Research article Special Issues

Kolmogorov algorithm for isochronous Hamiltonian systems

  • We present a Kolmogorov-like algorithm for the computation of a normal form in the neighborhood of an invariant torus in 'isochronous' Hamiltonian systems, i.e., systems with Hamiltonians of the form H=H0+εH1 where H0 is the Hamiltonian of N linear oscillators, and H1 is expandable as a polynomial series in the oscillators' canonical variables. This method can be regarded as a normal form analogue of a corresponding Lindstedt method for coupled oscillators. We comment on the possible use of the Lindstedt method itself under two distinct schemes, i.e., one producing series analogous to those of the Birkhoff normal form scheme, and another, analogous to the Kolomogorov normal form scheme in which we fix in advance the frequency of the torus.

    Citation: Rita Mastroianni, Christos Efthymiopoulos. Kolmogorov algorithm for isochronous Hamiltonian systems[J]. Mathematics in Engineering, 2023, 5(2): 1-35. doi: 10.3934/mine.2023035

    Related Papers:

    [1] Tiziano Penati, Veronica Danesi, Simone Paleari . Low dimensional completely resonant tori in Hamiltonian Lattices and a Theorem of Poincaré. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021029
    [2] Marco Sansottera, Veronica Danesi . Kolmogorov variation: KAM with knobs (à la Kolmogorov). Mathematics in Engineering, 2023, 5(5): 1-19. doi: 10.3934/mine.2023089
    [3] Chiara Caracciolo . Normal form for lower dimensional elliptic tori in Hamiltonian systems. Mathematics in Engineering, 2022, 4(6): 1-40. doi: 10.3934/mine.2022051
    [4] Hitoshi Ishii . The vanishing discount problem for monotone systems of Hamilton-Jacobi equations: a counterexample to the full convergence. Mathematics in Engineering, 2023, 5(4): 1-10. doi: 10.3934/mine.2023072
    [5] G. Gaeta, G. Pucacco . Near-resonances and detuning in classical and quantum mechanics. Mathematics in Engineering, 2023, 5(1): 1-44. doi: 10.3934/mine.2023005
    [6] Simone Paleari, Tiziano Penati . Hamiltonian lattice dynamics. Mathematics in Engineering, 2019, 1(4): 881-887. doi: 10.3934/mine.2019.4.881
    [7] Federico Bernini, Simone Secchi . Existence of solutions for a perturbed problem with logarithmic potential in $\mathbb{R}^2$. Mathematics in Engineering, 2020, 2(3): 438-458. doi: 10.3934/mine.2020020
    [8] Dario Bambusi, Beatrice Langella . A $C^\infty$ Nekhoroshev theorem. Mathematics in Engineering, 2021, 3(2): 1-17. doi: 10.3934/mine.2021019
    [9] Jorge E. Macías-Díaz, Anastasios Bountis, Helen Christodoulidi . Energy transmission in Hamiltonian systems of globally interacting particles with Klein-Gordon on-site potentials. Mathematics in Engineering, 2019, 1(2): 343-358. doi: 10.3934/mine.2019.2.343
    [10] Panayotis G. Kevrekidis . Instabilities via negative Krein signature in a weakly non-Hamiltonian DNLS model. Mathematics in Engineering, 2019, 1(2): 378-390. doi: 10.3934/mine.2019.2.378
  • We present a Kolmogorov-like algorithm for the computation of a normal form in the neighborhood of an invariant torus in 'isochronous' Hamiltonian systems, i.e., systems with Hamiltonians of the form H=H0+εH1 where H0 is the Hamiltonian of N linear oscillators, and H1 is expandable as a polynomial series in the oscillators' canonical variables. This method can be regarded as a normal form analogue of a corresponding Lindstedt method for coupled oscillators. We comment on the possible use of the Lindstedt method itself under two distinct schemes, i.e., one producing series analogous to those of the Birkhoff normal form scheme, and another, analogous to the Kolomogorov normal form scheme in which we fix in advance the frequency of the torus.



    The motivation for this paper stemmed from an exchange of e-mails and discussions that the authors of refs. [2] and [3] (C. E. and Eleni Christodoulidi) had with Antonio Giorgilli, Sergej Flach and Giancarlo Benettin several years ago. The discussion was on the meaning of Lindstedt series as a method to construct invariant tori in perturbed oscillator systems like the celebrated problem of Fermi-Pasta-Ulam (see Galgani [10] in the present volume). We, as many others, are fond of recognizing the deep influence of Antonio's thought on our way of viewing the role of perturbative methods in the understanding of Hamiltonian dynamical systems. Since this is a paper to appear in the special issue in his honor, our presentation below is, in turn, strongly shaped by what we hope could be a framework of discussion interesting to our inspiring and beloved teacher A. Giorgilli.

    In various contexts in the literature, the use of the term 'Lindstedt series' for isochronous Hamiltonian systems often refers to one of two distinct methods, both applicable to the perturbative study of the dynamics around systems with elliptic equilibria. The difference between these two methods can be conveniently explained with the help of the following example: consider a 'Henon-Heiles' type Hamiltonian

    H=12(p2x+p2y)+12ω20,1x2+12ω20,2y2+εP3(x,y) (1.1)

    where, contrary to the actual Hénon-Heiles model ([9]) (where ω0,1=ω0,2=1), we first assume that the frequencies (ω0,1,ω0,2) satisfy no resonance condition. P3 can be any polynomial cubic in x,y.

    We are interested in constructing perturbative series solutions in the model (1.1) under the form:

    x(t)=x0(t)+εx1(t)+ε2x2(t)+,y(t)=y0(t)+εy1(t)+ε2y2(t)+, (1.2)

    where we adopt some periodic form for the functions x0(t),y0(t) and compute iteratively all subsequent functions xi(t),yi(t), i=1,2,. We further want to secure that the iterative scheme used to compute the functions xi(t),yi(t) preserves the quasi-periodic character of the solutions, i.e., produces no secular terms (of the form tsin(ωt), etc) for some frequencies ω obtained as discussed below. An elementary remark in this context is that the nonlinear coupling of the oscillators implies that quasi-periodic orbits in the above model are expected to evolve, in general, with frequencies ω1,ω2 different from those of the unperturbed oscillators, ω0,1,ω0,2. As it is well known (see, for example, [13]), recognition of this fact implies to introduce formal series also for the frequencies:

    ω1(A1,A2)=ω0,1+εω1,1(A1,A2)+ε2ω2,1(A1,A2)+,ω2(A1,A2)=ω0,2+εω1,2(A2,A2)+ε2ω2,2(A1,A2)+. (1.3)

    The quantities ωi,1(A1,A2),ωi,2(A1,A2) are functions depending on two parameters A1,A2, called hereafter the 'amplitudes' of the oscillations in x and y respectively. They enter into the calculation through the choice made for the zero-th order terms x0(t), y0(t), since the iterative procedure starts by setting

    x0(t)=A1cos(ω1t+ϕx0),y0(t)=A2cos(ω2t+ϕy0) (1.4)

    where the initial phases ϕx0,ϕy0 can be arbitrary.

    The above are common elements of the point of departure for both versions discussed below of the Lindstedt method. However, at this stage emerges an important bifurcation in the way we define the iterative scheme by which the functions xi(t),yi(t),ωi,1,ωi,2 are to be computed. We discuss two distinct possibilities, referred to below as (ⅰ) a Lindstedt scheme 'analogous to the Birkhoff series', or (ⅱ) a Lindstedt scheme 'analogous to the Kolmogorov series'.

    As typical in perturbation theory, the formal difference between the above two schemes actually reflects a real (physical) difference in the way we interpret the meaning of the series (1.3). In summary, the difference can be posed as follows (see section 2 for details):

    (ⅰ) in the scheme called below 'analogous to Birkhoff', we seek to construct a quasi-periodic solution valid for any value of the amplitudes A1,A2 within a suitably defined open domain around the origin. Thus, the series (1.3) in this scheme are meant to answer the question of what are the values of the frequencies ω1,ω2 under which the motion takes place for any given and pre-selected sets of values of the amplitudes A1,A1 in the above domain. The reader is referred to [15] where a clear exposition of the method is given in the framework of special solutions of the three-body problem computed via Lindstedt series.

    (ⅱ) in the scheme called below 'analogous to Kolmogorov', instead, we fix in advance the values of the frequencies ω1,ω2 (see [13] for a clear exposition of the method in the context of the forced anharmonic oscillator); this is called by some authors a 'torus fixing method'. A relevant remark in the context of this last method is that the series (1.3) are actually purported to answer the question reverse to the one posed in (ⅰ) above. That is, the question now is: with given and pre-selected values of the frequencies ω1,ω2, invert the series (1.3) and compute which are the corresponding amplitudes A1,A2 for which we obtain quasi-periodic trajectories with the frequencies ω1,ω2. Thus, in method (ⅰ) the series are parameterized by the amplitudes A1,A2, which can be selected at the beginning of the construction, while in method (ⅱ) the solutions are parameterized by the frequencies ω1,ω2, which are the parameters to select at the beginning of the construction. Also, in the latter case the series inverse to (1.3) turn out to have the form (in the cubic case)

    ε2A21=i=1Ci,1(ω1ω0,1)i,ε2A22=i=1Ci,2(ω2ω0,2)i, (1.5)

    for some constant coefficients Ci,1,Ci,2 computable from the series (1.3). Thus, with all frequencies of the problem fixed in advance, establishing the convergence of the inverse series (1.5) suffices to answer the question posed at (ⅱ).

    The question of the convergence of the series is, of course, crucial, and related to the kind, and pattern of accumulation in the series terms, of small divisors appearing at successive perturbative steps. As regards the kind of divisors, we can readily see that:

    - in scheme (ⅰ) we obtain divisors of the form k1ω0,1+k2ω0,2, with (k1,k2)Z2, |k1|+|k2|0. This follows from the kind of linear (non-homogeneous) equation to solve iteratively. Deferring details to the example treated in Section 2, we briefly recall that in scheme (ⅰ) we introduce the parametrization (modulo two unimportant phases) φ1=ω1t, φ2=ω2t, and after introducing the series expressions (1.2) and (1.3) to the equations of motion and separate terms of like orders we arrive at equations (to be solved iteratively) of the form:

    (ω0,1φ1+ω0,2φ2)2xi+xi=Θ1,i(φ1,φ2)(ω0,1φ1+ω0,2φ2)2yi+yi=Θ2,i(φ1,φ2)i=1,2, (1.6)

    where the functions Θ1,i(φ1,φ2), Θ2,i(φ1,φ2) contain trigonometric terms in the angles φ1,φ2 (see [15], section 4).

    - In scheme (ⅱ), instead, we obtain divisors of the form k1ω1+k2ω2, i.e., depending on the (fixed) pre-selected new frequencies ω1,ω2. This follows from the fact that the linear non-homogeneous equations to solve are now of the form (see [13]):

    ¨xi+ω21xi=Φ1,i(φ1,φ2)¨yi+ω22yi=Φ2,i(φ1,φ2)i=1,2, (1.7)

    again with functions Φ1,i(φ1,φ2), Φ2,i(φ1,φ2) containing trigonometric terms in the angles φ1=ω1t,φ2=ω2t. Note that since the divisors depend on the new frequencies ω1,ω2, choosing non-resonant values for the latter permits the formal construction to proceed; this, even when the unperturbed frequencies ω0,1,ω0,2 are, instead, resonant.

