Spectrum of Discrete Second-Order Difference Operator with Sign-Changing Weight and Its Applications
Abstract
Let T > 1 be an integer, and let𝕋 = {1,2, …, T}. We discuss the spectrum of discrete linear second-order eigenvalue problems Δ2u(t − 1) + λm(t)u(t) = 0, t ∈ 𝕋, u(0) = u(T + 1) = 0, where λ ≠ 0 is a parameter, m : 𝕋 → ℝ changes sign and m(t) ≠ 0 on 𝕋. At last, as an application of this spectrum result, we show the existence of sign-changing solutions of discrete nonlinear second-order problems by using bifurcate technique.
1. Introduction
Theorem A. Problem (4) has an infinite sequence of simple eigenvalues
This result has been extended to one-dimensional p-Laplacian operator by Anane et al. [2] and to the high-dimensional case by Hess and Kato [3], Bongsoo and Brown [4], and Afrouzi and Brown [5]. Meanwhile, these spectrum results have been used to deal with several nonlinear problems; see, for example, [4–7] and the references therein.
However, these two results do not give any information of the eigenfunctions of the linear eigenvalue problems (6) and (7) or (6) and (8).
Theorem B. Problem (9) has exactly T real and simple eigenvalues λk, k ∈ 𝕋, which satisfies
Furthermore, when m(t) ≡ 1, Agarwal et al. [10] generalized the results of Theorem B to the dynamic equations on time scales with Sturm-Liouville boundary condition. Moreover, under the assumption that the weight functions are not changing sign, several important results on linear Hamiltonian difference systems have also been established by Shi and Chen [11], Bohner [12], and the references therein.
However, there are few results on the spectrum of discrete second-order linear eigenvalue problems when m(t) changes its sign on 𝕋. In 2008, Shi and Yan [13] discussed the spectral theory of left definite difference operators when m(t) may change its sign. However, they provided no information about the sign of the eigenvalues and no information about the corresponding eigenfunctions. Recently, Ma et al. [14] obtained that (1) and (2) have two principal eigenvalues λ1,− < 0 < λ1,+ and they studied some corresponding discrete nonlinear problems.
It is the purpose of this paper to establish the discrete analogue of Theorem A for the discrete problems (1) and (2). More precisely, we will prove the following.
Theorem 1. Let n be the number of elements in 𝕋+. Let (H0) hold and let ν ∈ {+, −}. Then
2. Preliminaries
Recall that 𝕋 = {1, …, T}. Let , . Then X is a Banach space under the norm . Let Y = {u ∣ u : 𝕋 → ℝ}. Then Y is a Banach space under the norm ∥u∥Y = maxt∈𝕋 | u(t)|.
Definition 3 (see [22].)Suppose that a function . If y(t) = 0, then t is a zero of y. If y(t) = 0 and y(t − 1)y(t + 1) < 0 for some t ∈ {2, …, T − 1}, then t is a simple zero of y. If y(t)y(t + 1) < 0 for some t ∈ {1, …, T − 1}, then y has a node at the point
Lemma 4 (Lagrange-type identities). For t ∈ 𝕋,
Proof. We write (15) for the two arguments in full, giving
Corollary 5. For t ∈ 𝕋,
Proof. Dividing (17) by μ − λ and making μ → λ for fixed λ, then we get the desired result.
Corollary 6. For t ∈ 𝕋, and complex λ,
3. Spectrum of (1) and (2)
Lemma 7. For t ∈ 𝕋, if (λ, u) is a solution of
Moreover, if λ > 0, then . If λ < 0, then .
Proof. It is easy to see that λ ≠ 0. Now, multiplying (22) by u(s), then we get that
Now, we prove some oscillatory properties.
Lemma 8. For t ∈ {1,2, …, T + 1}, the polynomial u(t, λ) has precisely t − 1 real and simple zeros.
Proof. This proof is divided into two steps.
Step 1 (each zero of u(t, λ) is real). Suppose on the contrary that u(t, λ) has a complex zero λ0; then . Furthermore,
Step 2 (all of the zeros of u(t, λ) are simple). Suppose on the contrary that u(t, λ) has a multiple zeros λ*, necessarily real. Then u(t, λ*) = 0 and u′(t, λ*) = 0. Moreover,
Lemma 9 (see [14].)Let n ∈ {1, …, T − 1}. Then (1) and (2) have two principal eigenvalues λ1,+ > 0 > λ1,− and the corresponding eigenfunctions do not change their sign.
Lemma 10. Let ν ∈ {+, −}. Then dim ker(L − λk,νmI) = 1, where I denotes the identity operator.
Proof. Suppose that φk,ν ∈ X and ϕk,ν ∈ X are two eigenfunctions corresponding to λk,ν. Then φk,ν(0) = ϕk,ν(0) = 0 and there exists a constant c ∈ ℝ such that φk,ν(1) = cϕk,ν(1) ≠ 0. Now, by the recurrence relation (15), we get that φk,ν(t) = cϕk,ν(t) for .
Lemma 11. For fixed real λ, the zeros of u(x, λ), 0 ≤ x ≤ T + 1 are simple.
Proof. Suppose that x′ is a zero of u(x, λ). Now, the proof can be divided into two cases.
Case 1. If x0 ∈ 𝕋, then by virtue of (15), we get that u(x0 + 1, λ)u(x0 − 1, λ) < 0.
Case 2. If x0 ∈ (t, t + 1) for some t ∈ 𝕋. Then (∂/∂x)u(x, λ) exists at x = x0 and is not zero by the definition of u(x, λ).
Using the same method used in [1, Page 102], we may prove the following.
Lemma 12. Let x(λ) be the zeros of u(t + 1, λ). Then x(λ) is a continuous function of λ.
Lemma 13. For λ > 0, the zeros, x(λ), of u(x, λ), x ∈ (0, T + 1] is a decreasing function of λ. For λ < 0, x(λ) is an increasing function of λ for x ∈ (0, T + 1].
Proof. Let the zero x(λ) occur in (t, t + 1]. Since u(x, λ) is linear in (t, t + 1], the location of this zero is given by
Now, we can set up the oscillatory characterization of the eigenvalues of (1) and (2).
Lemma 14. Let ν ∈ {+, −}. The sequence
Proof. To prove the times of changes of sign of (33), it is equivalent to find the number of zeros of u(x, λ), x ∈ (1, T + 1). We only deal with the case that ν = +; the case ν = − is similar.
By Lemma 12, for k = 1,2, 3, …, p − 1, there exist p − 1 functions xk,+(λ), which satisfy xk,+(λk,+) = T + 1 and are all decreasing function of λ; moreover, for fixed λ, there are the zeros of u(x, λ). Since u(0, λ) = 0 and u(1, λ) = 1, it follows that xk,+(λ)∈(1, T + 1) for λ > λk,+.
For 0 ≤ λ ≤ λ1,+, by Lemma 8, we get that u(x, λ) does not have a zero in (1, T + 1).
Let λ⋄ ∈ (λ1,+, λ2,+] be arbitrary. Since x1,+(λ) is continuous and decreasing, x1,+(λ) will intersect with λ = λ⋄ at (λ⋄, x1,+(λ⋄)). Moreover, x1,+(λ1,+) = T + 1 and λ⋄ > λ1,+, which implies that x1,+(λ⋄) < T + 1. Thus, for λ = λ2,+, (33) changes its sign exactly one time.
Now, we claim that for the same λ, xk,+(λ) and xk+1,+(λ) have no common zero for each k ∈ {1,2, …, p − 1}.
Suppose the contrary, then there exists λ* ∈ (λ2,+, λp,+) such that
Case 1. If there exists t0 ∈ 𝕋 such that , then by the definition of u(x, λ), we obtain that the signs of u(t0, λ*) and u(t0 + 1, λ*) are opposite. Without loss of generality, suppose that u(t0, λ*) > 0 and u(t0 + 1, λ*) < 0.
Now, consider the variation of when λ varies. Take ε > 0 sufficiently small, by the continuity of with respect to λ, for λ ∈ (λ* − ε, λ*), , and also u(t0, λ) > 0, u(t0 + 1, λ) < 0 since u(t, λ) is a continuous function of λ. However, for λ ∈ (λ* − ε, λ*), is the (k* + 1)th zero of u(x, λ), which implies that u(t0, λ) < 0, u(t0 + 1, λ) > 0, a contradiction.
Case 2. If there exists t0 ∈ 𝕋 such that , then we consider the sign of u(t0 − 1, λ*) and u(t0 + 1, λ*), and we can get the similar contradiction as in Case 1.
This claim implies that, for the same λ, xk,+(λ), k = 1,2, …, p − 1 do not intersect with each other. Thus, for λk−1,+ < λ⋄ ≤ λk,+, there are k − 1 functions: xj,+, j = 1,2, …, k − 1 which intersect with λ = λ⋄ in (1, T + 1) at k − 1 different points; that is, for λk−1,+ < λ⋄ ≤ λk,+, u(x, λ⋄) has exactly k − 1 zeros.
This completes the proof.
Lemma 15. Problems (1) and (2) have exactly n positive eigenvalues and exactly T − n negative eigenvalues.
Proof. First we show that ψp,+ changes its sign exactly n − 1 times.
