1 Introduction

This paper is devoted to the existence of positive solutions for the following third-order nonlocal boundary value problem (BVP for short):

{ x ( t ) + f ( t , x ( t ) ) = 0 , 0 < t < 1 , x ( 0 ) = i = 1 m 2 α i x ( ξ i ) , x ( 0 ) = x ( 1 ) = 0 ,
(1.1)

where 0< ξ 1 < ξ 2 << ξ m 2 <1, α i R + (i=1,2,,m2) and i = 1 m 2 α i =1. The problem (1.1) happens to be at resonance in the sense that the associated linear homogeneous BVP

{ x ( t ) = 0 , 0 < t < 1 , x ( 0 ) = i = 1 m 2 α i x ( ξ i ) , x ( 0 ) = x ( 1 ) = 0 ,
(1.2)

has nontrivial solutions. Clearly, the resonant condition is i = 1 m 2 α i =1. Third-order differential equations arise in a variety of different areas of applied mathematics and physics, e.g., in the deflection of a curved beam having a constant or varying cross section, a three-layer beam, electromagnetic waves or gravity-driven flows and so on [1].

Recently, the existence of positive solutions for third-order two-point or multi-point BVPs has received considerable attention; we mention a few works: [211] and the references therein. However, all of the papers on third-order BVPs focused their attention on the positive solutions with non-resonance cases. It is well known that the problem of the existence of positive solutions to BVPs is very difficult when the resonant case is considered. Only few papers deal with the existence of positive solutions to BVPs at resonance, and just to second-order BVPs [1215]. It is worth mentioning that Infante and Zima [13] studied the existence of positive solutions for the second-order m-point BVP

{ x ( t ) + f ( t , x ( t ) ) = 0 , 0 < t < 1 , x ( 0 ) = 0 , x ( 1 ) = i = 1 m 2 α i x ( η i )
(1.3)

by means of the Leggett-Williams norm-type theorem due to O’Regan and Zima [16], where 0< η 1 < η 2 << η m 2 <1, α i R + (i=1,2,,m2) and i = 1 m 2 α i =1.

However, third-order or higher-order derivatives do not have the convexity; to the best of our knowledge, no results are available for the existence of positive solutions for third-order or higher- order BVPs at resonance. The main purpose of this paper is to fill the gap in this area. Motivated greatly by the above-mentioned excellent works, in this paper we will investigate the third-order nonlocal BVP (1.1) at resonance, where 0< ξ 1 < ξ 2 << ξ m 2 <1, α i R + (i=1,2,,m2) and i = 1 m 2 α i =1. Some new existence results of at least one positive solution are established by applying the Leggett-Williams norm-type theorem due to O’Regan and Zima [16]. An example is also included to illustrate the main results.

2 Some definitions and a fixed point theorem

For the convenience of the reader, we present here the necessary definitions and a new fixed point theorem due to O’Regan and Zima.

Definition 2.1 Let X and Z be real Banach spaces. A linear operator L:domLXZ is called a Fredholm operator if the following two conditions hold:

  1. (i)

    KerL has a finite dimension, and

  2. (ii)

    ImL is closed and has a finite codimension.

Throughout the paper, we will assume that

1L is a Fredholm operator of index zero, that is, ImL is closed and dimKerL=codimImL<.

From Definition 2.1, it follows that there exist continuous projectors P:XX and Q:ZZ such that

ImP=KerL;KerQ=ImL;X=KerLKerP;Z=ImLImQ

and that the isomorphism

L | dom L Ker P :domLKerPImL

is invertible. We denote the inverse of L | dom L Ker P by K P :ImLdomLKerP. The generalized inverse of L denoted by K P , Q :ZdomLKerP is defined by K P , Q = K P (IQ). Moreover, since dimImQ=codimImL, there exists an isomorphism J:ImQKerL. Consider a nonlinear operator N:XZ. It is known (see [17, 18]) that the coincidence equation Lx=Nx is equivalent to

x=(P+JQN)x+ K P , Q Nx.