    As regards convergence, in the case (ⅰ) Poincaré ([19], Ch.Ⅸ) already emphasizes that the Lindstedt series with divisors depending on the original harmonic frequencies ω0,1,ω0,2 are divergent, exhibiting the well known asymptotic character associated with the series computed via a Birkhoff normal form (see [5] for a review). Indeed, as shown by example in section 2 below, it possible to construct Birkhoff series yielding the same individual solutions as those of the Lindstedt series of scheme (ⅰ). We note here that the series originally introduced by Lindstedt ([16,17,18]), albeit somewhat different in structure, exhibit the same divisors as those of the scheme (ⅰ) above, thus, according to Poincaré, they are only asymptotic. On the other hand, Eliasson ([6]) and Gallavotti ([11,12]) established the existence of convergent Lindstedt series by the 'torus fixing method' on the basis of the cancellations between terms with small divisors (see [14] for an instructive example). A proof of the convergence of scheme (ⅱ) is actually possible by diagrammatic methods via the following theorem [4]:

    Theorem 1.1 ([4]). Consider the N coupled oscillator equations

    ¨xj+ω2jxj+fj(x1,,xN;ε)+(ω20,jω2j)xj=0,j=1,,N (1.8)

    where ε is a real parameter, f(x,ε) is real analytic at x=0,ε=0 and at least quadratic in x, and the frequency vector ω is diophantine. Let

    x(0)j(t)=cjeiωjt+cjeiωjt,j=1,,N (1.9)

    be a solution of (1.8) for any choice of the complex constants cj and for ε=0. Let Γ(c)=max(|c1|,,|cN|,1). Then, there exists a positive constant η0 and a function η(ε,c) holomorphic in the domain |ε|Γ3(c)η0, real for real ε, such that the system

    ¨xj+ω2jxj+fj(x1,,xN,ε)+ηj(ε,c)xj=0,j=1,,N (1.10)

    admits a solution of the form

    x(t,ε,c)=νZNAνexp(iν(ωt)) (1.11)

    holomorphic in the domain |ε|Γ3(c)e3|ω||t|η0 and real for real ε,t. The constants Aν are O(ε), except for the constants A1,0,...,0, A0,1,...,0, A0,0,...,1, which are equal to c1, c2,,cN respectively.

    A similar proof in action-angle variables in the case N=2 is discussed in [1]. As a final introductory remark, the series construction in the isochronous case finds a plethora of applications in various fields of physics. We mention in particular, the use of the Lindstedt method for the computation of solutions lying on low-dimensional tori ('q-tori') in the celebrated Fermi-Pasta-Ulam (FPU) problem ([2,3]). The FPU model takes, in normal mode space, the form of N harmonic oscillators coupled with nonlinear terms:

    ¨Qk+Ω2kQ=εFk(Q1,,QN) (1.12)

    where the frequencies Ωk, k=1,,N are given in terms of the FPU normal mode spectrum Ωk=2sin(kπ/(2(N+1)), the function F can be cubic or quartic in the variables Qk, and the perturbation ε satisfies some scaling law with N.

    Flach and co-workers ([7,8]) emphasized the special role for dynamics played by solutions called 'q-breathers'. These are periodic orbits of the form Qq(t)=Aqcos(ωqt+ϕq), and Qk(t)=0 for kq. For the frequencies ωq we obtain series expressions of the form

    ωq=Ωq+Δωq(Aq;ε),Δωq=O(ε). (1.13)

    Then, for ε sufficiently small, the Lindstedt method (ⅱ) above allows to represent the q-breathers via the Fourier expansion

    Qq(t)=Aqcos(ωqt+ϕq)+m=0˜fq,m(Aq;ε)cos[m(ωqt+ϕq)]Qk(t)=m=0˜fk,m(Aq;ε)cos[m(ωqt+ϕq)],kq, (1.14)

    where ˜fk,m=O(εp(k,q,m)), with integer exponent p(k,q,m)1. The relevant point for the FPU problem is that the rules of propagation of the amplitude Aq in the series terms for all modes allows to find an analytic formula explaining the phenomenon of 'energy localization' observed for particular initial excitations in the FPU model. In [2] and [3], on the other hand, it was shown that the q-breathers constitute only the first member in the hierarchy of special FPU solutions that exhibit energy localization. More general members are the 'q-tori', i.e., special solutions with M<N incommensurable frequencies satisfying

    ωqi=Ωqi+Δωqi(Aq1,Aq2,,AqM;ε),Δωqi=O(ε), (1.15)

    where Rq=(q1,qM){1,2,...,N}M. The corresponding Fourier representation of these special solutions can again been computed using Lindstedt series, and it obtains the form

    Qk(t)=Akcos(ωkt+ϕk)+|m|=0˜fk,m(A;ε)cos[m(ωt+ϕ)],kRqQk(t)=0+|m|=0˜fk,m(A;ε)cos[m(ωt+ϕ)],kRq (1.16)

    with m(m1,,mM)ZM, A(Aq1,,AqM), ω(ωq1,,ωqM), ϕ(ϕq1,,ϕqM), and ˜fk,m=O(ε) for all k=1,N. Furthermore, the propagation of the amplitudes Ak in the series terms allows to interpret a variety of complex localization profiles encountered for particular initial mode excitations in the FPU problem (see the corresponding theorems in [3]).

    We mentioned already that for the Lindstedt series analogous to the Birkhoff ones there exists a Birkhoff normal form yielding the same solutions as those recovered by the Lindstedt method via an indirect approach, i.e., one based on a sequence of normalizing transformations involving canonical changes of variables. It is natural to ask whether this correspondence between a direct (Lindstedt) and indirect (normal form) method extends in the case of the torus-fixing method as well. Due to the lack of a twist condition, the torus-fixing process in the isochronous case has to be dealt with using a technique based on 'counterterms' (see [11]), or a KAM algorithm 'with knobs' (see [20] in the present volume). We have developed a Kolmogorov algorithm using counterterms, which is able to recover the solutions of the direct Lindstedt method in both cases of full or low-dimensional tori. In the present paper we will discuss in some detail the proposed Kolmogorov algorithm for the full-dimensional case only, deferring the low-dimensional case to a different publication. Section 2 gives an elementary example allowing to fix all principle ideas in this comparison. Section 3 gives the general algorithm for the Kolmogorov method in the isochronous case. This is similar as the algorithm 'with knobs' discussed by Sansottera and Danesi ([20]) in the present volume. We point out, however, some differences in our viewpoints regarding what the algorithm actually achieves to compute and how it should be implemented in the context of isochronous models with given initial unperturbed frequencies ω0. Some results and detailed proofs are deferred to the Appendix.

    In order to illustrate the methods discussed above, we will consider an elementary example stemming from the following one-degree of freedom Hamiltonian with a even power dependence on the canonical variables

    H(x,p)=H0+εH1=ω02(p2+x2)+εx44. (2.1)

    Using the harmonic oscillator action-angle variables (J,q) with x=2Jsin(q), p=2Jcos(q), we obtain

    H(q,J)=ω0J+3ε8J2ε2J2cos(2q)+ε8J2cos(4q). (2.2)

    Let J0 be the label of a given torus (periodic orbit) of the harmonic oscillator model H0. Consider a real neighborhood Dε={J=J0+pwith|p|<Dε}, where Dε=O(ε). We will illustrate four different perturbative methods to treat the dynamics in the phase-space neighborhood T×Dε: these are ⅰ) a Birkhoff normal form construction, with ⅱ) its analog in terms of Lindstedt series, ⅲ) a Lindstedt series exhibiting the torus-fixing property of the Kolmogorov method, and, finally ⅳ) the normal form analogue of ⅲ). In the next section, we will give the general formulas for the construction of the Kolmogorov normal form in the case of n-degree of freedom Hamiltonians with isochronous integrable part H0.

    Setting J=J0+p the Hamiltonian takes the form (apart from a constant)

    H(q,p)=ω0p+3εJ208+3J0ε4p+3ε8p2εJ202cos(2q)εJ0pcos(2q)=ε2p2cos(2q)+εJ208cos(4q)+εJ04pcos(4q)+ε8p2cos(4q). (2.3)

    A 'Birkhoff normal form' for the Hamiltonian (2.3) can be computed by introducing a canonical transformation eliminating the angle q. Writing the Hamiltonian (2.3) as

    H(q,p)=Z0+εf1,

    where Z0=ω0p and f1=k,lck,lpkeilq we will define a Lie generating function χ(1)(q,p) bringing this Hamiltonian to normal form up to terms O(ε). In the standard procedure, it is sufficient to set χ(1)=X(1), where X(1) solves the homological equation LX(1)Z0+εf1=εζ1, with ζ1=<f1>q and LX denoting the Poisson bracket operator LX={,X}. However, comparing the result with the one obtained by the corresponding Lindstedt method (see next subsection) requires a small modification in the definition of χ(1). Consider the canonical transformation (q,p)(q(r),p(r)) obtained after r normalization steps: we require that the canonical transformation be such that the initial condition q(r)=p(r)=0 in the new variables be mapped to the initial condition (q(q(r)=0,p(r)=0),p(q(r)=0,p(r)=0))=(0,0)+O(εr+1) in the original variables. It is easy to see that such a requirement of control on the initial condition can be fulfilled by setting

    χ(n)(q,p)=X(n)(q,p)+K(n)q+S(n)p,n=1,,r

    where K(n), S(n) are constants possible to compute at every step by requiring that

    (q(q(n)=0,p(n)=0),p(q(n)=0,p(n)=0))=(0,0)+O(εn+1).

    Since {ω0p,K(n)q}=ω0K(n), this procedure will only alter the normal form at the n-th step by a constant, adding, however, some trigonometric terms to the remainder at every step. As an example, we can readily verify the following formulas for the first step (and analogously for subsequent steps)

    χ(1)(q,p)=X(1)(q,p)+K(1)q+S(1)p,

    where

    X(1)(q,p)=l0kck,lilω0pkeilq,K(1)=l0c0,lω0,S(1)=l0c1,lilω0.

    Omitting details, the formulas obtained after two normalization steps as above are the following: the Hamiltonian takes the form

    H(2)(q(2),p(2))=expLε2χ(2)expLεχ(1)H|q=q(2)p=p(2)=(Z0+εZ1+ε2Z2+i3εifi)|q=q(2)p=p(2).

    For simplicity in the notation, from now on we omit superscripts from the variables q and p unless explicitly required, adopting, instead, the convention that the symbols (˜q,˜p) in any function of the form F(r)(˜q,˜p) imply the new canonical variables computed after r normalization steps. Then, up to order 2 in ε we obtain:

    H(2)(˜q,˜p)=Z(2)+R(2)=ω0˜p+3εJ04˜p+3ε˜p2869ε2J2064ω0˜p51ε2J064ω0˜p217ε264ω0˜p3+R(2)

    where the remainder R(2) is O(ε3). The Hamiltonian Z(2) can now be used to analytically compute ε2precise solutions to the equations of motion in the variables (˜q,˜p). The equations of motion are

    {˙˜q=H(2)˜p=ω0+3εJ04+3ε4˜p69ε2J2064ω051ε2J032ω0˜p51ε264ω0˜p2˙˜p=H(2)˜q=0;

    fixing the initial condition ˜p(0)=˜q(0)=0 yields the solution

    ˜p(t)=0,˜q(t)=ωt:=φB,ω=ω0+3εJ0469ε2J2064ω0. (2.4)

    This can be back-transformed to the solution in the original variables. We have

    q=expLε2χ(2)expLεχ(1)˜q,p=expLε2χ(2)expLεχ(1)˜p. (2.5)

    Substituting the solutions (˜q(t),˜p(t)) in the previous expression, we find

    q(t)=φBεJ02ω0sin(2φB)+31ε2J2032ω20sin(2φB)+εJ016ω0sin(4φB)=ε2J2032ω20sin(4φB)ε2J2032ω20sin(6φB)+ε2J20512ω20sin(8φB),p(t)=3εJ208ω0+13ε2J3016ω20+εJ202ω0cos(2φB)33ε2J3032ω20cos(2φB)=εJ208ω0cos(4φB)+3ε2J3016ω20cos(4φB)+ε2J3032ω20cos(6φB),J(t)=J0+p(t). (2.6)

    Observe that q(0)=0 and J(0)=J0, as was required.

