1 Introduction

In this paper we study response solutions for almost-periodically forced quasilinear PDEs close to an elliptic fixed point.

The problem of response solutions for PDEs has been widely studied in many contexts, starting from the papers [24, 25], where the Author considers a periodically forced PDE with dissipation. In the presence of dissipation, of course there is no small divisors problem. However as soon as the dissipation is removed, small divisors appear even in the easiest possible case of a periodic forcing when the spacial variable is one dimensional.

The first results of this type in absence of dissipation were obtained by means of a KAM approach [16,17,18,19, 22, 28]. However, a more functional approach, via a combination of a Ljapunov-Schmidt reduction and a Newton scheme, in the spirit of [24, 25], was proposed by Craig–Wayne [14], and then generalized in many ways by Bourgain; see for instance [5,6,7] to mention a few. All the results mentioned above concern semi-linear PDEs and the forcing is quasi-periodic.

In more recent times, the Craig–Wayne–Bourgain approach has been fruitfully used and generalized in order to cover quasi-linear and fully nonlinear PDEs, again in the quasi-periodic case; see for instance [1, 2, 12, 15].

Regarding the almost-periodic case, most of the classical results are obtained via a KAM-like approach; see for instance [9, 10, 23]. A notable exception is [8], where the Craig–Wayne–Bourgain method is used. More recently there have been results such as [20, 26, 27], which use a KAM approach. We mention also [3, 4, 11, 29] which however are tailored for an autonomous PDE.

All the aforementioned results, concern semi-linear PDEs, with no derivative in the nonlinearity. Moreover they require a very strong analyticity condition on the forcing term. Indeed the difficulty of proving the existence of almost-periodic response solution is strongly related to the regularity of the forcing, since one can see an almost periodic function as the limit of quasi-periodic ones with an increasing number of frequencies. If such limit is reached sufficiently fast, the most direct strategy would be to iteratively find approximate quasi-periodic response solutions and then take the limit. This is the overall strategy of [23] and [20, 26, 27]. However this procedure works if one considers a sufficiently regular forcing term and a bounded nonlinearity, but becomes very delicate in the case of unbounded nonlinearities.

In the present paper we study the existence of almost-periodic response solutions, for a quasi-linear PDE on \(\mathbb {T}\). To the best of our knowledge this is the first result of this type.

Specifically we consider a quasi-linear Airy equation

$$\begin{aligned} \partial _t u + \partial _{xxx} u + Q(u, u_x, u_{xx}, u_{xxx}) + {{\mathfrak {f}}}( t, x) =0,\qquad x\in \mathbb {T}:= (\mathbb {R}/(2 \pi \mathbb {Z})) \end{aligned}$$
(1.1)

where Q is a Hamiltonian, quadratic nonlinearity and \({{\mathfrak {f}}}\) is an analytic forcing term with zero average w.r.t. x. We assume \({{\mathfrak {f}}}\) to be “almost-periodic” with frequency \(\omega \in \ell ^{\infty }\), in the sense of Definition 1.1.

We mention that in the context of reducibility of linear PDEs a problem of this kind has been solved in [21]. Our aim is to provide a link between the linear techniques of [21] and the nonlinear Craig–Wayne–Bourgain method. Note that such a link is nontrivial, and requires a delicate handling; see below.

The overall setting we use is the one of [1]. However their strategy is taylored for Sobolev regularity; the quasi-periodic analytic case has been covered in [13]. Unfortunately the ideas of [13] cannot be directly applied in the almost-periodic case. Roughly, it is well known that the regularity and the small-divisor problem conflict. Thus, in the almost-periodic case one expect this issue to be even more dramatic. Specifically, we were not able to define a “Sobolev” norm for almost-periodic functions, satisfying the interpolation estimates needed in the Nash-Moser scheme; this is why we cannot use the theorem of [13].

Let us now present our main result in a more detailed way.

First of all we note that (1.1) is an Hamiltonian PDE whose Hamiltonian is given by

$$\begin{aligned} H(u) := \frac{1}{2}\int _\mathbb {T}u_x^2 d x - \frac{1}{6} \int _\mathbb {T}G(u, u_x)\, d x - \int _{\mathbb {T}}F( t, x) u d x,\qquad {{\mathfrak {f}}}(t,x) = \partial _x F(t,x) \end{aligned}$$
(1.2)

where \(G(u, u_x)\) is a cubic Hamiltonian density of the form

$$\begin{aligned} G(u, u_x) := {\mathtt {c}}_3 u_x^3 + {\mathtt {c}}_2 u u_x^2 + {\mathtt {c}}_1 u^2 u_x + {\mathtt {c}}_0 u^3, \quad {\mathtt {c}}_0, \ldots , {\mathtt {c}}_3 \in {\mathbb {R}}\, \end{aligned}$$
(1.3)

and the symplectic structure is given by \(J=\partial _x\). The Hamiltonian nonlinearity \(Q(u, \ldots , u_{xxx})\) is therefore given by

$$\begin{aligned} Q(u, u_x, u_{xx}, u_{xxx}) = \partial _{xx}(\partial _{u_x} G(u, u_x)) - \partial _x (\partial _u G(u, u_x)) \end{aligned}$$
(1.4)

and the Hamilton equations are

$$\begin{aligned} \partial _t u = \partial _x \nabla _{ u} H(u). \end{aligned}$$

We look for an almost-periodic solution to (1.1) with frequency \(\omega \) in the sense below.

For \(\eta >0\), define the set of infinite integer vectors with finite support as

$$\begin{aligned} {\mathbb {Z}}^\infty _* := \Big \{ \ell \in {\mathbb {Z}}^{\mathbb {N}}: |\ell |_\eta := \sum _{i\in {\mathbb {N}}} i ^\eta |\ell _i| < \infty \Big \}. \end{aligned}$$
(1.5)

Note that \(\ell _i \ne 0\) only for finitely many indices \(i \in {\mathbb {N}}\). In particular \(\mathbb {Z}^\infty _*\) does not depend on \(\eta \).

Definition 1.1

Given \(\omega \in [1,2]^{\mathbb {N}}\) with rationally independent componentsFootnote 1 and a Banach space \((X,|\cdot |_X)\), we say that \(F(t):{\mathbb {R}}\rightarrow X\) is almost-periodic in time with frequency \(\omega \) and analytic in the strip \(\sigma >0\) if we may write it in totally convergent Fourier series

$$\begin{aligned} \begin{aligned} F(t)&= \sum _{ \ell \in {\mathbb {Z}}^\infty _*} F(\ell ) e^{{\mathrm{i}}\ell \cdot \omega t } \quad \text {such that} \quad F(\ell ) \in X,\;\forall \ell \in {\mathbb {Z}}^\infty _* \\&\quad \text {and} \quad |F|_\sigma := \sum _{ \ell \in {\mathbb {Z}}^\infty _*} | F(\ell ) |_X e^{\sigma |\ell |_\eta } <\infty . \end{aligned} \end{aligned}$$

We shall be particularly interested in almost-periodic functions where \(X={{\mathcal {H}}}_0({\mathbb {T}}_\sigma )\)

$$\begin{aligned} {{\mathcal {H}}}_0({\mathbb {T}}_\sigma ):= \Big \{ u= \sum _{j \in {\mathbb {Z}}\setminus \{0\}} u_j e^{{\mathrm{i}}j x},\; u_j={{\bar{u}}}_{-j} \in {\mathbb {C}}\,:\quad |u|_{{{\mathcal {H}}}({\mathbb {T}}_\sigma )}:= \sum _{j \in {\mathbb {Z}}\setminus \{ 0 \}}| u_j| e^{\sigma |j| } <\infty \Big \} \end{aligned}$$

is the space of analytic, real on real functions \({\mathbb {T}}_s\rightarrow {\mathbb {C}}\) with zero-average, where \({\mathbb {T}}_s := \{ \varphi \in {\mathbb {C}}: {\mathrm{Re}}(\varphi ) \in {\mathbb {T}}, \ |{\mathrm{Im}}(\varphi )| \le s \}\) is the thickened torus. We recall that a function \(u : {\mathbb {T}}_s \rightarrow {\mathbb {C}}\) is real on real if for any \(x \in {\mathbb {T}}\), \(u(x) \in {\mathbb {R}}\).

Of course we need some kind of Diophantine condition on \(\omega \). We give the following, taken from [9, 21].

Definition 1.2

Given \(\gamma \in (0, 1)\), we denote by \({\mathtt {D}_\gamma }\) the set of Diophantine frequencies

$$\begin{aligned} {\mathtt {D}_\gamma }:={\left\{ \omega \in [1,2]^{\mathbb {N}}\,:\; |\omega \cdot \ell |> \gamma \prod _{i\in {\mathbb {N}}}\frac{1}{(1+|\ell _i|^{2} {i}^{2})}, \quad \forall \ell \in {\mathbb {Z}}^\infty _*\setminus \{0\}\right\} }. \end{aligned}$$
(1.6)

We are now ready to state our main result.

Theorem 1.3

(Main Theorem) Fix \(\overline{\gamma }\). Assume that \({{\mathfrak {f}}}\) in (1.1) is almost-periodic in time and analytic in a strip S (both in time and space). Fix \(\overline{s}<S\). If \({{\mathfrak {f}}}\) has an appropriately small norm depending on \(S-\overline{s}\), namely

$$\begin{aligned} |{{\mathfrak {f}}}|_S := \sum _{ \ell \in {\mathbb {Z}}^\infty _*} | {{\mathfrak {f}}}(\ell ) |_{{{\mathcal {H}}}_0({\mathbb {T}}_S)} e^{S |\ell |_\eta } \le \epsilon (S-\overline{s})\ll 1, \qquad \epsilon (0)=0, \end{aligned}$$
(1.7)

then there is a Cantor-like set \({{\mathcal {O}}}^{(\infty )}\subseteq {\mathtt {D}_{\overline{\gamma }}}\) with positive Lebesgue measure, and for all \(\omega \in {{\mathcal {O}}}^{(\infty )}\) a solution to (1.1) which is almost-periodic in time with frequency \(\omega \) and analytic in a strip \(\overline{s}\) (both in time and space).

Remark 1.4

Of course the same result holds verbatim if we replace the quadratic polynomial Q by a polinomial of arbitrary degree. We could also assume that the coefficients \({\mathtt {c}}_j\) appearing in (1.4) depend on x and \(\omega t\). In that case Theorem 1.3 holds provided we further require a condition of the type \( \sup _{j} |\partial _x^2 {\mathtt {c}}_j|_S \le C\). Actually one could also take Q to be an analytic function with a zero of order two. However this leads to a number of long and non particularly enlightening calculations.

To prove Theorem 1.3 we proceed as follows. First of all we regard (1.1) as a functional Implicit Function Problem on some appropriate space of functions defined on an infinite dimensional torus; see Definition 2.1 below. Then in Sect. 3 we prove an iterative “Nash-Moser-KAM” scheme to produce the solution of such Implicit Function Problem. It is well known that an iterative rapidly converging scheme heavily relies on a careful control on the invertibility of the linearized operator at any approximate solution. Of course, in the case of a quasi-linear PDE this amounts to study an unbounded non-constant coefficients operator. To deal with this problem, at each step we introduce a change of variables \(T_n\) which diagonalizes the highest order terms of the linearized operator. An interesting feature is that \(T_n\) preserves the PDE structure. As in [13] and differently from the classical papers, at each step we apply the change of variables \(T_n\) to the whole nonlinear operator. This is not a merely technical issue. Indeed, the norms we use are strongly coordinate-depending, and the change of variable \(T_n\) that we need to apply are not close-to-identity, in the sense that \(T_n - \mathtt { Id}\) is not a bounded operator small in size.

In Sect. 4 we show how to construct the change of variables \(T_n\) satisfying the properties above. Then in order to prove the invertibility of the linearized operator after the change of variables \(T_n\) is applied, one needs to perform a reducibility scheme: this is done in Sect. 5. For a more detailed description of the technical aspects see Remark 3.2.

2 Functional Setting

As it is habitual in the theory of quasi-periodic functions we shall study almost periodic functions in the context of analytic functions on an infinite dimensional torus. To this purpose, for \(\eta ,s>0\), we define the thickened infinite dimensional torus \({\mathbb {T}}^\infty _s\) as

$$\begin{aligned} \varphi =(\varphi _i)_{i\in {\mathbb {N}}},\quad \varphi _i\in {\mathbb {C}}\,:\; {\mathrm{Re}}(\varphi _i)\in {\mathbb {T}}\;,\;|{\mathrm{Im}}(\varphi _i)|\le s \langle i \rangle ^\eta . \end{aligned}$$

Given a Banach space \((X, | \cdot |_X )\) we consider the space \({{\mathcal {F}}}\) of pointwise absolutely convergent formal Fourier series \({\mathbb {T}}^\infty _s \rightarrow X\)

$$\begin{aligned} u(\varphi ) = \sum _{\ell \in {\mathbb {Z}}^\infty _*} u(\ell ) e^{{\mathrm{i}}\ell \cdot \varphi },\quad u(\ell ) \in X \end{aligned}$$
(2.1)

and define the analytic functions as follows.

Definition 2.1

Given a Banach space \((X, | \cdot |_X )\) and \(s > 0\), we define the space of analytic functions \({\mathbb {T}}^\infty _s \rightarrow X\) as the subspace

$$\begin{aligned} {{\mathcal {H}}}( {\mathbb {T}}^\infty _s, X) := \Big \{ u(\varphi ) = \sum _{\ell \in {\mathbb {Z}}^\infty _*} u(\ell ) e^{{\mathrm{i}}\ell \cdot \varphi }\in {{\mathcal {F}}}\; \;:\quad | u |_{s} := \sum _{\ell \in {\mathbb {Z}}^\infty _*} e^{s |\ell |_\eta } | u(\ell )|_X < \infty \Big \}. \end{aligned}$$

We denote by \({{\mathcal {H}}}_s\) the subspace of \({{\mathcal {H}}}({\mathbb {T}}^\infty _s, {{\mathcal {H}}}_0({\mathbb {T}}_s))\) of the functions which are real on real. Moreover, we denote by \( {{\mathcal {H}}}({\mathbb {T}}^\infty _{s}\times \mathbb {T}_s)\), the space of analytic functions \({\mathbb {T}}^\infty _s \times {\mathbb {T}}_s \rightarrow {\mathbb {C}}\) which are real on real. The space \({{\mathcal {H}}}_s\) can be identified with the subspace of zero-average functions of \( {{\mathcal {H}}}({\mathbb {T}}^\infty _{s}\times \mathbb {T}_s)\). Indeed if \(u \in {{\mathcal {H}}}_s\), then

$$\begin{aligned} \begin{aligned}&u= \sum _{ \ell \in {\mathbb {Z}}^\infty _*} u(\ell ,x)e^{{\mathrm{i}}\ell \cdot \varphi }= \sum _{ (\ell ,j )\in {\mathbb {Z}}^\infty _*\times {\mathbb {Z}}\setminus \{0\}} u_j(\ell )e^{{\mathrm{i}}\ell \cdot \varphi +{\mathrm{i}}j x}, \\&\quad \text {with} \quad u_j(\ell ) = \overline{u_{- j}(- \ell )} \end{aligned} \end{aligned}$$

For any \(u\in {{\mathcal {H}}}({\mathbb {T}}^\infty _{s}\times \mathbb {T}_s)\) let us denote

$$\begin{aligned} (\pi _0 u)(\varphi ,x) := \langle u(\varphi ,\cdot ) \rangle _x := \frac{1}{2 \pi }\int _{\mathbb {T}}u(\varphi , x)\, d x,\qquad \pi _0^\perp := \mathbb {1}-\pi _0. \end{aligned}$$
(2.2)

Throughout the algorithm we shall need to control the Lipschitz variation w.r.t. \(\omega \) of functions in some \({{\mathcal {H}}}({\mathbb {T}}^\infty _s,X)\), which are defined for \(\omega \) in some Cantor set. Thus, for \({{\mathcal {O}}}\subset {{\mathcal {O}}}^{(0)}\) we introduce the following norm.

