Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter January 14, 2016

Periodic perturbations of Hamiltonian systems

  • Alessandro Fonda EMAIL logo , Maurizio Garrione and Paolo Gidoni

Abstract

We prove existence and multiplicity results for periodic solutions of Hamiltonian systems, by the use of a higher dimensional version of the Poincaré–Birkhoff fixed point theorem. The first part of the paper deals with periodic perturbations of a completely integrable system, while in the second part we focus on some suitable global conditions, so to deal with weakly coupled systems.

MSC 2010: 34C25; 47H15

1 Introduction

This paper provides some results on the existence of periodic solutions for Hamiltonian systems which may be considered as time-periodic perturbations of an autonomous system of the type

(HS)𝒥z˙=(z).

Here, :2N is a continuously differentiable function and 𝒥 is the standard symplectic matrix, i.e.,

𝒥=(0-𝕀N𝕀N0).

There is a large bibliography on this problem, mainly motivated by models from classical mechanics. Remarkably, we observe that the literature on this issue can be split into two quite disjoint streams.

One of the two currents has a more topological spirit, and aims at minimal regularity assumptions. This kind of results finds a fertile ground mainly in the planar case, where, among others, a powerful tool, the Poincaré–Birkhoff fixed point theorem, can be used to prove the existence of periodic solutions of the perturbed system. Indeed, the case N=1 is privileged by the fact that, for an autonomous planar system like (HS), any periodic orbit is always surrounded by an annulus of periodic orbits. Then, assuming that the periods of the corresponding solutions do not remain the same, the needed twist condition is naturally obtained, and the Poincaré–Birkhoff theorem applies (see, e.g., [29] and the references therein).

On the other hand, when N2, a more analytical approach has usually been followed, requiring some additional structural assumptions on the unperturbed system (HS). Usually, the system is assumed to be completely integrable, and more regularity is asked for the Hamiltonian function. Moreover, some rather restrictive nondegeneracy conditions are needed so to obtain the existence of periodic solutions of the perturbed system (see, e.g., [4, 9]).

The aim of this paper is to provide a common framework for the two kinds of approach depicted above. Using a recent result by the first author and Ureña [31], where an extension of the Poincaré–Birkhoff theorem to higher dimensional Hamiltonian systems has been proposed, we will be able, on one hand, to relax the usual structural assumptions on the Hamiltonian function and, on the other hand, to extend to higher dimensions some existence results already established in the planar case.

Before entering into details, we will now spend a few words on the framework where our results are to be settled.

The general framework

A classical approach to the study of the Hamiltonian system (HS) is the search for constants of motion, since they can be used for suitably transforming the system into a simpler one. The most remarkable case occurs when (HS) has N constants of motion which are independent and in involution: In this case, the system is said to be completely integrable, and one has a foliation of the space in N-dimensional surfaces, which are invariant for the flow.

The Liouville–Arnold theorem then assures that, when one of these surfaces is bounded and connected, it has to be an N-dimensional torus. Moreover, for any such invariant torus Γ, there exists an open neighborhood 𝒜 of Γ and a canonical transformation z=(x,y)(φ,I), mapping 𝒜 onto 𝕋N×𝒟 (where 𝕋=/2π and 𝒟 is an open subset of N), and reducing the Hamiltonian function to the simpler form (φ,I)=𝒦(I). The coordinates I=(I1,,IN)𝒟 are usually known as action variables, whereas the coordinates φ=(φ1,,φN)𝕋N are called angle variables.

Each value I=I0 is associated with an invariant torus Γ0=𝕋N×{I0}, where the dynamics of the system is completely described by the frequency vector ω0=𝒦(I0). When the components ω10,,ωN0 are rationally independent, the solutions are quasiperiodic and each orbit is a dense subset of the N-torus Γ0. Such tori are called nonresonant. Otherwise, we have a foliation in M-dimensional tori, where M<N is the rational rank of the components of ω0, and the orbits will be quasiperiodic with respect to these lower dimensional tori. A special case occurs when the components of ω0 are all pairwise commensurable. Then, all the solutions on the torus are periodic with the same period, and the N-torus Γ0 admits a foliation in invariant 1-tori, each one defined by the orbit of a solution.

Since for every general Hamiltonian system (HS) a constant of motion is always given by the Hamiltonian function , we immediately deduce that every planar Hamiltonian system is completely integrable. In higher dimensions, a classical example of a completely integrable system comes from the Kepler two-body problem, or even from every central force field [38]. On the contrary, if more than two bodies are involved, the system is not completely integrable any more. However, assuming the masses of the “planets” to be small compared to the mass of the “Sun”, the system may be seen as being decomposed in n independent two-body systems, with the addition of a small perturbative term accounting for the other interactions (cf. [18, 19] and references therein). Such problems of Celestial Mechanics have probably been the main stimulus in the development of integrability and of Hamiltonian perturbation theory.

As a matter of fact, completely integrable Hamiltonian systems are rare, and most often the Hamiltonian function is their unique constant of motion [8, 50]. Yet, generic Hamiltonian systems may be considered as perturbations of completely integrable systems [42, 47], usually called nearly integrable systems. A glance of this scenario was already grasped by Henri Poincaré [46], who referred to Hamiltonian perturbation theory as the Problème général de la Dynamique. The efforts made by Poincaré and, among many others, by Birkhoff, led to a broad development of the theory. We suggest [5, 7] for a detailed introduction to Hamiltonian perturbation theory and [26] for a friendly overview.

As we have seen, complete integrability reveals strong properties of the dynamics. A natural question is: How much of this structure is preserved under a small perturbation? In particular, one could wonder whether, near an invariant torus of the unperturbed system, it is possible to find periodic or quasiperiodic solutions for the perturbed system with the same frequency.

A series of positive results are known for a large family of nonresonant tori, those with a Diophantine frequency. These results are usually collected under the name of KAM theory, recalling its main contributors Kolmogorov, Arnold and Moser. We remark that, beyond a nondegeneracy assumption on the torus, strong smoothness of the perturbation is always needed, cf. [2, 35, 48]. While these strongly nonresonant tori survive under small perturbations, the same is not true for the other tori [11, 41, 51] and, in particular, for those made of periodic solutions. Still, some traces of these tori can be found.

For instance, in the planar case, after the pioneering papers [39, 40], the survival of two periodic solutions was obtained as a consequence of the Poincaré–Birkhoff theorem (see, e.g., [17] and [29], where an overview on the use of the Poincaré–Birkhoff theorem for this kind of problems can be found). The required twist condition is satisfied, in this case, under some rather weak nondegeneracy assumptions. A fainter kind of traces of an invariant torus is provided by the so called Aubry–Mather theory (cf. [43] and the references therein), showing the existence of a Cantor set, called cantorus, that preserves, in a generalized sense, the rotational properties of the original torus.

For higher dimensional Hamiltonian systems, a local approach to the problem has been proposed by Bernstein and Katok [9], who showed the survival under small perturbations of N+1 periodic solutions, requiring a convexity assumption on the Hamiltonian function (see also [4, 27, 52]). This result has been later refined by Chen [21], who replaced the convexity by a classical nondegeneracy assumption.

