Abstract

This paper studies the existence of multiple solutions of the second-order difference boundary value problem Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛))=0, 𝑛(1,𝑇), 𝑢(0)=0=𝑢(𝑇+1). By applying Morse theory, critical groups, and the mountain pass theorem, we prove that the previous equation has at least three nontrivial solutions when the problem is resonant at the eigenvalue 𝜆𝑘(𝑘2) of linear difference problem Δ2𝑢(𝑛1)+𝜆𝑢(𝑛)=0, 𝑛(1,𝑇), 𝑢(0)=0=𝑢(𝑇+1) near infinity and the trivial solution of the first equation is a local minimizer under some assumptions on 𝑉.

1. Introduction

Let , , and be the sets of real numbers, natural numbers, and integers, respectively. For any 𝑎,𝑏, 𝑎𝑏, define (𝑎,𝑏)={𝑎,𝑎+1,,𝑏}.

Consider the second-order difference boundary value problem (BVP)

Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛))=0,𝑛(1,𝑇),𝑢(0)=0=𝑢(𝑇+1),(1.1) where 𝑉𝐶2(,) and Δ denotes the forward difference operator defined by Δ𝑢(𝑛)=𝑢(𝑛+1)𝑢(𝑛), Δ2𝑢(𝑛)=Δ(Δ𝑢(𝑛)).

By a solution 𝑢 of the BVP (1.1), we mean a real sequence {𝑢(𝑛)}𝑇+1𝑛=0(=(𝑢(0),𝑢(1),,𝑢(𝑇+1))) satisfying the BVP (1.1). For 𝑢={𝑢(𝑛)}𝑇+1𝑛=0 with 𝑢(0)=0=𝑢(𝑇+1), we say that 𝑢0 if there exists at least one 𝑛(1,𝑇) such that 𝑢(𝑛)0. We say that 𝑢 is positive (and write 𝑢>0) if for all 𝑛(1,𝑇), 𝑢(𝑛)>0, and similarly, 𝑢 is negative (𝑢<0) if for all 𝑛(1,𝑇), 𝑢(𝑛)<0. The aim of this paper is to obtain the existence of multiple solutions of the BVP (1.1) and analyse the sign of solutions.

Recently, a few authors applied the minimax methods to examine the difference boundary value problems. For example, in [1], Agarwal et al. employed the Mountain Pass Lemma to study the following BVP:

Δ2𝑢(𝑛1)+𝑓(𝑛,𝑢(𝑛))=0,𝑛(1,𝑇),𝑢(0)=0=𝑢(𝑇+1)(1.2) and obtained the existence of multiple positive solutions, where 𝑓 may be singular at 𝑢=0. In [2], Jiang and Zhou employed the Mountain Pass Lemma together with strongly monotone operator principle, to study the following difference BVP:

Δ2𝑢(𝑛1)+𝑓(𝑛,𝑢(𝑛))=0,𝑛(1,𝑇),𝑢(0)=0=Δ𝑢(𝑇)(1.3) and obtained existence and uniqueness results, where 𝑓(1,𝑇)× is continuous. In [3], Cai and Yu employed the Linking Theorem and the Mountain Pass Lemma to study the following difference BVP:

Δ𝑝(𝑛)(Δ𝑢(𝑛1))𝛿+𝑞(𝑛)𝑢𝛿(𝑛)=𝑓(𝑛,𝑢(𝑛)),𝑛(1,𝑇),Δ𝑢(0)=𝐴,𝑢(𝑇+1)=𝐵(1.4) and obtained the existence of multiple solutions, where 𝛿>0 is the ratio of odd positive integers, {𝑝(𝑛)}𝑇+1𝑛=1 and {𝑞(𝑛)}𝑇𝑛=1 are real sequences, 𝑝(𝑛)0 for all 𝑛(1,𝑇+1), and 𝐴, 𝐵 are two given constants, 𝑓(1,𝑇)× is continuous.

Although applications of the minimax methods in the field of the difference BVP have attracted some scholarly attention in the recent years, efforts in applying Morse theory to the difference BVP are scarce. The main purpose of this paper is to develop a new approach to the BVP (1.1) by using Morse theory. To this end, we first consider the following linear difference eigenvalue problem:

Δ2𝑢(𝑛1)+𝜆𝑢(𝑛)=0,𝑛(1,𝑇),𝑢(0)=0=𝑢(𝑇+1).(1.5)

On the above eigenvalue problem, the following results hold; see [4].

Proposition 1.1. The eigenvalues of (1.5) are 𝜆=𝜆𝑙=4sin2𝑙𝜋2(𝑇+1),𝑙=1,2,,𝑇,(1.6) and the corresponding eigenfunction with 𝜆𝑙 is 𝜙𝑙(𝑛)=sin(𝑙𝜋𝑛/(𝑇+1)), 𝑙=1,2,,𝑇.

Remark 1.2. (1) The set of functions {𝜙𝑙(𝑛),𝑙=1,2,,𝑇} is orthogonal on (1,𝑇) with respect to the weight function 𝑟(𝑛)1, that is, 𝑇𝑛=1𝜙𝑙(𝑛),𝜙𝑗(𝑛)=0𝑙𝑗.(1.7) Moreover, for each 𝑙(1,𝑇), 𝑇𝑛=1sin2(𝑙𝜋𝑛/(𝑇+1))=(𝑇+1)/2.
(2) It is easy to see that 𝜙1 is positive and 𝜙𝑙 changes sign for each 𝑙(2,𝑇), that is, {𝑛𝜙𝑙(𝑛)>0} and {𝑛𝜙𝑙(𝑛)<0}.

For (1.1), we assume that

𝑉(0)=𝑉𝑉(0)=0,(1.8)()=lim|𝑡|𝑉(𝑡)𝑡=𝜆𝑘,(1.9) where 𝜆𝑘 is an eigenvalue of (1.5). Hence the BVP (1.1) has a trivial solution 𝑢0. And we say that BVP (1.1) is resonant at infinity if (1.9) holds.