    Two remarks are in order:

    (ⅰ) The divisors appearing in all series expressions obtained above depend on the (unique, in the case of 1DOF systems) unperturbed frequency of the model, i.e., the frequency ω0 of the linear oscillator. In the case of systems with N>1 degrees of freedom, the above series will produce divisors of the form mω0, where mZN, |m|0, and ω0 is the Nvector of the unperturbed frequencies of the problem. This implies that the method may formally proceed only when the vector ω0 is non-resonant. As regards the series convergence, this is guaranteed in an open domain in the 1DOF case. However, when N>1 the series are in general only asymptotic (see [5] for reviews).

    (ⅱ) The value of the initial datum for the action J0 can be chosen at the end of the process, i.e., the numerical value of J0 need not be fixed in advance. In fact, J0 (or, more generally, the vector J0) can be regarded as a parameter free to transfer in all iterative computations. It is, then, in terms of this parameter that we obtain an expression for the frequency(ies) ω on the torus with the initial conditions J(0)=J0, q(0)=0 as a function of J0. In our example this function is given up to O(ε2) in Eq (2.4).

    (ⅲ) Any choice is possible to impose on the initial phase q(0), altering slightly the definition of the constants S(n).

    Analytical solutions in the form of formal series obtained like (2.6) above will be hereafter called 'Birkhoff solutions'. This terminology emphasizes the fact that the solutions are obtained using a Birkhoff normal form procedure. It is notworthy that Poincaré's reference to the 'Lindstedt series' (Methodes nouvelles, Ch. Ⅸ) actually consists of the construction of solutions to coupled oscillator problems obtained 'indirectly', i.e., via transformations of the variables and the engineering of the corresponding Hamiltonian function. Hence, the more general term 'Poincaré-Lindstedt' series often encountered in literature. We now give, instead, the 'direct' (i.e., without transformations) series construction of the Birkhoff solution (2.6) by implementing, instead, the original method of Lindstedt in the framework of the canonical action-angle variables of the harmonic oscillator model.

    The Hamilton's equations of motion for the Hamiltonian (2.2) are

    {˙q(t)=HJ=ω0+3εJ4εJcos(2q)+εJ4cos(4q)(2.7a)˙J(t)=Hq=εJ2sin(2q)+εJ22sin(4q).(2.7b)

    We perform the following steps:

    Step 1: time re-parametrization. Set φ=ωt where ω is the (still unknown) frequency on the 1-torus (periodic orbit) corresponding to the solution with initial conditions J(0)=J0, q(0)=0.

    Step 2: frequency expansion. This is the key element of the Lindstedt method. We expand ω as

    ω=ω0εa1(J0)ε2a2(J0)+. (2.8)

    Note that the corrections ai i=1,2, are functions of the parameter J0, i.e., of the 'amplitude' (square) of the oscillations.

    Step 3: expansion of the solution. We write

    q(φ)=q0(φ)+εq1(φ)+ε2q2(φ)+,J(φ)=J0(φ)+εJ1(φ)+ε2J2(φ)+. (2.9)

    Note here a key element of the method, which is the fact that all functions qi,Ji, i=0,1, are considered functions of the phase φ=ωt rather than of the time t itself. This is the key point in the differentiation of the method presented here with respect to the 'torus-fixing' method presented in Subsection 2.3. The relevant fact is that the use of the angle φ (or, more generally, of a set of angles φTN, with φ=ωt in the N-DOF case), instead of time, allows to split the equations of motion in a sequence of linear non-homogeneous equations, whose iterative solution introduces divisors depending on the unperturbed frequencies ω0. This is made clear in the next step.

    Step 4: splitting of the equations of motion in powers of ε and iterative solution. In our example, from the definition of φ we have:

    ωdq(φ)dφ=dq(t)dt,ωdJ(φ)dφ=dp(t)dt.

    Substituting (2.8) and (2.9) in the equations of motion (2.7a) and (2.7b) leads to the following expressions (up to order 2 in ε)

    (ω0εa1ε2a2)(dq0(φ)dφ+εdq1(φ)dφ+ε2dq2(φ)dφ)=ω0+3εJ0(φ)4εJ0(φ)cos(2q0(φ))+εJ0(φ)4cos(4q0(φ))+3ε2J1(φ)4ε2J1(φ)cos(2q0(φ)) (2.10a)
    =+ε2J1(φ)4cos(4q0(φ))+2ε2J0(φ)q1(φ)sin(2q0(φ))ε2J0(φ)q1(φ)sin(4q0(φ)),(ω0εa1ε2a2)(dJ0(φ)dφ+εdJ1(φ)dφ+ε2dJ2(φ)dφ)=εJ0(φ)2sin(2q0(φ))+εJ0(φ)22sin(4q0(φ))2ε2J0(φ)J1(φ)sin(2q0(φ))=+ε2J0(φ)J1(φ)sin(4q0(φ))2ε2J0(φ)2q1(φ)cos(2q0(φ))+2ε2J0(φ)2q1(φ)cos(4q0(φ)). (2.10b)

    Now, in order to iteratively determine the functions qi(φ),Ji(φ), we compare the two sides of the equations of motion at equal orders. At order 0 we have

    ω0dq0(φ)dφ=ω0,ω0dJ0(φ)dφ=0q0(φ)=φ,J0(φ)=J0.

    Fixing the initial data as q0(0)=0 and J(0)=J0, we then arrive at q0(φ)=φ and J0(φ)=J0. At order 1, we now have

    ω0dq1(φ)dφa1=3J04J0cos(2φ)+J04cos(4φ)ω0dJ1(φ)dφ=J20sin(2φ)+J202sin(4φ). (2.11)

    As well known, a1 is determined by the requirement that no secular terms be present in the series. This yields a1=3J0/4. Then, the Cauchy problem with the previous differential equations and the initial constants q1(0)=J1(0)=0 yields the solution

    q1(φ)=J02ω0sin(2φ)+J016ω0sin(4φ),J1(φ)=3J208ω0+J202ω0cos(2φ)J208ω0cos(4φ).

    The process can be repeated at subsequent orders. We leave to the reader to verify the result at second order, leading eventually to

    q(t)=φεJ02ω0sin(2φ)+31ε2J2032ω20sin(2φ)+εJ016ω0sin(4φ)=ε2J2032ω20sin(4φ)ε2J2032ω20sin(6φ)+ε2J20512ω20sin(8φ),J(t)=J03εJ208ω0+13ε2J3016ω20+εJ202ω0cos(2φ)33ε2J3032ω20cos(2φ)=εJ208ω0cos(4φ)+3ε2J3016ω20cos(4φ)+ε2J3032ω20cos(6φ), (2.12)

    with φ=ωt and ω=ω0+3εJ0469ε2J2064ω0. Comparing the result with Eqs (2.4) and (2.6), we can see that the solutions q(t), J(t) of the two methods are equal. This is easy to justify by checking the structure of the l.h.s of Eq (2.10). After the expansion of the frequency, the differential equations to solve at subsequent steps all involve the operators ω0d/dφ. Hence, all divisors appearing in the series terms are in terms of the unperturbed frequency ω0.

    It was already pointed out that the Lindstedt series examined so far, as well as their 'indirect' (normal form) counterpart, produce, in general, series which are divergent and only have an asymptotic character.* On the other hand, Eliasson [6] and Gallavotti [12] established the existence of convergent Lindstedt series in nonlinear Hamiltonian systems satisfying the necessary conditions for the holding of the Kolmogorov-Arnold-Moser theorem. Gallavotti ([11]) presented a diagrammatic proof of the convergence of the Kolmogorov normal form also in the 'twistless' case, i.e., when the size of the Hessian matrix of the unperturbed Hamiltonian H0(J) is not limited from below (it is equal to zero in the 'isochronous' case). A constructive Kolmogorov algorithm able to deal also with the twistless case is presented, along with the demonstration of its convergence, in the present volume by Sansottera and Danesi [20]). The convergence of the Lindstedt series in this case is addressed, instead, in [4] (see remarks in the introduction).

    *The examples treated in this section are obvious exceptions, since, we deal, for simplicity, with 1DOF systems exhibiting no small divisors. Small divisors appear, instead, for N>1.

    As discussed in the introduction, when series constructions analogous to Kolmogorov are sought for in the isochronous case, an important point to address is the need for performing, at the final stage of the construction, a process involving series reversal. This reversal is necessary in order to explicitely compute the solutions q(t), J(t) whose initial conditions q0,J0 correspond to motion on a torus with given frequency vector ω. The key remark is that the value of J0, which parametrises the solutions, cannot be fixed in advance due to the lack of a twist condition allowing to compute the mapping ω(J0). We now discuss the application of the direct (Lindstedt) method analogous to Kolmogorov in the same example as in the previous two sections, aiming to illustrate the above points.

    Fixing the frequency of the torus in the isochronous case can be implemented as described in [13]: assume that we target a particular solution of the equations of motion (2.7a) and (2.7b) represented as a trigonometric series and evolving according to a given pre-selected frequency ω. Inverting the expansion (2.8) we obtain

    ω0=ω+εa1(J0)+ε2a2(J0)+ (2.13)

    Also, as before we expand the solution as

    q(t)=q0(t)+εq1(t)+ε2q2(t)+J(t)=J0(t)+εJ1(t)+ε2J2(t)+. (2.14)

    Note, however, that this time we perform no time-reparametrization, i.e., the solutions remain expressed as functions of the time t. Thus, replacing the above expressions in (2.7a) and (2.7b) the equations of motion lead now to the expressions (up to order 2 in ε)

    ˙q0(t)+ε˙q1(t)+ε2˙q2(t)=ω+εa1+3εJ0(t)4εJ0(t)cos(2q0(t))+εJ0(t)4cos(4q0(t))˙q0(t)+ε˙q1(t)+ε2˙q2(t)=+3ε2J1(t)4+ε2a2ε2J1(t)cos(2q0(t))+ε2J1(t)4cos(4q0(t))˙q0(t)+ε˙q1(t)+ε2˙q2(t)=+2ε2J0(t)q1(t)sin(2q0(t))ε2J0(t)q1(t)sin(4q0(t)), (2.15a)
    ˙J0(t)+ε˙J1(t)+ε2˙J2(t)=εJ0(t)2sin(2q0(t))+εJ0(t)22sin(4q0(t))2ε2J0(t)J1(t)sin(2q0(t))+ε2J0(t)J1(t)sin(4q0(t))2ε2J0(t)2q1(t)cos(2q0(t))+2ε2J0(t)2q1(t)cos(4q0(t)). (2.15b)

    Collecting now the terms of equal order we can compute iteratively all the functions qi(t),Ji(t), for i=0,1,2, setting, at order zero:

    q0(t)=ωt,J0(t)=J0

    which corresponds to the choice of the initial condition q0(0)=0 and J0(0)=J0. Then, at first order we have

    ˙q1(t)=a1+3J04J0cos(2ωt)+J04cos(4ωt),˙J1(t)=J20sin(2ωt)+J202sin(4ωt) (2.16)

    implying a1=3J0/4 and

    q1(t)=J02ωsin(2ωt)+J016ωsin(4ωt),J1(t)=3J208ω+J202ωcos(2ωt)J208ωcos(4ωt).

