Skip to main content

On locally superquadratic Hamiltonian systems with periodic potential

Abstract

In this paper, we study the second-order Hamiltonian systems

$$ \ddot{u}-L(t)u+\nabla W(t,u)=0, $$

where \(t\in \mathbb{R}\), \(u\in \mathbb{R}^{N}\), L and W depend periodically on t, 0 lies in a spectral gap of the operator \(-d^{2}/dt^{2}+L(t)\) and \(W(t,x)\) is locally superquadratic. Replacing the common superquadratic condition that \(\lim_{|x|\rightarrow \infty }\frac{W(t,x)}{|x|^{2}}=+\infty \) uniformly in \(t\in \mathbb{R}\) by the local condition that \(\lim_{|x|\rightarrow \infty }\frac{W(t,x)}{|x|^{2}}=+\infty \) a.e. \(t\in J\) for some open interval \(J\subset \mathbb{R}\), we prove the existence of one nontrivial homoclinic soluiton for the above problem.

1 Introduction and main results

Consider the second-order Hamiltonian systems

$$\begin{aligned} \ddot{u}-L(t)u+\nabla W(t,u)=0, \end{aligned}$$
(1.1)

where \(t\in \mathbb{R}\), \(u\in \mathbb{R}^{N}\), \(L\in C(\mathbb{R},\mathbb{R}^{N\times N})\) and \(W\in C^{1}(\mathbb{R}\times \mathbb{R}^{N},\mathbb{R})\) satisfies the following basic conditions:

  1. (W1)

    W is T-periodic in t and there exist constants \(C_{0}>0\) and \(p>2\) such that

    $$ \bigl\vert \nabla W(t,x) \bigr\vert \leq C_{0} \bigl(1+ \vert x \vert ^{p-1} \bigr), \quad\forall (t,x)\in \mathbb{R}\times \mathbb{R}^{N}. $$
  2. (W2)

    \(\nabla W(t,x)=o(|x|)\) as \(x\rightarrow 0\) uniformly in t and \(W(t,x)\geq 0\) for all \((t,x)\).

Usually, a solution u of system (1.1) is said to be homoclinic to 0 if \(u(t)\rightarrow 0\) as \(|t|\rightarrow \infty \). Furthermore, if \(u(t)\not \equiv 0\), then u is called a nontrivial homoclinic solution.

During the past two decades, there has been a remarkable amount of progress in the study of homoclinic motions of Hamiltonian systems, with many new ideas and methods being introduced; see, for instance, [2, 3, 5, 6, 8, 1215, 20, 2226] for results concerning the second-order systems, and [4, 7, 1719] for the first-order systems. For (1.1) with periodic potential, most results are obtained under the assumptions that \(L(t)\) is positive definite for all \(t\in \mathbb{R}\) and \(W(t,x)\) is globally superquadratic in x; see, e.g., [2, 3, 5, 8, 10, 12, 14, 22, 24, 25]. A common feature of this work is that the following assumption, which is originally due to Ambrosetti and Rabinowitz [1], is imposed on the nonlinearity:

  1. (AR)

    \(\mu >2\) such that \(0<\mu W(t,x)\leq (\nabla W(t,x),x)\) for all \((t,x)\in \mathbb{R}\times (\mathbb{R}^{N}\backslash \{ 0 \} )\).

It is well known that the crucial role of (AR) is to verify the mountain-pass geometry of the corresponding functional and also to ensure the boundedness of Palais–Smale (PS) sequences. It is also well known that many functions such as

$$ W(t,x)= \vert x \vert ^{2}\ln \bigl(1+ \vert x \vert \bigr) $$

do not satisfy (AR). In recent years, many more natural conditions than (AR) have been proposed in the study of periodic or homoclinic solutions for superquadratic Hamiltonian systems. Ding and Lee [8] consider (1.1) with periodicity. Instead of (AR), they assume that

  1. (S1)

    \(W(t,x)/|x|^{2}\rightarrow +\infty \) as \(|x|\rightarrow \infty \) uniformly in t;

  2. (S2)

    \(\widetilde{W}(t,x):=\frac{1}{2}(\nabla W(t,x),x)-W(t,x)>0\) if \(x\neq 0\), and there exist \(\varepsilon \in (0,1)\) and \(r>0\) such that \((\nabla W(t,x),x)\leq c\widetilde{W}(t,x)|x|^{2-\varepsilon }\) for all \(t\in \mathbb{R}\) and \(|x|\geq r\);

and prove the existence of infinitely many geometrically distinct solutions for both asymptotically quadratic and superquadratic cases. See also [12, 22, 25] for the related results.

If 0 lies in a spectral gap of the operator \(-d^{2}/dt^{2}+L(t)\), that is,

  1. (L)

    \(-\varLambda _{1}:=\sup [\sigma (A)\cap (-\infty,0)]<0< \varLambda _{2}:= \inf [\sigma (A)\cap (0,+\infty )]\), where \(A:=-d^{2}/dt^{2}+L(t)\) and σ denotes the spectrum,

then the negative space \(E^{-}\) of the quadratic form in the energy functional given by (2.5) is infinite dimensional. For this reason, we say that the problem is strongly indefinite. Up to now, few papers deal with this situation; see [2, 3, 10, 24]. Besides the conditions (L), (AR) and the restrictive assumption of \(W(t,x)\geq c|x|^{\mu }\) and \(|\nabla W(t,x)|\leq c|x|^{\mu -1}\) for all \((t,x)\), Arioli and Szulkin [2] prove the existence of one nontrivial homoclinic orbit by constructing subharmonics and passing to the limit. Recently, Chen [3] proves the existence of one nontrivial ground state homoclinic orbit under hypotheses (L), (W2), (S1) and

  1. (S3)

    \(\widetilde{W}(t,x)>0\) if \(x\in \mathbb{R}^{N}\backslash \{ 0 \} \), and there exist c, \(r>0\) and \(k<1\) such that

    $$ \frac{ \vert \nabla W(t,x) \vert ^{k}}{ \vert u \vert ^{k}}\leq c\widetilde{W}(t,x),\quad \vert x \vert \geq r. $$

Inspired by the works mentioned above and the recent paper [21], we are interested in the case where 0 lies in a gap of \(\sigma (A)\) and \(W(t,x)\) is locally superquadratic, i.e., it is allowed to be superquadratic at some \(t\in \mathbb{R}\) and asymptotically quadratic at other \(t\in \mathbb{R}\). The main ingredient is the observation that even in the strongly indefinite case, all Cerami sequences of the energy functional are bounded. Therefore, the existence of one homoclinic solution is proved by using the generalized linking theorem of Li and Szulkin (see [11]). Precisely, we further weaken (S1) to the following hypotheses:

  1. (W3)

    There exists an open interval \(J\subset \mathbb{R}\) such that \(\lim_{|x|\rightarrow \infty }\frac{W(t,x)}{|x|^{2}}=+\infty \) a.e. \(t\in J\).

