Hostname: page-component-745bb68f8f-5r2nc Total loading time: 0 Render date: 2025-02-06T08:54:23.739Z Has data issue: false hasContentIssue false

On the accumulation of separatrices by invariant circles

Published online by Cambridge University Press:  17 November 2021

A. KATOK
Affiliation:
Department of Mathematics, CNRS UMR 8088, CY Cergy Paris Université (University of Cergy-Pontoise), 2, av. Adolphe Chauvin, F-95302Cergy-Pontoise, France
R. KRIKORIAN*
Affiliation:
Department of Mathematics, CNRS UMR 8088, CY Cergy Paris Université (University of Cergy-Pontoise), 2, av. Adolphe Chauvin, F-95302Cergy-Pontoise, France
Rights & Permissions [Opens in a new window]

Abstract

Let f be a smooth symplectic diffeomorphism of ${\mathbb R}^2$ admitting a (non-split) separatrix associated to a hyperbolic fixed point. We prove that if f is a perturbation of the time-1 map of a symplectic autonomous vector field, this separatrix is accumulated by a positive measure set of invariant circles. However, we provide examples of smooth symplectic diffeomorphisms with a Lyapunov unstable non-split separatrix that are not accumulated by invariant circles.

Type
Original Article
Copyright
© The Author(s), 2021. Published by Cambridge University Press

1 Introduction

A theorem by Herman,‘Herman’s last geometric theorem’, cf. [Reference Herman9, Reference Fayad and Krikorian12], asserts that if a smooth orientation- and area-preserving diffeomorphism f of the 2-plane ${\mathbb R}^2$ (or the 2-cylinder ${\mathbb R}/{\mathbb Z}\times {\mathbb R}$ ) admits a Kolmogorov–Arnold–Moser (KAM) circle $\Sigma $ (by definition, a smooth invariant curve, isotopic in ${\mathbb R}^2\smallsetminus \{o\}$ to a circle centered at the origin in the case $f:{\mathbb R}^2\to {\mathbb R}^2$ or isotopic to ${\mathbb R}/{\mathbb Z}\times \{0\}$ in the cylinder case, on which the dynamics of f is conjugated to a Diophantine translation), then this KAM circle is accumulated by other KAM circles, the union of which has positive two-dimensional Lebesgue measure in any neighborhood of $\Sigma $ . In this paper, we investigate whether such a phenomenon holds if, instead of being a KAM circle, the invariant set $\Sigma $ is a separatrix of a hyperbolic fixed (or periodic) point of f.

More precisely, we consider the following situation (see Figure 1). Let $f:{\mathbb R}^2\to {\mathbb R}^2$ , $f:(x,y)\mapsto f(x,y)$ , $f(0,0)=(0,0)$ be a smooth diffeomorphism which is symplectic with respect to the usual symplectic form $\omega =dx\wedge dy$ ( $f^*\omega =\omega $ ). We assume that $o:=(0,0)$ is a hyperbolic fixed point of f (the matrix $Df(o)\in SL(2,{\mathbb R})$ has distinct real eigenvalues) and that there exists an f-invariant compact connected set $\Sigma \ni o$ such that $\Sigma \smallsetminus \{o\}$ is a non-empty connected one-dimensional manifold included in both the stable and unstable manifolds $W^{s}_{f}(o)$ , $W^{u}_{f}(o)$ associated to o:

$$ \begin{align*}\text{for all }\ (x,y)\in \Sigma,\quad \lim_{n\to\pm\infty} f^n(x,y)=o.\end{align*} $$

Note that because o is f-hyperbolic, $\Sigma $ is homeomorphic to a circle and $\Sigma \smallsetminus \{o\}$ coincides with one of the two connected components of $W^s_{f}(o)\smallsetminus \{o\}$ (respectively $W^u_{f}(o)\smallsetminus \{o\}$ ). We shall say that $\Sigma $ is a separatrix of f associated to the hyperbolic fixed point o or, without referring to the hyperbolic fixed point o, that $\Sigma $ is a separatrix of f.

Figure 1 A (non-split) separatrix.

Examples of such diffeomorphisms f can be obtained in the following way. Let $X_{0}$ be a smooth autonomous Hamiltonian vector field of the form

(1.1) $$ \begin{align} X_{0}=J\nabla H_{0},\quad J=\begin{pmatrix} 0 & -1\\ 1& 0\end{pmatrix}\end{align} $$

where $H_{0}:{\mathbb R}^2\to {\mathbb R}$ , of the form

$$ \begin{align*} H_{0}(x,y)=\lambda x y+O^3(x,y),\quad \lambda\in{\mathbb R}^*, \end{align*} $$

(we can assume without loss of generality $\lambda>0$ ) is a smooth function. The time-1 map $f_{0}:=\phi _{X_{0}}^1$ of $X_{0}$ is a Hamiltonian (in particular, symplectic) diffeomorphism of ${\mathbb R}^2$ admitting o as a hyperbolic fixed point. We assume that it has a separatrix $\Sigma \ni o$ of the form

$$ \begin{align*}\Sigma\smallsetminus\{o\}=\{\phi^t_{X_{0}}(p),\ t\in{\mathbb R}\}\quad \text{for some}\ p\in {\mathbb R}^2\smallsetminus\{o\}\ \text{such that}\ \lim_{t\to\pm\infty}\phi^t_{X_{0}}(p)=o.\end{align*} $$

We now consider a smooth time-dependent Hamiltonian vector field $Y:{\mathbb R}/{\mathbb Z}\times {\mathbb R}^2\to {\mathbb R}$ , $(t,(x,y))\mapsto Y(t,x,y)$ which is 1-periodic in t, symplectic with respect to $(x,y)$ , and tangent to $\Sigma \smallsetminus \{o\}$ :

$$ \begin{align*}\text{for all } t\in{\mathbb R}/{\mathbb Z},\text{ for all } (x,y)\in \Sigma,\quad \det(X_{0}(x,y),Y(t,x,y))=0.\end{align*} $$

One can for example choose $Y(t,x,y)=J\nabla F(t,x,y)$ , where $F:{\mathbb R}/{\mathbb Z}\times {\mathbb R}^2\to {\mathbb R}$ is a smooth time-dependent Hamiltonian that satisfies

$$ \begin{align*}\text{for all } t\in{\mathbb R}/{\mathbb Z}, \text{ for all } (x,y)\in \Sigma,\quad F(t,x,y)=F(t,0,0).\end{align*} $$

Note that because o is a hyperbolic fixed point of $X_{0}$ , one has for all t, $Y(t,o)=0$ . For $\varepsilon \in {\mathbb R}$ , define the 1-periodic in t symplectic vector field ${\mathbb R}^2\to {\mathbb R}^2$ as

(1.2) $$ \begin{align} X_{\varepsilon}^t(x,y):=X_{\varepsilon}(t,x,y)=X_{0}(x,y)+\varepsilon Y(t,x,y).\end{align} $$

For $\varepsilon $ small enough, the time-0-to-1 map,

(1.3) $$ \begin{align} f_{\varepsilon}=\phi^{1,0}_{X_{\varepsilon}},\end{align} $$

of the symplectic vector field $X_{\varepsilon }$ is a symplectic diffeomorphism of ${\mathbb R}^2$ admitting o as a hyperbolic fixed point and still $\Sigma $ as a separatrix. (If $X(t,z)$ is a time dependent vector field, the time-s-to-t map of X is defined by $\phi ^{t,s}_{X}(z(s))=z(t)$ for any $z(\cdot )$ solution of $\dot z(t)=X(z(t))$ . When X is time independent, the notation $\phi ^t_{X}$ stands for $\phi ^{t,0}_{X}$ .) Note that $f_{\varepsilon }$ is a Hamiltonian diffeomorphism (for more details on Hamiltonian diffeomorphisms, see [Reference Polterovich16]).

Here is the analogue of the aforementioned last geometric theorem of Herman.

Theorem A. For any $r\in {\mathbb N}^*$ , there exists $\varepsilon _{r}>0$ such that, for any $\varepsilon \in\, ]\! -\varepsilon _{r},\varepsilon _{r}[$ , there exists a set of $f_{\varepsilon }$ -invariant $C^r$ KAM circles accumulating the separatix $\Sigma $ and which covers a set of positive Lebesgue measure of ${\mathbb R}^2$ in any neighborhood of $\Sigma $ .

Let us clarify some points made in the preceding statement.

By a $C^r$ circle, $r\geq 0$ , we mean a $C^r$ non-self-intersecting closed curve (or equivalently, if $r\geq 1$ , a non-empty compact connected one-dimensional $C^r$ submanifold of ${\mathbb R}^2$ ) which is isotopic in ${\mathbb R}^2\smallsetminus \{o\}$ to the separatrix $\Sigma $ . Such a set $\Gamma $ is invariant by $f_{\varepsilon }$ if $f_{\varepsilon }(\Gamma )=\Gamma $ .

We say that a set $\mathcal {G}$ of $f_{\varepsilon }$ -invariant circles accumulates the set $\Sigma $ if for any $\xi>0$ , the set of $\Gamma \in \mathcal {G}$ such that $\mathrm {dist}(\Gamma ,\Sigma )<\xi $ is not empty, where $\mathrm {dist}$ denotes the Hausdorff distance,

$$ \begin{align*}\mathrm{dist}(A,B)=\max\Big(\sup_{a\in A}d(a,B),\sup_{b\in B}d(b,A)\Big) \end{align*} $$

(here $d(x,C)=\inf _{c\in C}\|x-c\|_{{\mathbb R}^2}$ ).

The $f_{\varepsilon }$ -invariant circles obtained in Theorem A are KAM circles: the restrictions of $f_{\varepsilon }$ on each of these curves are $C^r$ circle diffeomorphisms that are conjugated to Diophantine translations. A real number $\alpha $ is Diophantine if there exist positive constants $\kappa ,\tau $ such that, for any $(p,q)\in {\mathbb Z}\times {\mathbb N}^*$ , $|\alpha -(p/q)|\geq \kappa /q^{\tau }$ . The constants $\tau $ and $\kappa $ are respectively called the exponent and the constant of the Diophantine condition. The set of Diophantine numbers with fixed exponent $\tau>2$ has full Lebesgue measure if the constant is not specified and positive measure if the constant is also fixed (and small). In our case, the exponent of the Diophantine condition can be chosen to be independent of $\varepsilon $ (it depends only on $\lambda $ ).

Remark 1.1. However, and this is a difference with the situation of Herman’s last geometric theorem, the constants of these Diophantine numbers are arbitrarily small. Moreover, as these circles accumulate the separatrix, their $C^2$ -norm must explode.

Remark 1.2. The phase space ${\mathbb R}^2$ can be replaced by the cylinder ${\mathbb R}/{\mathbb Z}\times {\mathbb R}$ in the statement of the main theorem.

The smallness condition in Theorem A is indeed necessary as shown by the following theorem.

Let $\Delta _{\Sigma }$ be the bounded connected component of ${\mathbb R}^2\smallsetminus \Sigma $ .

Theorem B. There exists a smooth symplectic diffeomorphism $f:{\mathbb R}^2\to {\mathbb R}^2$ admitting a separatrix $\Sigma $ which is included in an open set W of $\Sigma \cup \Delta _{\Sigma }$ that contains no f-invariant circle in $W\smallsetminus \Sigma $ .

The situation described in Theorems A and B is not generic. Indeed, as Poincaré discovered, in general, the stable and unstable manifolds of a hyperbolic fixed or periodic point of a symplectic map intersect transversally (one usually refers to this phenomenon as the splitting of separatrices), a fact that forces the dynamics of f to be ‘quite intricate’. This was Poincaré’s key argument in his proof of the fact that the Three-body problem in Celestial Mechanics does not admit a complete set of independent commuting first integrals. Later, Smale [Reference Smale18] showed that this splitting of separatrices has an even more striking consequence on the dynamics of f, namely the existence of a horseshoe, that is, a uniformly hyperbolic f-invariant compact set (locally maximal) with positive topological entropy and on which the dynamics of f is ‘chaotic’ (isomorphic to a two-sided shift). By a result of the first author [Reference Katok13], in this situation, positive topological entropy is indeed equivalent to the existence of a horseshoe. A consequence of the splitting of a separatrix is thus the existence of a Birkhoff instability zone (open region without invariant circles) in the vicinity of this split separatrix (see [Reference Herman11] for a detailed exposition on the topic). In some sense, Theorem A shows that in the perturbative situation of equations (1.2)–(1.3) ( $\varepsilon $ small enough), the splitting of separatrices is essentially the only mechanism responsible for the creation of instability zones. However, in a ‘non-perturbative’ situation, Theorem B points in the opposite direction. Figures 4 and 7 illustrate the role that plays the smallness assumption in Theorem A (or its absence in Theorem B).

1.1 On the proofs of Theorems A and B

As suggests Remark 1.1, the invariant circles of Theorem A cannot be obtained directly via a classical KAM approach. However, the existence of the (non-split) separatrix $\Sigma $ allows to associate to each diffeomorphism $f_{\varepsilon }$ a regular diffeomorphism $\mathring {f}_{\varepsilon }$ , defined on a standard open annulus and preserving a finite probability measure, to which one can apply Moser’s or Rüssmann’s invariant (or translated) curve theorem [Reference Moser15, Reference Rüssmann17] (see §6). The thus obtained invariant curves for $\mathring {f}_{\varepsilon }$ yield invariant curves for $f_{\varepsilon }$ . The construction of the diffeomorphism $\mathring {f}_{\varepsilon }$ is done as follows. We first make preliminary reductions involving some Birkhoff and symplectic Sternberg-like normal forms (§2) to have a control on the dynamics in some neighborhood of the hyperbolic fixed point o3). This allows us to define in §4 a first return map $\hat f_{\varepsilon }$ for $f_{\varepsilon }$ , in a fundamental domain $\mathcal {F}_{\varepsilon }$ , the boundaries of which can be glued together to obtain an open abstract cylinder (or annulus). This abstract cylinder can be uniformized to become a standard annulus and the first return map $\hat f_{\varepsilon }$ then becomes a regular diffeomorphism $\bar f_{\varepsilon }$ of a standard annulus (preserving some probability measure). This is done in §5. We call normalization (see §5.3) the uniformization operation and we say that $\bar f_{\varepsilon }$ is the renormalization of $f_{\varepsilon }$ . The term renormalization in this paper has the same acceptation as in the theories of circle diffeomorphisms, holomorphic germs, or quasi-periodic cocycles; cf. [Reference Avila and Krikorian6, Reference Krikorian14, Reference Yoccoz23, Reference Yoccoz, Marmi and Yoccoz24]. The dynamics of $\bar f_{\varepsilon }$ is closely related to that of $f_{\varepsilon }$ in the sense that the existence of invariant curves for $\bar f_{\varepsilon }$ translates into a similar statement for $f_{\varepsilon }$ (see §7). The renormalized diffeomorphism $\bar f_{\varepsilon }$ has a large twist (this is reminiscent of the hyperbolicity of $f_{\varepsilon }$ at o) and we are thus led to rescale it to obtain the aforementioned diffeomorphism $\mathring {f}_{\varepsilon }$ which is now a small $C^r$ -perturbation of an integrable twist map (this is where the smallness assumption of Theorem A appears) with a controlled twist (see §6). The proof of Theorem A is completed in §8.

To prove Theorem B (cf. §9), we construct a symplectic diffeomorphism f (named $f_{\mathrm {pert}}$ in that section) so that the associated renormalized diffeomorphism $\bar {f}$ has an orbit accumulating the boundary of the aforementioned annulus: this prevents the existence of $\bar {f}$ -invariant curves close to this boundary and therefore of f-invariant curves close to the separatrix $\Sigma $ .

We note that the authors of [Reference Treschev and Zubelevich21] introduce the ‘separatrix map’ constructed by a gluing construction to investigate the size of the instability zones. Our approach here, which is focused on a renormalization point of view, is different. The technique we use to prove Theorem A might be useful to study the dynamics of symplectic twist maps with zero topological entropy. That is, to which extent are they integrable? Angenent, [Reference Angenent, McGehee and Meyer1], proves they are $C^0$ -integrable in the sense that, for any rotation number, one can find a $C^0$ -invariant curve with this rotation number. Can one prove $C^k$ -integrability? The word ‘integrable’ is meant in a broad sense. Additionally, the construction of Theorem B might give a hint to provide examples of smooth twist maps admitting isolated invariant circles with irrational rotation number (if they exist, these curves bound two instability zones). A modification of the example of Theorem B yields examples of such isolated invariant curves with rational rotation numbers. For the existence of curves with irrational rotation number in low regularity and related results, see [Reference Arnaud2Reference Avila and Fayad5].

2 Normal forms

The main result of this section is the following Sternberg-like symplectic normal form theorem (Proposition 2.1) that will allow us in §3 to control the long-time dynamics of $f_{\varepsilon }$ in a neighborhood of the hyperbolic point o. This will be useful when we shall define first return maps for $f_{\varepsilon }$ in a convenient fundamental domain, see §4.

Let $f_{\varepsilon }$ be defined by (1.2) and (1.3).

Proposition 2.1. For any $k\in {\mathbb N}^*$ large enough, there exists $\varepsilon _{k}>0$ for which the following holds. There exist a smooth family $(q_{\varepsilon ,k})_{\varepsilon \in I}$ ( $I\ni 0$ some open interval of ${\mathbb R}$ ) of polynomials $q_{\varepsilon ,k}(s)=\lambda s+O(s^2)\in {\mathbb R}[s]$ and a continuous family $(\Theta _{\varepsilon ,k})_{\varepsilon \in I}$ of symplectic $C^k$ -diffeomorphism of ${\mathbb R}^2$ such that $\Theta _{\varepsilon ,k}(o)=o$ , $D\Theta _{\varepsilon ,k}(o)=\mathrm {id}$ , and on a neighborhood $V_{k}$ of o one has, provided $\varepsilon \in\, ]\!-\!\varepsilon _{k},\varepsilon _{k}[$ :

(2.4) $$ \begin{align}\text{on }\ V_{k},\quad f_{\varepsilon,k}&\mathop{=}_{\mathrm{defin.}}\Theta_{\varepsilon,k}\circ f_{\varepsilon}\circ \Theta_{\varepsilon,k}^{-1} \qquad\qquad \end{align} $$
(2.5) $$ \begin{align} & \ \quad\qquad\qquad\qquad\qquad\qquad\qquad=\phi^1_{J\nabla Q_{\varepsilon,k}}\quad \text{where }\ Q_{\varepsilon,k}(x,y)=q_{\varepsilon,k}(xy) \end{align} $$

and

(2.6) $$ \begin{align}\text{on }\ V_{k},\quad (\Theta_{0,k})_{*}X_{0}=J\nabla Q_{0,k}.\end{align} $$

Note that o is still a hyperbolic fixed point of $f_{\varepsilon ,k}$ and that

$$ \begin{align*}\Sigma_{\varepsilon,k}:=\Theta_{\varepsilon,k}(\Sigma)\end{align*} $$

is still a separatrix for $f_{\varepsilon ,k}$ .

2.1 Reduction of Theorem A to Theorem 2.1

After applying Proposition 2.1, we are thus left with a family $(f_{\varepsilon ,k})$ of $C^k$ - symplectic diffeomorphisms, each $f_{\varepsilon ,k}$ being conjugated to $f_{\varepsilon }$ and admitting a separatrix $\Sigma _{\varepsilon ,k}$ . Because the conclusions of Theorem A are clearly invariant by conjugation, to prove Theorem A, we just need to prove that if $k\geq r$ and $\varepsilon $ is small enough, each separatrix $\Sigma _{\varepsilon ,k}$ is accumulated by a set of positive measure of KAM circles for $f_{\varepsilon ,k}$ . This is the content of Theorem 2.1 below that we shall apply to the family of $C^k$ -diffeomorphisms $f_{\varepsilon ,k}$ defined by (1.2), (1.3), and (2.4), but that holds for any family (that we still denote in what follows by $(f_{\varepsilon })_{\varepsilon \in I}$ to alleviate the notations) of symplectic $C^k$ -diffeomorphisms satisfying the following hypothesis.

Let $(f_{\varepsilon })_{\varepsilon \in I}$ , ( $I\ni 0$ open interval of ${\mathbb R}$ ) be a family of $C^k$ -symplectic diffeomorphisms of ${\mathbb R}^2$ that satisfies:

  1. (H1) each $f_{\varepsilon }$ has a (non-split) separatrix $\Sigma _{\varepsilon }$ associated to the hyperbolic point o;

  2. (H2) the map $I\ni \varepsilon \to f_{\varepsilon }-\text {id}\in C^k({\mathbb R}^2,{\mathbb R}^2)$ is continuous (the norm on $C^k$ is the usual $C^k$ -norm);

  3. (H3) on some neighborhood V of o, each $f_{\varepsilon }$ coincides with the time-1 map of a symplectic vector field $J\nabla Q_{\varepsilon }(x,y)$ where $Q_{\varepsilon }(x,y)=q_{\varepsilon }(xy)$ , $q_{\varepsilon }\in C^{k+1}({\mathbb R}^2)$

    $$ \begin{align*}q_{\varepsilon}(t)=\lambda t+O^2(t),\quad \lambda>0;\end{align*} $$
  4. (H4) on ${\mathbb R}^2$ , $f_{0}=\phi ^1_{X_{0}}$ , where $X_{0}=J\nabla H_{0}$ is a Hamiltonian vector field that coincides with $J\nabla Q_{0}$ on V.

Remark 2.1. On $V_{}$ , the orbits of $f_{\varepsilon }{}_{|\ V}= \phi ^1_{J\nabla Q_{\varepsilon }}$ are pieces of hyperbolae $\{xy=\mathrm {constant}\}$ (condition (H3)).

When $\varepsilon =0$ , for any $z\in \{xy=c\}\cap V$ , $N\in {\mathbb Z}$ such that $f_{0}^N(z)\in V$ , one has $f_{0}^N(z)\in \{xy=c\}\cap V$ (condition (H4)).

