diff --git "a/1NAyT4oBgHgl3EQfbff_/content/tmp_files/2301.00266v1.pdf.txt" "b/1NAyT4oBgHgl3EQfbff_/content/tmp_files/2301.00266v1.pdf.txt" new file mode 100644--- /dev/null +++ "b/1NAyT4oBgHgl3EQfbff_/content/tmp_files/2301.00266v1.pdf.txt" @@ -0,0 +1,9270 @@ +arXiv:2301.00266v1 [math.SG] 31 Dec 2022 +Action-angle coordinates and KAM theory for +singular symplectic manifolds +Eva Miranda +Arnau Planas +Laboratory of Geometry and Dynamical Systems, Department of +Mathematics & IMTech, Universitat Polit`ecnica de Catalunya, Barcelona +and CRM, Centre de Recerca Matem`atica, Bellaterra +Current address: UPC-Edifici P, Avinguda del Doctor Maran´on, 44-50, 08028, +Barcelona, Spain +Email address: evamiranda@upc.edu +Department of Mathematics, Universitat Polit`ecnica de Catalunya, +Barcelona +Email address: arnau.planas.bahi@gmail.com + +2020 Mathematics Subject Classification. +53D05, 53D20, 70H08, 37J35, 37J40 ; +37J39, 58D19 +Key words and phrases. amsbook, AMS-LATEX +Eva Miranda is supported by the Catalan Institution for Research and Advanced +Studies via an ICREA Academia Prize 2016 and ICREA Academia Prize 2021. +Eva Miranda is also supported by the Spanish State Research Agency, through the +Severo Ochoa and Mar´ıa de Maeztu Program for Centers and Units of Excellence +in R&D (project CEX2020-001084-M). Eva Miranda also acknowledges partial +support from the grant “Computational, dynamical and geometrical complexity in +fluid dynamics”, Ayudas Fundaci´on BBVA a Proyectos de Investigaci´on Cient´ıfica +2021. Both authors are supported by the project PID2019-103849GB-I00 of the +Spanish State Agency AEI /10.13039/501100011033. +Dedicated to the memory of Amelia Galcer´an Sorribes. + +Contents +Preface +vii +Part 1. +Introduction and preliminaries +1 +Chapter 1. +Introduction +3 +1.1. +Structure and results of this monograph +3 +Chapter 2. +A primer on singular symplectic manifolds +7 +2.1. +b-Poisson manifolds +7 +2.2. +On bm-Symplectic manifolds +10 +2.3. +Desingularizing bm-Poisson manifolds +12 +Chapter 3. +A crash course on KAM theory +15 +Part 2. +Action-angle coordinates and cotangent models +19 +Chapter 4. +An action-angle theorem for bm-symplectic manifolds +23 +4.1. +Basic definitions +23 +4.2. +On bm-integrable systems +24 +4.3. +Examples of bm-integrable systems +26 +4.4. +Looking for a toric action +29 +4.5. +Action-angle coordinates on bm-symplectic +manifolds +32 +Chapter 5. +Reformulating the action-angle coordinate via cotangent lifts +37 +5.1. +Cotangent lifts and Arnold-Liouville-Mineur in Symplectic Geometry +37 +5.2. +The case of bm-symplectic manifolds +38 +Part 3. +A KAM theorem for bm-symplectic manifolds +41 +Chapter 6. +A new KAM theorem +43 +6.1. +On the structure of the proof +43 +6.2. +Technical results +50 +6.3. +A KAM theorem on bm-symplectic manifolds +69 +Chapter 7. +Desingularization of bm-integrable systems +93 +Chapter 8. +Desingularization of the KAM theorem on bm-symplectic +manifolds +97 +Chapter 9. +Potential applications to Celestial mechanics +101 +9.1. +The Kepler Problem +101 +v + +vi +CONTENTS +9.2. +The Problem of Two Fixed Centers +102 +9.3. +Double Collision and McGehee coordinates +103 +9.4. +The restricted three-body problem +105 +Bibliography +107 + +Preface +I confess I envy the planets — +they’ve got their own orbits and +nothing stands on their way. +Intermezzo, +Mykhailo +Kotsi- +ubynsky. +This monograph explores classification and perturbation problems for inte- +grable systems on a class of Poisson manifolds called bm-Poisson manifolds. This +is the first class of Poisson manifolds for which perturbation theory is established +outside the symplectic category. Even if the class of bm-Poisson manifolds is not +ample enough to represent the wild set of Poisson manifolds, this investigation can +be seen as a first step for the study of perturbation theory for general Poisson man- +ifolds. In view of the work of the first author with Nest, the theorems established +in this monograph constitute more than a mild generalization in Poisson Geometry +and, this toy example, sets the path to consider KAM theory in the general realm +of Poisson manifolds. Reduction theorems and bm-symplectic manifolds have been +recently explored in [MM22]. This monograph contributes to the theory opening +the investigation of perturbation theory on these manifolds thus completing other +facets in the study of their dynamics as the recent work on the Arnold conjecture +[BMO22]. +Symplectic geometry has been the common language of physics as the position- +momentum tandem can be modelled over a cotangent bundle. Cotangent bundles +are naturally endowed with a symplectic form which is a non-degenerate closed +2-form. The symplectic form of the cotangent bundle is given by the differential of +the Liouville one-form. +bm-Poisson manifolds are manifolds that are symplectic away from a hyper- +surface along which they satisfy some transversality properties. They often model +problems on symplectic manifolds with boundary such as the study of their de- +formation quantization and celestial mechanics. As on the complementary of the +critical set the manifolds are symplectic, extending the investigation of Hamiltonian +dynamics to this realm is key to understand Hamiltonian Dynamics on compact- +ification of symplectic manifolds. Several regularization transformations used in +celestial mechanics (as McGehee or Moser regularization) provide examples of such +compactifications. +One of the interesting properties of bm-Poisson manifolds is that their investi- +gation can be achieved considering the language of bm-forms. That is to say, we can +work with forms that are symplectic away from the critical set and admit a smooth +extension as a form over a Lie algebroid generalizing De Rham forms as form over +the standard Lie algebroid of the tangent bundle of the manifold. +To consider +bm-forms the standard tangent bundle is replaced by the bm-tangent bundle. This +vii + +viii +PREFACE +allows us to mimic symplectic geometry by replacing the cotangent bundle by the +dual of the bm-tangent bundle. However, Poisson geometry leaves its footprint and +new invariants which can be identified as the modular class of the Poisson structure +arise already at the semilocal level. +Contrary to the initial expectations, several of the results for bm-symplectic +manifolds do not resemble the b-case so far. Considering these more general singu- +larities yields a better understanding of the general Poisson case and the different +levels of complexity. As an illustration of this phenomena: in the study of quan- +tization of those systems an interesting pattern makes the quantization radically +different in the even and odd case [GMW18b, GMW21] and the resulting model +is finite-dimensional in the b-case. Understanding how the different degrees m are +related is a hard task: The desingularization technique introduced by Guillemin- +Miranda-Weitsman in [GMW17] turned out to have important applications in the +investigation of complexity properties of toric bm-symplectic manifolds [GMW18a] +and to the study of the Arnold conjecture in this set-up [BMO22]. In this mono- +graph we explore a new facet of these manifolds: that of perturbation theory. +In the second part of the monograph we consider integrable systems on these +manifolds turn out to have associated generalized Hamiltonian actions of tori in +a neighbourhood of a Liouville torus. We use this generalized Hamiltonian group +action to prove existence of action-angle coordinates in a neighborhood of a Liouville +torus. The action-angle coordinate theorem that we prove gives a semilocal normal +form in the neighbourhood of a Liouville torus for the bm-symplectic structure +which depends on the modular weight of the connected component of the critical +set in which the Liouville torus is lying and the modular weights of the associated +toric action. This action-angle theorem allows us to identify a neighborhood of +the Liouville torus with the bm-cotangent lift of the action of a torus acting by +translations on itself. This interpretation of the action-angle theorem as cotangent +lift allows us to identify the modular weight as their only semilocal invariant. In +doing so, we compare this action-angle coordinate theorem with the classical action- +angle coordinate theorems for symplectic manifolds and an action-angle theorem +for folded symplectic manifolds ([CM22]). +In part 3 of the monograph we study perturbation theory in this new set-up +and examine some potential applications to physical systems. In particular, we +prove a KAM theorem for bm-Poisson manifolds which clearly refines and improves +the one obtained for b-Poisson manifolds in [KMS16a]. As an outcome of this +result together with the extension of the desingularization techniques of Guillemin- +Miranda-Weitsman to the realm of integrable systems, we obtain a KAM theorem +for folded symplectic manifolds where KAM theory has never been considered be- +fore. In the way, we also obtain a brand-new KAM theorem for symplectic manifolds +where the perturbation keeps track of a distinguished hypersurface. In celestial me- +chanics, this distinguished hypersurface can be the line at infinity or the collision +set. +Barcelona, December 2022, Eva Miranda and Arnau Planas + +Part 1 +Introduction and preliminaries + + +CHAPTER 1 +Introduction +Both symplectic and Poisson geometry emerge from the study of classical me- +chanics. Both are broad fields widely studied and with powerful results. But as +Poisson structures are far more general than the symplectic ones, most outstanding +results in symplectic geometry do not translate well to Poisson manifolds. Here is +where bm-Poisson structures come to play. bm-Poisson structures (or bm-symplectic +structures) lie somewhere between these two worlds. They extend symplectic struc- +tures but in a really controlled way. This is why fundamental results in symplectic +geometry still work in bm-symplectic geometry. However, an adaptation of these +theories like deformation or Moser theory requires some work (see [GMPS15a] +and others). +The study of bm-Poisson geometry sparked from the study of symplectic man- +ifold with boundary [Mel93a]. In the last years the interest in this field increased +after the classification result for b-Poisson structures obtained in [Rad02]. Later +on, [GMP14] translated these structures to the language of forms and started +applying symplectic tools to study them. A lot of papers in the following years +studied different aspects of these structures: [GMP10], [GMP14], [GMPS15b], +[GMW17], [MOT14] and [GLPR17] are some examples. +Inspired by the study of manifolds with boundary, we work on a pair of mani- +folds (M, Z) where Z is an hypersurface and call this pair b-manifold +In this context, [Sco16] generalized the b-symplectic forms by allowing higher +degrees of degeneracy of the Poisson structures. +The bm-symplectic structures +inherit most of the properties of b-symplectic structures. This booklet focuses on +different aspects of the investigation of bm-symplectic structures covering mainly +integrable systems and KAM theory. First, we present some preliminary notions +necessary to address the problem of perturbation. +We present an action-angle +theorem for bm-Poisson structures and state and prove the KAM theory equivalent +in manifolds with bm-symplectic structures. +1.1. Structure and results of this monograph +1.1.1. Part 1: Introduction and Preliminaries. In the preliminaries, we +give the basic notions that lead to the questions we are addressing in this booklet. +In the first part, we introduce the concept of b-Poisson manifolds or b-symplectic +manifolds, a class of Poisson manifold which is symplectic outside a critical hyper- +surface. It study comes motivated by the investigation of manifolds with boundary. +Next, we talk about a generalization of these structures, that allow a higher degree +of degeneracy of the structure: the bm-symplectic structures. These structures are +the main focus of our investigations. A key concept that will play an important +work in this book is the study of the desingularization of these singular structures. +3 + +4 +1. INTRODUCTION +Finally, we give a short introduction to KAM theory, a theory that will be gener- +alized in the setting of bm-manifolds in the last chapter. +Motivation comes from several examples of singular symplectic structures ap- +pearing naturally in classical problems of celestial mechanics which are discussed +on the last chapter of the monograph. We also describe the difficulties of finding +these examples, and the subtleties of dealing with these singular structures in the +exploration of conservative systems. +1.1.2. Part 2: Action-angle coordinates and cotangent models for bm- +integrable systems. In this Chapter we define the concept of bm-functions and +bm-integrable systems. We present several examples of bm-integrable systems that +come from classical mechanics. After all this we present a version of the action-angle +theorem for bm-symplectic manifolds. +Theorem A. Let (M, x, ω, F) be a bm-integrable system, where F = (f1 = +a0 log(x)+�m−1 +j=1 aj 1 +xj , f2, . . . , fn). Let m ∈ Z be a regular point, and such that the +integral manifold through m is compact. Let Fm be the Liouville torus through m. +Then, there exists a neighborhood U of Fm and coordinates (θ1, . . . , θn, σ1, . . . , σn) : +U → Tn × Bn such that: +(1) We can find an equivalent integrable system F = (f1 = a′ +0 log(x) + +�m−1 +j=1 a′ +j +1 +xj ) such that a′ +0, . . . , a′ +m−1 ∈ R, +(2) +ω|U = + + +m +� +j=1 +c′ +j +c +σj +1 +dσ1 ∧ dθn + + + +n +� +i=2 +dσi ∧ dθi +where c is the modular period and c′ +j = −(j − 1)a′ +j−1, also +(3) the coordinates σ1, . . . , σn depend only on fn, . . . fn. +1.1.3. Part 3: KAM theory on bm-symplectic manifolds and applica- +tions to Celestial Mechanics. In this chapter we provide several KAM theorem +for (singular) symplectic manifolds including bm-symplectic manifolds. +We begin by considering perturbation theory for bm-symplectic manifolds. Then +we give an outline of how to construct the bm-symplectomorphism that will be the +main character of the proof of the KAM theorem for bm-symplectic manifolds. After +this, we show some technical results that are needed for the proof. These technical +results even if quite similar to the standard KAM equivalents, have some subtleties +that need to be addressed. We end the chapter with the proof of the bm-KAM +theorem and several applications to establish KAM theorems in other singular sit- +uations (folded symplectic manifolds) and on symplectic manifolds with prescribed +invariant hypersurfaces. +The first KAM theorem is the following: +Theorem B. Let G ⊂ Rn, n ≥ 2 be a compact set. Let H(φ, I) = ˆh(I) + +f(φ, I), where ˆh is a bm-function ˆh(I) = h(I) + q0 log(I1) + �m−1 +i=1 +qi +Ii +1 defined on +Dρ(G), with h(I) and f(φ, I) analytic. Let ˆu = ∂ˆh +∂I and u = ∂h +∂I . Assume | ∂u +∂I |G,ρ2 ≤ +M, |u|ξ ≤ L. Assume that u is µ non-degenerate (| ∂u +∂I | ≥ µ|v| for some µ ∈ R+ and +I ∈ G. Take a = 16M. Assume that u is one-to-one on G and its range F = u(G) +is a D-set. Let τ > n − 1, γ > 0 and 0 < ν < 1. Let + +1.1. STRUCTURE AND RESULTS OF THIS MONOGRAPH +5 +(1) +(1.1) +ε := ∥f∥G,ρ ≤ +ν2µ2ˆρ2τ+2 +24τ+32L6M 3 γ2, +(2) +(1.2) +γ ≤ min(8LMρ2 +ν ˆρτ+1 , L +K′ ) +(3) +(1.3) +µ ≤ min(2τ+5L2M, 27ρ1L4Kτ+1, βντ+122τ+1ρτ +1), +where ˆρ := min +� +νρ1 +12(τ+2), 1 +� +. Define the set ˆG = ˆGγ := {I ∈ G− 2γ +µ |u(I) is τ, γ, c, ˆq− +Dioph.}. Then, there exists a real continuous map T : W ρ1 +4 (Tn) × ˆG → Dρ(G) an- +alytic with respect the angular variables such that +(1) For all I ∈ ˆG the set T (Tn ×{I}) is an invariant torus of H, its frequency +vector is equal to u(I). +(2) Writing T (φ, I) = (φ + Tφ(φ, I), I + TI(φ, I)) with estimates +|Tφ(φ, I)| ≤ 22τ+15ML2 +ν2ˆρ2τ+1 +ε +γ2 +|TI(φ, I))| ≤ 210+τL(1 + M) +ν ˆρτ+1 +ε +γ +(3) meas[(Tn ×G)\T (Tn× ˆG)] ≤ Cγ where C is a really complicated constant +depending on n, µ, D, diamF, M, τ, ρ1, ρ2, K and L. +Also, we obtain a way to associate a standard symplectic integrable system or +a folded integrable system to a bm-integrable system, depending on the parity of m. +This is done in such a way that the dynamics of the desingularized system are the +same than the dynamics of the original one. So it defines a honest desingularization +of the integrable system. +Theorem C. The desingularization transforms a bm-integrable system into an +integrable system on a symplectic manifold for even m. +For m odd, the desin- +gularization associates to it a folded integrable system. +The integrable systems +satisfy: +Xω +fj = Xωǫ +fjǫ. +This allows us to obtain two new KAM theorems using this desingularization +combined with the former bm-KAM theorem. The first of these theorems is a KAM +theorem for standard symplectic manifolds, where the perturbation has a particu- +lar expression. This result is more restrictive than the standard KAM theorem but +allow us to guarantee that the perturbations leave a given hypersurface invariant. +This means that the tori belonging to that hypersurface remain on the hypersur- +face after the perturbation. This can be interesting for a number of reasons and +situations such as problems in Celestial mechanics where it is convenient to keep +track of a particular hypersurface such as the line at infinity. The higher order +singularities allow to consider perturbations that are tangent to the hypersurface +up to a certain order. + +6 +1. INTRODUCTION +Theorem D. Consider a neighborhood of a Liouville torus of an integrable +system Fε as in 8.1 of a symplectic manifold (M, ωε) semilocally endowed with +coordinates (I, φ), where φ are the angular coordinates of the torus, with ωε = +c′dI1 ∧dφi +�n +j=1 dIj ∧dφj. Let H = (m−1)cm−1c′I1 +h(˜I)+R(˜I, ˜φ) be a nearly +integrable system where +� +˜I1 += +c′ Im+1 +1 +m+1 , +˜φ1 += +c′Im +1 φ1, +and +� ˜I += +(˜I1, I2, . . . , In), +˜φ += +(˜φ1, φ2, . . . , φn). +Then the results for the bm-KAM theorem 6.3 applied to Hsing = +1 +I2k−1 +1 ++ h(I) + +R(I, φ) hold also for this desingularized system. +The second one is a KAM theorem for folded-symplectic manifolds, where KAM +theory has not been considered to-date. +Theorem E. Consider a neighborhood of a Liouville torus of an integrable +system Fε as in 8.2 of a folded symplectic manifold (M, ωε) semilocally endowed +with coordinates (I, φ), where φ are the angular coordinates of the Torus, with +ωε = 2cI1dI1 ∧ dφ1 + �m +j=2 dIj ∧ dφj. Let H = (m − 1)cm−1cI2 +1 + h(˜I) + R(˜I, ˜φ) a +nearly integrable system with� +˜I1 += +2c Im+2 +1 +m+2 , +˜φ1 += +2cIm+1 +1 +φ1, +and +� ˜I += +(˜I1, I2, . . . , In), +˜φ += +(˜φ1, φ2, . . . , φn). +Then the results for the bm-KAM theorem 6.3 applied to Hsing = +1 +I2k +1 +h(I)+R(I, φ) +also hold for this desingularized system. +Last but not least, we illustrate the connection between bm-symplectic struc- +tures and classical mechanics by providing several examples. +Several potential +applications to celestial mechanics are discussed. + +CHAPTER 2 +A primer on singular symplectic manifolds +In this first chapter of the booklet we introduce basic notions on singular sym- +plectic structures, as well as some concepts on standard KAM theory. Those are +the two main pillars of this monograph. +Let M be a smooth manifold, a Poisson structure on M is a bilinear map +{·, ·} : C∞(M) × C∞(M) → C∞(M) which is skew-symmetric and satisfies both +the Jacobi identity and the Leibniz rule. It is possible to express {f, g} in terms +of a bivector field via the following equality {f, g} = Π(df ∧ dg) with Π a section +of Λ2(T M). Π is the associated Poisson bivector. We will use indistinctively +the terminology of Poisson structure when referring to the bracket or the Poisson +bivector. +A b-Poisson bivector field on a manifold M 2n is a Poisson bivector such that +the map +(2.1) +F : M → +2n +� +T M : p �→ (Π(p))n +is transverse to the zero section. Then, a pair (M, Π) is called a b-Poisson man- +ifold and the vanishing set Z of F is called the critical hypersurface. Observe +that Z is an embedded hypersurface. +This class of Poisson structures was studied by Radko [Rad02] in dimension +two and considered in numerous papers in the last years: [GMP10], [GMP14], +[GMPS15b], [GMW17], [MOT14] and [GLPR17] among others. +2.1. b-Poisson manifolds +Next, we recall classification theorem of b-Poisson surfaces as presented by Olga +Radko and the cohomological re-statement and proof given by Guillemin, Miranda +and Pires in [GMP14]. +In what follows, (M, Π) will be a closed smooth surface with a b-Poisson struc- +ture on it, and Z its critical hypersurface. +Let h be the distance function to Z as in [MOT14]1. +Definition 2.1. The Liouville volume of (M, Π) is the following limit: +V (Π) := limǫ→0 +� +|h|>ǫ ωn2. +The previous limit exists and it is independent of the choice of the defining +function h of Z (see [Rad02] for the proof). +Definition 2.2. For any (M, Π) oriented Poisson manifold, let Ω be a volume +form on it, and let uf denote the Hamiltonian vector field of a smooth function +1Notice the difference with [Rad02] where h is assumed to be a global defining function. +2For surfaces n = 1. +7 + +8 +2. A PRIMER ON SINGULAR SYMPLECTIC MANIFOLDS +f : M → R. The modular vector field XΩ is the derivation defined as follows: +f �→ Luf Ω +Ω +. +Definition 2.3. Given γ a connected component of the critical set Z(Π) of a +closed b-Poisson manifold (M, Π), the modular period of Π around γ is defined +as: +Tγ(Π) := period of XΩ|γ. +Remark 2.4. The modular vector field XΩ of the b-Poisson manifold (M, Z) +does not depend at Z on the choice of Ω because for different choices for volume +form the difference of modular vector fields is a Hamiltonian vector field. Observe +that this Hamiltonian vector field vanishes on the critical set as Π vanishes there +too. +Definition 2.5. Let Mn(M) = Cn(M)/ ∼ where Cn(M) is the space of dis- +joint oriented curves and ∼ identifies two sets of curves if there is an orientation- +preserving diffeomorphism mapping the first one to the second one and preserving +the orientations of the curves. +The following theorem classifies b-symplectic structures on surfaces using these +invariants: +Theorem 2.6 (Radko [Rad02]). Consider two b-Poisson structures Π, Π′ on +a closed orientable surface M. Denote its critical hypersurfaces by Z and Z′. These +two b-Poisson structures are globally equivalent (there exists a global orientation +preserving diffeomorphism sending Π to Π′) if and only if the following coincide: +• the equivalence classes of [Z] and [Z′] ∈ Mn(M), +• their modular periods around the connected components of Z and Z′, +• their Liouville volume. +An appropriate formalism to deal with these structures was introduced in +[GMP10]. +Definition 2.7. A b-manifold3 is a pair (M, Z) of a manifold and an embed- +ded hypersurface. +In this way, the concept of b-manifold previously introduced by Melrose is +generalized to consider additional geometric structures on the manifold. +Definition 2.8. A b-vector field on a b-manifold (M, Z) is a vector field +tangent to the hypersurface Z at every point p ∈ Z. +Definition 2.9. A b-map from (M, Z) to (M ′, Z′) is a smooth map φ : M → +M ′ such that φ−1(Z′) = Z and φ is transverse to Z′. +Observe that if x is a local defining function for Z and (x, x1, . . . , xn−1) are +local coordinates in a neighborhood of p ∈ Z then the C∞(M)-module of b-vector +fields has the following local basis +(2.2) +{x ∂ +∂x, ∂ +∂x1 +, . . . , +∂ +∂xn−1 +}. +3The ‘b’ of b-manifolds stands for ‘boundary’, as initially considered by Melrose (Chapter 2 +of [Mel93b]) for the study of pseudo-differential operators on manifolds with boundary. + +2.1. b-POISSON MANIFOLDS +9 +Figure 1. Artistic representation of a b-function on a b-manifold +near the critical hypersurface. +In contrast to [GMP10], in this monograph we are not requiring the existence +of a global defining function for Z and orientability of M. However, we require the +existence of a