}) on intervals whose right endpoint is $T$. Before proceeding it is helpful to rewrite (\ref{ivp}) as an equivalent integral equation. Assuming that $y$ is a smooth positive solution of (\ref{ivp}) on the interval $(a,T]$, we integrate once to get \[ |y'(t)|^py'(t)=k^{p+1}+\int_t^T\frac{\phi(s)}{y^{\lambda}(s)}\, ds,\, t\in (a,T]. \] Clearly, $y'$ is nonnegative, so \[ y'(t)=\Big( k^{p+1}+\int_t^T\frac{\phi}{y^{\lambda}}\, ds \Big)^{\frac{1}{p+1}},\, t\in (a,T]. \] A second integration yields the equation \begin{equation} y(t)=h-\int_t^T\Big(k^{p+1}+\int_{\tau}^T\frac{\phi}{y^{\lambda}}\, ds\Big)^{\frac{1}{p+1}}\, d\tau,\, t\in (a,T]. \label{inteq} \end{equation} Our proofs will be based upon this representation of the problem. Several straightforward comments regarding (\ref{inteq}) are collected in the following lemma. \begin{lemma} Suppose that $y\in C(a,T]$ is a positive function satisfying (\ref{inteq}). Then $y$ is increasing and concave down, $y\in C[a,T]\bigcap C^1(a,T]$, $|y'|^py'=(y')^{p+1}$ is differentiable a.e., and $y$ satisfies (\ref{ivp}) . \label{lem1} \end{lemma} Thus for any positive $y\in C(a,T]$ we define \[ Fy(t)= h-\int_t^T\Big(k^{p+1}+\int_{\tau}^T\frac{\phi}{y^{\lambda}}\, ds\Big)^{\frac{1}{p+1}}\, d\tau, \] and it is clear that positive solutions of (\ref{ivp}) correspond to fixed points of $F$. \begin{theorem} Suppose that $T>0$, $h>0$, and $k\geq 0$. Then there exists an $a \in [0,T)$ and a unique positive $y\in C[a,T]\bigcup C^1(a,T]$ such that $y$ is a solution of (\ref{ivp}) on $[a,T]$. Moreover we may assume that the interval $[a,T]$ is maximal in the sense that either $a=0$ or $y(a)=0$. \end{theorem} \paragraph{Proof:} We begin by proving existence and uniqueness over some interval $[a_0,T]$. Our primary tool will be the Contraction Mapping Theorem. The estimates below will be useful in later arguments, so they are proved in slightly greater generality than is needed for this theorem. Suppose that $y\in C[a_0,T]$ such that $00$ and $y(a)>0$. We know that $y'(t)$ is decreasing and nonnegative. Also \[ 0\leq y'(t)=\Big( k^{p+1}+\int_t^T\frac{\phi}{y^{\lambda}}\, ds\Big)^{\frac{1}{p+1}}\leq \Big( k^{p+1}+\big(\frac{1}{y(a)} \big)^\lambda \int_{a}^T\phi \, ds\Big)^{\frac{1}{p+1}} <\infty, \] because $\phi\in L^1_{{\rm loc}}(0,1)$. Therefore $ \lim_{t\to a^+}y'(t)$ exists and $y\in C^1[a,T]$. Using $h=y(a)>0$ and $k=y'(a)\geq 0$, we apply the local existence result above to extend $y$ to a unique positive solution of (\ref{ivp}) over some interval $[a-\delta,T]$, contradicting the assumption that $[a,T]$ is maximal. The theorem is proved. \vspace{.1in} Observe that for the case $k>0$ Theorem 1 follows from the standard existence and uniqueness theory for ordinary differential equations, because we can work on an interval where potential solutions are bounded away from the singularities. It was our preference to treat the case $k\geq 0$ as a whole. For the remainder of this section we focus primarily on the case $k=0$, and invite the interested reader to generalize. Before continuing we introduce some useful notation. We refer to the initial value problem (\ref{ivp}) with the additional restriction $k=0$ as (\ref{ivp})$_0$. Given $T\in(0,1)$, let $y_h$ denote the positive solution of (\ref{ivp})$_0$ satisfying $y(T)=h$. We define \[ \begin{array}{l} \displaystyle {\mathcal H}_T^-:=\{h>0: y_h \mbox{ has maximal interval } [0,T]\}, \\ \displaystyle h_T^-:=\inf {\mathcal H}_T^- \end{array} \] The $-$ sign is a reminder that we are currently working to the left of $T$. \begin{lemma} If $\int_0^T\big( \int_s^T \phi \, ds\big)^{\frac{1}{p+1}} \,dt =\infty$ then (\ref{ivp})$_0$ has no positive solution whose maximal interval is $[0,T]$, i.e. ${\mathcal H}_T^-$ is empty. \label{neccond} \end{lemma} \paragraph{Proof:} If $y\in C[0,T]$ is such a solution, then \begin{equation} \begin{array}{lll} \displaystyle y & = & \displaystyle h-\int_t^T \Big( \int_{\tau}^T \frac{\phi}{y^{\lambda}}\, ds\Big)^{\frac{1}{p+1}}\, d\tau \\ \displaystyle &\leq & \displaystyle h-\left (\frac{1}{h}\right ) ^{\frac{\lambda}{p+1}}\int_t^T\Big( \int_s^T \phi \, ds\Big) ^{\frac{1}{p+1}} \, d\tau, \end{array} \label{lem2ineq} \end{equation} because $y\leq h$ on $(0,T]$. It follows that $ \lim_{t\to 0^+}y(t)=-\infty$, a contradiction. \bigskip An immediate consequence of this lemma is that $\int_0^T\left ( \int_s^T \phi \, ds\right )^{\frac{1}{p+1}} \, dt <\infty$ is a necessary condition for the existence of a solution to (\ref{bvp}). Similarly, $\int_T^1\left ( \int_s^T \phi \, ds\right )^{\frac{1}{p+1}} \, dt <\infty$ is necessary. Hence \[ \int_0^1\Big( \int_s^T \phi \, ds\Big)^{\frac{1}{p+1}} \, dt <\infty \] is necessary. Since $\phi$ is locally integrable is suffices to write this condition using $T=1/2$. \begin{lemma} Let $K:=\int_0^T\left ( \int_s^T \phi \, ds\right )^{\frac{1}{p+1}} \, dt$. If $K<\infty$ then ${\mathcal H}_T^-$ is nonempty, and is bounded below by $K^{\frac{p+1}{p+1+\lambda}}$ \end{lemma} \paragraph{Proof:} Let $y$ be a solution of (\ref{ivp})$_0$. Apply inequality (\ref{est1}) with $m=\frac{h}{2}$, $k=0$, and $Fy=y$ to get \[ 0\leq h-y \leq \left (\frac{2}{h}\right )^{\frac{\lambda}{p+1}}K. \] For large $h$ we have $0\leq h-y\leq \frac{h}{4}$, so $y\geq \frac{3h}{4}>0$. Thus if $y\geq\frac{h}{2}$ on an interval, then $y\geq\frac{3h}{4}$ on that interval. Hence, for large $h$, the solution remains above $\frac{3h}{4}$ no matter how far it is extended. Therefore the solution must have maximal interval $[0,T]$. Substitute $t=0$ into inequality (\ref{lem2ineq}) to get \[ 0\leq y(0)\leq h-\left (\frac{1}{h}\right )^{\frac{\lambda}{p+1}}K, \] for $h\in{\mathcal H}_T^-$. This simplifies to \[ h\geq K^{\frac{p+1}{p+1+\lambda}} .