Parameter dependence. Let Y be a Banach space and \(\gamma \in (0, 1)\). If \(f : \Omega \rightarrow Y\), \(\Omega \subseteq [1, 2]^{\mathbb {N}}\) is a Lipschitz function we define

$$\begin{aligned} \begin{aligned} | f |_Y^{{\mathrm{sup}}}&:= \sup _{\omega \in \Omega } | f(\omega )|_Y, \quad | f |_Y^{{\mathrm{lip}}} := \sup _{\begin{array}{c} \omega _1, \omega _2 \in \Omega \\ \omega _1 \ne \omega _2 \end{array}} \frac{| f(\omega _1) - f(\omega _2) |_Y}{| \omega _1 - \omega _2|_\infty }, \\ | f |_Y^\Omega&:= | f |_Y^{{\mathrm{sup}}} + \gamma | f |_Y^{{\mathrm{lip}}}. \end{aligned} \end{aligned}$$
(2.3)

If \(Y = {{\mathcal {H}}}_s\) we simply write \(| \cdot |_\sigma ^{{\mathrm{sup}}}\), \(| \cdot |_\sigma ^{{\mathrm{lip}}}\), \(| \cdot |_\sigma ^{\Omega }\). If Y is a finite dimensional space, we write \(| \cdot |^{{\mathrm{sup}}}\), \(| \cdot |^{{\mathrm{lip}}}\), \(| \cdot |^{\Omega }\).

Linear operators. For any \(\sigma > 0\), \(m \in {\mathbb {R}}\) we define the class of linear operators of order m (densely defined on \(L^2({\mathbb {T}})\)) \({{\mathcal {B}}}^{\sigma , m}\) as

$$\begin{aligned} \begin{aligned}&{{\mathcal {B}}}^{\sigma , m} := \Big \{ {{\mathcal {R}}} : L^2({\mathbb {T}}) \rightarrow L^2({\mathbb {T}}) : \Vert {{\mathcal {R}}}\Vert _{{{\mathcal {B}}}^{\sigma , m}} < \infty \Big \} \quad \text {where} \\&\quad \Vert {{\mathcal {R}}}\Vert _{{{\mathcal {B}}}^{\sigma , m}} := \sup _{j' \in \mathbb {Z}\setminus \{0\}} \sum _{j \in \mathbb {Z}\setminus \{0\}} e^{\sigma |j - j'| } |R_j^{j'}| \langle j' \rangle ^{- m} . \end{aligned} \end{aligned}$$
(2.4)

and for \({{\mathcal {T}}}\in {{\mathcal {H}}}({\mathbb {T}}^\infty _\sigma , {{\mathcal {B}}}^{\sigma , m})\) we set

$$\begin{aligned} \Vert {{\mathcal {T}}}\Vert _{\sigma , m} := \sum _{\ell \in {\mathbb {Z}}^\infty _*} e^{\sigma |\ell |_\eta } \Vert {{\mathcal {T}}}(\ell )\Vert _{{{\mathcal {B}}}^{\sigma , m}}. \end{aligned}$$
(2.5)

In particular we shall denote by \(\Vert \cdot \Vert _{{\sigma , m}}^{\Omega }\) the corresponding Lipshitz norm. Moreover if \(m=0\) we shall drop it, and write simply \(\Vert \cdot \Vert _{{\sigma }}\) or \(\Vert \cdot \Vert _{{\sigma }}^{\Omega }\).

3 The Iterative Scheme

Let us rewrite (1.1) as

$$\begin{aligned} F_0(u) =0 \end{aligned}$$
(3.1)

where

$$\begin{aligned} F_0(u):= (\omega \cdot \partial _\varphi + \partial _{xxx})u + Q(u,u_x,u_{xx},u_{xxx}) + f(\varphi ,x) \end{aligned}$$
(3.2)

where we \({{\mathfrak {f}}}(t,x)=f(\omega t,x)\) and, as custumary the unknown u is a function of \((\varphi ,x)\in \mathbb {T}^\infty \times \mathbb {T}\).

We introduce the (Taylor) notation

$$\begin{aligned} \begin{aligned} L_0&:= (\omega \cdot \partial _\varphi + \partial _{xxx})=F'_0(0),\qquad f_0=F_0(0) = f(\varphi ,x),\qquad \\ Q_0(u)&=Q(u,u_x,u_{xx},u_{xxx}){\mathop {=}\limits ^{(1.4)}} \partial _{xx}\Big (3 {\mathtt {c}}_3 u_x^2 + 2 {\mathtt {c}}_2 u u_x + {\mathtt {c}}_1 u^2 \Big ) \\&\quad - \partial _x ({\mathtt {c}}_2 u_x^2 + 2 {\mathtt {c}}_1 u u_x + 3 {\mathtt {c}}_0 u^2 ) \end{aligned} \end{aligned}$$
(3.3)

so that (3.1) reads

$$\begin{aligned} f_0 + L_0 u + Q_0(u) =0. \end{aligned}$$

Note that \(Q_0\) is of the form

$$\begin{aligned} Q_0(u)= \sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} q^{(0)}_{i,j} (\partial _x^i u)(\partial _x^j u) \end{aligned}$$
(3.4)

with the coefficients \(q^{(0)}_{i,j} \) satisfying

$$\begin{aligned} \sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} |q^{(0)}_{i,j}|\le C, \end{aligned}$$
(3.5)

where the constant C depends clearly on \(|{\mathtt {c}}_0|, \ldots , |{\mathtt {c}}_3|\). In particular, this implies that for all \(u\in {{\mathcal {H}}}_s\) one has the following.

  1. Q1.

    \(|Q_0(u)|_{s-\sigma } \lesssim \sigma ^{-4}|u|_s^2\)

  2. Q2.

    \(|Q_0'(u)[h]|_{s-\sigma }\lesssim \sigma ^{-4}|u|_s |h|_s\)

We now fix the constants

$$\begin{aligned} \begin{aligned}&\mu > \max \left\{ 1, \frac{1}{\eta }\right\} , \\&\gamma _0<\frac{1}{2}\overline{\gamma },\qquad \gamma _n:=(1-2^{-n})\gamma _{n-1},\quad n\ge 1\\&\sigma _{-1}:={\frac{1}{8}\min \{(S-\overline{s}),1\}},\qquad \sigma _{n-1}=\frac{6\sigma _{-1}}{\pi ^2n^2},\quad n\ge 1,\\&s_0=S-\sigma _{-1},\qquad s_n=s_{n-1}-6\sigma _{n-1},\quad n\ge 1, \\&\varepsilon _n:= \varepsilon _0 e^{-\chi ^n},\quad \chi =\frac{3}{2}, \end{aligned} \end{aligned}$$
(3.6)

where \(\varepsilon _0\) is such that

$$\begin{aligned} e^{{\mathtt {C}}_0\sigma _{-1}^{-\mu }}|f|_S=e^{{\mathtt {C}}_0\sigma _{-1}^{-\mu }}|f_0|_S\ll \varepsilon _0. \end{aligned}$$
(3.7)

Introduce

$$\begin{aligned} {\mathtt {d}}(\ell ) := \prod _{i\in {\mathbb {N}}}(1+|\ell _i|^{5} \langle i \rangle ^{5}), \quad \forall \ell \in {\mathbb {Z}}^\infty _*. \end{aligned}$$
(3.8)

We also set \({{\mathcal {O}}}^{(-1)}:= {\mathtt {D}}_{\overline{\gamma }}\) and

$$\begin{aligned} {{\mathcal {O}}}^{(0)}:={\left\{ \omega \in {\mathtt {D}_{\overline{\gamma }}}:\quad |\omega \cdot \ell + j^3|\ge \frac{\gamma _0}{{{\mathtt {d}}}(\ell )}, \quad \forall \ell \in {\mathbb {Z}}^\infty _*,\quad j\in {\mathbb {N}},\; (\ell ,j)\ne (0,0) \right\} }.\qquad \end{aligned}$$
(3.9)

Proposition 3.1

There exists \(\tau ,\tau _1,\tau _2,\tau _3, {\mathtt {C}},\epsilon _0\) (pure numbers) such that for

$$\begin{aligned} \varepsilon _0 \le \sigma _0^{\tau } e^{-{\mathtt {C}}\sigma _0^{-\mu }} \epsilon _0, \end{aligned}$$
(3.10)

for all \(n\ge 1\) the following hold.

  1. 1.

    There exist a sequence of Cantor sets \({{\mathcal {O}}}^{(n)}\subseteq {{\mathcal {O}}}^{(n-1)}\), \(n\ge 1\) such that

    $$\begin{aligned} {{\mathbb {P}}} ({{\mathcal {O}}}^{(n-1)}\setminus {{\mathcal {O}}}^{(n)}) \lesssim \frac{\gamma _0}{n^2}. \end{aligned}$$
    (3.11)
  2. 2.

    For \(n\ge 1\), there exists a sequence of linear, invertible, bounded and symplectic changes of variables defined for \(\omega \in {{\mathcal {O}}}^{(n-1)}\), of the form

    $$\begin{aligned} T_{n} v(\varphi ,x)= (1+\xi ^{(n)}_x)v( \varphi +\omega \beta ^{(n)}(\varphi ), x+ \xi ^{(n)}(\varphi ,x) + p^{(n)}(\varphi ))\, \end{aligned}$$
    (3.12)

    satisfying

    $$\begin{aligned} | \xi ^{(n)}|_{s_{n-1}- \sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}},|\beta ^{(n)}|_{s_{n-1}- \sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}}, |p^{(n)}|_{s_{n-1}- \sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}} \lesssim \sigma _{n-1}^{-\tau _1} \varepsilon _{n-1} e ^{C \sigma _{n-1}^{-\mu }}, \end{aligned}$$
    (3.13)

    for some constant \(C>0\).

  3. 3.

    For \(n\ge 0\), there exists a sequence of functionals \(F_n(u) \equiv F_n(\omega , u(\omega ))\), defined for \(\omega \in {{\mathcal {O}}}^{(n-1)}\), of the form

    $$\begin{aligned} F_n(u)=f_n + L_n u +Q_n(u), \end{aligned}$$
    (3.14)

    such that

    1. (a)

      \(L_n\) is invertible for \(\omega \in {{\mathcal {O}}}^{(n)}\) and setting

      $$\begin{aligned} h_n:=-L_n^{-1}f_n, \end{aligned}$$
      (3.15)

      there exists \({{\mathtt {r}}}_n ={{\mathtt {r}}}_{n}(\varphi )\in {{\mathcal {H}}}({\mathbb {T}}^\infty _{s_{n-1}-3\sigma _{n-1}})\) such that

      $$\begin{aligned} \begin{aligned}&F_{n}(u) = {{\mathtt {r}}}_nT_{n}^{-1}F_{n-1}(h_{n-1}+T_k u),\qquad n\ge 1, \\&|{{\mathtt {r}}}_n-1|_{s_{n-1}-3\sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}} \le \sigma _{n-1}^{-\tau _2} e^{C\sigma _{n-1}^{-\mu }}\varepsilon _{n-1} \end{aligned} \end{aligned}$$
      (3.16)
    2. (b)

      \(f_n=f_n(\varphi ,x)\) is a given function satisfying

      $$\begin{aligned} |f_n|_{s_{n-1}-2\sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}}\lesssim \sigma _{n-1}^{-4} \varepsilon _{n-1}^2,\qquad n\ge 1 \end{aligned}$$
      (3.17)
    3. (c)

      \(L_n\) is a linear operator of the form

      $$\begin{aligned} L_n = \omega \cdot \partial _\varphi + (1+A_n)\partial _{xxx} + B_n(\varphi ,x)\partial _x + C_n(\varphi ,x) \end{aligned}$$
      (3.18)

      such that

      $$\begin{aligned} \frac{1}{2\pi }\int _\mathbb {T}B_n(\varphi ,x) dx = \overline{b}_n \end{aligned}$$
      (3.19)

      and for \(n\ge 1\)

      $$\begin{aligned} \begin{aligned} |A_n-A_{n-1}|^{{{\mathcal {O}}}^{(n-1)}}&\le \sigma _{n-1}^{-\tau _2} e^{C\sigma _{n-1}^{-\mu }}\varepsilon _{n-1},\\ |B_{n}-B_{n-1}|_{s_{n-1}-3\sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}}&\lesssim \sigma _{n-1}^{-\tau _2} e^{C\sigma _{n-1}^{- \mu }}\varepsilon _{n-1} \\ |C_{n}-C_{n-1}|_{s_{n-1}-3\sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}}&\lesssim \sigma _{n-1}^{-\tau _2} e^{C\sigma _{n-1}^{- \mu }}\varepsilon _{n-1} . \end{aligned} \end{aligned}$$
      (3.20)
    4. (d)

      \(Q_n\) is of the form

      $$\begin{aligned} Q_n(u)= \sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} q^{(n)}_{i,j} (\varphi ,x) (\partial _x^i u)(\partial _x^j u) \end{aligned}$$
      (3.21)

      with the coefficients \(q^{(n)}_{i,j} (\varphi ,x) \) satisfying (3.5) for \(n=0\), while for \(n\ge 1\)

      $$\begin{aligned} \begin{aligned}&\sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} |q^{(n)}_{i,j}|_{s_{n-1}-3\sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}}\le C\sum _{l=1}^{n}2^{-l}, \\&|q^{(n)}_{i,j}-q^{(n-1)}_{i,j}|_{s_{n-1}-3\sigma _{n-1}}^{{{\mathcal {O}}}^{(n-1)}}\lesssim \sigma _{n-1}^{-\tau _3} e^{C\sigma _{n-1}^{- \mu }}\varepsilon _{n-1}. \end{aligned} \end{aligned}$$
      (3.22)
  4. 4.

    Finally one has

    $$\begin{aligned} |h_n|_{s_n}^{{{\mathcal {O}}}^{(n)}}\le \varepsilon _n \end{aligned}$$
    (3.23)

Moreover, setting

$$\begin{aligned} {{\mathcal {O}}}^{(\infty )}:=\bigcap _{n\ge 0}{{\mathcal {O}}}^{(n)}, \end{aligned}$$
(3.24)

and

$$\begin{aligned} u_n = h_0 + \sum _{j=1}^n T_1 \circ \ldots \circ T_j h_j. \end{aligned}$$
(3.25)

then

$$\begin{aligned} u_\infty := \lim _{n\rightarrow \infty } u_n \end{aligned}$$

is well defined for \(\omega \in {{\mathcal {O}}}^{(\infty )}\), belongs to \({{\mathcal {H}}}_{\overline{s}}\), and solves \(F(u_\infty )=0\). Finally the \({{\mathcal {O}}}^{(\infty )}\) has positive measure; precisely

$$\begin{aligned} {{\mathbb {P}}}({{\mathcal {O}}}^{(\infty )}) = 1 - O(\gamma _0). \end{aligned}$$
(3.26)

From Proposition 3.1 our main result Theorem 1.3 follows immediately by noting that (3.7) and (3.10) follow from (1.7) for an appropriate choice \(\varepsilon (S-\overline{s})\).

Remark 3.2

Let us spend few words on the strategy of the algorithm. At each step we apply an affine change of variables translating the approximate solution to zero; the translation is not particularly relevant and we perform it only to simplify the notation. On the other hand the linear change of variables is crucial.

In (3.14) we denote by \(f_n\) the “constant term”, by \(L_n\) is the “linearized” term and by \(Q_n\) the “quadratic” part. In this way the approximate solution at the n-th step is \(h_n=-L_n^{-1}f_n\).

In a classical KAM algorithm, in order to invert \(L_n\) one typically applies a linear change of variables that diagonalizes \(L_n\); this, together with the translation by \(h_n\) is the affine change of variables mentioned above, at least in the classical KAM scheme.

Unfortunately, in the case of unbounded nonlinearities this cannot be done. Indeed in order to diagonalize \(L_n\) in the unbounded case, one needs it to be a pseudo-differential operator. On the other hand, after the diagonalization is performed, one loses the pseudo-differential structure for the subsequent step. Thus we chose the operators \(T_n\) in (3.12) in such a way that we preserve the PDE structure and at the same time we diagonalize the highest order terms.

In the [1]-like algorithm the Authors do not apply any change of variables, but they use the reducibility of \(L_n\) only in order to deduce the estimates. However such a procedure works only in Sobolev class. Indeed in the analytic case, at each iterative step one needs to lose some analyticity, due to the small divisors. Since we are studying almost-periodic solutions, we need the analytic setting to deal with the small divisors. As usual, the problem is that the loss of the analyticity is related to the size of the perturbation; in the present case, at each step \(L_n\) is a diagonal term plus a perturbation \(O(\varepsilon _0)\) with the same \(\varepsilon _0\) for all n.

A more refined approach is to consider \(L_n\) as a small variation of \(L_{n-1}\); however the problem is that such small variation is unbounded. As a consequence, the operators \(T_n\) are not “close-to-identity”. However, since \(F_n\) is a differential operator, then the effect of applying \(T_n\) is simply a slight modification of the coefficients; see (3.20) and (3.22). Hence there is a strong motivation for applying the operators \(T_n\). In principle we could have also diagonalized the terms up to order \(-k\) for any \(k\ge 0\); however the latter change of variables are close to the identity and they introduce pseudo-differential terms.

