74
Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit¨ at Giessen Arndtstr. 2, 35392 Giessen, Germany Andrzej Szulkin Department of Mathematics, Stockholm University 106 91 Stockholm, Sweden Contents 0 Introduction 2 1 Critical point theory 4 1.1 Basic critical point theory ........................... 4 1.2 Critical point theory for strongly indefinite functionals ........... 14 2 Periodic solutions 22 2.1 Variational setting for periodic solutions ................... 22 2.2 Periodic solutions near equilibria ....................... 30 2.3 Fixed energy problem .............................. 35 2.4 Superlinear systems ............................... 40 2.5 Asymptotically linear systems ......................... 43 2.6 Spatially symmetric Hamiltonian systems ................... 46 3 Homoclinic solutions 50 3.1 Variational setting for homoclinic solutions .................. 50 3.2 Existence of homoclinics ............................ 55 3.3 Multiple homoclinic solutions ......................... 60 3.4 Multibump solutions and relation to the Bernoulli shift ........... 63 References 67 1

staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Hamiltonian systems: periodic and homoclinicsolutions by variational methods

Thomas Bartsch

Mathematisches Institut, Universitat Giessen

Arndtstr. 2, 35392 Giessen, Germany

Andrzej Szulkin

Department of Mathematics, Stockholm University

106 91 Stockholm, Sweden

Contents

0 Introduction 2

1 Critical point theory 4

1.1 Basic critical point theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 41.2 Critical point theory for strongly indefinite functionals . . . . . . . . . . . 14

2 Periodic solutions 22

2.1 Variational setting for periodic solutions . . . . . . . . . . . . . . . . . . . 222.2 Periodic solutions near equilibria . . . . . . . . . . . . . . . . . . . . . . . 30

2.3 Fixed energy problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352.4 Superlinear systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

2.5 Asymptotically linear systems . . . . . . . . . . . . . . . . . . . . . . . . . 432.6 Spatially symmetric Hamiltonian systems . . . . . . . . . . . . . . . . . . . 46

3 Homoclinic solutions 503.1 Variational setting for homoclinic solutions . . . . . . . . . . . . . . . . . . 50

3.2 Existence of homoclinics . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553.3 Multiple homoclinic solutions . . . . . . . . . . . . . . . . . . . . . . . . . 60

3.4 Multibump solutions and relation to the Bernoulli shift . . . . . . . . . . . 63

References 67

1

Page 2: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

0 Introduction

The complex dynamical behavior of Hamiltonian systems has attracted mathematiciansand physicists ever since Newton wrote down the differential equations describing plan-

etary motions and derived Kepler’s ellipses as solutions. Hamiltonian systems can beinvestigated from different points of view and using a large variety of analytical and ge-

ometric tools. The variational treatment of Hamiltonian systems goes back to Poincarewho investigated periodic solutions of conservative systems with two degrees of freedom

using a version of the least action principle. It took however a long time to turn thisprinciple into a useful tool for finding periodic solutions of a general Hamiltonian system

(HS)

p = −Hq(p, q, t)

q = Hp(p, q, t)

as critical points of the Hamiltonian action functional

Φ(p, q) =

∫ 2π

0

p · q dt−∫ 2π

0

H(p, q, t) dt

defined on a suitable space of 2π-periodic functions (p, q) : R → RN × RN . The reasonis that this functional is unbounded from below and from above so that the classical

methods from the calculus of variations do not apply. Even worse, the quadratic form

(p, q) 7→∫ 2π

0

p · q dt

has infinite-dimensional positive and negative eigenspaces. Therefore Φ is said to be

strongly indefinite. For strongly indefinite functionals refined variational methods likeMorse theory or Lusternik-Schnirelmann theory still do not apply. These were originally

developed for the closely related problem of finding geodesics and extended to many otherordinary and partial differential equations, in particular to the second order Hamiltonian

systemq = −Vq(q, t)

where the associated Lagrangian functional

J(q) =1

2

∫ 2π

0

q2 dt−∫ 2π

0

V (q, t) dt

is not strongly indefinite.

A major breakthrough was the pioneering paper [78] of Rabinowitz from 1978 whoobtained for the first time periodic solutions of the first order system (HS) by the above

2

Page 3: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

mentioned variational principle. Some general critical point theory for indefinite function-als was subsequently developed in the 1979 paper [19] by Benci and Rabinowitz. Since

then the number of papers on variational methods for strongly indefinite functionals andon applications to Hamiltonian systems has been growing enormously. These methods

are not restricted to periodic solutions but can also be used to find heteroclinic or ho-moclinic orbits and to prove complex dynamics. In fact, they can even be applied to

infinite-dimensional Hamiltonian systems and strongly indefinite partial differential equa-tions having a variational structure.

The goal of this paper is to present an introduction to variational methods for stronglyindefinite functionals like Φ and its applications to the Hamiltonian system (HS). The

paper is divided into three parts. Part 1 is concerned with critical point theory, Part 2with periodic solutions, and Part 3 with homoclinic solutions of (HS). We give proofs

or sketches of proofs for selected basic theorems and refer to the literature for moreadvanced results. No effort is being made to be as general as possible. Neither did we

try to write a comprehensive survey on (HS). The recent survey [81] of Rabinowitz inVolume 1A of the Handbook of Dynamical Systems facilitated our task considerably. We

chose our topics somewhat complementary to those treated in [81] and concentrated on

the first order system (HS), though a certain overlap cannot and should not be avoided.As a consequence we do not discuss second order systems nor do we discuss convex

Hamiltonian systems where one can work with the dual action functional which is notstrongly indefinite. One more topic which we have not included - though it has recently

attracted attention of many researchers - is the problem of finding heteroclinic solutionsby variational methods. These and many more topics are being treated in a number of

well written monographs dealing with variational methods for Hamiltonian systems, inparticular [1, 5, 33, 52, 69, 73, 80]. Further references can be found in these books and in

Rabinowitz’ survey [81]. Naturally, the choice of the topics is also influenced by our ownresearch experience.

Restricting ourselves to variational methods we do not touch upon the dynamicalsystems approach to Hamiltonian systems which includes perturbation theory, normal

forms, stability, KAM theory, etc. An introduction to these topics can be found forinstance in the textbooks [47, 75]. Also we do not enter the realm of symplectic topology

and Floer homology dealing with Hamiltonian systems on symplectic manifolds. Here we

refer the reader to the monograph [74] and the references therein.We conclude this introduction with a more detailed description of the contents. In the

first part of the paper we consider pertinent results in critical point theory. Particularemphasis is put on a rather simple and direct approach to strongly indefinite functionals.

The second part is concerned with periodic solutions of (HS). We present a unifiedapproach, via a finite-dimensional reduction in order to show the existence of one solution,

and via a Galerkin-type method in order to find more solutions. Section 2.2 concerns theexistence of periodic solutions near equilibria (Lyapunov-type results) and in Section 2.3

3

Page 4: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

the fixed energy problem is considered (finding solutions of a priori unknown period whichlie on a prescribed energy surface). The remaining sections consider the existence and

the number of periodic solutions under growth conditions on the Hamiltonian and forspatially symmetric Hamiltonians.

The third part deals with homoclinic solutions for (HS) with time-periodic Hamilto-nian. Here we present a few basic existence and multiplicity results and discuss a relation

to the Bernoulli shift and complicated dynamics. The proofs are more sketchy than inPart 2 because we did not want to enter too much into technicalities which are more com-

plex than in the periodic case. Moreover, the subject of this part is still rapidly developingand has not been systematized in the same way as the periodic solution problem.

1 Critical point theory

1.1 Basic critical point theory

Let E be a real Hilbert space with an inner product 〈. , .〉 and Φ a functional in C1(E,R).

Via the Riesz representation theorem we shall identify the Frechet derivative Φ′(x) ∈ E∗with a corresponding element of E, and we shall write 〈Φ′(x), y〉 rather than Φ′(x)y. Our

goal here is to discuss those methods of critical point theory which will be useful in ourapplications to Hamiltonian systems. In particular, although most of the results presented

here can be easily extended to real Banach spaces, we do not carry out such extension asit will not be needed for our purposes.

Recall that xj is said to be a Palais-Smale sequence (a (PS)-sequence in short) if

Φ(xj) is bounded and Φ′(xj) → 0. The functional Φ satisfies the Palais-Smale condi-tion (the (PS)-condition) if each (PS)-sequence possesses a convergent subsequence. If

Φ(xj)→ c and Φ′(xj)→ 0, we shall sometimes refer to xj as a (PS)c-sequence. Φ satis-fies the Palais-Smale condition at the level c (the (PS)c-condition) if every (PS)c-sequence

has a convergent subsequence.We shall frequently use the following notation:

Φc := x ∈ E : Φ(x) ≤ c, Kc := x ∈ E : Φ(x) = c, Φ′(x) = 0.

One of the basic technical tools in critical point theory is the deformation lemma.

Below we state a version of it, called the quantitative deformation lemma. It is due to

Willem [101], see also [23], Theorem I.3.4 and [102], Lemma 2.3.A continuous mapping η : A× [0, 1]→ E, where A ⊂ E, is said to be a a deformation

of A in E if η(x, 0) = x for all x ∈ A. Denote the distance from x to the set B by d(x,B).

Lemma 1.1 Suppose Φ ∈ C1(E,R) and let c ∈ R, ε, δ > 0 and a set N ⊂ E be given. If

(1.1) ‖Φ′(x)‖ ≥ δ whenever d(x, E \N) ≤ δ and |Φ(x)− c| ≤ ε,

4

Page 5: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

then there exists an ε ∈ (0, ε), depending only on ε and δ, and a deformation η : E ×[0, 1]→ E such that:

(i) η(x, t) = x whenever |Φ(x)− c| ≥ ε;(ii) η(Φc+ε \N, 1) ⊂ Φc−ε and η(Φc+ε, 1) ⊂ Φc−ε ∪N ;

(iii) The mapping t 7→ Φ(η(x, t)) is nonincreasing for each x ∈ E.

Proof Since the argument is well known, we omit some details. A complete proof may

be found e.g. in [23], Theorem I.3.4 or [102], Lemma 2.3.A mapping V : E \K → E is said to be a pseudo-gradient vector field for Φ if V is

locally Lipschitz continuous and satisfies

(1.2) ‖V (x)‖ ≤ 2‖Φ′(x)‖, 〈Φ′(x), V (x)〉 ≥ ‖Φ′(x)‖2

for each x ∈ E \ K. It is well known and not difficult to prove that any Φ ∈ C1(E,R)has a pseudo-gradient vector field; see e.g. [23, 80, 102].

Let χ : E → [0, 1] be a locally Lipschitz continuous function such that

χ(x) =

0 if |Φ(x)− c| ≥ ε or d(x, E \N) ≥ δ1 if |Φ(x)− c| ≤ ε/2 and d(x, E \N) ≤ δ/2

and consider the Cauchy problem

dt= −1

2δχ(η(x, t))

V (η(x, t))

‖V (η(x, t))‖ , η(x, 0) = x.

Since the vector field above is locally Lipschitz continuous and bounded, η(x, t) is uniquely

determined and continuous for each (x, t) ∈ E×R. It is now easy to see that (i) is satisfied.Moreover,

(1.3)d

dtΦ(η(x, t)) =

⟨Φ′(η(x, t)),

dt

⟩≤ −δ

4χ(η(x, t))‖Φ′(η(x, t))‖

according to (1.2). Hence also (iii) holds.Let x ∈ Φc+ε \N and 0 < ε ≤ ε/2. In order to establish the first part of (ii) we must

show that Φ(η(x, 1)) ≤ c− ε. Since

‖η(x, t)− x‖ ≤∫ t

0

∥∥∥∥dη

ds

∥∥∥∥ ds ≤1

2δt,

d(η(x, t), E \ N) ≤ δ/2 whenever 0 ≤ t ≤ 1. We may assume Φ(η(x, 1)) ≥ c − ε/2(otherwise we are done). Then, according to (1.3) and the definition of χ,

Φ(η(x, 1)) = Φ(x) +

∫ 1

0

d

dtΦ(η(x, t)) dt ≤ Φ(x)− δ

4

∫ 1

0

‖Φ′(η(x, t))‖ dt ≤ c+ ε− δ2/4.

5

Page 6: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Hence Φ(η(x, 1)) ≤ c− ε if we choose ε ≤ minε/2, δ2/8.In order to prove the second part of (ii) it remains to observe that if x ∈ Φc+ε and

η(x, t) /∈ N , then d(η(x, t), E \N) ≤ δ/2 and therefore again Φ(η(x, 1)) ≤ c− ε. 2

We emphasize that the constant ε is independent of the functional Φ and the space Eas long as Φ satisfies (1.1). We shall make repeated use of this fact.

It is easy to see that if Φ satisfies (PS) and N is a neighbourhood of Kc, then thereexist ε, δ > 0 such that (1.1) holds.

Next we introduce the concept of local linking, due to Li and Liu [60]. Let Φ ∈C1(E,R) and denote the ball of radius r and center at the origin by Br. The correspondingsphere will be denoted by Sr. The function Φ is said to satisfy the local linking condition

at 0 if there exists a subspace F0 ⊂ E and α, ρ > 0 such that F0 and F⊥0 have positivedimension,

(1.4) Φ ≤ 0 on F0 ∩ Bρ, Φ ≤ −α on F0 ∩ Sρand

(1.5) Φ ≥ 0 on F⊥0 ∩ Bρ, Φ ≥ α on F⊥0 ∩ Sρ.

We shall denote the inner product of x and y in Rm by x ·y and we set |x| := (x ·x)1/2.For a symmetric matrix B we denote the Morse index of the quadratic form corresponding

to B by M−(B).

Theorem 1.2 Suppose Φ ∈ C1(Rm,R) satisfies the local linking conditions (1.4) and(1.5) for some F0 ⊂ Rm. Then Φ has a critical point x with |Φ(x)| ≥ α in each of the

following two cases:(i) There exists R > 0 such that Φ < 0 in Rm \BR;

(ii) Φ(x) = 12Bx ·x+ψ(x), where ψ′(x) = o(|x|) as |x| → ∞, B is a symmetric invertible

matrix and M−(B) > dimF0.

Proof We first consider case (ii) which is more difficult. If there exists a critical point xwith Φ(x) ≤ −α, we are done. If there is no such point, then there exists a pseudogradient

vector field V whose domain contains Φ−α, and since |Φ′(x)| is bounded away from 0 as|x| is large, |Φ′(x)| ≥ δ (where δ > 0) whenever x ∈ Φ−α. Hence the Cauchy problem

dt= −V (γ(x, t)), γ(x, 0) = x

has a solution for all x ∈ Φ−α, t ≥ 0 and

(1.6) Φ(γ(x, t)) = Φ(x) +

∫ t

0

d

dsΦ(γ(x, s)) ds ≤ −α −

∫ t

0

|Φ′(γ(x, s))|2 ds ≤ −δ2t.

6

Page 7: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Choose R > 0 such that |ψ′(x)| ≤ 12λ0|x| for all |x| ≥ R, where λ0 := inf |λj| :

λj is an eigenvalue of B. Let Rm = F+ ⊕ F−, with F± respectively being the posi-

tive and the negative space of B. For x ∈ Rm write x = x+ + x−, x± ∈ F±. By (1.6) andthe form of Φ, |γ(x, T )−| ≥ R for any x ∈ F0 ∩ Sρ provided T is large enough.

Let Sn and Dn+1 be the unit sphere and the unit closed ball in Rn+1. Recall that aspace X is called l-connected if any mapping from Sn to X, 0 ≤ n ≤ l, can be extended

to a mapping from Dn+1 to X (cf. [84], Section 1.8). We want to show that the setΦ−α ∩ x ∈ Rm : |x−| ≥ R is (k − 1)-connected if k < M−(B), possibly after choosing a

larger R. This will imply in particular that any homeomorphic image of Sk−1 containedin this set is contractible there. Let r(x, t) := (1 − t)x+ + x−, 0 ≤ t ≤ 1. Then r is a

strong deformation retraction of Φ−α∩x ∈ Rm : |x−| ≥ R onto F−\BR. To see this, weonly need to verify that r(x, t) ∈ Φ−α for all x, t. Suppose first Bx+ · x+ ≤ −1

2Bx− · x−.

Then |x+| ≤ C|x−| for some C; thus ψ((1− t)x+ + x−) = o(|x−|2) as |x−| → ∞ and

Φ(r(x, t)) =1

2(1− t)2Bx+ · x+ +

1

2Bx− · x− + ψ((1− t)x+ + x−)

≤ 1

4Bx− · x− + ψ((1− t)x+ + x−) ≤ −α

if R is large enough. Let now Bx+ · x+ ≥ −12Bx− · x−; then |x−| ≤ D|x+| and

d

dtΦ(r(x, t)) = −Φ′(r(x, t)) · x+ = −Bx+ · x+ − ψ′((1− t)x+ + x−) · x+ ≤ 0,

again provided R is large enough. Hence in this case Φ(r(x, t)) ≤ Φ(x) ≤ −α. Since

F− \ BR is homeomorphic to S l−1 × [R,∞), where l := M−(B), F− \ BR is (k − 1)-connected for any k < l. It follows that so is the set Φ−α ∩ x ∈ Rm : |x−| ≥ R.

The set γ(x, T ) : x ∈ F0 ∩ Sρ is contained in Φ−α ∩ x ∈ Rm : |x−| ≥ R andhomeomorphic to Sk−1, k < M−(B). Hence it can be contracted to a point x∗ in Φ−α ∩x ∈ Rm : |x−| ≥ R. Denote this contraction by γ0, let D0 := F0 ∩ Bρ, D := D0 × [0, 1]and define a mapping f : ∂D → Rm by setting

(1.7) f(x0, s) :=

x0, s = 0, x0 ∈ D0

γ(x0, 2sT ), 0 ≤ s ≤ 12, x0 ∈ ∂D0 = F0 ∩ Sρ

γ0(γ(x0, T ), 2s− 1), 12≤ s ≤ 1, x0 ∈ ∂D0

x∗, s = 1, x0 ∈ D0.

It is clear from the construction that Φ(f(x0, s)) ≤ 0 whenever (x0, s) ∈ ∂D and < 0 ifx0 ∈ ∂D. Let

(1.8) Γ := g ∈ C(D,Rm) : g|∂D = fand

(1.9) c := infg∈Γ

max(x0,s)∈D

Φ(g(x0, s)).

7

Page 8: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

We shall show that f(∂D) links F⊥0 ∩Sρ in the sense that if g ∈ Γ, then g(x0, s) ∈ F⊥0 ∩Sρfor some (x0, s) ∈ D. Assuming this, we see that the maximum in (1.9) is always ≥ α,

and hence c ≥ α. We claim c is a critical value. Otherwise Kc = ∅, so (1.1) holds forg with N = ∅, ε ∈ (0, c) because Φ satisfies the (PS)-condition as a consequence of

(ii). Let ε < ε and η be as in Lemma 1.1. Let g ∈ Γ be such that g(x0, s) ∈ Φc+ε forall (x0, s) ∈ D. Since Φ(f(x0, s)) ≤ 0 if (x0, s) ∈ ∂D and η(x, 1) = x for x ∈ Φ0, the

mapping (x0, s) 7→ η(g(x0, s), 1) is in Γ. But this is impossible according to the definitionof c because Φ(η(g(x0, s), 1)) ≤ c− ε for all (x0, s) ∈ D.

It remains to show that f(∂D) links F⊥0 ∩ Sρ. Write x = x0 + x⊥0 ∈ F0 ⊕ F⊥0 ,D⊥0 = F⊥0 ∩ Bρ. For g ∈ Γ we consider the map

G : (D0 ⊕D⊥0 )× [0, 1]→ Rm, G(x, s) = x⊥0 − g(x0, s).

If there is no linking, then G(x, s) 6= 0 for some g ∈ Γ and all x0 ∈ D0, x⊥0 ∈ ∂D⊥0 ,0 ≤ s ≤ 1. For x0 ∈ ∂D0 we have g(x0, s) = f(x0, s), hence Φ(g(x0, s)) < 0 and

G(x, s) 6= 0 (because Φ(x⊥0 ) ≥ 0). It follows that G(x, s) 6= 0 when x ∈ ∂(D0 ⊕D⊥0 ) andG is an admissible homotopy for Brouwer’s degree. Hence

deg(G(., 0), D0 ⊕D⊥0 , 0) = deg(G(., 1), D0 ⊕D⊥0 , 0).

Since G(x, 0) = x⊥0 − g(x0, 0) = x⊥0 − f(x0, 0) = x⊥0 − x0, the degree on the left-handside above is (−1)dimF0. On the other hand, f(x0, 1) = x∗, where x∗ is a point outside

D0⊕D⊥0 ; hence G(x, 1) 6= 0 for any x ∈ D0⊕D⊥0 and the degree is 0. This contradictioncompletes the proof of (ii).

In case (i) the argument is similar but simpler. Suppose there is no critical point x withΦ(x) ≤ −α. Since Φ ≤ 0 in Rm \BR, |γ(x0, T )| ≥ R for some T > 0 and all x0 ∈ F0 ∩ Sρ.It is obvious that the set γ(x0, T ) : x0 ∈ F0 ∩ Sρ (which is homeomorphic to a sphereof dimension ≤ m − 2) can be contracted to a point in Rm \ BR. Now we can proceed

as above. Note only that Φ satisfies the (PS)c-condition because any (PS)c-sequence lieseventually in BR. 2

We shall need the following extension of Theorem 1.2:

Theorem 1.3 Suppose Φ ∈ C1(Rm,R) satisfies the local linking conditions (1.4) and

(1.5) for some F0 ⊂ Rm. If there exist subspaces F ⊃ F ⊃ F0, F 6= F , and R > 0 suchthat Φ < 0 on F \ BR and Φ|F has no critical point x ∈ Φ−α, then Φ′(x) = 0 for some x

with α ≤ Φ(x) ≤ maxx∈F∩BR+1Φ(x).

Proof This time we obtain γ by solving the Cauchy problem

dt= −χ(γ(x, t))V (γ(x, t)), γ(x, 0) = x,

8

Page 9: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

where χ ∈ C∞(Rm, [0, 1]) is such that χ = 1 on BR, χ = 0 on Rm\BR+1 and V : F∩Φ−α →F is a pseudogradient vector field for Φ|F . Now we proceed as in the proof of case (i) above

and obtain T such that γ(x0, T ) ∈ F ∩ (BR+1 \BR) whenever x0 ∈ F0 ∩ Sρ. This set canbe contracted to a point in F ∩ (BR+1 \BR), hence we obtain a map f ∈ C(∂D, F ∩ BR+1)

as in (1.7). Since there exists g ∈ Γ ∩ C(D, F ∩ BR+1) where Γ is as in (1.8) it followsthat

c ≤ max(x0,s)∈D

Φ(g(x0, s)) ≤ maxx∈F∩BR+1

Φ(x).

2

An infinite-dimensional version of the linking theorems (in a setting which correspondsto Theorems 1.2 and 1.3) may be found in [62]. However, we shall only make use of the

finite-dimensional versions stated above.

If the functional Φ is invariant with respect to a representation of some symmetrygroup, then Φ usually has multiple critical points. In order to exploit such symmetries,

we introduce index theories.

Let E be a Hilbert space and

(1.10) Σ := A = C ∩O ⊂ E : C is closed, O is open and − A = A.

Intersections of an open and a closed set (of a topological space) are called locally closed.

Thus Σ consists of the locally closed symmetric subsets of E. Let A ∈ Σ, A 6= ∅. Thegenus of A, denoted γ(A), is the smallest integer k such that there exists an odd mapping

f ∈ C(A,Rk\0). If such a mapping does not exist for any k, then γ(A) := +∞. Finally,γ(∅) = 0. Equivalently, γ(A) = 1 if A 6= ∅ and if there exists an odd map A→ +1,−1;γ(A) ≤ k if A can be covered by k subsets A1, . . . , Ak ∈ Σ such that γ(Aj) ≤ 1.

Proposition 1.4 The two definitions of genus given above are equivalent for A ∈ Σ.