defining function in a neighborhood of each point of Z. By relaxing +this condition, the normal bundle of Z need not be trivial. +Given (M, Z) a b-manifold, [GMP10] shows that there exists a vector bundle, +denoted by bT M whose smooth sections are b-vector fields. This bundle is called +the b-tangent bundle of (M, Z). +The b-cotangent bundle bT ∗M is defined using duality. A b-form is a section +of the b-cotangent bundle. Around a point p ∈ Z the C∞(M)-module of these +sections has the following local basis: +(2.3) +{ 1 +xdx, dx1, . . . , dxn−1}. +In the same way we define a b-form of degree k to be a section of the bundle +�k(bT ∗M), the set of these forms is denoted bΩk(M). Denoting by f the distance +function4 to the critical hypersurface Z, we may write the following decomposition +as in [GMP10] for any ω ∈b Ωk(M) : +(2.4) +ω = α ∧ df +f + β, with α ∈ Ωk−1(M) and β ∈ Ωk(M). +This decomposition allows to extend the differential of the de Rham complex +d to bΩ(M) by setting dω = dα ∧ df +f + dβ. +Degree 0 functions are called b-functions and and near Z can be written as +c log |x| + g, +where c ∈ R, g ∈ C∞, and x is a local defining function. +The associated cohomology is called b-cohomology and it is denoted by bH∗(M). +Definition 2.10. A b-symplectic form on a b-manifold (M 2n, Z) is defined +as a non-degenerate closed b-form of degree 2 (i.e., ωp is of maximal rank as an +element of Λ2( bT ∗ +p M) for all p ∈ M). +The notion of b-symplectic forms is dual to the notion of b-Poisson structures. +The advantage of using forms rather than bivector fields is that symplectic tools +can be ‘easily’ exported. +4Originally in [GMP10] f stands for a global function, but for non-orientable manifolds we +may use the distance function instead. + +10 +2. A PRIMER ON SINGULAR SYMPLECTIC MANIFOLDS +Radko’s classification theorem [Rad02] can be translated into this language. +This translation was already formulated in [GMP10]: +Theorem 2.11 (Radko’s theorem in b-cohomological language, [GMP14]). +Let S be a closed orientable surface and let ω0 and ω1 be two b-symplectic forms on +(S, Z) defining the same b-cohomology class (i.e.,[ω0] = [ω1]). Then there exists a +diffeomorphism φ : S → S such that φ∗ω1 = ω0. +2.2. On bm-Symplectic manifolds +2.2.1. Basic definitions. By relaxing the transversality condition allowing +higher order singularities ([Arn89] and [AA81]) we may consider other symplectic +structures with singularities as done by Scott [Sco16] with bm-symplectic struc- +tures. +Let m be a positive integer a bm-manifold is a b-manifold (M, Z) together +with a bm-tangent bundle attached to it. The bm-tangent bundle is (by Serre-Swan +theorem [Swa62]) a vector bundle, bmT M whose sections are given by, +Γ(bmT M) = {v ∈ Γ(T M) : v(x) +vanishes to order m at Z}, +where x is a defining function for the critical set Z in a neighborhood of each +connected component of Z and can be defined as x : M \ Z → (0, ∞), x ∈ C∞(M) +such that: +• x(p) = d(p) a distance function from p to Z for p : d(p) ≤ 1/2 +• x(p) = 1 on M \ {p ∈ M such that d(p) < 1}.5 +(This definition of x allows us to extend the construction in [Sco16] to the non- +orientable case as in [MOT14].) We may define the notion of a bm-map as a map +in this category (see [Sco16]). +The sections of this bundle are referred to as bm-vector fields and their flows +define bm-maps. In local coordinates, the sections of the bm-tangent bundle are +generated by: +(2.5) +{xm ∂ +∂x, ∂ +∂x1 +, . . . , +∂ +∂xn−1 +}. +Proceeding mutatis mutandis as in the b-case one defines the bm-cotangent +bundle (bmT ∗M), the bm-de Rham complex and the bm-symplectic structures. +A Laurent Series of a closed bm-form ω is a decomposition of ω in a tubular +neighborhood U of Z of the form +(2.6) +ω = dx +xm ∧ ( +m−1 +� +i=0 +π∗(αi)xi) + β +with π : U → Z the projection of the tubular neighborhood onto Z, αi a closed +smooth de Rham form on Z and β a de Rham form on M. +In [Sco16] it is proved that in a neighborhood of Z, every closed bm-form ω +can be written in a Laurent form of type (2.6) having fixed a (semi)local defining +function. +bm-Cohomology is related to de Rham cohomology via the following theorem: +5Then a bm-manifold will be a triple (M, Z, x), but for the sake of simplicity we refer to it +as a pair (M, Z) and we tacitly assume that the function x is fixed. + +2.2. ON bm-SYMPLECTIC MANIFOLDS +11 +Theorem 2.12 (bm-Mazzeo-Melrose, [Sco16]). Let (M, Z) be a bm-manifold, +then: +(2.7) +bmHp(M) ∼= Hp(M) ⊕ (Hp−1(Z))m. +The isomorphism constructed in the proof of the theorem above is non-canonical +(see [Sco16]). +The Moser path method can be generalized to bm-symplectic structures (see +[MS21] for the generalization from surfaces in [Sco16] to general manifolds): +Theorem 2.13 (Moser path method). Let ωt be a path of bm-symplectic +forms defining the same bm-cohomology class [ωt] on (M 2n, Z) with M 2n closed +and orientable then there exist a bm-symplectomorphism ϕ : (M 2n, Z) −→ (M 2n, Z) +such that ϕ∗(ω1) = ω0. +An outstanding consequence of Moser path method is a global classification +of closed orientable bm-symplectic surfaces `a la Radko in terms of bm-cohomology +classes. +Theorem 2.14 (Classification of closed orientable bm-surfaces, [Sco16]). +Let ω0 and ω1 be two bm-symplectic forms on a closed orientable connected bm- +surface (S, Z). Then, the following conditions are equivalent: +• their bm-cohomology classes coincide [ω0] = [ω1], +• the surfaces are globally bm-symplectomorphic, +• the Liouville volumes of ω0 and ω1 and the numbers +� +γ +αi +for all connected components γ ⊆ Z and all 1 ≤ i ≤ m coincide (where +αi are the one-forms appearing in the Laurent decomposition of the two +bm-forms of degree 2, ω0 and ω1). +Definition 2.15. The numbers [αi] = +� +γ αi are called modular weights for the +connected components γ ⊂ Z. +A relative version of Moser’s path method is proved in [GMW17]. As a corol- +lary we obtain the following local description of a bm-symplectic manifold: +Theorem 2.16 (bm-Darboux theorem, [GMW17]). Let ω be a bm-symplectic +form on (M, Z) and p ∈ Z. Then we can find a coordinate chart (U, x1, y1, . . . , xn, yn) +centered at p such that on U the hypersurface Z is locally defined by x1 = 0 and +ω = dx1 +xm +1 +∧ dy1 + +n +� +i=2 +dxi ∧ dyi. +Remark 2.17. For the sake of simplicity sometimes we will omit any explicit +reference to the critical set Z and we will talk directly about bm-symplectic struc- +tures on manifolds M implicitly assuming that Z is the vanishing locus of Πn where +Π is the Poisson vector field dual to the bm-symplectic form. +Next, we present two lemmas that allow us to talk about bm-symplectic struc- +tures and bm-Poisson as two different presentations of the same geometrical struc- +ture on a b-manifold. The lemma below shows that they are dual to each other +and, thus, in one-to-one correspondence. + +12 +2. A PRIMER ON SINGULAR SYMPLECTIC MANIFOLDS +Lemma 2.18. Let ω be a bm-symplectic and Π its dual vector field, then Π is a +bm-Poisson structure. +Proof. The quickest way to do this is to take the inverse, which is a bivector +field, and observe that it is a Poisson structure (because dω = 0 implies [Π, Π] = 0). +To see that it is bm-Poisson it is enough to check it locally for any point along the +critical set. Take a point p on the critical set Z and apply the bm-Darboux theorem +to get ω = dx1/xm +1 ∧ dy1 + � +i>1 dxi ∧ dyi This means that in the new coordinate +system +Π = xm +1 +∂ +∂x1 +∧ +∂ +∂y1 ++ +� +i>1 +∂ +∂xi +∧ ∂ +∂yi +and thus Π is a bm-Poisson structure. +□ +Conversely, +Lemma 2.19. Let Π be bm-Poisson and ω its dual vector field, then ω is a +bm-symplectic structure. +Proof. If Π transverse `a la Thom on Z with singularity of order m then +because of Weinstein’s splitting theorem we can locally write +Π = xm +1 +∂ +∂x1 +∧ +∂ +∂y1 ++ +� +i>1 +∂ +∂xi +∧ ∂ +∂yi +now its inverse is ω = dx1/xm +1 ∧ dy1 + � +i>1 dxi ∧ dyi which is a bm-symplectic +form. +□ +Hence we have a correspondence from bm-symplectic structures to bm-Poisson +structures. +2.3. Desingularizing bm-Poisson manifolds +In [GMW17] Guillemin, Miranda and Weitsman presented a desingularization +procedure for bm-symplectic manifolds proving that we may associate a family of +folded symplectic or symplectic forms to a given bm-symplectic structure depending +on the parity of m. Namely, +Theorem 2.20 (Guillemin-Miranda-Weitsman, [GMW17]). Let ω be a +bm-symplectic structure on a closed orientable manifold M and let Z be its critical +hypersurface. +• If m = 2k, there exists a family of symplectic forms ωǫ which coincide with +the bm-symplectic form ω outside an ǫ-neighborhood of Z and for which +the family of bivector fields (ωǫ)−1 converges in the C2k−1-topology to the +Poisson structure ω−1 as ǫ → 0 . +• If m = 2k + 1, there exists a family of folded symplectic forms ωǫ which +coincide with the bm-symplectic form ω outside an ǫ-neighborhood of Z. +As a consequence of Theorem 2.20, any closed orientable manifold that supports +a b2k-symplectic structure necessarily supports a symplectic structure. +In [GMW17] explicit formulae are given for even and odd cases. Let us refer +here to the even-dimensional case as these formulae will be used later on. + +2.3. DESINGULARIZING BM-POISSON MANIFOLDS +13 +Let us briefly recall how the desingularization is defined and the main result in +[GMW17]. Recall that we can express the b2k-form as: +(2.8) +ω = dx +x2k ∧ +�2k−1 +� +i=0 +xiαi +� ++ β. +This expression holds on a ǫ-tubular neighborhood of a given connected com- +ponent of Z. This expression comes directly from equation 2.6, to see a proof of +this result we refer to [Sco16]. +Definition 2.21. Let (S, Z, x), be a b2k-manifold, where S is a closed orientable +manifold and let ω be a b2k-symplectic form. Consider the decomposition given by +the expression (2.8) on an ǫ-tubular neighborhood Uǫ of a connected component of +Z. +Let f ∈ C∞(R) be an odd smooth function satisfying f ′(x) > 0 for all x ∈ [−1, 1] +and satisfying outside that +(2.9) +f(x) = +� +−1 +(2k−1)x2k−1 − 2 +for +x < −1, +−1 +(2k−1)x2k−1 + 2 +for +x > 1. +Let fǫ(x) be defined as ǫ−(2k−1)f(x/ǫ). +The fǫ-desingularization ωǫ is a form that is defined on Uǫ by the following +expression: +ωǫ = dfǫ ∧ +�2k−1 +� +i=0 +xiαi +� ++ β. +This desingularization procedure is also known as deblogging in the literature. +Remark 2.22. Though there are infinitely many choices for f, we will assume +that we choose one, and assume it fixed through the rest of the discussion. +It +would be interesting to discuss the existence of an isotopy of forms under a change +of function f. +Remark 2.23. Because ωǫ can be trivially extended to the whole S in such a +way that it agrees with ω (see [GMW17]) outside a neighborhood of Z, we can +talk about the fǫ-desingularization of ω as a form on S. + + +CHAPTER 3 +A crash course on KAM theory +The last part of this monograph is entirely dedicated to prove a KAM theorem +for bm-symplectic structures and to find applications. So the aim of this section is +to give a quick overview of the traditional KAM theorem. The setting of the KAM +theorem is a symplectic manifold with action-angle coordinates and an integrable +system in it. The theorem says that under small perturbations of the Hamiltonian +”most” of the Liouville tori survive. +Consider Tn×G ⊂ Tn×Rn with action-angle coordinates in it (φ1, . . . , φn, I1, . . . , In) +and the standard symplectic form ω in it. And assume the Hamiltonian function +of the system is given by h(I) a function only depending on the action coordinates. +Then the Hamilton equations of the system are given by +ιXhω = dh +where Xh is the vector field generating the trajectories. Because h does not de- +pend on φ the angular variables the system is really easy to solve, and the equations +are given by +x(t) = (φ(t), I(t)) = (φ0 + ut, I0), +where u = ∂h/∂I is called the frequency vector. These motions for a fixed +initial condition are inside a Liouville torus, and are called quasi-periodic. +The KAM theorem studies what happens to such systems when a small per- +turbation is applied to the Hamiltonian function, i.e. we consider the evolution of +the system given by the Hamiltonian h(I) + R(I, φ), where we think of the term +R(I, φ) as the small perturbation in the system. With this in mind, the Hamilton +equations can be written as +˙φ = u(I) + ∂ +∂I R(I, φ), ˙I = − ∂ +∂φR(I, φ), +Another important concept to have in mind is the concept of rational depen- +dency. A frequency u is rationally dependent if ⟨u, k⟩ = 0 for some k ∈ Zn, if there +exists no k satisfying the condition then the vector u is called rationally indepen- +dent. There is a stronger concept of being rationally independent and that is the +concept of being Diophantine. A vector u is γ,τ-diophantine if ⟨u, k⟩ ≥ +γ +|k|τ +1 for all +k ∈ Zn \ {0}. γ > 0 and τ > n − 1. +The KAM theorem states that the Liouville tori with frequency vector satisfying +the diophantine condition survive under the small perturbation R(I, φ). There are +conditions relating the size of the perturbation with γ and τ. Also, the set of tori +satisfying the Diophantine condition has measure 1 − Cγ for some constant C. +Now we give a proper statement of the theorem as was given in [DG96]. +15 + +16 +3. A CRASH COURSE ON KAM THEORY +Theorem 3.1 (Isoenergetic KAM theorem). Let G ⊂ Rn, n > 2, a compact, +and let H(φ, I) = h(I) + f(φ, I) real analytic on Dρ(G). +Let ω = ∂h/∂I, and +assume the bounds: +���� +∂2h +∂I2 +���� +G,ρ2 +≤ M, +|ω|G ≤ L +and +|ωn(I)| ≥ l∀I ∈ G. +Assume also that ω is µ-isoenergetically non-degenerate on G. For a = 16M/l2, +assume that the map Ω = Ωω,h,a is one-to-one on G, and that its range F = Ω(G) +is a D-set. Let τ > n − 1, γ > 0 and 0 < ν < 1 given, and assume: +ε := ∥f∥G,ρ ≤ ν2l6µ2ˆρ2τ+2 +24τ+32L6M 3 · γ2, +γ ≤ min +�8LMρ2 +νlˆρτ+1 , l +� +, +where we write ρ := min +� +νρ1 +12(τ+2), 1 +� +. Define the set +ˆG = ˆGγ := +� +I ∈ G − 2γ +µ : ω(I)isτ, γ − Diophantine +� +. +Then, there exists a real continuous map T : W ρ1 +4 (Tn) × ˆG → Dρ(G), analytic +with respect to the angular variables, such that: +(1) For every I ∈ ˆG, the set T (Tn × {I}) is an invariant torus of H, its +frequency vector is colinear to ω(I) and its energy is h(I). +(2) Writing +T (φ, I) = (φ + Tφ(φ, I), I + TI(φ, I)), +one has the estimates +|Tφ| ˆ +G,( ρ1 +4 ,0),∞ ≤ 22τ+15L2M +ν2l2ˆρ2τ+1 +ε +γ2 , +|TI| ˆ +G,( ρ1 +4 ,0) ≤ 2τ+16L3M +νl3µˆρτ+1 +ε +γ +(3) meas[(Tn ×G)\T (Tn × ˆG)] ≤ Cγ, where C is a very complicated constant +depending on n, τ, diamF, D, ˆρ, M, L, l, µ. +Remark 3.2. This version of the KAM theorem is the isoenergetic one, this +version ensures that the energy of the Liouville Tori identified by the diffeomorphism +after the perturbation remains the same as before the perturbation. Our version of +the bm-KAM is not isoenergetic for the sake of simplifying the computations. +Also, we should outline that the KAM theorem has already been explored in +singular symplectic manifolds before. In [KMS16a] the authors proved a KAM +theorem for b-symplectic manifolds, for a particular kind of perturbations. +Theorem 3.3 (KAM Theorem for b-Poisson manifolds). Let Tn × Bn +r be en- +dowed with standard coordinates (ϕ, y) and the b-symplectic structure. Consider a +b-function +H = k log |y1| + h(y) +on this manifold, where h is analytic. Let y0 be a point in Bn +r with first component +equal to zero, so that the corresponding level set Tn × {y0} lies inside the critical +hypersurface Z. +Assume that the frequency map +˜ω : Bn +r → Rn−1, +˜ω(y) := ∂h +∂˜y (y) + +3. A CRASH COURSE ON KAM THEORY +17 +has a Diophantine value ˜ω := ˜ω(y0) at y0 ∈ Bn and that it is non-degenerate at y0 +in the sense that the Jacobian ∂˜ω +∂˜y (y0) is regular. +Then the torus Tn × {y0} persists under sufficiently small perturbations of H +which have the form mentioned above, i.e. they are given by ǫP, where ǫ ∈ R and +P ∈b C∞(Tn × Bn +r ) has the form +P(ϕ, y) = k′ log |y1| + f(ϕ, y) +f(ϕ, y) = f1( ˜ϕ, y) + y1f2(ϕ, y) + f3(ϕ1, y1). +More precisely, if |ǫ| is sufficiently small, then the perturbed system +Hǫ = H + ǫP +admits an invariant torus T . +Moreover, there exists a diffeomorphism Tn → T close1 to the identity taking +the flow γt of the perturbed system on T to the linear flow on Tn with frequency +vector +�k + ǫk′ +c +, ˜ω +� +. +1By saying that the diffeomorphism is “ǫ-close to the identity” we mean that, for given H, P +and r, there is a constant C such that ∥ψ − Id∥ < Cǫ. + + +Part 2 +Action-angle coordinates and +cotangent models + +In this part, we consider the semilocal classification for any bm-Poisson manifold +in a neighbourhood of an invariant compact submanifold. The compact subman- +ifolds under consideration are the compact invariant leaves of the distribution D +generated by the Hamiltonian vector fields Xfi of an integrable system. An in- +tegrable system is given by a set of n functions on a 2n-dimensional symplectic +manifold which we can order in a map F = (f1, . . . , fn). Historically, integrable +systems were introduced to actually integrate Hamiltonian systems XH using the +first-integrals fi and, classically, we identify H = f1. It turns out that in the sym- +plectic context the compact regular orbits of the distribution D coincide with the +fibers F −1(F(p)) for any point p on these orbits/fibers. The fact that the orbit +coincides with the connected fiber is part of the magic of symplectic duality. +The same picture is reproduced for singular symplectic manifolds of bm-type +or bm-Poisson manifolds as we will see in this chapter. +The study of action-angle coordinates has interest from this geometrical point +of view of the classification of geometric structures in a neighbourhood of a compact +submanifold of a bm-Poisson manifold. It also has interest from a dynamical point +of view as these compact submanifolds now coincide with invariant subsets of the +Hamiltonian system under consideration. +From a geometric point of view, the existence of action-angle coordinates deter- +mines a unique geometrical model for the bm-Poisson (or bm-symplectic) structure +in a neighbourhood of the invariant set. From a dynamical point of view, the exis- +tence of action-angle coordinates provides a normal form theorem that can be used +to study stability and perturbation problems of the Hamiltonian systems (as we +will see in the last chapter of this monograph). +An important ingredient that makes our action-angle coordinate theorem brand- +new from the symplectic perspective is that the system under consideration is more +general than Hamiltonian, it is bm-Hamiltonian as the first-integrals of the system +can be bm-functions which are not necessarily smooth functions. Dynamically, this +means that we are adding to the set of Hamiltonian invariant vector fields, the +modular vector field of the integrable system. +In contrast to the standard action-angle coordinates for symplectic manifolds, +our action-angle theorem comes with m additional invariants associated with the +modular vector field which can be interpreted in cohomological terms as the pro- +jection of the bm-cohomology class determined by the modular vector field on the +first cohomology group of the critical hypersurface under the Mazzeo-Melrose cor- +respondence. +The strategy of the proof of the action-angle coordinate systems is the search +of a toric action (so this takes us back to the motivation of the use of symmetries +in this monograph). In contrast to the symplectic case, it is not enough that this +action is Hamiltonian as then a direction of the Liouville torus would be missing. +We need the toric action to be bm-Hamiltonian. The structure of this proof looks +like the one in [KMS16a] but encounters serious technical difficulties as in order +to check that the natural action to be considered is bm-Hamiltonian we need to go +deeper inspired by [Sco16] in the relation between the geometry of the modular +vector field and the coefficients of the Taylor series ci of one of the first-integrals. +This allows us to understand new connections between the geometry and analysis +of bm-Poisson structures not explored before. + +21 +Once we prove the existence of this bm-Hamiltonian action the proof looks very +close to the one in [KMS16a]. +In the second chapter of this part we re-state the action-angle theorem as a +cotangent lift theorem with the following mantra: +Every integrable system on a bm-Poisson manifold looks like a bm-cotangent lift +in a neighborhood of a Liouville torus. + + +CHAPTER 4 +An action-angle theorem for bm-symplectic +manifolds +4.1. Basic definitions +4.1.1. On bm-functions. The definition of the analogue of b-functions in the +bm-setting is somewhat delicate. The set of bmC∞(M) needs to be such that for all +the functions f ∈bm C∞(M), its differential df is a b-form, where d is the bm-exterior +differential. Recall that a form in bmΩk(M) can be locally written as +α ∧ dx +xm + β +where α ∈ Ωk−1(M) and β ∈ Ωk(M). Recall also that +d +� +α ∧ dx +xm + β +� += dα ∧ dx +xm + dβ. +We need df to be a well-defined bm-form of degree 1. Let f = g +1 +xk−1 , then +df = dg +1 +xk−1 − g k−1 +xk dx. This from can only be a bm-form if and only if g only +depends on x. If f = g log(x), then dg log(x) + g 1 +xdx, which imposes dg = 0 and +hence g to be constant. +With all this in mind, we make the following definition. +Definition 4.1. The set of bm-functions is defined recursively according to the +formula +bmC∞(M) = x−(m−1)C∞(x) + bm−1C∞(M) +with C∞(x) the set of smooth functions in the defining function x and +bC∞(M) = {g log |x| + h, g ∈ R, h ∈ C∞(M)}. +Remark 4.2. A bmC∞(M)-function can be written as +f = a0 log x + a1 +1 +x + . . . + am−1 +1 +xm−1 + h +where ai, h ∈ C∞(M). +Remark 4.3. From this chapter on we are only considering bm-manifolds +(M, x, Z) with x defined up to order m. +I.e. +we can think of x as a jet of a +function that coincides up to order m to some defining function. This is the orig- +inal viewpoint of Scott in [Sco16] which we adopt from now on. The difference +with respect to the other chapters is that we do not fix an specific function. +Definition 4.4. We say that two bm-integrable systems F1, F2 are equivalent +if there exists ϕ, a bm-symplectomorphism, i.e. a diffeomorphism preserving both +ω and the critical set Z (“up to order m”1), such that ϕ ◦ F1 = F2. +1I.e. it preserves the jet x +23 + +24 +4. AN ACTION-ANGLE THEOREM FOR bm-SYMPLECTIC MANIFOLDS +Remark 4.5. The Hamiltonian vector field associated to a bm-function f is a +smooth vector field. Let us compute it locally using the bm-Darboux theorem: +Π = xm +1 +∂ +∂x1 +∧ +∂ +∂y1 ++ +m +� +i=2 +∂ +∂xi +∧ ∂ +∂yi +and f = a0 log x1 + +m−1 +� +i=1 +ai +1 +xi +1 ++ h. +Then if we compute +df += +c1 +���� +a0 +1 +x1 ++ +m−1 +� +i=1 +ci +� +�� +� +(a′ +i − (i − 1)ai−1) 1 +xi +1 +dx1 +− +cm +� +�� +� +(m − 1)am−1 +1 +xm +1 dx1 + dh += +m +� +i=1 +ci +xi +1 +dx1 + dh. +Then, +(4.1) +Xf = Π(df, ·) = +m +� +i=1 +cixm−i +1 +∂ +∂y1 ++ Π(dh, ·), +we obtain a smooth vector field. +4.2. On bm-integrable systems +In this section we present the definition of a bm-integrable system as well as +some observations about these objects. +Definition 4.6. Let (M 2n, Z, x) be a bm-manifold, and let Π be a bm-Poisson +structure on it. F = (f1, . . . , fn)2 is a bm-integrable system3 if: +(1) df1, . . . , dfn are independent on a dense subset of M and in all the points +of Z where independent means that the form df1 ∧ . . . ∧ dfn is non-zero as +a section of Λn(bmT ∗(M)), +(2) the functions f1, . . . , fn Poisson commute pairwise. +Definition 4.7. The points of M where df1, . . . , dfn are independent are called +regular points. +The next remarks will lead us to a normal form for the first function f1. +Remark 4.8. Note that df1, . . . , dfn are independent on a point if and only if +Xf1, . . . , Xfn are independent at that point. This is because the map +bmT M →bm T ∗M : u �→ ωp(u, ·) +is an isomorphism. +Remark 4.9. The condition of df1, . . . , dfn being independent must be under- +stood as df1 ∧ . . . ∧ dfn being a non-zero section of �n( bmT ∗M). +2fi are bm-functions. +3In this monograph we only consider integrable systems of maximal rank n. + +4.2. ON bm-INTEGRABLE SYSTEMS +25 +Remark 4.10. By remark 4.8 the vector fields Xf1, . . . , Xfn have to be in- +dependent. +This implies that one of the f1, . . . , fn has to be a singular (non- +smooth) bm-function with a singularity of maximal degree. +If we write fi = +c0,i log(x1) + �m−1 +j=1 +cj,i +xj +1 + ˜f1 +Xfi = +m +� +j=1 +xm−j +1 +ˆcj,i +∂ +∂y1 ++ X ˜ +fi +where ˆcj,i(x) = d(cj,i) +dx +− (j − 1)cj−1,i. If there is no bm-function with a singularity +of maximum degree all the terms in the ∂/∂y1 direction become 0 at Z. And hence +Xf1, . . . , Xfn cannot have maximal rank at Z. +Lemma 4.11. Let F = (f1, . . . , fn) a bm-integrable system. If f1 has a singular- +ity of maximal degree, there exists an equivalent integrable system F ′ = (f ′ +1, . . . , f ′ +n) +where f ′ +1 has a singularity of maximal degree and no other f ′ +i has singularity of +any degree. +Proof. Let fi = c0,i log(x1) + +m−1 +� +j=1 +cj,1 +xj +1 +� +�� +� +ζi(x1) ++ ˜fi = ζi(x1) + ˜fi. By remark 4.104, +Xfi = +m +� +i=1 +xm−j +1 +ˆcj,i +� +�� +� +gi(x1) +∂ +∂y1 ++ X ˜ +fi = gi(x1) ∂ +∂y1 ++ X ˜ +fi. +Note that gi(x1) = gi(0) = ˆcm,i at Z. Let us look at the distribution given by the +Hamiltonian vector fields Xfi = gi(x1) ∂ +∂y1 +X ˜ +fi. This distribution is the same that +the one given by: +(4.2) +{Xf1, Xf2 − g2(x1) +g1(x1)Xf1, . . . , Xfn − gn(x1) +g1(x1) Xf1}. +Observe that for i > 1, Xfi − gi(x1) +g1(x1)Xf1 = X ˜ +fi + g2(x1) +g1(x1)X ˜ +f1. Also g1(x1) is different +from 0 close to Z because at Z g1(x1) = ˆcm,1. Since the distribution given by these +vector fields is the same, an integrable system that has Hamiltonian vector fields 4.2 +would be equivalent to F. From the expression 4.2 it is clear that the new vector +fields commute. And it is also true that this new vector fields are Hamiltonian. Let +us take F ′ the set of functions that have as Hamiltonian vector fields 4.2. +□ +From now on we will assume the integrable system to have only one singular +function and this function to be f1. +Remark 4.12. Because we asked Xf1, . . . , Xfn to be linearly independent at +all the points of Z and using the previous remarks cm := cm,1 ̸= 0 at all the points +of Z. +4Here have used the bm-Darboux theorem to do the computations. + +26 +4. AN ACTION-ANGLE THEOREM FOR bm-SYMPLECTIC MANIFOLDS +Furthermore, we can assume f1 to have a smooth part equal to zero as sub- +tracting the smooth part of f1 to all the functions gives an equivalent system. Also, +we can assume that cm is 1 because dividing all the functions of the bm-integrable +system by cm also gives us an equivalent system. +As a summary, we can assume f1 = a0 log(x)+a11/x+. . .