3.1 The Zero-th Step

Item 1., 2. are trivial for \(n=0\) while item 3.(b), (c), (d) amount to the definition of \(F_0\), see (3.2), (3.3), (3.4). Regarding item 3.(a) the invertibility of \(L_0\) follows from the definition of \({{\mathcal {O}}}^{(0)}\). Indeed, consider the equation

$$\begin{aligned} L_0 h_0 = -f_0 \end{aligned}$$
(3.27)

with

$$\begin{aligned} \langle f_0(\varphi ,\cdot ) \rangle _{x} =0 \end{aligned}$$

we have the following result.

Lemma 3.3

(Homological equation) Let \(s > 0\), \(0<\sigma <1\), \(f_0 \in {{\mathcal {H}}}_{s+\sigma }\), \(\omega \in {{\mathcal {O}}}^{(0)}\) (see (1.6)). Then there exists a unique solution \(h_0 \in {{\mathcal {H}}}_s\) of (3.27) . Moreover one has

$$\begin{aligned} | h_0 |_s^{{{\mathcal {O}}}^{(0)}} \lesssim \gamma ^{-1}{\mathrm{exp}}\Big (\frac{\tau }{\sigma ^{\frac{1}{\eta }}} \ln \Big (\frac{\tau }{\sigma } \Big ) \Big ) | f |_{s + \sigma }. \end{aligned}$$

for some constant \(\tau =\tau (\eta ) > 0\).

Remark 3.4

Note that from Lemma 3.3 above it follows that there is \({\mathtt {C}}_0\) such that a solution \(h_0\) of (3.27) actually satisfies

$$\begin{aligned} | h_0 |_s^{{{\mathcal {O}}}^{(0)}} \lesssim e^{{\mathtt {C}}_0 {\sigma ^{-\mu }} } | f |_{s + \sigma }. \end{aligned}$$
(3.28)

where we recall that by (3.6), \(\mu > \max \{ 1, \frac{1}{\eta }\}\). Of course the constant \({\mathtt {C}}_0\) is correlated with the correction to the exponent \(\frac{1}{\eta }\).

From Lemma 3.3 and (3.27) it follows that \(h_0\) is analytic in a strip \(s_0\) (where \(S=s_0+\sigma _{-1}\) is the analyticity of f, to be chosen). Moreover, by Lemma 3.3 the size of \(h_0\) is

$$\begin{aligned} |h_0|_{s_0}^{{{\mathcal {O}}}^{(0)}}\sim e^{{\mathtt {C}}_0\sigma _{-1}^{-\mu }} |f_0|_{S} \end{aligned}$$
(3.29)

proving item 4. for \(|f_0|_S\) small enough, which is true by (3.7).

3.2 The \( n+1\)-th Step

Assume now that we iterated the procedure above up to \(n\ge 0\) times. This means that we arrived at a quadratic equation

$$\begin{aligned} F_n(u)=0,\qquad F_n(u) = f_n + L_n u + Q_n(u). \end{aligned}$$
(3.30)

Defined on \({{\mathcal {O}}}^{(n-1)}\) (recall that \({{\mathcal {O}}}^{(-1)}={\mathtt {D}_\gamma }\)).

By the inductive hypothesis (3.22) we deduce that for all \(0<s-\sigma < s_{n-1}- 3\sigma _{n-1}\) one has

$$\begin{aligned} |Q_n(u)|_{s-\sigma }^{{{\mathcal {O}}}^{(n-1)}}&\lesssim \sigma ^{-4}(|u|_s^{{{\mathcal {O}}}^{(n-1)}})^2 \end{aligned}$$
(3.31a)
$$\begin{aligned} |Q_n'(u)[h]|_{s-\sigma }^{{{\mathcal {O}}}^{(n-1)}}&\lesssim \sigma ^{-4}|u|_s^{{{\mathcal {O}}}^{(n-1)}} |h|_s^{{{\mathcal {O}}}^{(n-1)}} \end{aligned}$$
(3.31b)

Moreover, again by the inductive hypothesis, we can invert \(L_n\) and define \(h_n\) by (3.15). Now we set

$$\begin{aligned} F_{n+1}(v) = {{\mathtt {r}}}_{n+1} T_{n+1}^{-1}F_n(h_n+ T_{n+1}v ) \end{aligned}$$
(3.32)

where

$$\begin{aligned} T_{n+1} v(\varphi ,x)= (1+\xi ^{(n+1)}_x)v( \varphi +\omega \beta ^{(n+1)}(\varphi ), x+ \xi ^{(n+1)}(\varphi ,x) + p^{(n+1)}(\varphi ))\, \end{aligned}$$
(3.33)

and \(r_{n+1}\) are to be chosen in order to ensure that \(L_{n+1}:= F'_{n+1}(0)\) has the form (3.18) with \(n\rightsquigarrow n+1\).

Of course by Taylor expansion we can identify

$$\begin{aligned} \begin{aligned} f_{n+1}&= {{\mathtt {r}}}_{n+1}T_{n+1}^{-1}(f_n + L_n(h_n) + Q_n(h_n))= {{\mathtt {r}}}_{n+1}T_{n+1}^{-1} Q_n(h_n),\\ L_{n+1}&= {{\mathtt {r}}}_{n+1}T_{n+1}^{-1}(L_n + Q'_n(h_n)) T_{n+1}\,\\ Q_{n+1}(v)&= {{\mathtt {r}}}_{n+1}(T_{n+1}^{-1} (Q_n( h_n + T_{n+1} v) - Q_n(h_n)- Q'_n(h_n) T_{n+1} v)) \\&= {{\mathtt {r}}}_{n+1}T_{n+1}^{-1}Q_n(T_{n+1} v). \end{aligned} \end{aligned}$$
(3.34)

Remark 3.5

Note that the last equality in (3.34) follows from the fact that the nonlinearity Q in (1.1) is quadratic. In the general case, the last term is controlled by the second derivative, and thus one has to assume a bound of the type (3.31) for \(Q''\).

In Sect. 4 we prove the following

Proposition 3.6

Assuming that

$$\begin{aligned} \varepsilon _n \le \sigma _n^{\tau _1 +1}e^{-C\sigma _n^{-\mu }} \end{aligned}$$
(3.35)

for some \(C>0\), there exist \(\xi ^{(n+1)}\), \(\beta ^{(n+1)}\), \(p^{(n+1)}\) and \( {{\mathtt {r}}}_{n+1}\in {{\mathcal {H}}}({\mathbb {T}}^\infty _{s_n - \sigma _n} \times {\mathbb {T}}_{s_n - \sigma _n})\), defined for all \(\omega \in {{{\mathcal {O}}}^{(n)}}\) and satisfying

$$\begin{aligned} | \xi ^{(n+1)}|^{{{\mathcal {O}}}^{(n)}}_{s_n- \sigma _n},|\beta ^{(n+1)}|^{{{\mathcal {O}}}^{(n)}}_{s_n-\sigma _n}, |p^{(n+1)}|^{{{\mathcal {O}}}^{(n)}}_{s_n-\sigma _n}, |{{\mathtt {r}}}_{n+1}-1|^{{{\mathcal {O}}}^{(n)}}_{s_n-{\sigma _n}} \lesssim \sigma _n^{-\tau _1} \varepsilon _n e ^{C \sigma _n^{- \mu }}\, \end{aligned}$$
(3.36)

such that (3.33) is well defined and symplectic as well as its inverse, and moreover

$$\begin{aligned} {{\mathtt {r}}}_{n+1}T_{n+1}^{-1}(L_n + Q'_{n}(h_n))T_{n+1} = \omega \cdot \partial _\varphi + (1+A_{n+1})\partial _{xxx} + B_{n+1}(\varphi ,x)\partial _x + C_{n+1}(\varphi ,x)\nonumber \\ \end{aligned}$$
(3.37)

and (3.19) and (3.20) hold with \(n\rightsquigarrow n+1\).

The assumption (3.35) follows from (3.10), provided that we choose the constants \(\tau ,{\mathtt {C}}\) and \(\epsilon _0\) appropriately.

We now prove (3.21) and (3.22) for \(n\rightsquigarrow n+1\), namely the following result.

Lemma 3.7

One has

$$\begin{aligned} Q_{n+1}(v)={{\mathtt {r}}}_{n+1} T_{n+1}^{-1}Q_n(T_{n+1} v)= {{\mathtt {r}}}_{n+1} \sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} q^{(n+1)}_{i,j} (\varphi ,x) (\partial _x^i v)(\partial _x^j v) \end{aligned}$$
(3.38)

with the coefficients \(q^{(n+1)}_{i,j} (\varphi ,x) \) satisfying

$$\begin{aligned} \begin{aligned}&\sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} |q^{(n+1)}_{i,j}|^{{{\mathcal {O}}}^{(n)}}_{s_{n}-3\sigma _{n}}\le C\sum _{l=1}^{n+1}2^{-l}, \\&|q^{(n+1)}_{i,j}-q^{(n)}_{i,j}|^{{{\mathcal {O}}}^{(n)}}_{s_{n}-3\sigma _{n}}\lesssim \sigma _{n}^{-\tau _3}e^{C\sigma _{n}^{-\mu }}\varepsilon _{n}. \end{aligned} \end{aligned}$$
(3.39)

Proof

By construction

$$\begin{aligned} Q_{n+1}(u)={{\mathtt {r}}}_{n+1} \sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} T_{n+1}^{-1}[q^{(n)}_{i,j} (\varphi ,x) (\partial _x^i T_{n+1}v)(\partial _x^j T_{n+1} v)]. \end{aligned}$$
(3.40)

Now we first note that

$$\begin{aligned} \partial _x (T_{n+1}v) = \xi ^{(n+1)}_{xx} v(\theta ,y) + (1+\xi _x)^2v_{y}(\theta ,y) \end{aligned}$$

where

$$\begin{aligned} (\theta ,y)=(\varphi +\omega \beta ^{(n+1)}(\varphi ), x+ \xi ^{(n+1)}(\varphi ,x) + p^{(n+1)}(\varphi )). \end{aligned}$$

Hence the terms \(\partial _x^i T_{n+1}v\) are of the form

$$\begin{aligned}&\partial _x^i T_{n+1}v = \partial _y^i v(\theta ,y) + \sum _{l=0}^{i} g_{l,i}(\varphi ,x)\partial _y^l v(\theta ,y),\nonumber \\&|g_{l,i}|_{s_n-2\sigma _n}^{{{\mathcal {O}}}^{(n)}} \lesssim \sigma _n^{-(i+2)}|\xi ^{(n+1)}|_{s_n-\sigma _n}^{{{\mathcal {O}}}^{(n)}} \end{aligned}$$
(3.41)

Inserting (3.41) into (3.40) we get

$$\begin{aligned} \begin{aligned} q^{(n+1)}_{l,m}&= {{\mathtt {r}}}_{n+1}\left( T_{n+1}^{-1} q^{(n)}_{l,m} +\sum _{j=0}^4T_{n+1}^{-1}(q_{l,j}^{(n)}g_{m,j}) + \sum _{i=0}^4T_{n+1}^{-1}(q_{i,m}^{(n)}g_{l,i})\right. \\&\quad +\left. \sum _{\begin{array}{c} 0\le i\le 2,\,0\le j\le 3 \\ 0\le i+j\le 4 \end{array}} T_{n+1}^{-1}(q^{(n)}_{i,j}g_{l,i}g_{m,j}) \right) \end{aligned} \end{aligned}$$
(3.42)

so that

$$\begin{aligned} q^{(n+1)}_{i,j}= T_{n+1}^{-1}(q^{(n)}_{i,j} + O(\xi _{n+1})),\qquad |T^{-1}_{n+1}O(\xi _{n+1})|_{s_{n}-3\sigma _n}^{{{\mathcal {O}}}^{(n)}} \lesssim \sigma _n^{-\tau _3}\varepsilon _ne^{C\sigma _n^{\mu }}. \end{aligned}$$
(3.43)

In order to obtain the bound (3.43) we used the first line of (3.22) to control the sums appearing in (3.42).

Finally, since

$$\begin{aligned} T_{n+1}^{-1}(q)- q:= (1+{{\widetilde{\xi }}}^{(n+1)}_x)q( \varphi ,x)- q(\theta ,y) \end{aligned}$$

the bound follows. \(\square \)

Now, by (3.31a) and (3.34) \(f_{n+1}=f_{n+1}(\varphi ,x)\) satisfies

$$\begin{aligned} |f_{n+1}|_{s_{n}-2\sigma _{n}}^{{{\mathcal {O}}}^{(n)}}\lesssim \sigma _{n}^{-4} \varepsilon _{n}^2. \end{aligned}$$
(3.44)

In Sect. 5 we prove the existence of a Cantor set \({{\mathcal {O}}}^{(n+1)}\) where item 3.(a) of the iterative lemma holds with \(n\rightsquigarrow n+1\).

Proposition 3.8

Assume that

$$\begin{aligned} 2^{n}\sigma _n^{-\tau }e^{{\mathtt {C}}\sigma _n^{-\mu }}\varepsilon _n\ll 1, \end{aligned}$$
(3.45)

with \(\tau \ge \tau _2\). Setting \(\lambda _3^{(n+1)}:=1+A_{n+1}\), there exist Lipschitz functions

$$\begin{aligned} \Omega ^{(n+1)} (j)= \lambda ^{(n+1)}_3 j^3 + \lambda ^{(n+1)}_1 j + r^{(n+1)}_j \end{aligned}$$
(3.46)

satisfying

$$\begin{aligned} |\lambda _1^{(n+1)}-\lambda _1^{(n)}|^{{{\mathcal {O}}}^{(n)}}, \sup _{j\in \mathbb {Z}\setminus \{0\}}|r_j^{(n+1)}-r_j^{(n)}|^{{{\mathcal {O}}}^{(n)}}\lesssim \sigma _{n}^{-\tau }\varepsilon _{n}e^{C\sigma _{n}^{- \mu }} \end{aligned}$$
(3.47)

such that setting

$$\begin{aligned} {{\mathcal {E}}}^{(n+1)}&:=\Big \{\omega \in {{\mathcal {O}}}^{(n)}:\; |\omega \cdot \ell + \Omega ^{(n+1)}(j)-\Omega ^{(n+1)}(h)|\nonumber \\&\qquad \ge \frac{2\gamma _{n+1}|j^3 - h^3|}{{{\mathtt {d}}}(\ell )}, \ \forall (\ell ,h,j)\ne (0,h,h) \Big \} \end{aligned}$$
(3.48)

for \(\omega \in {{\mathcal {E}}}^{(n+1)}\) there exists an invertible and bounded linear operator \(M^{(n+1)}\)

$$\begin{aligned} \Vert M^{(n+1)}-\mathtt { Id}\Vert _{s_{n}-5\sigma _{n}}^{{{\mathcal {E}}}^{(n+1)}}\le \sigma _{0}^{-\tau }e^{{\mathtt {C}}\sigma _{0}^{-\mu }}\varepsilon _{0} \end{aligned}$$
(3.49)

such that

$$\begin{aligned} (M^{(n+1)})^{-1} L_{n+1} M^{(n+1)} = D_{n+1}= {\mathrm{diag}}{\left( \omega \cdot \ell + \Omega ^{(n+1)}(j) \right) }_{(\ell ,j)\in {\mathbb {Z}}^\infty _*\times \mathbb {Z}\setminus \{0\}} \end{aligned}$$
(3.50)

The assumption (3.45) follows from (3.10), provided that we choose the constants \(\tau ,{\mathtt {C}}\) and \(\epsilon _0\) appropriately.

Remark 3.9

Note that in the context of [13] Proposition 3.8 is much simpler to prove, because in order to diagonalize the linearized operator one uses tame estimates coming from the Sobolev regularity on the boundary of the domain. Then the smallness conditions are much simpler to handle. Here we have to strongly rely on the fact that \(L_{n+1}\) is a “small” unbounded perturbation of \(L_n\) in order to show that the operators \(M^{(n)}\) and \(M^{(n+1)}\) are close to each other. This is a very delicate issue; see Lemma 5.2 and Sect. 5.3, which are probably the more technical parts of this paper.

