This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Kolmogorov’s Theorem for Degenerate Hamiltonian Systems with Continuous Parameters

Jiayin Dua,∗ 111 E-mail address : [email protected] , Yong Lia,b 222 E-mail address : [email protected], Hongkun Zhangc 333E-mail address : [email protected]
 *Corresponding author

aCollege of Mathematics and Statistics, and Center for Mathematics and Interdisciplinary Sciences, Northeast Normal University, Changchun, 130024, P. R. China.
bCollege of Mathematics, Jilin University, Changchun, 130012, P. R. China.
cDepartment of Mathematics and Statistics, University of Massachusetts, Amherst, 01003, USA
Abstract

In this paper, we investigate Kolmogorov type theorems for small perturbations of degenerate Hamiltonian systems. These systems are index by a parameter ξ\xi as H(y,x,ξ)=ω(ξ),y+εP(y,x,ξ,ε)H(y,x,\xi)=\langle\omega(\xi),y\rangle+\varepsilon P(y,x,\xi,\varepsilon) where ε>0\varepsilon>0. We assume that the frequency map, ω\omega, is continuous with respect to ξ\xi. Additionally, the perturbation function, P(y,x,,ε)P(y,x,\cdot,\varepsilon), maintains Hölder continuity about ξ\xi. We prove that persistent invariant tori retain the same frequency as those of the unperturbed tori, under certain topological degree conditions and a weak convexity condition for the frequency mapping. Notably, this paper presents, to our understanding, pioneering results on the KAM theorem under such conditions-with only assumption of continuous dependence of frequency mapping ω\omega on the parameter.

keywords:
Hamiltonian system, invariant tori, frequency-preserving, Kolmogorov’s theorem, degeneracy, continuous parameter.
MSC:
[2020] 37J40 , 70H08 , 70K43
journal: a

1 Introduction

As a conservation law of energy, Hamiltonian systems are frequently considered to describe models arose in celestial mechanics or the motion of charged particles in magnetic fields, see [9, 24, 38].

The classical KAM theory, as presented by Arnold, Kolmogorov, and Moser [1, 20, 26], posits that under the Kolmogorov non-degenerate condition, most invariant tori of an integrable Hamiltonian system can withstand small perturbations. While these tori might undergo minor deformations, they transform into other invariant tori that retain the original frequency

Numerous methods have been explored to study the persistence of invariant tori and the preservation of toral frequency within Hamiltonian systems under certain non-degenerate conditions. For instance, the KAM approach was used in [2, 4, 12, 22, 30, 35]. The direct method using Lindstedt series can be found in references [8, 13, 15], while renormalization group techniques were discussed in [5, 16]. Notably, the study presented in [10] introduced the idea of partial preservation of unperturbed frequencies and delved into the persistence problem on a specified smooth sub-manifold for real analytic Hamiltonian systems, particularly under the Rüssmann-like non-degenerate condition. For insights under analogous conditions, see also [37].

Yet, in the context of persistence, two fundamental questions emerge that warrant attention:

Q1: In the event of a failure in the Kolmogorov non-degenerate condition, can the invariant tori with the same frequency still be preserved under small perturbations?

Q2: If the regularity of the frequency mapping diminishes to mere continuity, can the aforementioned result withstand small perturbations?

To shed light on these questions, we review previous findings and offer a more comprehensive overview.

1.1 Degeneracy

Consider the real analytic nearly integrable Hamiltonian system

H(y,x,ε)=h(y)+εP(y,x,ε),\displaystyle H(y,x,\varepsilon)=h(y)+\varepsilon P(y,x,\varepsilon), (1.1)

where xx is the angle variable in the standard torus 𝕋n\mathbb{T}^{n}, nn refers to the dimension; yy is the action variable in a bounded closed region GnG\subset\mathbb{R}^{n}, and ε>0\varepsilon>0 is a small parameter.

A fundamental assumption in historical research is the Kolmogorov non-degenerate condition. However, if we assume that there exists a y0Gy_{0}\in G such that,

det2h(y0)y2=0,\displaystyle\det\frac{\partial^{2}h(y_{0})}{\partial y^{2}}=0,

then the Kolmogorov condition is not satisfied. The spatial solar system serves as a prominent example of this situation, as detailed in [14]. Naturally, a question arises: does the persistence result still stand under these conditions? This question has been a primary motivation for this research.

In fact, even under weaker non-degenerate conditions, KAM tori might not preserve their frequencies. As demonstrated in [6, 34, 36], under the Brjuno non-degenerate condition and Rüssmann non-degenerate condition, the presumption of an unchanged frequency may not necessarily hold true. This is because the frequency of perturbed tori can undergo slight variations. Similar observations are noted in [3, 7, 10, 11, 18, 32, 39]. Consequently, deriving conditions that assure the persistence of frequencies for KAM tori in the context of a degenerate Hamiltonian becomes rather challenging. Furthermore, the issue of the perturbed invariant tori maintaining a consistent frequency has seldom been tackled for degenerate systems.

1.2 Regularity

On the matter of regularity, it’s worth noting the distinctions in the studies of various researchers. Kolmogorov [20] and Arnold [1] focused on real analytic Hamiltonian systems. In contrast, Moser [26] illustrated that Hamiltonian systems don’t necessarily need to be analytic; a high, albeit finite, level of regularity for the Hamiltonian suffices. This regularity requirement was later reduced to C5C^{5} in work by [33]. Further important contributions on this topic can be found in [4, 21, 19, 35]. Moreover, the scenario where the frequency mapping has Lipschitz continuous parameters has been explored in [29]. A subsequent question of interest is: what are the implications when the regularity of the frequency mapping is merely continuous with respect to its parameters?

More precisely, we consider a family of Hamiltonian systems under small perturbations:

H(y,x,ξ,ε)=ω(ξ),y+εP(y,x,ξ,ε),H(y,x,\xi,\varepsilon)=\langle\omega(\xi),y\rangle+\varepsilon P(y,x,\xi,\varepsilon), (1.2)

where (y,x)G×𝕋n(y,x)\in G\times\mathbb{T}^{n} and ξ\xi is a parameter in a bounded closed region OnO\subset\mathbb{R}^{n}. The function ω()\omega(\cdot) is continuous with respect to ξ\xi on OO. The function P(,,ξ,ε)P(\cdot,\cdot,\xi,\varepsilon) is real analytic with respect to yy and xx, and P(y,x,,ε)P(y,x,\cdot,\varepsilon) is Hölder continuous with respect to the parameter ξ\xi with Hölder index β\beta, for some 0<β<10<\beta<1. Additionally, ε>0\varepsilon>0 is a small parameter.

It’s important to note that in the conventional KAM iteration process, the regularity of the frequency mapping concerning the parameters must be at least Lipschitz continuous. This ensures that the parameter domain remains intact. However, when the regularity of the frequency mapping is less stringent than Lipschitz continuous, the traditional method of parameter excavation becomes infeasible. This necessitates the exploration of novel approaches to address the issue.

1.3 Our work

Regarding regularity, when the frequency mapping is continuous with respect to parameters, we prove that the perturbed invariant tori retain the same Diophantine frequency as their unperturbed counterparts for Hamiltonian systems as described in (1.2), see Theorem 1. For the degeneracy problem, persistence results under the highly degenerate Hamiltonian system (1.1) are proved in Theorem 2.

We establish sufficient conditions based on the topological degree condition (A0)\rm(A0) and the weak convexity condition (A1)\rm(A1) for frequency mapping. Detailed descriptions of these conditions are provided in Section 2. In deriving our primary results, we employ the quasi-linear KAM iteration procedure as in [10, 17, 23, 31]. Notably, we introduce a parameterized family Hamiltonian systems to counteract frequency drift. Specifically, we adjust the action variable to maintain constant frequency for the highly degenerate Hamiltonian system (1.1). It’s also noteworthy that the weak convexity condition proposed in this paper is necessary regardless of the smoothness level of the frequency mapping, as evidenced by Proposition 1.

It should be pointed out that the KAM-type theorems associated with parameter family are due to Moser [25], Pöschel[28]. However, our results are different from theirs: a Diophantine frequency can be given in advance, but Moser’s systems need to be modified in KAM iteration and hence cannot be given beforehand; in Pöschel’s approach, the frequency set need to be dug out in KAM process. Our method is to find a parameter in the family of systems by translating parameter. Of course, it does not work generally. As pointed out in our paper, the weak convexity condition (A1) is indispensable. To our knowledge, this setting seems to be first.

The rest of this paper is organized as follows. In Section 2, we state our main results (Theorems 1, 2, 3 and 4). We will describe the quasi-linear iterative scheme, show the detailed construction and estimates for one cycle of KAM steps in Section 3. In Section 4, we complete the proof of Theorem 1 by deriving an iteration lemma and showing the convergence of KAM iterations. In Section 5, we prove Theorem 2 , which covers the analytic situation, and is also a special case of Theorem 1. We also prove Theorem 4 by directly computing. Finally, the proof of Theorem 3 can be found in Appendix B.

2 Main results

To state our main results we need first to introduce a few definitions and notations.

(1)

Given a domain DG×𝕋nD\subset G\times\mathbb{T}^{n}, we let D¯\bar{D}, D\partial D denote the closure of DD and the boundary of DD, respectively. Do:=D¯DD^{o}:=\bar{D}\setminus\partial D refers to the interior.

(2)

We shall use the same symbol |||\cdot| to denote an equivalent vector norm and its induced matrix norm, absolute value of functions, etc, and use ||D|\cdot|_{D} to denote the supremum norm of functions on a domain DD.

(3)

For the perturbation function P(y,x,ξ)P(y,x,\xi), which is analytic about yy and xx and Hölder continuous about ξ\xi with Hölder index β\beta, 0<β<10<\beta<1, we define its norm as follows

|P|D=PD+PCβ|\|P\||_{D}=\|P\|_{D}+\|P\|_{C^{\beta}}

where

PCβ=supξζ,ξ,ζO|P(y,x,ξ)P(y,x,ζ)||ξζ|β,(y,x)D.\|P\|_{C^{\beta}}=\sup_{\xi\neq\zeta,~{}\xi,\zeta\in{O}}\frac{|P(y,x,\xi)-P(y,x,\zeta)|}{|\xi-\zeta|^{\beta}},~{}~{}~{}\forall(y,x)\in{D}. (2.3)
(4)

For any two complex column vectors ξ\xi, η\eta in the same space, ξ,η\langle\xi,\eta\rangle always stands for ξη\xi^{\top}\eta.

(5)

idi_{d} is the unit map, and IdI_{d} is the unit matrix.

(6)

For a vector value function ff, DfDf denotes the Jacobian matrix of ff, and Jf=detDfJ_{f}=detDf its Jacobian determinant.

(7)

All Hamiltonian in the sequel are endowed with the standard symplectic structure.

(8)

As pointed out in [30], the real analyticity of the Hamiltonian H(y,x)H(y,x) about yy and xx on G×𝕋nG\times{\mathbb{T}^{n}} implies that the analyticity extends to a complex neighbourhood D(s,r)D(s,r) of G×𝕋nG\times{\mathbb{T}^{n}}, where D(s,r)D(s,r) is defined for some 0<s,r<10<s,r<1, with

D(s,r):={(y,x):dist(y,G)<s,|Imx|<r}.D(s,r):=\{(y,x):dist(y,G)<s,|\textrm{Im}x|<r\}.
(9)

For δ>0\forall\delta>0, y0Gy_{0}\in G, let

Bδ(y0)\displaystyle B_{\delta}(y_{0}) =:{yG:|yy0|<δ},\displaystyle=:\{y\in G:|y-y_{0}|<\delta\},
B¯δ(y0)\displaystyle\bar{B}_{\delta}(y_{0}) =:{yG:|yy0|δ}.\displaystyle=:\{y\in G:|y-y_{0}|\leq\delta\}.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded and open domain. We first give the definition of the degree for fC2(Ω¯,n)f\in C^{2}(\bar{\Omega},\mathbb{R}^{n}), see [27].

Definition 2.1.

If fC2(Ω¯,n)f\in C^{2}(\bar{\Omega},\mathbb{R}^{n}), and pnf(Ω)p\in\mathbb{R}^{n}\setminus f(\partial\Omega), let ς=infxΩf(x)p\varsigma=\inf_{x\in\partial\Omega}\left\|f(x)-p\right\|.

(1)

Denote Nf={x|xΩ,Jf(x)=0}N_{f}=\left\{x|x\in\Omega,J_{f}(x)=0\right\}. If pf(Nf)p\notin f(N_{f}), then

deg(f,Ω,p):=xf1(p)sign(Jf(x)),\displaystyle\deg(f,\Omega,p):=\sum_{x\in{f^{-1}(p)}}sign\left(J_{f}(x)\right),

setting deg(f,Ω,p)=0\deg\left(f,\Omega,p\right)=0 if f1(p)=Øf^{-1}\left(p\right)={\O}.

(2)

If pf(Nf)p\in f(N_{f}), by Sard theorem (see [27]), deg(f,Ω,p)\deg\left(f,\Omega,p\right) is defined as deg(f,Ω,p):=deg(f,Ω,p1)\deg(f,\Omega,p):=\deg(f,\Omega,p_{1}), for some (and all) p1f(Nf)p_{1}\notin f(N_{f}) such that p1p<ς7\left\|p_{1}-p\right\|<\frac{\varsigma}{7}.

We next extend the definition of degree to the general continuous mapping, see [27].

Definition 2.2.

(Brouwer’s Degree) Let gC(Ω¯,n)g\in C(\bar{\Omega},\mathbb{R}^{n}), and png(Ω)p\in\mathbb{R}^{n}\setminus g(\partial\Omega), i.e., ς=:infxΩg(x)p>0\varsigma_{*}=:\inf_{x\in\partial\Omega}\left\|g(x)-p\right\|>0. Let

S=:{f|fC2(Ω¯,n),maxxΩ¯{g(x)f(x)<ς}}.S=:\left\{f\Big{|}f\in C^{2}\left(\bar{\Omega},\mathbb{R}^{n}\right),\max_{x\in\bar{\Omega}}\left\{\left\|g(x)-f(x)\right\|<\varsigma_{*}\right\}\right\}.

Define the degree of gg as :

deg(g,Ω,p):=deg(f,Ω,p),fS.\deg{(g,\Omega,p)}:=\deg{(f,\Omega,p)},~{}~{}~{}\forall f\in S.

We are now ready to state our assumptions. Mainly we consider (1.2), i.e., for any ε>0\varepsilon>0 small enough, we consider the parameterized family of perturbed Hamiltonian equations

{H:G×𝕋n×O1,H(y,x,ξ)=ω(ξ),y+εP(y,x,ξ,ε).\left\{\begin{array}[]{ll}H:G\times\mathbb{T}^{n}\times O\rightarrow\mathbb{R}^{1},\\ H(y,x,\xi)=\langle\omega(\xi),y\rangle+\varepsilon P(y,x,\xi,\varepsilon).\end{array}\right.

First, we make the following assumptions:

(A0)

Fix ξ0Oo\xi_{0}\in O^{o} such that

deg(ω(),Oo,ω(ξ0))0.\displaystyle\deg\left(\omega(\cdot),O^{o},\omega(\xi_{0})\right)\neq 0. (2.4)
(A1)

There are σ>0\sigma>0, 0<Lβ0<L\leq\beta, (β\beta was defined in (2.3)), such that

|ω(ξ)ω(ξ)|σ|ξξ|L,ξ,ξO.\displaystyle\left|\omega(\xi)-\omega(\xi_{*})\right|\geq\sigma\left|\xi-\xi_{*}\right|^{L},~{}~{}~{}~{}\forall\xi,\xi_{*}\in O. (2.5)
(A2)

For the given ξ0Oo\xi_{0}\in O^{o}, ω(ξ0)\omega(\xi_{0}) satisfies the Diophantine condition

|k,ω(ξ0)|>γ|k|τ,kn{0},\displaystyle\left|\langle k,\omega(\xi_{0})\rangle\right|>\frac{\gamma}{|k|^{\tau}},~{}~{}~{}k\in{\mathbb{Z}^{n}\setminus{\{0\}}}, (2.6)

where k=(k1,,kn)k=(k_{1},\cdots,k_{n}), |k|=|k1|++|kn||k|=\left|k_{1}\right|+\cdots+\left|k_{n}\right|, γ>0\gamma>0 and τ>n1\tau>n-1.

