Abstract

Let 𝑇>5 be an integer, 𝕋={1,2,,𝑇}. We are concerned with the global structure of positive solutions set of the discrete second-order boundary value problems Δ2𝑢(𝑡1)+𝑟𝑚(𝑡)𝑓(𝑢(𝑡))=0,𝑡𝕋,𝑢(0)=𝑢(𝑇+1)=0, where 𝑟 is a parameter, 𝑚𝕋 changes its sign and 𝑚(𝑡)0for𝑡𝕋.

1. Introduction

Let 𝑇>5 be an integer, 𝕋={1,2,,𝑇}. In this paper, we are concerned with the global structure of positive solutions set of the discrete second-order boundary value problems Δ2𝑢(𝑡1)+𝑟𝑚(𝑡)𝑓(𝑢(𝑡))=0,𝑡𝕋,(1.1)𝑢(0)=𝑢(𝑇+1)=0,(1.2) where 𝑟 is a parameter, 𝑚𝕋 changes its sign, and 𝑚(𝑡)0 for 𝑡𝕋.

The boundary value problems with sign-changing weight arise from a selection-migration model in population genetics, see Fleming [1]. 𝑚 changes sign corresponds to the fact that an allele 𝐴1 holds an advantage over a rival allele 𝐴2 at the same points and is at a disadvantage at others. The parameter 𝑟 corresponds to the reciprocal of the diffusion. So, the existence and multiplicity of positive solutions with sign-changing weight in continuous case has been studied by many authors, see, for example, [28] and the references therein.

For the discrete case, there are many literature dealing with difference equations similar to (1.1) subject to various boundary value conditions. We refer to [917] and the references therein. However, there are few papers to discuss the existence of positive solutions to (1.1) and (1.2) if 𝑚(𝑡) changes sign on 𝕋. Maybe the main reason is the spectrum of the following linear eigenvalue problems Δ2𝑢(𝑡1)+𝜆𝑚(𝑡)𝑢(𝑡)=0,𝑡𝕋,𝑢(0)=𝑢(𝑇+1)=0(1.3) is not clear when 𝑚 changes its sign on 𝕋.

In 2008, Shi and Yan [18] investigated the spectrum of the second-order boundary value problems of difference equations [],Δ𝑝(𝑡1)Δ𝑢(𝑡1)+𝑞(𝑡)𝑢(𝑡)=𝜆𝑚(𝑡)𝑢(𝑡),𝑡𝕋,𝑝(0)Δ𝑢(0)𝑝(𝑇)Δ𝑢(𝑇)=𝐾𝑢(0)𝑢(𝑇)(1.4) where 𝐾=𝑘11𝑘12𝑘21𝑘22 is a symmetric and positive definite 2×2 matrix. 𝑝{0,1,,𝑇}, 𝑞𝕋, and 𝑚𝕋 satisfy(A1)𝑚𝕋 changes its sign, and 𝑚(𝑡)0 for 𝑡𝕋.(A2)𝑝(𝑡1)0,𝑞(𝑡)0,𝑝(𝑡1)+𝑞(𝑡)>0,for𝑡𝕋, and 𝑝(0)+𝑘110, 𝑝(𝑇)>0.

They proved that (1.4) has 𝑇 eigenvalues 𝜆1𝜆2𝜆𝑇 which have 𝑇 corresponding linearly independent and orthogonal eigenfunctions.

However, it's easy to see that (1.3) is not included in (1.4) under the condition (A2). Also, Shi and Yan [18] provided no information about the sign of the eigenvalues and no information about the corresponding eigenfunctions.

In this paper, we will show that (1.3) has two principal eigenvalues 𝜆𝑚,<0<𝜆𝑚,+, and the corresponding eigenfunctions we denote by 𝜓𝑚, and 𝜓𝑚,+ don't change their sign on 𝕋={0,1,,𝑇,𝑇+1}. Based on this result, using Rabinowitz's global bifurcation theorem [19], we will discuss the global structure of positive solutions set of (1.1) and (1.2).

The assumptions we are interested in this paper are as follows:(H1)𝑓𝐶(,) with 𝑠𝑓(𝑠)>0 for 𝑠0;(H2)𝑓0=lim|𝑠|0(𝑓(𝑠)/𝑠)(0,);(H3)𝑓=lim|𝑠|+(𝑓(𝑠)/𝑠)=0;(H3)𝑓=.

Our main result is the following.