Lemma 3.10

(Homological equation) Set

$$\begin{aligned} {{\mathcal {U}}}^{(n+1)}:={\left\{ \omega \in {{\mathcal {O}}}^{(n)}:\quad |\omega \cdot \ell + \Omega ^{(n+1)}(j)|\ge \gamma _{n+1} \frac{|j|^3}{{{\mathtt {d}}}(\ell )}, \quad \forall (\ell ,j)\ne (0,0) \right\} } \end{aligned}$$
(3.51)

For \(\omega \in {{\mathcal {O}}}^{(n+1)}:= {{\mathcal {U}}}^{(n+1)}\cap {{\mathcal {E}}}^{(n+1)}\) one has

$$\begin{aligned} h_{n+1}:= - L_{n+1}^{-1}f_{n+1} \in {{\mathcal {H}}}_{s_{n+1}} \end{aligned}$$
(3.52)

and one has

$$\begin{aligned} | h_{n+1} |_{s_{n+1}}^{{{\mathcal {O}}}^{(n+1)}} \lesssim {\mathrm{exp}} \Big ({\tau }{\sigma _{n}^{-\frac{1}{\eta }}} \ln \Big (\frac{\tau }{\sigma _n} \Big ) \Big ) | f_{n+1} |_{s_{n+1} + \sigma _n }^{{{\mathcal {O}}}^{(n)}}. \end{aligned}$$

Proof

The result follows simply by using the definition of \({{\mathcal {O}}}^{(n+1)}\) and applying Lemma A.7. \(\square \)

Of course from Lemma 3.10 it follows that,

$$\begin{aligned} |h_{n+1}|_{s_{n+1}}^{{{\mathcal {O}}}^{(n+1)}} \lesssim \sigma _{n}^{-4}e^{C\sigma _{n}^{- \mu }}\varepsilon _{n}^2 \end{aligned}$$
(3.53)

Now we want to show inductively that

$$\begin{aligned} \sigma _{n}^{-4}e^{C\sigma _{n}^{- \mu }}\varepsilon _{n}^2 \le \varepsilon _0 e^{-\chi ^n+1},\qquad \chi =\frac{3}{2} \end{aligned}$$
(3.54)

for \(\varepsilon _0\) small enough.

By the definition of \(\varepsilon _n\) in (3.6), (3.54) is equivalent to

$$\begin{aligned} \varepsilon _0 \lesssim \sigma _0^4 n^{-8} e^{ \chi ^{n}(2-\chi )-C'{n^{\mu }}} \end{aligned}$$
(3.55)

Since the r.h.s. of (3.55) admits a positive minimum, we can regard it as a smallness condition on \(\varepsilon _0\), which is precisely (3.10).

We now prove (3.11) with \(n\rightsquigarrow n+1\). We only prove the bound for the set \({{\mathcal {E}}}^{(n)} \setminus {{\mathcal {E}}}^{(n+1)}\). The other one can be proved by similar arguments (it is actually even easier). Let us start by writing

$$\begin{aligned}&{{\mathcal {E}}}^{(n)} \setminus {{\mathcal {E}}}^{(n+1)} = \bigcup _{(\ell , j, j') \ne (0, j, j)} {{\mathcal {R}}}(\ell , j, j'), \nonumber \\&{{\mathcal {R}}}(\ell , j , j') := \Big \{ \omega \in {{\mathcal {E}}}^{(n)} : |\omega \cdot \ell + \Omega ^{(n+1)}(j) - \Omega ^{(n+1)}(j')| < \frac{2 \gamma _{n+1} |j^3 - j'^3|}{{\mathtt {d}}(\ell )} \Big \}, \nonumber \\&\forall (\ell , j, j' ) \in {\mathbb {Z}}^\infty _* \times ({\mathbb {Z}}\setminus \{ 0 \})\times ({\mathbb {Z}}\setminus \{ 0 \}), \quad (\ell , j, j') \ne (0, j, j). \end{aligned}$$
(3.56)

Lemma 3.11

Denote \(|\ell |_1\) as in (1.5) with \(\eta \rightsquigarrow 1\). For any \((\ell , j, j') \ne (0, j, j)\) such that \(|\ell |_1 \le n^{2}\), one has that \({{\mathcal {R}}}(\ell , j , j') = \emptyset \).

Proof

Let \((\ell , j, j') \in {\mathbb {Z}}^\infty _* \times ({\mathbb {Z}}\setminus \{ 0 \})\times ({\mathbb {Z}}\setminus \{ 0 \})\), \((\ell , j, j') \ne (0, j, j)\), \(|\ell |_1 \le n^{2}\). If \(j = j'\), clearly \(\ell \ne 0\) and \({{\mathcal {R}}}(\ell , 0, 0) = \emptyset \) because \(\omega \in {\mathtt {D}_{\overline{\gamma }}}\) with \(\overline{\gamma }>2 \gamma _{n+1}\); recall (3.6). Hence we are left to analyze the case \(j \ne j'\).

By (3.47), for any \(j, j' \in {\mathbb {Z}}\setminus \{ 0 \}\), \(j \ne j'\)

$$\begin{aligned}&\Big | \Big ( \Omega ^{(n+1)}(j) - \Omega ^{(n+1)}(j') \Big ) - \Big (\Omega ^{(n)}(j) - \Omega ^{(n)}(j') \Big ) \Big | {\lesssim } \, \sigma _{n}^{-\tau }\varepsilon _{n}e^{C\sigma _{n}^{- \mu }} |j^3 - j'^3|.\qquad \end{aligned}$$
(3.57)

Therefore, for any \(\omega \in {{\mathcal {E}}}^{(n)}\)

$$\begin{aligned} |\omega \cdot \ell + \Omega ^{(n+1)}(j) - \Omega ^{(n+1)}(j')|&\ge |\omega \cdot \ell + \Omega ^{(n)}(j) - \Omega ^{(n)}(j')| \nonumber \\&\quad - \Big | \Big ( \Omega ^{(n+1)}(j) - \Omega ^{(n+1)}(j') \Big ) - \Big (\Omega ^{(n)}(j) - \Omega ^{(n)}(j') \Big ) \Big | \nonumber \\&\ge \frac{2 \gamma _n |j^3 - j'^3|}{{\mathtt {d}}(\ell )} - C \sigma _{n}^{-\tau }\varepsilon _{n}e^{C\sigma _{n}^{- \mu }} |j^3 - j'^3| \nonumber \\&\ge \frac{2 \gamma _{n+1}|j^3 - j'^3|}{{\mathtt {d}}(\ell )} \end{aligned}$$
(3.58)

where in the last inequality we used (3.6) and the fact that, by (A.4) one has

$$\begin{aligned} \begin{aligned} \sigma _{n}^{-\tau }\varepsilon _{n}e^{C\sigma _{n}^{- \mu }}{{\mathtt {d}}(\ell )}&\le \sigma _{n}^{-\tau }\varepsilon _{n}e^{C\sigma _{n}^{- \mu }} (1+n^2)^{C(1)n}\le \gamma _0 2^{-n}. \end{aligned} \end{aligned}$$

The estimate (3.58) clearly implies that \({{\mathcal {R}}}(\ell , j, j') = \emptyset \) for \(|\ell |_1 \le n^2\). \(\square \)

Lemma 3.12

Let \({{\mathcal {R}}}(\ell , j, j') \ne \emptyset \). Then \(\ell \ne 0\), \(|j^3 - j'^3| \lesssim \Vert \ell \Vert _1\) and \({{\mathbb {P}}}\Big ({{\mathcal {R}}}(\ell , j, j')\Big ) \lesssim \frac{\gamma _{n+1}}{{\mathtt {d}}(\ell )}\)

Proof

The proof is identical to the one for Lemma 6.2 in [21], simply replacing \(j^2\) with \(j^3\). \(\square \)

By (3.56) and collecting Lemmata 3.11, 3.12, one obtains that

$$\begin{aligned} \begin{aligned} {{\mathbb {P}}}\Big ( {{\mathcal {E}}}^{(n)} \setminus {{\mathcal {E}}}^{(n+1)}\Big )&\lesssim \sum _{\begin{array}{c} |\ell |_1 \ge n^2 \\ |j|, |j'| \le C \Vert \ell \Vert _1 \end{array}} \frac{\gamma _{n+1}}{{\mathtt {d}}(\ell )} \lesssim \gamma _{n+1} \sum _{|\ell |_1 \ge n^2} \frac{\Vert \ell \Vert _1^2}{{\mathtt {d}}(\ell )} \\&\lesssim \gamma _{n+1} n^{- 2} \sum _{\ell \in {\mathbb {Z}}^\infty _*} \frac{|\ell |_1^3}{{\mathtt {d}}(\ell )} \lesssim \gamma _{n+1} n^{- 2}. \end{aligned} \end{aligned}$$
(3.59)

where in the last inequality we used Lemma A.8. Thus (3.11) follows.

We now study the convergence of the scheme. Precisely we show that the series (3.25) converges totally in \({{\mathcal {H}}}_{\overline{s}}\) . Note that

$$\begin{aligned} |T_i u|_{\overline{s}}^{{{\mathcal {O}}}^{(\infty )}} \le (1+2^{-i})|u|_{\overline{s}+\sigma _i}^{{{\mathcal {O}}}^{(\infty )}} \le 2|u|_{\overline{s}+\sigma _i}^{{{\mathcal {O}}}^{(\infty )}}. \end{aligned}$$
(3.60)

Thus, using (3.60) into (3.25) we get

$$\begin{aligned} |u_n|_{\overline{s}}^{{{\mathcal {O}}}^{(\infty )}} \le |h_0|_{\overline{s}}^{{{\mathcal {O}}}^{(\infty )}} + \sum _{j=1}^n 2^j |h_j|_{\overline{s}+ (\sigma _1+\ldots +\sigma _j)}^{{{\mathcal {O}}}^{(\infty )}} \end{aligned}$$
(3.61)

Now since

$$\begin{aligned} \overline{s} + \sum _{n=1}^{\infty }\sigma _n = \overline{s} + \frac{6\sigma _{-1}}{\pi ^2} \sum _{n\ge 1} \frac{1}{n^2} = s_\infty \le s_j \end{aligned}$$
(3.62)

we deduce that \(u_\infty \in {{\mathcal {H}}}_{\overline{s}}\). Finally by continuity

$$\begin{aligned} F(u_\infty ) = \lim _{n\rightarrow \infty } F(u_n) =\lim _{n\rightarrow \infty } T_1^{-1}T_2^{-1}\ldots T_n^{-1} F_n(h_n)=0. \end{aligned}$$

so the assertion follows since (recall \(\overline{s}:= s_\infty - \sum _{n\ge 1} \sigma _n\) and (3.62))

$$\begin{aligned} |T_1^{-1}T_2^{-1}\ldots T_n^{-1} F_n(h_n)|_{\overline{s}}^{{{\mathcal {O}}}^{(\infty )}} \le 2^n \sigma _n^{-4}\varepsilon _n^2 . \end{aligned}$$

We finally conclude the proof of Proposition 3.1 by showing that (3.26) holds.

First of all, reasoning as in Lemma 3.12 and using Lemma A.8, we see that

$$\begin{aligned} {{\mathbb {P}}}({{\mathcal {O}}}^{(0)}) = 1-O(\gamma _0) \end{aligned}$$

Then

$$\begin{aligned} {{\mathbb {P}}}({{\mathcal {O}}}^{(\infty )}) = {{\mathbb {P}}}({{\mathcal {O}}}^{(0)}) - \sum _{n\ge 0}{{\mathbb {P}}}({{\mathcal {O}}}^{(n)} \setminus {{\mathcal {O}}}^{(n+1)}) \end{aligned}$$

so that (3.26) follows by (3.11). \(\square \)

4 Proof of Proposition 3.6

In order to prove Proposition 3.6, we start by dropping the index n, i.e. we set \({{\mathcal {L}}} \equiv L_n\) (see (3.18)) and \({{\mathcal {Q}}} \equiv Q'_n(h_n)\) (see (3.34)).

More generally, we consider a Hamiltonian operator of the form

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}^{(0)}&= {{\mathcal {L}}} + {{\mathcal {Q}}} \\ {{\mathcal {L}}}&: = \omega \cdot \partial _\varphi + \lambda _3 \partial _x^3 + a_1(\varphi , x ) \partial _x + a_0(\varphi , x), \\ {{\mathcal {Q}}}&:= d_3(\varphi , x) \partial _x^3 + d_2(\varphi , x) \partial _x^2 + d_1(\varphi , x) \partial _x + d_0(\varphi , x) \end{aligned} \end{aligned}$$
(4.1)

defined for all \(\omega \in \Omega \subseteq {\mathtt {D}_\gamma }\) and \(\lambda _3\), \(a_0, a_1, d_0 \ldots , d_3\) satisfy the following properties.

  1. 1.

    There is \(\delta _0\) small enough such that

    $$\begin{aligned} |\lambda _3 -1|^{\Omega } \le \delta _0 \end{aligned}$$
    (4.2)
  2. 2.

    There is \(\rho >0\) such that \(a_i\in {{\mathcal {H}}}({\mathbb {T}}^\infty _\rho \times {\mathbb {T}}_\rho )\) and

    $$\begin{aligned} | a_i |_{\rho }^{\Omega } \le \delta _0,\qquad i=0,1 \end{aligned}$$
    (4.3)

    and moreover

    $$\begin{aligned} \lambda _1 := \frac{1}{2 \pi } \int _{\mathbb {T}}a_1(\varphi , x)\, d x \end{aligned}$$
    (4.4)

    i.e. it does not depend on \(\varphi \).

  3. 3.

    \(d_0 \ldots , d_3 \in {{\mathcal {H}}}({\mathbb {T}}^\infty _\rho \times {\mathbb {T}}_\rho )\) (note that by the Hamiltonian structure \(d_2 = \partial _x d_3\)) and they satisfy the estimate

    $$\begin{aligned} | d_i |_{\rho }^{\Omega } \lesssim \delta , \end{aligned}$$
    (4.5)

    for some \(\delta \ll \min \{\delta _0,\rho \}\).

Let us now choose \(\zeta \) such that \(0<\zeta \ll \rho \) and

$$\begin{aligned} \zeta ^{- \tau '} e^{2 {\mathtt {C}}_0 \zeta ^{- \mu }} \delta \ll 1. \end{aligned}$$
(4.6)

for some \(\tau '>0\). We shall conjugate \({{\mathcal {L}}}^{(0)}\) to a new operator \(\frac{1}{{\mathtt {r}}}{{\mathcal {L}}}_+\) with \({{\mathtt {r}}}={{\mathtt {r}}}(\varphi )\) an explicit function with

$$\begin{aligned} {{\mathcal {L}}}_+ = \omega \cdot \partial _\varphi +\lambda _3^+\partial _x^3 + a_1^+(\varphi ,x)\partial _x + a_0^+(\varphi ,x) \end{aligned}$$
(4.7)

with the coefficients satisfying

$$\begin{aligned} |\lambda _3^+ -\lambda _3|^{\Omega } \lesssim \delta \end{aligned}$$
(4.8)

and

$$\begin{aligned} | a_i^+ - a_i |_{\rho -2\zeta }^{\Omega } \le \zeta ^{- \tau '} e^{2 {\mathtt {C}}_0 \zeta ^{- \mu }} \delta ,\qquad \lambda _1 := \frac{1}{2 \pi } \int _{\mathbb {T}}a_1(\varphi , x)\, d x . \end{aligned}$$
(4.9)

This will allow us to conclude the proof of Proposition 3.6.