Remark 2.2. The intersection $\Sigma _{\varepsilon }\cap V_{}$ is the union

$$ \begin{align*}\Sigma_{\varepsilon}\cap V_{}=(W^s_{f_{\varepsilon}}(o)\cap V_{})\cup (W^{u}_{f_{\varepsilon}}(o)\cap V_{})\end{align*} $$

and

$$ \begin{align*}W^{s}_{f_{\varepsilon}}(o)\cap V_{}=({\mathbb R}\times\{0\}) \cap V_{},\quad W^{u}_{f_{\varepsilon}}(o)\cap V_{}=(\{0\}\times {\mathbb R}) \cap V_{}.\end{align*} $$

One then has the following.

Theorem 2.1. There exists $k_{0}\in {\mathbb N}$ for which the following holds. Let $k\geq k_{0}+2$ and let $(f_{\varepsilon })_{\varepsilon \in I}$ be a family of $C^k$ -symplectic diffeomorphisms of ${\mathbb R}^2$ satisfying the previous conditions (H1)–(H4). Then, there exists $\varepsilon _{1}>0$ such that, for any $\varepsilon \in\, ]\!-\!\varepsilon _{1},\varepsilon _{1}[$ , the diffeomorphism $f_{\varepsilon }$ admits a set of positive Lebesgue measure of invariant $C^{k-k_{0}-2}$ -circles in any neighborhood of the separatrix $\Sigma _{\varepsilon }$ .

Moreover, if $k-k_{0}-2\geq k_{1}$ ( $k_{1}$ depending only on $\lambda $ ), these circles are KAM circles.

We shall give the proof of Theorem 2.1 in §8.

The proof of Proposition 2.1 occupies the rest of this section. It will be based on a first reduction obtained by performing some steps of Birkhoff normal forms (Proposition 2.3) and then the application of various Sternberg-like normal forms (Corollary 2.4 and Proposition 2.5).

2.2 Birkhoff normal form for the time-periodic vector field $X_{\varepsilon }^t$

A preliminary step in Sternberg’s classical linearization theorem is to first conjugate the considered system (diffeomorphism or vector field) defined in the neighborhood of the hyperbolic fixed point o to a system which is tangent to an integrable model to some high enough order. This is what we do in this subsection and in a symplectic framework (see Proposition 2.3) by using Birkhoff normal form techniques.

2.2.1 Periodically forced vector fields

Let $X:{\mathbb R}\times {\mathbb R}^2\ni (t,x)\mapsto X^{t}(z):= X(t,z)\in {\mathbb R}^2$ be a smooth time-dependent symplectic vector field: for each t, the 1-form $i_{X_{t}}\omega $ is closed (and hence locally exact). For $t,s\in {\mathbb R}$ , we denote by $\phi ^{t,s}_{X}$ the flow of X between times s and t when it is defined (see page 3 for the definition of $\phi ^{t,s}_{X}$ ). If $t\mapsto g^{t}(\cdot )$ is a one-parameter family of symplectic diffeomorphisms, one has

(2.7) $$ \begin{align} g^{t}\circ \phi_{X}^{t,s}\circ (g^{s})^{-1}=\phi^{t,s}_{\tilde X},\end{align} $$

where $\tilde X:(t,z)\mapsto \tilde X^{t}(z):=\tilde X(t,z)$ is the smooth time-dependent symplectic vector field

(2.8) $$ \begin{align}\tilde X^{t}=\partial_{t}g^{t}\circ (g^{t})^{-1}+(g^{t})_{*}X^{t}. \end{align} $$

Conversely, if (2.8) is satisfied, then so is (2.7). Note that if $g^t$ depends 1-periodically on t, then (2.7) yields the more classical conjugation equation

$$ \begin{align*}g\circ \phi^{1,0}_{X}\circ g^{-1}=\phi_{\tilde X}^{1,0}, \end{align*} $$

where $g=g^0=g^1$ ( $g^t$ is 1-periodic in t).

Assume now that $X^{t}$ depends 1-periodically in t and, in a smooth way, on a small parameter $\varepsilon \in {\mathbb R}$ ; we furthermore assume that it is of the form

(2.9) $$ \begin{align} X_{\varepsilon}^t(z)=J\nabla H_{\varepsilon}^t(z),\end{align} $$

where ( $z=(z_{1},z_{2})\in {\mathbb R}^2$ )

(2.10) $$ \begin{align} H_{\varepsilon}^{t}(z)=\lambda_{\varepsilon}(t)z_{1}z_{2}+O^3(z),\quad \int_{{\mathbb T}}\lambda_{\varepsilon}(t)\,dt>0,\quad \lambda_{0}(t)=\lambda\in{\mathbb R}^*_{+},\end{align} $$

$H_{\varepsilon }:{\mathbb R}/{\mathbb Z}\times {\mathbb R}^2\to {\mathbb R}$ , $H_{\varepsilon }:(t,z)\mapsto H_{\varepsilon }(t,z):=H_{\varepsilon }^{t}(z)$ being a smooth function. Assume also that, for some $j\in {\mathbb N}^*$ ,

$$ \begin{align*}g^t_{\varepsilon}(z)=\phi^{1}_{J\nabla G_{\varepsilon}^t}(z)=\text{id}+O^j(z),\quad G_{\varepsilon}^t(z)=O^{j+1}(z),\end{align*} $$

where $G:I\times {\mathbb R}/{\mathbb Z}\times ({\mathbb R}^2,o)\ni (\varepsilon ,t,z)\mapsto G_{\varepsilon }(t,z):=G_{\varepsilon }^{t}(z)\in {\mathbb R}$ is a smooth function. Then, one has

$$ \begin{align*}\partial_{t}g_{\varepsilon}^{t}\circ (g_{\varepsilon}^{t})^{-1}=J\nabla \partial_{t}G_{\varepsilon}^t+O^{j+1}(z), \end{align*} $$
$$ \begin{align*}(g_{\varepsilon}^{t})_{*}X_{\varepsilon}^{t}=J\nabla H_{\varepsilon}^t\circ (g_{\varepsilon}^{t})^{-1}=J\nabla H_{\varepsilon}^{t}+J\nabla\{G_{\varepsilon}^{t},H_{\varepsilon}^t\}+O^{j+1}(z), \end{align*} $$

(here $\{A,B\}$ denotes the Poisson bracket $\{A,B\}={\langle }\nabla A,J\nabla B{\rangle } $ ) so that $\tilde X_{\varepsilon }^t$ defined by (2.8) is of the form

(2.11) $$ \begin{align} \tilde X_{\varepsilon}^t= J\nabla \tilde H_{\varepsilon}^t,\end{align} $$

with

(2.12) $$ \begin{align}\tilde H_{\varepsilon}^t&=H_{\varepsilon}^t+\partial_{t}G_{\varepsilon}^{t}+\{G_{\varepsilon}^{t},H_{\varepsilon}^{t}\}+O^{j+2}(z) \end{align} $$
(2.13) $$ \begin{align} & \kern1.65pc =H_{\varepsilon}^t+\partial_{t}G_{\varepsilon}^t+\{G_{\varepsilon}^t,H_{2,\varepsilon}^t\}+O^{j+2}(z), \end{align} $$

where we have denoted $H_{2,\varepsilon }^t(z_{1},z_{2})=\lambda _{\varepsilon }(t)z_{1}z_{2}$ .

If in the preceding equation, one chooses $G^{t}_{\varepsilon }=G_{\varepsilon ,2}^t$ with $G_{\varepsilon ,2}^t(z)=a_{\varepsilon ,0}(t)z_{1}z_{2}$ , where $a_{\varepsilon ,0}$ is the 1-periodic function defined by

$$ \begin{align*}a_{\varepsilon,0}(t)=-\int_{0}^t\bigg(\lambda_{\varepsilon}(s)-\int_{{\mathbb T}}\lambda_{\varepsilon}(u)\,du\bigg)\,ds, \end{align*} $$

one has

$$ \begin{align*} \tilde H_{\varepsilon}^t(z)=\bar\lambda_{\varepsilon}z_{1}z_{2}+O^3(z),\end{align*} $$

where $\bar \lambda _{\varepsilon }=\int _{{\mathbb T}}\lambda _{\varepsilon }(t)\,dt$ . In other words, performing a change of coordinates (2.8) on $X_{\varepsilon }^t$ with $g^t_{\varepsilon }=g_{\varepsilon ,2}^t=\phi ^{1}_{J\nabla G^t_{\varepsilon ,2}}$ , we can assume that in (2.10), $\lambda _{\varepsilon }(t)$ does not depend on t:

(2.14) $$ \begin{align} H_{\varepsilon}^t(z)=\lambda_{\varepsilon}z_{1}z_{2}+O^3(z),\quad \lambda_{\varepsilon}\in{\mathbb R}^*_{+}\end{align} $$

(we write $\lambda _{\varepsilon }$ in place of $\bar \lambda _{\varepsilon }$ ).

2.2.2 Birkhoff normal form

Having put $H_{\varepsilon }^t$ under the form (2.14), we now eliminate by successive conjugations (2.8) non-diagonal higher-order terms in z from $H_{\varepsilon }^t$ (note that they depend on t).

The following lemma describes this elimination procedure.

Lemma 2.2. Let $j\in {\mathbb N}$ , $j\geq 2$ . Assume that, for some polynomials $q_{\varepsilon }(s)=\lambda s +O(s^2)\in {\mathbb R}[s]$ of degree $\leq [ j/2]$ depending smoothly on $\varepsilon $ ,

$$ \begin{align*}H_{\varepsilon}^t(z)=q_{\varepsilon}(z_{1}z_{2})+O^{j+1}(z).\end{align*} $$

Then, there exist a smooth family $(\tilde q_{\varepsilon })_{\varepsilon }$ of polynomials $\tilde q_{\varepsilon }(s)=\lambda s+O(s^2)\in {\mathbb R}[s]$ of degree $\leq [(j+1)/2]$ and a smooth family of smooth maps $G_{\varepsilon }:{\mathbb R}/{\mathbb Z}\times ({\mathbb R}^2,o)\ni (t,z)\to G_{\varepsilon }(t,z)=G^t_{\varepsilon }(z)\in {\mathbb R}^2$ such that on a neighborhood of o,

(2.15) $$ \begin{align} \begin{cases}G_{\varepsilon}^t(z)=O^{j+1}(z),\\ H_{\varepsilon}^t(z)+\partial_{t}G_{\varepsilon}^t(z)+\{G_{\varepsilon}^t,H_{\varepsilon}^t\}(z)=\tilde q_{\varepsilon}(z_{1}z_{2})+O^{j+2}(z).\end{cases}\end{align} $$

Moreover, if for $\varepsilon =0$ , $H_{0}^t$ does not depend on t, one can choose $G_{0}^t$ to be independent of t.

Proof. See Appendix A.

Let now $X_{\varepsilon }^{t}$ be the family of vector fields of (1.2).

Proposition 2.3. For any $N\geq 1$ there exist an open neighborhood $V_{N}$ of o, a smooth two-parameters family $(b_{\varepsilon }^t)_{\varepsilon \in I,t\in {\mathbb R}/{\mathbb Z}}$ (I some open interval containing 0) of smooth symplectic diffeomorphisms satisfying $b_{\varepsilon }^{t}(o)=o$ , $Db_{\varepsilon }^t(o)=\mathrm {id}$ and a smooth family of polynomials $q_{\varepsilon ,N}(s)=\lambda s+O(s^2)$ of degree $\leq [(N+1)/2]$ , such that, for any $\varepsilon \in I$ , $t\in {\mathbb R}/{\mathbb Z}$ , $(x,y)\in V_{N}$ one has

$$ \begin{align*}X_{\varepsilon}^{(1),t }&\mathop{=}_{\mathrm{defin.}}\ (\partial_{t}b^t_{\varepsilon})\circ (b^t_{\varepsilon})^{-1}+(b^t_{\varepsilon})_{*}X_{\varepsilon}^t\\ &\ \, =J\nabla Q_{\varepsilon,N}+O^{N+1}(x,y) \quad \text{with }Q_{\varepsilon,N}(x,y)=q_{\varepsilon,N}(xy),\end{align*} $$

and for $\varepsilon =0$ , $b^t_{0}$ is independent of t.

Proof. Applying the preceding Lemma 2.2 and relations (2.8)–(2.12) inductively (starting from (2.14)), we thus construct polynomials $q_{\varepsilon ,j}$ of degree $\leq [ j/2]$ ( $j\geq 2$ ) and functions $G_{\varepsilon ,j}^t=O^{j+1}(z)$ such that if one defines

$$ \begin{align*}b^t_{\varepsilon}=g^t_{\varepsilon,N}\circ\cdots\circ g^t_{\varepsilon,2}=\text{id}+O^2(z),\quad g^t_{\varepsilon,j}=\phi^1_{J\nabla G^t_{\varepsilon,j}}=\text{id}+O^{j}(z),\end{align*} $$

one has

$$ \begin{align*}\tilde X_{\varepsilon}^{t}:&=\partial_{t}b_{\varepsilon}^{t}\circ (b_{\varepsilon}^{t})^{-1}+(b_{\varepsilon}^{t})_{*}X_{\varepsilon}^{t}\\ &=J\nabla Q_{\varepsilon,N}+O^{N+1}(z)\quad \text{with } Q_{\varepsilon,N}(z)=q_{\varepsilon,N}(z_{1}z_{2}), \end{align*} $$

all depending on $\varepsilon $ being smooth. Moreover, if $X_{0}^t$ is independent of t, the diffeomorphism $b_{0}^t$ is independent of t.

Remark 2.3. Note that because $b_{0}^t\equiv b_{0}$ is independent of t, the vector field

$$ \begin{align*}X_{0}^{(1)}=(b_{0})_{*}X_{0}\end{align*} $$

is autonomous.

2.3 Symplectic Sternberg theorem for the autonomous vector field $X_{0}^{(1)}$

We shall need a symplectic version of the famous theorem by S. Sternberg (on smooth linearization of hyperbolic germs of smooth vector fields, see [Reference Sternberg19]), as proved in [Reference Banyaga, de la Llave and Wayne7] or [Reference Chaperon8] (see also [Reference Sternberg20]). We follow here the exposition of [Reference Banyaga, de la Llave and Wayne7].

Let $Z_{i}$ , $i=1,2$ , be two symplectic smooth autonomous vector fields such that, for some $\lambda \in {\mathbb R}^*$ and $N\in {\mathbb N}$ , one has

(2.16) $$ \begin{align}\begin{cases} Z_{i}(x,y)=-\lambda x\dfrac{\partial}{\partial x}+\lambda y\dfrac{\partial}{\partial y}+O^2(x,y)\quad (i=1,2),\\[9pt] Z_{1}(x,y)-Z_{2}(x,y)=O^{N+1}(x,y). \end{cases} \end{align} $$

Theorem 2.2. [Reference Banyaga, de la Llave and Wayne7, Theorem 1.2]

There exist positive constants $A,B$ for which the following holds. Let $m\in {\mathbb N}^*$ large enough and $N=[(m+B)/A]+1\geq 1$ . If (2.16) is satisfied, then there exists a $C^{m}$ symplectic change of coordinates such that on a neighborhood of o,

(2.17) $$ \begin{align} \begin{cases}(S_{0})_{*}Z_{1}=Z_{2},\\ S_{0}(o)=0,\qquad \ \ DS_{0}(o)=\mathrm{id}.\end{cases} \end{align} $$

We apply the preceding theorem to the case $Z_{1}=X_{0}^{(1)}$ and $Z_{2}=J\nabla Q_{0,N}$ ( $X_{0}^{(1)}$ , $Q_{0,N}$ given by Proposition 2.3 when $\varepsilon =0$ ). In view of Proposition 2.3, the condition (2.16) is satisfied and we hence get a symplectic diffeomorphism $S_{0}$ satisfying $S_{0}(o)=0$ , $DS_{0}(o)=\text {id}$ , and such that on a neighborhood of o,

$$ \begin{align*}(S_{0})_{*}X_{0}^{(1)}=J\nabla Q_{0,N}.\end{align*} $$

For each value of $t\in {\mathbb R}/{\mathbb Z}$ and $\varepsilon \in I$ , the diffeomorphism $(S_{0}\circ b^t_{\varepsilon })$ fixes the origin and its derivative at the origin is the identity. It can thus be extended as a symplectic $C^{m}$ -diffeomorphism $R^t_{\varepsilon }$ of ${\mathbb R}^2$ (cf. Lemma B.1). Note that the dependence of $R_{\varepsilon }^t$ with respect to t is smooth and 1-periodic ( $t\in {\mathbb R}/{\mathbb Z}$ ). We now define on ${\mathbb R}/{\mathbb Z}\times {\mathbb R}^2$ the time-periodic vector field $X_{\varepsilon }^{(2)}:(t,(x,y))\in ({\mathbb R}/{\mathbb Z})\times {\mathbb R}^2\to {\mathbb R}^2$ by

(2.18) $$ \begin{align} X_{\varepsilon}^{(2),t}\mathop{=}_{\mathrm{defin.}}(\partial_{t}R^t_{\varepsilon})\circ (R^t_{\varepsilon})^{-1}+(R^t_{\varepsilon})_{*}X_{\varepsilon}^t,\end{align} $$

and we observe that on a neighborhood of o,

$$ \begin{align*}(R^t_{0})_{*}X_{0}=X_{0}^{(2)}=J\nabla Q_{0,N}.\end{align*} $$

Because the conjugacy relation (2.18) is equivalent to (see §2.2.1)

$$ \begin{align*}\text{for all } t,s,\quad R^t_{\varepsilon}\circ \phi^{t,s}_{X_{\varepsilon}}\circ (R_{\varepsilon}^s)^{-1}=\phi^{t,s}_{X^{(2)}_{\varepsilon}},\end{align*} $$

we get by taking $t=1$ , $s=0$ , and setting $R_{\varepsilon }:=R^1_{\varepsilon }=R^0_{\varepsilon }$ , the following corollary.

Corollary 2.4. If $m\in {\mathbb N}^*$ is large enough and $N=[(m+B)/A]+1$ , there exists a smooth family $(R_{\varepsilon })$ of $C^m$ symplectic diffeomorphisms of ${\mathbb R}^2$ such that $R_{\varepsilon }(o)=o$ , $DR_{\varepsilon }(o)=\mathrm {id}$ , and on a neighborhood of o,

(2.19) $$ \begin{align}f_{\varepsilon}^{(1)}:&\mathop{=}_{\mathrm{defin.}}R_{\varepsilon}\circ f_{\varepsilon}\circ (R_{\varepsilon})^{-1} \end{align} $$
(2.20) $$ \begin{align} &\quad \ \, \kern3pc =\phi^1_{J\nabla Q_{\varepsilon,N}}+O^{N+1}(x,y).\end{align} $$

Moreover,

(2.21) $$ \begin{align} (R_{0})_{*}X_{0}=J\nabla Q_{0,N}.\end{align} $$

Note that the last equation shows that

(2.22) $$ \begin{align} f_{0}^{(1)}=\phi^1_{J\nabla Q_{0,N}}.\end{align} $$

2.4 Symplectic Sternberg normal form for the diffeomorphism $f_{\varepsilon }^{(1)}$

Theorem 2.2 has a version for smooth germs of symplectic diffeomorphisms which are hyperbolic at the origin. This is theorem 1.1 of [Reference Banyaga, de la Llave and Wayne7]. In our paper, we shall need a parametric version of that result, which is not explicitly stated in [Reference Banyaga, de la Llave and Wayne7] but that can be checked after close examination of the proof.

Proposition 2.5. There exist constants $A_{1},B_{1}$ depending on $\lambda \in {\mathbb R}^*$ such that the following holds. Let $m\in {\mathbb N}^*$ large enough and $N=[m/2]-3$ . If $(g_{1,\varepsilon })_{\varepsilon \in I}$ and $(g_{2,\varepsilon })_{\varepsilon \in I}$ ( $I\ni 0$ some open interval of ${\mathbb R}$ ) are two continuous (with respect to $\varepsilon \in I$ ) families of $C^m$ symplectic diffeomorphisms such that

(2.23) $$ \begin{align} \begin{cases}\text{for all } \varepsilon,\quad g_{1,\varepsilon}(o)=g_{2,\varepsilon}(o)=o,\\[-2pt] Dg_{1,0}(o)=\mathrm{diag}(\lambda,\lambda^{-1}) \quad \mbox{(is hyperbolic)},\\[-2pt] g_{1,\varepsilon}(x,y)=g_{2,\varepsilon}(x,y)+O^{N+1}(x,y),\\[-2pt] g_{1,0}=g_{2,0}, \end{cases}\end{align} $$

then, there exists a continuous family $(S^{(1)}_{\varepsilon })_{\varepsilon }$ (with respect to $\varepsilon \in I$ small enough) of $C^k$ symplectic diffeomorphisms such that $S^{(1)}_{\varepsilon }(o)=o$ , $DS^{(1)}_{\varepsilon }(o)=\mathrm {id}$ with $k=[NA_{1}-B_{1}]-1$ , and

$$ \begin{align*}\begin{cases}S^{(1)}_{\varepsilon}\circ g_{1,\varepsilon}\circ (S^{(1)}_{\varepsilon})^{-1}=g_{2,\varepsilon},\\[-2pt] S^{(1)}_{0}=\mathrm{id}.\end{cases}\end{align*} $$

2.5 Proof of Proposition 2.1

It will be a consequence of Corollary 2.4 and Proposition 2.5.