  2. (W4)

    \(\widetilde{W}(t,x)\geq 0\) for all \((t,x)\), and there exist \(C_{1}>0\), \(\delta \in (0,\varLambda _{0})\)\((\varLambda _{0}:=\min \{ \varLambda _{1},\varLambda _{2} \} )\), \(\sigma \in (0,1)\) and \(k\in (1,2/(1-\sigma )]\) such that

    $$ \frac{ \vert \nabla W(t,x) \vert }{ \vert x \vert }\geq \varLambda _{0}-\delta\quad \text{implies}\quad \biggl( \frac{ \vert \nabla W(t,x) \vert }{ \vert x \vert ^{\sigma }} \biggr)^{k}\leq C_{1} \widetilde{W}(t,x). $$

Our main result reads as follows.

Theorem 1.1

Assume that (L) and (W1)–(W4) are satisfied. Then system (1.1) has at least one nontrivial homoclinic solution.

Remark 1.1

(i) Comparing with the results of [2, 3, 10, 24], one advantage of Theorem 1.1 is that the globally superquadratic condition (cf. (AR) or (S1)) is replaced by the local one (W3). Thus our result applies to more general situations. Typical examples, which match our assumptions (W1)–(W4), but satisfying none of (AR), (S1), (S2) and (S3), are the following:

$$\begin{aligned} W(t,x)= \biggl(\sin \frac{2\pi t}{T}+ \biggl\vert \sin \frac{2\pi t}{T} \biggr\vert \biggr) \bigl(2 \bigl( \vert x \vert ^{2}-1 \bigr)\ln \bigl(1+ \vert x \vert \bigr)+ \vert x \vert \bigl(2- \vert x \vert \bigr) \bigr) \end{aligned}$$
(1.2)

and

$$\begin{aligned} W(t,x)= \vert x \vert ^{2+\alpha (t)} \biggl[1- \frac{1}{\ln (e+ \vert x \vert ^{2})} \biggr], \end{aligned}$$
(1.3)

where \(\alpha \in C(\mathbb{R},\mathbb{R})\), \(F\subset \mathbb{R}\) is a closed set such that \(\alpha (t)=0\) for \(t\in F\) and \(\alpha (t)\in (0,2)\) for \(t\in \mathbb{R}\backslash F\). One can easily check this fact for (1.2) by noting that

$$\begin{aligned} &\nabla W(t,x)= \biggl(\sin \frac{2\pi t}{T}+ \biggl\vert \sin \frac{2\pi t}{T} \biggr\vert \biggr)4x\ln \bigl(1+ \vert x \vert \bigr), \\ &\widetilde{W}(t,x)= \biggl(\sin \frac{2\pi t}{T}+ \biggl\vert \sin \frac{2\pi t}{T} \biggr\vert \biggr) \bigl( \vert x \vert ^{2}+2 \ln \bigl(1+ \vert x \vert \bigr)-2 \vert x \vert \bigr), \end{aligned}$$

and for (1.3) by noting that

$$\begin{aligned} &\nabla W(t,x)= \bigl(2+\alpha (t) \bigr) \vert x \vert ^{\alpha (t)}x \biggl[1- \frac{1}{\ln (e+ \vert x \vert ^{2})} \biggr]+ \frac{2 \vert x \vert ^{2+\alpha (t)}x}{(e+ \vert x \vert ^{2}) [\ln (e+ \vert x \vert ^{2}) ]^{2}}, \\ &\widetilde{W}(t,x)=\frac{\alpha (t)}{2} \vert x \vert ^{\alpha (t)+2} \biggl[1- \frac{1}{\ln (e+ \vert x \vert ^{2})} \biggr]+ \frac{ \vert x \vert ^{4+\alpha (t)}}{(e+ \vert x \vert ^{2}) [\ln (e+ \vert x \vert ^{2}) ]^{2}}. \end{aligned}$$

In addition, we point out that the function of (1.3) is asymptotically quadratic for \(t\in F\) and superquadratic for \(t\in \mathbb{R}\backslash F\).

(ii) Another advantage of this paper is that our argument is simpler. In [3], Chen discusses a family of perturbed functions

$$ \varphi _{\lambda }(u)=\frac{1}{2} \bigl\Vert u^{+} \bigr\Vert ^{2}-\lambda \biggl( \frac{1}{2} \bigl\Vert u^{-} \bigr\Vert ^{2}+ \int _{\mathbb{R}} W(t,x)\,dt \biggr),\quad \lambda \in [1,2] $$

and apply a variant generalized weak linking theorem for strongly indefinite functionals developed by Schechter and Zou (see [16]). This approach is not very satisfactory, since working with a family of perturbed functionals makes things unnecessary complicated. In the present paper we will prove Theorem 1.1 by directly applying the usual variational method to the energy functional φ. The key point in our proof is that, although φ may has unbounded (PS) sequences, we can prove that all Cerami sequences of φ are bounded (see Lemma 2.4 below), and hence Theorem 1.1 follows directly from the generalized linking theorem (see [11]).

(iii) Homoclinics for locally superquadratic Hamiltonian systems has been studied in Wang [22] for periodic case. However, this paper only deals with the definite case, which is much simpler than the strongly indefinite case considered in the present paper.

Notation: “→’ and “”, respectively, denote the strong convergence and the weak convergence. C and \(C_{i}\)\((i=1,2,\ldots )\) denote various positive constants which may vary from place to place.