Proof If f : A → Rk \ 0 is as in the first definition, then the sets Aj := x ∈ A :

fj(x) 6= 0, j = 1, . . . , k, cover A, are open in A and −Aj = Aj, hence Aj ∈ Σ. The mapfj/|fj| : Aj → +1,−1 shows that γ(Aj) ≤ 1.

Suppose γ(A) ≤ k in the sense of the second definition. Since A ∩ Aj ∈ Σ, we mayassume Aj ⊂ A, A = C ∩O, Aj = Cj ∩Oj and Cj ⊂ C, Oj ⊂ O, where C, Cj, O, Oj are

as in the definition of Σ. If fj : Aj → +1,−1 is odd, we may extend it to a continuousmap fj : Oj → R. This is a consequence of Tietze’s theorem because Aj is a closed subset

of Oj. Replacing f(x) by 12(f(x)− f(−x)) we may assume that the extension is also odd.

Let πj : A → [0, 1], j = 1, . . . , k, be a partition of unity subordinated to the covering

O1, . . . , Ok of A. Replacing πj(x) by 12(πj(x) + πj(−x)) we may assume that all πj are

even. Now the map

f : A→ Rk, f(x) = (π1(x)f1(x), . . . , πk(x)fk(x))

9

Page 10: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

is well defined, continuous, odd, and satisfies f(A) ⊂ Rk \ 0. 2

The above definitions of genus do not need to coincide for arbitrary subsets A = −Awhich are not locally closed.

Proposition 1.5 Let A,B ∈ Σ.

(i) If there exists an odd mapping g ∈ C(A,B), then γ(A) ≤ γ(B).

(ii) γ(A ∪ B) ≤ γ(A) + γ(B).

(iii) There exists an open neighbourhood N ∈ Σ of A such that γ(A) = γ(N).

(iv) If A is compact and 0 /∈ A, then γ(A) <∞.

(v) If U is an open bounded neighbourhood of 0 in Rl such that U ∈ Σ, then γ(∂U) = l.

In particular, γ(S l−1) = l, where S l−1 is the unit sphere in Rl.

(vi) If X is a subspace of codimension m in E and γ(A) > m, then A ∩X 6= ∅.

(vii) If 0 /∈ A and i(A) ≥ 2, then A is an infinite set.

A proof of this classical result may be found e. g. in [80, 85] if Σ contains only closedsets. This restriction is however not needed; see Proposition 1.7 below.

Let G be a compact topological group. A representation T of G in a Hilbert space E

is a family Tgg∈G of bounded linear operators Tg : E → E such that Te = id (where eis the unit element of G and id the identity mapping), Tg1g2 = Tg1Tg2 and the mappping

(g, x) 7→ Tgx is continuous. T is an isometric representation if each Tg is an isometry.

A set A ⊂ E is called T -invariant if TgA = A for all g ∈ G. When there is no risk ofambiguity we shall say A is G-invariant or simply invariant. The set

O(x) := Tgx : g ∈ G

will be called the orbit of x and

EG := x ∈ E : Tgx = x for all g ∈ G

the set of fixed points of the representation T . Obviously, EG is a closed subspace of Eand O(x) = x if and only if x ∈ EG.

Let

(1.11) Σ := A ⊂ E : A is locally closed and TgA = A for all g ∈ G.

Note that the definition (1.11) of Σ coincides with (1.10) if G = Z/2 ≡ 1,−1 and

T±1x = ±x. A mapping f : E → R is said to be T -invariant (or simply invariant) if

10

Page 11: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

f(Tgx) = x for all g ∈ G and x ∈ E. If T and S are two (possibly different) representationsof G in E and F , then a mapping f : E → F is equivariant with respect to T and S (or

equivariant) if f(Tgx) = Sgf(x) for all g ∈ G, x ∈ E. Finally, if f : E → F , we set

(1.12) fG(x) :=

G

Sg−1f(Tgx) dg,

where the integration is performed with respect to the normalized Haar measure. It iseasy to see that fG is equivariant. As a special case, for G = Z/2 acting via the antipodal

map on E and F we have fG(x) = 12(f(x)− f(−x)), so fG is odd. If G acts trivially on

F (i.e., S±1x = x) we obtain fG(x) = 12(f(x) + f(−x)), so fG is even.

If Φ ∈ C1(E,R) is invariant with respect to an isometric representation T of G, thenit is easy to see that Φ′(Tgx) = TgΦ

′(x) for all x ∈ E, g ∈ G. Hence x is a critical point

of Φ if and only if so are all y ∈ O(x). The set O(x) will be called a critical orbit of Φ.In what follows we restrict our attention to isometric representations of G = Z/p,

where p ≥ 2 is a prime, and G = S1 = R/2πZ. If G = S1, we do not distinguish betweenθ ∈ R and the corresponding element of G, and we may also identify this element with

eiθ. The same applies for G = Z/p ⊂ S1, where we identify the elements of G with roots

of unity, again represented as eiθ.Next we define an index i : Σ → N0 ∪ ∞ for G = S1 and G = Z/p, p a prime

number. In the case p = 2 we recover the genus. For A ∈ Σ, A 6= ∅, we define i(A) = 1if there exists a continuous map f : A → G ⊂ C \ 0 such that f(Tθx) = einθf(x) for

some n ∈ N and all x, θ (n/p /∈ N if G = Z/p). And i(A) ≤ k if A can be covered by ksets A1, . . . , Ak ∈ Σ such that i(Aj) ≤ 1. If such a covering does not exist for any k, then

i(A) := +∞. Finally, we set i(∅) := 0. We have a version of Proposition 1.4 for G = S1.

Proposition 1.6 If G = S1, then i(A) is the smallest integer k for which there exists a

mapping f ∈ C(A,Ck \ 0) such that f(Tθx) = einθf(x) for some n ∈ N and all x, θ.

The proof is similar to that of Proposition 1.4. Note only that (1.12) needs to be usedand if fj(Tθx) = einjθfj(x), then f(x) = (f1(x)n/n1 , . . . , fk(x)n/nk) where n is the least

common multiple of n1, . . . , nk. The corresponding version for G = Z/p, p an odd prime,requires spaces lying between Ck \ 0 and Ck+1 \ 0; see [10, Proposition 2.9].

The above definition is due to Benci [17, 18] in the case G = S1 and to Krasnoselski [56]for G = Z/p. Benci used in fact mappings f ∈ C(A,Ck \ 0) as in Proposition 1.6. Let

us also remark that a different, cohomological index, has been introduced by Fadell andRabinowitz [36] for G = Z/2 and G = S1, and by Bartsch [10, Example 4.5] for G = Z/p.While the geometrical indexes of Krasnoselski and Benci are much more elementary, thecohomological indexes have some additional properties (which will not be needed here).

Since we only consider isometric representations, it is easy to see that the orthogonal

complement E := (EG)⊥ is invariant. In order to formulate the properties of the index

11

Page 12: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

for G = S1 and G = Z/p we set

dG := 1 + dim G =

1 for G = Z/p;2 for G = S1.

Proposition 1.7 Suppose G = S1 or G = Z/p, where p is a prime, and let A,B ∈ Σ.

(i) If there exists an equivariant mapping g ∈ C(A,B), then i(A) ≤ i(B).

(ii) i(A ∪B) ≤ i(A) + i(B).

(iii) There exists an open neighbourhood N ∈ Σ of A such that i(A) = i(N).

(iv) If A is compact and A ∩ EG = ∅, then i(A) <∞.

(v) If U is an open bounded invariant neighbourhood of 0 in a finite-dimensional invari-

ant subspace X of E, then i(∂U) = 1dG

dimX.

(vi) If X is an invariant subspace of E with finite codimension and if i(A) > 1dG

codimEX,

then A ∩ (EG ⊕X) 6= ∅.

(vii) If A ∩ EG 6= ∅, then i(A) = +∞. If A ∩ EG = ∅ and i(A) ≥ 2, then A containsinfinitely many orbits.

Proof (i) Let i(B) = k < ∞ (otherwise there is nothing to prove) and B1, . . . , Bk be

a covering of B as in the definition of the index i(B). Then g−1(B1), . . . , g−1(Bk) is acovering of A as in the definition of i(A), hence i(A) ≤ k.

(ii), (iii) and (iv) are obvious.

(v) It follows easily from Proposition 1.4 or 1.6 that i(∂U) ≤ 1dG

dimX if G = Z/2or G = S1, respectively. In the Z/p-case for p ≥ 3 we may identify X with Cl and take

the covering Aj := z ∈ Cl : zj 6= 0, p arg(zj) 6= 0 mod 2π, Bj := z ∈ Cl : zj 6=0, p arg(zj) 6= π mod 2π, j = 1, . . . , k, of Cl \ 0 in order to see that i(Cl \ 0) ≤2l = dimX.

The reverse inequality is a consequence of the Borsuk-Ulam theorem. A proof for

G = S1 may be found in [73], Theorem 5.4, and for G = Z/p in [9].

(vi) Let Y be the orthogonal complement of X in E. Then Y is invariant and dimY =codimEX. Suppose A ∩ (EG ⊕ X) = ∅ and let f(x) = PY x where PY denotes the

orthogonal projector onto Y . Then f : A → Y \ 0. If G = Z/2 or G = S1 this impliesi(A) ≤ 1

dGdim Y by Propositions 1.4, 1.6, respectively. If G = Z/p, p ≥ 3, we identify Y

with Cm and write f = (f1, . . . , fm). It follows from the Peter-Weyl theorem (see [73],

Theorem 5.1, where the case G = S1 is considered) that fj(Tθx) = einjθf(x) (nj 6≡ 0 mod

12

Page 13: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

p). Let gj(x) := fj(x)n/nj , where n is the least common multiple of n1, . . . , nk. Theng : A → Cm \ 0 and g(Tθx) = einθg(x), so i(A) ≤ i(Cm \ 0) ≤ 2m = codimEX, a

contradiction.

(vii) Suppose A ∩ EG 6= ∅ and there exists a covering A1, . . . , Ak of A as in the

definition. Then Aj ∩ EG 6= ∅ for some j. For each x ∈ Aj ∩ EG we have f(Tθx) = f(x).So if fj(Tθx) = einθfj(x), with n as before, then fj(x) = 0. Thus there is no mapping

fj : Aj → G ⊂ C \ 0 as required in the definition of index, hence i(A) = +∞. IfA∩EG = ∅ and A consists of k orbits O(x1), . . . ,O(xk), then we let nj ≥ 1 be the largest

integer such that 2π/nj ∈ G and T2π/njxj = xj. If G = Z/p then all nj = 1. We definef : A→ G by setting f(Tθxj) = einθ, where n is the least common multiple of n1, . . . , nk.

2

It is easy to prove an equivariant version of the deformation lemma 1.1 for invariant

functionals Φ : E → R. One simply observes that if V is a pseudo-gradient vector fieldfor Φ then

(1.13) VG(x) :=

G

T−1g V (Tgx) dg

is an equivariant pseudo-gradient vector field for Φ. Integrating VG as in the proof of

Lemma 1.1 yields an equivariant deformation η.

Theorem 1.8 Suppose f ∈ C1(Sn−1,R) is invariant with respect to a representation ofZ/p in Rn without nontrivial fixed points. Then f has at least n Z/p-orbits of critical

points.

Theorem 1.9 Suppose f ∈ C1(S2n−1,R) is invariant with respect to a representation ofS1 in R2n without nontrivial fixed points. Then f has at least n S1-orbits of critical points.

Proof (outline) The proofs of Theorems 1.8 and 1.9 are standard. Set M = SdG·n−1 and

suppose f has finitely many critical orbits Oj = O(xj), j = 1, . . . , k. We may assumethat the critical values cj = f(xj) are ordered: c1 ≤ · · · ≤ ck. Then using the properties

of the index and an equivariant deformation lemma for functionals defined on manifoldsone sees that i(f cj) ≤ j. The result follows from i(M) = n.

We present a simpler proof which works if f ′ is locally Lipschitz continuous. Let η bethe negative gradient flow of f on M and consider the sets

Aj := x ∈M : η(x, t)→ Oj as t→∞.

and

Bj :=

j⋃

i=1

Ai for j = 0, . . . , k.

13

Page 14: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Then B0 = ∅, Bk = M , and it is not difficult to see that Bj−1 is an open subset of Bj.Consequently all Aj are locally closed. Using the flow η one constructs an equivariant

map fj : Aj → Oj. This implies i(Aj) ≤ i(Oj) = 1, and therefore n = i(M) = i(Bk) ≤ k.This proof does not need the equivariant deformation lemma, and it produces directly a

covering of M as in the definition of the index. 2

Remark 1.10 One can define index theories satisfying properties 1.7 (i)-(iv), (vii) for

arbitrary compact Lie groups G. However, properties 1.7 (v), (vi) which are important forapplications and computations, cannot be extended in general, except for a very restricted

class of groups. This has been investigated in detail in [10]. In certain applications therepresentation of G in E is of a special form which allows to obtain similar results as

above. In order to formulate this, call a finite-dimensional representation space V ∼= Rnof the compact Lie group G admissible if every equivariant map O → V k−1, O ⊂ V k

a bounded open and invariant neighbourhood of 0 in V k, has a zero on ∂O. Clearlythe antipodal action of Z/2 on R is admissible as are the nontrivial representations of

Z/p or S1 in R2. Let E =⊕∞

j=1 Ej be the Hilbert space sum of the finite-dimensionalHilbert spaces Ej such that each Ej is isomorphic to V as a representation space of G.

For instance, E = L2(S1, V ) with the representation of G given by (Tgx)(t) = Tg(x(t))has this property. The same is true for subspaces like H1(S1, V ) or H1/2(S1, V ). For an

invariant, locally closed set A ⊂ E let i(A) = 1 if A 6= ∅ and there exists a continuous

equivariant map A→ SV = v ∈ V : ‖v‖ = 1. And let i(A) ≤ k if A can be covered byA1, . . . , Ak ∈ Σ with i(A) ≤ 1. Proposition 1.7 can be extended to this index theory. See

[6, 16] for applications to Hamiltonian systems.

1.2 Critical point theory for strongly indefinite functionals

As will be explained in Section 2.1, functionals naturally corresponding to Hamiltonian

systems are strongly indefinite. This means that they are of the form Φ(z) = 12〈Lz, z〉 −

ψ(z) where L : E → E is a selfadjoint Fredholm operator with negative and positive

eigenspace both infinite-dimensional, and the same is true for the Hessian Φ′′(z) of acritical point z of Φ. In order to study such functionals it will be convenient to use a

variant of the Palais-Smale condition that allows a reduction to the finite-dimensional

case and leads to simpler proofs. We shall also present two useful critical point theoremswhich apply when the Palais-Smale condition does not hold. These will be needed for the

existence of homoclinic solutions.First we introduce certain sequences of finite-dimensional subspaces and replace the

Palais-Smale condition by another one which is adapted to these sequences.Let Enn≥1 be a sequence of finite-dimensional subspaces such that En ⊂ En+1 for

14

Page 15: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

all n and

E =

∞⋃

n=1

En.

Let Pn : E → En denote the orthogonal projection. Then xj is called a (PS)∗-sequence

for Φ (with respect to En) if Φ(xj) is bounded, each xj ∈ Enj for some nj, nj → ∞and PnjΦ

′(xj) → 0 as j → ∞. Φ is said to satisfy the (PS)∗-condition if each (PS)∗-sequence has a convergent subsequence. It is easy to see that Kc is compact for each cif (PS)∗ holds. Indeed, let xj ∈ Kc, then we can find nj ≥ j such that ‖yj − xj‖ ≤ 1/j,

Φ(yj) → c and PnjΦ′(yj) → 0, where yj := Pnjxj. Hence yj, and therefore also xj,

has a convergent subsequence. We shall repeatedly use the notation

Φn := Φ|En and An := A ∩ En.

Observe that Φ′n(x) = PnΦ′(x) for all x ∈ En.

The condition (PS)∗ (in a slightly different form) has been introduced independently

by Bahri and Berestycki [7, 8], and Li and Liu [60].

Lemma 1.11 If Φ satisfies (PS)∗ and N is a neighbourhood of Kc, then there exist

ε, δ > 0 and n0 ≥ 1 such that ‖Φ′n(x)‖ ≥ δ whenever d(x, E \N) ≤ δ, |Φ(x)− c| ≤ ε andn ≥ n0.

Proof If the conclusion is false, then we find a sequence xj such that xj ∈ Enj for somenj ≥ j, d(xj, E \N)→ 0, Φnj (xj)→ c and Φ′nj (xj)→ 0. Hence xj is a (PS)∗-sequence.

Passing to a subsequence, xj → x ∈ Kc. However, since Kc is compact, the sequence xjis bounded away from Kc and therefore x /∈ Kc, a contradiction. 2

Next we introduce the notion of limit index in order to deal with symmetric functionals.

As in Section 1.1 we consider the groups G = Z/p, where p is a prime, or G = S1 and their

isometric representations in E. The group Z/2 always acts via the antipodal map (i.e.,T±1x = ±x) so that obviously EG = 0. The reason for going beyond the usual index is

that we need to distinguish between certain infinite-dimensional sets having i(A) =∞; inparticular, we need to compare different spheres of infinite dimension and codimension.

Let En be a sequence of subspaces as above and suppose in addition that each En isG-invariant and EG ⊂ En for some n. Let dn be a sequence of integers and

E := En, dn∞n=1.

The limit index of A ∈ Σ with respect to E , iE(A), is defined by

iE(A) := lim supn→∞

(i(An)− dn).

15

Page 16: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Clearly iE(A) = ∞ if A ∩ EG 6= ∅. The limit index, in a somewhat different form, hasbeen introduced by Y. Q. Li [63], see also [92]. A special case is the limit genus, γE(A).

We note that iE(A) can take the values +∞ or −∞ and if En = E, dn = 0 for all n, theniE(A) = i(A), and similarly for the genus.

Remark 1.12 The limit index is patterned on the notion of limit relative categoryintroduced by Fournier et al. in [42]. Recall that if Y is a closed subset of X, then a

closed set A ⊂ X is said to be of category k in X relative to Y , denoted catX,Y (A) = k,if k is the least integer such that there exist closed sets A0, . . . , Ak ⊂ X, A0 ⊃ Y ,

which cover A, all Aj, 1 ≤ j ≤ k, are contractible in X and there exists a deformationh : A0 × [0, 1] → X with h(A0, 1) ⊂ Y and h(Y, t) ⊂ Y for all t ∈ [0, 1]. If Y = ∅(and A0 = ∅), then catX(A) = catX,∅(A) is the usual Lusternik-Schnirelman category ofA in X. For X ⊂ E and using the above notation for subsets of E, the limit relative

category cat∞X,Y (A) is by definition equal to lim supn→∞ catXn,Yn(An). Note that unlikefor the limit index, the limit category is necessarily a nonnegative integer. Note also that

if D is the unit closed ball and S its boundary in an infinite-dimensional Hilbert space,then catD,S(D) = catS(S) = 0 while cat∞D,S(D) = cat∞S (S) = 1.

Below we formulate some properties of iE which automatically hold for γE . As beforeE is the orthogonal complement of EG. It follows from the invariance of En that the

dimension of En = En ∩ E is even except when G = Z/2. Recall the notation dG =1 + dim G.

Proposition 1.13 Let A,B ∈ Σ.

(i) If for almost all n there exists an equivariant mapping gn ∈ C(An, Bn), then iE(A) ≤iE(B).

(ii) iE(A ∪B) ≤ iE(A) + i(B) if iE(A) 6= −∞.

(iii) Let l ∈ Z, R > 0. If Y is an invariant subspace of E such that dimYn = (dn + l)dGfor almost all n, then iE(Y ∩ SR) = l.

(iv) Let m ∈ Z. If X is an invariant subspace of E such that codimEnXn = (dn +m)dG

for almost all n and if iE(A) > m, then A ∩ (EG ⊕X) 6= ∅.

Proof (i) It follows from (i) of Proposition 1.7 that i(An)− dn ≤ i(Bn)− dn. So passing

to the limit as n→∞ we obtain the conclusion.

(ii) i(An ∪Bn)− dn ≤ i(An)− dn + i(Bn) ≤ (i(An)− dn) + i(B). Now we can pass to

the limit again.

(iii) This follows from (v) of Proposition 1.7.

16

Page 17: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

(iv) There exists a number n such that EG ⊂ En, codimEnXn = (dn + m)dG and

i(An) > dn +m. So by (vi) of Proposition 1.7, ∅ 6= An ∩ (EG⊕Xn) ⊂ A∩ (EG⊕X). 2

Recall that if x is a critical point of an invariant functional Φ, then so are all y ∈ O(x).

We have the following results concerning the existence of critical orbits.

Theorem 1.14 Suppose that Φ ∈ C1(E,R) is G-invariant, satisfies (PS)* and Φ(0) = 0.Moreover, suppose there exist numbers ρ > 0, α < β < 0, integers m < l, and invariant

subspaces X, Y ⊂ E such that:

(i) EG ⊂ En for almost all n;

(ii) codimEnXn = (dn +m)dG and dimYn = (dn + l)dG for almost all n;

(iii) Φ|Y ∩Sρ ≤ β;

(iv) Φ|EG⊕X ≥ α and Φ|EG ≥ 0.

Then Φ has at least l −m distinct critical orbits O(xj) such that O(xj) ∩ EG = ∅. The

corresponding critical values can be characterized as

cj = infiE(A)≥j

supx∈A

Φ(x), m+ 1 ≤ j ≤ l,

and are contained in the interval [α, β].

Proof It is clear that A ∈ Σ : i(A) ≥ j + 1 ⊂ A ∈ Σ : i(A) ≥ j, hence cm+1 ≤cm+2 ≤ . . . ≤ cl. According to (iii) of Proposition 1.13, iE(Y ∩ Sρ) ≥ l, hence by (iii),

cl ≤ β. Suppose iE(A) ≥ m+ 1. Then A ∩ (EG ⊕X) 6= ∅ by (iv) of Proposition 1.13 andit follows from (iv) that cm+1 ≥ α. Moreover, (iv) implies Kcj ∩ EG = ∅.

Suppose c := cj = . . . = cj+p for some p ≥ 0. The proof will be complete if we canshow that i(Kc) ≥ p + 1 (because either all cj are distinct and Kcj 6= ∅, or i(Kcj ) ≥ 2

for some j and Kcj contains infinitely many orbits according to (vii) of Proposition 1.7).By (iii) of Proposition 1.7 there exists a neighbourhood N ∈ Σ such that i(N) = i(Kc),

and for this N we may find ε, δ > 0 and n0 ≥ 1 such that the conclusion of Lemma 1.11holds. It follows from Lemma 1.1 that we can find an ε > 0 such that for each n ≥ n0

there exists a deformation ηn : En × [0, 1]→ En with ηn(Φc+εn , 1) ⊂ Φc−ε

n ∪Nn. Moreover,

using (1.13) we may assume that ηn(., t) is equivariant for each t. So by (i) and (ii) ofProposition 1.13 and the definition of c,

(1.14) j + p ≤ iE(Φc+ε) ≤ iE(Φ

c−ε ∪N) ≤ iE(Φc−ε) + i(N) < j + i(N).

Hence i(Kc) = i(N) > p. 2

Applying Theorem 1.14 to −Φ we immediately obtain the following result which will

be more convenient in our applications:

17

Page 18: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Corollary 1.15 Suppose that Φ ∈ C1(E,R) is G-invariant, satisfies (PS)* and Φ(0) = 0.Moreover, suppose there exist numbers ρ > 0, 0 < α < β, integers m < l, and invariant

subspaces X, Y ⊂ E such that:

(i) EG ⊂ En for almost all n;

(ii) codimEnXn = (dn +m)dG and dimYn = (dn + l)dG for almost all n;

(iii) Φ|Y ∩Sρ ≥ α;

(iv) Φ|EG⊕X ≤ β and Φ|EG ≤ 0.

Then Φ has at least l −m distinct critical orbits O(xj) such that O(xj) ∩ EG = ∅. The

corresponding critical values can be characterized as

cj = supiE (A)≥j

infx∈A

Φ(x), m + 1 ≤ j ≤ l,

and are contained in the interval [α, β].

