Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article

Normal forms, invariant manifolds and Lyapunov theorems

  • Received: 23 November 2022 Revised: 11 May 2023 Accepted: 22 May 2023 Published: 26 June 2023
  • Primary 05C38, 15A15; Secondary 05A15, 15A18

  • We present an approach to Lyapunov theorems about a center for germs of analytic vector fields based on the Poincaré–Dulac and Birkhoff normal forms. Besides new proofs of three Lyapunov theorems, we prove their generalization: if the Poincaré–Dulac normal form indicates the existence of a family of periodic solutions, then such a family really exists. We also present new proofs of Weinstein and Moser theorems about lower bounds for the number of families of periodic solutions; here, besides the normal forms, some topological tools are used, i.e., the Poincaré–Hopf formula and the Lusternik–Schnirelmann category on weighted projective spaces.

    Citation: Henryk Żołądek. Normal forms, invariant manifolds and Lyapunov theorems[J]. Communications in Analysis and Mechanics, 2023, 15(2): 300-341. doi: 10.3934/cam.2023016

    Related Papers:

    [1] Erlend Grong, Irina Markina . Harmonic maps into sub-Riemannian Lie groups. Communications in Analysis and Mechanics, 2023, 15(3): 515-532. doi: 10.3934/cam.2023025
    [2] Wenmin Gong . A short proof of cuplength estimates on Lagrangian intersections. Communications in Analysis and Mechanics, 2023, 15(2): 50-57. doi: 10.3934/cam.2023003
    [3] Vladimir Rovenski . Generalized Ricci solitons and Einstein metrics on weak K-contact manifolds. Communications in Analysis and Mechanics, 2023, 15(2): 177-188. doi: 10.3934/cam.2023010
    [4] Xiaotian Hao, Lingzhong Zeng . Eigenvalues of the bi-Xin-Laplacian on complete Riemannian manifolds. Communications in Analysis and Mechanics, 2023, 15(2): 162-176. doi: 10.3934/cam.2023009
    [5] Richard Cushman . Normalization and reduction of the Stark Hamiltonian. Communications in Analysis and Mechanics, 2023, 15(3): 457-469. doi: 10.3934/cam.2023022
    [6] Richard Pinčák, Alexander Pigazzini, Saeid Jafari, Özge Korkmaz, Cenap Özel, Erik Bartoš . A possible interpretation of financial markets affected by dark volatility. Communications in Analysis and Mechanics, 2023, 15(2): 91-110. doi: 10.3934/cam.2023006
    [7] Jonas Schnitzer . No-go theorems for r-matrices in symplectic geometry. Communications in Analysis and Mechanics, 2024, 16(3): 448-456. doi: 10.3934/cam.2024021
    [8] Xin Qiu, Zeng Qi Ou, Ying Lv . Normalized solutions to nonautonomous Kirchhoff equation. Communications in Analysis and Mechanics, 2024, 16(3): 457-486. doi: 10.3934/cam.2024022
    [9] Shengbing Deng, Qiaoran Wu . Existence of normalized solutions for the Schrödinger equation. Communications in Analysis and Mechanics, 2023, 15(3): 575-585. doi: 10.3934/cam.2023028
    [10] Xueqi Sun, Yongqiang Fu, Sihua Liang . Normalized solutions for pseudo-relativistic Schrödinger equations. Communications in Analysis and Mechanics, 2024, 16(1): 217-236. doi: 10.3934/cam.2024010
  • We present an approach to Lyapunov theorems about a center for germs of analytic vector fields based on the Poincaré–Dulac and Birkhoff normal forms. Besides new proofs of three Lyapunov theorems, we prove their generalization: if the Poincaré–Dulac normal form indicates the existence of a family of periodic solutions, then such a family really exists. We also present new proofs of Weinstein and Moser theorems about lower bounds for the number of families of periodic solutions; here, besides the normal forms, some topological tools are used, i.e., the Poincaré–Hopf formula and the Lusternik–Schnirelmann category on weighted projective spaces.



    In his memoir [21], Lyapunov stated several theorems about analytic families of periodic solutions of some analytic autonomous vector fields near equilibrium point. Below, we distinguish three of them.

    First, consider an analytic planar vector field with a singular point having a pair of pure imaginary eigenvalues. Thus, we can assume an autonomous differential system of the form

    ˙x=ωy+, ˙y=ωx+. (1.1)

    with real analytic right-hand sides.

    Here, one looks for a first integral in the form of a formal power series:

    F=ω2(x2+y2)+. (1.2)

    In general, such a formal first integral does not exist; one finds

    ˙F=c2k+1(x2+y2)k+1+, (1.3)

    where c2k+10 is the so-called Poincaré–Lyapunov focus quantity and we have a weak focus of order 2k1 (see Section 2 below). Otherwise, we have the following result attributed to Lyapunov [21, Section 39] and Poincaré [27, Chapter 11].

    Theorem 1. (Lyapunov–Poincaré) If all focus quantities vanish, then the formal integral (1.2) is, in fact, analytic, and all solutions near x=y=0 are periodic.

    In [21, Section 40] the latter situation is generalized to the following system:

    ˙x=Ax+, (1.4)

    x(Rn,0), where the matrix A has a pair λ1,2=±iω of pure imaginary eigenvalues and all of its other eigenvalues lie in the half-plane {Reλ<0}. Again, one looks for a formal first integral of the form

    F=ω2(x21+x22)+, (1.5)

    where x1,2 are the variables associated with λ1,2, and there are again obstacles to such an integral occurring.

    Theorem 2. (Lyapunov) If all of these obstacles vanish, then the first integral is analytic and system (1.4) has an analytic 1–parameter family of periodic solutions.

    The last situation considered in [21, Section 42]* is the case when the matrix A in system (1.4) has the pure imaginary eigenvalues

    *In fact, Lyapunov considered a more general situation where, besides ±iωj, there are also eigenvalues in the left half-plane.

    ±iω1,,±iωm, ωj>0,

    and one of the frequencies, say, ω1, is such that none of the other frequencies is an integer multiple of it; thus,

    ωj/ω1Z,  j1. (1.6)

    Theorem 3. (Lyapunov) In this case, there exists a family of periodic solutions

    x=ϕ(t;c), c(R+,0),

    of period T(c)2π/ω1, depending analytically on c and such that ϕ(t;0)0.

    The latter theorem has attracted the attention of specialists in Hamiltonian dynamics.

    I have somehow overlooked this result and have learned about it only recently on the occasion of reviewing the Ph.D. thesis of D. Strzelecki in the Nicolas Copernicus University in Toruń.

    Recall that an autonomous analytic Hamiltonian system with m=n2 degrees of freedom takes the form

    ˙qj=Hpj, ˙pj=Hqj,  j=1,,n; (1.7)

    the corresponding vector field is usually denoted as

    XH=Hpi/qiHpi/pi.

    Assume that it has equilibrium point q=p=0 with the eigenvalues ±iω1,,±iωm,ωj>0. Assuming H(0)=0, the leading part of the Taylor expansion of the Hamiltonian function is

    H2=12ϵjωj(q2j+p2j), (1.8)

    where qj,pj are suitable canonical variables, i.e., with the Poisson brackets {pi,qj}=δij, and ϵj=±1 are well-defined signs.

    We have ˙f={f,H} in the case of a general Hamiltonian system with the symplectic structure defined by a Poisson bracket.

    In particular, for zj=qj+ipj and vj=qjipj, we have {zj,zk}={vj,vk}=0 and {zj,vk}=2iδjk. Thus ˙zj={zj,vk}H/vk=2iH/vk.

    Also, the resonant monomials g=zkvl that form the Birkhoff theorem satisfy {g,H2}=0.

    In the case of (1.8) the corresponding linear system is diagonalizable in the complex variables. The non-diagonalizable case is somewhat special and we consider it only in Example 7 in Section 5.

    Schmidt [28] studied 1–parameter families of periodic solutions for such systems with two degrees of freedom in the cases of resonant frequencies ω1 and ω2. His analysis was mainly focused on definite Hamiltonians, i.e., when ϵ1=ϵ2; but, he also considered the situation near the Lagrangian libration point in the restricted three-body problem, where the Hamiltonian function is indefinite. He did not refer to the third Lyapunov theorem. Anyway, his results agree with what is presented below.

    Weinstein [35] has applied the Lusternik–Schnirelmann category to prove the following.

    Theorem 4. (Weinstein) Assume that H2 is positive definite, i.e., all ϵj=1 in Eq. (1.8). Then, system (1.7) has at least m=n2 1–parameter families of periodic solutions.

    Next, Moser tried in [23, Theorem 4] to specify Weinstein's result by assuming that H2 is positive on a linear subspace E associated with one frequency ω, i.e., all solutions in E of the linear system have the period 2π/ω and these are the only such solutions. He claimed that there exists at least 12dimE of periodic solution of the period 2π/ω of the full Hamiltonian system. But his own example [23, Example 2] contradicts his statement; see also Example 4 below. Note also that the change HH means the reversion of the time, and it does not influence the periodic property of solutions.

    We have the following specification of Theorem 4. Again assume an analytic Hamiltonian system (1.6) with the eigenvalues ±iωj,ωj>0.

    Following [35, Proof of Theorem 2.1] we consider the following equivalence relation on the set {ω1,,ωm} of frequencies

    ωiωj iff ωi/ωjQ. (1.9)

    For each equivalence class Cν we have a linear subspace Eν that is invariant for the linear part of the system.

    Choose some equivalence class C1 and let E1 be the corresponding linear subspace. Let us order the frequencies from C1 as follows:

    ˜ω1:=ω1==ωk1>˜ω2:=ωk1+1==ωk1+k2>>˜ωr:=ωk1++kr1+1==ωk1++kr. (1.10)

    Theorem 5. Let thinspace H be analytic and such that H2|E1 is definite (say, positive), i.e., ϵ1==ϵk1++kr=1 in Eq. (1.8); then, there exist the following:

    at least k1 1–parameter families of periodic solutions to system (1.7) with periods 2π/˜ω1,

    at least k1+k2 1–parameter families of periodic solutions to system (1.7) with periods 2π/˜ω1 or 2π/˜ω2,

    at least k1++kr 1–parameter families of periodic solutions to system (1.7) with periods 2π/˜ω1, or 2π/˜ω2,, or 2π/˜ωr.

    Next, this subject was raised up by specialists in the nonlinear functional analysis, including Szulkin [32] and several other groups [10,15,25,26]. The principal result here is that we assume the ordering (1.10) of all frequencies (not only from C1).

    Theorem 6. (Szulkin) If, for a given l,

    (i) ˜ωj/˜ωlN for all jl in Eq. (1.10), and

    (ii) ϵk1++kl1+1++ϵk1++kl0 for ϵi's from Eq. (1.8), then there exists a sequence {γn(t)} of periodic solutions to the equation ˙x=XH(x) of non-constant periodic solutions tending to γ(t)0 of periods tending to 2π/˜ωl.