A rather different type of problem arises when one looks for the existence and multiplicity of periodic solutions when only the global behavior of the nonlinearity is assumed to be known. In this case, the approach is no longer perturbative, and it usually combines topological and variational methods.

In this respect, there is a large literature in the planar case, mainly motivated by some models involving scalar second order differential equations, where the Poincaré–Birkhoff theorem has been successfully applied (see, e.g., [15, 25, 30, 34, 36], or again the review in [29]). The twist condition is generated by assuming a difference between the growth of the nonlinearity near a given periodic solution and at infinity, producing a gap in the rotation numbers of the corresponding solutions in the phase plane. A sharp use of the Poincaré–Birkhoff theorem then ensures that the larger this gap, the larger the number of solutions found. Furthermore, the same strategy applies also to the search for subharmonic solutions (see, for instance, [14, 25]).

Incidentally, the twist geometry has sometimes been recovered by detecting, in the unperturbed system, an annulus of periodic orbits displaying a gap between the periods of the boundary orbits. This picture displays the same features already discussed when considering completely integrable systems. A quite common way of producing this geometry is to require the strict monotonicity of the period function associated with system (HS), a feature which has been studied by many authors (see, e.g., [22, 33, 45]) and which ensures its nondegeneracy.

The first multiplicity results extending the Poincaré–Birkhoff philosophy to higher dimensions are due to Amann and Zehnder [3], who introduced a twist condition between zero and infinity. A different perspective was followed by Conley and Zehnder [24], where the existence of N+1 periodic solutions was proved for systems whose Hamiltonian function is 2π-periodic in the first N variables, and asymptotically quadratic in the other N ones. These pioneering results have been generalized in several directions, in a long series of papers (cf. [1, 20] and the references therein). See also [16, 44], where a further extension of the Poincaré–Birkhoff theorem in higher dimensions involving a monotone twist has been exploited.

The main tool and an overview of our results

Let us now recall the result in [31], which will be our main tool in the search for periodic solutions. Consider the Hamiltonian system

(1.1)𝒥ζ˙=ζ(t,ζ),

where the continuous function :×N×N is also continuously differentiable in ζ=(ξ,η)2N. Writing ξ=(ξ1,,ξN) and η=(η1,,ηN), the Hamiltonian function is assumed to be T-periodic in t, and 2π-periodic in each variable ξ1,,ξN. Let DN be a convex body, i.e., a compact, convex set with nonempty interior. For every y¯D, we denote the normal cone by

𝒩D(y¯)={vN:v,y-y¯0 for every yD}.

Moreover, let 𝔹 be an invertible symmetric matrix.

Theorem 1.1

Theorem 1.1 ([31])

If every solution ζ(t)=(ξ(t),η(t)) of (1.1) departing with η(0)D is defined for every t[0,T] and satisfies

(1.2)ξ(T)-ξ(0),𝔹v>0for every v𝒩D(η(0)){0},

then system (1.1) has at least N+1 geometrically distinct T-periodic solutions

ζ1(t)=(ξ1(t),η1(t)),,ζN+1(t)=(ξN+1(t),ηN+1(t))

such that ηk(0)D for every k=1,,N+1.

We recall that two solutions of system (1.1) are geometrically distinct if one of them cannot be obtained just by adding suitable integer multiples of 2π to some components ξi(t) of the other one.

We now briefly describe the main results of this paper, obtained by the use of Theorem 1.1.

The first part deals with small time-dependent perturbations of completely integrable systems. In Section 2, taking an invariant torus made of periodic solutions of the unperturbed system, and assuming a rather weak nondegeneracy condition, we prove the survival of N+1 periodic solutions for the perturbed system. Our main theorem thus improves some previous results of Bernstein and Katok [9] and Chen [21] in two directions: First, the Hamiltonian function is assumed to be only once continuously differentiable and, second, our nondegeneracy assumption does not even imply the invertibility of the frequency function. Moreover, it is shown that the nondegeneracy extends also to nearby tori, so that other families of periodic solutions can coappear.

In Section 3, still dealing with completely integrable systems, we gradually abandon the local point of view and move to a large scale perspective. Assuming a twist-type condition on the product of N planar annuli, which is shown to persist for small perturbations, we thus obtain the survival of N+1 periodic solutions, generalizing the planar result in [29].

In Section 4, we deal with weakly coupled systems with a T-periodic forcing term, depending on some parameters. We impose suitable conditions at zero and infinity for each of the N equations, producing a gap in the rotation numbers of the uncoupled systems. Using Theorem 1.1, we then prove the existence of N+1 periodic solutions having period T, and a number of subharmonic solutions which increases with the width of the gap. As an application, we can deal with weakly coupled systems of pendulum-like equations, generalizing the main result in [32].

Notation

In all the paper, , denotes the Euclidean scalar product in N, with its associated norm . We write (x0,r) for the open ball centered at x0 with radius r>0, and [x0,r] for the closed ball.

2 Periodic perturbations of completely integrable systems

Let us consider a completely integrable Hamiltonian system on 𝕋N×𝒟, where 𝕋N is the N-dimensional torus (/2π)N, and 𝒟 is an open subset of N. The continuously differentiable Hamiltonian function :𝕋N×𝒟 can be written in the form (φ,I)=𝒦(I). We recall that I=(I1,,IN)𝒟 are the action variables, while φ=(φ1,,φN)𝕋N are the angle variables.

For every I*𝒟, the torus 𝒯*=𝕋N×{I*} is invariant for the flow, and its evolution in time is determined by the associated frequency vector

ω*=(ω1*,,ωN*)=𝒦(I*).

We are interested in the case when the dynamics on the torus 𝒯* consists of a family of periodic orbits with minimal period T*. This happens if and only if there exist N integers a1,,aN such that

T*ωi*=2πai for every i=1,,N,

and T* is the minimum positive real number with such a property. The integers ai count the number of rotations made by each periodic solution around the i-th component of the torus in a period T*; the sign of ai describes the sense of rotation.

A standard approach to study such a system, defined on 𝕋N×𝒟, is to consider its canonical lift to N×𝒟. The Hamiltonian system then becomes

(CI){ξ˙=𝒦(η),η˙=0,

where ξ=(ξ1,,ξN)N and η=(η1,,ηN)𝒟. To be more precise, denoting by 𝕀N the identity on N and by PN:N𝕋N the standard projection on the torus, the map (PN,𝕀N):N×N𝕋N×N is a local change of variables which transforms (ξ,η) into (φ,I). Each translation of 2π in the ξi coordinate for system (CI) corresponds to a single rotation in the φi coordinate for the original system.

Let us now consider a general nearly integrable Hamiltonian system on 𝕋N×𝒟, with time-dependent Hamiltonian function 𝒦:×𝕋N×𝒟, sufficiently close to 𝒦. The canonical lift then leads to the Hamiltonian system on N×𝒟 given by

(CIper){ξ˙=ηK(t,ξ,η),η˙=-ξK(t,ξ,η).