Then, we have the following main results:

Theorem 1.

Consider Hamiltonian system (1.2). Assume that (A0){\rm(A0)} , (A1){\rm(A1)} and (A2){\rm(A2)} hold. Then there exists a sufficiently small ε0>0\varepsilon_{0}>0, for any 0<ε<ε00<\varepsilon<\varepsilon_{0}, there exist ξεO\xi_{\varepsilon}\in O and a symplectic transformation Ψ\Psi_{*} such that

H(Ψ(y,x,ξε),ε)=e+ω(ξ0),y+h¯(y,ξε)+P(y,x,ξε,ε)H(\Psi_{*}(y,x,\xi_{\varepsilon}),\varepsilon)=e_{*}+\langle\omega(\xi_{0}),y\rangle+\bar{h}_{*}(y,\xi_{\varepsilon})+P_{*}(y,x,\xi_{\varepsilon},\varepsilon)

where ee_{*} is a constant, h¯(y,ξε)=O(|y|2)\bar{h}_{*}(y,\xi_{\varepsilon})=O(|y|^{2}), P=O(|y|2)P_{*}=O(|y|^{2}). Thus the perturbed Hamiltonian system H(y,x,ξε,ε)H(y,x,\xi_{\varepsilon},\varepsilon) admits an invariant torus with frequency ω(ξ0)\omega(\xi_{0}).

Remark 2.1.

It should be emphasized that we deal with the degenerate Hamiltonian system in which the frequency mapping is continuous about parameters and the perturbation is Hölder continuous about parameters in this theorem. It seems to be the first version in KAM theory.

In the following, we will give some examples to state that conditions (A0){\rm(A0)} and (A1){\rm(A1)} are indispensable, especially for condition (A1){\rm(A1)}. See below for a counter example:

Proposition 1.

Consider the Hamiltonian system (1.2), for n=2n=2, with

ω(ξ)=(ω1(ξ1),ω2(ξ2)),εP=P0(ε)y2,\omega(\xi)=(\omega_{1}(\xi_{1}),\omega_{2}(\xi_{2}))^{\top},~{}~{}~{}\varepsilon P=P_{0}(\varepsilon)y_{2},

where

ω1(ξ1)=ω¯1+ξ1,ξ1(1,1),\displaystyle\omega_{1}(\xi_{1})=\bar{\omega}_{1}+\xi_{1},~{}~{}\xi_{1}\in(-1,1),
ω2(ξ2)={ω¯2+exp{1(ξ2+12)2},ξ2(1,12),ω¯2,ξ2[12,12],ω¯2exp{1(ξ212)2},ξ2(12,1),\displaystyle\omega_{2}{(\xi_{2})}=\left\{\begin{array}[]{lll}\bar{\omega}_{2}+\exp\{-\frac{1}{(\xi_{2}+{\frac{1}{2}})^{2}}\},{}{}{}&\xi_{2}\in(-1,-\frac{1}{2}),\\ \bar{\omega}_{2},{}{}{}&\xi_{2}\in[-\frac{1}{2},\frac{1}{2}],\\ \bar{\omega}_{2}-\exp\{-\frac{1}{(\xi_{2}-{\frac{1}{2}})^{2}}\},{}{}{}&\xi_{2}\in(\frac{1}{2},1),\end{array}\right.

ω¯=(ω¯1,ω¯2)\bar{\omega}=(\bar{\omega}_{1},\bar{\omega}_{2})^{\top} satisfies Diophantine condition (2.6), and

P0(ε)={0,ε=0,εsin1ε,ε0,+{0}.P_{0}(\varepsilon)=\left\{\begin{array}[]{lll}0,&&\varepsilon=0,\\ \varepsilon^{\ell}\sin\frac{1}{\varepsilon},&&\varepsilon\neq 0,\,\ell\in\mathbb{Z}^{+}\setminus\{0\}.\end{array}\right.

Then condition (A1)\rm(A1) fails for any parameter ξ(1,1)\xi\in(-1,1). Moreover, Theorem 1 fails.

See Appendix A for the complete proof.

Remark 2.2.

This counter example implies that (A1)\rm(A1) is necessary no matter how smooth the frequency mapping ω(ξ)\omega(\xi) is.

Nevertheless, one asks what happens to the frequency mapping in the analytic situation. As a special case of our Theorem 1, we also obtain the Kolmogorov’s theorem for analytic Hamiltonian systems under degenerate conditions. This is stated in the following theorem.

Theorem 2.

Consider the real analytic Hamiltonian system (1.1). Fix ξ0G\xi_{0}\in G such that (A0)\rm(A0), (A1)\rm(A1) and (A2)\rm(A2) hold for ω(ξ)=h(ξ)\omega(\xi)=\nabla h(\xi), O=GO=G, and L>0L>0. Then there exist a sufficiently small positive constant ε>0\varepsilon^{\prime}>0 such that if 0<ε<ε0<\varepsilon<\varepsilon^{\prime}, there exists yεGy_{\varepsilon}\in G such that Hamiltonian system (1.1) at y=yεy=y_{\varepsilon} admits an invariant torus with frequency h(ξ0)\nabla h(\xi_{0}).

This theorem is proved in Section 5.1.

Next, we will give an example that satisfies conditions (A0)\rm(A0)-(A1)\rm(A1). For simplicity we use the action variable yy as the parameter ξ\xi.

Theorem 3.

Consider the Hamiltonian system (1.1) with

h(y)=ω,y+12l+2|y|2l+2,h(y)=\langle\omega,y\rangle+\frac{1}{2l+2}|y|^{2l+2},

where yGny\in G\subset\mathbb{R}^{n}, ll is a positive integer, ωn{0}\omega\in\mathbb{R}^{n}\setminus{\{0\}} satisfies the Diophantine condition (2.6). Then the Hamiltonian system (1.1) admits an invariant torus with frequency ω\omega for any small enough perturbation.

The proof can be found in Appendix B.

Proposition 2.

If h(y)=ω,y+12l+1|y|2l+1h(y)=\langle\omega,y\rangle+\frac{1}{2l+1}\left|y\right|^{2l+1} in Hamiltonian system (1.1), ωn{0}\omega\in\mathbb{R}^{n}\setminus{\{0\}} satisfies the Diophantine condition (2.6), then the system may not admit torus with frequency ω\omega.

The proof can be found in Appendix C.

Above results imply that condition (A0){\rm(A0)} is indispensable for n>1n>1 case. Furthermore, we also prove that for n=1n=1, the persistence results in Theorem 3 hold under some weaker conditions, provided that the frequency satisfies Diophantine condition (A2){\rm(A2)}.

Theorem 4.

Consider Hamiltonian (1.1) with

h(y)=ωy+g(y),εP(y,x,ε)=εP(y),\displaystyle h(y)=\omega y+g(y),~{}~{}~{}~{}\varepsilon P(y,x,\varepsilon)=\varepsilon P(y),

where yG=[1,1]1y\in G=[-1,1]\subset\mathbb{R}^{1}, ω\omega satisfies Diophantine condition (2.6).

(1)

If g(y)C2+1g(y)\in C^{2\ell+1}, g(0)==g2(0)=0g^{\prime}(0)=\cdots=g^{2\ell}(0)=0, g2+1(0)0g^{2\ell+1}(0)\neq 0, \ell is a positive integer, then the perturbed system admits at least two invariant tori with frequency ω\omega for the small enough perturbation satisfying εP(y)sign(g2+1(0))<0\varepsilon P^{\prime}(y)\,sign(g^{2\ell+1}(0))<0; conversely, if εP(y)sign(g2+1(0))>0\varepsilon P^{\prime}(y)\,sign(g^{2\ell+1}(0))>0, the unperturbed invariant torus with frequency ω\omega will be destroyed.

(2)

If g(y)C2+2g(y)\in C^{2\ell+2}, g(0)==g2+1(0)=0g^{\prime}(0)=\cdots=g^{2\ell+1}(0)=0, g2+2(0)0g^{2\ell+2}(0)\neq 0, \ell is a positive integer, then the perturbed system admits an invariant tori with frequency ω\omega for any small enough perturbation.

Remark 2.3.

We don’t know whether the results in Theorem 4 can be extended to higher dimensions or not.

3 KAM step

In this section, we will describe the quasi-linear iterative scheme, show the detailed construction and estimates for one cycle of KAM steps, which is essential to study the KAM theory, see [10, 17, 22, 23, 30]. It should be pointed out that in our KAM iteration, we present a new way to move parameters; while in the usual KAM iteration, one has to dig out a decreasing series of parameter domains, see [10, 17, 23, 29, 30, 31, 32].

3.1 Description of the 0-th KAM step.

Given an integer m>L+1m>L+1, where LL was defined as in (A1)\rm(A1). Denote ρ=12(m+1)\rho=\frac{1}{2(m+1)}, and let η>0\eta>0 be an integer such that (1+ρ)η>2(1+\rho)^{\eta}>2. We define

γ=ε14(n+m+2).\gamma=\varepsilon^{\frac{1}{4(n+m+2)}}. (3.7)

Consider the perturbed Hamiltonian (1.2). We first define the following 0-th KAM step parameters:

r0=r,γ0=γ,e0=0,h¯0=0,μ0=ε18η(τ+1)(m+1),\displaystyle r_{0}=r,~{}~{}~{}~{}~{}\gamma_{0}=\gamma,~{}~{}~{}~{}~{}e_{0}=0,~{}~{}~{}~{}~{}\bar{h}_{0}=0,~{}~{}~{}~{}~{}\mu_{0}=\varepsilon^{\frac{1}{8\eta(\tau+1)(m+1)}}, (3.8)
s0=sγ016(M+2)K1τ+1,O0={ξO||ξξ0|<dist(ξ0,O)},\displaystyle s_{0}=\frac{s\gamma_{0}}{16(M^{*}+2)K_{1}^{\tau+1}},~{}O_{0}=\{\xi\in O|~{}|\xi-\xi_{0}|<dist(\xi_{0},\partial O)\},
D(s0,r0):={(y,x):dist(y,G)<s0,|Imx|<r0},\displaystyle D(s_{0},r_{0}):=\{(y,x):dist(y,G)<s_{0},|\textrm{Im}x|<r_{0}\},

where 0<s0,γ0,μ010<s_{0},\gamma_{0},\mu_{0}\leq 1, τ>n1\tau>n-1, M>0M^{*}>0 is a constant defined as in Lemma 3.3, and

K1=([log1μ0]+1)3η.\displaystyle K_{1}=([\log\frac{1}{\mu_{0}}]+1)^{3\eta}.

Therefore, we can write

H0\displaystyle H_{0} =:H(y,x,ξ0)=N0+P0,\displaystyle=:H(y,x,\xi_{0})=N_{0}+P_{0},
N0\displaystyle N_{0} =:N0(y,ξ0,ε)=e0+ω(ξ0),y+h¯0,\displaystyle=:N_{0}(y,\xi_{0},\varepsilon)=e_{0}+\langle\omega(\xi_{0}),y\rangle+\bar{h}_{0},
P0\displaystyle P_{0} =:εP(y,x,ξ0,ε).\displaystyle=:\varepsilon P(y,x,\xi_{0},\varepsilon).

We first prove an important estimate.

Lemma 3.1.
|P0|D(s0,r0)γ0n+m+2s0mμ0.|\|P_{0}\||_{D(s_{0},r_{0})}\leq\gamma_{0}^{n+m+2}s_{0}^{m}\mu_{0}. (3.9)
Proof.

Using the fact γ0n+m+2=ε14\gamma_{0}^{n+m+2}=\varepsilon^{\frac{1}{4}} and [log1μ0]+1<1μ0[\log\frac{1}{\mu_{0}}]+1<\frac{1}{\mu_{0}}, we have

s0m=smεm4(n+m+2)16m(M+2)mK1m(τ+1)>smεm4(n+m+2)μ03ηm(τ+1)16m(M+2)msmεm4(n+m+2)+3816m(M+2)m.\displaystyle s_{0}^{m}=\frac{s^{m}\varepsilon^{\frac{m}{4(n+m+2)}}}{16^{m}(M^{*}+2)^{m}K_{1}^{m(\tau+1)}}>\frac{s^{m}\varepsilon^{\frac{m}{4(n+m+2)}}\mu_{0}^{3\eta m(\tau+1)}}{16^{m}(M^{*}+2)^{m}}\geq\frac{s^{m}\varepsilon^{\frac{m}{4(n+m+2)}+\frac{3}{8}}}{16^{m}(M^{*}+2)^{m}}.

Moreover, let ε0>0\varepsilon_{0}>0 be small enough so that

ε01818η(τ+1)(m+1)|P|D(s0,r0)16m(M+2)msm1,\varepsilon_{0}^{\frac{1}{8}-\frac{1}{8\eta(\tau+1)(m+1)}}|\|P\||_{D(s_{0},r_{0})}\frac{16^{m}(M^{*}+2)^{m}}{s^{m}}\leq 1, (3.10)

using the fact that μ0=ε18η(τ+1)(m+1)\mu_{0}=\varepsilon^{\frac{1}{8\eta(\tau+1)(m+1)}}, we get

γ0n+m+2s0mμ0\displaystyle\gamma_{0}^{n+m+2}s_{0}^{m}\mu_{0} smεm4(n+m+2)+38+14+18η(τ+1)(m+1)16m(M+2)msmε14+38+14+18η(τ+1)(m+1)16m(M+2)m\displaystyle\geq\frac{s^{m}\varepsilon^{\frac{m}{4(n+m+2)}+\frac{3}{8}+\frac{1}{4}+\frac{1}{8\eta(\tau+1)(m+1)}}}{16^{m}(M^{*}+2)^{m}}\geq\frac{s^{m}\varepsilon^{\frac{1}{4}+\frac{3}{8}+\frac{1}{4}+\frac{1}{8\eta(\tau+1)(m+1)}}}{16^{m}(M^{*}+2)^{m}}
=ε78smε18η(τ+1)(m+1)16m(M+2)m,\displaystyle=\varepsilon^{\frac{7}{8}}\frac{s^{m}\varepsilon^{\frac{1}{8\eta(\tau+1)(m+1)}}}{16^{m}(M^{*}+2)^{m}}, (3.11)

and by (3.10) and 0<ε<ε00<\varepsilon<\varepsilon_{0},

ε1818η(τ+1)(m+1)|P|D(s0,r0)16m(M+2)msm1,\displaystyle\varepsilon^{\frac{1}{8}-\frac{1}{8\eta(\tau+1)(m+1)}}|\|P\||_{D(s_{0},r_{0})}\frac{16^{m}(M^{*}+2)^{m}}{s^{m}}\leq 1,

i.e.,

ε18|P|D(s0,r0)smε18η(τ+1)(m+1)16m(M+2)m.\displaystyle\varepsilon^{\frac{1}{8}}|\|P\||_{D(s_{0},r_{0})}\leq\frac{s^{m}\varepsilon^{\frac{1}{8\eta(\tau+1)(m+1)}}}{16^{m}(M^{*}+2)^{m}}. (3.12)

Then by (3.1) and (3.12),

|P0|D(s0,r0)=ε78ε18|P|D(s0,r0)ε78smε18η(τ+1)(m+1)16m(M+2)mγ0n+m+2s0mμ0,\displaystyle|\|P_{0}\||_{D(s_{0},r_{0})}=\varepsilon^{\frac{7}{8}}\varepsilon^{\frac{1}{8}}|\|P\||_{D(s_{0},r_{0})}\leq\varepsilon^{\frac{7}{8}}\frac{s^{m}\varepsilon^{\frac{1}{8\eta(\tau+1)(m+1)}}}{16^{m}(M^{*}+2)^{m}}\leq\gamma_{0}^{n+m+2}s_{0}^{m}\mu_{0},

which implies (3.9).

The proof is complete. ∎

3.2 Induction from ν\nu-th KAM step

3.2.1 Description of the ν\nu-th KAM step

We now define the ν\nu-th KAM step parameters:

rν=rν12+r04,sν=18μν12ρsν1,μν=8mμν11+ρ,r_{\nu}=\frac{r_{\nu-1}}{2}+\frac{r_{0}}{4},~{}~{}~{}s_{\nu}=\frac{1}{8}\mu_{\nu-1}^{2\rho}s_{\nu-1},~{}~{}~{}\mu_{\nu}=8^{m}\mu_{\nu-1}^{1+\rho},

where ρ=12(m+1)\rho=\frac{1}{2(m+1)}.