    We see that the first assumption is like in the third Lyapunov theorem. But, the assumption (ii) is new; it replaces the definiteness assumption of H2 from Theorems 4 and 5.

    Finally, we recall the following non-Hamiltonian version of the third Lyapunov theorem due to Moser [23, Theorem 2].

    Theorem 7. (Moser) Assume that an analytic system (1.4) with only purely imaginary eigenvalues has an analytic first integral F(x) such that D2F(0) is positive definite. Then, this system has at least one 1–parameter family of periodic solutions.

    The aim of this paper is to take a somehow different look at this center problem. Our approach is rather qualitative and less topological, like in [1] and [18]. It turns out that much can be deduced from the Poincaré–Dulac normal form and its Hamiltonian version, i.e., the Birkhoff normal form (Schmidt also used the Birkhoff normal form). Important is the use of some versions of the standard Poincaré return map combined with the analyticity assumption of the vector fields. We will reprove most of above theorems and we shall prove a new result.

    Assume that the matrix A in system (1.4) has pure imaginary eigenvalues: ±iω1,,±iωm, ωj>0. We can reduce it, via a formal change of variables, to the Poincaré–Dulac normal form (see Theorem 9 in the next section). It may happen that the latter formal system admits a formal 1–parameter family of periodic solutions with the period 2π/ωl.

    More precisely, one can consider the situation with only one equivalence class for the frequencies, i.e., ωj/ωk=pj/pk, with relatively prime positive integers pi, and that ω1ωl. Let Zj be the formal complex variables associated with the Poincaré–Dulac normal form. For the new variables RR,WjC and ΘS1, defined by Zl=ReiΘ and Zj=WjeipjΘ/pl, one obtains a formal differential system:

    dRdΘ=Π(R,W,ˉW), dWjdΘ=Λj(R,W,ˉW)

    (called the return system). The analogue of the center conditions in this case means that the latter system has a formal curve Γ of non-isolated equilibrium points. All of these notions are made precise in Section 7 (Eqs. (7.1)–(7.3), Lemma 3 and Definition 4).

    The next statement can be regarded as a fourth Lyapunov theorem.

    Theorem 8. In this case, there exists a continuous family of periodic solutions to system (1.4) with the period 2π/ωl.

    The paper is organized as follows. In Section 2, we define the Poincaré –Dulac and Birkhoff normal forms, with their principal properties. In Section 3, we present and discuss theorems about invariant manifolds. Section 4 is devoted to various proofs of the three Lyapunov theorems; they are included in this paper (which is treated as a sort of review) because some of the Lyapunov theorems are not widely known, and because their proofs are (we hope) illuminating. In Section 5, we present some examples of Hamiltonian systems with explicit families of periodic solutions. In Section 6, we prove the existence of some additional invariant submanifolds following from the normal forms. In Section 7, we prove Theorem 8. Section 8 is devoted to the discussion of situations with a first integral; there, we give proofs of Theorems 4, 5, 7 and an introduction to the functional analytic method. In Section 9, we present our approach, based only on the Birkhoff normal form, of two examples from the celestial mechanics: geostationary orbits of satellites and libration points in the restricted four-body problem. The last section contains appendices about the topology of weighted projective spaces and the Lusternik–Schnirelmann category.

    Consider the analytic system (1.4), but in (Cn,0) (not in (Rn,0). Thus, we can assume that the matrix A is in the Jordan form with eigenvalues λ1,,λn.

    Definition 1. We say that the latter system of eigenvalues satisfies the resonant relation of type (j;k),j=1,,n,k=(k1,,kn)Zn0, if

    λj=(k,λ)=k1λ1++knλn. (2.1)

    Theorem 9. (Poincaré–Dulac [13,27]) There exists a change xX, defined by formal power series, which transforms system (1.4) to the following Poincaré–Dulac normal form:

    ˙Xj=λjXj+kaj;kXk,j=1,,n, (2.2)

    where the summation runs over multi-indices k such that the resonant relations of type (j;k) hold and Xk=Xk11Xknn.

    The proof can be found in [3,17,37].

    Remark 1. In the case when there are multiple eigenvalues λj and the corresponding Jordan cell is not diagonal, we have the resonant relations (j;k), where k=(0,,0,1,0,0) with 1 in the (j+1)th entry. They imply non-diagonal entries in A.

    Remark 2. In the so-called Poincaré domain, i.e., when the convex hull of the set {λ1,,λn}C is separated from 0C, the Poincaré–Dulac normal form is polynomial and the normalizing change is analytic. We refer the reader to [30] for a short proof.

    Remark 3. In the general, situation the Poincaré–Dulac normal form is divergent. Indeed, in the example

    ˙x1=x21, ˙x2=x2x21, (2.3)

    due to Euler (see [37]), the Poincaré–Dulac normal form is

    ˙X1=X21+a3X31+, ˙X2=X2(1+b1X1+b2X21+).

    But, the so-called center manifold Wc, corresponding to the zero eigenvalue and formally defined by X2=0, in the original system (2.3) takes the form

    Wc={x2=(m1)!xm1}.

    Consider now an autonomous analytic Hamiltonian system (1.7) with m=n2 degrees of freedom near an equilibrium point q=p=0 with the eigenvalues ±iω1,,±iωm,ωj>0. Assuming H(0)=0, the leading part of the Taylor expansion of the Hamiltonian function is like in Eq. (1.8), i.e., H2=12ϵjωj(q2j+p2j), where qj,pj are suitable canonical variables, i.e., with the Poisson brackets {pi,qj}=δij, and ϵj=±1 are well-defined signs.

    It is natural to introduce the complex variables

    zj=qj+ipj, vj=qjipj, (2.4)

    with the Poisson brackets {zj,vk}=2iδjk. Then, we have

    H2=12ϵjωjzjvj (2.5)

    and

    ˙zj=λjzj+, ˙vj=λjvj+,  λj=iϵjωj. (2.6)

    Of course, in the real domain (where pj and qk are real), we have vj=ˉzj.

    The following result from [8] is a Hamiltonian analogue of the Poincaré–Dulac theorem.

    Theorem 10. (Birkhoff) There exists a formal canonical change (z,v)(Z,V) which reduces the Hamiltonian function to the following one:

    H2(Z,V)+ak;lZkVl,

    where summation runs over the pairs (k;l)=(k1,,kn;l1,,ln)Zn0×Zn0 such that the resonant relations

    (kl,λ)=(kjlj)λj=0

    hold and |k|+|l|3.

    Definition 2. The singular point x=0 of system (1.4) is hyperbolic if none of the eigenvalues of the matrix A is imaginary. Thus, the eigenvalues are divided into two groups:

    {λ1,,λp} with Reλj<0 and {λp+1,,λn} with Reλj>0; (3.1)

    i.e., separated by the imaginary axis.

    The principal result about invariant manifolds is the Hadamard–Perron theorem for hyperbolic singular points. These manifolds are the local stable manifold (of dimension p)

    Ws={x:gt(x)0 as t+} (3.2)

    and the local unstable manifold (of dimension q=np)

    Wu={x:gt(x)0 as t}, (3.3)

    where {gt} is the phase flow generated by the vector field (1.4); gt(x0) is the solution x(t) obeying the initial condition x(0)=x0. In the case of a finitely smooth vector field, the proof of the Hadamard–Perron theorem is rather involved. But, in the analytic case, it is quite easy.

    Theorem 11. (Holomorphic Hadamard–Perron) Assume that the vector field (1.4) is analytic. Assume also that the set {λ1,,λn}C of eigenvalues of the matrix A can be separated by a straight line through λ=0 into two groups: {λ1,,λp} on one side of and {λp+1,,λn} on the other side. Then, there exist local analytic invariant manifolds for (1.4) tangent to the corresponding linear subspaces invariant for A.

    It follows that, if the singular point x=0 is hyperbolic, then the invariant manifolds Ws and Wu are real analytic.

    Proof. We present a rather novel proof, following a method from [30]. Another proof can be found in [17]; note also that, in the case n=2 and p=1, this result was proved by Briot and Bouquet [9] (see also [37]).

    Let EsCp and EuCq be the linear subspaces associated with the division of the eigenvalues set, and let (x,y) be the coordinates associated with the splitting Cn=EsEu. We can assume that the matrix A has the form AsAu, where As=diag(μ1,,μp)+As1 and Au=diag(ν1,,νq)+Au1, where As,u1 are off-diagonal with small entries. We have

    ˙x=Asx+ϕ(x,y),  ˙y=Auy+ψ(x,y).

    We look for the manifold Ws as a graph of a map F:EsEu,Ws={(x,F(x))}; the case of Wu is treated analogously. The invariance condition, i.e.,

    ˙yDF˙x|y=F(x)0,

    leads to the equation

    AuF(x)DF(x)Asx=DF(x)ϕ(x,F(x))ψ(x,F(x)). (3.4)

    This equation takes the form

    LF=T(F),

    where L and T are operators, linear and nonlinear. We treat these operators as acting on the Banach space F consisting of power series F=j,kfj,kxkej with the norm

    F=|fj;k|ρ|k|,

    where ρ is the 'radius of convergence' (suitably chosen). We have L=L0+L1, where

    L0xkej=[νj(k,μ)]xkej

    and L1 is small with respect to L0. Because νj and μi are separated by the line , the operator L is invertible.

    Moreover, if Fd={F=j,|k|=dfj;kxkej} are the subspaces of homogeneous maps of degree d2, then

    L1|Fd<C1/d

    for some constant C1. On the other hand, the operators T|Fd:FdF>d have the Lipschitz constants bounded by C2dρ, where the factor d arises from DF and C2 is some constant. It follows that the operator

    P=L1T

    is Lipschitz-continuous with a small Lipschitz constant. Therefore, the fixed-point problem F=P(F) has a unique solution.

    Assume now the non-hyperbolic situation. Thus, we have the splittings Rn=EsEuEc and A=AsAuAc, where the eigenvalues of As (respectively, Au) lie in the left (respectively, right) half-plane and the eigenvalues of Ac lie on the imaginary axis. The next theorem is usually cited without proof (which is quite technical, see [3,17]).

    Theorem 12. (Center Manifold) There exist smooth invariant manifolds Ws, Wu and Wc tangent to the subspaces Es,Eu and Ec, respectively. The manifolds Ws and Wu are analytic; Wc is only infinitely smooth in general, but its Taylor series is defined uniquely.

    Remark 4. The Euler example from Remark 3 demonstrates that the center manifold can be non-analytic. If Wc={y=F(x)}, then the analogue of Eq. (3.4) is

    LF=x2ddxF+x2,

    where L=Id. Here L, although invertible, is not dominating in comparison to x2ddx; we have x2ddxxk/xk=kρ.

    In [17] and [37], one can find an explanation of the non-analyticity of Wc in terms of the so-called Stokes phenomena.

    Next, the center manifold can be non-unique. Indeed, the real Euler example is the so-called saddle-node singularity; on the left from the unstable manifold Wu={x=0}, we have two hyperbolic sectors separated by the left part of Wc (which is unique), but, on the right of Wu, we have a parabolic sector, and any of its phase curves can serve as the right part of Wc. This non-uniqueness is explained by the method of the proof of the existence of Wc; there, one extends the vector field to the whole Rn and Wc (which is a fixed point of some functional operator) depends on this extension.

    Below, we present other examples with non-analytic invariant manifolds.

    Example 1. Let z=x1+ix2,v=x1ix2 and y be coordinates in C3. The system is as follows:

    ˙z=iz+z2v, ˙v=iv+zv2, ˙y=y+zv. (3.5)

    One finds that

    Wc={y=12(n1)!(2zv)n}.

    The non-analyticity of the center manifold is explained via the nonlinear Stokes phenomena as follows. In the variables u=zv and y, we have

    ˙u=2u2, ˙y=y+u,

    i.e., a saddle-node similar to the Euler example. In a sectorial domain around {(u,y):Im(u)=0, Re(u)>0}, the phase portrait near the saddle, with a unique center manifold and unique stable manifold; but, in a sectorial domain near {(u,y):Im(u)=0,Re(u)<0}, the phase portrait is like near the node with many center-ype manifolds.

    Example 2. Let z1=x1+ix2,v1=x1ix2 and z2=x3+ix4,v2=x3ix4 be coordinates in C4. The system is as follows:

    ˙z1=iω1z1+z21v1, ˙v1=iω1v1+z1v21, ˙z2=iω2z2+z1v1, ˙v2=iω2v2+z1v1. (3.6)

    It has two invariant planes:

    W1={z1=v1=0}

    (analytic) and

    W2={z2=12(n1)!(2z1v1iω2)n, v2=12(n1)!(2iz1v1ω2)n},

    which is non-analytic.

    We finish this section with another theorem about invariant manifolds.

    Definition 3. Let f be a smooth diffeomorphism of a manifold M, and let LM be a smooth invariant submanifold for f,f(L)L. We say that f is normally hyperbolic at N if there is a splitting

    NL=TM/TNEsEu

    of the normal bundle to L into sub-bundles Es and Eu that are invariant with respect to Df and the following estimates hold:

    the bundle Es is contracted more strongly than the bundle TL, i.e.,

    supLDf|Es<{infL(Df|TL)1}1;

    the bundle Eu is expanded more strongly than the bundle TL, i.e.,

    {infL(Df|Eu)1}1>supLDf|TL.

    We refer the reader to [16] for the proof of the following statement.

    Theorem 13. (Normal Hyperbolicity) If f is normally hyperbolic on L, then there exists a neighborhood U of f in the functional space of C1–diffeomorphisms such that any gU has a unique invariant submanifold Lg close to L at which it is normally hyperbolic.

    Eq. (1.1) with the variables z=x+iy,v=xiy takes the form

    ˙z=iωz+, ˙v=iωv+. (4.1)

    The Poincaré–Dulac normal form here is as follows:

    ˙Z=Z{iω+(ak+ibk)(ZV)k}, ˙V=V{iω+(akibk)(ZV)k}. (4.2)

    It is easy to find that the first nonzero coefficient ak is related to the first nonzero Poincaré–Lyapunov quantity:

    c2k+1=ωak. (4.3)

    Next, after passing to the polar coordinates r,θ (in the real domain), i.e., with z=reiθ and v=ˉz=reiθ, we get the equation

    drdθ=f(r,θ),

    where f is a convergent series in powers of r with trigonometric polynomials in θ as coefficients. We define the Poincaré return map

    P(r0)=ϕ(2π;r0), (4.4)

    where ϕ(θ;r0) is the solution to the latter equation dr/dθ with the initial condition ϕ(0;r0)=r0. The Poincaré map is analytic and has a Taylor expansion of the form

    P(r)=r+d2k+1r2k+1+,

    where

    d2k+1=(2π/ω)ak (4.5)

    provided the Poincaré–Lyapunov quantity c2k+10.

    The assumption of Theorem 1 means formally that either

    (i) the Poincaré–Dulac normal form equals

    ˙Z=iZM(ZV), ˙V=iVM(ZV), (4.6)

    where M=ω+bk(ZV)k is a formal integrating factor and F=ZV is the formal first integral; or (equivalently),

    (ii) formally,

    P=Id. (4.7)

    Moreover, the assumptions of the holomorphic Hadamard–Perron theorem for system (4.1) are satisfied. Therefore, we have two holomorphic invariant lines, W+ and W, formally defined by {V=0} and {W=0}, respectively.

    Below, we present two geometrical proofs of Theorem 1.

    First proof. (Note that this proof was not found in the literature). Since the Poincaré map is analytic and formally obeys Eq. (4.7), it is the identity. Therefore, the real phase portrait is of the center type and the equilibrium point is surrounded by a family of closed phase curves. Moreover, we have a first integral ˜G(r,θ) defined via the initial condition ˜G(r,θ)=r20 if r=ϕ(θ;r0).

    But, this function is not correct because its value for negative r's should be compatible with the 'square root' of the Poincaré map Q:(R,0)(R,0),Q(r0)=r0+, such that P=QQ on (R+,0). But, we can take a function defined initially on (R+,0) as G(r0)=12(r20+Q2(r0)). Thus,

    G(r,θ)=G(r0) (4.8)

    if r=ϕ(θ,r0). This first integral is holomorphic near the circle {r=0,0θ2π}C×R, where R is a ring around the circle R/2πZC/2πZ.

    In fact, we can extend it to a holomorphic first integral in Z,V. For this, one first defines it as G(x)=12(x2+Q2(x)) on the disc Σ={(x,0):xC,|x|<ε}C2x,y, or Σ={(z,ˉz):zC,|z|<ε}C2z,v, where Q is the extension of the square root of the return map. Outside of the disc Σ, this function is defined by the condition of being constant on the complex phase curves (Riemann surfaces). In particular, it is single-valued (see the second proof) and equals zero on the complex separatrices of the singular point.

    Second proof (due to Moussu [24]). Using the holomorphic Hadamard–Perron theorem, we can assume that the coordinate lines are invariant, i.e.,

    ˙z=z(iω+), ˙v=v(iω+).

    Then, the phase curves outside of W={z=0} are graphs of functions,

    v=φ(z;v0),

    which satisfy the initial value problem

    dvdz=v(iω+)z(iω+), φ(z0;v0)=v0,

    where z00 is fixed. The analytic continuation of solutions φ along the circle {z=eiαz0:0α2π} defines the monodromy map M:DD,D={z0}×{|v0|<ε}, as

    M(v0)=φ(e2πiz0;v0). (4.9)

    This map is holomorphic and we have

    M(v0)=v0+ekvk+10+O(vk+20), ek=(4πz0/ω)ak. (4.10)

    The center conditions from Theorem 1 mean that M=Id.

    Now, we can define a first integral F(z,v) by putting

    F(z0,v0)=z0v0 (4.11)

    on D and continuing it analytically as constant on the phase curves. Because the monodromy is trivial, the function F is single-valued outside of the line {z=0}. It is also bounded there. Hence, it is holomorphic.

    The function F has a Morsean critical point at z=v=0. By the Morse lemma [37], there exist analytic coordinates Z=z+,V=V+ such that F=ZV.

    The vector field is parallel to the analytic Hamiltonian vector field XF=FVZFZV, i.e., it equals ΦXF, where Φ=iω+ is an analytic function (the orbital factor).

    Finally, one can reduce Φ to an analytic function of one variable: Φ(Z,V)=iM(ZV); we refer the reader to [30] for the proof.

    Final remark. The integrating factor M from Eq. (4.6) is also analytic. If it is constant, then the center is isochronous and the period does not depend on the periodic solution.

    Recall that the spectrum of the linear part consists of λ1,2=±iω and of μ1,,μm (m1) in the left half-plane. Therefore, the resonant relations from Definition 1 are as follows:

    λ1=(p+1)λ1+pλ2, λ2=qλ1+(q+1)λ2,μj=pλ1+pλ2+(l,μ), |l|1.

    In the Poincaré–Dulac normal form, we have

    ˙Z=Z{iω+(ap+ibp)(ZV)p}

    and the corresponding equation for ˙V. Moreover,

    ˙Yj=YrFr(Y,ZV), j=1,,m.

    We see that the variables Z,V are formally separated from Yj and the center condition from Theorem 2 reads as

    ap=0, p=1,2,.

    Moreover, we have the stable invariant manifold Ws={Z=V=0} (which is analytic) and a formal center manifold Wc={Y=0}, about its analytic properties we cannot judge at this moment (see Example 1 above).

    But, the eigenvalue λ1=iω is separated from the remaining eigenvalues μ1,,μm,λ2 by a line in C through 0, and, analogously, λ2 is separated from the other eigenvalues. By the analytic Hadamard–Perron theorem, we have analytic invariant hypersurfaces: W+ corresponding to λ2,μ1,,μm and W corresponding to λ1,μ1,,μm; of course, Ws=W+W. We also have two invariant lines L+ and L corresponding to the eigenvalues λ1 and λ2, respectively; Wc is 'spanned' by L+ and L.

    We are ready to present the following:

    Proof of Theorem 2. We would like to follow the second proof of Theorem 1.

    Thus, we can assume the following analytic system:

    ˙z=z(iω+), ˙v=v(iω+), ˙y=Y(z,v,y);

    so, W+={z=0} and W={v=0}. The phase curves outside of W+ are graphs of functions of z.

    Let us choose a hypersurface of initial conditions with the form

    Σ={z0}×{|v|<ε}×{|y|<ε}

    (point×disc×ball) of complex dimension 1+m2. The solutions to the equations for dvdz and for dydz with the initial conditions v(z0)=v0 and y(z0)=y0 are of the form

    v=φ(z;v0,y0)=z0v0/z+, y=χ(z;v0,y0).

    We also have the monodromy map M:ΣΣ:

    M(v0,y0)=(φ(e2πiz0;v0,y0),χ(e2πiz0;v0,z0)).

    In contrast to the Poincaré–Lyapunov case, this map is far from the identity. It has the form

    M:(v,y)(v+,By+),

    where the spectrum of the matrix B lies in the open unit disc; B is a strong contraction. Therefore, we need an additional argument.

    On the one hand, we can use the normal hyperbolicity theorem, which ensures the existence of a 2dimensional invariant manifold Vc=Vcz0, which is the center manifold for M. For this, we treat M as a perturbation of the monodromy map M0 associated with a polynomial truncation of the Poincaré–Dulac normal form (when Vc={y=0}. The manifold Vc is unique.

    Let us construct the first integral F, beginning with its definition F0=F|Σ on Σ. This function should have the property

    MF0=F0;

    then, we will use the analytic continuation. Let

    Σ={z0}×{|v|<ε}×{|y|=ε}

    (point×disc×sphere). We put

    F0|Σ=z0v.