    Note that the integration constants in (2.16) were set as q1(0)=J1(0)=0, consistent with our choice of initial condition. Repeating the procedure at second order yields

    q(t)=ωtεJ02ωsin(2ωt)+εJ016ωsin(4ωt)+19ε2J2032ω2sin(2ωt)+ε2J2064ω2sin(4ωt)=ε2J2032ω2sin(6ωt)+ε2J20512ω2sin(8ωt),J(t)=J03εJ208ω+εJ202ωcos(2ωt)εJ208ωcos(4ωt)+17ε2J3032ω2=21ε2J3032ω2cos(2ωt)+3ε2J3032ω2cos(4ωt)+ε2J3032ω2cos(6ωt),ω=ω0+3εJ0469ε2J2064ω. (2.17)

    It is instructive to compare the solutions (2.17) above with those (Eq (2.12)) found in the case of the original Lindstedt method. We have the following remarks:

    (ⅰ) From second order and beyond the coefficients in front of the harmonics cos(mωt),sin(mωt) for the same m are not all equal in the two solutions.

    (ⅱ) Most importantly, the structure of the divisors in the two solutions is different. In the torus fixing case (Eq (2.17)), all divisors involve the corrected frequency, ω, instead of the original frequency ω0 of the linear oscillator. This is not a problem for the method to proceed, since this frequency is known in advance. In NDOF systems, in general, with the 'torus fixing method' we obtain divisors of the form mω. Thus, the method can proceed even when the original frequencies of the N linear oscillators ω0 are resonant, as long as we impose a non-resonant detuning in the adopted values for the final frequencies ω on the torus.

    (ⅲ) On the other hand, the value of the initial condition J0 leading to motion on the torus with frequency ω remains uknown up to the end of the construction. From the point of view of the symbolic implementation of the method in the computer, J0 is a symbol whole powers have to be carried on along with the remaining powers of trigonometric monomials in all series terms and at all iterative steps.

    (ⅳ) At the end of the process, however, J0 can be estimated by reversing the series (2.13), given that the functions an(J0) are monomials of degree n in J0. In an analogous way, in the NDOF case we will end up with N series equations of the form

    ω0,jωj=n=1εnan,j(J0,1,,J0,N) (2.18)

    where the functions an,j(J0,1,,J0,N) have all been specified iteratively up to a maximum truncation order in n, and they are polynomial in J0(J0,1,,J0,N). Thus, the series (2.18) can be formally inverted, yielding

    J0,j=1εn=1Pn,j(ωω0) (2.19)

    where the functions Pn,j(ωω0) are polynomial of degree n in the differences (ωω0). Note that for the inverse series to converge we must require the difference |(ωω0)| to be smaller than ε, a fact which limits how far we can detune ω from ω0 in order to be able to specify the corresponding initial condition J0. An obvious choice is |(ωω0)|=O(ε2). At any rate, we emphasize that the convergence of the inverse series (2.18) is an open problem, crucial to the applications.

    In the lack of a known answer to this problem, we can adopt two different attitudes: we may fix J0 and claim that our series answer the question of what was the value of ω0 in the original Hamiltonian corresponding to the motion on the particular torus with frequency ω (see [20]). On the other hand, in real world applications the only certainty we have is about our initial model (!), i.e., ω0, so, we must devise a method to propose values of ω for which the method will probably turn to converge. In the case of the FPU problem, a numerical procedure to move within the space of the parameters ω so at to choose plausible values was proposed in [3].

    We finally arrive at the here proposed Kolmogorov normal form algorithm yielding solutions equivalent to those discussed in the previous subsection. This is implemented by the following steps:

    Step 1: substitution of the frequency series into the Hamiltonian. In our example, we substitute ω0 in the Hamiltonian (2.3) with the series (2.13). This yields (up to second order, apart from constants)

    H(q,p)=(ω+εa1+ε2a2+)p+3εJ208+3J0ε4p+3ε8p2εJ202cos(2q)εJ0pcos(2q)=ε2p2cos(2q)+εJ208cos(4q)+εJ04pcos(4q)+ε8p2cos(4q)+. (2.20)

    Step 2: normalization. To set the Hamiltonian into Kolmogorov normal form up to second order, we fix the value of the constants a1,a2 and we perform a sequence of Lie transformations aiming to give the Hamiltonian the form (in the transformed variables) H(2)(q,p)=Z(2)(q,p)+R(2)(q,p) where the remainder R(2) is O(ε3), while Z(2)(q,p) has the form Z(2)(q,p)=ωp+εR1(q,p)+ε2R2(q,p) with both Ri(q,p)=O(||p||2) i=1,2. To this end:

    - First order: we fix a1 so that the linear term εa1p acts as counterterm for the term ε(3J0/4)p. This provides, in the twistless case, a process by which the frequency ω can be kept fixed (in the usual twist case, instead, this would have been accomplished by exploiting the Hessian matrix of H0). Formally, we require that

    h(0)1,1q=a1pJ0pcos(2q)+J04pcos(4q)+3J0p4q=0, (2.21)

    leading to a1=3J04. Now, we insert this expression for a1 in the Hamiltonian (2.20), leading to

    H(q,p)=ωp+3εJ208+3ε8p2εJ202cos(2q)εJ0pcos(2q)ε2p2cos(2q)=+εJ208cos(4q)+εJ04pcos(4q)+ε8p2cos(4q)+ε2a2p. (2.22)

    We can now eliminate the O(ε) trigonometric terms in the Hamiltonian with the usual procedure. Namely, we define a generating function X(1)(q) used to eliminate terms not depending on the action p. These are

    h(0)1,0=3J208J202cos(2q)+J208cos(4q),

    leading to LX(1)(q)(ωp)+h(0)1,0=h(0)1,0q, that is

    X(1)(q)=J2032ωsin(4q)J204ωsin(2q).

    Note that, similarly as in the Birkhoff case (Subsection 2.1), here too we have to fix the initial conditions so that the relation q(0)=p(0)=0 is preserved between variables before and after the canonical transformation. This is achieved by setting the generating function as χ(1)1(q)=X(1)(q)+K(1)q, where K(1)=3J20/(8ω) is a constant. The general rules for the determination of the constants K(j) will be discussed in Section 3.

    Using the generating function χ(1)1(q) we obtain the intermediate Hamiltonian

    ˆH(1)=exp(Lεχ(1)1(q))H,

    given by

    ˆH(1)=ωp+3ε8p2εJ0pcos(2q)ε2p2cos(2q)+εJ04pcos(4q)+ε8p2cos(4q)=17ε2J3064ω+ε2a2p35ε2J2064ωp+ε2J302ωcos(2q)+7ε2J20p8ωcos(2q)11ε2J3032ωcos(4q)=7ε2J20p16ωcos(4q)+ε2J308ωcos(6q)+ε2J20p8ωcos(6q)ε2J3064ωcos(8q)ε2J20p64ωcos(8q). (2.23)

    We will now eliminate the trigonometric terms linear in the momentum in ˆH(1). These are

    ˆh(1)1,1=J0pcos(2q)+14J0pcos(4q)

    and can be eliminated by a generating function of the form χ(1)2(q,p)=˜χ(1)2(q,p)+S(1)p satisfying the homological equation Lχ(1)2(q,p)(ωp)+ˆh(1)1,1=0. As before, the value of the constant S(1) is fixed by the requirement that the initial phase of the solution of q be preserved to zero by the transformation. We find S(1)=0 (all constants S(i)=0 when the Hamiltonian has an even symmetry). Then

    ˜χ(1)2(q,p)=J0p2ωsin(2q)+J0p16ωsin(4q).

    The new Hamiltonian is

    H(1)=exp(Lεχ(1)2(q,p))ˆH(1)

    and it is in Kolmogorov normal form up to order ε.

    Repeating the same procedure at second order we arrive at the formulas

    a2=69J2064ω,χ(2)1=17J30q64ω2+J304ω2sin(2q)11J30128ω2sin(4q)+J3048ω2sin(6q)J30512ω2sin(8q),χ(2)2=37J20p64ω2sin(2q)7J20p64ω2sin(4q)+J20p64ω2sin(6q)J20p512ω2sin(8q), (2.24)

    leading to the Hamiltonian H(2)(˜q,˜p)=Z(2)(˜q,˜p)+R(2)(˜q,˜p) with the Kolmogorov normal form part

    Z(2)(˜q,˜p)=ω˜p+3ε˜p28ε˜p22cos(2˜q)+ε˜p28cos(4˜q)51ε2J0˜p264ω+37ε2J0˜p232ωcos(2˜q)=7ε2J0˜p216ωcos(4˜q)+3ε2J0˜p232ωcos(6˜q)ε2J0˜p264ωcos(8˜q)

    ((˜q,˜p) denote again the new variables after new normalization steps).

    Step 3: calculation of the solution on the torus. Using the compact notation

    Z(2)(˜q,˜p)=ω˜p+εR1(˜q,˜p)+ε2R2(˜q,˜p),

    where Rj(˜q,˜p)=O(||˜p||2) j=1,2, the equations of motions under the Hamiltonian Z(2) are

    {˙˜q=Z(2)˜p=ω+εR1(˜q,˜p)˜p+ε2R2(˜q,˜p)˜p=ω+O(||˜p||)˙˜p=Z(2)˜q=εR1(˜q,˜p)˜qε2R2(˜q,˜p)˜q=O(||˜p||2).

    The torus ˜p(t)=0, ˜q(t)=ωt (where we chose ˜q(0)=0) is a solution of this system. This can be back-transformed in the original variables using

    q=expLε2χ(2)2expLεχ(1)2˜q,p=expLε2χ(2)2expLε2χ(2)1expLεχ(1)2expLεχ(1)1˜p. (2.25)

    Substituting the solution (˜q(t)=ωt,˜p(t)=0) in the previous expressions, we readily deduce the solution in the original variables:

    q(t)=ωtεJ02ωsin(2ωt)+εJ016ωsin(4ωt)+19ε2J2032ω2sin(2ωt)+ε2J2064ω2sin(4ωt)=ε2J2032ω2sin(6ωt)+ε2J20512ω2sin(8ωt),p(t)=3εJ208ω+εJ202ωcos(2ωt)εJ208ωcos(4ωt)+17ε2J3032ω2=21ε2J3032ω2cos(2ωt)+3ε2J3032ω2cos(4ωt)+ε2J3032ω2cos(6ωt),J(t)=J0+p(t). (2.26)

    Also, using the computed expressions of a1 and a2 (Eqs (2.21) and (2.24)), we obtain the relation between the torus frequency and the original frequency

    ω=ω0+3εJ0469ε2J2064ω.

    These expressions are identical to the ones found by the Lindstedt method of Subsection 2.3 (see (2.17)).

    In order to better visualize the differences between the methods (ⅰ) and (ⅱ) (i.e., respectively, 'analogous to Birkhoff' and 'analogous to Kolmogorov') we report, in the Tables 1 and 2 below, the series terms corresponding to the solutions for q(t) and J(t) as obtained by the two methods up to order ε2.