2 Proof of Theorem 1.1

Let \(A:=-d^{2}/dt^{2}+L(t)\). Then A is self-adjoint in \(L^{2}(\mathbb{R},\mathbb{R}^{N})\) with domain \(\mathcal{D}(A)=H^{2}(\mathbb{R},\mathbb{R}^{N})\). Let \(\{ \mathcal{E}(\lambda ):-\infty \leq \lambda \leq +\infty \} \), \(|A|\) and \(|A|^{1/2}\), respectively, be the spectral family, the absolute value of A and the square root of \(|A|\). Take \(U:=\mathrm{id}-\mathcal{E}(0)-\mathcal{E}(0-)\). Then U commutes with A, \(|A|\) and \(|A|^{1/2}\), and \(A=U|A|\) is the polar decomposition of A (see [9, Theorem IV 3.3]). Set

$$ E=\mathcal{D} \bigl( \vert A \vert ^{1/2} \bigr),\qquad E^{-}= \mathcal{E}(0-)E \quad\text{and} \quad E^{+}= \bigl[\mathrm{id}-\mathcal{E}(0) \bigr]E. $$

For every \(u\in E\), we see that

$$\begin{aligned} u^{-}:=\mathcal{E}(0-)u\in E^{-}, \qquad u^{+}:= \bigl[\mathrm{id}-\mathcal{E}(0) \bigr]u \in E^{+}, \qquad u=u^{-}+u^{+} \end{aligned}$$
(2.1)

and

$$\begin{aligned} Au^{-}=- \vert A \vert u^{-}, \qquad Au^{+}= \vert A \vert u^{+}. \end{aligned}$$
(2.2)

Define on E the inner product and the norm

$$\begin{aligned} (u,v)= \bigl( \vert A \vert ^{1/2}u, \vert A \vert ^{1/2}v \bigr)_{L^{2}},\qquad \Vert u \Vert = \bigl\Vert \vert A \vert ^{1/2}u \bigr\Vert _{2}, \end{aligned}$$
(2.3)

where \((\cdot,\cdot )_{L^{2}}\) and \(\|\cdot \|_{s}\) denote the inner product of \(L^{2}(\mathbb{R},\mathbb{R}^{N})\) and the norm of \(L^{s}(\mathbb{R},\mathbb{R}^{N})\) (\(2\leq s\leq +\infty \)), respectively. Then E is a Hilbert space. By (L), \(E=H^{1}(\mathbb{R},\mathbb{R}^{N})\) with equivalent norms and E is continuously embedded in \(L^{s}(\mathbb{R},\mathbb{R}^{N})\) for \(2\leq s\leq +\infty \). It is easy to check that E has the following decomposition \(E=E^{-}\oplus E^{+}\) orthogonal with respect to both \((\cdot,\cdot )_{L^{2}}\) and \((\cdot,\cdot )\). Furthermore, it follows from the definitions of \(\varLambda _{1}\), \(\varLambda _{2}\) and \(\varLambda _{0}\) that

$$\begin{aligned} \varLambda _{1} \bigl\Vert u^{-} \bigr\Vert _{2}^{2}\leq \bigl\Vert u^{-} \bigr\Vert ^{2}, \qquad\varLambda _{2} \bigl\Vert u^{+} \bigr\Vert _{2}^{2}\leq \bigl\Vert u^{+} \bigr\Vert ^{2}, \qquad\varLambda _{0} \Vert u \Vert _{2}^{2} \leq \Vert u \Vert ^{2} \end{aligned}$$
(2.4)

for all \(u\in E\).

We shall apply the generalized linking theorem of Li and Szulkin to prove Theorem 1.1. First we introduce some notations. Let E be a real Hilbert space with \(E=E^{-}\oplus E^{+}\) and \(E^{-}\bot E^{+}\). For \(\varphi \in C^{1}(E,\mathbb{R}^{N})\), φ is said to be weakly sequentially lower semi-continuous if \(u_{n}\rightharpoonup u\) implies \(\varphi (u)\leq \liminf_{n\rightarrow \infty }\varphi (u_{n})\), and \(\varphi '\) is said to be weakly sequentially continuous if \(\lim_{n\rightarrow \infty }\langle \varphi '(u_{n}),v\rangle = \langle \varphi '(u),v\rangle \) for any \(v\in E\).

Theorem 2.1

(see [11, Theorem 2.1])

Let \((E,\|\cdot \|)\)be a real Hilbert space with \(E=E^{-}\oplus E^{+}\), and let \(\varphi \in C^{1}(E,\mathbb{R})\)of the form

$$ \varphi (u)=\frac{1}{2} \bigl( \bigl\Vert u^{+} \bigr\Vert ^{2}- \bigl\Vert u^{-} \bigr\Vert ^{2} \bigr)- \psi (u), $$

where \(u=u^{-}+u^{+}\in E^{-}\oplus E^{+}\). Suppose that

  1. (i)

    \(\psi \in C^{1}(E,\mathbb{R})\)is bounded from below, weakly sequentially lower semi-continuous and \(\psi '\)is weakly sequentially continuous;

  2. (ii)

    there exist \(e\in E^{+}\)with \(\|e\|=1\)and \(r>\rho >0\)such that \(\alpha:=\inf \varphi (S_{\rho }^{+})>\sup \varphi (\partial Q)\), where \(S_{\rho }^{+}= \{ u\in E^{+}:\|u\|=\rho \} \)and \(Q= \{ v+se:v\in E^{-},s\geq 0,\|v+se\|\leq r \} \).

Then, for some \(c>\alpha \), there is a sequence \((u_{n})\subset E\)such that

$$ \varphi (u_{n})\rightarrow c,\qquad \bigl\Vert \varphi '(u_{n}) \bigr\Vert \bigl(1+ \Vert u_{n} \Vert \bigr)\rightarrow 0. $$

Such a sequence is called a Cerami sequence on level c, or a \((C)_{c}\)sequence.