The Hamiltonian function K:×N×𝒟 is assumed to be continuous, T-periodic in the first variable, 2π-periodic in each variable ξi, and continuously differentiable in ζ=(ξ,η).

We now fix an I0𝒟 and introduce some kind of nondegeneracy condition at I0. Usually, in the literature (see, e.g., [4, 9, 21]), it is assumed that 𝒦 is twice continuously differentiable, and that

(2.1)det(𝒦′′(I0))0.

Here, we only ask 𝒦 to be once continuously differentiable, and that there exists an invertible symmetric N×N matrix 𝔹 such that

(2.2)0cl{ρ]0,+[:minI-I0=ρ𝒦(I)-𝒦(I0),𝔹(I-I0)>0},

where clA denotes the closure of a set A. Notice that (2.1) implies (2.2), taking 𝔹=𝒦′′(I0). On the other hand, the function 𝒦(I)=I-I0α satisfies (2.2) with 𝔹=𝕀, but not (2.1) if α>2. Moreover, we observe that (2.2) does not even require the local invertibility of 𝒦. An easy example, with N=1, is provided by the function 𝒦(I)=0If(s)ds with

f(s)={ω0+|s|sin(1s)if s0,ω0if s=0.

Clearly, this function 𝒦 is only once continuously differentiable at I0=0, and 𝒦=f is not invertible, but our nondegeneracy condition (2.2) is still satisfied, with 𝔹 being the identity on .

We will show that the nondegeneracy condition (2.2) extends by continuity to a neighborhood 𝒰 of I0. As a consequence, we will prove that for every I*𝒰 as above, if there exist two positive integers m* and n* satisfying

(2.3)T*=m*Tn*,

then the perturbed system (CIper) has at least N+1 geometrically distinct m*T-periodic solutions. These solutions stay near the corresponding solutions of the unperturbed problem, and their projections on 𝕋N×𝒟 will maintain the same rotational properties of 𝒯*.

Here is our main result.

Theorem 2.1

Suppose that there exist I0D and an invertible symmetric N×N matrix B such that (2.2) holds. Then, for every σ>0 there exists an open neighborhood UD of I0, with the following property: Given any positive integer m¯, there exists ε>0 such that if

(2.4)ξK(t,ξ,η)+ηK(t,ξ,η)-𝒦(η)<εfor every (t,ξ,η)[0,T]×[0,2π]N×𝒟,

then for every I*U being associated with an invariant torus of periodic solutions for (CI) with frequency vector ω*=(ω1*,,ωN*) and minimal period T* satisfying (2.3) for suitable positive integers m*m¯ and n*, system (CIper) has at least N+1 geometrically distinct m*T-periodic solutions

(ξ1(t),η1(t)),,(ξN+1(t),ηN+1(t))

with

(2.5)ξk(t)-ξk(0)-t𝒦(I*)+ηk(t)-I*σ

for every t[0,m*T] and k=1,,N+1. Moreover, for each solution (ξk(t),ηk(t)), its projection on TN×D makes exactly (ωi*/2π)m*T rotations around the i-th component of the torus in a period m*T for every i=1,,N.

Proof.

We can assume, without loss of generality, the function 𝒦 to be defined on the whole space N. Indeed, after replacing the set 𝒟 by a smaller open set, containing I0, where 𝒦 is bounded, we can construct a continuously differentiable extension of 𝒦 on N. The solutions we are interested in will nevertheless be contained in the smaller set, where 𝒦 has not been modified. Similarly, for our purposes we can assume without loss of generality that the Hamiltonian system (CIper) is defined on ×N×N.

Let us fix any σ>0 such that [I0,σ]𝒟. By assumption (2.2), there exist >0 and ρ1]0,σ/4] such that

η-I0=ρ1𝒦(η)-𝒦(I0),𝔹(η-I0)4.

By continuity, there is an open neighborhood 𝒰 of I0, contained in [I0,ρ1], such that for every I*𝒰,

(2.6)η-I*=ρ1𝒦(η)-𝒦(I*),𝔹(η-I*)2.

For any arbitrary I*𝒰, with frequency vector ω*=(ω1*,,ωN*)=𝒦(I*), let us define

K*(t,ξ,η)=K(t,ξ+ω*t,η)-ω*,η,

and consider the Hamiltonian system

(2.7)𝒥ζ˙=ζK*(t,ζ).
Claim

For any fixed positive real numbers m¯ and c¯, there exists ε>0 such that if (2.4) holds, then for every I*𝒰, every solution ζ(t)=(ξ(t),η(t)) of (2.7) with initial point satisfying η(0)-I*ρ1 will be such that

(2.8)ξ(t)-ξ(0)-t[𝒦(η(0))-ω*]+η(t)-η(0)c¯for every t[0,m¯T].

Proof of the claim.

Arguing by contradiction, assume that there is a sequence (Iλ*)λ𝒰 with ωλ*=𝒦(Iλ*), and a sequence (Kλ)λ of Hamiltonian functions as above (in particular, they are T-periodic in t), such that, writing

Kλ*(t,ξ,η)=Kλ(t,ξ+ωλ*t,η)-ωλ*,η,

one has that

ξKλ*(t,ξ,η)+ηKλ*(t,ξ,η)-𝒦(η)+ωλ*1λfor every (t,ξ,η)×N×𝒟,

and, accordingly, a sequence (ζλ)λ with ζλ=(ξλ,ηλ), solving 𝒥ζ˙λ=ζKλ*(t,ζλ), such that ηλ(0)-Iλ*ρ1, while (2.8) does not hold, i.e., for every λ there exists tλ[0,m¯T] for which

(2.9)ξλ(tλ)-ξλ(0)-tλ[𝒦(ηλ(0))-ωλ*]+ηλ(tλ)-ηλ(0)>c¯.

Since the Hamiltonians Kλ* are 2π-periodic in the variables ξ1,,ξN, we can assume that ξλ(0)[0,2π]N. Hence, passing to a subsequence, ζλ(0) converges to some point ζ[0,2π]N×[I0,2ρ1]. Moreover, for a subsequence, Iλ* converges to some I, and ωλ*=𝒦(Iλ*) converges to ω=𝒦(I). Finally, for a subsequence, tλ will converge to some t[0,m¯T]. By a lemma of Kamke (cf. [49]), for a further subsequence (ζλl)l we have uniform convergence on [0,m¯T] to the solution of

{ξ˙=𝒦(η)-ω,η˙=0,

given by

{ξ(t)=ξ(0)+t(𝒦(η(0))-ω),η(t)=η(0).

On the other hand, passing to the limit in (2.9) yields

ξ(t)-ξ(0)-t[𝒦(η(0))-ω]+η(t)-η(0)c¯>0,

which is a contradiction, since the left-hand side is equal to zero. The claim is thus proved. ∎

We can now conclude the proof of Theorem 2.1. Let m¯ be a fixed positive integer, and choose c¯ such that

c¯min{T𝔹ρ1,σ4}.