+am−21/xm−2+ +1/xm−1 and f2, . . . , fn to be smooth, a0 ∈ R and a1, . . . , am−2 ∈ C∞(x). +Also we are going to state lemma 3.2 in [GMPS17], because we are going to +use it later in this section. The result states that if we have a toric action on a bm- +symplectic manifold (which we will prove in a neighbourhood of a Liouville torus), +then we can assume the coefficients a2, . . . , am−2 to be constants. More precisely +Lemma 4.13. There exists a neighborhood of the critical set U = Z × (−ε, ε) +where the moment map µ : M → t∗ is given by +µ = a1 log |x| + +m +� +i=2 +ai +x−(i−1) +i − 1 ++ µ0 +with ai ∈ t0 +L and µ0 is the moment map for the TL-action on the symplectic leaves +of the foliation. +4.3. Examples of bm-integrable systems +The following example illustrates why it is necessary to use the definition of bm- +function as considered above. There are natural examples of changes of coordinates +in standard integrable systems on symplectic manifolds that yield bm-symplectic +manifolds but do not give well-defined bm-integrable systems. +Example 4.14. Consider a time change in the two body problem, to obtain +a b2-integrable system. In the classical approach to solve the 2-body problem the +following two conserved quantities are obtained: +f1 += +µy2 +2 + +l2 +2µr2 − k +r , +f2 += +l, +with symplectic form ω = dr ∧ dy + dl ∧ dα, where r is the distance between the +two masses and l is the angular momentum. We also know that l is constant along +the trajectories. Because l is a constant of the movement, we can do a symplectic +reduction on its level sets. The form on the symplectic reduction becomes dr ∧ dy. +To simplify the notation, we will use x instead of r. Then ω = dx ∧ dy. With +hamiltonian function given by f = µ +2 y2 + +l +2µ +1 +x2 − k 1 +x. Hence, the equations are: +˙x += +∂f +∂y , +˙y += +− ∂f +∂x. +Doing a time change τ = x3t then dx +dτ = +1 +x3 dx +dt . With this time coordinate, the +equations become: +˙x += +1 +x3 +∂f +∂y , +˙y += +− 1 +x3 +∂f +∂x. +These equations can be viewed as the motion equations given by a b3-symplectic +form ω = +1 +x3 dx ∧ dy. +Let us check that this is actually a bm-integrable system. + +4.3. EXAMPLES OF bm-INTEGRABLE SYSTEMS +27 +• All the functions Poisson commute is immediate because we only have +one. +• df = µydy +( k +x2 − l +µ +1 +x3 )dx is a b3-form because the term with dx does not +depend on y. +• All the functions are independent, this is true because df does not vanish +as a b3-form. +Example 4.15. In the paper [Mar19] the author builds an action of SL(2, R) +over (P, ωP ) where P = {ξ ∈ C|i(¯ξ − ξ) > 0} is the complex semi-plane, with +moment map JP (ξ) = +R +ξim ((|ξ|2 + 1), 2ξr, ±(|ξ|2 + 1)), where the ± sign depends +on the choice of the hemisphere projected by the stereographic projection. From +now on we will take the sign +. Also the symplectic form ωP has the following +expression: +ωP = ± R +ξ2 +im +dξr ∧ dξim +In order to simplify the notation we identify P with the real half-plane P = +{x, y ∈ R2|y > 0}. With this identification, the moment map becomes Jp(x, y) = +R +y (x2 + y2 + 1, 2x, x2 + y2 + 1). Obviously, this moment map does not give an +integrable system. The symplectic form writes as: +ωP = R +y2 dy ∧ dx. +This form can be viewed as a b2-form if we extend P including the line {y = 0} +as its singular set. +Let us consider only one of the components of JP as bm- +function and let us see if it gives a bm-integrable system. First we will try with +f1 = R +y (x2 + y2 + 1) and then f2 = R +y (2x). +(1) f1 = R +y (x2 + y2 + 1) We have to check three things to see if this gives a +b2-integrable system. +(a) All the functions Poisson commute is immediate because we only +have one. +(b) All the functions are bm-functions. This point does not hold because +df1 = R +y2 (2xydx + (y2 − x2 − 1)dy) and the first component makes no +sense as a section of Λ1(b2T ∗M). +(c) All the functions are independent. In this case, we need to check that +df1 does not vanish, but since it is not a bm-form it makes no sense +to be a non-zero section of Λ1(b2T ∗M). +(2) f2 = R +y (2x) +(a) Same as before. +(b) All the functions are bm-functions. This point does not hold because +df2 = 2R +y dx − 2Rx +y2 dy and the first component makes no sense as a +section of Λ1(b2T ∗M). +(c) Same as before. +Example 4.16. Toric actions give natural examples of integrable systems where +the component functions are given by the moment map. In the case of surfaces: +S1-actions on surfaces give natural examples of bm-integrable systems. Only torus +and spheres admit circle actions. + +28 +4. AN ACTION-ANGLE THEOREM FOR bm-SYMPLECTIC MANIFOLDS +In the picture below two integrable systems on the 2-sphere depending on the +degree m. On the right the image of the moment map that defines the integrable +system. The action is by rotations along the central axis. +Namely consider the sphere S2 as a bm-symplectic manifold having as critical +set the equator: +(S2, Z = {h = 0}, ω = dh +hm ∧ dθ), +with h ∈ [−1, 1] and θ ∈ [0, 2π). +For m = 1: The computation ι ∂ +∂θ ω = − dh +h = −d(log |h|), tells us that the +function µ(h, θ) = log |h| is the moment map and defines a b-integrable system. +For higher values of m: ι ∂ +∂θ ω = − dh +hm = −d(− +1 +(m−1)hm−1 ), and the moment +map is µ(h, θ) = − +1 +(m−1)hm−1 which defines a bm-integrable system. +µ, m = 1 +µ, m = 2 +Figure 1. Integrable systems associated to the moment map of +an S1-action by rotations on a bm-symplectic 2-sphere S2. +Example 4.17. Consider now as b2-symplectic manifold the 2-torus +(T2, Z = {θ1 ∈ {0, π}}, ω = +dθ1 +sin2 θ1 +∧ dθ2) +with standard coordinates: θ1, θ2 ∈ [0, 2π). Observe that the critical hypersurface +Z in this example is not connected. It is the union of two disjoint circles. Consider +the circle action of rotation on the θ2-coordinate with fundamental vector field +∂ +∂θ2 . +As the following computation holds, +ι +∂ +∂θ2 ω = − dθ1 +sin2 θ1 += d +�cos θ1 +sin θ1 +� +. +The fundamental vector field of the S1-action defines b2C∞-integrable system given +by the function − cos θ1 +sin θ1 . +Example 4.18. The former example can be made general to produce examples +of bm-integrable systems on a bm-symplectic manifold for any integer m +(T2, Z = {θ1 ∈ {0, π}}, ω = +dθ1 +sinm θ1 +∧ dθ2). +Then +ι +∂ +∂θ2 ω = − +dθ1 +sinm θ1 += d +� +| cos θ1| +cos θ1 +2F1 +� 1 +2, 1−m +2 ; 3−m +2 ; sin2(θ1) +� +(1 − m) sinm−1 θ1 +� +, +with 2F1 the hypergeometric function. + +4.4. LOOKING FOR A TORIC ACTION +29 +µ +Figure 2. Integrable system given by an S1-action on a b2-torus +T2 and its associated moment map. +Thus, the associated S1-action has as bmC∞-Hamiltonian the function +−| cos θ1| +cos θ1 +2F1 +� 1 +2, 1−m +2 ; 3−m +2 ; sin2(θ1) +� +(1 − m) sinm−1 θ1 +which defines a bm-integrable system. +Now we give a couple of examples of bm-integrable systems. +Example 4.19. This example uses the product of bm-integrable systems on a +bm-symplectic manifold with an integrable system on a symplectic manifold. Given +(M 2n1 +1 +, Z, x, ω1) a bm-symplectic manifold with f1, . . . , fn1 a bm-integrable system +and (M 2n2 +2 +, ω2) a symplectic manifold with g1, . . . , gn2 an integrable system. Then +(M1×M2, Z×M2, x, ω1+ω2) is a bm-symplectic manifold and (f1, . . . , fn1, g1, . . . , gn2) +is a bm-integrable system on the higher dimensional manifold. +In particular by combining the former examples of bm-integrable systems on +surfaces and arbitrary integrable systems on symplectic manifolds we obtain exam- +ples of bm-integrable systems in any dimension. +Example 4.20. (From integrable systems on cosymplectic manifolds +to bm-integrable systems:) +Using the extension theorem (Theorem 50) of [GMP14] we can extend any +integrable system (f2, . . . , fn) to an integrable system in a neighbourhood of a +cosymplectic manifold (Z, α, ω) by just adding a bm-function f1 to the integrable +system so that the new integrable system is (f1, f2, . . . , fn) and considering the +associated bm-symplectic form: +(4.3) +˜ω = p∗α ∧ dt +tm + p∗ω. +(t is the defining function of Z). +4.4. Looking for a toric action +In this section we pursue the proof of action-angle coordinates for bm-integrable +systems by recovering a torus group action. This action is associated to the Hamil- +tonian vector fields associated to Xfi. +This is the same strategy used for b-integrable systems in [KMS16a]- One of +the main difficulties is to prove that the coefficients a1, . . . , an can be considered + +30 +4. AN ACTION-ANGLE THEOREM FOR bm-SYMPLECTIC MANIFOLDS +as constant functions. This makes it more difficult to prove the existence of a Tn- +action in the general bm-case than in the b-case, but once we have it we can use +the results in [GMW17] to assume that the coefficients a1, . . . , an are constant +functions. +In this section we provide some preliminary material that will be needed later: +Proposition 4.21. Let (M, Z, x, ω) be a bm-symplectic manifold such that Z +is connected with modular period k. Let π : Z → S1 ≃ R/kZ be the projection +to the base of the corresponding mapping torus. Let γ : S1 = R/kZ → Z be any +loop such that π ◦ γ is positively oriented and has constant velocity 1. Then the +following are equal: +(1) The modular period of Z, +(2) +� +γ ιLω, +(3) The value am−1 for any bmC∞(M)-function +f = a0 log(x) + +m−1 +� +j=1 +aj +1 +xj + h +such that the hamiltonian vector field Xf has 1-periodic orbits homotopic +in Z to some γ. +Proof. Let us first prove that (1)=(2) and then that (2)=(3). +(1)=(2) Let us denote by Vmod the modular vector field. Recall from [GMW17] +that ιL(Vmod) is the constant function 1. Let s : [0, k] → Z be the trajec- +tory of the modular vector field. Because the modular period is k, s(0) +and s(k) are in the same leaf L. Let ˆs : [0, k + 1] → Z a smooth extension +of s such that s|[k,k+1] is a path in L joining ˆs(k) = s(k) to ˆs(k+1) = s(0). +This way ˆs becomes a loop. Then, +k = +� k +0 +1dt = +� +S +ιLω = +� +ˆs +ιLω = +� +γ +ιLω +(2)=(3) Let r : [0, 1] �→ Z be the trajectory of Xf the hamiltonian vector field of +f. Recall that Xf satisfies +ιXf ω = +m +� +j=1 +cj +dx +xi + dh. +Let xm ∂ +∂x be a generator of the linear normal bundle L. We know that +Xf is 1-periodic and its trajectory is homotopic to γ. Hence, +k = +� +r ιLω += +� 1 +0 +ιxm ∂ +∂x ω(Xf|r(t))dt += +� 1 +0 +−( +m +� +j=1 +ci +dx +xi + dh) · (xm ∂ +∂x)|r(t)dt += +−cm = −am−1 +□ + +4.4. LOOKING FOR A TORIC ACTION +31 +We will also need a Darboux-Carath´eodory theorem for bm-symplectic mani- +folds: +Theorem 4.22 (Darboux-Carath´eodory (bm-version)). Let +(M 2n, x, Z, ω) +be a bm-symplectic manifold and m be a point on Z. Let f1, . . . , fn be a bm-integrable +system. Then there exist bm-functions (q1, . . . , qn) around m such that +ω = +n +� +i=1 +dfi ∧ dqi +and the vector fields {Xfi, Xqj}i,j commute. If f1 is not smooth (recall that f1 = +a0 log(x) + �m−1 +j=1 aj 1 +xi with an ̸= 0 on Z and a0 ∈ R) the qi can be chosen to be +smooth functions, and (x, f2, . . . , fn, q1, . . . , qn) is a system of local coordinates. +Proof. The first part of this proof is exactly as in [KMS16a]. Assume now +f1 = a0 log(x) + +m−1 +� +j=1 +aj +1 +xi . We modify the induction requiring also that µi (in +addition to be in Ki) is also in T ∗M ⊆b T ∗M. We can also ask this extra condition +while asking µi(Xfi) = 1, we only have to check that Xfi does not vanish in T M. +This is clear because Xfi does not vanish at bT M and +0 = {fn, fi} = +� m +� +i=1 +˜ai +dx +xi +� +(Xfi) = +� +dx +xm +m +� +i=1 +aixi +� +(Xfi). +All the terms in the last expression vanish except for the one of degree m. +Then dx/xm is in the kernel of Xfi, hence Xfi does not vanish on T M and the +qi can be chosen to be smooth. +{Xx, Xf2, . . . , Xfn, Xq1, . . . Xqn} commute because {Xfi, Xqi}i,j commute. Then +dx ∧ df2 . . . ∧ dfn ∧ dq1 ∧ . . . ∧ dqn +is a non-zero section of �n(bT M). And hence +(x, f2, . . . , fn−1, q1, . . . , qn) +are local coordinates. +□ +Before proceeding with the proof of the action-angle coordinates, we need to +prove that in a neighbourhood of a Liouville torus the fibration is semilocally trivial: +Lemma 4.23 (Topological Lemma). Let m ∈ Z be a regular point of a bm- +integrable system (M, x, Z, ω, F). Assume that the integral manifold Fm through m +is compact. Then there exists a neighborhood U of Fm and a diffeomorphism +φ : U ≃ Tn × Bn +which takes the foliation F to the trivial foliation {Tn × {b}}b∈Bn. +Proof. We follow the steps of [LGMV08]. In this case, the only extra step +that must be checked is that the foliation given by the bm-hamiltonian vector fields +of F = (f1, f2, . . . , fn) is the same as the one given by the level sets of ˜F := +(x, f2, . . . , fn). In our case f1 = a0 log(x) + �m−1 +u=1 ai 1 +xi , where a0 ∈ R, ai ∈ C∞(x), +am−1 = 1. Hence the foliations are the same. Then as in [LGMV08], we take an + +32 +4. AN ACTION-ANGLE THEOREM FOR bm-SYMPLECTIC MANIFOLDS +Figure 4. Fibration by Liouville tori: The middle fiber of the +point p ∈ Z in magenta, the neighbouring Liouville tori in blue. +arbitrary Riemannian metric on M and this defines a canonical projection ψ : U → +Fm. Let us define φ := ψ × ˜F. We obtain the commutative diagram (Figure 3). +U +Tn × Bn +Bn +φ +˜ +F +p +Figure 3. Commutative diagram of the construction of the iso- +morphism of bm-integrable systems. +which provides the necessary equivalence of bm-integrable systems. +□ +4.5. Action-angle coordinates on bm-symplectic +manifolds +In a neighbourhood of one of our Liouville tori all we can assume about the +form of our bm-symplectic structure is that is given by the Laurent series defined +in [Sco16]. +That is to say, we can assume that in a tubular neighborhood U of Z +ω = +m−1 +� +j=1 +dx +xi ∧ π∗(αi) + β, +where π : U → Z is the projection of the tubular neighborhood onto Z, αi are +closed smooth de Rham forms on Z and β a de Rham form on M of degree 2. +In [RBM, MM22] normal forms are given for group actions in a neighbour- +hood of the orbit. Below we provide a normal for the integrable system in a neigh- +bourhood of an orbit of the torus action associated to the integrable system. This +theorem is finer than the bm-symplectic slice theorem provided in [MM22] as it +also gives information about the first integrals. +One of the non-trivial steps of the proof is to associate a toric action to the +integrable system. The connection to normal forms of group actions will become +even more evident when we discuss the associated cotangent models. + +4.5. ACTION-ANGLE COORDINATES ON bm-SYMPLECTIC +MANIFOLDS +33 +Theorem A (Action-angle coordinates for bm-symplectic manifolds). Let (M, x, ω, F) +be a bm-integrable system, where F = (f1 = a0 log(x) + �m−1 +j=1 aj 1 +xj , . . . , fn) with +aj for j > 1 functions in x. Let m ∈ Z be a regular point and let us assume that the +integral manifold of the distribution generated by the Xfi through m is compact. +Let Fm be the Liouville torus through m. Then, there exists a neighborhood U of +Fm and coordinates (θ1, . . . , θn, σ1, . . . , σn) : U → Tn × Bn such that: +(1) We can find an equivalent integrable system F = (f1 = a′ +0 log(x) + +�m−1 +j=1 a′ +j +1 +xj , . . . , fn) such that the coefficients a′ +0, . . . , a′ +m−1 of f1 are con- +stants ∈ R, +(2) +ω|U = + + +m +� +j=1 +c′ +j +c +σj +1 +dσ1 ∧ dθ1 + + + +n +� +i=2 +dσi ∧ dθi +where c is the modular period and c′ +j = −(j − 1)a′ +j−1, also +(3) the coordinates σ1, . . . , σn depend only on f1, . . . fn. +Proof. The idea of this proof is to construct an equivalent bm-integrable +system whose fundamental vector fields define a Tn-action on a neighborhood of +Tn × {0}. It is clear that all the vector fields Xf1, . . . , Xfn define a torus action on +each Liouville tori Tn × {b} where b ∈ Bn, but this does not guarantee that their +flow defines a toric action on all Tn × Bn. The proof is structured in three steps. +The first one is the uniformization of the periods, i.e. we define an Rn-action on a +neighborhood of Tn × {0} such that the lattice defined by its kernel at every point +is constant. This allows to induce an actual action of a torus (as the periods are +constant) of rank n: A Tn action by taking quotients. The second step consists +in checking that this action is actually bm-Hamiltonian. And in the final step we +apply theorem 4.22 to obtain the expression of ω. +(1) Uniformization of periods. +Let Φs +XF be defined as the joint flow by the Hamiltonian vector fields +of the action: +(4.4) +Φ : Rn × (Tn × Bn) +→ +(Tn × Bn) +((s1, . . . , sn), (x, b)) +�→ +Φs1 +Xf1 ◦ · · · ◦ Φsn +Xfn ((x, b)) +this defines an Rn-action on Tn×Bn. For each b ∈ Bn at a single orbit +Tn × {b} the kernel of this action is a discrete subgroup of Rn. We will +denote the lattice given by this kernel Λb. Because the orbit is compact, +the rank of Λb is maximal i.e. n. This lattice is known as the period lattice +of Tn × {b} as we know by standard arguments in group theory that the +lattice has to be of maximal rank so as to have a torus as a quotient. In +general we can not assume that Λb does not depend on b. The process of +uniformization of the periods modifies the action 4.4 in such a way that +Λb = Zn for all b. Let us consider the following Hamiltonian vector field +�n +i=1 kiXfi. The bm-function that generates this Hamiltonian vector field +is: +k1 + +a0 log(x) + +m−1 +� +j=1 +aj +1 +xj + + + +n +� +i=2 +kifi + +34 +4. AN ACTION-ANGLE THEOREM FOR bm-SYMPLECTIC MANIFOLDS +where recall that am−1 is constant equal 1. Observe that the coefficient +multiplying 1/xm−1 is k1. By proposition 4.21 k1 = c the modular period. +In this case c = [αm]. +Hence, for b ∈ Bn−1 × {0} the lattice Λb is contained in Rn−1 × cZ ⊆ +Rn. Pick (λ1, . . . , λn) : Bn → Rn such that: +• (λ1(b), . . . , λn(b)) is a basis of Λb for all b ∈ Bn, +• λn +i vanishes along Bn−1 × {0} at order m for i < n and λi is equal +to c along Bn−1 × {0}. +In the previous points, λj +i denotes the j-th component of λi. The first +condition can be satisfied by using the implicit function theorem. That +is because Φ(λ, m) = m is regular with respect to the s coordinates. The +second condition is automatically true because Λb ⊆ Rn−1×cZ. We define +the uniformed flow as: +(4.5) +˜Φ : Rn × (Tn × Bn) +→ +(Tn × Bn) +((s1, . . . , sn), (x, b)) +�→ +Φ(�n +i=1 siλi, (x, b)) +(2) The Tn-action is bm-Hamiltonian. The objective of this step is to find +bm-functions σ1, . . . , σn such that Xσi are the fundamental vector fields +of the Tn-action Yi = �n +j=1 λj +iXfj. +By using the Cartan formula for a bm-symplectic form, we obtain: +LYiLYiω += +LYi(d(ιYiω) + ιYidω) += +LYi(d(− �n +j=1 λj +idfi)) += +−LYi(�n +j=1 dλj +i ∧ dfj) = 0 +Note that λj +i are constant on the level sets of F as Φ(λ, m) = m and +the level sets of F are invariant by Φ. +Recall that if Y is a complete periodic vector field and P is a bivector +such that LY LY P = 0, then LY P = 0. So, the vector fields Yi are Poisson +vector fields. To show that each ιYiω has a bmC∞ primitive we will see +that [ιYiω] = 0 in the bm-cohomology. +One one hand, if i > 1, ιYiω vanishes at Z. This holds because Yi has +not any component ∂/∂Y . +Recall Proposition 6 from [GMP14]: +Proposition 4.24. If ω ∈b Ω(M) with ω|Z = 0, then ω ∈ Ω(M). +In a similar way for bm-forms we have, +Proposition 4.25. If ω ∈bm Ω(M) with ω|Z vanishing up to order +m, then ω ∈ Ω(M). +Thus as ιYiω vanishes at Z, the bm-forms ιYiω are indeed smooth. +Thus we can now apply the standard Poincar´e lemma and as these forms +are closed they are locally exact. This proves that all the vector fields Yi +with i > 1 are indeed Hamiltonian. +On the other hand, the fact that ιY1ω = cdf1 is obvious. +Then, because we have a toric action that is Hamiltonian, we can use +lemma 3.2 in [GMPS17], and we get an equivalent system such that ai +are all constant and moreover ⟨a′ +i, X⟩ = αi(Xω). Note that by dividing by +a′ +m−1, we can still assume a′ +m−1 = 1 to be consistent with our notation, +but we then have to multiply f1 · c in the next step. + +4.5. ACTION-ANGLE COORDINATES ON bm-SYMPLECTIC +MANIFOLDS +35 +(3) Apply Darboux-Carath´eodory theorem. +The construction above gives us some candidates σ1 = cf1, σ2, . . . , σn +for the action coordinates. +We now apply the Darboux-Carath´eodory theorem and express the +form in terms of x: +ω = + + +m +� +j=1 +c cj +xj dx ∧ dq1 + + + +n +� +i=2 +dσi ∧ dqi. +Since the vector fields Xσi = +∂ +∂qi are fundamental fields of the Tn- +action the flow 4.5 gives a linear action on the qi coordinates. +Observe that the coordinate system is only defined in U. It may not +be valid at points outside U that may be in the orbit of points in U. Let +us see that the charts can be extended to these points. +Define U′ the union of all tori that intersect U. We will see that the +coordinates are valid at U′. +Let {pi, θj} be the extension of {σi, qj}. It is clear that {pi, θj} = δij +by its construction in the Darboux-Carath´eodory theorem. +To see that {θi, θj} = 0 we take the flows by Xpk and extend the +expression to the whole U′: +Xpk({θi, θj}) = {{θi, θj}, pk} = {θi, δij} − {θj, δjk} = 0. +The fact that ω is preserved is obvious because Xpk are hamiltonian +vector fields and thus they preserve the bm-symplectic forms. Moreover, +t, θ1, p2, θ2, . . . , pn, θn are independent on U′ and hence are a coordinate +system in a