Now we define the functional \(\varphi:E\rightarrow \mathbb{R}\) by

$$\begin{aligned} \varphi (u)=\frac{1}{2} \int _{\mathbb{R}} \bigl[ \vert \dot{u} \vert ^{2}+ \bigl(L(t)u,u \bigr) \bigr]\,dt- \int _{\mathbb{R}} W(t,u)\,dt. \end{aligned}$$
(2.5)

In view of (L) and (W1)–(W2), \(\varphi \in C^{1}(E,\mathbb{R})\) and

$$ \bigl\langle \varphi '(u),v \bigr\rangle = \int _{\mathbb{R}} \bigl[(\dot{u},\dot{v})+ \bigl(L(t)u,v \bigr) \bigr] \,dt- \int _{\mathbb{R}} \bigl(\nabla W(t,u),v \bigr)\,dt,\quad u,v\in E. $$

Combining (2.1)–(2.3), we have

$$\begin{aligned} \varphi (u)=\frac{1}{2} \bigl( \bigl\Vert u^{+} \bigr\Vert ^{2}- \bigl\Vert u^{-} \bigr\Vert ^{2} \bigr)- \int _{ \mathbb{R}} W(t,u)\,dt \end{aligned}$$
(2.6)

and

$$ \bigl\langle \varphi '(u),v \bigr\rangle = \bigl(u^{+},v \bigr)- \bigl(u^{-},v \bigr)- \int _{\mathbb{R}} \bigl( \nabla W(t,u),v \bigr)\,dt, \quad\forall u,v \in E. $$

A standard argument shows that the critical points of φ are homoclinic solutions of (1.1) (see [5, 15]).

Let

$$ \psi (u)= \int _{\mathbb{R}}W(t,u) \,dt. $$

Obviously, \(\psi \geq 0\) and it follows from Fatou’s lemma that ψ is weakly sequentially lower semi-continuous. By (W1) and (W2), for any \(\varepsilon >0\), there exists \(C_{\varepsilon }>0\) such that

$$\begin{aligned} \bigl\vert \nabla W(t,x) \bigr\vert \leq \varepsilon \vert x \vert +C_{\varepsilon } \vert x \vert ^{p-1}, \qquad \bigl\vert W(t,x) \bigr\vert \leq \varepsilon \vert x \vert ^{2}+C_{\varepsilon } \vert x \vert ^{p} \end{aligned}$$
(2.7)

for all \((t,x)\in \mathbb{R}\times \mathbb{R}^{N}\). Since \(u_{n}\rightharpoonup u\) yields \(u_{n}\rightarrow u\) in \(L^{s}_{\mathrm{loc}}(\mathbb{R},\mathbb{R}^{N})\) for all \(s\in [1,\infty ]\), it is easy to check that \(\psi '\) is weakly sequentially continuous. Thus (i) of Theorem 2.1 is satisfied.

Next we study the linking structure of φ. Without loss of generality, we may suppose that \(J\subset \mathbb{R}\) is bounded. Choose \(e\in C_{0}^{\infty }(J,\mathbb{R}^{N})\) such that

$$\begin{aligned} \bigl\Vert e^{+} \bigr\Vert ^{2}- \bigl\Vert e^{-} \bigr\Vert ^{2}= \int _{\mathbb{R}} \bigl[ \vert \dot{e} \vert ^{2}+ \bigl(L(t)e,e \bigr) \bigr]\,dt= \int _{J} \bigl[ \vert \dot{e} \vert ^{2}+ \bigl(L(t)e,e \bigr) \bigr]\,dt\geq 1. \end{aligned}$$
(2.8)

Lemma 2.1

Let (L) and (W1)–(W2) be satisfied. Then there is \(\rho >0\)such that \(\alpha:={\inf \varphi (S^{+}_{\rho })> 0}\), where \(S^{+}_{\rho }= \{ u\in E^{+}:\|u\|=\rho \} \).

Proof

Since, by (2.7) and the Sobolev embedding inequality,

$$ \psi (u)\leq \varepsilon \Vert u \Vert _{2}^{2}+C_{\varepsilon } \Vert u \Vert _{p}^{p}=o \bigl( \Vert u \Vert ^{2} \bigr) \quad\text{as }n\rightarrow \infty, $$

the conclusion follows from the form of φ (see (2.6)). □

Lemma 2.2

Let (L) and (W2)–(W3) be satisfied. Then \(\sup \varphi (E^{-}\oplus \mathbb{R}^{+}e^{+})<+\infty \)and there exists \(R_{e}>0\)such that

$$ \varphi (u)\leq 0,\quad u\in E^{-}\oplus \mathbb{R}^{+}e^{+}, \Vert u \Vert \geq R_{e}. $$

Proof

It is sufficient to show that \(\varphi (u)\rightarrow -\infty \) as \(\|u\|\rightarrow \infty \), \(u\in E^{-}\oplus \mathbb{R}^{+}e^{+}\). Arguing indirectly, assume that, for some sequence \(\{ v_{n}+\theta _{n} e^{+} \} \subset E^{-}\oplus \mathbb{R}^{+}e^{+}\) with \(\|v_{n}+\theta _{n}e^{+}\|\rightarrow \infty \) as \(n\rightarrow \infty \), there is \(C_{2}>0\) such that

$$\begin{aligned} \varphi \bigl(v_{n}+\theta _{n}e^{+} \bigr)\geq -C_{2},\quad \forall n\in \mathbb{N}. \end{aligned}$$
(2.9)

Let \(w_{n}=(v_{n}+\theta _{n}e^{+})/\|v_{n}+\theta _{n}e^{+}\|=w_{n}^{-}+s_{n} e^{+}\). Then \(\|w_{n}\|=1\), and going if necessary to a subsequence, we may assume that

$$\begin{aligned} \begin{aligned} &s_{n}\rightarrow s_{0},\qquad w_{n}^{-}\rightharpoonup w^{-}\quad \text{in }E, w_{n}^{-}(t)\rightarrow w^{-}(t) \text{ a.e. }t\in \mathbb{R}, \\ &w_{n}^{-}\rightarrow w^{-}\quad\text{in }L^{2}\bigl(J,\mathbb{R}^{N}\bigr) \quad\text{and}\quad \dot{ \widehat{w_{n}^{-}}}\rightharpoonup \dot{ \widehat{w^{-}}}\quad\text{in }L^{2}\bigl(\mathbb{R}, \mathbb{R}^{N}\bigr). \end{aligned} \end{aligned}$$
(2.10)

By (2.9), we have

$$\begin{aligned} -\frac{C_{2}}{ \Vert v_{n}+\theta _{n}e^{+} \Vert ^{2}}\leq \frac{\varphi (v_{n}+\theta _{n} e^{+})}{ \Vert v_{n}+\theta _{n}e^{+} \Vert ^{2}}= \frac{s_{n}^{2}}{2} \bigl\Vert e^{+} \bigr\Vert ^{2}- \frac{1}{2} \bigl\Vert w_{n}^{-} \bigr\Vert ^{2}- \int _{\mathbb{R}} \frac{W(t,v_{n}+\theta _{n}e^{+})}{ \Vert v_{n}+\theta _{n}e^{+} \Vert ^{2}}\,dt. \end{aligned}$$
(2.11)