\] \vspace{.1in} Before characterizing ${\mathcal H}_T^-$ and $h_T^-$ further we need a comparison lemma. \begin{lemma} Let $T\in (0,1)$ and any let $y_1,y_2\in C[a_0,T]$ be solutions of (\ref{ivp}) with $0h_T^-$ then $h\in {\mathcal H}_T^-$, and if $h0$. Suppose that $h_1\in {\mathcal H}_T^-$ and $h_2>h_1$. Let $y_{h_1}$ and $y_{h_2}$ represent the corresponding solutions of (\ref{ivp})$_0$. Then Lemma \ref{comparison} shows that $y_{h_2}>y_{h_1}$ on any common interval of definition, and therefore $h_2\in {\mathcal H}_T^-$. The lemma follows. \vspace{.1in} Given Lemma \ref{comparison} and its consequences, it is reasonable to guess that the positive solution of (\ref{ivp})$_0$ satisfying $y(T)=h_T^-$ will have maximal interval $[0,T]$ and will satisfy $y(0)=0$. This is indeed the case as the following lemmas will show. \begin{lemma} Let $T\in (0,1)$ and assume that $\int_0^T\left ( \int_s^T \phi \, ds\right )^{\frac{1}{p+1}} \, dt<\infty$. Let $y_1,y_2\in C[a_0,T]$ be solutions of (\ref{ivp})$_0$ and $00$ there is a $\delta >0$ such that if $|y_1(T)-y_2(T)|<\delta$, then $|y_1(t)-y_2(t)|<\epsilon$ for all $t\in [a_0,T]$. \label{condep} \end{lemma} \paragraph{Proof:} The proof will follow from an extension of estimate (\ref{est2}) to this special case. Notice that we can substitute $y'$ for $(Fy)'$. Also, Lemma \ref{comparison} shows that $y_2(t)-y_1(t)$ is positive and decreasing, so $\sup_{[t,T]}|y_2-y_1|=(y_2(t)-y_1(t))$ on any $[t,T]\subset [a_0,T]$. Thus \[ 0\leq y_1'(t)-y_2'(t) \leq \big(\frac{\lambda}{p+1}\big) \big(\frac{1}{M}\big)^{\frac{\lambda}{p+1}-\lambda} \big(\frac{1}{m}\big)^{\lambda +1}\Big(\int_t^T \phi \, ds\Big)^{\frac{1}{p+1}}(y_2(t)-y_1(t)) . \] Let $C:=\left (\frac{\lambda}{p+1}\right )\left (\frac{1}{M}\right )^{\frac{\lambda}{p+1}-\lambda}\left (\frac{1}{m}\right )^{\lambda +1}$, $P(t):=C\left(\int_t^T \phi \, ds\right)^{\frac{1}{p+1}}$ and $w:=y_1-y_2$. We restate the inequality above as \[ w'(t)+P(t)w(t)\leq 0. \] Thus \[ \left ( e^{P(t)}w(t)\right )'\leq 0, \] and so \[ e^{P(T)}w(T)-e^{P(t)}w(t)\leq 0. \] Notice that $P(T)=0$ and $w(T)=y_1(T)-y_2(T)$, so \[ e^{-P(t)}(y_1(T)-y_2(T))\leq y_1(t)-y_2(t)\leq 0. \] The assumption $\int_0^T\left ( \int_s^T \phi \, ds\right )^{\frac{1}{p+1}} \, dt<\infty$ guarantees that $P(t)$ is bounded even in the case $a_0=0$. Therefore the lemma follows. \vspace{.1in} It is important to observe that the choice of $\delta$ in the previous proof depends upon $\epsilon,m,M$ and $K$, but does not depend upon the interval $[a_0,T]$. \begin{lemma} Let $T\in (0,1)$ and assume that $\int_0^T\left ( \int_s^T \phi \, ds\right )^{\frac{1}{p+1}} \, dt<\infty$. Then $h_T^-\in {\mathcal H}_T^-$, and the solution, $y$, of (\ref{ivp})$_0$ with $y(T)=h_T^-$ has maximal interval $[0,T]$ and satisfies $y(0)=0$. \end{lemma} \paragraph{Proof:} Suppose that $y$ has a maximal interval $[a,T]$ with $a>0$. Recall that $y(a)=0$ in this case. Let $h_a^-$ be defined as before, and choose $a_1\in (a,T)$ such that $0h_T^-$ we know that $y_{\delta}$ has maximal interval $[0,T]$, and by Lemma \ref{condep}, we know that $y(a_1)0$, so, by Lemma \ref{comparison}, we know that $y_{\delta}<{\overline y}$ on any common interval of definition. However, since ${\overline y}(a)0$. Let $m=\frac{\alpha}{2}, M=h_T^-$ and $\epsilon =\frac{\alpha}{4}$, and choose $\delta$ as in Lemma \ref{condep}. Let $y_{\delta}$ represent the solution of (\ref{ivp})$_0$ with $y_{\delta}(T)=h_T^--\delta$. By Lemma \ref{condep} we know that if $\frac{\alpha}{2}\leq y_{\delta}\leq y\leq h_T^-$ on some interval $[a_0,T]$, then $|y_{\delta}-y|\leq \frac{\alpha}{4}$ and thus $\frac{3\alpha}{4}\leq y_{\delta}\leq y\leq h_T^-$. It follows that $y_{\delta}$ is bounded below by $\frac{3\alpha}{4}$ on any subinterval of $[0,T]$, and so its maximal interval is $[0,T]$. However, this contradicts the fact that $y_{\delta}(T)=h_T^--\delta\not\in {\mathcal H}_T^-$. The theorem is proved. \vspace{.1in} It is easy to show, by Lemma \ref{comparison}, that $h_T^-$ is the unique $h\in{\mathcal H}_T^-$ such that the associated solution satisfies $y(0)=0$. We refer to the solution, $y$, of (\ref{ivp})$_0$ such that $y(T)=h_T^-$ as the {\it left half solution} on $[0,T]$. By identical arguments, which we omit, we introduce the quantity $h_T^+$ and the {\it right half solution} on the interval $[T,1]$. The remaining lemmas in this section will prove that $h_T^-=h_T^+$ for exactly one $T$. This will allow us to join the two half solutions to create the unique positive solution of (\ref{bvp}). \begin{lemma} $h_T^-$ is a monotone increasing function in $T$. \end{lemma} \paragraph{Proof:} Let $T_1,T_2\in (0,1)$ such that $T_20=y_{2}'(T_2)$. By Lemma \ref{comparison} we know that $y_2-y_1$ is strictly decreasing, which implies that $y_1(0)h_{T_m}^-$, A similar contradiction arises if $y_n(t)=y_m(t)$ for any $t\in (0,T_m]$, so $y_m(t)0$ we have \[ 0\leq {\overline y}_n'(t)\leq y_n'(t_0)= \Big(\int_{t_0}^T\frac{\phi}{y_{n}^{\lambda}}\, ds \Big)^{\frac{1}{p+1}} \leq \left ( \frac{1}{y_1(t_0)}\right )^{\frac{\lambda}{p+1}}\Big(\int_{t_0}^T \phi \, ds \Big)^{\frac{1}{p+1}}<\infty \] Thus, by the Arzela-Ascoli theorem, me may assume, without loss of generality, that $\{{\overline y}_n\}$ converges uniformly to $f$ on compact subsets of $(0,T]$. Thus $f\in C(0,T]$ and for $t\in (0,T)$ \begin{eqnarray*} \displaystyle f(t) &=& \lim_{n\to\infty}y_n(t)\\ &=&\lim_{n\to\infty}\Big( h_{T_n}^- -\int_{t}^{T_n}\Big( \int_{\tau}^{T_n}\frac{\phi}{y_n^{\lambda}}\, ds \Big)^{\frac{1}{p+1}} \, d\tau\Big)\\ &=&h-\int_{t}^{T}\Big( \int_{\tau}^{T}\frac{\phi}{f^{\lambda}}\, ds \Big)^{\frac{1}{p+1}}\, d\tau. \end{eqnarray*} Hence $f$ is a solution of (\ref{ivp})$_0$ with maximal interval $[0,T]$. Thus $h\in{\mathcal H}_{T}^-$ and $h\geq h_T^-$. But we already proved that $h\leq h_T^-$ so $h=h_T^-$, and thus $ \lim_{n\to\infty} h_{T_n}^-=h_T^-$. Moreover, $f(t)\equiv y(t)$. If $\{T_n\}$ is a sequence converging to $T$ from the right, then an analagous argument holds. In fact the argument is simpler, because there is no need to extend the functions to a common interval. We omit this part of the proof. The lemma is proved. \begin{lemma} $ \lim_{T\to 