We first choose N so that $k=[NA_{1}-B_{1}]-1$ and we define m by $N=[(m+B)/A]+1$ . If k is large enough, m will satisfy the assumption of Corollary 2.4. We then apply Proposition 2.5 to $g_{1,\varepsilon }=f_{\varepsilon }^{(1)}$ , $g_{2,\varepsilon }=\phi ^1_{J\nabla Q_{\varepsilon ,N}}$ , which satisfies (2.23) (note that (2.20) is satisfied). This provides us with a continuous family $(S_{\varepsilon }^{(1)})_{\varepsilon }$ of $C^k$ symplectic diffeomorphisms defined in a fixed neighborhood of o such that $S^{(1)}_{\varepsilon }(o)=o$ , $DS^{(1)}_{\varepsilon }(o)=\text {id}$ , and on a neighborhood of o,

(2.24) $$ \begin{align} \begin{cases}S^{(1)}_{\varepsilon}\circ f_{\varepsilon}^{(1)}\circ (S^{(1)}_{\varepsilon})^{-1}=\phi^1_{J\nabla Q_{\varepsilon,N}},\\[-2pt] S^{(1)}_{0}=\text{id}.\end{cases}\end{align} $$

We can extend these $S^{(1)}_{\varepsilon }$ as symplectic $C^k$ diffeomorphisms $S^{(2)}_{\varepsilon }$ of ${\mathbb R}^2$ which depend continuously on $\varepsilon $ (cf. Lemma B.1). We then define

$$ \begin{align*}\Theta_{\varepsilon,k}=S_{\varepsilon}^{(2)}\circ R_{\varepsilon}\end{align*} $$

and we observe that on a neighborhood of o,

$$ \begin{align*}\begin{cases} \Theta_{\varepsilon,k}\circ f_{\varepsilon}\circ \Theta_{\varepsilon,k}^{-1}=\phi^1_{J\nabla Q_{\varepsilon,N}},\\ (\Theta_{{0},k})_{*}X_{0}=J\nabla Q_{0,N}; \end{cases}\end{align*} $$

indeed, the first equality comes from (2.19 and the first equation of (2.24), while the second is a consequence of (2.21) and the second equation of (2.24).

To conclude the proof, we rename $q_{\varepsilon ,N}$ , $Q_{\varepsilon ,N}$ as $q_{\varepsilon ,k}$ , $Q_{\varepsilon ,k}$ .

Note: From now on, and until the end of §8, we shall work in the setting of Theorem 2.1 with a family of $C^k$ symplectic diffeomorphisms satisfying conditions (H1)–(H4).

3 Dynamics in a neighborhood of the origin

The purpose of this section is to estimate the time spent by the orbits of the flow $\Phi ^t_{J\nabla Q_{\varepsilon }}$ in the neighborhood $V_{}$ of the hyperbolic point o.

To do that, we perform one more change of coordinates.

Let us define the following diffeomorphisms $\Xi _{1},\Xi _{2}$

(3.25) $$ \begin{align}\begin{cases} \text{for all } (x,y)\in\ {\mathbb R}^*_{+}\times{\mathbb R},\quad \Xi_{1}(x,y)=(\ln x,xy),\\ \text{for all } (x,y)\in\ {\mathbb R}\times{\mathbb R}^*_{+}, \quad \Xi_{2}(x,y)=(-\ln y,xy). \end{cases} \end{align} $$

Because $d(\ln x)\wedge d(xy)= d(-\ln y)\wedge d(xy)=dx\wedge dy$ , we see that $\Xi _{i}$ , $i=1,2$ , are symplectic.

Let $I_{1},I_{2}\subset {\mathbb R}^*_{+}$ be some open intervals such that $I_{1}\times \{0\}$ and $\{0\}\times I_{2}$ are both contained in $V_{}$ .

Lemma 3.1. Let $(x_{*},y_{*})\in (I_{1}\times {\mathbb R})\cap V_{}$ and $\bar t_{I_{2}}(x_{*},y_{*})=\inf \{t>0:\phi _{J\nabla Q_{\varepsilon }}^t(x_{*},y_{*})\in ({\mathbb R}\times I_{2})\cap V_{}\}$ . Then the following hold.

  1. (1) There exists $c(I_{1},I_{2})\geq 1$ such that if $0<x_{*}y_{*}\lesssim _{I_{1},I_{2},\lambda } 1$ , one has

    (3.26) $$ \begin{align} c(I_{1},I_{2})^{-1} \frac{|\!\ln (x_{*}y_{*})|}{\lambda}\leq \bar t_{I_{2}}(x_{*},y_{*})\leq c(I_{1},I_{2}) \frac{|\!\ln (x_{*}y_{*})|}{\lambda}. \end{align} $$
  2. (2) For any $(x,y)$ in a neighborhood of $(x_{*},y_{*})$ and any t in a neighborhood of $\bar t_{I_{2}}(x_{*},y_{*})$ ,

    (3.27) $$ \begin{align} \Xi_{2}\circ\phi_{J\nabla Q_{\varepsilon}}^{t}\circ \Xi_{1}^{-1}:(u,v)\mapsto (u+\tau_{\varepsilon}^t(v),v),\end{align} $$
    with
    (3.28) $$ \begin{align} \tau_{\varepsilon}^t(v)=tq_{\varepsilon}'(v)-\ln v. \end{align} $$

Proof. (1) We evaluate $\bar t_{I_{2}}(x_{*},y_{*})$ . Because

$$ \begin{align*}\phi_{J\nabla Q_{\varepsilon}}^t(x_{*},y_{*})=(e^{-tq^{\prime}_{\varepsilon}(x_{*}y_{*})}x_{*},e^{tq^{\prime}_{\varepsilon}(x_{*}y_{*})}y_{*}),\end{align*} $$

we have $e^{tq_{\varepsilon }'(x_{*}y_{*})}y_{*}\in I_{2}$ if and only if

$$ \begin{align*}t\in \bigg]\frac{\ln((x_{*}y_{*})^{-1} \times x_{*}\min I_{2})}{q^{\prime}_{\varepsilon}(x_{*}y_{*})},\frac{\ln((x_{*}y_{*})^{-1}\times x_{*}\max I_{2})}{q^{\prime}_{\varepsilon}(x_{*}y_{*})}\bigg[. \end{align*} $$

Hence for $x_{*}y_{*}$ small enough,

$$ \begin{align*}\bigg|\bar t_{I_{2}}(x_{*},y_{*})-\frac{|\!\ln(x_{*}y_{*})|}{q^{\prime}_{\varepsilon}(x_{*}y_{*})}\bigg|\leq \frac{\max(|\!\ln(x_{*}\min I_{2} )| ,|\!\ln(x_{*}\max I_{2} )|) }{q^{\prime}_{\varepsilon}(x_{*}y_{*})}.\end{align*} $$

Because for $0<x_{*}y_{*}\lesssim 1$ one has $q^{\prime }_{\varepsilon }(x_{*}y_{*})\asymp \lambda $ , there exists $c(I_{1},I_{2})$ such that if $x_{*}y_{*}$ small enough (how small depends on $I_{1},I_{2},\lambda $ ), the inequality (3.26) is satisfied.

(2) We write

$$ \begin{align*}\Xi_{2}\circ\phi_{J\nabla Q_{\varepsilon}}^t\circ \Xi_{1}^{-1}=\Xi_{2}\circ\Xi_{1}^{-1}\circ \Xi_{1}\circ\phi_{J\nabla Q_{\varepsilon}}^t\circ \Xi_{1}^{-1} =\Xi_{2}\circ\Xi_{1}^{-1}\circ \phi^t_{J\nabla\tilde Q_{\varepsilon}}, \end{align*} $$

with $\tilde Q_{\varepsilon }(u,v)=(Q_{\varepsilon }\circ \Xi _{1}^{-1})(u,v)=q_{\varepsilon }(v)$ . Because $\phi ^t_{J\nabla \tilde Q_{\varepsilon }}(u,v)=(u-tq_{\varepsilon }'(v),v)$ and $\Xi _{2}\circ \Xi _{1}^{-1}(u,v)=(u-\ln v,v)$ , we get (3.28).

4 Fundamental domains and first return maps

We construct in this section adapted fundamental domains $\mathcal {F}_{\varepsilon ,y_{*}}$ for the maps $(f_{\varepsilon })_{\varepsilon }$ satisfying conditions (H1)–(H4) of Theorem 2.1 and define their first return maps $\hat f_{\varepsilon }$ in $\mathcal {F}_{\varepsilon ,y_{*}}$ .

4.1 Fundamental domains

Let $V_{}$ be the domain of Theorem 2.1. One can choose $x_{*}>0$ such that, for any $\varepsilon $ small enough,

$$ \begin{align*}(x_{*},0)\in V_{}\quad \text{and}\quad f_{\varepsilon}^{-1}(x_{*},0)\notin V_{}.\end{align*} $$

For $y_{*}>0$ small enough, we define the vertical segment

$$ \begin{align*}L_{x_{*},y_{*}}:=\{(x_{*},ty_{*}),\ 0<t<1\}\end{align*} $$

and the domain

$$ \begin{align*}\mathcal{F}_{\varepsilon,x_{*},y_{*}}\end{align*} $$

as the interior of the contour defined by (see Figure 2)

  1. (a) the segment $[f_{\varepsilon }(x_{*},0),(x_{*},0)]$ ;

  2. (b) the transversal $\overline {L_{x_{*},y_{*}}}$ ;

  3. (c) the piece of hyperbola joining $(x_{*},y_{*})$ to $f_{\varepsilon }(x_{*},y_{*})$ (cf. Remark 2.1);

  4. (d) the curve $f_{\varepsilon }(\overline {L_{x_{*},y_{*}}})$ .

We shall often drop the index $x_{*}$ in the notations of $L_{x_{*},y_{*}}$ , $\mathcal {F}_{\varepsilon ,x_{*},y_{*}}$ and simply set

$$ \begin{align*}L_{y_{*}}:=L_{x_{*},y_{*}}\quad \text{and}\quad \mathcal{F}_{\varepsilon,y_{*}}=\mathcal{F}_{\varepsilon,x_{*},y_{*}}.\end{align*} $$

If $y_{*}$ is small enough, one has $\mathcal {F}_{\varepsilon ,y_{*}},\overline {L_{y_{*}}}\subset V_{}$ . We set

$$ \begin{align*}\tilde{\mathcal{F}}_{\varepsilon,y_{*}}=\mathcal{F}_{\varepsilon,y_{*}}\cup L_{y_{*}}.\end{align*} $$

Figure 2 Fundamental domain $\mathcal {F}_{\varepsilon ,y_{*}}$ for $f_{\varepsilon }$ and the first return map $\hat f_{\varepsilon }$ .

4.2 First return maps

Our aim in this subsection is to define the first return map of $f_{\varepsilon }$ in $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ .

Because $\Sigma _{\varepsilon }$ is a separatrix for $f_{\varepsilon }$ , we can define (see Remark 2.2)

$$ \begin{align*}N_{}(\varepsilon)\mathop{=}_{\mathrm{defin.}}\min\{n\in{\mathbb N}^*,f_{\varepsilon}^{-n}(]f_{\varepsilon}(x_{*}),x_{*}])\subset V_{}\}.\end{align*} $$

We note that if $\varepsilon $ is small enough, $N_{}(\varepsilon )$ is independent of $\varepsilon $ , so we shall denote it by $N_{}$ . Moreover, if $\varepsilon $ and $y_{*}$ are small enough,

(4.29) $$ \begin{align} N_{}\mathop{=}_{\mathrm{defin.}}\min\{n\in{\mathbb N}^*,f_{\varepsilon}^{-n}(\tilde{\mathcal{F}}_{\varepsilon,y_{*}})\subset V_{}\}.\end{align} $$

Lemma 4.1. There exists a constant $0<c_{*}<1$ such that, for $(x,y)\in \mathcal {F}_{\varepsilon ,c_{*}y_{*}}$ ,

(4.30) $$ \begin{align} \tilde n_{\varepsilon}(x,y):\mathop{=}_{\mathrm{defin.}}\min\{j\in{\mathbb N}^*, f_{\varepsilon}^j(x,y)\in f_{\varepsilon}^{-N_{}}(\tilde{\mathcal{F}}_{\varepsilon,y_{*}})\}<\infty. \end{align} $$

One has

(4.31) $$ \begin{align} \tilde n_{\varepsilon}(x,y)\asymp \ln(xy)/\lambda.\end{align} $$

Proof. Note that the domain $f_{\varepsilon }^{-N_{}}(\mathcal {F}_{\varepsilon ,y_{*}})\subset \tilde V_{}$ is the interior of the contour defined by:

  1. (a) the segment $[f_{\varepsilon }^{-(N_{}-1)}(x_{*},0),f_{\varepsilon }^{-N_{}}(x_{*},0)]\subset W^{u}_{f_{\varepsilon }}(o)\cap V_{}\subset \{0\}\times {\mathbb R}$ ;

  2. (b) the curve $f_{\varepsilon }^{-N_{}}(\overline {L_{y_{*}}})$ ;

  3. (c) a curve joining $f_{\varepsilon }^{-N_{}}(x_{*},y_{*})$ to $f_{\varepsilon }^{-(N_{}-1)}(x_{*},y_{*})$ ;

  4. (d) the curve $f_{\varepsilon }^{-(N_{}-1)}(\overline {L_{y_{*}}})$ ,

and

$$ \begin{align*}f_{\varepsilon}^{-N_{}}(\tilde{\mathcal{F}}_{\varepsilon,y_{*}})=f_{\varepsilon}^{-N_{}}(\mathcal{F}_{\varepsilon,y_{*}})\cup f_{\varepsilon}^{-N_{}}(L_{y_{*}}). \end{align*} $$

Note that the lines $f_{\varepsilon }^{-N_{}}(\overline {L_{y_{*}}})$ , $f_{\varepsilon }^{-(N_{}-1)}(\overline {L_{y_{*}}})$ are transversal to the segment $[f_{\varepsilon }^{-(N_{}-1)} (x_{*},0), f_{\varepsilon }^{-N_{}}(x_{*},0)]$ .

Now let $(x,y)\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ . We denote by $\mathcal {H}_{x,y}$ the hyperbola

$$ \begin{align*}\mathcal{H}_{x,y}:=\{ (x',y'),\ x'y'=xy\},\end{align*} $$

and if $z,z'\in \mathcal {H}_{x,y}$ , by $\mathcal {H}_{x,y}(z,z')$ , the arc of hyperbola of $\mathcal {H}_{x,y}$ between z and $z'$ which is open in z and closed in $z'$ . If $y>0$ is small enough, $\mathcal {H}_{x,y}$ intersects $f_{\varepsilon }^{-N_{}}(\tilde {\mathcal {F}}_{\varepsilon ,y_{*}})\subset V_{}$ in an arc of hyperbola of the form $\mathcal {H}_{x,y}(p,f_{\varepsilon }^{-1}(p))$ with $p\in f_{\varepsilon }^{-(N_{}-1)}({L_{y_{*}}})$ and $f_{\varepsilon }^{-1}(p)\in f_{\varepsilon }^{-N_{}}({L_{y_{*}}})$ . The sets $f^{-j}_{\varepsilon }(\mathcal {H}_{x,y}(p,f_{\varepsilon }^{-1}(p))$ , $j\geq 0$ form a partition of the semi-arc of parabola $\bigcup _{n\geq 0}\mathcal {H}_{x,y}(p,f_{\varepsilon ,k}^{-n}(p))$ which contains $(x,y)$ . In particular, there exists $j\geq 0$ (in fact $j\geq 1$ ) such that $(x,y)\in f^{-j}_{\varepsilon }(\mathcal {H}_{x,y}(p,f_{\varepsilon }^{-1}(p))$ or equivalently

$$ \begin{align*}f_{\varepsilon}^j(x,y)\in \mathcal{H}_{x,y}(p,f_{\varepsilon}^{-1}(p))\subset f^{-N_{}}_{\varepsilon}(\tilde{\mathcal{F}}_{\varepsilon,y_{*}}).\end{align*} $$

This proves (4.30).

To prove (4.31), we note that there exists an interval $I_{2}$ not containing 0 and depending only on $x_{*},y_{*}$ such that $f_{\varepsilon }^{-N_{}}(\tilde {\mathcal {F}}_{\varepsilon ,y_{*}})\subset {\mathbb R}\times I_{2}$ . We then use Lemma 3.1 and the fact that $|\bar t_{I_{2}}(x,y)-\tilde n_{\varepsilon }(x,y)|\leq 1$ .

We now define

(4.32) $$ \begin{align} n_{\varepsilon}=N_{}+\tilde n_{\varepsilon}.\end{align} $$

By (4.31), one has

(4.33) $$ \begin{align} n_{\varepsilon}(x,y)\asymp \ln(xy)/\lambda.\end{align} $$

The map $\hat f_{\varepsilon }:\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}\to \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , defined by

(4.34) $$ \begin{align} \hat f_{\varepsilon}=f_{\varepsilon}^{n_{\varepsilon}},\end{align} $$

is the first return map of $f_{\varepsilon }$ in $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ (for points starting in $ \tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}$ ). Note that $\hat f_{\varepsilon }$ is not $C^k$ on the whole domain $\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}$ .

4.3 Estimates on first return maps

We denote for $a\in {\mathbb R}$

$$ \begin{align*}T_{a}:(u,v)\mapsto (u+a,v)\end{align*} $$

and we recall the definition (3.25) of the symplectic diffeomorphisms $\Xi _{1},\Xi _{2}$ .

We observe that there exist open sets $W_{1}\subset {\mathbb R}^*_{+}\cap {\mathbb R}$ , $W_{2}\subset {\mathbb R}\times {\mathbb R}_{+}^*$ such that, for any $\varepsilon $ and $y_{*}>0$ small enough,

$$ \begin{align*}\tilde{\mathcal{F}}_{\varepsilon,y_{*}}\subset W_{1}\subset V_{},\quad f_{0}^{-N_{}}(\tilde{\mathcal{F}}_{\varepsilon,y_{*}})\subset W_{2}\subset V_{}.\end{align*} $$

Lemma 4.2. There exists a $C^k$ function $\sigma _{0,N_{}}\in C^k({\mathbb R}^*_{+},{\mathbb R})$ such that on $\Xi _{2}(W_{2})$ , one has

(4.35) $$ \begin{align} \Xi_{1}\circ f_{0}^{N_{}}\circ \Xi_{2}^{-1}=T_{\sigma_{0,N_{}}}. \end{align} $$

Proof. From condition (H4), one can write on ${\mathbb R}^2$

$$ \begin{align*}f_{0}=\phi^1_{J\nabla H_{0}}\end{align*} $$

and hence

$$ \begin{align*}f_{0}^{N_{}}=\phi^{N_{}}_{J\nabla H_{0}} \quad\text{where}\ H_{0}\ |_{V_{}}=Q_{0}.\end{align*} $$

If $(u,v)\in \Sigma _{2}(W_{2})$ and $(\tilde u,\tilde v)=\Xi _{1}(f_{0}^{N_{}}(\Xi _{2}^{-1}(u,v)))$ , one then has

$$ \begin{align*}Q_{0}(\Xi_{1}^{-1}(\tilde u,\tilde v))=Q_{0}(f_{0}^{N_{}}(\Xi_{2}^{-1}(u,v)))=Q_{0}(\Xi_{2}^{-1}(u,v))\end{align*} $$

and hence

$$ \begin{align*}q_{0}(\tilde v)=q_{0}(v)\end{align*} $$

and thus $\tilde v=v$ . Because the map $(u,v)\mapsto (\tilde u, \tilde v)$ is symplectic, this forces $\tilde u=u+\tilde \sigma _{0,N_{}}(v)$ for some $C^k$ function $\tilde \sigma _{0,N_{}}$ ; this function can be extended as a $C^k$ function $\sigma _{0,N_{}}: {\mathbb R}\to {\mathbb R}$ .

Recall the definition (4.34) of $\hat f_{\varepsilon }$ .

Lemma 4.3. There exists a continuous family $(\hat \eta _{\varepsilon })_{\varepsilon }$ of $C^k$ symplectic diffeomorphisms defined on ${\mathbb R}^2$ and a neighborhood W of $ f_{\varepsilon }^{-1}(\tilde {\mathcal {F}}_{\varepsilon ,y_{*}})\cup \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}\cup f_{\varepsilon }(\tilde {\mathcal {F}}_{\varepsilon ,y_{*}})$ such that

(4.36) $$ \begin{align}\begin{cases}\lim_{\varepsilon\to 0} \|\hat \eta_{\varepsilon}-\mathrm{id}\|_{k}=0,\\ \hat\eta_{\varepsilon}(W\cap({\mathbb R}\times\{0\})\subset {\mathbb R}\times\{0\} \end{cases} \end{align} $$

and on a neighborhood of $\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}$ , one has

(4.37) $$ \begin{align} \Xi_{1}\circ \hat f_{\varepsilon}\circ \Xi_{1}^{-1}=\hat \eta_{\varepsilon}\circ T_{\hat l_{\varepsilon}},\end{align} $$

where

(4.38) $$ \begin{align}\hat l_{\varepsilon}(v)=\sigma_{0,N_{}}(v)+\hat n_{\varepsilon}(u,v)q_{\varepsilon}'(v)-\ln v\quad \text{with }\hat n_{\varepsilon}=\tilde n_{\varepsilon}\circ \Xi_{1}^{-1}.\end{align} $$

Proof. We write (we use (4.34), (4.32), (H3)):

$$ \begin{align*} \hat f_{\varepsilon}&=f_{\varepsilon}^{N_{}+\tilde n_{\varepsilon}}\\ &=f_{\varepsilon}^{N_{}}\circ \phi_{J\nabla Q_{\varepsilon}}^{\tilde n_{\varepsilon}}\\ &=\eta_{\varepsilon}\circ f_{0}^{N_{}}\circ \phi_{J\nabla Q_{\varepsilon}}^{\tilde n_{\varepsilon}},\end{align*} $$

with

(4.39) $$ \begin{align} \eta_{\varepsilon}=f_{\varepsilon}^{N_{}}\circ f_{0}^{-N_{}}.\end{align} $$

As a consequence, if we set

(4.40) $$ \begin{align} \hat \eta_{\varepsilon}=\Xi_{1}\circ \eta_{\varepsilon}\circ \Xi_{1}^{-1}\quad \text{and}\quad \hat n_{\varepsilon}=\tilde n_{\varepsilon}\circ \Xi_{1}^{-1}, \end{align} $$

we have, using (4.35),

$$ \begin{align*}\Xi_{1}\circ \hat f_{\varepsilon}\circ \Xi_{1}^{-1}&=\hat\eta_{\varepsilon}\circ (\Xi_{1}\circ f_{0}^{N_{}}\circ\Xi_{2}^{-1}) \circ (\Xi_{2}\circ \phi^{\tilde n_{\varepsilon}}_{J\nabla Q_{\varepsilon}} \circ \Xi_{2}^{-1}) \circ (\Xi_{2}\circ\Xi_{1}^{-1}) \\ &=\hat \eta_{\varepsilon}\circ T_{\sigma_{0,N_{}}}\circ \phi^{\tilde n_{\varepsilon}\circ \Xi_{1}^{-1}}_{J\nabla (Q_{\varepsilon}\circ \Xi_{2}^{-1})} \circ T_{-\ln v}\\ &=\hat \eta_{\varepsilon}\circ T_{\sigma_{0,N_{}}}\circ T^{\hat n_{\varepsilon}}_{q^{\prime}_{\varepsilon}}\circ T_{-\ln v}\\ &=\hat \eta_{\varepsilon}\circ T_{\sigma_{0,N_{}}+\hat n_{\varepsilon}q^{\prime}_{\varepsilon}-\ln v}, \end{align*} $$

which is (4.37) together with (4.38).