4.1 Elimination of the x-Dependence from the Highest Order Term

Consider an analytic function \(\alpha (\varphi , x)\) (to be determined) and let

$$\begin{aligned} {{\mathcal {T}}}_1 u (\varphi , x) := (1 + \alpha _x(\varphi , x)) ({{\mathcal {A}}} u)(\varphi , x), \quad {{\mathcal {A}}} u(\varphi , x) := u(\varphi , x + \alpha (\varphi , x)). \end{aligned}$$

We choose \(\alpha (\varphi , x)\) and \(m_3(\varphi )\) in such a way that

$$\begin{aligned} (\lambda _3 + d_3(\varphi , x)) \big ( 1 + \alpha _x(\varphi , x)\big )^{3} = m_3(\varphi ), \end{aligned}$$
(4.10)

which implies

$$\begin{aligned} \alpha (\varphi , x) := \partial _x^{- 1} \Big [ \frac{m_3(\varphi )^{\frac{1}{3}}}{\big (\lambda _3 + d_3(\varphi , x) \big )^{\frac{1}{3}}} - 1 \Big ], \quad m_3(\varphi ) := \Big (\frac{1}{2 \pi } \int _{\mathbb {T}}\frac{d x}{\big ( \lambda _3 + d_3(\varphi , x) \big )^{\frac{1}{3}}} \Big )^{- 3}.\nonumber \\ \end{aligned}$$
(4.11)

By (4.2), (4.5) and Lemma A.5 one has

$$\begin{aligned} | m_3 - \lambda _3 |_{\rho }^{\Omega },\, | \alpha |_{\rho }^{\Omega } \lesssim \delta \end{aligned}$$
(4.12)

Note that for any \(0<\zeta \ll \rho \) such that \(\delta \zeta ^{-1} \ll 1\), by Lemma A.1, \(x \mapsto x + \alpha (\varphi , x)\) is invertible and the inverse is given by \(y \mapsto y + {{\widetilde{\alpha }}}(\varphi , y)\) with

$$\begin{aligned} {{\widetilde{\alpha }}} \in {{\mathcal {H}}}({\mathbb {T}}^\infty _{\rho - \zeta } \times {\mathbb {T}}_{\rho - \zeta }), \quad | {{\widetilde{\alpha }}} |_{\rho - \zeta }^{\Omega } , | \alpha |_{\rho }^{\Omega } \lesssim \delta . \end{aligned}$$
(4.13)

A direct calculations shows that

$$\begin{aligned} \begin{aligned} {{\mathcal {A}}}^{- 1} u(\varphi , y) = u(\varphi , y + {{\widetilde{\alpha }}}(\varphi , y)), \quad {{\mathcal {T}}}_1^{- 1} = (1 + {{\widetilde{\alpha }}}_y) {{\mathcal {A}}}^{- 1} \end{aligned} \end{aligned}$$
(4.14)

and the following conjugation rules hold:

$$\begin{aligned} \begin{aligned}&{{\mathcal {T}}}_1^{- 1} \, a(\varphi , x) \, {{\mathcal {T}}}_1 = {{\mathcal {A}}}^{- 1} \, a(\varphi , x) \, {{\mathcal {A}}} = ({{\mathcal {A}}}^{- 1} a)(\varphi , y), \\&{{\mathcal {T}}}_1^{- 1} \partial _x {{\mathcal {T}}}_1 = \Big ( 1 + {{\mathcal {A}}}^{- 1}(\alpha _x) \Big ) \partial _y + (1 + {{\widetilde{\alpha }}}_y) {{\mathcal {A}}}^{- 1}(\alpha _{xx}) , \\&{{\mathcal {T}}}_1^{- 1} \omega \cdot \partial _\varphi {{\mathcal {T}}}_1 = \omega \cdot \partial _\varphi + {{\mathcal {A}}}^{- 1}(\omega \cdot \partial _\varphi \alpha ) \partial _y + (1 + {{\widetilde{\alpha }}}_y) {{\mathcal {A}}}^{- 1}(\omega \cdot \partial _\varphi \alpha _x). \end{aligned} \end{aligned}$$
(4.15)

Clearly one can get similar conjugation formulae for higher order derivatives, having expression similar to (3.41). In conclusion

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}^{(1)}&:= {{\mathcal {T}}}_1^{- 1} ({{\mathcal {L}}} + {{\mathcal {Q}}}) {{\mathcal {T}}}_1 \\&= \omega \cdot \partial _\varphi + {{\mathcal {A}}}^{- 1}\Big [ (\lambda _3 + q_3)(1 + \alpha _x)^3 \Big ] \partial _y^3\\&\quad + b_2(\varphi , y) \partial _y^2 + b_1(\varphi , y) \partial _y + b_0(\varphi , y)\\&=\omega \cdot \partial _\varphi + m_3(\varphi ) \partial _x^3 + b_1(\varphi , x) \partial _x + b_0(\varphi , x) \end{aligned} \end{aligned}$$
(4.16)

for some (explicitly computable) coefficients \(b_i\), where in the last equality we used (4.10) and the fact that \({{\mathcal {T}}}_1\) is symplectic, so that \(b_2(\varphi , x) = 2 \partial _x m_3(\varphi ) = 0\).

Furthermore, the estimates (4.2), (4.3), (4.12), (4.13), Corollary A.2 and Lemmata A.3, A.4 imply that for \(0 < \zeta \ll \rho \)

$$\begin{aligned} | b_i |_{\rho - 2\zeta }^{\Omega } \lesssim \delta _0, \quad | b_i - a_i |_{\rho -2\zeta }^{\Omega } \lesssim \zeta ^{- \tau } \delta , \quad \text {for some} \quad \tau > 0. \end{aligned}$$
(4.17)

4.2 Elimination of the \(\varphi \)-Dependence from the Highest Order Term

We now consider a quasi periodic reparametrization of time of the form

$$\begin{aligned} {{\mathcal {T}}}_2 u(\varphi , x) := u(\varphi + \omega \beta (\varphi ), x) \end{aligned}$$
(4.18)

where \(\beta : {\mathbb {T}}^\infty _{\rho -\zeta } \rightarrow {\mathbb {R}}\) is an analytic function to be determined. Precisely we choose \(\lambda _3^+ \in {\mathbb {R}}\) and \(\beta (\varphi )\) in such a way that

$$\begin{aligned} \lambda _3^+ \Big (1 + \omega \cdot \partial _\varphi \beta (\varphi ) \Big ) = m_3(\varphi ), \end{aligned}$$
(4.19)

obtaining thus

$$\begin{aligned} \lambda _3^+ := \int _{{\mathbb {T}}^\infty } m_3(\varphi )\,d \varphi , \quad \beta (\varphi ) := (\omega \cdot \partial _\varphi )^{- 1}\Big [ \frac{m_3}{\lambda _3^+} - 1\Big ] \end{aligned}$$
(4.20)

where we recall the definition A.3. By the estimates (4.12) and by Lemma 3.3, one obtains that for \(0 < \zeta \ll \rho \)

$$\begin{aligned} |\lambda _3^+ - \lambda _3|^{\Omega } \lesssim \delta , \quad | \beta |_{\rho - \zeta }^{\Omega } \lesssim e^{{\mathtt {C}}_0 \zeta ^{- \mu }} \delta . \end{aligned}$$
(4.21)

By Lemma A.1 and (4.6) we see that \(\varphi \mapsto \varphi + \omega \beta (\varphi )\) is invertible and the inverse is given by \(\vartheta \mapsto \vartheta + \omega {{\widetilde{\beta }}}(\vartheta )\) with

$$\begin{aligned} {{\widetilde{\beta }}} \in {{\mathcal {H}}}({\mathbb {T}}^\infty _{\rho - 2\zeta }), \quad | {{\widetilde{\beta }}} |_{\rho - 2\zeta }^{\Omega } \lesssim e^{{\mathtt {C}}_0 \zeta ^{- \mu }} \delta . \end{aligned}$$
(4.22)

The inverse of the operator \({{\mathcal {T}}}_2\) is then given by

$$\begin{aligned} {{\mathcal {T}}}_2^{- 1} u(\vartheta , x) = u(\vartheta + \omega {{\widetilde{\beta }}}(\vartheta ), x). \end{aligned}$$
(4.23)

so that

$$\begin{aligned} \begin{aligned} {{\mathcal {T}}}_2^{- 1} {{\mathcal {L}}}^{(1)} {{\mathcal {T}}}_2&= {{\mathcal {T}}}_2^{- 1} \big ( 1 + \omega \cdot \partial _\varphi \beta \big ) \omega \cdot \partial _\vartheta + {{\mathcal {T}}}_2^{- 1}(m_3) \partial _x^3 + {{\mathcal {T}}}_2^{- 1}(b_1) \partial _x + {{\mathcal {T}}}_2^{- 1}(b_0)\\&=:\frac{1}{{\mathtt {r}}} {{\mathcal {L}}}^{(2)} \end{aligned} \end{aligned}$$
(4.24)

where

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}^{(2)}&:= \omega \cdot \partial _\vartheta + \lambda _3^+ \partial _x^3 +c_1(\vartheta , x) \partial _x + c_0(\vartheta , x), \\ {{\mathtt {r}}}&:= \frac{1}{{{\mathcal {T}}}_2^{- 1}\big (1 + \omega \cdot \partial _\varphi \beta \big )} {\mathop {=}\limits ^{(4.19)}} \frac{\lambda _3^+}{{{\mathcal {T}}}_2^{- 1}(m_3)}, \\ c_i&:= {{\mathtt {r}}}{{{\mathcal {T}}}_2^{- 1}(b_i)}, \quad i = 1, 0. \end{aligned} \end{aligned}$$
(4.25)

Therefore by the estimates (4.12), (4.21), (4.22) and by applying Corollary A.2, Lemma A.5, and (4.6), one gets

$$\begin{aligned} \begin{aligned} | {{\mathtt {r}}} - 1 |_{\rho - \zeta }^{\Omega }&\lesssim \delta \,\\ | c_i - a_i |_{\rho - \zeta }^{\Omega }&{\lesssim }\, \zeta ^{- \tau }e^{{\mathtt {C}}_0 \zeta ^{- \mu }} \delta , \quad i = 0,1. \end{aligned} \end{aligned}$$
(4.26)

4.3 Time Dependent Traslation of the Space Variable

Let \(p : {\mathbb {T}}^\infty _{\rho -2\zeta } \rightarrow {\mathbb {R}}\) be an analytic function to be determined and let

$$\begin{aligned} {{\mathcal {T}}}_3 u(\varphi , x) := u(\varphi , x + p(\varphi )), \quad \text {with inverse} \quad {{\mathcal {T}}}_3^{- 1} u(\varphi , y) = u(\varphi , y - p(\varphi )). \end{aligned}$$
(4.27)

Computing explicitly

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}^{(3)}&:= {{\mathcal {T}}}_3^{- 1} {{\mathcal {L}}}^{(2)} {{\mathcal {T}}}_3 = \omega \cdot \partial _\varphi + \lambda _3^+ \partial _x^3 + a_1^+(\varphi , x) \partial _x + a_0^+(\varphi , x), \\ a_1^+&:= \omega \cdot \partial _\varphi p + {{\mathcal {T}}}_3^{- 1}(c_1), \quad a_0^+ := {{\mathcal {T}}}_3^{- 1}(c_0), \end{aligned} \end{aligned}$$
(4.28)

and by (4.4) one has

$$\begin{aligned} \begin{aligned} \frac{1}{2 \pi } \int _{\mathbb {T}}{{\mathcal {T}}}_3^{- 1}(c_1)(\varphi , y)\, d y&= \frac{1}{2 \pi } \int _{\mathbb {T}}c_1(\varphi , x)\, d x \\&= \frac{1}{2 \pi } \int _{\mathbb {T}}a_1(\varphi , x)\, d x + \frac{1}{2 \pi } \int _{\mathbb {T}}(c_1 - a_1)(\varphi , x)\, d x \\&= \lambda _1 + \frac{1}{2 \pi } \int _{\mathbb {T}}(c_1 - a_1)(\varphi , x)\, d x. \end{aligned} \end{aligned}$$
(4.29)

We want to choose \(p(\varphi )\) in such a way that the x-average of \(d_1\) is constant. To this purpose we define

$$\begin{aligned} p(\varphi ) := (\omega \cdot \partial _\varphi )^{- 1}\Big [ \langle (c_1 - a_1)\rangle _{\varphi , x} - \frac{1}{2 \pi } \int _{\mathbb {T}}(c_1 - a_1)(\varphi , x)\, d x \Big ] \end{aligned}$$
(4.30)

where for any \(a : {\mathbb {T}}^\infty _\sigma \times {\mathbb {T}}_\sigma \rightarrow {\mathbb {C}}\), \(\langle a \rangle _{\varphi , x}\) is defined by

$$\begin{aligned} \langle a \rangle _{\varphi , x} := \frac{1}{(2 \pi )} \int _{\mathbb {T}}\int _{{\mathbb {T}}^\infty } a(\varphi , x)\, d \varphi \, d x \end{aligned}$$

(recall the definition A.3). By (4.26) and Lemma 3.3 one gets

$$\begin{aligned} | p |_{\rho - 2\zeta }^\Omega \lesssim \zeta ^{- \tau } e^{2 {\mathtt {C}}_0 \zeta ^{- \mu }} \delta {\mathop {\ll }\limits ^{(4.6)}}\zeta . \end{aligned}$$
(4.31)

Moreover

$$\begin{aligned} \lambda _1^+ := \frac{1}{2 \pi } \int _{\mathbb {T}}d_1(\varphi , x)\, d x = \lambda _1 + \langle (c_1-a_1) \rangle _{\varphi , x}. \end{aligned}$$
(4.32)

Finally using (4.26), (A.2) (with \(\Phi _\alpha = {{\mathcal {T}}}_3^{- 1}\)), (4.31), one gets

$$\begin{aligned} | a^+_i - a_i |_{\rho - 2\zeta }^\Omega \lesssim \zeta ^{- \tau '}e^{2{\mathtt {C}}_0 \zeta ^{- \mu }} \delta , \end{aligned}$$
(4.33)

for some \(\tau '>0\).

4.4 Conclusion of the Proof

We start by noting that \({{\mathcal {T}}}:={{\mathcal {T}}}_3\circ {{\mathcal {T}}}_2\circ {{\mathcal {T}}}_1\) has the form (3.33) with \(p^{(n+1)}=p\), \(\beta ^{(n+1)}=\beta \) and \(\xi ^{(n+1)}(\varphi ,x) = \alpha (\varphi +\omega \beta (\varphi ),x+p(\varphi ))\). Hence, setting \( {{\mathtt {r}}}:={{\mathtt {r}}}_{n+1}\), \(\rho := s_n-\sigma _n\), \(\delta := \sigma _n^{-4}\varepsilon _n\), \(\delta _0:=2\varepsilon _0\) and \(\zeta := \sigma _n\) we denote

$$\begin{aligned} 1+A_{n+1} = \lambda _3^+,\quad , B_{n+1}(\varphi ,x):= a_1^{+}(\varphi ,x),\quad C_{n+1}=a_0^+(\varphi ,x), \end{aligned}$$

and thus Proposition 3.6 follows. \(\square \)

5 Proof of Proposition 3.8

In order to prove Proposition 3.8, we start by considering a linear Hamiltonian operator defined for \(\omega \in {{\mathcal {O}}} \subseteq {\mathtt {D}}_\gamma \) of the form

$$\begin{aligned} {{\mathcal {L}}} = {{\mathcal {L}}}(\lambda _3, a_1, a_0) := \omega \cdot \partial _\varphi + \lambda _3 \partial _x^3 + a_1(\varphi , x) \partial _x + a_0(\varphi , x). \end{aligned}$$
(5.1)

We want to show that, for any choice of the coefficients \(\lambda _3,a_1,a_0\) satisfying some hypotheses (see below), it is possible to reduce \({{\mathcal {L}}}\) to constant coefficients. Moreover we want to show that such reduction is “Lipshitz” w.r.t. the parameters \(\lambda _3,a_1,a_0\), in a sense that will be clarified below.

Regarding the coefficients, we need to require that

$$\begin{aligned} \begin{aligned}&a_i := \sum _{k = 0}^{m} a_i^{(k)}, \quad | a_i^{(k)} |_{\rho _k}^{{\mathcal {O}}} \lesssim \delta _k, \quad \forall k = 0, \ldots , m,\ \ i=0,1, \\&|\lambda _3 -1|^{{\mathcal {O}}}\lesssim \delta _0 , \\&\lambda _1 \equiv \lambda _1(a_1) = \sum _{k = 0}^{m} \lambda _1^{(k)}, \quad \lambda _1^{(k)} := \frac{1}{2 \pi } \int _{\mathbb {T}}a_1^{(k)}(\varphi , x)\, d x = {\mathrm{const}}. \end{aligned} \end{aligned}$$
(5.2)

for some \(0<\ldots<\rho _m< \ldots < \rho _0\) and \(0<\ldots \ll \delta _m \ll \ldots \ll \delta _0\ll 1\) so that there is a third sequence \(\zeta _i\) such that \(0< \zeta _i < \rho _i\) and

$$\begin{aligned} \sum _{i \ge 0} \zeta _i^{- \tau } { e}^{C \zeta _i^{- \mu }} \delta _i \lesssim \delta _0 , \end{aligned}$$
(5.3)

for some \(\tau ,C>0\).

5.1 Reduction of the First Order Term

We consider an operator \({{\mathcal {L}}}\) of the form (5.1) satisfying the hypotheses above. We start by showing that it is possible to reduce it to constant coefficients up to a bounded reminder, and that such reduction is “Lipshitz” w.r.t. the parameters \(\lambda _3,a_1,a_0\).