Theorem 1.1. Assume that (H1), (H2) hold. (i)If (H3) holds, then there exist 0<𝜆𝜆𝑚,+/𝑓0 and 𝜆𝑚,/𝑓0𝜆<0 such that (1.1), (1.2) has at least one positive solution for 𝑟(,𝜆)(𝜆,).(ii)If (H3) holds, then there exist 𝜌𝜆𝑚,/𝑓0<0 and 𝜌𝜆𝑚,+/𝑓0>0 such that (1.1), (1.2) has at least one positive solution for 𝑟(𝜌,0)(0,𝜌).

The rest of the paper is arranged as follows. In Section 2, the existence of two principal eigenvalues of (1.3), and some properties of these two eigenvalues will be discussed. In Section 3, we will prove our main result.

2. Existence of Two Principal Eigenvalues to (1.3)

Recall that 𝕋={1,2,,𝑇} and 𝕋={0,1,,𝑇+1}. Let 𝑋={𝑢𝕋𝑢(0)=𝑢(𝑇+1)=0}. Then 𝑋 is a Banach space under the norm 𝑢𝑋=max𝑡𝕋|𝑢(𝑡)|. Let 𝑌={𝑢𝑢𝕋}. Then 𝑌 is a Banach space under the norm 𝑢𝑌=max𝑡𝕋|𝑢(𝑡)|.

It's well known that the operator 𝜒𝑋𝑌, 𝜒(0,𝑢(1),𝑢(2),,𝑢(𝑇),0)=(𝑢(1),𝑢(2),,𝑢(𝑇))(2.1) is a homomorphism.

Define the operator 𝐿𝑋𝑌 by 𝐿𝑢(𝑡)=Δ2𝑢(𝑡1),𝑡𝕋.(2.2)

In this section, we will discuss the existence of principal eigenvalues for (1.3) with 𝑚𝕋 that changes its sign. The main idea we will use aries from [20, 21].

Theorem 2.1. Equation (1.3) has two principal eigenvalues 𝜆𝑚,+ and 𝜆𝑚, such that 𝜆𝑚,<0<𝜆𝑚,+, and the corresponding eigenfunctions, we denoted by 𝜓𝑚,+ and 𝜓𝑚, do not change sign on 𝕋.

Proof. Consider, for fixed 𝜆, the eigenvalue problems 𝑢𝐿𝑢𝜆𝑚(𝑡)𝑢(𝑡)=𝜇𝑢(𝑡),𝑡𝕋,(0)=𝑢(𝑇+1)=0.(2.3) By Kelley and Peterson [22, Theorem  7.6], for fixed 𝜆, (2.3) has 𝑇-simple eigenvalues 𝜇𝑚,1(𝜆)<𝜇𝑚,2(𝜆)<<𝜇𝑚,𝑇(𝜆),(2.4) and the corresponding eigenfunction 𝜓𝑚,𝑘(𝜆,𝑡) has exactly 𝑘1 simple generalized zeros.
Thus, 𝜆 is a principal eigenvalue of (1.3) if and only if 𝜇𝑚,1(𝜆)=0.
On the other hand, let 𝑆𝑚,𝜆=𝑇𝑡=0||||Δ𝜙(𝑡)2𝜆𝑇𝑡=1𝑚(𝑡)𝜙(𝑡)2𝜙𝑋,𝑇𝑡=1𝜙(𝑡)2=1.(2.5) Clearly, 𝑆𝑚,𝜆 is bounded below. Then 𝜇𝑚,1(𝜆)=inf𝑆𝑚,𝜆.
For fixed 𝜙𝑋, 𝜆𝑇𝑡=0|Δ𝜙(𝑡)|2𝜆𝑇𝑡=1𝑚(𝑡)𝜙2(𝑡) is an affine and so concave function. As the infimum of any collection of concave functions is concave, it follows that 𝜆𝜇𝑚,1(𝜆) is a concave function. Also, by considering test functions 𝜙1,𝜙2𝑋 such that 𝑇𝑡=1𝑚(𝑡)𝜙21(𝑡)<0 and 𝑇𝑡=1𝑚(𝑡)𝜙22(𝑡)>0, it is easy to see that 𝜇𝑚,1(𝜆) as 𝜆±. Thus, 𝜆𝜇𝑚,1(𝜆) is an increasing function until it attains its maximum and is a decreasing function thereafter.
Since 𝜇1(0)>0, 𝜆𝜇𝑚,1(𝜆) must have exactly two zeros. Thus, (1.3) has exactly two principal eigenvalues, 𝜆𝑚,+>0 and 𝜆𝑚,<0, and the corresponding eigenfunctions we denoted by 𝜓𝑚,+ and 𝜓𝑚, don't change sign on 𝕋.