We now focus our attention on those I*𝒰 whose associated invariant torus is composed of periodic solutions for (CI) with minimal period T*, such that there exist two positive integers m* and n* with m*m¯ and T*=m*T/n*. We observe that every m*T-periodic solution of (2.7) corresponds to an m*T-periodic solution (ξ(t),η(t)) of (CIper), such that every ξi(t) makes exactly (ωi*/2π)m*T turns around the origin in the time m*T. We will apply Theorem 1.1 to system (2.7).

Let D=[I*,ρ1], and let ζ(t)=(ξ(t),η(t)) be a solution of (2.7) with η(0)D, i.e., η(0)-I*=ρ1. Then, by (2.6) and (2.8), we get

ξ(m*T)-ξ(0),𝔹(η(0)-I*)=ξ(m*T)-ξ(0)-m*T[𝒦(η(0))-𝒦(I*)],𝔹(η(0)-I*)
+m*T[𝒦(η(0))-𝒦(I*)],𝔹(η(0)-I*)
-T𝔹ρ1𝔹ρ1+2m*Tm*T>0.

We can therefore apply Theorem 1.1, so to get N+1 geometrically distinct m*T-periodic solutions of (2.7),

ζ1(t)=(ξ1(t),η1(t)),,ζN+1(t)=(ξN+1(t),ηN+1(t)),

such that ηk(0)D for every k=1,,N+1. Moreover, by (2.8), we have that ηk(t)-I*c¯σ/2 for every t[0,m*T]. On the other hand, a continuity argument can be used, taking smaller values for c¯ and ε, to infer that ξk(t)-ξk(0)-t𝒦(I*)σ/2 for every t[0,m*T]. So, (2.5) holds, as well, and the proof is thus completed. ∎

Notice that, taking m¯ sufficiently large, it is possible to find an arbitrarily large number of values I*𝒰 for which the assumptions of Theorem 2.1 are satisfied, thus assuring the survival of N+1 subharmonic solutions from each of the corresponding invariant tori. This scenario may be compared with Birkhoff–Lewis type results [10, 13, 23], showing the existence of a family of periodic solutions with large period, accumulating towards an elliptic equilibrium. Such behavior has been observed also in the framework of Hamiltonian PDEs [6, 12].

A simple case is given by the choice I*=I0, when I0 is associated with an invariant torus 𝒯0 of periodic solutions for (CI) with frequency vector ω0 and minimal period T0.

Corollary 2.2

Suppose that there exists I0D and an invertible symmetric N×N matrix B such that (2.2) holds, and that there exist two positive integers m0 and n0 satisfying T0=m0T/n0. Then, for every σ>0 there exists ε>0 such that if

ξK(t,ξ,η)+ηK(t,ξ,η)-𝒦(η)<εfor every (t,ξ,η)[0,T]×[0,2π]N×𝒟,

then system (CIper) has at least N+1 geometrically distinct m0T-periodic solutions

(ξ1(t),η1(t)),,(ξN+1(t),ηN+1(t))

with the same rotational properties of the torus T0 and such that

ξk(t)-ξk(0)-t𝒦(I0)+ηk(t)-I0σ

for every t[0,m0T] and k=1,,N+1.

3 Twist conditions for weakly coupled period annuli

In the previous section, we have described the local phenomenon of the survival of some periodic solutions of system (CI) for the perturbed system (CIper). We now turn our attention to finding some conditions at a larger scale which guarantee the existence of multiple periodic solutions.

We still consider system (CIper) as a perturbation of system (CI), but we now look for periodic solutions (ξ(t),η(t)) starting with η(0) in some rectangle

D=[α1,β1]××[αN,βN],

contained in 𝒟. We denote the faces of this rectangle by

i-={ηD:ηi=αi},i+={ηD:ηi=βi}.
Theorem 3.1

Suppose that there exist N couples of real numbers ωi-<ωi+ such that for every i=1,,N, either

(3.1)𝒦ηi(η){ωi+for every ηi-,ωi-for every ηi+,
(3.2)𝒦ηi(η){ωi-for every ηi-,ωi+for every ηi+.

Let ω*=(ω1*,,ωN*) be the frequency vector associated with a torus T* of periodic solutions of system (CI), with minimal period T*. If

ω*Ω=]ω1-,ω1+[××]ωN-,ωN+[,

and there exist two positive integers m* and n* such that (2.3) holds, then there exists ε>0 such that every perturbed system (CIper) satisfying (2.4) has at least N+1 geometrically distinct m*T-periodic solutions

(ξ1(t),η1(t)),,(ξN+1(t),ηN+1(t)),

preserving the same rotational properties of T*.

Proof.

By the Poincaré–Miranda theorem (cf. [28, 37]), there exists an I*D such that ω*=𝒦(I*). We consider the Hamiltonian system

(3.3)𝒥ζ˙=ζK*(t,ζ)

with K*(t,ξ,η)=K(t,ξ+ω*t,η)-ω*,η.

Let us pick any ρ>0 such that

ρ<dist(D,N𝒟)andρ<m*Tdist(ω*,NΩ).

By the same argument used in the claim within the proof of Theorem 2.1, there exists ε1>0 such that if (2.4) holds with ε]0,ε1[, then every solution ζ(t)=(ξ(t),η(t)) of (3.3) with initial point η(0)D remains in N×𝒟 for t[0,m*T], and satisfies

ξ(t)-ξ(0)-t[𝒦(η(0))-ω*]+η(t)-η(0)<ρ

for every t[0,m*T]. Assume that η(0)D; we analyze four different cases.

If ηi(0)=αi for some i{1,,N}, and condition (3.1) holds, then

ξi(m*T)-ξi(0)>m*T[ωi+-ωi*]-ρ>0.

The same is true if ηi(0)=βi and (3.2) holds.

If ηi(0)=αi and condition (3.2) holds, then

ξi(m*T)-ξi(0)<m*T[ωi--ωi*]+ρ<0,

and the same is true if ηi(0)=βi and (3.1) holds.

Let us define the N×N diagonal matrix 𝔹 with, for each i=1,,N, 𝔹ii=-1 when (3.1) holds, and 𝔹ii=+1 when (3.2) is true. The estimates above ensure us that system (3.3) satisfies all the assumptions of Theorem 1.1, and the conclusion easily follows. ∎

Let us now describe a particular situation when Theorem 3.1 can be applied, generalizing the planar setting studied in [29]. We start by considering the autonomous Hamiltonian system

(3.4)𝒥z˙=(z),

where :2N is a continuously differentiable function of the special form

(x,y)=1(x1,y1)++N(xN,yN)

with x=(x1,,xN)N and y=(y1,,yN)N. Here we have used the notation z=(x,y).