    We define the following family of maps hα:ΣΣ,α0:

    hα(v0,y0)=(eiαφ(eiαz0;v0,y0),χ(eiα;v0,y0)).

    Since φ(eiαz0;v;y)eiα, the maps hα do not change much for v. But, they are contractions when acting on y. Moreover,

    h2π=M.

    If (v,y)=hα(v0,y0),(v0,y0)Σ, then we put

    F0(v,y)=z0v0.

    This defines a correct analytic function on ΣVc, which is extended to an analytic function on Σ. Next, the latter function F0 is extended to a neighborhood of the singular point, i.e., we get the first integral

    F(z,v,y)=zv+.

    with the critical locus along the stable manifold Ws={z=v=0}. This follows from the center conditions.

    By the parametric Morse lemma [37], there exist analytic coordinates Z=z+,V=v+,Y=y+ such that

    F=ZV.

    From the center conditions, it also follows that

    (i) M|Vc=Id|Vc,

    (ii) Wc=z0Vcz0 is an analytic center manifold, Wc={Y=0}, which supports a family of periodic solutions

    Z(t)=ceit, V(t)=deit, Y(t)0.

    The corresponding family

    z(t)=ceit+, v(t)=deit+, y(t)=

    depends analytically on the parameters c,d (defined by the initial conditions). In the real domain, we have the family z(t)=reit+,ˉz(t)=reit+,y(t)=, with the real parameter r; this is the convergent family from [21].

    Remark 5. We see that the existence of an analytic family of periodic solutions, under the assumptions of Theorem 2 (and also of Theorem 3), implies the uniqueness and analyticity of the center manifold Wc(C2,0). This surface is parametrized by complex parameters c,d and ζ=(1+δ)eit (from a neighborhood of the unit circle):

    z=cζ+, v=dζ1+, y=.

    (But, this situation is somewhat special In Example 5\ below, we encounter a 1–parameter family of periodic solutions spreading non-smooth surfaces.)

    Next, one can show that the Poincaré–Dulac normal form

    ˙Z=iZM(ZV),˙V=iVM(ZV), ˙Y=Ω(Y,ZV),

    where M is a real series and Ω is polynomial in Y (with series in VZ as coefficients), is analytic.

    Finally, note that, in Example 1, the formal center manifold does not support a family of periodic solutions; formally, it has an unstable focus.

    Recall that we have pure imaginary eigenvalues ±iω1,,±iωm,ωj>0. Let ˙zj=iωjzj+,˙vj=iωjvj+.

    Let us describe the Poincaré–Dulac (P–D) normal form for such systems. Assume first that the frequencies ωj are independent over Q. Then, the only resonant relations are

    iωj=k1iω1+k1(iω1)++(1+kj)(iωj)+kj(iωj)++kmiωm+km(iωm),iωj=k1iω1+k1(iω1)++kjiωj+(1+kj)(iωj)++kmiωm+km(iωm).

    It follows that the P–D normal form is

    ˙Zj=Zj(iωj+fj(Z1V1,,ZmVm)), ˙Vj=Vj(iωj+gj(Z1V1,,ZmVm)), (4.12)

    j=1,,m. If there are relations, like

    pω1=qω2, gcd(p,q)=1, (4.13)

    then the P–D normal form will contain additional monomial terms, like

    Zj(Zq1Vp2)rZj, Vj(Vq1Zp2)rVj, Vrq11Zrp2Z1

    (and some other in the case of (4.13)).

    Anyway, the sublattice (in Zm) of relations (k,ω)=0 is finitely generated and the P–D normal form has finitely many series (formal functional moduli) depending on finitely many monomials.

    Assume now that the frequency ω1 is such that no other frequency ωj is a multiple of it:

    ωj/ω1Z, j2. (4.14)

    Lemma 1. Under the which assumption, the P–D normal form is such that the equations

    Z2=V2==Zm=Vm=0

    define a formal invariant manifold (invariant surface) W1.

    In this case, the P–D normal form restricted to W1 is like in Eq. (4.2) with the center conditions ak=0,k=1,2,.

    Proof. The obstacles to the existence of such an invariant surface are implied by the following terms:

    Zk1Vl1Zj, Zk1Vl1Vj, (j2).

    They correspond to the resonant relations

    iωj=(kl)iω1, iωj=(kl)iω1.

    By condition (4.14), there are no such relations.

    Below we give two proofs of Theorem 3. Let us being with Lyapunov's argument. For this, the following standard is needed.

    Lemma 2. Consider the linear equation

    dxdθ=λx+a(θ),xC,

    where a(θ) is a periodic function with period 2π. If λ is not an integer multiple of i=1, then this equation has a unique periodic solution with the period 2π defined by

    x=ϕ(θ)=(e2πλ1)1θθ2πeλ(θα)a(α)dα. (4.15)

    Proof. The solutions of the initial value problem with the initial conditions x(0)=x0 are affine in x0. After the period of 2π, these solutions define the monodromy map x0M(x0), which is also affine:

    M(x0)=Λx0+A,

    where Λ=e2πλ. By the assumption, Λ1. Hence, the map M has a unique fixed point corresponding to a unique 2πperiodic solution.

    If Re(λ)<0, then this solution is as follows:§

    §This formula was used by Lyapunov in his proof of Theorem 2.

    x=ϕ(θ)=θeλ(θα)f(α)dα.

    After partition of the integration domain (,θ] into segments of length 2π, we get Eq. (4.15), which is valid for all λC2πZ.

    Lyapunov's proof of Theorem 3. Let z1=reiθ and v1=ˉz1=reiθ, where r,θ are polar coordinates. The phase curves of our system are defined by the equations

    r=f(r,θ,˜z,˜v),zj=i(ωj/ω1)zj+gj(r,θ,˜z,˜v),vj=i(ωj/ω1)vj+hj(r,θ,˜z,˜v), (4.16)

    j=2,3,, where ˜z=(z2,,zm),˜v=(v2,,vm) and =d/dθ. We can assume that f begins with quadratic terms in r,˜z,˜v. For simplicity, assume that the linearization matrix of the complex system is diagonal; the triangular case requires only slight modification of the proof. Thus, also gj and hj begin with quadratic terms.

    We look for 2πperiodic solutions to system (4.16) as power series with periodic coefficients:

    r=ϕ(1)(θ)c+ϕ(2)(θ)c2+zj=ψ(2)jc2+ψ(3)jc3+,vj=χ(2)jc2+χ(3)jc3+. (4.17)

    We set the following initial conditions:

    ϕ(1)(0)=1, ϕ(2)(0)=0,ϕ(3)(0)=0,; (4.18)

    thus, we have that r(0)=c and ϕ(1)(θ)1 (which is 2πperiodic). Substituting the Ansatz (4.17) into system (4.18), we obtain a recurrent system of equations for ϕ(k),ψ(k) and χ(k).

    We have

    dϕ(k)/dθ=Φ(k)(θ), k2,

    where the functions Φ(k)(θ) are defined inductively and should be 2πperiodic. Thus,

    ϕ(k)(θ)=θ0Φ(k)(α)dα, (4.19)

    where 2π0Φ(k)(α)dα should equal zero. The latter vanishing conditions are guaranteed by the assumptions of the theorem, i.e., the center conditions.

    The equations for ψ(k)j take the following form:

    dψ(k)j/dθ=λjψ(k)j+Ψ(k)j(θ),

    where λj=iωj/ω1 and Ψ(k)j(θ) are defined inductively and are 2πperiodic; analogous equations hold for χ(k)j, but with λj replaced with λj. The latter equations dψ(k)j/dθ are like the equation from Lemma 2 with λ=λjiZ. So, they have unique 2πperiodic solutions defined in Eq. (4.15); no center conditions are needed here.

    The series in Eq. (4.17) is convergent due to the analyticity of the right-hand sides of the equations of Eq. (4.16), the compactness of the domain [0,2π] of definition of this system and the simplicity of the integral formulas (4.15) and (4.19). In [21], suitable estimates are given.

    Second proof of Theorem 3. Recall that we have the formal invariant surface W1, defined in Lemma 2. We will show that it is analytic.

    For this, we use the Poincaré return map P defined by solutions after the period of 2π of system (4.16). We have

    P(r,z2,,zm)=(ρ,ζ1,,ζm), (4.20)

    where ρ=ρ(r,z2,,zm)=r+,ζj=ζj(r,z2,,zm)=μjzj+,μj=e2πiωj/ω11. The periodic orbits of the period 2π/ω1 correspond to the fixed points of the Poincaré map. The set of fixed points of P lies in the curve

    ζ2z2==ζmzm=0, (4.21)

    where ζjzj=μjzj+ with μj0. By the implicit function theorem, Eq. (4.21) defines a real analytic curve.

    By the center conditions, this curve is P–invariant and defines an analytic smooth surface supporting a family of periodic solutions.

    The following example demonstrates the importance of the condition (4.14).

    Example 3. The system

    ˙z1=iz1, ˙v1=iv1, ˙z2=2iz2+z21, ˙v2=2iv2+v21

    has a general solution with the form

    z1=c1eit, v1=d1eit, z2=(c2+c21t)e2it, v2=(d2+d21t)e2it.

    Only the solutions with c1=d1=0 are periodic, with a period of π (not 2π).

    Remark 6. (a) We see that the existence of a 1–parameter family of periodic solutions can be deduced from the P–D normal form. We need the following:

    (i) formal invariant surface, and

    (ii) vanishing of the series of Poincaré–Lyapunov coefficients after restriction of the P–D normal form to this surface.

    The following questions arise naturally in this context.

    1. Are there other 1–parameter families of periodic solutions?

    2. Are there many–parameter families of periodic solutions?

    Concerning the first question, there should exist other formal invariant surfaces with vanishing corresponding Poincaré–Lyapunov quantities.

    The answer to the second question can be positive in some resonant cases. For example, this holds for the following system:

    ˙zj=ipjzjM, ˙vj=ipjvjM, (4.22)

    with the integer pj and the integrating factor M=ω+. Here, the typical solution is periodic with the period 2πpω(1+), where p=gcd(p1,,pm).

    But, when

    ˙zj=ipjzjMj, ˙vj=ipvvjMj, (4.23)

    where the factors Mj=ω+ are independent, then the motions take place on the invariant 'tori' {z1v1=C1,,zmvm=Cm}. In the real domain, where vj=ˉzj, the motion on the generic torus is quasi-periodic, or periodic with a very long period.

    (b) Note that, in the case considered in this subsection, the system has two analytic invariant manifolds, W+ and W, corresponding to the partition of the eigenvalues' set into two subsets {iω1,,iωm} and {iω1,,iωm}. The system restricted to either of these submanifolds has analytic P–D normal form (the systems are in the Poincaré domain).

