1 Introduction

In this paper, we study the following scalar differential problem

$$\begin{aligned} \left\{ \begin{array}{ll} -\varphi (u')'= \lambda h(t)f(u),\quad t\in (0,1),\\ u(0)= 0 = u(1), \end{array} \right. \end{aligned}$$
(1)

where \(\varphi : {\mathbb {R}} \rightarrow {\mathbb {R}}\) is an odd increasing homeomorphism, \(\lambda \) is a positive real parameter, \(h \not \equiv 0\) on any subinterval in (0, 1). We also assume that

(A):

There exist an increasing homeomorphism \(\psi \) from \((0,\infty )\) onto \((0,\infty )\) and a function \(\gamma \) from \((0,\infty )\) into \((0,\infty )\) such that

$$\begin{aligned} \psi (\sigma )\le \frac{\varphi (\sigma x)}{\varphi (x)}\le \gamma (\sigma ),\quad \ \ \mathrm {for \ \ all}\quad \ \ \sigma >0, x \in {\mathbb {R}}/\{0\}. \end{aligned}$$
(H):

\(h:(0,1)\rightarrow [0,\infty )\) is locally integrable satisfying

$$\begin{aligned} \int _{0}^{\frac{1}{2}}\psi ^{-1}\left( \int _{s}^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \right) \mathrm{d}s+\int _{\frac{1}{2}}^{1}\psi ^{-1}\left( \int _{\frac{1}{2}}^{s}h(\tau )\mathrm{d}\tau \right) \mathrm{d}s<\infty . \end{aligned}$$
\((F_1)\):

\(f:[0,\infty )\rightarrow [0,\infty )\) is continuous.

\((F_2)\):

\(f(u)>0\), for \(u>0\).

We note that condition (A) on \(\varphi \) is more general than the condition given by Wang in [15, 16], which was first introduced by Xu and Lee [17]. Condition (H) indicates that the weight function h(t) may have a singularity at \(t=0\) and/or \(t=1\). For convenience, we introduce a new class of weight functions. For a bijection \(\iota : {\mathbb {R}}\rightarrow {\mathbb {R}}\), define \({\mathcal {H}}_\iota \) to be the subset of \(L_{loc}^{1}((0,1),[0,\infty ))\) given by

$$\begin{aligned} {\mathcal {H}}_\iota \,{=} \left\{ g \in L_{loc}^{1}((0,1),[0,\infty )) \;\Big |\; \int _{0}^{\frac{1}{2}} \iota ^{-1} \left( \int _{s}^{\frac{1}{2}} g(\tau )\mathrm{d} \tau \right) \mathrm{d}s + \int _{\frac{1}{2}}^{1} \iota ^{-1} \left( \int _{\frac{1}{2}}^{s} g(\tau )\mathrm{d} \tau \right) \mathrm{d}s < \infty \right\} . \end{aligned}$$

By the notation, condition (H) means \(h \in {\mathcal {H}}_\psi .\)

As an example of conditions (A) and (H), define \(\varphi : {\mathbb {R}}\rightarrow {\mathbb {R}}\) to be an odd function with

$$\begin{aligned} \varphi (x)=x^{2}+x,\quad \,\,x\ge 0. \end{aligned}$$

Then, \(\varphi \) satisfies (A) with functions \(\psi \) and \(\gamma \) given as

$$\begin{aligned} \psi (\sigma ) = \left\{ \begin{array}{ll} \sigma ^{2},\quad &{} \quad \mathrm{if} \ 0 <\sigma \le 1,\\ \sigma ,\quad &{} \quad \mathrm{if}\ \sigma >1, \end{array} \right. \end{aligned}$$

and

$$\begin{aligned} \gamma (\sigma ) = \left\{ \begin{array}{ll} 1, &{}\quad \mathrm{if}\ 0 <\sigma \le 1,\\ \sigma ^{2}, &{}\quad \mathrm{if}\ \sigma >1. \end{array} \right. \end{aligned}$$

Moreover, for \(h(t)=t^{-\frac{3}{2}}\), we can easily check h is singular at \(t=0\) and \(h\in {\mathcal {H}}_\psi \).

Due to a wide range of applications in mathematics and physics, p-Laplacian or more generalized Laplacian problems like (1) have been extensively investigated by many researchers in recent years. There is a vast literature dealing with existence, multiplicity and nonexistence of positive solutions for these problems under various assumptions on weight functions and nonlinearities (see [1,2,3,4,5, 7, 10,11,12,13,14,15,16,17,18] and the references therein). Among these works mentioned above, different asymptotic behaviors of the nonlinearities at 0 and/or \(\infty \) affect the number of positive solutions. Let us introduce notations

$$\begin{aligned} f_0:=\lim _{u\rightarrow 0}\frac{f(u)}{\varphi (u)},\ \ \ \ f_\infty :=\lim _{u\rightarrow \infty }\frac{f(u)}{\varphi (u)}. \end{aligned}$$

When \(\varphi (x)=\varphi _p(x):=|x|^{p-2}x\), \(p>1\), Agarwal et al.[1] and Sánchez [14] proved the multiplicity of positive solutions of problem (1) for \(\lambda \) belonging to some open interval if either \(f_0=f_\infty =0\) or \(f_0=f_\infty =\infty \). Later, Wang [16] extended the multiplicity results in [1, 14] to a scalar \(\varphi \)-Laplacian problem with \(h\in C[0,1]\). Recently, Xu and Lee [18] derived some explicit intervals for \(\lambda \) such that the singularly weighted \(\varphi \)-Laplacian problem (1) has at least one positive solution if \(0<f_0,f_\infty <\infty \). For more on existence of positive solutions of p-Laplcain or more generalized Laplacian systems, we refer to [5, 7, 11,12,13, 15, 17, 18] and the references therein for details.

Our goal is to extend the previous results of [1, 14, 16] to a singularly weighted \(\varphi \)-Laplacian problem (1) for the cases \(f_0=f_\infty =0\) and \(f_0=f_\infty =\infty \). We not only obtain the multiplicity results of positive solutions but also obtain the global results of positive solutions for problem (1) with respect to the parameter \(\lambda \). More precisely, our main results can be listed as follows.

Theorem 1

Assume that (A), (H), (\(F_1\)) and (\(F_2\)) hold. If \(f_0=f_\infty =0\), then there exist \(\lambda ^*\ge \lambda _*>0\) such that problem (1) has at least two positive solutions for \(\lambda >\lambda ^{*}\), one positive solution for \(\lambda \in [\lambda _*,\lambda ^{*}]\), and no positive solution for \(\lambda \in (0,\lambda _*).\)

Theorem 2

Assume that (A), (H), (\(F_1\)) and (\(F_2\)) hold. If \(f_0=f_\infty =\infty \), then there exist \(\lambda ^*\ge \lambda _*>0\) such that problem (1) has at least two positive solutions for \(\lambda \in (0,\lambda _*)\), one positive solution for \(\lambda \in [\lambda _*,\lambda ^{*}]\), and no positive solution for \(\lambda >\lambda ^{*}.\)

Remark 1

The condition \(f_0=\infty \) implies that \(f(0)\ge 0\). If \(f(0)>0\), then we can get \(\lambda _*=\lambda ^*\) in Theorem 2. The proof can be easily completed by arguments under to there in [10, 12].

For the proofs, we employ a newly developed solution operator for (1) introduced by Xu and Lee [17] and then we make use of the fixed point theorem of expansion/compression of a cone and Schauder’s fixed point theorem for the main results.

Our paper is organized as follows. In Sect. 2, we establish a solution operator for problem (1) and show some preliminary results. In Sects. 3 and 4, we prove the main results and give some examples, respectively.

2 Preliminaries

The main condition on the weight function h in problem (1) is that h belongs to \({\mathcal {H}}_\psi \)-class which includes singular functions specially on the boundary, i.e., h may not be integrable near the boundary, \(t=0\) and/or \(t=1\). In this case, the solution u may not be of \(C^{1}[0,1]\)-class. So by a solution to problem (1), we understand a function \(u \in C_0[0,1]\cap C^{1}(0,1)\) with \(\varphi (u')\) absolutely continuous and which satisfies problem (1).

Basic tools for proving our main results are the following two well-known fixed point theorems.

