Hindawi Publishing Corporation Abstract and Applied Analysis Volume 2013, Article ID 610906, 11 pages http://dx.doi.org/10.1155/2013/610906 Research Article Homoclinic Orbits for a Class of Nonperiodic Hamiltonian Systems with Some Twisted Conditions Qi Wang1 and Qingye Zhang2 1 2 Institute of Contemporary Mathematics, School of Mathematics and Information Science, Henan University, Kaifeng 475000, China Department of Mathematics, Jiangxi Normal University, Nanchang 330022, China Correspondence should be addressed to Qi Wang; mathwq@henu.edu.cn Received 4 January 2013; Accepted 1 April 2013 Academic Editor: Changbum Chun Copyright © 2013 Q. Wang and Q. Zhang. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited. By the Maslov index theory, we will study the existence and multiplicity of homoclinic orbits for a class of asymptotically linear nonperiodic Hamiltonian systems with some twisted conditions on the Hamiltonian functions. 1. Introduction and Main Results Consider the following first-order nonautonomous Hamiltonian systems π§Μ = π½π»π§ (π‘, π§) , (HS) where π§ : R → R2π, π½ = ( πΌ0π −πΌ0π ), π» ∈ πΆ1 (R × R2π, R) and ∇π§ π»(π‘, π§) denotes the gradient of π»(π‘, π§) with respect to π§. As usual we say that a nonzero solution π§(π‘) of (HS) is homoclinic (to 0) if π§(π‘) → 0 as |π‘| → ∞. As a special case of dynamical systems, Hamiltonian systems are very important in the study of gas dynamics, fluid mechanics, relativistic mechanics and nuclear physics. However it is well known that homoclinic solutions play an important role in analyzing the chaos of Hamiltonian systems. If a system has the transversely intersected homoclinic solutions, then it must be chaotic. If it has the smoothly connected homoclinic solutions, then it cannot stand the perturbation, and its perturbed system probably produces chaotic phenomena. Therefore, it is of practical importance and mathematical significance to consider the existence of homoclinic solutions of Hamiltonian systems emanating from 0. In the last years, the existence and multiplicity of homoclinic orbits for the first-order system (HS) were studied extensively by means of critical point theory, and many results were obtained under the assumption that π»(π‘, π§) depends periodically on π‘ (see, e.g., [1–12]). Without assumptions of periodicity, the problem is quite different in nature and there is not much work done so far. To the best of our knowledge, the authors in [13] firstly obtained the existence of homoclinic orbits for a class of first-order systems without any periodicity on the Hamiltonian function. After this, there were a few papers dealing with the existence and multiplicity of homoclinic orbits for the first-order system (HS) in this situation (see, e.g., [14–17]). In the present paper, with the Maslov index theory of homoclinic orbits introduced by Chen and Hu in [18], we will study the existence and multiplicity of homoclinic orbits for (HS) without any periodicity on the Hamiltonian function. To the best of the author’s knowledge, the Maslov index theory of homoclinic orbits is the first time to be used to study the existence of homoclinic solutions. We are mainly interested in the Hamiltonian functions of the form π» (π‘, π§) = −πΏ (π‘) π§ ⋅ π§ + π (π‘, π§) , (1) where πΏ is a 2π × 2π symmetric matrix valued function. We 2 assume that (πΏ 1 )πΏ ∈ πΆ(R, Rπ ), and there are πΌ, π > 0, π‘0 ≥ 0 and a constant matrix π, satisfying ππΏ (π‘) − π|π‘|πΌ πΌ2π ≥ 0, ∀ |π‘| ≥ π‘0 , (2) where πΌ2π is the identity map on R2π, and for a 2π × 2π matrix π, we say that π ≥ 0 if and only if inf π∈R2π ,|π|=1 ππ ⋅ π ≥ 0. (3) 2 Abstract and Applied Analysis In (πΏ 1 ), if π = ( πΌ0π πΌ0π ), then (πΏ 1 ) is similar to the condition (π 0 ) in [14]. But the restrictions on π (π‘, π§) will be different from [14], and we will give some examples in Remark 5. If π = πΌ 0 ±πΌ2π or π = ( π+π 0 −πΌπ−π ) in condition (πΏ 1 ), for example, it is quite different from the existing results as authors known. In short, condition (πΏ 1 ) means that the eigenvalues of πΏ(π‘) will tend to ±∞ with the speed no less than |π‘|πΌ . But (πΏ 1 ) does not contain all of these cases. For examples, let π = 1 and πΏ(π‘) = πΌ 2π‘ sin 2π‘ |π‘|πΌ ( cos sin 2π‘ − cos 2π‘ ), we have the eigenvalues of πΏ(π‘) are ±|π‘| , πΌ but there is no constant matrix π satisfying ππΏ(π‘) − π|π‘| πΌ2π ≥ 0, for all |π‘| ≥ π‘0 . Denote by πΉΜ the self-adjoint operator −J(π/ππ‘) + πΏ(π‘) on Μ = π»1 (R, R2π) if πΏ(π‘) is πΏ2 ≡ πΏ2 (R, R2π), with domain π·(πΉ) 1 2π Μ ⊂ π» (R, R ) if πΏ(π‘) is unbounded. Let bounded and π·(πΉ) Μ be the absolute value of πΉ, Μ and let |πΉ| Μ 1/2 be the square root |πΉ| Μ π·(πΉ) Μ is a Hilbert space equipped with the norm of |πΉ|. σ΅¨ σ΅¨ σ΅© σ΅© βπ§βπΉΜ = σ΅©σ΅©σ΅©σ΅©(πΌ + σ΅¨σ΅¨σ΅¨σ΅¨πΉΜσ΅¨σ΅¨σ΅¨σ΅¨) π§σ΅©σ΅©σ΅©σ΅©πΏ2 , Μ . ∀π§ ∈ π· (πΉ) (4) Μ 1/2 ), and define on πΈ the inner product and Let πΈ = π·(|πΉ| norm by Μ 1/2 π’, |πΉ| Μ 1/2 V) + (π’, V)2 , (π’, V)πΈ = (|πΉ| 2 βπ’βπΈ = (π’, π’)1/2 πΈ , (5) where (⋅, ⋅)πΏ2 denotes the usual inner product on πΏ2 (R, R2π). Then, πΈ is a Hilbert space. It is easy to see that πΈ is continuously embedded in π»1/2 (R, R2π), and we further have the following lemma. π§(π‘) σ³¨→ π΅(π‘)π§(π‘), for any π§ ∈ πΏ2 , we still denote this operator by π΅ and then πΎπ΅ : πΈ ⊂ πΏ2 → πΈ is a self-adjoint compact operator on πΈ and satisfies (πΎπ΅π’, V)πΈ = (π΅π’, V)πΏ2 , ∀π’, V ∈ πΈ. (9) Before presenting the conditions on π (π‘, π§), we need the concept of Maslov index for homoclinic orbits introduced by Chen and Hu in [18] which is equivalent to the relative Morse index. We will give a brief introduction of it by Definition 7, where for any π΅ ∈ B, we denote the associated index pair by (ππΉ (πΎπ΅), ππΉ (πΎπ΅)). Now we can present the conditions on π (π‘, π§) as