Since \(W\geq 0\), it follows that

$$ \frac{1}{2} \bigl\Vert w_{n}^{-} \bigr\Vert ^{2}\leq \frac{s_{n}^{2}}{2} \bigl\Vert e^{+} \bigr\Vert ^{2}+o(1), $$

and then

$$ \frac{1}{2}=\frac{1}{2} \bigl( \bigl\Vert w_{n}^{-} \bigr\Vert ^{2}+s_{n}^{2} \bigl\Vert e^{+} \bigr\Vert ^{2} \bigr)\leq s_{n}^{2} \bigl\Vert e^{+} \bigr\Vert ^{2}+o(1)=s_{0}^{2} \bigl\Vert e^{+} \bigr\Vert ^{2}+o(1), $$

which implies that \(s_{0}\neq 0\). We claim that

$$\begin{aligned} \bigl(w^{-}+s_{0}e^{+} \bigr)|_{J}\neq 0. \end{aligned}$$
(2.12)

Otherwise \((w^{-}+s_{0}e^{+})|_{J}=0\). Hence, using the T-periodicity of \(L(t)\), (2.10) and Lebesgue dominated convergence theorem, we obtain

$$\begin{aligned} \int _{J} \bigl(L(t) \bigl(w_{n}^{-}+s_{n}e^{+} \bigr), \bigl(w_{n}^{-}+s_{n}e^{+} \bigr) \bigr)\,dt&= \int _{J} \bigl(L(t) \bigl(w^{-}+s_{0}e^{+} \bigr), \bigl(w^{-}+s_{0}e^{+} \bigr) \bigr) \,dt+o(1) \\ &=o(1) \end{aligned}$$
(2.13)

and

$$\begin{aligned} \int _{J} \bigl\vert \dot{\widehat{w_{n}^{-}}}+s_{n} \dot{\widehat{e^{+}}} \bigr\vert ^{2}\,dt&= \int _{J} \bigl\vert \dot{\widehat{w_{n}^{-}}}+s_{0} \dot{\widehat{e^{+}}} \bigr\vert ^{2} \,dt+o(1) \\ &= \int _{J} \bigl\vert \dot{\widehat{w_{n}^{-}}}- \dot{\widehat{w^{-}}} \bigr\vert ^{2} \,dt+o(1) \\ &= \int _{J} \bigl( \bigl\vert \dot{\widehat{w_{n}^{-}}} \bigr\vert ^{2}- \bigl\vert \dot{\widehat{w^{-}}} \bigr\vert ^{2} \bigr) \,dt+o(1). \end{aligned}$$
(2.14)

Similar to (2.13), we also obtain

$$\begin{aligned} \int _{J} \bigl(L(t) \bigl(w_{n}^{-}-s_{n}e^{-} \bigr), \bigl(w_{n}^{-}-s_{n}e^{-} \bigr) \bigr)\,dt= \int _{J} \bigl(L(t) \bigl(w^{-}-s_{0}e^{-} \bigr), \bigl(w^{-}-s_{0}e^{-} \bigr) \bigr) \,dt+o(1). \end{aligned}$$
(2.15)

Now, combining (2.13)–(2.15), (2.11), (2.8) and using the fact \(e|_{\mathbb{R}\backslash J}\equiv 0\), we deduce that

$$\begin{aligned} 0\leq{} & 2 \int _{\mathbb{R}} \frac{W(t,v_{n}+\theta _{n}e^{+})}{ \Vert v_{n}+\theta _{n}e^{+} \Vert ^{2}}\,dt \\ ={}&s_{n}^{2} \bigl\Vert e^{+} \bigr\Vert ^{2}- \bigl\Vert w_{n}^{-} \bigr\Vert ^{2}+ \frac{2C_{2}}{ \Vert v_{n}+\theta _{n}e^{+} \Vert ^{2}} \\ ={}& \int _{\mathbb{R}} \bigl[ \bigl\vert \dot{\widehat{w_{n}^{-}}}+s_{n} \dot{\widehat{e^{+}}} \bigr\vert ^{2}+ \bigl(L(t) \bigl(w_{n}^{-}+s_{n}e^{+} \bigr), \bigl(w_{n}^{-}+s_{n}e^{+} \bigr) \bigr) \bigr]\,dt+o(1) \\ ={}& \int _{J} \bigl[ \bigl\vert \dot{\widehat{w_{n}^{-}}}+s_{n} \dot{\widehat{e^{+}}} \bigr\vert ^{2}+ \bigl(L(t) \bigl(w_{n}^{-}+s_{n}e^{+} \bigr), \bigl(w_{n}^{-}+s_{n}e^{+} \bigr) \bigr) \bigr]\,dt \\ &{}+ \int _{\mathbb{R}\backslash J} \bigl[ \bigl\vert \dot{\widehat{w_{n}^{-}}}+s_{n} \dot{\widehat{e^{+}}} \bigr\vert ^{2}+ \bigl(L(t) \bigl(w_{n}^{-}+s_{n}e^{+} \bigr), \bigl(w_{n}^{-}+s_{n}e^{+} \bigr) \bigr) \bigr]\,dt+o(1) \\ ={}& \int _{J} \bigl( \bigl\vert \dot{\widehat{w_{n}^{-}}} \bigr\vert ^{2}- \bigl\vert \dot{\widehat{w^{-}}} \bigr\vert ^{2} \bigr)\,dt+o(1)\\ &{}+ \int _{ \mathbb{R}} \bigl[ \bigl\vert \dot{\widehat{w_{n}^{-}}}-s_{n} \dot{\widehat{e^{-}}} \bigr\vert ^{2}+ \bigl(L(t) \bigl(w_{n}^{-}-s_{n}e^{-} \bigr), \bigl(w_{n}^{-}-s_{n}e^{-} \bigr) \bigr) \bigr]\,dt \\ &{}- \int _{J} \bigl[ \bigl\vert \dot{\widehat{w_{n}^{-}}}-s_{n} \dot{\widehat{e^{-}}} \bigr\vert ^{2}+ \bigl(L(t) \bigl(w_{n}^{-}-s_{n}e^{-} \bigr), \bigl(w_{n}^{-}-s_{n}e^{-} \bigr) \bigr) \bigr]\,dt \\ ={}& \int _{J} \bigl( \bigl\vert \dot{\widehat{w_{n}^{-}}} \bigr\vert ^{2}- \bigl\vert \dot{\widehat{w^{-}}} \bigr\vert ^{2} \bigr)\,dt- \bigl\Vert w_{n}^{-}-s_{n}e^{-} \bigr\Vert ^{2} \\ &{}- \int _{J} \bigl\vert \dot{\widehat{w_{n}^{-}}}-s_{n} \dot{\widehat{e^{-}}} \bigr\vert ^{2}\,dt- \int _{J} \bigl(L(t) \bigl(w^{-}-s_{0}e^{-} \bigr), \bigl(w^{-}-s_{0}e^{-} \bigr) \bigr) \,dt+o(1) \\ ={}&{-}\bigl\Vert w_{n}^{-}-s_{n}e^{-} \bigr\Vert ^{2}- \int _{J} \bigl[ \bigl\vert \dot{\widehat{w^{-}}}-s_{0} \dot{\widehat{e^{-}}} \bigr\vert ^{2}+ \bigl(L(t) \bigl(w^{-}-s_{0}e^{-} \bigr), \bigl(w^{-}-s_{0}e^{-} \bigr) \bigr) \bigr] \,dt+o(1) \\ ={}&{-} \bigl\Vert w_{n}^{-}-s_{n}e^{-} \bigr\Vert ^{2}-s_{0}^{2} \int _{J} \bigl[ \vert \dot{e} \vert ^{2}+ \bigl(L(t)e,e \bigr) \bigr]\,dt+o(1) \\ \leq{} &-s_{0}^{2}+o(1), \end{aligned}$$