Hence, for every i=1,,N, the functions i:2 are planar Hamiltonians, and we can consider the corresponding Hamiltonian systems

(${\mathrm{HS}_{i}}$)x˙i=yii(xi,yi),y˙i=-xii(xi,yi),

for each of which we assume the following:

  1. The planar system (${\mathrm{HS}_{i}}$) has a periodic solution (x¯i(t),y¯i(t)), which is non-constant and has minimal period T¯i>0.

  2. Each of such solutions has a corresponding planar open tubular neighborhood 𝒜i such that all the solutions of (${\mathrm{HS}_{i}}$) with initial point in 𝒜i are periodic, and their orbits are not contractible in 𝒜i.

  3. There exist two positive real numbers Ti-,Ti+, with Ti-<T¯i<Ti+, such that the periods of the solutions in 𝒜i cover the interval [Ti-,Ti+].

Let us define the set

𝒜={(x,y)2N:(xi,yi)𝒜i for every i=1,,N},

and consider the Hamiltonian system

(HSper)𝒥z˙=zH(t,z),

where H:×𝒜 is continuous, T-periodic in its first variable for some T>0, and has a continuous gradient with respect to its second variable z=(x,y).

For every i=1,,N, let us pick Ti]T-i,T+i[ for which there exist two positive integers mi,ni such that

Ti=miTni.

Denoting by a1,,aN the minimal positive integers such that

a1m1n1==aNmNnN,

we set

T*=a1T1==aNTN,

and define the frequency vector

ω*=2πT*(a1,,aN).

Moreover, we choose the two least positive integers m*, n* such that

T*=m*Tn*.
Theorem 3.2

In the above setting, there exists ε>0 such that every perturbed system (HSper), satisfying

(3.5)zH(t,z)-(z)<εfor every (t,z)[0,T]×𝒜,

has at least N+1 distinct m*T-periodic solutions

z1(t),,zN+1(t),

whose orbits lie in A. Moreover, for each solution zk(t), the number of rotations of the i-th component zik(t) along the annulus Ai in a period m*T is exactly equal to n*ai for every i=1,,N.

Proof.

By standard arguments (cf. [29]), each of the systems (${\mathrm{HS}_{i}}$) admits a canonical transformation in action-angle coordinates (φi,Ii). Without loss of generality we can assume that φ˙i(t)>0 for every t. The product of all such transformations is canonical, it reduces system (3.4) to the form (CI), and maps the set 𝒜 onto 𝕋N×𝒟, where 𝒟N is a product of open intervals.

For each i=1,,N, we define αi and βi as the values of the Ii-coordinate associated with two solutions of (${\mathrm{HS}_{i}}$) having periods Ti- and Ti+, in such a way that αi<βi, and we set

ωi-=2πTi+,ωi+=2πTi-.

Theorem 3.1 then applies, and the proof is readily completed. ∎

4 Weakly coupled pendulum-like systems

In this section, we consider a weakly coupled system of the type

(P){Jz˙1=A1H1(z1)+1(t,z1,,zN),Jz˙N=ANHN(zN)+N(t,z1,,zN),

where J is the 2×2 standard symplectic matrix, namely

J=(0-110),

and A1,,AN are positive real parameters. For every i=1,,N, we assume that Hi:2 is continuously differentiable, and i:×2N is continuous, T-periodic in t and continuously differentiable in (z1,,zN).

We assume that system (P) can be reduced to a Hamiltonian system by a linear change of variables. More precisely, there exist N invertible 2×2 matrices 𝕄1,,𝕄N, having positive determinant, such that the linear operator :2N2N, defined as

(4.1):(z1,,zN)(𝕄1z1,,𝕄NzN),

transforms system (P) into a Hamiltonian system. With such an assumption, we will say that (P) is a positive transformation of a Hamiltonian system.

Let us introduce the following notation for a closed cone in 2 determined by two angles ϑ1<ϑ2:

Θ(ϑ1,ϑ2)={(ρcosϑ,ρsinϑ):ρ0,ϑ1ϑϑ2}.

We are now ready to state the main theorem of this section.

Theorem 4.1

Let (P) be a positive transformation of a Hamiltonian system. For every i=1,,N, let the following assumptions hold:

  1. There exists Ci>0 such that

    Hi(w)Ci(w+1)for every w2.
  2. There exist ri>0 and mi>0 such that

    Hi(w),wmiw2 for every w[0,ri].
  3. For every σ>0, there exist Ri>0 and ϑ1i<ϑ2i, with ϑ2i-ϑ1i2π, such that

    (4.2)sup{Hi(w),ww2:wΘ(ϑ1i,ϑ2i)(0,Ri)}σ(ϑ2i-ϑ1i).

Then, for every fixed positive integers ν1,,νN, there exist A>0 and ε>0 such that if AiA and

(4.3)i(t,w1,,wN)ε for every t[0,T] and w1,,wN2

for every i=1,,N, then system (P) has at least N+1 distinct T-periodic solutions

zk(t)=(z1k(t),,zNk(t))

such that for every k=1,,N+1, each planar component zik(t), with i=1,,N, makes exactly νi clockwise rotations around the origin in the time interval [0,T[.

Some comments on the hypotheses of Theorem 4.1 are in order. Assumption (A1) is needed to ensure the global existence of the solutions to the Cauchy problems associated with (P). Concerning (A2), it will guarantee that the small amplitude planar components of the solutions do rotate around the origin, clockwise, with a least positive angular speed. Our hypothesis (A3), on the contrary, will ensure a small rotation number for large amplitude components. It could be compared with assumption (H) in [14, Theorem 4.1].

We now start the proof of Theorem 4.1. For a solution z(t) of system (P), whose i-th component is such that zi(t)=(xi(t),yi(t))2{0} for every t[0,T], we denote by Rot(zi(t);[0,T]) the standard clockwise winding number of the path tzi(t) around the origin, namely

Rot(zi(t);[0,T])=12π0TJz˙i(t),zi(t)zi(t)2dt.

Our first lemma concerns solutions z(t) whose i-th component zi(t) is small. We assume without loss of generality that Hi(0)=0, and consider the level set

Γih={w2:Hi(w)=h}.

By (A2), if h>0 is sufficiently small, then Γih is a strictly star-shaped Jordan curve around the origin. We will denote by Dih the bounded, closed and connected region of 2 with Dih=Γih.

Lemma 4.2

For any i=1,,N and every positive integer νi, if (A1) and (A2) hold, there exist three positive constants A¯i, ε¯i and h¯i such that, if AiA¯i, h]0,h¯i] and

(4.4)i(t,w1,,wN)ε¯ifor every t[0,T] and w1,,wN2,

then any solution z(t) of (P) with zi(0)Γih satisfies

Rot(zi(t);[0,T])>νi.

Proof.

Let i{1,,N} and νi be fixed. We can choose h>0 and r^]0,ri[, where ri is as in assumption (A2), in such a way that

(4.5)(0,r^)DihDi2hDi3h(0,ri).

We now claim that if (4.4) holds with a suitable choice of ε¯i, then for every solution z(t) of (P) with zi(0)Γi2h one has

h<Hi(zi(t))<3hfor every t[0,T].