Note that by (4.39), (4.40), Remark 2.2, and the fact that ${\mathbb R}\ni \varepsilon \mapsto f_{\varepsilon }\in C^k(V_{k},{\mathbb R}^2)$ is continuous, one has

$$ \begin{align*}\begin{cases}\lim_{\varepsilon\to 0}\|\hat \eta_{\varepsilon}-\text{id}\|_{C^k}=0,\\ \hat\eta_{\varepsilon}(W\cap ({\mathbb R}\times\{0\}))\subset {\mathbb R}\times\{0\}. \end{cases}\\[-4pc] \end{align*} $$

5 Renormalization

We define in this section a renormalization $\bar f_{\varepsilon }$ of the map $f_{\varepsilon }$ . The first return map $\hat f_{\varepsilon }$ of $f_{\varepsilon }$ in the fundamental domain $\mathcal {F}_{\varepsilon ,y_{*}}$ that we have constructed in §4 is not differentiable at every point (see (4.37), (4.38), and the fact that the integer valued function $\hat n_{\varepsilon }$ has, in general, discontinuity points). However, if one glues the ‘vertical’ boundaries of $\mathcal {F}_{\varepsilon ,y_{*}}$ by $f_{\varepsilon }$ , we obtain an abstract open annulus $\tilde F_{\varepsilon ,y_{*}}/f_{\varepsilon }$ (see §§5.1 and 5.2) and the map $\hat f_{\varepsilon }$ is now $C^k$ on it. We can uniformize this abstract annulus so that it becomes the standard (with the usual topology) open annulus ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c[$ (some $c>0$ ), see §5.3, and the map $\hat f_{\varepsilon }$ in these new coordinates turns into a $C^k$ diffeomorphism $\bar f_{\varepsilon }$ defined on (part of) this standard annulus. This is the (one should say ‘a’ instead of ‘the’ since the uniformizing/normalizing procedure is not unique) renormalized diffeomorphism associated to $f_{\varepsilon }$ . Uniformizing the annulus is equivalent to conjugating $f_{\varepsilon }$ to $(x,y)\mapsto (x+1,y)$ on a domain containing $\mathcal {F}_{\varepsilon ,y_{*}}$ . This procedure is, in a different context, the one described in [Reference Yoccoz, Marmi and Yoccoz24]. We shall often call the uniformization operation normalization in reference to the corresponding renormalization procedure defined for quasi-periodic cocycles, cf. [Reference Avila and Krikorian6, Reference Krikorian14].

5.1 Gluing

Let $\mathcal {F}$ be an open set of ${\mathbb R}^2$ , L a one-dimensional submanifold of ${\mathbb R}^2$ and f an orientation preserving smooth diffeomorphism from a neighborhood of $\mathcal {F}\cup L$ to a neighborhood of $f(\mathcal {F}\cup L)$ . We assume that:

  1. (1) $f(\mathcal {F}\cup L)\cap (\mathcal {F}\cup L)=\emptyset $ ;

  2. (2) $\mathcal {F}\cup L$ is a two-dimensional submanifold of ${\mathbb R}^2$ with boundary and this boundary is $\partial (\mathcal {F}\cup L)=L$ ; in particular, for any point $p\in L$ , there exists an open set $U_{p}$ , $p\in U_{p}\subset {\mathbb R}^2$ , and a smooth diffeomorphism $\varphi _{p}:U_{p}\to \varphi _{p}(U_{p})\subset {\mathbb R}^2$ such that $\varphi _{p}(U_{p}\cap L)=\varphi _{p}(U_{p})\cap ({\mathbb R}\times \{0\})$ and $\varphi _{p}(U_{p}\cap \mathcal {F})=\varphi _{p}(U_{p})\cap ({\mathbb R}\times {\mathbb R}^*_{+})$ ;

  3. (3) for any $p\in \mathcal {F}\cup L$ and $U_{p}$ , one has $U_{p}\cap f(\mathcal {F}\cup L)=\emptyset $ ;

  4. (4) for any $p\in L$ one has $f^{-1}(f(U_{p})\cap \mathcal {F})=\varphi _{p}^{-1}(\varphi _{p}(U_{p})\cap ({\mathbb R}\times {\mathbb R}^*_{-}))$ , for any of the previous chart $(U_{p},\varphi _{p})$ at p.

We define the topological space $(\mathcal {F}\cup L,\mathcal {T})$ as being the set $\mathcal {F}\cup L$ endowed with the following topology $\mathcal {T}$ : a subset S of $\mathcal {F}\cup L$ is an element of $\mathcal {T}$ (that is an open set) if for every $p\in S$ , there exists an open set $V\subset {\mathbb R}^2$ (contained in a neighborhood of $\mathcal {F}\cup L$ where f is defined) such that $V\cap f(\mathcal {F}\cup L)=\emptyset $ and $p\in (V\cup f(V))\cap (\mathcal {F}\cup L)\subset S$ .

We can then define the following differentiable structure on $(\mathcal {F}\cup L,\mathcal {T})$ as follows: (a) if $p\in \mathcal {F}$ , we define the local chart $C_{p}:=(W_{p},\text {id})$ , where $W_{p}$ is an open set of ${\mathbb R}^2$ such that $p\in W_{p}\subset \mathcal {F}$ ; and (b) if $p\in L$ , we define the local chart $C_{p}:=(W_{p},\psi _{p})$ where $W_{p}$ is the open set of $\mathcal {F}\cup L$ (see condition (3)), $W_{p}=(\mathcal {F}\cup L)\cap (U_{p}\cup f(U_{p}))$ (here $(U_{p},\varphi _{p})$ is the local chart for $p\in L$ as defined in (2)), and where $\psi _{p}$ is defined by (we use condition (4))

$$ \begin{align*}\begin{cases}\psi_{p}=\varphi_{p}\quad \text{on } U_{p}\cap (\mathcal{F}\cup L)=\varphi_{p}^{-1}((\varphi_{p}(U_{p})\cap ({\mathbb R}\times{\mathbb R}_{+})),\\ \psi_{p}=\varphi_{p}\circ f^{-1}\quad \text{on } f(U_{p})\cap \mathcal{F}=f\circ\varphi_{p}^{-1}((\varphi_{p}(U_{p})\cap ({\mathbb R}\times{\mathbb R}_{-}^*)). \end{cases} \end{align*} $$

We denote by $\mathcal {A}$ the collection of all these local charts $C_{p}$ and we set $(\mathcal {F}\cup L)/f= (\mathcal {F}\cup L,\mathcal {T},\mathcal {A})$ .

Remark 5.1. If we assume in addition that f preserves the standard symplectic form $dx\wedge dy$ on ${\mathbb R}^2$ , we can endow $(\mathcal {F}\cup L)/f$ with a symplectic form $\omega $ .

Remark 5.2. If $g:\mathcal {F}\to g(\mathcal {F})$ is a smooth diffeomorphism defined in a neighborhood of $\mathcal {F}$ , it induces a smooth diffeomorphism (that we still denote g) $g:(\mathcal {F}\cup L)/f\to (g(\mathcal {F})\cup g(L))/(g\circ f\circ g^{-1})$ .

Remark 5.3. If $\mathcal {F}=[0,1[{\kern2pt}{\times}{\kern2pt}]0,1[$ , $L={\kern2pt}]0,1[$ and $f=T_{1}:(x,y)\mapsto (x+1,y)$ , one sees that $(\mathcal {F}\cup L)/T_{1}$ is (diffeomorphic to) the standard open annulus $({\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,1[,\mathrm {canonical})$ endowed with its canonical differentiable structure.

5.2 The space $(\mathcal {F}_{\varepsilon ,y_{*}}\cup L_{y_{*}})/f_{\varepsilon }$

If $\varepsilon $ and $y_{*}$ are small enough, item (1) is satisfied and we can find charts $(p,U_{p})$ such that items (2)–(4) are satisfied. See Figure 3. We can then define the manifold $(\mathcal {F}_{\varepsilon ,y_{*}}\cup L_{y_{*}})/f_{\varepsilon }$ . We shall see that it is an annulus without boundary, cf. Lemma 5.3.

Figure 3 Gluing: $(\mathcal {F}_{\varepsilon ,y_{*}}\cup L_{y_{*}})/f_{\varepsilon }$ .

Note that if $0<c_{*}<1$ , the smaller set $\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}=\mathcal {F}_{\varepsilon ,c_{*}y_{*}}\cup L_{c_{*}y_{*}}$ is an open subset of $(\mathcal {F}_{\varepsilon ,y_{*}}\cup L_{y_{*}})/f_{\varepsilon }$ (which means that it belongs to $\mathcal {T}$ ) and it can be endowed with the topology and differentiable structure induced by the inclusion. We denote $(\mathcal {F}_{\varepsilon ,c_{*}y_{*}}\cup L_{c_{*}y_{*}})/f_{\varepsilon }$ the thus obtained submanifold of $(\mathcal {F}_{\varepsilon ,y_{*}}\cup L_{y_{*}})/f_{\varepsilon }$ . The following lemma is then tautological.

Lemma 5.1. The map $\hat f_{\varepsilon }$ induces a $C^k$ map $\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}/f_{\varepsilon }\to \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ .

We shall need in §6 the following lemma.

Lemma 5.2. There exists a probability measure with positive density $\pi _{\varepsilon ,y_{*}}$ on $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ which is $\hat f_{\varepsilon }$ invariant: for any measurable set $A\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ such that $\hat f_{\varepsilon }^{-1} (A)\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ , one has $\pi _{\varepsilon ,y_{*}}(A)=\pi _{\varepsilon ,y_{*}}(\hat f_{\varepsilon }^{-1}(A))$ .

Proof. We shall in fact construct this measure $\pi _{\varepsilon ,y_{*}}$ on the bigger set $\hat {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ ,

$$ \begin{align*}\hat{\mathcal{F}}_{\varepsilon,y_{*}}=\tilde{\mathcal{F}}_{\varepsilon,y_{*}}\cup \sigma(\tilde{\mathcal{F}}_{\varepsilon,y_{*}}),\end{align*} $$

where $\sigma :{\mathbb R}^2\to {\mathbb R}^2$ is the reflection $(x,y)\mapsto (x,-y)$ (it commutes with $f_{\varepsilon }$ in $V_{}$ , see condition (H3)). From Remark 5.1, there exists a symplectic form $\omega _{\varepsilon }$ on $\hat {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ . Note that the first return map $\hat f_{\varepsilon }$ is not defined on the whole set $\hat {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ but nevertheless

$$ \begin{align*}(\hat f_{\varepsilon})^*\omega_{\varepsilon}=\omega_{\varepsilon}\end{align*} $$

whenever this formula makes sense. The probability measure $\pi _{\varepsilon ,y_{*}}$ defined by

$$ \begin{align*}\pi_{\varepsilon,y_{*}}(A)=\int_{A}|\omega_{\varepsilon}|\Bigg/\int_{\mathcal{F}_{\varepsilon,y_{*}}}|\omega_{\varepsilon}|\end{align*} $$

is $\hat f_{\varepsilon }$ invariant.

5.3 Normalization of $f_{\varepsilon }$

We now uniformize the abstract annulus $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ . To do that, it is enough to normalize $f_{\varepsilon }$ in the sense of item 2 of the following lemma.

Lemma 5.3. (Normalization Lemma)

There exists a continuous family $(h_{\varepsilon })_{\varepsilon }$ of (not necessarily symplectic) $C^{k}$ -diffeomorphisms defined on a neighborhood of $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ such that for some $c_{}>0$ :

  1. (1) $h_{\varepsilon }$ sends $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ to the standard open annulus $( ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c_{}[,\mathrm {canonical})$ ;

  2. (2) $h_{\varepsilon }\circ f_{\varepsilon }\circ h_{\varepsilon }^{-1}=T_{1}:(x,y)\mapsto (x+1,y)$ ;

  3. (3) $h_{\varepsilon }([x_{*},f_{\varepsilon }(x_{*)}] \times \{0\}=[0,1[{\kern2pt}{\times}{\kern2pt}\{0\}$ .

Proof. Using condition (H3) and the change of coordinates (3.25) of §3, we see that on a neighborhood of $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , one has (we use the notation $(x,y)$ for $(u,v)$ )

$$ \begin{align*}\Xi_{1}\circ f_{\varepsilon}\circ \Xi_{1}^{-1}=T_{q^{\prime}_{\varepsilon}} :(x,y)\mapsto (x+q^{\prime}_{\varepsilon}(y),y).\end{align*} $$

If $g_{\varepsilon }$ is the (not necessarily symplectic) smooth diffeomorphism

(5.41) $$ \begin{align} g_{\varepsilon}:(x,y)\mapsto \bigg(\frac{x}{q^{\prime}_{\varepsilon}(y)},y\bigg),\end{align} $$

one has

(5.42) $$ \begin{align} g_{\varepsilon}\circ \Xi_{1}\circ f_{\varepsilon}\circ \Xi_{1}^{-1}\circ g_{\varepsilon}^{-1}=T_{1}.\end{align} $$

The set $\overline {(g_{\varepsilon }\circ \Xi _{1})(\mathcal {F}_{\varepsilon ,y_{*}})}$ is of the form

$$ \begin{align*}\overline{(g_{\varepsilon}\circ\Xi_{1})(\mathcal{F}_{\varepsilon,y_{*}})}=\{(x,y),\ y\in [0,c_{}],\ {\gamma}_{\varepsilon}(y)\leq x\leq {\gamma}_{\varepsilon}(y)+1\},\end{align*} $$

where $c_{}>0$ , ${\gamma }_{\varepsilon }:[0,c_{}]\to {\mathbb R}_{+}$ is $C^k$ , ${\gamma }_{\varepsilon }(0)=0$ , and the map ${\mathbb R}\ni \varepsilon \mapsto {\gamma }_{\varepsilon }\in C^k([0,c_{}],{\mathbb R})$ is continuous. This indeed follows from the definition of $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ in §4.1, the definition of $\Xi _{1}$ (3.25), and (5.41), (5.42). As a consequence, if we denote

(5.43) $$ \begin{align} j_{\varepsilon}:(x,y)\mapsto (x-\gamma_{\varepsilon}(y),y),\end{align} $$

we have

(5.44) $$ \begin{align}\begin{cases}j_{\varepsilon}\circ T_{1}=T_{1}\circ j_{\varepsilon},\\ j_{\varepsilon}( (g_{\varepsilon}\circ\Sigma_{1})(\mathcal{F}_{\varepsilon,y_{*}}))=\,]0,1[{\kern2pt}{\times}{\kern2pt}]0,c_{}[,\\ j_{\varepsilon}((g_{\varepsilon}\circ\Sigma_{1})(L_{y_{*}}))=\{0\}{\kern2pt}{\times}{\kern2pt}]0,c_{}[. \end{cases} \end{align} $$

By Remarks 5.2 and 5.3, the map

(5.45) $$ \begin{align} h_{\varepsilon}=j_{\varepsilon}\circ g_{\varepsilon}\circ \Xi_{1}\end{align} $$

is a diffeomorphism that sends $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ to the standard annulus $([0,1[{\kern2pt}{\times}{\kern2pt}]0,c_{}[)/T_{1}\simeq ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c_{}[$ and such that

$$ \begin{align*}h_{\varepsilon}\circ f_{\varepsilon}\circ h_{\varepsilon}^{-1}=T_{1}.\end{align*} $$

To conclude the proof, we notice (2) is an immediate consequence of the definition (5.45) of $h_{\varepsilon }$ .

Remark 5.4. Note that if $T_{a}(x,y)=(x+a(y),y)$ , one has

$$ \begin{align*}(h_{\varepsilon}\circ\Xi_{1}^{-1})\circ T_{a}\circ (h_{\varepsilon}\circ \Xi_{1}^{-1})^{-1}=T_{\tilde a},\quad \tilde a(y)=a(y)/q_{\varepsilon}'(y). \end{align*} $$

5.4 The renormalization $\bar f_{\varepsilon }$ of $f_{\varepsilon }$

There exists $\delta _{}\in{\kern2pt}]{\kern2pt}0,c_{}[$ such that the map

(5.46) $$ \begin{align} \bar f_{\varepsilon}\mathop{=}_{\mathrm{defin.}}h_{\varepsilon}\circ \hat{f}_{\varepsilon}\circ h_{\varepsilon}^{-1}:{\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,\delta_{}[{\kern2pt}\to {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c_{}[\end{align} $$

is well defined and is a $C^k$ diffeomorphism onto its image.

Proposition 5.4. One has

(5.47) $$ \begin{align} \bar f_{\varepsilon}=\bar\eta_{\varepsilon}\circ T_{l_{\varepsilon}},\end{align} $$

where $\bar \eta _{\varepsilon }$ is a $C^k$ diffeomorphism defined on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,\delta _{}[$ and $l_{\varepsilon }\in C^{k}(]0,c_{}[, {\mathbb R}/{\mathbb Z})$ ; they satisfy

(5.48) $$ \begin{align} l_{\varepsilon}(y)=\frac{\sigma_{0,N_{}}(y)}{q^{\prime}_{\varepsilon}(y)}-\frac{\ln y}{q^{\prime}_{\varepsilon}(y)}\!\!\mod {\mathbb Z},\end{align} $$
(5.49) $$ \begin{align}\lim_{\varepsilon\to0}\|\bar\eta_{\varepsilon}-\mathrm{id}\|_{C^{k}}=0,\end{align} $$
(5.50) $$ \begin{align} \bar\eta_{\varepsilon}:(x,y)\mapsto (x+a_{\varepsilon}(x,y),y+yb_{\varepsilon}(x,y)), \end{align} $$

where $a_{\varepsilon }\in C^k$ , $b_{\varepsilon }\in C^{k-1}$ are functions defined on ${\mathbb R}/{\mathbb Z}\times (0,\delta _{})$ .

Moreover, the map $\bar f_{\varepsilon }$ preserves a probability measure $\bar \pi _{\varepsilon ,y_{*}}$ with positive density defined on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c_{}[$ .

Proof. By (4.37) and Remark 5.4 after Lemma 5.3,

$$ \begin{align*}\bar f_{\varepsilon}&=(h_{\varepsilon}\circ\Xi_{1}^{-1})\circ \hat \eta_{\varepsilon}\circ (h_{\varepsilon}\circ\Xi_{1}^{-1})^{-1}\circ (h_{\varepsilon}\circ \Xi_{1}^{-1})\circ T_{\hat l_{\varepsilon}}\circ (h_{\varepsilon}\circ \Xi_{1}^{-1})^{-1}\\ &=\bar\eta_{\varepsilon}\circ T_{l_{\varepsilon}}, \end{align*} $$

where

(5.51) $$ \begin{align}\bar\eta_{\varepsilon}=(h_{\varepsilon}\circ \Xi_{1}^{-1})\circ \hat \eta_{\varepsilon}\circ (h_{\varepsilon}\circ \Xi_{1}^{-1})^{-1}\quad\text{and}\quad l_{\varepsilon}(y)=(1/q^{\prime}_{\varepsilon}(y))\hat l_{\varepsilon}(y).\end{align} $$

Because $\bar \eta _{\varepsilon }:=(h_{\varepsilon }\circ \Xi _{1}^{-1})\circ \hat \eta _{\varepsilon }\circ (h_{\varepsilon }\circ \Xi _{1}^{-1})^{-1}$ and $\bar f_{\varepsilon }$ are $C^{k}$ , the function ${l_{\varepsilon }}:{\kern2pt}]0,c_{}[{\kern2pt}\to {\mathbb R}/{\mathbb Z}$ is also $C^{k}$ and

$$ \begin{align*}l_{\varepsilon}(y)=(1/q^{\prime}_{\varepsilon}(y))\hat l_{\varepsilon}(y).\end{align*} $$

By (4.38) (remember that $\hat n_{\varepsilon }$ takes its value in ${\mathbb Z}$ ),

$$ \begin{align*}l_{\varepsilon}(y)&= \frac{\sigma_{0,N_{}}(y)}{q^{\prime}_{\varepsilon}(y)}+\hat n_{\varepsilon}(x,y)-\frac{\ln y}{q^{\prime}_{\varepsilon}(y)}\\ &=\frac{\sigma_{0,N_{}}(y)}{q^{\prime}_{\varepsilon}(y)}-\frac{\ln y}{q^{\prime}_{\varepsilon}(y)}\!\!\mod {\mathbb Z},\end{align*} $$

which is (5.48).

Equation (5.49) is a consequence of the definition of $\bar \eta _{\varepsilon }$ , cf. (5.51), the first equation of (4.36), and of the fact that ${\mathbb R}\ni \varepsilon \mapsto h_{\varepsilon }\in C^k$ is continuous (Lemma 5.3).