Remark 2.2. From the proof of Theorem 2.1, it is not difficult to see that the following results hold.(i)If 𝑚(𝑡)0 on 𝕋 and there exist at least one point 𝑡0𝕋 such that 𝑚(𝑡0)>0, then (1.3) has only one principal eigenvalue 𝜆𝑚,+>0.(ii)If 𝑚(𝑡)0 on 𝕋 and there exist at least one point 𝑡0𝕋 such that 𝑚(𝑡0)<0, then (1.3) has only one principal eigenvalue 𝜆𝑚,<0.

Now, we give some properties for the above principal eigenvalue(s).

Theorem 2.3. Let 𝑚𝕋 change its sign. Assume that there exists 𝑚𝕋 such that 𝑚(𝑡)𝑚(𝑡) for 𝑡𝕋. Then the followings hold. (i)If 𝑚 changes sign on 𝕋, then 𝜆𝑚,𝜆𝑚,, 𝜆𝑚,+𝜆𝑚,+;(ii)If 𝑚0, then 0<𝜆𝑚,+𝜆𝑚,+.

Proof. For convenience, we only prove the case (i). It can be seen that for 𝜆>0, 𝑆𝑚,𝜆𝑆𝑚,𝜆, which implies 𝜇𝑚,1(𝜆)𝜇𝑚,1(𝜆) and consequently, 𝜆𝑚,+𝜆𝑚,+.
On the other hand, for 𝜆<0, 𝑆𝑚,𝜆𝑆𝑚,𝜆, which implies 𝜇𝑚,1(𝜆)𝜇𝑚,1(𝜆) and consequently, 𝜆𝑚,𝜆𝑚,.

Suppose that 𝕋0={𝑎+1,𝑎+2,,𝑏1} is a strict subset of 𝕋, and 𝑚𝕋0 denote the restriction of 𝑚 on 𝕋0. Consider the linear eigenvalue problems Δ2𝑢(𝑡1)+𝜆𝑚(𝑡)𝑢(𝑡)=0,𝑡𝕋0,𝑢(𝑎)=𝑢(𝑏)=0.(2.6)

Then we get the following result.

Theorem 2.4. (i) If 𝑚(𝑡)>0 for 𝑡𝕋0, then (2.6) has only one positive principal eigenvalue 𝜆𝑚𝕋0,+ such that 0<𝜆𝑚,+<𝜆𝑚𝕋0,+.
(ii) If 𝑚(𝑡)<0 for 𝑡𝕋0, then (2.6) has only one principal eigenvalue 𝜆𝑚𝕋0, such that 𝜆𝑚𝕋0,<𝜆𝑚,.
(iii) If 𝑚 changes its sign on 𝕋0, then (2.6) has two principal eigenvalue 𝜆𝑚𝕋0,<0 and 𝜆𝑚𝕋0,+>0 such that 𝜆𝑚,+<𝜆𝑚𝕋0,+,𝜆𝑚𝕋0,<𝜆𝑚,.(2.7)