    Table 1.  Comparison between the series terms for the solution q(t) in the Lindstedt series obtained by the methods (ⅰ) and (ⅱ) up to order ε2.
    Method: O(1) O(ε) O(ε2)
    (ⅰ) 'Analogous to Birkhoff' ωt J02ω0sin(2ωt)+J016ω0sin(4ωt) 31J2032ω20sin(2ωt)J2032ω20sin(4ωt)J2032ω20sin(6ωt)+J20512ω20sin(8ωt)
    (ⅱ) 'Analogous to Kolmogorov' ωt J02ωsin(2ωt)+J016ωsin(4ωt) 19J2032ω2sin(2ωt)+J2064ω2sin(4ωt)J2032ω2sin(6ωt)+J20512ω2sin(8ωt)

     | Show Table
    DownLoad: CSV
    Table 2.  Comparison between the series terms for the solution J(t) in the Lindstedt series obtained by the methods (ⅰ) and (ⅱ) up to order ε2.
    Method: O(1) O(ε) O(ε2)
    (ⅰ) 'Analogous to Birkhoff' J0 3J208ω0+J202ω0cos(2ωt)J208ω0cos(4ωt) 13J3016ω2033J3032ω20cos(2ωt)+3J3016ω20cos(4ωt)+J3032ω20cos(6ωt)
    (ⅱ)'Analogous to Kolmogorov' J0 3J208ω+J202ωcos(2ωt)J208ωcos(4ωt) 17J3032ω221J3032ω2cos(2ωt)+3J3032ω2cos(4ωt)+J3032ω2cos(6ωt)

     | Show Table
    DownLoad: CSV

    From Table 1 above, we observe that the solutions q(t) and J(t) obtained by methods (ⅰ) and (ⅱ) have the same form up to order O(ε), except for the fact that in method (ⅰ) all divisors depend on the frequency ω0 rather than ω. On the other hand, the coefficients found by the two methods start differing from the order ε2 and beyond. Moreover, the series for the frequency ω obtained by the two methods are:

    ω=ω0+3εJ0469ε2J2064ω0,ω=ω0+3εJ0469ε2J2064ω,

    which differ again as regards their divisors.

    The effect of these differences on the precision of the two methods can be seen even in the simplest case of a system with one degree of freedom in which both series are convergent (the series 'analogous to Birkhoff' are, instead, divergent when N2). To this end, we report below a comparison between the numerical solutions and the analytical ones, obtained by the methods (ⅰ) and (ⅱ) in the example of the Hamiltonian (2.1). To produce the Lindstedt series (ⅱ) in this example we work as follows: fixing the initial and final frequency ω0 and ω, we reverse the series (2.8) according to

    dω=i0εiai(J0):=f(J0)J0=f1(dω), (2.27)

    where dω=ωω0 and f1 denotes the series inverse to f. Then, having specified J0 through the inverse series (2.27), we compute all numerical coefficients in the Lindstedt series (ⅰ) and (ⅱ) up to order ε4. We analyze three different cases:

    1) ε=1, ω0=1 and ω=1.002;

    2) ε=1, ω0=1 and ω=1.02;

    3) ε=1, ω0=1 and ω=1.2.

    Let us note that all results are rescalable to different choices of ε. Reversing the series (2.8) as prescribed by (2.27), we obtain, respectively, the following values for the amplitude (initial action datum): J0=0.0026769, 0.0277048,0.383509. Substituting these values of ε, ω, ω0 and J0 in the solutions q(t) and J(t) we obtain the solutions, as functions of t, produced by methods (ⅰ) and (ⅱ). At the same time, it is possible to integrate the equations of motion (2.7a) and (2.7b) to produce a numerical solution, starting from the initial conditions q(0)=0, J(0)=J0.

    Figure 1 shows the difference (in Log10 scale) between the semi-analytical solutions obtained by methods (ⅰ) and (ⅱ) as above and the numerical solution for the angle q(t). Since the only systematically growing error is due to differences in the frequency estimates, all errors between the numerical and analytical solutions in the angle q(t) grow linearly in time. We observe, however, that the Lindstedt method 'analogous to Kolmogorov' always produces more precise results than the one 'analogous to Birkhoff' with a difference in precision of about one order of magnitude when dω is of order 103 and raising up to two orders of magnitude when dω becomes of the order of unity.

    Figure 1.  Difference between the numerical solution q(t) and the one produced by the method 'analogous to Birkhoff' (in blue, (ⅰ)) or 'analogous to Kolmogorov' (in red, (ⅱ)). From left to right for the cases 1) ε=1, ω0=1 and ω=1.002, 2) ε=1, ω0=1 and ω=1.02, 3) ε=1, ω0=1 and ω=1.2.

    A particularity of the example treated above is the fact that the original Hamiltonian is analytic in the whole domain JR. This changes, however, in more general models in which the Hamiltonian is of the form

    H(q,J)=ω0J+εh(q,J;ε)

    where the development in powers of the variables J contains semi-integer powers, as is, for example, the case of polynomial nonlinear couplings containing odd terms in one or more of the oscillator variables xj,yj. Let us note that physical examples of such Hamiltonian systems, with oscillators non-linearly coupled through odd polynomial terms, are ubiquitous, and include the Fermi-Pasta-Ulam αmodel, the secular Hamiltonian in resonant cases of perturbed Keplerian Nbody problems, magnetic bottle Hamiltonian models, etc. In the next section, we will present a formal algorithm applicable to a generic form for the function H1. The sequence of normalizations in this algorithm is arranged so that the result agrees with the corresponding one obtained with the Lindstedt series. A particular example showing this agreement is presented in the Appendix.

    We now give the Kolmogorov algorithm for generic isochronous systems with Hamiltonian

    H(q,J)=ω0J+i1s3εi˜hi,s(q,J) (3.1)

    where qTn, JBRn, ˜hi,s=O(||J||s2). We assume that the Hamiltonian (3.1) has a half-integer power dependence on J, i.e., admits the expansion, for n, of the truncated series

    H(q,p;J0)=ω0J0+ω0p+i1s3εi˜hi,s(q,J0+p)=ω0J0+ω0p+i1s3nk=0εi˜h(k)i,s(q,J0+p)|p=0k!pk, (3.2)

    where ˜h(k) are the kth derivatives of ˜h with respect to p=JJ0.

    Apart from constants, the Hamiltonian can be written in the compact notation

    H(q,p;J0)=ω0p+i1εihi(q,p;J0). (3.3)

    The algorithm allows to compute quasi-periodic orbits with a frequency ω fixed in advance, given by

    ω=ω0i1εiai(J0), (3.4)

    where the parameters J0, whose values are to be specified in the end of the process, give the initial conditions for J of a trajectory on the torus with frequencies ω. To this end, along with the normalizing canonical transformation, the algorith computes 'on the go' the functions ai(J0) ('counter-terms').

    Substituting the series (3.4) in the Hamiltonian (3.3) we arrive at:

    H(q,p;J0)=ωp+i1εihi(q,p;J0)+i1εiai(J0)p. (3.5)

    In the following we use the notation

    εih(k)i(q,p;J0)=εi(h(k)i,0(q;J0)+h(k)i,1(q,p;J0)+j2h(k)i,j(q,p;J0)), (3.6)

    where h(k)i,0 is independent from p, h(k)i,1 linear in p and the remaining sum is O(||p||2). To facilitate reading, the indices (i,j,k) used in all subsequent expressions refer to

    h(k)i,j:i=degreeofthecorrespondingε,h(k)i,j:j=degreeofp,h(k)i,j:k=stepofthealgorithm. (3.7)

    We now have the following

    Proposition 1. Assume the vector ω is non-resonant. There exist Lie generating functions χ(r)1, χ(r)2 such that, after r normalization steps, the Hamiltonian (3.5) is given by the formal series:

    H(r)(q(r),p(r);J0)=(exp(Lεrχ(r)2)(exp(Lεrχ(r)1)H(r1)))|q(r1)=q(r)p(r1)=p(r)=ωp+εC1++εr1Cr1+εrCr=+εh(r)1(q,p;J0)++εrh(r)r(q,p;J0)=+εr+1h(r)r+1(q,p;J0)+εr+1ar+1(J0)p+,

    where h(r)j=i0h(r)j,ijr+1 and h(r)j=i2h(r)j,iO(||p||2)j=1,,r, with

    h(r)k,0=Ck=h(k1)k,0ωK(k)1kr,h(r)k,0=h(r1)k,0r+1k2r1,h(r)2r,0=2r1rj=01j!Ljχ(r)1(q)(h(r1)2rjr,j)h(r)k,0=k1r2j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,s)+1(k1r1)!L(k1)/r1χ(r)2(q,p)(h(r1)k(k1r1)r,0)k2r+1,kmr,mN,h(r)k,0=k1r1j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,s)k2r+1,k=mr,mN,h(r)k,1=01kr,h(r)k,1=k1r1j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,1+s)kr+1,kmr,mN,h(r)k,1=k2rj=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,1+s)h(r)k,i=+m1m!Lm1χ(r)2(q,p)(h(r1)r,1)kr+1,k=mr,mN,h(r)k,i=h(r1)k,i1kr,i2,h(r)k,i=k1rj=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,i+s)kr+1,i2.

    Given the Fourier series h(r1)r,0(q)=kZnc(r1)keikq and h(r1)r,1(q,p)=|j|=1kZnc(r)jkpjeikq, the generating functions χ(r)1, χ(r)2 are defined by

    χ(r)1(q)=X(r)(q)+K(r)q=kZn{0}(c(r1)kikωeikqc(r1)kkωkq),χ(r)2(q,p)=˜χ(r)2(q,p)+S(r)pχ(r)2(q,p)=Y(r)(q)p+S(r)p=|j|=1kZn{0}(c(r)jkikωeikq+ic(r)jkkω)pj.

    Proof of the proposition. The generic r-th iterative step of the algorithm is defined as follows: after r1 steps, the hamiltonian (3.5) has the form:

    H(r1)(q,p;J0)=ωp+εC1++εr1Cr1+εh(r1)1(q,p;J0)++εr1h(r1)r1(q,p;J0)+εrh(r1)r(q,p;J0)+εrar(J0)p+εr+1h(r1)r+1(q,p;J0)+εr+1ar+1(J0)p+, (3.8)

    where C1Cr1 are constants and h(r1)iO(||p||2) i=1,,r1. Taking into account the notation (3.7), we re-define the quantity h(k)i according to

    h(k)i,1h(k)i,1+aip=(ph(k)i,1+ai)pir. (3.9)

    We then have:

    First part of the proof: the r-th normalization step consists of two substeps, each involving a canonical transformation.

    First half step: we set

    ˆH(r)=exp(Lεrχ(r)1(q))H(r1)=j0(εr)jj!Ljχ(r)1(q)H(r1), (3.10)

    where the generating function χ(r)1 is defined as χ(r)1(q)=X(r)(q)+K(r)q, with K(r) an appropriate constant vector defined below, and X(r) is defined through the homological equation:

    {ωp,X(r)(q)}+h(r1)r,0(q)=h(r1)r,0(q), (3.11)

    where . denotes the mean over q. The function X(r)(q) eliminates all terms of order O(εr) depending only on the angles q in the Hamiltonian H(r1). Writing h(r1)r,0(q) in the Fourier form h(r1)r,0(q)=kZnc(r1)keikq from (3.11) we find:

    X(r)(q)=kZn{0}c(r1)kikωeikq. (3.12)

    We then impose the condition that the terms of order O(εr) linear in p have zero

    h(r1)r,1(q,p;J0)=0. (3.13)

    This specifies ar as a function of J0 via Eq (3.9).