which is a contradiction. Consequently, it follows from (2.11), (2.12), (W2)–(W3) and Fatou’s lemma that

$$\begin{aligned} 0&\leq \limsup_{n\rightarrow \infty } \biggl[\frac{s_{n}^{2}}{2} \bigl\Vert e^{+} \bigr\Vert ^{2}-\frac{1}{2} \bigl\Vert w_{n}^{-} \bigr\Vert ^{2}- \int _{\mathbb{R}} \frac{W(t,v_{n}+\theta _{n}e^{+})}{ \Vert v_{n}+\theta _{n}e^{+} \Vert ^{2}}\,dt \biggr] \\ &\leq \frac{s_{0}^{2}}{2} \bigl\Vert e^{+} \bigr\Vert ^{2}-\liminf_{n\rightarrow \infty } \int _{\mathbb{R}} \frac{W(t,v_{n}+\theta _{n}e^{+})}{ \vert v_{n}+\theta _{n}e^{+} \vert ^{2}} \bigl\vert w_{n}^{-}+s_{n}e^{+} \bigr\vert ^{2}\,dt \\ &\leq \frac{s_{0}^{2}}{2} \bigl\Vert e^{+} \bigr\Vert ^{2}- \int _{J} \liminf_{n \rightarrow \infty } \frac{W(t,v_{n}+\theta _{n}e^{+})}{ \vert v_{n}+\theta _{n}e^{+} \vert ^{2}} \bigl\vert w^{-}+s_{0}e^{+} \bigr\vert ^{2}\,dt \\ &=-\infty, \end{aligned}$$

a contradiction. □

Corollary 2.1

Let (L) and (W2)–(W3) be satisfied and \(\rho >0\)be given by Lemma 2.1. Then there exists \(r>\rho \)such that \(\sup \varphi (\partial Q)<0\), where \(Q= \{ v+se^{+}:v\in E^{-}, s\geq 0, \|v+se^{+}\|\leq r \} \).

Combining Lemma 2.1, Corollary 2.1 and Theorem 2.1, we have the following.

Lemma 2.3

Assume that (L) and (W1)–(W3) are satisfied. Then there exist a constant \(c>0\)and a sequence \((u_{n})\subset E\)such that

$$\begin{aligned} \varphi (u_{n})\rightarrow c,\qquad \bigl\Vert \varphi '(u_{n}) \bigr\Vert \bigl(1+ \Vert u_{n} \Vert \bigr)\rightarrow 0\quad \textit{as }n \rightarrow \infty. \end{aligned}$$
(2.16)

Lemma 2.4

Assume that (L) and (W1)–(W4) are satisfied. Then the sequence \((u_{n})\)obtained in Lemma 2.3is bounded.

Proof

Arguing by contradiction, suppose that \(\|u_{n}\|\rightarrow \infty \) as \(n\rightarrow \infty \) and set \(w_{n}=u_{n}/\|u_{n}\|\). Then \(\|w_{n}\|=1\). By (2.16), we have

$$\begin{aligned} &o(1)= \bigl\langle \varphi '(u_{n}),u_{n}^{+} \bigr\rangle = \bigl\Vert u_{n}^{+} \bigr\Vert ^{2}- \int _{\mathbb{R}} \bigl(\nabla W(t,u_{n}),u_{n}^{+} \bigr)\,dt, \\ &o(1)= \bigl\langle \varphi '(u_{n}),u_{n}^{-} \bigr\rangle =- \bigl\Vert u_{n}^{-} \bigr\Vert ^{2}- \int _{\mathbb{R}} \bigl(\nabla W(t,u_{n}),u_{n}^{-} \bigr)\,dt. \end{aligned}$$

Since \(\|w_{n}^{+}\|^{2}+\|w_{n}^{-}\|^{2}=\|w_{n}\|^{2}=1\), one has

$$\begin{aligned} 1+o(1)= \int _{\mathbb{R}} \frac{(\nabla W(t,u_{n}),u_{n}^{+}-u_{n}^{-})}{ \Vert u_{n} \Vert ^{2}}\,dt. \end{aligned}$$
(2.17)