Lemma 1

([6, 8], Fixed point theorem of expansion/compression of a cone) Let E be a Banach space and let K be a cone in E. Assume that \(\varOmega _1\) and \(\varOmega _2\) are open subsets of E with \(0\in \varOmega _1\), \(\overline{\varOmega _1}\subset \varOmega _2\). Assume that \(T: K\cap (\overline{\varOmega _2}{\setminus } \varOmega _1 ) \rightarrow K\) is completely continuous such that either

\(\Vert Tu\Vert \le \Vert u\Vert \), for \(u\in K\cap \partial \varOmega _1\) and \(\Vert Tu\Vert \ge \Vert u\Vert \), for \(u\in K\cap \partial \varOmega _2\) or

\(\Vert Tu\Vert \ge \Vert u\Vert \), for \(u\in K\cap \partial \varOmega _1\) and \(\Vert Tu\Vert \le \Vert u\Vert \), for \(u\in K\cap \partial \varOmega _2\),

Then, T has a fixed point in \(K\cap ( \overline{\varOmega _2}{\setminus } \varOmega _1)\).

Lemma 2

([9], Schauder’s fixed point theorem) Let X be a Banach space and let M be a closed, convex and bounded set in X. Assume that \(T: M \rightarrow M\) is completely continuous. Then, T has a fixed point in M.

To set up the solution operator for (1), let \(C_0[0,1]\) be the Banach space with norm \(\Vert u\Vert _\infty =\max _{t\in [0,1]}|u(t)|\) and define a cone K by taking

$$\begin{aligned} K=\left\{ u\in C_0[0,1] \;| \ u\; \mathrm{is\; concave \; on} \; [0,1]\right\} . \end{aligned}$$

Let us consider the following Poisson-type problem

$$\begin{aligned} -\varphi (w')'= & {} g(t),\quad \ \ t\in (0,1) \ \ \ \ \ \ \ \ \ \ \end{aligned}$$
(2)
$$\begin{aligned} w(0)= & {} w(1)=0, \end{aligned}$$
(3)

where \(\varphi \) satisfies (A) and \(g\in {\mathcal {H}}_\varphi \). Note from condition (A) that \({\mathcal {H}}_\psi \subset {\mathcal {H}}_\varphi \). The first step to derive a corresponding operator equation is usually to integrate the equation on the interval [0, s] in (2) for \(s>0.\) But as we remark, g may not be integrable on the interval [0, s]. It seems a small barrier at the start and we try the following process. Let w be a solution of (2), (3). Then, integrating both sides of (2) on the interval \([s,\frac{1}{2}]\) for \(s\in (0,\frac{1}{2}]\) and \([\frac{1}{2},s]\) for \(s\in [\frac{1}{2},1)\), respectively, we find that (2), (3) are equivalent to

$$\begin{aligned} \left\{ \begin{array}{ll} w'(s)=\varphi ^{-1}\left( a+\int _{s}^{\frac{1}{2}}g(\tau )\mathrm{d}\tau \right) ,\quad \; w(0)=0,\quad \; s\in (0,\frac{1}{2}],\\ w'(s)=\varphi ^{-1}\left( a-\int _{\frac{1}{2}}^{s}g(\tau )\mathrm{d}\tau \right) ,\quad \;w(1)=0,\quad \; s\in [\frac{1}{2},1), \end{array} \right. \end{aligned}$$
(4)

where \(a=\varphi (w'(\frac{1}{2}))\). The important step is to show the fact

$$\begin{aligned} \varphi ^{-1}\left( a+\int _{s}^{\frac{1}{2}}g(\tau )\mathrm{d}\tau \right) \in L^{1}(0,\frac{1}{2}] \end{aligned}$$

and it is not obvious. One may refer to Xu and Lee [17] for the proof. Now we may integrate both sides of (4) on the interval [0, t] for \(t\in [0,\frac{1}{2}]\) and on the interval [t, 1] for \(t\in [\frac{1}{2},1]\), respectively. And we get

$$\begin{aligned} w(t)= \left\{ \begin{array}{ll} \int _0^{t}\varphi ^{-1}\left( a+\int _{s}^{\frac{1}{2}}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s,\quad \ \ \ \ t\in [0,\frac{1}{2}],\\ \int _t^{1}\varphi ^{-1}\left( -a+\int _{\frac{1}{2}}^{s}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s,\quad \;\ t\in [\frac{1}{2},1]. \end{array} \right. \end{aligned}$$

To check \(w(\frac{1}{2}^{-})=w(\frac{1}{2}^{+})\), define for \(a\in {\mathbb {R}}\),

$$\begin{aligned} G(a)=\int _0^{\frac{1}{2}}\varphi ^{-1}\left( a+\int _{s}^{\frac{1}{2}}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s- \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( -a+\int _{\frac{1}{2}}^{s}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s. \end{aligned}$$
(5)

Then, the function \(G: {\mathbb {R}}\rightarrow {\mathbb {R}}\) is well defined and has a unique zero \(a=a(g)\) in \({\mathbb {R}}\) (See Xu and Lee [17] for the proof). This implies \(w(\frac{1}{2}^{-})=w(\frac{1}{2}^{+})\). Consequently, if \(\varphi \) satisfies (A) and \(g\in {\mathcal {H}}_\varphi \), then the solution w of (2), (3) can be represented by

$$\begin{aligned} w(t)= \left\{ \begin{array}{ll} \int _0^{t}\varphi ^{-1}\left( a(g)+\int _{s}^{\frac{1}{2}}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s,\quad \ \ \ \ t\in [0,\frac{1}{2}],\\ \int _t^{1}\varphi ^{-1}\left( -a(g)+\int _{\frac{1}{2}}^{s}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s,\quad \;\;t\in [\frac{1}{2},1], \end{array} \right. \end{aligned}$$
(6)

where \(a(g)\in {\mathbb {R}}\) uniquely satisfies

$$\begin{aligned} \int _0^{\frac{1}{2}}\varphi ^{-1}\left( a(g)+\int _{s}^{\frac{1}{2}}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s= \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( -a(g)+\int _{\frac{1}{2}}^{s}g(\tau )\mathrm{d}\tau \right) \mathrm{d}s. \end{aligned}$$

Replacing g(t) with \(\lambda h(t) f(u(t))\) in (2), (3), for \(\lambda >0\), \(u\in K\), we define

$$\begin{aligned} T_\lambda (u)(t) = \left\{ \begin{array}{ll} \int _0^t \varphi ^{-1}\left( a(\lambda h f(u))+ \int _s^{\frac{1}{2}}\lambda h(\tau ) f(u(\tau )) \mathrm{d}\tau \right) \mathrm{d}s,&{}\quad t\in [0,\frac{1}{2}],\\ \int _t^1 \varphi ^{-1}\left( -a(\lambda hf(u)) + \int _{\frac{1}{2}}^s \lambda h(\tau )f(u(\tau )) \mathrm{d}\tau \right) \mathrm{d}s, &{}\quad t\in [\frac{1}{2}, 1], \end{array} \right. \end{aligned}$$

where \(a(\lambda h f(u))\in {\mathbb {R}}\) uniquely satisfies

$$\begin{aligned}&\int _0^{\frac{1}{2}}\varphi ^{-1}\left( a(\lambda h f(u))+\int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad =\int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( -a(\lambda h f(u))+\int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s. \end{aligned}$$

One may show that \(T_\lambda : K \rightarrow K\) is completely continuous (See Lemma 11 in Xu and Lee [17] for details). Thus, we see that u is a positive solution of (1) if and only if

$$\begin{aligned} u=T_\lambda (u) \ \mathrm{on} \ K. \end{aligned}$$

We finally give some remarks and lemma for later use.

Remark 2

From condition (A), we get

$$\begin{aligned} \sigma x\le \varphi ^{-1} [ \gamma (\sigma )\varphi (x) ], \end{aligned}$$

and

$$\begin{aligned} \varphi ^{-1}[\sigma \varphi (x)]\le \psi ^{-1}(\sigma ) x,\quad \ \mathrm{for} \quad \ \sigma \ \mathrm{and} \ x>0. \end{aligned}$$

Remark 3

Let \(h\in L_{loc}^{1}((0,1),\mathbb {R_+})\). Then, for any fixed \(s\in (0,\frac{1}{2})\), we know \(\int _{s}^{\frac{1}{2}}h(\tau )\mathrm{d}\tau <\infty \). Applying \(\sigma =\int _{s}^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \) and \(x=\varphi ^{-1}(1)\) in Remark 2, we get

$$\begin{aligned} \varphi ^{-1}\left( \int _{s}^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \right) \le \varphi ^{-1}(1)\psi ^{-1}\left( \int _{s}^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \right) . \end{aligned}$$

This implies \({\mathcal {H}}_\psi \subset {\mathcal {H}}_\varphi \).

Lemma 3

[15] Let \(w\in C_{0}[0,1]\cap C^{1}(0,1)\) satisfy \(\varphi (w')'\le 0\) on (0, 1). Then, w is concave on [0, 1] and \(\min _{t\in [\frac{1}{4},\frac{3}{4}]}w(t)\ge \frac{1}{4}\Vert w\Vert _{\infty }\), where \(\Vert w\Vert _{\infty }\) is the supremum norm of w.