Indeed, set

C=max{Hi(w):w[0,ri]},ε¯i=h2CT,

and assume by contradiction that zi(0)Γi2h and there exists t1[0,T] such that h<Hi(zi(t))<3h for every t[0,t1[, and either Hi(zi(t1))=h or Hi(zi(t1))=3h. In view of (4.5),

|ddtHi(zi(t))|=|JHi(zi(t)),AiHi(zi(t))+i(t,z1,,zN)|
=|JHi(zi(t)),i(t,z1,,zN)|Cε¯i=h2T

for every t[0,t1], so that

|Hi(zi(t1))-Hi(zi(0))|h2Tt1<h,

a contradiction.

Consequently, if zi(0)Γi2h, we have that

r^<zi(t)rifor every t[0,T],

so that the rotation number of zi(t) around the origin is well defined. Writing zi(t) in polar coordinates, namely

zi(t)=(ρi(t)cosϑi(t),ρi(t)sinϑi(t)),

using (A2) and (4.4), we thus have

-ϑi(t)=Jz˙i(t),zi(t)zi(t)2=AiHi(zi(t))+i(t,z1,,zN),zi(t)zi(t)2Aimi-ε¯ir^.

Choosing finally

A¯i=2πr^νi+ε¯iTmir^T,

we easily conclude the proof. ∎

Now we need a control on the rotation number of the large planar components of the solutions.

Lemma 4.3

For any i=1,,N, let A¯i and ε¯i be as in Lemma 4.2, and assume that AiA¯i and (4.4) holds. Then, there exists R¯i>0 such that any solution z(t) of (P) with zi(0)R¯i satisfies

Rot(zi(t);[0,T])<1.

Proof.

Fix σ=1/(2AiT) and let Ri>0 and ϑ1i<ϑ2i, with ϑ2i-ϑ1i2π, be as in (A3). Choose R^iRi such that

R^i>2ε¯iTϑ2i-ϑ1i.

In view of assumption (A1), there exists R¯iR^i such that if zi(0)R¯i, then zi(t)R^i for every t[0,T]. In particular, the rotation number of zi(t) is well defined. Let us assume, by contradiction, that zi(0)R¯i and Rot(zi(t);[0,T])1. Then, writing

zi(t)=(ρi(t)cosϑi(t),ρi(t)sinϑi(t)),

as long as ϑi(t)Θ(ϑ1i,ϑ2i), since ρi(t)R^iRi, we can use (4.2) and (4.4) to obtain

-ϑi(t)=AiHi(zi(t))+i(t,z1,,zN),zi(t)zi(t)2
Ai12AiT(ϑ2i-ϑ1i)+ε¯iR^i<ϑ2i-ϑ1iT.

Consequently, the time needed to clockwise cross the sector Θ(ϑ1i,ϑ2i) is greater than T, a contradiction. ∎

Proof of Theorem 4.1.

For any i{1,,N}, let A¯i>0 and ε¯i>0 be as in Lemma 4.2, and set

A=max{A¯i:i=1,,N},ε=min{ε¯i:i=1,,N}.

Take AiA and assume that (4.3) holds. Then, take R¯i as in Lemma 4.3 for every i=1,,N, and consider the annulus 𝒜i=(0,R¯i)Dih¯i. Recall that, taking h¯i>0 sufficiently small, the inner boundary of 𝒜i is star-shaped. Then, by Lemmas 4.2 and 4.3, for every solution z(t) of (P), if zi(0) belongs to the inner boundary of 𝒜i, then zi(t) makes more than νi clockwise rotations around the origin in the time T, while if zi(0)=R¯i, it makes less than one clockwise turn in the same time.

We now use the fact that (P) is a positive transformation of a Hamiltonian system, and consider the linear transformation defined in (4.1). Being all the matrices 𝕄i invertible with positive determinant, the set

𝒜=(𝒜1××𝒜N)

is thus of the type 𝒜~1××𝒜~N, where each 𝒜~i is a planar annulus with star-shaped boundaries with respect to the origin. Since the change of variables preserves the above described rotational properties of the solutions, we can apply [31, Theorem 8.2] to the Hamiltonian system obtained from (P) through the change of variables given by . We thus obtain at least N+1 distinct T-periodic solutions

z~k(t)=(z~1k(t),,z~Nk(t))

such that for every k=1,,N+1, each component z~ik(t), with i=1,,N, makes exactly νi clockwise rotations around the origin in the time interval [0,T[. Setting

zk(t)=(𝕄1-1z~1k(t),,𝕄N-1z~Nk(t)),

we obtain the solutions of (P) we are looking for, and the proof is thus completed. ∎

Remark 4.4

Theorem 4.1 exploits a gap between the rotation numbers of the solutions at zero and at infinity. With reference to the assumption at infinity, another possibility could be to replace (A3) with the requirement that for some i{1,,N}, the system Jz˙i=Hi(zi) has a homoclinic orbit surrounding the origin (in the spirit of [32, Theorem 3.3]). Indeed, by continuity, small perturbations of trajectories next to the homoclinic would have small rotation number, since the homoclinic spends an infinite time to rotate around the origin. In this setting, assuming moreover (A2), it would then be possible to construct the gap which allows to apply [31, Theorem 8.2], taking a level curve of Hi sufficiently near the homoclinic orbit as outer boundary of the required annulus in the i-th planar component. The same line of thought can be also adapted when the homoclinic is replaced by heteroclinics. One could also combine assumptions at infinity like (A3) for some indices i1,,ir{1,,N} and existence of homoclinics for the other indices i{1,,N}{i1,,ir}. We omit the details for briefness.

As a particular case, we can deal with a system of scalar second order equations like

(4.6){x¨1+A12f1(x1)=𝒲x1(t,x1,,xN),x¨N+AN2fN(xN)=𝒲xN(t,x1,,xN),

where the continuous function 𝒲:×N is T-periodic in t, and continuously differentiable with respect to (x1,,xN). Indeed, we can write the equivalent system

{-y˙i=Aifi(xi)-1Ai𝒲xi(t,x1,,xn),x˙i=Aiyi,  i=1,,N,

which is in the form (P), with zi=(xi,yi), taking

Hi(xi,yi)=12yi2+Fi(xi),

where Fi is a primitive of fi and

i(t,x1,y1,,xN,yN)=-1Ai(𝒲xi(t,x1,,xn)0).

Notice that (4.6) is a positive transformation of a Hamiltonian system, with the linear function in (4.1) given by

𝕄i=(100Ai),i=1,,N.

As a consequence, we have the following statement, where, for simplicity, we only consider the case ν1==νN=1.

Corollary 4.5

Assume that the continuous functions fi:RR satisfy

lim infs0fi(s)s>0,lims+fi(s)s=0.

Moreover, for every i=1,,N, let Ki>0 be such that

(4.7)|𝒲xi(t,x1,,xn)|Kifor every t[0,T] and x1,,xN.