Definition 2.2 Let X be a real Banach space. A nonempty closed convex set P is said to be a cone provided that

  1. (1)

    λxP for all xP and λ0, and

  2. (2)

    x,xP implies x=θ.

Note that every cone PX induces a partial order ≤ in X by defining xy if and only if yxP. The following property is valid for every cone in a Banach space.

Lemma 2.3 ([19])

Let P be a cone in X. Then for everyuP{θ}, there exists a positive numberσ(u)such thatx+uσ(u)xfor allxP.

Let γ:XP be a retraction, that is, a continuous mapping such that γ(x)=x for all xP. Set

Ψ=P+JQN+ K P , Q N

and

Ψ γ =Ψγ.

Our main results are based on the following theorem due to O’Regan and Zima.

Theorem 2.4 ([16])

Let C be a cone in X and let Ω 1 , Ω 2 be open bounded subsets of X with Ω ¯ 1 Ω 2 andC( Ω ¯ 2 Ω 1 )θ. Assume that 1is satisfied and if the following assumptions hold:

(H1) QN:XZis continuous and bounded, and K P , Q :XXis compact on every bounded subset of X;

(H2) LxλNxfor allxC Ω 2 domLandλ(0,1);

(H3) γ maps subsets of Ω ¯ 2 into bounded subsets of C;

(H4) d B ([I(P+JQN)γ] | Ker L ,KerL Ω 2 ,θ)0, where d B stands for the Brouwer degree;

(H5) there exists u 0 C{θ}such thatxσ( u 0 )Ψ(x)forxC( u 0 ) Ω 1 , whereC( u 0 )={xC:μ u 0 xfor some μ>0}andσ( u 0 )is such thatx+ u 0 σ( u 0 )xfor allxC;

(H6) (P+JQN)γ( Ω 2 )C;

(H7) Ψ γ ( Ω ¯ 2 Ω 1 )C,

then the equationLx=Nxhas a solution in the setC( Ω ¯ 2 Ω 1 ).

3 Main results

For simplicity of notation, we set

h i (s)={ 2 ξ i s ξ i 2 s s 2 2 , 0 s ξ i , ξ i 2 ( 1 s ) 2 , ξ i s 1 ,

where i=1,2,,m2, and

G(t,s)={ ( 1 s ) ( s 2 2 s + 3 t 2 ) 6 ( t s ) 2 2 + 4 t 3 6 t 2 + 23 2 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i h i ( s ) , 0 s t 1 , ( 1 s ) ( s 2 2 s + 3 t 2 ) 6 + 4 t 3 6 t 2 + 23 2 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i h i ( s ) , 0 t s 1 .

It is easy to check that G(t,s)0, t,s[0,1], and since 0 h i (s) 1 2 s(1s), we get

1 12 δ i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i h i (s)0,s[0,1]

for every δ(0, 2 ξ 1 2 ( 3 2 ξ 1 ) 3 ]. We also let δ:=min{ 2 ξ 1 2 ( 3 2 ξ 1 ) 3 , 1 max t , s [ 0 , 1 ] G ( t , s ) ,1}.

We can now state our result on the existence of a positive solution for the BVP (1.1).

Theorem 3.1 Assume that f:[0,1]×[0,+)R is continuous and

  1. (1)

    there exists a constant M(0,) such that f(t,x)>δx for all (t,x)[0,1]×[0,M];

  2. (2)

    there exist r(0,M), t 0 [0,1], a(0,1], b(0,1) and continuous functions g:[0,1][0,), h:(0,r][0,) such that f(t,x)g(t)h(x) for (t,x)[0,1]×(0,r], h ( x ) x a is non-increasing on (0,r] with

    h ( r ) ( r ) a 0 1 G( t 0 ,s)g(s)ds 1 b b a .
  3. (3)

    f = lim inf x + min t [ 0 , 1 ] f ( t , x ) x <0.

Then the resonant BVP (1.1) has at least one positive solution on[0,1].