Now, suppose that at ν\nu-th step, we have arrived at the following real analytic Hamiltonian:

Hν\displaystyle H_{\nu} =Nν+Pν,\displaystyle=N_{\nu}+P_{\nu},~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{} (3.13)
Nν\displaystyle N_{\nu} =eν+ω(ξ0),y+h¯ν(y,ξ),\displaystyle=e_{\nu}+\langle\omega(\xi_{0}),y\rangle+\bar{h}_{\nu}(y,\xi),

defined on D(sν,rν)D(s_{\nu},r_{\nu}) and

|Pν|D(sν,rν)γ0n+m+2sνmμν.\left|\left\|P_{\nu}\right\|\right|_{D(s_{\nu},r_{\nu})}\leq\gamma_{0}^{n+m+2}s_{\nu}^{m}\mu_{\nu}. (3.14)

The equation of motion associated to HνH_{\nu} is

{y˙ν=xνHν,x˙ν=yνHν.\left\{\begin{array}[]{ll}\dot{y}_{\nu}=-\partial_{x_{\nu}}H_{\nu},\\ \dot{x}_{\nu}=~{}~{}\partial_{y_{\nu}}H_{\nu}.\end{array}\right. (3.15)

Except for additional instructions, we will omit the index for all quantities of the present KAM step (at ν\nu-th step) and use ++ to index all quantities (Hamiltonian, domains, normal form, perturbation, transformation, etc.) in the next KAM step (at (ν+1)(\nu+1)-th step). To simplify the notations, we will not specify the dependence of PP, P+P_{+} etc. All the constants c1c_{1}-c6c_{6} below are positive and independent of the iteration process, and we will also use cc to denote any intermediate positive constant which is independent of the iteration process.

Define

r+\displaystyle r_{+} =r2+r04,\displaystyle=\frac{r}{2}+\frac{r_{0}}{4},
s+\displaystyle s_{+} =18αs,α=μ2ρ=μ1m+1,\displaystyle=\frac{1}{8}\alpha s,~{}~{}~{}~{}~{}\alpha=\mu^{2\rho}=\mu^{\frac{1}{m+1}},
μ+\displaystyle\mu_{+} =8mc0μ1+ρ,c0=max{1,c1,c2,,c6},\displaystyle=8^{m}c_{0}\mu^{1+\rho},~{}~{}~{}~{}~{}c_{0}=\max\{1,c_{1},c_{2},\cdots,c_{6}\},
K+\displaystyle K_{+} =([log1μ]+1)3η,\displaystyle=([\log\frac{1}{\mu}]+1)^{3\eta},
D^\displaystyle\hat{D} =D(s,r++78(rr+)),\displaystyle=D(s,r_{+}+\frac{7}{8}(r-r_{+})),
D~\displaystyle\tilde{D} =D(12s,r++68(rr+)),\displaystyle=D(\frac{1}{2}s,r_{+}+\frac{6}{8}(r-r_{+})),
D(s)\displaystyle D(s) ={yCn:|y|<s},\displaystyle=\{y\in C^{n}:|y|<s\},
Di8α\displaystyle D_{\frac{i}{8}\alpha} =D(i8αs,r++i18(rr+)),i=1,2,,8,\displaystyle=D(\frac{i}{8}\alpha s,r_{+}+\frac{i-1}{8}(r-r_{+})),~{}~{}i=1,2,\cdots,8,
D+\displaystyle D_{+} =D18α=D(s+,r+),\displaystyle=D_{\frac{1}{8}\alpha}=D(s_{+},r_{+}),
O+\displaystyle O_{+} ={ξ:dist(ξ,O)μ1L},\displaystyle=\{\xi:dist(\xi,O)\leq\mu^{\frac{1}{L}}\},
Γ(rr+)\displaystyle\Gamma(r-r_{+}) =0<|k|K+|k|3τ+5e|k|rr+8.\displaystyle=\sum_{0<|k|\leq K_{+}}\left|k\right|^{3\tau+5}e^{-|k|\frac{r-r_{+}}{8}}.

3.2.2 Construct a symplectic transformation

We will construct a symplectic coordinate transformation Φ+\Phi_{+}:

Φ+:(y+,x+)D(s+,r+)Φ+(y+,x+)=(y,x)D(s,r)\displaystyle\Phi_{+}:(y_{+},x_{+})\in D(s_{+},r_{+})\rightarrow\Phi_{+}(y_{+},x_{+})=(y,x)\in D(s,r) (3.16)

such that it transforms the Hamiltonian (3.13) into the Hamiltonian of the next KAM cycle (at (ν+1)(\nu+1)-th step), i.e.,

H+=HΦ+=N++P+,H_{+}=H\circ\Phi_{+}=N_{+}+P_{+}, (3.17)

where N+N_{+} and P+P_{+} have similar properties as NN and PP respectively on D(s+,r+)D(s_{+},r_{+}), and the equation of motion (3.15) is changed into

{y˙+=x+H+,x˙+=y+H+.\left\{\begin{array}[]{ll}\dot{y}_{+}=-\partial_{x_{+}}H_{+},\\ \dot{x}_{+}=~{}~{}\partial_{y_{+}}H_{+}.\end{array}\right. (3.18)

In the following, we prove (3.18). Let Φ+(y+,x+):=(Φ+1(y+,x+),Φ+2(y+,x+))\Phi_{+}(y_{+},x_{+}):=\left(\Phi_{+}^{1}(y_{+},x_{+}),\Phi_{+}^{2}(y_{+},x_{+})\right), by (3.16), we have

(y˙x˙)=((y+Φ+1)y˙+(x+Φ+1)x˙+(y+Φ+2)y˙+(x+Φ+2)x˙+)=DΦ+(y˙+x˙+),\displaystyle\left(\begin{array}[]{c}\dot{y}\\ \dot{x}\end{array}\right)=\left(\begin{array}[]{cc}(\partial_{y_{+}}\Phi_{+}^{1})\dot{y}_{+}&(\partial_{x_{+}}\Phi_{+}^{1})\dot{x}_{+}\\ (\partial_{y_{+}}\Phi_{+}^{2})\dot{y}_{+}&(\partial_{x_{+}}\Phi_{+}^{2})\dot{x}_{+}\end{array}\right)=D\Phi_{+}\left(\begin{array}[]{c}\dot{y}_{+}\\ \dot{x}_{+}\end{array}\right),

by (3.16) and (3.17), we get

(y+Hx+H)\displaystyle\left(\begin{array}[]{c}\partial_{y_{+}}H\\ \partial_{x_{+}}H\end{array}\right) =(yHy+yxHy+xyHx+yxHx+x)=(y+Φ+1y+Φ+2x+Φ+1x+Φ+2)(yHxH)\displaystyle=\left(\begin{array}[]{cc}\partial_{y}H\partial_{y_{+}}y&\partial_{x}H\partial_{y_{+}}x\\ \partial_{y}H\partial_{x_{+}}y&\partial_{x}H\partial_{x_{+}}x\end{array}\right)=\left(\begin{array}[]{cc}\partial_{y_{+}}\Phi_{+}^{1}&\partial_{y_{+}}\Phi_{+}^{2}\\ \partial_{x_{+}}\Phi_{+}^{1}&\partial_{x_{+}}\Phi_{+}^{2}\end{array}\right)\left(\begin{array}[]{cc}\partial_{y}H\\ \partial_{x}H\end{array}\right)
=DΦ+(yHxH).\displaystyle=D\Phi_{+}^{\top}\left(\begin{array}[]{cc}\partial_{y}H\\ \partial_{x}H\end{array}\right).

Then this together with (3.15) yields

(y˙+x˙+)\displaystyle\left(\begin{array}[]{c}\dot{y}_{+}\\ \dot{x}_{+}\end{array}\right) =DΦ+1(y˙x˙)=DΦ+1J(yHxH)=DΦ+1J(DΦ+1)(y+Hx+H)\displaystyle=D\Phi_{+}^{-1}\left(\begin{array}[]{c}\dot{y}\\ \dot{x}\end{array}\right)=D\Phi_{+}^{-1}J\left(\begin{array}[]{cc}\partial_{y}H\\ \partial_{x}H\end{array}\right)=D\Phi_{+}^{-1}J(D\Phi_{+}^{-1})^{\top}\left(\begin{array}[]{c}\partial_{y_{+}}H\\ \partial_{x_{+}}H\end{array}\right)
=J(y+Hx+H),\displaystyle=J\left(\begin{array}[]{c}\partial_{y_{+}}H\\ \partial_{x_{+}}H\end{array}\right),

where JJ is the standard symplectic matrix, i.e.,

J=(0IdId0).\displaystyle J=\left(\begin{array}[]{cc}0&-I_{d}\\ I_{d}&0\end{array}\right).

This finishes the proof of (3.18).

Next, we show the detailed construction of Φ+\Phi_{+} and the estimates of P+P_{+}.

3.2.3 Truncation

Consider the Taylor-Fourier series of PP:

P=kZn,ıZ+npkıyıe1k,x,P=\sum_{k\in Z^{n},~{}\imath\in Z_{+}^{n}}p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle},

and let RR be the truncation of PP of the form

R=|k|K+,|ı|mpkıyıe1k,x.R=\sum_{|k|\leq K_{+},~{}|\imath|\leq m}p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle}.

Next, we will prove that the norm of PRP-R is much smaller than the norm of PP by selecting truncation appropriately, see the below lemma.

Lemma 3.2.

Assume that

(H1):K+tnetrr+16𝑑tμ.\displaystyle\textbf{{(H1)}}:\int_{K_{+}}^{\infty}t^{n}e^{-t\frac{r-r_{+}}{16}}dt\leq\mu.

Then there is a constant c1c_{1} such that

|PR|Dα\displaystyle\left|\left\|P-R\right\|\right|_{D_{\alpha}} c1γ0n+m+2smμ2,\displaystyle\leq c_{1}\gamma_{0}^{n+m+2}s^{m}\mu^{2}, (3.19)
|R|Dα\displaystyle\left|\left\|R\right\|\right|_{D_{\alpha}} c1γ0n+m+2smμ.\displaystyle\leq c_{1}\gamma_{0}^{n+m+2}s^{m}\mu. (3.20)
Proof.

Denote

I\displaystyle I =|k|>K+,ıZ+npkıyıe1k,x,\displaystyle=\sum_{|k|>K_{+},~{}\imath\in Z_{+}^{n}}p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle},
II\displaystyle II =|k|K+,|ı|>mpkıyıe1k,x.\displaystyle=\sum_{|k|\leq K_{+},~{}|\imath|>m}p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle}.

Then

PR=I+II.\displaystyle P-R=I+II.

To estimate II, we note by (3.14) that

|ıZ+npkıyı||P|D(s,r)e|k|rγ0n+m+2smμe|k|r,\displaystyle\left|\sum_{\imath\in Z_{+}^{n}}p_{k\imath}y^{\imath}\right|\leq\left|P\right|_{D(s,r)}e^{-|k|r}\leq\gamma_{0}^{n+m+2}s^{m}\mu e^{-|k|r}, (3.21)

where the first inequality has been frequently used in [10, 11, 17, 22, 30, 31, 32, 35] and the detailed proof see [35]. This together with (H1) yields

|I|D^\displaystyle\left|I\right|_{\hat{D}} |k|>K+|ıZ+npkıyı|e|k|(r+8+7r8)|k|>K+|P|D(s,r)e|k|rr+8\displaystyle\leq\sum_{|k|>K_{+}}\left|\sum_{\imath\in Z_{+}^{n}}p_{k\imath}y^{\imath}\right|e^{|k|(\frac{r_{+}}{8}+\frac{7r}{8})}\leq\sum_{|k|>K_{+}}\left|P\right|_{D(s,r)}e^{-|k|\frac{r-r_{+}}{8}}
γ0n+m+2smμκ=K+κneκrr+8γ0n+m+2smμK+tnetrr+16𝑑t\displaystyle\leq\gamma_{0}^{n+m+2}s^{m}\mu\sum_{\kappa=K_{+}}^{\infty}\kappa^{n}e^{-\kappa\frac{r-r_{+}}{8}}\leq\gamma_{0}^{n+m+2}s^{m}\mu\int_{K_{+}}^{\infty}t^{n}e^{-t\frac{r-r_{+}}{16}}dt (3.22)
γ0n+m+2smμ2.\displaystyle\leq\gamma_{0}^{n+m+2}s^{m}\mu^{2}.

It follows from (3.14) and (3.22) that

|PI|D^|P|D(s,r)+|I|D^2γ0n+m+2smμ.\displaystyle\left|P-I\right|_{\hat{D}}\leq|P|_{D(s,r)}+|I|_{\hat{D}}\leq 2\gamma_{0}^{n+m+2}s^{m}\mu.

For |p|=m+1|p|=m+1, let \int be the obvious antiderivative of pyp\frac{\partial^{p}}{\partial y^{p}}. Then the Cauchy estimate of PIP-I on DαD_{\alpha} yields

|II|Dα\displaystyle\left|II\right|_{D_{\alpha}} =|pyp|k|K+,|ı|>mpkıyıe1k,xdy|Dα\displaystyle=\left|\int\frac{\partial^{p}}{\partial y^{p}}\sum_{|k|\leq K_{+},~{}|\imath|>m}p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle}dy\right|_{D_{\alpha}}
||pyp(PI)|dy|Dα\displaystyle\leq\left|\int\left|\frac{\partial^{p}}{\partial y^{p}}(P-I)\right|dy\right|_{D_{\alpha}}
|csm+1|PI|D^dy|Dα\displaystyle\leq\left|\frac{c}{s^{m+1}}\int\left|P-I\right|_{\hat{D}}dy\right|_{D_{\alpha}}
2csm+1γ0n+m+2smμ(αs)m+1\displaystyle\leq 2\frac{c}{s^{m+1}}\gamma_{0}^{n+m+2}s^{m}\mu(\alpha s)^{m+1}
cγ0n+m+2smμ2.\displaystyle\leq c\gamma_{0}^{n+m+2}s^{m}\mu^{2}.

Thus,

|PR|Dα=|I+II|Dαcγ0n+m+2smμ2,\displaystyle\left|P-R\right|_{D_{\alpha}}=\left|I+II\right|_{D_{\alpha}}\leq c\gamma_{0}^{n+m+2}s^{m}\mu^{2}, (3.23)

and therefore,

|R|Dα|PR|Dα+|P|D(s,r)cγ0n+m+2smμ.\displaystyle\left|R\right|_{D_{\alpha}}\leq|P-R|_{D_{\alpha}}+|P|_{D(s,r)}\leq c\gamma_{0}^{n+m+2}s^{m}\mu. (3.24)

Next, we estimate PRCβ\left\|P-R\right\|_{C^{\beta}}. In view of the definition of Cβ\|\cdot\|_{C^{\beta}}, for y,xDα\forall y,x\in D_{\alpha}, we have

PRCβ\displaystyle\left\|P-R\right\|_{C^{\beta}} =supξζ|P(x,y,ξ)R(x,y,ξ)(P(x,y,ζ)R(x,y,ζ))||ξζ|β\displaystyle=\sup_{\xi\neq\zeta}\frac{\left|P(x,y,\xi)-R(x,y,\xi)-(P(x,y,\zeta)-R(x,y,\zeta))\right|}{\left|\xi-\zeta\right|^{\beta}}
supξζ||pyp(P(x,y,ξ)R(x,y,ξ)(P(x,y,ζ)R(x,y,ζ)))|dy||ξζ|β\displaystyle\leq\sup_{\xi\neq\zeta}\frac{\left|\int\right|\frac{\partial^{p}}{\partial y^{p}}(P(x,y,\xi)-R(x,y,\xi)-(P(x,y,\zeta)-R(x,y,\zeta)))\left|dy\right|}{|\xi-\zeta|^{\beta}}
supξζ|csm+2|P(x,y,ξ)P(x,y,ζ)|dy||ξζ|β\displaystyle\leq\sup_{\xi\neq\zeta}\frac{\left|\frac{c}{s^{m+2}}\int\left|P(x,y,\xi)-P(x,y,\zeta)\right|dy\right|}{|\xi-\zeta|^{\beta}}
supξζcsm+1|P(x,y,ξ)P(x,y,ζ)||ξζ|β(αs)m+1\displaystyle\leq\sup_{\xi\neq\zeta}\frac{c}{s^{m+1}}\frac{|P(x,y,\xi)-P(x,y,\zeta)|}{|\xi-\zeta|^{\beta}}(\alpha s)^{m+1}
cμPCβcγ0n+m+2smμ2,\displaystyle\leq c\mu\|P\|_{C^{\beta}}\leq c\gamma_{0}^{n+m+2}s^{m}\mu^{2}, (3.25)

where the third inequality follows from Cauchy estimate and the last inequality follows from (3.14).