Let us consider the following T + 2 ordered polynomials:
Observation 1. Consider
Observation 2. For j ∈ {2, …, T}, denoted by Γ(j), the number of the elements in the set
Next, using the result of first step and Lemma 14, it follows that p = n.
4. Application
As an application, we consider the existence of sign-changing solutions of the discrete nonlinear boundary value problems (12), (13).
- (H1)
f ∈ C([0, ∞), (−∞, +∞)) with f(0) = f(s1) = f(s2) = 0, 0 < s1 ≤ s2; f(s) > 0 for s ∈ (0, s1)∪(s2, +∞), f(s) < 0 for s ∈ (s1, s2);
- (H2)
there exist f0, f∞ ∈ (0, ∞) such that
()
for simplicity, we give some notations at first.
- (1)
u has exactly k − 1 simple generalized zeros in (1, T + 1);
- (2)
νu(1) > 0.
Define . They are disjoint in X. Finally, let and let Ψk = ℝ × Sk.
Theorem 16. Suppose that (H0), (H1), and (H2) hold. Assume that f0 < f∞. Then
- (i)
if
() -
problems (12) and (13) have at least four sign-changing solutions , , , and ;
- (ii)
if
() -
problems (12) and (13) have at least four sign-changing solutions , , , and .
Moreover, for r ∈ (λk,+/f∞, λk,+/f0], there also exist at least two sign-changing solutions and ; meanwhile, for r ∈ [λl,−/f0, λl,−/f∞), there also exist at least two sign-changing solutions and .
Theorem 17. Suppose that (H0), (H1), and (H2) hold. Assume that f∞ < f0. Then
- (i)
if
() -
problems (12) and (13) have at least four sign-changing solutions , , , and ;
- (ii)
if
() -
problems (12) and (13) have at least four sign-changing solutions , , , and .
Moreover, for r ∈ (λk,+/f0, λk,+/f∞], there exist at least two sign-changing solutions and ; meanwhile, for r ∈ [λl,−/f∞, λl,−/f0), there also exist at least two sign-changing solutions and .
-
f0 ∈ (0, ∞), f∞ = ∞,
then we obtain the following result.
Theorem 18. Let (H0), (H1), and (2) hold. Then (12) and (13) have a sign-changing solution in , (k = 1,2, …, n) if and only if r ≠ 0. Moreover, for r ∈ (−∞, λl,−/f0), there exist at least two solutions and , and there also exist at least two solutions and for r ∈ (λk,+/f0, +∞).
Lemma 19. Suppose that (r, u) is a nontrivial solution of (12) and (13); then there exists k0 ∈ {1,2, …, max{n, T − n}} such that
Proof. Suppose on the contrary that for every k ∈ {1,2, …, max{n, T − n}}
Lemma 20. Suppose that (H0), (H1), and (H2) hold. Then for ,
Proof. Suppose on the contrary that there exists such that
Let e : 𝕋 → (0, ∞), such that
Problem (70) is equivalent to
There exists such that
In the following we will investigate other sign-changing solutions of problems (1) and (2).
Lemma 21. Suppose that (H0), (H1), and (H2) hold. Then for , we have
Proof. It is similar to the proof of Lemma 20, so we omit it.
Lemma 22. Suppose (H0), (H1), and (H2) hold. Then
Proof. Firstly, we will prove . Take Λ ⊂ ℝ as an interval such that Λ∩{(λj,+/f∞) | j ∈ {1,2, …, n}} = {λk,+/f∞} and ℳ is a neighborhood of (λk,+/f∞, ∞) whose projection on ℝ lies in Λ and whose projection on X is bounded away from 0. Then by [25, Theorem 1.6 and Corollary 1.8], we have that either
- (1)
is bounded in ℝ × E in which case meets{(λ, 0) | λ ∈ ℝ}; or
- (2)
is unbounded.
Moreover if (2) occurs and has a bounded projection on ℝ, then meets where .
Obviously Lemma 21 implies that (1) does not occur. So is unbounded.
Lemma 19 guarantees that is a component of solutions of (80) in which meets (λk,+/f∞, ∞). Therefore there is no j ∈ {1,2, …, n}∖{k} such that also meets (λj,+/f∞, ∞). Otherwise, there will exist such that y has a multiple zero point t0 ∈ (0, T + 1); that is, y(t0) = 0 and y(t0 − 1)y(t0 + 1) ≥ 0. However this contradicts with Lemma 19, and consequently is unbounded. Thus
Conflict of Interests
The authors declare that there is no conflict of interests regarding the publication of this paper.
Acknowledgments
This work is supported by NSFC (nos. 11061030, 11326127, and 11361054), SRFDP (no. 20126203110004), and Gansu Provincial National Science Foundation of China (no. 1208RJZA258).