Ground State Solutions for the Periodic Discrete Nonlinear Schrödinger Equations with Superlinear Nonlinearities
Abstract
We consider the periodic discrete nonlinear Schrödinger equations with the temporal frequency belonging to a spectral gap. By using the generalized Nehari manifold approach developed by Szulkin and Weth, we prove the existence of ground state solutions of the equations. We obtain infinitely many geometrically distinct solutions of the equations when specially the nonlinearity is odd. The classical Ambrosetti-Rabinowitz superlinear condition is improved.
1. Introduction
We consider (7) as a nonlinear equation in the space l2 of two-sided infinite sequences. Note that every element of l2 automatically satisfies (6).
As it is well known, the operator L is a bounded and self-adjoint operator in l2. The spectrum σ(L) is a union of a finite number of closed intervals, and the complement ℝ∖σ(L) consists of a finite number of open intervals called spectral gaps. Two of them are semi-infinite (see [2]). If T = 1, then finite gaps do not exist. However, in general, finite gaps exist, and the most interesting case in (7) is when the frequency ω belongs to a finite spectral gap. Let us fix any spectral gap and denote it by (α, β).
DNLS equation is one of the most important inherently discrete models. DNLS equation plays a crucial role in the modeling of a great variety of phenomena, ranging from solid state and condensed matter physics to biology (see [1, 3–6] and references therein). In the past decade, solitons of the periodic DNLS have become a hot topic. The existence of solitons for the periodic DNLS equations with superlinear nonlinearity [7–10] and with saturable nonlinearity [11–13] has been studied, respectively. If ω is below or above the spectrum of the difference operator −Δ + εn, solitons were shown by using the Nehari manifold approach and a discrete version of the concentration compactness principle in [14]. If ω is a lower edge of a finite spectral gap, the existence of solitons was obtained by using variant generalized weak linking theorem in [10]. If ω lies in a finite spectral gap, the existence of solitons was proved by using periodic approximations in combination with the linking theorem in [8] and the generalized Nehari manifold approach in [9], respectively. The results were extended by Chen and Ma in [7]. In this paper, we employ the generalized Nehari manifold approach instead of periodic approximation technique to obtain the existence of a kind of special solitons of (7), which called ground state solutions, that is, nontrivial solutions with least possible energy in l2. We should emphasize that the results are obtained under more general super nonlinearity than the classical Ambrosetti-Rabinowitz superlinear condition [8, 9, 15].
This paper is organized as follows. In Section 2, we first establish the variational framework associated with (7) and transfer the problem on the existence of solutions in l2 of (7) into that on the existence of critical points of the corresponding functional. We then present the main results of this paper and compare them with existing ones. Section 3 is devoted to the proofs of the main results.
2. Preliminaries and Main Results
-
(V1 ω ∈ (α, β),
-
f1 fn ∈ C(ℝ, ℝ) and fn+T(u) = fn(u), and there exist a > 0 and p ∈ (2, ∞) such that
() -
f2 fn(u) = o(|u|) as u → 0,
-
f3 lim |u|→∞Fn(u)/u2 = ∞, where Fn(u) is the primitive function of fn(u), that is,
() -
f4 u ↦ fn(u)/|u| is strictly increasing on (−∞, 0) and (0, ∞).
In this paper, we also consider the multiplicity of solutions of (7).
Now, we are ready to state the main results.
Theorem 1. Suppose that conditions (V1), (f1)–(f4) are satisfied. Then, one has the following conclusions.
Theorem 2. Suppose that conditions (V1), (f1)–(f4) are satisfied and fn is odd in u. If either σ = 1 and β ≠ ∞ or σ = −1 and α ≠ −∞, then (7) has infinitely many pairs of geometrically distinct solutions.
In what follows, we always assume that σ = 1. The other case can be reduced to σ = 1 by switching L to −L and ω to −ω.