Proof Consider the Banach spaces X=Z=C[0,1] with x= max t [ 0 , 1 ] |x(t)|.

Let L:domLXZ and N:XZ with

domL= { x X : x C [ 0 , 1 ] , x ( 0 ) = i = 1 m 2 α i x ( ξ i ) , x ( 0 ) = x ( 1 ) = 0 }

be given by (Lx)(t)= x (t) and (Nx)(t)=f(t,x(t)) for t[0,1]. Then

KerL= { x dom L : x ( t ) c  on  [ 0 , 1 ] }

and

ImL= { y Z : i = 1 m 2 α i 0 1 h i ( s ) y ( s ) d s = 0 } .

Clearly, dimKerL=1 and ImL is closed. It follows from Z 1 =ZImL that

Z 1 = { y 1 Z : y 1 = 12 2 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i ( s ) y ( s ) d s , y Z } .

In fact, for each yZ, we have

i = 1 m 2 α i 0 1 h i ( s ) ( y ( s ) y 1 ( s ) ) d s = ( 1 12 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i ( s ) d s ) i = 1 m 2 α i 0 1 h i ( s ) y ( s ) d s = 0 ,

which shows that y y 1 ImL, which together with Z 1 ImL={θ} implies that Z= Z 1 ImL. Note that dim Z 1 =1 and thus codimImL=1. Therefore, L is a Fredholm operator of index zero.

Next, define the projections P:XX by

Px= 0 1 x(s)ds

and Q:ZZ by

Qy= 12 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i (s)y(s)ds.

Clearly, ImP=KerL, KerQ=ImL and KerL={xX: 0 1 x(s)ds=0}. Note that for yImL, the inverse K P of L P is given by

( K P y)(t)= 0 1 k(t,s)y(s)ds,

where

k(t,s)={ ( 1 s ) ( s 2 2 s + 3 t 2 ) 6 ( t s ) 2 2 , 0 s t 1 , ( 1 s ) ( s 2 2 s + 3 t 2 ) 6 , 0 t s 1 .

Considering that f can be extended continuously to [0,1]×R, it is easy to check that QN:XZ is continuous and bounded, and K P , Q N:XX is compact on every bounded subset of X, which ensures that (H1) of Theorem 2.4 is fulfilled.

Define the cone of nonnegative functions

C= { x X : x ( t ) 0 ,  on  [ 0 , 1 ] } .

Let

Ω 1 = { x X : b x < | x ( t ) | < r , t [ 0 , 1 ] }

and

Ω 2 = { x X : x < M } .

Clearly, Ω 1 and Ω 2 are bounded and open sets and

Ω ¯ 1 = { x X : b x | x ( t ) | r , t [ 0 , 1 ] } Ω 2 .

Moreover, C( Ω ¯ 2 Ω 1 )θ. Let J=I and (γx)(t)=|x(t)| for xX. Then γ is a retraction and maps subsets of Ω ¯ 2 into bounded subsets of C, which means that (H3) of Theorem 2.4 holds.

Let 0< r =min{ r 1 , r 2 }, where r 1 and r 2 will be defined in the following proof.

Suppose that there exist x 0 C Ω 2 domL and λ 0 (0,1) such that L x 0 = λ 0 N x 0 . Then

x 0 (t)+ λ 0 f ( t , x 0 ( t ) ) =0.

Let x 0 ( t 1 )= x 0 = max t [ 0 , 1 ] | x 0 (t)|=M. Now, we verify that t 1 0 and t 1 1.

First, we show t 1 0. Suppose, on the contrary, that x 0 (t) achieves maximum value M only at t=0. Then x 0 (0)= i = 1 m 2 α i x 0 ( ξ i ) in combination with i = 1 m 2 α i =1 yields that max 1 i m 2 { x 0 ( ξ i )}M, which is a contradiction.