Let

𝑊𝜙=span1,𝜙2,,𝜙𝑘1,𝑊0𝜙=span𝑘,𝑊+𝜙=span𝑘+1,𝜙𝑘+2,,𝜙𝑇.(1.10) Let 𝐺(𝑡)=𝑡0𝐺(𝑠)𝑑𝑠=𝑉(𝑡)(𝜆𝑘/2)𝑡2. By (1.9) we have

lim|𝑡|𝐺(𝑡)𝑡=0.(1.11) Assume that the following conditions on 𝐺(𝑡) hold.

(𝐺±) If 𝑢𝑚 such that 𝑣𝑚/𝑢𝑚1, then there exist 𝛿>0 and 𝑀 such that

±𝑇𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚(𝑛)𝛿,𝑚𝑀,(1.12) where 𝑢𝑚=𝑣𝑚+𝑤𝑚, 𝑣𝑚𝑊0, 𝑤𝑚𝑊=𝑊+𝑊.

The main result of this paper is as follows.

Theorem 1.3. Let (1.8), (1.9) hold and (𝑉1)𝑉(𝑡)>0 for all 𝑡, (𝑉2)𝑉(0)<𝜆1hold. Then the BVP (1.1) has at least three nontrivial solutions, with one positive solution and one negative solution, in each of the following cases: (i)(G+) and 𝑘2; (ii)(G) and 𝑘3.

To the author’s best knowledge, only Bin et al. [5] deal with the existence and multiplicity of nontrivial periodic solutions for asymptotically linear resonant difference problem by the aid of Su [6]. In [5], 𝐺 satisfies

||𝐺||(𝑧)𝑐1|𝑧|𝑠+𝑐2,(1.13)lim𝑣inf𝑣𝑊01𝑣2𝑠𝐺(𝑧)4𝛽2,𝛿𝑇(1.14) where 𝑐1>0, 𝑐2>0, 𝑠(0,1),𝛽=𝑐1𝑇(1𝑠)/2, 𝛿>0. In [5], the authors obtained the existence of one nontrivial periodic solution. Notice that (1.13) implies that (1.11) holds; however, (𝐺±) is not covered by (1.14). In fact, conditions (1.13) and (1.14) are borrowed from [6]. The conditions in Theorem 1.3 coincide with the assumptions of Theorem 1 in [7]. The aim of this paper is to develop a new approach to study the discrete systems by using Morse theory, minimax theorems, and some analysis technique. We wish to have some breakthrough points with the aid of the method of discretization.

The remaining part of this paper proceeds as follows. In the next section, we establish the variational framework of the BVP (1.1) and collect some results which will be used in the proof of Theorem 1.3. In Section 3, we give the proof of Theorem 1.3. Finally, in Section 4, we give an example to illustrate our main result and summarize conclusions and future directions.

2. Variational Framework and Auxiliary Results

Let

𝐸=𝑢𝑢={𝑢(𝑛)}𝑇+1𝑛=0.with𝑢(0)=0=𝑢(𝑇+1)(2.1)𝐸 can be equipped with the norm and the inner product , as follows:

𝑢=𝑇𝑛=0||||Δ𝑢(𝑛)20𝑥0200𝑑1/2,𝑢𝐸,𝑢,𝑣=𝑇𝑛=0(Δ𝑢(𝑛),Δ𝑣(𝑛)),𝑢,𝑣𝐸,(2.2) where || denotes the Euclidean norm in and (,) denotes the usual scalar product in . It is easy to see that (𝐸,,) is a Hilbert space. Consider the functional defined on 𝐸 by

1𝐽(𝑢)=2𝑇𝑛=0|Δ𝑢(𝑛)|2𝑇𝑛=1𝑉(𝑢(𝑛)).(2.3) We claim that if 𝑢𝐸 is a critical point of 𝐽, then 𝑢 is precisely a solution of the BVP (1.1). Indeed, for every 𝑢,𝑣𝐸, we have

𝐽(𝑢),𝑣=𝑇𝑛=0(Δ𝑢(𝑛),Δ𝑣(𝑛))𝑇𝑛=1𝑉(𝑢(𝑛)),𝑣(𝑛)=𝑇𝑛=1Δ2𝑢(𝑛1)+𝑉.(𝑢(𝑛)),𝑣(𝑛)(2.4)

So, if 𝐽(𝑢)=0, then we have

𝑇𝑛=1Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛)),𝑣(𝑛)=0.(2.5) Since 𝑣𝐸 is arbitrary, we obtain

Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛))=0,𝑛(1,𝑇).(2.6) Therefore, we reduce the problem of finding solutions of the BVP (1.1) to that of seeking critical points of the functional 𝐽 in 𝐸.

According to Proposition 1.1 and Remark 1.2, 𝐸 can be decomposed as 𝐸=𝑊𝑊0𝑊+. For all 𝑢𝐸, denote 𝑢=𝑤0+𝑤++𝑤 with 𝑤0𝑊0, 𝑤+𝑊+, and 𝑤𝑊, then we have the following Wirtinger type inequalities:

𝜆1𝑇𝑛=1(𝑢(𝑛),𝑢(𝑛))𝑢2𝜆𝑇𝑇𝑛=1𝜆(𝑢(𝑛),𝑢(𝑛)),𝑢𝐸,(2.7)1𝑇𝑛=1(𝑤(𝑛),𝑤(𝑛))𝑤2𝜆𝑇𝑘1𝑛=1(𝑤(𝑛),𝑤(𝑛)),𝑤𝑊𝜆,(2.8)𝑇𝑘+1𝑛=1𝑤+(𝑛),𝑤+𝑤(𝑛)+2𝜆𝑇𝑇𝑛=1𝑤+(𝑛),𝑤+(𝑛),𝑤+𝑊+,(2.9) see [4] for details.