Remark 3. In [8], the author considered (7) with fn defined by
Remark 4. In [9], the author assumed that fn satisfies the following condition: there exists θ ∈ (0,1) such that
Remark 5. In [7], it is shown that (7) has at least a nontrivial solution u ∈ l2 if f satisfies (V1), (f2), (f3), and the following conditions:
-
(B1) Fn(u) ≥ 0 for any u ∈ ℝ and Hn(u): = (1/2)fn(u)u − Fn(u) > 0 if u ≠ 0,
-
(B2) Hn(u) → ∞ as |u | → ∞, and there exist r0 > 0 and γ > 1 such that |fn(u)|γ/|u|γ ≤ c0Hn(u) if |u | ≥ r0, where c0 is a positive constant,
In our paper, we use (9) and (f4) instead of (B1) and (B2).
3. Proofs of Main Results
We assume that (V1) and (f1)–(f4) are satisfied from now on.
Lemma 6. Fn(u) > 0 and (1/2)fn(u)u > Fn(u) for all u ≠ 0.
Proof. By (f2) and (f4), it is easy to get that
To continue the discussion, we need the following proposition.
Proposition 7 (see [16], [17].)Let u, s, v ∈ ℝ be numbers with s ≥ −1 and w : = su + v ≠ 0. Then,
Lemma 8. If u ∈ ℳ, then
Proof. We rewrite J by
Since u ∈ ℳ, we have
Together with Proposition 7, we know that
The proof is complete.
Lemma 9. (a) There exists α > 0 such that , where Sα : = {u ∈ E+ : ∥u∥ = α}.
(b) for every u ∈ ℳ.
Proof. (a) By (f1) and (f2), it is easy to show that for any ε > 0, there exists cε > 0 such that
The first inequality is a consequence of Lemma 8 since for every u ∈ ℳ, there is s > 0 such that .
(b) For u ∈ ℳ, by (25), we have
Lemma 10. Let 𝒲 ⊂ E+∖{0} be a compact subset. Then, there exists R > 0 such that J ≤ 0 on E(u)∖BR(0) for every u ∈ 𝒲, where BR(0) denotes the open ball with radius R and center 0.
Proof. Suppose by contradiction that there exist u(k) ∈ 𝒲 and w(k) ∈ E(u(k)), k ∈ ℕ, such that J(w(k)) > 0 for all k and ∥w(k)∥ → ∞ as k → ∞. Without loss of generality, we may assume that ∥u(k)∥ = 1 for k ∈ ℤ. Then, there exists a subsequence, still denoted by the same notation, such that u(k) → u ∈ E+. Set v(k) = w(k)/∥w(k)∥ = s(k)u(k) + v(k)−. Then,
Lemma 11. For each u ∈ E+∖{0}, the set consists of precisely one point which is the unique global maximum of .
Proof. By Lemma 8, it suffices to show that . Since , we may assume that u ∈ S+. By Lemma 10, there exists R > 0 such that J ≤ 0 on E(u)∖BR(0) provided that R is large enough. By Lemma 9 (a), J(tu) > 0 for small t > 0. Moreover, J ≤ 0 on . Hence, .
Let v(k)⇀v in . Then, as k → ∞ for all n after passing to a subsequence if necessary. Hence, . Let φ(v) = ∑n∈ℤ χnFn(vn). Then,
According to Lemma 11, for each u ∈ E+∖{0}, we may define the mapping , , where is the unique point of .
Lemma 12. J is coercive on ℳ; that is, J(u) → ∞ as ∥u∥ → ∞, u ∈ ℳ.
Proof. Suppose, by contradiction, that there exists a sequence {u(k)} ⊂ ℳ such that ∥u(k)∥ → ∞ and J(u(k)) ≤ d for some d ∈ [c, ∞). Let v(k) = u(k)/∥u(k)∥. Then, there exists a subsequence, still denoted by the same notation, such that v(k)⇀v and for every n as k → ∞.
