Abstract
Based on quantitative “kam theory”, we state and prove two theorems about the continuation of maximal and whiskered quasi-periodic motions to slightly perturbed systems exhibiting proper degeneracy. Next, we apply such results to prove that, in the three-body problem, there is a small set in phase space where it is possible to detect both such families of tori. We also estimate the density of such motions in proper ambient spaces. Up to our knowledge, this is the first proof of co-existence of stable and whiskered tori in a physical system.
Similar content being viewed by others
Avoid common mistakes on your manuscript.
1 Overview
1.1 Two kam Theorems for Properly Degenerate Hamiltonian Systems We deal with Hamiltonians which meet the demand of being close-to-be-integrable [see, e.g., Gallavotti (1986)], but, in addition, with the number of degrees of freedom of perturbing term being possibly larger than the one of the unperturbed part. Such kind of Hamiltonians often arise in problems of celestial mechanics and are referred to as “properly degenerate”, after (Arnold 1963). We denote them as
where the coordinates \((I, \varphi )=(I_1, \ldots , I_n, \varphi _1, \ldots , \varphi _n)\) are of “action-angle” kind (after a possible application of the Liouville–Arnold theorem to the unperturbed term), while (for our needs) the \((p, q)=(p_1, \ldots , p_m, q_1, \ldots , q_m)\) are “rectangular”, namely, take value in a small ball (say, of radius \((\varepsilon _0)\)) about some point (say, the origin). The symplectic form is standard:
We work in the real-analytic framework, which means that we assume that H admits a holomorphic extension on a complex neighborhood of the real “phase space” (namely, the domain)
where \(V\subset {{\mathbb {R}}}^n\) is bounded, open and connected, \(({{\mathbb {T}}}={{\mathbb {R}}}/(2\pi {\mathbb Z}))\) is the “flat torus”, \(B^{2m}_\varepsilon \) is the 2m-dimensional ball around 0 of radius \(\varepsilon \), relatively to some norm in \({\mathbb {R}}^{2m}\).
In this framework, we presentFootnote 1 two “kam theorems” which deal with different situations. A basic assumption, common to both statements, and often referred to as “Kolmogorov condition”, is:
- (A\({}_1\)):
-
the map \(I\rightarrow \partial _I H_0(I)\) is a diffeomorphism of V.
However, due to the proper degeneracy mentioned above, such assumption is to be reinforced with some statement concerning the perturbing term, or, more precisely, its Lagrange average
with respect to the \(\varphi \)-coordinates. Such extra-assumption will be different in the two statements; therefore, we quote them below.
The first result is a revisitation of the so-called Fundamental Theorem by V.I. Arnold, Arnold (1963). Such theorem has been already studied, generalized and extended in previous works (Chierchia and Pinzari 2010; Pinzari 2018). Here, we deal with the situation (not considered in the aforementioned papers) where \(P_\textrm{av}\) admits a “Birkhoff Normal Form” (bnf hereafter) about \((p, q)=(0, 0)\) of highFootnote 2 order; say s. As expected, a higher order of bnf allows to improve the measure of the “Kolmogorov set”, namely the set given by the union of all kam tori. We shall proveFootnote 3 the following
Theorem 1.1
Assume \((A_1)\) above and the following conditions:
- (A\({}_2\)):
-
\( P_\textrm{av} (I, p, q)=\sum _{j=1}^{s}\mathcal{P} _j(r;I) +\textrm{O}_{2 s+1}(p,q;I)\), with \(r_i:=\frac{p_i^2+q_i^2}{2}\) and \(\mathcal{P} _j(r;I)\) being a polynomial of degree j in \(r=(r_1, \cdots , r_{m})\), for some \(2\le s\in {{\mathbb {N}}}\).
- (A\({}_3\)):
-
the \(m\times m\) matrix \(\beta (I)\) of the coefficients of the second-order term \(\mathcal{P} _2(r;I)=\frac{1}{2}\sum _{i,j=1}^{m} {\beta }_{ij}(I)r_i r_j\) is non-degenerate: \(|\det {\beta }(I)|\ge {\, \mathrm const\,}>0\) for all \(I\in V\).