Corollary 1.16 If the hypotheses of Theorem 1.14 or Corollary 1.15 are satisfied withl ∈ Z fixed and m ∈ Z arbitrarily small, then Φ has infinitely many geometrically distinct

critical orbits O(xj) such that O(xj) ∩ EG = ∅. Moreover, cj → −∞ in Theorem 1.14and cj →∞ in Corollary 1.15 as j → −∞.

Proof It suffices to consider the case of Theorem 1.14. The value cj is defined for allj ≤ l, j ∈ Z and since the sequence cj is nondecreasing, either cj → −∞ and we are

done, or cj → c ∈ R as j → −∞. In the second case Kc is nonempty and compactaccording to (PS)∗. Let N ∈ Σ be a neighbourhood of Kc such that i(N) = i(Kc) < ∞and let ε > 0 be as in Lemma 1.1. Since c + ε ≥ cj0 for some j0 and c − ε < cj for allj ≤ l, we have (cf. (1.14))

j0 ≤ iE(Φc+ε) ≤ iE(Φ

c−ε) + i(N) = −∞,

a contradiction. 2

Remark 1.17 Proposition 1.13, Theorem 1.14 and Corollaries 1.15-1.16 are valid if

G = Z/2 and T±1x = ±x (i.e., Φ is even). For this G, i(A) is just the genus γ(A). IfG = Z/p and p ≥ 3, then l −m is necessarily an even integer.

We conclude this part with a critical point theorem which needs tools from algebraictopology.

18

Page 19: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Theorem 1.18 Let M be a compact differentiable manifold and Φ : E × M → R aC1-functional defined on the product of the Hilbert space E and M . Suppose Φ satisfies

(PS)∗, there exist numbers ρ > 0, α < β ≤ γ and subspaces W , Y , where E = W ⊕ Y ,Wn ⊂ W , Yn ⊂ Y , dimWn ≥ 1, such that:

(i) Φ|(W∩Sρ)×M ≤ α;

(ii) Φ|Y×M ≥ β;

(iii) Φ|(W∩Bρ)×M ≤ γ.

Then Φ possesses at least cupl(M) + 1 critical points.

Here cupl(M) denotes the cuplength of M with respect to singular cohomology theorywith coefficients in an arbitrary field. For a proof we refer to Fournier et al. [42]. The

argument there uses the limit relative category (see Remark 1.12) and is in the spirit ofTheorem 1.14. In particular, the numbers cj are defined by minimaxing over sets A ⊃ D

with cat∞C,D(A) ≥ j, where (C,D) := (E×M, (W ∩Sρ)×M). An important role is playedby the inequality cat∞C,D((W ∩ Bρ)×M) ≥ cupl(M) + 1. A related result can be found in

[90].

For applications to homoclinic solutions one has to deal with functionals where neither

the (PS)- nor the (PS)∗-condition holds. We present two abstract critical point theoremswhich are helpful in this case. The proofs involve again a reduction to a finite-dimensional

situation.

Theorem 1.19 Let E be a separable Hilbert space with the orthogonal decomposition E =

E+ ⊕ E−, z = z+ + z−, and suppose Φ ∈ C1(E,R) satisfies the hypotheses:

(i) Φ(z) = 12

(‖z+‖2 − ‖z−‖2) − ψ(z) where ψ ∈ C1(E,R) is bounded below, weakly

sequentially lower semicontinuous with ψ ′ : E → E weakly sequentially continuous;

(ii) Φ(0) = 0 and there are constants κ, ρ > 0 such that Φ(z) > κ for every z ∈ Sρ∩E+;

(iii) there exists e ∈ E+ with ‖e‖ = 1, and R > ρ such that Φ(z) ≤ 0 for z ∈ ∂M whereM = z = z− + ζe : z− ∈ X−, ‖z‖ ≤ R, ζ ≥ 0.

Then there exists a sequence zj in E such that Φ′(zj) → 0 and Φ(zj) → c for some

c ∈ [κ,m], where m := sup Φ(M)].

The theorem is due to Kryszewski and Szulkin [58]. Some compactness is hidden in

condition (i) where the weak topology is used. In the applications the concentration-compactness method, see [65], can sometimes be used in order to obtain an actual critical

19

Page 20: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

point. Of course, if the Palais-Smale condition holds then there exists a critical point atthe level c.

Proof (outline) Let P± : E → E± be the orthogonal projections. We choose a Hilbert

basis ekk∈ of E− and define the norm

|||u||| := max

‖P+u‖ ,

∞∑

k=1

|〈u, ek〉|2k

.

The topology induced on E by this norm will be denoted by τ . On subsets u ∈ E :

‖P−u‖ ≤ R this topology coincides with the weak× strong product topology (E−, w)×(E+, ‖ · ‖) on E. In particular, for a ‖ · ‖-bounded sequence uj in E we have uj u if

and only if uj → u with respect to ||| · |||. Given a finite-dimensional subspace F ⊂ E+,‖ · ‖-bounded subsets of E− ⊕ F are ||| · |||-precompact.

We prove the theorem arguing indirectly. Suppose there exists α > 0 with ‖Φ′(u)‖ ≥α for all u ∈ Φm

κ := u ∈ E : κ ≤ Φ(u) ≤ m. Then we construct a deformation

h : I × Φm → Φm, I = [0, 1], with the properties:

(h1) h : I × Φm → Φm is continuous with respect to the ‖ · ‖-topology on Φm, and withrespect to the τ -topology;

(h2) h(0, u) = u for all u ∈ Φm;

(h3) Φ(h(t, u)) ≤ Φ(u) for all t ∈ I, u ∈ Φmκ ;

(h4) each (t, u) ∈ I ×Φm has a τ -open neighbourhood W such that the set v− h(s, v) :

(s, v) ∈ W is contained in a finite-dimensional subspace of E;

(h5) h(1,Φm) ⊂ Φκ.

This leads to a contradiction as follows. Since M is τ -compact, by (h1) and (h4) thereexists a finite-dimensional subspace F ⊂ E containing the set v−h(s, v) : (s, v) ∈ I×M,hence h(I × (M ∩ F )) ⊂ F . Since

h(I × ∂M) ⊂ F ∩ Φκ ⊂ F \ (Sρ ∩ E+)

a standard argument using the Brouwer degree yields h(1,M ∩ F ) ∩ Sρ ∩ E+ 6= ∅. (The

sets F ∩ ∂M and F ∩ Sρ ∩ E+ link in F .) Now condition (ii) of the theorem impliesh(1,M ∩ F ) 6⊂ Φκ, contradicting (h5).

It remains to construct a deformation h as above. For each u ∈ Φmκ we choose a pseudo-

gradient vector w(u) ∈ E, that is ‖w(u)‖ ≤ 2 and 〈Φ′(u), w(u)〉 > ‖Φ′(u)‖ (this definition

differs somewhat from (1.2)). By condition (i) of the theorem there exists a τ -open

neighbourhood N(u) of u in E such that 〈Φ′(v), w(u)〉 > ‖Φ′(u)‖ for all v ∈ N(u) ∩ Φmκ .

20

Page 21: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

If Φ(u) < κ we set N(u) = v ∈ E : Φ(v) < κ. As a consequence of condition (i) of thetheorem this set is τ -open. Let (πj)j∈J be a τ -Lipschitz continuous partition of unity of

Φm subordinated to the covering N(u), u ∈ Φm. This exists because the τ -topology ismetric. Clearly, the πj : E → [0, 1] are also ‖ · ‖-Lipschitz continuous. For each j ∈ Jthere exists uj ∈ Φm with supp πj ⊂ N(uj). We set wj = w(uj) and define the vector field

f : Φm → E, f(u) :=m− κα

j∈Jπj(u)wj .

This vector field is locally Lipschitz continuous and τ -locally Lipschitz continuous. It is

also τ -locally finite-dimensional. Thus we may integrate it and obtain a flow η : [0,∞)×Φm → Φm. It is easy to see that the restriction of η to [0, 1]× Φm satisfies the properties

(h1)-(h5). 2

Remark 1.20 A sequence zj is called a Cerami sequence if Φ(zj) is bounded and(1 + ‖zj‖)Φ′(zj) → 0. This definition has been introduced by Cerami in [22]. Note in

particular that if zj is as above, then 〈Φ′(zj), zj〉 → 0 which does not need to be thecase for an (a priori unbounded) (PS)-sequence. It has been shown in [59] that under the

hypotheses of Theorem 1.19 a stronger conclusion holds: there exists a Cerami sequence

zj such that Φ(zj)→ c ∈ [κ,m].

The next result of this section deals with Z/p-invariant functionals Φ ∈ C1(E,R). Asa substitute for the (PS)- or (PS)∗-condition we introduce the concept of (PS)-attractor.

Given an interval I ⊂ R, we call a set A ⊂ E a (PS)I-attractor if for any (PS)c-sequencezj with c ∈ I, and any ε, δ > 0 one has zj ∈ Uε(A ∩ Φc+δ

c−δ) provided j is large enough.

Here Uε(F ) denotes the ε-neighbourhood of F in E.

Theorem 1.21 Let E be a separable Hilbert space with an isometric representation ofthe group G = Z/p, where p is a prime, such that E /p = 0. Let E = E+ ⊕ E−,

z = z+ +z−, be an orthogonal decomposition and E± be Z/p-invariant. Let Φ ∈ C1(E,R)be a Z/p-invariant functional satisfying the following conditions:

(i) Φ(z) = 12

(‖z+‖2 − ‖z−‖2) − ψ(z) where ψ ∈ C1(E,R) is bounded below, weaklysequentially lower semicontinuous with ψ ′ : E → E weakly sequentially continuous;

(ii) Φ(0) = 0 and there exist κ, ρ > 0 such that Φ(z) > κ for every z ∈ Sρ ∩ E+;

(iii) there exists a strictly increasing sequence of finite-dimensional Z/p-invariant sub-spaces Fn ⊂ E+ such that sup Φ(En) <∞ where En := E−⊕Fn, and an increasing

sequence of real numbers Rn > 0 with sup Φ(En \BRn) < inf Φ(Bρ);

(iv) for any compact interval I ⊂ (0,∞) there exists a (PS)I-attractor A such that

inf‖z+ − w+‖ : z, w ∈ A, z+ 6= w+ > 0.

21

Page 22: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Then Φ has an unbounded sequence cj of positive critical values.

Proof (outline) Let τ be the topology on E introduced in the proof of Theorem 1.19.

For c ∈ R we consider the set M(c) of maps g : Φc → E satisfying:

(P1) g is τ -continuous and equivariant;

(P2) g(Φa) ⊂ Φa for all a ≥ inf Φ(Bρ)− 1 where ρ is from condition (ii);

(P3) each u ∈ Φc has a τ -open neigbourhood W ⊂ E such that the set (id− g)(W ∩ Φc)is contained in a finite-dimensional linear subspace of E.

Let i be the Z/p-index from Section 1.1 and set

i0(c) := ming∈M(c)

i(g(Φc) ∩ Sρ ∩ E+) ∈ N0 ∪ ∞.

Clearly i0 is nondecreasing and i0(c) = 0 for c ≤ κ where κ is from (ii). i0 is a kind ofpseudoindex in the sense of Benci’s paper [18]. Now we define the values

ck := infc > 0 : i0(c) ≥ k .One can show that i0(c) is finite for every c ∈ R and can only change at a critical level of

Φ. In order to see the latter, given an interval [c, d] without critical values one needs toconstruct maps g ∈ M(d) with g(Φd) ⊂ Φc. Such a map can be obtained as time-1-map

of a deformation as in the proof of Theorem 1.19. Of course one has to make sure thatthe deformation is equivariant which is the case if the vector field is equivariant. This

can be easily achieved, see (1.13). Given a finite-dimensional subspace Fn ⊂ E+ fromcondition (iii) one next proves that i0(c) ≥ dimFn for any c ≥ Φ(E− ⊕ Fn). This is a

consequence of the properties of the index stated in Proposition 1.7. No extension of theindex to infinite dimensions is needed.

Details of the proof, of a slightly more general result in fact, can be found in [12] forp = 2 and in [13] for p an odd prime. 2

2 Periodic solutions

2.1 Variational setting for periodic solutions

In this section we reformulate the problem of existence of 2π-periodic solutions of the

Hamiltonian system

(2.1) z = JHz(z, t)

in terms of the existence of critical points of a suitable functional and we collect some

basic facts about this functional. When looking for periodic solutions of (2.1) we shall

always assume that the Hamiltonian H = H(z, t) satisfies the following conditions:

22

Page 23: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

(H1) H ∈ C(R2N × R,R), Hz ∈ C(R2N × R,R2N) and H(0, t) ≡ 0;

(H2) H is 2π-periodic in the t-variable;

(H3) |Hz(z, t)| ≤ c(1 + |z|s−1) for some c > 0 and s ∈ (2,∞).

We note that it causes no loss of generality to assume H(0, t) ≡ 0. Occasionally we shallneed two additional conditions:

(H4) Hzz ∈ C(R2N × R,R4N2);

(H5) |Hzz(z, t)| ≤ d(1 + |z|s−2) for some d > 0 and s ∈ (2,∞).

Clearly, (H5) implies (H3).

Let E := H1/2(S1,R2N ) be the Sobolev space of 2π-periodic R2N -valued functions

(2.2) z(t) = a0 +

∞∑

k=1

(ak cos kt+ bk sin kt), a0, ak, bk ∈ R2N

such that∑∞

k=1 k(|ak|2 + |bk|2) <∞. Then E is a Hilbert space with an inner product

(2.3) 〈z, w〉 := 2πa0 · a′0 + π∞∑

k=1

k(ak · a′k + bk · b′k),

where a′k, b′k are the Fourier coefficients of w. It is well known that the Sobolev embedding

E → Lq(S1,R2N ) is compact for any q ∈ [1,∞) (see e.g. [2]) but z ∈ E does not imply z isbounded. There is a natural action of R on Lq(S1,R2N) and E given by time translation:

(Tθz)(t) := z(t + θ) for θ, t ∈ R.

Since the functions z are 2π-periodic in t, T induces an isometric representation of G =S1 ≡ R/2πZ. In the notation of Section 1.1, we have O(z1) = O(z2) if and only if

z2(t) = z1(t+ θ) for some θ and all t ∈ R. Let

Φ(z) :=1

2

∫ 2π

0

(−Jz · z) dt−∫ 2π

0

H(z, t) dt

and

ψ(z) :=

∫ 2π

0

H(z, t) dt.

Proposition 2.1 If H satisfies (H1)-(H3), then Φ ∈ C1(E,R) and Φ′(z) = 0 if and onlyif z is a 2π-periodic solution of (2.1). Moreover, ψ ′ is completely continuous in the sense

that ψ′(zj) → ψ′(z) whenever zj z. If, in addition, H satisfies (H4) and (H5), then

Φ ∈ C2(E,R) and ψ′′(z) is a compact linear operator for each z.

23

Page 24: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Proof We only outline the argument. The details may be found e.g. in [80], AppendixB or [102], Appendix A and Lemma 2.16. Although the results in [102] concern elliptic

partial differential equations, the proofs are easy to adapt to our situation.Let s′ = s/(s− 1) be the conjugate exponent. By (H3),

Hz : Ls(S1,R2N)→ Ls′(S1,R2N)

is a continuous mapping, and using this one shows that ψ ∈ C1(E,R) and

〈ψ′(z), w〉 =

∫ 2π

0

Hz(z, t) · w dt.

Moreover, it follows by the compact embedding of E into Ls(S1,R2N) that ψ′ is completelycontinuous.

Since −Jz · w = z · Jw, the bilinear form (z, w) 7→∫ 2π

0(−Jz · w) dt is (formally)

selfadjoint. According to (2.2) and (2.3),

∫ 2π

0

(−Jz · w) dt = π∞∑

k=1

k(−Jbk · a′k + Jak · b′k),

hence this form is continuous in E and the quadratic form z 7→∫ 2π

0(−Jz · z) dt is of class

C1. Now it is easy to see that Φ′(z) = 0 if and only if z is a 2π-periodic solution of (2.1).

Moreover, by elementary regularity theory, z ∈ C1(S1,R2N).If (H4) and (H5) are satisfied, then, referring to the arguments in [80], [102] again, we

see that ψ ∈ C2(E,R) and

〈ψ′′(z)w, y〉 =

∫ 2π

0

Hzz(z, t)w · y dt.

Since ψ′ is completely continuous, ψ′′(z) is a compact linear operator. 2

Note that complete continuity of ψ′ implies weak continuity of ψ (i.e., ψ(zj) → ψ(z)whenever zj z).

Remark 2.2 If system (2.1) is autonomous, i. e. H = H(z), then Φ(Tθz) = Φ(z) for allθ ∈ R. Thus Φ is T -invariant. Two 2π-periodic solutions z1, z2 of an autonomous system

are geometrically distinct if and only if O(z1) 6= O(z2). When H = H(z), we shall writeH ′(z) instead of Hz(z).

Let z(t) = ak cos kt± Jak sin kt. Then

∫ 2π

0

(−Jz · z) dt = ±2πk|ak|2 = ±‖z‖2.

24

Page 25: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

It follows that E has the orthogonal decomposition E = E+ ⊕ E0 ⊕ E−, where

E0 = z ∈ E : z ≡ a0 ∈ R2N,

E± =

z ∈ E : z(t) =

∞∑

k=1

ak cos kt± Jak sin kt, ak ∈ R2N

and if z = z0 + z+ + z−, then∫ 2π

0

(−Jz · z) dt = ‖z+‖2 − ‖z−‖2.

Hence

(2.4) Φ(z) =1

2

∫ 2π

0

(−Jz · z) dt−∫ 2π

0

H(z, t) dt =1

2‖z+‖2 − 1

2‖z−‖2 − ψ(z).

Remark 2.3 Φ is called the action functional and the fact that z is a 2π-periodic solutionof (2.1) if and only if Φ′(z) = 0 is the least action (or the Euler-Maupertuis) principle.

However, although the solutions z are critical points (or extremals) of Φ, they can never beminima (or maxima). Indeed, let a ∈ R2N and zj = a cos jt±Ja sin jt. Then Φ(zj)→ ±∞,

so Φ is unbounded below and above. Moreover, using (2.4) and the weak continuity of ψ,it is easy to see that Φ has neither local maxima nor minima.

The first to develop a variational method for finding periodic solutions of a Hamiltonian

system as critical points of the action functional was Rabinowitz [78]. Among other he hasshown that a star-shaped compact energy surface necessarily carries a closed Hamiltonian

orbit (see Section 2.3 for a discussion of this problem).

Lemma 2.4 Suppose (H1)-(H3) are satisfied.(i) If H(z, t) = 1

2A0(t)z · z + G0(z, t) and (G0)z(z, t) = o(|z|) uniformly in t as z → 0,

then ψ′0(z) = o(‖z‖) and ψ0(z) = o(‖z‖2) as z → 0, where ψ0(z) :=∫ 2π

0G0(z, t) dt.

(ii) If H(z, t) = 12A∞(t)z ·z+G∞(z, t) and (G∞)z(z, t) = o(|z|) uniformly in t as |z| → ∞,

then ψ′∞(z) = o(‖z‖) and ψ∞(z) = o(‖z‖2) as ‖z‖ → ∞, where ψ∞(z) :=∫ 2π

0G∞(z, t) dt.

Proof Since H(0, t) = 0, ψ0(0) = ψ∞(0) = 0. It follows from (H3) that for each ε > 0there is a c1(ε) > 0 such that |(G0)z(z, t)| ≤ ε|z| + c1(ε)|z|s−1. Hence by the Sobolev

inequality,

|〈ψ′0(z), w〉| ≤∫ 2π

0

(ε|z|+ c1(ε)|z|s−1)|w| dt ≤ (ε‖z‖+ c2(ε)‖z‖s−1)‖w‖.

Taking the supremum over ‖w‖ ≤ 1 and letting z → 0 we see that ψ ′0(z) = o(‖z‖) as

z → 0. Since

ψ0(z) =

∫ 1

0

d

dsψ0(sz) ds =

∫ 1

0

〈ψ′0(sz), z〉 ds,

25

Page 26: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

ψ0(z) = o(‖z‖2).Similarly, for each ε > 0 there is a c3(ε) such that |(G∞)z(z, t)| ≤ ε|z|+ c3(ε). So

|〈ψ′∞(z), w〉| ≤ (ε‖z‖+ c4(ε))‖w‖

and ψ′∞(z) = o(‖z‖) as ‖z‖ → ∞. Since

|ψ∞(z)| =∣∣∣∣∫ 1

0

〈ψ′∞(sz), z〉 ds∣∣∣∣ ≤

1

2ε‖z‖2 + c4(ε)‖z‖,

ψ∞(z) = o(‖z‖2) as ‖z‖ → ∞. 2

Let now

(2.5) En :=

z ∈ E : z(t) =

n∑

k=1

(ak cos kt + bk sin kt)

and En := E0 ⊕ En.

Then En ⊂ En+1 for all n,

E =

∞⋃

n=1

En

and

(2.6) E0n = E0, E±n = E±n = E± ∩ En.

Moreover, ES1= E0 (so ES1 ⊂ En for all n) and each subspace En is S1-invariant (or

more precisely, T -invariant). When using limit index theories we shall have E = En, dn,where

dn · dG = 2N(1 + n) = dimE0 +1

2dim En = dimE0 + dim E+

n .

The orthogonal projection E → En will be denoted by Pn.

Proposition 2.5 Suppose (H1)-(H3) are satisfied and let zj be a sequence such thatzj ∈ Enj for some nj, nj →∞ and PnjΦ

′(zj)→ 0 as j →∞. Then zj has a convergent

subsequence in each of the following two cases:

(i) H(z, t) = 12A∞(t)z · z +G∞(z, t), where (G∞)z(z, t) = o(|z|) as |z| → ∞ and z = 0 is

the only 2π-periodic solution of the linear system

z = JA∞(t)z.

(ii) Φ(zj) is bounded above and there exist µ > max2, s− 1 and R > 0 such that

0 < µH(z, t) ≤ z ·Hz(z, t) for all |z| ≥ R.

So in particular, (PS)∗ holds if H satisfies one of the conditions above.

26

Page 27: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

It follows upon integration that (ii) implies

(2.7) H(z, t) ≥ a1|z|µ − a2

for some a1, a2 > 0. Hence H grows superquadratically and Hz superlinearly as |z| → ∞.Note also that µ ≤ s according to (H3).

Proof Let zj ∈ Enj , nj → ∞ and PnjΦ′(zj) → 0 as j → ∞. Since PnjΦ

′(zj) =

z+j − z−j − Pnjψ

′(zj) → 0 and ψ′ is completely continuous, it follows that if zj isbounded, then zj → z after passing to a subsequence. Moreover, Φ′(z) = 0. Hence it

remains to show that zj must be bounded.Suppose (i) is satisfied and let

〈B∞z, w〉 :=

∫ 2π

0

A∞(t)z · w dt, L∞z := z+ − z− −B∞z.

Then B∞ : E → E is a compact linear operator (cf. Proposition 2.1). Since L∞z = 0 ifand only if z = JA∞(t)z, L∞ is invertible and it follows that (PnL∞|En)−1 is uniformly

bounded for large n. Hence

(PnjL∞)−1PnjΦ′(zj) = zj − (PnjL∞)−1Pnjψ

′∞(zj)→ 0

and zj is bounded because ψ′∞(z) = o(‖z‖) as ‖z‖ → ∞.

Suppose now (ii) holds. Below c1, c2, . . . will denote different constants whose exactvalues are insignificant. Since zj ∈ Enj ,

(2.8)

c1‖zj‖+ c2 ≥ Φ(zj)−1

2

⟨PnjΦ

′(zj), zj⟩

=

∫ 2π

0

(1

2zj ·Hz(zj, t)−H(zj, t)

)dt

≥(µ

2− 1)∫ 2π

0

H(zj, t) dt

≥ c3‖zj‖µµ − c4,

where the last inequality follows from (2.7). Since E0 is finite dimensional, ‖z0j ‖ ≤ c5‖zj‖µ

and by (2.8),

(2.9) ‖z0j ‖ ≤ c6‖zj‖1/µ + c7.