Proof. Consider the following problems: Δ2𝑢(𝑡1)𝜆𝑚(𝑡)𝑢(𝑡)=𝜇𝑢(𝑡),𝑡𝕋0,𝑢(𝑎)=𝑢(𝑏)=0.(2.8) Let 𝜇𝑚𝕋0,𝑘 denote the 𝑘th eigenvalue of (2.8), and 𝜓𝑚𝕋0,𝑘 the corresponding eigenfunction which has exactly 𝑘1 generalized zeros in (𝑎,𝑏). Let 𝐸={𝑢{𝑎,,𝑏}𝑢(𝑎)=𝑢(𝑏)=0} be a Banach space under the norm 𝑢𝐸=max{|𝑢(𝑡)|𝑡{𝑎,𝑎+1,,𝑏1,𝑏}}. Let 𝑆𝑚𝕋0,𝜆=𝑏1𝑡=𝑎||||Δ𝜙(𝑡)2𝜆𝑏1𝑡=𝑎+1𝑚(𝑡)𝜙(𝑡)2𝜙𝐸,𝑏1𝑡=𝑎+1𝜙(𝑡)2=1.(2.9) We get 𝜇𝑚𝕋0,1=inf𝑆𝑚𝕋0,𝜆. Similar to the proof of Theorem 2.1, we get the following assertions.(i)If 𝑚(𝑡)>0 for 𝑡𝕋0, then (2.6) has only one principal eigenvalue 𝜆𝑚𝕋0,+>0.(ii)If 𝑚(𝑡)<0 for 𝑡𝕋0, then (2.6) has only one negative principal eigenvalue 𝜆𝑚𝕋0,<0.(iii)If 𝑚 changes its sign on 𝕋0, then (2.6) has two principal eigenvalue 𝜆𝑚𝕋0,<0 and 𝜆𝑚𝕋0,+>0.
Now, we prove that the inequalities in (i), (ii), and (iii) hold. For convenience, suppose that 𝑏1𝑡=𝑎+1𝜓2𝑚𝕋0,1(𝜆,𝑡)=1.
Let 𝜓𝑚𝕋0,1 denote the extension of 𝜓𝑚𝕋0,1 by zero on 𝕋, that is, 𝜓𝑚𝕋0,1𝜓(𝜆,𝑡)=𝑚𝕋0,1(𝜆,𝑡),𝑡𝕋0,0,𝑡𝕋𝕋0.(2.10) Then, 𝜇𝑚𝕋0,1(𝜆)=𝑏1𝑡=𝑎|||Δ𝜓𝑚𝕋0,1|||(𝜆,𝑡)2𝜆𝑏1𝑡=𝑎+1𝑚(𝑡)𝜓2𝑚𝕋0,1=(𝜆,𝑡)𝑇𝑡=1|||Δ𝜓𝑚𝕋0,1|||(𝜆,𝑡)2𝜆𝑇𝑡=1𝑚(𝑡)𝜓2𝑚𝕋0,1(𝜆,𝑡)>inf𝑣𝑋𝑇𝑡=1||||Δ𝑣(𝜆,𝑡)2𝜆𝑇𝑡=1𝑚(𝑡)𝑣2(𝜆,𝑡),𝑇𝑡=1𝑣2(𝜆,𝑡)=1=𝜇𝑚,1(𝜆),(2.11) which implies the desired results.

3. The Proof of the Main Result

First, we deal with the case 𝑟>0.

Recall that 𝐿𝑋𝑌, 𝐿𝑢(𝑡)=Δ2𝑢(𝑡1).(3.1) Let 𝜁,𝜉𝐶(,) be such that 𝑓(𝑢)=𝑓0𝑢+𝜁(𝑢),𝑓(𝑢)=𝑓𝑢+𝜉(𝑢).(3.2) Clearly, lim|𝑢|0𝜁(𝑢)𝑢=0,lim|𝑢|𝜉(𝑢)𝑢=0.(3.3) Let ̃𝜉(𝑢)=max0𝑠𝑢||||𝜉(𝑠).(3.4) Then, ̃𝜉 is nondecreasing and lim|𝑢|̃𝜉(𝑢)𝑢=0.(3.5) Let us consider 𝐿𝑢𝜆𝑚(𝑡)𝑟𝑓0𝑢𝜆𝑚(𝑡)𝑟𝜁(𝑢)=0,(3.6) as a bifurcation problem from the trivial solution 𝑢0.

Equation (3.6) can be converted to the equivalent equation 𝑢(𝑡)=𝜆𝐿1𝑚()𝑟𝑓0𝑢()+𝑚()𝑟𝜁(𝑢())(𝑡).(3.7) Further, we note that 𝐿1[𝑚()𝜁(𝑢())]=𝑜(𝑢) for 𝑢 near 0 in 𝑋, since 𝐿1[]𝑚()𝜁(𝑢())=max𝑡𝕋|||||𝑇𝑠=1|||||𝐺(𝑡,𝑠)𝑚(𝑠)𝜁(𝑢(𝑠))𝐶max𝑠𝕋||||(𝑚(𝑠)𝜁𝑢()),(3.8) where 1𝐺(𝑡,𝑠)=(𝑇+1(𝑇+1𝑡)𝑠,0𝑠𝑡𝑇+1,𝑇+1𝑠)𝑡,0𝑡𝑠𝑇+1.(3.9)

The results of Rabinowitz [19] for (3.6) can be stated as follows: from (𝜆𝑚,+/𝑟𝑓0,0), there emanates an unbounded continua 𝒞+ of positive solutions in ×𝑋.