follows. For notational simplicity, we set π΅0 (π‘) = ∇π§2 π (π‘, 0), and in what follows, the letter π will be repeatedly used to denote various positive constants whose exact value is irrelevant. Besides, for two 2π × 2π symmetric matrices π1 and π2 , π1 ≤ π2 means that π2 − π1 is semipositive definite. (π 1 ) π ∈ πΆ2 (R × R2π, R), and there exists a constant π > 0 such that σ΅¨σ΅¨ 2 σ΅¨ σ΅¨σ΅¨∇π§ π (π‘, π§)σ΅¨σ΅¨σ΅¨ ≤ π, ∀ (π‘, π§) ∈ R × R2π. (10) σ΅¨ σ΅¨ (π 0 ) ∇π§ π (π‘, 0) ≡ 0 and π΅0 ∈ B. (π ∞ ) There exists some π 0 > 0 and continuous symmetric matrix functions π΅1 , π΅2 ∈ B with ππΉ (πΎπ΅1 ) = ππΉ (πΎπ΅2 ) and ππΉ (πΎπ΅2 ) = 0 such that π΅1 (π‘) ≤ ∇π§2 π (π‘, π§) ≤ π΅2 (π‘) , ∀π‘ ∈ R, |π§| > π 0 . (11) Then, we have our first result. Lemma 1. Suppose that πΏ satisfies (πΏ 1 ). Then πΈ is compactly embedded in πΏπ (R, R2π) with the usual norm β ⋅ βπΏπ for any 1 ≤ π ∈ (2/(1 + πΌ), ∞). Theorem 2. Assume that (πΏ 1 ), (π 1 ), (π 0 ), and (π ∞ ) hold. If This lemma is similar to Lemmas 2.1–2.3 in [13], and we will prove it in Section 3. Define the quadratic form Q on πΈ by then (HS) has at least one nontrivial homoclinic orbit. Moreover, if ππΉ (πΎπ΅0 ) = 0 and |ππΉ (πΎπ΅1 ) − ππΉ (πΎπ΅0 )| ≥ π, the problem possesses at least two nontrivial homoclinic orbits. Q (π’, V) = ∫ ((−π½π’,Μ V) + (πΏ (π‘) π’, V)) ππ‘, R ∀π’, V ∈ πΈ. (6) It is easy to check that Q(π’, V) is a bounded quadratic form on πΈ, and, hence, there exists a unique bounded self-adjoint operator πΉ : πΈ → πΈ such that (πΉπ’, V)πΈ = Q (π’, V) , ∀π’, V ∈ πΈ. (7) (πΎπ’, V)πΈ = (π’, V)πΏ2 , ∀π’ ∈ πΏ , V ∈ πΈ. (12) Condition (π ∞ ) is a two-side pinching condition near the ± ) as follows. infinity, and we can relax (π ∞ ) to condition (π ∞ ± ) There exist some π 0 > 0 and a continuous symmetric (π ∞ matrix function π΅∞ ∈ B with ππΉ (πΎπ΅∞ ) = 0 such that ±∇π§2 π (π‘, π§) ≥ ±π΅∞ (π‘) , ∀π‘ ∈ R, |π§| > π 0 . (13) Then, we have the following results. Besides, define a linear operator πΎ : πΏ2 → πΈ by 2 ππΉ (πΎπ΅1 ) ∉ [ππΉ (πΎπ΅0 ) , ππΉ (πΎπ΅0 ) + ππΉ (πΎπ΅0 )] , (8) In view of Lemma 1, we know that πΉ is a Fredholm operator and πΎ is a compact operator. Denote by B the set of all uniformly bounded symmetric 2π × 2π matric functions. That is to say, π΅ ∈ B if and only if π΅π (π‘) = π΅(π‘) for all π‘ ∈ R, and π΅(π‘) is uniformly bounded in π‘ as the operator on R2π. For any π΅ ∈ B, it is easy to see that π΅ determines a bounded self-adjoint operator on πΏ2 , by + − Theorem 3. Assumes (πΏ 1 ), (π 1 ), (π 0 ), (π ∞ ) (or (π ∞ )), and ππΉ (πΎπ΅0 ) = 0 hold. If ππΉ (πΎπ΅∞ ) ≥ ππΉ (πΎπ΅0 ) + 2 (or ππΉ (πΎπ΅∞ ) ≤ ππΉ (πΎπ΅0 ) − 2), then (HS) has at least one nontrivial homoclinic orbit. + − ) (or (π ∞ )), Theorem 4. Suppose that (πΏ 1 ), (π 1 ), (π 0 ), (π ∞ and ππΉ (πΎπ΅0 ) = 0 are satisfied. If in addition π is even in π§ and ππΉ (πΎπ΅∞ ) ≥ ππΉ (πΎπ΅0 ) + 2 (or ππΉ (πΎπ΅∞ ) ≤ ππΉ (πΎπ΅0 ) − 2), then (HS) has at least |ππΉ (πΎπ΅∞ ) − ππΉ (πΎπ΅0 )| − 1 pairs of nontrivial homoclinic orbits. Abstract and Applied Analysis 3 Remark 5. Lemma 1 shows that π(π΄), the spectrum of π΄, consists of eigenvalues numbered by (counted in their multiplicities): ⋅ ⋅ ⋅ ≤ π −2 ≤ π −1 ≤ 0 < π 1 ≤ π 2 ≤ ⋅ ⋅ ⋅ (14) with π ±π → ±∞ as π → ∞. Let π΅0 (π‘) ≡ π΅0 and π΅∞ (π‘) ≡ π΅∞ , with the constants π΅0 , π΅∞ satisfying π π < π΅0 < π π+1 , and π π+π < π΅∞ < π π+π+1 for some π ∈ Z and π ≥ 1 (or π ≤ −1). Define 1 1 π (π‘, π§) = πΏ (|π§|) π΅0 |π§|2 + (1 − πΏ (|π§|)) π΅∞ |π§|2 , 2 2 Definition 7. For any bounded self-adjoint Fredholm operator F and a compact self-adjoint operator T on H, the relative Morse index pair (πF (T), πF (T)) is defined by πF (T) = ind (πF |H− (F−T) ) , (15) where πΏ is a smooth cutoff function satisfying πΏ(|π§|) = |π§|<1, { 1, 0, |π§|>2. By Proposition 12 that, it is easy to verify π satisfies all the conditions in Theorem 2. Furthermore, let the constant π΅∞ satisfy π π+π < π΅∞ < π π+π+1 for some π ∈ Z and π ≥ 2 (or π ≤ −2). Define 1 1 π (π‘, π§) = πΏ (|π§|) π΅0 |π§|2 + (1 − πΏ (|π§|)) π΅∞ |π§|2 . 2 2 (see [19, Lemma 2.7]), where πF : H → H− (F) and πF−T : H → H− (F − T) are the respective projections. Then, by Fredholm operator theory, πF |H− (F−T) : H− (F − T) → H− (F) is a Fredholm operator. Here and in the sequel, we denote by ind(⋅) the Fredholm index of a Fredholm operator. πF (T) = dim H0 (F − T) . (18) 2.2. Saddle Point Reduction. In this subsection, we describe the saddle point reduction in [20–22]. Recall that H is a real Hilbert space, and π΄ is a self-adjoint operator with domain π·(π΄) ⊂ H. Let Φ ∈ πΆ1 (H, R), with ΦσΈ (π) = 0. Assume that (1) there exist real numbers πΌ < π½ such that πΌ, π½ ∉ π(π΄), and that π(π΄)∩[πΌ, π½] consists of at most finitely many eigenvalues of finite multiplicities; (16) (2) ΦσΈ is Gateaux differentiable in H, which satisfies Then π satisfies all the conditions in Theorems 3 and 4. However, it is easy to see that some conditions of the main results in [13–15, 17] do not hold for these examples. σ΅©σ΅© πΌ + π½ σ΅©σ΅©σ΅©σ΅© π½ − πΌ σ΅©σ΅© σΈ πΌσ΅© ≤ , σ΅©σ΅©πΦ (π’) − σ΅©σ΅© 2 σ΅©σ΅©σ΅© 2 ± ) is Remark 6. Note that the assumption ππΉ (πΎπ΅∞ ) = 0 in (π ∞ + ) with not essential for our main results. For the case of (π ∞ ππΉ (πΎπ΅∞ ) =ΜΈ 0, let π΅Μ∞ = π΅∞ − ππΌ2π with π > 0 small enough, where πΌ2π is the identity map on R2π; then, ππΉ (πΎπ΅Μ∞ ) = + ππΉ (πΎπ΅∞ ) and ππΉ (πΎπ΅Μ∞ ) = 0, and, hence, (π ∞ ) holds for π΅Μ∞ . Therefore, Theorems 3 and 4 still hold in this case. While for − ) with ππΉ (πΎπ΅∞ ) =ΜΈ 0, if we replace ππΉ (πΎπ΅∞ ) by the case of (π ∞ ππΉ (πΎπ΅∞ )+ππΉ (πΎπ΅∞ ) in Theorems 3 and 4, then similar results hold. Indeed, let π΅Μ∞ = π΅∞ + ππΌ2π with π > 0 small enough such that ππΉ (πΎπ΅Μ∞ ) = ππΉ (πΎπ΅∞ ) + ππΉ (πΎπ΅∞ ) and ππΉ (πΎπ΅Μ∞ ) = − 0, then this case is also reduced to the case of (π∞ ) for π΅Μ∞ with ππΉ (πΎπ΅Μ∞ ) = 0. ∀π’ ∈ H, (19) and without loss of generality, we may assume that πΌ = −π½, π½ > 0; (3) Φ ∈ πΆ2 (π, R), π = π·(|π΄|1/2 ), with the norm 1/2 σ΅© σ΅©2 βπ§βπ = (σ΅©σ΅©σ΅©σ΅©|π΄|1/2 π§σ΅©σ΅©σ΅©σ΅©H + π2 βπ§β2H ) , (20) where π > 0 small and −π ∉ π(π΄). Consider the solutions of the following equation: π΄π§ = ΦσΈ (π§) , π§ ∈ π· (π΄) . (21) Let 2. Preliminaries π½ In this section, we recall the definition of relative Morse index and saddle point reduction and give the relationship between them. For this propose, the notion of spectral flow will be used. 2.1. Relative Morse Index. Let H be a separable Hilbert space; for any self-adjoint operator π΄ on H, there is a unique π΄invariant orthogonal splitting H = H+ (π΄) ⊕ H− (π΄) ⊕ H0 (π΄) , π0 = ∫ ππΈπ , π+ = ∫ −π½ +∞ π½ π− = ∫ −π½ −∞ ππΈπ , (22) ππΈπ , where {πΈπ } is the spectral resolution of π΄, and let H∗ = π∗ H, ∗ = 0, ±. (23) Decompose the space π as follows: (17) where H0 (π΄) is the null space of π΄, π΄ is positive definite on H+ (π΄) and negative definite on H− (π΄), and ππ΄ denotes the orthogonal projection from H to H− (π΄). For any bounded self-adjoint Fredholm operator F and a compact self-adjoint operator T on H, πF − πF−T is compact π = π0 ⊕ π− ⊕ π+ , (24) where π∗ = |π΄ π |−1/2 H∗ , ∗ = 0, ±, and π΄ π = π΄ + ππΌ. For each π’ ∈ H, we have the decomposition π’ = π’+ + π’0 + π’− , (25) 4 Abstract and Applied Analysis where π’∗ ∈ H∗ and ∗ = 0, ±; let π§ = π§+ + π§0 + π§− , with σ΅¨ σ΅¨−1/2 π§∗ = σ΅¨σ΅¨σ΅¨π΄ π σ΅¨σ΅¨σ΅¨ π’∗ , ∗ = 0, ±. (26) Define a functional π on H as follows: π (π’) = 1 σ΅©σ΅© σ΅©σ΅©2 σ΅© σ΅©2 σ΅© σ΅©2 (σ΅©π’ σ΅© + π+ σ΅©σ΅©σ΅©π’0 σ΅©σ΅©σ΅© − π− σ΅©σ΅©σ΅©π’0 σ΅©σ΅©σ΅© − βπ’β2 ) 2 σ΅© +σ΅© (27) − Φπ (π§) , ∞ 0 where π+ = ∫0 ππΈπ , π− = ∫−∞ ππΈπ , and Φπ (π§) = (π/2) βπ§βH + Φ(π§). The Euler equation of this functional is the system σ΅¨ σ΅¨−1/2 π’± = ±σ΅¨σ΅¨σ΅¨π΄ π σ΅¨σ΅¨σ΅¨ π± ΦσΈ π (π§) , σ΅¨ σ΅¨−1/2 π± π’0 = ±σ΅¨σ΅¨σ΅¨π΄ π σ΅¨σ΅¨σ΅¨ π± π0 ΦσΈ π (π§) . (28) (29) Thus, π§ = π§+ + π§0 + π§− is a solution of (21) if and only if π’ = π’+ + π’0 + π’− is a critical point of π. The implicit function can be applied, yielding a solution 𧱠(π§0 ) for fixed π§0 ∈ π0 , such that 𧱠∈ πΆ1 (π0 , π± ). Since dim π0 is finite, all topologies on π0 are equivalent, and we choose β ⋅ βH as it norm. We have σ΅¨ σ΅¨1/2 π’± (π§0 ) = σ΅¨σ΅¨σ΅¨π΄ π σ΅¨σ΅¨σ΅¨ 𧱠(π§0 ) ∈ πΆ1 (H0 , H) , 1 (π΄ (π§ (π₯) , π§ (π₯))) − Φ (π§ (π₯)) , 2 (32) (33) between the critical points of the πΆ2 -function π ∈ πΆ2 (H0 , R) with the solutions of the operator equation π§ ∈ π· (π΄) . (34) Moreover, the functional π satisfies πσΈ σΈ (π₯) = [π΄ − ΦσΈ σΈ (π§ (π₯))] π§σΈ (π₯) = π΄π0 − π0 ΦσΈ σΈ (π§ (π₯)) π§σΈ (π₯) . Proof. Note that ≤ 1 σ΅©σ΅© σΈ + σ΅© σ΅©σ΅©Φ (π§ + π§− + π₯) + π(π§+ + π§− + π₯)σ΅©σ΅©σ΅© 2 σ΅©πΏ σ΅© √π½ − π ≤ πΆπ + π σ΅©σ΅© + σ΅© − σ΅©π§ + π§ + π₯σ΅©σ΅©σ΅©πΏ2 √π½ − π σ΅© ≤ σ΅© + σ΅© σ΅© − σ΅© πΆπ + π σ΅©σ΅©σ΅©π’ (π₯)σ΅©σ΅©σ΅©πΏ2 σ΅©σ΅©σ΅©π’ (π₯)σ΅©σ΅©σ΅©πΏ2 ( + + βπ₯βπΏ2 ) . √π½ − π √π½ √π½ (38) (39) Therefore, 2√π½ (πΆπ + π) σ΅© − σ΅© σ΅©σ΅© + σ΅©σ΅© βπ₯βπΏ2 . σ΅©σ΅©π’ (π₯)σ΅©σ΅©πΏ2 + σ΅©σ΅©σ΅©π’ (π₯)σ΅©σ΅©σ΅©πΏ2 ≤ π½ − 2πΆπ − 3π (40) Next, since σΈ σ΅¨ σ΅¨−1/2 (π’± ) (π₯) = ±σ΅¨σ΅¨σ΅¨π΄ π σ΅¨σ΅¨σ΅¨ π± ΦσΈ π σΈ πσΈ (π₯) = π΄ (π§ (π₯)) − ΦσΈ (π§ (π₯)) = π΄π₯ − π0 Φ (π§ (π₯)) , (37) σΈ × (π§+ + π§− + π₯) ((π§+ ) (π₯) + (π§− ) (π₯) + πΌ) , (41) 1 π (π₯) = (π΄ (π§ (π₯) , π§ (π₯))) − Φ (π§ (π₯)) , 2 σΈ π½ σ³¨→ ∞. (36) (31) Theorem 8. Under assumptions (1), (2), and (3), there is a oneone correspondence π΄π§ = ΦσΈ (π§) , Moreover, one has σ΅©σ΅© ± σΈ σ΅©σ΅© σ΅©σ΅©(π’ ) (π₯)σ΅©σ΅© σ³¨→ 0, σ΅©σ΅© σ΅©σ΅©πΏ2 ∀π₯ ∈ H0 . From ∇π§ π (π‘, 0) = 0, |∇π§2 π | ≤ πΆπ , we have ΦσΈ (0) = 0 and βΦσΈ (π§)βπΏ2 ≤ πΆπ βπ§βπΏ2 . Since β|π΄ π |−1/2 π± β ≤ 1/√(π½ − π), we have σ΅©σ΅© ± σ΅©σ΅© σ΅©σ΅©π’ (π₯)σ΅©σ΅©πΏ2 where π§(π₯) = π(π₯)+π₯, π(π₯) = π§+ (π₯)+π§− (π₯) ∈ π·(π΄). Then, we have the following theorem due to Amann [20], Chang [21], and Long [22]. π₯ σ³¨σ³¨→ π§ = π§ (π₯) = π§+ (π₯) + π§− (π₯) + π₯ 2√π½ (πΆπ + 1) σ΅©σ΅© ± σ΅©σ΅© βπ₯βπΏ2 , σ΅©σ΅©π’ (π₯)σ΅©σ΅©πΏ2 ≤ π½ − 2πΆπ − 3π σ΅¨ σ΅¨−1/2 π’± (π₯) = ±σ΅¨σ΅¨σ΅¨π΄ π σ΅¨σ΅¨σ΅¨ π± ΦσΈ π (π§+ + π§− + π₯) . where π’0 (π§0 ) = |π΄ π |1/2 π§0 , and let π§0 = π₯ we have π (π₯) = Lemma 9. Assume that π ∈ πΆ2 (R × R2π, R), |∇π§2 π | ≤ πΆπ , for all (π‘, π§) ∈ R × R2π and ∇π§ π (π‘, 0) ≡ 0; then one has (30) which solves the system (28). Let π (π§0 ) = π (π’+ (π§0 ) + π’− (π§0 ) + π’0 (π§0 )) , Since H0 is a finite dimensional space, for every critical point π₯ of π in H0 , the Morse index and nullity are finite, and we denote them by (ππ− (π₯), ππ0 (π₯)). Now, let the Hilbert space H be πΏ2 (R, R2π), and the oper∞ ator π΄ be πΉΜ = −π½(π/ππ‘)+πΏ, Φ(π’) = ∫−∞ π (π‘, π’). Then we have π = πΈ. For π ∈ πΆ2 (R×R2π, R) and |∇π§2 π | ≤ πΆπ , for all (π‘, π§) ∈ Μ we have π΄ and R × R2π, let −πΌ = π½ ≥ 2(πΆπ + 1) and π½ ∉ π(πΉ); Φ satisfying the previous conditions. Thus, from Theorem 8, we can solve our problems on the finite dimensional space. Similar to Lemma 2.2 and Remark 2.3 in [23], we have the following estimates. (35) where πΌ is the identity map on H0 , we have σ΅© σ΅©σ΅© + σΈ σ΅©σ΅© σ΅© σ΅©σ΅©(π’ ) (π₯)σ΅©σ΅© + σ΅©σ΅©σ΅©(π’− )σΈ (π₯)σ΅©σ΅©σ΅© σ΅©σ΅©πΏ2 σ΅©σ΅© σ΅©σ΅©πΏ2 σ΅©σ΅© ≤ 2√π½ (πΆπ + π) , π½ − 2πΆπ − 3π (42) ∀π₯ ∈ H0 . Abstract and Applied Analysis 5 Remark 10. For π§(π₯), we also have that there is a constant πΆ > 0 dependent of πΆπ , but independent of π½, such that πΆ σ΅©σ΅© ± σ΅©σ΅© βπ₯βπΏ2 , σ΅©σ΅©π§ (π₯)σ΅©σ΅©π ≤ √π½ πΆ σ΅©σ΅© σΈ ± σ΅©σ΅© σ΅©σ΅©π§ (π₯)σ΅©σ΅© ≤ σ΅©π √π½ , σ΅© ∀π₯ ∈ H0 . (43) If π satisfies the condition (π 1 ), then for any homoclinic orbit π§ of (HS), ∇π§2 π (⋅, π§) ∈ B, and, hence, we have the associated index pair (ππΉ (πΎπ΅), ππΉ (πΎπ΅)). For notation simplicity, in what follows, we set ππΉ (π§) = ππΉ (πΎ∇π§2 (π (π‘, π§))) , (44) ππΉ (π§) = ππΉ (πΎ∇π§2 (π (π‘, π§))) . Theorem 11. Let π ∈ πΆ2 (R × R2π, R) satisfying |∇π§2 π | ≤ πΆπ , for all (π‘, π§) ∈ R × R2π and ∇π§ π (π‘, 0) ≡ 0. For each critical point π₯ of π in H0 , π§(π₯) is a homoclinic orbit of (HS) and one has ππ− (π₯) = dim (πΈ− (H0 )) + ππΉ (πΎ∇π§2 π (π‘, π§ (π₯))) = dim (πΈ (H0 )) + ππΉ (π§ (π₯)) , (π₯) = ππΉ (πΎ∇π§2 π (π‘, π§ (π₯)) (45) = ππΉ (π§ (π₯)) , 0 where dim(πΈ− (H0 )) is the dimension of the space ∫−π½ ππΈπ (H0 ). This theorem shows the relations between the relative Morse index and the Morse index of the saddle point reduction, and it will play an important role in the proof of our main results. The proof of this theorem will be postponed to the next subsection where the notion of spectral flow will be used. 2.3. The Relationship between ππΉ (π), Spectral Flow, and the Morse Index of Saddle Point Reduction. It is well known that the concept of spectral flow was first introduced by Atiyah et al. in [24] and then extensively studied in [19, 25–28]. Here, we give a brief introduction of the spectral flow as introduced in [18]. Let H be a separable Hilbert space as defined before, and {Fπ | π ∈ [0, 1]} be a continuous path of self-adjoint Fredholm operators on the Hilbert space H. The spectral flow of Fπ represents the net change in the number of negative eigenvalues of Fπ as π runs from 0 to 1, where the counting follows from the rule that each negative eigenvalue crossing to the positive axis contributes +1 and each positive eigenvalue crossing to the negative axis contributes −1, and for each crossing, the multiplicity of eigenvalue is taken into account. In the calculation of spectral flow, a crossing operator introduced in [28] will be used. Take a πΆ1 path {Fπ π |∈ [0, 1]} and let Pπ be the projection from H to H0 (Fπ ). When eigenvalue crossing occurs at Fπ , the operator π F P : H0 (Fπ ) σ³¨→ H0 (Fπ ) ππ π π A crossing occurring at Fπ is called simple crossing if dim H0 (Fπ ) = 1. As indicated in [19], the spectral flow Sf(Fπ ) will remain the same after a small disturbance of Fπ , that is, Sf(Fπ ) = Sf(Fπ + πid) for π > 0 and small enough, where id is the identity map on H. Furthermore, we can choose suitable π such that all the eigenvalue crossings occurring in Fπ , 0 ≤ π ≤ 1 are regular [28]. Thus, without loss of generality, we may assume that all the crossings are regular. Let D be the set containing all the points in [0, 1] at which the crossing occurs. The set D contains only finitely many points. The spectral flow of Fπ is = ∑ sign πΆπ [Fπ ] − dim H− (πΆπ [F0 ]) π∈D∗ − Pπ sign πΆπ [Fπ ] = dim H+ (πΆπ [Fπ ]) − dim H− (πΆπ [Fπ ]) . (47) Sf (Fπ , 0 ≤ π ≤ 1) − ππ0 is called a crossing operator, denoted by πΆπ [Fπ ]. As mentioned in [28], an eigenvalue crossing at Fπ is said to be regular if the null space of πΆπ [Fπ ] is trivial. In this case, we define (46) (48) + dim H+ (πΆπ [F1 ]) , where D∗ = D ∩ (0, 1). In what follows, the spectral flow of Fπ will be simply denoted by Sf(Fπ ) when the starting and end points of the flow are clear from the contents. And πFπ will be simply denoted by ππ . Proposition 12 (see [18, Proposition 3]). Suppose that, for each π ∈ [0, 1], Fπ − F0 is a compact operator on H then ind (π0 |H− (F1 ) ) = −Sf (Fπ ) . (49) Thus, from Definition 7, πF0 (T) = −Sf (Fπ , 0 ≤ π ≤ 1) , (50) where Fπ = F − πT, T is a compact operator. Moreover, if π(T) ⊂ [0, ∞) and 0 ∉ ππ (T), from the definition of spectral flow, we have πF0 (T) = −Sf (Fπ , 0 ≤ π ≤ 1) = ∑ πF (πT) π∈[0,1) (51) = ∑ dim H0 (F − πT) . π∈[0,1) The proof of Theorem 11 is the direct consequence of the aformentioned Proposition 12 and Theorem 3.2 in [19], so we omit it here. − can be transformed into the case Remark 13. The case of π ∞ + − + case follows from the π ∞ of π ∞ . More concretely, the π ∞ Μ π§) = −π (−π‘, π§). If π§(π‘) case by applying to the function π (π‘, 6 Abstract and Applied Analysis Μ is a homoclinic solution of πΉπ§(π‘) = ∇π§ π (π‘, π§(π‘)), let π§Μ(π‘) = π§(−π‘), it is easy to check that π§Μ(π‘) is a homoclinic solution of Μ π§Μ(π‘)), and this is a one-one correspondence Μπ§(π‘) = ∇π§ π (π‘, πΉΜ between the two systems. By the definition of spectral flow and it is catenation property [19], we have ππΉ (−π΅∞ (−π‘)) − ππΉ (−π΅0 (−π‘)) = ππΉ (π΅0 (π‘)) − ππΉ (π΅∞ (π‘)). Thus, we only consider + from now on. the case of π ∞ From (55) and (57), we have βπ§ − π§π βπΏ2 < π; thus, πΎ has a finite π-net in πΏ2 , and so the embedding πΈ σ³¨ → πΏ2 is compact. Step 2 (the case of π > 2). Since πΈ is continuously embedded in π»1/2 , hence, by the Sobolev embedding theorem, πΈ is continuously embedded in πΏπ , for all π > 2. For any π > 2, by the HoΜlder inequality we have π−1 Proof of Lemma 1. Recall the operator πΉΜ = −J(π/ππ‘) + πΏ(π‘), Μ = π»1 (R, R2π) if πΏ(π‘) is bounded and with domain π·(πΉ) 1 Μ is a Hilbert Μ π·(πΉ) ⊂ π» (R, R2π) if πΏ(π‘) is unbounded. π·(πΉ) Μ space equipped with the norm βπ§βπΉΜ = β(πΌ + |πΉ|)π§βπΏ2 , for all Μ Recall the Hilbert space πΈ = π·(|πΉ| Μ 1/2 ), with the π§ ∈ π·(πΉ). inner product and norm by σ΅¨ σ΅¨1/2 σ΅¨ σ΅¨1/2 (π’, V)πΈ = (σ΅¨σ΅¨σ΅¨σ΅¨πΉΜσ΅¨σ΅¨σ΅¨σ΅¨ π’, σ΅¨σ΅¨σ΅¨σ΅¨πΉΜσ΅¨σ΅¨σ΅¨σ΅¨ V) + (π’, V)2 , 2 βπ’βπΈ = Step 1 (the case of π = 2). For π > 0, from (πΏ 1 ) and (53) we have |π‘|>π |π§|2 ππ‘ ≤ π|π |πΌ−2 ∫ |π‘|>π β¨πΏ (π‘) π§, ππ π§β©R2π ππ‘ (54) ≤ π|π |πΌ−2 βπ§β2πΈ . For any π > 0, from (54), we can choose π 0 large enough, such that ∫ |π‘|>π 0 |π§|2 ππ‘ < π2 , 4 ∀π§ ∈ πΎ. (55) On the other hand, by the definition of β ⋅ βπΈ , we have ∫ 2 |π‘|≤π 0 Step 3 (the case of 1 ≤ π ∈ (2/(1 + πΌ), 2)). First, we have (πΌ/(2 − π)) ⋅ π > 1; so we can choose πΌπ satisfying πΌπ ∈ (0, πΌ) and (πΌπ /(2 − π) ⋅ π) > 1. Denote that π = πΌπ /(2 − π). For π > 0 and π§ ∈ πΈ, denote that πΈπ 1 (π§) = {π‘; |π‘| ≥ π and |π‘|π |π§(π‘)| > 1} and πΈπ 2 (π§) = {π‘; |π‘| ≥ π and |π‘|π |π§(π‘)| ≤ 1}. Then, from (53), ∫ πΈπ 1 (π§) πβπ§β2πΈ . |π§| ππ‘ ≤ βπ§βπΈ ≤ πΆ, ∀π§ ∈ πΎ. (56) Thus, by the Sobolev compact embedding theorem, there exist π§1 , π§2 , . . . , π§π ∈ πΎ, such that for any π§ ∈ πΎ, there is π§π satisfying π2 σ΅©π σ΅©σ΅© σ΅©σ΅©π§ − π§π σ΅©σ΅©σ΅©πΏπ ((−π ,π ),R2π ) < . 