Now we collect some results on Morse theory and the minimax methods.

Let 𝐸 be a real Hilbert space and 𝐽𝐶1(𝐸,). Denote

𝐽𝑐={𝑢𝐸𝐽(𝑢)𝑐},𝒦𝑐={𝑢𝐸𝐽(𝑢)=0,𝐽(𝑢)=𝑐}(2.10) for 𝑐. The following is the definition of the Palais-Smale condition ((PS) condition).

Definition 2.1. The functional 𝐽 satisfies the (PS) condition if any sequence {𝑢𝑚}𝐸 such that {𝐽(𝑢𝑚)} is bounded and 𝐽(𝑢𝑚)0 as 𝑚 has a convergent subsequence.

In [8], Cerami introduced a weak version of the (PS) condition as follows.

Definition 2.2. The functional 𝐽 satisfies the Cerami condition ((𝐶) condition) if any sequence {𝑢𝑚}𝐸 such that {𝐽(𝑢𝑚)} is bounded and (1+𝑢𝑚)𝐽(𝑢𝑚)0 as 𝑚 has a convergent subsequence.

If 𝐽 satisfies the (PS) condition or the (𝐶) condition, then 𝐽 satisfies the following deformation condition which is essential in Morse theory (cf. [9, 10]).

Definition 2.3. The functional 𝐽 satisfies the (𝐷𝑐) condition at the level 𝑐 if for any 𝜖>0 and any neighborhood 𝒩 of 𝒦𝑐, there are 𝜖>0 and a continuous deformation 𝜂[0,1]×𝐸𝐸 such that(1)𝜂(0,𝑢)=𝑢 for all 𝑢𝐸;(2)𝜂(𝑡,𝑢)=𝑢 for all 𝑢𝐽1([𝑐𝜖,𝑐+𝜖]);(3)𝐽(𝜂(𝑡,𝑢)) is nonincreasing in 𝑡 for any 𝑢𝐸;(4)𝜂(1,𝐽𝑐+𝜖𝒩)𝐽𝑐𝜖. 𝐽 satisfies the (𝐷) condition if 𝐽 satisfies the (𝐷𝑐) condition for all 𝑐.

Let 𝑢0 be an isolated critical point of 𝐽 with 𝐽(𝑢0)=𝑐, and let 𝑈 be a neighborhood of 𝑢0, the group

𝐶𝑞𝐽,𝑢0=𝐻𝑞𝐽𝑐𝑈,𝐽𝑐𝑢𝑈0,𝑞,(2.11) is called the 𝑞th critical group of 𝐽 at 𝑢0, where 𝐻𝑞(𝐴,𝐵) denotes the 𝑞th singular relative homology group of the pair (𝐴,𝐵) over a field 𝐹, which is defined to be quotient 𝐻𝑞(𝐴,𝐵)=𝑍𝑞(𝐴,𝐵)/𝐵𝑞(𝐴,𝐵), where 𝑍𝑞(𝐴,𝐵) is the 𝑞th singular relative closed chain group and 𝐵𝑞(𝐴,𝐵) is the 𝑞th singular relative boundary chain group.

Let 𝒦={𝑢𝐸𝐽(𝑢)=0}. If 𝐽(𝒦) is bounded from below by 𝑎 and 𝐽 satisfies the (𝐷𝑐) condition for all 𝑐𝑎, then the group

𝐶𝑞(𝐽,)=𝐻𝑞(𝐸,𝐽𝑎),𝑞,(2.12) is called the 𝑞th critical group of 𝐽 at infinity [11].

Assume that #𝒦< and 𝐽 satisfies the (𝐷) condition. The Morse-type numbers of the pair (𝐸,𝐽𝑎) are defined by

𝑀𝑞=𝑀𝑞(𝐸,𝐽𝑎)=𝑢𝒦dim𝐶𝑞(𝐽,𝑢),(2.13) and the Betti numbers of the pair (𝐸,𝐽𝑎) are

𝛽𝑞=dim𝐶𝑞(𝐽,).(2.14) By Morse theory [12, 13], the following relations hold:

𝑞𝑗=0(1)𝑞𝑗𝑀𝑗𝑞𝑗=0(1)𝑞𝑗𝛽𝑗,𝑞,𝑞=0𝑀𝑞=𝑞=0𝛽𝑞.(2.15)

Thus, if 𝐶𝑞(𝐽,)0, for some 𝑘, then there must exist a critical point 𝑢 of 𝐽 with 𝐶𝑞(𝐽,𝑢)0, which can be rephrased as follows.

Proposition 2.4. Let 𝐸 be a real Hilbert space and 𝐽𝐶2(𝐸,). Assume that #𝒦< and that 𝐽 satisfies the (𝐷) condition. If there exists some 𝑞 such that 𝐶𝑞(𝐽,)0, then 𝐽 must have a critical point 𝑢 with 𝐶𝑞(𝐽,𝑢)0.

In order to prove our main result, we need the following result about the critical group on 𝐶𝑞(𝐽,).

Proposition 2.5. Let the functional 𝐽𝐸 be of the form 1𝐽(𝑢)=2𝐴𝑢,𝑢+𝑄(𝑢),(2.16) where 𝐴𝐸𝐸 is a self-adjoint linear operator such that 0 is isolated in 𝜎(𝐴), the spectrum of 𝐴. Assume that 𝑄𝐶1(𝐸,) satisfies lim𝑢𝑄(𝑢)𝑢=0.(2.17) Denote 𝑉=ker𝐴, 𝑊=𝑉=𝑊+𝑊, where 𝑊+ (𝑊) is the subspace of 𝐸 on which 𝐴 is positive (negative) definite. Assume that 𝜇=dim𝑊, 𝜈=dim𝑉0 are finite and that 𝐽 satisfies the (𝐷) condition. Then 𝐶𝑞(𝐽,)𝛿𝑞,𝑘±𝐹,𝑞,(2.18) provided that 𝐽 satisfies the angle conditions at infinity. (𝐴𝐶±): there exist 𝑀>0 and 𝛼(0,1) such that ±𝐽(𝑢),𝑣0for𝑢=𝑣+𝑤,𝑢𝑀,𝑤𝛼𝑢,(2.19) where 𝑘+=𝜇, 𝑘=𝜇+𝜈, 𝑣𝑉, and 𝑤𝑊.