By the Holder and Sobolev inequalities,

(2.10)

‖z+j ‖2 = 〈Φ′(zj), z+

j 〉+

∫ 2π

0

Hz(zj, t) · z+j dt

≤ c8‖z+j ‖+ c9

∫ 2π

0

|zj|s−1|z+j | dt

≤ c8‖z+j ‖+ c10‖zj‖s−1

µ ‖z+j ‖

27

Page 28: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

(here we have used that µ > s− 1) and a similar inequality holds for z−j . Hence

‖z±j ‖ ≤ c8 + c10‖zj‖s−1µ

and by (2.8),‖z±j ‖ ≤ c11 + c12‖zj‖(s−1)/µ.

This and (2.9) combined imply zj is bounded. 2

Remark 2.6 (a) If H(z, t) = 12A(t)z · z + G(z, t), where A is a symmetric 2N × 2N

matrix with periodic entries and G satisfies the superlinearity condition of Proposition2.5, it is easy to see by an elementary computation that so does H, possibly with a smaller

µ > max2, s− 1 and a larger R.

(b) In some problems the growth restriction (H3) may be removed and the conditionµ > max2, s− 1 can be replaced by µ > 2. For this purpose one introduces a modified

Hamiltonian HK such that HK(z, t) = H(z, t) for |z| ≤ K and HK(z, t) = C|z|s for|z| ≥ K + 1 and some convenient C, s. Then the modified functional satisfies (PS)∗,one can apply a suitable variational method to obtain one or more solutions which areuniformly bounded independently ofK. Hence these solutions satisfy the original equation

for K large. We shall comment on that when appropriate.

Let now H satisfy (H1)-(H5). For each fixed z we have 〈Φ′′(z)w,w〉 = ‖w+‖2 −‖w−‖2 − 〈ψ′′(z)w,w〉, and since ψ′′(z) is a compact linear operator, it is easy to see thatthe quadratic forms ±Φ′′(z) have infinite Morse index. However, it is possible to define a

certain relative index which will always be finite. Let A be a symmetric 2N×2N constantmatrix and

(2.11) 〈Lz, w〉 :=

∫ 2π

0

(−Jz − Az) · w dt.

It follows from (2.2) and (2.3) that

(2.12) 〈Lz, z〉 = −2πAa0 · a0 + π

∞∑

k=1

k

((−Jbk −

1

kAak) · ak + (Jak −

1

kAbk) · bk

).

The restriction of this quadratic form to a subspace corresponding to a fixed k ≥ 1 isrepresented by the (4N × 4N)-matrix πkTk(A), where

Tk(A) :=

(− 1kA −J

J − 1kA

).

Let M+(·) and M−(·) respectively denote the number of positive and negative eigenvalues

of a symmetric matrix (counted with their multiplicities) and let M 0(·) be the dimension

28

Page 29: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

of the nullspace of this matrix. Then M 0(Tk(A)) = 0 and M±(Tk(A)) = 2N for all klarge enough. Indeed, a simple computation shows that the matrix

(0 −JJ 0

)

has the eigenvalues ±1, each of multiplicity 2N , so by a simple perturbation argument,M±(Tk(A)) = 2N for almost all k (cf. [3], Section 12, [4], Section 2). Therefore the

following numbers are well defined and finite:

i−(A) := M+(A)−N +

∞∑

k=1

(M−(Tk(A))− 2N),(2.13)

i+(A) := M−(A)−N +∞∑

k=1

(M+(Tk(A))− 2N)(2.14)

and

i0(A) := M0(A) +

∞∑

k=1

M0(Tk(A)).(2.15)

Clearly, i±(A), i0(A) are finite and i−(A) + i+(A) + i0(A) = 0. The quantity i−(A) is a

relative Morse index in the sense that it provides a measure for the difference between thenegative parts of the quadratic forms z 7→

∫ 2π

0(−Jz −Az) · z dt and z 7→

∫ 2π

0(−Jz · z) dt.

It is easy to see that i0(A) = dimN(L) is the number of linearly independent 2π-periodic solutions of the linear system

(2.16) z = JAz.

Hence in particular i0(A) ≤ 2N . Moreover, i0(A) = 0 if and only if σ(JA) ∩ iZ = ∅.Indeed, it follows from (2.12) (or by substituting (2.2) into (2.16)) that (2.16) has anontrivial 2π-periodic solution if and only if either A is singular (so 0 ∈ σ(JA)) or

(2.17) −kJbk = Aak and kJak = Abk

for some (ak, bk) 6= (0, 0) and k ≥ 1. (2.17) is equivalent to

JA(ak − ibk) = ik(ak − ibk).

Hence ±ik ∈ σ(JA) and (2.16) has a nontrivial 2π-periodic solution if and only if σ(JA)∩iZ 6= ∅. We also see that Pn commutes with L, hence if E = E+(L) ⊕ E0(L) ⊕ E−(L)is the orthogonal decomposition corresponding to the positive, zero and negative part

of the spectrum of L and E±n (L) := E±(L) ∩ En, E0n(L) := E0(L) ∩ En, then En =

E+n (L)⊕E0

n(L)⊕E−n (L) is an orthogonal decomposition into the positive, zero and negative

part of Ln := PnL|En . Note that E0(L) = N(L) and E0n(L) = E0(L) for almost all n.

In one of our applications we shall need a slight extension of Proposition 2.5.

29

Page 30: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Corollary 2.7 Suppose H is as in (a) of Remark 2.6 and A is a constant matrix. Ifzj is a sequence such that Φ(zj) is bounded above, zj = w+

j + w0j + w−j ∈ E+

mj(L) ⊕

E0(L) ⊕ E−nj (L), mj, nj → ∞ and (P+mj

+ P 0 + P−nj)Φ′(zj) → 0 as j → ∞ (P±n and P 0

denote the orthogonal projections E → E±n (L), E → E0(L)), then zj has a convergent

subsequence.

The proof follows by inspection of the argument of (ii) in Proposition 2.5. Note in

particular that (2.8) holds with G(z, t) = H(z, t)− 12Az ·z replacing H and E0(L) replacing

E0. Moreover, since

(2.18) 〈±Lnz, z〉 ≥ ε‖z‖2 for some ε > 0 and all z ∈ E±n (L)

(ε independent of n), also (2.10) can be easily adapted.

Remark 2.8 a) If A is a nonconstant matrix with 2π-periodic entries, the definitionsof i±(A) and i0(A) no longer make sense. Therefore we need some other quantities to

measure the size of the positive and the negative part of L. Assume for simplicity that

the operator L corresponding to A is invertible and let M−n (L) be the Morse index of the

quadratic form z 7→ 〈Lz, z〉 restricted to En. Let

j−(A) := limn→∞

(M−n (L)− (1 + 2n)N)

and j+(A) := j−(−A). It can be shown that j±(A) are well defined and finite, and

j±(A) = i±(A) whenever A is a constant matrix ([57], Sections 7 and 5). See also thereferences below.

b) Morse-type indices for Hamiltonian systems have been introduced by Amann andZehnder [3, 4] and Benci [18]. In [3, 4] and [18] computational formulas for these indices are

also discussed. Our definitions of i±(A), i0(A) follow Li and Liu [61] (more precisely, theindices i±(A) as defined here differ from those in [61] by N). The number j−(A) equals the

Conley-Zehnder (or Maslov) index of the fundamental solution γ : [0, 2π]→ Sp(2N) of theequation z(t) = JA(t). Here Sp(2N) denotes the group of symplectic 2N × 2N -matrices.

Recall that a matrix C is symplectic if C tJC = J . See the books of Abbondandolo [1],

Chang [23, Section IV.1] and in particular Long [68, 69] for a comprehensive discussionof the Conley-Zehnder index.

2.2 Periodic solutions near equilibria

The first existence and multiplicity results for periodic solutions of (2.1) are concernedwith solutions near an equilibrium. The classical results of Lyapunov [72], Weinstein [98],

and Moser [76] have been very influential for the development of the theory and can beproved using the basic variational methods from Section 1.1.

30

Page 31: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

We consider the autonomous Hamiltonian system

(2.19) z = JH ′(z)

where the Hamiltonian H : R2N → R is of class C2. Since the vector field H ′ is of classC1, each initial value problem has a unique solution z = z(t) defined on some maximal

interval I. Furthermore,

(2.20)d

dtH(z(t)) = H ′(z(t)) · z(t) = −Jz(t) · z(t) = 0,

hence H(z(t)) is constant for all t ∈ I. Throughout this section we assume that H has 0 ascritical point. We first consider the case where 0 is nondegenerate. The constant function

z ≡ 0 is then an isolated, hyperbolic stationary solution of (2.19). It is well known thatperiodic orbits near 0 can only exist if JH ′′(0) has purely imaginary eigenvalues. The

Lyapunov center theorem states that if JH ′′(0) has a pair of purely imaginary eigenvalues±iω which are simple, and if no integer multiples ±ikω are eigenvalues of JH ′′(0) then

(2.19) has a one-parameter family of periodic solutions emanating from the equilibriumpoint. More precisely, let E(±iω) ⊂ R2N be the two-dimensional eigenspace associated

to ±iω and let σ ∈ ±2 be the signature of the quadratic form Q(z) = 12H ′′(0)z · z on

E(±iω). Then for each ε > 0 small enough there exists a periodic solution of (2.19) onthe energy surface H = H(0) + σε2 with period converging to 2π/ω as ε→ 0.

If ±iω is not simple, or if integer multiples are eigenvalues then there may be noperiodic solutions near 0 as elementary examples show; see [76, 27]. In order to formulate

a sufficient condition we assume that all eigenvalues of JH ′′(0) which are of the form ikω,k ∈ Z, are semisimple, i.e., their geometric and algebraic multiplicities are equal. Let

Eω ⊂ R2N be the generalized eigenspace of JH ′′(0) corresponding to the eigenvalues ofthe form ±ikω, k ∈ N. Let Q : R2N → R, Q(z) = 1

2H ′′(0)z · z, be the quadratic part of H

at 0 and let σ = σ(ω) ∈ Z be the signature of the quadratic form Q|Eω on Eω. Observethat σ is automatically an even integer.

Theorem 2.9 If σ 6= 0 then one of the following statements hold.

(i) There exists a sequence of non-constant Tk-periodic solutions zk of (2.19) which lie

on the energy surface H = H(0) with zk → 0 and Tk → 2π/ω as k → ∞. Theperiod Tk is not necessarily minimal.

(ii) For ε > 0 small enough there are at least |σ/2| non-constant periodic solutions

zεj , j = 1, . . . , |σ/2|, of (2.19) on the energy surface H = H(0) + σε2 with (notnecessarily minimal) period T εj . These solutions converge towards 0 as ε → 0.

Moreover, T εj → 2π/ω as ε→ 0.

31

Page 32: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

This theorem is due to Bartsch [11]. It generalizes the Weinstein-Moser theorem[98, 76] which corresponds to the case where Q|Eω is positive or negative definite, hence

|σ| = dim Eω. Observe that the energy surfaces H = c are not necessarily compact for cclose to H(0).

Proof (outline) We may assume H(0) = 0. The τ -periodic solutions of (2.19) correspond

to 2π-periodic solutions of

(2.21) z =τ

2πJH ′(z).

These in turn correspond to critical points of the action functional

A(z) =

∫ 2π

0

Jz(t) · z(t) dt

restricted to the surface z ∈ E : ψ(z) = 2πH(0) + λ where E = H1/2(S1,R2N )) andψ(z) =

∫ 2π

0H(z(t))dt are as in Section 2.1. The period appears as Lagrange multiplier in

this approach. After performing a Lyapunov-Schmidt reduction of the equation

A′(z) =τ

2πψ′(z)

near τ = 2π/ω and z ≡ 0, one is left with the problem of finding critical points of a

functionA0(v) = A(v + w(v))

constrained to the level set v ∈ V : ψ0(v) = λ. Here V is the kernel of the linearization

E 3 z 7→ A′′(0)z − τ

2πψ′′(0)z ∈ E,

w : V ⊃ U → V ⊥ ⊂ E is defined on a neighborhood U of 0 in V , and ψ0(v) := ψ(v+w(v)).

Thus V ∼= Eω and one checks that A0 and ψ0 are of class C1 and that ψ′′0 (0) exists. Infact:

〈ψ′′0 (0)v, w〉 =

∫ 2π

0

H ′′(0)v(t) · w(t) dt ,

hence we can apply the Morse lemma to ψ0 near 0. After a change of coordinates ψ0 looks

(in the sense of the Morse lemma) near 0 like the non-degenerate quadratic form

q : V → R, v 7→ 1

2H ′′(0)v · v.

Therefore the level surfaces ψ−10 (λ) look locally like the level surfaces q−1(λ). If q is positive

definite, hence σ = dim V = dim Eω (which is just the situation of the Weinstein-Moser

theorem), one can conclude the proof easily upon observing that the functionals A, ψ and,

32

Page 33: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

hence A0 and ψ0 are invariant under the representation of S1 = R/2πZ in E induced bythe time shifts. Moreover, ψ−1

0 (λ) ∼= q−1(λ) is diffeomorphic to the unit sphere SV of V

for λ > 0 small. By Theorem 1.9 any C1-functional SV → R which is invariant under theaction of S1 has at least 1

2dimV = 1

2dimEω S

1-orbits of critical points. If q is negative

definite then ψ−10 (λ) is diffeomorphic to the unit sphere SV of V for λ < 0 close to 0, and

one obtains |σ/2| critical orbits on these levels.

This elementary argument from S1-equivariant critical point theory does not work if qis indefinite. Instead one looks at the flow ϕλ on Σλ := ψ−1

0 (λ) which is essentially induced

by the negative gradient ofA0|Σλ. Since A0 and ψ0 are only of class C1 the gradient vectorfield is of class C0, so it may not be integrable and has to be replaced by a pseudogradient

vector field which leaves Σλ invariant for all λ. Next one observes that the hypersurfacesΣλ undergo a surgery as λ passes H(0) = 0. If 2n+ (respectively 2n−) is the maximal

dimension of a subspace of V on which q is positive (respectively negative) definite thenΣε is obtained from Σ−ε upon replacing a handle of type B2n+ × S2n−−1 by S2n+ ×B2n− .

It is this change in the topology of Σλ near 0 which forces the existence of stationaryorbits of ϕ near the origin. In order to analyze the influence of this surgery on the flow

ϕλ one has to use methods from equivariant Conley index theory and Borel cohomology.

The difference |n+ − n−| = |σ|/2 is a lower bound for the number of stationary S1-orbitsof ϕλ on Σλ if λ > 0 is small and σ · (λ−H(0)) > 0. 2

It is also possible to parametrize the nontrivial periodic orbits near an equilibrium by

their period. The following result is due to Fadell and Rabinowitz [36].

Theorem 2.10 If σ 6= 0 then one of the following statements hold.

(i) There exists a sequence of nonconstant periodic orbits zk → 0 of (2.19) with (notnecessarily minimal) period T = 2π/ω.

(ii) There exist integers k, l ≥ 0 with k + l ≥ |σ|/2, and there exists ε > 0 such thatfor each τ ∈ (T − ε, T ) (2.19) has at least k periodic orbits zτj , j = 1, . . . , k with

(not necessarily minimal) period τ . And for each τ ∈ (T, T + ε) (2.19) has at leastl periodic orbits zτj , j = k + 1, . . . , k + l with (not necessarily minimal) period τ .

Moreover, zτj → 0 as τ → T = 2π/ω.

The proof uses a cohomological index theory. The integers k, l (and thus the direction

of the bifurcating solutions with the period as parameter) are not determined by H ′′(0)

– unlike case (ii) in Theorem 2.9.Now we consider the case of a degenerate equilibrium. Suppose first that 0 is an

isolated critical point of H, so there are no stationary orbits of (2.19) in a neighbourhoodof the origin. Let iω be an eigenvalue of JH ′′(0) and let Fω ⊂ R2N be the generalized

eigenspace of JH ′′(0) corresponding to ±iω. Thus Fω ⊂ Eω does not contain generalized

33

Page 34: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

eigenvectors of JH ′′(0) corresponding to multiples ±ikω with |k| ≥ 2. Let σ1 = σ1(ω) bethe signature of the quadratic form Q|Fω. Since we allow H ′′(0) to have a nontrivial kernel

we also need the critical groups Cq(H, 0) = Hq(H0, H0 \ 0) associated to 0 ∈ R2N asa critical point of the Hamiltonian H. Here H∗ denotes the Cech (or Alexander-Spanier)

cohomology with coefficients in an arbitrary field.

Theorem 2.11 If σ1 6= 0 and Cq(H, 0) 6= 0 for some q ∈ Z, then there exists a sequencezk of nonconstant periodic orbits of (2.19) with (not necessarily minimal) period Tn such

that ‖zn‖L∞ → 0 and Tn → 2π/ω.

The result has been proved by Szulkin in [91] using Morse theoretic methods. It is

unknown whether the solutions obtained in Theorems 2.9–2.11 lie on connected branchesof periodic solutions. Continua of periodic solutions however do exist under stronger hy-

potheses when degree theoretic methods apply. We state one such result in this direction.

Theorem 2.12 If σ1 6= 0 and the local degree deg(∇H, 0) of ∇H at the isolated criticalpoint 0 is non-trivial, then there exists a connected branch of periodic solutions of (2.19)

near 0.

For a proof see the paper [27] by Dancer and Rybicki. They work in the space

W 1(S1,R2N) and apply a degree for S1-gradient maps to the bifurcation equations as-sociated to y = λH ′(y). 2π-periodic solutions y(s) of this equation correspond to 2π/λ-

periodic solutions of (2.19). The degree theory allows a classical Rabinowitz type argu-ment yielding a global continuum of solutions in R ×W 1(S1,R2N) that bifurcates from

(λ0, 0) with λ0 = ω.Observe that σ =

∑∞k=1 σk where σk = σk(Ω) = σ1(kω) is the signature of Q|Fk, Fk

the generalized eigenspace of JH ′′(0) corresponding to ±ikω. Also observe that the localdegree can be expressed in terms of the critical groups as

deg(H ′, 0) =∞∑

q=0

dimCq(H, 0).

Thus the hypotheses of Theorem 2.11 are weaker than those of Theorem 2.12. Corre-

spondingly, the conclusion is also weaker.Since the above results require only local conditions on the Hamiltonian near a station-

ary point they immediately generalize to Hamiltonian systems on a symplectic manifold(W,Ω). The last result that we state in this section deals with periodic orbits near a

manifold M of equilibria. This result can in general not be reduced to the special caseW = R2N with the standard symplectic structure Ω =

∑Nk=1 dpi ∧ dqi because the man-

ifold M need not lie in a symplectic neighbourhood chart. We therefore state it in the

general setting.

34

Page 35: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Theorem 2.13 Let (W,Ω) be a symplectic manifold and let H : W → R be a smoothHamiltonian. Suppose there exists a compact symplectic submanifold M ⊂ H−1(c) ⊂ W

which is a Bott-nondegenerate manifold of minima of H. Then there exists a sequence ofnon-constant periodic trajectories of the Hamiltonian flow associated to H which converge

to M .

The result is due to Ginzburg and Kerman [45]. It clearly applies to (2.19) where

W = R2N and Ω is as above. Compared with the Weinstein-Moser theorem where M is apoint, Theorem 2.13 does not yield periodic orbits on all energy surfaces close to M , and

neither does it yield a multiplicity result. We refer to [45] and the references therein forfurther results on periodic orbits of Hamiltonian flows near manifolds of equilibria.

2.3 Fixed energy problem

Let H ∈ C2(R2N ,R) and suppose D := z ∈ R2N : H(z) ≤ 1 is a compact subset ofR2N such that H ′(z) 6= 0 for all z ∈ S := H−1(1). Then S is a compact hypersurface of

class C2 and we may assume without loss of generality that 0 is in the interior of D. We

consider the autonomous Hamiltonian system (2.19). If z(t0) ∈ S then z(t) ∈ S for allt because H(z(t)) is constant along solutions of (2.19) (see (2.20)). Since S is compact,

z(t) exists for all t ∈ R.We will be interested in the existence of closed Hamiltonian orbits on S, i.e., the sets

Orb(z) := z(t) : t ∈ R, where z = z(t) is a periodic solution of (2.19) with z(t) ∈ S.Here we use the notation Orb(z) for closed orbits in order to distinguish them from S1-

orbits O(z) defined in Section 2.1. If H ∈ C2(R2N ,R) is another Hamiltonian such thatS = H−1(c) for some c and H ′(z) 6= 0 on S, then H ′(z) and H ′(z) are parallel and nowhere

zero on S. It follows that two solutions z and z of the corresponding Hamiltonian systemsare equivalent up to reparametrization if z(t0) = z(t0) for some t0, t0 ∈ R. In particular,

the orbits Orb(z) and Orb(z) coincide (see [80] for a detailed argument). Consequently,closed orbits depend only on S and not on the particular choice of a Hamiltonian having

the properties given above. It is also possible to define closed orbits without referringto any Hamiltonian: given a compact surface S of class C2, one may look for periodic

solutions of the system z = JN(z), where N(z) is the unit outer normal to S at z.

Throughout this section we assume that S satisfies the following condition:

(S) S is a compact hypersurface of class C2 in R2N , S bounds a starshaped neighbourhood

of the origin and all z ∈ S are transversal to S.

It follows from (S) that for each z ∈ R2N \ 0 there exists a unique α(z) > 0 such

that z/α(z) ∈ S. Let α(0) := 0 and

(2.22) H(z) := α(z)4.

35

Page 36: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Clearly α(sz) = s4α(z) for all s ≥ 0, hence H is positively homogeneous of degree 4.Moreover, S = H−1(1), H ∈ C2(R2N ,R) and (by Euler’s identities) H ′(z) ·z = 4H(z) 6= 0

whenever z 6= 0. In particular, H ′(z) 6= 0 on S.Suppose z is a periodic solution of (2.19) with the Hamiltonian H given by (2.22). If

Orb(z) ⊂ H−1(λ) then z(t) := λ−1/4z(t/√λ) is a periodic solution of (2.19) on H−1(1).

If z has minimal period 2π then z has minimal period T := 2π√λ. On the other hand,

given a periodic solution z of (2.19) on H−1(1) with minimal period T then z(t) :=(T/2π)1/2z(T t/2π) is a periodic solution of (2.19) on H−1((T/2π)2) with minimal period

2π. One easily checks that z = z and ˜z = z. Given two periodic solutions z1, z2 of (2.19) on

H−1(1) with minimal period T and having the same orbit Orb(z1) = Orb(z2) ⊂ H−1(1)

then there exists θ ∈ R with z2(t) = z1(t + θ) for all t. The corresponding solutionsz1, z2 with minimal period 2π then satisfy z2(t) = z1(t + θ) for some θ ∈ R, hence they

are not geometrically distinct in the sense of Remark 2.2. If z1, z2 have different orbitsOrb(z1),Orb(z2) ⊂ H−1(1) then z1, z2 are geometrically distinct.

We summarize the above considerations in the following

Theorem 2.14 Let S be a hypersurface satisfying (S) and let H be defined by (2.22).

A periodic solution z(t) of (2.19) on H−1(λ) yields a periodic solution λ−1/4z(t/√λ) on

S = H−1(1). Moreover, there is a one-to-one correspondence between closed orbits on S

and geometrically distinct periodic solutions of (2.19) with minimal period 2π.

We emphasize the importance of the assumption on the minimality of the period. If

z is a solution of (2.19) with minimal period T and Orb(z) ⊂ S, then z(t) covers Orb(z)k times as t goes from 0 to kT . A corresponding solution zk(t) := (kT/2π)1/2z(kT t/2π)

has minimal period 2π/k, hence zk and zm are geometrically distinct if k 6= m, yet

Orb(zk) = Orb(zm) = Orb(z).

Now we state the first main result of this section. It is due to Rabinowitz [78] and, ifS bounds a convex neighbourhood of the origin to Weinstein [99].

Theorem 2.15 Let S be a hypersurface satisfying (S). Then S contains a closed Hamil-tonian orbit.

Proof By Theorem 2.14 it suffices to show that (2.19) with H given by (2.22) has a2π-periodic solution z 6= 0. Let r be the largest and R the smallest number such that

(2.23) r ≤ |z| ≤ R for all z ∈ S.

Then

(2.24)|z|4R4≤ H(z) ≤ |z|

4

r4, for all z ∈ R2N .