Lemma 5.1

There exists a symplectic invertible operator \({{\mathcal {M}}} = {\mathrm{exp}}({{\mathcal {G}}})\), with \({{\mathcal {G}}} \equiv {{\mathcal {G}}}(\lambda _3, a_1)\) and an operator \({{\mathcal {R}}}_0 \equiv {{\mathcal {R}}}_0(\lambda _3, a_1, a_0)\) satisfying

$$\begin{aligned} \begin{aligned} {{\mathcal {G}} }&= \sum _{i = 0}^{m } {{\mathcal {G}}}^{(i)}, \quad \Vert {{\mathcal {G}}}^{(i)} \Vert _{\rho _i, - 1}^{{\mathcal {O}}} \lesssim \delta _i, \\ {{\mathcal {R}}}_0&= \sum _{i = 0}^{m } {{\mathcal {R}}}_0^{(i)}, \quad \Vert {{\mathcal {R}}}_0^{(i)} \Vert _{\rho _i - \zeta _i}^{{\mathcal {O}}} \lesssim \zeta _i^{- \tau } e^{C \zeta _i^{- \mu }} \delta _i \end{aligned} \end{aligned}$$
(5.4)

for some \(C , \tau \gg 1\), such that

$$\begin{aligned} {{\mathcal {L}}}_0 := {{\mathcal {M}}}^{- 1} {{\mathcal {L}}} {{\mathcal {M}}} = \omega \cdot \partial _\varphi + \lambda _3 \partial _x^3 + \lambda _1 \partial _x + {{\mathcal {R}}}_0. \end{aligned}$$
(5.5)

Proof

We look for \({{\mathcal {G}}}\) of the form

$$\begin{aligned} {{\mathcal {G}}}= \pi _0^\bot g(\varphi , x) \partial _x^{- 1} \end{aligned}$$

and we choose the function \(g (\varphi , x)\) where \(g = g(\lambda _3, a_1)\) in order to solve

$$\begin{aligned} \begin{aligned} 3 \lambda _3 \partial _x g(\varphi , x) + a_1(\varphi , x) = \lambda _1. \end{aligned} \end{aligned}$$
(5.6)

By (5.2), one obtains that

$$\begin{aligned} g := \frac{1}{3 \lambda _3} \partial _x^{- 1}\Big [ \lambda _1 - a_1\Big ] \end{aligned}$$
(5.7)

and therefore

$$\begin{aligned} \begin{aligned}&g = \sum _{i = 0}^{m} g_i, \quad g_i := \frac{1}{3 \lambda _3} \partial _x^{- 1} \Big [ \lambda _1^{(i)} - a_1^{(i)}\Big ], \\&| g_i |_{\rho _i}^{{\mathcal {O}}}\lesssim \delta _i, \quad i = 0, \ldots , m. \\ \end{aligned} \end{aligned}$$
(5.8)

Of course we can also write the operator \({{\mathcal {G}}} := \pi _0^\bot g(\varphi , x) \partial _x^{- 1} = \sum _{i = 0}^{m } {{\mathcal {G}}}_i\) where \({{\mathcal {G}}}_i := \pi _0^\bot g_i(\varphi , x) \partial _x^{- 1}\) and one has

$$\begin{aligned} \begin{aligned} \Vert {{\mathcal {G}}}_i \Vert _{\rho _i, - 1}^{{\mathcal {O}}}\lesssim \delta _i, \quad i = 0, \ldots , m . \end{aligned} \end{aligned}$$
(5.9)

Again by (5.2), defining \({{\mathcal {P}}} := a_1 \partial _x + a_0\), one has that \({{\mathcal {P}}} = \sum _{i = 0}^{m} {{\mathcal {P}}}_i\), where \({{\mathcal {P}}}_i := a_1^{(i)} \partial _x + a_0^{(i)}\) satisfies

$$\begin{aligned} \Vert {{\mathcal {P}}}_i \Vert _{\rho _i , 1}^{{\mathcal {O}}}\lesssim \delta _i. \end{aligned}$$
(5.10)

Therefore

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}_0&= {{\mathcal {M}}}^{- 1} {{\mathcal {L}}} {{\mathcal {M}}} = e^{- {{\mathcal {G}}}} \omega \cdot \partial _\varphi e^{{\mathcal {G}}} + \lambda _3 e^{- {{\mathcal {G}}}} \partial _x^3 e^{{\mathcal {G}}} + e^{- {{\mathcal {G}}}} {{\mathcal {P}}}e^{{\mathcal {G}}} \\&= \omega \cdot \partial _\varphi + \lambda _3 \partial _x^3 + \Big (3 \lambda _3 g_x + a_1\Big ) \partial _x + {{\mathcal {R}}}_0 \\&{\mathop {=}\limits ^{(5.6)}} \omega \cdot \partial _\varphi + \lambda _3 \partial _x^3 + \lambda _1 \partial _x + {{\mathcal {R}}}_0 \end{aligned} \end{aligned}$$
(5.11)

where

$$\begin{aligned} \begin{aligned} {{\mathcal {R}}}_0&:= \Big ( e^{- {{\mathcal {G}}}} \omega \cdot \partial _\varphi e^{{\mathcal {G}}} - \omega \cdot \partial _\varphi \Big ) + \lambda _3 \Big ( e^{- {{\mathcal {G}}}} \partial _x^3 e^{{\mathcal {G}}} -\partial _x^3 - 3 g_x \partial _x \Big ) + \Big ( e^{-{{\mathcal {G}}}} {{\mathcal {P}}} e^{{\mathcal {G}}} - {{\mathcal {P}}} \Big ) + a_0. \end{aligned}\nonumber \\ \end{aligned}$$
(5.12)

Then (5.3), (5.9), (5.10) guarantee that the hypotheses of Lemmata A.10-A.11 are verified. Hence, we apply Lemma A.10-(ii) to expand the operator \(e^{-{{\mathcal {G}}}} {{\mathcal {P}}} e^{{\mathcal {G}}} - {{\mathcal {P}}}\), Lemma A.11-(ii) to expand \(e^{- {{\mathcal {G}}}} \partial _x^3 e^{{\mathcal {G}}} - \partial _x^3 - 3 g_x \partial _x\) and Lemma A.11-(iii) to expand \(e^{- {{\mathcal {G}}}} \omega \cdot \partial _\varphi e^{{\mathcal {G}}} - \omega \cdot \partial _\varphi \). The expansion of the multiplication operator \(a_0\) is already provided by (5.2). Hence, one obtains that there exist \(C, \tau \gg 1\) such that (5.4) is satisfied. \(\square \)

We now consider a “small modification” of the operator \({{\mathcal {L}}}\) in the following sense. We consider an operator

$$\begin{aligned} {{\mathcal {L}}}^+={{\mathcal {L}}}(\lambda _3^+,a_1^+,a_0^+) : = \omega \cdot \partial _\varphi + \lambda _3^+ \partial _x^3 + a_1^+(\varphi ,x) \partial _x +a_0^+(\varphi ,x) \end{aligned}$$
(5.13)

with

$$\begin{aligned} \frac{1}{2\pi } \int _{\mathbb {T}}a_1^+(\varphi , x)\, d x =:\lambda _1^+= {\mathrm{const}}, \quad |a_i^+ - a_i |_{\rho _{m+1}},\, |\lambda _3^+ - \lambda _3|\lesssim \delta _{m+1}. \end{aligned}$$
(5.14)

Of course we can apply Lemma 5.1 and conjugate \({{\mathcal {L}}}^+\) to

$$\begin{aligned} {{\mathcal {L}}}_0^+ : = \omega \cdot \partial _\varphi + \lambda _3^+ \partial _x^3 + \lambda _1^+ \partial _x + {{\mathcal {R}}}_0^+ \end{aligned}$$
(5.15)

with \({{\mathcal {R}}}_0^+\) a bounded operator. We want to show that \({{\mathcal {L}}}_0^+\) is “close” to \({{\mathcal {L}}}_0\), namely the following result.

Lemma 5.2

One has

$$\begin{aligned} \begin{aligned} |\lambda _1^+ - \lambda _1| \lesssim \delta _{m+1},\qquad \Vert {{\mathcal {R}}}_0^+ - {{\mathcal {R}}}_0 \Vert _{\rho _{m+1} - \zeta _{m+1}} \lesssim \zeta _{m+1}^{- \tau }e^{C \zeta _{m+1}^{- \mu }} \delta _{m+1}. \end{aligned} \end{aligned}$$
(5.16)

Proof

The first bound follows trivially from (5.14). Regarding the second bound one can reason as follows. As in Lemma 5.1, er can define \({{\mathcal {G}}}^+ := \pi _0^\bot g^+(\varphi , x) \partial _x^{- 1}\) with

$$\begin{aligned} g^+ := \frac{1}{3 \lambda _3^+} \partial _x^{- 1}\Big [ \lambda _1^+ - a_1^+\Big ] \end{aligned}$$
(5.17)

so that

$$\begin{aligned} \begin{aligned} \Vert {{\mathcal {G}}}^+ - {{\mathcal {G}}} \Vert _{\rho _{m+1}, -1} \lesssim \delta _{m+1}. \end{aligned} \end{aligned}$$
(5.18)

Defining \({{\mathcal {P}}}^+ := a_1^+ \partial _x + a_0^+\) and recalling that \({{\mathcal {P}}} := a_1 \partial _x + a_0\), by (5.14), one gets

$$\begin{aligned} \Vert {{\mathcal {P}}}^+ - {{\mathcal {P}}} \Vert _{\rho _{m+1}, 1} \lesssim \delta _{m+1}. \end{aligned}$$
(5.19)

The estimate on \({{\mathcal {R}}}_0^+ - {{\mathcal {R}}}_0\) follows by applying Lemmata A.13A.14, and by the estimates (5.14), (5.19), (5.18). \(\square \)

5.2 Reducibility

We now consider an operator \({{\mathcal {L}}}_0\) of the form

$$\begin{aligned} {{\mathcal {L}}}_0 \equiv {{\mathcal {L}}}_0(\lambda _1, \lambda _3, {{\mathcal {P}}}_0) := \omega \cdot \partial _\varphi + {{\mathcal {D}}}_0 + {{\mathcal {P}}}_0 \end{aligned}$$
(5.20)

with \({{\mathcal {P}}}_0\) a bounded operator and

$$\begin{aligned} {{\mathcal {D}}}_0 \equiv {{\mathcal {D}}}_0(\lambda _1, \lambda _3) := {{\mathrm{i}}}\, {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _{0}(j), \qquad \Omega _{ 0}(j) := - \lambda _3 j^3 + \lambda _1 j, \quad j \in {\mathbb {Z}}\setminus \{ 0 \},\nonumber \\ \end{aligned}$$
(5.21)

and we show that, under some smallness conditions specified below it is possible to reduce it to constant coefficients, and that the reduction is “Lipschitz” w.r.t. the parameters \(\lambda _1, \lambda _3, {{\mathcal {P}}}_0\).

In order to do so, we introduce three sequences \( 0<\ldots< \rho _{m }< \ldots < \rho _0\), \(0<\ldots \ll \delta _{m } \ll \ldots \ll \delta _0\) and \(1\ll N_0 \ll N_1 \ll \cdots \) and we assume that setting \(\Delta _i=\rho _i-\rho _{i+1}\) one has

$$\begin{aligned}&\sum _{i \ge 0} \Delta _i^{- \tau } { e}^{C \Delta _i^{- \mu }} \delta _i \lesssim \delta _0, \end{aligned}$$
(5.22)
$$\begin{aligned}&e^{- N_k\Delta _k} \delta _k+ e^{C \Delta _k^{- \mu }} \delta _k^2 \ll 2^{-k} \delta _{k + 1}, \end{aligned}$$
(5.23)
$$\begin{aligned}&\delta _k \ll (1+N_k)^{-C N_k^{\frac{1}{1+\eta }}} \end{aligned}$$
(5.24)

and

$$\begin{aligned} \begin{aligned}&|\lambda _3-1|^{{\mathcal {O}}}, |\lambda _1|^{{\mathcal {O}}} \le \delta _0, \\&{{\mathcal {P}}}_0 := \sum _{i = 0}^{m} {{\mathcal {P}}}_0^{(i)}, \quad \Vert {{\mathcal {P}}}_0^{(i)} \Vert _{\rho _i }^{{\mathcal {O}}} \le \delta _i ,\quad i = 0, \ldots , m, \end{aligned} \end{aligned}$$
(5.25)

for some \(\tau ,C>0\).

We have the following result.

Lemma 5.3

Fix \(\gamma \in [\gamma _0/2,2\gamma _0]\). For \(k=0,\ldots ,m\) there is a sequence of sets \({{\mathcal {E}}}_k\subseteq {{\mathcal {E}}}_{k-1}\) and a sequence of symplectic maps \(\Phi _k\) defined for \(\omega \in {{\mathcal {E}}}_{k+1}\) such that setting \({{\mathcal {L}}}_0\) as in (5.20) and for \(k\ge 1\),

$$\begin{aligned} {{\mathcal {L}}}_{k}:= \Phi _{k - 1}^{- 1} {{\mathcal {L}}}_{k - 1} \Phi _{k - 1}, \end{aligned}$$
(5.26)

one has the following.

  1. 1.

    \({{\mathcal {L}}}_k\) is of the form

    $$\begin{aligned} {{\mathcal {L}}}_k := \omega \cdot \partial _\varphi + {{\mathcal {D}}}_k + {{\mathcal {P}}}_k \end{aligned}$$
    (5.27)

    where

    • The operator \({{\mathcal {D}}}_k\) is of the form

      $$\begin{aligned} \begin{aligned} {{\mathcal {D}}}_k := {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _k(j), \quad \Omega _k(j) = \Omega _0 (j) + r_k(j) \end{aligned} \end{aligned}$$
      (5.28)

      with \(r_0(j)=0\) and for \(k\ge 1\), \(r_k(j)\) is defined for \(\omega \in {{{\mathcal {E}}}}_0={{\mathcal {O}}}\) and satisfies

      $$\begin{aligned} \begin{aligned} \sup _{j \in {\mathbb {Z}}\setminus \{ 0 \}}|r_k(j) - r_{k - 1}(j)|^{{\mathcal {O}}} \le \delta _{k - 1} \sum _{i=1}^{k - 1}2^{-i} . \end{aligned} \end{aligned}$$
      (5.29)
    • The operator \({{\mathcal {P}}}_k\) is such that

      $$\begin{aligned} \begin{aligned}&\text {for} \quad 0 \le k \le m, \quad {{\mathcal {P}}}_k = \sum _{i = k}^{m} {{\mathcal {P}}}_k^{(i)},\quad \Vert {{\mathcal {P}}}_k^{(i)}\Vert _{\rho _i}^{{{{\mathcal {E}}}}_k} \le \delta _i \sum _{j=1}^{k} 2^{-j}, \quad \forall i = k, \ldots , m. \\ \end{aligned}\nonumber \\ \end{aligned}$$
      (5.30)
  2. 2.

    One has \(\Phi _{k - 1} := {\mathrm{exp}}(\Psi _{k - 1})\), such that

    $$\begin{aligned} \begin{aligned} \Vert \Psi _{k - 1} \Vert _{\rho _k}^{{{{\mathcal {E}}}}_k} \lesssim e^{C \Delta _{k-1}^{- \mu }} \Vert {{\mathcal {P}}}_{k - 1}^{(k - 1)}\Vert _{\rho _{k - 1}}^{{{{\mathcal {E}}}}_{k - 1}} \lesssim e^{C \Delta _{k-1}^{- \mu }} \delta _{k - 1} \end{aligned} \end{aligned}$$
    (5.31)
  3. 3.

    The sets \({{\mathcal {E}}}_k\) are defined as

    $$\begin{aligned} \begin{aligned} {{{\mathcal {E}}}}_k&:= \Big \{\omega \in {{{\mathcal {E}}}}_{k - 1} \ : \ |\omega \cdot \ell + \Omega _{k - 1}(j) - \Omega _{ k - 1}(j')| \ge \frac{\gamma |j^3 - j'^3|}{{\mathtt {d}}(\ell )}, \\&\qquad \forall (\ell , j, j') \ne (0, j, j), \quad |\ell |_\eta \le N_{k - 1} \Big \}. \end{aligned} \end{aligned}$$
    (5.32)

Proof

The statement is trivial for \(k=0\) so we assume it to hold up to \(k<m\) and let us prove it for \(k + 1\). For any \(\Phi _k := {\mathrm{exp}}(\Psi _k)\) one has

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}_{k + 1}= \Phi _k^{- 1} {{\mathcal {L}}}_k \Phi _k = \omega \cdot \partial _\varphi + {{\mathcal {D}}}_k + \omega \cdot \partial _\varphi \Psi _k + [{{\mathcal {D}}}_k, \Psi _k]+ \Pi _{N_k}{{\mathcal {P}}}^{(k)}_k + {{\mathcal {P}}}_{k + 1} \end{aligned} \end{aligned}$$
(5.33)

where the operator \({{\mathcal {P}}}_{k + 1}\) is defined by

$$\begin{aligned} \begin{aligned} {{\mathcal {P}}}_{k + 1}&:= \Pi _{N_k}^\bot {{\mathcal {P}}}_k^{(k)} + \sum _{p \ge 2} \frac{{\mathrm{Ad}}_{\Psi _k}^p(\omega \cdot \partial _\varphi + {{\mathcal {D}}}_k)}{p!} + \sum _{i = k + 1}^{m} e^{- \Psi _k} {{\mathcal {P}}}^{(i)}_k e^{\Psi _k}\\&\quad + \sum _{p \ge 1} \frac{{\mathrm{Ad}}^p_{\Psi _k}({{\mathcal {P}}}^{(k)}_k)}{p!}. \end{aligned} \end{aligned}$$
(5.34)

Then we choose \(\Psi _k\) in such a way that

$$\begin{aligned} \begin{aligned}&\omega \cdot \partial _\varphi \Psi _k + [{{\mathcal {D}}}_k, \Psi _k]+ \Pi _{N_k} {{\mathcal {P}}}^{(k)}_k = {{\mathcal {Z}}}_k, \\&{{\mathcal {Z}}}_k := {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} ({{\mathcal {P}}}^{(k)}_k)_j^j(0), \end{aligned} \end{aligned}$$
(5.35)

namely for \(\omega \in {{{\mathcal {E}}}}_{k + 1}\) we set

$$\begin{aligned} ( \Psi _k)_j^{j'}(\ell ) := {\left\{ \begin{array}{ll} \dfrac{({{\mathcal {P}}}_k^{(k)})_j^{j'}(\ell )}{{{\mathrm{i}}}\big (\omega \cdot \ell + \Omega _{k}(j) - \Omega _k(j') \big )}, &{}\quad \forall (\ell , j, j') \ne (0, j, j), \quad |\ell |_\eta \le N_k, \\ 0 &{}\quad \text {otherwise.} \end{array}\right. }\nonumber \\ \end{aligned}$$
(5.36)

Therefore,

$$\begin{aligned} | ( \Psi _k)_j^{j'}(\ell )| \lesssim {\mathtt {d}}(\ell ) |({{\mathcal {P}}}_k^{(k)})_j^{j'}(\ell )|, \quad \forall \omega \in {{{\mathcal {E}}}}_{k+1}. \end{aligned}$$
(5.37)

and by applying Lemma A.6, using the induction estimate (5.30), one obtains

$$\begin{aligned} \Vert \Psi _k \Vert _{\rho _k - \zeta }^{{{{\mathcal {E}}}}_{k+1} } \lesssim e^{C \zeta ^{- \mu }} \Vert {{\mathcal {P}}}_k^{(k)} \Vert _{\rho _k}^{{{{\mathcal {E}}}}_{k} } {\mathop {\lesssim }\limits ^{(5.30)}} e^{C \zeta ^{- \mu }} \delta _k, \end{aligned}$$
(5.38)

for any \(\zeta <\rho _k\).

We now define the diagonal part \({{\mathcal {D}}}_{k + 1}\).

For any \(j \in {\mathbb {Z}}\setminus \{ 0 \}\) and any \(\omega \in {{{\mathcal {E}}}}_k\) one has \(|({{\mathcal {P}}}_k^{(k)})_j^j(0)| \lesssim \Vert {{\mathcal {P}}}_k^{(k)}\Vert _{\rho _k}^{{{{\mathcal {E}}}}_{k} } {\mathop {\le }\limits ^{(5.30)}} \delta _k \). The Hamiltonian structure guarantees that \({{\mathcal {P}}}_k^{(k)}(0)_j^j\) is purely imaginary and by the Kiszbraun Theorem there exists a Lipschitz extension \(\omega \in {{\mathcal {O}}} \rightarrow {{\mathrm{i}}}z_k(j)\) (with \(z_k(j)\) real) of this function satisfying the bound \(|z_k(j)|^{{\mathcal {O}}}\lesssim \delta _k\). Then, we define

$$\begin{aligned} \begin{aligned} {{\mathcal {D}}}_{k + 1}&:= {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _{k + 1}(j), \\ \Omega _{k + 1}(j)&:= \Omega _k(j) + z_k(j) = \Omega _0(j) + r_{k + 1}(j), \quad \forall j \in {\mathbb {Z}}\setminus \{ 0 \}, \\ r_{k + 1}(j)&:= r_k(j) + z_k(j) \\ \end{aligned} \end{aligned}$$
(5.39)

and one has

$$\begin{aligned} |r_{k + 1}(j) - r_k(j)|^{{{\mathcal {O}}}}= |z_k(j)|^{{{\mathcal {O}}}} \le \Vert {{\mathcal {P}}}_k^{(k)} \Vert _{\rho _k}^{{{{\mathcal {E}}}}_{k} } {\mathop {\le }\limits ^{(5.30)}} \delta _k \sum _{j=1}^k 2^{-j} \end{aligned}$$
(5.40)

which is the estimate (5.29) at the step \(k + 1\).

We now estimate the remainder \({{\mathcal {P}}}_{k + 1}\) in (5.34). Using (5.35) we see that

$$\begin{aligned} \begin{aligned} {{\mathcal {P}}}_{k + 1}&= \Pi _{N_k}^\bot {{\mathcal {P}}}_k^{(k)} + \sum _{p \ge 2} \frac{{\mathrm{Ad}}_{\Psi _k}^{p - 1}({{\mathcal {Z}}}_k - \Pi _{N_k}{{\mathcal {P}}}_k^{(k)})}{p!} + \sum _{i = k + 1}^{m} e^{- \Psi _k} {{\mathcal {P}}}^{(i)}_k e^{\Psi _k} \\&\quad + \sum _{p \ge 1} \frac{{\mathrm{Ad}}^p_{\Psi _k}({{\mathcal {P}}}^{(k)}_k)}{p!}. \\ \end{aligned} \end{aligned}$$
(5.41)

Denote

$$\begin{aligned} \begin{aligned} {{\mathcal {P}}}_{k + 1}&= \sum _{i = k + 1}^{m} {{\mathcal {P}}}_{k + 1}^{(i)} \quad \text {where} \\ {{\mathcal {P}}}_{k + 1}^{(k + 1)}&:= \Pi _{N_k}^\bot {{\mathcal {P}}}_k^{(k)} + \sum _{p \ge 2} \frac{{\mathrm{Ad}}_{\Psi _k}^{p - 1}({{\mathcal {Z}}}_k - \Pi _{N_k}{{\mathcal {P}}}_k^{(k)})}{p !} + e^{- \Psi _k} {{\mathcal {P}}}^{(k + 1)}_{k } e^{\Psi _k} \\&\quad + \sum _{p \ge 1} \frac{{\mathrm{Ad}}^p_{\Psi _k}({{\mathcal {P}}}^{(k)}_k)}{p!}, \\ {{\mathcal {P}}}_{k + 1}^{(i)}&:= e^{- \Psi _k} {{\mathcal {P}}}^{(i)}_k e^{\Psi _k}, \quad i = k + 2, \ldots , m. \end{aligned} \end{aligned}$$
(5.42)

Estimate of \({{\mathcal {P}}}_{k + 1}^{(i)}\), \(i = k + 2, \ldots , m\). By the induction estimate, one has

$$\begin{aligned} \begin{aligned} \Vert e^{- \Psi _k}{{\mathcal {P}}}_k^{(i)} e^{\Psi _k} \Vert _{\rho _i}^{{{{\mathcal {E}}}}_{k+1} }&\le \Vert {{\mathcal {P}}}_k^{(i)} \Vert _{\rho _i}^{{{{\mathcal {E}}}}_{k} } + \Vert {{\mathcal {P}}}_k^{(i)}-e^{- \Psi _k}{{\mathcal {P}}}_k^{(i)} e^{\Psi _k} \Vert _{\rho _i}^{{{{\mathcal {E}}}}_{k+1} } \\&\lesssim \delta _i \sum _{j=1}^{k}2^{-j} + \Vert \Psi _k\Vert _{\rho _i}^{{{{\mathcal {E}}}}_{k+1} } \Vert {{\mathcal {P}}}^{(i)}_k\Vert _{\rho _i}^{{{{\mathcal {E}}}}_{k+1} } {\mathop {\lesssim }\limits ^{(5.23)}} \delta _i \sum _{j=1}^{k+1}2^{-j}. \end{aligned} \end{aligned}$$
(5.43)

Estimate of \({{\mathcal {P}}}_{k + 1}^{(k + 1)}\). We estimate separately the four terms in the definition of \({{\mathcal {P}}}_{k + 1}^{(k + 1)}\) in (5.42). By Lemma A.9-(ii), one has

$$\begin{aligned} \Vert \Pi _{N_k}^\bot {{\mathcal {P}}}_k^{(k)} \Vert _{\rho _{k + 1}}^{{{\mathcal {E}}}_k} \lesssim e^{- N_k\Delta _k} \Vert {{\mathcal {P}}}_k^{(k)} \Vert _{\rho _k}^{{{\mathcal {E}}}_k} \lesssim e^{- N_k\Delta _k} \delta _k. \end{aligned}$$
(5.44)

By applying (A.7) and the estimate of Lemma A.9-(iii), one obtains

$$\begin{aligned} \begin{aligned} \Big \Vert \sum _{p \ge 2} \frac{{\mathrm{Ad}}_{\Psi _k}^{p - 1}({{\mathcal {Z}}}_k - \Pi _{N_k} {{\mathcal {P}}}_k^{(k)})}{p!} \Big \Vert _{\rho _{k + 1}}^{{{\mathcal {E}}}_{k+1}}&\le \sum _{p \ge 2} \frac{C^{p - 1}}{p!} (\Vert \Psi _k \Vert _{\rho _{k + 1}}^{{{\mathcal {E}}}_{k+1}})^{p - 1} \Vert {{\mathcal {P}}}_k^{(k)} \Vert _{\rho _k}^{{{\mathcal {E}}}_{k}} \\&{\lesssim } \Vert \Psi _k \Vert _{\rho _{k + 1}}^{{{\mathcal {E}}}_{k+1}} \Vert {{\mathcal {P}}}_k^{(k)} \Vert _{\rho _k}^{{{\mathcal {E}}}_{k}} \lesssim e^{C \Delta _k^{- \mu }} \delta _k^2 \end{aligned} \end{aligned}$$
(5.45)

and similarly

$$\begin{aligned} \Big \Vert \sum _{m \ge 1} \frac{{\mathrm{Ad}}^m_{\Psi _k}({{\mathcal {P}}}^{(k)}_k)}{m!} \Big \Vert _{\rho _{k + 1}}^{{{\mathcal {E}}}_{k+1}} \lesssim e^{C \Delta _k^{- \mu }} \delta _k^2. \end{aligned}$$
(5.46)

In conclusion we obtained

$$\begin{aligned} \Vert {{\mathcal {P}}}_{k + 1}^{(k + 1)} \Vert _{\rho _{k + 1}}^{{{\mathcal {E}}}_{k+1}} \le C' e^{- N_k\Delta _k} \delta _k + C'e^{C\Delta _k^{- \mu }} \delta _k^2 + \delta _{k + 1}\sum _{j=1}^{k}2^{-j} \end{aligned}$$
(5.47)

where \(C'\) is an appropriate constant and the last summand is a bound for the term \(e^{- \Psi _k} {{\mathcal {P}}}^{(k + 1)}_{k } e^{\Psi _k}\), which can be obtained reasoning as in (5.43). Thus we obtain

$$\begin{aligned} \Vert {{\mathcal {P}}}_{k+1}^{(k+1)}\Vert _{\rho _{k+1}}^{{{{\mathcal {E}}}}_{k+1}} \le \delta _{k + 1} \sum _{j=1}^{k+1} 2^{-j} \end{aligned}$$
(5.48)

provided

$$\begin{aligned} C' e^{- N_k\Delta _k} \delta _k+ C' e^{C\Delta _k^{- \mu }} \delta _k^2 + \delta _{k + 1}\sum _{j=1}^{k}2^{-j} \le \delta _{k + 1} \sum _{j=1}^{k + 1}2^{-j}, \end{aligned}$$

which is of course follows from (5.23). \(\square \)

Now that we reduced \({{\mathcal {L}}}_0\) to the form \({{\mathcal {L}}}_m = \omega \cdot \partial _\varphi + {{\mathcal {D}}}_m + {{\mathcal {P}}}_m\) we can apply a “standard” KAM scheme to complete the diagonalization. This is a super-exponentially convergent iterative scheme based on iterating the following KAM step.

Lemma 5.4

(The \((m+1)\)-th step) Following the notation of Lemma 5.3 we define

$$\begin{aligned} {{\mathcal {E}}}_{m+1}&:= \Big \{\omega \in {{{\mathcal {E}}}}_{m} : |\omega \cdot \ell + \Omega _{m}(j) - \Omega _{ m}(j')| \ge \frac{\gamma |j^3 - j'^3|}{{\mathtt {d}}(\ell )}, \\&\qquad \forall (\ell , j, j') \ne (0, j, j), \quad |\ell |_\eta \le N_{m} \Big \} \end{aligned}$$

and fix any \(\zeta \) such that

$$\begin{aligned} e^{- N_m\zeta } \delta _m + e^{C\zeta ^{- \mu }} \delta _m^2\ll \delta _{m+1} \end{aligned}$$
(5.49)

Then there exists a change of variables \(\Phi _{m} := {\mathrm{exp}}(\Psi _{m})\), such that

$$\begin{aligned} \begin{aligned} \Vert \Psi _{m} \Vert _{\rho _m-\zeta }^{{{{\mathcal {E}}}}_{m+1}} \lesssim e^{C \zeta ^{- \mu }} \delta _{m} \end{aligned} \end{aligned}$$
(5.50)

which conjugates \({{\mathcal {L}}}_m\) to the operator

$$\begin{aligned} {{\mathcal {L}}}_{m+1}= \omega \cdot \partial _\varphi + {{\mathcal {D}}}_{m+1} + {{\mathcal {P}}}_{m+1}. \end{aligned}$$

The operator \({{\mathcal {D}}}_{m+1}\) is of the form (5.28) and satisfies (5.29), with \(k\rightsquigarrow m+1\), while the operator \({{\mathcal {P}}}_{m+1}\) is such that

$$\begin{aligned} \Vert {{\mathcal {P}}}_{m+1}\Vert _{\rho _m-\zeta }^{{{{\mathcal {E}}}}_{m+1}} \le \delta _{m+1 } . \end{aligned}$$
(5.51)

Proof

We reason similarly to Lemma 5.3 i.e. we fix \(\Psi _{m}\) in such a way that

$$\begin{aligned} \begin{aligned}&\omega \cdot \partial _\varphi \Psi _m + [{{\mathcal {D}}}_m, \Psi _m]+ \Pi _{N_m} {{\mathcal {P}}}_m = {{\mathcal {Z}}}_m, \\&{{\mathcal {Z}}}_m := {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} ({{\mathcal {P}}}_m)_j^j(0), \end{aligned} \end{aligned}$$
(5.52)

so that we obtains

$$\begin{aligned} \Vert \Psi _m \Vert _{\rho _m - \zeta }^{{{{\mathcal {E}}}}_{m+1} } \lesssim e^{C \zeta ^{- \mu }} \Vert {{\mathcal {P}}}_m \Vert _{\rho _m}^{{{{\mathcal {E}}}}_{m} } \lesssim e^{C \zeta ^{- \mu }} \delta _m, \end{aligned}$$
(5.53)

for any \(\zeta <\rho _m\).

Now, for any \(j \in {\mathbb {Z}}\setminus \{ 0 \}\) and any \(\omega \in {{{\mathcal {E}}}}_{m}\) one has \(|({{\mathcal {P}}}_m)_j^j(0)| \lesssim \Vert {{\mathcal {P}}}_m\Vert _{\rho _m}^{{{{\mathcal {E}}}}_{m} } {\le } 2\delta _m \). The Hamiltonian structure guarantees that \({{\mathcal {P}}}_m(0)_j^j\) is purely imaginary and by the Kiszbraun Theorem there exists a Lipschitz extension \(\omega \in {{\mathcal {O}}} \rightarrow {{\mathrm{i}}}z_m(j)\) (with \(z_m(j)\) real) of this function satisfying the bound \(|z_m(j)|^{{\mathcal {O}}}\lesssim \delta _m\). Then, we define

$$\begin{aligned} \begin{aligned} {{\mathcal {D}}}_{m + 1}&:= {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _{m + 1}(j), \\ \Omega _{m + 1}(j)&:= \Omega _m(j) + z_m(j) = \Omega _0(j) + r_{m + 1}(j), \quad \forall j \in {\mathbb {Z}}\setminus \{ 0 \}, \\ r_{m + 1}(j)&:= r_m(j) + z_m(j) \\ \end{aligned} \end{aligned}$$
(5.54)

and (5.29), with \(k\rightsquigarrow m+1\).

In order to obtain the bound 5.51 we start by recalling that

$$\begin{aligned} {{\mathcal {P}}}_{m + 1} := \Pi _{N_m}^\bot {{\mathcal {P}}}_m + \sum _{p \ge 2} \frac{{\mathrm{Ad}}_{\Psi _m}^{p - 1}({{\mathcal {Z}}}_m - \Pi _{N_m}{{\mathcal {P}}}_m)}{p !} + \sum _{p \ge 1} \frac{{\mathrm{Ad}}^p_{\Psi _m}({{\mathcal {P}}}_m)}{p!}, \end{aligned}$$
(5.55)

so that reasoning as in (5.47) we obtain

$$\begin{aligned} \Vert {{\mathcal {P}}}_{m + 1} \Vert _{\rho _{m}-\zeta }^{{{\mathcal {E}}}_{m+1}} \le C' e^{- N_m\zeta } \delta _m + C'e^{C\zeta ^{- \mu }} \delta _m^2 \end{aligned}$$
(5.56)

and by (5.49) the assertion follows. \(\square \)

We now iterate the step of Lemma 5.4, using at each step a smaller loss of analyticity, namely at the p-th step we take \(\zeta _p\) with

$$\begin{aligned} \sum _{p\ge m+1}\zeta _p =\zeta , \end{aligned}$$

so that we obtain the following standard reducibility result; for a complete proof see [21].

Proposition 5.5

For any \(j \in {\mathbb {Z}}\setminus \{ 0 \}\), the sequence \(\Omega _k(j) = \Omega _0(j) + r_k(j)\), \(k \ge 1\) provided in Lemmata 5.35.4, and defined for any \(\omega \in {{\mathcal {O}}}\) converges to \(\Omega _\infty (j) = \Omega _0(j) + r_\infty (j)\) with \(|r_\infty (j) - r_k(j)|^{{\mathcal {O}}} \lesssim \delta _k\). Defining the Cantor set

$$\begin{aligned} \begin{aligned} {{{\mathcal {E}}}}_\infty&:= \Big \{ \omega \in {{\mathcal {O}}} : |\omega \cdot \ell + \Omega _\infty (j) - \Omega _\infty (j')| \ge \frac{2 \gamma |j^3 - j'^3|}{{\mathtt {d}}(\ell )}, \quad \forall (\ell , j, j') \ne (0, j, j) \Big \} \end{aligned}\nonumber \\ \end{aligned}$$
(5.57)

and

$$\begin{aligned} {{\mathcal {L}}}_\infty := \omega \cdot \partial _\varphi + {{\mathcal {D}}}_\infty , \quad {{\mathcal {D}}}_\infty := {{\mathrm{i}}}\, {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _\infty (j), \end{aligned}$$
(5.58)

one has \({{{\mathcal {E}}}}_\infty \subseteq \cap _{k \ge 0} {{{\mathcal {E}}}}_k\).

Defining also

$$\begin{aligned} {{\widetilde{\Phi }}}_k := \Phi _0 \circ \ldots \circ \Phi _k \quad \text {with inverse} \quad {{\widetilde{\Phi }}}_k^{- 1} = \Phi _k^{- 1} \circ \ldots \circ \Phi _0^{- 1}, \end{aligned}$$
(5.59)

the sequence \({{\widetilde{\Phi }}}_k\) converges for any \(\omega \in {{{\mathcal {E}}}}_\infty \) to a symplectic, invertible map \(\Phi _\infty \) w.r.t. the norm \(\Vert \cdot \Vert _{\rho _m - 2\zeta }^{{{{\mathcal {E}}}}_\infty }\) and \(\Vert \Phi _\infty ^{\pm 1} - {\mathrm{Id}} \Vert _{\rho _m - 2\zeta }^{{{{\mathcal {E}}}}_\infty } \lesssim \delta _0 \). Moreover for any \(\omega \in {{{\mathcal {E}}}}_\infty \), one has that \(\Phi _\infty ^{- 1} {{\mathcal {L}}}_0 \Phi _\infty = {{\mathcal {L}}}_\infty \).

5.3 Variations

We now consider an operator

$$\begin{aligned} \begin{aligned} {{\mathcal {L}}}_0^+&\equiv {{\mathcal {L}}}_0 (\lambda _1^+, \lambda _3^+, {{\mathcal {P}}}_0^+) = \omega \cdot \partial _\varphi + {{\mathcal {D}}}_0^+ + {{\mathcal {P}}}_0^+, \\ {{\mathcal {D}}}_0^+&:= \lambda _3^+ \partial _x^3 + \lambda _1^+ \partial _x = {{\mathrm{i}}}\,{\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _0^+(j), \\ \Omega _0^+(j)&:= - \lambda _3^+ j^3 + \lambda _1^+ j, \quad j \in {\mathbb {Z}}\setminus \{ 0 \}. \end{aligned} \end{aligned}$$
(5.60)

such that

$$\begin{aligned} |\lambda _1^+ - \lambda _1|^{{{\mathcal {O}}}^+},\,|\lambda _3^+ - \lambda _3|^{{{\mathcal {O}}}^+},\, \Vert {{\mathcal {P}}}_0^+ - {{\mathcal {P}}}_0 \Vert _{\rho _{m+1}}^{{{\mathcal {O}}}^+} \le \delta _{m+1} \end{aligned}$$
(5.61)

where \({{\mathcal {L}}}\), \(\lambda _1\), \(\lambda _3\), \({{\mathcal {P}}}_0\) are given in (5.21) and \({{{\mathcal {O}}}^+}\subseteq {{\mathcal {O}}}\). In other words, \({{\mathcal {L}}}_0^+\) is a small variation of \({{\mathcal {L}}}_0\) in (5.20) with also \(m\rightsquigarrow m+1\).

Of course we can apply Proposition 5.5 to \({{\mathcal {L}}}_0^+\); our aim is to compare the “final frequencies” of \({{\mathcal {L}}}_\infty ^+\) with those of \({{\mathcal {L}}}_\infty \).

To this aim, we first apply Lemma 5.3 with \({{\mathcal {L}}}_0\rightsquigarrow {{\mathcal {L}}}_0^+\) and \(\gamma \rightsquigarrow \gamma _+<\gamma \). In this way we obtain a sequence of sets \({{\mathcal {E}}}_k^+\subseteq {{\mathcal {E}}}_{k-1}^+\) and a sequence of symplectic maps \(\Phi _k^+\) defined for \(\omega \in {{\mathcal {E}}}_{k+1}^+\) such that setting \({{\mathcal {L}}}_0^+\) as in (5.60) and

$$\begin{aligned} {{\mathcal {L}}}_{k}:= \Phi _{k - 1}^{- 1} {{\mathcal {L}}}_{k - 1} \Phi _{k - 1}, \end{aligned}$$
(5.62)

one has

$$\begin{aligned} {{\mathcal {L}}}_k^+ := \omega \cdot \partial _\varphi + {{\mathcal {D}}}_k^+ + {{\mathcal {P}}}_k^+,\qquad k\le m+1, \end{aligned}$$
(5.63)

where

$$\begin{aligned} \begin{aligned}&{{\mathcal {D}}}_k^+ := {\mathrm{diag}}_{j \in {\mathbb {Z}}\setminus \{ 0 \}} \Omega _k^+(j), \quad \Omega _k^+(j) = \Omega _0^+ (j) + r_k^+(j) \end{aligned} \end{aligned}$$
(5.64)

The sets \({{\mathcal {E}}}_k^+\) are defined as \({{\mathcal {E}}}_0^+ := {{\mathcal {O}}^+}\) and for \(k \ge 1\)

$$\begin{aligned} \begin{aligned} {{\mathcal {E}}}_k^+&:= \Big \{\omega \in {{\mathcal {E}}}_{k - 1}^+ \ :\ |\omega \cdot \ell + \Omega _{k - 1}^+(j) - \Omega _{k - 1}^+(j')| \ge \frac{\gamma _+ |j^3 - j'^3|}{{\mathtt {d}}(\ell )}, \\&\qquad \forall (\ell , j, j') \ne (0, j, j), \quad |\ell _\eta | \le N_{k - 1} \Big \}. \end{aligned} \end{aligned}$$
(5.65)

Moreover one has \(\Phi _{k - 1}^+ := {\mathrm{exp}}(\Psi _{k - 1}^+)\), with

$$\begin{aligned} \Vert \Psi _{k - 1}^+ \Vert _{\rho _k}^{{{{\mathcal {E}}}}^+_k} \lesssim e^{C \Delta _{k-1}^{- \mu }} \delta _{k - 1} . \end{aligned}$$
(5.66)

The following lemma holds.

Lemma 5.6

For all \(k = 1, \ldots , m+1\) one has

$$\begin{aligned} \Vert {{\mathcal {P}}}_k^+ - {{\mathcal {P}}}_k \Vert _{\rho _{k}}^{{{\mathcal {E}}}_k \cap {{\mathcal {E}}}_k^+}&\le \delta _{m+1}, \end{aligned}$$
(5.67a)
$$\begin{aligned} |r_{k}^+(j) - r_k(j)|^{{{\mathcal {O}}}\cap {{\mathcal {O}}}^+}&\le \delta _{m+1} \end{aligned}$$
(5.67b)

and

$$\begin{aligned} \Vert \Psi _{k - 1}^+ - \Psi _{k - 1} \Vert _{\rho _k}^{{{\mathcal {E}}}_k \cap {{\mathcal {E}}}_k^+} \lesssim \delta _{m+1}, \end{aligned}$$
(5.68)

Proof

We procede differently for \(k=1,\ldots ,m\) and \(k=m+1\).

For the first case we argue by induction. Assume the statement to hold up to some \(k<m\). We want to prove

$$\begin{aligned} \Vert \Psi _k^+ - \Psi _k \Vert _{\rho _{k+1}}^{{{\mathcal {E}}}_{k + 1} \cap {{\mathcal {E}}}_{k + 1}^+} \le \delta _{m+1}. \end{aligned}$$
(5.69)

By Lemma 5.3, one has for \(\omega \in {{\mathcal {E}}}_{k + 1}^+\)

$$\begin{aligned} ( \Psi _k^+)_j^{j'} (\ell ) := {\left\{ \begin{array}{ll} \dfrac{({({{\mathcal {P}}}_k^+)}^{(k)})_j^{j'}(\ell )}{{{\mathrm{i}}}\big (\omega \cdot \ell + \Omega _{k}^+(j) - \Omega _k^+(j') \big )}, &{}\quad \forall (\ell , j, j') \ne (0, j, j), \quad |\ell |_\eta \le N_k, \\ 0 &{}\quad \text {otherwise}, \end{array}\right. }\nonumber \\ \end{aligned}$$
(5.70)

and direct calculation shows that for \(\omega \in {{\mathcal {E}}}_{k + 1} \cap {{\mathcal {E}}}_{k + 1}^+\), one has

$$\begin{aligned} \left| {(\Omega _k^+(j) - \Omega _k^+(j')) -(\Omega _k(j) - \Omega _k(j'))} \right| \le \delta _{m+1}|j^3-j'^3| \end{aligned}$$
(5.71)

and hence

$$\begin{aligned} \begin{aligned} | (\Psi _k^+)_j^{j'}(\ell ) - (\Psi _k)_j^{j'}(\ell )|^{ {{\mathcal {E}}}_{k + 1} \cap {{\mathcal {E}}}_{k + 1}^+}&\lesssim \delta _{m+1} {\mathtt {d}}(\ell )^3 |({{\mathcal {P}}}_k^{(k)})_j^{j'}(\ell )|^{ {{\mathcal {E}}}_{k + 1} \cap {{\mathcal {E}}}_{k + 1}^+}\\&\quad + {\mathtt {d}}(\ell )^2 |({{\mathcal {P}}}_k^{(k)})_j^{j'}(\ell ) - (({{\mathcal {P}}}_k^+)^{(k)})_j^{j'}(\ell )|^{ {{\mathcal {E}}}_{k + 1} \cap {{\mathcal {E}}}_{k + 1}^+} . \end{aligned}\nonumber \\ \end{aligned}$$
(5.72)

Therefore, reasoning as in (5.37)–(5.38), one uses Lemma A.6, the smallness condition (5.23) and the induction estimate (5.67a) so that (5.69) follows.

Now, from the definition of \(r_{k+1}\) in (5.39) it follows

$$\begin{aligned} |r_{k+1}^+(j) - r_{k+1}(j)|^{{{{\mathcal {E}}}_{k+1} \cap {{\mathcal {E}}}_{k+1}^+}} \le \delta _{m+1}, \end{aligned}$$
(5.73)

and by Kiszbraun Theorem applied to \(r_{k+1}^+(j) - r_{k+1}(j)\), (5.67b) holds.

The estimate of \({{\mathcal {P}}}_{k + 1}^+ - {{\mathcal {P}}}_{k + 1}\) follows by explicit computation the difference by using the expressions provided in (5.41), using the induction estimates (5.30), (5.67a), the estimate (5.69) and by applying Lemma A.12.

For \(k=m+1\) the proof can be repeated word by word, the only difference being that \(\Psi _m\) is defined in (5.52) while \(\Psi _m^+\) is defined in (5.36) with \(k=m\). \(\square \)

5.4 Conclusion of the Proof

To conclude the proof of Proposition 3.8 we start by noting that, setting \({{\mathcal {O}}}\) appearing in (5.2) as \({{\mathcal {O}}}^{(n)}\) appearing in (3.11), the operator \(L_{n+1}\) appearing in (3.18) with of course \(n\rightsquigarrow n+1\) is of the form (5.1) with

$$\begin{aligned} \begin{aligned}&\lambda _3=1+A_{n+1},\\&a_1^{(k)}(\varphi ,x) = B_{k+1}(\varphi ,x) - B_k(\varphi ,x) ,\\&a_0^{(k)}(\varphi ,x) = C_{k+1}(\varphi ,x) - C_k(\varphi ,x). \end{aligned} \end{aligned}$$

Moreover from (3.20) we have

$$\begin{aligned} \delta _k = \sigma _k^{-\tau _2}e^{{\mathtt {C}}\sigma _k^{-\mu }}\varepsilon _k,\qquad \rho _k = s_k - 3\sigma _k \end{aligned}$$

where \(s_k\), \(\sigma _k\) and \(\varepsilon _k\) are defined in (3.6), so that \(L_{n+1}\) satisfies (5.2) with \(m=n\). Thus, fixing

$$\begin{aligned} \zeta _k=\sigma _k,\qquad 2\zeta =\sigma _k, \end{aligned}$$

the smallness conditions (5.3) follows by definition. Hence we can apply Lemma 5.1 to \(L_{n+1}\) obtaining an operator of the form (5.5). In particular the conjugating operator \({{\mathcal {M}}}\) satisfies

$$\begin{aligned} \Vert {{\mathcal {M}}}- \mathtt { Id}\Vert ^{{{\mathcal {O}}}}_{s_n - 3\sigma _n}\lesssim \sigma _0^{-\tau _2}e^{{\mathtt {C}}\sigma _0^{-\mu }}\varepsilon _0. \end{aligned}$$

We are now in the setting of Sect. 5.2 with

$$\begin{aligned} \rho _k = s_k - 4\sigma _k,\qquad \delta _k=\sigma _k^{-\tau _3}e^{2{\mathtt {C}}\sigma _k^{-\frac{1}{\eta }+}}\varepsilon _k \end{aligned}$$

for some \(\tau _3>0\). A direct calculation shows that the smallness conditions (5.22), (5.23), (5.24), (5.49) are satisfied provided we choose \(N_k\) appropriately, so that we can apply Proposition 5.5.

In conclusion we obtain an operator \(M_{n+1} ={{\mathcal {M}}}\circ \Phi _\infty \) (recall that \({{\mathcal {M}}}\) is constructed in Lemma 5.1) satisfying (3.49), (3.50), where \(\Omega ^{(n+1)}(j) := \Omega _{\infty }(j)\) and \({{\mathcal {E}}}^{(n+1)} ={{\mathcal {E}}}_{\infty }\). Note that in particular the functions \(\Omega ^{(n+1)}(j)\) turn out to be of the form (3.46).

Finally (3.47) follows from Lemmata 5.2 and 5.6 where \({{\mathcal {L}}}_+\) has the role of \(L_{n+1}\) while \({{\mathcal {L}}}\) has the role of \(L_n\). This means that here we are taking \(m\rightsquigarrow n-1\).