    Finally, we specify the constant vector K(r). To this end, we impose the condition that the solution in p has the form kAk(cos(kq)1), so that, at time t=0, we have p(0)=0. Writing (3.12) as

    X(r)(q)=kZn{0}[ic(r1)kkωcos(kq)+c(r1)kkωsin(kq)], (3.14)

    the poisson bracket of χ(r)1(q) with the i-th component pi (1in) of the vector p yields the following expressions for the i-th components K(r)i and ki of the vectors K(r) and k:

    {pi,X(r)(q)}+{pi,K(r)q}=kZn{0}[{pi,ic(r1)kkωcos(kq)}+{pi,c(r1)kkωsin(kq)}]K(r)i=kZn{0}[ic(r1)kkωkisin(kq)+c(r1)kkωkicos(kq)]K(r)i==kZn{0}ic(r1)kkωkisin(kq)kZn{0}c(r1)kkωki(cos(kq)1). (3.15)

    Hence

    K(r)i=kZn{0}c(r1)kkωki, (3.16)

    i.e.,

    K(r)=kZn{0}c(r1)kkωk. (3.17)

    Finally, we compute ˆH(r) in (3.10) as:

    ˆH(r)(q,p;J0)=ωp+εC1++εr1Cr1+εrCr+εˆh(r)1(q,p;J0)++εr1ˆh(r)r1(q,p;J0)+εrˆh(r)r(q,p;J0)+εr+1ˆh(r)r+1(q,p;J0)+, (3.18)

    where ˆh(r)j=i0ˆh(r)j,ijr and ˆh(r)j=i2ˆh(r)j,iO(||p||2) j=1,,r1, where:

    ˆh(r)k,0=Ck=h(k1)k,0ωK(k)1kr, (3.19)
    ˆh(r)k,0=h(r1)k,0r+1k2r1, (3.20)
    ˆh(r)k,0=k1rj=01j!Ljχ(r)1(q)(h(r1)kjr,j)k2r (3.21)
    ˆh(r)k,1=01kr1, (3.22)
    ˆh(r)r,1=h(r1)r,1, (3.23)
    ˆh(r)k,i=k1rj=01j!Ljχ(r)1(q)(h(r1)kjr,i+j)kr+1,i1, (3.24)
    ˆh(r)k,i=h(r1)k,i1kr,i2. (3.25)

    Second half step (of r-th): We compute

    H(r)=exp(Lεrχ(r)2(q,p))ˆH(r)=j0(εr)jj!Ljχ(r)2(q,p)ˆH(r), (3.26)

    with a generating function χ(r)2(q,p) linear in p. Setting χ(r)2(q,p)=˜χ(r)2(q,p)+S(r)p=Y(r)(q)p+S(r)p. The function ˜χ(r)2(q,p) satisfies the homological equation

    {ωp,˜χ(r)2(q,p)}+ˆh(r)r,1(q,p)=0. (3.27)

    Setting ˆh(r)r,1(q,p)=|j|=1kZnc(r)jkpjeikq the solution of (3.27) is

    ˜χ(r)2(q,p)=|j|=1kZn{0}c(r)jkikωpjeikq. (3.28)

    We finally compute the constant vector S(r) by the condition q(0)=0 at the time t=0. By the poisson bracket

    {qi,χ(r)2}={qi,˜χ(r)2(q,p)}+{qi,S(r)p}=|j|=1(kZn{0}[{qi,c(r)jkikωpjcos(kq)}+{qi,c(r)jkkωpjsin(kq)}]+{qi,S(r)jpj})=|j|=1(kZn{0}[c(r)jkikωjipji1icos(kq)+c(r)jkkωjipji1isin(kq)]+S(r)jjipji1i)=|j|=1kZn{0}[c(r)jkikωjipji1i(cos(kq)1)+c(r)jkkωjipji1isin(kq)] (3.29)

    we obtain

    S(r)j=kZn{0}ic(r)jkkω. (3.30)

    The Hamiltonian H(r) (Eq (3.26)) is:

    H(r)(q,p;J0)=ωp+εC1++εr1Cr1+εrCr+εh(r)1(q,p;J0)++εrh(r)r(q,p;J0)+εr+1h(r)r+1(q,p;J0)+, (3.31)

    where h(r)j=i0h(r)j,ijr+1 and h(r)j=i2h(r)j,iO(||p||2)j=1,,r, where:

    h(r)k,0=ˆh(r)k,01k2r, (3.32)
    h(r)k,0=k1rj=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,0)k2r+1, (3.33)
    h(r)k,1=01kr, (3.34)
    h(r)k,1=k1rj=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,1)kr+1,kmr(mN), (3.35)
    h(r)k,1=k1rj=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,1)+1m!Lmχ(r)2(q,p)(ωp)kr+1,k=mr(mN). (3.36)
    h(r)k,i=ˆh(r)k,i1kr,i2, (3.37)
    h(r)k,i=k1rj=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,i)kr+1,i2. (3.38)

    Equation (3.36), using homological equation (3.27), can be written equivalently as:

    h(r)k,1=m2j=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,1)+m1m!Lm1χ(r)2(q,p)(ˆh(r1)r,1)kr+1,k=mr(mN) (3.39)

    where, from (3.36) we have

    m1j=01j!Ljχ(r)2(q,p)(ˆh(r)mrjr,1)+1m!Lmχ(r)2(q,p)(ωp)=m2j=01j!Ljχ(r)2(q,p)(ˆh(r)mrjr,1)+1(m1)!Lm1χ(r)2(q,p)(ˆh(r)r,1)+1m!Lmχ(r)2(q,p)(ωp)=m2j=01j!Ljχ(r)2(q,p)(ˆh(r)mrjr,1)+1m!Lm1χ(r)2(q,p)(ˆh(r)r,1+Lχ(r)2(q,p)(ωp))+m1m!Lm1χ(r)2(q,p)(ˆh(r)r,1).

    Second part of the proof: using the formulas (3.19)–(3.25) and (3.32)–(3.38), we can express each term in the function h(r) in terms of the functions h(r1) instead of ˆh(r). From Eq (3.32) we have

    h(r)k,0=ˆh(r)k,0=Ck=h(k1)k,0ωK(k)1kr, (3.40)
    h(r)k,0=ˆh(r)k,0=h(r1)k,0r+1k2r1, (3.41)
    h(r)2r,0=ˆh(r)2r,0=2r1rj=01j!Ljχ(r)1(q)(h(r1)2rjr,j)=h(r1)2r,0+Lχ(r)1(q)(h(r1)r,1), (3.42)

    where we used Eqs (3.19), (3.20) and (3.21), respectively. Using (3.25) we have

    h(r)k,i=ˆh(r)k,i=h(r1)k,i1kr,i2. (3.43)

    Equation (3.38) can now be written in the form

    h(r)k,i=k1r1j=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,i)+1k1r!L(k1)/rχ(r)2(q,p)(ˆh(r)kk1rr,i) (3.44)

    kr+1,i2. The indices i,j,k in the first term of the above equation satisfy the relation

    kjrkk1rr+rr+1, (3.45)

    where we have used the inequalities j(k1)/r1 and k1k1. Then Eq (3.45) ensures that the first term of (3.44) satisfies the definition (3.24). For the second term in (3.44), we have the following useful (also in the sequel) remark: we can write k=mr+f, where 0fr1 and mN. Thus

    kk1rr=mr+fmr+f1rr=ff1rr. (3.46)

    Then

    kk1rr=ff1rr=riff=0 (3.47)

    and

    1kk1rr=ff1rrr1if1fr1. (3.48)

    We can conclude that 1k(k1)/rrr, implying that the definition (3.25) holds for the second term in (3.44). We can then write (3.44) as

    Observe that, due to (3.47) and (3.48), we have that if j=(k1)/r, then 0(kjr1)/r=(k(k1)/rr1)/r(r1)/r=0, i.e., s=0. This allows to join all terms in a single sum.

    h(r)k,i=k1r1j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,i+s)+1k1r!L(k1)/rχ(r)2(q,p)(h(r1)kk1rr,i)=k1rj=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,i+s), (3.49)

    kr+1,i2.

    Equation (3.33) can be analyzed similarly as above, by splitting it in three different parts

    h(r)k,0=k1r2j=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,0)+1(k1r1)!L(k1)/r1χ(r)2(q,p)(ˆh(r)k(k1r1)r,0)+1k1r!L(k1)/rχ(r)2(q,p)(ˆh(r)kk1rr,0) (3.50)

    k2r+1. We study separately the relations satisfied by the indices (j,k) of each of the three terms in the previous equality. Following (3.45), we have that

    kjrkk1rr+2r2r+1,

    (since j(k1)/r2). Then, the definition (3.21) holds for the first term. For the second and third terms, we have different formulas according to whether or not k is a multiple of r.

    (ⅰ) First case: k is not a multiple of r, i.e.,

    k=mr+fwith1fr1andmN.

    From (3.48) we have that 1kk1rrr1 and, consequently, r+1kk1rr+r2r1. Then, the definitions (3.19) and (3.20) hold, respectively, for the third and second term of (3.50). Thus, we can write Eq (3.50) as

    h(r)k,0=k1r2j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,s)+1(k1r1)!L(k1)/r1χ(r)2(q,p)(h(r1)k(k1r1)r,0),

    k2r+1, kmr, mN.

    (ⅱ) Second case: k is a multiple of r, i.e.,

    k=mrwithmN.

    From (3.47) we now have that kk1rr=r and, consequently, kk1rr+r=2r. Then, the definitions (3.19) and (3.21) hold, respectively, for the third and the second term of (3.50). Thus, we can write Eq (3.50) as

    h(r)k,0=k1r1j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,s), (3.51)

    k2r+1, k=mr, mN. By the same argument, we can write Eq (3.35) as:

    h(r)k,1=k1r1j=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,1)+1k1r!L(k1)/rχ(r)2(q,p)(ˆh(r)kk1rr,1) (3.52)

    kr+1,kmr(mN). In view of the inequalities (3.45) and (3.48) we then readily find that the definitions (3.24) and (3.22) hold, respectively, for the first and second part of the previous equation, i.e., (3.52) leads to

    h(r)k,1=k1r1j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,1+s), (3.53)

    kr+1,kmr(mN). Finally, recalling again (3.45) and (3.47), we can write Eq (3.36) as§

    §In the last equality, due to (3.47), we have that if j=(k1)/r, then (kjr1)/r=(k(k1)/rr1)/r=(r1)/r=0, i.e., s=0. Thus, also here we can join the terms in a single sum.

    h(r)k,1=k1r1j=01j!Ljχ(r)2(q,p)(ˆh(r)kjr,1)+1k1r!L(k1)/rχ(r)2(q,p)(ˆh(r)kk1rr,1)+1m!Lmχ(r)2(q,p)(ωp)h(r)k,1=k1r1j=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,1+s)+1k1r!L(k1)/rχ(r)2(q,p)(h(r1)kk1rr,1)+1m!Lmχ(r)2(q,p)(ωp)=k1rj=0kjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)kjrsr,1+s)+1m!Lmχ(r)2(q,p)(ωp), (3.54)

    kr+1,k=mr(mN) (where we used the definitions (3.24) and (3.23) for the first and second part of the sum). As before, we can then write the previous equation as

    m1j=0mrjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)mrjrsr,1+s)+1m!Lmχ(r)2(q,p)(ωp)=m2j=0mrjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)mrjrsr,1+s)+1(m1)!Lm1χ(r)2(q,p)(h(r1)r,1)+1m!Lmχ(r)2(q,p)(ωp)=m2j=0mrjr1rs=01j!s!Ljχ(r)2(q,p)Lsχ(r)1(q)(h(r1)mrjrsr,1+s)+1m!Lm1χ(r)2(q,p)(h(r1)r,1+Lχ(r)2(q,p)(ωp))+m1m!Lm1χ(r)2(q,p)(h(r1)r,1). (3.55)

    Use of Eq (3.23) and the homological equation (3.27) then concludes the proof.