36

Page 37: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

The functional

Φ(z) =1

2

∫ 2π

0

(−Jz · z) dt−∫ 2π

0

H(z) dt =1

2‖z+‖2 − 1

2‖z−‖2 − ψ(z)

is S1-invariant. Moreover, H ′(z) · z = 4H(z) > 0, hence Φ satisfies (PS)∗ according to

Proposition 2.5. Let En, En be given by (2.5) and let 2dn = 2N(1 + n), Y = E+ andX = E+

1 ⊕ E− (cf. (2.6)). Then ES1= E0 ⊂ En for all n, codimEn

Xn = 2(dn − 2N) and

dim Yn = 2(dn −N). Hence (i) and (ii) of Corollary 1.15 are satisfied, with l = −N andm = −2N . By (2.24) and Lemma 2.4, Φ|Y ∩Sρ ≥ α for some α, ρ > 0. Since

(2.25) Φ(z) ≤ 1

2‖z+‖2 − 1

2‖z−‖2 − 1

R4

∫ 2π

0

|z|4 dt

and dimE+1 < ∞, Φ(z) → −∞ as ‖z‖ → ∞, z ∈ X. Finally, Φ|E0 ≤ 0 because

H ≥ 0. It follows that also (iii) and (iv) of Corollary 1.15 hold. Hence (2.19) has at

least N geometrically distinct 2π-periodic solutions z 6= 0, and by Theorem 2.14, the

hypersurface S carries a closed Hamiltonian orbit. 2

It is not necessary to exploit the S1-symmetry in order to show the existence of one2π-periodic solution z 6= 0. However, the argument presented here will be needed below.

Remark 2.16 Since the above proof gives no information on the minimal period of theN geometrically distinct 2π-periodic solutions, we do not know whether they correspond

to distinct closed orbits on S. System (2.19) has in fact infinitely geometrically distinct2π-periodic solutions. Indeed, we may replace X = E+

1 ⊕ E1 by X = E+r ⊕ E− for any

positive integer r and use Corollary 1.16 (see also Theorem 2.19). On the other hand, if

z = (p1, . . . , pN , q1, . . . , qN) and

S =

z ∈ R2N :

N∑

j=1

1

2αj(p

2j + q2

j ) = 1

,

where α1, . . . , αN are rationally independent positive numbers, then it is easy to see that

S has exactly N distinct closed orbits.

In an answer to a conjecture of Weinstein, Viterbo [96] has generalized Theorem 2.15

to all compact hypersurfaces admitting a so-called contact structure. Subsequently hisproof has been simplified (and a more general result obtained) by Hofer and Zehnder [51].

Struwe [86], building upon the work of Hofer and Zehnder, proved that given an interval[a, b] of regular values of H such that the hypersurfaces Sc := H−1(c) ⊂ R2N , c ∈ [a, b],

are compact, the set c ∈ [a, b] : Sc carries a closed Hamiltonian orbit has full measure

b − a. It has been shown by counterexamples of Ginzburg [43] and Herman [49] that in

37

Page 38: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

general a compact hypersurface may not have any closed Hamiltonian orbit (see also [44]and the references there).

In view of Remark 2.16 it is natural to ask whether each S satisfying (S) must nec-essarily have N distinct closed Hamiltonian orbits. We shall show that this is indeed the

case under an additional geometric conditon.Denote the tangent hyperplane to S at w by Tw(S), suppose S satisfies (S) and let ρ

be the largest number such that

(2.26) Tw(S) ∩ z ∈ R2N : |z| < ρ = ∅ for all w ∈ S.

Then ρ is the minimum of the distances from Tw(S) to the origin over all w ∈ S. It followsfrom (S) that ρ is well-defined; moreover, if r is as in (2.23) and S bounds a convex set,

then ρ = r.

Theorem 2.17 Let S be a hypersurface satisfying (S) and suppose R2 < 2ρ2, where R,

ρ are as in (2.23), (2.26). Then S contains at least N distinct closed Hamiltonian orbits.

Proof It follows from the proof of Theorem 2.15 that (2.19) (with H given by (2.22))has at least N geometrically distinct 2π-periodic solutions. According to Theorem 2.14

it suffices to show that these solutions have minimal period 2π.Recall from the proof of Theorem 2.15 that l = −N and m = −2N , so invoking

Corollary 1.15 we have

cj = supiE (A)≥j

infz∈A

Φ(z), −2N + 1 ≤ j ≤ −N.

Since codimEnXn = 2(dn − 2N), it follows from (iv) of Proposition 1.13 that if iE(A) ≥

−2N + 1, then A ∩ (EG ⊕X) = A ∩ (E+1 ⊕ E0 ⊕ E−) 6= ∅. Hence

(2.27) cj ≤ supΦ(z) : z ∈ E+1 ⊕ E0 ⊕ E−.

Let z ∈ E+1 ⊕ E0 ⊕ E−. Using (2.25), the fact that ‖z+‖ = ‖z+‖2 for z+ ∈ E+

1 and theHolder inequality, we obtain

Φ(z) ≤ 1

2‖z+‖2 − 1

2‖z−‖2 − 1

R4‖z‖4

4 ≤1

2‖z+‖2

2 −1

R4‖z‖4

4

≤ 1

2‖z‖2

2 −1

R4‖z‖4

4 ≤√π

2‖z‖2

4 −1

R4‖z‖4

4 ≤πR4

8.

This and (2.27) imply

(2.28) cj ≤πR4

8.

38

Page 39: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

By the definition of ρ and the homogeneity of H,

ρ|H ′(z)| ≤ z ·H ′(z) = 4H(z) = 4 = 4H(z)3/4

whenever z ∈ S. By the homogeneity again,

(2.29) ρ|H ′(z)| ≤ 4H(z)3/4 for all z ∈ R2N .

Let now z = z(t) be a 2π-periodic solution of (2.19). Since H(z(t)) is constant,

Φ(z) =1

2

∫ 2π

0

(−Jz · z) dt−∫ 2π

0

H(z) dt(2.30)

=

∫ 2π

0

(1

2H ′(z) · z −H(z)

)dt =

∫ 2π

0

H(z) dt = 2πH(z).

Suppose z has minimal period 2π/m and write z = z + z, z ∈ E0, z ∈ E+ ⊕ E−. ByWirtinger’s inequality,

‖z‖2 ≤1

m‖ ˙z‖2,

and it follows using (2.30), (2.29) that

2πH(z) = Φ(z) =1

2

∫ 2π

0

(−J ˙z · z) dt−∫ 2π

0

H(z) dt

≤ 1

2‖ ˙z‖2‖z‖2 − 2πH(z) ≤ 1

2m‖z‖2

2 − 2πH(z)

=1

2m

∫ 2π

0

|H ′(z)|2 dt− 2πH(z) ≤ 8

mρ2

∫ 2π

0

H(z)3/2 dt− 2πH(z)

= 2π

(8H(z)3/2

mρ2−H(z)

).

Hence

H(z) ≥ m2ρ4

16

and

Φ(z) = 2πH(z) ≥ πm2ρ4

8.

Since R2 < 2ρ2,

Φ(z) >πm2R4

32.

On the other hand, if the solution z corresponds to cj, −2N + 1 ≤ j ≤ −N , then

Φ(z) ≤ πR4/8 according to (2.28). Hence m = 1 and z has minimal period 2π. 2

39

Page 40: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Theorem 2.17 is due to Ekeland and Lasry [35] (see also [33]) in the case of S boundinga compact strictly convex region and R2 < 2r2, and by Berestycki et al. [20] in the more

general case considered here. We would also like to mention a result by Girardi andMatzeu [46] showing that if S satisfies (S), then the condition R2 < 2ρ2 may be replaced

by R2 <√

3ρr in Theorem 2.17.It has been a longstanding conjecture (see, e.g., [33], p. 235) that if S bounds a

compact strictly convex set, then the minimal number of distinct closed Hamiltonianorbits such S must carry is N . Ekeland and Lassoued [34] and Szulkin [89] have shown

that S carries at least 2 such orbits if its Gaussian curvature is positive everywhere. In arecent work Liu, Long and Zhu [67] have shown that if in addition S is symmetric about

the origin, the number of such orbits is at least N , and for general (possibly nonsymmetric)S as above, Long and Zhu [71] have shown the existence of at least

[N2

]+ 1 closed orbits

([a] denotes the integer part of a). They also make a new conjecture that[N2

]+ 1 (and

not N) is the lower bound for the number of closed Hamiltonian orbits. See also Long’s

book [69] for a detailed discussion.In the case N = 2, Hofer, Wysocki and Zehnder [50] proved that if S bounds a strictly

convex set then there are either two or infinitely many closed Hamiltonian orbits on S.

In [32] Ekeland has shown that a generic S bounding a compact convex set and havingpositive Gaussian curvature carries infinitely many closed Hamiltonian orbits. This result

has been partially generalized by Viterbo [97] to hypersurfaces satisfying a conditionsimilar to (S). The question of existence of infinitely many closed orbits is extensively

discussed in [33, 69] where many additional references may be found.

2.4 Superlinear systems

Throughout this section we assume that H satisfies (H1)-(H3), H(z, t) = 12Az ·z+G(z, t),

where A is a symmetric 2N × 2N matrix, Gz(z, t) = o(z) uniformly in t as z → 0, and

there exist µ > max2, s− 1 and R > 0 such that

(2.31) 0 < µG(z, t) ≤ z ·Gz(z, t) for all |z| ≥ R.

Recall from (2.7) that the last condition implies G (and hence H) is superquadratic and

Hz superlinear.

Theorem 2.18 Suppose H satisfies the hypotheses given above and σ(JA) ∩ iZ = ∅.Then the system (2.1) has a 2π-periodic solution z 6= 0.

Proof Let

Φ(z) =1

2

∫ 2π

0

(−Jz − Az) · z dt−∫ 2π

0

G(z, t) dt

and let En be given by (2.5). Denote the linear operator corresponding to the quadratic

part of Φ by L (cf. (2.11)). Since σ(JA) ∩ iZ = ∅, L is invertible, E = E+(L) ⊕ E−(L)

40

Page 41: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

and En = E+n (L) ⊕ E−n (L) (see the discussion and notation preceding Corollary 2.7). It

follows using (2.18) and Lemma 2.4 that there exist α, ρ > 0 such that

(2.32) Φ ≤ 0 on E−n (L) ∩ Bρ, Φ ≤ −α on E−n (L) ∩ Sρ

and

(2.33) Φ ≥ 0 on E+n (L) ∩ Bρ, Φ ≥ α on E+

n (L) ∩ Sρ.

for all n. If Φ has a critical point z ∈ Φ−α, then z 6= 0, so z is a solution of (2.1) we werelooking for. Suppose no such z exists. We shall complete the proof by showing that in

this case Φ′(z) = 0 for some z with Φ(z) ≥ α.We claim that Φmn := Φ|E+

m(L)⊕E−n (L) has no critical point z ∈ Φ−αmn whenever m,n ≥ n0

and n0 is large enough. Indeed, otherwise there is a sequence zj ⊂ Φ−α such thatzj ∈ E+

mj(L) ⊕ E−nj (L), mj, nj → ∞ and Φ′mjnj (zj) = 0. By Corollary 2.7, zj → z after

passing to a subsequence, so Φ(z) ≤ −α and Φ′(z) = 0, a contradiction. Hence we maychoose n0 so that Φn0n has no critical point in Φ−α for any n ≥ n0. Let z = w+ + w− ∈E+n0+1(L)⊕ E−n (L). Then

(2.34) Φ(z) =1

2〈Lw+, w+〉+

1

2〈Lw−, w−〉 −

∫ 2π

0

G(z, t) dt,

and since G(z, t) ≥ a1|z|µ − a2 according to (2.7), it follows that Φ(z) ≤ 0 whenever

|z| ≥ R. Moreover, since n0 is fixed, R does not depend on n. If n ≥ n0 + 1, then byCorollary 1.3 (with En corresponding to Rm, F0 = E−n (L), F = E+

n0(L)⊕E−n (L) and F =

E+n0+1 ⊕ E−n (L)) there exists zn ∈ En such that Φ′n(zn) = 0 and α ≤ Φ(zn) ≤ supBR+1

Φ.Applying Proposition 2.5 to the sequence zn we obtain a critical point z with Φ(z) ≥ α.

2

Next we prove that the autonomous system

(2.35) z = JH ′(z) = J(Az +G′(z))

with superquadratic Hamiltonian has infinitely many geometrically distinct T -periodicsolutions for any T > 0. Since there is a one-to-one correspondence between T -periodic

solutions for the system z = JH ′(z) and 2π-periodic solutions for z = λJH ′(z), whereλ = T/2π (this can be easily seen by substituting τ = t/λ), we may assume without loss

of generality that T = 2π.

Theorem 2.19 Suppose H(z) = 12Az · z + G(z) satisfies (H1), (H3), H(z) ≥ 0 for all

z ∈ R2N , G satisfies (2.31) and G′(z) → 0 as z → 0. Then the system (2.35) has asequence zj of nonconstant 2π-periodic solutions such that ‖zj‖∞ →∞.

41

Page 42: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

If one can show that for each T > 0 the system (2.35) has a nonconstant T -periodicsolution zT 6= 0, then the number of geometrically distinct nonconstant 2π-periodic solu-

tions is in fact infinite. Indeed, let zk = z2π/k, then zk and zl may coincide for some k 6= l,yet the sequence zk will contain infinitely many distinct elements. However, the result

stated above shows much more: the solutions zj have amplitude which goes to infinitywith j.

Proof of Theorem 2.19 We verify the hypotheses of Corollary 1.15. By Proposition2.5, Φ satisfies (PS)∗, and obviously, ES1

= E0 ⊂ En. Let 2dn = 2N(1 + n), X =

(E+r (L)⊕E0(L)⊕E−(L))∩E, where r is a positive integer, and Y = E+(L)∩E (we use the

notation of the preceding proof). Employing (2.14) and recalling that M+(Tk(A)) = 2N

for large k, we have for n, r large enough (n ≥ r),

dimYn =n∑

k=1

M+(Tk(A)) = 2nN +n∑

k=1

(M+(Tk(A))− 2N)(2.36)

= 2dn + i+(A)−M−(A)−N =: 2(dn + l)

and

codimEnXn =

n∑

k=r+1

M+(Tk(A)) = 2Nn− 2Nr = 2dn − 2N(r + 1) =: 2(dn +m).

It follows from (2.33) that Φ|Y ∩Sρ ≥ α and from (2.34) with z = w+ + w0 + w− ∈E+r (L) ⊕ E0(L) ⊕ E−(L) that Φ(z) → −∞ whenever ‖z‖ → ∞, z ∈ E0 ⊕ X. Hence

Φ|E0⊕X ≤ β for some β. Moreover, Φ|E0 ≤ 0 because H ≥ 0. We have verified the

hypotheses of Corollary 1.15 for all r large enough. Since m → −∞ as r → ∞, weconclude from Corollary 1.16 that (2.35) has a sequence zj of nonconstant 2π-periodic

solutions such that Φ(zj)→∞.It remains to show that ‖zj‖∞ →∞. By (H3),

cj = Φ(zj) = Φ(zj)−1

2〈Φ′(zj), zj〉 =

∫ 2π

0

(1

2zj ·H ′(zj)−H(zj)

)dt(2.37)

≤ c

∫ 2π

0

(1 + |z|s) dt ≤ 2πc(1 + ‖z‖s∞),

and the conclusion follows because cj →∞. 2

The assumption H ≥ 0 is not necessary. Below we show how the proof of Theorem2.19 can be modified in order to remove it.

Corollary 2.20 The conclusion of Theorem 2.19 remains valid without the condition

H ≥ 0.

42

Page 43: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Proof Since G satisfies (2.31), so does H according to Remark 2.6; hence H is boundedbelow. Let

α := maxz∈E0

Φ(z)

and let r0 < r < n be positive integers. Let X = (E+r (L)⊕E0(L)⊕E−(L))∩ E as before

and Y = Er0(L)⊥ ∩ E+(L) ∩ E. Then dimYn = 2(dn + l) and we still have m < l if

r − r0 is large enough. Define S = Y ∩ z ∈ E : ‖z‖s = 1 and note that S is radiallyhomeomorphic to the unit sphere in Y . We claim that Φ|Y ∩S ≥ α > α for all large

r0. Assuming this for the moment, we find r0 such that the condition above is satisfied.

Since Φ|E0 ≤ α, we can easily see by modifying the argument of Theorem 1.14 that theconclusion of Corollary 1.15 (and hence also of Corollary 1.16) holds.

It remains to prove the claim. Arguing by contradiction, we find rj → ∞ and zj ∈Erj+ (L) ∩ E such that ‖zj‖s = 1 and Φ(zj) ≤ α. Hence

α ≥ Φ(zj) =1

2〈Lzj, zj〉 −

∫ 2π

0

G(zj) dt

≥ ε‖zj‖2 − c(‖zj‖ss + 1) = ε‖zj‖2 − 2c,

so zj is bounded in E. Passing to a subsequence, zj z in E and zj → z in Ls(S1,R2N).

It follows that ‖z‖s = 1; in particular, z 6= 0. On the other hand, zj ∈ Erj (L)⊥ ∩ E+(L)implies zj 0, a contradiction. A somewhat different argument will be given in the proof

of Theorem 2.25. 2

The first result on the existence of a nontrivial periodic solution of (2.1) is due toRabinowitz [78]. Theorem 2.18 may be found in [62]. The result contained there is in

fact more general: the case σ(JA)∩ iZ 6= ∅ is allowed if G has constant sign for small |z|.Also some Hamiltonians not satisfying the requirement µ > s − 1 are allowed; for this

purpose a truncation argument indicated in Remark 2.6 is employed. Other extensionsof Theorem 2.18 are due to Felmer [38] and Long and Xu [70]. Corollary 2.19 is due to

Rabinowitz [79] (in [79] no growth restriction (H3) is needed; again, this is achieved bytruncation).

An interesting question concerning the autonomous system (2.35) is whether one can

find solutions with prescribed minimal period. Results in this direction, mainly for convexHamiltonians, can be found in Ekeland’s book [33, Section IV.5] and in Long’s book [69,

Chapter 13].

2.5 Asymptotically linear systems

In this section we assume that in addition to (H1)-(H3) H satisfies the conditions

(2.38) H(z, t) =1

2A0z · z +G0(z, t), where (G0)z(z, t) = o(z) uniformly in t as z → 0

43

Page 44: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

and(2.39)

H(z, t) =1

2A∞z · z +G∞(z, t), where (G∞)z(z, t) = o(|z|) uniformly in t as |z| → ∞.

Here A0, A∞ are 2N × 2N constant matrices. We assume for simplicity that the sys-

tem (2.1) is nonresonant at the origin and at infinity, that is, σ(JA0)∩iZ = σ(JA∞)∩iZ =∅. This terminology is justified by the fact that the systems z = JA0z and z = JA∞zhave no other 2π-periodic solutions than z = 0.

As a first result in this section we give a sufficient condition for the existence of a

nontrivial 2π-periodic solution of (2.1).

Theorem 2.21 Suppose H satisfies (H1)-(H3), (2.38), (2.39) and σ(JA0) ∩ iZ =

σ(JA∞) ∩ iZ = ∅. If i−(A0) 6= i−(A∞), then the system (2.1) has a 2π-periodic solu-tion z 6= 0.

Proof Suppose i−(A0) < i−(A∞). The same argument applied to −Φ will give theconclusion for i−(A0) > i−(A∞). Let L0 and L∞ be given by (2.11), with respectively

A = A0 and A = A∞. As in the proof of Theorem 2.18, we see that (2.32) and (2.33)

are satisfied, with L0 replacing L. Suppose Φ has no other critical points than 0. ThenΦn = Φ|En has no critical points with |Φn(z)| ≥ α provided n ≥ n0 and n0 is large

enough. For otherwise we find zj ∈ Enj such that nj → ∞ and Φ′nj (zj) = 0. Accordingto Proposition 2.5, zj → z after passing to a subsequence, hence z is a critical point and

|Φ(z)| ≥ α which is impossible. Fix n ≥ n0. Now we invoke Theorem 1.2. It follows fromLemma 2.4 that Φ′n(z) = PnL∞|En(z) + o(‖z‖) as ‖z‖ → ∞. If n is large enough, then

(with 2dn = 2N(1 + n) as in the proof of Theorem 2.19)

dimE−n (L0) = M+(A0) +

n∑

k=1

M−(Tk(A0)) = 2dn + i−(A0)−N,

and similarly,

M−(PnL∞|En) = dimE−n (L∞) = 2dn + i−(A∞)−N.Since i−(A0) < i−(A∞), Theorem 1.2 with F0 = E−n (L0) and B = PnL∞|En implies that

Φn has a critical point z such that |Φn(z)| ≥ α. This contradiction completes the proof.2

As in the preceding section, we now turn our attention to the autonomous case.

Theorem 2.22 Suppose H = H(z) satisfies (H1), (H3), (2.38), (2.39) and σ(JA0)∩iZ =

σ(JA∞) ∩ iZ = ∅. If H ≥ 0 and i−(A0) < i−(A∞), then the system (2.1) has at least12(i−(A∞)− i−(A0)) geometrically distinct nonconstant 2π-periodic solutions.

44

Page 45: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Proof We verify the assumptions of Corollary 1.15. Φ satisfies (PS)∗, ES1= E0 ⊂ En

and Φ|E0 ≤ 0. Let X = E−(L∞)∩E and Y = E+(L0)∩E. Since H ≥ 0 and σ(JA0)∩iZ =

σ(JA∞) ∩ iZ = ∅, A0 and A∞ are positive definite and M+(A0) = M+(A∞) = 2N .Therefore (cf. (2.36))

dimYn = 2dn + i+(A0)−N =: 2(dn + l)

and

codimEnXn =

n∑

k=1

M+(Tk(A∞)) = 2dn + i+(A∞)−N =: 2(dn +m).

Since i+(A0) = −i−(A0) and i+(A∞) = −i−(A∞), 12(i−(A∞)− i−(A0)) = l−m. Finally,

Φ|Y ∩Sρ ≥ α > 0 according to (2.33) and since E−(L∞) = E0 ⊕ X, it is easy to see that

Φ|E0⊕X ≤ β for an appropriate β > α. Corollary 1.15 yields at least l −m geometricallydisitnct 2π-periodic solutions with Φ ≥ α. Since Φ > 0 these solutions cannot be constant.

2

In the next theorem we drop the hypothesis H ≥ 0 and require H to be even in z.

Theorem 2.23 Suppose H = H(z) satisfies (H1), (H3), (2.38), (2.39) and σ(JA0)∩iZ =

σ(JA∞) ∩ iZ = ∅. If H(−z) = H(z) for all z ∈ R2N , then the system (2.1) has at least12|i−(A∞)− i−(A0)| geometrically distinct 2π-periodic solutions z 6= 0.

Proof We only sketch the argument. Since Φ is an even functional, in view of Remark1.17 we may apply Theorem 1.14 if i−(A0) > i−(A∞) and Corollary 1.15 if i−(A0) <

i−(A∞) in order to get |i−(A∞)− i−(A0)| pairs of nonzero 2π-periodic solutions (here weuse genus instead of index and disregard the S1-symmetry). For a critical value c the

set Kc consists of critical S1-orbits, some of them may correspond to constant solutions,

the other ones are homeomorphic to S1. So if Kc contains a nonconstant solution, thenγ(Kc) ≥ 2. On the other hand, if γ(Kc) > 2, it is easy to see that Kc contains infinitely

many geometrically distinct critical orbits. Hence the number of nonzero geometricallydistinct critical orbits is at least 1

2|i−(A∞)− i−(A0)|. 2

Remark 2.24 (a) The argument of Theorem 2.23 does not guarantee the existence ofnonconstant solutions.

(b) If H = H(z, t) is even in z and satisfies the other assumptions of the abovetheorem, then the same argument asserts the existence of at least |i−(A∞)− i−(A0)| pairs

of nontrivial 2π-periodic solutions (see [4] or [92]). Hamiltonian systems with spatialsymmetries will be further discussed in Section 2.6 below.