Similarly, we get

RCβ<PRCβ+PCβcγ0n+m+2smμ.\displaystyle\|R\|_{C^{\beta}}<\|P-R\|_{C^{\beta}}+\|P\|_{C^{\beta}}\leq c\gamma_{0}^{n+m+2}s^{m}\mu. (3.26)

It follows from (3.23), (3.24), (3.2.3) and (3.26) that (3.19) and (3.20) hold.

The proof is complete. ∎

3.2.4 Homological Equation

As usual, we shall construct a symplectic transformation as the time 1-map ϕF1\phi_{F}^{1} of the flow generated by a Hamiltonian FF to eliminate all resonant terms in RR, i.e., all terms

pkıyıe1k,x,0<|k|K+,|ı|m.\displaystyle p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle},~{}~{}~{}~{}0<|k|\leq K_{+},|\imath|\leq m.

To do so, we first construct a Hamiltonian FF of the form

F=0<|k|K+,|ı|mfkıyıe1k,x,F=\sum_{0<|k|\leq K_{+},|\imath|\leq m}f_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle}, (3.27)

satisfying the equation

{N,F}+R[R]=0,\{N,F\}+R-[R]=0, (3.28)

where [R]=1(2π)nTnR(y,x)𝑑x[R]=\frac{1}{(2\pi)^{n}}\int_{T^{n}}R(y,x)dx is the average of the truncation RR.

Substituting (3.27) into (3.28) yields that

0<|k|<K+,|ı|m1k,ω(ξ0)+yh¯fkıyıe1k,x\displaystyle-\sum_{0<|k|<K_{+},|\imath|\leq m}\sqrt{-1}\left\langle k,\omega(\xi_{0})+\partial_{y}\bar{h}\right\rangle f_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle}
+0<|k|<K+,|ı|mpkıyıe1k,x=0.\displaystyle+\sum_{0<|k|<K_{+},|\imath|\leq m}p_{k\imath}y^{\imath}e^{\sqrt{-1}\langle k,x\rangle}=0.

By comparing the coefficients above, we then obtain the following quasi-linear equations:

1k,ω(ξ0)+yh¯fkı=pkı,|ı|m,0<|k|K+.\sqrt{-1}\left\langle k,\omega(\xi_{0})+\partial_{y}\bar{h}\right\rangle f_{k\imath}=p_{k\imath},~{}~{}~{}|\imath|\leq m,~{}~{}~{}0<|k|\leq K_{+}. (3.29)

We declare that the quasi-linear equations (3.29) is solvable under some suitable conditions. The details can be seen in the following lemma:

Lemma 3.3.

Assume that

(H2):max|i|2|yih¯yih¯0|D(s)μ012,\displaystyle\textbf{{(H2)}}:max_{|i|\leq 2}\left|\left\|\partial_{y}^{i}\bar{h}-\partial_{y}^{i}\bar{h}_{0}\right\|\right|_{D(s)}\leq\mu_{0}^{\frac{1}{2}},
(H3):2s<γ0(M+2)K+τ+1,\displaystyle\textbf{{(H3)}}:2s<\frac{\gamma_{0}}{(M^{*}+2)K_{+}^{\tau+1}},

where

M=max|i|2,yD(s)|yih¯0(ξ0,y)|.\displaystyle M^{*}=\max_{|i|\leq 2,y\in D(s)}\left|\partial_{y}^{i}\bar{h}_{0}(\xi_{0},y)\right|.

Then the quasi-linear equations (3.29) can be uniquely solved on D(s)D(s) to obtain a family of functions fkıf_{k\imath} which are analytic in yy, and satisfy the following properties:

|yifkı|D(s)c2|k|(|i|+1)τ+|i|γ0n+m+1|i|sm|ı|μe|k|r,\left|\left\|\partial_{y}^{i}f_{k\imath}\right\|\right|_{D(s)}\leq c_{2}|k|^{(|i|+1)\tau+|i|}\gamma_{0}^{n+m+1-|i|}s^{m-|\imath|}\mu e^{-|k|r}, (3.30)

for all |ı|m,0<|k|K+,|i|2|\imath|\leq m,0<|k|\leq K_{+},|i|\leq 2, where c2c_{2} is a constant.

Proof.

For yD(s)\forall y\in D(s), by (H2),(H3)\textbf{\text{(H2)}},\textbf{(H3)}, we have

|yh¯|D(s)=|(yh¯yh¯0)+yh¯0|D(s)(1+M)|y|<(1+M)s<γ02|k|τ+1\left|\partial_{y}\bar{h}\right|_{D(s)}=\left|(\partial_{y}\bar{h}-\partial_{y}\bar{h}_{0})+\partial_{y}\bar{h}_{0}\right|_{D(s)}\leq(1+M^{*})|y|<(1+M^{*})s<\frac{\gamma_{0}}{2|k|^{\tau+1}}

and

yh¯Cβ=supξξ,ξ,ξO|yh¯(y,ξ)yh¯(y,ξ)||ξξ|βμ012|y|<|s|<γ02|k|τ+1,\left\|\partial_{y}\bar{h}\right\|_{C^{\beta}}=\sup_{\xi_{*}\neq\xi_{**},~{}\xi_{*},\xi_{**}\in{O}}\frac{\left|\partial_{y}\bar{h}(y,\xi_{*})-\partial_{y}\bar{h}(y,\xi_{**})\right|}{|\xi_{*}-\xi_{**}|^{\beta}}\leq\mu_{0}^{\frac{1}{2}}|y|<|s|<\frac{\gamma_{0}}{2|k|^{\tau+1}},

which imply that

|yh¯|D(s)<γ02|k|τ+1.\left|\left\|\partial_{y}\bar{h}\right\|\right|_{D(s)}<\frac{\gamma_{0}}{2|k|^{\tau+1}}. (3.31)

It follows from (3.31) and (A2){\rm(A2)} that

|k,ω(ξ0)+yh¯(y)|D(s)>γ0|k|τγ02|k|τ=γ02|k|τ.\left|\left\|\left\langle k,\omega(\xi_{0})+\partial_{y}\bar{h}(y)\right\rangle\right\|\right|_{D(s)}>\frac{\gamma_{0}}{|k|^{\tau}}-\frac{\gamma_{0}}{2|k|^{\tau}}=\frac{\gamma_{0}}{2|k|^{\tau}}. (3.32)

Hence

Lk=1k,ω(ξ0)+yh¯(y)L_{k}=\sqrt{-1}\left\langle k,\omega(\xi_{0})+\partial_{y}\bar{h}(y)\right\rangle (3.33)

is invertible, and

fkı=Lk1pkı,f_{k\imath}=L_{k}^{-1}p_{k\imath}, (3.34)

for all yD(s)y\in D(s), 0<|k|K+0<|k|\leq K_{+}, |ı|m|\imath|\leq m. Let 0<|k|K+0<|k|\leq K_{+}. We note by the first inequality of (3.21) and Cauchy estimate that

|pkı|\displaystyle\left|\left\|p_{k\imath}\right\|\right| |yıP|D~e|k|rγ0n+m+2sm|ı|μe|k|r,|ı|m,\displaystyle\leq\left|\left\|\partial_{y}^{\imath}P\right\|\right|_{\tilde{D}}e^{-|k|r}\leq\gamma_{0}^{n+m+2}s^{m-|\imath|}\mu e^{-|k|r},~{}~{}~{}~{}|\imath|\leq m, (3.35)

and by (3.32) and (3.33) that

|yiLk1|D(s)\displaystyle\left|\left\|\partial_{y}^{i}L_{k}^{-1}\right\|\right|_{D(s)} c2|k||i||Lk1|D(s)|i|+1c2|k|(|i|+1)τ+|i|γ0|i|+1,|i|2.\displaystyle\leq c_{2}\left|k\right|^{|i|}\left\|\left|L_{k}^{-1}\right\|\right|_{D(s)}^{|i|+1}\leq c_{2}\frac{|k|^{(|i|+1)\tau+|i|}}{\gamma_{0}^{|i|+1}},~{}~{}~{}~{}|i|\leq 2. (3.36)

So, by (3.34), (3.35) and (3.36), we get

|yifkı|D(s)\displaystyle\left|\left\|\partial_{y}^{i}f_{k\imath}\right\|\right|_{D(s)} c2|k|(|i|+1)τ+|i|γ0|i|+1γ0n+m+2sm|ı|μe|k|r\displaystyle\leq c_{2}\frac{|k|^{(|i|+1)\tau+|i|}}{\gamma_{0}^{|i|+1}}\gamma_{0}^{n+m+2}s^{m-|\imath|}\mu e^{-|k|r}
=c2|k|(|i|+1)τ+|i|γ0n+m+1|i|sm|ı|μe|k|r,|i|2.\displaystyle=c_{2}|k|^{(|i|+1)\tau+|i|}\gamma_{0}^{n+m+1-|i|}s^{m-|\imath|}\mu e^{-|k|r},~{}~{}~{}~{}|i|\leq 2.

The proof is complete. ∎

Next, we apply the above transformation ϕF1\phi_{F}^{1} to Hamiltonian HH, i.e.,

HϕF1=\displaystyle H\circ\phi_{F}^{1}= (N+R)ϕF1+(PR)ϕF1\displaystyle(N+R)\circ\phi_{F}^{1}+(P-R)\circ\phi_{F}^{1}
=\displaystyle= (N+R)+{N,F}+01{(1t){N,F}+R,F}ϕFt𝑑t\displaystyle(N+R)+\{N,F\}+\int_{0}^{1}\left\{(1-t)\{N,F\}+R,F\right\}\circ\phi_{F}^{t}dt
+(PR)ϕF1\displaystyle+(P-R)\circ\phi_{F}^{1}
=\displaystyle= N+[R]+01{Rt,F}ϕFt𝑑t+(PR)ϕF1\displaystyle N+[R]+\int_{0}^{1}\left\{R_{t},F\right\}\circ\phi_{F}^{t}dt+\left(P-R\right)\circ\phi_{F}^{1}
=\displaystyle= :N¯++P¯+,\displaystyle:\bar{N}_{+}+\bar{P}_{+},

where

N¯+=N+[R]=e++ω(ξ),y+j=0νp01j(ξ),y+h¯+(y,ξ),\displaystyle\bar{N}_{+}=N+[R]=e_{+}+\langle\omega(\xi),y\rangle+\left\langle\sum_{j=0}^{\nu}p_{01}^{j}(\xi),y\right\rangle+\bar{h}_{+}(y,\xi), (3.37)
e+=e+p00ν,\displaystyle e_{+}=e+p_{00}^{\nu}, (3.38)
h¯+=h¯(y,ξ)+[R]p00νp01ν(ξ),y,\displaystyle\bar{h}_{+}=\bar{h}(y,\xi)+[R]-p_{00}^{\nu}-\left\langle p_{01}^{\nu}(\xi),y\right\rangle, (3.39)
P¯+=01{Rt,F}ϕFt𝑑t+(PR)ϕF1,\displaystyle\bar{P}_{+}=\int_{0}^{1}\{R_{t},F\}\circ\phi_{F}^{t}dt+(P-R)\circ\phi_{F}^{1}, (3.40)
Rt=(1t)[R]+tR.\displaystyle R_{t}=(1-t)[R]+tR.

3.2.5 Translation

In this subsection, we will construct a translation so as to keep the frequency unchanged. It should be pointed out that we present a new way to move parameters, but in the usual KAM iteration, one has to dig out a decreasing series of parameter domains, in which the Diophantine condition doesn’t hold, see [10, 17, 23, 29, 30, 31, 32].

Consider the translation

ϕ:xx,yy,ξ~ξ~+ξ+ξ,\phi:x\rightarrow x,~{}~{}~{}~{}~{}y\rightarrow y,~{}~{}~{}~{}~{}\tilde{\xi}\rightarrow\tilde{\xi}+\xi_{+}-\xi,

where ξ+\xi_{+} is to be determined. Let

Φ+=ϕF1ϕ.\Phi_{+}=\phi_{F}^{1}\circ\phi.

Then

HΦ+\displaystyle H\circ\Phi_{+} =N++P+,\displaystyle=N_{+}+P_{+},
N+\displaystyle N_{+} =N¯+ϕ=e++ω(ξ+),y+j=0νp01j(ξ+),y+h¯+(y,ξ+),\displaystyle=\bar{N}_{+}\circ\phi=e_{+}+\left\langle\omega(\xi_{+}),y\right\rangle+\left\langle\sum_{j=0}^{\nu}p_{01}^{j}(\xi_{+}),y\right\rangle+\bar{h}_{+}(y,\xi_{+}), (3.41)
P+\displaystyle P_{+} =P¯+ϕ.\displaystyle=\bar{P}_{+}\circ\phi. (3.42)

3.2.6 Frequency-preserving

In this subsection, we will show that the frequency can be preserved in the iteration process. Recall the topological degree condition (A0)\rm(A0) and the weak convexity condition (A1)\rm(A1). The former ensures that the parameter ξ+\xi_{+} can be found in the parameter set to keep the frequency unchanged at this KAM step. The later assures that the distance between ξ+\xi_{+} and ξ\xi is smaller than the distance between ξ\xi and ξν1\xi_{\nu-1}, i.e., the sequence of parameters is convergent after infinite steps of iteration. The following lemma is crucial to our arguments.

Lemma 3.4.

Assume that

(H4):|j=0νp01j|D(s,r)<μ012.\displaystyle\textbf{({H4})}:\left|\left\|\sum_{j=0}^{\nu}p_{01}^{j}\right\|\right|_{D(s,r)}<\mu_{0}^{\frac{1}{2}}.

There exists ξ+Bcμ1/L(ξ)Oo\xi_{+}\in B_{c\mu^{{1}/{L}}}(\xi)\subset O^{o} such that

ω(ξ+)+j=0νp01j(ξ+)=ω(ξ0).\displaystyle\omega(\xi_{+})+\sum_{j=0}^{\nu}p_{01}^{j}(\xi_{+})=\omega(\xi_{0}). (3.43)
Proof.

The proof will be completed by an induction on ν\nu. We start with the case ν=0\nu=0. It is obvious that ω(ξ0)=ω(ξ0)\omega(\xi_{0})=\omega(\xi_{0}). Now assume that for some ν>0\nu>0 we have got

ω(ξi)+j=0i1p01j(ξi)=ω(ξ0),ξiBcμi11/L(ξi1)Oo,i=1,,ν.\displaystyle\omega(\xi_{i})+\sum_{j=0}^{i-1}p_{01}^{j}(\xi_{i})=\omega(\xi_{0}),~{}~{}\xi_{i}\in B_{c\mu_{i-1}^{{1}/{L}}}(\xi_{i-1})\subset O^{o},~{}~{}~{}i=1,\cdots,\nu. (3.44)

We need to find ξ+\xi_{+} near ξ\xi such that

ω(ξ+)+j=0νp01j(ξ+)=ω(ξ0).\displaystyle\omega(\xi_{+})+\sum_{j=0}^{\nu}p_{01}^{j}(\xi_{+})=\omega(\xi_{0}). (3.45)

In view of the property of topological degree, (H4) and (A0)\rm(A0), we have

deg(ω()+j=0νp01j(),Oo,ω(ξ0))=deg(ω(),Oo,ω(ξ0))0,\displaystyle\deg\left(\omega(\cdot)+\sum_{j=0}^{\nu}p_{01}^{j}(\cdot),O^{o},\omega(\xi_{0})\right)=\deg\left(\omega(\cdot),O^{o},\omega(\xi_{0})\right)\neq 0,

i.e., there exists at least a ξ+Oo\xi_{+}\in O^{o} such that (3.43) holds.