    The focus of the present paper is twofold.

    ⅰ) We emphasize the differences between two distinct methods by which Lindstedt series can be computed in nonlinearly coupled oscillator Hamiltonian models. In particular, we discussed a Lindstedt method called 'analogous to the Birkhoff series', and another called 'analogous to the Kolmogorov series'. In the first case, series expansions are defined in open domains in the action space around the origin, and the frequencies are represented as (series) polynomial functions of the action variables. In the second method, instead, the frequencies along any required torus solution must be fixed in advance ('torus fixing'), and the series allow (after reversion) to determine a posteriori the amplitudes (or values of the actions) for which the motion takes place with a given set of frequencies. We explain how this difference in the physical interpretation of what the series are meant to compute leads also to formal differences in the way the series terms are computed, as well as to real differences in the convergence and precision properties of the series.

    ⅱ) We propose and give the formal structure of an algorithm viewed as an analogue of the Kolmogorov scheme within the framework of the 'torus fixing' method, but properly dealing with the lack of the twist property in the Hamiltonian appearing in the kernel of the associated perturbative scheme. In particular, we demonstrate how the lack of the twist condition can be compensated by a suitable introduction of counter-terms (in the spirit of the procedures proposed already in [1,11,12,20]). In fact, we demonstrate how to compute such counter-terms in a way leading to precisely the same frequency corrections as those obtained by the Lindstedt method called 'analogous to Kolmogorov'.

    Our formal algorithm given in Section 3 above is accompanied by particular examples offering some intuition into the subtleties of each examined method. On the other hand, as stressed in the introduction, the convergence of the Kolmogorov series computed by the algorithm presented in Section 3 can only be inferred by an indirect argument, namely, their equivalence with the Lindstedt series of Subsection 2.3. Thus, we point out the interesting open question of a direct proof of the convergence of our hereby presented KAM algorithm for isochronous systems (see also [20]).

    The authors declare no conflict of interest.

    Consider the following one-degree of freedom Hamiltonian, with odd power dependence on the variable x

    H(x,p)=ω02(p2+x2)εx33. (A.1)

    We pass to action-angle variables (J,q) through the transformation x=2Jsin(q), p=2Jcos(q), obtaining

    H(q,J)=ω0Jε2J3/2sin(q)+ε32J3/2sin(3q). (A.2)

    Finally, we define the translation J=J0+p leading to (apart from constants)

    H(q,p)=ω0pε2(J0+p)3/2sin(q)+ε32(J0+p)3/2sin(3q). (A.3)

    Since the variable J=J0+p appears in the above Hamiltonian in half-integer powers, we expand the Hamiltonian in powers of the variable p up to the same order as the maximum normalization order in ε. This proves to be sufficient since higher powers of p only influence the process at powers (in the book keeping order) higher than the maximum normalization order. In particular, the following lemma can be easily proved:

    Lemma 1. LetH(q,J)=ω0J+εh(q,J), where h=O(Jk2) k3. Then, for all n1 (with n= number of normalization steps) χ(n)1=O(J(k2)n+220) and χ(n)2=O(J(k2)n20).

    Proof. For the proof, see the appendix B.

    We will now illustrate the method by computing the direct (Lindstedt) and indirect (Kolmogorov) series up to order 3 in ε in the example above (the reason for reaching order 3 instead of 2 will become clear below). Starting from the Hamiltonian (A.3) we perform an expansion in p leading to (apart from constants)

    H(q,p)=ω0pεJ3/202sin(q)3εJ022psin(q)3ε82J0p2sin(q)+ε162J3/20p3sin(q)=+εJ3/2032sin(3q)+εJ022psin(3q)+ε82J0p2sin(3q)ε482J3/20p3sin(3q). (A.4)

    Following the same procedure as in subsection 2.3, we start with the equations of motion under the Hamiltonian (A.2)

    {˙q(t)=HJ=ω03εJ22sin(q)+εJ22sin(3q)(A.5a)˙J(t)=Hq=εJ3/22cos(q)εJ3/22cos(3q).(A.5b)

    Replacing, as before, the expressions

    ω0=ω+εa1+ε2a2+ε3a3,q(t)=q0(t)+εq1(t)+ε2q2(t)+ε3q3(t),J(t)=J0(t)+εJ1(t)+ε2J2(t)+ε3J3(t), (A.6)

    into the equations of motion and performing an expansion up to order 3 in ε (having fixed ω) we compare terms of like orders in ε in the l.h.s and r.h.s of (A.5a) and (A.5b). At order zero we have

    ˙q0(t)=ω,˙J0(t)=0q0(t)=ωt,J0(t)=J0,

    where we fix the initial phase q0(0)=0 and J0(0)=J0. At order one, we find

    ˙q1(t)=a13J022sin(ωt)+J022sin(3ωt),˙J1(t)=J3/202cos(ωt)J3/202cos(3ωt). (A.7)

    Since no constant terms arise in ˙q1(t), we have that a1=0. Then

    q1(t)=22J03ω+3J022ωcos(ωt)J062ωcos(3ωt),J1(t)=J3/202ωsin(ωt)J3/2032ωsin(3ωt)

    yielding the constants q1(0)=J1(0)=0. At order two we now get

    ˙q2(t)=a25J06ω+J0ωcos(ωt)+3J08ωcos(2ωt)J0ωcos(3ωt)+J02ωcos(4ωt)J024ωcos(6ωt),˙J2(t)=2J203ωsin(ωt)+4J203ωsin(2ωt)2J20ωsin(3ωt)+2J203ωsin(4ωt); (A.8)

    To compensate for the constant term in ˙q2(t) we now set a2=5J0/(6ω). Then,

    q2(t)=J0ω2sin(ωt)+3J016ω2sin(2ωt)J03ω2sin(3ωt)+J08ω2sin(4ωt)J0144ω2sin(6ωt),J2(t)=5J206ω22J203ω2cos(ωt)2J203ω2cos(2ωt)+2J203ω2cos(3ωt)J206ω2cos(4ωt). (A.9)

    In a similar way, at order 3 we find a3=0, and the solutions

    q(t)=ωt22εJ03ω+3εJ022ωcos(ωt)εJ062ωcos(3ωt)+ε2J0ω2sin(ωt)+3ε2J016ω2sin(2ωt)q(t)=ε2J03ω2sin(3ωt)+ε2J08ω2sin(4ωt)ε2J0144ω2sin(6ωt)382ε3J3/2081ω3+ε3J3/202ω3cos(ωt)q(t)=ε3J3/2022ω3cos(2ωt)+145ε3J3/201442ω3cos(3ωt)2ε3J3/203ω3cos(4ωt)+ε3J3/20162ω3cos(5ωt)q(t)=+ε3J3/20182ω3cos(6ωt)ε3J3/20482ω3cos(7ωt)+ε3J3/2012962ω3cos(9ωt),J(t)=J0+εJ3/202ωsin(ωt)εJ3/2032ωsin(3ωt)+5ε2J206ω22ε2J203ω2cos(ωt)2ε2J203ω2cos(2ωt)J(t)=+2ε2J203ω2cos(3ωt)ε2J206ω2cos(4ωt)+7ε3J5/2042ω3sin(ωt)82ε3J5/209ω3sin(2ωt)J(t)=+13ε3J5/2082ω3sin(3ωt)42ε3J5/209ω3sin(4ωt)+7ε3J5/20722ω3sin(5ωt),ω=ω05ε2J06ω. (A.10)

    Starting from the Hamiltonian (A.4), we perform canonical transformations in order to bring the Hamiltonian into Kolmogorov normal form, i.e., H(q,p)=ωp+εR1(q,p)+ε2R2(q,p)+ε3R3(q,p) where Ri(q,p)=O(||p||2) i=1,2,3. Substituting, as in Section 2, the expression ω0=ω+εa1+ε2a2+ε3a3 in the Hamiltonian (A.4), we have

    H(q,p)=ωp+εa1pεJ3/202sin(q)3εJ022psin(q)3ε82J0p2sin(q)=+ε162J3/20p3sin(q)+εJ3/2032sin(3q)+εJ022psin(3q)+ε82J0p2sin(3q)=ε482J3/20p3sin(3q)+ε2a2p+ε3a3p. (A.11)

    At first order, we have

    h(0)1,1q=a1p3J022psin(q)+J022psin(3q)q=0a1=0; (A.12)

    implying that a1=0 for the corresponding counterterm in the Hamiltonian (A.11). In order to eliminate the terms constant in the actions (depending only in the angle q) εh1,0, given by

    h(0)1,0(q)=J3/202sin(q)+J3/2032sin(3q),

    we define the generating function X(1)(q) such that LX(1)(q)(ωp)+h(0)1,0=h(0)1,0q. Hence

    X(1)=J3/202ωcos(q)J3/2092ωcos(3q).

    In order to fix the initial value of p at zero, we then set (as in Section 2) χ(1)1(q)=X(1)(q)+K(1)q; in this case, the constant K(1)=0. We can now determine the intermediate Hamiltonian

    ˆH(1)=exp(Lεχ(1)1(q))H.

    Denoting by εˆh(1)1,1 the terms of ˆH(1) of order one in ε and linear in p, we have

    ˆh(1)1,1=3J0p22sin(q)+J0p22sin(3q).

    These terms are eliminated by the generating function χ(1)2(q,p)=˜χ(1)2(q,p)+S(1)p satisfying the homological equation Lχ(1)2(q,p)(ωp)+ˆh(1)1,1=0. In order to fix the initial value of q at zero, we readily find

    S(1)=22J03ω

    and

    χ(1)2(q,p)=22J0p3ω+3J0p22ωcos(q)J0p62ωcos(3q).

    Thus, the Hamiltonian

    H(1)=exp(Lεχ(1)2(q,p))ˆH(1)

    is now in Kolmogorov normal form up to first order in ε (it is easy to check that the transformed Hamiltonian contains the normal form terms 3p282J0sin(q)+p282J0sin(3q)+p3162J3/20sin(q)p3482J3/20sin(3q)).