(c) As we have mentioned in Section 2.4, there is a one-to-one correspondence between2π-periodic solutions of the system z = JH ′(z) and T -periodic solutions of z = λJH ′(z),

45

Page 46: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

where λ = 2π/T . Hence, in view of the results of that section, there exist nonconstantT -periodic solutions of any period T whenever H is autonomous and superquadratic.

Here the situation is different. If H ≥ 0, then i−(λA0) = i−(λA∞) = N for all smallλ > 0. Thus Theorem 2.22 gives no nonconstant solutions of small period T . This is

not surprising, for if H ′ is Lipschitz continuous with Lipschitz constant M , then eachnonconstant periodic solution must have period T ≥ 2π/M according to Theorem 4.3 in

[20]. On the other hand, it is easy to give examples where i−(λA∞) − i−(λA0) → ∞ asλ → ∞. So the number of geometrically distinct T -periodic solutions will go to infinity

with T . However, there may be no solutions of arbitrarily large minimal period (see [92],Remark 6.3 for an example).

There is an extensive literature concerning the existence of one or two nontrivial

solutions of (2.1) in the framework of Theorem 2.21. Usually the argument is based on aninfinite-dimensional Morse theory and it is possible to weaken the nonresonance conditions

at zero and infinity. Also, it is not necessary to have constant matrices A0 and A∞. Thefirst results related to Theorem 2.21 may be found in Amann and Zehnder [3, 4]. For other

results and more references, see e.g. Abbondandolo [1], Chang [23], Guo [48], Izydorek[53], Kryszewski and Szulkin [57], Li and Liu [61], Szulkin and Zou [93]. Theorem 2.22 is

due to Amann and Zehnder [4] and Benci [18]. It has been extended by Degiovanni and

Olian Fannio [28], see also [92]. While the proof in [28] uses a cohomological index theory(like the one in [36]) and a variant of Benci’s pseudoindex [18], the argument in [92] is

based on a relative limit index (which is a generalization of the limit index iE). Anotherextension, using Conley index theory, has been carried out by Izydorek [54]. Some other

aspects of the problem (an estimate of the number of T -periodic solutions in terms of theso-called twist number) are discussed in Abbondandolo [1]. However, in all results related

to Theorem 2.22 we know of, the assumptions on H are rather restrictive. This is brieflydiscussed in Remark 6.4 of [92]. Theorem 2.23 is due to Benci [18]. For results about

solutions of the autonomous equation (2.1) with prescribed minimal period we refer againto Long’s book [69, Section 13.3].

2.6 Spatially symmetric Hamiltonian systems

In this section we consider the non-autonomous Hamiltonian system (2.1) when H is

invariant with respect to certain group representations in R2N . More precisely, we considertwo different kinds of symmetries:

• A compact Lie groupG acts on R2N via an orthogonal and symplectic representation;

the standard example is the antipodal action of Z/2 (i.e., H is even in z).

• The infinite group Zk acts on R2N via space translations; the standard example

is Z2N leading to a Hamiltonian system on the torus T 2N := R2N/Z2N . Another

46

Page 47: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

example is ZN acting via translation of the q-variables which leads to a Hamiltoniansystem on the cotangent space T ∗TN of the N -dimensional torus.

In the first case we may think of G as a closed subgroup of O(2N) ∩ Sp(2N). We shall

always assume that H : R2N×R→ R satisfies the hypotheses (H1)−(H3) from Section 2.1.First we treat the compact group case and require:

(S) The compact group G acts on R2N via an orthogonal and symplectic representationT such that the action is fixed point free on R2N \ 0 (i.e., (R2N)G = 0). H is

invariant with respect to T : H(Tgz, t) = H(z, t) for all g ∈ G, z ∈ R2N , t ∈ R.

By an orthogonal and symplectic representation we mean that the matrix of Tg is in

O(2N)∩Sp(2N) for all g ∈ G. Clearly, if (S) holds and z(t) is a periodic solution of (2.1)then so is Tgz(t) for every g ∈ G. Thus one has to count G-orbits of periodic solutions

and not just periodic solutions.

Theorem 2.25 Suppose (H1) − (H3) and (S) hold for G of prime order. If H is su-perquadratic in the sense of (2.31) then the system (2.1) has a sequence of 2π-periodic

solutions zj such that ‖zj‖∞ →∞.

If H is invariant with respect to an orthogonal symplectic representation of a moregeneral compact Lie group G then one can apply Theorem 2.25 provided there exists a

subgroup G1 ⊂ G of prime order having 0 as the only fixed point. This may or may notbe the case. It is always the case for G = S1 or more generally, for G = (S1)k a torus,

acting without nontrivial fixed points. A general existence result in this direction worksfor admissible group actions; see Remark 1.10.

Proof We want to apply Corollary 1.16 to the usual action functional

Φ(z) =1

2

∫ 2π

0

(−Jz · z) dt−∫ 2π

0

H(z, t) dt.

Although the proof of Corollary 2.20 could be used here with minor changes, we provide

a slightly different argument as we have mentioned earlier. By Proposition 2.5 the (PS)∗-condition holds.

Recall the spaces En ⊂ E from (2.5). We choose k0 ∈ N and set Y := E⊥k0∩E+. Now

we claim that for k0 large enough, there exists ρ, α > 0 so that Φ satisfies condition (iii)from Corollary 1.15, that is,

(2.40) Φ(z) ≥ α for z ∈ Y with ‖z‖ = ρ.

In order to see this we first observe that

‖z‖ ≥√k0‖z‖2 holds for z ∈ Y ,

47

Page 48: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

and that there exists c1 > 0 with

|H(z, t)| ≤ c1(|z|s + 1)

by (H3). Using the continuous embedding E → L2s−2 we obtain

‖z‖ss ≤ ‖z‖2 · ‖z‖s−12s−2 ≤

c2√k0

‖z‖ · ‖z‖s−1 for z ∈ Y .

This implies

Φ(z) ≥ 1

2‖z‖2 − c1(‖z‖ss − 2π) ≥ 1

2‖z‖2 − c1c2√

k0

‖z‖s − 2c1π

for every z ∈ Y . Setting ρ = (√k0

sc1c2)1/(s−2) we therefore have for z ∈ Y with ‖z‖ = ρ:

Φ(z) ≥(

1

2− 1

s

)( √k0

sc1c2

)2/(s−2)

− 2c1π > 0

provided k0 is large. Thus we may fix k0 ∈ N so that (2.40) holds.Next we define Xk := E−+Ek for k ∈ N. Then sup Φ(Xk) <∞ because Φ(z)→ −∞

for z ∈ Xk with ‖z‖ → ∞ as we have seen earlier. In order to apply Corollary 1.16 withX = Xk it remains to check the dimension condition from Corollary 1.16. Recall that

dG = 1 for G = Z/p. Setting dn = 2N(1 + n) we have

dim Yn = 2N(n− k0) = dn + l

with l = −2N(k0 + 1) and

codimEn(Xk ∩ En) = 2N(n− k) = dn +m(k)

with m(k) = −2Nk − 2N (En = En here because EG = 0). Clearly l −m(k) →∞ ask →∞. Now the theorem follows from Corollary 1.16 and (2.37). 2

Comparing Theorem 2.25 with Theorem 2.18 and 2.19 we see that from the variationalpoint of view the spatial symmetry condition (S) has the same effect as the S1-symmetry of

the autonomous problem. This is also true for asymptotically linear Hamiltonian systems– as we already observed in Remark 2.24(b). We state one multiplicity result in this

setting.

Theorem 2.26 Suppose H satisfies (H1) − (H3) and is asymptotically quadratic in the

sense of (2.38) and (2.39). Suppose moreover that σ(JA0)∩ iZ = σ(JA∞)∩ iZ = ∅, andlet i−(A0) and i−(A∞) be the Morse indices defined in (2.13). If (S) holds for G of prime

order or for G ∼= S1 then the system (2.1) has at least 1dG|i−(A∞) − i−(A0)| G-orbits of

nontrivial 2π-periodic solutions.

48

Page 49: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Proof The result follows from Theorem 1.14 or Corollary 1.15; cf. also the proof ofTheorem 2.23. 2

Theorem 2.26 is also true for G = (Z/p)k a p-torus and dG = 1, or G = (S1)k a torusand dG = 2, or if G acts freely on R2N \ 0 and dG = 1 + dim(G). G acting freely means

that Tgz = z for some z 6= 0 implies that g is the identity. Such actions exist only for avery restricted class of Lie groups. It does not help much if a subgroup G1 of G acts freely

(or without nontrivial fixed points if G1 is a torus or p-torus), because a G-orbit consistsof several G1-orbits. So the multiple G1-orbits of periodic solutions may correspond to

just one G-orbit of periodic solutions.There are various extensions of Theorem 2.26 when the linearized equations at 0 or at

∞ have nontrivial 2π-periodic solutions, mostly for even Hamiltonians; see for instance

[54] and the references therein.

Now we consider the spatially periodic case. The classical result is due to Conley and

Zehnder [24] and deals with the case of Z2N , that is H is periodic in all variables. Thenperiodic solutions appear in Z2N -orbits.

Theorem 2.27 Suppose H ∈ C1(R2N × R) is 2π-periodic in all variables. Then (2.1)has at least 2N + 1 distinct distinct Z2N -orbits of 2π-periodic solutions.

Proof We consider the decomposition E = E+⊕E0⊕E− from Section 2.1 and observethat

Φ(z + 2πk) = Φ(z) for every z ∈ E, k ∈ Z2N .

Setting M = T 2N = E0/2πZ2N we obtain an induced C1-functional

Ψ : (E+ ⊕ E−)×M → R, Ψ(z+, z−, z0 + 2πZ2N ) = Φ(z+ + z− + z0).

Critical points of Ψ correspond to Z2N -orbits of 2π-periodic solutions of (2.1). The con-

clusion follows from Theorem 1.18 and the fact that cupl(T 2N) = 2N + 1. More precisely,

we let W = E− and Y = E+. Since H is bounded, β as in (ii) of Theorem 1.18 existsand taking ρ large enough, we also find α < β and γ. Finally, it is easy to see that

(PS)∗-sequences are bounded (cf. Proposition 2.5), consequently, the (PS)∗-condition issatisfied. 2

Using Theorem 1.18 one can also treat more general periodic symmetries, for instancewhen H(p, q, t) is invariant under ZN acting on the q-variables by translations. Then one

needs some condition on the behavior of H(p, q, t) as |p| → ∞. Results in this directionhave been obtained by a number of authors, see [23, 37, 42, 40, 66, 90].

If the periodic solutions are non-degenerate then Conley and Zehnder [24] used Morse

theoretic arguments to prove:

49

Page 50: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Theorem 2.28 Suppose H ∈ C2(R2N × R) is 2π-periodic in all variables and all 2π-periodic solutions of (2.1) are non-degenerate. Then (2.1) has at least 22N distinct Z2N -

orbits of 2π-periodic solutions.

The Morse theoretic arguments involve in particular the Conley-Zehnder index; see

Section 2.1, in particular Remark 2.8. Theorems 2.27 and 2.28 are special cases of theArnold conjecture. This states that a Hamiltonian flow on a compact symplectic manifold

M has at least cat(M) periodic solutions. If all periodic solutions are non-degeneratethen it has at least

∑dim Mi=0 dim Hi(M) critical points where Hi(M) denotes the i-th

homology group of M with coefficients in an arbitrary field. Theorems 2.27 and 2.28correspond to the case M = T 2N where cat(M) = 2N+1 and dim Hi(M) =

(2Ni

), so that∑dim M

i=0 dim Hi(M) = 22N . The interested reader can find results and many referencesconcerning the Arnold conjecture in the book [52] and in the paper [41].

We conclude this section with a theorem on Hamiltonian systems where the Hamilto-

nian is both even and spatially periodic in the z-variables:

(2.41) H(z + 2πk, t+ 2π) = H(z, t) = H(−z, t) for all z ∈ R2N , k ∈ Z2N , t ∈ R.

It follows that Hz(z, t) = 0 for all z ∈ (πZ)2N , hence modulo the Z2N -action (2.1) has at

least 22N stationary solutions z(t) ≡ ci with ci ∈ 0, π, i = 1, . . . , 2N . Thus the Arnoldconjecture is trivially satisfied if (2.41) holds. A natural question to ask is whether there

exist 2π-periodic solutions in addition to the 22N trivial equilibria.

Theorem 2.29 Suppose H ∈ C2(R2N × R) satisfies (2.41) and suppose all 2π-periodic

solutions of (2.1) are non-degenerate. For z ∈ πZ2N let Az(t) := Hzz(z, t) and letj−(Az) ∈ Z be the Conley-Zehnder index. Then (2.1) has at least k = max |j−(Az)| −Npairs ±z1, . . . ,±zk of 2π-periodic solutions which lie on different Z2N -orbits.

Theorem 2.29 has been proved in [14], an extension to Hamiltonian systems on T ∗TN ,

where H is even in z and periodic in the q-variables can be found in [15].

3 Homoclinic solutions

3.1 Variational setting for homoclinic solutions

Up to now we have been concerned with periodic solutions of Hamiltonian systems. In

this part we turn our attention to homoclinic solutions of the system

(3.1) z = JHz(z, t).

We assume H satisfies the following hypotheses:

50

Page 51: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

(H1) H ∈ C(R2N × R,R), Hz ∈ C(R2N × R,R2N) and H(0, t) ≡ 0;

(H2) H is 1-periodic in the t-variable;

(H3) |Hz(z, t)| ≤ c(1 + |z|s−1) for some c > 0 and s ∈ (2,∞);

(H4) H(z, t) = 12Az · z + G(z, t), where A is a constant symmetric 2N × 2N -matrix,

σ(JA) ∩ iR = ∅ and Gz(z, t)/|z| → 0 uniformly in t as z → 0.

Note that the hypotheses (H1)-(H3) are the same as (H1)-(H3) in Section 2.1 except that

the period is normalized to 1 and not 2π (which is slightly more convenient here).Let z0 be a 1-periodic solution of (3.1). A solution z is said to be homoclinic (or

doubly asymptotic) to z0 if z 6≡ z0 and |z(t) − z0(t)| → 0 as |t| → ∞. It has been shownby Coti Zelati, Ekeland and Sere in [25] that if H satisfies (H1)-(H3), then in many cases

a symplectic change of variables will reduce the problem of finding homoclinics to z0 tothat of finding solutions homoclinic to 0 for a system (3.1), with a new Hamiltonian

H satisfying (H1)-(H4). Therefore in what follows we only consider solutions which arehomoclinic to 0 (i.e., z 6≡ 0 and z(t) → 0 as |t| → ∞), or homoclinic solutions for short.

Recall that if (H4) holds, then 0 is called a hyperbolic point.Let E := H1/2(R,R2N) be the Sobolev space of functions z ∈ L2(R,R2N) such that

(1 + ξ2)1/2|z(ξ)|2 dξ <∞,

where z is the Fourier transform of z. E is a Hilbert space with an inner product

(z, w) :=

(1 + ξ2)1/2z(ξ) · w(ξ) dξ.

The Sobolev embedding E → Lq(R,R2N) is continuous for any q ∈ [2,∞) (see e.g. [2]

or Section 10 in [87]) but not compact. Indeed, let zj(t) := z(t − j), where z 6≡ 0; thenzj 0 in E as j → ∞ but zj 6→ 0 in Lq. However, the embedding E → Lqloc(R,R2N) is

compact.Now we introduce a more convenient inner product. It follows from (H4) that −iξJ−A

is invertible and (−iξJ − A)−1 is uniformly bounded with respect to ξ ∈ R. Using thisfact, the equality −Jz · w = z · Jw and Plancherel’s formula it can be shown that the

mapping L : E → E given by

(Lz, w) =

(−Jz − Az) · w dt

is bounded, selfadjoint and invertible (see Section 10 in [87] for the details). It is also

shown in [87] that the spectrum σ(−J(d/dt) − A) is unbounded below and above in

H1(R,R2N); hence E = E+ ⊕ E−, where E± are L-invariant infinite-dimensional spaces

51

Page 52: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

such that the quadratic form (Lz, z) is positive definite on E+ and negative definite onE−. Therefore we can define a new equivalent inner product in E by setting

〈z, w〉 := (Lz+, w+)− (Lz−, w−),

where z±, w± ∈ E±. If ‖ . ‖ denotes the corresponding norm, we have

(3.2)

(−Jz − Az) · z dt = (Lz, z) = ‖z+‖2 − ‖z−‖2.

Let

Φ(z) :=1

2

(−Jz − Az) · z dt−∫

G(z, t) dt

and

ψ(z) :=

G(z, t) dt.

Then

(3.3) Φ(z) =1

2‖z+‖2 − 1

2‖z−‖2 − ψ(z).

Proposition 3.1 If H satisfies (H1)-(H4), then Φ ∈ C1(E,R) and z is a homoclinicsolution of (3.1) if and only if z 6= 0 and Φ′(z) = 0. Moreover, ψ′ and Φ′ are weakly

sequentially continuous.

Proof We outline the argument. By (H3) and (H4), |Gz(z, t)| ≤ c(|z|+ |z|s−1) for some

constant c > 0. Hence ψ ∈ C1(E,R) and

〈ψ′(z), w〉 =

Gz(z, t) · w dt

according to Lemma 3.10 in [102] (although in [102] E is the Sobolev space H 1(RN),

an inspection of the proof shows that the argument remains valid in our case). Having

this, it is easy to see from (3.2) and (3.3) that Φ ∈ C1(E,R) and Φ′(z) = 0 if and onlyif z ∈ E and z is a weak solution of (3.1). Since z ∈ Lq(R,R2N) for all q ∈ [2,∞),

Gz(z(.), .) ∈ L2(R,R2N); hence z ∈ H1(R,R2N). In particular, z is continuous by theSobolev embedding theorem, and consequently, z ∈ C1(R,R2N), i.e., z is a classical

solution of (3.1).It is well known that if z ∈ H1(R,R2N), then z(t) → 0 as |t| → ∞. For the reader’s

convenience we include a proof. Let t ≤ u ≤ t + 1. Given ε > 0, there is R such that if|t| ≥ R, then ‖z‖L2((t,t+1), 2N ) < ε and ‖z‖L2((t,t+1), 2N ) < ε. Hence |z(u)| < ε for some

u ∈ [t, t+ 1] and, using Holder’s inequality,

|z(t)| ≤ |z(u)|+ |z(u)− z(t)| ≤ ε+

∫ u

t

|z(τ)| dτ ≤ ε+ ‖z‖L2((t,t+1), 2N ) < 2ε.

52

Page 53: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Finally, weak sequential continuity follows from the compact embedding E → Lqloc,2 ≤ q < ∞. Indeed, if zj z in E, then zj → z in L2

loc(R,R2N) ∩ Lsloc(R,R2N), and

invoking the argument of Lemma 3.10 in [102] again, we see that 〈ψ ′(zj), w〉 → 〈ψ′(z), w〉for all w ∈ E, i.e., ψ′(zj) ψ′(z). Clearly, Φ′(zj) Φ′(z) as well. 2

It will be important in what follows that the functional Φ is invariant with respect tothe representation of the group Z of integers given by

(3.4) (Taz)(t) := z(t + a), a ∈ Z

(this is an immediate consequence of the periodicity of G). Moreover, since the linear

operator L is Z-invariant, so are the subspaces E±.

It follows from the Z-invariance of Φ that Φ′ is Z-equivariant; hence if z = z(t) is ahomoclinic, so are all Taz, a ∈ Z. Therefore Φ cannot satisfy the Palais-Smale condition

at any critical level c 6= 0. Setting O(z) := Taz : a ∈ Z (cf. Section 1.1), we call twohomoclinic solutions z1, z2 geometrically distinct if O(z1) 6= O(z2).

Lemma 3.2 If H satisfies (H1)-(H4), then ψ′(z) = o(‖z‖) and ψ(z) = o(‖z‖2) as z → 0.

We omit the argument which is exactly the same as in Lemma 2.4. Next we turn ourattention to Palais-Smale sequences.

Proposition 3.3 If H satisfies (H1)-(H4) and there are µ > max2, s − 1 and δ > 0

such that

(3.5) δ|z|µ ≤ µG(z, t) ≤ z ·Gz(z, t) for all z, t,

then each (PS)-sequence zj for Φ is bounded. Moreover, if Φ(zj)→ c, then c ≥ 0, and

if c = 0, then zj → 0.

Proof As in (2.8), we have

(3.6) c1‖zj‖+ c2 ≥ Φ(zj)−1

2〈Φ′(zj), zj〉 ≥

(µ2− 1)∫

G(zj, t) dt ≥ c3‖zj‖µµ.

Since for each ε > 0 there is C(ε) such that |Gz(z, t)| ≤ ε|z| + C(ε)|z|s−1 (by (H3) and(H4)), we see as in (2.10) that

(3.7) ‖z±j ‖2 ≤ αj‖z±j ‖+ c4ε‖zj‖ ‖z±j ‖+ c5(ε)‖zj‖s−1µ ‖z±j ‖,

where αj := ‖Φ′(zj)‖ → 0. Hence choosing ε small enough, we have ‖zj‖ ≤ c6 +c7‖zj‖s−1µ ,

and taking (3.6) into account,

‖zj‖ ≤ c8 + c9‖zj‖(s−1)/µ.

53

Page 54: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

It follows that zj is bounded.Now we obtain from (3.6) that if Φ(zj)→ c, then

c ≥ c3 lim supj→∞

‖zj‖µ,

so c ≥ 0. If c = 0, then zj → 0 in Lµ and passing to a subsequence, zj z in E. Hencez = 0. Letting j →∞ in (3.7) we see that zj → 0 also in E. 2

Condition (3.5) is rather restrictive. Other conditions (always including the inequality0 < µG(z, t) ≤ z · Gz(z, t) for all z 6= 0 and some µ > 2) which imply boundedness of

(PS)-sequences may be found e.g. in [6] and [31].We shall need the following result which is a special case of P.L. Lions’ vanishing

lemma (see e.g. [65] or [102]):

Lemma 3.4 Let r > 0 be given and let zj be a bounded sequence in E. If

(3.8) limj→∞

supa∈

∫ a+r

a−r|zj|2 dt = 0,

then zj → 0 in Lq(R,R2N ) for all q ∈ (2,∞).

Proof In [65, 102] the case of E = H1(RN ) has been considered. Below we adapt theargument of [102, Lemma 1.21] to our situation.

Let q ∈ (2, 4). By Holder’s inequality,

‖zj‖qLq((a−r,a+r), 2N )≤ ‖zj‖q−2

L2((a−r,a+r), 2N )‖zj‖2

Lp((a−r,a+r), 2N ),

where p satisfies (q − 2)/2 + 2/p = 1. Hence

‖zj‖qLq((a−r,a+r), 2N )≤ sup

b∈

(‖zj‖q−2

L2((b−r,b+r), 2N )

)‖zj‖2

Lp((a−r,a+r), 2N )(3.9)

≤ C supb∈

(‖zj‖q−2

L2((b−r,b+r), 2N )

)‖zj‖2

H1/2((a−r,a+r), 2N ).

Here we may use the norm in H1/2((a− r, a+ r),R2N) given by

‖z‖2H1/2((a−r,a+r), 2N ) = ‖z‖2

L2((a−r,a+r), 2N ) +

∫ a+r

a−r

∫ a+r

a−r

|z(t)− z(s)|2(t− s)2

dsdt

(see [2], Theorem 7.48). Covering R by intervals (an − r, an + r), n ∈ Z, in such a waythat each t ∈ R is contained in at most 2 of them and taking the sum with respect to n

in (3.9), we obtain

‖zj‖qq ≤ 2C supa∈

(‖zj‖q−2

L2((a−r,a+r), 2N )

)|zj|2E,

54

Page 55: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

where

|zj|2E := ‖zj‖22 +

|z(t)− z(s)|2(t− s)2

dsdt.

According to Theorem 7.12 in [64], the norms | . |E and ‖ . ‖ are equivalent. Since zj is

bounded in E, it follows that zj → 0 in Lq.If q ≥ 4, we can choose q0 ∈ (2, 4) and p > q. Then by Holder’s inequality, ‖zj‖qq ≤

‖zj‖(1−λ)qq0 ‖zj‖λqp , where (1−λ)q/q0+λq/p = 1, and the conclusion follows because ‖zj‖q0 →

0 and ‖zj‖p is bounded. 2

Proposition 3.5 Suppose (H1)-(H4) are satisfied and zj is a bounded (PS)-sequence

such that Φ(zj)→ c > 0. Then (3.1) has a homoclinic solution.