3 Proof of Main Results

In this section, we prove Theorems 1 and 2. For this, we need to give some lemmas which will play a crucial role.

Lemma 4

Assume that (A), (H), (\(F_1\)) and (\(F_2\)) hold. If \(f_0=f_\infty =0\), then there exists \(\lambda _0>0\) such that (1) has at least two positive solutions for \(\lambda >\lambda _0\).

Proof

For any \(r>0\), define

$$\begin{aligned} {\hat{m}}_r:=\min \{f(x)\ | \ \frac{r}{4}\le x \le r\}. \end{aligned}$$

We see that \({\hat{m}}_r>0\), by (\(F_2\)). Let \(u\in K\cap \partial B_r\); then, by Lemma 3, for \(t\in [\frac{1}{4},\frac{3}{4}]\),

$$\begin{aligned} r=\Vert u\Vert _\infty \ge u(t)\ge \frac{1}{4}\Vert u\Vert _\infty =\frac{r}{4}, \end{aligned}$$

and

$$\begin{aligned} f(u(r))\ge {\hat{m}}_r. \end{aligned}$$
(7)

For simplicity, denote \(a_{\lambda ,u}\triangleq a(\lambda hf(u))\). Then, for \(u\in K\cap \partial B_r\), we get

$$\begin{aligned} 2T_\lambda (u)\left( \frac{1}{2}\right)= & {} \int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( a_{\lambda ,u}+\int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad + \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( -a_{\lambda ,u}+\int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s. \end{aligned}$$

If \(a_{\lambda ,u}\ge 0\), then

$$\begin{aligned}&\int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( a_{\lambda ,u}+\int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad \ge \int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( \int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s, \end{aligned}$$

and by the definition of \(a_{\lambda ,u}\),

$$\begin{aligned}&\int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( -a_{\lambda ,u}+\int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad = \int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( a_{\lambda ,u}+\int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \ge 0. \end{aligned}$$

Thus,

$$\begin{aligned} 2T_\lambda (u)\left( \frac{1}{2}\right) \ge \int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( a_{\lambda ,u}+\int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s. \end{aligned}$$

If \(a_{\lambda ,u}< 0\), then \(-a_{\lambda ,u}> 0\) and

$$\begin{aligned}&\int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( -a_{\lambda ,u}+\int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad \ge \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( \int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s, \end{aligned}$$

and by the same argument, we get

$$\begin{aligned} 2T_\lambda (u)\left( \frac{1}{2}\right) \ge \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( \int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s. \end{aligned}$$

Thus, we obtain

$$\begin{aligned}&2T_\lambda (u)(\frac{1}{2})\\&\quad \ge \min \left\{ \int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( \int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s, \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( \int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \right\} . \end{aligned}$$

By using (7), we get

$$\begin{aligned}&2\Vert T_\lambda (u)\Vert _\infty \ge 2T_\lambda (u)(\frac{1}{2})\\&\quad \ge \min \left\{ \int _{0}^{\frac{1}{2}}\varphi ^{-1}\left( \int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s, \int _{\frac{1}{2}}^{1}\varphi ^{-1}\left( \int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \right\} \\&\quad \ge \min \left\{ \int _{0}^{\frac{1}{4}}\varphi ^{-1}\left( \int _{s}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s, \int _{\frac{3}{4}}^{1}\varphi ^{-1}\left( \int _{\frac{1}{2}}^{s}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \right\} \\&\quad \ge \min \left\{ \int _{0}^{\frac{1}{4}}\varphi ^{-1}\left( \int _{\frac{1}{4}}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s, \int _{\frac{3}{4}}^{1}\varphi ^{-1}\left( \int _{\frac{1}{2}}^{\frac{3}{4}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \right\} \\&\quad \ge \min \left\{ \int _{0}^{\frac{1}{4}}\varphi ^{-1}\left( \lambda {\hat{m}}_r\int _{\frac{1}{4}}^{\frac{1}{2}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s, \int _{\frac{3}{4}}^{1}\varphi ^{-1}\left( \lambda {\hat{m}}_r \int _{\frac{1}{2}}^{\frac{3}{4}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s \right\} \\&\quad = \frac{1}{4}\varphi ^{-1}\left( \lambda {\hat{m}}_r \min \{\int _{\frac{1}{4}}^{\frac{1}{2}} h(\tau )\mathrm{d}\tau ,\int _{\frac{1}{2}}^{\frac{3}{4}} h(\tau )\mathrm{d}\tau \}\right) \triangleq \frac{1}{4}\varphi ^{-1}(\lambda {\hat{m}}_r \varGamma ). \end{aligned}$$

Define

$$\begin{aligned} p(r):= \frac{\varphi (8r)}{{\hat{m}}_r \varGamma }; \end{aligned}$$

then \(p:(0,\infty )\rightarrow (0,\infty )\) is continuous. Since \(f_0=f_\infty =0\), we get

$$\begin{aligned} \lim _{r\rightarrow 0}p(r)=\lim _{r\rightarrow \infty }p(r)=0. \end{aligned}$$

Thus, there exists \(r^{*}\in (0,\infty )\) such that \(p(r^{*})=\inf \{p(r) \ | \ r>0\}\triangleq \lambda _0\). Now for any \(\lambda >\lambda _0\), there exist \(r_1\), \(r_2>0\) such that \(0<r_1<r^{*}<r_2<\infty \) with \(p(r_1)=p(r_2)=\lambda \). Therefore, if \(u\in K\cap \partial B_{r_1}\), then for any \(\lambda >\lambda _0\),

$$\begin{aligned} 2\Vert T_\lambda (u)\Vert _\infty \ge 2T_\lambda (u)(\frac{1}{2})\ge \frac{1}{4}\varphi ^{-1}\left( \frac{\varphi (8r_1)}{{\hat{m}}_{r_1}\varGamma } {\hat{m}}_{r_1}\varGamma \right) =2r_1=2\Vert u\Vert _\infty , \end{aligned}$$

i.e.,

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \ge \Vert u\Vert _\infty ,\quad \ \ \mathrm{for}\quad \ u\in K\cap \partial B_{r_1},\quad \ \lambda >\lambda _0. \end{aligned}$$
(8)

Similarly,

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \ge \Vert u\Vert _\infty ,\quad \ \ \mathrm{for}\quad \ u\in K\cap \partial B_{r_2},\quad \ \lambda >\lambda _0. \end{aligned}$$
(9)

For any \(\lambda >\lambda _0\), we may choose \(\epsilon \ (=\epsilon (\lambda ))>0\) sufficiently small so that

$$\begin{aligned} \psi ^{-1}(\lambda \epsilon )\varUpsilon \le 1, \end{aligned}$$

where

$$\begin{aligned} \varUpsilon =\max \left\{ \int _{0}^{\frac{1}{2}}\psi ^{-1}\left( \int _{s}^{\frac{1}{2}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s, \int _{\frac{1}{2}}^{1}\psi ^{-1}\left( \int _{\frac{1}{2}}^{s} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s \right\} . \end{aligned}$$

Since \(f_0=0\), there exists \(r_{3}^{*} \ (=r_{3}^{*}(\epsilon ))>0\) such that for \(0<x\le r_{3}^{*}\),

$$\begin{aligned} f(x)\le \epsilon \varphi (x). \end{aligned}$$

Take \(0<r_3<\min \{r_1,r_{3}^{*}\}\). Then, for \(u\in K\cap \partial B_{r_3}\), we get

$$\begin{aligned} f(u(t))\le \epsilon \varphi (\Vert u(t)\Vert )\le \epsilon \varphi (r_3). \end{aligned}$$
(10)

Since \(f_\infty =0\), we define a function \({\hat{f}}:[0,\infty )\rightarrow [0,\infty )\) by

$$\begin{aligned} {\hat{f}}(t):=\max \{f(x)\ | \ 0\le x\le t\}. \end{aligned}$$

By Lemma 2.8 in Wang [15], we have

$$\begin{aligned} {\hat{f}}_\infty :=\lim _{t\rightarrow \infty }\frac{{\hat{f}}(t)}{\varphi (t)}=f_\infty =0. \end{aligned}$$

Since \({\hat{f}}_\infty =0\), then for \(\epsilon \) given above, there exists \(r_{4}^{*} \ (=r_{4}^{*}(\epsilon ))>0\) such that for \(t\ge r_{4}^{*}\),

$$\begin{aligned} {\hat{f}}(t)\le \epsilon \varphi (t). \end{aligned}$$

Take \(r_4>\max \{r_2,r_{4}^{*}\}\). Then, for \(u\in K\cap \partial B_{r_4}\), we get

$$\begin{aligned} f(u(t))\le {\hat{f}}(r_4)\le \epsilon \varphi (r_4). \end{aligned}$$
(11)