Remark 2.6. Conditions (2.16) and (2.17) imply that 𝐽 is asymptotically quadratic. Bartsch and Li [11] introduced the notion of critical groups at infinity and proved that if 𝐽 satisfied some angle properties at infinity, the critical groups can be completely figured out. Proposition 2.5 is a slight improvement of [11, Proposition 3.10] by Su and Zhao [7]. There are many other papers considering concrete problems by computing the critical groups at infinity with different methods, for example, see [1417].

We will use the Mountain Pass Lemma (cf. [12, 18]) in our proof.

Let 𝐵𝜌 denote the open ball in 𝐸 about 0 of radius 𝑟 and let 𝜕𝐵𝜌 denote its boundary.

Theorem 2.7 (mountain pass lemma). Let 𝐸 be a real Banach space and 𝐽𝐶1(𝐸,) satisfying the (PS) condition. Suppose 𝐽(0)=0 and that (𝐽1) there are constants 𝜌>0,𝑎>0 such that 𝐽|𝜕𝐵𝜌𝑎>0, (𝐽2) there is a 𝑢0𝐸𝐵𝜌 such that 𝐽(𝑢0)0,then 𝐽 possesses a critical value 𝑐𝑎. Moreover 𝑐 can be characterized as 𝑐=infΓsup𝑠[0,1]𝐽((𝑠)),(2.20) where ([]Γ=𝐶0,1,𝐸)(0)=0,(1)=𝑢0.(2.21)

Definition 2.8 (mountain pass point). An isolated critical point 𝑢 of 𝐽 is called a mountain pass point, if 𝐶1(𝐽,𝑢)0.

The following result is useful in computing the critical group of a mountain pass point; see [13, 19] for details.

Theorem 2.9. Let 𝐸 be a real Hilbert space. Suppose that 𝐽𝐶2(𝐸,) has a mountain pass point 𝑢, and that 𝐽(𝑢) is a Fredholm operator with finite Morse index, satisfying 𝐽𝑢0𝐽0,0𝜎𝑢0𝐽dimker𝑢0=1,(2.22) then 𝐶𝑞𝐽,𝑢0𝛿𝑞,1𝐹,𝑞.(2.23)

3. Proof of Theorem 1.3

We give the proof of Theorem 1.3 in this section. Firstly, we prove that the functional 𝐽 satisfies the (𝐶) condition (Lemma 3.1) and compute the critical group 𝐶𝑞(𝐽,) (Lemma 3.2). Then, we employ the cut-off technique and the Mountain Pass Lemma to obtain two critical points 𝑢+,𝑢 of 𝐽 and compute the critical groups 𝐶𝑞(𝐽,𝑢+) and 𝐶𝑞(𝐽,𝑢) (Lemmas 3.3 and 3.4). Finally, we prove Theorem 1.3.

Rewrite the functional 𝐽 as

1𝐽(𝑢)=2𝑇𝑛=0|Δ𝑢(𝑛)|2𝜆𝑘2𝑇𝑛=1|𝑢(𝑛)|2𝑇𝑛=1𝐺(𝑢(𝑛)),𝑢𝐸.(3.1)

Lemma 3.1. Let (1.8) and (1.9) hold. If 𝐺 satisfies (𝐺±), then the functional 𝐽 satisfies the (𝐶) condition.