Proof Suppose first zj is vanishing in the sense that (3.8) is satisfied. It is clear that

(3.7) holds for µ = s. Since αj = ‖Φ′(zj)‖ → 0 and zj → 0 in Ls according to Lemma3.4, it follows from (3.7) with ε appropriately small that zj → 0 in E; hence Φ(zj) → 0.

This contradiction shows that zj cannot be vanishing. Therefore there exist δ > 0 andaj such that, up to a subsequence,

(3.10)

∫ aj+r

aj−r|zj|2 dt ≥ δ

for almost all j. Choosing a larger r if necessary we may assume aj ∈ Z. Let zj(t) :=zj(aj + t). It follows from the Z-invariance of Φ that zj is a bounded (PS)-sequence

and Φ(zj) → c. Hence zj z in E and zj → z in L2loc after passing to a subsequence.

Moreover, since

(3.11)

∫ r

−r|zj|2 dt =

∫ aj+r

aj−r|zj|2 dt,

z 6= 0. According to Proposition 3.1, Φ′ is weakly sequentially continuous. Thus Φ′(z) = 0and the conclusion follows. 2

3.2 Existence of homoclinics

Our first result in this section asserts that if Gz is superlinear, then (3.1) has at least 1homoclinic.

Theorem 3.6 Suppose H satisfies (H1)-(H4) and G satisfies (3.5) with µ > max2, s−1. Then (3.1) has a homoclinic solution.

55

Page 56: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Proof According to (3.3), the functional Φ corresponding to (3.1) has the form requiredin Theorem 1.19. Moreover, G ≥ 0 and therefore ψ ≥ 0. Let zj z. Then zj → z

in L2loc(R,R2N) and zj → z a.e. in R after passing to a subsequence, so it follows from

Fatou’s lemma that ψ is weakly sequentially lower semicontinuous. Moreover, ψ ′ is weakly

sequentially continuous according to Proposition 3.1. Hence (i) of Theorem 1.19 holds,and so does (ii) because Φ(z) = 1

2‖z‖2 + o(‖z‖2) whenever z → 0, z ∈ E+.

We shall verify (iii). Let z0 ∈ E+, ‖z0‖ = 1. Since there exists a continuous projectionfrom the closure of Rz0 ⊕ E− in Lµ(R,R2N) to Rz0 and G(z, t) ≥ δµ−1|z|µ,

Φ(z− + ζz0) ≤ ζ2

2− 1

2‖z−‖2 − δµ−1‖z− + ζz0‖µµ ≤

ζ2

2− 1

2‖z−‖2 − δ0ζ

µ‖z0‖µµ

for some δ0 > 0, and it follows that Φ(z−+ ζz0) ≤ 0 whenever ‖z−+ ζz0‖ is large enough.

Obviously, Φ(z−) ≤ 0 for all z− ∈ E−.We have shown that the hypotheses of Theorem 1.19 are satisfied. Hence there exists

a (PS)-sequence zj such that Φ(zj)→ c > 0 and it remains to invoke Propositions 3.3and 3.5. 2

Next we turn our attention to asymptotically linear systems. Suppose σ(JA)∩iR = ∅,let λ1 be the smallest positive and λ−1 the largest negative λ such that σ(J(A+λI))∩iR 6=∅ and set λ0 := minλ1,−λ−1. Then

(3.12)

λ1 = inf‖z‖2 : z ∈ E+, ‖z‖2 = 1λ−1 = − inf‖z‖2 : z ∈ E−, ‖z‖2 = 1‖z‖2 ≥ λ0‖z‖2 for all z ∈ E

(see Section 10 of [87] for a detailed argument). We shall need the following two additional

assumptions on G:

(H5) G(z, t) = 12A∞(t)z · z + F (z, t), where A∞(t)z · z ≥ λ|z|2 for some λ > λ1 and

Fz(z, t)/|z| → 0 uniformly in t as |z| → ∞;

(H6) G(z, t) ≥ 0 and 12Gz(z, t) · z−G(z, t) ≥ α(|z|), where α(0) = 0 and α(|z|) is positive

and bounded away from 0 whenever z is bounded away from 0.

A simple example of a function G which satisfies (H5) and (H6) is given by G(z, t) =

a(t)B(|z|), where a is 1-periodic, a(t) ≥ a0 > 0 for all t, B ∈ C1(R,R), B(0) = B′(0) = 0,B′(s)/s is strictly increasing, tends to 0 as s→ 0 and to λ > λ1/a0 as s→∞. That (H6)

holds follows from the identity

1

2B′(s)s− B(s) =

∫ s

0

(B′(s)

s− B′(σ)

σ

)σ dσ.

56

Page 57: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Lemma 3.7 If H satisfies (H1)-(H6), then each Cerami sequence zj (see Remark 1.20for a definition) is bounded.

Proof Suppose zj is unbounded and let wj := zj/‖zj‖. We may assume taking a

subsequence that wj w. We shall obtain a contradiction by showing that wj is

neither vanishing (in the sense that (3.8) holds) nor nonvanishing.Assume first wj is nonvanishing. As in the proof of Proposition 3.5, we find aj ∈ Z

such that, passing to a subsequence, wj(t) := wj(aj + t) satisfy

∫ r

−r|wj|2 dt ≥ δ > 0

for j large. Passing to a subsequence once more, wj w in E and wj → w in L2loc and

a.e. in R. In particular, w 6= 0. Since ‖Φ′(zj)‖ = ‖Φ′(zj)‖, it follows that Φ′(zj)/‖zj‖ → 0and therefore

‖zj‖−1〈Φ′(zj), v〉 = 〈w+j , v〉 − 〈w−j , v〉 −

A∞(t)wj · v dt−∫

Fz(zj, t) · v|zj|

|wj| dt→ 0

for all v ∈ C∞0 (R,R2N). Since |Fz(z, t)| ≤ a|z| for some a > 0, Fz(z, t) = o(|z|) as|z| → ∞ and supp v is bounded, we see by the dominated convergence theorem that the

last integral on the right-hand side above tends to 0. Consequently, letting j → ∞, weobtain

˙w = J(A+ A∞(t))w

which contradicts the fact that the operator −J(d/dt)− (A+ A∞(t)) has no eigenvalues

in L2(R,R2N) ([31], Proposition 2.2).

Suppose now wj is vanishing. Then

‖zj‖−1〈Φ′(zj), w+j 〉 = ‖w+

j ‖2 −∫

Gz(zj, t) · w+j

|zj||wj| dt→ 0

and

‖zj‖−1〈Φ′(zj), w−j 〉 = −‖w−j ‖2 −∫

Gz(zj, t) · w−j|zj|

|wj| dt→ 0.

Since ‖wj‖ = 1, ∫

Gz(zj, t) · (w+j − w−j )

|zj||wj| dt→ 1.

LetIj := t ∈ R : |zj(t)| ≤ ε,

57

Page 58: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

where ε > 0 has been chosen so that |Gz(z, t)| ≤ 12λ0|z| whenever |z| ≤ ε (such ε exists

according to (H4)). Since w+j and w−j are orthogonal in L2, it follows from (3.12) that

Ij

Gz(zj, t) · (w+j − w−j )

|zj||wj| dt ≤

1

2λ0‖wj‖2

2 ≤1

2

and therefore, since |Gz(z, t)| ≤ a0|z| for some a0 > 0,

1

4≤∫

\Ij

Gz(zj, t) · (w+j − w−j )

|zj||wj| dt ≤ a0

\Ij|wj|2 dt ≤ a0 meas (R \ Ij)(p−2)/p‖wj‖2/p

p

for almost all j (p > 2 arbitrary but fixed). As wj is vanishing, wj → 0 in Lp(R,R2N)according to Lemma 3.4 and consequently, meas (R \ Ij) → ∞. Let α0 := inf |z|>ε α(|z|).Then α0 > 0 and (H6) implies

Φ(zj)−1

2〈Φ′(zj), zj〉 =

(1

2Gz(zj, t)−G(zj, t)

)dt

≥∫

\Ij

(1

2Gz(zj, t)−G(zj, t)

)dt ≥

\Ijα0 dt→∞.

However, Φ(zj) is bounded and since zj is a Cerami sequence, 〈Φ′(zj), zj〉 → 0. There-

fore the left-hand side above must be bounded, a contradiction. 2

The idea of showing boundedness of zj by excluding both vanishing and nonvanish-ing of wj goes back to Jeanjean [55].

Theorem 3.8 Suppose H satisfies (H1)-(H6). Then (3.1) has a homoclinic solution.

Proof We shall use Theorem 1.19 again, this time together with Remark 1.20. Clearly,

(i) and (ii) still hold. So if we can show that also (iii) is satisfied, then in view of Remark1.20 there exists a Cerami sequence zj with Φ(zj) → c > 0. By Lemma 3.7, zj is

bounded, hence it is a (PS)-sequence as well and we can invoke Propositions 3.3 and 3.5in the same way as before.

It remains to verify (iii) of Theorem 1.19. According to (3.12) and since λ > λ1 thereexists z0 ∈ E+, ‖z0‖ = 1, such that

(3.13) 1 = ‖z0‖2 < λ‖z0‖22.

Since Φ|E− ≤ 0, it suffices to show that

(3.14) Φ(z− + ζz0) =ζ2

2− 1

2‖z−||2 −

G(z− + ζz0, t) dt ≤ 0

58

Page 59: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

whenever ‖z− + ζz0‖ is large enough. Assuming the contrary, we find z−j and ζj suchthat zj = z−j + ζjz0 satisfies ‖zj‖ → ∞ and the reverse inequality holds in (3.14). Setting

wj := zj/‖zj‖ = (z−j + ζjz0)/‖zj‖ = w−j + ηjz0, we obtain

η2j

2− 1

2‖w−j ‖2 −

G(zj, t)

‖zj‖2dt ≥ 0,

and since G ≥ 0,

(3.15)η2j

2− 1

2‖w−j ‖2 −

I

G(zj, t)

‖zj‖2dt ≥ 0,

where I is a bounded interval (to be specified). Passing to a subsequence, ηj → η ∈ [0, 1],w−j w− in E and w−j → w− in L2

loc and a.e. in R. It follows from (3.15) that ηj ≥ ‖w−j ‖,hence η > 0 because η2

j +‖w−j ‖2 = 1. In view of (3.13) and since z0 and w− are orthogonalin L2,

η2 − ‖w−‖2 −∫

A∞(t)w · w dt ≤ η2 − ‖w−‖2 − λ‖w‖22

= η2(1− λ‖z0‖22)− ‖w−‖2 − λ‖w−‖2

2

< 0,

hence there exists a bounded interval I such that

(3.16) η2 − ‖w−‖2 −∫

I

A∞(t)w · w dt < 0.

On the other hand,

I

G(zj, t)

‖zj‖2dt =

1

2

I

A∞(t)wj · wj dt+

I

F (zj, t)

|zj|2|wj|2 dt,

and since I is bounded, |F (z, t)| ≤ a|z|2 for some a > 0 and F (z, t) = o(|z|2) as |z| → ∞,

it follows from the dominated convergence theorem that the second integral on the right-hand side above tends to 0. Consequently, passing to the limit in (3.15) we obtain

η2 − ‖w−‖2 −∫

I

A∞(t)w · w dt ≥ 0,

a contradiction to (3.16). 2

While several results concerning the existence of homoclinic solutions for second order

systems (e.g. of Newtonian or Lagrangian type) may be found in the literature, much less

seems to be known about (3.1) under conditions similar to (H1)-(H4). The first paper

59

Page 60: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

to use modern variational methods for finding homoclinic solutions seems to be [25] ofCoti-Zelati, Ekeland and Sere. It has been shown there that if H satisfies (H1)-(H4),

G = G(z, t) is superquadratic in an appropriate sense and convex in z, then (3.1) has atleast 2 geometrically distinct homoclinics. If G = G(z), there is at least 1 homoclinic. The

convexity of G was used in order to reformulate the problem in terms of a dual functionalwhich is better behaved than Φ (see also a comment in Section 3.4). A result comparable to

our Theorem 3.6, but stronger in the sense that there is no growth restriction on G, is dueto Tanaka [95]. His proof is rather different from ours: he shows using a linking argument

and fine estimates that there is a sequence of 2πkj-periodic solutions zj of (3.1) whichtend to a homoclinic as j →∞. A somewhat different result has been obtained by Ding

and Willem [31]. Their function G is also superquadratic but they allow the matrix A tobe time-dependent (and 1-periodic) and moreover, they allow 0 to be the left endpoint of

a gap of the spectrum of −J(d/dt)−A(t) (more precisely, σ(−J(d/dt)−A(t))∩ (0, α) = ∅for some α > 0). See also Xu [103], where the superlinearity condition has been weakened

with the aid of a truncation argument. Theorem 3.8 is due to Szulkin and Zou [94];however, the argument presented here is simpler.

Finally we would like to mention that if A is time-dependent and |A(t)| → ∞ in an

appropriate sense as |t| → ∞, then it can be shown that in many cases Φ satisfies (PS)∗

and methods similar to those developed in Sections 2.4-2.6 become available, see e.g. Ding

[29]. Results concerning bifurcation of homoclinics may be found in Stuart [87] and Secchiand Stuart [82].

3.3 Multiple homoclinic solutions

In Section 2.4 we have seen that in the autonomous case if H is superquadratic, thenthe Hamiltonian system has infinitely many periodic solutions whose amplitude tends

to infinity. The proof relied in a crucial way on the S1-invariance of the correspondingfunctional. For homoclinics the situation is very different. Let z = (p, q) ∈ R2 and

(3.17) H(z) =1

2p2 − 1

2q2 +

1

4q4.

The corresponding Hamiltonian system reduces to a second order equation −q = V ′(q),where the potential V (q) = 1

4q4 − 1

2q2. It is easy to see that there exists a homoclinic z0

and

(3.18) S := ±z0(t− a) : a ∈ R

is the set of all homoclinics. So although S consists of infinitely (in fact uncountably)

many geometrically distinct Z-orbits (in the sense of Section 3.1), it contains only twohomoclinics which are really distinct, the reason for this being that Φ is invariant with

60

Page 61: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

respect to the representation (3.4) of R rather than Z. In particular, there are no homo-clinics of large amplitude.

Below we shall assume that, in addition to the Z-invariance, Φ is also invariant withrespect to a representation of Z/p (p a prime) in R2N and (S) of Section 2.6 is satisfied.

We also recall from Section 2.6 that if H is even in z, then (S) holds with G = Z/2 (herewe denote groups by G in order to distinguish them from functions G = G(z, t)). Our

aim is to show that there are infinitely many geometrically distinct homoclinics providedH is superquadratic. Note that since Z/p is finite, (3.1) has infinitely many homoclinics

which are geometrically distinct when both the representations of Z and Z/p are takeninto account if and only if it has infinitely many geometrically distinct homoclinics with

respect to the representation of Z only.

Theorem 3.9 Suppose H and G satisfy (H1)-(H4) and (3.5) with µ > max2, s − 1,there exist c0, ε0 > 0 such that

(3.19) |Gz(z + w, t)−Gz(z, t)| ≤ c0|w|(1 + |z|s−1) whenever |w| ≤ ε0

and (S) of Section 2.6 holds for G = Z/p, p a prime. Then (3.1) has infinitely many

geometrically distinct homoclinic solutions.

Clearly, (3.19) is satisfied if Hzz is continuous and |Hzz(z, t)| ≤ c(1 + |z|s−1) for some

c > 0.According to our comments in Section 2.6, if H is invariant with respect to an orthog-

onal and symplectic representation of a group G, if Z/p ⊂ G and (R2N ) /p = 0, thenthe conclusion of the above theorem remains valid. However, if the system is autonomous

or G is infinite, then already the existence of one homoclinic (which follows from Theorem3.6) implies that there are infinitely many homoclinics which are geometrically distinct in

the Z× Z/p-sense but not in the sense of a representation of the larger group.An important step in the proof of Theorem 3.9 is the following

Proposition 3.10 Suppose (H1)-(H4), (3.5) and (3.19) are satisfied and let zj bea (PS)c-sequence with c > 0. Then there exist (not necessarily distinct) homoclinics

w1, . . . , wk and sequences bmj (1 ≤ m ≤ k) of integers such that, passing to a subse-quence if necessary,

‖zj −k∑

m=1

Tbmj wm‖ → 0 andk∑

m=1

Φ(wm) = c.

Proof (outline) We shall only very briefly sketch the argument which is exactly the

same as in [58] (where a Schrodinger equation has been considered) or [30], see also

[26]. By Proposition 3.3, zj is bounded, and it is nonvanishing by the argument of

61

Page 62: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Proposition 3.5. Hence (3.10) is satisfied, and so is (3.11), where zj = Ta1jzj. It follows

that zj w1 6= 0 after passing to a subsequence and w1 is a homoclinic. Let v1j := zj−w1.

Then one shows that v1j is a (PS)-sequence such that Φ(v1

j )→ c−Φ(w1). This argumentis rather technical, and it is here (in the proof that Φ′(v1

j ) → 0 to be more precise) that

the condition (3.19) plays a role. Moreover, there exists α > 0 such that Φ(w) ≥ α for allcritical points w 6= 0. Indeed, otherwise we find a sequence of critical points wn 6= 0 with

Φ(wn)→ 0. But then, according to Proposition 3.3, wn → 0 which is impossible because(3.3) and the fact that ψ′(z) = o(‖z‖) as z → 0 imply w = 0 is the only critical point in

some neighbourhood of 0.Now we can repeat the same argument for v1

j and obtain after passing to a subsequence

again that v1j := Ta2

jv1j w2 and Φ(v2

j )→ c− Φ(w1)− Φ(w2), where v2j = v1

j − w2. Since

α > 0, after a finite number of steps

Φ(vkj )→ β := c−k∑

m=1

Φ(wm) ≤ 0.

But then β = 0 and vkj → 0. Since (up to a subsequence) Tbkj vkj = zj −

∑km=1 Tbmj wm,

where bmj = −(a1j + . . .+ amj ), the conclusion follows. 2

Proof of Theorem 3.9 Assuming that (3.1) has finitely many geometrically distinct

homoclinics, we shall show that the hypotheses of Theorem 1.21 are satisfied therebyobtaining a contradiction.

Since Tg ∈ O(2N)∩Sp(2N) for g ∈ G, the quadratic form (3.2) is G-invariant, hence Φis G-invariant. It has been shown in the proof of Theorem 3.6 that ψ is weakly sequentially

semicontinuous. This and Proposition 3.1 imply (i). We already know that (ii) holds.In order to verify (iii), we first note that there is an increasing sequence of G-invariant

subspaces Fn ⊂ E+ such that dimFn = n if p = 2 (in this case all subspaces are invariant)and dimFn = 2n if p > 2. Since there exists a continuous projection Lµ(R,R2N) → Fn,

we obtain

Φ(z) ≤ 1

2‖z+‖2 − 1

2‖z−‖2 − δµ−1‖z‖µµ ≤

1

2‖z+‖2 − δ0||z+‖µµ −

1

2‖z−‖2

for some δ0 > 0 and all z ∈ En. The right-hand side above tends to −∞ as ‖z‖ → ∞because dimFn <∞.

It remains to verify (iv). Choose a unique point in each Z-orbit of homoclinics anddenote the set of all such points by F . According to our assumption, F is finite. Changing

the bmj :s if necessay we may assume wm ∈ F in Proposition 3.10. For a given positiveinteger l, let

[F , l] :=

k∑

m=1

Tamwm : 1 ≤ k ≤ l, am ∈ Z, wm ∈ F.

62

Page 63: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

If I ⊂ (0,∞) is a compact interval and l is large enough, then [F , l] is a (PS)I-attractoraccording to Proposition 3.10. Finally, that

(3.20) inf‖z+ − w+‖ : z, w ∈ [F , l], z+ 6= w+ > 0

is a consequence of the result below. 2

Proposition 3.11 Let F be a finite set of points in E. Then (3.20) holds.

Here it is not assumed that F is a set whose points have any special property. The

proof is straightforward though rather tedious and may be found in [26], Proposition 1.55.In [26] this result is proved for z and w (and not z+ and w+), however, the argument is

exactly the same in our case.

Theorem 3.9, for p = 2 (even Hamiltonian), is due to Ding and Girardi [30]. They

have allowed A to be time-dependent and 0 to be the left endpoint of a gap of thespectrum of −J(d/dt) − A(t). A result similar to Theorem 3.9 but allowing much more

general (also infinite) groups has been obtained by Arioli and Szulkin [6]. Subsequentlythe superquadraticity condition in [6] has been weakened by Xu [103] by means of a

truncation argument mentioned in the preceding section.

3.4 Multibump solutions and relation to the Bernoulli shift

It is known by Melnikov’s theory that certain integrable Hamiltonian systems having 0 as

a hyperbolic point can be perturbed in such a way that the stable and unstable manifoldat 0 intersect transversally. This in turn implies that there exists a compact set, invariant

with respect to the Poincare mapping and conjugate to the Bernoulli shift (these notions

will be defined later), see Palmer [77], or [47], [100] for a more comprehensive account ofthe subject. However, in general it is not an easy task to decide whether the intersection

is transversal. In this section we shall see that sometimes under conditions which areweaker than transversality it is still possible to show the existence of an invariant set

which is semiconjugate to the Bernoulli shift.Consider the Hamiltonian system (3.1), with H satisfying (H1)-(H4) and suppose that

z0 is a homoclinic solution. Let χ ∈ C∞(R, [0, 1]) be a function such that χ(t) = 1for |t| ≤ 1/8 and suppχ ⊂ (−1/4, 1/4). If χj(t) := χ(t/j), one easily verifies that

χjz0 is a (PS)-sequence, Φ(χjz0) → Φ(z0) = c and supp(χjz0) ⊂ (−j/4, j/4). Letwj(t) := χj(t)z0(t) + χj(t − j)z0(t − j). Since χjz0 and χj(· − j)z0(· − j) have disjoint

supports, it follows that wj is a (PS)-sequence such that Φ(zj)→ 2c. One can thereforeexpect that under suitable conditions there is a large j and a homoclinic solution

z(t) = χj(t)z0(t) + χj(t− j)z0(t− j) + v(t) = z0(t) + z0(t− j) + v(t)

63

Page 64: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

such that ‖v‖∞ (and hence also ‖v‖∞) is small compared to ‖z0‖∞. We shall call this za 2-bump solution. In a similar way one can look for k-bump solutions with k > 2.

Suppose now (3.1) has a homoclinic solution z0 and for some ε reasonably small, sayε ≤ 1

2‖z0‖∞, there exists M > 0 such that for any k ∈ N and any sequence of integers

a1 < a2 < . . . < ak satisfying aj − aj−1 ≥M for all j, there is a homoclinic solution

(3.21) z(t) =k∑

j=1

z0(t− aj) + v(t),

where ‖v‖∞ ≤ ε (so z is a k-bump solution). We emphasize that M is independent

of k here. Existence of k-bump solutions for a class of superquadratic second order

Hamiltonian systems has been shown by Coti-Zelati and Rabinowitz [26]. However, in[26] M may depend on k. In [83] Sere has shown that under appropriate conditions on

H, there is M = M(ε) (independent of k) such that homoclinic solutions of the form(3.21) exist. More precisely, he has assumed that H ∈ C2(R2N ,R) satisfies (H1)-(H4), G

is convex in z and

δ1|z|s ≤ G(z, t) ≤ δ2|z|s, sG(z, t) ≤ z ·Gz(z, t)

for all z, t and some δ1, δ2 > 0, s > 2. Let s′ = s/(s− 1). Since G is convex, one can useClarke’s duality principle in order to construct a dual functional Ψ ∈ C1(Ls

′(R,R2N),R)

such that there is a one-to-one correspondence between critical points u 6= 0 of Ψ andhomoclinic solutions z of (3.1) ([25], see also [33]). The functional Ψ is better behaved

than Φ; in particular, it is not strongly indefinite and under the conditions specified aboveit has the mountain pass geometry near 0. This fact has been employed in [25] in order

to obtain a homoclinic.Let c be the mountain pass level for Ψ, or more precisely, let

c := infh∈Γ

maxτ∈[0,1]

Ψ(h(τ)),

where

Γ := h ∈ C([0, 1], Ls′(R,R2N)) : h(0) = 0, Ψ(h(1)) < 0.