Then, there exist positive numbers \(\varepsilon _*<\varepsilon _0\), \(C_*\) and \(c_*\) such that, for
one can find a set \(\mathcal{K} \subset \mathcal{P} _\varepsilon \) formed by the union of H-invariant n-dimensional Lagrangian tori, on which the H-motion is analytically conjugated to linear Diophantine quasi-periodic motions with frequencies \(({\omega }_1,{\omega }_2)\in {\mathbb R}^{n_1}\times {{\mathbb {R}}}^{n_2}\) with \({\omega }_1=O(1)\) and \({\omega }_2=O({{\mu }})\). The set \(\mathcal{K} \) has positive Liouville–Lebesgue measure and satisfies
The second result deals with lower-dimensional quasi-periodic motions, the so-called whiskered tori. These are n-dimensional quasi-periodic motions (in a phase space of dimension \(2n+2m\)), approached or reached at an exponential rate. For simplicity, in view of our application, we focus on the case \(m=1\). In addition, we allow a further degeneracy in the Hamiltonian: the unperturbed term \(H_0\) may possibly depend not on all the I’s, but only on a part of them.
Theorem 1.2
Let \(m=1\), and let \(H_0\) depend on the components \(I_1=(I_{11}, \ldots , I_{1n_1})\) of the \(I=(I_1, I_2)\)’s, with \(1\le n_1\le n:=n_1+n_2\). Assume \((A_1)\) with \(I_1\) replacing I and, in addition, that
- (A\({}_2'\)):
-
\( P_\textrm{av} (I, p, q; {\mu })=P_0 (I, pq; {\mu })+ P_1 (I,\varphi , p, q; {\mu })\) with \(\Vert P_1\Vert \le a \Vert P_0\Vert \);
- (A\({}_3'\)):
-
\(| \partial _{pq} P_0 |\ge {\, \mathrm const\,}>0\) and \(|\det \partial ^2_{I_2} P_0 |\ge {\, \mathrm const\,}>0\) if \(n_2\ne 0\).
Fix \(\eta >0\). Then, there exist positive numbers \(a_*\), \(\varepsilon _*<\varepsilon _0\), \(C_*\) and \(c_*\) such that, if
one can find a set \(\mathcal{K} \) formed by the union of H-invariant n-dimensional Lagrangian tori, on which the H-motion is analytically conjugated to linear Diophantine quasi-periodic motions with frequencies \(({\omega }_1,{\omega }_2)\in {{\mathbb {R}}}^{n_1}\times {{\mathbb {R}}}^{n_2}\) with \({\omega }_1=O(1)\) and \({\omega }_2=O({{\mu }})\). The projection \(\mathcal{K} _0\) of set \(\mathcal{K} \) on \(\mathcal{P} _0:=V\times {{\mathbb {T}}}^{n}\) has positive Liouville–Lebesgue measure and satisfies
Furthermore, for any \(\mathcal{T} \in \mathcal{K} \) there exist two \((n+1)\)-dimensional invariant manifolds \(\mathcal{W}_\textrm{u}\), \(\mathcal{W}_\textrm{s}\subset \mathcal{P} _{\varepsilon _*}\) such that \(\mathcal{T} =\mathcal{W}_\textrm{u}\cap \mathcal{W}_\textrm{s}\) and the motions on \(\mathcal{W}_\textrm{u}\), \(\mathcal{W}_\textrm{s}\) leave, approach \(\mathcal{T} \) at an exponential rate.
Before we go on with describing how we aim to use the theorems above, we premise some comment.