    In a similar way we can proceed at orders 2 and 3, obtaining the formulas:

    a2=5J06ω,χ(2)1=5J20q12ω2+5J2016ω2sin(2q)J2016ω2sin(4q)+J20144ω2sin(6q),χ(2)2=J0p2ω2sin(q)+13J0p32ω2sin(2q)J0p6ω2sin(3q)+J0p288ω2sin(6q) (A.13)

    and

    a3=0,χ(3)1=49J5/20962ω3cos(q)5J5/2062ω3cos(2q)+43J5/204322ω3cos(3q)+J5/2032ω3cos(4q)χ(3)1=29J5/202402ω3cos(5q)J5/20182ω3cos(6q)+7J5/201922ω3cos(7q)11J5/2051842ω3cos(9q),χ(3)2=107J3/20p1622ω3+295J3/20p1922ω3cos(q)47J3/20p362ω3cos(2q)+133J3/20p2882ω3cos(3q)χ(3)2=J3/20p182ω3cos(4q)+13J3/20p2882ω3cos(5q)J3/20p362ω3cos(6q)+7J3/20p11522ω3cos(7q)χ(3)2=J3/20p103682ω3cos(9q), (A.14)

    for the generating functions. The final Hamiltonian is

    H(3)(˜q,˜p)=ω˜p+εR1(q,p)+ε2R2(q,p)+ε3R3(q,p)

    where (˜q,˜p) indicate the new variables, and Rj(˜q,˜p)=O(||˜p||2) j=1,2,3. Thus, the torus ˜p(t)=0,˜q(t)=ωt is a solution for the equations of motion of this Hamiltonian. Using the Lie transformations, the solution in the original variables reads

    q(t)=ωt22εJ03ω+3εJ022ωcos(ωt)εJ062ωcos(3ωt)+ε2J0ω2sin(ωt)+3ε2J016ω2sin(2ωt)q(t)=ε2J03ω2sin(3ωt)+ε2J08ω2sin(4ωt)ε2J0144ω2sin(6ωt)382ε3J3/2081ω3+ε3J3/202ω3cos(ωt)q(t)=ε3J3/2022ω3cos(2ωt)+145ε3J3/201442ω3cos(3ωt)2ε3J3/203ω3cos(4ωt)+ε3J3/20162ω3cos(5ωt)q(t)=+ε3J3/20182ω3cos(6ωt)ε3J3/20482ω3cos(7ωt)+ε3J3/2012962ω3cos(9ωt),p(t)=εJ3/202ωsin(ωt)εJ3/2032ωsin(3ωt)+5ε2J206ω22ε2J203ω2cos(ωt)2ε2J203ω2cos(2ωt)J(t)=+2ε2J203ω2cos(3ωt)ε2J206ω2cos(4ωt)+7ε3J5/2042ω3sin(ωt)82ε3J5/209ω3sin(2ωt)J(t)=+13ε3J5/2082ω3sin(3ωt)42ε3J5/209ω3sin(4ωt)+7ε3J5/20722ω3sin(5ωt),J(t)=J0+p(t). (A.15)

    Recalling also the computed values of a1, a2 and a3 (Eqs (A.12)–(A.14)) we have also

    ω=ω05ε2J06ω.

    Thus, we obtain the same solutions as by the Lindstedt method (Eq (A.10)).

    We remark that also in this case the solutions produced by the Birkhoff normal form are equal to those produced by the Lindstedt method in the version 'analogous to Birkhoff'. However, if we compare these solutions with those produced by the KAM algorithm, we also note many differences from order 3 and beyond. For completeness, we report in the following the solutions with the Birkhoff method:

    q(t)=ωt22εJ03ω0+3εJ022ω0cos(ωt)εJ062ω0cos(3ωt)+ε2J0ω20sin(ωt)+3ε2J016ω20sin(2ωt)q(t)=ε2J03ω20sin(3ωt)+ε2J08ω20sin(4ωt)ε2J0144ω20sin(6ωt)832ε3J3/2081ω30+9ε3J3/2042ω30cos(ωt)q(t)=ε3J3/2022ω30cos(2ωt)+125ε3J3/201442ω30cos(3ωt)2ε3J3/203ω30cos(4ωt)+ε3J3/20162ω30cos(5ωt)q(t)=+ε3J3/20182ω30cos(6ωt)ε3J3/20482ω30cos(7ωt)+ε3J3/2012962ω30cos(9ωt),p(t)=εJ3/202ω0sin(ωt)εJ3/2032ω0sin(3ωt)+5ε2J206ω202ε2J203ω20cos(ωt)2ε2J203ω20cos(2ωt)p(t)=+2ε2J203ω20cos(3ωt)ε2J206ω20cos(4ωt)+31ε3J5/20122ω30sin(ωt)82ε3J5/209ω30sin(2ωt)p(t)=+97ε3J5/20722ω30sin(3ωt)42ε3J5/209ω30sin(4ωt)+7ε3J5/20722ω30sin(5ωt),ω=ω05ε2J06ω0. (A.16)

    Consider the Hamiltonian:

    H(q,J)=ω0J+εh(q,J),

    where h=O(Jk2) kN, k3. Introducing the translation J=J0+p, h=O((J0+p)k2), if k is even we obtain a finite expression in the powers of p. If k is odd, define f(p)=ε(J0+p)k2 and suppose we want to compute the normalization steps by the Kolmogorov algorithm up to order εn n1. The expansion in p=0 up to order n yields

    f(p)=ε(J0+p)k2=εJk20+εk2Jk210p++εn!k2(k21)(k2n+1)Jk2n0pn.

    Moreover, observe that χ(1)1εJk20 and χ(1)2εJk210. Now, we identify those terms whose Lie derivatives could modify the generating functions along the normalization process. We have Ln1εχ(1)1(εJk2n0pn). Thus the term of higher degree in p in the expansion of f(p), denoted by f(n)(p), influences h(1)n,1. In fact, since

    Ln1εχ(1)1(εJk2n0pn)=Ln2εχ(1)1{εJk2n0pn,εχ(1)1(q)}ε2Jkn0pn1=Ln3εχ(1)1{{εJk2n0pn,εχ(1)1(q)},εχ(1)1(q)}ε3J32kn0pn2(n3)timesεnJk2n0(Jk20)n1p=εnJ(k2)n20p, (B.1)

    after n step a contribution stemming from the term f(n)(p) will appear in the generating function χ(n)2. On the contrary, the term Ln1εχ(1)1(εf(n+1)(p)) contributes to h(1)n,2 (which is quadratic in the action p), while the term Lnεχ(1)1(εf(n+1)(p)) contributes to h(1)n+1,1, which is of order εn+1. From (9.1) it follows that h(n)n,1=O(J(k2)n/20). Hence, the dependence of χ(n)2 on the parameter J0 is in the power χ(n)2J(k2)n20. Also

    Ln1εχ(1)2(εJk210p)=Ln2εχ(1)2{εJk210p,εχ(1)2(q,p)}ε2Jk20p=(n2)timesεnJk210(Jk210)n1p=εnJ(k2)n20p,

    thus, again, h(n)n,1=O(J(k2)n/20). Finally

    Ln1εχ(1)1(εJk2n+10pn1)=Ln2εχ(1)1{εJk2n+10pn1,εχ(1)1(q)}ε2Jkn+10pn2=(n2)timesεnJk2n+10(Jk20)n1p=εnJ(k2)n+220p,

    implying χ(n)1J(k2)n+220. This concludes the proof of the lemma.



    [1] M. Bartuccelli, G. Gentile, Lindstedt series for perturbations of isochronous systems: a review of the general theory, Rev. Math. Phys., 14 (2002), 121–171. http://doi.org/10.1142/S0129055X02001120 doi: 10.1142/S0129055X02001120
    [2] H. Christodoulidi, C. Efthymiopoulos, T. Bountis, Energy localization on q-tori, long-term stability, and the interpretation of fermi-pasta-ulam recurrences, Phys. Rev. E, 81 (2010), 016210. http://doi.org/10.1103/PhysRevE.81.016210 doi: 10.1103/PhysRevE.81.016210
    [3] H. Christodoulidi, C. Efthymiopoulos, Low-dimensional q-tori in FPU lattices: dynamics and localization properties, Physica D, 261 (2013), 92–113. http://doi.org/10.1016/j.physd.2013.07.007 doi: 10.1016/j.physd.2013.07.007
    [4] L. Corsi, G. Gentile, M. Procesi, KAM theory in configuration space and cancellations in the Lindstedt series, Commun. Math. Phys., 302 (2011), 359–402. http://doi.org/10.1007/s00220-010-1131-7 doi: 10.1007/s00220-010-1131-7
    [5] C. Efthymiopoulos, A. Giorgilli, G. Contopoulos, Nonconvergence of formal integrals Ⅱ. Improved estimates for the optimal order of truncation, J. Phys. A: Math. Gen., 37 (2004), 10831. http://doi.org/10.1088/0305-4470/37/45/008 doi: 10.1088/0305-4470/37/45/008
    [6] L. H. Eliasson, Absolutely convergent series expansions for quasi periodic motions, Mathematical Physics Electronic Journal, 2 (1996), 1–33.
    [7] S. Flach, M. V. Ivanchenko, O. I. Kanakov, q-Breathers and the fermi-pasta-ulam problem, Phys. Rev. Lett., 95 (2005), 064102. http://doi.org/10.1103/PhysRevLett.95.064102 doi: 10.1103/PhysRevLett.95.064102
    [8] S. Flach, M. V. Ivanchenko, O. I. Kanakov, q-breathers in Fermi-Pasta-Ulam chains: existence, localization, and stability, Phys. Rev. E, 73 (2006), 036618. http://doi.org/10.1103/PhysRevE.73.036618 doi: 10.1103/PhysRevE.73.036618
    [9] M. Hénon, C. Heiles, The applicability of the third integral of motion: some numerical experiments, The Astronomical Journal, 69 (1964), 73–79.
    [10] L. Galgani, Foundations of physics in Milan, Padua and Paris. Newtonian trajectories from celestial mechanics to atomic physics, Mathematics in Engineering, 3 (2021), 1–24. http://doi.org/10.3934/mine.2021045 doi: 10.3934/mine.2021045
    [11] G. Gallavotti, Twistless KAM tori, Commun. Math. Phys., 164 (1994), 145–156. http://doi.org/10.1007/BF02108809 doi: 10.1007/BF02108809
    [12] G. Gallavotti, Twistless KAM tori, quasi flat homoclinic intersections, and other cancellations in the perturbation series of certain completely integrable Hamiltonian systems. A review, Rev. Math. Phys., 6 (1994), 343–411. https://doi.org/10.1142/S0129055X9400016X doi: 10.1142/S0129055X9400016X
    [13] A. Giorgilli, Small denominators and exponential stability: from Poincaré to the present time, Seminario Mat. e. Fis. di Milano, 68 (1998), 19. http://doi.org/10.1007/BF02925829 doi: 10.1007/BF02925829
    [14] A. Giorgilli, U. Locatelli, On classical series expansions for quasi-periodic motions, Mathematical Physics Electronic Journal, 3 (1997), 1–25.
    [15] A. Jorba, J. Masdemont, Dynamics in the center manifold of the collinear points of the restricted three body problem, Physica D, 132 (1999), 189–213. http://doi.org/10.1016/S0167-2789(99)00042-1 doi: 10.1016/S0167-2789(99)00042-1
    [16] A. Lindstedt, Ueber die allgemeine form der integrale des dreikörperproblems, Astron. Nachr., 105 (1883), 97–112. http://doi.org/10.1002/ASNA.18831050702 doi: 10.1002/ASNA.18831050702
    [17] A. Lindstedt, Ueber die integration einer gewissen differentialgleichung, Astron. Nachr., 104 (1883), 145–150. http://doi.org/10.1002/asna.18831041002 doi: 10.1002/asna.18831041002
    [18] A. Lindstedt, Sur les séries trigonométriques dans le problème des trois corps, Bulletin Astronomique, Serie I, 3 (1886), 217–221.
    [19] H. Poincaré, Les méthodes nouvelles de la mécanique céleste. Tome II, Méthodes de MM. Newcomb, Gyldén, Lindstedt et Bohlin, Paris: Gauthier-Villars et fil, 1893.
    [20] M. Sansottera, V. Danesi, Kolmogorov variation: Kam with knobs (à la Kolmogorov), 2021, arXiv: 2109.06345.
  • This article has been cited by:

    1. Marco Sansottera, Veronica Danesi, Kolmogorov variation: KAM with knobs (à la Kolmogorov), 2023, 5, 2640-3501, 1, 10.3934/mine.2023089
  • 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(1852) PDF downloads(137) Cited by(1)

Figures and Tables

Figures(1)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog