\documentclass[reqno]{amsart} \AtBeginDocument{{\noindent\small {\em Electronic Journal of Differential Equations}, Vol. 2004(2004), No. 120, pp. 1--13.\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 2004 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2004/120\hfil Multiple positive solutions] {Multiple positive solutions to fourth-order singular boundary-value problems in abstract spaces} \author[Yansheng Liu\hfil EJDE-2004/120\hfilneg] {Yansheng Liu} \address{Yansheng Liu \hfill\break Department of Mathematics, Shandong Normal University, Jinan, 250014, China} \email{ysliu6668@sohu.com} \date{} \thanks{Submitted September 22, 2004. Published October 14, 2004.} \thanks{Supported by the Rewarded Foundation for Outstanding Middle and Young Scientist, \hfill\break\indent Shandong Province, China} \subjclass[2000]{34G20, 34B15} \keywords{Banach space; singularity; positive solution; fourth order equation} \begin{abstract} We prove the existence of multiple positive solutions to singular boundary-value problems for fourth-order equations in abstract spaces. Our results improve and extend that obtained in \cite{o1,r1,w1}, even in the scalar case. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{example}[theorem]{Example} \section {Introduction} In this paper, we consider the following singular boundary-value problem (BVP) for fourth-order differential equations in a Banach space $E$: \begin{equation} x^{(4)}(t)=f(t, x(t)), \quad 00$ such that $\|u\|\leq N\|v\|$ if $\theta\leq u\leq v$, where $\theta$ denotes the zero element of $E$. Evidently, $C[I, E]$ is a Banach space with norm $\|x\|_c:= \max_{t\in I}\|x(t)\|$. Moreover, $$ C[I, P]:= \{x\in C[I, E]:\ x(t)\in P,\ t\in I\} $$ is a normal cone of $C[I, E]$ with the same normal constant $N$ as $P$ in $E$. A function $x$ is said to be a solution of \eqref{e1.1} subject to \eqref{e1.2} if $x\in C^2[I, E]\cap C^4[(0, 1), E]$ satisfies \eqref{e1.1} and boundary conditions \eqref{e1.2}; in addition, $x$ is said to be a positive solution if $x(t)>\theta$ for $t\in (0, 1)$ and $x$ is a solution of \eqref{e1.1} with \eqref{e1.2}. Let $u: (0, 1]\to E$ be continuous. The abstract generalized integral $\int_0^1 u(t)dt$ is said to be convergent if the limit $\lim_{\epsilon\to 0^+}\int_{\epsilon}^1u(t)dt$ exists. The convergency or divergency of other kinds of generalized integrals can be defined similarly. For a bounded subset $V$ of Banach space $E$, by $\alpha(V)$ we denote the Kuratowskii measure of noncompactness of $V$(for details, see \cite{d1,l1}). In this paper, the Kuratowskii measure of noncompactness of bounded set in $E$ and $C[I, E]$ are denoted by $\alpha(\cdot)$ and $\alpha_c(\cdot)$, respectively. To conclude this section, we list three lemmas which will be used in Section 3. \begin{lemma}[\cite{g4}] \label{lem2.1} If $V\subset C[J,E]$ is bounded and equicontinuous, then $\alpha(V(t))$ is continuous on $J$ and $\alpha_c(V)=\max\{\alpha(V(t))| t\in J\}$, where $V(t)=\{x(t)|x\in V\}$. \end{lemma} \begin{lemma}[\cite{n1}] \label{lem2.2} Let $K$ be a cone of a real Banach space $E$ and $B$: $K\to K$ a completely continuous operator. Assume that $B$ is order-preserving and positively homogeneous of degree $1$ and that there exist $v\in K\setminus \{\theta\}$, $\lambda>0$ such that $Bv\geq \lambda v$. Then $r(B)\geq\lambda$, where $r(B)$ denotes the spectral radius of $B$. \end{lemma} \begin{lemma}[Fixed point theorem of cone expansion and compression \cite{g4}] \label{lem2.3} Let $P$ be a cone of a real Banach space $E$ and $ P_{r,s}=\{x\in P: r\leq\|x\|\leq s\}$ with $s>r>0$. Suppose that $A$: $P_{r, s}\to P$ is a strict contraction such that one of the following two conditions is satisfied: \begin{itemize} \item[(i)] $Ax\not\leq x$ for $x\in P$, $\|x\|=r$ and $Ax\not\geq x$ for $x\in P$, $\|x\|=s$. \item[(ii)] $Ax\not\geq x$ for $x\in P$, $\|x\|=r$ and $Ax\not\leq x$ for $x\in P$, $\|x\|=s$. \end{itemize} Then, the operator $A$ has a fixed point $x\in P$ such that $r<\|x\|0, $$ where for $t\in (0, 1)$, $z(t):= \min\{t, 1-t\}$ and $$ f_{r, R}(t):= \sup\{\|f(t, x)\|:\ \frac{z(t)}{N}r\leq \|x\|\leq R,\ x\in P\}. $$ \item[(H1)] For every $[c, d]\subset (0, 1)$, and positive numbers $R_2> R_1> 0$, $f(t, x)$ is uniformly continuous on $[c, d]\times\overline{P_{R_2}}\setminus P_{R_1}$ with respect to $t$. Here $P_r=$: $\{x\in P:\|x\|< r\}$ for each $r>0$. \item[(H2)] For every $ t\in (0, 1)$ and every bounded subset $D\subset \overline{P_{R_2}}\setminus P_{R_1}$ $(R_2>R_1>0)$, we have $\alpha(f(t, D))\leq l\alpha(D)$, where $l$ is a constant with $l<15$. \item[(H3)] There exist $\varphi\in L[I, R^+]$, $[c, d]\subset [0, 1]$, and $\varphi^*\in P^*$ (here $P^*$ denotes the dual cone of $P$) with $\|\varphi^*\|=1$ such that $$ \liminf_{x\to\theta,\; x\in P}\varphi^*(f(t, x)) \geq \varphi (t) $$ uniformly with respect to $t\in [c, d]$ and $\int_c^ds(1-s)\varphi(s)ds > 0$. \item[(H4)] There exist functions $a \in C[I, R^+]$ and $b\in C[I, P]$ with $a(t)\not\equiv 0$ on every subinterval of $I$ such that $$ f(t, x)\geq a(t)x- b(t)\quad {for }\ t\in (0, 1)\ { and }\ x\in P\setminus\{\theta\}. $$ \item[(H5)] There exists a positive number $R$ such that $$ N\int_0^1s(1-s)f_{R,R}(s)ds < 8R, $$ where $f_{R,R}(s)$ is the same as in (H0).\end{itemize} Note that assumption (H3) is reasonable since $f(t, x)$ is singular at $x=\theta$. We assume that (H0) holds throughout the remainder of the paper. To overcome the difficulties arising from singularities, we define \begin{equation} Q:=\{x\in C[I, P]:\ x(t)\geq z(t)x(s)\geq \theta,\ \forall t, s\in I\}.\label{e3.1} \end{equation} It is easy to see that $Q$ is a nonempty (notice $t(1-t)\in Q$), convex, and closed subset of $C[I,E]$. Furthermore, $Q$ is a cone of the Banach space $C[I, E]$ and for every $x\in Q\setminus\{\theta\}$, by \eqref{e3.1} and the normality of the cone $P$, we have \begin{equation} \|x(t)\|\geq \frac{z(t)}{N}\|x\|_c > 0\quad \quad\mbox{for }t\in (0, 1).\label{e3.2} \end{equation} Therefore, $x$ is a positive solution of \eqref{e1.1}-\eqref{e1.2} provided that $x\in Q\setminus\{\theta\}$ is a solution of \eqref{e1.1}-\eqref{e1.2}. Define the operator $A$ on $Q\setminus\{\theta\}$ by \begin{equation} (Ax)(t):= \int_0^1 J(t, \tau)f(\tau, x(\tau))d\tau, \quad\forall x\in Q\setminus\{\theta\},\label{e3.3} \end{equation} where $J(t, \tau)=\int_0^1 G(t,s)G(s, \tau)ds\quad$ and \begin{equation} G(t, s)=\begin{cases} t(1-s),& 0\leq t\leq s\leq 1;\\ s(1-t),& 0\leq s\leq t\leq 1. \end{cases} \label{e3.4} \end{equation} Now we show the operator $A$ is well defined on $Q\setminus\{\theta\}$. First we claim that for each $x\in Q\setminus\{\theta\}$, $\int_0^1 G(s, \tau)f(\tau, x(\tau))d\tau $ is convergent. In fact, since $x\in Q\setminus\{\theta\}$, we can see by \eqref{e3.2} that $\|x\|_c\neq 0$ and $$ \frac{z(t)}{N}\|x\|_c\leq \|x(\tau)\| \leq \|x\|_c\quad \mbox{for each } \tau\in (0, 1). $$ This together with $G(s, \tau)\leq \tau(1-\tau)$ for all $ s, \tau\in I$ and (H0) implies that $\int_0^1 G(s, \tau)f(\tau, x(\tau))d\tau $ is convergent and $$ \int_0^1 \tau(1-\tau)\|f(\tau, x(\tau))\|d\tau < +\infty\quad \mbox{for each } x\in Q\setminus\{\theta\}. $$ The Lebesgue dominated convergence theorem yields that for every $x\in Q\setminus\{\theta\}$, $\int_0^1 G(s, \tau)f(\tau, x(\tau))d\tau $ is continuous in $s$ on $I$. Therefore, by \eqref{e3.3} we obtain that $Ax\in C^2[I, P]$ and \begin{equation} \begin{gathered} (Ax)^{(4)}(t)=f(t, x(t)), \quad 0\theta$ for $t\in (0, 1)$ and satisfies \eqref{e1.1}-\eqref{e1.2}. Now we show $x\in Q\setminus\{\theta\}$. To see this notice that \begin{equation} x(t)= \int_0^1 G(t, s)\int_0^1 G(s, \tau) f(\tau, x(\tau))d\tau ds\quad \mbox{for } t\in I.\label{e3.6} \end{equation} This and \begin{equation} \frac{G(t, s)}{G(\xi, s)}= \begin{cases} \frac{t(1-s)}{\xi (1-s)}\geq t\geq z(t), & t, \xi\leq s;\\[3pt] \frac{s(1-t)}{s(1-\xi)}\geq 1-t\geq z(t),& t, \xi\geq s;\\[3pt] \frac{t(1-s)}{s(1-\xi)}\geq t\geq z(t), & ts>\xi \end{cases} \label{e3.7} \end{equation} yield $$ x(t)\geq z(t)\int_0^1 G(\xi,s)\int_0^1 G(s, \tau)f(\tau, x(\tau))d\tau\quad \mbox{for } t, \xi\in I, $$ which implies $x\in Q\setminus\{\theta\}$. On the other hand, it is easy to see by \eqref{e3.6} that $Ax=x$. This completes the proof. \end{proof} Consequently, the existence of positive solution for \eqref{e1.2} is equivalent to that of fixed point of $A$ in $Q\setminus\{\theta\}$. By \eqref{e3.5} and the process similar to the proof of Lemma \ref{lem3.1}, we also obtain the following Lemma. \begin{lemma} \label{lem3.2} $A(Q\setminus\{\theta\})\subset Q$. \end{lemma} \begin{lemma} \label{lem3.3} For every pair of positive numbers $R_2$ and $R_1$ with $R_2>R_1>0$, $A: \overline{Q_{R_2}}\setminus Q_{R_1}\to Q$ is a strict set contraction, where $Q_r:= \{x\in Q:\ \|x\|_c0)$. \end{lemma} \begin{proof} First, under the assumptions for $R_2$ and $R_1$, (H0) guarantees that for each $x\in\overline{Q_{R_2}}\setminus Q_{R_1}$, \begin{equation} \int_0^1 \tau(1-\tau)\|f(\tau, x(\tau))\|d\tau \leq \int_0^1\tau(1-\tau)f_{R_1, R_2}(\tau)d\tau < +\infty \label{e3.8} \end{equation} which implies $A$: $\overline{Q_{R_2}}\setminus Q_{R_1}\to Q$ is bounded. Next we show $A$: $\overline{Q_{R_2}}\setminus Q_{R_1}\to Q$ is continuous. To see this from \eqref{e3.3} it follows that for $x\in\overline{Q_{R_2}}\setminus Q_{R_1}$ and $t_1, t_2\in I$, \begin{equation} \|(Ax)(t_1)-(Ax)(t_2)\|\leq \int_0^1 |G(t_1, s)-G(t_2, s)|ds\int_0^1 G(s, \tau)\|f(\tau, x(\tau))\|d\tau. \label{e3.9} \end{equation} This and \eqref{e3.8} yield that for every subset $V\subset \overline{Q_{R_2}}\setminus Q_{R_1}$, $(AV)(t)$ is equicontinuous on $I$, where $(AV)(t)= \{(Ax)(t): x\in V\}$, $t\in I$. Let $x_n, x\in \overline{Q_{R_2}}\setminus Q_{R_1}$ with $\|x_n-x\|_c\to 0 $ as $n\to +\infty$ . This implies $$ \|x_n(t)-x(t)\|\to 0\quad \mbox{as } n\to +\infty \quad\mbox{for } t\in I. $$ From Lebesgue dominated convergence theorem and \eqref{e3.8}, it follows that $$ \|(Ax_n)(t)-(Ax)(t)\|\to 0\quad \mbox{as } n\to +\infty. $$ Thus, $\{(Ax_n)(t)\}$ is relatively compact for every $t\in I$. From this and the equicontinuity of $\{Ax_n(t)\}$ by the Ascoli-Arzela theorem, we obtain that $\{Ax_n\}$ is a relatively compact subset of $Q$. Now it remains to show $\|Ax_n-Ax\|_c\to 0$ as $n\to +\infty$. In fact, if this is not true, then there is a constant $\epsilon_0>0$ and a subsequence $\{x_{n_i}\}$ of $\{x_n\}$ such that $\|Ax_{n_i}-Ax\|_c\geq \epsilon_0$ ($i=1,2,\dots$). However, the relative compactness of $\{Ax_n\}$ implies that $\{Ax_{n_i}\}$ contains a subsequence which converges in $C[I, P]$. Without loss of generality, we may assume that $\{Ax_{n_i}\}$ itself converges to $y$, that is, $\|Ax_{n_i}-y\|_c\to 0 $ as $i\to +\infty $. So we have $y=Ax$. This is a contradiction. Therefore, $A$ is continuous. Finally, we show $A$: $\overline{Q_{R_2}}\setminus Q_{R_1}\to Q$ is a strict set contraction, that is, there exists $k\in (0, 1)$ such that $\alpha_c(AV)\leq k\alpha_c(V)\quad$ for each $V\subset\overline{Q_{R_2}}\setminus Q_{R_1}$. Fix $V\subset\overline{Q_{R_2}}\setminus Q_{R_1}$. Let \begin{equation} (A_nx)(t):= \int_{1/n}^{1-(1/n)} J(t, s)f(s, x(s))ds\quad \mbox{for } x\in V.\label{e3.10} \end{equation} By \eqref{e3.8} we know \begin{equation} (A_nx)(t)\to (Ax)(t)\quad \mbox{as } n\to +\infty\quad \mbox{for each } x\in V \mbox{ and } t\in I.\label{e3.11} \end{equation} This implies $$ d_H((A_nV)(t), (AV)(t))\to 0\quad \mbox{as } n\to +\infty \quad \mbox{for each } t\in I, $$ where $d_H(\cdot, \cdot)$ denotes the Hausdorff metric. Thus, by the property of noncompactness measure, \begin{equation} \alpha((A_nV)(t))\to\alpha((AV)(t))\quad \quad\mbox{for }t\in I.\label{e3.12} \end{equation} In what follows, we estimate $\alpha((A_nV)(t))$ for each $t\in I$. Note that $x(s)\geq z(s)x(\tau)\geq \theta\quad$ for $s, \tau \in I$ and $x\in V.$ Thus, $$ \frac{R_1}{nN}\leq \frac{\|x\|_c}{nN}\leq \|x(s)\|\leq R_2\quad \quad\mbox{for } s\in[\frac{1}{n}, 1-\frac{1}{n}]. $$ By the definition of integration and \eqref{e3.4}, respectively, we have $$ \int_{1/n}^{1-(1/n)} J(t, s)f(s, x(s))ds\in(1-\frac{2}{n})\overline{co}(\{J(t,s)f(s, x(s)): s\in[\frac{1}{n}, 1-\frac{1}{n}]\} $$ and $$ J(t, s)=\int_0^1 G(t, \tau)G(\tau, s)d\tau\leq \int_0^1 \tau^2(1-\tau)^2d\tau =\frac{1}{30}\quad \mbox{for all } t, s\in I. $$ These, (H1), (H2), and Lemma \ref{lem2.1} guarantee that \begin{equation} \begin{aligned} \alpha((A_nV)(t)) &=\alpha (\{\int_{1/n}^{1-(1/n)} J(t, s)f(s, x(s))ds |\ x\in V\})\\ &\leq (1-\frac{2}{n})\alpha (\overline{co}\{J(t, s)f(s, x(s)) |\ s\in [\frac{1}{n},\ 1-\frac{1}{n}],\ x\in V\})\\ &\leq \alpha(\{J(t, s)f(s, x(s))|\ s\in [\frac{1}{n},\ 1-\frac{1}{n}],\ x\in V\})\\ &\leq \frac{1}{30}\alpha(\{f(s, x(s))|\ s\in [\frac{1}{n},\ 1-\frac{1}{n}],\ x\in V\})\\ &\leq \frac{1}{30}\alpha(f(I_n\times V(I_n))\leq \frac{1}{30}l\cdot\alpha V(I_n))\leq \frac{1}{15}l\alpha_c(V), \end{aligned}\label{e3.13} \end{equation} where $I_n= [\frac{1}{n}, 1-\frac{1}{n}]$ and $V(I_n)= \{x(s)$: $x\in V,\ s\in I_n\}$. Combining Lemma \ref{lem2.1} with \eqref{e3.9}, \eqref{e3.12}, and \eqref{e3.13} again, one can obtain $$\alpha_c(AV)= \max_{t\in I}\alpha ((AV)(t))\leq \frac{1}{15}l\cdot \alpha_c(V). $$ This implies $A$ is a strict set contraction with $k=\frac{1}{15}l< 1$ from $\overline{Q_{R_2}}\setminus Q_{R_1}$ to $Q$. \end{proof} Using (H4), we define an operator $L$ on $C[I, R]$ by $$ (Lu)(t):= \int_0^1 J(s, t)a(s)u(s)ds=\int_0^1 G(\tau, t)\int_0^1 G(s, \tau)a(s)u(s)dsd\tau $$ for $u\in C[I, R]$ where $J$ is given by \eqref{e3.4}, and $a$ is the same as in (H4). It is easy to see $L$: $C[I, R]\to C[I, R]$ is a completely continuous positive operator. Note that if $v(t)=t(1-t)$ on $I$, then $\|v\|_c=\frac{1}{4}$. By $G(t, \tau)\geq t\tau (1-t)(1-\tau)$ for $t, \tau\in I$, we know $$ (Lv)(t)\geq \int_0^1\tau ^2(1-\tau)^2d\tau\int_0^1 s^2(1-s)^2a(s)ds\cdot v(t)=\delta_0 v(t)\quad\quad\mbox{for }t\in I, $$ where $ \delta_0 =\frac{1}{30}\int_0^1 s^2(1-s)^2a(s)ds > 0$. From Lemma \ref{lem2.2} it follows that the spectral radius $r(L)\geq \delta_0>0$. So the well-known Krein-Rutman theorem \cite{n1} guarantees that there exists an $p\in C[I, R^+]$ with $p(t)\not\equiv 0$ on $I$ such that \begin{equation} (Lp)(t)=\int_0^1J(s,t)a(s)p(s)ds=r(L)p(t)\quad\quad\mbox{for }t\in I. \label{e3.14} \end{equation} From \eqref{e3.7} one deduces that \begin{align*} p(t) &=\frac{1}{r(L)}\int_0^1(\int_0^1 G(s, \tau)G(\tau, t)d\tau)a(s)p(s)ds\\ &\geq \frac{1}{r(L)}\int_0^1 z(s)\int_0^1 G(\xi, \tau)G(\tau, t)d\tau)a(s)p(s)ds\\ &\geq \frac{1}{r(L)}\int_0^1 z(s)a(s)p(s)ds\cdot J(\xi, t)\quad\mbox{for all } t, \xi\in I. \end{align*} Therefore, \begin{equation} p(t)\geq \delta J(\xi, t)\quad{\rm for\ all}\ t, \xi\in I, \label{e3.15} \end{equation} where $\delta:= \frac{1}{r(L)}\int_0^1 z(s)a(s)p(s)ds$. This and \eqref{e3.14} guarantees that \begin{equation} \int_0^1 p(t)a(t)dt\geq \delta r(L).\label{e3.16} \end{equation} Now we are ready to give the main result of the present paper. \begin{theorem} \label{thm3.1} Assume that (H0)--(H5) hold and $r(L)>1$. Then \eqref{e1.1} subject to \eqref{e1.2} has at least two positive solutions. \end{theorem} \begin{proof} Set \begin{equation} K:= \{x\in Q:\quad \int_0^1p(t)a(t)x(t)dt\geq \delta r(L)x(s),\quad \forall s\in I\},\label{e3.17} \end{equation} where $Q$ is given by \eqref{e3.1}, $a(t)$ is given by (H4), $p(t)$, $r(L)$, and $\delta$ are given by \eqref{e3.14} and \eqref{e3.15}. By \eqref{e3.16} it is easy to see $K\setminus\{\theta\}\neq\emptyset$ and $K$ is also a cone of $C[I, E]$. We now prove that the operator $A$ defined by \eqref{e3.3} maps $Q\setminus\{\theta\}$ into $K$. In fact, for $x\in Q\setminus\{\theta\}$, it follows from \eqref{e3.3}, \eqref{e3.15}, and \eqref{e3.2} that \begin{align*} \int_0^1p(t)a(t)(Ax)(t)dt &=\int_0^1p(t)a(t)\int_0^1 J(t, s)f(s, x(s))dsdt\\ &=\lim_{n\to +\infty} \int_0^1 p(t)a(t)\int_{1/n}^{1-(1/n)} J(t, s)f(s, x(s))dsdt\\ &= \lim_{n\to +\infty} \int_{1/n}^{1-(1/n)} f(s, x(s))ds\int_0^1 J(t, s)a(t)p(t)dt\\ &= r(L)\lim_{n\to +\infty} \int_{1/n}^{1-(1/n)} p(s)f(s, x(s))ds \\ &\geq \delta r(L)\lim_{n\to +\infty} \int_{1/n}^{1-(1/n)} J(\tau, s)f(s, x(s))ds\\ &= \delta r(L)\lim_{n\to +\infty} \int_0^1 J(\tau, s)f(s, x(s))ds \\ &= \delta r(L)(Ax)(\tau)\quad \mbox{for all } \tau\in I, \end{align*} which implies $A(Q\setminus\{\theta\})\subset K$. Consequently, we obtain by Lemma \ref{lem3.2} and Lemma \ref{lem3.3} that $A: \overline{K_{R_2}}\setminus K_{R_1}\to K$ is a strict set contraction for every pair of positive numbers $R_2$ and $R_1$ with $R_2> R_1>0$, where $K_{R_1}= \{x\in K: \|x\|_c< R_1\}$. Choose a positive number $R_0$ with $R_0 > R$ such that $$ R_0> \frac{N\|b\|_c}{30\delta(r(L) - 1)}\int_0^1 a(t)p(t)dt. $$ We proceed to prove \begin{equation} Ax\not\leq x\quad\mbox{for all } x\in\partial K_{R_0}. \label{e3.18} \end{equation} Suppose, on the contrary, there exists an $x_0\in\partial K_{R_0}$ such that $Ax_0\leq x_0$. Therefore, $$ x_0(t)\geq (Ax_0)(t) = \int_0^1 J(t,s)f(s, x_0(s))ds\quad\mbox{for all } t\in I. $$ Multiply by $p(t)a(t)$ and integrate from $0$ to $1$ to obtain \begin{align*} & \int_0^1 p(t)a(t)x_0(t)dt \geq \int_0^1 p(t)a(t)\int_0^1 J(t, s)f(s, x_0(s))dsdt\\ &\geq \int_0^1 p(t)a(t)\int_0^1 J(t, s)a(s)x_0(s)dsdt - \int_0^1\int_0^1 J(t, s)a(t)p(t)b(s)dsdt\\ &= \int_0^1(\int_0^1 J(t, s)a(t)p(t)dt)a(s)x_0(s)ds -\int_0^1\int_0^1 J(t, s)a(t)p(t)b(s)dsdt\\ &= r(L)\int_0^1 p(s)a(s)x_0(s)ds - \int_0^1\int_0^1 J(t, s)a(t)p(t)b(s)dsdt. \end{align*} This and \eqref{e3.17} yield \begin{align*} \int_0^1\int_0^1 J(t, s)a(t)p(t)b(s)dsdt &\geq (r(L) -1)\int_0^1 p(s)a(s)x_0(s)ds\\ &\geq (r(L)-1)\delta x_0(\tau)\geq \theta\quad\mbox{for all } \tau\in I. \end{align*} The normality of the cone $P$ and $|J(t, s)|\leq \frac{1}{30}$ for all $t, s\in I$ guarantee that $$ \frac{N\|b\|_c}{30}\int_0^1 a(t)p(t)dt\geq \delta(r(L) - 1)R_0. $$ This is a contradiction with the selection of $R_0$. Consequently, \eqref{e3.18} holds. In what follows we show \begin{equation} Ax\not\geq x\quad\mbox{for all } x\in\partial K_R. \label{e3.19} \end{equation} If this is false, then there exists an $x_1\in\partial K_R$ such that $x_1\leq Ax_1$, that is, $$ \theta\leq x_1(t)\leq (Ax_1)(t)\quad\quad\mbox{for }t\in I. $$ Since $x_1\in K\subset Q$, we get $$ x_1(t)\geq z(t)x_1(\tau)\geq \theta\quad\quad\mbox{for }t, \tau\in I. $$ As a result, $$\frac{z(t)}{N}R = \frac{z(t)}{N}\|x_1\|_c \leq \|x_1(t)\| \leq R\quad\quad\mbox{for }t\in I. $$ This implies $$ \|f(t, x_1(t))\| \leq f_{R, R}(t)\quad\quad\mbox{for }t\in (0, 1). $$ Combining the above with (H5) we know \begin{align*} \|x_1(t)\|&\leq N\|(Ax_1)(t)\|\leq N\|\int_0^1 J(t, s)f(s, x_1(s))ds\|\\ &\leq N \int_0^1 J(t, s)\|f(s, x_1(s))\|ds\\ &\leq N\int_0^1 J(t, s)f_{R, R}(s)ds\\ &\leq N\int_0^1 G(t, \tau)(\int_0^1 s(1-s)f_{R, R}(s)ds)d\tau\\ &\leq \frac{N}{8}\int_0^1 s(1-s)f_{R, R}(s)ds < R\quad\quad\mbox{for }t\in I. \end{align*} This is a contradiction. Then \eqref{e3.19} follows. Finally, we prove that there exists a positive number $R'$ with $R' < R$ such that \begin{equation} Ax\not\leq x\quad\quad\mbox{for }x\in \partial K_{R'}.\label{e3.20} \end{equation} In fact, by (H3), given $\epsilon\in (0, \int_c^d J(\frac{1}{2},s)\phi(s)ds)$, there exists an $R''> 0$ such that \begin{equation} \phi^*(f(t, x(t))\geq \phi(t) -\epsilon\quad\quad\mbox{for }t\in [c, d] \mbox{ and } x\in P_{R''}\setminus\{\theta\}.\label{e3.21} \end{equation} Choose \begin{equation} R':= \min\big\{\frac{R}{2},\ R'', \ \int_c^d J(\frac{1}{2}, s)\phi(s)ds- \epsilon\big\}. \label{e3.22} \end{equation} Now we are ready to prove \eqref{e3.20} holds. Suppose, on the contrary, there exists an $x_2\in \partial K_{R'}$ such that $Ax_2\leq x_2$. Then by \eqref{e3.2} we know $$ \frac{z(t)}{N}R'\leq \|x_2(t)\| \leq R'\leq R''\quad\quad\mbox{for }t\in (0, 1). $$ This and \eqref{e3.21} guarantee that \begin{align*} \phi^*(x_2(t))&\geq \int_0^1 J(t, s)\phi^*(f(s, x_2(s))ds\\ &\geq \int_c^d J(t, s)[\phi(s) - \epsilon]ds \\ &> \int_c^d J(t, s)\phi(s)ds - \frac{\epsilon}{2}\quad\quad\mbox{for }t\in (0, 1). \end{align*} Consequently, $$ \|x_2\|_c\geq \phi^*(x_2(\frac{1}{2}))\geq \int_c^d J(\frac{1}{2}, s)\phi(s)ds - \frac{\epsilon}{2} > R'. $$ This is a contradiction with $x_2\in \partial K_{R'}$. Then the conclusion follows from Lemma \ref{lem2.3}. \end{proof} \begin{remark} \label{rmk3.2}\rm If $r(L)>1$ is replaced with $\int_0^1 s^2(1-s)^2a(s)ds > 30$ in Theorem \ref{thm3.1}, the conclusion of Theorem \ref{thm3.1} also holds. In fact, by $v(t)=t(1-t)\in Q$ and $G(t, s)\geq ts(1-t)(1-s)$ for $t, s\in I$, and using the definition of the operator $L$, one can obtain $$ Lv\geq \frac{1}{30}\int_0^1 s^2(1-s)^2a(s)ds\cdot v. $$ Then by Lemma \ref{lem2.2}, $r(L)>1$ follows. \end{remark} For the next theorem we replace (H4) by \begin{itemize} \item[(H'4)] There exists an $\psi^*\in P^*$ with $\|\psi^*\| = 1$ and a subinterval $[c', d']\subset (0, 1)$ such that $$ \lim_{\|x\|\to +\infty,\; x\in P}\frac{\psi^*(f(t, x))}{\|x\|} > \mu $$ uniformly with respect to $t\in [c', d']$, where $$ \mu= N\Big(\min\{c', 1-d'\}\cdot\int_{c'}^{d'} J(\frac{1}{2}, s)ds\Big)^{-1}. $$ \end{itemize} \begin{theorem} \label{thm3.2} Assume that (H0)-(H3), (H'4), and (H5) hold. Then \eqref{e1.1} subject to \eqref{e1.2} has at least two positive solutions. \end{theorem} \begin{proof} Consider the operator $A$ in $Q\setminus\{\theta\}$, where $Q$ is defined by \eqref{e3.1}. From the proof of Theorem \ref{thm3.1}, it is not difficult to see that \eqref{e3.19} and \eqref{e3.20} hold for $x\in\partial Q_{R'}$ and $x\in\partial Q_{R}$, respectively; where $R'$ is given by \eqref{e3.22} and $R$ is given by (H5). It remains to show that there exists a positive number $R_0$ with $R_0> R$ such that \eqref{e3.18} holds for $x\in\partial Q_{R_0}$. By (H'4), there exist an $\epsilon>0$ and an $R_1>0$ such that \begin{equation} \psi^*(f(t, x))\geq (\mu+\epsilon)\|x\|\quad\quad\mbox{for }t\in[c', d'],\ x\in P, \mbox{ and }\|x\|\geq R_1.\label{e3.23} \end{equation} Choose $$ R_0\geq \max\big\{R+1,\ \frac{NR_1}{\min\{c', 1-d'\}}\big\}. $$ We proceed to prove \eqref{e3.18} holds for $x\in\partial Q_{R_0}$. If this is false, then there exists an $x_0\in\partial Q_{R_0}$ such that $Ax_0\leq x_0$. By \eqref{e3.1} we know $$ x_0(t)\geq z(t)x_0(s)\geq \min\{c',1-d'\}\cdot x_0(s)\geq \theta\quad\quad \mbox{for }t\in[c', d'] \mbox{ and } s\in I. $$ >From the normality of the cone $P$ it follows that $$ N\|x_0(t)\| \geq \min\{c',\ 1-d'\}\cdot\|x_0\|_c = \min\{c',\ 1-d'\}\cdot R_0\quad\quad\mbox{for }t\in[c',\ d'], $$ that is, $\|x_0(t)\|\geq R_1$ for $t\in [c,\ d']$. This, \eqref{e3.23}, and \eqref{e3.3} guarantee that \begin{align*} R_0 &\geq \psi^*(x_0(\frac{1}{2}))\\ &\geq \int_0^1 J(\frac{1}{2}, s)\psi^*(f(s, x_0(s))ds\\ &\geq \int_{c'}^{d'} J(\frac{1}{2}, s)(\mu+\epsilon)\|x_0(s)\|ds \\ &\geq \frac{\min\{c', 1-d'\}}{N}\int_{c'}^{d'} J(\frac{1}{2}, s)(\mu+\epsilon)R_0ds > R_0, \end{align*} which is a contradiction. Then the statement of Theorem \ref{thm3.2} follows. \end{proof} \begin{corollary} \label{coro0} Suppose (H0), (H1), (H2), and one of the following conditions are satisfied: \begin{itemize} \item[(i)] (H3) and (H5). \item[(ii)] (H4), (H5), and $r(L)>1$. \item[(iii)] (H'4) and (H5). \end{itemize} Then \eqref{e1.1} subject to \eqref{e1.2} has at least one positive solution. \end{corollary} \begin{remark} \label{rmk3.3} \rm By the same method used above, we can study the existence of multiple positive solutions of second order nonlinear singular boundary-value problems in scalar or in abstract space. \end{remark} \section{Examples} \begin{example} \label{ex1} \rm Consider the boundary-value problem consisting of a finite system of fourth-order scalar nonlinear differential equations. \begin{equation} \begin{gathered} x_n^{(4)}(t)=\frac{1}{\sqrt{t(1-t)}}\big(x_n^{\frac{3}{2}} + \frac{1}{\max_{1\leq i\leq m}|x_i|}\big), \quad 00$$ and $$ \int_0^1 t(1-t)f_{1, 1}(t)dt\leq \int_0^1[\sqrt{t(1-t)} + \frac{1}{\sqrt{t(1-t)}}]dt < (\frac{1}{8} + \pi) < 8. $$ Therefore, by Theorem \ref{thm3.1} or Theorem \ref{thm3.2}, our conclusion follows. \end{proof} \begin{example} \label{ex2} \rm Consider the boundary-value problem consisting of an infinite system of fourth order scalar nonlinear differential equations. \begin{equation} \begin{gathered} x_n^{(4)}(t)=f_n(t, x(t)), \quad t\in (0, 1); \\ x_n(0)=x_n(1)=x_n''(0)=x_n''(1)=0,\quad (n=1, 2,\dots ). \end{gathered} \label{e4.2} \end{equation} where \begin{gather*} f_1(t,x)=\frac{1}{\sqrt{t(1-t)}}\big(x_1 + \frac{x_2}{2} + (\sum_{i\geq 1}|x_i|)^2 + b_1\big),\\ f_2(t,x)=\frac{1}{\sqrt{t(1-t)}}\big( \frac{x_2}{2} + \frac{x_3}{3}+ \frac{1}{\sum_{i\geq 1}|x_i|} + b_2\big),\\ f_n(t, x)= \frac{1}{\sqrt{t(1-t)}}\big(\frac{x_n}{n} + \frac{x_{n+1}}{n+1} + b_n\big),\quad n= 3, 4, \dots;\\ x=(x_1, x_2, \dots),\quad b_i\geq 0 \quad(i=1, 2, \dots),\quad \sum_{i\geq 1}b_i\leq 1. \end{gather*} \textbf{Claim:} System \eqref{e4.2} has at least two positive solutions $x^*(t)= (x_1^*(t), x_2^*(t),\dots )$ and $x^{**}(t)= (x_1^{**}(t), x_2^{**}(t),\dots )$ such that $$ 0< \sum_{1\leq i,\; t\in[0, 1]}|x_i^*(t)| < 1 < \sum_{1\leq i,\;t\in[0, 1]}|x_i^{**}(t)|. $$ \end{example} \begin{proof} Let $E= l^{1}$ with norm $\|x\|= \sum_{i\geq 1}|x_i|$ and $$ P=\{x=(x_1, x_2, \dots ): x_n\geq 0 \quad\mbox{for }n=1, 2, \dots \}. $$ Then $P$ is a normal cone in $E$ and the normal constant is $N=1$. System \eqref{e4.2} can be regarded as a boundary-value problem form \eqref{e1.1} with \eqref{e1.2}, where $x=(x_1, x_2, \dots )$, $$ f(t, x)= (f_1(t, x_1, \dots ),\dots, f_n(t, x_1, \dots ),\dots,). $$ Evidently, $f\in C[ (0,1)\times P\setminus\{\theta\}, P]$ and is singular at $t=0$, $t=1$, and $ x=\theta =(0,0,\dots, )$. Note that for $t\in (0, 1)$ and $R\geq r>0$, $$ f_{r, R}(t)\leq \frac{1}{\sqrt{t(1-t)}}\big( 2R + R^2 + \frac{1}{rt(1-t)} + \|b\|\big) $$ So, (H0) is satisfied. In addition, (H1) is obvious. As in \cite[Example 2.1.2]{g4}, one can see (H2) is satisfied with $l=0$. Choosing $\phi^*=\psi^*=(1, 1, 0,\dots, 0, \dots)$, we know that (H3) and (H'4) holds for \eqref{e4.1}. From $$ \int_0^1 s(1-s)f_{1,1}(s)ds\leq \int_0^1 ((3+\|b\|)\sqrt{s(1-s)} + \frac{1}{\sqrt{s(1-s)}})ds \leq (\frac{4}{8} +\pi) < 8, $$ it follows that (H5) is satisfied. By Theorem \ref{thm3.2}, our conclusion follows. \end{proof} \begin{thebibliography}{00} \bibitem{a1} A. R. Aftabizadeh, \emph{Existence and uniqueness theorems for fourth-order boundary value problem}, J. Math. Anal. Appl., 116 (1986), 415-426. \bibitem{a2} R. P. Agarwal, \emph{Some new results on two point boundary value problems for higher order differential equations}, Funkcial. Ekvac., 29 (1986), 197-212. \bibitem{d1} K. Deimling, \emph{Ordinary differential equations in Banach spaces}, LNM 886. New York, Berlin: Springer-Verlag, 1987. \bibitem{e1} P. W. Eloe and J. Henderson, \emph{Nonlinear boundary value problems and a priori bounds on solutions}, SIAM J. Math. Anal., 15 (1984), 642-647. \bibitem{g1} A. Granas, R. B. Guenther, and J. W. Lee, \emph{Nonlinear boundary value problems for some classes of ordinary differential equations}, Rocky Mountain J. Math., 10 (1980),35-58. \bibitem{g3} J. R. Graef and B. Yang, \emph{On a nonlinear boundary value problem for fourth order equations}, Appl. Anal., 72 (1999), 439-448. \bibitem{g4} D. Guo, V. Lakshmikantham and X. Liu. \emph{Nonlinear Integral Equations in Abstract Spaces}, Kluwer Academic Publishers, 1996. \bibitem{g2} C. P. Gupta, \emph{Existence and uniqueness theorems for the bending of an elastic beam equation}, Appl. Anal., 26 (1988), 289-304. \bibitem{j1} L. K. Jackson, \emph{Existence and uniqueness of solutions of boundary value problems for Lipschitz equations}, J. Differential Equations, 32 (1979), 76-90. \bibitem{l1} V. Lakshmikantham and S. Leela, \emph{Nonlinear differential equations in abstract spaces}, Pergamon, Oxford, 1981. \bibitem{m1} R. Ma, J. Jihui and F. Shengmao, \emph{The method of lower and upper solutions for fourth-order two-point boundary value problems}, J. Math. Anal. Appl., 215 (1997),415-422. \bibitem{m2} R. H. Martin, \emph{Nonlinear operators and differential equations in Banach spaces}, John Wiley and Sons, New York, 1976. \bibitem{n1} R. D. Nussbaum, \emph{Eigenvectors of nonlinear positive operators and the linear Krein-Rutman theorem}. In \emph{Fixed Point Theory} LNM 886. New York, Berlin: Springer-Verlag, 1980, 309-330. \bibitem{o1} D. O'Regan, \emph{Solvability of some fourth (and higher) order singular boundary value problems}, J. Math. Anal. Appl., 161 (1991),78-116. \bibitem{r1} D. O'Regan, \emph{Fourth (and higher) order singular boundary value problems}, Nonlinear Anal., 14 (1990), 1001-1038. \bibitem{w1} Z. Wei, \emph{Positive solutions of singular boundary value problems of fourth order differential equations}, Acta Mathematica Sinica, 42 (1997), 715-722. \end{thebibliography} \end{document}