It is clear that any solution of (3.6) of the form (1,𝑢) yields a solution 𝑢 of (1.1) and (1.2). So, we focus on the shape of 𝒞+ under the conditions (H1)–(H3) or (H1)–(H3), and we will show that 𝒞+ crosses the hyperplane {1}×𝑋 in ×𝑋.

Lemma 3.1. Suppose that (H1)–(H3) hold. Let 𝐽=[𝑎,𝑏] be a given compact interval in (0,). Then for all 𝑟𝐽, there exists 𝑀𝐽>0 such that all possible positive solution 𝑢 of (1.1) and (1.2) satisfy 𝑢𝑋𝑀𝐽.

Proof. Suppose on the contrary that there exists a sequence {𝑦𝑛} of positive solutions for (3.6) with {𝜇𝑛}𝐽 and 𝑦𝑛𝑋. Let 𝛼(0,1/𝑏𝑄), where 𝑄=𝑇𝑠=1𝐺(𝑠,𝑠)|𝑚(𝑠)|. Then, by (H3), there exists 𝑢𝛼>0 such that 𝑢>𝑢𝛼 implies 𝑓(𝑢)<𝛼𝑢.
Let 𝐾𝛼=max𝑢[0,𝑢𝛼]𝑓(𝑢) and let 𝐴𝑛={𝑡𝕋𝑦𝑛(𝑡)𝑢𝛼} and 𝐵𝑛={𝑡𝕋𝑦𝑛(t)>𝑢𝛼}. Then we have 𝑦𝑛(𝑡)=𝜇𝑛𝑇𝑠=1𝑦𝐺(𝑡,𝑠)𝑚(𝑠)𝑓𝑛(𝑠)=𝜇𝑛𝐴𝑛𝑦𝐺(𝑡,𝑠)𝑚(𝑠)𝑓𝑛(𝑠)+𝜇𝑛𝐵𝑛𝑦𝐺(𝑡,𝑠)𝑚(𝑠)𝑓𝑛(𝑠)𝜇𝑛𝐾𝛼𝑄+𝜇𝑛𝐵𝑛𝑦𝐺(𝑡,𝑠)𝑚(𝑠)𝑓𝑛(𝑠)(3.10) for 𝕋𝑡. Thus, 1𝜇𝑛𝐾𝛼𝑄𝑦𝑛𝑋+𝐵𝑛𝑓𝑦𝐺(𝑡,𝑠)𝑚(𝑠)𝑛(𝑠)𝑦𝑛𝑋.(3.11) On 𝐵𝑛,𝑦𝑛(𝑠)>𝑢𝛼 implies 𝑓(𝑦𝑛(𝑠))/𝑦𝑛𝑋<𝑓(𝑦𝑛(𝑠))/𝑦𝑛(𝑠)<𝛼. Thus, 1𝜇𝑛𝐾𝛼𝑄𝑦𝑛𝑋+𝛼𝐵𝑛||||𝐾𝐺(𝑠,𝑠)𝑚(𝑠)𝛼𝑄𝑦𝑛𝑋+𝛼𝑄.(3.12) Since 0<𝑎𝜇𝑛𝑏 for all 𝑛, we have 1/𝜇𝑛1/𝑏 for all 𝑛, and, thus, 1𝑏𝐾𝛼𝑄𝑦𝑛𝑋+𝛼𝑄.(3.13) By the fact 𝑦𝑛𝑋 as 𝑛, we get 1𝑏1𝛼𝑄<𝑏.(3.14) This contradiction completes the proof.

Lemma 3.2. Suppose that (H1)–(H3) hold. Then Proj𝒞+[𝜆𝑚,+/𝑟𝑓0,+).

Proof. Assume on the contrary that sup{𝜆(𝜆,𝑦)𝒞+}<, then, there exists a sequence {(𝜇𝑛,𝑦𝑛)}𝒞+ such that lim𝑛𝑦𝑛𝑋=,𝜇𝑛<𝐶0(3.15) for some positive constant 𝐶0 independent of 𝑛, since 𝒞+ is unbounded. On the other hand, 𝜇𝑛>0 for all 𝑛, since (0,0) is the only solution of (3.6) for 𝜆=0 and 𝒞+({0}×𝑋)=. Meanwhile, {(𝜇𝑛,𝑦𝑛)} satisfy Δ2𝑦𝑛(𝑡1)+𝜇𝑛𝑓𝑦𝑚(𝑡)𝑛(𝑡)𝑦𝑛𝑦(𝑡)𝑛𝑦(𝑡)=0,𝑡𝕋,𝑛(0)=𝑦𝑛(𝑇+1)=0.(3.16) By (H3), there exist a positive constant 𝐿𝑓>0 such that 𝑓(𝑢)𝐿𝑓𝑢. Define a function 𝜒𝕋[0,) by 𝐿𝜒(𝑡)=𝑓,𝑚(𝑡)>0,0,𝑚(𝑡)<0.(3.17) Then, 𝑚(𝑡)(𝑓(𝑦𝑛(𝑡))/𝑦𝑛(𝑡))𝜒(𝑡)𝑚(𝑡). Let 𝛾+ be the principal eigenvalue of linear eigenvalue problem Δ2𝑣(𝑡1)+𝛾𝜒(𝑡)𝑚(𝑡)𝑣(𝑡)=0,𝑡𝕋,𝑣(0)=𝑣(𝑇+1)=0.(3.18) Then, by Remark 2.2, 𝛾+>0. Subsequently, by Theorem 2.3, we know that 𝛾+𝜇𝑛.(3.19) This combine with Lemma 3.1, lim𝑛𝜇𝑛=, which contradicts (3.15). Thus, Proj𝒞+𝜆𝑚,+𝑟𝑓0,.(3.20)

Now, by Lemma 3.2, 𝒞+ crosses the hyperplane {1}×𝑋 in ×𝑋, and, then, Theorem 1.1(i) holds. To obtain Theorem 1.1(ii), we need to prove the following Lemma.

Lemma 3.3. Suppose that (H1), (H2), and (H3) hold. Then Proj𝒞+(0,𝜆𝑚,+/𝑟𝑓0).

Proof. Let {(𝜇𝑛,𝑦𝑛)}𝒞+ be such that |𝜇𝑛|+𝑦𝑛𝑋 as 𝑛. Then, Δ2𝑦𝑛(𝑡1)+𝜇𝑛𝑦𝑚(𝑡)𝑓𝑛(𝑦𝑡)=0,𝑡𝕋,𝑛(0)=𝑦𝑛(𝑇+1)=0.(3.21) If {𝑦𝑛} is bounded, say, 𝑦𝑛𝑀1, for some 𝑀1 independent of 𝑛, then we may assume that lim𝑛𝜇𝑛=.(3.22) Note that 𝑓𝑦𝑛(𝑡)𝑦𝑛(𝑡)inf𝑓(𝑠)𝑠0<𝑠𝑀1>0.(3.23) Then, there exist two constants 𝑀3>0,𝑀2>0 such that 0<𝑀2<𝑓𝑦𝑛(𝑡)𝑦𝑛(𝑡)<𝑀3.(3.24) Define two functions 𝜒1,𝜒2𝕋(0,) by 𝜒1=𝑀2𝑀,𝑚(𝑡)>0,3𝜒,𝑚(𝑡)<0.2=𝑀3𝑀,𝑚(𝑡)>0,2,𝑚(𝑡)<0.(3.25) Let 𝜂,𝜂 be the positive principal eigenvalue of the linear eigenvalue problems Δ2𝑣(𝑡1)+𝜂𝜒2(Δ𝑡)𝑚(𝑡)𝑣(𝑡)=0,𝑡𝕋,(3.26)𝑣(0)=𝑣(𝑇+1)=0,(3.27)2𝑣(𝑡1)+𝜂𝜒1(𝑡)𝑚(𝑡)𝑣(𝑡)=0,𝑡𝕋,(3.28)𝑣(0)=𝑣(𝑇+1)=0,(3.29) respectively.
Combining (3.22) and (3.24) with the relation Δ2𝑦𝑛(𝑡1)+𝑟𝜇𝑛𝑓𝑦𝑚(𝑡)𝑛(𝑡)𝑦𝑛𝑦(𝑡)𝑛𝑦(𝑡)=0,𝑡𝕋,𝑛(0)=𝑦𝑛(𝑇+1)=0,(3.30) using Theorem 2.3, we get 𝜂𝑟𝜇𝑛𝜂𝑟.(3.31) This contradicts (3.22). So, {𝑦𝑛𝑋} is bounded uniformly for all 𝑛.
Now, taking {(𝜇𝑛,𝑦𝑛)}𝒞+ be such that 𝑦𝑛𝑋+,as𝑛+.(3.32) We show that lim𝑛𝜇𝑛=0.
Suppose on the contrary that, choosing a subsequence and relabeling if necessary, 𝜇𝑛𝑏0 for some constant 𝑏0>0. By (3.32), there exists 𝑡0𝕋 such that 𝑦𝑛(𝑡0)=𝑦𝑛𝑋 and 𝑦𝑛(𝑡0)+ as 𝑛+. Thus, 𝜇𝑛𝑓𝑦𝑛𝑡0𝑦𝑛𝑡0+,as𝑛+.(3.33)
Now, the proof can be divided into two cases.
Case 1 (𝑚(𝑡0)>0). Consider the following linear eigenvalue problems Δ2𝑣(𝑡1)+𝛼𝑚(𝑡)𝑣(𝑡)=0,𝑡=𝑡0,𝑣𝑡0𝑡1=𝑣0+1=0.(3.34) By Theorem 2.4, (3.34) has a positive principal eigenvalue 𝛼+, and 𝜇𝑛𝑓𝑦𝑛𝑡0𝑦𝑛𝑡0𝛼+,(3.35) which contradicts (3.33).Case 2 (𝑚(𝑡0)<0). Since (𝜇𝑛,𝑦𝑛) is a solution of (3.7), we get 0Δ2𝑦𝑛𝑡01=𝜇𝑛𝑡𝑟𝑚0𝑓𝑦𝑛𝑡0>0.(3.36) This is a contradiction.
Thus, lim𝑛𝜇𝑛=0.