Since \(T_\lambda (u)\in C_0[0,1]\cap C^1(0,1)\) for \(u\in K \cap \partial B_{r_j}\) (\(j=3,4\)), there exists a unique \(\sigma _j\in (0,1)\) such that

$$\begin{aligned} T_\lambda (u)(\sigma _j)=\max _{t\in [0,1]}T_\lambda (u)(t),\quad \ \ \ \ T_{\lambda }(u)^{'}(\sigma _j)=0. \end{aligned}$$

We first consider the case \(\sigma _j\in (0,\frac{1}{2}]\).

$$\begin{aligned} 0=T_{\lambda }(u)^{'}(\sigma _j)=\varphi ^{-1}\left( a_{\lambda ,u}+\int _{\sigma _j}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) . \end{aligned}$$

Since \(\varphi \) is an odd homeomorphism, \(a_{\lambda ,u}=-\int _{\sigma _j}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \). Applying (10), (11) and Remark 2 with \(\sigma =\lambda \epsilon \), \(x=\varphi ^{-1}\left( \varphi (r_j)\int _s^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \right) \) and then \(\sigma =\int _s^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \), \(x=r_j\) consecutively, we obtain

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty= & {} T_\lambda (u)(\sigma _j)=\int _0^{\sigma _j}\varphi ^{-1}\left( a_{\lambda ,u}+\int _s^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad =\int _0^{\sigma _j}\varphi ^{-1}\left( -\int _{\sigma _j}^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau +\int _s^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s\\&\quad = \int _0^{\sigma _j}\varphi ^{-1}\left( \int _s^{\sigma _j}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \\&\quad \le \int _0^{\frac{1}{2}}\varphi ^{-1}\left( \int _s^{\frac{1}{2}}\lambda h(\tau )f(u(\tau ))\mathrm{d}\tau \right) \mathrm{d}s \\&\quad \le \int _0^{\frac{1}{2}}\varphi ^{-1}\left( \lambda \epsilon \varphi (r_j)\int _s^{\frac{1}{2}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s \\&\quad \le \psi ^{-1}(\lambda \epsilon ) \int _0^{\frac{1}{2}}\varphi ^{-1}\left( \varphi (r_j)\int _s^{\frac{1}{2}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s \\&\quad \le \psi ^{-1}(\lambda \epsilon ) \left[ \int _0^{\frac{1}{2}}\psi ^{-1}\left( \int _s^{\frac{1}{2}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s\right] r_j. \end{aligned}$$

Similarly, for the case \(\sigma _j\in [\frac{1}{2},1)\), we get

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \le \psi ^{-1}(\lambda \epsilon ) [\int _0^{\frac{1}{2}}\psi ^{-1}\left( \int _s^{\frac{1}{2}} h(\tau )\mathrm{d}\tau \right) \mathrm{d}s] r_j. \end{aligned}$$

Combining the above two inequalities and using the choice of \(\epsilon \), we get

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \le \psi ^{-1}(\lambda \epsilon )\varUpsilon r_j \le r_j=\Vert u\Vert _\infty ,\quad \ \ \mathrm{for}\quad \ u\in K\cap \partial B_{r_j} (j=3,4). \end{aligned}$$
(12)

Combining (8), (9) and (12), we conclude that problem (1) has at least two positive solutions \(u_1,u_2\) with \(r_3\le \Vert u_1\Vert _\infty \le r_1<r_2\le \Vert u_2\Vert _\infty \le r_4\), for \(\lambda >\lambda _0\). \(\square \)

Lemma 5

Assume that (A), (H) and (\(F_1\)) hold. If \(f_0=f_\infty =0\), then there exists \({\bar{\lambda }}>0\) such that if (1) has a positive solution at \(\lambda \), then \(\lambda \ge {\bar{\lambda }}\).

Proof

Let \(f_0=f_\infty =0<\infty \), then there exist positive numbers \(\beta _1,\beta _2,R_1,R_2\) such that \(R_1<R_2\), \(\beta _1>f_0\), \(\beta _2>f_\infty \),

$$\begin{aligned} f(x)\le \beta _1 \varphi (x),\quad \ \ \mathrm{for}\quad \ \ 0\le x\le R_1, \end{aligned}$$

and

$$\begin{aligned} f(x)\le \beta _2 \varphi (x),\quad \ \ \mathrm{for}\quad \ \ x\ge R_2. \end{aligned}$$

Let

$$\begin{aligned} \beta =\max \left\{ \beta _1,\beta _2,\max \{\frac{f(x)}{\varphi (x)}\ | \ R_1\le x\le R_2\}, \min \{\beta \ | \ \beta >0, \frac{\psi (\frac{1}{\varUpsilon })}{\beta }<\lambda _0\}\right\} . \end{aligned}$$

Thus, we have

$$\begin{aligned} f(x)\le \beta \varphi (x),\quad \ \ \mathrm{for}\quad \ \ x\in [0,\infty ). \end{aligned}$$
(13)

Indeed, on the contrary, suppose that (1) has a positive solution v for \(0<\lambda <{\bar{\lambda }}\), where \({\bar{\lambda }}=\frac{\psi (\frac{1}{\varUpsilon })}{\beta }\). Since \(v(t)=T_\lambda (v)(t)\) for \(t\in [0,1]\), applying the same argument in the proof of Lemma 4 with aid of (13) and Remark 2 with \(\sigma =\lambda \beta \), \(x=\varphi ^{-1}\left( \varphi (\Vert u\Vert _\infty )\int _s^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \right) \) and \(\sigma =\int _s^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \), \(x=\Vert v\Vert _\infty \) consecutively, we get for \(0<\lambda <{\bar{\lambda }}\),

$$\begin{aligned} \Vert v\Vert _\infty =\Vert T_\lambda (v)\Vert _\infty \le \psi ^{-1}(\lambda \beta )\varUpsilon \Vert v\Vert _\infty <\Vert v\Vert _\infty , \end{aligned}$$

which is a contradiction. \(\square \)

Lemma 6

Assume that (A), (H), (\(F_1\)) and \(f_\infty =0\) hold. If (1) has a positive solution at \(\lambda ={\hat{\lambda }}\), then (1) has at least one positive solution for \(\lambda >{\hat{\lambda }}\).

Proof

Let \({\hat{u}}\) be a positive solution of (1) at \(\lambda ={\hat{\lambda }}\) and let \(\lambda >{\hat{\lambda }}\) be fixed. Let us consider the modified problem

$$\begin{aligned} \left\{ \begin{array}{ll} -\varphi (u')'= \lambda h(t)f^*(u),\quad t\in (0,1),\\ u(0)= 0 = u(1), \end{array} \right. \end{aligned}$$
(14)

where \(f^*:[0,\infty )\rightarrow [0,\infty )\) is defined by \(f^*(u)=f(\gamma (u))\) with

$$\begin{aligned} \gamma (u)= \left\{ \begin{array}{ll} u,\quad \ \ \mathrm{if }\quad \ u\ge {\hat{u}},\\ {\hat{u}},\quad \ \ \mathrm{if }\quad \ 0\le u< {\hat{u}}. \end{array} \right. \end{aligned}$$

First, let us show that (14) has at least one positive solution. Since \(f_\infty =0\), we can easily get

$$\begin{aligned} f_\infty ^*:=\lim _{u\rightarrow \infty }\frac{f^*(u)}{\varphi (u)}=\lim _{u\rightarrow \infty }\frac{f(u)}{\varphi (u)}=f_\infty =0. \end{aligned}$$

Define \(T_\lambda ^*\) to be the same as \(T_\lambda \) replacing f by \(f^*\). Then, \(T_{\lambda }^*: K\rightarrow K\) is also completely continuous. For fixed \(\lambda >{\hat{\lambda }}\), we can choose \(\epsilon \ (=\epsilon (\lambda ))>0\) so that \(\psi ^{-1}(\lambda \epsilon )\varUpsilon \le 1\). Since \(f_\infty ^*=0\), there exists \(r \ (=r(\epsilon ))>0\) such that for \(x\ge r\), we have

$$\begin{aligned} f^*(x)\le \epsilon \varphi (x). \end{aligned}$$

Here we consider two cases.

Case 1\(f^*\) is bounded on \([0,\infty )\).

Then, there exists \(M>0\) such that for \(x\ge 0\), we have

$$\begin{aligned} f^*(x)\le M. \end{aligned}$$

Choose \(R> \max \{\Vert {\hat{u}}\Vert _\infty ,\varphi ^{-1}(\lambda M)\varUpsilon \}\). Then, for \(u\in K\cap B_R\), we have

$$\begin{aligned} 0\le u(t)\le \Vert u\Vert _\infty \le R, \end{aligned}$$

and

$$\begin{aligned} f^*(u(t))\le M. \end{aligned}$$

Applying the similar arguments in Lemma 4, we can get

$$\begin{aligned} \Vert T_\lambda ^*(u)\Vert _\infty \le \varphi ^{-1}(\lambda M)\varUpsilon \le R,\quad \ \ \mathrm{for}\quad \ u\in K\cap B_R. \end{aligned}$$

Case 2\(f^*\) is unbounded on \([0,\infty )\).

Then, there exists \(R> \max \{\Vert {\hat{u}}\Vert _\infty ,r\}\) such that for \(x\in [0,R]\), we have

$$\begin{aligned} f^*(x)\le f^*(R). \end{aligned}$$

Thus, if \(u\in K\cap B_R\), we have

$$\begin{aligned} 0\le u(t)\le \Vert u\Vert _\infty \le R, \end{aligned}$$

and

$$\begin{aligned} f^*(u(t))\le f^*(R)\le \epsilon \varphi (R), \end{aligned}$$

for \(t\in [0,1].\) Applying arguments under to there in the proof of Lemma 4, we can get

$$\begin{aligned} \Vert T_\lambda ^*(u)\Vert _\infty \le \psi ^{-1}(\lambda \epsilon ) \varUpsilon R\le R,\quad \ \ \mathrm{for}\quad \ u\in K\cap B_R. \end{aligned}$$

By Lemma 2, we conclude that (14) has a solution \(u\in K\cap B_R\).

Next, we show that if u is a solution of (14), then \(u(t)\ge {\hat{u}}(t)\), for \(t\in [0,1]\).

If it is true, then (14) and (1) are equivalent and the proof can be done. By contradiction, assume that \(u(t) \not \ge {\hat{u}}(t)\), for \(t\in [0,1]\). Then, by the boundary values of u and \({\hat{u}}\), there exist \(T_1\), \(T_2\in (0,1)\) such that

$$\begin{aligned} u(t)-{\hat{u}}(t)<0\ \quad \mathrm{for}\quad \ t\in (T_1,T_2),\quad \ \mathrm{and} \quad \ u(T_1)-{\hat{u}}(T_1)=u(T_2)-{\hat{u}}(T_2)=0. \end{aligned}$$

For \(t\in (T_1,T_2)\), we have

$$\begin{aligned} -\varphi (u^{'}(t))^{'}=\lambda h(t)f^*(u(t))=\lambda h(t)f({\hat{u}}(t))\ge {\hat{\lambda }}h(t)f({\hat{u}}(t))=-\varphi ({\hat{u}}^{'}(t))^{'}, \end{aligned}$$

i.e.,

$$\begin{aligned} \varphi (u^{'}(t))^{'} \le \varphi ({\hat{u}}^{'}(t))^{'}. \end{aligned}$$
(15)

Since \(u-{\hat{u}}\in C_0[T_1,T_2]\), there exists \(t_0 \in (T_1,T_2)\) and \(0<\delta <T_2-t_0\) such that

$$\begin{aligned} u(t_0)-{\hat{u}}(t_0)=\min _{t\in [T_1,T_2]}\{u(t)-{\hat{u}}(t)\}, \end{aligned}$$

and

$$\begin{aligned} \ u'(t_0)-{\hat{u}}'(t_0)=0,\quad \ \ u'(t)-{\hat{u}}'(t)>0,\quad \ t\in (t_0. t_0+\delta ). \end{aligned}$$

Integrating both sides of (15) from \(t_0\) to \(t\in (t_0,t_0+\delta )\), we get

$$\begin{aligned} \varphi (u'(t))-\varphi (u'(t_0))\le \varphi ({\hat{u}}'(t))-\varphi ({\hat{u}}'(t_0)). \end{aligned}$$

Since \(\varphi \) is increasing, we have \(u'(t)\le {\hat{u}}'(t)\), \(t\in (t_0,t_0+\delta )\), which is a contradiction. \(\square \)

Lemma 7

Assume that (A), (H), (\(F_1\)) and \(f_\infty =0\) hold. Let I be a compact interval of \((0,\infty )\). Then, there exists a constant \(b_I>0\) such that all possible solutions u of (1) with \(\lambda \in I\) satisfy \(\Vert u\Vert _{\infty }<b_I\).

Proof

Suppose to the contrary that there exists a sequence \(\{u_n\}\) of positive solutions of (1) at \(\lambda =\lambda _n\) with \(\{\lambda _n\}\subset I=[\alpha ,\beta ]\subset (0,\infty )\) and \(\Vert u_n\Vert _\infty \rightarrow \infty \) as \(n \rightarrow \infty \). Choose \(\epsilon >0\) sufficiently small so that \(\psi ^{-1}(\beta \epsilon )\varUpsilon <1\). Using the fact that \(f_\infty =0\), we may define a function \({\hat{f}}:[0,\infty )\rightarrow [0,\infty )\) by

$$\begin{aligned} {\hat{f}}(t):=\max \{ f(x) \ | \ 0\le x\le t\}. \end{aligned}$$

Then, by Lemma 2.8 in Wang [15], we have

$$\begin{aligned} {\hat{f}}_\infty := \lim _{t\rightarrow \infty }\frac{{\hat{f}}(t)}{\varphi (t)}=f_\infty =0. \end{aligned}$$

Thus, for \(\epsilon \) given above, there exists \(r \ (=r(\epsilon ))>0\) such that for \(t\ge r\),

$$\begin{aligned} {\hat{f}}(t)\le \epsilon \varphi (t). \end{aligned}$$

From the assumption, we can get \(\Vert u_n\Vert _\infty >r\), for sufficiently large n. Thus, for sufficiently large n, we have

$$\begin{aligned} f(u_n(t))\le {\hat{f}}(\Vert u_n\Vert _\infty )\le \epsilon \varphi (\Vert u_n\Vert _\infty ). \end{aligned}$$
(16)

Since \(v_n(t)=T_{\lambda _n}(v_n(t))\), for \(t\in [0,1]\), applying the same argument in the proof of Lemma 4 with aid of (16) and Remark 2 with \(\sigma =\lambda _n \epsilon \),

\(x=\varphi ^{-1}\left( \varphi (\Vert v_n\Vert _\infty )\int _s^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \right) \) and \(\sigma =\int _s^{\frac{1}{2}}h(\tau )\mathrm{d}\tau \), \(x=\Vert v_n\Vert _\infty \) consecutively, and using the fact \(\psi ^{-1}(\beta \epsilon )\varUpsilon <1\), we get for \(\lambda _n\in I\) with sufficiently large n,

$$\begin{aligned} \Vert v_n\Vert _\infty =\Vert T_{\lambda _n}(v_n)\Vert _\infty \le \psi ^{-1}(\lambda _n \epsilon )\varUpsilon \Vert v_n\Vert _\infty \le \psi ^{-1}(\beta \epsilon )\varUpsilon \Vert v_n\Vert _\infty < \Vert v_n\Vert _\infty , \end{aligned}$$

which is a contradiction. \(\square \)

Proof of Theorem 1

Define

$$\begin{aligned}&\lambda ^*:=\inf \{{\tilde{\lambda }} \ | \ (1)\ \mathrm{has \ at \ least \ two \ positive \ solutions \ for} \ \lambda >{\tilde{\lambda }}\}. \end{aligned}$$
(17)
$$\begin{aligned}&\lambda _*:=\inf \{\lambda \ | \ (1) \ \mathrm{has \ at \ least \ one \ positive \ solution}\}. \end{aligned}$$
(18)

By Lemmas 4 and 5, \(\lambda ^*\) and \(\lambda _*\) are both well-defined and \(\lambda ^*\ge \lambda _*\ge {\bar{\lambda }}>0\). By the definitions of \(\lambda ^*\) and \(\lambda _*\), and Lemma 6, we get that (1) has at least two positive solutions for \(\lambda >\lambda ^*\), one positive solution for \(\lambda \in (\lambda _*,\lambda ^*]\), and no positive solution for \(\lambda \in (0,\lambda _*)\). Finally, it is enough to show that (1) has at least one positive solution at \(\lambda =\lambda _*\). By the definition of \(\lambda _*\) and Lemma 5, we can choose a sequence \(\{\lambda _n\}\) with \({\bar{\lambda }}\le \lambda _*<\lambda _n\le 2\lambda _*\) such that \(\lambda _n \rightarrow \lambda _*\) as \(n\rightarrow \infty \). Then, by Lemma 7, take \(I=[{\bar{\lambda }},2\lambda _*]\); there exists \(b_I>0\) such that the corresponding solutions \(u_n\) satisfying \(\Vert u_n\Vert _\infty <b_I\), i.e., \(\{u_n\}\) is bounded. Since \(T_{\lambda _n}\) is completely continuous, we know that \(\{T_{\lambda _n}(u_n)\}\) is equicontinuous. So is \(\{u_n\}\), since \(u_n = T_{\lambda _n}(u_n).\) By the Ascoli–Arzel\(\grave{a}\) theorem, \(\{u_n\}\) is relatively compact. Hence, there exists a convergent subsequence \(\{u_n\}\) (still denoted by \(\{u_n\}\)) and \(u_*\in K\) such that \(u_n \rightarrow u_*\) as \(n\rightarrow \infty \). Since \(u_n=T_{\lambda _n}(u_n)\), by the Lebesgue Dominated Convergence Theorem, we get \(u_*=T_{\lambda _*}(u_*)\) which implies that \(u_*\) is a solution of (1) at \(\lambda =\lambda _*\). Moreover, by \(f_0=0\) and applying a similar argument as in the proof of Lemma 7, we see that \(u_*\not \equiv 0\). Consequently \(u_*\) is a positive solution of (1) at \(\lambda =\lambda _*\). \(\square \)

Lemma 8

Assume that (A), (H), (\(F_1\)) and (\(F_2\)) hold. If \(f_0=f_\infty =\infty \), then there exists \(\lambda _0>0\) such that (1) has at least two positive solutions for \(\lambda \in (0,\lambda _0)\).

Proof

For any \(r>0\), define

$$\begin{aligned} {\hat{M}}_r = \max \{f(x) \ | \ 0\le x\le r\}. \end{aligned}$$

Then, by (\(F_2\)), \({\hat{M}}_r>0\). Let \(u\in K\cap \partial B_r\), then for \(t\in [0,1]\),

$$\begin{aligned} 0\le u(t)\le \Vert u\Vert _\infty =r, \end{aligned}$$

and

$$\begin{aligned} f(u(t))\le {\hat{M}}_r. \end{aligned}$$
(19)

Since \(T_\lambda (u)\in C_0[0,1]\cap C^1 (0,1)\) for \(u\in K\cap \partial B_r\), there exists a unique \(\sigma \in (0,1)\) such that

$$\begin{aligned} T_\lambda (u)(\sigma )=\max _{t\in [0,1]}T_\lambda (u(t))\quad \ \ \mathrm{and}\quad \ \ T_\lambda (u)'(\sigma )=0. \end{aligned}$$

We also consider the two cases \(\sigma \in (0,\frac{1}{2}]\) and \(\sigma \in [\frac{1}{2},1).\) By the same argument as in the proof of Lemma 4 with aid of (19), we get

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \le \varphi ^{-1}(\lambda {\hat{M}}_r)\varUpsilon . \end{aligned}$$

Define

$$\begin{aligned} q(r):=\frac{\varphi (\frac{r}{\varUpsilon })}{{\hat{M}}_r}; \end{aligned}$$

then \(q:(0,\infty )\rightarrow (0,\infty )\) is continuous. Since \(f_0=f_\infty =\infty \), we get

$$\begin{aligned} \lim _{r\rightarrow 0}q(r)=\lim _{r\rightarrow \infty }q(r)=0. \end{aligned}$$

Therefore, we may choose \(r^*\in (0,\infty )\) satisfying \(q(r^*)=\sup \{ q(r) \ | \ r>0\}\triangleq \lambda _0,\) and for any \(\lambda \in (0,\lambda _0)\), there exist \(r_1\), \(r_2>0\) such that \(0<r_1<r^*<r_2<\infty \) with \(q(r_1)=q(r_2)=\lambda \). Therefore, if \(u\in K\cap \partial B_{r_1}\), then for \(\lambda \in (0,\lambda _0)\), we obtain

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \le \varphi ^{-1}\left( \frac{\varphi (\frac{r_1}{\varUpsilon })}{{\hat{M}}_{r_1}}{\hat{M}}_{r_1}\right) \varUpsilon =r_1=\Vert u\Vert _\infty , \end{aligned}$$
(20)

for \( u\in K\cap \partial B_{r_1}.\) Similarly,

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \le \Vert u\Vert _\infty , \end{aligned}$$
(21)

for \(u\in K\cap \partial B_{r_2}.\) Let \(\lambda \in (0,\lambda _0)\) and take \(M=\frac{\gamma (32)}{\lambda \varGamma }>0\). Then, by the fact that \(f_0=\infty \), there exists \(r_M>0\) such that for \(0 \le x\le r_M\), we have

$$\begin{aligned} f(x)\ge M\varphi (x). \end{aligned}$$

If \(u\in K\) with \(\Vert u\Vert _\infty \le r_M\), then by Lemma 3, we get

$$\begin{aligned} 0\le u(t)\le \Vert u\Vert _\infty \le r_M, \end{aligned}$$

and

$$\begin{aligned} f(u(t))\ge M\varphi (u(t))\ge M\varphi (\frac{1}{4}\Vert u\Vert _\infty ), \end{aligned}$$

for \(t\in [\frac{1}{4},\frac{3}{4}].\) Take \(0<r_3<\min \{r_1,r_M\}\). Then, for \(u\in K\cap \partial B_{r_3}\), we get

$$\begin{aligned} f(u(t))\ge M\varphi (u(t))\ge M\varphi (\frac{1}{4}\Vert u\Vert _\infty ), \end{aligned}$$
(22)

for \(t\in [\frac{1}{4},\frac{3}{4}].\) Since \(f_\infty =\infty \), for M given above, we may choose \(R_M>0\) such that \(x\ge R_M\) implies \(f(x)\ge M\varphi (x).\) If \(u\in K\) with \(\Vert u\Vert _\infty \ge 4R_M\), then by Lemma 3

$$\begin{aligned} u(t)\ge \min _{t\in [\frac{1}{4},\frac{3}{4}]}u(t)\ge \frac{1}{4}\Vert u\Vert _\infty \ge R_M, \end{aligned}$$

and

$$\begin{aligned} f(u(t))\ge M\varphi (u(t))\ge M\varphi (\frac{1}{4}\Vert u\Vert _\infty ), \end{aligned}$$

for \(t\in [\frac{1}{4},\frac{3}{4}].\) Take \(r_4>\max \{r_2,4R_M\}\). Then, for \(u\in K\cap \partial B_{r_4}\), we get

$$\begin{aligned} f(u(t))\ge M\varphi (u(t))\ge M\varphi (\frac{1}{4}\Vert u\Vert _\infty ), \ \end{aligned}$$
(23)

for \(t\in [\frac{1}{4},\frac{3}{4}].\) We also consider the two cases \(a_{\lambda ,u}\ge 0\) and \(a_{\lambda ,u}<0\). Applying the same argument as in the proof of Lemma 4 with aid of (22), (23) and the definition of M, we get

$$\begin{aligned} 2\Vert T_\lambda (u)\Vert _\infty \ge 2T_\lambda (u)(\frac{1}{2})\ge \frac{1}{4}\varphi ^{-1}\left( \lambda M\varphi (\frac{1}{4}\Vert u\Vert _\infty )\varGamma \right) \ge \frac{1}{4}\varphi ^{-1}\left( \gamma (32)\varphi (\frac{1}{4}\Vert u\Vert _\infty )\right) . \end{aligned}$$

Applying Remark 2 with \(\sigma =32\) and \(x=\frac{1}{4}\Vert u\Vert _\infty \), we get

$$\begin{aligned} 2\Vert T_\lambda (u)\Vert _\infty \ge \frac{1}{4}\times 32\times \frac{1}{4}\Vert u\Vert _\infty =2\Vert u\Vert _\infty . \end{aligned}$$

This implies that

$$\begin{aligned} \Vert T_\lambda (u)\Vert _\infty \ge \Vert u\Vert _\infty ,\quad \ \ \mathrm{for}\quad \ u\in K\cap \partial B_{r_j} (j=3,4). \end{aligned}$$
(24)

Combining (20), (21) and (24), we conclude that (1) has at least two positive solutions \(u_1\) and \(u_2\) with \(r_3\le \Vert u_1\Vert _\infty \le r_1<r_2\le \Vert u_2\Vert _\infty \le r_4\), for \(\lambda \in (0,\lambda _0)\). \(\square \)

Lemma 9

Assume that (A), (H) and (\(F_1\)) hold. If \(f_0=f_\infty =\infty \), then there exists \({\bar{\lambda }}>0\) such that if (1) has a positive solution at \(\lambda \), then \(\lambda \le {\bar{\lambda }}\).

Proof

Let \(f_0=f_\infty =\infty \), then there exist positive numbers \(\eta _1\), \(\eta _2\), \(r_1^{'}\), \(r_2^{'}\) such that \(r_1^{'}<r_2^{'}\), \(0<\eta _1<f_0\), \(0<\eta _2<f_\infty \),

$$\begin{aligned} f(x)\ge \eta _1\varphi (x),\quad \ \ 0\le x\le r_1^{'}, \end{aligned}$$

and

$$\begin{aligned} f(x)\ge \eta _2\varphi (x),\quad \ \ x\ge r_2^{'}. \end{aligned}$$

Let

$$\begin{aligned} \eta _3=\min \left\{ \eta _1,\eta _2, \min \{\frac{f(x)}{\varphi (x)} \ | \ \frac{r_1^{'}}{4}\le x\le r_2^{'}\}, \max \{\eta \ | \ \eta>0, \frac{\gamma (32)}{\eta \varGamma }>\lambda _0\}\right\} >0. \end{aligned}$$

Then, we have

$$\begin{aligned} f(x)\ge \eta _3\varphi (x),\quad \ \ 0\le x\le r_1^{'}, \end{aligned}$$
(25)

and

$$\begin{aligned} f(x)\ge \eta _3\varphi (x),\quad \ \ x\ge \frac{r_1^{'}}{4}. \end{aligned}$$
(26)

Suppose to the contrary that (1) has a positive solution v for \(\lambda >{\bar{\lambda }}\), where \({\bar{\lambda }}=\frac{\gamma (32)}{\eta _3 \varGamma }\). If \(\Vert v\Vert _\infty \le r_1^{'}\), by (25) and Lemma 3,

$$\begin{aligned} f(v(t))\ge \eta _3 \varphi (v(t))\ge \eta _3 \varphi (\frac{1}{4}\Vert v\Vert _\infty ), \end{aligned}$$
(27)

for \(t\in [\frac{1}{4},\frac{3}{4}].\) If \(\Vert v\Vert _\infty > r_1^{'}\), then by Lemma 3 and (26), we get

$$\begin{aligned} v(t)\ge \min _{t\in [\frac{1}{4},\frac{3}{4}]}v(t)\ge \frac{1}{4}\Vert v\Vert _\infty > \frac{r_1^{'}}{4}, \end{aligned}$$

and

$$\begin{aligned} f(v(t))\ge \eta _3 \varphi (v(t))\ge \eta _3 \varphi (\frac{1}{4}\Vert v\Vert _\infty ), \end{aligned}$$
(28)

for \(t\in [\frac{1}{4},\frac{3}{4}].\) Since \(v=T_\lambda (v),\) applying the same argument as in the proof of Lemma 4 with the aid of (27), (28), and Remark 2 with \(\sigma =32\), \(x=\frac{1}{4}\Vert v\Vert _\infty \), we get

$$\begin{aligned} \Vert v\Vert _\infty= & {} \Vert T_\lambda (v)\Vert _\infty \ge \frac{1}{8}\varphi ^{-1}\left( \lambda \eta _3 \varphi (\frac{1}{4}\Vert v\Vert _\infty )\varGamma \right) \\&\quad > \frac{1}{8}\varphi ^{-1}\left( \gamma (32) \varphi (\frac{1}{4}\Vert v\Vert _\infty )\right) \ge \frac{1}{8}\times 32\times \frac{1}{4}\Vert v\Vert _\infty =\Vert v\Vert _\infty , \end{aligned}$$

for \(\lambda >{\bar{\lambda }}.\) This is a contradiction. \(\square \)

Lemma 10

Assume that (A), (H), (\(F_1\)) and \(f_0=\infty \) hold. If (1) has a positive solution at \(\lambda ={\hat{\lambda }}\), then (1) has at least one positive solution for \(\lambda \in (0,{\hat{\lambda }})\).

Proof

Let \({\hat{u}}\) be a positive solution of (1) at \(\lambda ={\hat{\lambda }}\) and let \(\lambda \in (0,{\hat{\lambda }})\) be fixed. Let us consider the modified problem

$$\begin{aligned} \left\{ \begin{array}{ll} -\varphi (u')'=\lambda h(t)f^*(u),\quad t\in (0,1),\\ u(0)=0=u(1), \end{array} \right. \end{aligned}$$
(29)

where \(f^*:[0,\infty )\rightarrow [0,\infty )\) is defined by \(f^*(u)=f(\gamma (u))\) with

$$\begin{aligned} \gamma (u)= \left\{ \begin{array}{ll} {\hat{u}},&{}\quad \mathrm{if}\ u>{\hat{u}},\\ u,&{}\quad \mathrm{if}\ 0\le u\le {\hat{u}}. \end{array} \right. \end{aligned}$$

First, we show that (29) has at least one positive solution. Define \(T_\lambda ^{*}\) to be the same as \(T_\lambda \) replacing f by \(f^*\). Then, \(T_\lambda ^{*}:K\rightarrow K\) is completely continuous. By the fact that \(f^*\) is bounded, we may choose \(R>0\) satisfying \(\Vert T_\lambda ^{*}(u)\Vert _\infty \le R\), for \(u\in K\). Thus, we obtain

$$\begin{aligned} \Vert T_\lambda ^{*}(u)\Vert _\infty \le \Vert u\Vert _\infty , \end{aligned}$$
(30)

for \(u\in K\cap \partial B_R.\) Since \(f_0=\infty \), we can easily get

$$\begin{aligned} f_0^{*}:= \lim _{u\rightarrow 0}\frac{f^*(u)}{\varphi (u)}=\lim _{u\rightarrow 0}\frac{f(u)}{\varphi (u)}=f_0=\infty . \end{aligned}$$

For fixed \(\lambda \in (0,{\hat{\lambda }})\), we take \(M=\frac{\gamma (32)}{\lambda \varGamma }>0\). Then, by the fact that \(f_0^{*}=\infty \), there exists \(r_M^{*}>0\) such that

$$\begin{aligned} f^{*}(x)\ge M\varphi (x), \end{aligned}$$

for \(0\le x\le r_M^{*}.\) Take \(0<r<\min \{R,r_M^{*}\}\). Then, for \(u\in K\cap \partial B_r\), we get

$$\begin{aligned} 0\le u(t)\le \Vert u\Vert _\infty =r<r_M^{*}, \end{aligned}$$

and

$$\begin{aligned} f^{*}(u(t))\ge M\varphi (u(t))\ge M\varphi (\frac{1}{4}\Vert u\Vert _\infty ), \end{aligned}$$

for \(t\in [\frac{1}{4},\frac{3}{4}].\) Applying the same argument as in the proof of Lemma 8, we can get

$$\begin{aligned} \Vert T_\lambda ^{*}(u)\Vert _\infty \ge \Vert u\Vert _\infty ,\quad \ \mathrm{for}\quad \ u\in K\cap \partial B_r. \end{aligned}$$
(31)

Combining (30) and (31), we conclude that (29) has at least one solution u with \(r\le \Vert u\Vert _\infty \le R\). This implies that u must be a positive solution.

Next, we show that if u is a solution of (29), then \(0\le u(t)\le {\hat{u}}(t)\), \(t\in [0,1]\).

If it is true, then (29) and (1) are equivalent and the proof can be done. Clearly, \(u(t)\ge 0\), for \(t\in [0,1]\). Applying arguments under to there in the proof of Lemma 6, we can show that \(u(t)\le {\hat{u}}(t)\), \(t\in [0,1]\). \(\square \)

Lemma 11

Assume that (A), (H), (\(F_1\)) and \(f_\infty =\infty \) hold. Let I be a compact interval of \((0,\infty )\). Then, there exists a constant \(b_I>0\) such that all possible solutions u of (1) with \(\lambda \in I\) satisfy \(\Vert u\Vert _{\infty }<b_I\).

Proof

Suppose to the contrary that there exists a sequence \(\{u_n\}\) of positive solutions of (1) at \(\lambda =\lambda _n\) with \(\{\lambda _n\}\subset I=[\alpha ,\beta ]\subset (0,\infty )\) and \(\Vert u_n\Vert _\infty \rightarrow \infty \) as \(n\rightarrow \infty \). Take \(M=\frac{2\gamma (32)}{\alpha \varGamma }\). Since \(f_\infty =\infty \), there exists \(R_M>0\) such that for \(x\ge R_M\), we have

$$\begin{aligned} f(x)\ge M\varphi (x). \end{aligned}$$

From the assumption, we can get \(\Vert u_n\Vert _\infty \ge 4R_M\), for sufficiently large n. Thus, by Lemma 3, we have

$$\begin{aligned} u_n(t)\ge \min _{t\in [\frac{1}{4},\frac{3}{4}]}u_n(t)\ge \frac{1}{4}\Vert u_n\Vert _\infty \ge R_M, \end{aligned}$$

and

$$\begin{aligned} f(u_n(t))\ge M\varphi (u_n(t))\ge M\varphi (\frac{1}{4}\Vert u_n\Vert _\infty ), \end{aligned}$$
(32)

for \(t\in [\frac{1}{4},\frac{3}{4}]\) and sufficiently large n. Since \(u_n=T_{\lambda _n}(u_n)\), applying the same argument in Lemma 4 with aid of (32) and by the definition of M and Remark 2 with \(\sigma =32\), \(x=\frac{1}{4}\Vert u_n\Vert _\infty \), we get for \(\lambda _n\in I\) with sufficiently large n,

$$\begin{aligned} \Vert u_n\Vert _\infty= & {} \Vert T_{\lambda _n}(u_n)\Vert _\infty \ge \frac{1}{8}\varphi ^{-1}\left( \lambda _n M\varphi (\frac{1}{4}\Vert u_n\Vert _\infty )\varGamma \right) \\\ge & {} \frac{1}{8}\varphi ^{-1}\left( \alpha M\varphi (\frac{1}{4}\Vert u_n\Vert _\infty )\varGamma \right) = \frac{1}{8}\varphi ^{-1}\left( 2\gamma (32)\varphi (\frac{1}{4}\Vert u_n\Vert _\infty )\right) \\> & {} \frac{1}{8}\varphi ^{-1}\left( \gamma (32)\varphi (\frac{1}{4}\Vert u_n\Vert _\infty )\right) \ge \frac{1}{8} \times 32 \times \frac{1}{4}\Vert u_n\Vert _\infty =\Vert u_n\Vert _\infty . \end{aligned}$$

This is a contradiction. \(\square \)

Proof of Theorem 2

Define

$$\begin{aligned}&\lambda ^*:=\sup \{\lambda \ | \ (1) \ \mathrm{has \ at \ least \ one \ positive \ solution}\}. \end{aligned}$$
(33)
$$\begin{aligned}&\lambda _*:=\sup \{{\tilde{\lambda }} \ | \ (1) \ \mathrm{has \ at \ least \ two \ positive \ solutions \ for} \ \lambda \in (0,{\tilde{\lambda }})\}. \end{aligned}$$
(34)

By Lemmas 8 and 9, both \(\lambda _*\) and \(\lambda ^*\) are well-defined and \(0<\lambda _*\le \lambda ^*\le {\bar{\lambda }}\). By the definitions of \(\lambda _*\), \(\lambda ^*\) and by Lemma 10, we conclude that (1) has at least two positive solutions for \(\lambda \in (0,\lambda _*)\), one positive solution for \(\lambda \in [\lambda _*,\lambda ^*)\), and no positive solution for \(\lambda >\lambda ^*\). Finally, it is enough to show that (1) has at least one positive solution at \(\lambda =\lambda ^*\). By the definition of \(\lambda ^*\) and Lemma 9, we can choose a sequence \(\{\lambda _n\}\) with \(\frac{\lambda ^*}{2}\le \lambda _n< \lambda ^*\le {\bar{\lambda }}\) such that \(\lambda _n \rightarrow \lambda ^*\) as \(n\rightarrow \infty \). By Lemma 11 and applying similar argument in the proof of Theorem 1, we have (1) has a solution \(u^*\) at \(\lambda =\lambda ^*\). Moreover, using the fact \(f_0=\infty \) and applying similar argument to there in the proof of Lemma 11, we see that \(u^*\) is a positive solution of (1) at \(\lambda =\lambda ^*\). \(\square \)

4 Applications

In this section, we give some examples illustrating our main results.

Example 4.1

Consider the following scalar \(\varphi \)-Laplacian problem

$$\begin{aligned} \left\{ \begin{array}{ll} \varphi (u')'+ \lambda t^{-\frac{3}{2}}f(u)=0,\quad t\in (0,1),\\ u(0)= 0 = u(1), \end{array} \right. \end{aligned}$$
(35)

where \(\varphi (x)=|x|x+x\), \(x\in {\mathbb {R}}\), and

$$\begin{aligned} f(u) = \left\{ \begin{array}{ll} u^{3}, &{}\quad \mathrm{if}\ 0\le u<1,\\ u,&{}\quad \mathrm{if}\ u\ge 1. \end{array} \right. \end{aligned}$$

We easily see that \(\varphi \) is an odd increasing homeomorphism. Define functions \(\psi \) and \(\gamma \) by

$$\begin{aligned} \psi (\sigma ) =\left\{ \begin{array}{ll} \sigma ^{2}, &{}\quad \mathrm{if}\ 0<\sigma \le 1,\\ \sigma , &{}\quad \mathrm{if}\ \sigma >1, \end{array} \right. \end{aligned}$$

and

$$\begin{aligned} \gamma (\sigma ) =\left\{ \begin{array}{ll} 1, &{}\quad \mathrm{if} \ 0<\sigma \le 1,\\ \sigma ^{2}, &{}\quad \mathrm{if}\ \sigma >1. \end{array} \right. \end{aligned}$$

Then, \(\psi , \gamma :(0,\infty )\rightarrow (0,\infty )\) and \(\psi \) is an increasing homeomorphism with

$$\begin{aligned} \psi ^{-1}(\sigma ) =\left\{ \begin{array}{ll} \sigma ^{\frac{1}{2}}, &{}\quad \mathrm{if}\ 0<\sigma \le 1,\\ \sigma , &{}\quad \mathrm{if}\ \sigma >1. \end{array} \right. \end{aligned}$$

We may see that (35) satisfies assumptions \((A), \ (H), \ (F_1)\) and \((F_2)\) (see Xu and Lee [17] for details) and exactly obtain

$$\begin{aligned} f_0= & {} \lim _{u\rightarrow 0}\frac{f(u)}{\varphi (u)}=\lim _{u\rightarrow 0}\frac{u^{3}}{u^{2}+u}=0,\\ f_\infty= & {} \lim _{u\rightarrow \infty }\frac{f(u)}{\varphi (u)}=\lim _{u\rightarrow \infty }\frac{u}{u^{2}+u}=0. \end{aligned}$$

Consequently, by Theorem 1, we see that there exists \(\lambda ^{*}\ge \lambda _{*}>0\) such that (35) has at least two positive solutions for \(\lambda >\lambda ^{*}\), one positive solution for \(\lambda \in [\lambda _*,\lambda ^{*}]\), and no positive solution for \(\lambda \in (0,\lambda _*)\).

Example 4.2

Consider the following scalar \(\varphi \)-Laplacian problem

$$\begin{aligned} \left\{ \begin{array}{ll} \varphi (u')'+ \lambda t^{-\frac{5}{4}}f(u)=0,\quad t\in (0,1),\\ u(0)= 0 = u(1), \end{array} \right. \end{aligned}$$
(36)

where \(\varphi (x)=x^{\frac{1}{3}}\), \(x\in {\mathbb {R}}\), and

$$\begin{aligned} f(u) = \left\{ \begin{array}{ll} u^{\frac{1}{4}}, &{}\quad \mathrm{if}\ 0\le u<1,\\ u^{2}, &{}\quad \mathrm{if}\ u\ge 1. \end{array} \right. \end{aligned}$$

Then, \(\varphi \) is an odd increasing homeomorphism. By the homogeneity of \(\varphi \), taking \(\psi (\sigma )=\gamma (\sigma )\equiv \varphi (\sigma )\). We can easily check that (36) satisfies the assumptions \((A), \ (H), \ (F_1)\) and \((F_2)\) (see Xu and Lee [17] for details) and exactly obtain

$$\begin{aligned} f_0= & {} \lim _{u\rightarrow 0}\frac{f(u)}{\varphi (u)}=\lim _{u\rightarrow 0}\frac{u^{\frac{1}{4}}}{u^{\frac{1}{3}}}=\lim _{u\rightarrow 0}u^{-\frac{1}{12}}=\infty ,\\ f_\infty= & {} \lim _{u\rightarrow \infty }\frac{f(u)}{\varphi (u)}=\lim _{u\rightarrow \infty }\frac{u^{2}}{u^{\frac{1}{3}}}=\lim _{u\rightarrow \infty }u^{\frac{5}{3}}=\infty . \end{aligned}$$

Consequently, by Theorem 2, we see that there exists \(\lambda ^{*}\ge \lambda _{*}>0\) such that (36) has at least two positive solutions for \(\lambda \in (0,\lambda _*)\), one positive solution for \(\lambda \in [\lambda _*,\lambda ^{*}]\), and no positive solution for \(\lambda >\lambda ^{*}\).