Proof. We only prove the case where (𝐺+) holds. Let {𝑢𝑚}𝐸 such that 𝐽𝑢𝑚𝑢𝑐,1+𝑚𝐽𝑢𝑚0as𝑚.(3.2) Then for all 𝜑𝐸, we have 𝐽𝑢𝑚,𝜑=𝑢𝑚,𝜑𝜆𝑘𝑇𝑛=1𝑢𝑚(𝑛),𝜑(𝑛)𝑇𝑛=1𝐺𝑢𝑚.(𝑛),𝜑(𝑛)(3.3) Denote 𝑢𝑚=𝑣𝑚+𝑤+𝑚+𝑤𝑚 with 𝑣𝑚𝑊0, 𝑤+𝑚𝑊+ and 𝑤𝑚𝑊. Since 𝐸 is a finite-dimensional Hilbert space, it suffices to show that {𝑢𝑚} is bounded. Suppose that {𝑢𝑚} is unbounded. Passing to a subsequence we may assume that 𝑢𝑚 as 𝑚.
By (1.11), for any 𝜖>0, there exists 𝑏 such that ||𝐺||(𝑡)𝜖|𝑡|+𝑏,𝑡.(3.4) Let 𝜑=𝑤+𝑚 in (3.3). Then by (2.7), (2.9), and (3.4), we have 𝑐1𝑤+𝑚2𝜆=1𝑘𝜆𝑘+1𝑤+𝑚2𝑤+𝑚2𝜆𝑘𝑇𝑛=1𝑤+𝑚(𝑛),𝑤+𝑚=(𝑛)𝑇𝑛=1𝑢𝐺𝑚(𝑛),𝑤+𝑚(𝑢𝑛)+𝐽𝑚,𝑤+𝑚𝑤+𝑚+𝑇𝑛=1𝜖||𝑢𝑚||||𝑤(𝑛)+𝑏+𝑚||𝑤(𝑛)+𝑚+𝜖𝜆1𝜆𝑘+1𝑢𝑚𝑤+𝑚+𝑏𝑇𝜆𝑘+1𝑤+𝑚=𝑐2𝑤+𝑚+𝑐3𝑢𝑚𝑤+𝑚,(3.5) where 𝑐1𝜆=1𝑘𝜆𝑘+1>0,𝑐2𝑏=1+𝑇𝜆𝑘+1,𝑐3=𝜖𝜆1𝜆𝑘+1.(3.6) And since 𝜖>0 is arbitrary, we have 𝑤+𝑚𝑢𝑚0as𝑚.(3.7) Similarly, let 𝜑=𝑤𝑚 in (3.3), by (2.8) and (3.4) we get 𝑐4𝑤𝑚2𝜆=1𝑘𝜆𝑘1𝑤𝑚2𝑤𝑚2𝜆𝑘𝑇𝑛=1𝑤𝑚(𝑛),𝑤𝑚=(𝑛)𝑇𝑛=1𝐺𝑢𝑚(𝑛),𝑤𝑚(𝑛)+𝐽𝑢𝑚,𝑤𝑚𝑤𝑚𝑇𝑛=1𝜖||𝑢𝑚||||𝑤(𝑛)+𝑏𝑚||𝑤(𝑛)𝑚𝜖𝜆1𝑢𝑚𝑤𝑚𝑏𝑇𝜆1𝑤𝑚=𝑐5𝑤𝑚𝑐6𝑢𝑚𝑤𝑚,(3.8) where 𝑐4=𝜆𝑘𝜆𝑘11>0,𝑐5𝑏=1+𝑇𝜆1,𝑐6=𝜖𝜆1.(3.9) And hence we also have 𝑤𝑚𝑢𝑚0as𝑚.(3.10) By (3.7) and (3.10), we have 𝑤𝑚𝑢𝑚𝑣0,𝑚𝑢𝑚1as𝑚.(3.11) By (𝐺+), there exist 𝛿>0 and 𝑀 such that 𝑇𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚(𝑛)𝛿,𝑚𝑀.(3.12) This implies that 𝐽𝑢𝑚,𝑣𝑚=𝑇𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚(𝑛)𝛿,𝑚𝑀,(3.13) and hence 𝐽𝑢𝑚𝑢𝑚𝐽𝑢𝑚𝑣𝑚||𝐽𝑢𝑚,𝑣𝑚||𝛿,𝑚𝑀,(3.14) which is a contradiction to (3.2). Thus {𝑢𝑚} is bounded. The proof is complete.

Lemma 3.2. Let (1.8) and (1.9) hold. Then (1)𝐶𝑞(𝐽,)𝛿𝑞,𝑘𝐹 provided that (𝐺+) holds; (2)𝐶𝑞(𝐽,)𝛿𝑞,𝑘1𝐹 provided that (𝐺) holds.

Proof. We only prove the case (1). Define a bilinear function 𝑎(𝑢,𝑣)=𝜆𝑘𝑇𝑛=1(𝑢(𝑛),𝑣(𝑛)),𝑢,𝑣𝐸.(3.15) Then by (2.7) we have ||||𝜆𝑎(𝑢,𝑣)𝑘𝜆1𝑢𝑣.(3.16) And hence there exists a unique continuous bounded linear operator 𝐾𝐸𝐸 such that 𝐾𝑢,𝑣=𝜆𝑘𝑇𝑛=1(𝑢(𝑛),𝑣(𝑛)).(3.17) Since 𝐾𝑢,𝑢 for all 𝑢𝐸, we can conclude that 𝐾 is a self-adjoint operator and 1𝐽(𝑢)=2(𝐼𝐾)𝑢,𝑢𝑇𝑛=1𝐺(𝑢(𝑛)).(3.18) Then 𝐽 has the form (2.16) with 𝑄(𝑢)=𝑇𝑛=1𝐺(𝑢(𝑛)),(3.19) and (1.11) implies that (2.17) holds. Let 𝐴=𝐼𝐾. Then ker𝐴=𝑊0=span{𝜙𝑘}. Next we show that (𝐺+) implies that the angle condition (𝐴𝐶) at infinity holds.
If not, then for any 𝑚 and each 𝛼𝑚=1/𝑚, there exists 𝑢𝑚=𝑣𝑚+𝑤𝑚𝑊0(𝑊+𝑊) with 𝑣𝑚𝑊0, 𝑤𝑚𝑊+𝑊 such that 𝑢𝑚𝑤𝑚,𝑚1𝑚𝑢𝑚𝑢,(3.20)𝐽𝑚,𝑣𝑚>0.(3.21) On the other hand, (3.20) implies 𝑢𝑚𝑣,𝑚𝑢𝑚1as𝑚.(3.22) Thus, by (𝐺+) there exist 𝛿>0 and 𝑀 such that 𝑇𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚(𝑛)𝛿,𝑚𝑀.(3.23) Therefore, 𝐽𝑢𝑚,𝑣𝑚=𝑇𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚(𝑛)𝛿,𝑚𝑀,(3.24) which is a contradiction to (3.21). Consequently (𝐴𝐶) holds and by Lemma 3.1 and Proposition 2.5, 𝐶𝑞(𝐽,)=𝛿𝑞,𝑘𝐹. Similarly, we can prove that (2) holds.

In order to obtain a mountain pass point, we need the following lemmas.

Lemma 3.3. Let 𝑉+𝑉(𝑡)=𝑉(𝑡),𝑡0,0,𝑡0,𝑉(𝑡)=(𝑡),𝑡0,0,𝑡0,(3.25) and 𝑉±(𝑡)=𝑡0𝑉±(𝑠)𝑑𝑠. If lim|𝑡|𝑉(𝑡)𝑡=𝛼𝜆1,(3.26) then the functional 𝐽±1(𝑢)=2𝑇𝑛=0||||Δ𝑢(𝑛)2𝑇𝑛=1𝑉±(𝑢(𝑛))(3.27) satisfies the (PS) condition.

Proof. We only prove the case (𝐽+). Let {𝑢𝑚}𝐸 such that 𝐽𝑢𝑚𝑐,𝐽+𝑢𝑚0(3.28) as 𝑚. Since 𝐸 is a finite-dimensional space, it suffices to show that {𝑢𝑚} is bounded in 𝐸. Suppose that {𝑢𝑚} is unbounded. Passing to a subsequence we may assume that 𝑢𝑚 and for each 𝑛, either |𝑢𝑚(𝑛)| or {𝑢𝑚(𝑛)} is bounded.
Noticing that for all 𝜑𝐸, 𝐽+𝑢𝑚,𝜑=𝑢𝑚,𝜑𝑇𝑛=1𝑉+𝑢𝑚.(𝑛),𝜑(𝑛)(3.29)
Denote 𝑤𝑚=𝑢𝑚/𝑢𝑚, for a subsequence, 𝑤𝑚 converges to some 𝑤 with 𝑤=1. By (3.29), we have 𝐽+𝑢𝑚,𝜑𝑢𝑚=𝑤𝑚,𝜑𝑇𝑛=1𝑉+𝑢𝑚(𝑛)𝑢𝑚.,𝜑(𝑛)(3.30) If |𝑢𝑚(𝑛)|, then lim𝑚𝑉+𝑢𝑚(𝑛)𝑢𝑚𝑤(𝑛)𝑚(𝑛)=𝛼𝑤+(𝑛),(3.31) where 𝑤+(𝑛)=max{𝑤(𝑛),0} with 𝑛(1,𝑇). If {𝑢𝑚(𝑛)} is bounded, then lim𝑚𝑉+𝑢𝑚(𝑛)𝑢𝑚=0,𝑤(𝑛)=0.(3.32) Since 𝑤0, there is an 𝑛 for which |𝑢𝑚(𝑛)|. So passing to the limit in (3.30), we have 𝑇𝑛=0(Δ𝑤(𝑛),Δ𝜑(𝑛))𝛼𝑇𝑛=1𝑤+(𝑛),𝜑(𝑛)=0.(3.33) This implies that 𝑤0 satisfies Δ2𝑤(𝑛1)+𝛼𝑤+(𝑛)=0,𝑛(1,𝑇),𝑤(0)=0=𝑤(𝑇+1).(3.34)
Now, we claim that 𝑤(𝑛)>0,𝑛(1,𝑇).(3.35)
In fact, let 𝑤𝑛0=min{𝑤(𝑛)𝑛(1,𝑇)}.(3.36) We only need to prove 𝑤(𝑛0)>0. If not, assume that 𝑤(𝑛0)0. Then by (3.34), we have Δ2𝑤(𝑛01)=0 and hence 𝑤(𝑛01)=𝑤(𝑛0)=𝑤(𝑛0+1). By induction, it is easy to get 𝑤(𝑛)=0 for all 𝑛(1,𝑇) which is a contradiction to 𝑤0 and hence (3.35) holds.
On the other hand, by Proposition 1.1 and Remark 1.2, we see that only the eigenfunction corresponding to the eigenvalue 𝜆1 is positive, which is a contradiction to 𝛼𝜆1. The proof is complete.

Lemma 3.4. Under the conditions of Theorem 1.3, the functional 𝐽+ has a critical point 𝑢+>0 and 𝐶𝑞(𝐽+,𝑢+)𝛿𝑞,1𝐹; the functional 𝐽 has a critical point 𝑢<0 and 𝐶𝑞(𝐽,𝑢)𝛿𝑞,1𝐹.