At last, we deal with the case 𝑟<0.
Let us consider 𝐿𝑢𝜆𝑟𝑚(𝑡)𝑓0𝑢𝜆𝑟𝑚(𝑡)𝜁(𝑢)=0(3.37) as a bifurcation problem from the trivial solution 𝑢0. The results of Rabinowitz [19] for (3.37) can be stated as follows: from (𝜆𝑚,/𝑟𝑓0,0), there emanates an unbounded continua 𝒞 of positive solutions in ×𝑋.
It is clear that any solution of (3.37) of the form (1,𝑢) yields a solution 𝑢 of (1.1) and (1.2). Now, our proofs focus on the shape of 𝒞. It will be proved that when (H1)–(H3) hold, then Proj𝒞(,𝜆𝑚,/𝑟𝑓0), and when (H1)–(H3) hold, Proj𝒞(𝜆𝑚,/𝑟𝑓0,0), that is, 𝒞 crosses the hyperplane {1}×𝑋 in ×𝑋. Since the proof is similar to the case 𝑟>0, we omit it.

Remark 3.4. As an application of Theorem 1.1, let us consider nonlinear discrete problem with indefinite weight Δ2𝑢𝑓(𝑡1)+𝑟𝑚(𝑡)(𝑢(𝑡))=0,𝑡𝕋,𝑢(0)=0,𝑢(3)=0,(3.38) where 𝕋={1,2}, 𝑚𝕋 with 𝑚(1)=1 and 𝑚(2)=1, and [],𝑓(𝑠)=𝑠,𝑠1,12𝑠1+𝑠4,𝑠(,1)(1,).(3.39) Then, 𝑓0𝑓=1,=0.(3.40)
To find the principal eigenvalues of linear eigenvalue problem Δ2𝑢(𝑡1)+𝜆𝑚(𝑡)𝑢(𝑡)=0,𝑡𝕋,(3.41)𝑢(0)=0,𝑢(3)=0,(3.42) we rewrite (3.41) to the recursive sequence 𝑢𝑢(𝑡+1)=2𝜆𝑚(𝑡)(𝑡)𝑢(𝑡1).(3.43) This together with the initial value condition 𝑢(0)=0,𝑢(1)=1(3.44) imply that 𝑢𝑢(2)=2𝜆𝑚(1)(1)𝑢(0)=2𝜆,𝑢(3)=2𝜆𝑚(2)𝑢(2)𝑢(1)=3𝜆2.(3.45) The last equation together with the boundary value condition 𝑢(3)=0 imply that 𝜆𝑚,=3,𝜆𝑚,+=3.(3.46) Thus by Theorem 1.1(i), (3.38), has a positive solution if 𝑟(,3)(3,).

Acknowledgments

The authors are grateful to the anonymous referee for their valuable suggestions. R. Ma is supported by NSFC (11061030).