    Here, we discuss Theorem 3 (as well as the Weinstein- and Moser-type theorems) in the context of Hamiltonian systems. Again, we assume the imaginary eigenvalues ±iω1,,±iωm,ωj>0, where m=n2 is the number of degrees of freedom. The Hamiltonian functions considered in this section are of the form

    H=H2+H3+, (5.1)

    where H2 is like in Eq. (2.5) and the functions Hj are homogeneous of degree j. Such an H2 is obvious in the case that all frequencies ωj are pairwise different. The opposite case is discussed below.

    Recall that, in the third Lyapunov theorem, there are two assumptions: the existence of a formal invariant surface W1 associated with the eigenvalues ±iω1, and the vanishing of the series of the Poincaré–Lyapunov focus quantities after restriction of the system to W1 (see Remark 6 above). But, in the Hamiltonian case, the situation is simpler.

    Proposition 1. If there exists a formal invariant surface W1, then all of the focus quantities for the restriction of the system to W1 vanish.

    Proof. Indeed, H restricted to W1 is a formal first integral for the restricted system. So, in the real domain, all phase curves of that system are closed.

    People have tried to improve the third Lyapunov theorem in the Hamiltonian case. Recall that Weinstein ([35, Theorem 2.1] and Theorem 4 above) proved that, if H2 is positive definite, i.e., all ϵj=1, then there exist at least m=n2 families of periodic solutions.

    Moreover, Moser, in [23, Theorem 4], claimed that, if H2 restricted to the maximal linear invariant subspace E supporting periodic solutions of period T=2π/ω of the linearized system is positive definite, then 'on each energy surface H=ε2, the number of periodic orbits is at least 12dimE'. Recall that that result is wrong, but that some version of Moser's statement is valid provided that H2 is positive definite (Theorem 5).

    Of course, the number m from the Weinstein theorem cannot be improved because the Hamiltonian system

    ˙zj=iωjzjMj(|zj|2), j=1,,m,

    with different factors Mj, has m families of periodic solutions with independently varying periods Tj=Tj(|zj|2)2π/ωj. Compare with Remark 6 above.

    Recall also Example 3 above, where we observe 1:2 resonance between the frequencies and only one family of periodic solutions; but, there, the system is not Hamiltonian.

    Example 4. In the case of the Hamiltonian

    In [35], we find another Hamiltonian H=12|z1|2|z2|+12Im(z21z2) (due to Siegel [29]), also with only one family of periodic solutions.

    H=12{|z1|2k|z2|2+zk1z2+ˉzk1ˉz2}, k2, (5.2)

    from [23, Example 2], we deal with the resonance, but with indefinite quadratic part. The corresponding equations for are

    Of course, the plane is invariant and supports a family of periodic solutions with the period

    But, the Lyapunov-type function satisfies

    which is positive in the domain It prevents the existence of other periodic solutions.

    Surprisingly, this example contradicts Moser's original statement [23, Theorem 4]. Indeed, we have two invariant subspaces of the linear system supporting periodic solutions: and We have that and But, only periodic solutions from are 'extended' to periodic solutions of the whole system.

    We complete this example by analysis of some return map associated with the polar coordinates in the plane. Thus, we put

    (5.3)

    We have

    Next, gives

    We get the system

    (5.4)

    which we call the return system. The twisted Poincaré map is defined via the evaluation of solutions at the 'time' with the initial value at the 'time' The periodic solutions with the period and outside of correspond to the singular points of the latter system (5.4) with Thus, we have

    i.e., and

    (5.5)

    The latter equation does not have nontrivial solutions.

    The following example demonstrates that, in the case of a positively definite Hamiltonian, the situation is quite different.

    Example 5. (a) Consider the positive definite Hamiltonian

    (5.6)

    generating the system

    (5.7)

    Of course, the plane is invariant and has a a family of periodic solutions with the period

    (b) To look for periodic solutions with the period , use the polar-type coordinates

    (5.8)

    compare with Eq. (5.3). We get the return system

    (5.9)

    Again, the closed phase curves of the period outside of correspond to the singular points of the vector field given by Eq. (5.9). We find that

    (5.10)

    Therefore, we have three families of periodic solutions. The first family spans the invariant plane but none of the two other families of closed phase curves span a smooth surface. These real surfaces (in take the form

    and are singular; in , we get one complex singular surface

    The number of families of periodic solutions is greater than in the Weinstein theorem.

    (c) To explain the this difference we use the fact that the Hamiltonian system in the Birkhoff normal form is invariant with respect to the action. In the present case, the action is as follows:

    (5.11)

    it is symplectic and generated by a corresponding Hamiltonian function (the momentum map), namely, (which is a first integral for system (5.7)). The symplectic reduction procedure (see [4]) says that, for each level surface (which is invariant), the quotient space with respect to the action acquires a symplectic structure and we get a Hamiltonian vector field in this quotient space. The singular points of the latter vector field correspond to periodic solutions to system (5.7).

    In this case, the hypersurface is a 3–sphere and the quotient space turns out to be the weighted projective space, (see Section 10.1). The weighted projective space is homeomorphic with the standard complex projective line The natural charts in are and thus, and .

    In these charts, we get the systems

    (5.12)

    Note that, in the chart on , we have that so, for , we find that is separated from zero and the first term on the right-hand side of the equation for is dominating. It follows that the latter system(s) (5.12) has three singular points: (with index and (both with index . As the Euler characteristic , this agrees with the Poincaré–Hopf formula.

    Note also that the equilibrium points of correspond to the critical points of as a function on This amounts to critical circles of the function , restricted to . With and the Lagrange multiplier, we get the equations

    (5.13)

    One solution is , and For the other solutions, we have thus, and

    (d) Finally, the Hamiltonian system (5.7) is completely integrable, with two independent first integrals (in involution). By the Liouville–Arnold theorem [3], the common level surfaces which are compact, are tori (if smooth). There are also the action–angle variables, which are our objectives.

    Let and We have the differential equations

    The levels of the first integrals take the form

    This leads to

    where Finally, we arrive at the system

    (5.14)
    (5.15)

    We have an oval in the plane defined by Eq. (5.14). This is one of the circles generating the invariant torus. The corresponding Liouville–Arnold angle is

    (5.16)

    where is a path in from to ( is the left root of the equation and is the corresponding period. It is the period of the solution to Eq. (5.14).

    Next, the second generating circle is parametrized by the angle The corresponding solution to Eq. (5.15) takes the form We have the second period The second Liouville–Arnold angle equals

    (5.17)

    One can find corresponding action variables from the condition For this, one has to solve some first order PDE; we omit the details.

    Consider now the situation with resonance and, more generally, with resonances. Here, non-diagonal cells for the linear systems associated with homogeneous quadratic Hamiltonians are expected. Canonical forms of such Hamiltonians are given in [3, Appendix 6].

    Probably, the following easy statement was a motivation of Moser's work [23].

    Proposition 2. In the case of resonance, i.e., with and of definite (positive definite or negative definite), the corresponding linearization matrix is diagonalizable.

    Proof. If there existed a pair of Jordan cells, then the general solution of the corresponding linear system would be unbounded. But, this would contradict the compactness of the level hypersurfaces

    It is natural to look for families of periodic solutions in the cases with a positive definite i.e., with a diagonalizable linear part. How many such families should exist?

    Consider first the positive definite and a homogeneous perturbation in the Birkhoff normal form. Thus, we have

    (5.18)

    where and is a homogeneous polynomial depending only on the monomials , Here, we have some freedom. Namely, we can apply linear changes of the variables which leave invariant. The corresponding group is ; this group preserves the Hermitian product whose real part is the scalar product and whose imaginary part defines the symplectic form. Using these changes, one can reduce the our situation to the case with invariant complex planes.

    (Note that, in the case with the indefinite quadratic part, , where and , the corresponding group is and the Birkhoff resonant terms depend on and Also, here, invariant planes can be found for

    Example 6. A general quartic Hamiltonian with the form of Eq. (5.16) for with two invariant complex planes is of the form

    (5.19)

    with real constants The corresponding system becomes

    We study the twisted Poincaré map using the substitutions and It leads to the equations

    (we do not write equations for and

    The periodic solutions with period the outside of are given by and Thus, either (1) or (2) Let

    In case (1), we get the equation

    and, in case (2), we get

    It follows that, for generic values of the parameters, besides the solutions and (corresponding to periodic solutions of the Hamiltonian system in the invariant complex planes), we have four additional solutions. Thus, the total number of periodic solutions in the generic case is six.

    On the other hand, by analogy with the previous example, the symplectic reduction, with and leads to Hamiltonian vector field on with One can check that the additional two pairs of singular points of have opposite indices.

    The above agrees with the following result of van Straten [34].

    Theorem 14. We have

    (5.20)

    where, on the left-hand side stands the generating function for the numbers of periodic modes for Hamiltonian systems with a generic Hamiltonian function in the Birkhoff normal form given by Eq. (5.18).

    For , the right-hand side of Eq. (5.20) becomes which agrees with Example 6; in fact, there are no such examples in [34].

    Finally, we have the following Hamiltonian systems without periodic solutions for the expected period.

    Example 7. Assume the following Poisson brackets for the complex variables :

    The Hamiltonian function equals

    and generates the system

    Note that, for , we get a nilpotent linear system with invariant plane that supports a family of periodic solutions; this Hamiltonian can be found also in [3,7]. But, we assume that ; hence, the planes and are not invariant.

    One should look for periodic solutions for the period Hence, we make the natural change (look at the previous examples): with We get the equations

    The equations for the equilibrium points are given by (i.e., and i

    We refer also to previous works [12,33] devoted to bifurcations of Hamiltonian systems in the 1:1 resonant case.

    In the last example from [22, Example 9.2], the linear part is diagonal.

    Example 8. Let

    It generates the system

    One finds that

    which excludes the existence of nontrivial periodic solutions.

    Again, assume an analytic system (1.4) with the eigenvalues Recall the equivalence relation (1.9) on the set of frequencies. For each equivalence class , we have a linear subspace that is invariant for the linear part of system (1.4). In fact, we can say more.

    Proposition 3. For each class , system (1.4) has an invariant analytic submanifold tangent to at the origin.

    Proof. First, it is rather obvious that there exists such an invariant manifold at the formal level. To prove its analyticity, we use the Poincaré return map (4.20), introduced in the second proof of Theorem 3.

    Let be the linear eigenfunctions associated with one equivalence class, say, and let be the remaining eigenfunctions.

    We put and get Eq. (4.16) for the phase curves, as graphs of functions of . We have that and . Let The solutions to Eq. (4.16) after 'time' define the Poincaré return map (like in Eq. 4.20)).

    The linear part of this map has the form

    for in fact, are is not a root of unity.

    Then, the equations define an analytic submanifold . It generates the submanifold , where is the 2–parameter family of diffeomorphisms defined by solutions to Eq. (4.6).

    Therefore, our problem is reduced to the case when there is only one equivalence class, i.e., that all of the frequencies are rationally related. Let us order these frequencies like in Eq. (1.10), i.e.,

    Proposition 4. In this situation, there exists a series of analytic invariant submanifolds (of dimensions tangent at the origin to linear subspaces) associated with corresponding groups of frequencies.