Next, we estimate |ξ+ξ||\xi_{+}-\xi|. (3.20) in Lemma 3.2 implies that

p01jCβ<cμj,j=0,1,,ν,\displaystyle\left\|p_{01}^{j}\right\|_{C^{\beta}}<c\mu_{j},~{}~{}~{}j=0,1,\cdots,\nu,

i.e.,

|p01j(ξ+)p01j(ξ)|<cμj|ξ+ξ|β,ξ+,ξO.\displaystyle\left|p_{01}^{j}(\xi_{+})-p_{01}^{j}(\xi)\right|<c\mu_{j}\left|\xi_{+}-\xi\right|^{\beta},~{}~{}\forall\xi_{+},\xi\in{O}. (3.46)

According to (3.44) and (3.45), we get

ω(ξ+)ω(ξ)+j=0ν1(p01j(ξ+)p01j(ξ))=p01ν(ξ+).\displaystyle\omega(\xi_{+})-\omega(\xi)+\sum_{j=0}^{\nu-1}\left(p_{01}^{j}(\xi_{+})-p_{01}^{j}{(\xi)}\right)=-p_{01}^{\nu}(\xi_{+}). (3.47)

This together with (A1)\rm(A1) and (3.46) yields

|p01ν(ξ+)|\displaystyle\left|p_{01}^{\nu}(\xi_{+})\right| =|ω(ξ+)ω(ξ)+j=0ν1(p01j(ξ+)p01j(ξ))|\displaystyle=\left|\omega(\xi_{+})-\omega(\xi)+\sum_{j=0}^{\nu-1}(p_{01}^{j}(\xi_{+})-p_{01}^{j}{(\xi)})\right|
|ω(ξ+)ω(ξ)|j=0ν1|p01j(ξ+)p01j(ξ)|\displaystyle\geq\left|\omega(\xi_{+})-\omega(\xi)\right|-\sum_{j=0}^{\nu-1}\left|p_{01}^{j}(\xi_{+})-p_{01}^{j}{(\xi)}\right|
σ|ξ+ξ|Lcj=0ν1μj|ξ+ξ|β\displaystyle\geq\sigma\left|\xi_{+}-\xi\right|^{L}-c\sum_{j=0}^{\nu-1}\mu_{j}\left|\xi_{+}-\xi\right|^{\beta}
(σcj=0ν1μj)|ξ+ξ|L,\displaystyle\geq(\sigma-c\sum_{j=0}^{\nu-1}\mu_{j})\left|\xi_{+}-\xi\right|^{L}, (3.48)

where the last inequality follows from 0<Lβ0<L\leq\beta in (A1)\rm(A1). Then, by (3.48) and (3.20) in Lemma 3.2, we have

|ξ+ξ|L\displaystyle|\xi_{+}-\xi|^{L} p01ν(ξ+)σcj=0ν1μj<cμσcj=0ν1μj<2cμσ,\displaystyle\leq\frac{p_{01}^{\nu}(\xi_{+})}{\sigma-c\sum_{j=0}^{\nu-1}\mu_{j}}<\frac{c\mu}{\sigma-c\sum_{j=0}^{\nu-1}\mu_{j}}<\frac{2c\mu}{\sigma},

which implies ξ+Bcμ1/L(ξ)\xi_{+}\in B_{c\mu^{{1}/{L}}}(\xi). From ξOo\xi\in O^{o} in (3.44) and the fact ε\varepsilon is small enough (i.e., μ\mu is small enough), we have Bcμ1/L(ξ)OoB_{c\mu^{{1}/{L}}}(\xi)\subset O^{o}.

The proof is complete. ∎

3.2.7 Estimate on N+N_{+}

Now, we give the estimate on the next step N+N_{+}.

Lemma 3.5.

There is a constant c3c_{3} such that the following holds:

|ξ+ξ|c3μ1L,\displaystyle\left|\xi_{+}-\xi\right|\leq c_{3}\mu^{\frac{1}{L}}, (3.49)
|e+e|c3smμ,\displaystyle\left|e_{+}-e\right|\leq c_{3}s^{m}\mu, (3.50)
|h¯+h¯|D(s)c3smμ.\displaystyle\left|\left\|\bar{h}_{+}-\bar{h}\right\|\right|_{D(s)}\leq c_{3}s^{m}\mu. (3.51)
Proof.

It is obvious by ξ+Bcμ1/L(ξ)\xi_{+}\in B_{c\mu^{{1}/{L}}}(\xi) in Lemma 3.4 that (3.49) holds. It follows from (3.38) and (3.39) that (3.50) and (3.51) hold. ∎

3.2.8 Estimate on Φ+\Phi_{+}

Recall that FF is as in (3.27) with the coefficients and its estimate given by Lemma 3.3. Then, we have the following estimate on FF.

Lemma 3.6.

There is a constant c4c_{4} such that for all |j|+|i|2|j|+|i|\leq 2,

|xjyiF|D^c4γ0n+m+1sm|i|μΓ(rr+).\displaystyle\left|\left\|\partial_{x}^{j}\partial_{y}^{i}F\right\|\right|_{\hat{D}}\leq c_{4}\gamma_{0}^{n+m+1}s^{m-|i|}\mu\Gamma(r-r_{+}). (3.52)
Proof.

By (3.27) and (3.30), we have

|xjyiF|D^\displaystyle\left|\left\|\partial_{x}^{j}\partial_{y}^{i}F\right\|\right|_{\hat{D}} |ı|m,0<|k|K+|k|j|yi(fkıyı)|D(s)e|k|(r++78(rr+))\displaystyle\leq\sum_{|\imath|\leq m,0<|k|\leq K_{+}}|k|^{j}\left|\left\|\partial_{y}^{i}(f_{k\imath}y^{\imath})\right\|\right|_{D(s)}e^{|k|(r_{+}+\frac{7}{8}(r-r_{+}))}
c40<|k|K+|k|τ+|j|γ0n+m+1sm|i|μe|k|rr+8\displaystyle\leq c_{4}\sum_{0<|k|\leq K_{+}}|k|^{\tau+|j|}\gamma_{0}^{n+m+1}s^{m-|i|}\mu e^{-|k|\frac{r-r_{+}}{8}}
c4γ0n+m+1sm|i|μΓ(rr+).\displaystyle\leq c_{4}\gamma_{0}^{n+m+1}s^{m-|i|}\mu\Gamma(r-r_{+}).

The proof is complete. ∎

Lemma 3.7.

Assume that

(H5):c4sm1μΓ(rr+)<18(rr+),\displaystyle\textbf{{(H5)}}:c_{4}s^{m-1}\mu\Gamma(r-r_{+})<\frac{1}{8}(r-r_{+}),
(H6):c4smμΓ(rr+)<18αs.\displaystyle\textbf{{(H6)}}:c_{4}s^{m}\mu\Gamma(r-r_{+})<\frac{1}{8}\alpha s.

Then the following holds.

(1)

For all 0t10\leq t\leq 1, the mappings

ϕFt\displaystyle\phi_{F}^{t} :D14αD12α,\displaystyle:D_{\frac{1}{4}\alpha}\rightarrow D_{\frac{1}{2}\alpha}, (3.53)
ϕ\displaystyle\phi :OO+,\displaystyle:O\rightarrow O_{+}, (3.54)

are well defined.

(2)

Φ+:D+D(s,r).\Phi_{+}:D_{+}\rightarrow D(s,r).

(3)

There is a constant c5c_{5} such that

|ϕFtid|D~\displaystyle\left|\left\|\phi_{F}^{t}-id\right\|\right|_{\tilde{D}} c5μΓ(rr+),\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}),
|DϕFtId|D~\displaystyle\left|\left\|D\phi_{F}^{t}-Id\right\|\right|_{\tilde{D}} c5μΓ(rr+),\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}),
|D2ϕFt|D~\displaystyle\left|\left\|D^{2}\phi_{F}^{t}\right\|\right|_{\tilde{D}} c5μΓ(rr+).\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}).
(4)
|Φ+id|D~\displaystyle\left|\left\|\Phi_{+}-id\right\|\right|_{\tilde{D}} c5μΓ(rr+),\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}),
|DΦ+Id|D~\displaystyle\left|\left\|D\Phi_{+}-Id\right\|\right|_{\tilde{D}} c5μΓ(rr+),\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}),
|D2Φ+|D~\displaystyle\left|\left\|D^{2}\Phi_{+}\right\|\right|_{\tilde{D}} c5μΓ(rr+).\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}).
Proof.

(1)  (3.54) immediately follows from (3.49) and the definition of O+O_{+}. To verify (3.53), we denote ϕF1t\phi_{F_{1}}^{t}, ϕF2t\phi_{F_{2}}^{t} as the components of ϕFt\phi_{F}^{t} in the yy, xx planes, respectively. Let XF=(Fy,Fx)X_{F}=(F_{y},-F_{x})^{\top} be the vector field generated by FF. Then

ϕFt=id+0tXFϕFu𝑑u,0t1.\displaystyle\phi_{F}^{t}=id+\int_{0}^{t}X_{F}\circ\phi_{F}^{u}du,~{}~{}~{}~{}0\leq t\leq 1. (3.55)

For any (y,x)D14α(y,x)\in D_{\frac{1}{4}\alpha}, we let t=sup{t[0,1]:ϕFt(y,x)Dα}t_{*}=\sup\{t\in[0,1]:\phi_{F}^{t}(y,x)\in D_{\alpha}\}. Then for any 0tt0\leq t\leq t_{*}, by (y,x)D14α(y,x)\in D_{\frac{1}{4}\alpha}, (3.52) in Lemma 3.6, (H5) and (H6), we can get the following estimates:

|ϕF1t(y,x)|D14α\displaystyle\left|\left\|\phi_{F_{1}}^{t}(y,x)\right\|\right|_{D_{\frac{1}{4}\alpha}} |y|+0t|FxϕFu|D14αdu\displaystyle\leq\left|y\right|+\int_{0}^{t}\left|\left\|F_{x}\circ\phi_{F}^{u}\right\|\right|_{D_{\frac{1}{4}\alpha}}du
14αs+c4smμΓ(rr+)\displaystyle\leq\frac{1}{4}\alpha s+c_{4}s^{m}\mu\Gamma(r-r_{+})
<38αs,\displaystyle<\frac{3}{8}\alpha s,
|ϕF2t(y,x)|D14α\displaystyle\left|\left\|\phi_{F_{2}}^{t}(y,x)\right\|\right|_{D_{\frac{1}{4}\alpha}} |x|+0t|FyϕFu|D14αdu\displaystyle\leq|x|+\int_{0}^{t}\left|\left\|F_{y}\circ\phi_{F}^{u}\right\|\right|_{D_{\frac{1}{4}\alpha}}du
r++18(rr+)+c4sm1μΓ(rr+)\displaystyle\leq r_{+}+\frac{1}{8}(r-r_{+})+c_{4}s^{m-1}\mu\Gamma(r-r_{+})
<r++28(rr+).\displaystyle<r_{+}+\frac{2}{8}(r-r_{+}).

Thus, ϕFtD12αDα\phi_{F}^{t}\in D_{\frac{1}{2}\alpha}\subset D_{\alpha}, i.e. t=1t_{*}=1 and (1) holds.

(2) It follows from (1) that (2) holds.

(3) Using (3.52) in Lemma 3.6 and (3.55), we immediately have

|ϕFtid|D~c5μΓ(rr+).\displaystyle\left|\left\|\phi_{F}^{t}-id\right\|\right|_{\tilde{D}}\leq c_{5}\mu\Gamma(r-r_{+}).

By (3.52) in Lemma 3.6, (3.55) and Gronwall Inequality, we get

|DϕFtId|D~\displaystyle\left|\left\|D\phi_{F}^{t}-Id\right\|\right|_{\tilde{D}} |0tDXFϕFλDϕFλ𝑑λ|D~\displaystyle\leq\left|\left\|\int_{0}^{t}DX_{F}\circ\phi_{F}^{\lambda}D\phi_{F}^{\lambda}d\lambda\right\|\right|_{\tilde{D}}
0t|DXFϕFλ|D~|DϕFλId|D~𝑑λ+0t|DXFϕFλ|D~𝑑λ\displaystyle\leq\int_{0}^{t}\left|\left\|DX_{F}\circ\phi_{F}^{\lambda}\right\|\right|_{\tilde{D}}\left|\left\|D\phi_{F}^{\lambda}-Id\right\|\right|_{\tilde{D}}d\lambda+\int_{0}^{t}\left|\left\|DX_{F}\circ\phi_{F}^{\lambda}\right\|\right|_{\tilde{D}}d\lambda
c5μΓ(rr+).\displaystyle\leq c_{5}\mu\Gamma(r-r_{+}).

It follows from the induction and a similar argument that we have the estimates on the 2-order derivatives of ϕFt\phi_{F}^{t}, i.e.,

|D2ϕFt|D~c5μΓ(rr+).\displaystyle\left|\left\|D^{2}\phi_{F}^{t}\right\|\right|_{\tilde{D}}\leq c_{5}\mu\Gamma(r-r_{+}).

(4) now follows from (3).

The proof is complete. ∎

3.2.9 Estimate on P+P_{+}

In the following, we estimate the next step P+P_{+}.

Lemma 3.8.

Assume (H1)-(H6). Then there is a constant c6c_{6} such that,

|P+|D+c6γ0n+m+2smμ2(Γ2(rr+)+Γ(rr+)).\left|\left\|P_{+}\right\|\right|_{D_{+}}\leq c_{6}\gamma_{0}^{n+m+2}s^{m}\mu^{2}(\Gamma^{2}(r-r_{+})+\Gamma(r-r_{+})). (3.56)

Moreover, if

(H7):μρ(Γ2(rr+)+Γ(rr+))1\displaystyle\textbf{{(H7)}}:\mu^{\rho}(\Gamma^{2}(r-r_{+})+\Gamma(r-r_{+}))\leq 1

then,

|P+|D+γ0n+m+2s+mμ+.\left|\left\|P_{+}\right\|\right|_{D_{+}}\leq\gamma_{0}^{n+m+2}s_{+}^{m}\mu_{+}. (3.57)
Proof.

By (3.19) and (3.20) in Lemma 3.2, (3.52) in Lemma 3.6 and Lemma 3.7 (3), we have that, for all 0t10\leq t\leq 1,

|{Rt,F}ϕFt|D14α\displaystyle\left|\left\|\{R_{t},F\}\circ\phi_{F}^{t}\right\|\right|_{D_{\frac{1}{4}\alpha}} cγ0n+m+2smμ2Γ2(rr+),\displaystyle\leq c\gamma_{0}^{n+m+2}s^{m}\mu^{2}\Gamma^{2}(r-r_{+}),
|(PR)ϕF1|D14α\displaystyle\left|\left\|(P-R)\circ\phi_{F}^{1}\right\|\right|_{D_{\frac{1}{4}\alpha}} cγ0n+m+2smμ2Γ(rr+).\displaystyle\leq c\gamma_{0}^{n+m+2}s^{m}\mu^{2}\Gamma(r-r_{+}).

So, by (3.40),

|P¯+|D14αcγ0n+m+2smμ2(Γ2(rr+)+Γ(rr+)).\left|\left\|\bar{P}_{+}\right\|\right|_{D_{\frac{1}{4}\alpha}}\leq c\gamma_{0}^{n+m+2}s^{m}\mu^{2}\left(\Gamma^{2}(r-r_{+})+\Gamma(r-r_{+})\right).

By (H7), we see that,

|P+|D+\displaystyle\left|\left\|P_{+}\right\|\right|_{D_{+}} 8mc0μ1+ρs+mμ12ρmm+1γ0n+m+2(μρ(Γ2(rr+)+Γ(rr+)))\displaystyle\leq 8^{m}c_{0}\mu^{1+\rho}s_{+}^{m}\mu^{1-2\rho-\frac{m}{m+1}}\gamma_{0}^{n+m+2}\left(\mu^{\rho}\left(\Gamma^{2}(r-r_{+})+\Gamma(r-r_{+})\right)\right)
γ0n+m+2s+mμ+,\displaystyle\leq\gamma_{0}^{n+m+2}s_{+}^{m}\mu_{+},

which implies (3.57).