Since 0 is a strict local minimum of Ψ and Ψ is unbounded below, c > 0.

Theorem 3.12 (Sere [83]) Suppose that H, G satisfy the hypotheses above and there is

c′ > c such that the set Kc′ := u ∈ Ψc′ : Ψ′(u) = 0 is countable. Then for each ε > 0there exists M = M(ε) such that to every choice of integers a1 < a2 < . . . < ak satisfying

aj − aj−1 ≥ M there corresponds a homoclinic solution z of (3.1) given by (3.21).

The countability of Kc′ is a sort of nondegeneracy condition. A similar (but stronger)

condition has also been employed in [26]. Consider the Hamiltonian system with H given

64

Page 65: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

by (3.17). If z0 is a homoclinic, u0 corresponds to z0 and c = Ψ(u0), then Kc′ is uncountable(see (3.18)). On the other hand, there are no multibump solutions in this case. Therefore

in general it is necessary to assume some kind of nondegeneracy. Note also that sinceautonomous systems are invariant with respect to time-translations by a for any a ∈ R,

the countability condition can never be satisfied in this case. However, as has been shownby Bolotin and Rabinowitz [21], autonomous systems may have multibumps.

The proof of Theorem 3.12, in particular the construction of M independent of k,is lengthy and very technical. Therefore we omit it and refer the reader to [83]. To

our knowledge no multibump results are known for first order Hamiltonian systems withnonconvex G (except when a reduction to a second order system like in [26] can be made).

From now on we assume that (3.1) has k-bump solutions of the form (3.21) for any kand M is independent of k. Choose a ≥M and let

zj(t) := z0(t− aj), j ∈ Z.

We claim that, given any sequence sj of 0’s and 1’s, there exists a solution

z(t) =∑

j∈

sjzj(t) + v(t)

of (3.1) such that ‖v‖∞ ≤ ε. Note that if sj = 1 for infinitely many j ′s, then z hasinfinitely many bumps and is not a homoclinic. By our asssumption, for any positive

integer m we can find a solution

zm(t) =m∑

j=−msjzj(t) + vm(t)

with ‖vm‖∞ ≤ ε. Since 0 is a hyperbolic point and z0(t)→ 0 as |t| → ∞, z0 decays to 0

exponentially (this follows by exponential dichotomy [77], see also [25] or [31]). Therefore‖zm‖∞ is bounded uniformly in m, and by (3.1), the same is true of ‖zm‖∞. Hence zm,

and a posteriori also vm, are uniformly bounded in H1loc(R,R2N). Since the embedding

H1loc(R,R2N) → L∞loc(R,R2N) is compact and ‖vm‖∞ ≤ ε, vm → v in L∞loc(R,R2N) for

some v with ‖v‖∞ ≤ ε. It follows that the corresponding function z is a weak solution of(3.1) and z 6= 0. Moreover, z ∈ H1

loc(R,R2N ) ∩ C2(R,R2N) (recall H is of class C2).

Let now

Σ2 := 0, 1 =s = sjj∈ : sj ∈ 0, 1

be the set of doubly infinite sequences of 0’s and 1’s, endowed with the metric

d(s, s) :=∑

j∈

2−|j||sj − sj|.

65

Page 66: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

The space (Σ2, d) is easily seen to be compact, totally disconnected and perfect (it is infact homeomorphic to the Cantor set). The mapping σ ∈ C(Σ2,Σ2) given by

(σ(s))j = sj+1

is called the Bernoulli shift on two symbols. It is often considered as a prototype of a

chaotic map. In particular, it has a countable infinity of periodic orbits, an uncountableinfinity of nonperiodic orbits, a dense orbit, and it exhibits sensitive dependence on initial

conditions. The details may be found e.g. in Wiggins [100], Chapter 2.Let

Z :=z ∈ L∞(R,R2N) : z(t) =

j∈

sjzj(t) + v(t), sj ∈ 0, 1, ‖v‖∞ ≤ ε

and Ij := [a(j − 12), a(j + 1

2)]. In Z we introduce a metric d by setting

(3.22) d(z, 0) =∑

j∈

2−|j|(sj‖zj‖∞ + ‖v‖L∞(Ij , 2N )

).

Since z0 decays exponentially, the topology induced by d coincides with the L∞loc-topologyon Z as one readily verifies. Using this, the compactness of Σ2 and of the embedding

H1loc(R,R2N) → L∞loc(R,R2N), it follows that the set

X := z ∈ Z : z is a solution of (3.1)

is compact.

As before, let Ta : X → X be the mapping given by (Taz)(t) = z(t + a) and letfa : R2N → R2N be the Poincare (or time-a) mapping defined by fa(z0) = z(a) where

z = z(t) is the unique solution of (3.1) satisfying the initial condition z(0) = z0. Finally,let Ev : X → R2N be the evaluation mapping, Ev(z) := z(0), and let I := Ev(X). Since

Ev is continuous and injective, it is a homeomorphism between X and I, and it is easyto verify that the diagram

(3.23)

Ifa−−−→ I

Ev

xxEv

XTa−−−→ X

is commutative. Note in particular that the set I is invariant with respect to the Poincaremapping fa.

66

Page 67: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

Theorem 3.13 There exists a continuous surjective mapping g : I → Σ2 such that thediagram

Ifa−−−→ I

g

yyg

Σ2σ−−−→ Σ2

is commutative.

If a continuous surjective mapping g as above exists, we shall say that fa : I → I is

semiconjugate to σ : Σ2 → Σ2, and fa will be called conjugate to σ if g is a homeomor-phism.

Proof For z =∑

j∈

sjzj + v ∈ X we define ϕ(z) = s = sjj∈ . Then ϕ is a continouos

mapping from X onto Σ2 (cf. (3.22)) and the diagram

(3.24)

XTa−−−→ X

ϕ

yyϕ

Σ2σ−−−→ Σ2

commutes. Now the conclusion follows from (3.23) and (3.24) upon setting g := ϕ(Ev)−1.

2

Remark 3.14 It follows from the definition of topological entropy h(.) (see e.g. [47],

Definition 5.8.3 or [88], Definition 5.8.4) and the uniform continuity of g that h(fa) ≥ h(σ)(cf. [88], Exercise 5.8.1.B). It is well known that h(σ) > 0. Hence fa|I , and therefore also

the time-1-mapping f1, has positive entropy. More precisely, h(σ) = log 2 and h(f1) ≥(log 2)/a according to [88], Example 5.8.1 and Theorem 5.8.4. The same conclusion about

the entropy may also be found in [83], where a different (in a sense, dual) argument hasbeen used.

The approach presented in this section is taken from a work in progress by W. Zouand the second author. They study a certain second order system and hope to show

that, in addition to the result of Theorem 3.13, to each m-periodic sequence s ∈ Σ2 therecorresponds a z ∈ X which has period ma.

References

[1] A. Abbondandolo, Morse Theory for Hamiltonian Systems, Chapman & Hall/CRC

Research Notes in Mathematics 425, Boca Raton, FL 2001.

67

Page 68: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[2] R.A. Adams, Sobolev Spaces, Academic Press, New York 1975.

[3] H. Amann and E. Zehnder, Nontrivial solutions for a class of nonresonance problems

and applications to nonlinear differential equations, Ann. Scuola Norm. Sup. Pisa (4)7 (1980), 539–603.

[4] H. Amann and E. Zehnder, Periodic solutions of asymptotically linear Hamiltonian

systems, Manuscr. Math. 32 (1980), 149–189.

[5] A. Ambrosetti and V. Coti-Zelati, Periodic Solutions of Singular Lagrangian Systems,Birkhauser 1993.

[6] G. Arioli and A. Szulkin, Homoclinic solutions of Hamiltonian systems with symme-

try, J. Diff. Eq. 158 (1999), 291–313.

[7] A. Bahri and H. Berestycki, Existence of forced oscillations for some nonlinear dif-ferential equations, Comm. Pure Appl. Math. 37 (1984), 403–442.

[8] A. Bahri and H. Berestycki, Forced vibrations of superquadratic Hamiltonian systems,

Acta Math. 152 (1984), 143–197.

[9] T. Bartsch, A simple proof of the degree formula for Z/p-equivariant maps, Math. Z.

212 (1993), 285–292.

[10] T. Bartsch, Topological Methods for Variational Problems with Symmetries, Springer-Verlag, Berlin Heidelberg 1993.

[11] T. Bartsch, A generalization of the Weinstein-Moser theorems on periodic orbits of

a Hamiltonian system near an equilibrium, Ann. Inst. H. Poincare, Anal. non lin. 14(1997), 691–718.

[12] T. Bartsch and Y.H. Ding, Homoclinic solutions of an infinite-dimensional Hamilto-

nian system, Math. Z. 240 (2002), 289–310.

[13] T. Bartsch and Y.H. Ding, Deformation theorems on non-metrizable vector spacesand applications to critical point theory, preprint.

[14] T. Bartsch and Z.-Q. Wang, Periodic solutions of even Hamiltonian systems on the

torus, Math. Z. 224 (1997), 65–76.

[15] T. Bartsch and Z.-Q. Wang, Periodic solutions of spatially periodic, even Hamiltoniansystems, J. Diff. Eq. (1997), 103–128.

[16] T. Bartsch and M. Willem, Periodic solutions of non-autonomous Hamiltonian sys-

tems with symmetries, J. Reine Angew. Math. 451 (1994), 149–159.

68

Page 69: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[17] V. Benci, A geometrical index for the group S1 and some applications to the studyof periodic solutions of ordinary differential equations, Comm. Pure Appl. Math. 34

(1981), 393–432.

[18] V. Benci, On critical point theory of indefinite functionals in the presence of symme-

tries, Trans. Amer. Math. Soc. 274 (1982), 533–572.

[19] V. Benci and P. Rabinowitz, Critical point theorems for indefinite functionals, Invent.

Math. 52 (1979), 241–273.

[20] H. Berestycki, J.M. Lasry, G. Mancini and B. Ruf, Existence of multiple periodic

orbits on star-shaped Hamiltonian surfaces, Comm. Pure Appl. Math. 38 (1985),253–289.

[21] S.V. Bolotin and P.H. Rabinowitz, A variational construction of chaotic trajectoriesfor a Hamiltonian system on a torus, Boll. Unione Mat. (8) 1 (1998), 541–570.

[22] G. Cerami, Un criterio di esistenza per i punti critici su varieta illimitate, Istit.

Lombardo Accad. Sci. Lett. Rend. A 112 (1978), 332–336.

[23] K.C. Chang, Infinite Dimensional Morse Theory and Multiple Solution Problem,

Birkhauser, Boston 1993.

[24] C. Conley and E. Zehnder, The Birkhoff-Lewis fixed point theorem and a conjecture

of V. Arnold, Invent. Math. 73 (1983), 33–49.

[25] V. Coti-Zelati, I. Ekeland and E. Sere, A variational approach to homoclinic orbits

in Hamiltonian systems, Math. Ann. 288 (1990), 133–160.

[26] V. Coti-Zelati and P.H. Rabinowitz, Homoclinic orbits for second order Hamiltonian

systems possessing superquadratic potentials, J. Amer. Math. Soc. 4 (1991), 693–727.

[27] N. Dancer and S. Rybicki, A note on periodic solutions of autonomous Hamiltonian

systems emanating from degenerate stationary solutions, Diff. Int. Eq. 12 (1999),

147–160.

[28] M. Degiovanni and L. Olian Fannio, Multiple periodic solutions of asymptotically

linear Hamiltonian systems, Nonl. Analysis 26 (1996), 1437–1446.

[29] Y.H. Ding, Infinitely many homoclinic orbits for a class of Hamiltonian systems with

symmetry, Chinese Ann. Math. Ser. B 19 (1998), 167–178.

[30] Y.H. Ding and M. Girardi, Infinitely many homoclinic orbits of a Hamiltonian system

with symmetry, Nonl. Anal. 38 (1999), 391–415.

69

Page 70: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[31] Y.H. Ding and M. Willem, Homoclinic orbits of a Hamiltonian system, Z. Angew.Math. Phys. 50 (1999), 759–778.

[32] I. Ekeland, Une theorie de Morse pour les systemes hamiltoniens convexes, Ann. Inst.H. Poincare, Anal. non lin. 1 (1984), 19–78.

[33] I Ekeland, Convexity Methods in Hamiltonian Mechanics, Springer-Verlag, Berlin1990.

[34] I. Ekeland and L. Lassoued, Multiplicite des trajectoires fermees de systemes hamil-toniens convexes, Ann. Inst. H. Poincare, Anal. non lin. 4 (1987), 307–335.

[35] I. Ekeland and J.M. Lasry, On the number of periodic trajectories for a Hamiltonianflow on a convex energy surface, Ann. Math. 112 (1980), 283–319.

[36] E.R. Fadell and P.H. Rabinowitz, Generalized cohomological index theories for Liegroup actions with an application to bifurcation questions for Hamiltonian systems,

Inv. Math. 45 (1978), 139–174.

[37] P. Felmer, Periodic solutions of spatially periodic Hamiltonian systems, J. Diff. Eq.

98 (1992), 143–168.

[38] P. Felmer, Periodic solutions of “superquadratic” Hamiltonian systems, J. Diff. Eq.

102 (1993), 188–207.

[39] P. Felmer and Z.-Q. Wang, Multiplicity for symmetric indefinite functionals and

application to Hamiltonian and elliptic systems, Top. Meth. Nonlin. Anal. 12 (1998),207–226.

[40] A. Fonda and J. Mawhin, Multiple periodic solutions of conservative systems withperiodic nonlinearity, in: Differential Equations and Applications, Vol I and II, Ohio

Univ. Pr., Athens OH, 1989.

[41] K. Fukaya and K. Ono, Arnold conjecture and Gromov-Witten invariants, Topology

38 (1999), 933–1048.

[42] G. Fournier, D. Lupo, M. Ramos and M. Willem, Limit relative category and critical

point theory, in: Dynamics Reported 3 (New Series), C.K.R.T. Jones, U. Kirchgraberand H.O. Walther eds., Springer-Verlag, Berlin 1994, pp. 1-24.

[43] V.L. Ginzburg, An embedding S2n−1 → R2n, 2n−1 ≥ 7, whose Hamiltonian flow hasno periodic trajectories, Internat. Math. Res. Notices 1995, 83–98.

[44] V.L. Ginzburg and B.Z. Gurel, A C2-smooth counterexample to the Hamiltonian

Seifert conjecture in R4, Ann. Math. 158 (2003), 953–976.

70

Page 71: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[45] V.L. Ginzburg and E. Kerman, Periodic orbits of Hamiltonian flows near symplecticextrema, Pac. J. Math. 206 (2002), 69–91.

[46] M. Girardi and M. Matzeu, Solutions of minimal period for a class of nonconvexHamiltonian systems and applications to the fixed energy problem, Nonl. Analysis 10

(1986), 371–382.

[47] J. Guckenheimer and P. Holmes, Nonlinear Oscillations, Dynamical Systems, andBifurcations of Vector Fields, Springer-Verlag, New York 1983.

[48] Y.X. Guo, Nontrivial periodic solutions for asymptotically linear Hamiltonian systems

with resonance, J. Diff. Eq. 175 (2001), 71–87.

[49] M. Herman, Examples of compact hypersurfaces in R2p, 2p ≥ 6, Preprint, 1994.

[50] H. Hofer, K. Wysocki and E. Zehnder, The dynamics on three-dimensional strictlyconvex energy surfaces, Ann. Math. 148 (1998), 197–289.

[51] H. Hofer and E. Zehnder, Periodic solutions on hypersurfaces and a result by C.

Viterbo, Inv. Math. 90 (1987), 1–9.

[52] H. Hofer and E. Zehnder, Symplectic Invariants and Hamiltonian Dynamics,

Birkhauser, Basel 1994.

[53] M. Izydorek, A cohomological Conley index in Hilbert spaces and applications tostrongly indefinite problems, J. Diff. Eq. 170 (2001), 22–50.

[54] M. Izydorek, Equivariant Conley index in Hilbert spaces and applications to strongly

indefinite problems, Nonl. Analysis 51 (2002), 33–66.

[55] L. Jeanjean, On the existence of bounded Palais-Smale sequences and application toa Landesman-Lazer type problem set on RN , Proc. Roy. Soc. Edinb. 129A (1999),

787–809.

[56] M. A. Krasnoselski, Topological Methods in the Theory of Nonlinear Integral Equa-tions, Pergamon Press, Oxford 1964.

[57] W. Kryszewski and A. Szulkin, An infinite dimensional Morse theory with applica-

tions, Trans. Amer. Math. Soc. 349 (1987), 3181–3234.

[58] W. Kryszewski and A. Szulkin, Generalized linking theorem with an application to asemilinear Schrodinger equation, Adv. in Diff. Equations 3, 441–472 (1998).

[59] G. Li and A. Szulkin, An asymptotically periodic Schrodinger equation with indefinite

linear part, Comm. Contemp. Math. 4 (2002), 763–776.

71

Page 72: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[60] S.J. Li and J.Q. Liu, Some existence theorems on multiple critical points and theirapplications, Kexue Tongbao 17 (1984), 1025–1027 (in Chinese).

[61] S.J. Li and J.Q. Liu, Morse theory and asymptotic linear Hamiltonian system, J. Diff.Eq. 78 (1989), 53–73.

[62] S.J. Li and A. Szulkin, Periodic solutions for a class of nonautonomous Hamiltoniansystems, J. Diff. Eq. 112 (1994), 226–238.

[63] Y.Q. Li, A limit index theory and its applications, Nonl. Anal. 25 (1995), 1371–1389.

[64] E.H. Lieb and M. Loss, Analysis, Amer. Math. Soc., Providence, R.I. 1997.

[65] P.L. Lions, The concentration compactness principle in the calculus of variations.The locally compact case. Part II, Ann. IHP - Analyse Nonl. 1 (1984), 223–283.

[66] J.Q. Liu, A generalized saddle point theorem, J. Diff. Eq. 82 (1989), 372–385.

[67] C.G. Liu, Y. Long and C. Zhu, Multiplicity of closed characteristics on symmetricconvex hypersurfaces in R2n, Math. Ann. 323 (2002), 201–215.

[68] Y. Long, Maslov-type index, degenerate critical points, and asymptotically linearHamiltonian systems, Science in China (Series A) 33 (1990), 1409–1419.

[69] Y. Long, Index Theory for Symplectic Paths with Applications, Birkhauser, Basel2002.

[70] Y. Long and X. Xu, Periodic solutions for a class of nonautonomous Hamiltoniansystems, Nonl. Anal. 41 (2000), 455–463.

[71] Y. Long and C. Zhu, Closed characteristics on compact convex hypersurfaces in R2n,Ann. Math. 155 (2002), 317–368.

[72] A. M. Lyapunov, Probleme general de la stabilite du mouvement, Ann. Fac. Sci.Toulouse 2 (1907), 203–474.

[73] J. Mawhin and M. Willem, Critical Point Theory and Hamiltonian Systems, Springer-

Verlag, New York 1989.

[74] D. McDuff and D. Salamon, Introduction to Symplectic Topology, Oxford University

Press, New York 1998.

[75] K.R. Meyer and G.R. Hall, Introduction to Hamiltonian Dynamical Systems and the

N-Body Problem, Springer-Verlag 1992.

72

Page 73: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[76] J. Moser, Periodic orbits near an equilibrium and a theorem by A. Weinstein, Comm.Pure Appl. Math. 29 (1976), 727–747.

[77] K.J. Palmer, Exponential dichotomies and transversal homoclinic points, J. Diff. Eq.55 (1984), 225–256.

[78] P.H. Rabinowitz, Periodic solutions of Hamiltonian systems, Comm. Pure Appl.Math. 31 (1978), 157–184.

[79] P. H. Rabinowitz, Periodic solutions of large norm of Hamiltonian systems, J. Diff.Eq. 50 (1983), 33–48.

[80] P.H. Rabinowitz, Minimax Methods in Critical Point Theory with Applications toDifferential Equations, CBMS 65, Amer. Math. Soc., Providence, R.I. 1986.

[81] P.H. Rabinowitz, Variational methods for Hamiltonian systems, in B. Hasselblattand A. Katok (eds.), Handbook of dynamical systems. Volume 1A. Elsevier 2002,

1091–1127.

[82] S. Secchi and C.A. Stuart, Global bifurcation of homoclinic solutions of Hamiltoniansystems, Discrete Cont. Dyn. Syst. 9 (2003), 1493–1518.

[83] E. Sere, Looking for the Bernoulli shift, Ann. IHP - Analyse Nonl. 10 (1993), 561–590.

[84] E.H. Spanier, Algebraic Topology, Springer-Verlag, New York 1966.

[85] M. Struwe, Variational Methods, Springer-Verlag, Berlin 1990.

[86] M. Struwe, Existence of periodic solutions of Hamiltonian systems on almost every

energy surface, Bol. Soc. Bras. Mat. 20 (1990), 49–58.

[87] C.A. Stuart, Bifurcation into spectral gaps, Bull. Belg. Math. Soc., Supplement, 1995.

[88] W. Szlenk, An Introduction to the Theory of Smooth Dynamical Systems, PolishScientific Publishers, Warsaw and Wiley, New York 1984.

[89] A. Szulkin, Morse theory and existence of periodic solutions of convex Hamiltonian

systems, Bull. Soc. Math. France 116 (1988), 171–197.

[90] A. Szulkin, A relative category and applications to critical point theory for strongly

indefinite functionals, Nonl. Anal. 15 (1990), 725–739.

[91] A. Szulkin, Bifurcation for strongly indefinite functionals and a Liapunov type theo-

rem for Hamiltonian systems, Diff. Int. Eq. 7 (1994), 217–234.

73

Page 74: staff.math.su.se · Hamiltonian systems: periodic and homoclinic solutions by variational methods Thomas Bartsch Mathematisches Institut, Universit at Giessen Arndtstr. 2, 35392 Giessen,

[92] A. Szulkin, Index theories for indefinite functionals and applications. In: Topologicaland Variational Methods for Nonlinear Boundary Value Problems, P. Drabek ed.,

Pitman Research Notes in Mathematics Series 365, Harlow, Essex 1997, pp. 89–121.

[93] A. Szulkin and W. Zou, Infinite dimensional cohomology groups and periodic solutions

of asymptotically linear Hamiltonian systems, J. Diff. Eq. 174 (2001), 369–391.

[94] A. Szulkin and W. Zou, Homoclinic orbits for asymptotically linear Hamiltonian

systems, J. Func. Anal. 187 (2001), 25–41.

[95] K. Tanaka, Homoclinic orbits in a first order superquadratic Hamiltonian system:

convergence of subharmonic orbits, J. Diff. Eq. 94 (1991), 315–339.

[96] C. Viterbo, A proof of the Weinstein conjecture in R2n, Ann. IHP - Analyse Nonl. 4

(1987), 337–356.

[97] C. Viterbo, Equivariant Morse theory for starshaped Hamiltonian systems, Trans.

Amer. Math. Soc. 311 (1989), 621–655.

[98] A. Weinstein, Normal modes for nonlinear Hamiltonian systems, Inv. Math. 20(1973), 47–57.

[99] A. Weinstein, Periodic orbits for convex Hamiltonian systems, Ann. Math. 108(1978), 507–518.

[100] S. Wiggins, Global Bifurcations and Chaos, Springer-Verlag, New York 1988.

[101] M. Willem, Lectures on Critical Point Theory, Trabalho de Mat. 250, Fundacao

Univ. Brasılia, Brasılia 1983.

[102] M. Willem, Minimax Theorems, Birkhauser, Boston 1996.

[103] X. Xu, Homoclinic orbits for first order Hamiltonian systems possessing super-quadratic potentials, Nonl. Anal. 51 (2002), 197–214.

74