    Proof. It essentially repeats the previous proof. We find other submanifolds are obtained inductively.

    Let be the eigenfunctions associated with the eigenvalues and be the remaining eigenvalues (associated with We put and get the equations

    for phase curves. Their solutions after the 'time' define the Poincaré map

    where the coefficient is a root of unity. The equations define a submanifold which is spread to the submanifold

    From now on, we focus our attention on the case with only one equivalence class for the frequencies. Thus, we can write

    (6.1)

    Proposition 5. In the case of Eq. (6.1), the Poincaré–Dulac normal form is invariant with respect to the following action of the circle :

    (6.2)

    Proof. Recall that a term is resonant in the P–D normal form if and only if the resonant relation holds. But, the result of change, given by Eq. (6.2), on is as follows:

    Only such terms are invariant.

    Recall that the assumptions of the three Lyapunov's theorems are two-fold. One has an assumption about the eigenvalues of the linear part; it implies the existence of a formal invariant smooth surface (it is rather irrelevant in Theorem 1). The second assumption means that the focus quantities of the system restricted to the formal invariant surface vanish; it amounts to the existence of a family of periodic solutions at a formal level. We underline that this vanishing condition admits a precise definition in terms of the P–D normal form; but, there are other approaches.

    To provide a precise assumption of Theorem 8, we definitely need the P–D theorem. It roughly states that the P–D normal form predicts the existence of a 1–parameter family of periodic solutions for the period at a formal level. Consider the equivalence class of frequencies which contains . By Proposition 3, we can focus our attention on this class, i.e., that we have the situation of Eq. (1.10). By Proposition 4, there is an invariant submanifold associated with the frequencies such that Therefore, we can assume that

    (7.1)

    i.e., besides property (1.10).

    Assume the Poincaré–Dulac normal form (with the coordinates . Following Examples 4 and 5, introduce the change

    (7.2)

    Lemma 3. We have the following return system for phase curves:

    (7.3)

    where the right-hand sides do not depend on and the dots mean nonlinear terms.

    Proof. We have that and where the right-hand sides do not depend on due to the P–D normal form.

    Definition 4. We say that the vector field given by (1.4) satisfies a center condition if system (7.3) has a formal curve staring at of (non-isolated) equilibrium points.

    Theorem 8 states that, in this case, there exists a genuine continuous 1–parameter family of periodic solutions to system (1.4).

    Proof of Theorem 8. Analogously to the change given by Eq. (7.2), introduce the change

    We get a system of equations (the return system) with analytic right-hand sides. The periodic solutions of the period correspond to the fixed points of the twisted Poincaré map which is defined via solutions to the latter system and after the 'time'

    The equation has a zero locus along an analytic curve which is formally defined as the curve from Definition 4 (after the change

    Remark 7. The statement of Theorem 8 can be generalized to the case when system (7.3) has a formal singular locus of higher dimension, say, . Then, the equation has an analytic zero locus of dimension it corresponds to a parameter family of periodic solutions for the period

    The assumptions of Theorems 1, 2, 3 and 8 involve an infinite number of conditions, related to either the vanishing of all focus quantities after restriction to a formal invariant plane or the existence of a curve of non-isolated singular points for system (7.3). These phenomena have infinite codimension.

    In this section, we avoid the infinite number of conditions by imposing an additional restriction to the class of considered vector fields and the existence of an analytic first integral with a regular leading part. This restriction is also of infinite codimension, but it is natural in some applications.

    Theorem 15. Assume that analytic system (1.4) has pure imaginary eigenvalues and that it has a first integral such that is definite on a linear subspace (of dimension corresponding to one pair of eigenvalues such that, for any other frequency , one has Then, there exists at least one 1–parameter family of periodic solutions for the period

    Proof. By Proposition 4, we reduce the situation to one equivalence class of frequencies containing thus, we can assume that the class and in Eq. (1.10). By Proposition 5, we can assume and there is a definite first integral , say, with positive definite .

    Assume first that the P–D normal form is analytic. So, let the system be in this form; denote the corresponding vector field (with variables Recall that such a system is invariant with respect to the corresponding action (Proposition 6). By averaging,

    (8.1)

    we get that the first integral can be chosen also as invariant:

    The level hypersurfaces are diffeomorphic with a dimensional sphere. They are invariant for the vector field Like in [23], we obtain vector fields on the quotient varieties . The vector fields ' are diffeomorphic with the complex projective space:

    So, the Euler characteristic

    By the Poincaré–Hopf theorem [14], (where denotes the indices of at the isolated singular points , and the vector field has at least one singular point (if its singular points are isolated). This amounts to the existence of a periodic solution of on This implies the existence of a parameter family of periodic solutions like in the thesis of Theorem 15.

    Let us present our approach without the assumption of analyticity of the P–D normal form.

    Suppose that there are no families of periodic solutions as above. We shall construct a real vector field on the projective space , which is a P–D approximation of the vector field

    We take an invariant approximation of the first integral we take the average given by Eq. (8.1).

    We project onto the tangent spaces of the level hypersurfaces along the radii, such that the projection is equivariant; we get a regular vector field in the P–D normal form.

    Next, we quotient and by and obtain a vector field on without singular points. This contradicts the Poincaré–Hopf theorem.

    In the proof of Theorem 15 (in the previous section), we have used, essentially, the Poincaré–Hopf formula. But, that formula does not give us much information about the singular points of the quotient vector field on the quotient variety it is because there is no restriction for the values of the indices. In particular, we cannot get any bound for the number of singular points of

    Probably, the only topological tool that could be used in this problem is the Lusternik–Schnirelmann category. The corresponding theorem (Theorem 17 from Section 10.2) provides the estimate from below for the number of critical points of a sufficiently smooth function on a manifold based on the category of this manifold.

    Weinstein [35] used this idea in the case of germs of Hamiltonian vector fields generated by Hamiltonian functions such that are definite, and in fact, positive definite. He skillfully constructed a function on the level hypersurface which has a critical locus at the set of periodic phase curves of in His construction is not direct and involves many technical details.

    Below, we propose a more direct proof.

    Proof of Theorem 4. Recall that we have a Hamiltonian vector field in with nonzero pure imaginary eigenvalues of the linear part and with a positive definite Thus, where

    with (compare with Eq. (2.5)). Here, we can reduce to the Birkhoff normal form, which is an analogue of the P–D normal form. Like in the proof of Theorem 15 (in the previous section), we can reduce the situation to one equivalence class of frequencies, i.e., we assume Eq. (7.1).

    Assume first that the Birkhoff normal form is analytic. So, let be in this form.

    Proposition 6. The corresponding action of on defines a symmetry of the vector field This action is symplectic; in fact, it is Hamiltonian with

    as the momentum map.

    Proof. We have to prove the second statement. Note that the phase flow generated by the is and coincides with the action

    Now, we apply the symplectic reduction. Recall that, in the case of the general (non-abelian) symmetry group, the symplectic reduction concerns only the zero level of the momentum map (see [4]). Here, the group is abelian and the symplectic reduction works for any level of the momentum function; in fact, it is the simplest case of the symplectic reduction.

    So, we take the manifolds

    which are diffeomorphic with , and their quotients

    of dimension The varieties are equipped with a natural symplectic structure and support fields obtained from Each vector field is Hamiltonian with the Hamiltonian function where is the projection and

    contains higher-order terms.

    Proposition 7. The periodic orbits of the in of period for some correspond to the critical points of the function on

    Proof. The periodic orbits from this proposition are somewhat distinguished. They are defined via the return vector fields of type (7.3), i.e., they are associated with a fixed choice of the frequency and for (We do not consider possible orbits of a very long period.) The singular points of correspond to such periodic orbits.

    But, the singular points of , when considered on , are the equilibrium points of They are the critical points of , i.e., of , on Here, there is some subtlety associated with the fact that is singular, but one can use the local charts in defined in Eq. (8.2) below.

    Since is a sphere, its quotient is the weighted projective space By the Lusternik–Schnirelmann theorem (Theorem 16 in Section 10.2), the number of critical points of the function is bounded from below by the category of But, is homologically 'similar' to (see Section 10.1) and (see Theorem 17 from Section 10.2).

    The category of quotient spaces was studied also by Weinstein in [36].

    This gives the estimate for the number of isolated periodic orbits.

    Now, we present our argumentation without the assumption of the analyticity of the Birkhoff normal form. We have essentially two tools at our disposal (we have used them in the proof of Theorem 8). Again, we assume only one equivalence class of the frequencies.

    The first tool is the Birkhoff normal form, which we assume only to be formal.

    The other is the Poincaré return map (as defined in the second proof of Theorem 3, see Eq. (4.20)), or, rather, the twisted Poincaré map (as defined in the proof of Theorem 8 and in Examples 4–6).

    In fact, we will use a collection of twisted Poincaré maps via the substitutions

    as well as the solutions after the 'time' of corresponding differential systems. The variables and parametrize corresponding twisted Poincaré sections.

    As before, we assume that with relatively prime positive integers such that The twisted Poincaré map differs from the standard Poincaré map (where in the coordinates such that In the standard case we have that and, in the twisted case, we have that

    If the system (1.7) were in an analytic Poincaré–Dulac–Birkfoff normal form with an invariant first integral , then the variables and when restricted to , would form a local chart

    (8.2)

    in the quotient variety or an affine chart

    Recall that the 1–parameter families of periodic solutions of system (1.7), whose corresponding vector field we denote by of period correspond to 1–dimensional curves of fixed points of the twisted Poincaré map (see the above proof of Theorem 7). They are analytic curves defined by the analytic equation We assume that this equation has a zero locus of dimension otherwise, we have infinitely many 1–parameter families of such periodic solutions.

    From the analytic geometry, we know that (germs of) analytic sets are analytically equivalent (diffeomorphic) with algebraic sets. This holds in the case of complex analytic sets and in the case of real analytic sets. For example, a germ of an analytic curve is defined by a corresponding Puiseux series. The proof uses the Weierstrass preparation theorem (see [37]). Moreover, if such a germ is defined by a system of germs of analytic functions, then an approximation of these functions by polynomials (sufficiently long jets) define an algebraic set locally diffeomorphic with the original germ.

    We apply this to the germs of curves defined by the equation These equations admit the approximation where the twisted Poincaré map is defined via a Hamiltonian vector field generated by a polynomial approximation of in the Birkhoff normal form. Then, the curve will correspond to a curve The number of the latter curves is estimated like above.

    Recall also that, in the original version of the work [23, Theorem 4], Moser assumed only that where is the linear subspace associated with one frequency for the linear system, but his example from [23, Example 2] (or our Example 4 above) provides a contradiction to this restricted version.

    It seems that the crucial point in Moser's approach is his lemma (Lemma 1) from [23]. There, he somehow associates with any small vector a solution to the Hamiltonian system and claims that some of them are periodic; probably, therein lies the flaw of his argument. Finally, he uses the Lusternik–Schnirelmann category for a suitable function on

    Our approach is more direct. In part, it repeats the above proof of the Weinstein theorem.