Then, there exists A¯>0 such that if AiA¯ for every i=1,,N, system (4.6) has at least N+1 distinct periodic solutions

xk(t)=(x1k(t),,xNk(t))

with minimal period T. Moreover, for every k=1,,N+1, each component xik(t), with i=1,,N, has exactly two simple zeros in the interval [0,T[.

Proof.

First, we notice that (A1) is fulfilled, in view of the growth assumption on the nonlinearities. Let us now check (A2). We know that there exist αi>0 and βi>0 such that

0<|s|<βifi(s)sαi.

Then, if (xi,yi)βi,

Hi(xi,yi),(xi,yi)(xi,yi)2=xifi(xi)+yi2xi2+yi2min{αi,1}>0,

as desired.

We now verify (A3). Fix σ]0,π[, and take ϑ1i=0, ϑ2i=σ/2. Writing

zi=(xi,yi)=(ρicosϑi,ρisinϑi),

we have that if ziΘ(0,σ/2), then

Hi(zi),zizi2=(ρicosϑi)fi(ρicosϑi)+(ρisinϑi)2ρi2
sin2ϑi+|fi(ρicosϑi)ρicosϑi|σ24+|fi(ρicosϑi)ρicosϑi|.

Taking Ri>0 large enough, if ziΘ(0,σ/2)(0,Ri), then

Hi(zi),zizi2σ24+σ24=σ(ϑ2i-ϑ1i).

The proof is thus completed, noticing that it suffices to choose Ai large enough in order to make i(t,z1,,zN) as small as desired. ∎

As an example, Corollary 4.5 directly applies to the following system of N coupled pendulums,

{x¨1+A12sinx1=𝒲x1(t,x1,,xN),x¨N+AN2sinxN=𝒲xN(t,x1,,xN),

where 𝒲xi(t,x1,,xN) is continuous and bounded for i=1,,N, and the constants A1,,AN are large enough. We are thus able to recover the results obtained in [32], by the use of the Poincaré–Birkhoff theorem, for a single equation modeling a forced pendulum having a very small length.

Funding statement: The authors have been partially supported by the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA – INdAM).

References

[1] Abbondandolo A., Morse Theory for Hamiltonian Systems, CRC Res. Notes Math. 425, Chapman & Hall/CRC, Boca Raton, 2001. 10.1201/9781482285741Search in Google Scholar

[2] Albrecht J., On the existence of invariant tori in nearly-integrable Hamiltonian systems with finitely differentiable perturbations, Regul. Chaotic Dyn. 12 (2007), 281–320. 10.1134/S1560354707030033Search in Google Scholar

[3] Amann H. and Zehnder E., Nontrivial solutions for a class of non-resonance problems and applications to nonlinear differential equations, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 7 (1980), 539–603. Search in Google Scholar

[4] Ambrosetti A., Coti Zelati V. and Ekeland I., Symmetry breaking in Hamiltonian systems, J. Differential Equations 67 (1987), 165–184. 10.1016/0022-0396(87)90144-6Search in Google Scholar

[5] Arnold V. I., Kozlov V. V. and Neishtadt A. I., Mathematical Aspects of Classical and Celestial Mechanics, Encyclopaedia Math. Sci. 3, Springer, Berlin, 2006. 10.1007/978-3-540-48926-9Search in Google Scholar

[6] Bambusi D. and Berti M., A Birkhoff-Lewis type theorem for some Hamiltonian PDEs, SIAM J. Math. Anal. 37 (2005), 83–102. 10.1137/S0036141003436107Search in Google Scholar

[7] Benettin G., The elements of Hamiltonian perturbation theory, Hamiltonian Systems and Fourier Analysis, Adv. Astronom. Astrophys. 9, Cambridge Scientific Publishers, Cambridge (2005), 1–98. Search in Google Scholar

[8] Benettin G., Galgani L. and Giorgilli A., Poincaré’s nonexistence theorem and classical perturbation theory for nearly integrable Hamiltonian systems, Advances in Nonlinear Dynamics and Stochastic Processes (Florence 1985), World Scientific, Singapore (1985), 1–22. Search in Google Scholar

[9] Bernstein D. and Katok A., Birkhoff periodic orbits for small perturbations of completely integrable Hamiltonian systems with convex Hamiltonians, Invent. Math. 88 (1987), 225–241. 10.1007/BF01388907Search in Google Scholar

[10] Berti M., Biasco L. and Valdinoci E., Periodic orbits close to elliptic tori and applications to the three-body problem, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 3 (2004), 87–138. 10.2422/2036-2145.2004.1.06Search in Google Scholar

[11] Bessi U., An analytic counterexample to the KAM theorem, Ergodic Theory Dynam. Systems 20 (2000), 317–333. 10.1017/S0143385700000146Search in Google Scholar

[12] Biasco L. and Di Gregorio L., A Birkhoff–Lewis type theorem for the nonlinear wave equation, Arch. Ration. Mech. Anal. 196 (2010), 303–362. 10.1007/s00205-009-0240-ySearch in Google Scholar

[13] Birkhoff G. D. and Lewis D. C., On the periodic motions near a given periodic motion of a dynamical system, Ann. Mat. Pura Appl. (4) 12 (1934), 117–133. 10.1007/BF02413852Search in Google Scholar

[14] Boscaggin A., Subharmonic solutions of planar Hamiltonian systems: A rotation number approach, Adv. Nonlinear Stud. 11 (2011), 77–103. 10.1515/ans-2011-0104Search in Google Scholar

[15] Boscaggin A. and Garrione M., Resonance and rotation numbers for planar Hamiltonian systems: Multiplicity results via the Poincaré–Birkhoff theorem, Nonlinear Anal. 74 (2011), 4166–4185. 10.1016/j.na.2011.03.051Search in Google Scholar

[16] Boscaggin A. and Ortega R., Monotone twist maps and periodic solutions of systems of Duffing type, Math. Proc. Cambridge Philos. Soc. 157 (2014), 279–296. 10.1017/S0305004114000310Search in Google Scholar

[17] Buttazzoni P. and Fonda A., Periodic perturbations of scalar second order differential equations, Discrete Contin. Dyn. Syst. 3 (1997), 451–455. 10.3934/dcds.1997.3.451Search in Google Scholar

[18] Celletti A., Stability and Chaos in Celestial Mechanics, Springer, Berlin, 2010. 10.1007/978-3-540-85146-2Search in Google Scholar

[19] Celletti A. and Chierchia L., KAM stability and celestial mechanics, Mem. Amer. Math. Soc. 187 (2007). 10.1090/memo/0878Search in Google Scholar

[20] Chang K. C., Methods in Nonlinear Analysis, Springer, Berlin, 2005.Search in Google Scholar

[21] Chen W. F., Birkhoff periodic orbits for small perturbations of completely integrable Hamiltonian systems with nondegenerate Hessian, Twist Mappings and Their Applications, IMA Vol. Math. Appl. 44, Springer, Berlin (1992), 87–94. 10.1007/978-1-4613-9257-6_5Search in Google Scholar