neighbourhood of the torus. +□ +Remark 4.26. In the proof we have seen that there exists an equivalent inte- +grable system where the coefficients of the singular function are indeed constant. +From now on, when considering a bm-integrable system we are going to make this +assumption. +Remark 4.27. By means of the desingularization transformation we may ob- +tain an action-angle coordinate theorem for folded manifolds as we do in Part 3 +for the KAM theorem for folded symplectic manifolds. This folded action-angle +theorem is a particular case of the one obtained in [CM22]. + + +CHAPTER 5 +Reformulating the action-angle coordinate via +cotangent lifts +The action-angle theorem for symplectic manifolds (also known as action-angle +coordinate theorem) can be reformulated in terms of a cotangent lift. +Recall that given a Lie group action on any manifold its cotangent lifted action +is automatically Hamiltonian. By considering the action of a torus on itself by +translations this action can be lifted to its cotangent bundle and give a semilocal +normal form theorem as the Arnold-Liouville-Mineur theorem for symplectic man- +ifolds. If we now replace this cotangent lift to the cotangent bundle to a lift to the +bm-cotangent bundle we obtain the semilocal normal form of the main theorem of +this chapter. +Let start recalling the symplectic and b-symplectic case following [KM17]. +5.1. Cotangent lifts and Arnold-Liouville-Mineur in Symplectic +Geometry +Let G be a Lie group and let M be any smooth manifold. Given a group action +ρ : G × M −→ M, we define its cotangent lift as the action on T ∗M given by +ˆρg := ρ∗ +g−1 where g ∈ G. We then have a commuting diagram +T ∗M +T ∗M +M +M +ˆ +ρg +ˆπ +π +ρg +Figure 1. Commutiative diagram of the construction of the iso- +morphism of bm-integrable systems. +where π is the canonical projection from T ∗M to M. +The cotangent bundle T ∗M is a symplectic manifold endowed with the exact +symplectic form given by the differential of the Liouville one-form ω = −dλ. The +Lioville one-form can be defined intrinsically: +(5.1) +⟨λp, v⟩ := ⟨p, (πp)∗(v)⟩ +with v ∈ T (T ∗M), p ∈ T ∗M. +A standard argument (see for instance [GS90]) shows that the cotangent lift ˆρ +is Hamiltonian with moment map µ : T ∗M → g∗ given by +⟨µ(p), X⟩ := ⟨λp, X#|p⟩ = ⟨p, X#|π(p)⟩, +37 + +38 +5. ACTION-ANGLE COORDINATES AND COTANGENT LIFTS +where p ∈ T ∗M, X is an element of the Lie algebra g and we use the same symbol +X# to denote the fundamental vector field of X generated by the action on T ∗M +or M. This construction is known as the cotangent lift. +In the special case where the manifold M is a torus Tn and the group is Tn +acting by translations, we obtain the following explicit structure: Let θ1, . . . , θn be +the standard (S1-valued) coordinates on Tn and let +(5.2) +θ1, . . . , θn +� +�� +� +=:θ +, t1, . . . , tn +� +�� +� +=:t +be the corresponding chart on T ∗Tn, i.e. +we associate to the coordinates (5.2) +the cotangent vector � +i tidθi ∈ T ∗ +θ Tn. The Liouville one-form is given in these +coordinates by +λ = +n +� +i=1 +tidθi +and its negative differential is the standard symplectic form on T ∗Tn: +(5.3) +ωcan = +n +� +i=1 +dθi ∧ dti. +Denoting by τβ the translation by β ∈ Tn on Tn, its lift to T ∗Tn is given by +ˆτβ : (θ, t) �→ (θ + β, t). +The moment map µcan : T ∗Tn → t∗ of the lifted action with respect to the canonical +symplectic form is +(5.4) +µcan(θ, t) = +� +i +tidθi, +where the θi on the right hand side are understood as elements of t∗ in the obvious +way. Even simpler, if we identify t∗ with Rn by choosing the standard basis +∂ +∂θi +of t then the moment map is just the projection onto the second component of +T ∗Tn ∼= Tn × Rn. Note that the components of µ naturally define an integrable +system on T ∗Tn. +We can rephrase the Arnold-Liouville-Mineur theorem in terms of the symplec- +tic cotangent model: +Theorem 5.1. Let F = (f1, . . . , fn) be an integrable system on the symplectic +manifold (M, ω). Then semilocally around a regular Liouville torus the system is +equivalent to the cotangent model (T ∗Tn)can restricted to a neighbourhood of the +zero section (T ∗Tn)0 of T ∗Tn. +5.2. The case of bm-symplectic manifolds +Let us start by introducing the twisted bm-cotangent model for torus actions. +This model has additional invariants: the modular vector field of the connected +component of the critical set and the modular weights of the associated toric action. +Consider T ∗Tn be endowed with the standard coordinates (θ, t), θ ∈ Tn, t ∈ Rn and +consider again the action on T ∗Tn induced by lifting translations of the torus Tn. +We will now view this action as a bm-Hamiltonian action with respect to a suitable +bm-symplectic form. In analogy to the classical Liouville one-form we define the +following non-smooth one-form away from the hypersurface Z = {t1 = 0} : + +5.2. THE CASE OF bm-SYMPLECTIC MANIFOLDS +39 +� +cc1 log |t1| + +m +� +i=2 +cci +t−(i−1) +1 +−(i − 1) +� +dθ1 + +n +� +i=2 +tidθi. +When differentiating this form we obtain a bm-symplectic form on T ∗Tn which +we call (after a sign change) the twisted bm-symplectic form on T ∗Tn with +invariants (cc1, . . . , ccm): +(5.5) +ωtw,c := + + +m +� +j=1 +cj +c +tj +1 +dt1 ∧ dθ1 + + + +n +� +i=2 +dti ∧ dθi, +where c is the modular period. The moment map of the lifted action is then given +by +(5.6) +µtw,q0,...,qm−1) := (q0 log |t1| + +m +� +i=2 +qit−(i−1) +1 +, t2, . . . , tn), +where we are identifying t∗ with Rn and cj = −(j − 1)qj−1. +We call this lift together with the bm-symplectic form 5.5 the twisted bm- +cotangent lift with modular period c and invariants (c1, . . . , cm). Note that the +components of the moment map define a bm-integrable system on (T ∗Tn, ωtw,(cc1,...,ccm)). +The model of twisted bm-cotangent lift allows us to express the action-angle +coordinate theorem for bm-integrable systems in the following way: +Theorem 5.2. Let F = (f1, . . . , fn) be a bm-integrable system on the bm- +symplectic manifold (M, ω). +Then semilocally around a regular Liouville torus +T, which lies inside the critical hypersurface Z of M, the system is equivalent to +the cotangent model (T ∗Tn)tw,(cc1,...,ccm) restricted to a neighbourhood of (T ∗Tn)0. +Here c is the modular period of the connected component of Z containing T and the +constants (c1, . . . , cm) are the invariants associated to the integrable system and its +associated toric action. + + +Part 3 +A KAM theorem for bm-symplectic +manifolds + +The KAM theorem explains how integrable systems behave under small per- +turbations. +More precisely, it studies how an integrable system in action-angle +coordinates responds to a small perturbation on its Hamiltonian. The trajectories +of an integrable system in action-angle coordinates can be seen as linear trajectories +over a torus. The KAM theorem finds a way to transform these original trajectories +to other linear trajectories over some transformed torus. The KAM theorem states +that most of these tori, and the linear solutions of the system on these tori, survive +if the perturbation is small enough. +In this part, we give a new KAM theorem for bm-symplectic manifolds with +detailed proof. This is contained in the first chapter of this part. Moreover, we +devote three more chapters to applications: +(1) Desingularization of bm-integrable systems. We present a way to use +the desingularization of bm-symplectic manifolds presented in [GMW17] +to construct standard smooth integrable systems from bm-integrable sys- +tems. This desingularized integrable system is uniquely defined. +(2) Desingularization of the KAM theorem on bm-symplectic man- +ifolds. In this section we use the desingularization of bm-integrable sys- +tems in conjunction with the KAM theorem for bm-symplectic manifolds +to deduce the original KAM theorem as well as a completely new KAM +theorem for folded symplectic forms. +(3) Potential applications to Celestial mechanics. We overview a list of +motivating examples from Celestial mechanics where regularization trans- +formations give rise to bm-symplectic forms. We discuss some potential +applications of perturbation theory in this set-up. + +CHAPTER 6 +A new KAM theorem +The objective of this chapter is to give a construction of KAM theory in the +setting of bm-symplectic manifolds and with bm-integrable systems. The core of the +chapter is the construction of the proper statement and the proof of the equivalent +of the KAM theorem on bm-symplectic manifolds. +This chapter is divided different sections: +(1) On the structure of the proof. On this section we are going to present +the main ideas that are going to appear in the proper statement and proof +of the main theorem. The idea of the theorem is to build a sequence of +bm-symplectomorphisms such that its limit transforms the hamiltonian to +only depend on the action coordinates. +(2) Technical results and definitions. On this section we present some +technical results and definitions that are key for the proof of the main +theorem. +(3) KAM theorem on bm-symplectic manifolds. +On this section we +present the statement and the proof of the main result of this chap- +ter. +The proof is structured in 6 parts. +In the first part we define +the parameters that are going to be used to define the sequence of bm- +symplectomorphisms. In the second part we build precisely this sequence +of bm-symplectomorphisms. In the third part we see that the sequence of +frequency maps of the transformed Hamiltonian functions at every step +converges. In the fourth part we see that the sequence of bm-symplectomorphisms +converges. In the fifth part we obtain results on the stability of the trajec- +tories under the original perturbation. In the sixth part, we find bounds +to explain how close the invariant tori are from the unperturbed. Finally, +we obtain a bound for the measure of the set of invariant tori. +6.1. On the structure of the proof +The first thing we do is to reduce our study to the case the perturbation is not +a bm-function but an analytic one. This is because any purely singular perturbation +only affects the component in the direction of the modular vector field and can be +easily controlled. +The idea of the proof is really similar to the classical KAM case. We want +to build a diffeomorphism such that its transformed hamiltonian only depends on +the action coordinates. But it is not possible to build this diffeomorphism in one +step. What we do, as it is done in the classical case, it is to build a sequence of +diffeomorphisms such that the part of the hamiltonian depending on the angular +variables decreases at every step. The idea is to remove the first K terms of its +Fourier expression at every step while making K rapidly increase. This is done by +43 + +44 +6. A NEW KAM THEOREM +assuming the diffeomorphism comes as the flow at time 1 generated by a Hamil- +tonian function. In this way one can use the Lie Series in conjunction with the +Fourier series to find the expression for the hamiltonian function that generates +our diffeomorphism. The final diffeomorphism will be the composition of all the +diffeomorphisms obtained at each step. One of the main difficulties of the proof, as +in the classical case, is to prove that these diffeomorphisms converge and to prove +some bounds of its norm. +We also note that for our bm-symplectic setting, the diffeomorphisms we con- +sider leave the defining function of the critical set invariant up to order m, this will +have an important role later. Also observe in particular that the critical set can +not be transformed by any perturbation given by a bm-function. +Next we give some technical definitions and results. We define the norms we are +going to use to do all the estimates. We set the notation for the proof and the state- +ment of the theorem. We define the notion of non-resonance for a neighborhood +of the critical set of the bm-symplectic manifold. We study the set of all possible +non-resonant vectors. And we state the inductive lemma, which gives us estimates +and constructions for every step of our sequence of diffeomorphisms. +After all this discussion we are in conditions to properly state the bm-version of +the KAM theorem. One important difference to the classical KAM theorem is that +we have to guarantee that at Z the set of non-resonant vectors does not become +the whole set of frequencies. This condition can be understood as the perturbation +being smaller than some constant multiplied by the inverse of the modular period. +The proof of the theorem is done in six different steps by following the structure +on [DG96]. Since we are going to use the inductive lemma at every step, first we +define the parameters and sets to which we are going apply such lemma. Then we +check that we can actually apply the lemma and obtain some extra estimates for +the results of the lemma. After this we see that the sequence of frequency vectors +converges. We do the same with the sequence of canonical transformations. Then +we get some bounds for the size of the components of the final diffeomorphism. +Next we characterize the tori that survive by the perturbation. Finally we give +some estimates for the measure of the set of these tori. +Note that our version of the bm-KAM theorem improves the one in [KMS16a] +in several ways. Firstly it is applicable to bm-symplectic structures not only for +b-symplectic. Also we give several estimates that are not obtained in [KMS16a], +this estimates have sense in a neighborhood of the critical set Z, while [KMS16a] +only studied the behavior at Z. Finally the type of perturbation we consider is far +more general, since we do not have any condition of the form of the perturbation +but only on its size. +6.1.1. Reducing the problem to an analytical perturbation. In the +standard KAM, we assume to have an analytic Hamiltonian h(I) depending only +on the action coordinates and we add to it a small analytical perturbation R(φ, I). +This perturbed system receives the name of nearly integrable system. And then find +a new coordinate system such that h(I) + R(φ, I) = ˜h(˜I) where most of the quasi- +periodic orbits are preserved and can be mapped to the unperturbed quasi-periodic +orbits by means of the coordinate change. +In our setting we may assume h(I) to not be analytical and be a bm-function. +Also the perturbation R(φ, I) may as well be considered a bm-function. +In the + +6.1. ON THE STRUCTURE OF THE PROOF +45 +following lines we justify without loss of generality that actually we can assume the +perturbation to be analytical. +Let us state this more precisely. Let (M, x, Z, ω, F) be a bm-manifold with a +bm integrable system F on it. Consider action angle coordinates on a neighborhood +of Z. Then we can assume the expressions: +ω = + + +m +� +j=1 +cj +Ij +1 + + dI1 ∧ dφ1 + +n +� +i=2 +dIi ∧ dφi, and +F = (q′ +0 log I1 + +m−1 +� +i=1 +q′ +i +1 +Ii +1 ++ h(I), f2, . . . , fn) +where h, f2, . . . , fn are analytical. +Let the Hamiltonian function of our system be the first component of the +moment map ˆh′ = q′ +0 log I1 + �m−1 +i=1 q′ +i +1 +Ii +1 + h = ζ′ + h, where ζ′ := q′ +0 log I1 + +�m−1 +i=1 q′ +i +1 +Ii +1 . Note that dζ′ = �m +i=1 ˆq′ +i +1 +I′ +1 , where ˆq′ +i = −(i − 1)q′ +i−1. Note that by +the result of the previous chapter cj/ˆq′ +j = K the modular period. In particular +cm/ˆq′ +m = K. +The hamiltonian system given by ˆh′ can be easily solved by φ = φ0 +u′t, I = I0 +where u′ is going to be defined in the following sections. Consider now a pertur- +bation of this system: +ˆH′ = ˆh′(I) = ˆR(I, φ), where ˆR is a bm-function ˆR(I, φ) = +Rζ(I1)+R(I, φ) where Rζ(I1) = (r0 log I1 +�m−1 +i=1 ri 1 +Ii +1 ) is the singular part. Then +we can consider the perturbations Rζ(I1) and R(I, φ) separately. This way, we may +consider Rζ(I) as part of ˆh′(I). Then we have a new hamiltonian +ˆh(I) = (q′ +0 + r0) log I1 + +m−1 +� +i=1 +(q′ +i + ri) 1 +Ii +1 ++ h = q0 log I1 + +m−1 +� +i=1 +qi +1 +Ii +1 ++ h. +Now, instead of the identity Kˆq′ +j = cj we will have K(ˆqj − ˆrj) = cj, which +implies K +� +1 − +ˆrj +ˆq′ +j+ˆrj +� += cj +ˆqj . In particular +K +� +1 − +ˆrm +ˆq′m + ˆrm +� += cm +ˆqm +Let us define K′ = K +� +1 − +ˆrm +ˆq′m+ˆrm +� +. So from now on we assume ˆh = q0 log I1 + +�m−1 +i=1 qi 1 +Ii +1 + h, that the perturbation R(φ, I) is analytical, and we have the condi- +tion cm +ˆqm = K′. Observe that this system with only the singular perturbation is still +easy to solve in the same way that the system previous to this perturbation was. +6.1.2. Looking for a bm-symplectomorphism. Assume we have a Hamil- +tonian function H = ˆh(I) + R(φ, I) in action-angle coordinates. Where ˆh(I) is the +singular component of the bm-integrable system, i.e. +(6.1) +ˆh(I) = h(I) + q0 log(I1) + +m−1 +� +i=1 +qi +1 +Ii +1 +, + +46 +6. A NEW KAM THEOREM +where h(I) is analytical1. Assume also that the bm-symplectic form ω2 in these +coordinates is expressed as: +(6.2) +ω = + + +m +� +j=1 +cj +Ij +1 + + dI1 ∧ dφ1 + +n +� +i=2 +dIi ∧ dφi. +And finally, the expression for the frequency vector is: +ˆu = ∂ˆh +∂I = +∂(h(I) + q0 log(I1) + �m−1 +i=1 qi 1 +Ii +1 ) +∂I += +� +u1 + +m +� +i=1 +ˆqi +Ii +1 +, u2, . . . , un +� +, +where ˆq1 = q0 and ˆqi−1 = −iqi if i ̸= 0. +The objective is to follow the steps of the usual KAM construction (the steps +followed are highly inspired in [DG96]) replacing the standard symplectic form for +ω and taking as hamiltonian the bm-function ˆh. +Remark 6.1. The objective of the construction is to find a diffeomorphism +(actually a bm-symplectomorphism) ψ such that H ◦ ψ = h(˜I). This is done induc- +tively, by taking H ◦ ψ = H ◦ φ1 ◦ . . .◦ φq ◦ . . ., while trying to make R(φ, I) smaller +at every step. +Let us focus in one single step +Recall the classical formula: +Lemma 6.2. See [DG96]. +f ◦ φt = +∞ +� +j=0 +tj +j!Lj +W f, +Lj +W f = {Lj−1 +W f, W} +Where W is the Hamiltonian that generates the flow φt, and {·, ·} is the correspond- +ing Poisson bracket. +We will denote rk(H, W, t) = �∞ +j=k +tj +j! Lj +W H. +1If another component of the moment map is chosen to be the hamiltonian of the system, +the result still holds: the computations can be replicated assuming ˆh(I) = h(I). +2In classical KAM, ω is used to denote the frequency vector ∂h +∂I . We need ω to denote the +bm-symplectic form so we are going to use u to denote the frequency vector. + +6.1. ON THE STRUCTURE OF THE PROOF +47 +(6.3) +H ◦ φ = H ◦ φ|t=1 += +∞ +� +j=0 +tj +j! Lj +W H +���� +ˆh+R +������ +t=1 += +ˆh + R{ˆh + R, W} + r2(H, W, 1) += +ˆh + R + {ˆh, W} + {R, W} + r2(ˆh, W, 1) ++r2(R, W, 1) += +ˆh + +R + {ˆh, W} +� +�� +� +We want to cancel +this term as +fast as we can ++r2(ˆh, W, 1) + r2(R, W, 1) +We want {ˆh, W} +R≤k = 0, equivalently {W, ˆh} = R≤k, where R≤k means the +Fourier expression of R up to order K: +R≤k = +� +k∈Rn +|k|1≤K +Rk(I)eik·φ +Let us impose the condition {W, ˆh} = R≤K. Let us write the expression of the +Poisson bracket associated to the bm-symplectic form. +{W, ˆh} += +� +1 +�m +j=1 +cj +Ij +1 +� � +∂W +∂φ1 +∂ˆh +∂I1 +− ∂W +∂I1 +∂ˆh +∂φ1 +� ++ +n +� +i=2 +� +∂W +∂φi +∂ˆh +∂Ii +− ∂W +∂Ii +∂ˆh +∂φi +� +Because ˆh depends only on I, +∂ˆh +∂φi = 0 for all i. Moreover, the singular part of +the bm-function only depends on I1 and hence its derivatives with respect to the +other variables are also 0. Using that ∂ˆh +∂I = u + �m +i=1 +ˆqi +Ii +1 the previous expression +can be simplified: +{W, ˆh} += + +u1 + �m +i=1 +ˆqi +Ii +1 +�m +j=1 +cj +Ij +1 + + ∂W +∂φ1 ++ +n +� +i=2 +∂W +∂φi +ui +To expand the expression further we develop W in its Fourier expression: W = +� +k∈Rn +|k|1≤K +Wk(I)eikφ. The Fourier expansion is added up to order K, because it is +only necessary for the expressions to agree up to order K. With this notations the +condition becomes: +{W, ˆh}≤K += + +u1 + �m +i=1 +ˆqi +Ii +1 +�m +j=1 +cj +Ij +1 + + +∂ +∂φ1 + + + + +� +k∈Rn +|k|1≤K +Wk(I)eikφ + + + + + +48 +6. A NEW KAM THEOREM ++ +n +� +j=2 +uj +∂ +∂φj + + + + +� +k∈Rn +|k|1≤K +Wk(I)eikφ + + + + += + +u1 + �m +i=1 +ˆqi +Ii +1 +�m +j=1 +cj +Ij +1 + + + + + + +� +k∈Rn +|k|1≤K +Wk(I)eikφik1 + + + + ++ +n +� +j=2 +uj + + + + +� +k∈Rn +|k|1≤K +Wk(I)eikφikj + + + + += +� +k∈Rn +|k|1≤K +Wk(I)eikφ · + +ik1 + +u1 + �m +i=1 +ˆqi +Ii +1 +�m +j=1 +cj +Ij +1 + + + +n +� +j=2 +ikjuj + + += R≤K +Then, it is possible to make the two sides of the equation equal by imposing +the condition term by term: +(6.4) +Wk(I) += +Rk(I) +1 +i +� +k1 +� +u1+�m +i=1 +ˆ +qi +Ii +1 +�m +j=1 +cj +Ij +1 +� ++ �n +j=2 kjuj +� += +Rk(I) +1 +i +� +k1 +� +u1+�m +i=1 +ˆ +qi +Ii +1 +�m +j=1 +cj +Ij +1 +� ++ ¯k¯u +�, +where we adopted the notation �n +j=2 kjuj = ¯k¯u. +Remark 6.3. Observe that the expression 6.4 has no sense when k = ⃗0 and +hence {W, h}0 = R03 can not be solved. Let W0(I) = 0, then {h, W}≤K = R≤K − +R0. +Plugging the results above into the equation 6.3, one obtains: +H ◦ φ = ˆh + R0 + R≥K + r2(ˆh, W, 1) + r1(R, W, 1) +With this construction the diffeomorphism φ is found. But this is only the +first of many steps. If q denotes the number of the iteration of this procedure, in +general, we obtain: +3The zero term of the Fourier series can be seen as the angular average of the function + +6.1. ON THE STRUCTURE OF THE PROOF +49 +(6.5) +H(q) = H(q−1) �� φ(q) += +ˆh(q−1) + R(q−1) +0 ++ R(q−1) +≥K ++r2(h(q−1), W (q), 1) + r1(R(q−1), W (q), 1), +and at every step: +(6.6) +�ˆh(q) = ˆh(q−1) + R(q−1) +0 +R(q) = R(q−1) +>K ++ r2(ˆh(q−1), W (q), 1) + r1(R(q−1), W (q), 1) +6.1.3. On the change of the defining function under +bm-symplectomorphisms. Note that since we are considering bm-manifolds it +only makes sense to consider I1 up to order m, see [Sco16]. When talking about +defining functions we are interested in [I1], its jet up to order m. By definition bm- +maps preserve I1 up to order m and bm-vector fields X are such that LX(I1) = g·Im +1 +for g ∈ C∞(M). +Lemma 6.4. Let φt be the integral flow of X a bm-vector field, then φt is a +bm-map. +Proof. We want +I1 ◦ φt = I1 + Im +1 · g +for some g ∈ C∞(M). We will use 6.2. +I1 ◦ φt = +∞ +� +j=0 +tj +j!Lj +XI1 = I1 + LX(I1) + +∞ +� +j=2 +tj +j!Lj +XI1 += I1 + Im +1 + +∞ +� +j=2 +tj +j!Lj +XI1. +On the other hand, let us prove by induction Lk +XI1 = g(k)Im +1 . The first case is +obvious, assume the case k holds and let us prove the case k + 1. +Lk+1 +X +I1 += +{Lk +XI1, X} += +{g(k)Im +1 , X} += +(LXg(q))Im +1 + g(k) · mIm−1 +1 +LXI1 += +(LXg(k) + g(k) · m · Im−1 +1 +· g)Im +1 += +g(k+1)Im +1 +where g(k+1) = LXg(k) + g(k) · m · Im−1 +1 +· g. +□ +Lemma 6.5. The Hamiltonian vector flow of some smooth hamiltonian function +h is a bm-vector field. +Proof. At each point of Z the following identity holds LXhI1 = Im +1 +∂f +∂φ1 . The +result can be extended at a neighborhood of Z. +□ +Observe that combining the two previous results we get that the hamiltonian +flow of a function preserves I1 up to order m. + +50 +6. A NEW KAM THEOREM +6.2. Technical results +As the non-singular part of our functions we will be considering analytic func- +tions on T × G, G ⊂ Rn. +The easiest way to work with these functions is to +consider them as holomorphic functions on some complex neighborhood. Let us +define formally this neighborhood. +Wρ1(Tn) := {φ : ℜφ ∈ Tn, |ℑφ|∞ ≤ ρ1}, +Vρ2(G) := {I ∈ Cn : |I − I′| ≤ ρ2 for some I′ ∈ G}, +Dρ(G) := Wρ1(Tn) × Vρ2(G), +where | · |∞ denotes the maximum norm and | · |2 denotes de Euclidean norm. +Now it is necessary to clarify the norms that are going to be used on these sets. +Definition 6.6. Let f be an action function (only depending on the I-coordinates), +and F an action vector field. +|f|G,η := supI∈Vη(G) |f(I)|, +|f|G := |f|G,0 +|F|G,η,p := supI∈Vη(G) |F(I)|p, +|F|G,η := |F|G,η,2 +Now, assume f(I, φ) to be an action-angle function written using its Fourier ex- +pansion as � +k∈Zn fk(I)eik·φ, and F to be an action-angle vector field. +|f|G,ρ := sup(φ,I)∈Dρ(G) |f(I)|, +∥f∥G,ρ := � +k∈Zn |fk|G,ρ2e|k|1ρ1 +|F|G,ρ,p := � +k∈Zn |Fk|G,ρ2,pe|k|1ρ1, +∥F∥G,ρ = ∥F∥G,ρ,2 +Lemma 6.7 (Cauchy Inequality). +���� +∂f +∂φ +���� +G,(ρ1−δ1,ρ2),1 +≤ +1 +eδ1 +∥f∥G,ρ +���� +∂f +∂I +���� +G,(ρ1,ρ2−δ2),∞ +≤ 1 +δ2 +∥f∥G,ρ +Definition 6.8. If Df = ( ∂f +∂φ, ∂f +∂I ), +∥Df∥G,ρ,c := max +� +∥∂f +∂φ∥G,ρ,1, c∥∂f +∂I ∥G,ρ,∞ +� +Definition 6.9. To simplify our notation, let us define: +A(I1) = +�m +j=1 +ˆqj +Ij +1 +�m +j=1 +cj +Ij +1 +and +B(I1) = +1 +�m +j=1 +cj +Ij +1 +. +Remark 6.10. With this notation, equation 6.4 can be written as: +Wk(I) = +Rk(I) +i(k1B(I1)u1 + ¯k¯u + k1A(I1)) +Observe that A(I1) and B(I1) are analytic (holomorphic on the complex ex- +tended domain) where the denominator does not vanish. We can assume that this +does not happen by shrinking the domain G in the direction of I1. Observe, in +particular, that when I1 → 0, A(I1) → ˆqm/cm = 1/K′ the inverse of the modular +period and B(I1) → 0. In this way, the norms of A(I1) and B(I1) are bounded and +well defined. We will denote these norms by KA and KB respectively. Also, since + +6.2. TECHNICAL RESULTS +51 +A(I1) and B(I1) are analytic, their derivatives will also be bounded, and we will +denote the norms of these derivatives by KA′ and KB′. +To further simplify the notation in the following computations we introduce +the definition: +Definition 6.11. +¯ +A = +� +A +0 +� +and +¯B = +� +B +0 +0 +Idn−1,n−1 +� +Remark 6.12. With this notation, equation 6.4 can be written as: +(6.7) +Wk(I) = +Rk(I) +i(k ¯B(I1)u + k ¯ +A(I1)) +Definition 6.13. Having fixed ω, a bm-symplectic form (as in equation 6.2) +and ˆh a bm-function (as in equation 6.1) as a hamiltonian. Given an integer K and +α > 0, F ⊂ Rn (or Cn) the space of frequencies is said to be α, K-non-resonant +with respect to (c1, . . . , cm) and (ˆq1, . . . , ˆqm) if +|k ¯B(I1)u + k ¯ +A(I1)| ≥ α, ∀k ∈ Z \ {0}, |k|1 ≤ K, ∀u ∈ F. +We are going to use the notation α, K, c, ˆq-non-resonant. +Remark 6.14. The non-resonance condition is established on u = ∂h/∂I, not +on ˆu = ∂ˆh/∂I, because our non-resonance condition already takes into account the +singularities. In this way we can use the analytic character of u. +Remark 6.15. If +�� ∂u +∂I +�� +G,ρ2 is bounded by M ′, then +�� ∂ +∂I +� ¯Bu + ¯ +A +��� +G,ρ2 is also +bounded: +(6.8) +�� ∂ +∂I +� ¯Bu + ¯ +A +��� +G,ρ2 +≤ +��� ∂ ¯ +B +∂I u + ¯B ∂u +∂I + ∂ ¯ +A +∂I +��� +G,ρ2 +≤ +KB′|u|G,ρ2 + KBM ′ + KA =: M. +Remark 6.16. When we consider the standard KAM theorem, the frequency +vector u is relevant because the solution to the Hamilton equations of the unper- +turbed problem has the form: +I = I0, +φ = φ0 + ut. +Let us see what plays the role of u in our bm-KAM theorem. Let us find the +coordinate expression of the solution to ιXˆhω = dˆh, where ω is a bm-symplectic +form in action-angle coordinates. +Xˆh = ˙I1 +∂ +∂I1 ++ . . . + ˙In +∂ +∂In +, +where ˙I1, . . . , ˙In are the functions we want to find. +dˆh = + + +m +� +j=1 +ˆqi +1 +Ij +1 + + dI1 + dh, +and hence, + +52 +6. A NEW KAM THEOREM +Xˆh = Π(dˆh, ·) = +�m +i=1 +ˆqi +Ii +1 +�m +i=1 +cj +Ij +1 +∂ +∂φi ++ Xh. +Hence φ = φ0 + ( ¯Bu + ¯ +A +� �� � +u′ +)t. So the frequency vector that we are going to be +concerned about is going to be u′ instead of ˆu = +∂ +∂I ˆh. +Lemma 6.17. If u is one-to-one from G to its image then u′ = ¯Bu + ¯ +A is +also one-to-one from G′ to its image in a neighborhood of Z, while at Z it is the +projection of u such that the first coordinate is sent to ˆqm +cm = 1/K′ the inverse of the +modular period, were G′ ⊆ G. +Proof. Because +u′ = + + +1 +�m +j=1 +cj +Ij +1 +u1 + +�m +j=1 +ˆqj +Ij +1 +�m +j=1 +cj +I1 +, u2, . . . , un + + , +and B is invertible outside I1 = 0, shrinking G if necessary in the first dimension +the map is one-to-one. But at the critical set {I1 = 0}, u′ is a projection of u where +the first component is sent to the constant value ˆqm +cm = +1 +K′ . +□ +Lemma 6.18. If u(G) is α, K, c, ˆq-non-resonant, then u(Vρ2(G)) is α +2 , K, c, ˆq- +non-resonant, assuming that ρ2 ≤ +α +2MK and +�� ∂u +∂I +�� +G,ρ2 ≤ M ′ +Proof. Fix k ∈ Z \ {0}, we want to bound |k ¯B(I1)v + k ¯ +A(I1)| where v ∈ +u(Vρ2(G)) as a function on v. Given v ∈ u(Vρ2(G)) we ask whether there is any +bound for the distance to some v′ ∈ u(G). +v ∈ u(Vρ2(G)) ⇒ v = u(x), x ∈ Vρ2(G) ⇒ ∃y ∈ G such that |x − y| ≤ ρ2. +Take v′ = u(y). +|v − v′| ≤ |x − y| +���� +∂u +∂I +���� +G,ρ2 +≤ ρ2M ′ ≤ ρ2M/KB ≤ +α +2MK M/KB = +α +2KKB +. +Where we used equation 6.8 in the third inequality. +|k1B(I1)v1 + ¯k¯v + k1A(I1)| +≥ +|k1B(I1)v′ +1 + ¯k ¯v′ + k1A(I1)| +� +�� +� +≥α +−|k1B(I1)(v1 − v′ +1) + ¯k(¯v − ¯v′)| +≥ +α − KB |k · (v − v′)| +� +�� +� +≤Kα/(2KKB) +≥ +α − α/2 = α/2 +□ +Proposition 6.19. Let ˆh(I) be a bm-function as in equation 6.1. Assume h(I) +and R(φ, I) be real analytic on Dρ(G), u(G) = ∂h +∂I (G) is α, K, c, ˆq-non-resonant. +Assume also that | ∂ +∂I u|G,ρ2 ≤ M ′ and ρ2 ≤ +α +2MK . Let c > 0 given. Then R0(φ, I), + +6.2. TECHNICAL RESULTS +53 +W≤K(φ, I) given by the previous construction are both real analytic on Dρ(G) and +the following bounds hold +(1) ||DR0||G,ρ,c ≤ ||DR||G,ρ,c +(2) ||D(R − R0)||G,ρ,c ≤ ||DR0||G,ρ,c +(3) ||DW||G,ρ,c ≤ 2A +α ||DR0||G,ρ,c +Where A = 1 + 2Mc +α +Proof. Inequalities 1 and 2 are obvious because of the Fourier expression. +Let us prove inequality 3. +Let us expand R(φ, I) and W(φ, I) in their Fourier +expression: +R = +� +k∈Rn +Rk(I)eik·φ, +W = +� +k∈Rn +Wk(I)eik·φ. +We will bound this expression finding term-by-term bounds. +∂R +∂φ = +� +k∈Rn +Rk(I)eik·φik. +Hence, if we denote [ ∂R +∂φ ]k the k-th term of the Fourier expansion of ∂R +∂φ , we +have: +�∂R +∂φ +� +k += Rkik. +Let us compute the derivative of Wk with respect to the I variables: +∂Wk +∂I += +∂ +∂I +� +Rk +i(k ¯B(I1)u + k ¯ +A(I1)) +� += +∂Rk/∂I +i(k ¯B(I1)u + k ¯ +A(I1))) − Rki ∂ +∂I (k ¯B(I1)u + k ¯ +A(I1))) +[i(k ¯B(I1)u + k ¯ +A(I1)))]2 += +∂Rk/∂I +i(k ¯B(I1)u + k ¯ +A(I1))) + Rkik ∂ +∂I ( ¯B(I1)u + ¯ +A(I1))) +[(k ¯B(I1)u + k ¯ +A(I1)))]2 += +∂Rk/∂I +i(k ¯B(I1)u + k ¯ +A(I1))) + +[ ∂Rk +∂φ ]k ∂ +∂I ( ¯B(I1)u + ¯ +A(I1))) +[(k ¯B(I1)u + k ¯ +A(I1)))]2 +. +Then, we take norms (| · |G,ρ2,∞) on each side of the equation. +���� +∂Wk +∂I +���� +G,ρ2,∞ +≤ +2 +α +���� +∂Rk +∂I +���� +G,ρ2,∞ ++ 4M +α2 +���� +�∂Rk +∂φ +� +k +���� +G,ρ2,∞ +≤ +2 +α +���� +∂Rk +∂I +���� +G,ρ2,∞ ++ 4M +α2 +���� +�∂Rk +∂φ +� +k +���� +G,ρ2,1 +. +Taking the supremum at the whole domain: +���� +∂Wk +∂I +���� +G,ρ2,∞ +≤ +2 +α +���� +∂Rk +∂I +���� +G,ρ2,∞ ++ 4M +α2 +���� +�∂Rk +∂φ +� +k +���� +G,ρ2,1 +. +Moreover, + +54 +6. A NEW KAM THEOREM +∂W(I) +∂φ += +∂ +∂φ +� � +k∈Rn +Wk(I)eik·φ +� += +∂ +∂φ +� � +k∈Rn +ikWk(I)eik·φ +� +. +Hence, the k-th term of the Fourier series of ∂W +∂φ is +�∂W +∂φ +� +k += Wkik = +Rk +i(k ¯B(I1)u + k ¯ +A(I1)))ik += +1 +i(k ¯B(I1)u + k ¯ +A(I1))) +�∂R +∂φ +� +k +. +Taking norms (∥ · ∥G,ρ,1) at each side: +���� +∂W +∂φ +���� +G,ρ,1 +≤ 2 +α +���� +∂W +∂φ +���� +G,ρ,1 +. +Then, +∥DW∥G,ρ,c += +max +����� +∂W +∂φ +���� +G,ρ,1 +, c +���� +∂W +∂I +���� +G,ρ,∞ +� +≤ +max +� +2 +α +���� +∂R +∂φ +���� +G,ρ,1 +, c 2 +α +���� +∂R +∂I +���� +G,ρ2,∞ ++ c4M +α2 +���� +∂R +∂φ +���� +G,ρ2,1 +� +≤ +max +� +2 +α +���� +∂R +∂φ +���� +G,ρ,1 +, 2 +α ∥DR∥G,ρ2,c + c4M +α2 ∥DR∥G,ρ2,c +� += +max +� +2 +α +���� +∂R +∂φ +���� +G,ρ,1 +, 2 +α +� +1 + 2M +α c +� +∥DR∥G,ρ2,c +� +≤ +2 +α +� +1 + 2M +α c +� +∥DR∥G,ρ2,c +≤ +2 +αA ∥DR∥G,ρ2,c, +where A is as desired. +□ +Recall the Cauchy inequalities, see [P¨os93]: +(6.9) +���� +∂f +∂φ +���� +G,(ρ1,ρ2),1 +≤ +1 +eδ1 +∥f∥G,ρ +���� +∂f +∂I +���� +G,(ρ1,ρ2−δ2),∞ +≤ +1 +δ2 +∥f∥G,ρ + +6.2. TECHNICAL RESULTS +55 +Lemma 6.20. Let f, g be analytic functions on Dρ(G), where 0 < δ = (δ1, δ2) < +ρ = (ρ1, ρ2) and c > 0. Define ˆδc := min(cδ1, δ2). The following inequalities hold: +(1) ∥Df∥G,ρ−δ,c ≤ +c +ˆδc ∥f∥G,ρ +(2) ∥{f, g}∥G,ρ ≤ 2 +c∥Df∥G,ρ,c · ∥Dg∥G,ρ,c +(3) ∥D(f>K)∥G,(ρ−δ1,ρ2),c ≤ e−Kδ1∥Df∥G,ρ,c +Proof. Let us prove each point separately. +(1) Using the Cauchy inequalities one obtains the following: +���� +∂f +∂φ +���� +G,ρ−δ,1 += +���� +∂f +∂φ +���� +G,(ρ1−δ1,ρ2−δ2),1 +≤ +���� +∂f +∂φ +���� +G,(ρ1−δ1,ρ2),1 +≤ +1 +eδ1 +∥f∥G,ρ, +���� +∂f +∂I +���� +G,ρ−δ,∞ += +���� +∂f +∂I +���� +G,(ρ1−δ1,ρ2−δ2),∞ +≤ +���� +∂f +∂I +���� +G,(ρ1,ρ2−δ2),∞ +≤ 1 +δ1 +∥f∥G,ρ. +Putting the two inequalities inside the definition of the norm: +∥Df∥G,ρ−δ,c += +max +����� +∂f +∂φ +���� +G,ρ−δ,1 +, c +���� +∂f +∂I +���� +G,ρ−δ,∞ +� +≤ +max +� 1 +eδ1 +∥f∥G,ρ, c +δ2 +∥f∥G,ρ +� +≤ +max +� 1 +eδ1 +c +c, c +δ2 +� +∥f∥G,ρ +≤ +max +� c +eˆδc +, c +ˆδc +� +∥f∥G,ρ, +where the last inequality holds because ˆδc = min(cδ1, δ2). +(2) Let us find the expression of {f, g} for a bm-symplectic structure. {f, g} = +ω(Xf, Xg) where Xf and Xg are such that ιXf ω = df and ιXgω = dg. +Let restrict the computations only to f. +df = +n +� +i=1 +∂f +∂φ1 +dφ1, +Xf = +n +� +i=1 +ai +∂ +∂φi ++ +n +� +i=1 +bi +∂ +∂φi +. +Where ai and bi are coefficients to be determined by imposing the +following condition: +ιXf ω = + + +m +� +j=1 +cj +Ij +1 + + (a1dI1 − b1dφ1) + +n +� +i=2 +(aidIi − bidφi) = df. +Then, solving for the coefficients the following expressions are ob- +tained: + +56 +6. A NEW KAM THEOREM +a1 = +1 +��m +j=1 +cj +Ij +1 +� ∂f +∂φ1 +and +ai = ∂f +∂φi +for i ̸= 1, +b1 = − +1 +��m +j=1 +cj +Ij +1 +� ∂f +∂φ1 +and +bi = − ∂f +∂φi +for i ̸= 1. +Hence, the expression for the hamiltonian vector fields becomes: +Xf = +1 +��m +j=1 +cj +Ij +1 +� +� ∂f +∂φ1 +∂ +∂φ1 +− ∂f +∂I1 +∂ +∂I1 +� ++ +n +� +i=1 +� ∂f +∂φi +∂ +∂φi +− ∂f +∂Ii +∂ +∂Ii +� +, +Xg = +1 +��m +j=1 +cj +Ij +1 +� +� ∂g +∂φ1 +∂ +∂φ1 +− ∂g +∂I1 +∂ +∂I1 +� ++ +n +� +i=1 +� ∂g +∂φi +∂ +∂φi +− ∂g +∂Ii +∂ +∂Ii +� +. +Then the Poisson bracket applied to the two functions: +{f, g} = ω(Xf, Xg) += +1 +��m +j=1 +cj +Ij +1 +� +� ∂f +∂I1 +∂g +∂φ1 +− ∂f +∂φ1 +∂g +∂I1 +� ++ +n +� +i=2 +� ∂f +∂Ii +∂g +∂φi +− ∂f +∂φi +∂g +∂Ii +� +. +And hence the norm of the Poisson bracket becomes: +∥{f, g}∥G,ρ += +������� +1 +��m +j=1 +cj +Ij +1 +� +� ∂f +∂I1 +∂g +∂φ1 +− ∂f +∂φ1 +∂g +∂I1 +� ++ +n +� +i=2 +� ∂f +∂Ii +∂g +∂φi +− ∂f +∂φi +∂g +∂Ii +������ +G,ρ +≤ +����� +n +� +i=1 +� ∂f +∂Ii +∂g +∂φi +− ∂f +∂φi +∂g +∂Ii +������ +G,ρ +Where we assumed +����m +j=1 +cj +Ij +1 +��� ≥ 1. This assumption makes sense, +because we are interested in the behaviour close the critical set Z. Close +enough to the critical set this expression holds. Then, +∥{f, g}∥G,ρ +≤ +n +� +i=1 +���� +∂f +∂Ii +���� +G,ρ +���� +∂g +∂��i +���� +G,ρ ++ +n +� +i=1 +���� +∂f +∂φi +���� +G,ρ +���� +∂g +∂Ii +���� +G,ρ +≤ +���� +∂f +∂I +���� +G,ρ,∞ +���� +∂g +∂I +���� +G,ρ,1 ++ +���� +∂f +∂I +���� +G,ρ,1 +���� +∂g +∂I +���� +G,ρ,∞ +≤ +1 +c |Df∥G,ρ,c∥Dg∥G,ρ,c + 1 +c|Df∥G,ρ,c∥Dg∥G,ρ,c + +6.2. TECHNICAL RESULTS +57 +≤ +2 +c ∥Df∥G,ρ,c∥Dg∥G,ρ,c. +(3) Lastly, +∥D(f>K)∥G,(ρ1−δ1,ρ2),1 += max +����� +∂f>K +∂φ +���� +G,(ρ1−δ1,ρ1),1 +, c +���� +∂f>K +∂I +���� +G,(ρ1−δ1,ρ1),∞ +� +. +We will proceed by bounding each term separately. On one hand: +���� +∂f +∂φ +���� +G,(ρ1,ρ2),1 += +����� +� +k∈Zn +ikfk(I)eikφ +����� +G,(ρ1,ρ2),1 +≥ +� +k∈Zn +k ∥fk(I)∥G,ρ2,1 e|k|1ρ1 +≥ +� +k∈Zn +|k|1>K +k ∥fk(I)∥G,ρ2,1 e|k|1(ρ1+δ1−δ1) +≥ +eKδ1 +� +k∈Zn +|k|1>K +k ∥fk(I)∥G,ρ2,1 e|k|1(ρ1−δ1) += +eKδ1 +���� +∂f>K +∂φ +���� +G,(ρ1−δ1,ρ2),1 +. +On the other hand: +���� +∂f +∂I +���� +G,(ρ1,ρ2),∞ += +����� +� +k∈Zn +∂fk(I) +∂I +eikφ +����� +G,(ρ1,ρ2),∞ +≥ +� +k∈Zn +���� +∂fk(I) +∂I +���� +G,ρ2,∞ +e|k|1ρ1 +≥ +� +k∈Zn +|k|1>K +���� +∂fk(I) +∂I +���� +G,ρ2,∞ +e|k|1(ρ1+δ1−δ1) +≥ +eKδ1 +� +k∈Zn +|k|1>K +���� +∂fk(I) +∂I +���� +G,ρ2,∞ +e|k|1(ρ1−δ1) +≥ +eKδ1 +���� +∂f>K +∂I +���� +G,(ρ1−δ1,ρ2),∞ +. +Hence ∥D(f>k)∥G,(ρ1−δ1,ρ2),c ≤ e−Kδ1∥Df∥G,ρ,c. + +58 +6. A NEW KAM THEOREM +□ +Now we define a norm that indicates how close a map Φ is to the identity. +Definition 6.21. Let x = (φ, I) ∈ C2n, then +|x|c := max(|φ|1, c|I|∞) +Definition 6.22. For a map Υ : Dρ(G) → C2n its norm and the norm of its +derivative its defined as: +|Υ|G,ρ,c := +sup +x∈Dρ(G) +|Υ(x)|c, +|DΥ|G,ρ,c := +sup +x∈Dρ(G) +|DΥ(x)|c, +where |DΥ(x)|c = sup +y∈R2n +|y|c=1 +|DΥ(x) · y|c +Lemma 6.23. If Υ is analytic on Dρ(G), then |DΥ|G,ρ−δ,C ≤ |Υ|G,ρ,c +ˆδc +Proof. Observe that if we consider ∥.∥ any norm on Cn and a matrix A of +size n × n, and ∥A∥ defines the induced norm of matrices i.e. +∥A∥ = sup +y∈C2n +∥y∥=1 +∥A · y∥ +then one has that ∥(∥a1∥′, . . . , ∥an∥′)∥ ≤ ∥A∥ where aj denotes the j-th row of A. +Also note that ∥ · ∥′ can be a any norm consider the infinity norm. This can be +easily proven in the following way: +∥A · y∥ = +������� + + + +a1 · y +... +an · y + + + +������� +≤ +������� + + + +∥a1∥′∥y∥′ +... +∥an∥′∥y∥′ + + + +������� +Where ∀y ∈ Cn such that ∥y∥ = 1. Let aj be the rows of DΥ(x), +aj = +�∂Υj +∂φ , ∂Υj +∂I +� +, +and be ∥aj∥′ its norm. With this property in mind we proceed as follows: +|DΥ|G,ρ−δ,c += +sup +x∈Dρ−δ(G) +|DΥ(x)|c +≤ +sup +x∈Dρ−δ(G) +|(|a1|∞, . . . , |an|∞)|c +≤ +��� +� +supx∈Dρ−δ ∥DΥ1∥∞ , . . . , supx∈Dρ−δ ∥DΥ2n∥∞ +���� +c += +��� +� +∥DΥ1∥G,ρ−δ,∞ , . . . , ∥DΥ2n∥G,ρ−δ,∞ +���� +c +≤ +��� +� +1 +δ1 ∥Υ1∥G,ρ , . . . , 1 +δ1 ∥Υ2n∥G,ρ +���� +c + +6.2. TECHNICAL RESULTS +59 +≤ +1 +ˆδc +���∥Υ1∥G,ρ , . . . , ∥Υ2n∥G,ρ +��� +c += +1 +ˆδc supx∈Dρ(G) |Υ1, . . . , Υ2n|c = +1 +ˆδc supx∈Dρ(G) |Υ|c += +1 +ˆδc |Υ|G,ρ,c +□ +Lemma 6.24. Let W be an analytic function on Dρ(G), ρ > 0 and let Φt be its +Hamiltonian flow at time t (t > 0). Let δ = (δ1, δ2) > 0 and c > 0 given. Assume +that ∥DW∥G,ρ,c ≤ ˆδc. Then, Φt maps Dρ−tδ(G) into Dρ(G) and one has: +(1) |Φt − Id|G,ρ−tδ,c ≤ t∥DW∥G,ρ,c, +(2) Φ(Dρ(G)) ⊃ Dρ−tδ(G) for ρ′ ≤ ρ − tδ, +(3) Assuming that ∥DW∥G,ρ,c < ˆδc/2e, for any given function f analytic on +Dρ(G), and for any integer m ≥ 0, the following bound holds: +∥rm(f, W, t)∥G,ρ−tδ +≤ +∞ +� +l=0 +� +1 +�l+m +m +� · +�2e∥DW∥G,ρ,c +ˆδc +�l� +tm +m!∥Lm +Wf∥G,ρ += γm +�2e∥DW∥G,ρ,c +ˆδc +� +· tm∥Lm +Wf∥G,ρ, +where for 0 ≤ x ≤ 1 we define +γm(x) := +∞ +� +l=0 +l! +(l + m)!xl +Proof. During the proof we are going to denote Φs(φ0, I0) by (φ(s), I(s)). +Let us find the coordinate expression of the hamiltonian flow for the expression +6.2 of a bm-symplectic form. Recall that the equation for the hamiltonian flow is +d +dsφi(s) = {φi, W} and +d +dsIi(s) = {Ii, W}. +{φi, W} = +1 +��m +j=1 +cj +Ij +1 +� +�∂φi +∂I1 +· ∂W +∂φ1 +− ∂φi +∂φ1 +· ∂W +∂I1 +� ++ +n +� +j=2 +�∂φi +∂Ij +· ∂W +∂φj +− ∂φi +∂φj +· ∂W +∂Ij +� +. +Hence, +d +dsφi(s) = − +1 +��m +j=1 +cj +Ij +1 +� ∂W +∂I1 +if i = 1 and d +dsφi(s) = −∂W +∂Ii +if i ̸= 1. +On the other side, +{Ii, W} = +1 +��m +j=1 +cj +Ij +1 +� +� ∂Ii +∂I1 +· ∂W +∂φ1 +− ∂Ii +∂φ1 +· ∂W +∂I1 +� ++ +n +� +j=2 +� ∂Ii +∂Ij +· ∂W +∂φj +− ∂Ii +∂φj +· ∂W +∂Ij +� +. + +60 +6. A NEW KAM THEOREM +Hence, +d +dsIi(s) = +1 +��m +j=1 +cj +Ij +1 +� ∂W +∂φ1 +if i = 1 and d +dsIi(s) = ∂W +∂φi +if i ̸= 1. +(1) Assume now that 0 < s0 ≤ t. Then, +|φ(s0) − φ0|∞ +≤ +s0 sup0 0, A = 1 + 2Mc +α . +Assume that ρ2 ≤ +α +2MK , ∥DR∥G,ρ,c ≤ αˆδc +74A. Then, there exists a real analytic map +Φ : Dρ− δ +2 (G) → Dρ(G), such that H ◦ Φ = ˆh + ˜R,with: +(1) ∥D ˜R∥G,ρ−δ,c ≤ e−Kδ1∥DR∥G,ρ,c + 14A +αˆδc ∥DR∥2 +G,ρ,c, +(2) |Φ − Id|G,ρ− δ +2 ,c ≤ 2A +α ∥DR∥G,ρ,c, +(3) Φ(Dρ′(G)) ⊃ Dρ′− δ +2 (G) for ρ′ ≤ ρ − δ +2 +Proof. Recall that +�� ∂ +∂I u +�� +G,ρ2 ≤ M ′ implies +�� ∂ +∂I ( ¯Bu + ¯ +A) +�� +G,ρ2 ≤ M by equa- +tion 6.8. By equation 6.6 +R(q) = R(q−1) +>K ++ r2(ˆh(q−1), W (q), 1) + r1(R(q−1), W (q), 1). +To simplify the notation we are going to omit the index of the iteration: +(6.11) +˜R = R>K + r2(ˆh,W, 1) + r1(R, W, 1). +Where W is defined in terms of its Fourier expressions by equation 6.7: +Wk(I) = +Rk(I) +i(k ¯B(I1)u + k ¯ +A(I1)) +By proposition 6.19: ∥DW∥G,ρ,c ≤ 2A +α ∥DR∥G,ρ,c ≤ 2A +α +αˆδc +74A = +ˆδc +37. And Φ is +defined as in lemma 6.20: Φ : Dρ− δ +2 (G) → Dρ(G). +(1) Differentiating equation 6.11 we obtain: +D ˜R = DR>K + Dr2(ˆh, W, 1) + Dr1(R, W, 1). +Taking norms at every side of the expression: +∥D ˜R∥G,ρ−δ,c += +∥DR>K + Dr2(ˆh, W, 1) + Dr1(R, W, 1)∥G,ρ−δ,c +≤ +∥DR>K∥G,ρ−δ,c + ∥Dr2(ˆh, W, 1)∥G,ρ−δ,c ++∥Dr1(R, W, 1)∥G,ρ−δ,c +≤ +e−Kδ1∥DR∥G,ρ,c ++ 2c +ˆδc +� +∥r2(ˆh, W, 1)∥G,ρ− δ +2 ,c + ∥r1(R, W, 1)∥G,ρ− δ +2 ,c +� +Let us further develop the two last terms of the previous expression, +by using lemma 6.24: +∥r2(ˆh, W, 1)∥G,ρ− δ +2 ,c +≤ +γ2 +� +2e∥DW∥G,ρ,c +ˆδc/2 +� +∥L2 +W h∥G,ρ +≤ +γ2 +� +4e∥DW∥G,ρ,c +ˆδc +� +∥{{h, W}, W}∥G,ρ, +∥r1(ˆh, W, 1)∥G,ρ− δ +2 ,c +≤ +γ1 +� +2e∥DW∥G,ρ,c +ˆδc/2 +� +∥L1 +W R∥G,ρ +≤ +γ1 +� +4e∥DW∥G,ρ,c +ˆδc +� +∥{R, W}∥G,ρ. + +62 +6. A NEW KAM THEOREM +Then, using the second statement of lemma 6.20 and that {W, h} = +R≤K: +∥{R, W}∥G,ρ ≤ 2 +c∥DR∥G,ρ,c∥DW∥G,ρ,c, and +|{{h, W}, W}∥G,ρ += +∥{R≤K, W}∥G,ρ +≤ +2 +c∥DR≤K∥G,ρ,c∥DW∥G,ρ,c +≤ +2 +c∥DR∥G,ρ,c∥DW∥G,ρ,c. +Moreover, it is easy to see that γ1(x) = +− log(1−x) +x +and γ2(x) = +x+(1−x) log(1−x) +x2 +. Observe that these functions are monotonously increas- +ing in x. Recall that ∥DW∥G,ρ,c ≤ 2A +α ∥DR∥G,ρ,c. Then, +∥r1(ˆh, W, 1)∥G,ρ− δ +2 ,c ++∥r2(ˆh, W, 1)∥G,ρ− δ +2 ,c +≤ +γ1 +� +4e∥DW∥G,ρ,c +ˆδc +� +∥{R, W}∥G,ρ ++γ2 +� +4e∥DW∥G,ρ,c +ˆδc +� +∥{{h, W}, W}∥G,ρ +≤ +γ1 +� +4e∥DW∥G,ρ,c +ˆδc +� +2 +c∥DR∥G,ρ,c∥DW∥G,ρ,c ++γ2 +� +4e∥DW∥G,ρ,c +ˆδc +� +2 +c∥DR∥G,ρ,c∥DW∥G,ρ,c +≤ +γ1 +� +4e∥DW∥G,ρ,c +ˆδc +� +2 +c +2A +α ∥DR∥2 +G,ρ,c ++γ2 +� +4e∥DW∥G,ρ,c +ˆδc +� +2 +c +2A +α ∥DR∥2 +G,ρ,c +≤ +2 +c[γ1( 4e +37) + γ2( 4e +37)] 2A +α ∥DR∥2 +G,ρ,c += +4A +αc [γ1( 4e +37) + γ2( 4e +37)]∥DR∥2 +G,ρ,c. +Moreover γ1( 4e +37) + γ2( 4e +37) ≈ 1.741 . . . < 7 +4. +Then, +∥D ˜R∥G,ρ−δ,c +≤ +e−Kδ1∥DR∥G,ρ,c + 2c +ˆδc +4A +αc +7 +4∥∥2 +G,ρ,c +≤ +e−Kδ1∥DR∥G,ρ,c + 14A +ˆδcα ∥DR∥2 +G,ρ,c, +as we wanted to prove. +(2) Direct from lemma 6.24: +|Φ − Id|G,ρ. δ +2 ,c ≤ ∥DW∥G,ρ,c ≤ 2A +α ∥DR∥G,ρ,c +(3) Also direct from lemma 6.24: +Φ(Dρ(G)) ⊃ Dρ′− δ +2 (G), for ρ′ ≤ ρ − δ/2 +□ +Definition 6.26. ∆c,ˆq(k, α) = {J ∈ Rn such that |k ¯B(I1)J + k ¯ +A(I1)| < α} +Lemma 6.27. With the previous definitions we have the following bounds. +Outside of Z: +meas (F ∩ ∆c,ˆq(k, α)) ≤ (diamF)n−1 2α +|k|2,ω +. + +6.2. TECHNICAL RESULTS +63 +At Z: +meas (F ∩ ∆c,ˆq(k, α)) +� += 0 +if α ≤ |k1| +K′ +≤ (diamF)n +if α > |k1| +K′ +Proof. It is important to understand the geometry of the set ∆c,ˆq(k, α). Re- +call that k ¯B(I1)J = k1B(I1)J1 + ¯k ¯J, hence this part of the expression can be inter- +preted as the scalar product of the vector J with the vector (k1B(I1), k2, . . . , kn). +Then the set {J ∈ Rn such that |k ¯B(I1)J| < α} is the space between two hyper- +planes orthogonal to (k1B(I1), k2, . . . , kn). Adding the term k ¯ +A(I1) only applies +a transition to the previous set. Let us find what is the separation between the +hyperplanes. Assume J is parallel to (k1B(I1), k2, . . . , kn) with lengths a: +J = a(k1B(I1), k2, . . . , kn) +|k|2,ω +, +where |k|2,ω = +� +B(I1)2k2 +1 + k2 +2 + . . . k2n. Then, +J · (B(I1), k1, . . . , kn) += +c(B(I1)k2 +1 + k2 +2 + . . . k2 +n) +1 +|k|2,ω += +a|k|2,ω ≤ α ⇔ a ≤ +α +|k|2,ω . +And finally, +meas (F ∩ ∆c,ˆq(k, α)) ≤ (diamF)n−1 2α +|k|2,ω +. +The previous formula can not be applied if when we are at Z and k = (k1, 0, . . . , 0). +(B(I1)k1, k2, . . . , kn) +|k ¯B(I1)J + k ¯ +A(I1)| < α +|k ¯B(I1)J| < α +−k ¯ +A(I1) +Figure 1. Graphical representation of the set ∆c,ˆq(α) +At Z, +∆c,ˆq(K, α) = {J ∈ Rn such that | ¯K ¯J + k1 +ˆqm +cm +| < α}. +And if k = (k1, 0, . . . , 0) then +∆c,ˆq(K, α) = {J ∈ Rn such that |k1 +ˆqm +cm +| < α}. + +64 +6. A NEW KAM THEOREM +Then +∆c,ˆq(k, α) = +� Rn +if |k1| < α cm +ˆqm = αK′, +{∅} +if |k1| ≥ α cm +ˆqm = αK′. +Using this last identity, the statement we wanted to prove is immediate. +□ +Definition 6.28. G − b := {I ∈ G such that Ub(I) ⊂ G}, where Ub(I) is the +ball of radius b centered at I. +Definition 6.29. F is a D-set if meas[(F − b1) \ (F − b2)] ≤ D(b2 − b1). +Lemma 6.30. Let F ⊂ Rn be a D-set for d ≥ 0, τ > 0, β ≥ 0 and k ≥ 0 an +integer. Consider the set +F(d, β, K) := (F − d) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, β +|k|τ +1 +� +. +Then, outside of Z: +(1) If d′ ≥ d, β′ ≥ β, k′ ≥ k, then +meas[F(d, β, k) \ F(d′, β′, k′)] ≤ +D(d′ − d) + 2(diamF)n−1 + + + + +� +k∈Zn\{0} +|k|1≤K +β′ − β +|k|τ +1|k|2,ω ++ +� +k∈Zn\{0} +0<|k|1≤K +β′ +|k|τ +1|k|2,ω + + + + +(2) For every b ≥ 0 +meas[F(d, β, K) \ (F(d, β, K) − b)] ≤ (D + 2n+1(dim F)n−1Kn)b +And inside of Z, if we assume β ≤ +1 +K′ , the equation 1 holds adding only the terms +¯k ̸= 0 and 2 holds without any change. +Proof. Recall that +∆c,ˆq +� +k, β +|k|τ +1 +� += +� +J ∈ Rn such that +��k ¯B(I1)J + k ¯ +A(I1) +�� < +β +|k|τ +1 +� +. +First we will prove the results outside of Z and then +(1) Let us expand the expression of meas[F(d, β, k) \ F(d′, β′, k′)]: + +(F − d) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, β +|k|τ +1 +� + + \ + +(F − d) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, β +|k|τ +1 +� + + . +Now we use the following property on the previous expression: +(A \ B) \ (C \ D) += +[(A \ B) \ C] ∪ [(A \ B) ∩ D] +⊂ +(A \ C) ∪ [(A \ B) ∩ D] += +(A \ C) ∪ (A ∩ (D \ B)), +where the last equality holds true because D ⊃ B. Using this property +we have that meas[F(d, β, k) \ F(d′, β′, k′)] is included in + +6.2. TECHNICAL RESULTS +65 +[(F − d) \ (F − d′)] +∪ + +(F − d) ∩ + + + + + + +� +k∈Zn\{0} +|k|1≤K′ +∆c,ˆq +� +k, β′ +|k|1 +� + + + + +\ + + + + +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, β +|k|1 +� + + + + + + + + . +And this expression is equivalent to: +[(F − d) \ (F − d′)] +∪ +� +k∈Zn\{0} +|k|1≤K +� +(F − d) ∩ +� +∆c,ˆq +� +k, β′ +|k|τ +1 +� +\∆c,ˆq +� +k, β +|k|τ +1 +��� +∪ +� +k∈Zn\{0} +K<|k|1≤K′ +� +(F − d) ∩ ∆c,ˆq +� +k, β′ +|k|τ +1 +�� +. +Now, using lemma 6.27 we obtain: +meas(F(d, β, K) \ F(d′, β′, K′)) ≤ +≤ D(d′ − d) + (diamF)n−1 + + + + +� +k∈Zn\{0} +|k|1≤K +2(β′ − β) +|k|τ +1|k|2,ω ++ +� +k∈Zn\{0} +K<|k|1≤K′ +2β′ +|k|τ +1|k|2,ω + + + + +(2) Observe that: +F(d, β, K) − b += + +(F − d) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, β +|k|τ +1 +� + + − b +⊃ +(F − (d + b)) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, β +|k|τ +1 ++ b|k|2,ω +� +. +Then, +meas[(F(d, β, K)) \ (F(d, β, K) − b)] +≤ +meas +�� +(F − d) \ � +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, +β +|k|τ +1 +�� +\ +� +(F − (d + b)) \ � +k∈Zn\{0} +|k|1≤K +∆c,ˆq +� +k, +β +|k|τ +1 +��� +≤ +meas [(F − d) \ (F − (d + b))∪ + +66 +6. A NEW KAM THEOREM +� +k∈Zn\{0} +|k|1≤K +� +(F − d) ∩ +� +∆c,ˆq +� +k, +β +|k|τ +1 + b|k|2,ω +��� +\ +� +∆c,ˆq +� +k, +β +|k|τ +1 +��� +≤ +Db + � +k∈Zn\{0} +|k|1≤K +(diamF)n−1 2b|k|2,ω +|k|2,ω +≤ +Db + 2nKn(diamF)n−1 · 2 = Db + 2n+1Kn(diamF)n−1, +where in the last inequality we used that the number of vectors k such +that |k|1 ≤ K is less or equal than 2nKn. +The previous identities worked outside of Z. Let us understand the set F(d, β, K) +when we are ate Z. +F(d, β, K) +:= +(F − d) \ � +k∈Zn\{0} +|k|1≤K +∆cˆq(k, +β +|k|τ +1 ) += +(F − d) \ + + + + + +� +k∈Zn\{0} +|k|1≤K +¯k̸=0 +∆cˆq(k, +β +|k|τ +1 ) + + + +∪ + + + +� +k∈Zn\{0} +|k|1≤K +¯k=0 +∆cˆq(k, +β +|k|τ +1 ) + + + + + += +(F − d) \ + + + + + +� +k∈Zn\{0} +|k|1≤K +¯k̸=0 +∆cˆq(k, +β +|k|τ +1 ) + + + +∪ + +� +k1∈Z\{0} +|k|1≤ +β +|k1|τ K′ +Rn + + + +. +Note that if for some k1 ∈ Z \ {0}, |k|1 ≥ +β +|k|τ +1 K′, we take out all the possible +frequencies. Then seems natural to ask |k|1 ≥ +β +|k|τ +1 K′ for all k1 ∈ Z \ {0}, which +holds if and only if |k1|1+τ ≥ βK′ for all k1 ∈ Z \ {0} or simply β ≤ +1 +K′ which we +assumed. Then +F(d, β, K) := (F − d) \ +� +k∈Zn\{0} +|k|1≤K +¯k̸=0 +∆cˆq(k, β +|k|τ +1 +). +Hence we can replicate the proof of 1 only with the terms ¯k ̸= 0. And the bound +of 2 can be slightly improved by using that the number of vectors k ∈ Zn \ {0} +such that |k|1 ≤ K and |¯k| ̸= 0 is bounded by 2nKn − K, but since it is not a big +improve, for the sake of simplicity we assume the bound 2 at Z. +□ +Lemma 6.31. Let G ⊂ Rn be compact. +u, ˜u : G → Rn maps of class C2. +|˜u − u| ≤ ε. +Assume that u is one-to-one on G, let F = u(G). +Consider the +following bounds: +���� +∂u +∂I +���� +G +≤ M, +���� +∂u +∂I (I) · v +���� ≥ µ|v| +∀v ∈ Rn, ∀I ∈ G, + +6.2. TECHNICAL RESULTS +67 +���� +∂˜u +∂I +���� +G +≤ ˜ +M, +���� +∂˜u +∂I2 +���� +G +≤ ˜ +M2, +���� +∂˜u +∂I (I)v +���� ≥ ˜µ|v| +∀v ∈ Rn, ∀I ∈ G, +˜µ < µ and ˜ +M < M. Assume ε ≤ ˜mu2/(4 ˜ +M2). Then, given a subset ˜F ⊂ F − 4Mε +˜µ +and writing ˜G = (˜u)−1( ˜F ), the map ˜u is one-to-one from ˜G to ˜F and +˜G ⊂ G − 2ǫ +˜µ , +u( ˜G) ⊃ ˜F − ε. +Moreover, +|(˜u)−1 − u−1| ˜ +F ≤ ε +µ +Proof. The statement is not any different than the classical one, so we are +not going to prove it in here. A proof can be found in [DG96]. +□ +Lemma 6.32 (Inductive lemma). Let G ⊂ Rn be a compact. +H(φ, I) = ˆh(I) + R(φ, I) +where