    Proof of Theorem 5. Much of his proof repeats the corresponding part of the proof of Theorem 4 in the previous section. In particular, we can assume that the Hamiltonian function is in an analytic Birkhoff normal form. Recall that we assume only one equivalence class among the frequencies.

    Thus, we have the additional first integral which defines the momentum map for the action of the group

    Recall that, by Proposition 5, we have the filtration of submanifolds that are invariant for the vector field such that is associated with one frequency. This filtration induces corresponding filtrations in and in

    where denotes weighted projective spaces with complex dimensions

    We have to show that as a function on has at least critical points in But, this follows from Theorems 16 and 17 in Section 10.2.

    Topological theorems of nonlinear functional analysis are well suited to deal with nonlinear PDEs. But, some specialists began to use them to prove (or reprove) some qualitative results in the theory of ODEs; see [22] for an example.

    In particular, Szulkin [32] developed a functional analytic approach to the question of the accumulation of small-amplitude periodic solutions to Hamiltonian systems. His method was followed in papers by Rybicki with various collaborators [10,15,25,26,31].

    The first trick in the analysis of periodic solutions for the period which usually varies with variation of the solution, is to normalize this period to This is done by replacing the Hamiltonian system with the following family of Hamiltonian systems:

    (8.3)

    where is a parameter chosen (for a given periodic solution such that the corresponding solution to Eq. (8.3) has the period . One says that a family of periodic solutions to Eq. (8.3) emanates from , when , with close to , denotes solutions to with the period

    Next, it is not difficult to show that the periodic solutions to system (8.3), where correspond to the stationary paths of the following functional (see also [23]):

    (8.4)

    This functional is defined on the periodic path The author considers as a functional on the Sobolev space using complex vector-variables and , we have that and

    The first part of the functional (8.4) takes the form

    With the splitting of the above Hilbert space into subspaces generated by negative, zero and positive modes, we have that In particular, the subspace is invariant for

    Assume that with We have the quadratic functional

    if like in Eq. (1.8), then It turns out (see [32]) that the variational derivative is a compact operator in , and that the functional satisfies the so-called Palais–Smale condition (which we do not define here).

    Consider the operator and its restrictions to the subspace , which are as follows:

    We see that the operator is non-invertible if and only if for some this is the case when the linear system has a family of periodic solutions for the period

    Consider the splitting where and The subspace is invariant for and has a quadratic form. The operator is a bounded self-adjoint operator in . Let be subspaces of (invariant for and be corresponding orthogonal projections. We have the Morse indices i.e., the numbers of negative eigenvalues. One defines another index

    (It has a cohomological interpretation; some cohomology groups , which we do not define, vanish for and are equal to otherwise; see [32, Proposition 2.3].)

    An abstract statement, i.e., [32, Theorem 2.3], says the following:

    Proposition 8. If (i) is invertible for where is a neighborhood of and (ii) the function has a jump at then either (a) the path is a non-isolated critical path of , or, (b) for all 's in from one side of , the functional has a critical path near

    The quantity is explicitly calculated in terms of the matrix we have

    (see [32, Proposition 3.3]).

    Assume the situation like in Eq. (1.10), i.e., that we have different frequencies, (of multiplicity (of multiplicity etc., possibly belonging to different equivalence classes. Let us fix It turns out that, if for all then jumps at if and only if jumps.

    Next, jumps at if and only if

    (8.5)

    where is the invariant subspace for corresponding to the eigenvalues (see [32, Proposition 3.6]). Condition (8.5) means that, if is like in Eq. (1.8) and , then for from Eq. (1.8).

    This yields the Szulkin's Theorem 6 from Introduction.

    We see that the first assumption of Theorem 6, i.e., is like in the third Lyapunov theorem. But, the assumption, i.e., inequality (8.5), is new; it replaces the definiteness assumption of from Theorems 4 and 5. Note also that, in this topological approach, one does not assume the analyticity of the Hamiltonian function. But, when one additionally assumes the analyticity of the Hamiltonian function, then the statement of Theorem 6 can be strenghtened: there exists at least one continuous family of periodic solutions.

    At this moment, the method to reprove this result using the proposed approach is unknown by the author. But, in all examples considered by the author, my tools seem to be more effective. In fact, examples of where only the Szulkin theorem works are not known by the author.

    Geostationary orbits of celestial objects, like GPS satellites, are determined by the property by which they are fixed with respect to an Earth's observer. The following definitions were taken from Strzelecki's thesis [31].

    Regarding the position and momentum framework of , associated with the moving object, we get a Hamiltonian system with the following Hamiltonian:

    (9.1)

    where is the Earth's angular velocity and the gravity potential (generated by an ellipsoid of rotation) equals

    where is the quadratic Legendre polynomial, and is a constant evaluated experimentally. The corresponding Hamiltonian equations

    (9.2)

    imply the following second-order Newtonian equations:

    here, and are the Coriolis and the centrifugal forces caused by the rotation of the coordinate frame.**

    **Such systems were already considered by Lyapunov [21, Section 45]

    The function (9.1) is invariant with respect to the simultaneous rotations in the and planes. This action is symplectic, and the corresponding momentum map is the vertical component of the classical angular momentum, The symplectic reduction (see Section 10.1) means the introduction of the coordinates where is the angle between the vectors and Thus, , . On the hypersurface , where , we have

    Next, implies and So, we get the reduced differential system on :

    (9.3)

    The equilibrium points are found as follows. First, since depends on We have i.e., , and we choose Next, from Eq. (9.2), we get and Thus, implies , and we get the following equation for :

    for it has a unique positive solution.

    Let us linearize system (9.3) at this equilibrium point. First, Next, Finally, for a small , we have that So, the linearization, with , , is as follows:

    with positive constants It follows that there is a pair of pure imaginary eigenvalues and a pair of real eigenvalues . Since the system is Hamiltonian, we have a 1–parameter family of periodic solutions for the period on the 2–dimensional center manifold .

    But, for the Hamiltonian system with three degrees of freedom (before the symplectic reduction), we have, in fact, invariant tori, with the additional angle The movement on these tori is quasi-periodic; we have

    In Strzelecki's approach [31], the same result was obtained using the so-called equivariant Conley index.

    In the restricted four-body problem, we deal with three heavy bodies, (from the sun), (from Jupiter) and (from a planet), and one light body (an asteroid). The heavy bodies lie in the vertices of an equilateral triangle which rotates with constant angular velocity (equal about their center of mass and in a fixed plane. The asteroid moves in the gravity field generated by the principal bodies and does not influence their movement. ††

    ††Recall that periodic solutions near the triangular libration point for the restricted three-body problem, i.e., with only and were studied by Schmidt [28]. I would like to thank one of the reviewers for this reference

    As in the previous section, one passes to the coordinate–momentum frame wherein the heavy bodies rest. Thus, we have the Hamiltonian function like in Eq. (9.1), but with and

    (9.4)

    here, is the center of mass of the system. We assume two degrees of freedom (the case with three degrees of freedom is only slightly more complex; see [15]), i.e., that lies in the plane so, and

    It is useful to rewrite the Hamiltonian as

    where

    is the effective potential. The critical points of are defined by and where are critical points of By putting (and , the quadratic part of the expansion of the Hamiltonian at such a critical point becomes

    One finds (see [20]) that the characteristic polynomial of the corresponding linear system becomes

    Note that the characteristic polynomial of the Hessian matrix equals

    If then we have the following relations between zeros of the polynomials and

    (9.5)

    We have the following generic possibilities for the zeros of (in [15], these cases are expressed in terms of

    (i) i.e., and then,

    (ii) i.e., and then, and

    (iii) i.e., and then, and

    (iv) i.e., and then, and .

    We note also the following non-generic case of resonance (considered in [15]):

    (v) i.e., and then, and

    Using only the above information about the eigenvalues , we can state the following.

    (1) In cases (i) and (ii), there are no families of periodic solutions.

    (2) In case (iii), there is one 1–parameter family of periodic solutions for the period

    (3) In case (iv), there exists at least one 1–parameter family of periodic solutions for the period moreover, if is not an integer, then there is another 1–parameter family for the period

    Analogues of these statements can be found in [15, Theorem 3.1]. But, in the same theorem, we find the following statement:

    If in case (v), additionally, the index (the authors call it the Brouwer index) is nonzero, then there exists a branch of closed orbits.

    We shall justify this using a refined version of the Birkhoff normal form in the resonant case:

    where and In the case of one degree of freedom, this result is due to Baider and Sanders [5] (some coefficients in the series vanish); the above case is a generalization of [7, Eq. (3.4)]. The corresponding differential system becomes

    As usual, when looking for periodic phase curves, we make the substitution . We get the system

    with equilibrium points defined by and The sufficient condition for the existence of a nontrivial solution to the this system is that the exponent is odd. But, this amounts to the property that the index of at this singular point is nonzero; equivalently, the index of at is nonzero.

    In the case of the potential given by Eq. (9.4) for the planar restricted four-body problem, the authors of [15] referred to works of other authors when looking for libration points (see [2,6,19] for example). One assumes (normalization); then, the sides of the equilateral triangle equal

    One draws three circles of radius with centers at and and three straight lines which extend the sides of . Then, one obtains 16 bounded regions, but only seven of them contain libration points (critical points of These (open) domains are as follows: the triangle ; three circular sectors and with vertices at and , respectively; three circular triangles and opposite to the corresponding sides of (see [15, Theorem 4.1(i)]).

    Moreover, in the case of equal masses, there are 10 critical points of all nondegenerate: the minimum and three saddle points in , one local maximum in each and one saddle in each (see [15, Lemma 4.2]). In the case of non-equal masses, one calculates the indices of along the boundaries of the above domains (called the Brouwer degree) and finds that, in each of the above seven domains, there exists a libration point with a nonzero index of

    By the above analysis of the eigenvalues using the relations of Eq. (9.5) between and the eigenvalues of , one concludes the existence of a 1–parameter family of periodic solutions for each of the seven above libration points.

    It would be interesting to study periodic solutions in the resonance case, i.e., when and (

    Finally, we note that the restricted four-body problem was studied from the points of view of the Lyapunov stability of the libration points, i.e., for case (iv); the existence of a family of periodic solutions does not imply it. Here, one uses the KAM theorem and fourth-order Birkhoff normal form. In [20], the generic Lyapunov stability for case (iv) was established for the following situations:

    when and dominate over and the libration points lie near

    when and the libration points lie in a symmetry line of ;

    when dominates over and and the libration points lie near the circle with its center at and radius

    Let be relatively prime positive integers, called the weights. Consider the following action of on

    By definition, the quotient of by is the weighted projective space,

    (10.1)

    Its complex dimension is and real dimension is

    From the algebraic geometry point of view (see [11]), it is the projective spectrum S of the graded polynomial algebra S such that These spaces are important from the complex geometry point of view. In particular, the morphism defines the formal isomorphism

    But, we are more interested in their topological properties. Consider the standard sphere Each orbit intersects this sphere along a circle. Therefore, we have

    where the action of the circle is like in Eq. (6.2), i.e.,

    Let us find the representation of our weighted projective space as a kind of CW complex (see [14]). This complex consists of 'cells' of (real) dimensions

    The 'cell' of maximal dimension is defined by in We rotate the variable to get i.e., We have the following description of this 'cell':

    where and is the stabilizer of the set The above upper semi-sphere is diffeomorphic with the standard dimensional unit ball, and its quotient is singular but with the singular locus of (real) codimension (complex codimension is This cell defines the fundamental cycle of

    The boundary of is the weighted projective space Note that now the weights can be not relatively prime; but, if , then the spaces and are isomorphic (see [11]).