We now claim that if $\bar \eta _{\varepsilon }(x,y)=(x+a_{\varepsilon }(x,y), y+\bar b_{\varepsilon }(x,y))$ , one has for any y,

(5.52) $$ \begin{align} \bar b_{\varepsilon}(x,0)=0.\end{align} $$

Indeed, because

$$ \begin{align*}\bar \eta_{\varepsilon}:=(h_{\varepsilon}\circ \Xi_{1}^{-1})\circ \hat \eta_{\varepsilon}\circ (h_{\varepsilon}\circ \Xi_{1}^{-1})^{-1},\end{align*} $$

equality (5.52) is a consequence of the second equation of (4.36), of item (3) of Lemma 5.3, and of the fact that $\Xi _{1}({\mathbb R}_{+}^*\times \{0\})={\mathbb R}_{+}^*\times \{0\}$ .

To prove (5.50), we thus notice that equality (5.52) gives us for $\bar b_{\varepsilon }$ a decomposition

$$ \begin{align*}\begin{cases}\bar b_{\varepsilon}(x,y)=yb_{\varepsilon}(x,y),\\ b_{\varepsilon}\in C^{k-1} .\end{cases}\end{align*} $$

Finally, to conclude the proof of the proposition, we observe that because the map $\hat f_{\varepsilon }:\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}/f_{\varepsilon }\to \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ preserves the probability measure $\pi _{\varepsilon ,y_{*}}$ , cf. Lemma 5.2, the diffeomorphism $\bar f_{\varepsilon }:{\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,\delta _{}[{\kern2pt}\to {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c_{}[$ preserves the probability measure $\bar \pi _{\varepsilon ,y_{*}}=(h_{\varepsilon })_{*}\pi _{\varepsilon ,y_{*}}$ defined on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c_{}[$ (in the sense that if $A\subset {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c_{}[{\kern2pt}$ is a Borelian set such that $\bar f_{\varepsilon }^{-1}(A)\subset {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c_{}[$ , one has $\bar \pi _{\varepsilon ,y_{*}}(A)=\bar \pi _{\varepsilon ,y_{*}}(\bar f_{\varepsilon }^{-1}(A))$ ).

6 Applying the translated curve theorem

We apply in this section Rüssmann’s (or Moser’s) translated curve theorem to some rescaled version $\mathring {f}_{\varepsilon ,n}$ of the renormalization $\bar f_{\varepsilon }$ of $f_{\varepsilon }$ defined in §5.4.

6.1 The translated curve theorem

Let $\psi : {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[{\kern2pt}\to {\mathbb R}/{\mathbb Z}\times {\mathbb R}$ ( $\ln e=1$ ) be a $C^k$ diffeomorphism defined on the annulus (or cylinder) ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[$ . We say that the graph $\mathrm {Gr}_{{\gamma }}:=\{(x,\gamma (x)): x\in {\mathbb R}/{\mathbb Z}\}$ of a continuous map $\gamma :{\mathbb R}/{\mathbb Z}\to {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[{\kern2pt}$ is translated by $\psi $ if for some $t\in {\mathbb R}$ ,

(6.53) $$ \begin{align}\psi(\mathrm{Gr}_{\gamma})=\mathrm{Gr}_{t+\gamma} \end{align} $$

and invariant if $t=0$ . If $\mathrm {Gr}_{{\gamma }}$ satisfies (6.53), there exists an orientation preserving homeomorphism of the circle $g:{\mathbb R}/{\mathbb Z}\to {\mathbb R}/{\mathbb Z}$ such that $\psi (x,\gamma (x))=\psi (g(x),t+\gamma (g(x))$ . If $t=0$ (respectively $t\ne 0$ ), we define (respectively with a clear abuse of language) the rotation number of ( $\psi $ on) the invariant (respectively translated) graph $\mathrm {Gr}_{{\gamma }}$ as the rotation number of the circle diffeomorphism g. We say that $\psi $ has the intersection property if for any continuous ${\gamma }:{\mathbb R}/{\mathbb Z}\to {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[$ , the curve $\mathrm {Gr}_{{\gamma }}:=\{(x,{\gamma }(x)):x\in {\mathbb R}/{\mathbb Z}\}$ intersects its image $\psi (\mathrm {Gr}_{{\gamma }})$ . Note the following important fact: If $\psi $ has the intersection property, any translated graph by $\psi $ is invariant.

We state the translated curve theorem by Rüssmann [Reference Rüssmann17] (which implies the invariant curve theorem by Moser [Reference Moser15]):

Theorem 6.1. (Rüssmann, [Reference Rüssmann17])

There exists $k_{0}\in {\mathbb N}$ for which the following holds. Let $k\geq k_{0}$ , $C,\mu>0$ , and $l:{\mathbb R}/{\mathbb Z}\to {\mathbb R}$ a $C^k$ map satisfying the twist condition,

(6.54) $$ \begin{align} \min_{y} |\partial_{y}l(y)|>\mu>0 \quad \text{and}\quad \|l\|_{C^{k_{0}}}\leq C,\end{align} $$

and define

$$ \begin{align*}\psi_{0}:(x,y)\mapsto (x+l(y),y).\end{align*} $$

There exists $\varepsilon _{0}=\varepsilon _{0}(C,\mu )>0$ such that for any $C^k$ diffeomorphism

$$ \begin{align*}\psi:{\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[{\kern2pt}\to{\mathbb R}/{\mathbb Z}\times{\mathbb R}\end{align*} $$

satisfying

(6.55) $$ \begin{align} \|\psi-\psi_{0}\|_{C^{k_{0}}}<\varepsilon_{0},\end{align} $$

the diffeomorphism $\psi $ admits a set of positive Lebesgue measure of $C^{k-k_{0}}$ translated graphs contained in $({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]e^{-3/4},e^{-1/4}[$ . Moreover, all these translated graphs have Diophantine rotation numbers (they are in a fixed Diophantine class $DC(\kappa ,\tau )$ (the exponent is $\tau $ and the constant $\kappa $ ) that can be prescribed in advance once $\mu $ is fixed ( $k_{0}$ then depends on $\tau $ and $\varepsilon _{0}$ on $\kappa $ and $\tau $ )).

6.2 The rescaled diffeomorphism $\mathring {f}_{\varepsilon ,n}$

Let $\bar f_{\varepsilon }$ be the renormalized map defined in §5.4 and define $u_{\varepsilon },v_{\varepsilon }$ by

$$ \begin{align*}\bar f_{\varepsilon}(x,y)=(x+u_{\varepsilon}(x,y),y+v_{\varepsilon}(x,y)).\end{align*} $$

Because $\bar f_{\varepsilon }=\bar \eta _{\varepsilon }\circ T_{l_{\varepsilon }}$ (cf. (5.47)), one has using (5.50):

$$ \begin{align*}u_{\varepsilon}(x,y)&=l_{\varepsilon}(y)+a_{\varepsilon}(x+l_{\varepsilon}(y),y),\\ v_{\varepsilon}(x,y)&=yb_{\varepsilon}(x+l_{\varepsilon}(y),y). \end{align*} $$

Figure 4 The diffeomorphism $\bar f_{\varepsilon }$ on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}[e^{-(n+1)},e^{-n}]$ .

Now, let $n\in {\mathbb N}^*$ large enough so that

(6.56) $$ \begin{align} ]e^{-(n+1)},e^{-n}[{\kern2pt}{\subset}{\kern2pt}]0,\delta_{}[\end{align} $$

(the $\delta _{}$ of (5.46)) and introduce the rescaled $C^{k}$ diffeomorphism $\mathring {f}_{\varepsilon ,n}$ defined on the annulus ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[$ by

(6.57) $$ \begin{align} \mathring{f}_{\varepsilon,n}\mathop{=}_{\mathrm{defin.}}\Lambda_{e^n}\circ \bar f_{\varepsilon}\circ \Lambda_{e^n}^{-1}, \end{align} $$

where $\Lambda _{e^n}:(x,y)\mapsto (x,e^n y)$ . Let us denote

$$ \begin{align*}\mathring{f}_{\varepsilon,n}(x,y)=(x+u_{\varepsilon,n}(x,y), y+v_{\varepsilon,n}(x,y)).\end{align*} $$

A computation shows that:

(6.58) $$ \begin{align} \begin{cases}u_{\varepsilon,n}(x,y)&=l_{\varepsilon,n}(y)+a_{\varepsilon}(x+l_{\varepsilon,n}(y),e^{-n}y),\\ v_{\varepsilon,n}(x,y)&=yb_{\varepsilon}(x+l_{\varepsilon,n}(y),e^{-n}y), \end{cases}\end{align} $$

where

(6.59) $$ \begin{align} l_{\varepsilon,n}(y)=l_{\varepsilon}(e^{-n}y). \end{align} $$

We can now state the following important proposition the proof of which occupies the next subsection.

Proposition 6.1. Assume that $k\geq k_{0}+2$ (k is the regularity in conditions (H1)–(H4) and $k_{0}$ is the one of Theorem 6.1). There exists $\varepsilon _{1}>0$ such that the following holds. If $|\varepsilon |\leq \varepsilon _{1}$ and $n\gg 1$ , $\mathring {f}_{\varepsilon ,n}$ admits a set of positive Lebesgue measure of invariant $C^{k-k_{0}-2}$ -graphs in ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[$ .

6.3 Proof of Proposition 6.1

6.3.1 Twist condition for $l_{\varepsilon ,n}$

Lemma 6.2. There exist $C,\mu>0$ such that, for any $\varepsilon $ small enough and any $n\gg 1$ , the map $l_{\varepsilon ,n}$ satisfies the twist condition (6.54) provided $k\geq k_{0}+1$ .

Proof. Using (5.48), (6.59), we have

$$ \begin{align*}l_{\varepsilon,n}(y)&=l_{\varepsilon}(e^{-n}y)\\ &=\frac{\sigma_{0,N_{}}(e^{-n}y)}{q^{\prime}_{\varepsilon}(e^{-n}y)}+\frac{n}{q^{\prime}_{\varepsilon}(e^{-n}y)}-\frac{\ln y}{q^{\prime}_{\varepsilon}(e^{-n}y)}\!\!\mod {\mathbb Z}\\ &=\frac{\sigma_{0,N_{}}(0)+n}{\lambda}-\frac{\ln y}{\lambda}+\theta_{\varepsilon,n}(y)\!\!\mod {\mathbb Z}, \end{align*} $$

where

$$ \begin{align*} \|\theta_{\varepsilon,n}\|_{C^{k-1}([e^{-1},1])}=O(e^{-n});\end{align*} $$

this last inequality is a consequence of the fact that $q_{\varepsilon }(s)=\lambda s+O(s^2)$ is continuous with respect to $\varepsilon $ (cf. condition H2) and of the fact that $\sigma _{0,N_{}}$ is $C^k$ (cf. Lemma 4.2). In particular, for some $C_{k}>0$ (depending on $\lambda $ ),

$$ \begin{align*}\| l_{\varepsilon,n} \|_{C^{k-1}}\leq C_{k}\end{align*} $$

and because $\partial _{y} l_{\varepsilon ,n}(y)=-1/ (\lambda y)+\partial _{y}\theta _{\varepsilon ,n}(y)$ and $y\in{\kern2pt}]e^{-1},1[$ ,

$$ \begin{align*}|\partial_{y} l_{\varepsilon,n}(y)|\geq 1/(2\lambda) .\end{align*} $$

Hence (6.54) holds uniformly in $\varepsilon ,n$ with $C=C_{k_{0}+1}$ and $\mu =1/(2\lambda )$ as soon as n is large enough.

6.3.2 $\mathring {f}_{\varepsilon ,n}$ is close to a twist

We observe that from (5.49), (5.50), (6.58), and Lemma 6.2, one has uniformly in n,

(6.60) $$ \begin{align} \lim_{\varepsilon\to0}\max(\|u_{\varepsilon,n}-l_{\varepsilon,n}\|_{C^{k-2}},\|v_{\varepsilon,n}\|_{C^{k-2}})=0.\end{align} $$

In particular, if n is large enough, inequality (6.55) is satisfied if $k\geq k_{0}+2$ with $\psi =\mathring {f}_{\varepsilon ,n}$ and $\psi _{0}:(x,y)\mapsto (x+l_{\varepsilon ,n}(y),y)$ .

We see from §§6.3.1 and 6.3.2 that, if

$$ \begin{align*}|\varepsilon|\leq \varepsilon_{1}\mathop{=}_{\mathrm{defin.}} \varepsilon_{0}(C_{k_{0}+1},1/(2\lambda))\end{align*} $$

and $n\gg 1$ , the assumptions of Theorem 6.1 are then satisfied by $\mathring {f}_{\varepsilon ,n}$ with $k-2$ in place of k. Under these conditions, there thus exists a set $\mathring {\mathcal {G}}_{\varepsilon ,n}$ of $C^{k-k_{0}-2}\ \mathring {f}_{\varepsilon ,n}$ -translated graphs, the union of which covers a set of positive Lebesgue measure in $({\mathbb R}/{\mathbb T}){\kern2pt}{\times}{\kern2pt}]e^{-3/4},e^{-1/4}[$ . We just have to check that these translated graphs are indeed invariant.

6.3.3 $\mathring {f}_{\varepsilon ,n}$ -translated graphs are invariant

Let $\mathring {{\gamma }}\subset ({\mathbb R}/{\mathbb T}){\kern2pt}{\times}{\kern2pt}]e^{-3/4},e^{-1/4}[$ be a $\mathring {f}_{\varepsilon ,n}$ -translated graph: $\mathring {f}_{\varepsilon ,n}(\mathring {{\gamma }})=\mathring {{\gamma }}+(0,t)$ for some $t\in {\mathbb R}$ . We shall prove that $t=0$ . We can without loss of generality assume that $t\geq 0$ (the case $t\leq 0$ is treated in a similar way).

Formula (6.60) shows that if $n\gg 1$ , one has $\mathring {f}_{\varepsilon ,n}(\mathring {{\gamma }})\subset ({\mathbb R}/{\mathbb T}){\kern2pt}{\times}{\kern2pt}]e^{-1},1[$ . From the conjugation relation (6.57), we see that (cf. (6.56))

$$ \begin{align*}\bar {\gamma}:=\Lambda_{e^n}^{-1}(\mathring{{\gamma}})\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]e^{-n-3/4}, e^{-n-1/4}[\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,\delta[\end{align*} $$

is a $\bar f_{\varepsilon }$ -translated graph such that

$$ \begin{align*}\bar f_{\varepsilon}(\bar {\gamma})=\bar{\gamma}+(0,e^{-n}t)\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]e^{-(n+1)},e^{-n}[\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,\delta[.\end{align*} $$

Let A be the open domain of $({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c[$ between $({\mathbb R}/{\mathbb Z})\times \{0\}$ and $\bar {\gamma }$ . Because $t\geq 0$ , one has $A\subset \bar f_{\varepsilon }(A)\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c[$ .

Assume by contradiction that $t>0$ ; then the set $\bar f_{\varepsilon }(A)\smallsetminus A$ contains a non-empty open set. We have seen (cf. Proposition 5.4) that $\bar f_{\varepsilon }$ preserves a probability measure $\bar \pi _{\varepsilon ,y_{*}}$ with positive density defined on $({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c[$ , so $\bar \pi _{\varepsilon ,y_{*}}(\bar f_{\varepsilon }(A)\smallsetminus A)>0$ . However, this contradicts the invariance of $\bar \pi _{\varepsilon ,y_{*}}$ by $\bar f_{\varepsilon }$ .

The proof of Proposition 6.1 is complete.□

6.4 Invariant curves for $\bar f_{\varepsilon }$

We can now state the following.

Theorem 6.2. Let $k\geq k_{0}+2$ and $|\varepsilon |\leq \varepsilon _{1}$ . There exists $\nu _{}\in{\kern2pt}]0,\delta _{}[$ such that, for any $\nu \in{\kern2pt}]0,\nu _{}[$ , there exists a set $\bar {\mathcal {G}}_{\varepsilon ,\nu }$ of $C^{k-k_{0}-2}$ , $\bar f_{\varepsilon }$ -invariant graphs contained in $({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]e^{-1}\nu ,\nu [$ such that

$$ \begin{align*}\mathrm{Leb}_{{\mathbb R}^2}\bigg(\bigcup_{\bar{\gamma}\in\bar{\mathcal{G}}_{\varepsilon,\nu}} \bar{\gamma}\bigg)>0.\end{align*} $$

Proof. We choose n so that

(6.61) $$ \begin{align} ]e^{-(n+1)},e^{-n}[{\kern2pt}{\subset}{\kern2pt}]0,\nu[\end{align} $$

and we observe that when $\nu \to 0$ , one has $n\to \infty $ . Define

$$ \begin{align*} \mathring{f}_{\varepsilon,n}=\Lambda_{e^n}\circ \bar f_{\varepsilon}\circ \Lambda_{e^n}^{-1}.\end{align*} $$

By Proposition 6.1, there exists $\nu _{1}>0$ such that if $\nu \in{\kern2pt}]0,\nu _{1}[$ (n satisfying (6.61) is then large enough), the diffeomorphism $\mathring {f}_{\varepsilon ,n}$ admits $C^{k-k_{0}-2}$ -invariant curves in ${\mathbb T}{\kern2pt}{\times}{\kern2pt}]e^{-1},1[$ covering a set of positive Lebesgue measure; hence, $\bar f_{\varepsilon ,k}$ has $C^{k-k_{0}-2}$ -invariant curves in ${\mathbb T}{\kern2pt}{\times}{\kern2pt}]e^{-1}\nu ,\nu [$ covering a set of positive Lebesgue measure.

We shall denote

$$ \begin{align*}\bar{\mathcal{G}}_{\varepsilon}=\bigcup_{\nu\in{\kern2pt}]0,\nu_{1}[}\bar{\mathcal{G}}_{\varepsilon,\nu}.\end{align*} $$

Remark 6.1. For all $\bar {\gamma }\in \bar {\mathcal {G}}_{\varepsilon ,\nu }$ , the rotation number of the circle diffeomorphism $\bar f_{\varepsilon }\ |_{\bar {\gamma }}$ is Diophantine in a fixed Diophantine class $DC(\kappa ,\tau )$ (see the comment at the end of the statement of Theorem 6.1).

7 Invariant curves for $f_{\varepsilon }$

We define

$$ \begin{align*}r=k-k_{0}-2\end{align*} $$

and assume that $|\varepsilon |\leq \varepsilon _{1}$ .

Let $\bar {\gamma }\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,\delta [$ , $\bar {\gamma }\in \bar {\mathcal {G}}_{\varepsilon }$ be a $C^r$ invariant graph for $\bar f_{\varepsilon }:({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,\delta [{\kern2pt}\to ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c[$ . Note that there exists $\delta _{1}>0$ such that $\bar {\gamma }\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]\delta _{1},\delta [$ .

We can view $\bar {\gamma }$ as an invariant graph sitting in $([0,1[{\kern2pt}{\times}{\kern2pt}]0,c[)/T_{1}$ (recall $T_{1}(x,y)=(x+1,y)$ ). In particular, one can find a $C^r$ , 1-periodic function

$$ \begin{align*}\bar z:{\mathbb R}\to ([0,1[{\kern2pt}{\times}{\kern2pt}]0,\delta[)/T_{1}\end{align*} $$

such that, for all t, $({d}/{dt})\bar z(t)\ne 0$ and

$$ \begin{align*}\bar \gamma= \bar z([0,1[),\quad \bar z(0)\in \{0\}{\kern2pt}{\times}{\kern2pt}]0,c[,\quad \lim_{t\to 1-}\bar z(t)=T_{1}(\bar z(0))\in \{1\}{\kern2pt}{\times}{\kern2pt}]0,c[.\end{align*} $$

Let

$$ \begin{align*}\hat \gamma=h_{\varepsilon}^{-1}(\bar\gamma),\end{align*} $$

where $h_{\varepsilon }$ was defined in Lemma 5.3. Because $\bar f_{\varepsilon }=h_{\varepsilon }\circ \hat f_{\varepsilon }\circ h_{\varepsilon }^{-1}$ (cf. (5.46)), we see that

$$ \begin{align*}\hat {\gamma}\subset h_{\varepsilon}^{-1}(({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]\delta_{1},\delta[)\subset \tilde{\mathcal{F}}_{\varepsilon,c_{*}y_{*}}\end{align*} $$

is a $C^r$ compact, connected, one-dimensional submanifold (without boundary) of $\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}/f_{\varepsilon }$ , which is invariant by $\hat f_{\varepsilon }:\tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}/f_{\varepsilon }\to \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon }$ . Moreover, the function

$$ \begin{align*}\hat z\mathop{=}_{\mathrm{defin.}} h_{\varepsilon}^{-1}\circ\bar z:{\mathbb R}\to \tilde{\mathcal{F}}_{\varepsilon,c_{*}y_{*}}/f_{\varepsilon}\end{align*} $$

is a $C^r$ , 1-periodic function and

$$ \begin{align*}\hat \gamma= \hat z([0,1[),\quad \hat z(0)\in L_{y_{*}},\quad \lim_{t\to 1-}\hat z(t)=f_{\varepsilon}(\hat z(0))\in f_{\varepsilon}(L_{y_{*}}). \end{align*} $$

The main result of this section is the following proposition.

Proposition 7.1. The set

$$ \begin{align*}\hat \Gamma=\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat {\gamma})\subset{\mathbb R}^2 \end{align*} $$

is an invariant $C^r$ curve for $f_{\varepsilon }$ : it is a compact, connected, one-dimensional $C^r$ submanifold of ${\mathbb R}^2$ which is invariant by $f_{\varepsilon }$ .

We give the proof of this proposition in §7.2.