ˆh is defined as in 6.1 in the domain Dρ(G),and R(φ, I) analytic on the same +domain. Let ˆu = ∂ˆh +∂I and u = ∂h +∂I . Assume that | ∂ +∂I u|G,ρ2 ≤ M ′ and |u|G ≤ L. Also, +assume that u is non-degenerate: +���� +∂u +∂I v +���� ≥ µ|v| +∀I ∈ G. +Let ˜ +M > M, ˜L > L and ˜µ < µ. Assume u is one-to-one on G and denote F = u(G). +Assume τ > 0, 0 < β ≤ 1 and K given. Assume also that +F ∩ ∆c,ˆq +� +K, β +|k|τ +1 +� += ∅, +∀k ∈ Zn, |k|1 ≤ K, k ̸= 0. +Denote ǫ := ∥DR∥G,ρ,c, η := |R0|G,ρ2 and ξ := +�� ∂R0 +∂I +�� +G,ρ2. +(1) ρ2 ≤ +β +2MKτ+1 +(2) ǫ ≤ min +� +βˆδc +74AKτ , ˜µ2(ρ2−δ2) +4 ˜ +M +� +(3) ξ ≤ min +� +( ˜ +M − M)δ2/R, (µ − ˜µ)ρ2 +� +Then there exists a real canonical transformation +Φ : Dρ− δ +2 (G) → Dρ(G) +and a decomposition H ◦ Φ = ˜ˆh(I) + ˜R(φ, I). Writing ˜u = +∂ +∂I ˜h one has. +(1) |˜u − u|G,ρ2 = ξ, +|˜h − h|G,ρ2 = η, +(2) ˜ǫ := ∥D ˜R∥G,ρ−δ,c ≤ e−Kδ1ǫ + 14AKτ +βˆδc +ǫ2, +(3) ˜η := | ˜R0|G,ρ2− δ2 +2 ≤ 7AKτ +cβ +ǫ2, +(4) |Φ − Id|G,ρ− δ +2 ,c ≤ 2AKτ +β +ǫ, +(5) +�� ∂ +∂I ˜u +�� +G,ρ2 ≤ ˜ +M ′, |˜u|G ≤ ˜L, +(6) | ∂˜u +∂I v| ≥ ˜µ|v| +∀I ∈ G, +(7) Given a subset ˜F ⊂ F − 4Mǫ +˜µ , ˜G(˜u)−1( ˜F) the map ˜u is one-to-one from +˜G to ˜F, ˜G ⊂ G − 2ǫ +˜µ , u( ˜G) ⊃ ˜F − ǫ. Moreover |˜u−1 − u−1| ˜ +F ≤ ǫ/µ. + +68 +6. A NEW KAM THEOREM +Proof. The set u(I) is β/Kτ, K-non-resonant with respect to ω. This implies +that +(6.12) +|k1B(I1)u1 + ¯k¯u + A(I1)u1| ≥ β/Kτ. ≥ +β +|k|τ +1 +≥ β +Kτ . +Then ρ2 ≤ β/Kτ +2MK = +β +2MKτ+1 , ∥DR∥G,ρ,c ≤ β/Kτ ˆδc +74A += +βˆδc +74AKτ . We apply the iterative +lemma (Theorem 6.25) to obtain Φ : Dρ− δ +2 (G) → Dρ(G), such that H ◦ Φ = ˜h + ˜R +where ˜h = h + R0. +We have taken out the points that are not β/Kτ, K-non-resonant with respect +to ω. Because of conditions 1 and 2 we can apply the Iterative lemma. Now let us +prove each of the points in the statement. +(1) We know by definition that ˜u = ∂˜h +∂I = ∂(h+R0) +∂I += ∂h +∂I + R0 +∂I , hence: +|˜u − u|G,ρ2 = |∂h +∂I + ∂R0 +∂I − ∂h +∂I |G,ρ2 = |∂R0 +∂I |G,ρ2 = ξ +˜h = h + R0 ⇒ |˜h − h|G,ρ2 = |h + R0 − h|G,ρ2 = |R0|G,ρ2 = η +(2) By the iterative lemma: +∥D ˜R∥G,ρ−δ,c +≤ +e−Kδ1∥DR∥G,ρ,c + 14A +αˆδc ∥DR∥G,ρ,c +≤ +e−Kδ1ε + 14A +αˆδc ε2 += +e−Kδ1ε + 14AKτ +βˆδc +ε2, +where we have used that α = +β +Kτ . +(3) At this point we use an inequality used in the proof of the iterative Lemma +(theorem 6.25). +| ˜R0|G,ρ2−δ2/2 +≤ +|r2(h, W, 1) + r1(R, W, 1)|G,ρ2−δ2/2 +≤ +7A +αc ∥DR∥2 +G,ρ,c = 7AKτ +β +ε2. +(4) Also using the the iterative Lemma: +|Φ − id|G,ρ−δ/2,c ≤ 2A +α ∥DR∥G,ρ,c = 2AKτ +β +∥DR∥G,ρ,c. +(5) Recall that | ∂ +∂I Aω˜u|G,ρ2−δ2 ≤ ˜ +M, |˜u|G ≤ ˜L, ˜h = h + R0, | ∂ +∂I Aωu|G,ρ2 ≤ +M, |u|G ≤ L. Note that A(I1) ≤ m · maxj(qj)/ minj(cj) and B(I1) ≤ +1/ minj(cj). Hence A(I1) + B(I1) ≤ maxj(qj)/ minj(cj) + 1/ minj(cj) := +R, and we have that |Aω| ≤ R. +| ∂ +∂I Aω˜u|G,ρ2−δ2 += +| ∂ +∂I Aω˜u + ∂ +∂I Aωu − ∂ +∂I Aωu|G,ρ2−δ2 +≤ +| ∂ +∂I Aω(˜u − u)|G,ρ2−δ2 + | ∂ +∂I Aωu|G,ρ2−δ2 +≤ +| ∂ +∂I AωR0|G,ρ2−δ2 + M +≤ +|Aω|G,ρ2|R0|G,ρ +δ2 ++ M +≤ +|Aω|G,ρ2·ξ +δ2 ++ M +≤ +Rξ +δ2 + M +≤ +( ( ˜ +M−M)δ2 +R +)R +δ2 ++ M ≤ ˜ +M − M + M = ˜ +M, +where ξ ≤ ( ˜ +M − M)δ2/R. + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +69 +(6) We know | ∂u +∂I (I)v| ≥ µ|v| for all I ∈ G, then | ∂u +∂I (I)v|G ≥ µ|v|. We want +to find | ∂˜u +∂I (I)v|G ≥ µ′|v| if µ′ < µ. +| ∂˜u +∂I v|G += +|( ∂˜u +∂I + ∂u +∂I − ∂u +∂I )v|G += +|( ∂2R0 +∂I2 + ∂u +∂I )v|G +≥ +−| ∂2R0 +∂I2 v|G + | ∂u +∂I v|G +≥ +µ|v| − | ∂2R0 +∂I2 |G|v| +≥ +µ|v| − | ∂R0 +∂I |G 1 +δ2 |v| +≥ +µ|v| − ξ +ρ2 |v| = (µ − ξ/ρ2)|v| ≥ µ′|v|, +where we have used that | ∂2R0 +∂I2 |G ≤ | ∂R0 +∂I | 1 +ρ2 , and also that µ′ < µ − ξ/ρ2, +hence ξ ≤ (µ − µ′)ρ2. +(7) To apply lemma 6.31 we only need to check that ε ≤ +˜µ2 +˜ +M2 . +˜ +M2 can be +chosen such that | ∂2u +∂I2 |G ≤ ˜ +M2. Note that | ∂2u +∂I2 |G ≤ | ∂2u +∂I2 |G,ρ2−δ2. +| ∂u +∂I |G,ρ2−δ2 ≤ ˜ +M +⇒ +| ∂2u +∂I2 |G,ρ2−δ2(ρ2 − δ2) ≤ | ∂u +∂I |G,ρ2−δ2 ≤ ˜ +M +⇒ +| ∂2u +∂I2 |G,ρ2−δ2 ≤ +˜ +M +ρ2−δ2 = ˜ +M2 +⇒ +| ∂2u +∂I2 |G ≤ ˜ +M2 +Then ε ≤ +˜ +M +4 ˜ +M2 if and only if ε ≤ µ2/(4 +˜ +M +(ρ2−δ2)) if and only if ε ≤ µ2(ρ2−δ2) +4 ˜ +M +which it is assumed in the statement. +□ +6.3. +A KAM theorem on bm-symplectic manifolds +Theorem B ( A bm-KAM theorem). Let G ⊂ Rn, n ≥ 2 be a compact set. Let +H(φ, I) = ˆh(I)+f(φ, I), where ˆh is a bm-function ˆh(I) = h(I)+q0 log(I1)+�m−1 +i=1 +qi +Ii +1 +defined on Dρ(G), with h(I) and f(φ, I) analytic. Let ˆu = ∂ˆh +∂I and u = ∂h +∂I . Assume +| ∂u +∂I |G,ρ2 ≤ M, |u|G ≤ L. Assume that u is µ non-degenerate (| ∂u +∂I v| ≥ µ|v| for some +µ ∈ R+ and I ∈ G. Take a = 16M. Assume that u is one-to-one on G and its range +F = u(G) is a D-set. Let τ > n − 1, γ > 0 and 0 < ν < 1. Let +(1) +(6.13) +ε := ∥f∥G,ρ ≤ +ν2µ2ˆρ2τ+2 +24τ+32L6M 3 γ2, +(2) +(6.14) +γ ≤ min(8LMρ2 +ν ˆρτ+1 , L +K′ ) +(3) +(6.15) +µ ≤ min(2τ+5L2M, 27ρ1L4Kτ+1, βντ+122τ+1ρτ +1), +where ˆρ := min +� +νρ1 +12(τ+2), 1 +� +. Define the set ˆG = ˆGγ := {I ∈ G− 2γ +µ |u(I) is τ, γ, c, ˆq− +Dioph.}. Then, there exists a real continuous map T : W ρ1 +4 (Tn) × ˆG → Dρ(G) an- +alytic with respect the angular variables such that +(1) For all I ∈ ˆG the set T (Tn ×{I}) is an invariant torus of H, its frequency +vector is equal to u(I). + +70 +6. A NEW KAM THEOREM +(2) Writing T (φ, I) = (φ + Tφ(φ, I), I + TI(φ, I)) with estimates +|Tφ(φ, I)| ≤ 22τ+15ML2 +ν2ˆρ2τ+1 +ε +γ2 +|TI(φ, I))| ≤ 210+τL(1 + M) +ν ˆρτ+1 +ε +γ +(3) meas[(Tn ×G)\T (Tn× ˆG)] ≤ Cγ where C is a really complicated constant +depending on n, µ, D, diamF, M, τ, ρ1, ρ2, K and L. +Proof. This proof, as the one in [DG96] is going to be structured in six +sections. First we define the parameters used in each iteration while building the +diffeomorphism. After that, we prove that we can apply the inductive lemma 6.32 +and we exhibit some bound that hold using the results of the inductive lemma. Next, +we find that the sequence of frequency vectors and the sequence of diffeomorphisms +that we built actually converges. Then we find estimates of the components of the +canonical transformation that we have built. Then we find a way to identify the +invariant tori and finally we give a bound for the measure of the set of invariant +tori. +(1) Choice of parameters +We are going to make iterative use of proposition 6.32. So we need +to properly define all the parameters in the statement for every iteration. +Let: + + + +Mq += +(2 − 1 +2q )M, +Lq += +(2 − 1 +2q )L, +µq += +(1 + 1 +2q ) µ +2 . +Note that Mq, Lq monotonically increase from M to 2M and L to 2L +when q → ∞. On the other hand µq monotonically decreases from µ to +µ/2. Also, let: +� K0 += +0, +Kq += +K · qq−1, q ≥ 1, +where K is the minimum natural number greater or equal than 1/ˆρ +and greater or equal than ( +νβ +µ22τ+12 )1/τ. Moreover, β := γ/L ≤ 1, and + + + + + +ρ(q) += +(ρ(q) +1 , ρ(q) +2 ), +ρ(q) +1 += +(1 + +1 +2νq ) ρ1 +4 , +ρ(q) +2 += +νβ +32MKτ+1 +q+1 . +Notice that ρ(q) +1 +decreases monotonically from ρ1/2 to ρ1/4. Also, ρ(q) +2 +decreases to 0. We also denote: + + + + + + + +δ(q) +1 += +ρ(q−1) +1 +− ρ(q) +1 , +δ(q) +2 += +ρ(q−1) +2 +− ρ(q) +2 , +cq += +δ(q) +2 +δ(q) +1 . + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +71 +Note that +δ(q) +1 += +� +1 + +1 +2ν(q−1 +� ρ1 +4 − +� +1 + +1 +2νq +� ρ1 +4 += +� +1 +2ν(q−1) − +1 +2νq +� ρ1 +4 += +1−1/2ν +2ν(q−1) +ρ1 +4 . +Also, since 0 < ν < 1 then ν/2 ≤ 1 − 1/2ν ≤ ν. Plugging this in the +previous equation we obtain: +(6.16) +νρ1 +2ν(q−1)8 ≤ δ(q) +1 +≤ +νρ1 +2ν(q−1)4. +Also, +δ(q) +2 += +νβ +32MKτ+1 +q +− +νβ +32MKτ+1 +q+1 += +νβ +32M(K2q−1)τ+1 − +νβ +32M(K2q)τ+1 += +νβ +32M(K2q−1)τ+1 +� +1 − +1 +2τ+1 +� +. +Also, since τ > 0 then 1/2 ≤ (1 − 1/2τ+1) ≤ 1. Using this in the +previous equation: +(6.17) +νβ +64MKτ+1 +q +≤ δ(q) +2 +≤ +νβ +32MKτ+1 +q +. +Using equations 6.16 and 6.17 we find bounds for cq + + + + + + + + + + + +cq +≤ +� +νβ +32MKτ+1 +q +� +� +νρ1 +2ν(q−1) +� += +β2ν(q−1) +4MKτ+1 +q +ρ1 , +cq +≥ +� +νβ +64MKτ+1 +q +� +� +νρ1 +2ν(q−1)4 +� = +β2ν(q−1) +16MKτ+1 +q +ρ1 . +Then, we also define +� βq += +(1 − +1 +2νq )β, +β′ +q += +βq+βq+1 +2 +. +Observe that both βq and β′ +q tend to β. Also observe that β′ +q ≥ ν +4β, +because: +β′ +q += +βq+βq+1 +2 += +(1− +1 +2νq )+� +1− +1 +2ν(q+1) +� +2 +β += +� +1 − +� 1+ 1 +2ν +2νq +� +1 +2 +� +β +≥ +� +1 − (1 − 1/2ν) 1 +2 +� +β ≥ ν +4β. +As K is the minimal natural number such that K ≥ 1/ˆρ then K ≤ +2/ˆρ. Hence ˆρ ≤ 2 +K . Also +1 +ˆρτ+1 ≥ +�K +2 +�τ+1 +. +Recall that ˆρ = min( +νρ1 +12(τ+2), 1) and, in particular, ˆρ ≤ νρ1 and ˆρ ≤ 1. + +72 +6. A NEW KAM THEOREM +By definition γ ≤ 8LMρ2 +ν ˆρτ+1 . And because β = γ/L: +βL ≤ 8LMρ2 +ν ˆρτ+1 ≤ 8LMρ2Kτ+1 +ν +. +Because we assumed ε ≤ +ν2µ2 ˆρ2τ+2 +24τ+32L6M3 γ2 then, using that γ = Lβ and +ˆρ ≤ 2/K: +(6.18) +ε ≤ ν2µ2 � 2 +K +�2τ+2 +24τ+32L6M 3 ≤ +ν2µ2β2 +24τ+30L4M 3K2τ+2 . +Also using again the assumption that ε ≤ +ν2µ2 ˆρ2τ+2 +24τ+32L6M3 γ2 we want to +prove that +(6.19) +ε ≤ +ν3ρ1β2 +22τ+22MK2τ+1 . +It is enough to check that: +ν2µ2ˆρ2τ+2L2β2 +24τ+32L6M 3 +≤ +ν3ρ1β2 +22τ+22MK2τ+1 +where we used γ = Lβ. Now observing that ˆρ ≤ νρ1 it suffices to see +ν2µ2ν2τ+2ρ2τ+2 +1 +L2β2 +24τ+32L6M 3 +≤ +ν3ρ1β2 +22τ+22MK2τ+1 , +which simplifies to +µ2ρ2τ+2 +1 +24τ+10L4M 2 ≤ +1 +K2τ+1 . +Using that K ≥ 1/(νρ1) is enough to check that +µ2ρ2τ+1 +1 +ν2τ+2 +22τ+12L4M 2 ≤ (νρ1)2τ+1, +which holds if and only if µ ≤ 2τ+5L2M as we assumed. +(2) Induction +Let us take G0 = G. Now the goal is to construct a decreasing se- +quence of compact sets Gq ⊂ G and a sequence of real analytic canonical +transformations +Φ(q) : Dρ(q)(Gq) → Dρ(q−1)(Gq−1), +q ≥ 1. +Denoting Ψ(q) = Φ1 ◦· · ·◦Φ(q) the transformed Hamiltonian functions +will be noted by H(q) = H ◦ Ψ(q) = ˆh(q)(I) + R(q)(φ, I). Moreover, u(q) = +∂h(q) +∂I +and ˆu(q) = ∂ˆh(q) +∂I . +We are going to show that the following bounds hold for all q ≥ 0: +(a) εq := ∥DR(q)∥Gq,ρ(q),cq+1 ≤ +8ε +νρ12(2τ+2)q , +(b) ηq := |R(q) +0 |Gq,ρ(q) +2 +≤ +ε +2(2τ+3)q and ξq := | ∂R(q) +0 +∂I |Gq,ρ(q) +2 +≤ 4MKτ+1ε +νβ2(τ+2)q , +(c) | ∂2h(q) +∂I2 |Gq,ρ(q) +2 +≤ Mq, +|u(q)| ≤ Lq +∀I ∈ Gq, +(d) u(q) is µq-non-degenerate on Gq, + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +73 +(e) u(q) is one-to-one on Gq, and u(q)(Gq) = Fq where we define: +Fq := (F − βq) \ +� +k∈Zn\{0} +|k|1≤K +∆cq,ˆq(K, βq +|k|τ +1 +) +To prove this we proceed by induction. For q = 0: + + + +G0 = G, +h(0) = h, ˆh(0) = ˆh, +R(0) = f. +Using the definitions from the previous point: +� +ρ(0) +1 += +(1 + 1) ρ1 +4 = ρ1/2, +ρ(0) +2 += +νβ +32MKτ+1 ≤ ρ2 +2 , +where in the last inequality we have used that β ≤ +8Mρ2Kτ+1 +ν +and +hence ρ2 ≥ +βν +8MKτ+1 . +Then, +ε0 = ∥Df∥G,ρ(0),c1 = ∥Df∥G,ρ(1)+δ(1). +Now, let us use that |DΥ|G,ρ−δ,c ≤ 2|Υ|G,ρ,c +ˆδc +while having in mind that +ˆδ(1) +c1 = min(c1δ(1) +1 , δ(1) +2 ). Then, +∥Df∥G,ρ(1),c1 ≤ c1|f|G,ρ(0) +ˆδc1 +≤ |f|G,ρ(0) +δ(1) +1 +≤ |f|G,ρ(0)8 +νρ1 += 8ε +νρ1 +, +where we have used δ(1) +1 +≥ +νρ1 +8·2ν(1−1) = νρ1 +8 . +This proves the first step of +the induction for 2a). +Let us prove now the base case for 2b). On one side η0 = |R(0) +0 |G0,ρ2(0) ≤ +ε +2(2τ+3)0 = ε, which holds because |R(0) +0 |G0,ρ(0) +2 +≤ |R(0)|G,ρ(0) = |f|G,ρ(0) = +ǫ. On the other hand ξ0 = | ∂R(0) +0 +∂I |G0,ρ(0) +2 +≤ | ∂R(0) +0 +∂I |G0,ρ2−ρ2/2 ≤ +1 +ρ2/2∥R0∥G,ρ ≤ +2ε +ρ2 ≤ +ε +ρ(0) +2 , where we used that ρ2(0) ≤ ρ2/2 = ρ2 − ρ2/2. +The base case of 2c) is immediate because | ∂2h(0) +∂I2 |G0,ρ(0) +2 +≤ | ∂2h +∂I2 |G,ρ2 = +M = M0 and also |u(0)|G0 = |u|G ≤ L = L0. +The base case of 2d holds because u(0) = u is µ non-degenerate in +G = G0. +The base case of 2e holds because u(0) = u is one-to-one in G0 = G +by hypothesis. u(0)(G0) = F0 where F0 = (F − β0) \ {∅} = F because +K0 = 0 and β0 = 0. +For q ≥ 1, we assume the statements hold for q − 1 and we prove +it for q. Let us apply proposition 6.32 (Inductive Lemma) to H(q−1) = +hq−1 + Rq−1 with Kq instead of K. + +74 +6. A NEW KAM THEOREM +We have to be careful with the condition F ∩ ∆c,ˆq(k, +β +|k|τ +1 ) = ∅ ∀k ∈ +Zn, |k|1 ≤ Kq, k ̸= 0 and with the definition +Fq−1 := (F − βq−1) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, βq−1 +|k|τ +1 +), +because the resonances have to be removed up to order Kq, not Kq−1. +Let us define +F ′ +q−1 := (F − βq−1) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, β′ +q−1 +|k|τ +1 +) +where we simply replaced βq−1 for β′ +q−1 because ∆c,ˆq(k, βq−1 +|k|τ +1 ) makes +no sense when q = 1, βq−1 = 0. +Accordingly let us define G′ +q−1 := (u(q−1))−1(F ′ +q−1). The conditions in +proposition 6.32 are going to be satisfied with F ′ +q−1, β′ +q−1, Kq, Mq−1, Lq−1, +µq−1,ρ(q−1), δ(q), cq, Mq, Lq, µq replacing F, β, K, M, L, µ, ρ, δ, c, ˜ +M, ˜L, ˜µ. +And also a = 16M ≥ 8Mq. +We are now going to check that 1, 2 and 3 are satisfied so we can +apply proposition 6.32. +– 1 We want to see that ρ(q−1) +2 +≤ +β′ +q−1 +2MqKτ+1 +q +. +By definition ρq−1 +2 += +νβ +32MKτ+1 +q +≤ +4β′ +q−1 +32MKτ+1 +q +≤ +β′ +q−1 +8MqKτ+1 +q +≤ +βq−1 +2MqKτ+1 +q +, where we used that +Mq ≥ M. +– 2 We want to see that εq−1 ≤ min +� +βq−1 ˆρ(q) +c +74AqKτ +q−1 , +µτ +q (ρ(q−1) +2 +−δ(q−1) +c +) +4Mq +� +, +where Aq := 1 + +2Mq−1cqKτ +q +β′ +q−1 +. +Notice that: +Aq +:= +1 + +2Mq−1cqKτ +q +β′ +q−1 +≤ +1 + +8Mq−1cqKτ +q +νβ +≤ +1 + +8Mq−1β2ν(q−1)Kτ +q +4MKτ+1 +q +ρ1νβ += +1 + 2Mq−12ν(q−1) +MKqρ1ν += +1 + 2Mq−12ν(q−1) +MK2q−1ρ1ν +≤ +1 + +4M2q−1 +MK2q−1ρ1ν += +1 + +4 +Kρ1ν ≤ 1 + 4 = 5 +First, we check that εq−1 ≤ +βq−1 ˆρ(q) +c +74AqKτ +q−1 . +By induction hypothesis we know that εq−1 ≤ +8ε +νρ12(2τ+2)(q−1) . Hence +it is enough to see +8ε +νρ12(2τ+2)(q−1) ≤ β′ +q−1δ(q) +2 +75 · 5Kτq +. + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +75 +Notice that +β′ +q−1δ(q) +2 +379Kτq +≥ +νβ +4 +νβ +64MKτ+1 +q +1 +37Kτq += +ν2β2 +4·64·370 +1 +MK2τ+1 +q += +ν2β2 +4·64·370·M2(q−1)(2τ+1)K2τ+1 . +And this holds if the following is true: +8ε +νρ1 +≤ +ν2β2 +4 · 64 · 379 · K2τ+1M ⇔ ε ≤ +ν3β2ρ1 +212135MK2τ+1. +This is true because in the previous section we have seen that ε ≤ +ν3ρ1β2 +22τ+30MK2τ+1 +Let us now prove that ε ≤ +µ2 +q(ρ(q−1) +2 +−δ(q−1) +2 +) +2Mq +. First of all observe that +(ρ(q−1) +2 +− δ(q) +2 ) = ρ(q) +2 . So, what we want to prove is equivalent to +proving εq−1 ≤ +µ2 +qρ(q) +2 +2Mq . +On the other hand, we know that εq−1 ≤ +8ε +νρ12(2τ+2)(q−1) . And observe +also that +µ2 +qρ(q) +2 +2Mq +≥ +(µ/2)2 +νβ +32MKτ+1 +2M += +µ2νβ +28M2Kτ+1 . +If are able to check that +8ε +νρ12(2τ+2)(q−1) ≤ +µνβ +28M2Kτ+1 we would be fine. +The previous equation holds if and only if the following holds, +ε ≤ µν2βρ12(2τ+2)(q−1) +211M 2Kτ+1 +. +If we knew beforehand that ε ≤ µν2βρ12−(2τ+2) +211M2Kτ+1 += +µν2βρ1 +22τ+13M2Kτ+1 we +would be done. +But we also know that +ǫ ≤ +ν2µ2β2 +22τ+30L4M 3K2τ+2 . +Then it is enough to check that +ν2µ2β2 +22τ+30L4M 3K2τ+2 ≤ +µν2βρ1 +22τ+13M 2Kτ+1 . +And this holds because µ ≤ 27ρ1L4Kτ+1 +– 3 Lastly we want to see that +ξq−1 ≤ min((Mq − Mq−1)δ(q) +2 +R , (µq−1 − µq)ρ(q−1) +2 +). +Observe that R does not depend on q because at each iteration ˆh(q) +singular part is not modified. ˆh(q) = ˆh(q) + R(q) +0 +and R0 is analytic +depending only on the action coordinates. By induction hypothesis, +we know that +ξq−1 = |∂R(q) +0 +∂I +|Gq,ρ(q) +2 +≤ 4MKτ+1ε +νβ2(τ+2)q . +We are going to check the two different inequalities separately + +76 +6. A NEW KAM THEOREM +(a) ξq−1 ≤ (Mq − Mq−1) δ(q) +2 +R . Note that Mq = (2 − +1 +2q )M, then +Mq − Mq−1 = M +2q . +δ(q) +2 +≥ +νβ +64M(K2q−1)τ+1 ≥ +νβ +64M(K2q)τ+1 += +νβ +64MKτ+1 +1 +2qτ+q . +We deduce +(Mq − Mq−1)δ(q) +2 +≥ +νβ +64Kτ+1 +1 +2τq+2q . +Hence we only need to check that +4MKτ+1ε +νβ2(τ+2)q ≤ +νβ +64Kτ+1 +1 +2τq+2q . +The previous condition holds if and only if +4MKτ+1ε +νβ +≤ +νβ +26Kτ+1 ⇔ ε ≤ +ν2β2 +2K2τ+2M . +On the other hand, let us use again that ε ≤ +ν2µ2β2 +22τ+30L4M3K2τ+2 . +If we apply the condition µ ≤ 2τ+6L2M in the last expression +we obtain: +ε ≤ ν2β222τ+12L4M 2 +22τ+30L4M 3K2τ+2 = +ν2β2 +28K2τ+2M . +(b) ξq−1 ≤ (µq−1 − µq)ρ(q−1) +2 +. +Observe that +µq = (1 + 1 +2q )µ +2 , +(µq−1 − µq) = ((1 + +1 +2q−1 ) − (1 + 1 +2q ))µ +2 = ( +1 +2q−1 − 1 +2q )µ +2 += +�2 − 1 +2q +� µ +2 = 1 +2q +µ +2 = +µ +2q+1 +Also, +ρ(q−1) +2 += +νβ +32MKτ+1 +q += +νβ +32M(K2q−1)τ+1 +≥ +νβ +32MKτ+12q(τ+1) . +Then, +(µq−1 − µq)ρ(q−1) +2 +≥ +µ +2q+1 +νβ +32MKτ+12q(τ+1) . +Then we only have to check that +4MKτ+1ε +νβ2τq+2q−2 +≤ +µ +2q+1 +νβ +32MKτ+12q(τ+1) += +µ +2τq+2q+1 +νβ +32MKτ+1 . +Which holds if and only if + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +77 +MKτ+1ε +νβ2−2 +≤ µ +2 +νβ +32MKτ+1. +Then, +ε ≤ µ2−2 +2 +ν2β2 +32M 2K2τ+2 = +µν2β2 +28M 2K2τ+2 . +But we know +ε +≤ +ν2µ2β2 +2τ+30L4M3K2τ+2 +≤ +ν2µ2τ+5L4Mβ2 +2τ+30L4M3K2τ+2 += +ν2µβ2 +225M2K2τ+2 +≤ +µν2β2 +28M2K2τ+2 +as we wanted. +In the second inequality we used that µ ≤ +2τ+5L4M. +So, finally, we can apply the inductive lemma 6.32 with the pa- +rameters mentioned previously in this section. Hence we obtain a +canonical transformation Φ(q) and a transformed hamiltonian H(q) = +h(q) + R(q). The new domains Gq ⊂ G′ +q−1 are going to be specified +in the following lines. So now we are going to prove 2a,2b,2c,2d,2e. +– 2a. We want to see εq := ∥DR(q)∥Gq,ρ(q),cq+1 ≤ +8ε +νρ12(2τ+2)q. +By the second result of proposition 6.32 we have: +(6.20) +εq ≤ e−Kqδ(q) +1 εq−1 + 14AqKτ +q +β′ +q−1δ(q) +2 +ε2 +q−1. +Now we are going to bound each term of the right hand of the +expression at a time. +Recall that δ(q) +1 +≥ +νρ1 +82ν(q−1) . +Kqδ(q) +1 +≥ +K2q−1 νρ1 +8 2−ν(q−1) += +νρ1 +8 K2(1−ν)(q−1) +≥ +12(τ+2) +8 +2(1−ν)(q−1) += +(3/2τ + 3)2(1 − ν)(q − 1) +≥ +(2τ + 3) 3 +4 ≥ (2τ + 3) ln 2, +where we used that K ˆρ ≥ 1 and hence K ≥ 12(τ+2) +νρ1 +. So we +conclude that e−Kqδ(q) +1 +≤ +1 +22τ+3 , and we have bounded the first +term of 6.20. Let us bind the second one. +On one hand, we have that +14AqKτ +q +β′ +q−1 +≤ 14 · 5Kτ +q +νβ +4 +≤ 29K2 +q +νβ +where we have used that β′ +q ≥ νβ +4 and Aq ≤ 5. + +78 +6. A NEW KAM THEOREM +Now we are going to apply that εq−1 ≤ +8ε +νρ12(2τ+2)(q−1) , δ(q) +2 +≥ +νβ +64MKτ+1 +q +and ǫ ≤ +ν3ρ1β2 +22τ+22MK2τ+1 to obtain +14AqKτ +q +β′ +q−1δ(q) +2 εq−1 += +14AqKτ +q +β′ +q−1 +1 +δ(q) +2 εq−1 +≤ +29Kτ +q +νβ +64MKτ+1 +q +νβ +8ε +νρ12(2τ+2)(q−1) +≤ +218MK2τ+1 +q +ν3β2ρ12(2τ+2)(q−1) +ν3ρ1β2 +22τ+22MK2τ+1 +≤ +2182(q−1)(2τ+1)−(2τ+2)(q−1)−(2τ+22) += +2(1−q)2−2τ−4 = +1 +22τ+32q−1 . +This gives us the bound of the second term of 6.20. Now we +put both bounds together: +εq ≤ +1 +22τ+3 εq−1 + +1 +22τ+3 +1 +2q−1 εq−1 ≤ +1 +22τ+2 εq���1. +That implies εq ≤ +ǫ +2(2τ+2)(q−1) as we wanted. Because we can +assume νρ1 ≤ 1. +– 2b +Let us write σ(q) +2 += ρ(q−1) +2 +− δ(q) +2 /2 = ρ(q) +2 ++ δ(q) +2 /2 ≥ ρ(q) +2 , then +ηq = |R(q) +0 |Gq,ρ(q) +2 +≤ |R(q) +0 |Gq,σ(q) +2 . +By the inductive lemma 6.32: +ηq +≤ +7AqKτ +q +cqβ′ +q−1 ε2 +q−1 +≤ +7AqKτ +q +β′ +q−1 ε2 +q−1 +δ(q) +1 +δ(q) +2 += +14AqKτ +q +β′ +q−1δ(q) +2 ε2 +q−1 +δ(q) +1 +2 +≤ +1 +22τ+32q−1 εq−1 +δq) +1 +2 +≤ +1 +2 +δ(q) +1 +22τ+32q−1 +ε +νρ12(2τ+2)(q−1) +≤ +1 +2 +νρ1 +4·2ν(q−1) +1 +22τ+32q−1 +8ε +νρ12(2τ+2)(q−1) +≤ +ε +2(2τ+3)q . +For the second part we only need to apply Cauchy inequalities: +ξq ≤ +2 +δ(q) +2 +|Rq +0|Gq,ρ(q) +2 +≤ +2 +δ(q) +2 +ε +2(2τ+3)q . +– 2c and 2d are direct from lemma 6.32. +– 2e We need to consider again the results from lemma 6.32 +with Fq as F. We have to check the condition Fq ⊂ F ′ +q−1 − +4Mq−1εq−1 +µq +. Let us define dq := βq−βq−1 +2Kτ+1 +q +. Using that F ′ +q−1 := +(F − βq−1) \ � +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k. +β′ +q−1 +|k|τ +1 ) we have +F ′ +q−1 − dq ⊃ (F − (βq−1 + dq)) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, β′ +q−1 +|k|τ +1 ++ |k|dq). + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +79 +Moreover, + + + + + + + + + + + +βq−1 + dq +≤ +βq, and +β′ +q−1 +|k|τ +1 + |k|dq += +β′ +q−1+|k|τ +1 |k| +βq−βq−1 +2Kτ+1 +q +|k|τ +1 +≤ +β′ +q−1+Kτ+1 +q +βq−βq−1 +2Kτ+1 +q +|k|τ +1 += +β′ +q−1+ +βq +2 − +βq−1 +2 +|k|τ +1 += +βq +|k|τ +1 . +Now if we see that 4Mq−1εq−1 +µq +≤ dq we will have the inclusion +we want. Observe that 4Mq−1 +µq +≤ 4·2M +µ/2 = 16M +µ . So, it is enough +to check that 16M +µ εq−1 ≤ dq. +εq−1 +≤ +8ε +νρ12(2τ+2)q +≤ +8ν3ρ1β2 +νρ12(2τ+2)q2(2τ+22)MK2τ+1 +≤ +8ν2β2 +2(2τ+2)q+2τ+20MK2τ+1 +≤ +8ν2β +2(βq−βq−1) +ν +2(τ+1)q+(τ+1)q+2τ+20MK2τ+1 += +8νβ2(βq−βq−1) +2(τ+1)+(τ+1)q+2τ+20MKτ+1 +q +Kτ += +8νβ2 +2(τ+1)+(τ+1)q+2τ+19MKτ +(βq−βq−1) +2Kτ+1 +q += +νβ +2(τ+1)+(τ+1)q+2τ+15MKτ dq +≤ +νβ +23τ+16MKτ dq. +Hence, it is enough to prove the following: +16M +µ +νβ +23τ+16MKτ dq ≤ dq. +Wich holds if and only if +16M +µ +νβ +2τ+16MKτ ≤ 1 ⇔ Kτ ≥ +νβ +µ2τ+12 , +which we assumed when choosing K. +(3) Convergence of diffeomorphisms +Now we are going to prove the convergence of the successive maps +u(q) : Gq → Fq +i.e. we want to see that exist proper sets G∗, F ∗ and an analytical +map u∗ such that u(q) : Gq → Fq converge to u∗ : G∗ → F ∗. +Let us use lemma 6.32 as before. +For q ≥ 1 we obtain +|u(q) − u(q−1)|Gq ≤ ξq +and +|(u(q))−1 − (u(q−1))−1|Fq ≤ εq +µq +. +Now, because the following two inequalities hold +� +ξq +≤ +4MKτ+1ε +νβ2(τ+2)q +εq +µq +≤ +8ε +νρ122τ+2q +1 +(1+ 1 +2q ) µ +2 = +8ε2q−1 +νβ2(2τ+2)q(2q+1)µ + +80 +6. A NEW KAM THEOREM +the sequences uq and (u(q))−1 converge to maps u∗ and Υ respectively. +These maps are defined on the following sets: +G∗ +:= +� +q≥0 Gq, +F ∗ +:= +� +q≥0 Fq = (F − β) \ � +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, +β +|k|τ +1 ). +The second equality holds because F ∗ is a compact for being the +intersection of compact sets. We can now deduce that +|u∗ − u(q)|G∗ +≤ +� +s≥q |u(q) − u(q−1)|G∗ +≤ +� +s≥q |u(q) − u(q−1)|G +≤ +� +s≥q ξq. +with the same argument we see that |Υ−(u(q))−1|F ∗ ≤ . . . ≤ � +s≥q +εq +µq . +The next steps are going to be to prove that Gq ⊂ Gq−1 − 2εq−1 +µq−1 and +Fq ⊂ Fq−1 − 4Mq−1εq−1 +µq−1 +. If we check it and take the limit we would have: +G∗ ⊂ Gq − +� +s≥q +2εq +µq +and +F ∗ ⊂ Fq − +� +s≥q +4Mqεq +µq +. +Let us first check Fq ⊂ Fq−1 − 4Mq−1εq−1 +µq−1 +. Let us define x := 4Mq−1 +µq−1 . +Fq−1 − x +⊃ +(F − (βq−1 + x)) \ � +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, βq−1 +|k|τq + |k|x) +⊃ +(F − (βq−1 + x)) \ � +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, +βq−1+Kτ+1 +q +x +|k|τq +). +To have the inclusion we want, we have to check that: +(a) βq−1 + x ≤ βq. +(b) +βq−1+Kτ+1 +q +x +|k|τ +1 +≤ +βq +|k|τ +1 ⇔ βq−1 + Kτ+1 +q +x. +Since the second one implies the first we will only check the second +one. +βq−1 + Kτ+1 +q +x += +βq−1 + Kτ+1 +q +4Mq−1εq−1 +µq−1 +≤ +βq−1 + Kτ+1 +q +16Mεq−1 +µ +≤ +βq−1 + Kτ+1 +q +dq += +βq−1 + Kτ+1 +q +βq−βq−1 +2Kτ+1 +q += +βq−1 − βq−1/2 + βq/2 += +βq−1+βq +2 += +βq +Where we have used that 16Mεq/µ ≤ dq and that βq is monotonically +increasing with q. +The inclusion Gq ⊂ Gq−1 − 2εq−1 +µq−1 is given as a result of the lemma +6.32. +So we proved what we wanted. We are now going to see that u∗ is +one-to-one on G∗ and that u∗(G∗) = F ∗. + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +81 +Tale I ∈ G∗, we have that u(q)(I) ∈ Fq for every q. Hence u∗(I) ∈ F ∗, +and we deduce that u∗(G∗) ⊂ F ∗. +With the same argument, we see +Υ(F ∗) ⊂ G∗. Let us prove that Υ(u∗(I)) = I. +|Υ(u∗(I)) − I| +≤ +|Υ(u∗(I)) − (u(q))−1(u∗(I)) ++(u(q))−1(u∗(I)) − (u(q))−1(u(q)(I))| +≤ +|Υ(u∗(I)) − (u(q))−1(u∗(I))| ++|(u(q))−1(u∗(I)) − (u(q))−1(u(q)(I))| +≤ +|Υ − (u(q))−1|F ∗ + +1 +µq |u∗ − u(q)|G∗. +Where to bound the second term we used the mean value theorem, +i.e. |u(q)(x) − u(q)(y)|Gq ≤ | ∂ +∂I u(q)|Gq|x − y|, and the fact that because +of the µq-non-degeneracy, | ∂u(q) +∂I | ≥ µq|v|, ∀v ∈ Rn and ∀I′ ∈ Gq. Note +that we can use the mean value theorem because u∗(I) − u(q)(I) belongs +to Fq because 4Mqεq +µq +≥ ξq. Let us prove this inequality. If we want to see +4Mqεq +µq +≥ ξq, it is enough to see 4Mεq +µ +≥ ξq. +ξq +≤ +2 +δ(q) +2 |R(q) +0 |Gq,σ(q) +2 +≤ +2 +δ(q) +2 +δ(q) +1 +εq−1 +2 +1 +22τ+32q−1 += +1 +cq +1 +22τ+22q−1 εq−1 ≤ 4M +µ εq−1 +The last inequality holds true if and only if +µ +≤ +β2ν(q−1)22τ+32q−1 +Kτ+1 +q +ρ14 += +β2ν(q−1)22τ+32q−1 +Kτ+12(τ+1)(q−1)ρ14 +≤ +β22τ+3 +Kτ+1ρ14 += +β22τ+1 +Kτ+1ρ1 +≤ +β22τ+1 +( +1 +νρ1 )τ+1ρ1 = βντ+122τ+1ρt +1au +as we assumed in the statement of the theorem. +Since the bound +obtained tends to 0, we have Υ(u∗(I)) = I and hence u∗ is one-to-one. +Analogously we obtain u∗(Υ(J)) = J +∀J ∈ F ∗. Finally, u∗ is one-to-one +and u∗(G∗) = F ∗. Note