Letting \(\varOmega _{n}= \{ t\in \mathbb{R}:|\nabla W(t,u_{n})|\leq ( \varLambda _{0}-\delta )|u_{n}| \} \), we obtain using the relation \((w^{+},w^{-})=0\), (2.4) and the Hölder inequality

$$\begin{aligned} \int _{\varOmega _{n}} \frac{ \vert \nabla W(t,u_{n}) \vert \vert u_{n}^{+}-u_{n}^{-} \vert }{ \Vert u_{n} \Vert ^{2}}\,dt& \leq \int _{\varOmega _{n}}\frac{ \vert \nabla W(t,u_{n}) \vert }{ \vert u_{n} \vert } \bigl\vert w_{n}^{+}-w_{n}^{-} \bigr\vert \vert w_{n} \vert \,dt \\ &\leq (\varLambda _{0}-\delta ) \Vert w_{n} \Vert _{2}^{2} \\ &\leq (\varLambda _{0}-\delta )\frac{1}{\varLambda _{0}} \Vert w_{n} \Vert ^{2} \\ &\leq 1-\frac{\delta }{\varLambda _{0}}. \end{aligned}$$
(2.18)

Moreover, it follows from (2.16) that

$$\begin{aligned} \int _{\mathbb{R}} \widetilde{W}(t,u_{n})\,dt=\varphi (u_{n})- \frac{1}{2} \bigl\langle \varphi '(u_{n}),u_{n} \bigr\rangle \leq C, \quad\forall n\in \mathbb{N}. \end{aligned}$$

Combining this with (W4) and Hölder’s inequality, we deduce that

$$\begin{aligned} & \biggl\vert \int _{\mathbb{R}\backslash \varOmega _{n}} \frac{(\nabla W(t,u_{n}),u_{n}^{+}-u_{n}^{-})}{ \Vert u_{n} \Vert ^{2}}\,dt \biggr\vert \\ &\quad\leq \frac{1}{ \Vert u_{n} \Vert ^{1-\sigma }} \biggl[ \int _{\mathbb{R} \backslash \varOmega _{n}} \biggl( \frac{ \vert \nabla W(t,u_{n}) \vert }{ \vert u_{n} \vert ^{\sigma }} \biggr)^{k} \,dt \biggr]^{\frac{1}{k}} \bigl\Vert w_{n}^{+}-w_{n}^{-} \bigr\Vert _{\frac{k(1+\sigma )}{k-1}} \Vert w_{n} \Vert ^{\sigma }_{\frac{k(1+\sigma )}{k-1}} \\ &\quad\leq \frac{C}{ \Vert u_{n} \Vert ^{1-\sigma }} \biggl( \int _{\mathbb{R} \backslash \varOmega _{n}}\widetilde{W}(t,u_{n})\,dt \biggr)^{\frac{1}{k}}=o(1). \end{aligned}$$
(2.19)

Hence, by (2.17), (2.18) and (2.19),

$$\begin{aligned} 1+o(1)= \int _{\mathbb{R}} \frac{(\nabla W(t,u_{n}),u_{n}^{+}-u_{n}^{-})}{ \Vert u_{n} \Vert ^{2}}\,dt\leq 1- \frac{\delta }{\varLambda _{0}}+o(1), \end{aligned}$$

a contradiction. □

Proof of Theorem 1.1

According to Lemmas 2.3 and 2.4, there is a bounded \((C)_{c}\) sequence \((u_{n})\) with \(c>0\). Since \((u_{n})\) is bounded, there exists \(M>0\) such that

$$\begin{aligned} \sqrt{\varLambda _{0}} \Vert u_{n} \Vert _{2}\leq \Vert u_{n} \Vert \leq M,\quad \forall n\in \mathbb{N}. \end{aligned}$$
(2.20)

By (2.7), for \(\varepsilon =\frac{c\varLambda _{0}}{2M^{2}}\), there is \(C_{3}>0\) such that

$$\begin{aligned} \bigl\vert \widetilde{W}(t,x) \bigr\vert \leq \frac{c\varLambda _{0}}{2M^{2}} \vert x \vert ^{2}+C_{3} \vert x \vert ^{p},\quad \forall (t,x)\in \mathbb{R}\times \mathbb{R}^{N}. \end{aligned}$$
(2.21)

If \((u_{n})\) is vanishing, that is, for each \(r>0\), \({\lim_{n\rightarrow \infty }\sup_{a\in \mathbb{R}} \int _{a-r}^{a+r}|u_{n}|^{2}\,dt=0}\), then, by [25, Lemma 2.3], \(u_{n}\rightarrow 0\) in \(L^{s}(\mathbb{R},\mathbb{R}^{N})\) for \(2\leq s\leq +\infty \). Hence, using (2.20) and (2.21), we deduce that

$$\begin{aligned} c+o(1)&=\varphi (u_{n})-\frac{1}{2} \bigl\langle \varphi '(u_{n}),u_{n} \bigr\rangle \\ &= \int _{\mathbb{R}} \widetilde{W}(t,u_{n})\,dt \\ &\leq \frac{c\varLambda _{0}}{2M^{2}} \int _{\mathbb{R}} \vert u_{n} \vert ^{2} \,dt+C_{3} \int _{\mathbb{R}} \vert u_{n} \vert ^{p}\,dt \\ &\leq \frac{c}{2}+o(1), \end{aligned}$$

a contradiction.

Hence \((u_{n})\) is nonvanishing, i.e., there are r, \(\sigma >0\) and \((a_{n})\subset \mathbb{Z}\) such that

$$ \liminf_{n\rightarrow \infty } \int _{a_{n}-r}^{a_{n}+r} \vert u_{n} \vert ^{2}\,dt \geq \frac{\sigma }{2}. $$

Setting \(\tilde{u}_{n}(t)=u_{n}(t+a_{n}T)\), we have

$$\begin{aligned} \liminf_{n\rightarrow \infty } \int _{-r}^{r} \vert \tilde{u}_{n} \vert ^{2}\,dt\geq \frac{\sigma }{2}. \end{aligned}$$
(2.22)

Noticing L and W are T-periodic in t, we get \(\|\tilde{u}_{n}\|=\|u_{n}\|\), \(\varphi (\tilde{u}_{n})\rightarrow c\) and \(\|\varphi '(\tilde{u}_{n})\|(1+\|\tilde{u}_{n}\|)\rightarrow 0\) as \(n\rightarrow \infty \). Passing to a subsequence, we assume that \(\tilde{u}_{n}\rightharpoonup \tilde{u}\), \(\varphi '(\tilde{u})=0\) and \(\tilde{u}\neq 0\) by (2.22). This completes the proof. □

References

  1. Ambrosetti, A., Rabinowitz, P.: Dual variational methods in critical point theory and applications. J. Funct. Anal. 14, 349–381 (1973)