0 0 2 (57) (58) thus, the embedding πΈ σ³¨ → πΏπ is compact, for all π > 2. (52) Let πΎ ⊂ πΈ be a bounded set. We will show that πΎ is precompact in πΏπ for 1 ≤ π ∈ (2/(1 + πΌ), ∞). We divide the proof into three steps. ∫ R (π’, π’)1/2 πΈ , where (⋅, ⋅)πΏ2 denotes the usual inner product on πΏ2 (R, R2π). From (πΏ 1 ), there is a matrix πΏ 0 such that π(πΏ(π‘) − πΏ 0 ) ≥ 0 for Μ πΉΜ0 +πΏ 0 , with πΉΜ0 = −J(π/ππ‘)+πΏ 0 all π‘ ∈ R. We have πΏ(π‘) = πΉ− 1 and π·(πΉΜ0 ) = π» . Thus, for any π§ ∈ πΈ, σ΅¨σ΅¨σ΅¨(πΏ (π‘) π§, ππ π§) σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨ πΏ2 σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ σ΅¨ Μ ππ π§) σ΅¨σ΅¨σ΅¨ + σ΅¨σ΅¨σ΅¨(πΉΜ0 π§, ππ π§) σ΅¨σ΅¨σ΅¨ + σ΅¨σ΅¨σ΅¨(πΏ 0 π§, ππ π§) σ΅¨σ΅¨σ΅¨ (53) ≤ σ΅¨σ΅¨σ΅¨(πΉπ§, σ΅¨ σ΅¨ σ΅¨ σ΅¨ πΏ2 σ΅¨ σ΅¨ πΏ2 σ΅¨ σ΅¨ πΏ 2 σ΅¨σ΅¨ σ΅¨ ≤ π−1 ∫ |π§|π ππ‘ ≤ βπ§βπΏ2 βπ§βπΏ2(π−1) ≤ πΆβπ§βπΏ2 βπ§βπΈ , 3. Proof of Our Main Results π |π§|π ππ‘ = ∫ πΈπ 1 (π§) ≤∫ πΈπ 1 (π§) ≤ ≤ π (|π‘|π |π§|) |π‘|−ππ ππ‘ |π§|2 |π‘|πΌπ ππ‘ |π |πΌ−πΌπ σ΅¨σ΅¨ σ΅¨ σ΅¨σ΅¨(πΏ (π‘) π§, ππ π§) 2 σ΅¨σ΅¨σ΅¨ σ΅¨ πΏ σ΅¨ (59) π βπ§β2πΈ , |π |πΌ−πΌπ and so ∫ |π‘|≥π |π§|π ππ‘ =∫ πΈπ 1 (π§) ≤ |π§|π ππ‘ + ∫ πΈπ 2 (π§) π πΌ−πΌπ |π | βπ§β2πΈ + |π§|π ππ‘ (60) 2 , (ππ − 1) π ππ−1 ∀π§ ∈ πΈ. Let πΎ ⊂ πΈ be a bounded set. For any π > 0, from (60), choose π 0 > 0 large enough, such that ∫ |π‘|≥π 0 |π§|π ππ‘ < ππ , 4 ∀π§ ∈ πΎ. (61) On the other hand, by the Sobolev compact embedding theorem, there are π§1 , π§2 , . . . , π§π ∈ πΎ, such that for any π§ ∈ πΎ, there exists π§π satisfying ππ σ΅©σ΅© σ΅©π σ΅©σ΅©π§ − π§π σ΅©σ΅©σ΅©πΏπ ((−π ,π ),R2π ) < . 0 0 2 (62) From (61) and (62), we have σ΅© σ΅©σ΅© σ΅©σ΅©π§ − π§π σ΅©σ΅©σ΅©πΏπ < π; (63) that is to say, πΎ has a finite π-net in πΏπ , and the embedding πΈ σ³¨ → πΏπ is compact. The proof of the lemma is complete. Abstract and Applied Analysis 7 Consider the homoclinic orbits of the linear Hamiltonian systems π§Μ (π‘) = π½π΅ (π‘) π§ (π‘) , π§ (π‘) σ³¨→ 0, ∀π‘ ∈ R, |π‘| σ³¨→ ∞, the saddle point reduction, choose a suitable number π½, which is used in the projection for the saddle point reduction in Section 2.2. Let π½ (64) π = ∫ ππΈπ , −π½ 2 ( πΌ0π −πΌ0π ), and π΅(π‘) is a where π§(π‘) : R → R2π, π½ = continuous symmetric matrix function. Denote by π the set of homoclinic orbits of linear systems (64); then, π is a linear subspace of πΏ2 (R, R2π) and we have the following lemma. Lemma 14. The dimension of the solution space π will be less than or equal to π. Thus for any homoclinic orbit π§(π‘) of (HS), if π satisfies (π 1 ), one has 0 ≤ ππΉ (π§) ≤ π. (65) Proof. As usual, we define the symplectic groups on R2π by ππ (2π) = {π ∈ L (R2π) , | ππ π½π = π½} , (66) where L(R2π) is the set of all 2π × 2π real matrices, and ππ denotes the transpose of π. Let π(π‘) be the fundamental solution of (64); then, π(π‘) is a path in ππ(2π). Let π§(π‘) be a nontrivial homoclinic orbit of (64); that is to say, π§(0) =ΜΈ 0 and satisfies π§ (π‘) = π (π‘) π§ (0) , lim π (π‘) π§ (0) = 0. (67) π‘→∞ π‘→∞ (68) lim π (π‘) π§0 = 0, lim π (π‘) π½π§0 = 0. (69) π‘→∞ Since π(π‘) is a path in ππ(2π), ππ (π‘)π½π(π‘) = π½, for all π‘ ∈ R, thus 0 = lim (π½π (π‘) π§0 , π (π‘) π½π§0 )R2π π‘→∞ = − lim (π§0 , ππ (π‘) π½π (π‘) π½π§0 )R2π π‘→∞ By Theorem 8, we have a functional π(π₯) with π₯ ∈ π, whose critical points give rise to solutions of (HS). Lemma 15. (1) π satisfies (PS) condition; (2) π»π (π, π; R) ≅ πΏπ,π R, π = 0, 1, . . . for −π ∈ R large enough, where π = dim(πΈ− (π)) + ππΉ (πΎπ΅1 ). Proof. Assume that there is a sequence {π₯π } ⊂ π, satisfying πσΈ (π₯π ) → 0(π → ∞). That is, σ΅©σ΅©σ΅©πΉπ§π − πΎ∇π§ π (π‘, π§π )σ΅©σ΅©σ΅© σ³¨→ 0, (72) σ΅© σ΅©πΈ where π§π = π§(π₯π ) defined in Section 2.1. Since π is a finite dimensional space, and from the definition of π§π , it is enough to prove that {π§π } is bounded in πΈ. For each π ∈ (0, 1), define πΆπ ∈ B by 1 2 { { {∫0 ∇π§ π (π‘, π π§π ) ππ , πΆπ (π‘) = { { { π΅ (π‘) , { 1 σ΅¨ π σ΅¨σ΅¨ σ΅¨σ΅¨π§π (π‘)σ΅¨σ΅¨σ΅¨ ≥ 0 , π π σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨π§π (π‘)σ΅¨σ΅¨σ΅¨ < 0 . π (70) = (π§0 , π§0 )R2π , which contradicts π§0 =ΜΈ 0. Since π½ is an isomorphism on R2π, we have dim π0 ≤ π. And from the definition of ππΉ (π§) in the last part of Section 2.2, we have completed the proof. Before the proof of Theorem 2, we need the following lemma. Since π satisfies condition (π 1 ), performing on (HS) (73) It is easy to verify that {πΆπ } satisfies ≤ πΆπ (π‘) ≤ π΅2 (π‘) + π (π ⋅ πΌ − π΅2 (π‘)) , Then, we have dim π = dim(π0 ). We claim that π½π§0 ∉ π0 if π§0 ∈ π0 and π§0 =ΜΈ 0. We prove it indirectly. Assume that π§0 , π½π§0 ∈ π0 with π§0 =ΜΈ 0; that is to say, π‘→∞ π = ππΏ (R, R ) . π΅1 (π‘) − π (π΅1 (π‘) + π ⋅ πΌ) Denote by π0 the subset of R2π satisfying π0 = {π§ ∈ R2π | lim π (π‘) π§ = 0} , (71) 2π ∀π‘ ∈ R, (74) where π is the constant in condition (π 1 ) and πΌ is the identity map on R2π. Since π΅1 ≤ π΅2 , ππΉ (π΅1 ) = ππΉ (π΅2 ) and ππΉ (πΎπ΅1 ) = ππΉ (πΎπ΅2 ) = 0, we can choose π small enough, such that for each π ∈ N+ , ππΉ (πΎπΆπ ) = ππΉ (πΎπ΅1 ) and ππΉ (πΎπΆπ ) = 0. Thus πΉ − πΎπΆπ is reversible on πΈ, and there is a constant πΏ > 0, such that σ΅© σ΅©σ΅© + (75) σ΅©σ΅©(πΉ − πΎπΆπ )π§σ΅©σ΅©σ΅©πΈ ≥ πΏβπ§βπΈ , ∀π§ ∈ πΈ, π ∈ N . On the other hand, for π ∈ (0, 1), there is a constant π > 0 depending on π, such that for each π ∈ N+ , σ΅¨ σ΅¨ σ΅¨π σ΅¨σ΅¨ (76) σ΅¨σ΅¨∇π§ π (π‘, π§π (π‘)) − πΆπ π§π (π‘)σ΅¨σ΅¨σ΅¨ ≤ π σ΅¨σ΅¨σ΅¨π§π (π‘)σ΅¨σ΅¨σ΅¨ , ∀π‘ ∈ R. Choose π > (1−πΌ)/(1+πΌ) in (76); that is, 1+π ∈ (2/(1+πΌ), 2), we have σ΅©2 σ΅©σ΅© σ΅©σ΅©(πΉπ§π − πΎ∇π§ π (π‘, π§π )) − (πΉ − πΎπΆπ ) π§π σ΅©σ΅©σ΅©πΈ σ΅© σ΅©2 = σ΅©σ΅©σ΅©πΎ (∇π§ π (π‘, π§π ) − πΆπ π§π )σ΅©σ΅©σ΅©πΈ σ΅©2 σ΅© ≤ σ΅©σ΅©σ΅©∇π§ π (π‘, π§π ) − πΆπ π§π σ΅©σ΅©σ΅©πΏ2 (77) σ΅¨σ΅¨ σ΅¨ σ΅¨σ΅¨∇π§ π (π‘, π§π ) − πΆπ π§π σ΅¨σ΅¨σ΅¨ σ΅¨σ΅¨ σ΅¨σ΅¨1+π ≤ π∫ σ΅¨σ΅¨π§π σ΅¨σ΅¨ ππ‘ σ΅¨σ΅¨ σ΅¨σ΅¨π R σ΅¨σ΅¨π§π σ΅¨σ΅¨ σ΅© σ΅©1+π ≤ πσ΅©σ΅©σ΅©π§π σ΅©σ΅©σ΅©πΏ1+π . 