    In this way, we successively construct the lower-dimensional cells, as well as cells in the singular locus of the action of the group on By the dimensional argument, the boundaries of these 'cells' are zero in the homological sense; so, the only nonzero homology groups have even dimensions. We distinguish the dimensional cycles corresponding to the hyperplane sections respectively. The intersection of any collection of of the above cycles is nonempty and, hence, (see Theorem 17 below).

    Recall [14] that the homological length of a manifold is the maximal number such that there exist homology classes such that their intersection is a nonzero cycle.‡‡

    ‡‡In fact, in [14], one finds the definition of the cohomological length (or the cup length) as the maximal number such that there exist cohomology classes such that These cohomology classes are Poincaré-dual to the cycles

    But, in the proof of Theorem 17 below given in [14], only intersections of the cycles are used.

    Finally, we note that, in our analysis, we can replace the standard sphere with any manifold of the form

    for positive constants

    By definition (see [14]), the Lusternik–Schnirelmann category of with respect to is the smallest number for which there exist closed subsets of such that

    (i) and

    (ii) each is contractible in If is connected, then we define

    In this work, we have used the following properties of this notion.

    Theorem 16. (Lusternik–Schnirelmann) Let be a compact closed (without boundary) connected manifold, and let be a smooth function on with isolated critical points (or bifurcational points). Then,

    Theorem 17. If is a manifold, then is bounded from below by 1 plus the homological length of

    For the proofs, we refer the reader to a previous monograph [14].

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Supported by Polish National Science Centre OPUS Grant No. 2017/25/B/ST1/00931.

    The authors declare that there is no conflict of interest.



    [1] A. A Andronov, E. A. Leontovich, A. G. Gordon, A. G. Maier, Qualitative Theory of Second–Order Dynamical Systems, Halsted Press, New York, 1973. https://doi.org/10.1137/1017026
    [2] R. F. Arenstorf, Central configurations of four bodies with one inferior mass, Celestial Mech., 28 (1982), 9–15. https://doi.org/10.1007/BF01230655 doi: 10.1007/BF01230655
    [3] V. I. Arnold, Geometrical Methods in the Theory of Ordinary Differential Equations, Springer, New York, 1983. https://doi.org/10.1007/978-1-4612-1037-5
    [4] V. I. Arnold, V. V. Kozlov, A. I. Neishtadt, Mathematical Aspects of the Mathematical and Celestial Mechanics, Encyclopaedia of Math. Sci., Dynamical Systems, 3, Springer, New York, 1988. https://doi.org/10.2307/3619341
    [5] A. Baider, J. A. Sanders, Unique normal forms: The nilpotent Hamiltonian case, J. Differ. Equations, 92 (1991), 282–304. https://doi.org/10.1016/0022-0396(91)90050-J doi: 10.1016/0022-0396(91)90050-J
    [6] J. F. Barros, E. Leandro, Bifurcations and enumeration of classes of relative equilibria in the planar restricted four-body problem, SIAM J. Math. Anal., 46 (2014), 1185–1203. https://doi.org/10.1137/130911342 doi: 10.1137/130911342
    [7] W. Barwicz, M. Wiliński, H. Żołądek, Birkhoff normalization, bifurcations of Hamiltonian vector fields and the Deprits formula, J. Fixed Point Theory Appl. 13 (2013), 587–610. https://doi.org/10.1007/s11784-013-0136-1
    [8] G. D. Birkhoff, Dynamical Systems, Amer. Math. Soc., Providence, 1927. https://doi.org/10.1016/B978-044450871-3/50149-2
    [9] C. A. Briot, J. C. Bouquet, Recherches sur les propriétés des fonctions définies par des équations différentiells, J. Éc. Polytech. 36 (1856), 133–198.
    [10] E. N. Dancer, S. Rybicki, A note on periodic solutions of autonomous Hamiltonian systems emanating from degenerate stationary solutions, Differ. Integr. Equations, 12 (1999), 147–160. https://doi.org/10.57262/die/1367265626 doi: 10.57262/die/1367265626
    [11] I. Dolgachev, Weighted projective varieties, Lect. Notes in Math., 956 (1982), Springer, Berlin, 34–71. https://doi.org/10.1007/BFB0101508 doi: 10.1007/BFB0101508
    [12] J. J. Duistermaat, The monodromy in the Hamiltonian Hopf bifurcation, Z. angew. Math. Phys. 49 (1998), 156–161. https://doi.org/10.1007/s000330050086
    [13] H. Dulac, Sur les cycles limites, Bull. Soc. Math. France, 51, 1923. https://doi.org/10.24033/BSMF.1031
    [14] A. Fomenko, D. Fuchs, Homotopical Topology, Graduate Texts in Math., 273, Springer, New York, 2016. https://doi.org/10.1007/978-3-319-23488-5
    [15] A. Gołȩbiewska, E. Pérez-Chavela, S. Rybicki, A. Ureña, Bifurcation of closed orbits from equilibria of Newtonian systems with Coriolis forces, J. Differ. Equations, 338 (2022), 441–473. https://doi.org/10.1016/j.jde.2022.08.004 doi: 10.1016/j.jde.2022.08.004
    [16] M. Hirsch, C. Pugh, M. Shub, Invariant manifolds, Lect. Notes Math. 583, Springer, New York, 1977. https://doi.org/10.1007/BFb0092042
    [17] Yu. Ilyashenko, S. Yakovenko, Lectures on Analytic Differential Equations, Graduate Studies in Math. 86, Amer. Math. Soc., Providence, 2008. https://doi.org/10.1090/gsm/086
    [18] L.G. Khazin, E.E. Shnol, Stability of Critical Equilibrium States, Nonlinear Science: Theory and Applications, Manchester Un-ty Press, Manchester, 1991.
    [19] E. Leandro, On the central configurations of the planar restricted four-body problem, J. Differ. Equations, 226 (2006), 323–351. https://doi.org/10.1006/j.jde.2005.10.015 doi: 10.1006/j.jde.2005.10.015
    [20] A. Ligȩza, H. Żołądek, Qualitative analysis of some libration points in the restricted four-body problem, Rus. J. Nonlin. Dyn., 17 (2021), 369–390. https://doi.org/10.20537/nd210402 doi: 10.20537/nd210402
    [21] A. M. Lyapunov, The General Problem of the Stability of Motion, Taylor & Francis Group, London, 1992. https://doi.org/10.1115/1.2901415
    [22] J. Mawhin, J. Willem, Critical Point Theory and Hamiltonian Systems, Springer, New York, 1989. https://doi.org/10.1007/971-1-4757-2061-7
    [23] J. Moser, Periodic orbits near an equilibrium and a theorem by Alan Weinstein, Comm. Pure Appl. Math. 29 (1976), 724–747. https://doi.org/10.1016/s0304-0208(08)71098-3
    [24] R. Moussu, Une démonstration géometrique d'un théorème de Lyapunov–Poincaré, in: Bifurcation, Ergodic Theory and Applications (Dijon, 1981), Asterisque (1982), 98–99.
    [25] E. Pérez-Chavela, S. Rybicki, D. Strzelecki, Symmetric Lyapunov center theorem, Calculus Variations PDEs, 56 (2017), art. 26. https://doi.org/10.1007/s00526-017-1120-1 doi: 10.1007/s00526-017-1120-1
    [26] E. Pérez-Chavela, S. Rybicki, D. Strzelecki, Symmetric Lyapunov center theorem for minimal orbit, J. Differential Equations, 265 (2018), 752–778. https://doi.org/10.1016/j.jde.2018.03.009 doi: 10.1016/j.jde.2018.03.009
    [27] H. Poincaré, Mémoire sur les Courbes Définies par une Équation Différentielle, in: Œuvres de Henri Poincaré, 1, Gauthier–Villars, Paris, 1951.
    [28] D. S. Schmidt, Periodic solutions near a resonant equilibrium of a Hamiltonian system, Celestial Mech. 9 (1974) 81–103. https://doi.org/10.1007/BF01236166
    [29] C. L. Siegel, Vorlesungen über Himmelsmechanik, Springer, Berlin, 1956. https://doi.org/10.1007/978-3-642-94671-4
    [30] E. Stróżyna, H. Żołądek, Analytic properties of the complete normal form for the Bogdanov–Takens singularity, Nonlinearity, 34 (2021), 3046–3082. https://doi.org/10.1088/1361-6544/abe51d doi: 10.1088/1361-6544/abe51d
    [31] D. Strzelecki, Periodic solutions of symmetric Hamiltonian systems, Arch. Rational Mech. Anal. 237 (2020), 921–950. https://doi.org/10.1007/s00205-020-01522-6
    [32] A. Szulkin, Bifurcation of strongly indefinite functionals and a Liapunov type theorem for Hamiltonian systems, Differential Integral Equations, 7 (1994), 217–234. https://doi.org/10.57262/die/1369926976 doi: 10.57262/die/1369926976
    [33] J. C. van der Meer, Bifurcation at non-semisimple 1: -1 resonance, J. Appl. Math. Phys. 37 (1986), 425–437. https://doi.org/10.1007/BF00946761
    [34] A. van Straten, A note on the number of periodic orbits near a resonant equilibrium point, Nonlinearity, 2 (1989), 445–458. https://doi.org/10.1007/BF02570469 doi: 10.1007/BF02570469
    [35] A. Weinstein, Normal modes for nonlinear Hamiltonian systems, Invent. Math. 20 (1973), 47–57. https://doi.org/10.1007/BF01405263
    [36] A. Weinstein, Symplectic V–manifolds, periodic orbits of Hamiltonian systems and the volume of some Riemannian manifolds, Comm. Pure Appl. Math. 30 (1977), 265–271. https://doi.org/10.1002/cpa.3160300207
    [37] H. Żołądek, The Monodromy Group, Monografie Matematyczne, 67 Birkhäuser, Basel, 2006. https://doi.org/10.1007/3-7643-7536-1
  • This article has been cited by:

    1. Henryk Żoła̧dek, An example in Hamiltonian dynamics, 2024, 16, 2836-3310, 431, 10.3934/cam.2024020
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1331) PDF downloads(84) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog