• Aucun résultat trouvé

A quantitative KAM theorem

N/A
N/A
Protected

Academic year: 2021

Partager "A quantitative KAM theorem"

Copied!
21
0
0

Texte intégral

(1)

HAL Id: hal-01514613

https://hal.archives-ouvertes.fr/hal-01514613v2

Preprint submitted on 21 Aug 2017

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

A quantitative KAM theorem

Thibaut Castan

To cite this version:

(2)

A quantitative KAM theorem

Thibaut Castan

Observatoire de Paris, IMCCE 77, Avenue Denfert-Rochereau, 75014 Paris

thibaut.castan@obspm.fr

Keywords: KAM Theorem, Hamiltonian Systems, Stability, Quasi-periodic Motions, Perturbation Theory MSC 2010: 37C55, 37J25, 37J40, 70H08, 70H09, 70H14, 70H15, 70H20, 70K43

Abstract. We revisit Pöschel’s 2001 version of the KAM theorem so as to find an explicit quantitative bound for the size of the allowed perturbation. Our theorem is applied to the plane planetary problem in a pair of other papers.

In 1963, Arnold [1] proved that Kolmogorov’s theorem [7] could be applied to the plane planetary three-body problem, thus showing the existence of quasi-periodic solutions over an infinite time interval. This theorem relies on a smallness condition on the ratio of masses between the planets and the star. Hénon, in a letter to Arnold (see [8] pages 11-12), derived an explicit, necessary condition on this ratio for Arnold’s scheme to apply: Hénon picked three of the necessary inequalities in Arnold’s proof, namely those which depend only on the dimension of the phase space (and not specifically on the perturbing function), and evaluated them numerically in the case of the circular restricted three-body problem in the plane. Hénon thus showed that the ratio had to be less than 10−320. The analogous computation in the plane (full) planetary three-body problem leads to the condition that the mass ratio should be less than 10−1130.

Since then, proofs of the invariant torus theorem have become simpler and their hypotheses are now somewhat less restrictive. Quantitative conditions for the KAM stability have been established using computer-assisted proofs, for similar systems (obtained by truncating the plane circular restricted three-body problem) notably in the works of Celletti-Chierchia [3, 4, 5], Giorgilli-Locatelli [9] and Robutel [14]. In another line of thought, quantitative results on the stability of the three-body problems over exponentially long time were also com-puted by Niederman [12] and later improved by the author [2].

This paper is the part of a study aiming at determining a sufficient condition for quasiperiodic solutions to exist in the plane planetary three-body problem. This study was split into three independent articles. A first one is dedicated to the complex singularities and the analyticity width of the perturbation, yielding an essential estimate of the disturbing function on complex domains of analyticity. The second one is the present paper, a technical part dedicated to the statement of a quantitative KAM theorem, where all the constants are explicit. Finally, a last one is dedicated to the application of this paper’s theorem to the plane planetary three-body problem, and the different issues that arise in such an application, such as the presence of degeneracies.

The aim of this paper is thus to state a quantitative KAM theorem, in which every involved constant is explicit. Such KAM theorems were stated before as they were a necessary step to obtain KAM stability for different systems. Notably, the work of Celletti and Chierchia [3, 4] and of Celletti, Giorgilli and Locatelli [6] are prominent examples of such statements. Here we prove a version of the KAM theorem of our own in order to fit our specific goal as close as possible (which, given the intricacies of the computations, is not useless). In this purpose, we have chosen to revisit the KAM theorem of Pöschel in [13] so as to derive explicit hypotheses depending on the parameters of the system, such as the analyticity widths in the actions and in the angles. Pöschel’s theorem is a version of the KAM theorem with parameters, using the frame-work of Moser [10]. As often in KAM theory, a given statement is well suited to some purposes, and less to some others. For example, Pöschel’s theorem deals directly with the Hamiltonian, whereas Celletti and Chierchia use a Lagrangian approach developed by Moser, Salamon and Zehnder [16], probably leading to

(3)

better constants since they avoid using Cauchy’s inequality at least once at each step of the induction. So we clearly do not aim at getting optimal estimates, but to get through the whole argument. Also, the relatively small number of hypotheses in Pöschel’s theorem make it possible to emphasize the competition between the different parameters, a pleasant feature for the application we have in mind.

1

Hamiltonian with parameter

1.1

Parameterizing the Hamiltonian

Consider, for n ≥ 2 the following analytic Hamiltonian:

H(p, q) = h(p) + f (p, q, ), (p, q) ∈ D × Tn,   1, (1) where D ⊂ Rn. The Hamilton equations of h leads to a quasi-periodic motion of frequency ω(p) = h0

(p) ∈ Rn

for p ∈ D. By assumption, the perturbation f (p, q, ) has small norm compared to the Hamiltonian. Assume that the unperturbed Hamiltonian is non-degenerate on the set D, that is det(h00(p)) = det∂ω

∂p(p)

 6= 0. Calling Ω = h0(D), the frequency map h0 : D → Ω is a local diffeomorphism between D and the frequency set Ω ⊂ Rn. The approach of Möser [10] consists in expanding the Hamiltonian h around one particular

frequency ω, and work with a linear Hamiltonian parameterized by that frequency. Let p = p0+ I, with I ∈ B = D − p0, the Hamiltonian h can be written:

h(p) = h(p0) + hh0(p0), Ii +

Z 1

0

(1 − t)hh00(p0+ tI)I, Iidt (2)

Since the frequency map is a local diffeomorphism, it is equivalent to work not only with the action p, but with the two variables (ω, I). Fixing the action p0 (and therefore ω), for I sufficiently small, one can write

equation (2) as follows:

h(p) = e(ω) + hω, Ii + Ph(I; ω), where Ph(I; ω) =

Z 1

0

(1 − t)hh00((h0)−1(ω) + tI)I, Iidt.

From the action-angle coordinates, we defined new coordinates (ω, I, θ) ∈ Ω × B × Tn, where we wrote θ instead of q not to be mistaken. The term Ph is of order 2 in the action I, and we now consider it as a new

part of the perturbation, being as small as wanted when considering a sufficiently small ball in the action around the origin.

Hence, one can write H = N + P , where N is called the normal form and is defined by N = e(ω) + hω, Ii, and

P = Ph(I; ω) + P(I, θ; ω), (3)

with P(I, θ; ω) = f ((h0)−1(ω) + I, θ, ).

The equations of motion of the family of Hamiltonian under normal form N are straightforward to compute. Indeed, their associated vector fields are of the form

XN = n

X

j=1

ωj∂θj, ω ∈ Ω.

The motion is quasi-periodic, and takes place on a specific torus {0} × Tn for every ω ∈ Ω. These tori can be seen as a trivial embedding of Tn over the set Ω on the phase space given by the function

Φ0: Tn× Ω → B × Tn, (θ, ω) 7→ (0, θ) (4)

For a generic Hamiltonian, the perturbation P will limit the existence of these tori. However, under hypothe-ses made clear further in the paper, one can show that almost all of these tori (in the sense of the Lebesgue measure) survive a perturbation, this is the main result of the KAM theorem.

(4)

1.2

Sets of analyticity and other definitions

Let Ω be the set of initial frequencies we are considering. Let τ > n − 1, γ > 0 and consider the set of Diophantine vectors: D(γ, τ ) =  ω ∈ Rn: ∀k ∈ Zn, |k · ω| ≥ γ |k|τ 1  , (5)

where |.|1is the l1-norm |k|1= |k1| + ... + |kn| for k ∈ Zn. Let Ωγ,τ = Ω ∩ D(γ, τ ), it is as well a Cantor set.

Finally, for β > 0, let

Ωβγ = Ωγ,τ \ {ω ∈ Ωγ,τ : ∃ ω0 ∈ Rn\ Ω, |ω − ω0| < β} .

The last set is the set of vectors in Ωγ,τ that are at least at a distance β to the boundary of Ω. We will fix

later the needed value for the constant β. Now let us define the various domains we use in the theorem: the sets we consider for the analytic version of the KAM thorem are complex polydiscs around some real set. For the frequencies, define

Oh= {ω ∈ Cn, |ω − Ωβγ| < h}. (6)

Call Tn

C= (T × R)

n the complex extension of the n−torus. The action-angle variable will take values in

Dr,s= {(I, θ) ∈ Cn× TnC, ∀i ∈J1, nK : |Ii| < r, |=(θi)| < s} . (7) Finally, we define the set

Dr,s,h= Dr,s× Oh. (8)

The norms on the vectors and matrices of Cn are the sup norm: ||I|| = sup

1≤i≤n

|Ii|, ||M || = sup 1≤i,j≤n

|Mi,j|.

Define the norms with indices as follows: for a function f defined respectively on Dr,s, Oh, Dr,s,h, we have

||f ||r,s= sup Dr,s |f | , ||f ||h= sup Oh |f |, ||f ||r,s,h= sup Dr,s,h |f |.

In the same spirit, for vector valued functions, we have: ||f ||r,s= sup Dr,s ||f ||, ||f ||h= sup Oh ||f ||, ||f ||r,s,h= sup Dr,s,h ||f ||,

where || · || is the sup norm. Finally, to state the theorem we will need the following Lipschitz norm on the frequencies:

|f |L= sup ω6=ω0

|f (ω) − f (ω0)|

|ω − ω0| ,

where | · | represent the supremum norm.

2

Quantitative KAM theorem

2.1

Statement of the theorem

In this section, we give the explicit statement of the KAM theorem. We chose τ = n in the following, guaranteeing the set of Diophantine vectors to be of positive measure for γ small enough, as well as leading to simpler hypotheses. However it is not hard to adapt the proof without this choice if required.

We introduce first some definitions to be able to state the theorem.

Let r, s, h > 0, ν = n + 1, δ ≥ 8(3n + 2), σ = s/20, and K0= min(A ∩ B+), where A = {K ∈ N : Kσ ≥ (2n + ν)log 2} , B =  K ∈ N : 2Kn+νσνe−Kσ ≤1 δ  , (9) B+= {K ∈ B : ∀m ∈ N, K + m ∈ B} .

(5)

Call C1= 4ν(8n(3n + 2)2C0+ n(n + 1)/2 + 8νn!)2with C0= 3π6 n 2

pn(2n)!

2n . Define the important value:

 = min γrσ ν 4νC 1 ,hr δ , γr 2K0νδ  . (10)

Define the following (exponentially convergent) series for ν ≥ 2:

Sν= ∞ X i=0 ni2ν  i+2−2(3 2) i , Tν= +∞ X i=0 ni4ν  i+1−(3 2) i , (11) and finally µ = exp 9n + 6 2δ  , ξ = exp  12(3n + 2)C0  γrσν  . (12)

The quantitative statement of the KAM theorem of Pöschel is the following:

Theorem 1. Let H = N + P be a Hamiltonian, such that P is real analytic on the set Dr,s × Oh and

kP kr,s,h = 0 ≤ . Then there exists a Lipschitz continuous map ϕ : Ωγ,τ → Ωhγ0, with h0 =

δ0

r , and a Lipschitz continuous family of real analytic torus embeddings Φ : Tn× Ωγ,τ → B × Tn close to the trivial

embedding Φ0 such that for each ω ∈ Ωγ,τ, the embedded tori are Lagrangian and

XH|ϕ(ω)◦ Φ = Φ0· XN.

Φ is real analytic on the set {θ ∈ Cn : |=θ

i| < s/2} for each ω, and the following inequalities on Φ and ϕ

hold: kW (Φ − Φ0)k ≤ 1 2ξ log ξ (13) kϕ − Idk ≤ 2h0µ log µ, (14) (15) where W = diag(r−1Id, s−1Id). As for their Lipschitz constant, we have:

kW (Φ − Φ0)kL≤ 2 3Tνξ log ξ (16) kϕ − IdkL≤ 4 3Sνµ log µ. (17)

It is important to notice that the Lipschitz estimates allow us to control the size of the the set ϕ(Ωγ,τ).

Indeed, with these estimates, one can prove that its complement is of size O(α) (see [13]). Besides, every embedded torus is Lagrangian and is close to its associated unperturbed torus.

2.2

Sketch of the proof

Let us describe succinctly the general scheme of the proof, that will consist in an iteration of a KAM step. At each step, we consider an analytic Hamiltonian H = N + P with N under normal form: N = e(ω) + hω, Ii. We want to find an analytic transformation F on the variables (I, θ, ω), close to the identity, such that H ◦ F = N++ P+, where N+is again under normal form and P+verifies ||P+|| ≤ C||P ||κ, for some constants

κ > 1 and C.

The composition of H with the function F implies some loss of analyticity, related to the norm of P . However, if the initial perturbation is small enough, the presence of the power κ will ensure that the norm of the perturbation will rapidly converge to zero after an infinite amount of steps, while the Hamiltonian stays analytic.

We are looking at a transformation F of the form

(6)

The transformation Φ acts on the action-angle variables, and depends on the parameter ω, while ϕ only changes the frequencies ω. In the following, the derivatives of these two functions are defined in the following way: Φ0 = (∂IΦ, ∂θΦ), and ϕ0= ∂ωϕ.

To build the transformation F , instead of considering the whole perturbation P , we first consider its linear part in the action by truncating its Taylor expansion:

Q(I, θ; ω) = P (0, θ; ω) + ∂IP (0, θ; ω) · I. (19)

The difference P − Q is then of order two in the actions, and will be part of the new perturbation at the end of our step. Next, we truncate the Fourier series of Q in the angle at some order K:

R(I, θ; ω) = X

k∈Zn,|k| 1≤K

Qk(I; ω) exp(ı2πθ). (20)

The remainder Q − R is the remainder of the Fourier series, and can be considered as small as wanted by taking K large enough, the function Q being analytic (see appendix A). Assume now that ω is fixed. Let F be a Hamiltonian affine in the actions, and XF = X its associated Hamiltonian vector field. Call Φt= ΦtX

the flow associated to the previous vector field, and Φ = Φtt=1its time-1 map.

Call ¯H = N + R, we have: ¯ H ◦ Φ = N + {N, F } + Z 1 0 (1 − t) {{N, F } , F } ◦ Φtdt + R + Z 1 0 {R, F } ◦ Φtdt = N + {N, F } + R + Z 1 0 {(1 − t) {N, F } + R, F } ◦ Φtdt.

We want F such that

N + {N, F } + R = N+. (21)

The term under the integral is small and therefore added to the perturbation P − R. The average of R not being necessarily zero, we divide R into two parts: R = ¯R + ˜R where

¯ R = 1 (2π)n Z Tn Rdθ. Formally, F is defined by F = X k∈Zn\{0},|k|1<K Rk ıhk, ωiexp(ık · θ),

where the Rk are the Fourier coefficients of the Hamiltonian R. As for the term ¯R, we simply add it to the

unperturbed Hamiltonian. Observe that since R was affine in the actions, N+ = N + ¯R remains affine in the

actions. Define

N+= e+(ω) + hω + v(ω), Ii.

The new frequency vector is given by ω+ = ω + v(ω), and if v is small enough and analytic, then there

exists a diffeomorphism ϕ close to the identity such that ϕ(ω+) = ω. The map N+ can then be written

N+= (N + ¯R) ◦ ϕ. The total transformation F = (Φ, ϕ) is now completely defined. As for the perturbation,

computing the remainders of the transformation gives:

P+=

Z 1

0

(1 − t) ¯R + tR, F ◦ Φtdt + (P − R) ◦ Φ.

If the perturbation is small enough, then P+will be even smaller, and we can iterate the scheme an infinite

(7)

3

Quantitative KAM step and its proof

Before proving the quantitative KAM theorem, we prove a quantitative KAM step, that will be iterated an infinite amount of times so as to obtain the complete theorem. Again, we follow almost completely the work of Pöschel in his proof. Recall as well that we chose τ = n as Diophantine constant for the sake of simplicity. 3.0.1 Statement of the KAM step

Proposition 1. Assume that kP kr,s,h ≤  where  is defined in (10). Then there exists a real analytic

transformation

F = (Φ, ϕ) : Dηr,s−5σ× Oh/4→ Dr,s× Oh

with η = r 

γrσν such that H ◦ F = N++ P+ with

kP+kηr,s−5σ,h/4≤ 8(3n + 2)2nC0 2 γrσν +  n(n + 1) 2 η 2+ 4νn!Kne−Kσ. (22) Moreover, 2kW (Φ − Id)k, kW (Φ0− Id)W−1k ≤ 8(3n + 2)C0 γrσν (23) kϕ − Idk, 4hkϕ0− Idk ≤ (6n + 4) r (24)

uniformly on Dηr,s−5σ× Oh and Oh/4 respectively, with the weight matrix

W = diag(r−1Id, σ−1Id).

Observations on the statement The different conditions (10) on  arise from different parts of the proof. The first condition in the minimum is a limit due to the analyticity widths in the actions and on the angles. This limit is necessary to obtain an exponential decrease of the bound on the norm of the perturbation at each step. The second condition is related to the transformation on the frequency vector. To be able to invert the map giving the new frequency ω + v(ω), it is essential to have enough analyticity width in the frequencies compared to the size of v. The third condition is a condition on K0, the order of truncature of the Fourier series of the affine perturbation. K0 needs to be large enough to allow the remainder of the Fourier series of Q to be small enough, and then for this remainder to decrease exponentially while iterating our scheme. However, we want as well K0 to be small enough so that all the frequencies in Oh satisfy a non-resonance

condition of order K0.

3.0.2 Proof of the proposition

Implications of the hypotheses: Let 0≤ . Define h0=

δ0 r and K0=  ν r γ 2h0 

. Using the definitions of  and K0 in given by (9) and (10),these two constants satisfy h0 < h and K0 > K0. Indeed, the first

inequality is clear, and for the second one, using the fact that both K0and K0are integers, and the definition

of : K0ν ≤ γr 2δ ≤ γr 2δ0 ≤ γ 2h0 .

The definition of h0 means that if we consider a smaller perturbation, we will not use all the available

ana-lyticity width h corresponding to the frequencies. These definitions allow us to compute a crucial inequality using the definitions of section 2.1, and that will be useful later in the proof:

1 2νK n 0σνexp(−K0σ) ≤ 0 γr ≤ h0 γδ ≤ 1 2δKν 0 . (25)

(8)

The last two inequalities are straightforward given the definition of h0and K0. Regarding the first one, using the definition of B+: 0 γrσν = h0 γσνδ ≥ 2K0n+νe−Kσ γ h0≥ K n 0e−K0σ ν s 2h0 γ ×  ν r γ 2h0 ! ν ≥ Kn 0e−K0 σ  ν r γ 2h0 −1 ×  ν r γ 2h0 !ν ≥ 1 2νK n 0e−K0 σ.

We now use the definition of K0 to show the non-resonance condition that the frequency vectors of Oh0 must

satisfy. Indeed, let k such that 0 < |k| ≤ K0, and let ω ∈ Oh0, there exists ω ?∈ Ωβ

γ such that |ω − ω?| < h0,

and therefore, the following inequalities hold:

|hk, ω − ω?i| ≤ |k||ω − ω?| ≤ K0h0≤ γ 2Kτ 0 ≤ γ 2|k|τ.

Since ω? satisfies a Diophantine condition for the constant γ and τ = n, we get:

|hk, ωi| ≥ γ

2|k|n, ∀0 < |k| ≤ K0

The goal at each step of the KAM theorem is to make a change of variables that decreases the norm of the perturbation. Yet, instead of making the value of 0 decrease directly, we want the ratio E0 = γrσ0ν to

decrease. Letting r, and σ decrease in a polynomial way, but with the value of E decreasing exponentially, we will as well obtain an exponential decrease for 0.

First estimates: As introduced in the outline of the proof in section 2.2, we switch from the analytic perturbation P to another analytic perturbation R in two steps. First, consider the linearization Q of P in the actions around the origin, secondly, truncate the Fourier series of Q at order K0. R is a trigonometric

polynomial on the angles and affine in the action.

The size of the perturbation, by assumption, is smaller than 0on the set Dr,s×Oh0. Let us start by bounding

the linearized function Q:

kQk3r 4,s,h0 ≤ kP k3r4,s,h0+ 3r 4 kPIk3r4,s,h0 ≤ kP kr,s,h0+ 3nr 4 kP kr,s,h0 r/4 ≤ (3n + 1)0. (26) Now let η = r  0

γrσν. Using the definition of  and the fact that 0≤ , we obtain 8η ≤ 1. From the Taylor

expansion formula and Cauchy’s inequality (see appendix D), after integration, we get:

kP − Qk2ηr,s,h0 ≤ (2ηr) 2n(n + 1) 2 kP kr,s,h0 (1 − 2η)2r2 ≤ n(n + 1) 2 η 2 0. (27)

With lemma 1 of appendix A, we obtain the following estimates on the difference between Q and its truncation at the order K0: kR − Qk3r 4,s−σ,h0 ≤ 4 n n!K0nexp(−K0σ)kQk3r 4,s,h0 ≤ 4νn!Kn 0exp(−K0σ)0. (28) Hence: kRk3r 4,s−σ,h0 ≤ kR − Qk 3r 4,s−σ,h0+ kQk 3r 4,s,h0 ≤ (3n + 1 + 4νn!Kn 0e−K0σ)0.

(9)

Since K0∈ A ∩ B+, we have the following inequality: 4νn!K0ne−K0σ = 4νn!2K n+ν 0 σ νe−K0σ 2Kν 0σν ≤1 δ 4νn! 2((2n + ν) log 2)ν ≤ 1.