7.1 Preliminary results

We define the function $\hat Z:{\mathbb R}\to {\mathbb R}^2$

$$ \begin{align*}\text{for all } t\in{\mathbb R},\quad \hat Z(t)=f_{\varepsilon,k}^{[t]}(\hat z(t-[t]))\end{align*} $$

( $[t]$ denotes the integer part of t that is the unique integer such that $[t]\leq t<[t]+1$ ).

Lemma 7.2. The function $\hat Z:{\mathbb R}\to {\mathbb R}^2$ is $C^r$ .

Proof. Note that, for $t\in [0,1[$ , $\hat Z(t)=\hat z(t)$ . Also, the very definition of $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}/f_{\varepsilon ,k}$ shows that the function $\hat Z$ is $C^r$ on a neighborhood of $t=1$ . It is hence $C^r$ on $[0,2[$ and because for $j\in {\mathbb Z}$ , $\hat Z(t+j)=f_{\varepsilon }^j(\hat Z(t))$ , it is $C^r$ on ${\mathbb R}$ .

Let us set

$$ \begin{align*}\tau\mathop{=}_{\mathrm{defin.}}\inf\{t\geq 1,\ \hat Z(t)\in \tilde{\mathcal{F}}_{\varepsilon,y_{*}}\}.\end{align*} $$

Note that

(7.62) $$ \begin{align} 2\leq \tau<\infty.\end{align} $$

Indeed, the left-hand side inequality is a consequence of the fact that $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}\kern-1pt\cap\kern-1pt f_{\varepsilon }(\tilde {\mathcal {F}\kern-1pt}_{\varepsilon ,y_{*}})=\emptyset $ . For the right-hand side, we observe that because $\hat Z(0)=\hat z(0)\in \tilde {\mathcal {F}}_{\varepsilon ,c_{*}y_{*}}$ , one has (see (4.34)), $\hat Z(n_{\varepsilon }(\hat z(0)))=f_{\varepsilon }^{n_{\varepsilon }(\hat z(0))}(\hat z(0))\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , hence $\tau \leq n_{\varepsilon }(\hat z(0))<\infty $ .

Lemma 7.3. The map $\hat Z:[0,\tau [{\kern2pt}\to {\mathbb R}^2$ is injective.

Proof. Assume by contradiction that $\hat Z:[0,\tau [{\kern2pt}\to {\mathbb R}^2$ is not injective; then, there exists $m_{i}\in {\mathbb N}$ , $0\leq s_{i}<1$ ,

(7.63) $$ \begin{align} 0\leq m_{i}+s_{i}<\tau,\quad i=1,2,\quad \hat Z(s_{1}+m_{1})=\hat Z (s_{2}+m_{2}).\end{align} $$

Hence, $f_{\varepsilon }^{m_{1}}(\hat \gamma )\cap f_{\varepsilon }^{m_{2}}(\hat \gamma )\ne \emptyset $ and if $m:=m_{2}-m_{1}\geq 0$ , $f_{\varepsilon }^{m}(\hat {\gamma })\cap \hat {\gamma }\ne \emptyset $ . In particular, there exists $t\in [m,m+1[$ such that $\hat Z(t)\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ and then $t\geq \tau $ . As a consequence, $m> \tau -1$ and because $0\leq m<\tau $ ( $m_{1},m_{2}$ are both in the interval $[0,\tau [$ ), one has $m=[\tau ]$ , and hence $m_{2}=m=[\tau ]$ and $m_{1}=0$ . We then have from (7.63), $\hat Z(s_{2}+[\tau ])=\hat Z(s_{1})\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ (because $s_{1}\in [0,1[$ ) and hence by the definition of $\tau $ , $s_{2}+[\tau ]\geq \tau $ , which contradicts $m_{2}+s_{2}<\tau $ .

Lemma 7.4. If, for some $t\geq 1$ , $\hat Z(t)\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , then $\hat Z(t)\in \hat \gamma $ .

Proof. Indeed, writing $t=s+n$ , $s\in [0,1[$ , $n\in {\mathbb N}^*$ , one has $\hat Z(t)=f_{\varepsilon }^n(\hat z(s))$ . The integer $n\geq 1$ is thus a $m^{\textrm{th}}$ return time of $\hat z(s)$ in $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , $\hat Z(t)=\hat f_{\varepsilon }^{m}(\hat z(s))$ , and because $\hat \gamma $ is invariant by $\hat f_{\varepsilon }$ , it is readily seen by induction on m that $f_{\varepsilon }^n(\hat z(s))\in \hat \gamma $ .

Lemma 7.5. One has $\hat Z(\tau )=\hat z(0)$ .

Proof. From the definition of $\tau $ and Lemma 7.4, we have $\hat Z(\tau )\in \mathrm {closure}(\hat \gamma )\cap \mathrm {closure}(L_{y_{*}} \cup f_{\varepsilon }(L_{y_{*}})) $ and hence $\hat Z(\tau )\in \{\hat z(0),f_{\varepsilon }(\hat z(0))\}$ . To conclude, we observe that one cannot have $\hat Z(\tau )=f_{\varepsilon }(\hat z(0))$ because otherwise, one would have $\hat Z(\tau -1)=\hat z(0)\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , which contradicts the definition of $\tau $ (from (7.62) $\tau -1\geq 1$ ).

Lemma 7.6. The derivative of $\hat Z$ at $\tau $ is transverse to $L_{y_{*}}$ .

Proof. (1) If there exists a sequence $t_{n}\in {\mathbb R}$ , $\lim t_{n}=\tau $ , such that $Z(t_{n})\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , then from Lemma 7.4, one has $Z(t_{n})\in \hat \gamma $ and consequently $(d\hat Z/dt)(\tau )$ is tangent to $\hat \gamma $ , thus transverse to $L_{y_{*}}$ .

(2) Otherwise, there exists an open interval $I\subset {\mathbb R}$ , $I\ni \tau $ , such that, for all $t\in I\smallsetminus \{\tau \}$ , $\hat Z(t)\notin \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ and $f_{\varepsilon }(\hat Z(t))\in \mathcal {F}_{\varepsilon ,y_{*}}$ . From Lemma 7.4, one then has for all $t\in I\smallsetminus \{\tau \}$ , $\hat Z(t+1)=f_{\varepsilon }(\hat Z(t))\in \hat \gamma $ (see item (2) of §5.1), and hence $Df_{\varepsilon }(f_{\varepsilon }(\hat Z(\tau )))\cdot (d\hat Z/dt)(\tau )$ is tangent to $\hat {\gamma }$ and in particular transverse to $f_{\varepsilon }(L_{y_{*}})$ . This implies that $(d\hat Z/dt)(\tau )$ is transverse to $L_{y_{*}}$ .

Lemma 7.7. One has $\hat Z([\tau ,\tau +1[)=\hat Z([0,1[)$ .

Proof. We define $s_{*}=\sup \{s\geq 0:\text {for all } t\in [\tau , \tau +s[, \ \hat Z(t)\in \mathcal {F}_{\varepsilon ,y_{*}}\}$ . From Lemmata 7.4 and 7.6, one has: (a) $s_{*}>0$ ; (b) for any $t\in [\tau ,\tau +s_{*}[$ , $\hat Z(t)\in \hat \gamma $ ; and (c) $\hat Z(\tau +s_{*})\in f_{\varepsilon }(L_{y_{*}})\cap \mathrm {closure}( \hat \gamma )=f_{\varepsilon }(\hat z(0))=\hat Z(1)$ . In particular, $\hat Z(\tau +s_{*}-1)=f_{\varepsilon }^{-1}(\hat Z(1))=\hat Z(0)\in \tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ and by the definition of $\tau $ , this implies $s_{*}\geq 1$ . Now we notice that one cannot have $s_{*}>1$ because otherwise $\tau +1\in [\tau ,\tau +s_{*}[$ and by definition of $s_{*}$ , $\hat Z(\tau +1)\in \mathcal {F}_{\varepsilon ,y_{*}}$ ; however, $\hat Z(\tau +1)=f_{\varepsilon }(\hat Z(\tau ))$ and because $\hat Z(\tau )=\hat z(0)$ (Lemma 7.5), one has $\hat Z(\tau +1)=f_{\varepsilon }(\hat z(0))\notin \mathcal {F}_{\varepsilon ,y_{*}}$ . We have thus proven that $s_{*}=1$ . This implies that $\hat Z([\tau ,\tau +1[)=\hat Z([0,1[)$ .

7.2 Proof of Proposition 7.1

We first observe that

(7.64) $$ \begin{align} \hat\Gamma=\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat {\gamma})=\hat Z({\mathbb R})=\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat Z([0,\tau+1[)).\end{align} $$

Next we note the following.

  1. (1) One has $\hat Z ([0,\tau +1[)=\hat Z([0,\tau ])$ .

  2. (2) The set $\hat Z([0,\tau +1[)$ is $f_{\varepsilon }$ -invariant.

Item (1) is a consequence of

$$ \begin{align*}\hat Z ([0,\tau+1[)&=\hat Z ([0,\tau])\cup \hat Z ([\tau,\tau+1[)\\ &=\hat Z([0,\tau])\cup \hat Z([0,1[) \quad(\text{Lemma}\ {7.7})\\ &=\hat Z([0,\tau])\quad (1\leq \tau). \end{align*} $$

Item (2) follows from item (1) and

$$ \begin{align*}f_{\varepsilon}(\hat Z([0,\tau+1[))&=f_{\varepsilon}(\hat Z ([0,\tau]) )\\ &=\hat Z([1,\tau+1])\\ &=\hat Z([1,\tau])\cup \hat Z([\tau,\tau+1])\\ &=\hat Z([1,\tau])\cup \hat Z([0,1])\quad(\text{Lemma}\ {7.7})\\ &=\hat Z([0,\tau])\quad (1\leq \tau)\\ &=\hat Z([0,\tau+1[). \end{align*} $$

Item (2) and (7.64) yield

$$ \begin{align*}\hat \Gamma=\hat Z([0,\tau+1[).\end{align*} $$

This last identity shows that $\hat \Gamma $ is a connected, compact (cf. item (1)) subset of ${\mathbb R}^2$ which is $f_{\varepsilon }$ -invariant.

Let us prove that $\hat \Gamma $ is a one-dimensional submanifold of ${\mathbb R}^2$ . Because $\hat Z(\tau )=\hat Z(0)$ (Lemma 7.5), one has

$$ \begin{align*}\hat Z([0,\tau+1[)=\hat Z(]0,\tau+1[)=\hat Z(]0,\tau[)\cup \hat Z(]\tau-1,\tau+1[).\end{align*} $$

From Lemmata 7.2 and 7.3, the set $\hat Z(]0,\tau [)$ is a one-dimensional submanifold of ${\mathbb R}^2$ as well as the set $\hat Z(]\tau -1,\tau +1[)$ (note that $\hat Z(]\tau -1,\tau +1[)=f_{\varepsilon }(\hat Z([\tau -2,\tau [)$ ). The intersection of these two sets is $\hat Z(]\tau ,\tau +1[)$ and from Lemma 7.7, it is equal to $\hat Z(]0,1[)$ which is a one-dimensional submanifold of ${\mathbb R}^2$ . As a consequence, the union $\hat Z(]0,\tau [)\cup \hat Z(]\tau -1,\tau +1[)$ is one-dimensional submanifold of ${\mathbb R}^2$ .

This concludes the proof of Proposition 7.1

8 Proof of Theorem 2.1 (and hence of Theorem A)

As we have mentioned in §2.5, Theorem A follows from Theorem 2.1; we describe the proof of this latter result in this section.

Let $r=k-k_{0}-2$ , $|\varepsilon |\leq \varepsilon _{1}$ , and $\nu \leq \nu _{1}$ . Theorem 6.2 yields a set $\bar {\mathcal {G}}_{\varepsilon ,\nu }$ of $C^r$ , $\bar f_{\varepsilon }$ -invariant graphs contained in $({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]e^{-1}\nu ,\nu [$ , the union of which covers a set of positive Lebesgue measure.

In the previous section (cf. Proposition 7.1), for all $\nu \in{\kern2pt}]0,\nu _{1}[$ , we have associated to each $\bar f_{\varepsilon }$ -invariant graph $\bar {\gamma }\in \bar {\mathcal {G}}_{\varepsilon ,\nu }$ an $f_{\varepsilon }$ -invariant $C^r$ -curve:

(8.65) $$ \begin{align} \hat \Gamma=\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat{\gamma})\quad\text{where } \hat {\gamma}=h_{\varepsilon}^{-1}(\bar {\gamma}).\end{align} $$

We denote by $\hat {\mathcal {G}}_{\varepsilon ,\nu }$ the set of all such curves $\hat \Gamma $ .

To prove Theorem 2.1, we just have to prove that, for all $\nu \in{\kern2pt}]0,\nu _{1}[$ ,

(8.66) $$ \begin{align} (\text{Positive measure})\quad \mathrm{Leb}_{{\mathbb R}^2}\bigg(\bigcup_{\hat \Gamma\in\hat{\mathcal{G}}_{\varepsilon,\nu}}\hat \Gamma \bigg)>0 \end{align} $$

and

(8.67) $$ \begin{align}(\text{Accumulation})\quad\lim_{\nu\to 0}\sup_{\hat\Gamma\in\hat{\mathcal{G}}_{\varepsilon,\nu}} \mathrm{dist}(\hat \Gamma,\Sigma_{\varepsilon} )=0. \end{align} $$

8.1 Proof of (8.66) (positive measure)

This is a consequence of the inclusion (cf. (8.65))

$$ \begin{align*}h_{\varepsilon}^{-1}\bigg(\bigcup_{\bar{\gamma}\in\bar{\mathcal{G}}_{\varepsilon,\nu}}\bar {\gamma} \bigg)\subset \bigcup_{\hat \Gamma\in\hat{\mathcal{G}}_{\varepsilon,\nu}}\hat \Gamma \end{align*} $$

and of the fact that $\mathrm {Leb}_{2}(\bigcup _{\bar {\gamma }\in \bar {\mathcal {G}}_{\varepsilon ,\nu }}\bar {\gamma })>0$ (this is the content of Theorem 6.2).

8.2 Proof of (8.67) (accumulation)

Let $\bar {\gamma }\in \bar {\mathcal {G}}_{\varepsilon ,\nu }$ , $\bar {\gamma }\subset ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,\nu [$ . From the definition (5.45) of the diffeomorphism $h_{\varepsilon }$ , we see that, for some positive constant $C_{\lambda }$ depending on $\lambda $ (cf. condition (H3)),

$$ \begin{align*}\hat{\gamma}=h_{\varepsilon}^{-1}(\bar{\gamma})\subset\{(x,y)\in \tilde{\mathcal{F}}_{\varepsilon,y_{*}},\quad xy\in{\kern2pt}]0,C_{\lambda}\nu[ \}.\end{align*} $$

However,

$$ \begin{align*}\hat \Gamma=\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat {\gamma})=\bigg(\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat {\gamma})\cap V_{}\bigg)\cup \bigcup_{\substack{n\in{\mathbb Z}\\ f_{\varepsilon}^n(\hat {\gamma})\not\subset V_{}}}f_{\varepsilon}^n(\hat {\gamma}).\end{align*} $$

From condition (H3), one has

$$ \begin{align*}\bigcup_{n\in{\mathbb Z}}f_{\varepsilon}^n(\hat{\gamma})\cap V_{}\subset V_{}\cap \{(x,y),\quad xy\in{\kern2pt}]0,C_{\lambda}\nu[\},\end{align*} $$

and hence, using Remark 2.2,

(8.68) $$ \begin{align} \mathrm{dist}\bigg( \bigcup_{n\in{\mathbb Z}} f_{\varepsilon}^n(\hat {\gamma})\cap V_{} ,\Sigma_{\varepsilon}\cap V_{} \bigg)=o_{\nu}(1) \quad (\text{uniform in }\ \hat{\gamma}).\end{align} $$

Now, recalling the definition (4.29) of the integer $N_{}$ of §4.2, one has

$$ \begin{align*} \bigcup_{\substack{n\in{\mathbb Z}\\ f_{\varepsilon}^n(\hat {\gamma})\not\subset V_{}}}f_{\varepsilon}^n(\hat {\gamma})\subset \bigcup_{n=1}^{N_{}} f^{-n}_{\varepsilon}(\hat{\gamma}), \end{align*} $$

and using the fact that $\mathrm {dist}(\hat {\gamma },\Sigma _{\varepsilon }\cap [(x_{*},0),f_{\varepsilon }(x_{*},0)[)=o_{\nu }(1)$ , one can see that ( $N_{}$ is fixed)

(8.69) $$ \begin{align} \mathrm{dist}\bigg(\!\bigcup_{\substack{n\in{\mathbb Z} \\ f_{\varepsilon}^n(\hat{\gamma})\not\subset V_{}}} f_{\varepsilon}^n(\hat {\gamma}) ,\Sigma_{\varepsilon}\cap \bigcup_{j=1}^{N_{}} f_{\varepsilon}^{-1}([(x_{*},0),f_{\varepsilon}(x_{*},0)[)) \bigg)=o_{\nu}(1),\end{align} $$

where the previous limit is uniform in $\hat {\gamma }$ .

Equations (8.68) and (8.69) give

$$ \begin{align*}\mathrm{dist}(\hat \Gamma,\Sigma_{\varepsilon})=o_{\nu}(1).\end{align*} $$

8.3 KAM circles for $f_{\varepsilon }$

Let $\hat \Gamma $ be a $C^r$ invariant curve for $f_{\varepsilon }$ of the form (8.65) and $g_{\hat \Gamma }$ the restriction of $f_{\varepsilon }$ to $\hat \Gamma $ . The map $g_{\hat \Gamma }$ can be identified with a circle diffeomorphism. Similarly, the restriction of $\bar f_{\varepsilon }$ to the invariant curve $\bar {\gamma }$ yields a circle diffeomorphism $g_{\bar {\gamma }}$ .

Let $\hat \alpha $ and $\bar \alpha $ be the rotation numbers of $g_{\hat \Gamma }$ and $g_{\bar {\gamma }}$ .

Lemma 8.1. One has $\{1/\hat \alpha \}=\bar \alpha $ (here $\{\cdot \}$ denotes the fractional part).

Proof. We refer to the renormalization procedure defined in §§4 and 5. Let $\hat J$ be the arc $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}\cap \hat \Gamma $ . The restriction on $\hat J$ of $\hat f_{\varepsilon }$ , the first return map of $f_{\varepsilon }$ in $\tilde {\mathcal {F}}_{\varepsilon ,y_{*}}$ , defines a $C^r$ diffeomorphism of the abstract circle $\hat J/f_{\varepsilon }$ . Classical arguments show that the rotation number of this circle diffeomorphism is equal to $\{1/\hat \alpha \}$ . However, after normalization of $f_{\varepsilon }$ by $h_{\varepsilon }$ (cf. formula (5.46)), $\hat \Gamma $ is transported to $\bar {\gamma }$ and the $C^r$ diffeomorphism $\hat f_{\varepsilon }:\hat J/f_{\varepsilon }\to \hat J/f_{\varepsilon }$ to the circle diffeomorphism $\bar f_{\varepsilon }: \bar J/T_{1}\to \bar J/T_{1}$ , where $\bar J=h_{\varepsilon }(\hat J)\subset \bar {\gamma }$ is a fundamental domain of $\bar f\ {|\ \bar {\gamma }}$ . The rotation numbers of $\hat f_{\varepsilon }:\hat J/f_{\varepsilon }\to \hat J/f_{\varepsilon }$ and $\bar f_{\varepsilon }: \bar J/T_{1}\to \bar J/T_{1}$ are hence equal. However, the rotation number of $\bar f_{\varepsilon }: \bar J/T_{1}\to \bar J/T_{1}$ is (same argument as before) equal to $\{1/\bar \alpha \}$ .

Because $\bar \alpha $ can be chosen in a fixed Diophantine class $DC(\kappa ,\tau )$ (see Remark 6.1), the rotation number $\hat \alpha $ is Diophantine with the same exponent $\tau $ . By the Herman–Yoccoz theorem on linearization of $C^r$ -circle diffeomorphisms [Reference Herman10, Reference Yoccoz22], this implies that if r is large enough (depending on $\tau $ which is fixed), the diffeomorphism $g_{\hat \Gamma }$ is linearizable; in other words, $\hat \Gamma $ is a KAM curve. However, one has a priori no control on the Diophantine constant of $\hat \alpha $ .

This concludes the proof of Theorem 2.1, whence of Theorem A.□

9 Proof of Theorem B

We construct in §9.1 a symplectic diffeomorphism $f_{\mathrm {pert}}$ admitting a separatrix $\Sigma $ (see Figure 5) and depending on a (‘large’) parameter M. We renormalize $f_{\mathrm {pert}}$ like in §§4 and 5 to get a diffeomorphism $\bar {f}_{\mathrm {pert}}$ of an open annulus ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c[$ . We prove in Proposition 9.3 of §9.2 that this renormalized diffeomorphism $\bar f_{\mathrm {pert}}$ sends some graphs projecting on a fixed interval $J_{M}$ (see (9.81) on graphs which project on the whole circle and which are below the initial graphs we have started from, see Figure 6. We then iterate this procedure in §9.3 to find an orbit of $\bar f_{\mathrm {pert}}$ accumulating the boundary ${\mathbb R}/{\mathbb Z}\times \{0\}$ of the aforementioned annulus: this prevents the existence of $\bar {f}_{\mathrm {pert}}$ -invariant curves close to this boundary and therefore of $f_{\mathrm {pert}}$ -invariant curves close to the separatrix $\Sigma $ . The diffeomorphism $f_{\mathrm {pert}}$ is the searched for example of Theorem B.

Figure 5 The perturbed map $f_{\mathrm {pert}}$ .

Figure 6 The image of the graph ${\gamma }_{J_{M},y}$ by the diffeomorphism $\bar f_{\mathrm {pert}}$ .