Proof. We only prove the case of 𝐽+. Firstly, we prove that 𝐽+ satisfies the Mountain Pass Lemma and hence 𝐽+ has a nonzero critical point 𝑢+. In fact, 𝐽+𝐶1(𝐸,) and by Lemma 3.3 we see that 𝐽+ satisfies the (PS) condition. Clearly 𝐽+(0)=0. Thus we need to show that 𝐽+ satisfies (𝐽1) and (𝐽2). To verify (𝐽1), by (1.8) and (𝑉2), there exist 𝜌1>0 and 𝜌2>0 with 𝑉(0)<𝜌2<𝜆1 such that 1𝑉(𝑡)2𝜌2𝑡2(3.37) for |𝑡|𝜌1. So, for all 𝑢𝐸, if 𝑢𝜆1𝜌1, then for each 𝑛(1,𝑇), |𝑢(𝑛)|𝜌1 and 𝐽+1(𝑢)=2𝑢2𝑇𝑛=1𝑉+=1(𝑢(𝑛))2𝑢2𝑛𝑁11𝑉(𝑢(𝑛))2𝑢212𝜌2𝑛𝑁11(𝑢(𝑛),𝑢(𝑛))2𝑢212𝜌2𝑇𝑛=11(𝑢(𝑛),𝑢(𝑛))2𝑢212𝜌2𝜆1𝑢2,(3.38) where 𝑁1={𝑛(1,𝑇)𝑢(𝑛)0}. Let 𝜌=𝜆1𝜌11,𝑎=2𝜌12𝜆1𝜌2.(3.39) Then 𝐽+(𝑢)|𝜕𝐵𝜌𝑎>0 and hence (𝐽1) holds. For (𝐽2), by 𝑉()=𝜆𝑘(𝜆𝑘1,𝜆𝑘+1), we claim that there exist 𝛾>𝜆𝑘1(𝜆1), 𝑏 such that 𝑉(𝑡)𝛾2𝑡2+𝑏,𝑡.(3.40)
In fact, by assumption (1.9), there exist 𝑀>0 and 𝑏1 such that 𝑉(𝑡)(𝛾/2)𝑡2+𝑏1 for |𝑡|𝑀. Meanwhile, there exists 𝑏2 such that 𝑉(𝑡)(𝛾/2)𝑡2𝑏2 for |𝑡|𝑀 by virtue of the continuity of 𝑉. Let 𝑏=min{𝑏1,𝑏2}, we get the conclusion.
Thus, if we choose 𝑒span{𝜙1} with 𝑒>0 and 𝑒=1, then 𝐽+𝑡(𝑡𝑒)=22𝑇𝑛=1𝑡𝑉(𝑡𝑒(𝑛))22𝛾𝑡22=𝑡(𝑒,𝑒)𝑏𝑇22𝛾𝑡22𝜆1𝑏𝑇(3.41) as 0<𝑡+. Thus, we can choose a constant 𝑡 large enough with 𝑡>𝜌 and 𝑢0=𝑡𝑒𝐸 such that 𝐽+(𝑢0)0. (𝐽2) holds.
Therefore, by Theorem 2.7, 𝐽+ has a critical point 𝑢+0 and similar to the proof of Lemma 3.3, we can prove that 𝑢+>0. So 𝑢+ is also a critical point of 𝐽. In the following we compute the critical group 𝐶𝑞(𝐽+,𝑢+) by using Theorem 2.9.
Assume that 𝐽𝑢+𝑣,𝑣=𝑣,𝑣𝑇𝑛=1𝑉𝑢+(𝑛)𝑣(𝑛),𝑣(𝑛)0,𝑣𝐸,(3.42) and that there exists 𝑣00 such that 𝐽𝑢+𝑣0,𝑣=0,𝑣𝐸.(3.43) This implies that 𝑣0 satisfies Δ2𝑣0(𝑛1)+𝑉𝑢+(𝑣𝑛)0(𝑣𝑛)=0,𝑛(1,𝑇),0(0)=𝑣0(𝑇+1)=0.(3.44) Hence the eigenvalue problem Δ2𝑣(𝑛1)+𝜆𝑉𝑢+(𝑛)𝑣(𝑛)=0,𝑛(1,𝑇),𝑣(0)=𝑣(𝑇+1)=0(3.45) has an eigenvalue 𝜆=1. (𝑉1) implies that 1 must be a simple eigenvalue; see [4]. So, dimker(𝐽(𝑢0))=1. Since 𝐸 is a finite-dimensional Hilbert space, the Morse index of 𝑢+ must be finite and 𝐽(𝑢+) must be a Fredholm operator. By Theorem 2.9, 𝐶𝑞(𝐽+,𝑢+)𝛿𝑞,1𝐹. The proof is complete.

Remark 3.5. We can choose the neighborhood 𝑈 of 𝑢+ such that 𝑢>0 for all 𝑢𝑈. Therefore, 𝐶𝑞𝐽,𝑢+𝐶𝑞𝐽+,𝑢+𝛿𝑞,1𝐹.(3.46) Similarly, 𝐶𝑞(𝐽,𝑢)𝐶𝑞(𝐽,𝑢)𝛿𝑞,1𝐹.(3.47)

Now, we give the proof of Theorem 1.3.

Proof of Theorem 1.3. We only prove the case (i). By Lemma 3.2, 𝐶𝑞(𝐽,)𝛿𝑞,𝑘𝐹.(3.48) Hence by Proposition 2.4 the functional 𝐽 has a critical point 𝑢1 satisfying 𝐶𝑘𝐽,𝑢10.(3.49) Since 𝐽𝑉(0)𝑢,𝑢1(0)𝜆1𝑢2,(3.50) by (𝑉2) and 𝐽(0)=𝐽(0)=0, we see that 0 is a local minimum of 𝐽. Hence 𝐶𝑞(𝐽,0)𝛿𝑞,0𝐹.(3.51) By Remark 3.5, (3.49), (3.51), and 𝑘2 we get that 𝑢+, 𝑢, and 𝑢1 are three nonzero critical points of 𝐽 with 𝑢+>0 and 𝑢<0. The proof is complete.

4. An Example and Future Directions

To illustrate the use of Theorem 1.3, we offer the following example.

Example 4.1. Consider the BVP Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛))=0,𝑛(1,5),𝑢(0)=0=𝑢(6),(4.1) where 𝑉𝐶2(,) is defined as follows: 1𝑉(𝑡)=𝑡1021,|𝑡|1,2𝑡2+34𝑡4/3,|𝑡|10,astrictlyconvexfunction,otherwise.(4.2) It is easy to verify that 𝑉 satisfies (1.8), (1.9), (1.11), (𝑉1), and (𝑉2) with 𝑘=2. To verify the condition (𝐺+), note that 𝐺(𝑡)=𝑡1/3 for |𝑡|10, we claim that 5𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚(𝑛)+,as𝑚(4.3) which implies that (𝐺+) holds.
To this end, for any constant 𝑟>1, we introduce another norm in 𝐸(𝑇=5) as follows: 𝑢𝑟=5𝑛=1||||𝑢(𝑛)𝑟1/𝑟,𝑢𝐸.(4.4) Since 𝐸 is finite dimensional, there exist two constants 𝐶2𝐶1>0 such that 𝐶1𝑢𝑢𝑟𝐶2𝑢,𝑢𝐸.(4.5)
Now, by (𝐺+), for any 𝜖 small enough, it is easy to see that 𝑤𝑚𝑢𝜖𝑚(4.6) holds for 𝑚 large enough.
Set Ω1=||𝑢𝑛(1,5)𝑚||(𝑛)10,Ω2=(1,5)Ω1.(4.7) Since 𝑢𝑚, Ω1, for 𝑚 large enough. And for 𝑚 large enough, we have 5𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚=(𝑛)𝑛Ω1𝑢𝑚1/3(𝑛),𝑣𝑚+(𝑛)𝑛Ω2𝐺𝑢𝑚(𝑛),𝑢𝑚(𝑛)𝑛Ω2𝑢𝐺𝑚(𝑛),𝑤𝑚(𝑛)𝑛Ω1𝑢𝑚1/3(𝑛),𝑣𝑚(𝑢𝑛)𝑐𝑐𝜖𝑚=𝑛Ω1𝑢𝑚1/3(𝑛),𝑢𝑚(𝑛)𝑛Ω1𝑢𝑚1/3(𝑛),𝑤𝑚𝑢(𝑛)𝑐𝑐𝜖𝑚.(4.8) Here and below we denote by 𝑐 various positive constants. Since 𝑛Ω1𝑢𝑚1/3(𝑛),𝑢𝑚=(𝑛)5𝑛=1𝑢𝑚1/3(𝑛),𝑢𝑚(𝑛)𝑛Ω2𝑢𝑚1/3(𝑛),𝑢𝑚=𝑢(𝑛)𝑚4/34/3𝑐,𝑛Ω1𝑢𝑚1/3(𝑛),𝑤𝑚(𝑛)5𝑛=1||𝑢𝑚||(𝑛)1/3||𝑤𝑚||(𝑛)5𝑛=1||𝑢𝑚||(𝑛)4/31/45𝑛=1||𝑤𝑚||(𝑛)4/33/4=𝑢𝑚1/34/3𝑤𝑚4/3𝑢𝑐𝜖𝑚4/34/3.(4.9) Hence 5𝑛=1𝐺𝑢𝑚(𝑛),𝑣𝑚𝑢(𝑛)(1𝑐𝜖)𝑚4/34/3𝑢𝑐𝜖𝑚𝑐.(4.10) Since 𝜖 is small enough, we get (4.3) holds by the above and (4.5). Hence, by Theorem 1.3, BVP (4.1) has at least three nontrivial solutions.