8 Abstract and Applied Analysis For each π ∈ N, we consider the following problem As we claimed in the introduction, in (76) and (77), the letter π denotes different positive constants whose exact value is irrelevant. Thus, from (72), (75), (77), and Lemma 1, we have that {π§π } is bounded in πΈ, and π satisfies the (PS) conditions. And by Lemma 5.1 in Chapter II of [21], we have π»π (π, (π)πΌ ; R) ≅ πΏπ,π R, π = 0, 1, . . . , (78) for −πΌ ∈ R large enough. From Theorem 11, Lemmas 14 and 15, Theorem 2 is a direct consequence of Theorem 5.1 and Corollary 5.2 in Chapter II of [21]. In order to proof Theorems 3 and 4, we need the following lemma which is similar to Lemma 3.4 in [29] and Lemma 3.3 in [23]. Μ = ∇π§ π π (π‘, π§) , πΉπ§ π§ (π‘) σ³¨→ 0, π π (π‘, π§) ≡ π (π‘, π§) , ∀π‘ ∈ R, |π§| ≤ ππ . (3) For each π ∈ N, there exist some πΆπ > 0 and a constant πΎ with πΎπΌ2π > π΅∞ , ]πΉ (πΎπΎπΌ2π) = 0, such that σ΅¨σ΅¨ σ΅¨ σ΅¨σ΅¨∇π§ π π (π‘, π§) − πΎπ§σ΅¨σ΅¨σ΅¨ < πΆπ , ∀ (π‘, π§) ∈ R × R2π, (81) where πΌ2π is the identity map on R2π. Proof. Define π : [0, ∞) → R by 0, 0 ≤ π < 1, { { { { { { 1 {2 3 4 π (π ) = { 9 (π − 1) − 9 (π − 1) , 1 ≤ π < 2, { { { { 128 { {1 − , 2 ≤ π < ∞. 9 (12 + π 2 ) { πΎ π π (π‘, π§) = (1 − ππ (|π§|)) π (π‘, π§) + ππ (|π§|) |π§|2 , 2 π ∈ N. (83) As in [23, 29], we can check that π π satisfies (79)–(81) for each π ∈ N. (84) π½ ππ½ = ∫ ππΈπ , −π½ 2 (85) 2π ππ½ = ππ½ πΏ (R, R ) . Thus, for each π and such a π½ fixed, by Theorem 8, we have a functional ππ,π½ (π₯) , π₯ ∈ ππ½ , (86) whose critical points give rise to solutions of (HS)π . Similarly we have a functional ππΎ,π½ (π₯) , π₯ ∈ ππ½ , (87) whose critical points give rise to solutions of the following systems (HS)πΎ Μ = πΎπ§, πΉπ§ π§ (π‘) σ³¨→ 0, π§σΈ (π‘) σ³¨→ 0, π‘ σ³¨→ ∞. (HS)πΎ For notational simplicity, we denote ππ , ππΎ for ππ,π½ and ππΎ,π½ . Define ∞ −∞ It is easy to see that π ∈ πΆ2 ([0, ∞), R). Choose a sequence {ππ } of positive numbers such that π 0 < π1 < π2 < ⋅ ⋅ ⋅ < ππ < ⋅ ⋅ ⋅ → ∞ as π → ∞. For each π ∈ N, let ππ (π ) = π(π /ππ ) and (HS)π Let Φπ (π§) = ∫ (82) π½ ∉ π (π΄ 0 ) . π½ > max {2 (πΆ + 1) , 2 (πΎ + 1)} , (79) (2) For each π ∈ N, there is a πΆ > 0 independent of π, such that σ΅¨ σ΅¨σ΅¨ 2 σ΅¨σ΅¨∇π§ π π (π‘, π§)σ΅¨σ΅¨σ΅¨ ≤ πΆ, ∀π‘ ∈ R, π§ ∈ R2π, σ΅¨ σ΅¨ (80) 2 ∇π§ π π (π‘, π§) ≥ π΅∞ , ∀π‘ ∈ R, |π§| ≥ π 0 . π‘ σ³¨→ ∞, where π π is given in Lemma 16. Performing on (HS)π the saddle point reduction, we choose the number π½ which is used in the projection for the saddle point reduction in Section 2.2. First, we choose + ) hold; then, there Lemma 16. Assume that (π 1 ), (π 0 ), and (π ∞ 2 exists a sequence of functions π π ∈ πΆ (R × R2π, R), π ∈ N, satisfying the following properties. (1) There exists an increasing sequence of real numbers ππ → ∞ (π → ∞) such that π§σΈ (π‘) σ³¨→ 0, π π (π‘, π§) ππ‘. (88) For the functional ππ , similar to Lemma 15, we have the following lemma. Lemma 17. (1) ππ satisfies (PS) condition, and the critical point set of ππ is compact; (2) π»π (ππ½ , (ππ )πΌπ ; R) ≅ πΏπ,ππ½ R, π = 0, 1, . . . for −πΌπ ∈ R large enough, where ππ½ = dim(πΈ− (ππ½ )) + ππΉ (πΎπΎπΌ2π). Proof. The proof is similar to Lemma 15. From Theorem 11, we have σ΅©σ΅©σ΅©πσΈ (π₯) − πσΈ σΈ (0) π₯σ΅©σ΅©σ΅© = σ΅©σ΅©σ΅©π πΎ(∇ Φ (π§ (π₯)) − πΎπ₯)σ΅©σ΅©σ΅© σ΅©σ΅© π σ΅©σ΅©πΏ2 σ΅©σ΅© π½ σ΅©σ΅©πΏ2 π§ π π πΎ σ΅© σ΅© ≤ σ΅©σ΅©σ΅©σ΅©ππ½ (∇π§ π π (π§π (π₯)) − πΎπ§π (π₯))σ΅©σ΅©σ΅©σ΅©πΏ2 σ΅© σ΅© ≤ σ΅©σ΅©σ΅©(∇π§ π π (π§π (π₯)) − πΎπ§π (π₯))σ΅©σ΅©σ΅©πΏ2 . (89) Abstract and Applied Analysis 9 Similar to (76), for π ∈ (0, 1), there is some π > 0, such that σ΅¨ σ΅¨ σ΅¨σ΅¨ σ΅¨ (90) σ΅¨σ΅¨∇π§ π π (π‘, π§π ) − πΎπ§π σ΅¨σ΅¨σ΅¨ ≤ π σ΅¨σ΅¨σ΅¨π§π (π‘)σ΅¨σ΅¨σ΅¨ , ∀π‘ ∈ R. for π½ large enough, we have the functional π∞,π½ (denote by π∞ for simplicity) and the function π§(π₯); since ππΉ (πΎπ΅∞ ) = 0, we have the following decomposition: Choose π ∈ ((1 − πΌ)/(1 + πΌ), 1), similar to (77), and we have ππ½ = ππ½+ + ππ½− , σ΅©σ΅© σ΅©2 σ΅© σ΅©1+π σ΅©σ΅©∇π§ π π (π‘, π§π ) − πΎπ§π σ΅©σ΅©σ΅©πΏ2 ≤ πσ΅©σ΅©σ΅©π§π σ΅©σ΅©σ΅©πΏ1+π . (91) From Lemma 1, Remark 10, and (91), σ΅©σ΅© σ΅©2 σ΅© σ΅©1+π σ΅©σ΅©∇π§ π π (π‘, π§π ) − πΎπ§π σ΅©σ΅©σ΅©πΏ2 ≤ πσ΅©σ΅©σ΅©π§π σ΅©σ΅©σ΅©πΈ ≤ π (π½) βπ₯β1+π πΏ2 . (92) From (89) and (92), we have σ΅©σ΅© σΈ σ΅©2 σ΅©σ΅©ππ (π₯) − ππΎσΈ σΈ (0) π₯σ΅©σ΅©σ΅© 2 ≤ π (π½) βπ₯β1+π πΏ2 . σ΅© σ΅©πΏ (93) Now, for each π ∈ N, we assume that {π₯π } ⊂ ππ½ satisfying βππσΈ (π₯π )β → 0. By ]πΉ (πΎπΎπΌ2π) = 0, we have that ππΎσΈ σΈ (0) is invertible on ππ½ , since π < 1 the sequence {π₯π } must be bounded. Thus the (PS) condition for ππ holds. From the same reason, we have the compactness of the critical point set of ππ . And by Lemma 5.1 in Chapter II of [21], we have π»π (ππ½ , (ππ )πΌπ ; R) ≅ πΏπ,ππ½ R, π = 0, 1, . . . , (94) for −πΌπ ∈ R large enough. ((πΉΜ − π΅∞ (π‘)) π₯, π₯)πΏ2 ≤ −πΌβπ₯β2πΏ2 , ∀π₯ ∈ ππ½− . (99) From the uniform boundary of ∇π§2 π π (π‘, π§) and Remark 10, we can choose π½ large enough, such that πΌ σ΅© σ΅©σ΅© 2 σ΅©σ΅©(∇π§ π π (π‘, π§π ) π§πσΈ ± (π₯π ) π₯, π₯σ΅©σ΅©σ΅© 2 ≤ βπ₯βπΏ2 , σ΅©πΏ σ΅© 4 ∀π₯ ∈ πΏ2 , (100) where π₯π = ππ½ π§π , π§π (π₯π ) = π§π (π‘) defined in Theorem 8. Choose π > 0 small enough and π < πΌ/4, such that ππΉ (πΎπ΅∞ ) = ππΉ (πΎ(π΅∞ −π⋅πΌπ)). Since ππ½− is finite dimensional space, choose π large enough, such that ((∇π§2 π π (π‘, π§π ) − π΅∞ ) π₯, π₯)πΏ2 (πΌπ ) ≥ −π(π₯, π₯)πΏ2 , ∀π₯ ∈ ππ½− , (101) Proof. We prove it indirectly. Assume there exist π π and π§π satisfying the conditions, and βπ§π βπΏ∞ → ∞; that is, βπ§βπΉΜ → ∞. Since |∇π§ π π (π‘, π§)| < π|π§|, for all π‘ ∈ R, π§ ∈ R2π, we have βπ§π βπΉΜ ≤ πβπ§π βπΏ2 . Denote that π¦π = π§π /βπ§π βπΉΜ ; then, we have π¦π → π¦ in πΏ2 for some π¦ ∈ πΏ2 with βπ¦βπΏ2 > 0, and π πσΈ (π‘, π§π ) σ΅©σ΅© σ΅©σ΅© . σ΅©σ΅©π§π σ΅©σ΅©πΉΜ (95) Then, for any π > 0, there exists πΆπ > 0, satisfying σ΅¨ σ΅¨ σ΅¨σ΅¨ Μ σ΅¨ σ΅¨σ΅¨π¦π (π‘)σ΅¨σ΅¨σ΅¨ ≤ πΆπ σ΅¨σ΅¨σ΅¨π¦π (π‘)σ΅¨σ΅¨σ΅¨ , π‘ ∈ πΌπ , 1σ΅© σ΅© σ΅©σ΅© σ΅©σ΅© σ΅©σ΅©π¦σ΅©σ΅©πΏ2 (−π,π) > σ΅©σ΅©σ΅©π¦σ΅©σ΅©σ΅©πΏ2 > 0, 2 ∀π > π0 . (96) (97) Then, from the similar argument in [23], there is a subsequence throught which we may assume that {π¦π } converges in uniform norm to π¦, and π¦(π‘) =ΜΈ 0, for all π‘ ∈ πΌπ . Therefore |π§π (π‘)| → ∞ uniformly on πΌπ , and there is πΎ(π) depending on π, such that |π§π (π‘)| ≥ π 0 , for any π‘ ∈ πΌπ and π ≥ πΎ(π). Performing the saddle point reduction on the following systems: Μ = π΅∞ π§, πΉπ§ π§σΈ (π‘) σ³¨→ 0, π‘ σ³¨→ ∞, where πΌππ = R \ πΌπ , and from the definition of π π , ∇π§2 π π (π‘, π§π (π‘)) ≥ π΅∞ (π‘) , π‘ ∈ πΌπ , π ≥ πΎ (π) ; (HS)∞ (102) that is, ((∇π§2 π π (π‘, π§π ) − π΅∞ ) π₯, π₯)πΏ2 (πΌ ) ≥ 0, π ∀π₯ ∈ ππ½− , π ≥ πΎ (π) . (103) From (101) and (103), (∇π§2 π π (π‘, π§π ) π₯, π₯)πΏ2 ≥ (π΅∞ π₯, π₯)πΏ2 − π (π₯, π₯)πΏ2 , where πΌπ = [−π, π]. Since βπ¦βπΏ2 > 0, there is a π0 > 0, such that π§ (π‘) σ³¨→ 0, σΈ σΈ where π∞ (0) is positive definite on ππ½+ and negative definite on ππ½− . From Remark 10, and ππΉ (πΎπ΅∞ ) = 0, there exists πΌ > 0, such that for π½ large enough, π Lemma 18. There exist π > 0, such that for any π ∈ N, and π§ ∈ πΏ2 satisfies the systems (HS)π , if ππΉ (π§) ≤ ππΉ (πΎπ΅∞ ) − 1, one has βπ§βπΏ∞ ≤ π. Μ π= πΉπ¦ (98) (104) for π large enough. Thus we have (ππσΈ σΈ (π₯π ) π₯, π₯)πΏ2 = ((πΉΜ − ∇π§2 π π (π‘, π§π )) π₯, π₯)πΏ2 − (∇π§2 (π π (π‘, π§π )) (π§πσΈ + + π§πσΈ − ) π₯, π₯)πΏ2 (105) πΌ ≤ ((πΉΜ − π΅∞ ) π₯, π₯)πΏ2 + βπ₯β2πΏ2 + πβπ₯β2πΏ2 2 πΌ ≤ − βπ₯β2πΏ2 . 4 That is, ππ−π (π₯) ≥ ππ−∞ (0), from Theorem 11, ππ−π (π₯π ) = dim(πΈ− (H0 )) + ππΉ (π§π₯ (π₯π )), ππ−∞ (0) = dim(πΈ− (H0 )) + ππΉ (πΎπ΅∞ ), and thus ππΉ (π§π₯ ) ≥ ππΉ (πΎπ΅∞ ), which contradicts the assumption. 10 Abstract and Applied Analysis Proof of Theorem 3. As claimed in Remark 13, we can only + ). Note that π§ = 0 is a critical point of consider the case of (π ∞ ππ , and the morse index of 0 for ππ is ππ−π (0) = dim(πΈ− (H0 ))+ ππΉ (πΎπ΅0 ), since πΎ ⋅ πΌ2π > π΅∞ , we have ππΉ (πΎπΎ ⋅ πΌ2π) ≥ ππΉ (πΎπ΅∞ ) ≥ ππΉ (πΎπ΅0 ) . (106) From proposition (2) in Lemma 17, use the (ππ−π (0))th and (ππ−π (0) + 1)th Morse inequalities, and ππ has a nontrivial critical point π₯π with its morse index ππ−π (π₯π ) ≤ ππ−π (0) + 1; that is, ππΉ (π§π ) ≤ ππΉ (πΎπ΅0 ) + 1 ≤ ππΉ (πΎπ΅∞ ) − 1, then from Lemma 18, we have that {π§π } is bounded in πΏ∞ . Thus π§π is a nontrivial solution of (HS) for π large enough. The proof of Theorem 4 is similar to the proof of Theorem 3. Instead of Morse theory we make use of minimax arguments for multiplicity of critical points. Let π be a Hilbert space and assume π ∈ πΆ2 (π, R) is an even functional, satisfying the (PS) condition and π(0) = 0. Denote ππ = {π’ ∈| πβπ’β = π}. Lemma 19 (see [30, Corollary 10.19]). Assume that π and π are subspaces of π satisfying dim π = π > π = codim π. If there exist π > π > 0 and πΌ > 0 such that inf π (ππ ∩ π) ≥ πΌ, sup π (ππ ∩ π) ≤ 0, (107) then π has π−π pairs of nontrivial critical points {±π₯1 , ±π₯2 , . . . , ±π₯π−π }, so that π(π₯π ) ≤ π + π, for π = 1, 2, . . . , π − π. + ), since π is even, we First, we consider the case of (π ∞ have that π π is also even and satisfies Lemma 16. Let π = ππ½− , and π the positive space of ππσΈ σΈ (0) in ππ½ , and we have and dim π = πΈ− (ππ½ ) + ππΉ (πΎπ΅∞ ), codim π = πΈ− (ππ½ ) + ππΉ (πΎπ΅0 ), dim π > codimπ. So, ππ has π := ππΉ (πΎπ΅∞ ) − ππΉ (πΎπ΅0 ) pairs of nontrivial critical points {±π₯1 , ±π₯2 , . . . , ±π₯π } , (108) and π − 1 pairs of them satisfy π− (π₯π ) ≤ ππΉ (πΎπ΅0 ) + π < ππΉ (πΎπ΅∞ ) , π = 1, 2, . . . , π − 1. (109) Then, we can complete the proof. In order to prove the case − ), we need the following lemma. of (π ∞ Lemma 20 (see [21, Corollary II 4.1]). Assume that π and π are subspaces of π satisfying dim π = π > π = codimπ. If there exist π > 0, and πΌ > 0 such that inf π (π) > −∞, sup π (ππ ∩ π) ≤ −πΌ, (110) then π has π − π pairs of nontrivial critical points ±π’1 , ±π’2 , . . . , ±π’π−π so that π(π’π ) + ](π’π ) ≥ π + π − 1 for π = 1, 2, . . . , π − π. + ), and we omit it The proof is similar to the case of (π ∞ here. Acknowledgment This work is partially supported by NFSC (11126154, 11126146, and 11201196). References [1] G. Arioli and A. Szulkin, “Homoclinic solutions of Hamiltonian systems with symmetry,” Journal of Differential Equations, vol. 158, no. 2, pp. 291–313, 1999. [2] G. Chen and S. Ma, “Homoclinic orbits of superlinear Hamiltonian systems,” Proceedings of the American Mathematical Society, vol. 139, no. 11, pp. 3973–3983, 2011. [3] V. C. Zelati, I. Ekeland, and E. SeΜreΜ, “A variational approach to homoclinic orbits in Hamiltonian systems,” Mathematische Annalen, vol. 288, no. 1, pp. 133–160, 1990. [4] Y. Ding, “Multiple homoclinics in a Hamiltonian system with asymptotically or super linear terms,” Communications in Contemporary Mathematics, vol. 8, no. 4, pp. 453–480, 2006. [5] Y. Ding and M. Girardi, “Infinitely many homoclinic orbits of a Hamiltonian system with symmetry,” Nonlinear Analysis: Theory, Methods & Applications, vol. 38, no. 3, pp. 391–415, 1999. [6] Y. Ding and M. Willem, “Homoclinic orbits of a Hamiltonian system,” Zeitschrift fuΜr Angewandte Mathematik und Physik, vol. 50, no. 5, pp. 759–778, 1999. [7] H. Hofer and K. Wysocki, “First