The proof is complete. ∎

This completes one cycle of KAM steps.

4 Proof of Theorem 1

4.1 Iteration lemma

In this subsection, we will prove an iteration lemma which guarantees the inductive construction of the transformations in all KAM steps.

Let r0,s0,γ0,μ0,H0,N0,e0,h¯0,P0r_{0},s_{0},\gamma_{0},\mu_{0},H_{0},N_{0},e_{0},\bar{h}_{0},P_{0} be given at the beginning of Section 3 and let D0=D(s0,r0)D_{0}=D(s_{0},r_{0}), K0=0K_{0}=0, Φ0=id\Phi_{0}=id. We define the following sequence inductively for all ν=1,2,\nu=1,2,\cdots.

rν\displaystyle r_{\nu} =r0(1i=1ν12i+1),\displaystyle=r_{0}\left(1-\sum_{i=1}^{\nu}\frac{1}{2^{i+1}}\right),
sν\displaystyle s_{\nu} =18αν1sν1,\displaystyle=\frac{1}{8}\alpha_{\nu-1}s_{\nu-1},
αν\displaystyle\alpha_{\nu} =μν2ρ=μν1m+1,\displaystyle=\mu_{\nu}^{2\rho}=\mu_{\nu}^{\frac{1}{m+1}},
μν\displaystyle\mu_{\nu} =8mc0μν11+ρ,\displaystyle=8^{m}c_{0}\mu_{\nu-1}^{1+\rho},
Kν\displaystyle K_{\nu} =([log(1μν1)]+1)3η,\displaystyle=\left(\left[\log\left(\frac{1}{\mu_{\nu-1}}\right)\right]+1\right)^{3\eta},
D~ν\displaystyle\tilde{D}_{\nu} =D(12sν,rν+68(rν1rν)).\displaystyle=D\left(\frac{1}{2}s_{\nu},r_{\nu}+\frac{6}{8}\left(r_{\nu-1}-r_{\nu}\right)\right).
Lemma 4.1.

Denote

μ=μ0(M+2)m1K1(τ+1)(m1).\displaystyle\mu_{*}=\frac{\mu_{0}}{\left(M^{*}+2\right)^{m-1}K_{1}^{(\tau+1)(m-1)}}.

If ε\varepsilon is small enough, then the KAM step described on the above is valid for all ν=0,1,\nu=0,1,\cdots, resulting the sequences

Hν,Nν,eν,h¯ν,Pν,Φν,H_{\nu},N_{\nu},e_{\nu},\bar{h}_{\nu},P_{\nu},\Phi_{\nu},

ν=1,2,,\nu=1,2,\cdots, with the following properties:

(1)
|eν+1eν|\displaystyle\left|e_{\nu+1}-e_{\nu}\right| μ122ν,\displaystyle\leq\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}}, (4.58)
|eνe0|\displaystyle\left|e_{\nu}-e_{0}\right| 2μ12,\displaystyle\leq 2\mu_{*}^{\frac{1}{2}}, (4.59)
|h¯ν+1h¯ν|D(sν)\displaystyle\left|\left\|\bar{h}_{\nu+1}-\bar{h}_{\nu}\right\|\right|_{D(s_{\nu})} μ122ν,\displaystyle\leq\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}}, (4.60)
|h¯νh¯0|D(sν)\displaystyle\left|\left\|\bar{h}_{\nu}-\bar{h}_{0}\right\|\right|_{D(s_{\nu})} 2μ12,\displaystyle\leq 2\mu_{*}^{\frac{1}{2}}, (4.61)
|Pν|D(sν,rν)\displaystyle\left|\left\|P_{\nu}\right\|\right|_{D(s_{\nu},r_{\nu})} μ122ν,\displaystyle\leq\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}}, (4.62)
|ξν+1ξν|\displaystyle\left|\xi_{\nu+1}-\xi_{\nu}\right| (μ122ν)1L.\displaystyle\leq(\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}})^{\frac{1}{L}}. (4.63)
(2)

Φν+1:D~ν+1D~ν\Phi_{\nu+1}:\tilde{D}_{\nu+1}\rightarrow\tilde{D}_{\nu} is symplectic, and

|Φν+1id|D~ν+1μ122ν.\displaystyle\left|\left\|\Phi_{\nu+1}-id\right\|\right|_{\tilde{D}_{\nu+1}}\leq\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}}. (4.64)

Moreover, on Dν+1D_{\nu+1},

Hν+1=HνΦν+1=Nν+1+Pν+1.H_{\nu+1}=H_{\nu}\circ\Phi_{\nu+1}=N_{\nu+1}+P_{\nu+1}.
Proof.

The proof amounts to the verification of (H1)-(H7) for all ν\nu. For simplicity, we let r0=1r_{0}=1. It follows from ε\varepsilon small enough that μ0\mu_{0} is small. So, we see that (H2), (H4)-(H7) hold for ν=0\nu=0. From (3.8), (H3) holds for ν=0\nu=0. According to the definition of μν\mu_{\nu}, we see that

μν=(8mc0)(1+ρ)ν1ρμ0(1+ρ)ν.\mu_{\nu}=(8^{m}c_{0})^{\frac{(1+\rho)^{\nu}-1}{\rho}}\mu_{0}^{(1+\rho)^{\nu}}. (4.65)

Let ζ1\zeta\gg 1 be fixed and μ0\mu_{0} be small enough so that

μ0<(18mc0ζ)1ρ<1.\mu_{0}<(\frac{1}{8^{m}c_{0}\zeta})^{\frac{1}{\rho}}<1. (4.66)

Then

μ1\displaystyle\mu_{1} =8mc0μ01+ρ<1ζμ0<1,\displaystyle=8^{m}c_{0}\mu_{0}^{1+\rho}<\frac{1}{\zeta}\mu_{0}<1,
μ2\displaystyle\mu_{2} =8mc0μ11+ρ<1ζμ1<1ζ2μ0,\displaystyle=8^{m}c_{0}\mu_{1}^{1+\rho}<\frac{1}{\zeta}\mu_{1}<\frac{1}{\zeta^{2}}\mu_{0},
\displaystyle\vdots
μν\displaystyle\mu_{\nu} =8mc0μν11+ρ<<1ζνμ0.\displaystyle=8^{m}c_{0}\mu_{\nu-1}^{1+\rho}<\cdots<\frac{1}{\zeta^{\nu}}\mu_{0}. (4.67)

Denote

Γν=Γ(rνrν+1).\Gamma_{\nu}=\Gamma(r_{\nu}-r_{\nu+1}).

We notice that

rνrν+1r0=12ν+2.\frac{r_{\nu}-r_{\nu+1}}{r_{0}}=\frac{1}{2^{\nu+2}}. (4.68)

Since

Γν\displaystyle\Gamma_{\nu} 1t3τ+5et2ν+5\displaystyle\leq\int_{1}^{\infty}t^{3\tau+5}e^{-\frac{t}{2^{\nu+5}}}
(3τ+5)!2(ν+5)(3τ+5),\displaystyle\leq(3\tau+5)!2^{(\nu+5)(3\tau+5)},

it is obvious that if ζ\zeta is large enough, then

μνρΓνi<μνρ(Γνi+Γν)1,i=1,2,\displaystyle\mu_{\nu}^{\rho}\Gamma_{\nu}^{i}<\mu_{\nu}^{\rho}(\Gamma_{\nu}^{i}+\Gamma_{\nu})\leq 1,~{}~{}~{}i=1,2,

which implies that (H7) holds for all ν1\nu\geq 1, and

μνΓνμν1ρμ01ρζ(1ρ)ν.\mu_{\nu}\Gamma_{\nu}\leq\mu_{\nu}^{1-\rho}\leq\frac{\mu_{0}^{1-\rho}}{\zeta^{(1-\rho)\nu}}. (4.69)

By (4.68) and (4.69), it is easy to verify that (H5), (H6) hold for all ν1\nu\geq 1 as ζ\zeta is large enough and ε\varepsilon is small enough.

By (3.20) in Lemma 3.2 and (4.67), we have

|j=0νp01j|D(sν,rν)<cj=0νμj<cj=0ν1ζjμ0<cμ012,\displaystyle\left|\left\|\sum_{j=0}^{\nu}p_{01}^{j}\right\|\right|_{D(s_{\nu},r_{\nu})}<c\sum_{j=0}^{\nu}\mu_{j}<c\sum_{j=0}^{\nu}\frac{1}{\zeta^{j}}\mu_{0}<c\mu_{0}^{\frac{1}{2}},

which implies (H4).

To verify (H3), we observe by (4.65) and (4.67) that

14(M+2)μν12ρKν+1τ+1<12ν+2,\frac{1}{4}\left(M^{*}+2\right)\mu_{\nu-1}^{2\rho}K_{\nu+1}^{\tau+1}<\frac{1}{2^{\nu+2}},

as ζ\zeta is large enough. Then

2(M+2)sνKν+1τ+1\displaystyle 2\left(M^{*}+2\right)s_{\nu}K_{\nu+1}^{\tau+1} sν14(M+1)μν12ρKν+1τ+1\displaystyle\leq\frac{s_{\nu-1}}{4}\left(M^{*}+1\right)\mu_{\nu-1}^{2\rho}K_{\nu+1}^{\tau+1} (4.70)
s02ν+2<γ02ν+2<γ0,\displaystyle\leq\frac{s_{0}}{2^{\nu+2}}<\frac{\gamma_{0}}{2^{\nu+2}}<\gamma_{0}, (4.71)

which verifies (H3) for all ν1\nu\geq 1.

Let ζ1ρ2\zeta^{1-\rho}\geq 2 in (4.66), (4.67). We have that for all ν1\nu\geq 1

c0μν\displaystyle c_{0}\mu_{\nu} μ02νμ122ν,\displaystyle\leq\frac{\mu_{0}}{2^{\nu}}\leq\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}}, (4.72)
c0μνΓν\displaystyle c_{0}\mu_{\nu}\Gamma_{\nu} μ01ρ2νμ122ν,\displaystyle\leq\frac{\mu_{0}^{1-\rho}}{2^{\nu}}\leq\frac{\mu_{*}^{\frac{1}{2}}}{2^{\nu}}, (4.73)
c0sνm1μν\displaystyle c_{0}s_{\nu}^{m-1}\mu_{\nu} μ01+2ρ(m1)s0m12ν+3μ2ν.\displaystyle\leq\frac{\mu_{0}^{1+2\rho(m-1)}s_{0}^{m-1}}{2^{\nu+3}}\leq\frac{\mu_{*}}{2^{\nu}}. (4.74)

The verification of (H2) follows from (4.72)(\ref{eq30}) and an induction application of (3.51)(\ref{equ4}) in Lemma 3.5 for all ν=0,1,.\nu=0,1,\cdots.

Since (1+ρ)η>2\left(1+\rho\right)^{\eta}>2, we have

12ν+6([log1μ]+1)η\displaystyle\frac{1}{2^{\nu+6}}\left(\left[\log\frac{1}{\mu}\right]+1\right)^{\eta} 12ν+6((1(1+ρ)ν)log(8mc0)(1+ρ)νlogμ0)η\displaystyle\geq\frac{1}{2^{\nu+6}}\left(\left(1-\left(1+\rho\right)^{\nu}\right)\log\left(8^{m}c_{0}\right)-\left(1+\rho\right)^{\nu}\log\mu_{0}\right)^{\eta}
12ν+6(1+ρ)ην(logμ0)η1.\displaystyle\geq-\frac{1}{2^{\nu+6}}\left(1+\rho\right)^{\eta\nu}\left(\log\mu_{0}\right)^{\eta}\geq 1.

It follows from the above that

log(n+1)!+(ν+6)nlog2+3nηlog([log1μ]+1)12(ν+6)([log1μ]+1)3η\displaystyle\log\left(n+1\right)!+\left(\nu+6\right)n\log 2+3n\eta\log\left(\left[\log\frac{1}{\mu}\right]+1\right)-\frac{1}{2^{\left(\nu+6\right)}}\left(\left[\log\frac{1}{\mu}\right]+1\right)^{3\eta}
log(n+1)!+(ν+6)nlog2+3nηlog(log1μ+2)(log1μ)2η\displaystyle\leq\log\left(n+1\right)!+\left(\nu+6\right)n\log 2+3n\eta\log\left(\log\frac{1}{\mu}+2\right)-\left(\log\frac{1}{\mu}\right)^{2\eta}
log1μ,\displaystyle\leq-\log\frac{1}{\mu},

as μ\mu is small, which is ensured by making ε\varepsilon small. Thus,

Kν+1tnet2ν+6𝑑t(n+1)!2(ν+6)nKν+1neKν+12ν+6μ,\int_{K_{\nu+1}}^{\infty}t^{n}e^{-\frac{t}{2^{\nu+6}}}dt\leq\left(n+1\right)!2^{\left(\nu+6\right)n}K_{\nu+1}^{n}e^{-\frac{K_{\nu+1}}{2^{\nu+6}}}\leq\mu,

i.e. (H1) holds.

Above all, the KAM steps described in Section 3 are valid for all ν\nu, which give the desired sequences stated in the lemma.

Now, (4.58) and (4.60) follow from Lemma 3.5, (4.72) and (4.74); by adding up (4.58), (4.60) for all ν=0,1,\nu=0,1,\cdots, we can get (4.59), (4.61); (4.62) follows from (3.57) in Lemma 3.8 and (4.72); (4.63) follows from (3.49) in Lemma 3.5 and (4.72); (2)(2) follows from Lemma 3.7.

The proof is complete. ∎

4.2 Convergence

The convergence is standard. For the sake of completeness, we briefly give the framework of proof. Let

Ψν=Φ1Φ2Φν,ν=1,2,.\displaystyle\Psi^{\nu}=\Phi_{1}\circ\Phi_{2}\circ\cdots\circ\Phi_{\nu},~{}~{}~{}~{}\nu=1,2,\cdots.

By Lemma 4.1, we have

Dν+1\displaystyle D_{\nu+1} Dν,\displaystyle\subset D_{\nu},
Ψν\displaystyle\Psi^{\nu} :D~νD~0,\displaystyle:\tilde{D}_{\nu}\rightarrow\tilde{D}_{0},
H0Ψν\displaystyle H_{0}\circ\Psi^{\nu} =H=Nν+Pν,\displaystyle=H=N_{\nu}+P_{\nu},
Nν\displaystyle N_{\nu} =eν+ω(ξν)+j=0ν1p01j(ξν),y+h¯ν(y,ξν),\displaystyle=e_{\nu}+\left\langle\omega\left(\xi_{\nu}\right)+\sum_{j=0}^{\nu-1}p_{01}^{j}\left(\xi_{\nu}\right),y\right\rangle+\bar{h}_{\nu}(y,\xi_{\nu}),

ν=0,1,,\nu=0,1,\cdots, where Ψ0=id\Psi_{0}=id. Using (4.64) and the identity

Ψν=id+i=1ν(ΨiΨi1),\displaystyle\Psi^{\nu}=id+\sum_{i=1}^{\nu}\left(\Psi^{i}-\Psi^{i-1}\right),

it is easy to verify that Ψν\Psi^{\nu} is uniformly convergent and denote the limitation by Ψ\Psi^{\infty}.

In view of Lemma 4.1, it is obvious to see that eνe_{\nu}, h¯ν\bar{h}_{\nu}, ξν\xi_{\nu} converge uniformly about ν\nu, and denote its limitation by ee_{\infty}, h¯\bar{h}_{\infty}, ξ\xi_{\infty}. By Lemma 3.4, we have

ω(ξ1)+p010(ξ1)=ω(ξ0),\displaystyle\omega\left(\xi_{1}\right)+p_{01}^{0}\left(\xi_{1}\right)=\omega\left(\xi_{0}\right),
ω(ξ2)+p010(ξ2)+p011(ξ2)=ω(ξ0),\displaystyle\omega\left(\xi_{2}\right)+p_{01}^{0}\left(\xi_{2}\right)+p_{01}^{1}\left(\xi_{2}\right)=\omega\left(\xi_{0}\right),
\displaystyle~{}~{}~{}\vdots
ω(ξν)+p010(ξν)++p01ν1(ξν)=ω(ξ0).\displaystyle\omega\left(\xi_{\nu}\right)+p_{01}^{0}\left(\xi_{\nu}\right)+\cdots+p_{01}^{\nu-1}\left(\xi_{\nu}\right)=\omega\left(\xi_{0}\right). (4.75)

Taking limits at both sides of (4.75), we get

ω(ξ)+j=0p01j(ξ)=ω(ξ0).\omega\left(\xi_{\infty}\right)+\sum_{j=0}^{\infty}p_{01}^{j}\left(\xi_{\infty}\right)=\omega(\xi_{0}).

Then, on D(s02)D(\frac{s_{0}}{2}), NνN_{\nu} converge uniformly to

N\displaystyle N_{\infty} =e+ω(ξ0),y+h¯(y,ξ).¯\displaystyle=e_{\infty}+\left\langle\omega(\xi_{0}),y\right\rangle+\bar{h}_{\infty}\left(y,\xi_{\infty}\right).\underline{}

Hence, on D(s02,r02)D\left(\frac{s_{0}}{2},\frac{r_{0}}{2}\right),

Pν=H0ΨνNν\displaystyle P_{\nu}=H_{0}\circ\Psi^{\nu}-N_{\nu}

converge uniformly to

P=H0ΨN.P_{\infty}=H_{0}\circ\Psi^{\infty}-N_{\infty}.

Since

|Pν|Dνγ0n+m+2sνmμν,\displaystyle\left|\left\|P_{\nu}\right\|\right|_{D_{\nu}}\leq\gamma_{0}^{n+m+2}s_{\nu}^{m}\mu_{\nu},

by (4.72), we have that it converges to 0 as ν\nu\rightarrow\infty. So, on D(0,r02)D(0,\frac{r_{0}}{2}),

JP=0.\displaystyle J\nabla P_{\infty}=0.

Thus, for the given ξ0O\xi_{0}\in O, the Hamiltonian

H=N+P\displaystyle H_{\infty}=N_{\infty}+P_{\infty}

admits an analytic, quasi-periodic, invariant nn-torus 𝕋n×{0}\mathbb{T}^{n}\times\{0\} with the Diophantine frequency ω(ξ0)\omega(\xi_{0}), which is the corresponding unperturbed toral frequency.

5 Proof of Theorems 2 and 4

First, we briefly give the proof framework of Theorem 2 because it can follow the KAM step in Section 3, where we mainly point out the two major differences from the proof of Theorem 1. The first one is that the homotopy invariance and excision of topological degree are used to keep the frequency unchanged in the iteration process not by picking parameters because we consider a Hamiltonian not a family of Hamiltonian. The other one is that the transformation defines on a smaller domain because we see the action-variable as parameter and the translation of parameter is equivalent to the action-variable’s.

5.1 Proof of Theorem 2

In this section, we will describe the translation of action variable and state how the frequency can be preserved in the iterative process, which are different from subsection 3.2.5 and subsection 3.2.6.

Let ξ0G\xi_{0}\in{G} be fixed as statement (A0)\rm(A0). The Taylor expansion of Hamiltonian (1.1) about ξ0\xi_{0} reads

H(y,ξ0,x)=e0+ω0(ξ0),yξ0+h¯(yξ0)+εP(y,x,ε),H(y,\xi_{0},x)=e_{0}+\left\langle\omega_{0}(\xi_{0}),y-\xi_{0}\right\rangle+\bar{h}(y-\xi_{0})+\varepsilon P(y,x,\varepsilon),

where e0=h(ξ0)e_{0}=h(\xi_{0}), ω0(ξ0)=h(ξ0)\omega_{0}(\xi_{0})=\nabla h(\xi_{0}), h¯(yξ0)=O(|yξ0|2)\bar{h}(y-\xi_{0})=O(\left|y-\xi_{0}\right|^{2}). Using the transformation (yξ0)y(y-\xi_{0})\rightarrow y in the above, we have

H(y,ξ0,x)=e0+ω0(ξ0),y+h¯(y,ξ0)+εP(y,x,ξ0,ε),H(y,\xi_{0},x)=e_{0}+\langle\omega_{0}(\xi_{0}),y\rangle+\bar{h}(y,\xi_{0})+\varepsilon P(y,x,\xi_{0},\varepsilon), (5.76)

where (y,x)(y,x) lies in a complex neighborhood D(s,r)D(s,r). Denote

H0\displaystyle H_{0} =:H(y,x,ξ0)=N0+P0,\displaystyle=:H(y,x,\xi_{0})=N_{0}+P_{0},
N0\displaystyle N_{0} =:e0+ω0(ξ0),y+h¯0,\displaystyle=:e_{0}+\langle\omega_{0}(\xi_{0}),y\rangle+\bar{h}_{0},
P0\displaystyle P_{0} =:εP(y,x,ξ0,ε).\displaystyle=:\varepsilon P(y,x,\xi_{0},\varepsilon).

Now, suppose that at ν\nu-th step, we have arrived at the following real analytic Hamiltonian:

H\displaystyle H =N+P,\displaystyle=N+P,~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{} (5.77)
N\displaystyle N =e+ω(ξ),y+h¯(y,ξ).\displaystyle=e+\langle\omega(\xi),y\rangle+\bar{h}(y,\xi).

Next, we will construct a translation so as to keep the frequency unchanged. Consider the translation

ϕ:xx,yy+ξ+ξ,\phi:x\rightarrow x,~{}~{}~{}~{}~{}y\rightarrow y+\xi_{+}-\xi,

where ξ+\xi_{+} is to be determined. Let

Φ+=ϕF1ϕ.\Phi_{+}=\phi_{F}^{1}\circ\phi.

Then

HΦ+\displaystyle H\circ\Phi_{+} =N++P+,\displaystyle=N_{+}+P_{+},
N+\displaystyle N_{+} =N¯+ϕ=e++ω+(ξ+),y+h¯+(y),\displaystyle=\bar{N}_{+}\circ\phi=e_{+}+\langle\omega_{+}(\xi_{+}),y\rangle+\bar{h}_{+}(y), (5.78)
P+\displaystyle P_{+} =P¯+ϕ,\displaystyle=\bar{P}_{+}\circ\phi, (5.79)

where

e+\displaystyle e_{+} =e+ω(ξ),ξ+ξ+h¯(ξ+ξ)+[R](ξ+ξ),\displaystyle=e+\langle\omega(\xi),\xi_{+}-\xi\rangle+\bar{h}(\xi_{+}-\xi)+[R](\xi_{+}-\xi), (5.80)
ω+\displaystyle\omega_{+} =ω(ξ)+h¯(ξ+ξ)+[R](ξ+ξ),\displaystyle=\omega(\xi)+\nabla\bar{h}(\xi_{+}-\xi)+\nabla[R](\xi_{+}-\xi), (5.81)
h¯+\displaystyle\bar{h}_{+} =h¯(y+ξ+ξ)h¯(ξ+ξ)h¯(ξ+ξ),y+[R](y+ξ+ξ)\displaystyle=\bar{h}(y+\xi_{+}-\xi)-\bar{h}(\xi_{+}-\xi)-\left\langle\nabla\bar{h}(\xi_{+}-\xi),y\right\rangle+[R](y+\xi_{+}-\xi)
[R](ξ+ξ)[R](ξ+ξ),y.\displaystyle~{}~{}~{}~{}-[R](\xi_{+}-\xi)-\left\langle\nabla[R](\xi_{+}-\xi),y\right\rangle. (5.82)

As in subsection 3.2.6, we will show that the frequency can be preserved in the iteration process. The following lemma is crucial to our arguments.

Lemma 5.1.

There exists ξ+Bsμ1/L(ξ)\xi_{+}\in B_{s\mu^{{1}/{L}}}(\xi) such that

ω+(ξ+)=ω(ξ)==ω0(ξ0).\displaystyle\omega_{+}(\xi_{+})=\omega(\xi)=\cdots=\omega_{0}(\xi_{0}). (5.83)
Proof.

The proof will be completed by an induction on ν\nu. We begin with the case ν=0\nu=0. It is obvious that ω0(ξ0)=ω0(ξ0)\omega_{0}(\xi_{0})=\omega_{0}(\xi_{0}). Now suppose that for some ν0\nu\geq 0 we have got

ωi(ξi)\displaystyle\omega_{i}(\xi_{i}) =ωi1(ξi1),ξiBsi1μi11/L(ξi1),\displaystyle=\omega_{i-1}(\xi_{i-1}),~{}~{}~{}~{}~{}\xi_{i}\in B_{s_{i-1}\mu_{i-1}^{{1}/{L}}}\left(\xi_{i-1}\right),

where i=1,2,,ν.i=1,2,\cdots,\nu. Then, we need to find ξ+\xi_{+} near ξ\xi such that ω+(ξ+)=ω(ξ)\omega_{+}(\xi_{+})=\omega(\xi). In view of (5.81), we observe that

|ω+(y)ω(y)|=O(μ).\displaystyle\left|\omega_{+}(y)-\omega(y)\right|=O(\mu). (5.84)

We split

ω+(y)ω(ξ)\displaystyle\omega_{+}(y)-\omega(\xi) =(ω(y)ω(ξ))+(ω+(y)ω(y)).\displaystyle=(\omega(y)-\omega(\xi))+(\omega_{+}(y)-\omega(y)). (5.85)

Consider homotopy Ht(y):[0,1]×OnH_{t}(y):[0,1]\times O\rightarrow\mathbb{R}^{n},

Ht(y)\displaystyle H_{t}(y) =:(ω(y)ω(ξ))+t(ω+(y)ω(y)).\displaystyle=:\left(\omega(y)-\omega(\xi)\right)+t\left(\omega_{+}(y)-\omega(y)\right).

For any yOy\in\partial O, t[0,1]t\in[0,1], by (A1)\rm(A1), we have that

|Ht(y)|\displaystyle\left|H_{t}(y)\right| |ω(y)ω(ξ)||ω+(y)ω(y)||ω0(y)ω0(ξ0)|i=0ν|ωi+1(y)ωi(y)|\displaystyle\geq|\omega(y)-\omega(\xi)|-|\omega_{+}(y)-\omega(y)|\geq|\omega_{0}(y)-\omega_{0}(\xi_{0})|-\sum_{i=0}^{\nu}\left|\omega_{i+1}(y)-\omega_{i}(y)\right|
σ|yξ0|Li=0νγ0n+m+2sim1μi>σδL2,\displaystyle\geq\sigma|y-\xi_{0}|^{L}-\sum_{i=0}^{\nu}\gamma_{0}^{n+m+2}s_{i}^{m-1}\mu_{i}>\frac{\sigma\delta^{L}}{2},

where δ:=min{|yξ0|,yO}\delta:=\min\{|y-\xi_{0}|,\forall y\in\partial O\}.

So, it follows from homotopy invariance and (A0)\rm(A0) that

deg(H1(),Oo,0)=deg(H0(),Oo,0)0.\displaystyle\deg\left(H_{1}(\cdot),O^{o},0\right)=\deg\left(H_{0}(\cdot),O^{o},0\right)\neq 0.

We note by (A1)\rm(A1), (5.84) and (5.85) that for any yO\Bsμ1/L(ξ)y\in O\backslash B_{s\mu^{{1}/{L}}}(\xi),

|H1(y)|\displaystyle\left|H_{1}(y)\right| =|ω+(y)ω(ξ)||yξ|Lc1γ0n+m+2sm1μ\displaystyle=|\omega_{+}(y)-\omega(\xi)|\geq|y-\xi|^{L}-c_{1}\gamma_{0}^{n+m+2}s^{m-1}\mu
sLμc1γ0n+m+2sm1μsLμ2.\displaystyle\geq s^{L}\mu-c_{1}\gamma_{0}^{n+m+2}s^{m-1}\mu\geq\frac{s^{L}\mu}{2}.

Hence, by excision, we have that

deg(H1(),Bsμ1/L(ξ),0)=deg(H1(),Oo,0)0,\displaystyle\deg\left(H_{1}(\cdot),B_{s\mu^{{1}/{L}}}(\xi),0\right)=\deg\left(H_{1}(\cdot),O^{o},0\right)\neq 0,

i.e., there exists at least a ξ+Bsμ1/L(ξ)\xi_{+}\in B_{s\mu^{{1}/{L}}}(\xi), such that H1(ξ+)=0,H_{1}(\xi_{+})=0, i.e.,

ω+(ξ+)=ω(ξ),\omega_{+}(\xi_{+})=\omega(\xi),

which implies (5.83).

The proof is complete. ∎

In the following, we prove

ϕ:D18αD14α,\displaystyle\phi:D_{\frac{1}{8}\alpha}\rightarrow D_{\frac{1}{4}\alpha}, (5.86)

which is different from (3.54) in Lemma 3.7. Recall that m>L+1m>L+1 and α=μ1m+1\alpha=\mu^{\frac{1}{m+1}}, we have

csμ1L<18αs.\displaystyle cs\mu^{\frac{1}{L}}<\frac{1}{8}\alpha s. (5.87)

For (y,x)D18α\forall(y,x)\in D_{\frac{1}{8}\alpha}, we note by ξ+Bsμ1/L(ξ)\xi_{+}\in B_{s\mu^{{1}/{L}}}(\xi) in Lemma 5.1 and (5.87) that

|y+ξ+ξ|<|y|+|ξ+ξ|<18αs+csμ1L<14αs,\left|y+\xi_{+}-\xi\right|<|y|+|\xi_{+}-\xi|<\frac{1}{8}\alpha s+cs\mu^{\frac{1}{L}}<\frac{1}{4}\alpha s,

which implies (5.86).

Next, we prove Theorem 4 by a direct method.

5.2 Proof of Theorem 4

(1) The unperturbed motion of (1.1) is described by the equation

{y˙=0,x˙=h(y).\left\{\begin{array}[]{ll}\dot{y}=0,\\ \dot{x}=h^{\prime}(y).\end{array}\right.

The flow is x=h(y)t+x0,yGx=h^{\prime}(y)t+x_{0},y\in G, where x0x_{0} is an initial condition. Notice that

h′′(0)=0,\displaystyle h^{\prime\prime}(0)=0,

i.e., h(y)h(y) is degenerate at ξ0=0\xi_{0}=0. Obviously, by simple calculation, we get

deg(h(y)h(0),Bδ(0),0)=0,\displaystyle\deg\left(h^{\prime}(y)-h^{\prime}(0),B_{\delta}(0),0\right)=0,

i.e., (A0)\rm(A0) fails, then Theorem 3 is not applicable.

Note that the perturbed motion equation is

{y˙=0,x˙=h(y)+εP(y).\left\{\begin{array}[]{ll}\dot{y}=0,\\ \dot{x}=h^{\prime}(y)+\varepsilon P^{\prime}(y).\end{array}\right.

The flow is x=(h(y)+εP(y))t+x1,yGx=\left(h^{\prime}(y)+\varepsilon P^{\prime}(y)\right)t+x_{1},~{}~{}~{}y\in G, where x1x_{1} is an initial condition. To ensure the frequency is equal to h(0)h^{\prime}(0), we need to find a solution of the following equation in GG:

h(y)+εP(y)=h(0),\displaystyle h^{\prime}(y)+\varepsilon P^{\prime}(y)=h^{\prime}(0),

i.e.,

g(y)=εP(y).\displaystyle g^{\prime}(y)=-\varepsilon P^{\prime}(y). (5.88)

Notice that the Taylor expansion of g(y)g^{\prime}(y) at ξ0=0\xi_{0}=0 is

g(y)=g(0)+g′′(0)y++g2+1(0)y2+o(y2),g^{\prime}(y)=g^{\prime}(0)+g^{\prime\prime}(0)y+\cdots+g^{2\ell+1}(0)y^{2\ell}+o(y^{2\ell}),

then the equation (5.88) is equivalent to

g2+1(0)y2+o(y2)=εP(y),g^{2\ell+1}(0)y^{2\ell}+o(y^{2\ell})=-\varepsilon P^{\prime}(y),

which is solvable provided that εP(y)sign(g2+1(0))<0\varepsilon P^{\prime}(y)\,sign\left(g^{2\ell+1}(0)\right)<0. So the perturbed system admits at least two invariant tori with frequency ω=h(0)\omega=h^{\prime}(0) for the small enough perturbation satisfying εP(y)sign(g2+1(0))<0\varepsilon P^{\prime}(y)\,sign\left(g^{2\ell+1}(0)\right)<0. Conversely, if εP(y)sign(g2+1(0))>0\varepsilon P^{\prime}(y)\,sign\left(g^{2\ell+1}(0)\right)>0, the unperturbed invariant torus with frequency ω=h(0)\omega=h^{\prime}(0) will be destroyed.

(2) Note that h(y)h(y) is degenerate in ξ0=0\xi_{0}=0. Obviously, by simple calculation, we get

deg(h(y)h(0),Bδ(0),0)0.\displaystyle\deg\left(h^{\prime}(y)-h^{\prime}(0),B_{\delta}(0),0\right)\neq 0.

Then, by Theorem 3, the above persistence result hold. In addition, we can also directly prove this result. Similarly, we need to solve the following equation in GG:

h(y)+εP(y)=h(0),\displaystyle h^{\prime}(y)+\varepsilon P^{\prime}(y)=h^{\prime}(0),

i.e.,

g(y)=εP(y).\displaystyle g^{\prime}(y)=-\varepsilon P^{\prime}(y). (5.89)

Notice that the Taylor expansion of g(y)g^{\prime}(y) at ξ0=0\xi_{0}=0 is

g(y)=g(0)+g′′(0)y++g2+2(0)y2+1+o(y2+1),g^{\prime}(y)=g^{\prime}(0)+g^{\prime\prime}(0)y+\cdots+g^{2\ell+2}(0)y^{2\ell+1}+o\left(y^{2\ell+1}\right),

then the equation (5.89) is equivalent to

g2+2(0)y2+1+o(y2+1)=εP(y),g^{2\ell+2}\left(0\right)y^{2\ell+1}+o\left(y^{2\ell+1}\right)=-\varepsilon P^{\prime}(y),

whose solution always exists in GG for any small enough perturbation. Hence, the perturbed system admits an invariant torus with frequency ω=h(0)\omega=h^{\prime}(0) for any small enough perturbation.

6 Appendix A. Proof of Proposition 1

Proof.

Obviously, for ξ(1,1)×(1,1)\forall\xi\in(-1,1)\times(-1,1),

(ωω¯)(ξ)=(ωω¯)(ξ),(\omega-\bar{\omega})(-\xi)=-(\omega-\bar{\omega})(\xi),

and for ξ(1,1)×(1,1),\forall\xi\in\partial(-1,1)\times(-1,1),

(ωω¯)(ξ)0.(\omega-\bar{\omega})(\xi)\neq 0.

Using Borsuk’s theorem in [27], we have

deg(ω()ω¯,(1,1)×(1,1),0)0,\deg\left(\omega(\cdot)-\bar{\omega},(-1,1)\times(-1,1),0\right)\neq 0,

i.e.,

deg(ω(),(1,1)×(1,1),ω¯)0,\deg\left(\omega(\cdot),(-1,1)\times(-1,1),\bar{\omega}\right)\neq 0,

i.e., (A0)\rm(A0) holds. For ξ,ξ[12,12]\xi,\xi_{*}\in\left[-\frac{1}{2},\frac{1}{2}\right], and ξξ\xi\neq\xi_{*}, we have

ω(ξ)ω(ξ)=0,\omega(\xi)-\omega(\xi_{*})=0,

but

|ξξ|L>0,L>0,\left|\xi-\xi_{*}\right|^{L}>0,~{}~{}\forall L>0,

which shows that (A1)\rm(A1) fails. Note that the flow of unperturbed motion equation is

x=ω(ξ)t+x0,ξ(1,1)×(1,1),x=\omega(\xi)t+x_{0},~{}~{}~{}\xi\in(-1,1)\times(-1,1),

where x0x_{0} is an initial condition, and the flow of perturbed motion equation is

x=(ω(ξ)+(0,P0(ε)))t+x0,ξ(1,1)×(1,1).x=\left(\omega(\xi)+\left(0,P_{0}\left(\varepsilon\right)\right)^{\top}\right)t+x_{0},~{}~{}~{}\xi\in(-1,1)\times(-1,1).

In order to keep the frequency ω(0)=ω¯\omega(0)=\bar{\omega} unchanged, we have to solve the following equation

ω(ξ)+(0,P0(ε))=ω¯,\omega(\xi)+\left(0,P_{0}\left(\varepsilon\right)\right)^{\top}=\bar{\omega},

i.e.,

ω(ξ)ω¯(ξ)=(ξ1,ω2ω¯2)=(0,P0(ε)),\omega\left(\xi\right)-\bar{\omega}\left(\xi\right)=\left(\xi_{1},\omega_{2}-\bar{\omega}_{2}\right)^{\top}=-\left(0,P_{0}\left(\varepsilon\right)\right)^{\top},

which implies that the second component ξ2\xi_{2} of solution ξ\xi is discontinuous and alternately appears on (1,12)\left(-1,-\frac{1}{2}\right) and (12,1)\left(\frac{1}{2},1\right) as ε0+\varepsilon\rightarrow 0_{+}. So, this example shows that condition (A1)\rm(A1) is necessary no matter how smooth the frequency mapping ω(ξ)\omega(\xi) is. ∎

7 Appendix B. Proof of Theorem 3

Proof.

Notice that

h(y)h(0)=y|y|2l.\displaystyle\nabla h(y)-\nabla h(0)=y|y|^{2l}.

For 0<δ<10<\delta<1, Bδ(0)B_{\delta}(0) denotes the open ball centered at the origin with radius δ\delta. We have that h(y)h(0)\nabla h(y)-\nabla h(0) is odd and unequal to zero on Bδ(0)\partial B_{\delta}(0), i.e.,

h(y)h(0)=y|y|2l=(h(y)h(0)),h(y)h(0)0,yBδ(0).\nabla h(-y)-\nabla h(0)=-y|y|^{2l}=-(\nabla h(y)-\nabla h(0)),~{}~{}\nabla h(y)-\nabla h(0)\neq 0,~{}~{}\forall y\in\partial B_{\delta}(0).

It follows from Borsuk’s theorem in [27] that,

deg(h(y)h(0),Bδ(0),0)0.\displaystyle\deg(\nabla h(y)-\nabla h(0),B_{\delta}(0),0)\neq 0.

Obviously, there exist σ=minyBδ(0){(y+ϱ)|y+ϱ|2y|y|2}2ϱ2+1\sigma=\frac{\min_{y\in B_{\delta}(0)}\{{(y+\varrho)|y+\varrho|^{2\ell}-y|y|^{2\ell}}\}}{2\varrho^{2\ell+1}} and L=2l+1L=2l+1 such that

|h(y)h(y)|σ|yy|L,yBδ(0),yBδ(0)Bϱ(y),\displaystyle\left|\nabla h(y)-\nabla h(y_{*})\right|\geq\sigma\left|y-y_{*}\right|^{L},~{}~{}y_{*}\in B_{\delta}(0),y\in B_{\delta}(0)\setminus B_{\varrho}(y_{*}),

where ϱ>0\varrho>0, Bϱ(y)Bδ(0)B_{\varrho}(y_{*})\subset B_{\delta}(0). So, by Theorem 2, the perturbed system admits an invariant torus with frequency ω\omega for any small enough perturbation. ∎

8 Appendix C. Proof of Proposition 2

Proof.

Let εP=εy,ε>0\varepsilon P=\varepsilon y,\varepsilon>0. Notice that for yG1y\in G\subset\mathbb{R}^{1},

h(y)=ω+y2,h(0)=ω,h′′(y)|y=0=0,h^{\prime}(y)=\omega+y^{2\ell},~{}~{}h^{\prime}(0)=\omega,~{}~{}h^{\prime\prime}(y)\big{|}_{y=0}=0,

which implies that the Hamiltonian HH is degenerate at y=0y=0. By the definition of degree, we have for 0<δ<10<\delta<1

deg(h(y)h(0),Bδ(0),0)=0,\displaystyle\deg\left(\nabla h(y)-\nabla h(0),B_{\delta}(0),0\right)=0,

i.e., (A0)\rm(A0) fails. Then, Theorem 2 cannot be used to prove the persistence result of keeping frequency unchanged.

Note that the flow of unperturbed motion equation at y=0y=0 is

x=ωt+x0,x=\omega t+x_{0},

where x0x_{0} is an initial condition, and the flow of perturbed motion equation is

x=(ω+y2+ε)t+x0,yG.x=\left(\omega+y^{2\ell}+\varepsilon\right)t+x_{0},~{}~{}~{}y\in G.

In order to preserve frequency ω\omega, we need to solve y2+ε=0y^{2\ell}+\varepsilon=0 in GG, which has no real solution in GG. Hence, the persistence result of keeping frequency unchanged fails.

Acknowledgments

The second author (Li Yong) is supported by National Basic Research Program of China (Grant number [2013CB8-34100]), National Natural Science Foundation of China (Grant numbers [11571065], [11171132], and [12071175]), and Natural Science Foundation of Jilin Province (Grant number [20200201253JC]).

References

  • [1] Arnold, V.I.: Proof of a theorem of A. N. Kolmogorov on the preservation of conditionally periodic motions under a small perturbation of the Hamiltonian. Uspehi Mat. Nauk.18,5,13-40(1963). MR0163025
  • [2] Benettin, G., Galgant, L., Giorgilli, A., Strelcyn, J.: A proof of Kolmogorov’s theorem on invariant tori using canonical transformations defined by the Lie method. Nuovo Cimento.79B,2,201-223(1984). MR0743977
  • [3] Biasco, L., Chierchia, L., Treschev, D.: Stability of nearly integrable, degenerate Hamiltonian systems with two degrees of freedom. J. Nonlinear Sci.16,1,79-107(2006). MR2202903
  • [4] Bounemoura, A.: Positive measure of KAM tori for finitely differentiable Hamiltonians. J. Éc. polytech. Math.7,1113-1132(2020). MR4167789
  • [5] Bricmont, J., Gawedzki, K., Kupiainen, A.: KAM theorem and quantum field theory. Commun. Math. Phys.201,699-727(1999). MR1685894
  • [6] Brjuno, A.D.: Nondegeneracy conditions in the Kolmogorov theorem. Soviet Math. Dokl.45,1,221-225(1992). MR1171798
  • [7] Cheng, C.Q., Sun, Y.S.: Existence of KAM tori in degenerate Hamiltonian systems. J. Differential Equations.114,1,288-335(1994). MR1302146
  • [8] Chierchia, L., Falcolini, C.: A direct proof of a theorem by Kolmogorov in Hamiltonian systems. Ann. Scuola Norm. Sup. Pisa Cl. Sci.21,4,541-593(1994). MR1318772
  • [9] Chierchia, L.: Periodic solutions of the planetary NN-body problem. XVIIth International Congress on Mathematical Physics.269-280(2014). MR3204477
  • [10] Chow, S.N., Li, Y., Yi, Y.F.: Persistence of invariant tori on submanifolds in Hamiltonian systems. J. Nonlinear Sci.12,585-617(2002). MR1938331
  • [11] Cong, F.Z., Küpper, T., Li, Y., You, J.G.: KAM-type theorem on resonant surfaces for nearly integrable Hamiltonian systems. J. Nonlinear Sci.10,1,49-68(2000). MR1730569
  • [12] de la Llave, R., González, A., Jorba, À., Villanueva, J.: KAM theory without action-angle variables. Nonlinearity.18,2,855-895(2005). MR2122688
  • [13] Eliasson, L.H.: Absolutely convergent series expansions for quasi-periodic motions. Math. Phys. Elect. J.2,1-33(1996). MR1399458
  • [14] Féjoz, J.: Démonstration du ‘théorème d’Arnold’ sur la stabilité du système planétaire (d’après Herman). Ergodic Theory Dynam. Systems.24,5,1521-1582(2004). MR2104595
  • [15] Gallavotti, G.: Twistless KAM tori. Commun. Math Phys.164,145-156(1994). MR1288156
  • [16] Gallavotti, G., Gentile, G., Mastropietro, V.: Field theory and KAM tori. Math. Phys. Elect. J.1,1-13(1995). MR1359460
  • [17] Han, Y.C., Li, Y., Yi, Y.F.: Invariant tori in Hamiltonian systems with high order proper degeneracy. Ann. Henri Poincaré.10,1419-1436(2010). MR2639543
  • [18] Heinz, H.: Non-degeneracy conditions in KAM theory. Indag. Math. (N.S.).22,3-4,241-256(2011). MR2853608
  • [19] Herman, M.R.: Sur les courbes invariantes par les difféomorphismes de l’anneau. Vol. 1. (French) [On the curves invariant under diffeomorphisms of the annulus. Vol. 1] With an appendix by Albert Fathi. With an English summary. Astérisque, 103-104. Société Mathématique de France, Paris, 1983, pp. i+221. MR0728564
  • [20] Kolmogorov, A.N.: On conservation of conditionally periodic motions for a small change in Hamilton’s function. Dokl. Akad. Nauk SSSR.98,527-530(1954). MR0068687
  • [21] Koudjinan, C.E.: A KAM theorem for finitely differentiable Hamiltonian systems. J. Differential Equations.269,6,4720-4750(2020). MR4104457
  • [22] Li, Y., Yi, Y.F.: Persistence of invariant tori in generalized Hamiltonian systems. Ergodic Theory Dynam. Systems.22,4,1233-1261(2002). MR1926285
  • [23] Li, Y., Yi, Y.F.: A quasi-periodic Poincaré’s theorem. Math. Ann.326,649-690(2003). MR2003447
  • [24] Meyer, K.R.: Periodic solutions of the NN-body problem. Springer-Verlag, Berlin,1999. MR1736548
  • [25] Moser, J.: Convergent series expansions for quasi-periodic motions. Math. Ann.169,136-176(1967). MR0208078
  • [26] Moser, J.: On invariant curves of area-preserving mapping of an annulus. Nachr. Akad. Wiss. Göttingen.II,1-20(1962). MR0147741
  • [27] Motreanu, D., Motreanu, V.V., Papageorgiou, N.: Topological and variational methods with applications to nonlinear boundary value problems. Springer Science+Business Media, LLC,2014. MR3136201
  • [28] Pöschel, J.: On elliptic lower-dimensional tori in Hamiltonian systems. Math. Z.202,4,559-608(1989). MR1022821
  • [29] Pöschel, J.: A KAM theorem for some nonlinear partial differential equations. Ann. Sc. Norm. Super. Pisa Cl. Sci.23,1,119-148(1996). MR1401420
  • [30] Pöschel, J.: A lecture on the classical KAM theorem. Proc. Symp. Pure Math.69,707-732(2001). MR1858551
  • [31] Qian, W.C., Li, Y., Yang, X.: Multiscale KAM theorem for Hamiltonian systems. J. Differential Equations.266,1,70-86(2019). MR3870557
  • [32] Qian, W.C., Li, Y., Yang, X.: Melnikov’s conditions in matrices. J. Dynam. Differential Equations.32,4,1779-1795(2020). MR4171876
  • [33] Rüssmann H., Kleine Nenner I.: Über invariante Kurven differenzierbarer Abbildungen eines Kreisringes. Nachr. Akad. Wiss. Göttingen Math.-Phys. Kl. II, 67-105, 1970. MR0273156
  • [34] Rüssmann, H.: Invariant tori in non-degenerate nearly integrable Hamiltonian systems. Regul. Chaotic Dyn.6,2,119-204(2001). MR1843664
  • [35] Salamon, D.A.: The Kolmogorov-Arnold-Moser theorem. Math. Phys. Electron. J.10,1-37(2004). MR2111297
  • [36] Sevryuk, M.B.: KAM-stable Hamiltonians. J. Dynam. Control Systems.1,3,351-366(1995). MR1354540
  • [37] Sevryuk, M.B.: Partial perservation of frequencies in KAM theory. Nonlinearity.19,1099-1140(2006). MR2221801
  • [38] Wayne, C.E.: An introduction to KAM theory. Lectures in Appl. Math.31(1996). MR1363023
  • [39] Xu, J.X., You, J.G., Qiu, Q.J.: Invariant tori for nearly integrable Hamiltonian systems with degeneracy. Math. Z.226,3,375-387(1997). MR1483538