[22] Chicone C., The monotonicity of the period function for planar Hamiltonian vector fields, J. Differential Equations 69 (1987), 310–321. 10.1016/0022-0396(87)90122-7Search in Google Scholar

[23] Conley C. C. and Zehnder E., An index theory for periodic solutions of a Hamiltonian system, Geometric Dynamics, Lecture Notes in Math. 1007, Springer, Berlin (1983), 132–145. 10.1007/BFb0061415Search in Google Scholar

[24] Conley C. C. and Zehnder E., The Birkhoff–Lewis fixed point theorem and a conjecture of V. I. Arnold, Invent. Math. 73 (1983), 33–49. 10.1007/BF01393824Search in Google Scholar

[25] Ding T. and Zanolin F., Periodic solutions of Duffing’s equation with superquadratic potential, J. Differential Equations 97 (1992), 328–378. 10.1016/0022-0396(92)90076-YSearch in Google Scholar

[26] Dumas H. S., The KAM Story: A Friendly Introduction to the Content, History, and Significance of classical Kolmogorov–Arnold–Moser Theory, World Scientific, Singapore, 2014. 10.1142/8955Search in Google Scholar

[27] Ekeland I., A perturbation theory near convex Hamiltonian systems, J. Differential Equations 50 (1983), 407–440. 10.1016/0022-0396(83)90069-4Search in Google Scholar

[28] Fonda A. and Gidoni P., Generalizing the Poincaré–Miranda theorem: The avoiding cones condition, Ann. Mat. Pura Appl. (4), 10.1007/s10231-015-0519-6. 10.1007/s10231-015-0519-6Search in Google Scholar

[29] Fonda A., Sabatini M. and Zanolin F., Periodic solutions of perturbed Hamiltonian systems in the plane by the use of the Poincaré–Birkhoff theorem, Topol. Methods Nonlinear Anal. 40 (2012), 29–52. Search in Google Scholar

[30] Fonda A. and Sfecci A., Periodic solutions of weakly coupled superlinear systems, J. Differential Equations 260 (2016), 2150–2162. 10.1016/j.jde.2015.09.056Search in Google Scholar

[31] Fonda A. and Ureña A. J., On the higher dimensional Poincaré–Birkhoff theorem for Hamiltonian flows. 1. The indefinite twist, preprint 2013, www.dmi.units.it/~fonda/2013\_Fonda-Urena.pdf. Search in Google Scholar

[32] Fonda A. and Zanolin F., Periodic oscillations of forced pendulums with very small length, Proc. Roy. Soc. Edinburgh Sect. A 127 (1997), 67–76. 10.1017/S0308210500023519Search in Google Scholar

[33] Garijo A., Gasull A. and Jarque X., A note on the period function for certain planar vector fields, J. Difference Equ. Appl. 16 (2010), 631–645. 10.1080/10236190903241689Search in Google Scholar

[34] Hartman P., On boundary value problems for superlinear second order differential equations, J. Differential Equations 26 (1977), 37–53. 10.1016/0022-0396(77)90097-3Search in Google Scholar

[35] Herman M. R., Sur les courbes invariantes par les difféomorphismes de l’anneau, Astérisque 103, Société Mathématique de France, Paris, 1983. Search in Google Scholar

[36] Jacobowitz H., Periodic solutions of x′′+f(x,t)=0 via the Poincaré–Birkhoff theorem, J. Differential Equations 20 (1976), 37–52. 10.1016/0022-0396(76)90094-2Search in Google Scholar

[37] Kulpa W., The Poincaré–Miranda theorem, Amer. Math. Monthly 104 (1997), 545–550. 10.1080/00029890.1997.11990676Search in Google Scholar

[38] Lacomba E. A. and Llibre J., Integrals, invariant manifolds, and degeneracy for central force problems in n, J. Math. Phys. 33 (1992), 2138–2147. 10.1063/1.529633Search in Google Scholar

[39] Lazer A. C., Small periodic perturbations of a class of conservative systems, J. Differential Equations 13 (1973), 438–456. 10.1016/0022-0396(73)90003-XSearch in Google Scholar

[40] Loud W. S., Periodic solutions of x′′+cx+g(x)=εf(t), Mem. Amer. Math. Soc. 31 (1959). Search in Google Scholar

[41] MacKay R. S. and Percival I. C., Converse KAM: Theory and practice, Comm. Math. Phys. 98 (1985), 469–512. 10.1007/BF01209326Search in Google Scholar

[42] Markus L. and Meyer K. R., Generic Hamiltonian dynamical systems are neither integrable nor ergodic, Mem. Amer. Math. Soc. 144 (1974). 10.1090/memo/0144Search in Google Scholar

[43] Moser J., Recent developments in the theory of Hamiltonian systems, SIAM Rev. 28 (1986), 459–485. 10.1137/1028153Search in Google Scholar

[44] Moser J. and Zehnder E., Notes on Dynamical Systems, Courant Lect. Notes Math. 12, American Mathematical Society, Providence, 2005. 10.1090/cln/012Search in Google Scholar

[45] Pinsky M. A. and Zevin A. A., Monotonicity criteria for an energy-period function in planar Hamiltonian systems, Nonlinearity 14 (2001), 1425–1432. 10.1088/0951-7715/14/6/301Search in Google Scholar

[46] Poincaré H., Les Méthodes Nouvelles de la Méchanique Céleste, Gauthier–Villars, Paris, 1892. Search in Google Scholar

[47] Robinson R. C., Generic properties of conservative systems, Amer. J. Math. 92 (1970), 562–603. 10.1007/BFb0070136Search in Google Scholar

[48] Salamon D. A., The Kolmogorov–Arnold–Moser theorem, Math. Phys. Electron. J. 10 (2004), Article ID 3. Search in Google Scholar

[49] Sell G. R., Differential equations without uniqueness and classical topological dynamics, J. Differential Equations 14 (1973), 42–56. 10.1016/0022-0396(73)90075-2Search in Google Scholar

[50] Siegel C. L., Über die Existenz einer Normalform analytischer Hamiltonscher Differentialgleichungen in der Nähe einer Gleichgewichtslösung, Math. Ann. 128 (1954), 144–170. 10.1007/BF01360131Search in Google Scholar

[51] Treshchëv D. V., A mechanism for the destruction of resonance tori in Hamiltonian systems, Mat. Sb. 180 (1989), 1325–1346. 10.1070/SM1991v068n01ABEH001371Search in Google Scholar

[52] Willem M., Perturbations of nondegenerate periodic orbits of Hamiltonian systems, Periodic Solutions of Hamiltonian Systems and Related Topics, NATO ASI Series C 209, Springer, Dordrecht (1987), 261–265. 10.1007/978-94-009-3933-2_24Search in Google Scholar

Received: 2015-9-4
Revised: 2015-12-3
Accepted: 2015-12-3
Published Online: 2016-1-14
Published in Print: 2016-11-1

© 2016 by De Gruyter

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 5.5.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2015-0122/html
Scroll to top button