Morse theory has been proved very useful in proving the existence and multiplicity of solutions of operator equations with variational frameworks. However, it is well known that the minimax methods is also a useful tool for the same purpose. The advantage of the minimax methods is that it provides an estimate of the critical value. But it is hard to distinguish critical points obtained by this methods with those by other methods, if the local behavior of the critical points is not very well known. However, critical groups serve as a topological tool in distinguishing isolated critical points. Hence, in order to obtain multiple solutions by using Morse theory, it is crucial to describe critical groups clearly.

A natural question is: can we use the same methods in this paper to other BVPs? Noticing that the key conditions which guarantee the multiplicity of solutions of the BVP (1.1) are as follows:

(1)the BVP has a variational framework;(2)the eigenvalues of the corresponding linear BVP are nonzero and there is a one-sign eigenfunction,

hence, if the difference equation

Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛))=0,𝑛(1,𝑇)(4.11) subject to some other boundary value conditions satisfying (1) and (2), then we can obtain similar results to Theorem 1.3.

Example 4.2. Consider the BVP Δ2𝑢(𝑛1)+𝑉(𝑢𝑢(𝑛))=0,𝑛(1,𝑇),(0)=0=Δ𝑢(𝑇).(4.12) Let 𝐸=𝑢𝑢={𝑢(𝑛)}𝑇+1𝑛=0.with𝑢(0)=0=Δ𝑢(𝑇)(4.13) Then 𝐸 is a 𝑇-dimensional Hilbert space with the inner product 𝑢,𝑣=𝑇𝑛=0(Δ𝑢(𝑛),Δ𝑣(𝑛)).(4.14) Define the functional 𝐽 on 𝐸 by 𝐽(𝑢)=𝑇𝑛=012||||Δ𝑢(𝑛)2𝑇𝑛=1𝑉(𝑢(𝑛)).(4.15) It is easy to see that 𝑢 is a critical point of 𝐽 in 𝐸 if and only if 𝑢 is a solution of the BVP (4.12).
The eigenvalues of the linear BVP Δ2𝑢(𝑛1)+𝜆𝑢(𝑛)=0,𝑛(1,𝑇),𝑢(0)=0=Δ𝑢(𝑇)(4.16) are 𝜆=𝜆𝑙=4sin2𝑙𝜋2(2𝑇+1),𝑙=1,2,,𝑇,(4.17) and the corresponding eigenfunctions are 𝜙𝑙(𝑛)=sin𝑙𝜋𝑛2𝑇+1,𝑙=1,2,,𝑇.(4.18) Hence, 𝜆𝑙0 for all 𝑙(1,𝑇) and 𝜙1(𝑛)>0 for all 𝑛(1,𝑇). Therefore, the BVP (4.12) satisfies (1) and (2) and hence we can obtain similar results as in Theorem 1.3.

However, consider the following difference BVP:

Δ2𝑢(𝑛1)+𝑉(𝑢(𝑛))=0,𝑛(1,𝑇),𝑢(0)=𝑢(𝑇),Δ𝑢(0)=Δ𝑢(𝑇).(4.19) It is easy to verify that the variational functional of the BVP (4.19) is

𝐽(𝑢)=𝑇𝑛=110𝑥0200𝑑2||||Δ𝑢(𝑛)2𝑉(𝑢(𝑛)),𝑢𝐸1,(4.20) where

𝐸1=𝑢𝑢={𝑢(𝑛)}𝑇+1𝑛=0.with𝑢(0)=𝑢(𝑇),Δ𝑢(0)=Δ𝑢(𝑇+1)(4.21) But, 𝜆=0 is an eigenvalue of the linear BVP:

Δ2𝑢(𝑛1)+𝜆𝑢(𝑛)=0,𝑛(1,𝑇),𝑢(0)=𝑢(𝑇),Δ𝑢(0)=Δ𝑢(𝑇).(4.22) So, for the BVP (4.19), we need to find other techniques (e.g., dual variational methods if possible) to study the BVP (4.19).

Acknowledgments

The authors would like to express their thanks to the referees for helpful suggestions. This research is supported by Guangdong College Yumiao Project (2009).