0^+}h_T^- =0$. \end{lemma} \paragraph{Proof:} Let $y_1$ represent the left half solution of (\ref{ivp})$_0$ over $[0,\frac{1}{2}]$, and let $y_2$ represent the left half solution of (\ref{ivp})$_0$ over $[0,T]$ where $T\in (0,\frac{1}{2})$. Clearly, $y_1'(T)>0=y_2'(T)$. If $y_1(T)\leq h_T^-=y_2(T)$, then Lemma \ref{comparison} implies that $y_1(0)0$ such that $y(t)\geq At$ on $[0,T]$. Then \[ (y'(t))^{p+1} \leq \frac{1}{A^{\lambda}}\int_t^{T}\frac{\phi}{s^{\lambda}}\, ds. \] Thus \[ \lim_{t\to 0^+}\int_t^{T}\frac{\phi}{s^{\lambda}}\, ds\geq A^{\lambda}\lim_{t\to 0^+}(y'(t))^{p+1}=\infty. \] For the finite slope case we derive asymptotic formulae for $y$ that are similar to those in \cite{T}. \begin{theorem} Let $y$ be a left half solution of (\ref{ivp})$_0$ on $[0,T]$ with $y'(0)=A<\infty$. Then \[ y'(t) = \left ( \frac{1}{A}\right )^{\frac{\lambda}{p+1}}\Big( \int _t^T\frac{\phi}{s^{\lambda}}(1+o(1))\, ds \Big)^{\frac{1}{p+1}}, \] and \[ y(t)=\left ( \frac{1}{A}\right )^{\frac{\lambda}{p+1}}\int _0^t \Big( \int _{\tau}^T\frac{\phi}{s^{\lambda}}(1+o(1))\, ds \Big)^{\frac{1}{p+1}}\, d\tau\,. \] \end{theorem} \paragraph{Proof:} Substitute $\frac{1}{y^\lambda (s)}=\frac{1}{(As)^{\lambda}}(1+o(1))$. \vspace{.1in} For the remainder of this section we concentrate on the infinite slope case and assume $ \int_0^{1/2} \frac{\phi}{t^{\lambda}}\, dt =\infty$. The following theorem provides a general tool for comparing the boundary behavior of solutions. \begin{theorem} Let $\psi\in L^1_{{\rm loc}}(0,\delta)$ be a positive function satisfying $ \int_0^{\delta} \frac{\psi}{t^{\lambda}}\, dt =\infty$, $ \int_0^{\delta}\left ( \int_t^{\delta} \psi \, ds\right )^{\frac{1}{p+1}} \, dt<\infty$, and $ \lim_{t\to 0^+}\frac{\psi}{\phi}=1$. Let $y$ be the left half solution of (\ref{ivp})$_0$ on $[0,T]$, and let $z$ be a particular solution of \[ \begin{array}{l} \displaystyle (|z'|^pz')' + \frac{\psi}{z^\lambda}=0 \mbox{ in } (0,\delta),\\ z(0)=0. \end{array} \] Then $\displaystyle \lim_{t\to 0^+}\frac{z'}{y'}=\lim_{t\to 0^+}\frac{z}{y}=1$. \label{bdry} \end{theorem} \vspace{.1in} Before proving Theorem \ref{bdry} we provide an interesting application. \begin{corollary} Assume that $ \lim_{t\to 0^+}\frac{c t^{r}}{\phi}=1$ where $c > 0$ and $-p-2 < r \leq \lambda -1$. Let $y$ be the left half solution of (\ref{ivp})$_0$ on $[0,T]$, and let $z(t)=\gamma t^\rho$ such that $ \rho=\frac{r+p+2}{\lambda +p+1}$ and $ \gamma=\left (\frac{c}{\rho^{p+1}(1-\rho)}\right )^{\frac{1}{\lambda+p+1}}$. Then $ \lim_{t\to 0^+}\frac{z'}{y'}=\lim_{t\to 0^+}\frac{z}{y}=1$. \end{corollary} \paragraph{Proof:} The restrictions on $r$ guarantee that $\psi = ct^r$ satisfies $$ \int_0^1 \big(\int_{t}^{1/2}\psi \, ds \big)^{\frac{1}{p+1}}\, dt<\infty$$ and $ \int_0^{1/2} \frac{\psi}{\tau^{\lambda}} =\infty$. It is straight forward to check that $z(t)$ is a solution of $$\displaylines{ (|z'|^pz')' + \frac{\psi}{z^\lambda}=0 \quad \mbox{in }(0,\infty),\cr z(0)=0\,. }$$ Hence the result follows from Theorem \ref{bdry}. \vspace{.1in} The proof of Theorem \ref{bdry} depends upon several lemmas. \begin{lemma} Let $\psi\in L^1_{{\rm loc}}(0,\delta)$ be a positive function satisfying $ \int_0^{\delta} \frac{\psi}{t^{\lambda}}\, dt =\infty$, and $ \int_0^{\delta}\big( \int_t^{\delta} \psi \, ds \big)^{\frac{1}{p+1}} \, dt<\infty$. Let let $z_1$ and $z_2$ be solutions of $$\displaylines{ (|z'|^pz')' + \frac{\psi}{z^\lambda}=0\quad \mbox{in } (0,\delta),\cr z(0)=0. }$$ Then $\displaystyle \lim_{t\to 0^+}\frac{z_1'}{z_2'}=\lim_{t\to 0^+}\frac{z_1}{z_2}=1$. \end{lemma} \paragraph{Proof:} Assume that $z_1$ and $z_2$ are distinct solutions with $z_1(t)z(t)$ at some point in $(0,\delta)$. Then $y(t)>z(t)$ in some interval $(a_0,b_0)\subset (0,\delta)$ such that $z'(b_0)\geq y'(b_0)$. For $t\in (a_0,b_0)$ we have \begin{eqnarray*}{ll} y'(t) &=& \Big( \left ( y'(b_0)\right )^{p+1} + \int_t^{b_0}\frac{\phi}{y^{\lambda}} \, ds \Big)^{\frac{1}{p+1}} \\ & <& \Big( \left (z'(t_0)\right )^{p+1} + \int_t^{b_0}\frac{\psi}{z^{\lambda}}\, ds \Big)^{\frac{1}{p+1}}\\ &=& z'(t)\,. \end{eqnarray*} Thus $y-z$ is decreasing in $(a_0,b_0)$. Thus the maximal interval where $y>z$ must be $(0,b_0)$, and, by the same argument, $y-z$ is decreasing on $(0,b_0)$. But this implies $0=y(0)-z(0)< y(b_0)-z(b_0)\leq 0$, a contradiction. Hence $z(t)\geq y(t)$ in $[0,\delta]$. The case $\psi\leq\phi$ can be argued similarly so the lemma is proved. \vspace{.1in} \paragraph{\bf Proof of Theorem 5:} Let $\epsilon > 0$ be given. Observe that for $c>0$ we have that $w=cy$ is a left half solution on $[0,T]$ of the problem \begin{equation} \begin{array}{c} \displaystyle (|w'|^pw')'+\frac{c^{\lambda + p + 1}\phi_1}{w^{\lambda}}=0 , \\[7pt] w(T)=h,\, w'(T)=0, \end{array} \label{scaledivp} \end{equation} Also, without loss of generality, we may assume that $(1-\epsilon)^{\lambda +p+1}\phi\leq\psi\leq(1+\epsilon)^{\lambda +p+1}\phi$ in $(0,\delta)$. Therefore, the previous lemma and the observation about (\ref{scaledivp}) show that $ \liminf_{t\to 0^+} \frac{z}{(1-\epsilon)y}\geq 1$ and $ \limsup_{t\to 0^+} \frac{z}{(1+\epsilon)y}\leq 1$. Hence $ \lim_{t\to 0^+} \frac{z}{y}= 1$. By further restricting the size of the interval $(0,\delta)$ we may now assume that $ (1-\epsilon)\leq \frac{\phi z^{\lambda}}{\psi y^{\lambda}}\leq (1+\epsilon)$. We have \begin{eqnarray*} y'(t) &=&\Big( (y'(\delta))^{p+1}+\int_t^{\delta}\frac{\phi}{y^{\lambda}}\, ds\Big)^{\frac{1}{p+1}}\\ &=& \Big( (y'(\delta))^{p+1}+\int_t^{\delta}\frac{\psi}{z^{\lambda}}\frac{\phi z^{\lambda}}{\psi y^{\lambda}}\, ds \Big)^{\frac{1}{p+1}}. \end{eqnarray*} Thus \begin{eqnarray*} \big( (y'(\delta))^{p+1}+\frac{1}{(1+\epsilon)}\int_t^{\delta} \frac{\psi}{z^{\lambda}}\, ds\big)^{\frac{1}{p+1}} &\leq& y'(t)\\ &\leq&\big( (y'(\delta))^{p+1}+\frac{1}{(1-\epsilon)} \int_t^{\delta}\frac{\psi}{z^{\lambda}}\, ds \big)^{\frac{1}{p+1}}. \end{eqnarray*} Divide this inequality through by $ z'(t)=\big( (z'(\delta))^{p+1}+\int_t^{\delta}\frac{\psi}{z^{\lambda}}\, ds \big)^{\frac{1}{p+1}}$ and let $t\to 0^+$ to get \[ \left ( \frac{1}{1+\epsilon}\right)^{\frac{1}{p+1}}\leq \liminf_{t\to 0^+} \frac{y'(t)}{z'(t)} \leq \limsup_{t\to 0^+} \frac{y'(t)}{z'(t)}\leq \left ( \frac{1}{1-\epsilon}\right)^{\frac{1}{p+1}}. \] Hence $ \lim_{t\to 0^+}\frac{z'(t)}{y'(t)}=1$, and Theorem \ref{bdry} is proved. \vspace{.1in} In \cite{T} Taliaferro applies a result similar to Theorem \ref{bdry} to obtain a precise description of boundary behavior for a more general collection of functions $\phi$ than those described in Corollary 2. Taliaferro assumes that $ \lim_{t\to 0^+}\frac{tf''(t)}{f'(t)}=R$, where $ f(t):=\int_t^T \psi/s^\lambda \, ds$ and where $\psi$ is a smooth function such that $ \lim_{t\to 0^+} \psi/\phi=1$. We note that if $tf''(t)/f'(t)\equiv R$ then one can show that $\psi$ is of the form $ct^r$, so Corollary 2 is applicable. Taliaferro's more general condition implies that for any $\epsilon>0$ $\psi$ is bounded between some $ct^{r-\epsilon}$ and $ct^{r+\epsilon}$ in some neighborhood of $0$. Thus $\psi$ still behaves much like $ct^r$. Our methods can be used to find corresponding estimates on the boundary behavior of the solution, but we have not generalized Taliaferro's argument and results to this case. \begin{thebibliography}{99} \bibitem{B} J. Baxley, Some singular nonlinear boundary value problems, { Siam J. Math. Anal.} {\bf 22} (1991), 463-479. \bibitem{BM} Baxley, J. V. and J. C. Martin, Positive solutions of singular nonlinear boundary value problems, { Journal of Computational and Applied Mathematics}, to appear. \bibitem{CF} A. J. Callegari and M. B. Friedman, An analytic solution of a nonlinear singular boundary-value problem in the theory of viscous fluids, {J. Math. Analysis Applic.} {\bf 21}, (1968), 510-529 \bibitem{CN} A.J. Callegari and A. Nachman, Some singular nonlinear differential equations arising in boundary layer theory, {J. Math. Anlysis Applic.} {\bf 64}, (1978), 96-105. \bibitem{GOW} J. Gatica, V. Oliker, and P. Waltman, Singular nonlinear boundary value problems for second-order ordinary differential equations, {Journal of Differential Equations} {\bf 79} (1989), 62-78. \bibitem{GHW} J. Gatica, G. Hernandez, and P. Waltman, Radially symmetric solutions of a class of singular elliptic equations, { Proceedings of the Edinburgh Mathematical Society} {\bf 33} (1990), 169-180. \bibitem{S} J. Shin, A singular nonlinear differential equation arising in the Homann Flow, {Journal Math. Anal. Appl.} {\bf 212} (1997), 443-451. \bibitem{T} S. Taliaferro, A nonlinear singular boundary-value problem, {Nonlinear Analysis: Theory, Methods and Applicaitons} {\bf 3} (1979), 897-904. \end{thebibliography} \bigskip \noindent{\sc Robert M. Houck} (e-mail: houck@mthcsc.wfu.edu)\newline {\sc Stephen B. Robinson} (e-mail: robinson@mthcsc.wfu.edu)\newline Department of Mathematics and Computer Science \newline Wake Forest University \newline Winston-Salem, NC 27109, USA \end{document}