Indeed, the right term depends only on n, and is decreasing with this variable. Since it takes a value less than 1 for n = 1, the result follows directly. Hence, we have

kRk3r

4,s−σ,h0 ≤ (3n + 2)0. (29)

Solving the cohomological equation: We would now like to solve the equation (21) in F . Letting ˆ

N = N+− N , we can write:

{F, N } + ˆN = R. Recall that in the outline, we wanted ˆN = ¯R = (2π)1n

R

TnRdθ. With the work done previously, we get the

following bound on ˆN :

k ˆN k3r

4,h0≤ kRk3r4,s−σ,h0 ≤ (3n + 2)0.

Since the Fourier series of R only contains terms of indices k such that |k|1≤ K0, we can apply the theorem

of Rüssmann 2 of appendix B, and solve the remainder of the cohomological equation (21). The norm of the Hamiltonian F solving this equation therefore verifies:

kF k3r 4,s−2σ,h0 ≤ C0kRk3r 4,s−σ,h0 γσn ≤ (3n + 2)C00 γσn .

We multiplied C0 by a factor 2, in order to get rid of this same factor in the Diophantine condition (5)

satisfied on Oh0. With Cauchy’s inequality, we get:

kFθkr 2,s−3σ,h0 ≤ (3n + 2)C00 γσν , kFIkr2,s−3σ,h0 ≤ 4(3n + 2)C00 γrσn .

Estimates on the transformation Φ: After obtaining the estimates on the derivatives of F , we can deduce some estimates on the vector field associated to F , and then on the time-1 map Φ. On the domain Dr 2,s−3σ,h0, we get: kFθkr 2,s−3σ,h0≤ r  0 γrσν r  0 γrσν(3n + 2)C0r ≤ (3n + 2)C0 2ν(8n(3n + 2)2C 0+ n(n + 1)/2 + 8νn!) ηr ≤ ηr ≤ r 8, kFIkr 2,s−3σ,h0≤ 4(3n + 2)C0 4ν(8n(3n + 2)2C 0+ n(n + 1)/2 + 8νn!)2 σ ≤ σ.

With these two inequalities, the time-1 map Φ is well-defined on the domains: Φ = Φt t=1: Dr

4,s−4σ,h0−→ Dr2,s−3σ,h0,

Φ = Φt t=1: Dηr,s−5σ,h0−→ D2ηr,s−4σ,h0.

Only considering the first domain is not enough to prove the KAM step. Indeed, the estimates of the difference P − Q requires to lose a lot of analyticity on the actions to keep this term small. Writing Φ = (U, V ), F being linear in the actions implies that V is independent of I. The Jacobian of F is

Φ0=UI Uθ 0 Vθ

 .

(10)

On the set Dr

8,s−5σ,h0, and hence on the set Dηr,s−5σ,h0, the following inequalities are satisfied

kU − Idk ≤ kFθk ≤ (3n + 2)C00 γσν , kV − Idk ≤ kFIk ≤ 4(3n + 2)C00 γrσn , kUI − Idk ≤ 8(3n + 2)C00 γrσν , kUθk ≤ (3n + 2)C00 γσν+1 , kVθ− Idk ≤ 4(3n + 2)C00 γrσν ,

whence the estimates (23) on Φ in proposition 1.

Estimates on the new perturbation: After the transformation, the Hamiltonian takes the form H = N++ P+. We want to obtain the bound (22) on the norm of P+. First, consider {R, F }:

k{R, F }kr 2,s−3σ,h0 ≤ n(kRIkr2,s−3σ,h0kFθkr2,s−3σ,h0+ kFIkr2,s−3σ,h0kRθkr2,s−3σ,h0) ≤ n 4(3n + 2)0 r (3n + 2)C00 γσν + (3n + 2)0 σ 4(3n + 2)C00 γrσn  ≤ 8(3n + 2)2nC0 2 0 γrσν (30) The same inequality stays true for { ˆN , F }. Hence:

Z 1 0 {(1 − t) ˆN + tR, F } ◦ XFtdt ηr,s−5σ,h 0 ≤ {(1 − t) ˆN + tR, F } r 2,s−4σ,h0 ≤ 8(3n + 2)2nC 0 2 0 γrσν. (31)

It remains to find the bound of the term induced by P − R: k(P − R) ◦ Φkηr,s−5σ,h0 ≤ kP − Rk2ηr,s−4σ,h0

≤ kP − Qk2ηr,s−4σ,h0+ kQ − Rk2ηr,s−4σ,h0

≤ (n(n + 1)/2η2+ 4νn!Kne−Kσ)

0. (32)

In the end, the estimate on the norm of P+ is the following:

kP+k ηr,s−5σ,h0 ≤ 8(3n + 2)2nC0 2 0 γrσν + (n(n + 1)/2η 2+ 4νn!Kne−Kσ) 0= +0. (33)

Exponential decrease: As explained before, we are interested in the exponential decrease of the ratio E = 0

γrσν. Let us already think about the iterative step, and choose the variables we will use at the next

step. First, let σ+ = σ/2, h+ = h0/4ν and K+ = 4K0; regarding the actions, since we have to lose much

more analyticity width and we let r+= ηr. With these definitions, let us compute E+:

E+=  + 0 γr+σ+ν = 2ν+ 0 γηrσν ≤ 8(3n + 2)2nC0 2ν η 2 0 (γrσν)2 + 2 ν(n(n + 1)/2η2+ 4νn!Kn 0e−K0 σ) 0 γηrσν ≤ 2ν8(3n + 2)2nC 0 E2 η + (n(n + 1)/2η 2+ 8νn!Kn 0e−K0σ) E η. Using the fundamental inequality (25), we have 2νE0≥ K0ne−K0

σ. Observe as well that E = η2, hence:

E+≤ 2ν(8(3n + 2)2nC 0+ n(n + 1)/2 + 8νn!)E 3 2 =pC1E 3 2, (34) i.e. C1E+ ≤ (C1E) 3

2. The scheme converges exponentially fast if E < C−1

1 , therefore if 0 ≤ γrσ ν C1 . The

initial condition on , and on 0 ≤  shows that we have even better: C1E ≤ 4−ν, hence the exponential

(11)

Change in the frequencies: It remains to deal with the function ϕ, which controls the frequency shift when adding the mean of the linearized perturbation over the angles. We use lemma 2 of appendix C, so as to make explicit the domain on which this map is well-defined. Let v = ˆNI = [RI], the new frequency vector

is defined by ω+= ω + v(ω). Computing the norm of v gives: kvkh0 = kNIkh0 ≤ 3n + 2 3r 4 − η 0≤ (6n + 4) 0 r ≤ (6n + 4)h0 δ ≤ h0 4 ,

because of the hypothesis δ ≥ 8(3n + 2). Applying lemma 2 of appendix C, we obtain the inverse map ϕ : Oh0 4 → Oh0, ω+ → ω, satisfying: kϕ − Idkh0 4 ≤(6n + 4)0 r , (35) kDϕ − Idkh0 4 ≤(3n + 2)0 2h0r . (36)

In this configuration, we let N+= (N + ˆN ) ◦ ϕ, and we obtained the new Hamiltonian H+ = N++ P+.

4

End of the proof of theorem 1

Soundness of the iteration:

To iterate the KAM step, the hypotheses at a step j + 1 need to be fulfilled knowing that they are at a step j. We therefore use the new value of each variables obtained after one KAM step, and check if they satisfy the hypotheses of the KAM step. Recall that after a step j, we have: Kj+1= 4Kj, σj+1= σj/2, ηj =

q j γrjσνj

, rj+1= ηjrj, hj+1= hj/4ν.

Regarding the equalities that need to be fulfilled, we first check that Kj+1 still belongs to the set A ∩ B+

defined in (9) using σj+1(we call the sets Aj+1and B+,j+1 referring to the value of σ at step j + 1):

• The equality Kj+1σj+1= 2Kjσj shows that Kj+1belongs again to the set Aj+1.

• Let us check that Kj+1belongs to B+,j+1 as well:

2Kj+1n+νσj+1ν exp(−Kj+1σj+1) = 2(4n+νKjn+ν) σν j 2ν  exp(−2Kjσj) ≤ 4 n+ν 2ν exp(−Kjσj) δ = 2 2n+νexp(−Kjσj) δ . The condition to have Kj+1∈ B+,j+1 is therefore the following:

22n+νexp(−Kjσj) ≤ 1.

Since Kj∈ Aj, it is satisfied.

We need now to check that the new size of the perturbation is less than the new value of  as defined in (10), using the analyticity widths at step j + 1:

• In the KAM step, we defined the limit value  = −. Define now

+= min γrj+1σ ν j+1 4νC 1 ,hj+1rj+1 δ , γ 2Kν j+1δ ! .

With the hypotheses, we have in fact

+= min ηj 2ν, ηj 4ν, 1 4ν  −=ηj − 4ν .

(12)

+ is therefore the new limit of the application of the KAM step. We have to check the condition  j+1≤ +. j+1= γrj+1σj+1ν Ej+1 ≤ ηj 2νγrjσ ν j p C1E 3 2 j ≤ ηj 2νγrjσ ν j p C1  3 2 j (γrjσjν) 3 2 ≤ηjj √ C1 2ν r  j γrjσνj . By assumption, j ≤ − ≤ γrjσνjC 1. Hence j+1 ≤ ηj −/4ν = +. The condition  ≤ γrσν 4νC 1 that we imposed

in the definition of , corresponding to the control of the transformation among the actions and the angles, allows us to iterate the KAM step here.

It remains to check that the fundamental inequality (25) holds at step j + 1. • We have:

Kj+1n exp(−Kj+1σj+1) = 4n(Kjnexp(−Kjσj)) exp(−Kjσj)

≤ 2ν − γrjσνj 4nexp(−Kjσj) ≤ 2ν4 ν+ ηj ηj 2νγr j+1σj+1ν 4nexp(−Kjσj) ≤ 2νE j+122n+νexp(−Kjσj).

Since Kj∈ Aj, we have: Kj+1n exp(−Kj+1σj+1) ≤ 2νEj+1. The first inequality of (25) is verified.

• We have to verify that the quantity j hjrj is decreasing: j+1 hj+1rj+1 =γσ ν j+1Ej+1 hj+1 = 2 νγσν jEj+1 hj = 2νEj+1 Ej j hjrj ≤ 2νpC1Ej j hjrj ≤ j hjrj .

The last inequality of (25) holding straightforwardly, we can now iterate the KAM step an infinite amount of time. The scheme is well-defined, and we can now compute the size of the transformations.

Transformations involved and their estimates:

The initial Hamiltonian is H = N + P . At each KAM step, we define two transformations: Φjwhich modifies

the action-angle coordinates, and ϕj which modifies the frequencies. We let sj+1 = sj− 5σj, with s0 = s,

r0= r, η0= η and ηj =

q j γrjσν

j

=pEj. Define F0= Id, and for j > 0:

Fj+1: Dj+1× Oj+1→ Dj× Oj

(I, θ, ω) 7→ (Φj+1(I, θ; ω), ϕj+1(ω)),

with

Dj = {I ∈ Cn: kIk < rj} × {θ ∈ Tn: |=(θi)| ≤ sj} ,

Oj =ω ∈ Rn: |ω − Ωβγ| < hj .

Call Fj = F0◦ ... ◦ Fj−1. We then have:

Fj = (Φj, ϕj) : Dj× Oj→ D0× O0

(13)

Preliminaries: The map F = limj→+∞Fj transforms a torus associated to a frequency vector belonging

to the set Ωγ,τ to a deformed torus where the motion has frequencies belonging to the set Ωβγ. The action

p0 on the first torus are entirely and uniquely determined by the frequency vector. The uniqueness comes

from the hypothesis of non-degeneracy of the unperturbed Hamiltonian. In order to be precise, we define the following mapping:

Ψ : Tn× Ωγ,τ → D × Tn.

Define as well the map:

Ξ : B × Tn× Ωγ,τ → D × Tn (37)

(I, θ, ω) 7→ ((h0)−1(ω) + I, θ). (38) Assume Φ : {0} × Tn→ B × Tn and ϕ : Ωβ

γ → Ωγ,τ exists as a limit of Φj and ϕj.

Then, one can define Ψ as follows:

Ψ : Tn× Ωγ,τ → D × Tn

(θ, ω) 7→ Ξ(Φ(0, θ), ϕ(ω)) The KAM theorem shows that, on T?× Ω

γ,τ, H ◦ Ψ = N0, where N0= limj→∞Nj.

Estimates on the transformations: In order to simplify the formulas, we introduce the weight matrix Wj = diag(r−1j Id, σ

−1

j Id). Recall the size of the transformation obtained previously in proposition 1 on the

set Dj× Oj: kWj(Φj− Id)k ≤ 4(3n + 2)C0j γrjσνj , kϕj− Idk ≤ (6n + 4)j rj , Wj(Φ0j− Id)Wj−1 ≤ 8(3n + 2)C0j γrjσνj , ϕ0j− Id ≤ (3n + 2)j 2hjrj .

We can estimate the norm of the difference between to consecutive transformations Fj.

W0(Φj+1− Φj) = W0(Φj◦ Φj− Φj) ≤ 2n W0(Φj)0Wj−1 kWj(Φj− Id)k ≤ 2nξjkWj(Φj− Id)k W0(Φj+1− Φj) ≤ ξj 8(3n + 2)nC0j γrjσjν , (39)

it is well-defined when j goes to infinity if the variable ξj= kW0DΦjWj−1k does not increase too fast on Dj.

In the same way, we compute: ϕj+1− ϕj = ϕj◦ ϕj− ϕj ≤ n (ϕj)0 kϕj− Idk ≤ nµjkϕj− Idk ϕj+1− ϕj ≤ µj n(6n + 4)j rj , (40)

where again it is necessary to check the increase of µj =

(ϕj)0 on Oj.

On Dj× Oj, we have in fact (Φj)0 = (Φ0)0· ... · (Φj−1)0, where the differentials are estimated on different

points, that are not important to explicit in our case, as we have a bound on their whole set of definition. With the decrease of the variables r and σ, we get WjWj+1−1

(14)

norm of the product between two matrices, as well as the fact that the matrices Wj are diagonal, we have ξj = W0(Φj)0Wj−1 = W0(Φ0)0· ... · (Φj−1)0Wj−1 ≤ W0(Φ0)0W0−1 W0W1−1 2n × ... × 2n Wj−1(Φj−1)0Wj−1−1 Wj−1Wj−1 ≤ 1 2 j (2n)j−1 j−1 Y i=0  1 +8(3n + 2)C0i γriσνi  ≤n j−1 2 j−1 Y i=0  1 +8(3n + 2)C0i γriσiν  . (41) Likewise, for µj: n1−jµj≤ j−1 Y i=0  1 + 5i 2hiri  ≤ j−1 Y i=0  1 +(3n + 2)i 2hiri  . (42)

Since the variables i decreases exponentially fast towards 0, and that the terms hi and ri do not decrease

as fast, the products in the formulas will converge when j goes to infinity. We can bound them using the estimates we obtained in the KAM step. First, recall that 8(3n+2)C0i < γriσiν, whence, using the logarithm

for i ≥ 0: log 2n1−jξj ≤ j−1 X i=0 log  1 +8(3n + 2)C0i γriσiν  ≤ j−1 X i=0 8(3n + 2)C0i γriσνi ≤ j−1 X i=0 8(3n + 2)C0Ei.

Using the exponential decrease of Ej:

Ej≤ p C1E 3 2 j−1≤ ... ≤ C 1 2 Pj−1 i=0(32) i 1 E (3 2) j 0 ≤ (C1E0)( 3 2) j −1E 0≤ 4−ν( 3 2) j +νE 0. Finally: 2n1−jξj ≤ exp ∞ X i=0 8(3n + 2)C0E04ν−ν( 3 2) i ! ≤ exp 8(3n + 2)C0E0 ∞ X i=0 4ν−ν(32) i ! ≤ exp (12(3n + 2)C0E0) = exp  12(3n + 2)C0 0 γr0σ0ν  ≡ ξ. (43)

In the same way, we get for µj:

n1−jµj≤ exp 3n + 2 2 ∞ X i=0 i rihi ! = exp 3n + 2 2 ∞ X i=0 γEiσiν hi ! ≤ exp 3n + 2 2 ∞ X i=0 γEiσ0ν2 νi h0 ! ≤ exp 3n + 2) 2 γσν 0 h0 ∞ X i=1 E02ν(i+2−2( 3 2) i ! ≤ exp 3(3n + 2) 2 0 r0h0  ≤ exp 3(3n + 2) 2δ  ≡ µ. (44)

(15)

With these computations, we can continue towards our aim of estimating Fj for all j ≥ 0, using (39) and (43). W0(Φj− Φ0) ≤ j−1 X i=0 W0(Φi+1− Φi) ≤ j−1 X i=0 ξi 8(3n + 2)nC0i γriσiν ≤ 4(3n + 2)C0ξ ∞ X i=0 niEi  ≤ 4(3n + 2)C0E0ξ ∞ X i=0  ni4−ν(32) i +ν ≤ 6(3n + 2)C0E0ξ = 1 2ξ log ξ (45)

As well, using (40) and (44), for all j ≥ 0,

h−10 (ϕj− Id) ≤ j−1 X i=0 h−10 (ϕi+1− ϕi) ≤ j−1 X i=0 µ(6n + 4)n j i h0ri ≤ 2µ log µ. (46)

Therefore, with these uniform bounds, we can let j go to infinity. The transformation F is well-defined on T?× Ω

γ,τ. The set Ωβγ, defined while constructing ϕj, depends on Ωγ,τ and on h0. More precisely, recall that

for all ω0 ∈ Ωβ

γ, there exists ω ∈ Ωγ,τ such that |ω − Ωβγ| < h0. Therefore, we can let β = h0so that the set

Oh⊂ Ω.

First conclusion on the transformation: Before computing the Lipschitz norm of the transformation, we are going to give some conclusion on the transformation we built. First, we have the relation H ◦Ξ◦Fj−Nj=

Pj on the set D

j× Oj for all j ≥ 1. With this equality, we can describe the difference between the vector

field associated to H and the one associated to Nj. Formally, by derivation on the action-angle coordinates,

we have

t(H ◦ Ξ ◦ Fj− Nj)0 =t(Ξ ◦ Φj)0· ∇H ◦ Ξ ◦ Fj− ∇Nj.

For a fixed ω, the map Ξ is constant and linear in the actions and the angles, which simplifies the computation. Therefore, using Cauchy’s inequality on the action coordinates, we can bound the previous derivative.

t(Φj)0· ∇H ◦ Ξ ◦ Fj− ∇Nj 0,sj−σj ≤ max  j rj ,j σj  .

Using the weighted matrix Wj which was useful to give the estimates on (Φj)0, and the symplectic matrix J ,

J =  0 Id −Id 0  , Wj× J =  0 σ−1 j −rj−1 0  ,

and multiplying on the left our relation by the latter matrix, we obtain:

kWj× J ×t(Φj)0· ∇H ◦ Ξ ◦ Fj− Wj× J × ∇Njk0,sj−σj ≤

j

rjσj

.

The map Φj being symplectic, it satisfies JtΦ0 = (Φ0)−1J . Hence, multiplying the last inequality by W0Φ0Wj−1, we get: W0 J ∇H ◦ Ξ ◦ Fj− Φ0J ∇Nj  0,sj−σj ≤ kW0Φ0Wj−1k0,sj−σj j rjσj ≤ n j−1 2 ξ j rjσj

Looking at the vector field, this inequality becomes:

kW0k × kXH◦ Ξ ◦ Fj− (Φj)0· XNk0,sj−σj ≤

nj−1 2 ξ

j

(16)

Letting j go to infinity, we get the equality of these two vector fields on the set T?× Ω

γ,τ, i.e., for some

ω ∈ Ωγ,τ, with N = (e(ω) + hω, Ii,

XH◦ Ψ = Φ0· XN (47)

Observe that we wrote Ψ instead of Φ, it expresses the fact that we consider the "origin" of the action I at the point ϕ(ω).

Lipschitz norm of the transformation: On the Cantor set Ωh0

γ ⊂ Ω, the formulas of the derivatives of

(ϕj− Id) converge. Indeed, the existence of a uniform constant for ξ

j/(2n2)j and µj, and the exponentially

fast convergence of the terms j rjhmj

for all m implies that the norm of every derivative of the mapping Fj

(more precisely its difference to (Φ0, Id)) converges, whatever the order of the derivative.

We compute the Lipschitz norm of the transformation F to finish proving the theorem. First we evaluate the derivative of the map ϕ − Id with respect to ω. Although the map ϕ is defined on a Cantor set, it is possible to extend it in a Lipschitz way, and even in a C1 map, using Whitney’s extension theorem (see article [17]

or the statement of the theorem 4 in appendix D.3). We will not extend further on these notions, and will just compute an estimate on its norm.

k(ϕj− Id)0khj 2 ≤ j−1 X i=0 k(ϕi+1)0− (ϕi)0khi 2 ≤ j−1 X i=0 k(ϕi+1− ϕi)0khi 2 ≤ j−1 X i=0 2kϕi+1− ϕik hi hi ≤ 4(3n + 2)µ j−1 X i=0 ni i rihi .

As done before, we can sum these terms:

k(ϕj)0− I nkhj 2 ≤ 4(3n + 2) 0 r0h0 µ j−1 X i=0 ni2ν  i+2−2(3 2) i ≤4 3µ log µ ∞ X i=0 ni2ν  i+2−2(3 2) i ≡ 4 3µ log µ × Sν. Hence, letting j goes to infinity, we obtain the Lipschitz norm:

kϕ − IdkL≤

4

3Sνµ log µ. (48)

As for Φ − Φ0, computing the Lipschitz estimate in the exact same way, we get:

kW0(Φ − Φ0)kL ≤

2

3Tνξ log ξ, (49) where we recover the expression of Tν given in (11):

Tν ≡ +∞ X i=0 ni4ν  i+1−(3 2) i .

5

Estimates in the initial actions

After giving the estimates on the map F = (Φ, ϕ), we determine some estimates on the map Ψ = Ξ ◦ F where Ξ was defined in (38). Indeed, using this transformation we shift the torus back to its original place around the action p0.

For the sake of simplicity, call g the Legendre transform of the unperturbed Hamiltonian h: g(ω) = supp(ω · p−

h(p)); we then have g0= (h0)−1.

Call Ψ0(I, θ, ω) = Ξ(Φ0(θ, ω), ω), we are interested in the difference Ψ − Ψ0:

(17)

where Φ(I, θ) = (Φ1(I, θ), Φ2(I, θ)). Hence, the norm of Ψ − Ψ0 on the set T?× Ωγ,τ satisfies: kW0(Ψ − Ψ0)k ≤ 1 rkg 0◦ ϕ − g0k + kW 0(Φ − Φ0)k ≤ n sup Ωγ,τ kg00kkϕ − Idk r + kW0(Φ − Φ0)k, kW0(Ψ − Ψ0)k ≤ sup Ωγ,τ kg00k × 2h0µ log µ + 1 2ξ log ξ. (50) We can also compute an estimate on the Lipschitz norm of Ψ with respect to ω. The estimate on the Lipschitz norm of Φ − Φ0 being known, we are interested in the map Υ(ω) = g0(ϕ(ω)) − g0(ω). Let ω, ω0 ∈ Ωγ,τ, we

have: |Υ(ω) − Υ(ω0)| = Z 1 0 [g00(ω + t(ϕ − Id)(ω)) · (ϕ − Id)(ω)] dt − Z 1 0 [g00(ω0+ t(ϕ − Id)(ω0)) · (ϕ − Id)(ω0)] dt . (51)

In order to compute this norm, we need to add some intermediate terms under the integral. For the sake of simplicity, we use v = ϕ − Id, we have

g00(ω + tv(ω)) · v(ω) − g00(ω + tv(ω)) · v(ω) =

g00(ω + tv(ω)) · [v(ω) − v(ω0)] + [g00(ω + tv(ω)) − g00(ω0+ tv(ω0))] · v(ω0). The first term of this sum is bounded by

|g00(ω + tv(ω)) · [v(ω) − v(ω0)] | ≤ n sup

Ωγ,τ

|g00| × |ϕ − Id|L× |ω − ω0|.

As for the second term, we write:

|g00(ω + tv(ω)) − g00(ω0+ tv(ω0))| = Z 1 0 g(3)((1 − s)(ω + tv(ω)) − s(ω0+ tv(ω0))ds · |(ω + tv(ω)) − (ω0+ tv)| ≤ n sup Ωγ,τ g (3) × (|ω − ω 0| + t|v(ω) − v(ω0)|) ≤ n sup Ωγ,τ g (3) × (1 + t|ϕ − Id|L) |ω − ω 0|.

Injecting these bounds in inequality (51), we obtain

|Υ|L≤ n sup Ωγ,τ |g00| × |ϕ − Id|L+ n2sup Ωγ,τ g (3) × |ϕ − Id|L×  1 + |ϕ − Id|L 2  . (52) Finally, |Ψ − Ψ0|L≤ |Υ|L+ |Φ − Φ0|L. (53)

I would like to thank the reviewer for his critical remarks and his precious advices that were very helpful to me.

(18)

A

Remainder of the truncated Fourier series

Let Asbe the set of functions defined on Tnthat are bounded and analytic on the set Tns = {θ ∈ TnC, |=θ| < s}.

Let f ∈ As, for θ ∈ Tns, we can write

f (θ) = X

k∈Zn

fkeık·θ.

For all k ∈ Zn, we have |f

k| ≤ |f |se−|k|s. Indeed, this result is straightforward using the fact that f is

2π-periodic in each variable, and analytic and bounded on its set of definition. Let us consider the truncation of order K ∈ N of f :

TKf =

X

k≤K

fkeık·θ.

Lemma 1. Let s > 0 and σ < s. If f ∈ As, and Kσ ≥ n − 1 then

|f − TKf |s−σ≤ 4nn!Kne−Kσ|f |s, 0 ≤ σ ≤ s Proof. We have: |f − TKf |s−σ≤ X k∈Zn,|k|1>K |fk| exp(|k|1(s − σ)) ≤ |f |s X k∈Zn,|k| 1>K exp(−|k|σ) ≤ 4n|f | s X l∈N,l>K ln−1exp(−lσ),

where we used the fact that the number of k ∈ Znsuch that |k|

1= l is less than 4nln−1. As for the last sum,

since the general term is strictly decreasing, it can be bounded by the incomplete gamma function: X l∈N,l>K ln−1exp(−lσ) ≤ Z ∞ K xn−1exp(−xσ)dx ≤ 1 σn Z ∞ Kσ xn−1exp(−x)dx ≤(n − 1)! σn exp(−Kσ) n−1 X k=0 (Kσ)k k! ≤ (n − 1)! σn exp(−Kσ)n × (Kσ) n−1 ≤ n!K n−1 σ exp(−Kσ) ≤ n!K nexp(−Kσ).

Injecting this result in the previous inequation, the lemma is proved.

B

Rüssman optimal estimate on the cohomological equation

We recall here the result obtained by Rüssmann in [15], which gives an optimal estimate on the norm of the solution for generic analytic functions. Let n ≥ 1 and s > 0, define:

Tns = {θ ∈ T n C, ∀i ∈J1, nK : |=θi| < s} As = {f : Tns → C, f C -analytic} As 0 = {f ∈ A s, s.t. Z Tn f = 0}.

Writing |f |s= supTns|f |, we have the following theorem:

Theorem 2 (Rüssmann). Let ω ∈ Dγ,τ a Diophantine vector, and g ∈ As0. Then the equation

(19)

has a unique solution f inS

0<σ<sA s−σ

0 , and we have the following bound on the norm of f for 0 < σ < s:

|f |s−σ≤ C0 γστ|g|s, where C0= 3π2 6 n 2 √ τ Γ(2τ ) 2τ .

C

Inversion of analytic map close to the identity

Recall that we defined the set Oh as the open complex neighborhood of radius h of the subset of frequencies

Ωγ,τ for some γ > 0. Using as usual the sup-norm for maps and vectors, Pöschel proves the following lemma

on the inversion of the frequency vector:

Lemma 2. Let f : Oh⊂ Cn → Cn be analytic, such that |f − Id| ≤ δ ≤ h/4 on the set Oh. Then f has an

analytic inverse g on Oh/4, and it satisfies:

|g − Id|h/4,

h 4|g

0− Id| ≤ δ

See the appendix of Pöschel’s paper [13] for the proof.

D

Classical formulas for analytic multivariate functions

In this section, we recall Taylor’s theorem and Cauchy’s formula for multivariate functions (see [11] for more details on the latter). As well, we state the Whitney extension theorem (for the demonstration, see [17]). Define the following notations for α ∈ Nn and x ∈ Rn with n > 0:

|α| = α1+ ... + αn,

α! = α1!...αn!,

xα= xα11 ...xαn n .

Introduce as well for an analytic function f :

f(α)= ∂

|α|f

∂xα11...∂ αn xn

D.1

Taylor expansion of analytic function

Theorem 3. Let f : Rn → R be a function analytic at the point a ∈ Rn. Then, for k ≥ 0, there exists a

function Rk : Rn→ R such that:

f (x) − X |α|≤k f(α) α! (x − a) α= R k(x) = o((x − a)k).

Moreover, on a closed ball B around a, we have for x ∈ B: Rk(x) =

X

|β|=k+1

Rk,β(x)(x − a)β,

with the bound

max x∈B|Rk,β(x)| ≤ 1 β!|α|=|β|max  max x0∈Bf (α)(x0)  .

(20)

D.2

Cauchy Formula

Let a ∈ Cn and ρ = (ρ1, ..., ρn) with ρi> 0. Define the polydisc with center a and radius ρ:

P (a, ρ) = {z ∈ Cn, s.t. |zi− ai| < ρi for i ∈J1, nK}

Proposition 2 (Cauchy’s formula). Let Ω be an open set in Cn, f a function holomorphic on Ω, a ∈ Ω and

let ρ = (ρ1, ..., ρn) with ρi> 0 be such that P (a, ρ) ⊂ Ω. Then, for z ∈ P (a, ρ), we have

f (z) = 1 (2πı)n Z |ζ1−a1|=ρ1 ... Z |ζn−an|=ρn f (ζ1, ..., ζn) (ζ1− z1)...(ζn− zn) dζ1...dζn.

Corollary 1 (Cauchy’s inequality). If f is holomorphic on Ω and P (a, ρ) ⊂ Ω, we have

|f(α)(a)| ≤ sup |ζi−ai|=ρj |f (ζ)| ! α!ρ−α.

D.3

Whitney theorem

Define, for some function f defined on Rn, some m ∈ N, and some α ∈ Nn such that |α| ≤ n, the following functions: fα(x) = X |k|≤m−|alpha| fk+α(x0) k! (x − x 0)k+ R α(x, x0).

Now let A be a closed subset of Rn. We will need some condition of smoothness in this set.

Definition 1 (Cm in the Whitney sense). Let f be a function defined in the set A and let m be a positive integer. f is said to be of class Cmin A in the Whitney sense if the functions fα (|α| ≤ m) are defined in A

and the remainders Rαare such that for any point x of A, any  > 0, there exists δ > 0 such that if x0 and

x00are any two points of A ∩ B(x, δ) then

|Rα(x0, x00)| ≤ |x0− x00|m−|α|.

With this definition, we have the following statement:

Theorem 4 (Whitney). Let A be a closed subset of Rn and let f be of class Cm (m finite or infinite) in the

Whitney sense. Then there is a function F of class Cm

(in the ordinary sense) in Rn such that

(1) F(α)(x) = fα(x) in A for |α| ≤ m,

(2) F (x) is analytic in Rn.

In particular, f = F A.

References

[1] V. Arnold. Small denominators and problems of stability of motion in classical and celestial mechanics. In A. B. Givental, B. A. Khesin, J. E. Marsden, A. N. Varchenko, V. A. Vassiliev, O. Y. Viro, and V. M. Zakalyukin, editors, Collected Works: Representations of Functions, Celestial Mechanics and KAM Theory, 1957–1965, pages 306–412. Springer Berlin Heidelberg, Berlin, Heidelberg, 2009.

[2] T. Castan. Estimates on stability over exponentially long times in the plane planetary problem. 2017. [3] A. Celletti and L. Chierchia. On the stability of realistic three-body problems. Comm. Math. Phys.,

186(2):413–449, 1997.

[4] A. Celletti and L. Chierchia. Kam stability for a three-body problem of the solar system. Zeitschrift für angewandte Mathematik und Physik ZAMP, 57(1):33–41, 2005.

(21)

[5] A. Celletti and L. Chierchia. KAM stability and celestial mechanics. Mem. Amer. Math. Soc., 187(878):viii+134, 2007.

[6] A. Celletti, A. Giorgilli, and U. Locatelli. Improved estimates on the existence of invariant tori for Hamiltonian systems. Nonlinearity, 13(2):397–412, 2000.

[7] A. N. Kolmogorov. On conservation of conditionally periodic motions for a small change in Hamilton’s function. Dokl. Akad. Nauk SSSR (N.S.), 98:527–530, 1954.

[8] J. Laskar. Michel Hénon and the Stability of the Solar System. ArXiv e-prints, November 2014. [9] U. Locatelli and A. Giorgilli. Invariant tori in the Sun-Jupiter-Saturn system. Discrete Contin. Dyn.

Syst. Ser. B, 7(2):377–398, 2007.

[10] J. Moser. Convergent series expansions for quasi-periodic motions. Mathematische Annalen, 169:136– 176, 1967.

[11] R. Narasimhan. Several complex variables. Chicago Lectures in Mathematics. University of Chicago Press, Chicago, IL, 1995. Reprint of the 1971 original.

[12] L. Niederman. Stability over exponentially long times in the planetary problem. Nonlinearity, 9(6):1703– 1751, 1996.

[13] J. Pöschel. A lecture on the classical KAM theorem. In Smooth ergodic theory and its applications (Seat-tle, WA, 1999), volume 69 of Proc. Sympos. Pure Math., pages 707–732. Amer. Math. Soc., Providence, RI, 2001.

[14] P. Robutel. Stability of the planetary three-body problem. II. KAM theory and existence of quasiperiodic motions. Celestial Mech. Dynam. Astronom., 62(3):219–261, 1995.

[15] H. Rüssmann. On optimal estimates for the solutions of linear partial differential equations of first order with constant coefficients on the torus. volume 38, pages 598–624. Springer Berlin Heidelberg, Berlin, Heidelberg, 1975.

[16] D. Salamon and E. Zehnder. Kam theory in configuration space. Commentarii Mathematici Helvetici, 64(1):84–132, Dec 1989.

[17] H. Whitney. Analytic extensions of differentiable functions defined in closed sets. Trans. Amer. Math. Soc., 36(1):63–89, 1934.

Références

Documents relatifs

G., Novikov, Approximate solution of the inverse problem of quantum scattering theory with fixed energy in dimension 2, (Russian) Tr. Novikov, Formulae and equations for

For the case when τ = 1 in (1.9), (1.11) or when s = 0 in (1.14) the history of such estimates in dimension d = 3 goes back to [29], [31], where such energy and regularity

This problem can be considered as the Gel’fand inverse boundary value problem for the Schr¨ odinger equation at zero energy (see [9], [16])... A variation of this result is presented

If the linearizations at unstable equilibria have at least one eigenvalue with positive real part, then almost global asymptotic stability turns out to be robust with respect

Previously, some stability estimates for inverse problems involving the stationary transport equation have been obtained by Romanov in [4].. He used different

of the perturbing function); a closer analysis of the proofs of the KAM stability theorems shows that in fact the estimates met in the proofs are.. not really bad

We consider the problem of existence and stability of solitary traveling waves for the one dimen- sional discrete non linear Schrödinger equation (DNLS) with cubic nonlinearity,

Moreover, in the zero range model, there are energy crossings between universal and efimovian states as a function of R t /a ho (see solid lines in Fig.2a); as we shall see, for