    Article  MathSciNet  Google Scholar 

  2. Arioli, G., Szulkin, A.: Homoclinic solutions for a class of systems of second order differential equations. Topol. Methods Nonlinear Anal. 6, 189–197 (1995)

    Article  MathSciNet  Google Scholar 

  3. Chen, G.-W.: Superquadratic or asymptotically quadratic Hamiltonian systems: ground state homoclinic orbits. Ann. Mat. Pura Appl. (4) 194, 903–918 (2015)

    Article  MathSciNet  Google Scholar 

  4. Coti Zelati, V., Ekeland, I., Séré, E.: A variational approach to homoclinic orbits in Hamiltonian systems. Math. Ann. 288(1), 133–160 (1990)

    Article  MathSciNet  Google Scholar 

  5. Coti-Zelati, V., Rabinowitz, P.: Homoclinic orbits for second order Hamiltonian systems possessing superquadratic potentials. J. Am. Math. Soc. 4, 693–727 (1991)

    Article  MathSciNet  Google Scholar 

  6. Ding, Y.: Existence and multiplicity results for homoclinic solutions to a class of Hamiltonian systems. Nonlinear Anal. 25, 1095–1113 (1995)

    Article  MathSciNet  Google Scholar 

  7. Ding, Y., Jeanjean, L.: Homoclinic orbits for a nonperiodic Hamiltonian system. J. Differ. Equ. 237, 473–490 (2007)

    Article  MathSciNet  Google Scholar 

  8. Ding, Y., Lee, C.: Homoclinics for asymptotically quadratic and superquadratic Hamiltonian systems. Nonlinear Anal. 71, 1395–1413 (2009)

    Article  MathSciNet  Google Scholar 

  9. Edmunds, D.E., Evans, W.D.: Spectral Theory and Differential Operators. Clarendon, Oxford (1987)

    MATH  Google Scholar 

  10. Kryszewski, W., Szulkin, A.: Generalized linking theorem with an application to semilinear Schrödinger equation. Adv. Differ. Equ. 3, 441–472 (1998)

    MATH  Google Scholar 

  11. Li, G., Szulkin, A.: An asymptotically periodic Schrödinger equation with indefinite linear part. Commun. Contemp. Math. 4, 763–776 (2002)

    Article  MathSciNet  Google Scholar 

  12. Liu, Z., Guo, S., Zhang, Z.: Homoclinic orbits for the second-order Hamiltonian systems. Nonlinear Anal., Real World Appl. 36, 116–138 (2017)

    Article  MathSciNet  Google Scholar 

  13. Papageorgiou, N.S., Rǎdulescu, V.D., Repovš, D.D.: Nonlinear Analysis: Theory and Methods. Springer Monographs in Mathematics. Springer, Cham (2019)

    Book  Google Scholar 

  14. Rabinowitz, P.H.: Homoclinic orbits for a class of Hamiltonian systems. Proc. R. Soc. Edinb., Sect. A 114, 33–38 (1990)

    Article  MathSciNet  Google Scholar 

  15. Rabinowitz, P.H., Tanaka, K.: Some results on connecting orbits for a class of Hamiltonian systems. Math. Z. 206, 473–499 (1991)

    Article  MathSciNet  Google Scholar 

  16. Schechter, M., Zou, W.: Weak linking theorems and Schrödinger equations with critical Sobolev exponent. ESAIM Control Optim. Calc. Var. 9, 601–619 (2003)

    Article  MathSciNet  Google Scholar 

  17. Séré, E.: Existence of infinitely many homoclinic orbits in Hamiltonian systems. Math. Z. 209(1), 27–42 (1992)

    Article  MathSciNet  Google Scholar 

  18. Séré, E.: Looking for the Bernoulli shift. Ann. Inst. Henri Poincaré, Anal. Non Linéaire 10, 561–590 (1993)

    Article  MathSciNet  Google Scholar 

  19. Szulkin, A., Zou, W.: Homoclinic orbits for asymptotically linear Hamiltonian systems. J. Funct. Anal. 187, 25–41 (2001)

    Article  MathSciNet  Google Scholar 

  20. Tang, X.: Infinitely many homoclinic solutions for a second-order Hamiltonian system. Math. Nachr. 289, 116–127 (2016)

    Article  MathSciNet  Google Scholar 

  21. Tang, X.H., Lin, X.Y., Yu, J.S.: Nontrivial solutions for Schrödinger equation with local super-quadratic conditions. J. Dyn. Differ. Equ. 31, 369–383 (2019)

    Article  Google Scholar 

  22. Wang, X.: Local super-quadratic conditions on homoclinic solutions for a second-order Hamiltonian system. Appl. Math. Lett. 75, 7–12 (2018)

    Article  MathSciNet  Google Scholar 

  23. Wu, D.-L., Yu, X.: New homoclinic orbits for Hamiltonian systems with asymptotically quadratic growth at infinity. Qual. Theory Dyn. Syst. 19(1), 22 (2020)

    Article  MathSciNet  Google Scholar 

  24. Yang, M.-H., Han, Z.-Q.: The existence of homoclinic solutions for second-order Hamiltonian systems with periodic potentials. Nonlinear Anal., Real World Appl. 12(5), 2742–2751 (2011)

    Article  MathSciNet  Google Scholar 

  25. Ye, Y.: Homoclinic solutions for second-order Hamiltonian systems with periodic potential. Bound. Value Probl. 2018, 186 (2018)

    Article  MathSciNet  Google Scholar 

  26. Ye, Y., Tang, C.-L.: Multiple homoclinic solutions for second-order perturbed Hamiltonian systems. Stud. Appl. Math. 132, 112–137 (2014)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The author would like to thank the handling editors and the anonymous reviewers.

Availability of data and materials

Not applicable.

Funding

This work is supported by the National Natural Science Foundation of China (No. 11971393) and the Science and Technology Research Program of Chongqing Municipal Education Commission (No. KJQN201900501).

Author information

Authors and Affiliations

Authors

Contributions

The author conceived of the study, drafted the manuscript, and approved the final manuscript.

Corresponding author

Correspondence to Yiwei Ye.

Ethics declarations

Competing interests

The author declares that there are no competing interests.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ye, Y. On locally superquadratic Hamiltonian systems with periodic potential. Bound Value Probl 2020, 146 (2020). https://doi.org/10.1186/s13661-020-01444-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13661-020-01444-y

MSC

Keywords