also that from the inductive lemma we obtain +|h(q) − h(q−1)|Gq,ρ(q−1) +2 +≤ ηq−1. Also, observe the following bound that we +are going to use in the next sections. +|u∗ − u(q)|G∗ ≤ +� +s≥q +4MKτ+1ε +νβ2(τ+2)s . +(4) Convergence of the canonical transformations +Let σ(q) = ρ(q−1) − δ(q) +2 /2. Observe that this definition implies that +σ(q) − ρ(q) = δ(q) +2 +and σ(q) − δ(q) +2 += ρ(q). +Observe that applying the +inductive lemma 6.32: +|Φ(q) − id|Gq,σ(q),cq +≤ +2Aq−1Kτ +q +β′ +q−1 +εq−1 + +82 +6. A NEW KAM THEOREM +≤ +2·5·4 +νβ +8ε +νρ12(2τ+2)(q−1) +≤ +29Kτ ε +ν2ρ1β2(τ+2)(q−1) +≤ +29Kτν3ρ1β2 +ν2ρ1β2(τ+2)(q−1)22τ+22MK2τ+1 +≤ +29νβ +2(τ+2)(q−1)22τ+20MKτ+1 += +νβ +26M(K2q−1)τ+1 +29 +2(q−1)22τ+14 +≤ +δ(q) +2 +1 +2(q−1)22τ+5 +≤ +δ(q) +2 +2(q−1)32, +where we have used that δ(q) +2 +≥ +νβ +84MKqP τ+1, ε ≤ +ν3ρ1β2 +22τ+20MK2τ+1 , β ≤ +8MKτ+1ρ2 +ν +and β′ +q−1 ≥ νβ +4 . +Now, recall that ˆδc = min(cδ1, δ2), then ˆδcq = min(cqδ(q) +1 , δ(q) +2 ) = +min(δ(q) +2 , δ(q) +2 ) = δ(q) +2 . +Now using that |DΥ|G,ρ−δ,c ≤ |Υ|G,ρ,c +ˆδc +, we can obtain: +|DΦ(q) − Id|Gq,ρ(q),cq += +|D(Φ(q)) − id|Gq,ρ(q),cq +≤ +|D(Φ(q)) − id|Gq,σ(q)−δ(q) +2 +,cq +≤ +|Φ(q)−id|Gq,σ(q),cq +ˆδcq +≤ +|Φ(q)−id|Gq,σ(q),cq +δ(q) +2 +≤ +2|Φ(q)−id|Gq,σ(q),cq +δ(q) +2 +≤ +2 +δ(q) +2 +δ(q) +2 +2(q−1)·32 ≤ +1 +2q−116 ≤ +1 +2(q−1)4 +Let x, y be such that the segment joining them is contained in Dρ(q)(Gq). +Using the mean value theorem one can deduce the following bound: +|Φq(x) − Φq(y)|cq ≤ |DΦ(q)|Gq,ρ(q),cq · |x − y|cq. +By 22, in particular |Φ(q)(x)−x|cq ≤ δq +2 and |Φ(q)(y)−y|cq ≤ δq +2. Then +the segment that join Φ(q)(x) and Φ(q)(y) is contained in Dρ(q−1)(Gq−1) = +Dρ(q)+δ(q), because Gq ⊂ Gq−1 − 2εq−1 +µq−1 and because ρ(q) − ρ(q−1) ≤ δ(q) +2 +because ρ(q) − ρ(q−1) = δ(q) +2 . +Therefore we can apply the mean value theorem once again: +|Φ(q−1)(Φ(q)(x)) − Φ(q−1)(Φ(q)(y))|cq−1 +≤ |DΦ(q−1)|Gq−1,ρq−1,cq−1|Φ(q)(x) − Φ(q)(y)|cq−1 +≤ 2τ+1−ν|DΦ(q−1)|Gq−1,ρq−1,cq−1|Φ(q)(x) − Φ(q)(y)|cq, +where we have used that cq−1/cq = +δ(q−1) +2 +/δ(q−1) +1 +δ(q) +2 +/δ(q) +1 += +δ(q−1) +2 +δ(q) +2 +δ(q) +1 +δ(q−1) +1 += +2τ+1 1 +2ν = 2τ+1−ν. +Using the previous bounds and iterating by q, we obtain the following: +|Ψ(q)(x) − Ψ(q)(y)|c1 + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +83 +≤ 2(τ+1−ν)(q−1)|DΦ(1)|G1,ρ(1),c1 · . . . · |DΦ(q)|Gq,ρ(q),cq|x − y|cq +≤ 2(τ+1−ν)(q−1)(1 + 1 +4)(1 + +1 +4·2) · . . . · (1 + +1 +4·2q−1 )|x − y|cq +≤ 2(τ+1−ν)(q−1)e1/2|x − y|cq ≤ 2(τ+1−ν)(q−1) · 2|x − y|cq. +Which holds for q ≥ 1 and for every x, y such that the segment joining +them is contained in Dρ(q)(Gq). Now, given q ≥ 2 and x ∈ Dρ(q)(Gq) let +y = Φ(q)(x): +|Ψ(q)(x) − Ψ(q−1)(x)|c1 += +|Ψ(q−1)(Φ(q)(x)) − Ψ(q−1)(x)|c1 +≤ +2(τ+1+ν)(q−2)2|Φ(q)(x) − x|cq−1 +≤ +2(τ+1+ν)(q−1)2|Φ(q)(x) − x|cq +≤ +2(τ+1+ν)(q−1)2δ(q) +2 +≤ +2(τ+1+ν)(q−1)2 +28Kτε +ν2ρ1β2(τ+2)(q−1) += +29Kτε +ν2ρ1β2(1+ν)(q−1) . +Which holds even for q = 1 by setting Ψ(0) = id by 22. Hence 25 +implies that Ψ(q) converges to a map +Ψ∗ : D(ρ1/4,0)(G∗) = W ρ1 +4 (Tn) × G∗ → Dρ(G). +And we deduce for every q ≥ 0 that +|Ψ∗ − Ψ(q)|G∗,( ρ1 +4 ,0),c1 ≤ +210Kτε +ν2ρ1β2(1+ν)q . +Moreover by taking the limit to the equation +H ◦ Ψ(q) = h(q) + R(q) +we see that H ◦ Ψ∗ = h∗(I) on D( ρ1 +4 ,0)(G∗). +(5) Stability estimates +Next we see that for q → ∞, the motions associated to the trans- +formed hamiltonian ˆH(q) = ˆh(q) + R(q) and the quasi-periodic motions of +ˆh(q) get closer and closer. +Let us denote +� x(q)(t) = (φ(q)(t), I(q)(t)) +the trajectory of H(q), +ˆx(q)(t) = (ˆφ(q)(t), ˆI(q)(t)) +the trajectory of ˆH(q) +corresponding to a given initial condition x(q)(0) = x∗ +0 = (φ∗ +0, I∗ +0) ∈ +Tn × Gq. Let +� +˜x(q)(t) +:= +(˜φ(q)(t), I∗ +0) = (φ∗ +0 + u(q)(I∗ +0))t, I∗ +0 , +ˆ˜x(q)(t) +:= +(ˆ˜φ(q)(t), I∗ +0) = (φ∗ +0 + u′(q)(I∗ +0))t, I∗ +0 +the corresponding trajectories of the integrable parts of h(q) and ˜h(q) +respectively. Recall that ˆh(q)(I) = h(q)(I)+ζ(q)(I1) = h(q)(I)+q0 log(I1)+ +�m−1 +i=1 qi 1 +Ii +1 and u′(q) = ¯Bu(q) + ¯ +A(I1). It is clear that ˜x(q)(t) and ˆ˜x(q)(t) +are defined for all t ∈ R. +Let us denote: + +84 +6. A NEW KAM THEOREM +Tq = inf{t > 0 : |I(q)(t)−I∗ +0| > δ(q+1) +2 +or |φ(q)(t)−˜φ(q)(t)|∞ > δ(q+1) +1 +}. +ˆTq = inf{t > 0 : |ˆI(q)(t) − I∗ +0| > δ(q+1) +2 +or |ˆφ(q)(t) − ˆ˜φ(q)(t)|∞ > δ(q+1) +1 +}. +Observe that x(q)(t) and ˆx(q)(t) are defined and belong do Dρ(q)(Gq), +for 0 ≤ t ≤ Tq and 0 ≤ t ≤ ˆTq respectively, because δ(q) ≤ ρ(q). Also +recall the Hamiltonian equations. Let us first state the motion equations +for our Hamiltonian function ˆH(q): +ιX ˆ +H(q) ω = d ˆH(q), +or +X ˆ +H(q) = Π(d ˆH(q), ·). +Let us write +X ˆ +H(q) = ˙ˆI(q) +1 +∂ +∂I1 ++ . . . ˙ˆI(q) +n +∂ +∂In ++ ˙ˆφ(q) +1 +∂ +∂φ1 ++ . . . + ˙ˆφ(q) +n +∂ +∂φn +. +Moreover +d ˆH(q) += +dˆh(q) + dR(q) += +dζ(q) + dh(q) + dR(q) += +�n +i=1 +∂ζ(q) +∂Ii + +n +� +i=1 +∂ζ(q) +∂φi +� +�� +� +=0 ++ �n +i=1 +∂h(q) +∂Ii ++ +n +� +i=1 +∂h(q) +∂φi +� +�� +� +=0 ++ �n +i=1 +∂R(q) +∂Ii ++ �n +i=1 +∂R(q) +∂φi . +Recall +ω = + + +m +� +j=1 +cj +Ij +1 + + dI1 ∧ dφ1 + +n +� +i=2 +dIi ∧ dφi, +Π = +1 +��m +j=1 +cj +IJ +1 +� ∂ +∂I1 +∧ +∂ +∂φ1 ++ +n +� +i=2 +∂ +∂Ii +∧ +∂ +∂φi +. +Then: + + + + + +˙ˆI(q) +j += +− ∂R(q) +∂φj (ˆx(q)(t)), +if j ̸= 1 and +˙ˆI(q) +1 += +− +1 +��m +i=1 +ci +Ii +1 +� ∂R(q) +∂φj (ˆx(q)(t)) = −B(I1) ∂R(q) +∂φ1 (ˆx(q)(t)). +Observe that +(6.21) +| ˙ˆI(q) +1 (t)| ≤ +���� +∂R(q) +∂φ1 +(ˆx(q)(t)) +���� . +Moreover, + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +85 + + + + + + + + + + + + + +˙ˆφ(q) +j += +ˆu(q) +j (ˆI(q)(t)) + ∂R(q) +∂Ij (ˆx(q)(t)) += +u(q) +j (ˆI(q)) + ∂R(q) +∂Ij (ˆx(q)(t)) +if j ̸= 1 +˙ˆφ(q) +1 += +(B(I1)u(q) +1 ++ A(I1)) +� +�� +� +u(q) +1 +(ˆI(q)(t)) + B(I1) ∂R(q) +∂I1 (ˆx(q)(t)), +where we have used that ˆu(q) +j += u′(q) +j +if j ̸= 1. Using 6.21 we obtain +| ˙ˆI(q)(t)| ≤ +���� +∂R(q) +∂φ +���� +Gq,ρ(q) +≤ εq. +Hence, +| ˙ˆφ(q) − u′(q)(I∗ +0 )|∞ += +|u′(q)(ˆI(q)(t)) + ¯B ∂R(q) +∂I1 (ˆx(q)(t)) − u′(q)(I∗ +0)|∞ +≤ +|u′(q)(ˆI(q))(ˆI(q)(t)) − u′(q)(I∗ +0)|∞ + | ∂R(q) +∂I (ˆx(q)(t))|∞ +≤ +M ′ +q|ˆI(q)(t) − I∗ +0| + ∥ ∂R(q) +∂I ∥Fq,ρ(q),∞ +≤ +Mq|ˆI(q)(t) − I∗ +0| + +εq +cq+1 +≤ +2Mδ(q+1) +2 ++ +εq +cq+1 ≤ 3Mδ(q+1) +2 +. +Where in the last bound we used that +(6.22) +εq +cq+1 +≤ Mδ(q+1) +2 +, +that holds because: +εq +cq+1 +≤ +16MKτ+1 +q+1 ρ1 +β2νq +8ε +νρ12(2τ+2)q +≤ +16MKτ+1 +q+1 ρ1 +βeνq +8 +νρ12(2τ+2)q +ν2µ2β2 +2τ+30L4M2K2τ+2 +≤ +27Kτ+1 +q+1 νµ2β +2(2τ+2)q+νq+τ+30L4M2K2τ+2 +≤ +νβµ2 +Kτ+1 +q+1 2νq+τ+23L4M2 += +µ2 +2νq+τ+17L4M +νβ +26MKτ+1 +q+1 +≤ +µ2 +2τ+17+νqL4M δ(q+1) +2 +≤ +22τ+12L4M2 +2τ+17+νqL4M δ(q+1) +2 +≤ +2τ−5−νqδ(q+1) +2 +≤ +2τ +25+νq Mδ(q+1) +2 +≤ +Mδ(q+1) +2 +if q is large enough. +Thus, since one of the inequalities defining ˆTq has to be an equality +for t = Tq we obtain, +δ(q+1) +2 += +|ˆI(q)(Tq) − I∗ +0| ≤ Tqεq, +or +δ(q+1) +1 += +|ˆφ(q)(Tq) − ˆ˜φ(q)(T1)|∞ ≤ Tq3Mδ(q+1) +2 +. +Hence, ˆTq ≥ min( δ(q+1) +2 +εq +, +δ(q+1) +1 +3Mδ(q+1) +2 +) ≥ +1 +3Mcq+1 , where we used again +6.22. + +86 +6. A NEW KAM THEOREM +Let us denote T ′ +q := +1 +3Mcq+1 , then ˆTq ≥ T ′ +q. This implies +|ˆx(q)(t) − ˆ˜x(q)(t)|cq+1 ≤ δ(q+1) +2 +for |t| ≤ T ′ +q. +Since ˆH(q) = ˆH◦Ψ(q) and Ψ(q) is canonical it turns out that Ψ(q)(ˆx(q)(t)) +is a trajectory of ˆH defined for t ≤ T ′ +q. It is important to observe that +for q big enough this trajectory remains near the torus Ψ(q)(Tn × {I∗ +0}). +Moreover T ′ +q tends to infinity when q → ∞. +(6) Invariant tori +Assume now that x∗ +0 ∈ Tn × G∗ and let us write +� x∗(t) += +(φ∗ +0 + u∗(I∗ +0 )t, I∗ +0) +ˆx∗(t) += +(φ∗ +0 + u′∗(I∗ +0)t, I∗ +0) +for t ∈ R. +Note that +|ˆ˜x(q)(t) − ˆx∗(t)|cq+1 +≤ +cq+1|u′(q)(I∗ +0 ) − u′∗(I∗ +0 )|∞|t| +≤ +cq+1|u′(q) − u′∗|G∗,∞|t|. +And observe that if |t| ≤ +δ(q+1) +1 +|u′(q)−u′∗|G∗,∞ =: T ′′ +q then, +|ˆ˜x(q)(t) − ˆx∗(t)|cq+1 +≤ +cq+1|u′(q) − u′∗|G∗,∞ +δ(q+1) +1 +|u′(q)−u′∗|G∗,∞ +≤ +δq+1 +2 +δq+1 +1 +δq+1 +1 += δq+1 +2 +. +Observe that +|u′∗ − u′(q)|G∗ = | ¯Bu∗ + ¯ +A − ¯Bu(q) − ¯ +A|G∗ += |B(u∗ − u(q))|G∗ ≤ |u∗ − u(q)|G∗, +close enough to Z. +Hence the bound obtained for |u∗−u(q)|G∗ also holds for |u′∗−u′(q)|G∗. +|u′∗ − u′(q)|G∗ ≤ +� +s≥q +4MKτ+1ε +νβ2(τ+2)s ≤ 8MKτ+1ε +νβ2(τ+2)q . +Using this bound, we see that T ′′ +q tends to infinity because +T ′′ +q ≥ +� νρ1 +8 · 2νq +� � νβ2(τ+2)q +8MKτ+1ε +� += +ν2βρ1 +64MKτ+1ε2(τ+2−ν)q. +Then +|ˆx(q)(t) − ˆx∗(t)|cq+1 ≤ |ˆx(q)(t) − ˆ˜x(q)(t)|cq+1 + |ˆ˜x(q)(t) − ˆx∗(t)|cq+1 ≤ 2δ(q+1) +2 +. +when t ≤ T ′′′ +q := min(T ′ +q, T ′′ +q ). +Next, we see that the trajectory Ψ(q)(x(q)(t)) is very close to Ψ∗(x∗(t)) +for large values of q. This is true because when |t| ≤ T ′′′ +q . +|Ψ(q)(ˆx(q) − Ψ∗(ˆx∗(t)))|c1 +≤ |Ψ(q)(ˆx(q)(t)) − Ψ(q)(ˆx∗(t))|c1 + |Ψ(q)(ˆx∗(t)) − Ψ∗(ˆx∗(t))|c1 +≤ 2(τ+1−ν)(q−1) · 2|ˆx(q)(t) − ˆx∗(t)|cq + |Ψ(q) − Ψ∗|G∗,(ρ1/4,0),c1 + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +87 +≤ 2(τ+1−ν)(q−1) · 4δ(q+1) +2 ++ |Ψ(q) − Ψ∗|G∗,(ρ1/4,0),c1 +≤ 2(τ+1−ν)(q−1) · 4δ(q+1) +2 ++ +210Kτε +ν2ρ1β2(1+ν)q +≤ +c1 +cq+1 +4δ(q+1) +2 +2(τ+1−ν) + +210Kτε +ν2ρ1β2(1+ν)q +≤ +c14 +2(τ+1−ν) +δ(q+1) +1 +δ(q+1) +2 +δ(q+1) +2 ++ +210Kτ ε +ν2ρ1β2(1+ν)q +≤ +c14 +2(τ+1−ν) δ(q+1) +1 ++ +210Kτ ε +ν2ρ1β2(1+ν)q +where we used that cq−1/cq = 2τ+1−ν then c1/cq+1 = 2(τ+1−ν)q. +The bound 28 tends to zero. So we deduce, for every fixed t, Ψ(q)(ˆx(q)(t)) +exits or q large enough and its limit is Φ∗(ˆx∗(t)). This fact and the con- +tinuity of the flow of ˆH imply that Ψ∗(ˆx∗(t)) is also a trajectory of ˆH, +which is defined for all t ∈ R. +This holds for every initial condition x∗ +0 = (φ∗ +0, I∗ +0) ∈ Tn × G∗ for this +reason Ψ∗(Tn × {I∗ +0}) is an invariant torus of ˆH, with frequency vector +u′∗(I∗ +0). Observe that the energy on the torus is ˆH(Ψ∗(φ∗ +0, I∗ +0)) = h∗(I∗ +0). +The preserved invariant tori are completely determined by the trans- +formed actions I∗ +0 ∈ G∗. We are now going to characterize the preserved +tori by the original action coordinates. +First, let us see that u( ˆG) ⊂ F ∗. Recall that: +∆c,ˆq(k, α) = {J ∈ R such that |k ¯Bu(I) + k ¯ +A| < α}, +ˆG = {I ∈ G − 2γ +µ such that |k ¯Bu(I) + k ¯ +A| < +β +|k|τ +1 +}. +With this definition is obvious that if I ∈ ˆG then u(I) is +β +|k|τ +1 , K, c, ˆq- +non-resonant. Hence u(I) /∈ ∆c,ˆq(k, +β +|k|τ +1 ) for all k ̸= 0. Then u( ˆG) ⊂ F ∗. +We want to find a correspondence between the invariant tori of ˆh and +the invariant tori of the perturbed system ˆH = ˆh + R, or in the new +coordinates ˆh∗. +Recall +u′ = ¯Bu + ¯ +A, +u′∗ = ¯Bu∗ + ¯ +A = ( +1 +�m +i=1 +ci +Ii +1 +u∗ +1 + +�m +i=1 +ˆqi +Ii +1 +�m +i=1 +ci +Ii +1 +, u∗ +2, . . . , u∗ +n). +Observe u′∗(0, I2, . . . , In) = ˆqm +cm = +1 +K′ the inverse of the modular period, +hence u′∗ and u′ are not one-to-one at Z because they project the first +component of u∗ and u to +1 +K′ . +Let us define I∗ +0 = (u∗)−1(u(I0)), recall that u and u∗ are indeed +one-to-one even though u′ and u′∗ are not, so I∗ +0 is properly defined. +With this definition u∗(I∗ +0) = u(I0) and this implies u′∗(I∗ +0 ) = u′(I0). +Now, let us define T (φ0, I0) = Ψ∗(φ0, I∗ +0). +We obtain 6.13 because the set T (Tn × {I0}) is an invariant torus +of the hamiltonian flow of ˆH with frequency vector u′∗(I∗ +0 ) because Tn × +{I∗ +0} is an invariant torus for the hamiltonian flow of ˆh∗. And we have +seen that u′∗(I∗ +0) = u′(I0). In a nutshell, the original frequencies (of the +unperturbed system) u(I0) for I0 ∈ ˆG are in F ∗ and hence can be seen + +88 +6. A NEW KAM THEOREM +Dρ(G) = Wρ1(Tn × Vρ1(G)) +Wρ1/4(Tn) × G∗ +Wρ1/4(Tn) × G +Wρ1/4(Tn) × ˆG +ˆG +u( ˆG) ⊂ F +F ∗ +Ψ∗ +u∗ +i +i +π +u| ˆ +G +T +i +(u∗)−1 +Figure 2. Diagram of the different maps and sets used in the proof. +as frequencies of the unperturbed system in the new coordinates u∗(I∗ +0). +Hence we can conclude that for this I0 ∈ ˆG its new (perturbed) solution +is also linear in a torus (φ0 + u′∗t, I∗ +0) ∈ Ψ∗(Tn × {I∗ +0}) = T (Tn × {I0}). +And the new frequency vector u′∗ is such that u′∗ = u′. +Let us now prove 6.14. Let us write, for (φ0, I∗ +0) ∈ W ρ1 +4 (Tn) × G∗. +Ψ∗(φ0, I∗ +0) = (φ0 + Ψ∗ +φ(φ0, I∗ +0), I∗ +0 + Ψ∗ +I(φ0, I∗ +0 )). +And for (φ0, I0) ∈ W ρ1 +4 (Tn)× ˆ +G. +T (φ0, I0) = (φ0 + Tφ(φ0, I0), I0 + TI(φ0, I0)). +Then, for (φ0, I0) ∈ W ρ1 +4 (Tn)× ˆ +G: +Tφ(φ0, I0) = Ψ∗ +φ(φ0, I∗ +0), +and +TI(φ0, I0) = Ψ∗ +I(φ0, I∗ +0) + I0 − I∗ +0 . +Let us bound the norms of these terms: +|Ψ∗ +φ(φ0, I∗ +0)|∞ +≤ +1 +c1 |Ψ∗ − id|G∗,( ρ1 +4 ,0),c1 +≤ +16MKτ+1ρ1 +β +210Kτ ε +ν2ρ1β +≤ +214MK2τ+1ε +ν2β2 +, +where we used that c1 ≥ +β +16MKτ+1ρ1 . Then, +Ψ∗ +I(φ0, I∗ +o) +≤ +|Φ∗ − id|G∗,( ρ1 +4 ,0,c1) +≤ +210kτ ε +ν2ρ1β . +Now it only remains the term I∗ +0 − I0: +|I∗ +0 − I0| ≤ |(u∗)(−1) − (u)(−1)|F ∗ ≤ +� +s≥0 +ξs + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +89 +≤ +� +s≥0 +4MKτ+1ε +νβ2(τ+2)s ≤ 8MKτ+1ε +νβ2(τ+2) . +Let us put everything together and use ˆρ ≤ νρ1, K ≤ 2/ˆρ and β = +γ/L. +|Ψ∗ +φ(φ0, I∗ +0)|∞ +≤ +214M( 2 +ˆ +ρ )2τ+1ε +ν2( γ +L )2 +≤ +22τ+15ML2 +ν2 ˆρ2τ+1 +ε +γ2 +|Ψ∗(φ0, I∗ +0)| + |I∗ +0 − I0| +≤ +210( 2 +ˆ +ρ )τε +ν ˆρ( γ +L ) ++ +8M( 2 +ˆ +ρ )τ+1ε +ν( γ +L )2(τ+2) += +210+τ Lε +ν ˆρτ+1γ + +8M2τ+1Lε +ν ˆρτ+1γ2(τ+2) +≤ +220+τ Lε+M2τ+4Lε +ν ˆρτ+1γ +≤ 210+τL(1+M) +ν ˆρτ+1 +ε +γ +(7) Estimate of the measure +Finally, we carry out the estimate of part 3. Let us write +ˆG∗ = (u∗)−1(u( ˆG)). +The invariant tori fill the set +T (Tn × ˆG) = Ψ∗(Tn × ˆG∗) +i.e. all the tori inside T (Tn × ˆG) are invariant although there are more of +them. Because Ψ(q) are hamiltonian transformations, in particular, they +preserve the volume: +meas[Ψ(q)(Tn × ˆG∗)] = meas(Tn × ˆG∗) = (2π)nmeas( ˆG∗). +Now, let us consider the measure of the limit: +meas[Ψ∗(Tn × ˆG∗)]. +To do this we use the superior limit of sets: +∞ +� +n=q +∞ +� +j=q +(Ψ(j)(Tn × ˆG∗)). +Because Ψ(j)(Tn × ˆG∗) are compact and we have the bound +|Ψ∗ − Ψ(q)|G∗,( ρ1 +4 ,0),c1 ≤ +210Kτε +ν2ρ1β2(1+ν)q , +�∞ +j=q(Ψ(j)(Tn × ˆG∗)) is also compact. All the measures are well-defined +and we can say that +meas[Ψ∗(Tn × ˆG∗)] ≥ (2π)nmeas( ˆG∗). +Then, to bound the measure of the complement of the invariant set it +is enough to bound the measure of G \ ˆG∗. +But first, we are going to define some auxiliary sets. Let ˜β = 2γM +µ , +˜βq = (1− +1 +2νq )˜β. Note that ˜β ≥ β if and only if µ ≤ 2ML and we assumed +µ ≤ 2τ+6L2M. +Then, for q ≥ 0 we define + +90 +6. A NEW KAM THEOREM +˜Fq = (F − ˜βq) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, +˜βq +|k|τ +1 +), +˜Gq = (u(q))−1( ˜Fq) +and +˜F ∗ = +� +q≥0 +˜Fq = (F − ˜β) \ +� +k∈Zn\{0} +|k|1≤K +∆c,ˆq(k, +˜β +|k|τ +1 +), +˜G∗ = +� +q≥0 +˜Gq. +In order to prove the bounds, we need to prove previously the inclu- +sions ˜G∗ ⊂ ˆG∗ and ˜G0 ⊂ G. +(a) G ⊃ ˜G0 = (u(0))−1( ˜F0) = (u)−1(F − ˜β), but we know u(G) = F. +(b) ˜G∗ ⊂ ˆG∗. Take I ∈ ˜G∗, then I ∈ ˜Gq∀q ≥ 0. Hence ∃J ∈˜˜Fq∀q such +that u(q)(I). Then ∃J ∈ ˜F ∗ such that u∗(J) = I. +If we check that J ∈ u( ˜G) we obtain (u∗)−1(J) = I ∈ ˆG∗ and we will +be done. We want ˜F ∗ ⊂ u( ˆG). Because we are taking out all the +resonances in ˜F ∗ it is enough to see (F − ˜β) ⊂ u(G − 2γ +µ ). We only +need to use that | ∂u +∂I |G,ρ2 ≤ M. Then F − ˜β ⊂ u(G − 2γ +µ ). This holds +if and only if +˜β +M ≤ 2γ +µ which is true because ˆβ ≤ 2γM +µ . +Then, we proceed as follows +meas(G \ ˆG∗) +≤ +meas(G \ ˜G∗) +≤ +meas( ˜G0 \ ˜G∗) +≤ +�∞ +q=1 meas( ˜Gq−1 \ ˜Gq). +For q ≥ 1 we obtain the following estimate: +meas( ˜Gq−1 \ ˜Gq) ≤ +1 +| det( ∂u(q−1) +∂I +(I))| +meas( +u(q−1)( ˜ +Gq−1) +� �� � +˜Fq−1 +\( +u(q−1)( ˜ +Gq) +� +�� +� +˜Fq − εq−1)). +Where we have used lemma 6.32. Also det( ∂u(q−1) +∂I +(I)) ≥ µn +q−1 because +of the µq−1-non-degeneracy condition all the eigenvalues have to be greater +than µq−1. +meas( ˜Gq−1 \ ˜Gq) +≤ +1 +µn +q−1 meas( ˜Fq−1 \ ( ˜Fq − εq−1)) +≤ +2n +µn meas( ˜Fq−1 \ ( ˜Fq − εq−1)). +Now, we are going to apply lemma 6.30 with +˜Fq−1 = F(˜βq−1, ˜βq−1, Kq − 1) +and ˜Fq = F(˜βq, ˜βq, Kq). +Applying the lemma: + +6.3. +A KAM THEOREM ON bm-SYMPLECTIC MANIFOLDS +91 +meas( ˜Fq−1 \ ˜Fq) ≤ D(˜βq − ˜βq−1) ++2(diamF)n−1 + + + + +� +k∈Zn\{0} +|k|1≤Kq−1 +˜βq − ˜βq−1 +|k|τ +1|k|2,ω ++ +� +k∈Zn\{0} +Kq−1≤|k|1≤Kq +˜βq +|k|τ +1|k|2,ω + + + + +and +meas( ˜Fq \ ( ˜Fq − εq)) ≤ (D + 2n+1(diamF)n−1Kn)εq. +Putting everything together (and using that ˜β0 = 0), we get +(6.23) +meas(G \ ˆG∗) +≤ +2n +µn + +D ˜β + 2(diamF)n−1 +� +k∈Zn\{0} +˜β +|k|τ +1|k|2,ω ++D +∞ +� +q=1 +εq−1 + 2n+1(diamF)n−1 +∞ +� +q=1 +Kn +q εq−1 +� +. +We now only have to check that the series in the previous expression +converge. Let us check that they converge at Z first and then outside of +Z. Recall that at Z we take the vectors ¯k ̸= 0. +� +k∈Zn\{0} +¯k̸=0 +1 +|k|τ +1 |k|2,ω +≤ +� +k∈Zn\{0} +¯k̸=0 +1 +|k|τ +1 |¯k| +≤ +� +¯k∈Zn−1\{0} +¯k̸=0 +� +kn∈Z +√n +(|¯k|1+|kn|)τ|¯k|1 +≤ +√n2n−1 �∞ +j=1 +� +kn∈Z +jn−3 +j+|kn|)τ +where we used that the number of vectors ¯k ∈ Zn−1 with |¯k|1 = j ≥ 1 +can be bounded by 2n−1jn−2. This series can be bounded by comparing +it to an integral: +� +kn∈Z +1 +(j + |kn|)τ ≤ 1 +jτ + 2 +� ∞ +0 +dx +(j + x)τ += 1 +jτ + +2 +(τ + 1)jτ−1 ≤ τ + 1 +τ − 1 +1 +jτ+1 . +Where we used that τ > 1 because n ≥ 2. Then +� +k∈Zn\{0} +¯k̸=0 +1 +|k|τ +1|k|2,ω +≤ +√n2n−1(τ + 1) +τ − 1 +∞ +� +j=1 +1 +jτ−n+2 +which converges by the condition τ > n − 1. +Now let us check that it converges outside of Z. +� +k∈Zn\{0} +1 +|k|τ +1 |k|2,ω += +� +k∈Zn\{0} +¯k̸=0 +1 +|k|τ +1 |k|2,ω + � +k∈Zn\{0} +¯k=0 +1 +|k|τ +1 |k|2,ω += +� +k∈Zn\{0} +¯k̸=0 +1 +|k|τ +1 |¯k| + � +k1∈Z +1 +|k|τ +1 |k2 +1B(I1)2| + +92 +6. A NEW KAM THEOREM +We have seen before that the first term converges. The second term: +� +k1∈Z +1 +|k1|τ|kτ +1B(I1)2| = +1 +B(I1)2 +� +k1∈Z +1 +kτ+2 +1 +, +which converges ∀I1 ̸= 0, i.e. outside of Z. +Now we go back to the expression 6.23. +The other terms of that +expression can be bounded simultaneously inside and outside Z. Now we +only have to check that the third series converges, because if the third +converges so does the second. We only have to check that �∞ +q=1 Kn +q εq−1 +converges. We will use that ε ≤ +8ε +νρ12(2τ+2)(q−1) . +�∞ +q=1 Kn +q εq−1 += +Kn �∞ +q=1 2n(q−1)εq−1 += +Kn �∞ +q=1 +8ε2n(q−1) +νρ12(2τ+2)(q−1) += +Kn 8ε +νρ1 +�∞ +q=1 +1 +2(2τ+2−n)(q−1) . +Which converges if and only if 2τ + 2 − n ≥ 1. And we are done +because 2τ ≥ n − 1 since τ ≥ n − 1 by hypothesis. +Putting everything together: +meas(G \ ˆG∗) +≤ 2n +µn + +D2 2γM +µ ++ 2(diamF)n−1 2γM +µ +√n2n−1(τ + 1) +τ − 1 +∞ +� +j=1 +1 +jτ−n+2 +D 8ε +νρ1 +∞ +� +q=1 +1 +2(2τ + 2)(q − 1) ++2n+1(diamF)n−1Kn 8ε +νρ1 +∞ +� +q=1 +1 +2(2τ+2−n)(q−1) +� +Now using that +ε ≤ +ν2µ2β2 +2τ+30L4M 3K2τ+1 ≤ 2τ−18 · 8MKτ+1ρ2 +LMK2τ+2 +γ ≤ 2τ−15ρ2 +LKτ+1 γ +We can write meas(G \ ˆG∗) ≤ C′γ where C′ depends only on n, µ, D, +diamF, M, τ, ρ1, ρ2, L, K and if we efine C = (2π)nC′. Hence, +meas[(Tn × G) \ T (Tn ˆG)] ≤ Cγ. +□ + +CHAPTER 7 +Desingularization of bm-integrable systems +In this chapter, we follow [GMW17], for the definition of the desingularization +of the bm-symplectic form. +Definition 7.1. The fǫ-desingularization ωǫ form of ω = dx +xm ∧ +��m−1 +i=0 xiαm−i +� ++ +β is: +ωǫ = dfǫ ∧ +�m−1 +� +i=0 +xiαm−i +� ++ β. +Where in the even case, fǫ(x) is defined as ǫ−(2k−1)f(x/ǫ). And f ∈ C∞(R) is an +odd smooth function satisfying f ′(x) > 0 for all x ∈ [−1, 1] and satisfying outside +that +(7.1) +f(x) = +� +−1 +(2k−1)x2k−1 − 2 +for +x < −1, +−1 +(2k−1)x2k−1 + 2 +for +x > 1. +And in the odd case, fǫ(x) = ǫ−(2k)f(x/ǫ). And f ∈ C∞(R) is an even smooth +positive function which satisfies: f ′(x) < 0 if x < 0, f(x) = −x2 + 2 for x ∈ [−1, 1], +and +(7.2) +f(x) = +� +−1 +(2k+2)x2k+2 − 2 +if k > 0, x ∈ R \ [−2, 2] +log(|x|) +if k = 0, x ∈ R \ [−2, 2]. +Remark 7.2. With the previous definition, we obtain smooth symplectic (in +the even case) or smooth folded symplectic (in the odd case) forms that agree +outside an ǫ-neighbourhood with the original bm-forms. Moreover, there is a con- +vergence result in terms of m. See [GMPS17] for the details. +To simplify notation, we introduce F m−i +ǫ +(x) = ( d +dxfǫ(x))xi, and hence F i +ǫ(x) = +( d +dxfǫ(x))xm−i. With this notation the desingularization ωǫ is written: +ωǫ = +m−1 +� +i=0 +F m−i +ǫ +(x)dx ∧ αm−i + β. +Definition 7.3. The desingularization for (M, ω, µ) is the triple (M, ωε, µǫ) +where ωε is defined as above and µε is: +µ �→ µǫ = +� +f1ǫ = +m +� +i=1 +ˆciGi +ǫ(x), f2(˜I, ˜φ), . . . , fn(˜I, ˜φ) +� +, +where +µ = +� +f1 = c0 log(x) + +m−1 +� +i=1 +ci +1 +xi , f2(I, φ) . . . , fn(I, φ) +� +93 + +94 +7. DESINGULARIZATION OF bm-INTEGRABLE SYSTEMS +Gi +ǫ(x) = +� x +0 +F i +ǫ(τ)dτ, +and ˆc1 = c0 and ˆci−1 = −ici if i ̸= 0. Also + + + + + + + + + + + +˜I = (˜I1, I2, . . . , In), +˜I1 += +� I1 +0 +��m +i=1 KˆciF i +ε(τ) +�m +i=1 +Kˆcj +τj +� +dτ +˜φ1 = (˜φ1, φ2, . . . , φn), +˜φ1 += + + +�m +i=1 K��ciF i +ε(I1) +�m +i=1 +Kˆcj +Ij +1 + + φ1 +Remark 7.4. Observe that with the last definition, when ǫ tends to 0, µǫ tends +to µ. +Theorem C. The desingularization transforms a bm-integrable system into an +integrable system for m even on a symplectic manifold. For m odd the desingular- +ization transforms it into a folded integrable system. The integrable systems are +such that: +Xω +fj = Xωǫ +fjǫ. +Proof. Let us first check the singular part, i.e. let us check that that Xω +f1 = +Xωǫ +f1ǫ. Let us compute the two equations that define each one of the vector fields. +We have to impose −df1 = ιXω +f1 ω and −df1ǫ = ιXωǫ +f1ǫ ωǫ. But observe first that we +can rewrite ω = �m +i=1 +1 +xi dx ∧ αi + β and ωǫ = �m +i=1 F i +ǫdx ∧ αi + β. The conditions +translate as: +− +m +� +i=1 +ˆci +1 +xi dx = ιXω +f1 +� m +� +i=1 +1 +xi dx ∧ αi + β +� +, +− +m−1 +� +i=0 +ˆciF i +ǫ(x)dx = ιXωǫ +f1ǫ +�m−1 +� +i=0 +F i +ǫ(x)dx ∧ αi + β +� +. +Since the toric action leaves the form ω invariant, in particular, the singular +set is invariant, and then Xωǫ +f1ǫ and Xω +f1 are in the kernel of dx. Moreover, since β +is a symplectic form in each leaf of the foliation and Xωǫ +f1ǫ and Xω +f1 are transversal +to this foliation, they are also in the kernel of β. +− +m−1 +� +i=0 +ˆci +1 +xi dx = +m−1 +� +i=0 +1 +xi dx ∧ αi(Xω +f1), +− +m−1 +� +i=0 +ˆciF i +ǫ(x)dx = +m−1 +� +i=0 +F i +ǫ(x)dx ∧ αi(Xωǫ +f1ǫ). +Then, the conditions over Xω +f1 and Xωǫ +f1ǫ are respectively: +−ˆci = αi(Xω +f1), +−ˆci = αi(Xωǫ +f1ǫ). +Then, the two vector fields have to be the same. +Let us now see Xω +fj = Xωǫ +fjǫ for j > 1. Assume now we have the bm-symplectic +form in action-angle coordinates ω = �m +i=1 +Kˆci +Ii +1 dI1 ∧ dφ1 + �n +i=1 dIi ∧ dφi. +The differential of the functions are + +7. DESINGULARIZATION OF bm-INTEGRABLE SYSTEMS +95 +df ε +i += +∂f ε +i +∂I1 dI1 + ∂f ε +i +∂φ1 dφ1 + �n +j=2 +� +∂f ε +i +∂Ij dIj + ∂f ε +i +∂φj dφj +� += +∂fi +∂I1 +��m +i=1 KˆciF i +ε(τ) +�m +i=1 +Kˆcj +τj +� +dI1 + ∂fi +∂φ1 +��m +i=1 KˆciF i +ε(τ) +�m +i=1 +Kˆcj +τj +� +dφ1 ++ �n +j=2 +� +∂f ε +i +∂Ij dIj + ∂f ε +i +∂φj dφj +� +. +On the other hand, the desingularized form is: +ωε = +m +� +j=1 +KˆciF j +ε (I1)dI1 ∧ dφ1 + +m +� +j=2 +dIj ∧ dφj. +Hence, one can see that the expression for both Xω +fj and Xωǫ +fjǫ is +Xω +fj = Xωǫ +fjǫ = +∂fi +∂I1 +�m +i=1 +Kˆci +Ii +1 +∂ +∂φ1 +− +∂fi +∂φ1 +�m +i=1 +Kˆci +Ii +1 +∂ +∂I1 ++ +n +� +j=2 +�∂f ε +i +∂Ij +dIj + ∂f ε +i +∂φj +dφj +� +□ +Remark 7.5. The previous lemma tells us that the dynamics of the desingu- +larized system are identical to the dynamics of the original bm-integrable system in +the bm-symplectic manifold. +Hence the desingularized bm-form goes to a folded symplectic form in the case +m = 2k + 1 and to symplectic for m = 2k. And the bm-integrable system goes to +a folded integrable system (see [CM22]) in the case m = 2k + 1 and to a standard +integrable system for n = 2k. + + +CHAPTER 8 +Desingularization of the KAM theorem on +bm-symplectic manifolds +The idea of this section is to recover some version of the classical KAM theorem +by “desingularizing the bm-KAM theorem”, as well as a new version of a KAM +theorem that works for folded symplectic forms. Observe that no KAM theorem +is known for folded symplectic forms. The best that is known is a KAM theorem +for presymplectic structures that was done in [AdlL12]. Desingularizing the KAM +means applying the bm-KAM in the bm-manifold and then translating the result to +the desingularized setting. +To be able to obtain proper desingularized theorems we need to identify which +integrable systems can be obtained