Next, we show t 1 1. It follows from x 0 (0)= x 0 (1)=0 that there is a constant η(0,1) such that x 0 (η)=0, and thus x 0 (t)= λ 0 η t f(s, x 0 (s))ds. By the condition (3) we have, for x r (M r >r>0), there exists σ(0,δ) such that

f(t,x)σx,t[0,1].
(3.1)

Suppose, on the contrary, that x 0 (1)=M. The step is divided into two cases:

Case 1. Assume that x 0 (t)>0 on [η,1]. Let r 1 = min t [ η , 1 ] x 0 (t). Then (3.1) yields

x 0 (1)= λ 0 η 1 f ( s , x 0 ( s ) ) ds λ 0 σ η 1 x 0 (s)ds>0,

which implies x 0 (t) is increasing close to 1. This together with x 0 (1)=0 induces x 0 (t)<0 (t close to 1), that is, x 0 (t) is decreasing close to 1, which contradicts x 0 (1)=M.

Case 2. Assume that x 0 (t) has zero points on [η,1]; we may choose η nearest to 1 with x 0 ( η )=0. Then there is a constant η( η ,1) such that x 0 (η)=0. Similar to the above arguments, we easily get a contradiction too.

Hence, we can choose t 1 (0,1) so that x 0 ( t 1 )=M. This gives x 0 ( t 1 )=0 and x 0 ( t 1 )0. By x 0 domL, we know x 0 (t)0 and x 0 (0)=0. Similarly, we also divide the part of the proof into two cases.

Case 1. If x 0 (t)>0 on [0, t 1 ], then there is a constant η(0, t 1 ) such that x 0 (η)=0. Thus we have

x 0 (t)= λ 0 η t f ( s , x 0 ( s ) ) ds.

Let r 2 = min t [ η , t 1 ] x 0 (t). Then it follows from (3.1) that

0 x 0 ( t 1 )= λ 0 η t 1 f ( s , x 0 ( s ) ) ds λ 0 σ η t 1 x 0 (s)ds>0,

which is a contradiction.

Case 2. If x 0 (t) has zero points on [0, t 1 ], we may choose η nearest to t 1 with x 0 ( η )=0. Then there is a constant η( η , t 1 ) such that x 0 (η)=0. Thus we have

x 0 (t)= λ 0 η t f ( s , x 0 ( s ) ) ds.

Let r 2 = min t [ η , t 1 ] x 0 (t). Then it follows from (3.1) that

0 x 0 ( t 1 )= λ 0 η t 1 f ( s , x 0 ( s ) ) ds λ 0 σ η t 1 x 0 (s)ds>0,

which is a contradiction. Therefore, (H2) of Theorem 2.4 holds.

Consider xKerL Ω 2 . Then x(t)c on [0,1]. Similar to [13], we define

H ( x , λ ) = x λ | x | 12 λ i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i ( s ) f ( s , | x | ) d s , x Ker L Ω 2  and  λ [ 0 , 1 ] .

Suppose H(x,λ)=0. Then in view of (1), we obtain

c = λ c + 12 λ i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i ( s ) f ( s , | c | ) d s λ c 12 λ i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i ( s ) δ | c | d s λ | c | ( 1 δ ) 0 .

Hence, H(x,λ)=0 implies c0. Furthermore, if H(M,λ)=0, then we have

0M(1λ)= 12 λ i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i (s)f(s,M)ds,

contradicting (3.1). Thus H(x,λ)=0 for x Ω 2 and λ[0,1]. Therefore,

d B ( H ( x , 0 ) , Ker L Ω 2 , θ ) = d B ( H ( x , 1 ) , Ker L Ω 2 , θ ) .

However,

d B ( H ( x , 0 ) , Ker L Ω 2 , θ ) = d B (I,KerL Ω 2 ,θ).

This gives

d B ( [ I ( P + J Q N ) γ ] | Ker L , Ker L Ω 2 , θ ) = d B ( H ( x , 1 ) , Ker L Ω 2 , θ ) 0,

which shows that (H4) of Theorem 2.4 holds.

Let x Ω ¯ 2 Ω 1 and t[0,1]. Then

From (1), we know that

( Ψ γ ) ( t ) 0 1 | x ( s ) | d s δ 0 1 G ( t , s ) | x ( s ) | d s = 0 1 ( 1 δ G ( t , s ) ) | x ( s ) | d s 0 .

Hence Ψ γ ( Ω 2 ¯ Ω 1 )C. Moreover, for x Ω 2 , we have

( P + J Q N ) γ x = 0 1 | x ( s ) | d s + 12 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i 0 1 h i ( s ) f ( s , | x ( s ) | ) d s 0 1 [ 1 12 δ 2 i = 1 m 2 α i ( 3 ξ i 2 2 ξ i 3 ) i = 1 m 2 α i h i ( s ) ] | x ( s ) | d s 0 ,

which shows that (P+JQN)γ( Ω 2 )P. These ensure that (H6), (H7) of Theorem 2.4 hold. It remains to show that (H5) is satisfied.

Taking u 0 (t)1 on [0,1] and σ( u 0 )=1, we confirm that

u 0 C{θ},C( u 0 )= { x C : x ( t ) > 0  on  [ 0 , 1 ] } .

Let xC( u 0 ) Ω 1 . Then we have x(t)>0, t[0,1], 0<x<r and x(t)bx, t[0,1]. Therefore, in view of (2), for all xC( u 0 ) Ω 1 , we obtain

( Ψ x ) ( t 0 ) = 0 1 x ( s ) d s + 0 1 G ( t 0 , s ) f ( s , x ( s ) ) d s b x + 0 1 G ( t 0 , s ) g ( s ) h ( x ( s ) ) d s = b x + 0 1 G ( t 0 , s ) g ( s ) h ( x ( s ) ) x a ( s ) x a ( s ) d s b x + h ( r ) ( r ) a 0 1 G ( t 0 , s ) g ( s ) x a ( s ) d s b x + h ( r ) ( r ) a 0 1 G ( t 0 , s ) g ( s ) b a x a d s b x + ( 1 b ) x = x .

That is, xσ( u 0 )Ψx for all xC( u 0 ) Ω 1 , which shows that (H5) of Theorem 2.4 holds.

Summing up, all the hypotheses of Theorem 2.4 are satisfied. Therefore, the equation Lx=Nx has a solution xC( Ω ¯ 2 Ω 1 ). And so, the resonant BVP (1.1) has at least one positive solution on [0,1]. □

4 An example

Consider the BVP

{ x ( t ) + f ( t , x ( t ) ) = 0 , 0 < t < 1 , x ( 0 ) = x ( 1 2 ) , x ( 0 ) = x ( 1 ) = 0 .
(4.1)

Here α 1 =1, ξ 1 = 1 2 , g(t)= 1 4 ( t 2 t1), h(x)= x + 1 and

f(t,x)={ 1 4 ( t 2 t 1 ) ( x 2 8 x + 15 ) x + 1 , t [ 0 , 1 ] , x [ 0 , 2 ] , 1 4 ( t 2 t 1 ) ( 2 1 2 x 1 ) x + 1 , t [ 0 , 1 ] , x ( 2 , + ) .

By a simple computation, we get

δ= 1 3 , 1 4 g(t) 5 16 ,t[0,1], x 2 8x+15x,x[0,3].

We may choose M=3, r= 5 2 , t 0 =0, a=1, b= 13 14 . It is easy to check

  1. (1)

    f(t,x)> 1 3 x for all (t,x)[0,1]×[0,3];

  2. (2)

    f(t,x)g(t)h(x) for (t,x)[0,1]×(0, 5 2 ], x + 1 x is non-increasing on (0, 5 2 ] with

    h ( r ) ( r ) a 0 1 G(0,s)g(s)ds 7 14 320 1 13 = 1 b b a ;
  3. (3)

    f = lim inf x + min t [ 0 , 1 ] f ( t , x ) x = 1 8 .

Thus, all the conditions of Theorem 3.1 are satisfied. Then the resonant problem (4.1) has at least one positive solution on [0,1].