First, we know that there exist δ > 0 and nk ∈ ℤ such that
Since for s ≥ 0, Lemma 8 implies that
Due to the periodicity of coefficients, both J and ℳ are invariant under T-translation. Making such shifts, we can assume that 1 ≤ nk ≤ T − 1 in (39). Moreover, passing to a subsequence if needed, we can assume that nk = n0 is independent of k. Next, we may extract a subsequence, still denoted by {v(k)}, such that for all n ∈ ℤ. In particular, for n = n0, inequality (39) shows that and hence v+ ≠ 0.
Since as k → ∞, it follows again from (f3) and Fatou’s lemma that
Lemma 13. (a) The mapping is continuous.
(b) The mapping is a homeomorphism between S+ and ℳ, and the inverse of m is given by m−1(u) = u+/∥u+∥, where S+ : = {u ∈ E+ : ∥u∥ = 1}.
(c) The mapping m−1 : ℳ ↦ S+ is the Lipschitz continuous.
Proof. (a) Let (u(k)) ⊂ E+∖{0} be a sequence with u(k) → u. Since , without loss of generality, we may assume that ∥u(k)∥ = 1 for all k. Then, . By Lemma 10, there exists R > 0 such that
(b) This is an immediate consequence of (a).
(c) For u, v ∈ ℳ, by (b), we have
Lemma 14. (a) and
(b) Ψ ∈ C1(S+, ℝ) and
(c) {wn} is a Palais-Smale sequence for Ψ if and only if {m(wn)} is a Palais-Smale sequence for J.
(d) w ∈ S+ is a critical point of Ψ if and only if m(w) ∈ ℳ is a nontrivial critical point of J. Moreover, the corresponding values of Ψ and J coincide and .
Proof. (a) We put , so we have u = (∥u+∥/∥w∥)w + u−. Let z ∈ E+. Choose δ > 0 such that wt : = w + tz ∈ E+∖{0} for |t | < δ and put . We may write with st > 0. From the proof of Lemma 13, the function t ↦ st is continuous. Then, s0 = ∥u+∥/∥w∥. By Lemma 11 and the mean value theorem, we have
(b) It follows from (a) by noting that since w ∈ S+.
(c) Let {wn} be a Palais-Smale sequence for Ψ, and let un = m(wn) ∈ ℳ. Since for every n ∈ ℤ, we have an orthogonal splitting ; using (b), we have
(d) By (57), Ψ′(w) = 0 if and only if J′(m(w)) = 0. The other part is clear.
Proof of Theorem 1. (1) We know that c > 0 by Lemma 9(a). If u0 ∈ ℳ satisfies J(u0) = c, then m−1(u0) ∈ S+ is a minimizer of Ψ and therefore a critical point of Ψ and also a critical point of J by Lemma 14. We shall show that there exists a minimizer u ∈ ℳ of J|ℳ. Let {w(k)} ⊂ S+ be a minimizing sequence for Ψ. By Ekeland’s variational principle, we may assume that Ψ(w(k)) → c and Ψ′(w(k)) → 0 as k → ∞. Then, J(u(k)) → c and J′(u(k)) → 0 as k → ∞ by Lemma 14(c), where u(k) : = m(w(k)) ∈ ℳ. By Lemma 12, {u(k)} is bounded, and hence {u(k)} has a weakly convergent subsequence.
First, we show that there exist δ > 0 and nk ∈ ℤ such that
From the periodicity of the coefficients, we know that J and J′ are both invariant under T-translation. Making such shifts, we can assume that 1 ≤ nk ≤ T − 1 in (58). Moreover, passing to a subsequence, we can assume that nk = n0 is independent of k.
Next, we may extract a subsequence, still denoted by {u(k)}, such that u(k)⇀u and for all n ∈ ℤ. Particularly, for n = n0, inequality (58) shows that , so u ≠ 0. Moreover, we have
Finally, we show that J(u) = c. By Lemma 6 and Fatou’s lemma, we have
(2) If β = ∞, by way of contradiction, we assume that (7) has a nontrivial solution u ∈ E. Then, u is a nonzero critical point of J in E. Thus, J′(u) = 0. But by Lemma 6,
This completes the proof of Theorem 1.
It is easy to see that ν(a) < ∞ for every a by Lemma 12.
Proof of Theorem 2. It is easy to see that mappings m, m−1 are equivariant with respect to the ℤ-action by Lemma 13; hence, the orbits 𝒪(u) ⊂ ℳ consisting of critical points of J are in 1-1 correspondence with the orbits 𝒪(w) ⊂ S+ consisting of critical points of Ψ by Lemma 14(d). Next, we may choose a subset ℱ ⊂ K such that ℱ = −ℱ and ℱ consists of a unique representative of ℤ-orbits. So, we only need to prove that the set ℱ is infinite. By contradiction, we assume that
Now, we claim that
Firstly, we show that
Next, we consider a pseudogradient vector field of Ψ [18]; that is, there exists a Lipschitz continuous map V: S+∖K → TwS+ and for all w ∈ S+∖K,
This completes the proof of Theorem 2.
Acknowledgments
The authors would like to thank the anonymous referees for their constructive comments and suggestions, which considerably improved the presentation of the paper. This work is supported by the Program for Changjiang Scholars and Innovative Research Team in University (no. IRT1226), the National Natural Science Foundation of China (no. 11171078), and the Specialized Fund for the Doctoral Program of Higher Education of China (no. 20114410110002).
Appendix
Here, we give a proof of (74). We state the discrete property of the Palais-Smale sequences. It yields nice properties of the corresponding pseudogradient flow.
Lemma A.1. Let d ≥ c. If , are two Palais-Smale sequences for Ψ, then either as k → ∞ or , where ϱ(d) depends on d but not on the particular choice of the Palais-Smale sequences.
Proof. Set and . Then, are the bounded Palais-Smale sequences for J. We fix p in (f2) and consider the following two cases.
(i) as k → ∞.
By a straightforward calculation and (32), for any ε > 0, there exist C1, C2 > 0, and k0 such that for all k ≥ k0,
This implies . Hence, . Similarly, . Therefore, as k → ∞. By Lemma 13(c), we have as k → ∞.
(ii) as k → ∞.
There exist δ > 0 and nk ∈ ℤ such that
If u1 ≠ 0 and u2 ≠ 0. Then, u1, u2 ∈ ℳ and w1 = m−1(u1) ∈ K, w2 = m−1(u2) ∈ K, w1 ≠ w2. Therefore,
If u1 = 0, then u2 ≠ 0 and
The proof is complete.
Lemma A.2. For every w ∈ S+, the limit exists and is a critical point of Ψ.
Proof. Fix w ∈ S+ and set d = Ψ(w). We distinguish two cases to finish the proof.
Case 1 (T+(w) < ∞). For 0 ≤ s < t < T+(w), by (72) and (73), we have
Case 2 (T+(w) = ∞). To prove that exists, we claim that for every ε > 0, there exists tε > 0 such that ∥η(tε, w) − η(t, w)∥ < ε for t ≥ tε. If not, then there exist 0 < ε0 < (1/2)ϱ(d) (ϱ(d) is the same number in Lemma A.1) and a sequence {tn}⊂[0, ∞) with tn → ∞ such that ∥η(tn, w) − η(tn+1, w)∥ = ε0 for every n. Choose the smallest such that . Let . By (72) and (73), we have
Lemma A.3. Let d ≥ c. Then, for every δ > 0, there exists ε = ε(δ) > 0 such that
- (a)
,
- (b)
for w ∈ Ψd+ε∖Nδ(Kd).
Proof. (a) According to (66), for ε > 0 small enough, it is easy to see that (a) is satisfied.
(b) Without loss of generality, we may assume that Nδ(Kd) ⊂ Ψd+1 and δ < ϱ(d + 1). Set
We claim that τ > 0. Indeed, if not, then there exists a sequence such that . By the ℤ-invariance of Ψ and assumption (66), we may assume for some w0 ∈ Kd after passing to a subsequence. Let . Then, and
Let
Then,
It follows that Ψ(η(t2, w)) ≤ d + ε − (δτ2/8M) < d and therefore , a contradiction again. This completes the proof.