as a desingularization of a bm-integrable system. +To simplify computations we are going to use a particular case of bm-integrable +systems, where f1 = +1 +Im−1 +1 +. We call these systems simple. Observe that by taking +a particular case of bm-integrable systems we will not get all the systems that can +be obtained by desingularizing a bm-integrable system, but some of them. +(1) Even case m = 2k. +F = (f1 = +1 +I2k−1 +1 +, f2, . . . , fn), ω = +1 +Im +1 dI1 ∧ dφ1 + �n +j=1 dIj ∧ dφj. +Observe that close to Z in the even case we can assume f(I1) = cI1 +for some c > (2 − +1 +22k−1 ). Then fε(I1) = +1 +ε(2k−1) +cI1 +ε += c′I, hence ωε = +c′dI1 ∧ dφ1 + �n +j=1 dIj ∧ dφj. Also F m +ε (I1) = c′, Gm +ε (I1) = c′I1. Then, +� ˜I1 += +� I1 +0 +c′ +1/τ m dτ = +� I1 +0 c′τ mdτ = c′ Im+1 +1 +m+1 , +˜φ1 += +c′ +1/Im +1 φ1 = c′Im +1 φ1 +(8.1) +F ε = ((m − 1)cm−1c′I1, f2(˜I, ˜φ), . . . fn(˜I, ˜φ)). +Hence, the systems in this form can be viewed as a desingularization +of a bm-integrable system. +Theorem D (Desingularized KAM for symplectic manifolds). Con- +sider a neighborhood of a Liouville torus of an integrable system Fε as +in 8.1 of a symplectic manifold (M, ωε) semilocally endowed with coor- +dinates (I, φ), where φ are the angular coordinates of the torus, with +ωε = c′dI1∧dφi+�n +j=1 dIj∧dφj. Let H = (m−1)cm−1c′I1+h(˜I)+R(˜I, ˜φ) +be a nearly integrable system where +� +˜I1 += +c′ Im+1 +1 +m+1 , +˜φ1 += +c′Im +1 φ1, +97 + +98 +8. bm-KAM DESINGULARIZATION +and +� ˜I += +(˜I1, I2, . . . , In), +˜φ += +(˜φ1, φ2, . . . , φn). +Then the results for the bm-KAM theorem 6.3 applied to Hsing = +1 +I2k−1 +1 ++ +h(I) + R(I, φ) hold for this desingularized system. +Remark 8.1. This theorem is not as general as the standard KAM, +but we also know extra information about the dynamics. For instance, +the perturbation of trajectories in tori inside of Z will be trajectories lying +inside of Z. In this sense, the theorem is new because it leaves invariant +an hypersurface of the manifold. +(2) Odd case m = 2k + 1. +F = (f1 = +1 +I2k +1 , f2, . . . , fn) and ω = +1 +I2k+1 +1 +dI1 ∧ dφ1 + �n +j=1 dIj ∧ dφj. +Before continuing we need the following notions defined in [CM22]. +Definition 8.2. A function f : M → R in a folded symplectic mani- +fold (M, ω) is folded if df|Z(v) = 0 for all v ∈ V = kerω|Z. +Definition 8.3. An integrable system in a folded symplectic manifold +(M, ω) with critical surface Z is a set of functions F = (f1, . . . , fn) such +that they define Hamiltonian vector fields which are independent (df1 ∧ +. . . ∧ dfn ̸= 0 in the folded cotangent bundle) on a dense subset of Z and +M, and commute with respect to ω. +Note that we need to prove that the desingularized functions in this +case are folded. +Observe that close to Z in the odd case we can assume f(I1) = −I2 +1+2. +Then fε(I1) = ε−(2k)f( I1 +ε ) = +1 +ε2k (−( I1 +ε )2 + 2) = cI2 +1 + +2 +ε2k . Then +ωε = 2cI1dI1 ∧ dφ1 + +n +� +j=1 +dIj ∧ dφj. +Also F m +ε (I1) = 2cI1, Gm +ε (I1) = cI2 +1. Then, +� +˜I1 += +� I1 +0 +2cτ +1/τ m dτ = 2c I(m+2) +1 +(m+2) , +˜φ1 += +2cIm+1 +1 +φ1 +Then the desingularized moment map becomes +(8.2) +F ε = ((m − 1)cm−1cI2 +1, f2(˜I, ˜φ), . . . fn(˜I, ˜φ)). +It is a simple computation to check that these functions are actually +folded and hence they form a folded integrable system. Note that the +systems of the form 8.2 can be viewed as a desingularization of a bm- +integrable system. Then, as we proceeded in the even case: +Theorem E (Desingularized KAM for folded symplectic manifolds). +Consider a neighborhood of a Liouville torus of an integrable system Fε as +in 8.2 of a folded symplectic manifold (M, ωε) semilocally endowed with +coordinates (I, φ), where φ are the angular coordinates of the Torus, with + +8. bm-KAM DESINGULARIZATION +99 +ωε = 2cI1dI1 ∧ dφ1 + �m +j=2 dIj ∧ dφj. Let H = (m − 1)cm−1cI2 +1 + h(˜I) + +R(˜I, ˜φ) a nearly integrable system with +� +˜I1 += +2c Im+2 +1 +m+2 , +˜φ1 += +2cIm+1 +1 +φ1, +and +� ˜I += +(˜I1, I2, . . . , In), +˜φ += +(˜φ1, φ2, . . . , φn). +Then the results for the bm-KAM theorem 6.3 applied to Hsing = +1 +I2k +1 ++ +h(I) + R(I, φ) hold for this desingularized system. +Remark 8.4. The last two theorems can be improved if we consider +bm-integrable systems not necessarily simple. + + +CHAPTER 9 +Potential applications to Celestial mechanics +All the theory developed in this monograph would not be fertile if we could not +envisage applications of perturbation theory to actual physical systems. We provide +several examples from Celestial mechanics and conclude with potential applications +of our KAM theory to detect periodic trajectories. +In this chapter we present several examples appearing in Celestial Mechanics +where singular symplectic forms show up. Some of these examples are contained +in [MDD+19]. +Most of the singularities appear as a consequence of applying +regularization techniques. We invite the reader to consult the book [Kna18] for a +pedagogical approach to the study of regularization. +This list of examples is of special relevance for this booklet as the theoretical +results that we obtain such as action-angle coordinates or KAM can be, de facto, +applied to the list of problems considered below. +Structures that are symplectic almost everywhere can arise as the result of +changes of coordinates which do not preserve the canonical symplectic structure. +For instance: For the Kepler problem given a configuration space R2 and phase +space T ∗R2, the traditional (canonical) Levi-Civita transformation described as +follows: identify R2 ∼= C so that T ∗R2 ∼= T ∗C ∼= C2 and treat (q, p) as complex +variables (q1 + iq2 := u, p1 + ip2 := v). Take the following change of coordinates +(q, p) = (u2/2, v/¯u), where ¯u denotes the complex conjugation of u. The resulting +coordinate change can easily be seen to preserve the canonical symplectic form. +However, this canonical change of coordinates can make the Hamiltonian equations +more complicated making more difficult to study the dynamics of the system. This +is why it is often interesting to consider other changes of coordinates where the +symplectic form is not preserved. Some of them induce new singular forms where +our geometrical and dynamical techniques can be applied. +Other examples are discussed in [DKM17]. +9.1. The Kepler Problem +In suitable coordinates in T ∗ � +R2 \ {0} +� +, the Kepler problem has Hamiltonian +(9.1) +H(q, p) = ∥p∥2 +2 +− +1 +∥q∥. +With the canonical Levi-Civita transformation (q, p) = (u2/2, v/¯u), this expression +becomes +(9.2) +H(u, v) = ∥v∥2 +2∥¯u∥2 − +1 +∥u∥2 . +101 + +102 +9. +POTENTIAL APPLICATIONS TO CELESTIAL MECHANICS +Changes of coordinates preserving the canonical symplectic form leads to more +complicated equations. So we propose a new reciipe: leave the momentum un- +changed and examine the transformation (q, p) = (u2/2, p) instead. This can result +in a simpler Hamiltonian. The transformation is not a symplectomorphism and the +symplectic form on T ∗R2 pulls-back under the transformation to a two-form which +is symplectic almost everywhere, but degenerates on a hypersurface of T ∗R2 +Namely, the Liouville one-form p1dq1 + p2dq2 = ℜ(pd¯q) pulls back to +θ = ℜ +� +pd +� ¯u2 +2 +�� += +ℜ (p¯ud¯u) += +p1(u1du1 − u2du2) + p2(u2du1 + u1du2) +and the associated 2-form −dθ yields a form that is almost everywhere symplectic +ω = u1du1 ∧ dp1 − u2du1 ∧ dp2 + u2du2 ∧ dp1 + u1du2 ∧ dp2. +In order to test the nature of this form we wedge the form with itself and we +find +ω ∧ ω = (u2 +1 − u2 +2)du1 ∧ dp1 ∧ du2 ∧ dp2 +which is degenerate along the hypersurface given by u1 = ±u2. +We now consider the restriction of the form to the critical set. It does not have +maximal rank so it is not a folded symplectic structure. This form is degenerately +folded and the folding hypersurface is not regular and is described by the equations +u1 = ±u2. +9.2. The Problem of Two Fixed Centers +We now regularize the problem of two fixed centers. +The problem of two fixed centers is associated to the motion of a satellite moving +in a gravitational potential generated by two fixed massive bodies. We assume also +that the motion of the satellite is restricted to the plane in R3 containing the two +massive bodies. +The Hamiltonian function in suitable coordinates reads: +(9.3) +H = p2 +2m − µ +r1 +− 1 − µ +r2 +where µ is the mass ratio of the two bodies (i.e. µ = +m1 +m1+m2 ). +The integrability of this problem was first proved by Euler via elliptic coordi- +nates, where the coordinate lines are confocal ellipses and hyperbola. +Explicitly, consider a coordinate system in which the two centers are placed at +(±1, 0), in which the (Cartesian) coordinates are given by (q1, q2). Then the elliptic +coordinates of the system are given by +q1 = sinh λ cos ν +(9.4) +q2 = cosh λ sin ν +(9.5) +for (λ, ν) ∈ R × S1. Thus lines of λ = c and ν = c are given by confocal hyperbola +and ellipses in the plane, respectively. Similar to the Levi-Civita transformation +this results in a double-branched covering with branch points at the centers of +attraction. +Pulling back the canonical symplectic structure ω = dq ∧ dp we find +(9.6) +ω = cosh λ cos ν(dλ ∧ dp1 + dν ∧ dp2) − sinh λ sin ν(dν ∧ dp1 + dλ ∧ dp2) + +9.3. DOUBLE COLLISION AND MCGEHEE COORDINATES +103 +which is degenerate along the hypersurface (λ, ν) satisfying cosh λ cos ν = sinh λ sin λ. +9.3. Double Collision and McGehee coordinates +In this section, we describe another example of b-symplectic structure appearing +quite naturally in physical dynamical systems. From this example, it would seem +natural that a collection of different examples for bm-symplectic models or even +bm-folded models would follow. +But one finds a major problem while pursuing +these examples. Understanding why this example does not extend to construct +bm-symplectic models of bm-folded for any m gives a general pattern. +First let us introduce the McGehee coordinate change for the problem of double +collision. +The system of two particles moving under the influence of the generalized po- +tential U(x) = −|x|−α, α > 0, where |x| is the distance between the two particles, +is studied by McGehee in [McG81]. We fix the center of mass at the origin and +hence can simplify the problem to the one of a single particle moving in a central +force field. +The equation of motion can be written as, +(9.7) +¨x = −∇U(x) = −α|x|−α−2x +where the dot represents the derivative with respect to time. In the Hamiltonian +formalism, this equation becomes +(9.8) +˙x += +y, +˙y += +−α|x|−α−2x. +To study the behavior of this system, the following change of coordinates is sug- +gested in [McG81]: +(9.9) +x += +rγeiθ, +y += +r−βγ(v + iw)eiθ +where the parameters β and γ are related with α as follows: +(9.10) +β += +α/2, +γ += +1/(1 + β). +Identifying the plane R2 with the complex plane C, we can write the symplectic +form of this problem as ω = ℜ(dx ∧ dy). +Remark 9.1. To check that a form ω is actually a bm-symplectic form, it is not +enough to check that the multi-vector field dual to ω∧ω is a section of �2n(bmT M) +which is transverse to the zero section. One has to check additionally that the +Poisson structure dual to ω itself is a proper section of �2(bmT M). +Proposition 9.2. Under the coordinate change (9.9), the symplectic form ω +is sent to a b-symplectic structure for α = 2. +Proof. The proof of this proposition is a straightforward computation. Ob- +serve that the change is not a smooth change, so we are not working with standard +De Rham forms. But, at the end of the computation it will become clear that the +form is a b-symplectic form and hence the computations are legitimate. If one does + +104 +9. +POTENTIAL APPLICATIONS TO CELESTIAL MECHANICS +the change of variables, we obtain: +(9.11) +¯y += +rβγ(v − iw)e−iθ. +dx += +γrγ−1eiθdr + rγeiθidθ. +d¯y += +r−βγ−1(−βγ)(v − iw)e−iθdr + r−βγe−iθdv ++rβγ(v − iw)e−iθ(−i)dθ. +By wedging the previous two forms, we obtain: +(9.12) +dx ∧ d¯y += +dr ∧ dv(γrγ−1−βγ) ++ +dr ∧ dw(γrγ−1−βγ) ++ +dr ∧ dθ(γrγ−1−βγ(−iv − w)) ++ +dθ ∧ dr(irγ−1−βγ(−βγ)(v − iw)) ++ +dθ ∧ dv(irγ−βγ) ++ +dθ ∧ dw(irγ−βγ(−i)). +Now we can take the real part of this form and use that γ − 1 − βγ = −αγ. In the +new coordinates, the form reads. +(9.13) +ω = ℜ(dx ∧ d¯y) += +γr−βγ+γ−1dr ∧ dv − γ(1 − β)r−βγ+γ−1wdr ∧ dθ +− +r−βγ+γdw ∧ dθ. +Moreover, we can use that γ(1 + β) = 1 to simplify the previous expression further +to: +(9.14) +ω = (dr ∧ dv + dr ∧ dw)γr−αγ + dr ∧ dθ(wr−αγ) + dθ ∧ dw(r−αγ+1). +In order to classify this structure, we wedge it with itself and look at the structure +of the form in the singular set. Wedging this form, we obtain +(9.15) +ω ∧ ω += +−γr−2βγ+2γ−1dr ∧ dv ∧ dθ ∧ dw += +−γr +2−3α +2+α dr ∧ dv ∧ dθ ∧ dw. +where we use (9.10). Let us set f(α) = 2−3α +2+α . This function does not take values +lower than −3 or higher than 1. When α = 2 this gives us a b-symplectic structure: +ω ∧ ω = −γrdr ∧ dv ∧ dθ ∧ dw. +The section of �4(bT M) given by the dual structure of ω ∧ ω is clearly transverse +to the zero section. +On the other hand if α = 2, then β = 1 and hence: +ω = γr−1dr ∧ ω ∧ dv, +and its dual Poisson structure is clearly also a proper section of �2(bT M). +□ +Remark 9.3. One may ask if for other values of α it is possible to obtain +other bm-symplectic structures for different m. For example for α = 6, as ω ∧ ω = +−γr−2dr ∧dv ∧dθ ∧dw, so it seems likely to obtain a b2-symplectic form. But from +the expression of ω it becomes clear that it is not a proper section of �2(b2T ∗M) + +9.4. POTENTIAL APPLICATIONS +105 +m1 = 1 − µ +m2 = µ +q +r2 = q − q2 +r1 = q − q1 +Center of mass +r +q1 +q2 +Figure 1. Scheme of the three-body problem. +9.4. The restricted three-body problem +In this last section of the monograph, we catch up with the circular planar +restricted three-body problem. +The restricted elliptic 3-body problem is a simplified version of the 3-body +problem. It describes the trajectory of body with negligible mass moving in the +gravitational field of two massive bodies called primaries, orbiting in elliptic Kep- +lerian motion. The restricted planar version assumes that all motion occurs in a +plane. +The associated Hamiltonian of the particle can be written as: +(9.16) +H(q, p) = ∥p∥2 +2 ++ +1 − µ +∥q − q1∥ + +µ +∥q − q2∥ = T + U +wit µ the reduced mass of the system. +As it was observed in [KMS16b], it is possible to associate a singular struc- +ture to this problem. Consider the symplectic form on T∗R2 in polar coordinates, +After making a change to polar coordinates (q1, q2) = (r cos α, r sin α) and the +corresponding canonical change of momenta we find the Hamiltonian function +(9.17) +H(r, α, Pr, Pα) = P 2 +r +2 + P 2 +α +2r2 + U(r cos α, r sin α) +where Pr, Pα are the associated canonical momenta and with potential energy: +U(r cos α, r sin α) +The McGehee change of coordinates is used to examine the behavior of orbits +near infinity, see also [DKdlRS19]: +(9.18) +r = 2 +x2 . +The corresponding change for the canonical momenta is easily seen to be +(9.19) +Pr = −x3 +4 Px. + +106 +9. +POTENTIAL APPLICATIONS TO CELESTIAL MECHANICS +The Hamiltonian is transformed to +(9.20) +H(r, α, Pr, Pα) = x6P 2 +x +32 ++ x4P 2 +α +8 ++ U(x, α). +By transforming the position coordinate (9.18) without modifying the momentum +associated to r, we are left with a simpler Hamiltonian, however, the pull-back of +the symplectic form is no longer symplectic, but exhibits a singularity of order 3 +and it is called b3-symplectic: +(9.21) +ω = 4 +x3 dx ∧ dPr + dα ∧ dPα. +Adding the line at infinity provides a description of the dynamics within the +critical set Z = {x = 0}. From the change of coordinates implemented, we might +think that the dynamics within Z may have no physical meaning, but its interplay +with the dynamics close to Z gives information about the behaviour of escape orbits +sometimes identified as singular periodic orbits (see [MO21] and [MOPS22]). +Given an autonomous Hamiltonian system of a symplectic manifold of dimen- +sion 2n, the level sets of the Hamiltonian function are often endowed with a contact +structure ( a contact structure is given by a one form α satisfying a condition of +type α ∧ (dα)n−1 ̸= 0). +In [MO18, MO21] applications of the b-apparatus are discussed in this con- +text. In particular, the notion of bm-contact structures is introduced by translating +the condition above for bm-forms. The classical notions in the contact realm such +as Reeb vector fields can also be introduced in this set-up. +By considering the Mc Gehee change as we did in the contact context, in +[MO21] it is proved: +Theorem 9.4. After the McGehee change, the Liouville vector field Y = p ∂ +∂p is +a b3-vector field that is everywhere transverse to the level sets of the Hamiltonian Σc +for c > 0 and the level-sets (Σc, ιY ω) for c > 0 are b3-contact manifolds. Topologi- +cally, the critical set of this contact manifold is a cylinder (which can be interpreted +as a subset of the line at infinity) and the Reeb vector field admits infinitely many +non-trivial periodic orbits on the critical set. +One of the possible applications of our KAM theorem would be to find new +periodic orbits of the restricted three body problem close to infinity by perturbing +the periodic orbits described above. This old technique of perturbation theory is +probably due to Poincar´e (Poincar´e’s continuation method, see [MO17]). +This +opens the door to new investigations which will be considered elsewhere. + +Bibliography +[AA81] +V. I. Arnol’d and V. I. Arnol’d, Singularity theory : selected papers / [edited by] v.i. +arnold, Cambridge University Press Cambridge [Cambridgeshire] ; New York, 1981 +(English). +[AdlL12] +Hassan Najafi Alishah and Rafael de la Llave, Tracing KAM tori in presymplectic +dynamical systems, J. Dynam. Differential Equations 24 (2012), no. 4, 685–711. +MR 3000600 +[Arn89] +V. I. Arnol’d, Poisson structures on the plane and other powers of volume forms, +Journal of Soviet Mathematics 47 (1989), no. 3, 2509–2516. +[BMO22] +Joaquim Brugu´es, Eva Miranda, and C´edric Oms, The arnold conjecture for singular +symplectic manifolds, arXiv:2212.01344 (2022). +[CM22] +Robert Cardona and Eva Miranda, Integrable Systems on Singular Symplectic Mani- +folds: From Local to Global, Int. Math. Res. Not. IMRN (2022), no. 24, 19565–19616. +MR 4523256 +[DG96] +Amadeu Delshams and Pere Guti´errez, Effective stability and KAM theory, J. Dif- +ferential Equations 128 (1996), no. 2, 415–490. MR 1398328 +[DKdlRS19] Amadeu Delshams, Vadim Kaloshin, Abraham de la Rosa, and Tere M. Seara, Global +instability in the restricted planar elliptic three body problem, Communications in +Mathematical Physics 366 (2019), no. 3, 1173–1228. +[DKM17] +Amadeu Delshams, Anna Kiesenhofer, and Eva Miranda, Examples of integrable and +non-integrable systems on singular symplectic manifolds, J. Geom. Phys. 115 (2017), +89–97. MR 3623614 +[GLPR17] +Marco Gualtieri, Songhao Li, ´Alvaro Pelayo, and Tudor Ratiu, The tropical momen- +tum map: a classification of toric log symplectic manifolds, Mathematische Annalen +367 (2017). +[GMP10] +Victor Guillemin, Eva Miranda, and Ana Pires, Codimension one symplectic folia- +tions and regular poisson structures, Bulletin of the Brazilian Mathematical Society, +New Series 42 (2010). +[GMP14] +Victor Guillemin, Eva Miranda, and Ana Rita Pires, Symplectic and poisson geom- +etry on b-manifolds, Advances in Mathematics 264 (2014), 864–896. +[GMPS15a] Victor Guillemin, Eva Miranda, Ana Rita Pires, and Geoffrey Scott, Toric actions +on b-symplectic manifolds, Int. Math. Res. Not. IMRN (2015), no. 14, 5818–5848. +MR 3384459 +[GMPS15b] +, Toric actions on b-symplectic manifolds, Int. Math. Res. Not. IMRN (2015), +no. 14, 5818–5848. MR 3384459 +[GMPS17] +, Convexity for Hamiltonian torus actions on b-symplectic manifolds, Math. +Res. Lett. 24 (2017), no. 2, 363–377. MR 3685275 +[GMW17] +Victor Guillemin, Eva Miranda, and Jonathan Weitsman, Desingularizing bm- +symplectic structures, International Mathematics Research Notices 2019 (2017), +no. 10, 2981–2998. +[GMW18a] +Victor W. Guillemin, Eva Miranda, and Jonathan Weitsman, Convexity of the mo- +ment map image for torus actions on bm-symplectic manifolds, Philos. Trans. Roy. +Soc. A 376 (2018), no. 2131, 20170420, 6. MR 3868423 +[GMW18b] +, On geometric quantization of b-symplectic manifolds, Adv. Math. 331 +(2018), 941–951. MR 3804693 +[GMW21] +, On geometric quantization of bm-symplectic manifolds, Math. Z. 298 (2021), +no. 1-2, 281–288. MR 4257086 +107 + +108 +BIBLIOGRAPHY +[GS90] +V. Guillemin and S. Sternberg, Symplectic techniques in physics, Cambridge Univer- +sity Press, 1990. +[KM17] +Anna Kiesenhofer and Eva Miranda, Cotangent models for integrable systems, +Comm. Math. Phys. 350 (2017), no. 3, 1123–1145. MR 3607471 +[KMS16a] +Anna Kiesenhofer, Eva Miranda, and Geoffrey Scott, Action-angle variables and a +KAM theorem for b-Poisson manifolds, J. Math. Pures Appl. (9) 105 (2016), no. 1, +66–85. MR 3427939 +[KMS16b] +, Action-angle variables and a KAM theorem for b-Poisson manifolds, J. +Math. Pures Appl. (9) 105 (2016), no. 1, 66–85. MR 3427939 +[Kna18] +Andreas Knauf, Mathematical physics: +classical mechanics, Unitext, vol. 109, +Springer-Verlag, Berlin, 2018, Translated from the 2017 second German edition by +Jochen Denzler, La Matematica per il 3+2. MR 3752660 +[LGMV08] +Camille Laurent-Gengoux, Eva Miranda, and Pol Vanhaecke, Action-angle coordi- +nates for integrable systems on poisson manifolds, International Mathematics Re- +search Notices 2011 (2008). +[Mar19] +Charles-Michel Marle, Projection st´er´eographique et moments, working paper or +preprint, June 2019. +[McG81] +Richard P McGehee, Double collisions for a classical particle system with nongravi- +tational interactions, Commentarii Mathematici Helvetici 56 (1981), no. 1, 524–557 +(English (US)). +[MDD+19] +E. Miranda, A. Delshams, R. Dempsey, C. Oms, and A. Planas, An invitation to +singular symplectic geometry, International journal of geometric methods in modern +physics (2019). +[Mel93a] +R. Melrose, The atiyah-patodi-singer index theorem, Research Notes in Mathematics, +CRC Press, 1993. +[Mel93b] +, The atiyah-patodi-singer index theorem, Research Notes in Mathematics, +CRC Press, 1993. +[MM22] +Anastasia Matveeva and Eva Miranda, Reduction theory for singular symplectic +manifolds and singular forms on moduli spaces, arXiv:2205.12919 (2022). +[MO17] +Kenneth R. Meyer and Daniel C. Offin, Introduction to Hamiltonian dynamical sys- +tems and the N-body problem, third ed., Applied Mathematical Sciences, vol. 90, +Springer, Cham, 2017. MR 3642697 +[MO18] +Eva Miranda and C´edric Oms, The geometry and topology of contact structures with +singularities, arXiv:1806.05638 (2018). +[MO21] +Eva Miranda and C´edric Oms, The singular weinstein conjecture, Advances in Math- +ematics 389 (2021), 107925. +[MOPS22] +Eva Miranda, C´edric Oms, and Daniel Peralta-Salas, On the singular Weinstein con- +jecture and the existence of escape orbits for b-Beltrami fields, Commun. Contemp. +Math. 24 (2022), no. 7, Paper No. 2150076, 25. MR 4476313 +[MOT14] +Ioan Marcut and Boris Osorno Torres, Deformations of log-symplectic structures, +Journal of the London Mathematical Society 90 (2014), no. 1, 197–212. +[MS21] +Eva Miranda and Geoffrey Scott, The geometry of E-manifolds, Rev. Mat. Iberoam. +37 (2021), no. 3, 1207–1224. MR 4236806 +[P¨os93] +J¨urgen P¨oschel, Nekhoroshev estimates for quasi-convex hamiltonian systems, Math- +ematische Zeitschrift 213 (1993), no. 1, 187–216. +[Rad02] +Olga Radko, A classification of topologically stable Poisson structures on a compact +oriented surface, J. Symplectic Geom. 1 (2002), no. 3, 523–542. MR 1959058 +[RBM] +A. Kiesenhofer R. Braddell and E. Miranda, A b-symplectic slice theorem, Bulletin +of the London Mathematical Society, published online in 2022. +[Sco16] +Geoffrey Scott, The geometry of bk manifolds, J. Symplectic Geom. 14 (2016), no. 1, +71–95. MR 3523250 +[Swa62] +Richard G. Swan, Vector bundles and projective modules, Transactions of the Amer- +ican Mathematical Society 105 (1962), 264–277. +