9.1 Construction of the example

We start with a smooth autonomous symplectic vector field of the form $X_{0}=J\nabla H_{0}$ , where $H_{0}:{\mathbb R}^2\to {\mathbb R}$ satisfies on some neighborhood V of $o=(0,0)$

$$ \begin{align*}H_{0}(x,y)=x y\quad \text{on} \ V \end{align*} $$

and has the property that $\Sigma =H_{0}^{-1}(H_{0}(0,0))$ is compact and connected. The set $\Sigma $ is a separatrix of

$$ \begin{align*}f\mathop{=}_{\mathrm{defin.}}\phi^1_{J\nabla H_{0}}\end{align*} $$

associated to the hyperbolic fixed point o.

Fixing $x_{*}>0$ small enough, we can define like in §4, for $y_{*}>0$ small enough, a fundamental domain $\tilde {\mathcal {F}}_{y_{*}}=\mathcal {F}_{y_{*}}\cup L_{y_{*}}\subset V$ , where $\mathcal {F}_{y_{*}}$ is defined by $(a)-(d)$ 4.1) with $\phi ^1_{J\nabla H_{0}}$ in place of $f_{\varepsilon }$ . We can even assume that $\phi _{J\nabla H_{0}}^{-j}(\tilde {\mathcal {F}}_{y_{*}})\subset V$ , $j=1,2$ . There exists $c_{*}>0$ such that the first return map,

$$ \begin{align*}\hat f:\tilde{\mathcal{F}}_{c_{*}y_{*}}\to \mathcal{F}_{y_{*}},\end{align*} $$

is well defined. We can renormalize $f=\phi ^1_{J\nabla H_{0}}$ like in §5 by first normalizing f (cf. Lemma 5.3):

(9.70) $$ \begin{align} h\circ f\circ h^{-1}=T_{1}, \end{align} $$

where

(9.71) $$ \begin{align} h:\mathcal{F}_{y_{*}}\to [0,1[{\kern2pt}{\times}{\kern2pt}]0,c[ \end{align} $$

is symplectic (see (5.45), (5.41), and the fact that we choose $q(s)=s$ ) and then setting (cf. (5.46)):

(9.72) $$ \begin{align} \bar f\mathop{=}_{\mathrm{defin.}}h\circ \hat f\circ h^{-1}:{\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,\delta[{\kern2pt}\to {\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}]0,c[.\end{align} $$

By (5.48) of Proposition 5.4, we have

(9.73) $$ \begin{align} \bar f=T_{l},\quad l(y)=\sigma(y)-\ln y\end{align} $$

for some smooth function $\sigma $ .

We can assume that $h(\tilde {\mathcal {F}}_{y_{*}})=[0,1[{\kern2pt}{\times}{\kern2pt}]0,c[$ and that $T^{-j}([0,1[{\kern2pt}{\times}{\kern2pt}]0,c[)\subset h^{-1}(V)$ , $j=1,2$ .

We now construct a symplectic perturbation $f_{\mathrm {pert}}:{\mathbb R}^2\to {\mathbb R}^2$ of f which admits $\Sigma $ as a separatrix. We shall need first the following lemma.

Lemma 9.1. There exist $b\in (0,1)$ and a non-empty compact interval $I\subset ]0,1[$ such that, for any $M>0$ , there exists a smooth function $\varphi _{M}:{\mathbb R}\to {\mathbb R}$ satisfying:

  1. (1) $\varphi _{M}\ |_{I}\leq -bM$ ;

  2. (2) $({b^{-1} M}/{|I|})\geq -\varphi _{M}' \ |_{ I}\geq ({M}/{|I|})$ ;

  3. (3) the map $s_{M}:{\mathbb R}\to {\mathbb R}$ , defined by

    $$ \begin{align*}s_{M}(t)=\int_{0}^te^{\varphi_{M}(u)}du,\end{align*} $$
    is an increasing smooth diffeomorphism of ${\mathbb R}$ that coincides with the identity on ${\mathbb R}\smallsetminus [0,1]$ .

Proof. See Appendix C.

Let $\chi :{\mathbb R}\to {\mathbb R}$ be a smooth function equal to 1 on $[-c/2,c/2]$ and to 0 on ${\mathbb R}\smallsetminus [-(3/4)c,(3/4)c]$ , and define

(9.74) $$ \begin{align} S_{M}(x,y)=(s_{M}(x)y)\chi (y)+xy(1-\chi(y)).\end{align} $$

The canonical (hence symplectic) mapping $g_{M}$ associated to $S_{M}$ :

(9.75) $$ \begin{align} g_{M}(x,y)=(\tilde x, \tilde y) \ \ \iff\ \ \begin{cases}x=\dfrac{\partial S_{M}}{\partial y}(\tilde x,y),\\[10pt] \tilde y=\dfrac{\partial S_{M}}{\partial \tilde x}(\tilde x,y), \end{cases} \end{align} $$

is equal to the identity on $({\mathbb R}\smallsetminus [0,1]){\kern2pt}{\times}{\kern2pt} [-c,c]$ and satisfies for $(x,y)\in [0,1[{\kern2pt}{\times}{\kern2pt}]0,c/2[$

(9.76) $$ \begin{align} \begin{cases}\tilde x=s_{M}^{-1}(x),\\ \tilde y=s^{\prime}_{M}\circ s_{M}^{-1}(x)y. \end{cases} \end{align} $$

The following symplectic perturbation of f:

$$ \begin{align*}f_{\mathrm{pert}}:&\mathop{=}_{\mathrm{defin.}}h^{-1}\circ (g_{M}\circ T_{1})\circ h\\[-2pt] & \, \, =(h^{-1}\circ g_{M}\circ h)\circ f \end{align*} $$

(recall h satisfies (9.70)) is thus defined on ${\mathbb R}^2$ and coincides with f outside $f^{-1}(\tilde {\mathcal {F}}_{y_{*}})$ . Moreover, because $g_{M}({\mathbb R}\times \{0\})={\mathbb R}\times \{0\}$ ,

$$ \begin{align*}\Sigma \ \ \text{is a separatrix for } f_{\mathrm{pert}}.\end{align*} $$

Now, because $\tilde {\mathcal {F}}_{y_{*}}$ is a fundamental domain for $f_{\mathrm {pert}}$ ( $f_{\mathrm {pert}}$ coincide with f on $\tilde {\mathcal {F}}_{y_{*}}$ ), for some $c_{\mathrm {pert}}>0$ small enough, the first return map

$$ \begin{align*} \hat f_{\mathrm{pert}}:f_{\mathrm{pert}}^{-1}(\tilde{\mathcal{F}}_{c_{\mathrm{pert}}y_{*}})\to f_{\mathrm{pert}}^{-1}(\tilde{\mathcal{F}}_{y_{*}}),\end{align*} $$

is well defined and satisfies

$$ \begin{align*}\hat f_{\mathrm{pert}}=(\hat f\circ f^{-1})\circ f_{\mathrm{pert}}.\end{align*} $$

In particular, on

$$ \begin{align*}[-1,0[{\kern2pt}{\times}{\kern2pt}]0,c/2[,\end{align*} $$

one has (cf. (9.72), (9.73))

(9.77) $$ \begin{align}\bar f_{\mathrm{pert}}\mathop{=}_{\mathrm{defin.}}h\circ \hat f_{\mathrm{pert}}\circ h^{-1}&=\bar f\circ T_{-1}\circ g_{M}\circ T_{1} \end{align} $$
(9.78) $$ \begin{align} & \hspace{7pc} =T_{l-1}\circ g_{M}\circ T_{1} \end{align} $$

and

$$ \begin{align*}\bar f_{\mathrm{pert}}:{\mathbb R}{\kern2pt}{\times}{\kern2pt}]0,c/2[{\kern2pt}\to {\mathbb R}{\kern2pt}{\times}{\kern2pt}]0,c[\quad \text{satisfies}\quad \bar f_{\mathrm{pert}}\circ T_{1}=T_{1}\circ \bar f_{\mathrm{pert}};\end{align*} $$

in particular, it defines a smooth map $({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c/2[{\kern2pt}\to ({\mathbb R}/{\mathbb Z}){\kern2pt}{\times}{\kern2pt}]0,c[$ .

Note that because $g_{M}$ is the identity outside $[0,1]{\kern2pt}{\times}{\kern2pt}[-c,c]$ , it admits a $T_{1}$ -periodization $\tilde g_{M}:{\mathbb R}{\kern2pt}{\times}{\kern2pt}[-c,c]\to {\mathbb R}{\kern2pt}{\times}{\kern2pt}[-c,c]$ (which means that $\tilde g_{M}$ and $g_{M}$ coincide on $[0,1]{\kern2pt}{\times} [-c,c]$ and $\tilde g_{M}$ commutes with $T_{1}$ ). This $\tilde g_{M}$ is defined by the same formula (9.75) as $g_{M}$ , where now the new function $\tilde s_{M}$ involved in (9.74) is the ${\mathbb Z}$ -periodization of $s_{M}$ . To simplify the notation, we shall continue to denote $\tilde g_{M}$ and $\tilde s_{M}$ by $g_{M}$ and $s_{M}$ .

Let

(9.79) $$ \begin{align} t:=t(x):=s_{M}^{-1}(x+1). \end{align} $$

Lemma 9.2. For $(x,y)\in [-1,0[{\kern2pt}{\times}{\kern2pt}]0,c/2[$ , the point $(\bar x,\bar y):=\bar f_{\mathrm {pert}}(x,y)$ satisfies with the notation (9.79)

(9.80) $$ \begin{align} \begin{cases} \bar x=t-1+\sigma(s_{M}'(t)\times y)-\ln(s_{M}'(t))-\ln y,\\ \ln \bar y=\ln(s_{M}'(t))+\ln y. \end{cases}\end{align} $$

Proof. Let $(x,y)\in [-1,0[{\kern2pt}{\times}{\kern2pt}]0,c/2[ $ ; with the notation $(x_{1},y_{1})=(g_{M}\circ T_{1})(x,y)=g_{M}(x+1,y)$ , one has from (9.78), $(\bar x,\bar y)=T_{l-1}(x_{1},y_{1})$ and from (9.76), (9.73),

$$ \begin{align*}\begin{cases}x_{1}= s_{M}^{-1}(x+1),\\ y_{1}=s_{M}'\circ s_{M}^{-1}(x+1)\times y, \end{cases}\text{and}\quad \begin{cases}\bar x=x_{1}-1+\sigma(y_{1})-\ln y_{1},\\ \bar y=y_{1}, \end{cases}\end{align*} $$

and hence (9.80).

9.2 Image of a piece of graph by $\bar f_{\mathrm {pert}}$

We take $M>0$ (from Lemma 9.1) large enough and we define

(9.81) $$ \begin{align} J_{M}=s_{M}(I)-1\subset [-1,0[, \end{align} $$

where I is the interval introduced in Lemma 9.1.

If $y:J\to {\mathbb R}_{+}^*$ , $x\to y(x)$ is a differentiable function, we denote by $\gamma _{J,y}$ its graph:

$$ \begin{align*}\gamma_{J,y}=\{(x,y(x)),\ x\in J\} \subset [-1,0[{\kern2pt}{\times}{\kern2pt}]0,c/2[.\end{align*} $$

Proposition 9.3. There exists a constant $y_{\mathrm {pert}}>0$ for which the following holds. Assume that $y:J_{M}\to ]0,y_{\mathrm {pert}}[$ , $x\mapsto y(x)$ is a differentiable function such that

$$ \begin{align*}\text{for all } x\in J_{M},\quad \bigg|\frac{d\ln y}{dx}+1\bigg|\leq 1/2.\end{align*} $$

Then, $\bar f_{\mathrm {pert}}({\gamma }_{J_{M},y})+({\mathbb Z},0)$ contains the graph ${\gamma\kern1pt }_{[-1,0[,\bar y}$ of a differentiable function $\bar y:[-1,0[{\kern2pt}\to {\mathbb R}_{+}^*$ (see Figure 6)

$$ \begin{align*}{\gamma}_{[-1,0[,\bar y}=\{(\bar x,\bar y(\bar x)),\ \bar x\in [-1,0[\},\end{align*} $$

such that

(9.82) $$ \begin{align} &\text{for all } \bar x\in [-1,0[,\quad \bigg|\frac{d\ln \bar y}{d\bar x}+1\bigg|\leq 1/2; \end{align} $$
(9.83) $$ \begin{align} &\sup_{\bar x\in [{-}1,0[} \ln\bar y(\bar x) \leq \sup_{x\in J_{M}}\ln y(x)-bM. \end{align} $$

Moreover, for some interval $J_{M}^{1}\subset J_{M}$ , one has

(9.84) $$ \begin{align} \gamma_{[{-}1,0[,\bar y}=\bar f_{\mathrm{pert}}(\gamma_{J_{1},y}).\end{align} $$

We prove this proposition in §9.2.2.

9.2.1 Preliminary results

If we introduce the variable

$$ \begin{align*}\varphi=\ln(s_{M}'(t))=\ln s_{M}'\circ s_{M}^{-1}(x+1)\quad (\text{recall}\ t=s_{M}^{-1}(x+1)),\end{align*} $$

we can write (9.80) as

(9.85) $$ \begin{align} (\bar x,\bar y)=f_{\mathrm{pert}}(x,y(x)) \ \ \iff\ \ \begin{cases} \bar x=t-1+\sigma(e^\varphi\times y(x))-\varphi-\ln y(x),\\ \ln \bar y=\varphi+\ln y(x). \end{cases}\end{align} $$

Note that the maps $I\ni t\mapsto \varphi =\ln s^{\prime }_{M}(t)\in \varphi (I)$ and $J_{M}\ni x\mapsto \varphi = \ln s_{M}'\circ s_{M}^{-1}(x+1) \in \varphi _{M}(I)$ are smooth diffeomorphisms. In particular, the maps $\varphi _{M}(I)\ni \varphi \mapsto \bar x$ and $\varphi _{M}(I)\ni \varphi \mapsto \ln y$ , $\varphi _{M}(I)\ni \varphi \mapsto \ln \bar y$ are well defined and smooth.

Lemma 9.4. For any $\varphi $ such that $t\in I$ , one has

$$ \begin{align*}\bigg |\frac{dt}{d\varphi}\bigg|\leq |I|/M\leq 1/4.\end{align*} $$

Proof. This follows from the identity (recall $\varphi =\ln (s_{M}'(t))$ , $s_{M}'=e^{\varphi _{M}}$ )

$$ \begin{align*}\frac{dt}{d\varphi}=\frac{1}{d\varphi/dt}=\frac{1}{\varphi_{M}'(t)}\end{align*} $$

and the estimates given by the second item of Lemma 9.1 (M is assumed to be large enough).

Lemma 9.5. One has

(9.86) $$ \begin{align}&\sup_{\varphi_{M}(I)}\bigg|\frac{d\bar x}{d\varphi}+1\bigg|\leq 1/4, \end{align} $$
(9.87) $$ \begin{align} &\hspace{-1pc}\sup_{\varphi_{M}(I)}\bigg|\frac{d\ln \bar y}{d\varphi}-1\bigg|\leq 1/4. \end{align} $$

Proof. Indeed, from (9.85),

$$ \begin{align*}\frac{d\bar x}{d\varphi}&=\frac{dt}{d\varphi}+e^\varphi\sigma'(e^\varphi\times y)\frac{dy}{d\varphi}-1-\frac{d\ln y}{d\varphi}\\ &=\frac{dt}{d\varphi}+ye^\varphi\sigma'(e^\varphi\times y)\frac{d\ln y}{d\varphi}-1-\frac{d\ln y}{d\varphi}\\ &=-1+A \end{align*} $$

with

$$ \begin{align*}A=\frac{dt}{d\varphi}+ye^\varphi\sigma'(e^\varphi\times y)\frac{d\ln y}{d\varphi}-\frac{d\ln y}{d\varphi}.\end{align*} $$

Note that (recall $x=s_{M}(t)-1$ , $s_{M}'=e^{\varphi _{M}}$ )

$$ \begin{align*}\frac{d\ln y}{d\varphi}=\frac{d\ln y}{dx}\frac{dx}{dt}\frac{dt}{d\varphi}=\frac{d\ln y}{dx}e^\varphi\frac{dt}{d\varphi}\end{align*} $$

so, by Lemma 9.4,

$$ \begin{align*}|A|\leq (|I|/M)+(ye^{-bM}\|\sigma'\|_{0}+1)e^{-bM}( |I|/M)\bigg|\frac{d\ln y}{dx}\bigg| \end{align*} $$

and if M is large enough,

(9.88) $$ \begin{align} |A|\leq 1/4.\end{align} $$

In a similar way,

$$ \begin{align*}\frac{d\ln \bar y}{d\varphi}=1+\frac{d\ln y}{dx}\frac{dx}{dt}\frac{dt}{d\varphi}=1+\frac{d\ln y}{dx}e^\varphi\frac{dt}{d\varphi}=1+B\end{align*} $$

with

(9.89) $$ \begin{align} |B|\leq 2e^{-bM}\times (1/4)\leq 1/4\quad (M\gg 1).\end{align} $$

9.2.2 Proof of Proposition 9.3

From (9.86) of Lemma 9.5, we see that the map $\varphi _{M}(I)\ni \varphi \mapsto \bar x\in {\mathbb R}$ is a diffeomorphism onto its image $\bar J_{M}\subset {\mathbb R}$ , and hence the maps $J_{M}\ni x\mapsto \bar x\in \bar J_{M}$ and $I\ni t\mapsto \bar x\in \bar J_{M}$ are diffeomorphisms. Note that from (9.86), one has

$$ \begin{align*}|\bar J_{M}|\geq (3/4)|\varphi_{M}(I)|\end{align*} $$

and from item (2) of Lemma 9.1, one has

(9.90) $$ \begin{align} |\bar J_{M}|\geq (3/4)(M/|I|)\times |I|>2;\end{align} $$

there thus exists an interval $J_{M}^{1}\subset J_{M}$ such that the map $J_{M}^{1}\ni x\mapsto \bar x\in n+ [-1,0[$ (for some $n\in {\mathbb Z}$ ) is a differentiable homeomorphism. Replacing $\bar y(\bar x)$ by $\bar y (\bar x+n)$ shows (9.84).

We now prove (9.82): for $\bar x\in [-1,0[$ ,

(9.91) $$ \begin{align}\bigg| \frac{d\ln \bar y}{d\bar x}+1\bigg|\leq 1/2. \end{align} $$

Indeed, let $I_{1}\subset I$ be the image of $[0,1[$ by $\bar J_{M}\ni \bar x\mapsto t\in I$ ; from Lemma 9.5, for any $\varphi \in \varphi _{M}(I_{1})$ , one has for some $A,B\in [0,1/4]$

$$ \begin{align*}\frac{d\bar x}{d\varphi}=-1+A,\quad \frac{d\ln \bar y}{d\varphi}=1+B,\end{align*} $$

so that

$$ \begin{align*}\bigg|\frac{d\ln \bar y}{d\bar x}+1\bigg|=\bigg| \bigg(\frac{d\ln\bar y}{d\varphi} \bigg/ \frac{d\bar x}{d\varphi}\bigg)+1\bigg|=\bigg|\frac{1+B}{-1+A}+1\bigg|\leq 1/2.\end{align*} $$

The preceding discussion shows that the map $\bar y:[-1,0[{\kern2pt}\ni \bar x\mapsto \bar y(\bar x)$ is a well-defined differentiable function, that its graph is included in $f_{\mathrm {pert}}(\gamma _{J_{M},y})+({\mathbb Z},0)$ , and that (9.82) holds.

There remains to prove (9.83). By the second equality of (9.85), if $(\bar x,\bar y(\bar x))=f_{\mathrm {pert}}(x,y)$ , one has

$$ \begin{align*}\ln\bar y(\bar x)\leq \ln y(x)-bM\leq \sup_{x\in J_{M}}\ln y-bM\end{align*} $$

and as a consequence, because the map $J_{M}\supset J_{M}^{1}\ni x\mapsto \bar x\in [-1,0[$ is a bijection, (9.83) holds.□

9.3 End of the proof of Theorem B

We shall prove that if M is large enough, the diffeomorphism $f_{\mathrm {pert}}$ constructed in §9.1 provides the searched for example of Theorem B.

Let M be large enough and $y_{0}\in{\kern2pt}]0,y_{\mathrm {pert}}[$ ; we define the function

$$ \begin{align*}y_{0}:[-1,0[{\kern2pt}\to {\mathbb R},\quad x\mapsto y_{0}e^{-x}.\end{align*} $$

Using inductively Proposition 9.3, we construct differentiable functions

$$ \begin{align*}y_{n}:[-1,0[{\kern2pt}\to {\mathbb R}\end{align*} $$

such that, for every $n\in {\mathbb N}^*$ ,

(9.92) $$ \begin{align} \text{for all } x\in J_{M},\quad \bigg|\frac{d\ln y_{n}}{dx}+1\bigg|\leq 1/2, \end{align} $$
(9.93) $$ \begin{align} \gamma_{[-1,0[,y_{n}}\subset \bar f_{\mathrm{pert}}(\gamma_{J_{M},y_{n-1}})+({\mathbb Z},0), \end{align} $$
(9.94) $$ \begin{align} \sup_{x\in [-1,0[} \ln y_{n}(x) \leq \sup_{x\in J_{M}}\ln y_{n-1}(x)-bM. \end{align} $$

Inclusion (9.93) implies the existence of a decreasing sequence of non-empty compact intervals $K_{n}\subset J_{M}$ such that

$$ \begin{align*}\gamma_{[-3/4,-1/4],y_{n}}=\bar f_{\mathrm{pert}}^n(\gamma_{K_{n},y_{0}})\!\!\mod ({\mathbb Z},0).\end{align*} $$

In particular, if $x_{\infty }\subset \bigcap _{n\in {\mathbb N}^*}K_{n}$ , one has

(9.95) $$ \begin{align} \text{for all } n\in{\mathbb N}^*,\quad \bar f_{\mathrm{pert}}^n((x_{\infty},y_{0}))\in \gamma_{[-3/4,-1/4],y_{n}}\subset \gamma_{[-1,0[,y_{n}}\!\!\mod ({\mathbb Z},0).\end{align} $$

From (9.94),

$$ \begin{align*}\sup_{x\in [-1,0[} y_{n}(x)\leq e^{-nbM}y_{0},\end{align*} $$

and hence, using (9.95), we see that $ \bar f_{\mathrm {pert}}^n((x_{\infty },y_{0}))$ accumulates ${\mathbb R}\times \{0\}$ :

(9.96) $$ \begin{align} \bar f_{\mathrm{pert}}^n((x_{\infty},y_{0}))\in [-1,0[{\kern2pt}{\times}{\kern2pt}]0,e^{-nbM}y_{0}[\!\!\mod ({\mathbb Z},0).\end{align} $$

As a consequence of (9.77) and of the fact that, for some constant $C>0$

$$ \begin{align*}\text{for all } \nu \in{\kern2pt}]0,c[,\quad h^{-1}([-1,0[{\kern2pt}{\times}{\kern2pt}]0,\nu[)\subset f_{\mathrm{pert}}^{-1} (\hat{\mathcal{F}}_{C\nu})\end{align*} $$

(this is owing to the fact that the diffeomorphism h given by (9.71) is indeed defined on a neighborhood of $\tilde {\mathcal {F}}_{y_{*}}$ ), one has

$$ \begin{align*}\hat f_{\mathrm{pert}}^n(h^{-1}(x_{\infty},y_{0}))\in f^{-1}_{\mathrm{pert}}( \hat{\mathcal{F}}_{Ce^{-nbM}y_{0}}).\end{align*} $$

Because $\hat f_{\mathrm {pert}}$ is the first return map of $f_{\mathrm {pert}}$ in $f^{-1}_{\mathrm {pert}}(\hat {\mathcal {F}}_{y_{*}})$ , there exists a sequence $(p_{n})_{n\in {\mathbb N}}\in {\mathbb N}^{\mathbb N}$ , $\lim _{n\to \infty } p_{n}=\infty $ such that

(9.97) $$ \begin{align} f_{\mathrm{pert}}^{p_{n}}(h^{-1}(x_{\infty},y_{0}))\in f^{-1}_{\mathrm{pert}}( \hat{\mathcal{F}}_{Ce^{-nbM}y_{0}}).\end{align} $$

However, this last fact prevents the existence of invariant circles in $\Delta _{\Sigma }$ accumulating the separatrix $\Sigma $ of $f_{\mathrm {pert}}$ . More precisely, let W be a neighborhood of $\Sigma $ in $\Sigma \cup \Delta _{\Sigma }$ (we recall that $\Delta _{\Sigma }$ is the bounded connected component of ${\mathbb R}^2\smallsetminus \Sigma $ ), such that

$$ \begin{align*} h^{-1}(x_{\infty},y_{0})\notin W. \end{align*} $$

We claim that $W\smallsetminus \Sigma $ does not contain any $f_{\mathrm {pert}}$ -invariant circle $\Gamma $ . Indeed, if this were not the case, the topological annulus $\mathcal {A}\subset W$ having $\Sigma $ and $\Gamma $ for boundaries would be $f_{\mathrm {pert}}$ -invariant (by topological degree theory). However, this is impossible because one would have at the same time

$$ \begin{align*} h^{-1}(x_{\infty},y_{0})\notin \mathcal{A}\quad \text{and}\quad f_{\mathrm{pert}}^{p_{n}}(h^{-1}(x_{\infty},y_{0}))\in \mathcal{A} \end{align*} $$

for some large $p_{n}$ (see (9.97)).

Remark 9.1. If we define the renormalization $\bar {\bar f}_{\mathrm {pert}}$ of $f_{\mathrm {pert}}$ by considering the first return map of $f_{\mathrm {pert}}$ in $\mathcal {F}_{y_{*}}$ instead of $f_{\mathrm {pert}}^{-1}(\mathcal {F}_{y_{*}})$ , as we have done to construct $\bar f_{\mathrm {pert}}$ , the dynamics of $\bar {\bar f}_{\mathrm {pert}}$ looks more like the one pictured in Figure 7. The comparison of this picture and that of Figure 4 illustrates the effect of the perturbative assumption in Theorem A.

Figure 7 The diffeomorphism $\bar {\bar f}_{\mathrm {pert}}$ on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}[e^{-(n+1)},e^{-n}]$ . Compare with Figures 4 and 6.

Acknowledgments

R.K. wishes to thank Bassam Fayad and the referee for their thorough reading of a preliminary version of this paper and their very useful comments. A.K. was partially supported by NSF grant DMS 1602409. R.K. was supported by a Chaire d’Excellence LABEX MME-DII, the project ANR BEKAM: ANR-15-CE40-0001 and an AAP project from CY Cergy Paris Université.

A Appendix. Proof of Lemma 2.2

We write for $j\geq 2$

$$ \begin{align*}H_{\varepsilon}^t(z)=\lambda_{\varepsilon} z_{1}z_{2}+\sum_{2\leq i\leq [j/2]}a_{\varepsilon,i}\times(z_{1}z_{2})^{i}+\sum_{\substack{i_{1},i_{2}\in{\mathbb N}\\ i_{1}+i_{2}=j+1 }}h_{\varepsilon,i_{1},i_{2}}(t)z_{1}^{i_{1}}z_{2}^{i_{2}}+O^{j+2}(z),\end{align*} $$

where $a_{\varepsilon ,i}\in {\mathbb R}$ and the $h_{\varepsilon ,i_{1},i_{2}}(\cdot )$ are smooth 1-periodic functions. We define

$$ \begin{align*}H_{\varepsilon,2}(z)=\lambda_{\varepsilon} z_{1}z_{2}.\end{align*} $$

We first observe that if $G_{\varepsilon }^t$ is a solution of

(A.98) $$ \begin{align} \begin{cases}G_{\varepsilon}^t(z)=O^{j+1}(z),\\ H_{\varepsilon}^t(z)+\partial_{t}G_{\varepsilon}^t(z)+\{G_{\varepsilon}^t,H_{\varepsilon,2}^t\}(z)=\tilde q_{\varepsilon}(z_{1}z_{2}),\end{cases}\end{align} $$

for some $\tilde q_{\varepsilon }(u)=\lambda _{\varepsilon }u+\sum _{2\leq i\leq [(j+1)]/2}\tilde a_{\varepsilon ,i}\times u^{i}$ , $\tilde a_{\varepsilon ,i}\in {\mathbb R}$ , then $G_{\varepsilon }^t$ solves (2.15). We then have to solve (A.98) for some $\tilde q_{\varepsilon }$ and some $G_{\varepsilon }^t$ of the form

$$ \begin{align*} &\tilde q_{\varepsilon}(u)=\lambda_{\varepsilon}u+\sum_{2\leq i\leq [(j+1)]/2}\tilde a_{\varepsilon,i}\times u^{i}\\ &G_{\varepsilon}^t(z)=\sum_{i_{1}+i_{2}=j+1}g_{\varepsilon,i_{1},i_{2}}(t)z_{1}^{i_{1}}z_{2}^{i_{2}}=O^{j+1}(z),\end{align*} $$

where the $g_{\varepsilon ,i_{1},i_{2}}(\cdot )$ are 1-periodic. This amounts to finding 1-periodic solutions to the equations

(A.99) $$ \begin{align} h_{\varepsilon,i_{1},i_{2}}(t)+\partial_{t}g_{\varepsilon,i_{1},i_{2}}(t)-\lambda_{\varepsilon}(i_{1}-i_{2})g_{\varepsilon,i_{1},i_{2}}(t)=0\quad & \text{if}\ i_{1}\ne i_{2}, \end{align} $$
(A.100) $$ \begin{align} h_{\varepsilon,i,i}(t)+\partial_{t}g_{\varepsilon,i,i}(t)=\tilde a_{\varepsilon,i} \quad &\text{if}\ i_{1}= i_{2}=i, \end{align} $$

for each couple $(i_{1},i_{2})\in {\mathbb N}^2$ such that $i_{1}+i_{2}=j+1$ . Note that in (A.100), this last equality occurs only if $j+1$ is even and $i=(j+1)/2$ . Equation (A.100) is then easily solved by setting

$$ \begin{align*}\tilde a_{\varepsilon,i}=\int_{{\mathbb R}/{\mathbb Z}}h_{\varepsilon,i,i}(t)\,dt,\quad g_{\varepsilon,i,i}(t)=-\int_{0}^t(h_{\varepsilon,i,i}(s)-\tilde a_{\varepsilon,i})\,ds.\end{align*} $$

Equation (A.99) always admits unique 1-periodic solutions of the form

$$ \begin{align*}\begin{cases}g_{\varepsilon,i_{1},i_{2}}(t)=e^{\lambda_{\varepsilon}(i_{1}-i_{2}) t}c_{\varepsilon,i_{1},i_{2}}-\int_{0}^t e^{(t-s)\lambda_{\varepsilon}(i_{1}-i_{2})}h_{\varepsilon,i_{1},i_{2}}(s)\,ds,\\ \text{where}\quad c_{\varepsilon,i_{1},i_{2}}=(e^{\lambda_{\varepsilon}(i_{1}-i_{2})}-1)^{-1}\int_{0}^1 e^{(1-s)\lambda_{\varepsilon}(i_{1}-i_{2})}h_{\varepsilon,i_{1},i_{2}}(s)\,ds. \end{cases}\end{align*} $$

In the preceding solutions, the dependence on $\varepsilon $ is smooth and if, for $\varepsilon =0$ , the functions $h_{0,i_{1},i_{2}}$ do not depend on t, we see that $g_{0,t_{1},t_{2}}$ is a constant.

This concludes the proof of Lemma 2.2.□

B Appendix. Extension of symplectic diffeomorphisms

Lemma B.1. Let $(\Theta _{\varepsilon })_{\varepsilon \in\, ]-\varepsilon _{0},\varepsilon _{0}[}$ be a smooth (or continuous) family of $C^k$ symplectic diffeomorphisms $C^1$ -close to the identity, defined on some open disk $D(o,\delta )$ of ${\mathbb R}^2$ , and such that $\Theta _{\varepsilon }(o)=o$ . Then, there exists $(\tilde \Theta _{\varepsilon })_{\varepsilon \in\, ]-\varepsilon _{0},\varepsilon _{0}[}$ , a smooth (or continuous) family of $C^k$ symplectic diffeomorphisms of ${\mathbb R}^2$ such that on $D(o,\delta /2)$ , one has $\tilde \Theta _{\varepsilon }=\Theta _{\varepsilon }$ .

Proof. We use the notation $\Theta _{\varepsilon }(x,y)=(\tilde x,\tilde y)$ . Because $\Theta _{\varepsilon }$ is symplectic, the 1-form $\tilde y d\tilde x-ydx$ is closed and defined on a disk $D(o,4\delta /5)$ of center o and radius $4\delta /5$ (we assume $\Theta _{\varepsilon } \ C^1$ -close to the identity so that we can use the implicit function theorem). It is hence locally exact and there exists a function $S_{\varepsilon }(y,\tilde y)$ such that $\tilde y d\tilde x-ydx=dS_{\varepsilon }$ . Now the function $F_{\varepsilon }(x,\tilde y)=-S_{\varepsilon }(y,\tilde y)+(\tilde x-x)\tilde y$ is defined on $D(0,3\delta /4)$ and satisfies $( y-\tilde y)dx+(\tilde x- x)d\tilde y=dF_{\varepsilon }$ or equivalently,

(B.101) $$ \begin{align} \Theta_{\varepsilon}(x,y)=(\tilde x, \tilde y)\ \ \iff\ \ \begin{cases}\tilde x=x+\partial_{\tilde y}F_{\varepsilon}(x,\tilde y),\\ y=\tilde y+\partial_{x}F_{\varepsilon}(x,\tilde y) .\end{cases}\end{align} $$

Note that we can choose $(F_{\varepsilon })_{\varepsilon }$ as a $C^k$ -family of $C^{k+1}$ -functions such that $F_{\varepsilon }(o)=0$ , $DF_{\varepsilon }(o)=0$ .

We can then choose $\chi :{\mathbb R}^2\to {\mathbb R}$ as a smooth function which is equal to 1 on $D(o,2\delta /3)$ and 0 outside $D(o,3\delta /4)$ , set

$$ \begin{align*}\tilde F_{\varepsilon}=\chi\times F_{\varepsilon},\end{align*} $$

and define $\tilde \Theta _{\varepsilon }$ by (B.101) with $F_{\varepsilon }$ replaced by $\tilde F_{\varepsilon }$ . The family of diffeomorphisms $(\tilde \Theta _{\varepsilon })_{\varepsilon }$ is a smooth (or continuous) family of exact symplectic $C^k$ -diffeomorphisms.

C Appendix. Proof of Lemma 9.1

Let $\chi :{\mathbb R}\to [0,1]$ be a smooth even function with support in $[-1/2,1/2]$ such that $\chi (0)=1$ and which is increasing on $[-1/2,0]$ . There exists $\alpha \in{\kern2pt}]0,1/4[$ such that, for all $x\in{\kern2pt}] {-}2\alpha ,2\alpha [ $ , one has $\chi (x)>1/2$ and

$$ \begin{align*}\beta_{\min}:=\min_{[-2\alpha,-\alpha]}\chi'>0,\quad \beta_{\max}:=\max_{[-2\alpha,-\alpha]}\chi'>0. \end{align*} $$

We define for $\rho \in{\kern2pt}]0,1/12]$ and $C_{M}>0$ ,

$$ \begin{align*}\varphi_{M}(x)=a(\rho,C_{M})\chi\bigg(\frac{x-1/3}{1/12}\bigg)-C_{M} \chi\bigg(\frac{x-2/3}{\rho}\bigg), \end{align*} $$

where $a(\rho ,C_{M})>0$ is chosen so that

$$ \begin{align*}\int_{0}^1e^{\varphi_{M}(u)}du=1.\end{align*} $$

Let $I_{}=(2/3)+{\kern2pt}] {-}2\alpha \rho ,-\alpha \rho [$ . For $x\in I_{}$ , one has

$$ \begin{align*}&\varphi_{M}(x)\leq -C_{M}/2=-C_{M}\alpha\beta_{\min}/(2\alpha\beta_{\min}),\\ &\varphi_{M}'(x)\leq -(C_{M}/\rho)\beta_{\min}=-(C_{M}\alpha\beta_{\min})/(\alpha\rho)=-C_{M}\alpha\beta_{\min}/|I_{}|,\\ &\varphi_{M}'(x)\geq -(C_{M}/\rho)\beta_{\max}=-(C_{M}\alpha\beta_{\max})/(\alpha\rho)=-(\beta_{\max}/\beta_{\min})C_{M}\alpha\beta_{\min}/|I_{}|. \end{align*} $$

Fixing $\rho $ (for example $\rho =1/12$ ) and taking

$$ \begin{align*} b^{-1}=\max\bigg(\frac{\beta_{\max}}{\beta_{\min}},2\alpha\beta_{\min}\bigg),\quad C_{M}=\frac{M}{\alpha\beta_{\min}}, \end{align*} $$

provides the first two items of Lemma 9.1.

Let us check the third item is satisfied. From the definition of $s_{M}$ , one has $s_{M}'(x)=e^{\varphi _{M}(x)}=1$ for $x\notin [0,1]$ . Because $s_{M}(0)=0$ , one has $s_{M}(x)=x$ for $x\leq 0$ . Similarly, because

$$ \begin{align*} s_{M}(1)=\int_{0}^1e^{\varphi_{M}(u)}du=1, \end{align*} $$

we have $s_{M}(x)=x$ for $x\geq 1$ .

Because in any case $s'(x)>0$ , this concludes the proof of Lemma 9.1.□

Footnotes

*

A preliminary version of this paper was discussed by the authors some months before Anatole Katok passed away in April 2018.

References

Angenent, S.. A remark on the topological entropy and invariant circles of an area preserving twist map. Twist Mappings and Their Applications (The IMA Volumes in Mathematics and its Applications, 44). Eds McGehee, R. and Meyer, K. R.. Springer, New York, 1992, pp. 15.Google Scholar
Arnaud, M.-C.. Three results on the regularity of the curves that are invariant by an exact symplectic twist map. Publ. Math. Inst. Hautes Études Sci. 109 (2009), 117.CrossRefGoogle Scholar
Arnaud, M.-C.. A non-differentiable essential irrational invariant curve for a ${C}^1$ symplectic twist map. J. Mod. Dyn. 5(3) (2011), 583591.CrossRefGoogle Scholar
Arnaud, M.-C.. Boundaries of instability zones for symplectic twist maps. J. Inst. Math. Jussieu 13(1) (2013), 1941.CrossRefGoogle Scholar
Avila, A. and Fayad, B.. Non-differentiable irrational curves for ${C}^1$ twist map. Ergod. Th. & Dynam. Sys. this issue, doi:10.1017/etds.2021.118.CrossRefGoogle Scholar
Avila, A. and Krikorian, R.. Reducibility or non-uniform hyperbolicity for quasi-periodic Schrödinger cocycles. Ann. of Math. (2) 164(3) (2006), 911940.CrossRefGoogle Scholar
Banyaga, A., de la Llave, R. and Wayne, C. E.. Cohomology equations near hyperbolic points and geometric versions of Sternberg linearization theorem. J. Geom. Anal. 6(4) (1996), 613649.CrossRefGoogle Scholar
Chaperon, M.. Géométrie differentielle et singularités des systèmes dynamiques. Astérisque 138–139 (1986), 444.Google Scholar
Fayad, B. and Krikorian, R.. Herman’s last geometric theorem. Ann. Sci. Éc. Norm. Supér. (4) 42(2) (2009), 193219.CrossRefGoogle Scholar
Herman, M.. Sur la conjugaison différentiable des difféomorphismes du cercle à des rotations. Publ. Math. Inst. Hautes Études Sci. 49 (1979), 5233.CrossRefGoogle Scholar
Herman, M. R.. Sur les courbes invariantes par les difféomorphismes de l’anneau. (Astérisque 1 (1983), 103104). With an appendix by Albert Fathi.Google Scholar
Herman, M. R.. Some open problems in dynamical systems. Proceedings of the International Congress of Mathematicians, Volume II (Berlin, 1998). Doc. Math. Extra Volume (1998), 797808.Google Scholar
Katok, A.. Lyapunov exponents, entropy and periodic orbits for diffeomorphisms. Publ. Math. Inst. Hautes Études Sci. 51 (1980), 137173.CrossRefGoogle Scholar
Krikorian, R.. Global density of reducible quasi-periodic cocycles on ${T}^1\times \mathrm{SU}(2)$ . Ann. of Math. (2) 154 (2001), 269326.CrossRefGoogle Scholar
Moser, J.. On invariant curves of area-preserving mappings of an annulus. Nachr. Akad. Wiss. Göttingen Math-Phys. Kl. II 1962 (1962), 120.Google Scholar
Polterovich, L.. The Geometry of the Group of Symplectic Diffeomorphisms (Lectures in Mathematics. ETH Zürich). Birkhäuser Verlag, Basel, 2001.CrossRefGoogle Scholar
Rüssmann, H.. Kleine Nenner I: Über invariante Kurven differenzierbarer Abbildungen eines Kreisringes. Nachr. Akad. Wiss. Gottingen, II Math.-Phys. Kl. 5 (1970), 67105.Google Scholar
Smale, S.. Differentiable dynamical systems. Bull. Amer. Math. Soc. (N.S.) 73 (1967), 747817.CrossRefGoogle Scholar
Sternberg, S.. On the structure of local homeomorphisms of Euclidean $n$ -space II. Amer. J. Math. 80 (1958), 623631.CrossRefGoogle Scholar
Sternberg, S.. The structure of local homeomorphism III. Amer. J. Math. 81 (1959), 578604.CrossRefGoogle Scholar
Treschev, D. and Zubelevich, O.. Introduction to the Perturbation Theory of Hamiltonian Systems (Springer Monographs in Mathematics). Springer-Verlag, Berlin, 2010.CrossRefGoogle Scholar
Yoccoz, J.-C.. Conjugaison différentiable des difféomorphismes du cercle dont le nombre de rotation vérifie une condition diophantienne. Ann. Sci. Éc. Norm. Supér. (4) 17(3) (1984), 333359.CrossRefGoogle Scholar
Yoccoz, J.-C.. Théorème de Siegel, nombres de Bruno et polynômes quadratiques, Petits diviseurs en dimension 1. Astérisque 231 (1995), 388.Google Scholar
Yoccoz, J.-C.. Analytic linearization of circle diffeomorphisms. Dynamical Systems and Small Divisors (Cetraro, 1998) (Lecture Notes in Mathematics/C.I.M.E. Foundation Subseries, 1784). Eds Marmi, S. and Yoccoz, J.-C.. Springer, Berlin, 2002, pp. 125173.Google Scholar
Figure 0

Figure 1 A (non-split) separatrix.

Figure 1

Figure 2 Fundamental domain $\mathcal {F}_{\varepsilon ,y_{*}}$ for $f_{\varepsilon }$ and the first return map $\hat f_{\varepsilon }$.

Figure 2

Figure 3 Gluing: $(\mathcal {F}_{\varepsilon ,y_{*}}\cup L_{y_{*}})/f_{\varepsilon }$.

Figure 3

Figure 4 The diffeomorphism $\bar f_{\varepsilon }$ on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}[e^{-(n+1)},e^{-n}]$.

Figure 4

Figure 5 The perturbed map $f_{\mathrm {pert}}$.

Figure 5

Figure 6 The image of the graph ${\gamma }_{J_{M},y}$ by the diffeomorphism $\bar f_{\mathrm {pert}}$.

Figure 6

Figure 7 The diffeomorphism $\bar {\bar f}_{\mathrm {pert}}$ on ${\mathbb R}/{\mathbb Z}{\kern2pt}{\times}{\kern2pt}[e^{-(n+1)},e^{-n}]$. Compare with Figures 4 and 6.