order elliptic systems and the existence of homoclinic orbits in Hamiltonian systems,” Mathematische Annalen, vol. 288, no. 3, pp. 483–503, 1990. [8] E. SeΜreΜ, “Looking for the Bernoulli shift,” Annales de l’Institut Henri PoincareΜ. Analyse Non LineΜaire, vol. 10, no. 5, pp. 561–590, 1993. [9] E. SeΜreΜ, “Existence of infinitely many homoclinic orbits in Hamiltonian systems,” Mathematische Zeitschrift, vol. 209, no. 1, pp. 27–42, 1992. [10] J. Sun, J. Chu, and Z. Feng, “Homoclinic orbits for first order periodic Hamiltonian systems with spectrum point zero,” Discrete and Continuous Dynamical Systems, vol. 33, no. 8, pp. 1241– 1257, 2013. [11] A. Szulkin and W. Zou, “Homoclinic orbits for asymptotically linear Hamiltonian systems,” Journal of Functional Analysis, vol. 187, no. 1, pp. 25–41, 2001. [12] K. Tanaka, “Homoclinic orbits in a first order superquadratic Hamiltonian system: convergence of subharmonic orbits,” Journal of Differential Equations, vol. 94, no. 2, pp. 315–339, 1991. [13] Y. H. Ding and S. J. Li, “Homoclinic orbits for first order Hamiltonian systems,” Journal of Mathematical Analysis and Applications, vol. 189, no. 2, pp. 585–601, 1995. [14] Y. Ding and L. Jeanjean, “Homoclinic orbits for a nonperiodic Hamiltonian system,” Journal of Differential Equations, vol. 237, no. 2, pp. 473–490, 2007. [15] Y. Ding and C. Lee, “Existence and exponential decay of homoclinics in a nonperiodic superquadratic Hamiltonian system,” Journal of Differential Equations, vol. 246, no. 7, pp. 2829–2848, 2009. [16] W. Qin, J. Zhang, and F. Zhao, “Homoclinic orbits for a class of nonperiodic Hamiltonian systems,” Abstract and Applied Analysis, vol. 2012, Article ID 769232, 20 pages, 2012. [17] J. Sun, H. Chen, and J. J. Nieto, “Homoclinic orbits for a class of first-order nonperiodic asymptotically quadratic Hamiltonian systems with spectrum point zero,” Journal of Mathematical Analysis and Applications, vol. 378, no. 1, pp. 117–127, 2011. Abstract and Applied Analysis [18] C. N. Chen and X. Hu, “Maslov index for homoclinic orbits of Hamiltonian systems,” Annales de l’Institut Henri PoincareΜ. Analyse Non LineΜaire, vol. 24, no. 4, pp. 589–603, 2007. [19] C. Zhu and Y. Long, “Maslov-type index theory for symplectic paths and spectral flow. I,” Chinese Annals of Mathematics B, vol. 20, no. 4, pp. 413–424, 1999. [20] H. Amann, “Saddle points and multiple solutions of differential equations,” Mathematische Zeitschrift, vol. 169, no. 2, pp. 127– 166, 1979. [21] K. C. Chang, Infinite-Dimensional Morse Theory and Multiple Solution Problems, Progress in Nonlinear Differential Equations and their Applications, BirkhaΜauser, Basel, Switzerland, 1993. [22] Y. Long, Index Theory for Symplectic Paths with Applications, vol. 207 of Progress in Mathematics, BirkhaΜuser, Basel, Switzerland, 2002. [23] Z. Liu, J. Su, and Z. Q. Wang, “A twist condition and periodic solutions of Hamiltonian systems,” Advances in Mathematics, vol. 218, no. 6, pp. 1895–1913, 2008. [24] M. F. Atiyah, V. K. Patodi, and I. M. Singer, “Spectral asymmetry and Riemannian geometry. III,” Mathematical Proceedings of the Cambridge Philosophical Society, vol. 79, no. 1, pp. 71–99, 1976. [25] S. E. Cappell, R. Lee, and E. Y. Miller, “On the Maslov index,” Communications on Pure and Applied Mathematics, vol. 47, no. 2, pp. 121–186, 1994. [26] A. Floer, “A relative Morse index for the symplectic action,” Communications on Pure and Applied Mathematics, vol. 41, no. 4, pp. 393–407, 1988. [27] J. Robbin and D. Salamon, “The Maslov index for paths,” Topology, vol. 32, no. 4, pp. 827–844, 1993. [28] J. Robbin and D. Salamon, “The spectral flow and the Maslov index,” The Bulletin of the London Mathematical Society, vol. 27, no. 1, pp. 1–33, 1995. [29] Z. Liu, J. Su, and Z. Q. Wang, “Solutions of elliptic problems with nonlinearities of linear growth,” Calculus of Variations and Partial Differential Equations, vol. 35, no. 4, pp. 463–480, 2009. [30] N. Ghoussoub, Duality and Perturbation Methods in Critical Point Theory, vol. 107 of Cambridge Tracts in Mathematics, Cambridge University Press, Cambridge, UK, 1993. 11 Advances in Operations Research Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Advances in Decision Sciences Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Mathematical Problems in Engineering Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Journal of Algebra Hindawi Publishing Corporation http://www.hindawi.com Probability and Statistics Volume 2014 The Scientific World Journal Hindawi Publishing Corporation http://www.hindawi.com Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 International Journal of Differential Equations Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Volume 2014 Submit your manuscripts at http://www.hindawi.com International Journal of Advances in Combinatorics Hindawi Publishing Corporation http://www.hindawi.com Mathematical Physics Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Journal of Complex Analysis Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 International Journal of Mathematics and Mathematical Sciences Journal of Hindawi Publishing Corporation http://www.hindawi.com Stochastic Analysis Abstract and Applied Analysis Hindawi Publishing Corporation http://www.hindawi.com Hindawi Publishing Corporation http://www.hindawi.com International Journal of Mathematics Volume 2014 Volume 2014 Discrete Dynamics in Nature and Society Volume 2014 Volume 2014 Journal of Journal of Discrete Mathematics Journal of Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Applied Mathematics Journal of Function Spaces Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Optimization Hindawi Publishing Corporation http://www.hindawi.com Volume 2014 Hindawi Publishing Corporation http://www.hindawi.com Volume 2014