-
(i)
The conditions involving \(\mu \) in (1) and (3) are not optimal. With a procedure similar to the one shown in Chierchia and Pinzari (2010, proof of Theorem 1.2, steps 1–4), one can show that they can be relaxed to, respectively
$$\begin{aligned} {\mu }<\frac{1}{{C_* (\log \varepsilon ^{-1})^{2b}}} ,\qquad {\mu }<\frac{1}{{C_* (\log (a \Vert P_0\Vert )^{-1})^{2b}}} \end{aligned}$$with some \(C_*\), \(b>0\).
-
(ii)
The careful bounds on the measure of the invariant sets provided in (2) and (4) are needed in view of our application. Indeed, we shall apply both the theorems above in order to prove that, in the three-body problem, closely to the co-planar, co-circular, outer retrograde configuration (see below for the exact definition), full-dimensional and “whiskered” quasi-periodic tori co-exist [the result was conjectured in Pinzari (2018)]. In the application, \(\varepsilon \) will correspond to the maximum eccentricity or inclination; a the semi-major axes ratio, and the use of a high-order bnf in Theorem 1.1 will be necessary because the size of the set goes to 0 with some power of \(\varepsilon \) (\(s=4\) will be enough for our application).
-
(iii)
Following Chierchia and Gallavotti (1994), Theorem 1.2 might be extended to prove the existence of “diffusion paths” and “whisker ladders”. We shall not do, as proving Arnold instability [in the sense of Arnold (1964)] for the system (5) below is not the purpose of this paper. We, however, remark that such kind of instability has been found for the four-body problem in a very similar framework (Clarke et al. 2022). We remark that proofs of chaos or Arnold instability in celestial mechanics are quite recent (Féjoz et al. 2014; Delshams et al. 2019), by the difficulty of overcoming the so-called problem of large gaps. See Guzzo et al. (2020) and references therein.
-
(iv)
Another important aspect in view of the application described above is a rather standard consequence of the proof of Theorem 1.2: If P (namely, \(P_1\)) has an equilibrium at \((p, q)=0\), then, along the motions of \(\mathcal K\), the coordinates (p, q) remain fixed at (0, 0) (rather than varying closely to it), namely
$$\begin{aligned} \mathcal{K}\subset V\times {{\mathbb {T}}}^{n}\times \{(0, 0)\} . \end{aligned}$$More generally, the stable and unstable invariant manifolds do not shift from the unperturbed ones:
$$\begin{aligned} \mathcal{W}_\textrm{s}\subset \mathcal{P} _{\varepsilon }\cap \big \{q=0\big \} ,\qquad \mathcal{W}_\textrm{u}\subset \mathcal{P} _{\varepsilon }\cap \big \{p=0\big \} . \end{aligned}$$
1.2 Application to the three-body problem We apply the results above to prove that, in a region of the phase space of the three-body problem, and under conditions that will be specified later, full dimensional and whiskered tori co-exist. We underline that the co-existence of such different kind of motions is not a mere consequence of the non-integrability of the system (as in such case the result would be somewhat expected) as it persists in two suitable integrable approximations of the system, close one to the other. Indeed, such motions will be found in a very small zone in the phase space of the three-body problem which simultaneously is in the neighborhood of an elliptic equilibrium of one of such approximations and in a hyperbolic one of the other. Such an occurrence is intimately related to the use of two different systems of coordinates, which are singular one with respect to the other, in the region of interest. The authors are not aware of the appearance of such phenomenon, previously.
After the “heliocentric reduction” of translational invariance, the three-body problem Hamiltonian with gravitational masses equal to \(m_0\), \({\mu }m_1\) and \({\mu }m_2\) and Newton constant \(\mathcal{G} \equiv 1\), takes the form of the two-particle system [see, e.g., Féjoz (2004), Laskar and Robutel (1995) for a derivation]:
with suitable values of \(\textrm{m}_i=m_i+\textrm{O}(\mu )\), \(\textrm{M}_i=m_0+\textrm{O}(\mu )\). We consider the system in the Euclidean space, namely we take, in (5), \(y^{\mathrm{(i)}}\), \(x^{\mathrm{(i)}}\in {\mathbb {R}}^3\), with \(x^{(1)}\ne x^{(2)}\).
We call Kepler maps the class of symplecticFootnote 4 coordinate systems \(\mathcal{C} =({\Lambda }_1, {\Lambda }_2, \ell _1, \ell _2, \mathrm y, \mathrm x)\) for the Hamiltonian (5), where \( \mathrm y=(y_1, \ldots , y_4), \mathrm x=(x_1, \ldots , x_{4})\), such that:
-
\(\Lambda _i=\textrm{m}_i \sqrt{\textrm{M}_i a_i}\), where \(a_i\) denotes the semi-major axis of the \(i^\textrm{th}\) instantaneousFootnote 5 ellipse;
-
\(\ell _1, \ell _2\in {\mathbb {T}}\) are conjugated to \({\Lambda }_1\), \({\Lambda }_2\). Such angles are defined in a different way according to the choice of \(\mathcal{C} \). In all known examples, they are related to the area spanned by the planet along the instantaneous ellipse.
Using a Kepler map, the Hamiltonian (5) takes the form
where \( \hat{\text {y}}, \hat{\text {x}}\) include the couples \( (\textrm{y}_i, \textrm{x}_i)\) suchFootnote 6 that nor \(\textrm{y}_i\) nor \(\textrm{x}_i\) is negligible. \( {{\hat{\text {y}}}}, \hat{\text {x}}\) are often called degenerate coordinates, because they do not appear in (6) when \(\mu \) is set to zero. In other words, \(\textrm{H}_\mathcal{C} \) is a properly degenerate close-to-be-integrable system, in the sense of the previous paragraph.
We call co-planar, co-circular, outer retrograde configuration the configuration of two planets in circular and co-planar motions, with the angular momentum of the outer planet having opposite verse to the resulting one. In Pinzari (2018) it has been pointed out that, under a careful choice of \(\mathcal{C} \) such configuration plays the rôle of an equilibrium for the \((\ell _1, \ell _2)\)-averaged perturbing function
But what matters more is that, closely to such equilibrium, there exist two such \(\mathcal{C} _i\)’s such that the Hamiltonian \(\textrm{H}_{\mathcal{C} _1}\) is suited to Theorem 1.1, while \(\textrm{H}_{\mathcal{C} _2}\) is suited to Theorem 1.2. This leads to the following result, which states co-existence of stable and whiskered quasi-periodic motions in the three-body problem. It will be made more precise (see Theorem 2.1) and proved along the paper.
Theorem A In the vicinity of the co-planar, co-circular, outer retrograde configuration, and provided that the masses of the planets and the semi-axes ratio are small, there exists a positive measure set \(\mathcal{K} _1\) made of 5-dimensional quasi-periodic motions \(\mathcal{T} _1\)’s “surrounding” (in a sense which will be specified) 3-dimensional quasi-periodic motions \(\mathcal{T} _2\)’s, each equipped with two invariant manifolds, called, respectively, unstable, stable manifold, where the motions are respectively asymptotic to the \(\mathcal{T} _2\)’s in the past, in the future.
We conclude with saying how this paper is organized.
-
In Sects. 2.1 and 2.2 we recall the main arguments of the discussion in Pinzari (2018), which lead to put the system (5) to a form suited to apply Theorems 1.1 and 1.2.
-
In Sects. 2.3 and 2.4 we check that the two domains where Theorems 1.1 and 1.2 apply have a non-empty intersection, and such intersection includes both families of tori. This check is subtle, because of the difference of the frameworks used.
-
In Sect. 3, we prove Theorems 1.1 and 1.2 via a carefully quantified kam theory.
2 Ellipticity and Hyperbolicity Closely to Co-planar, Co-circular, Outer Retrograde Configuration
Putting the system in a form suited to Theorem 1.1 requires identifying an elliptic equilibrium, while Theorem 1.2 calls for a hyperbolic one.
Denoting as \((\text {C}^{{(j)}}\,{:}{=}\,x^{{(j)}}\times y^{{(j)}}\) the angular momenta of the planets, we proceed to study motions evolving from initial data close to the manifold
The sub-fix “\({\pi }\)” recalls that \(\mathrm C^{(1)}\) and \(\mathrm C^{(2)}\) are opposite. In the two next sections, we recall material from Pinzari (2018), which highlights a sort of “double (elliptic, hyperbolic) nature” of \(\mathcal{M}_{\pi }\).
2.1 Ellipticity (with bnf)
BasicallyFootnote 7, the construction of the elliptic equilibrium—and of its associated bnf—proceeds as in Chierchia and Pinzari (2011). We briefly resume the procedure here.
We fix a domain \(\mathcal{D} _{{{c}}}\subset {{\mathbb {R}}}^{12}\) for impulse-position “Cartesian” coordinates
of two point masses relatively to a prefixed orthonormal frame \((k^{(1)}, k^{(2)}, k^{(3)})\) in \({{\mathbb {R}}}^3\). As a first step, we switch to a set of coordinates, well known in the literature, which we name jrd, after C. G. J. Jacobi, R. Radau and A. Deprit (Jacobi 1842; Radau 1868; Deprit 1983), who, at different stages, contributed to their construction.
We fix a region of phase space where the orbits \(t\rightarrow (x^{(j)}(t), y^{(j)}(t))\) generated by the unperturbed “Kepler” Hamiltonians
in (5) are ellipses with non-vanishing eccentricity. Then, we denote as \(\textrm{P}^{(j)}\) the unit vectors pointing in the directions of the perihelia; as \(a_j\) the semi-major axes; as \(\ell _j\) the “mean anomaly” of \(x^{(j)}\)(which, we recall, is defined as area of the elliptic sector from \(\textrm{P}^{(j)}\) to \(x^{(j)}\) “normalized at \(2{\pi }\)”); as \(\textrm{C}^{(j)}=x^{(j)}\times y^{(j)}\), \(j=1\), 2, the angular momenta of the two planets and \(\textrm{C}:=\textrm{C}^{(1)}+\textrm{C}^{(2)}\) the total angular momentum integral. We assume that the “nodes”
do not vanish, anytime. Such condition is equivalent to ask that the planes determined by the instantaneous ellipses and the \((k^{(1)}, k^{(2)})\) plane never pairwise coincide. As in previous works, we use the following notations. For three vectors u, v, w with u, \(v\perp \) w, we denote as \({\alpha }_{w}(u,v)\) the angle formed by u to v relatively to the positive (counterclockwise) orientation established by w. Then, the jrd coordinates are here denoted with the symbols
and defined via the formulae
The main point of jrd is that \(\textrm{Z}\), \(\zeta \) and \(\gamma \) are ignorable coordinates and \(\textrm{G}\) is constant along the motions of SO(3)-invariant systems. Therefore, most of motions of SO(3)-invariant systems are effectively described by the “reduced” coordinates \(\widehat{{jrd}}\). This strong property cannot be exploited in the case study of the paper, as the manifold \(\mathcal{M}_{\pi }\) in (7) is a singularity of the change (10). More generally, any co-planar or circularFootnote 8 configuration is so. Pretty similarly as in Chierchia and Pinzari (2011), we bypass such difficulty switching to new coordinates denoted as
where the \(\Lambda _j\)’s, \(\textrm{Z}\) and \(\zeta \) are the sameFootnote 9 as in (9), while
As in jrd, \((Z, \zeta )\) is a cyclic couple in SO(3)-invariant Hamiltonians but now no more cyclic coordinates but it appears. This leaves the system with 5 degrees of freedom and an extra-integral: the action \(\textrm{G}\) written using \({rps}_\pi \):
We denote as
the Hamiltonian (5) written in rps\({}_{\pi }\) coordinates, and worry about it.
We note that the manifold \(\mathcal{M}_\pi \) in (7) is now given by
Then we consider a neighborhood of \(\mathcal{M}_\pi \) of the form
where \(B^6_{\varepsilon _0}\) is the 6-ball centered at \(0\in {{\mathbb {R}}}^6\) with radius \(\varepsilon _0\); \({\mathbb T}\,{:}{=}\,{{\mathbb {R}}}/(2{\pi }{{\mathbb {Z}}})\) and \(\mathcal{L}\) is defined as
Here, \(0<{\Lambda }_-<{\Lambda }_+\) are arbitrarily taken (more conditions on such numbers will be specified in the course of the paper) and, for fixed positiveFootnote 10 numbers \(0<{\alpha }_-<{\alpha }_+<1\), \(k_\pm \) are constants depending on \(\alpha _\pm \) and the masses via
We now take \(0<\delta <1\) andFootnote 11 and assume
Then weFootnote 12 have
Proposition 2.1
[Pinzari (2018, Section III and Appendix A)] One can find \(\varepsilon _0>0\), depending only on \({\Lambda }_-\), \(\delta \), \({\alpha }_-\), \(m_1\), \(m_2\) such that the function \(\textrm{H}_{{rps}_{\pi }}\) in (13) is real-analyticFootnote 13 for \(({\Lambda },{\lambda },\eta ,\xi ,p,q)\in \mathcal{M}_{{rps}_\pi , \varepsilon _0}\). In addition, for any \(s\in {{\mathbb {N}}}\), there exists a positive number \({\alpha }^{\#}\) such that, if \({\alpha }_+<{\alpha }^{\#}\), there exists a positive number \(\varepsilon _1<\varepsilon _0\) and a real-analytic canonical transformation
which carries \((\overline{\eta }, \overline{\xi }, \overline{p}, \overline{q})=0\) to \((\eta , \xi , p, q)=0\) for all \((\overline{{\Lambda }}, \overline{{\lambda }})\in \mathcal{L}\times {\mathbb T}^2\), such that, if
then the averaged perturbing function
“is in Birkhoff Normal Form of order s”, namely:
where \({\Omega }({\Lambda })=({\Omega }_1({\Lambda }), {\Omega }_2({\Lambda }), {\Omega }_3({\Lambda }))\); \(\mathcal{P} _j(\overline{{\tau }};{\Lambda })\) are homogeneous polynomials of degree j in \(\overline{{\tau }}\,{:}{=}\,\left( \frac{\overline{\eta }_1^2+\overline{\xi }_1^2}{2},\ \frac{\overline{\eta }_2^2+\overline{\xi }_2^2}{2}\ ,\ \frac{\overline{p}^2+\overline{q}^2}{2}\right) \) and the determinant of the \(3\times 3\) matrix \(\textrm{T}({\Lambda })\) does not identically vanish. Moreover, \(\phi _{{bnf}}\) leaves \(\textrm{G}_{{rps}_{\pi }}\) unvaried, meaning that the function
is still a first integral to \(\overline{\textrm{H}}\).
2.2 Hyperbolicity
The hyperbolic character appears using a set of canonical coordinates, named perihelia reduction (p -coordinates). This is a further set of canonical coordinates
performing full reduction of SO(3) invariance for a n-particle system, which, in addition keeps regular for co-planar motions. The \(\text {{p}}\)-coordinates have been firstly introduced in Pinzari (2018), to which we refer for the proof of their canonical character. We remark that in (18), \(\textrm{G}\), \(\textrm{Z}\) and \(\zeta \) are the same as in jrd in (10). The coordinate \(\textrm{g}\), conjugated to \(\textrm{G}\), is not the same as in (10), but of course \((\textrm{Z}, \zeta , \textrm{g})\) are again ignorable and \(\textrm{G}\) is constant in SO(3) invariant systems. For the 3-body problem, namely, \(n=2\), the 8-plet \({\widehat{\text {{p}}}}\) is given by
with \(\Lambda _j\), \(\ell _j\), \(\textrm{G}_2\) as in (10). To define \(\Theta \), \(\textrm{g}\), \(\vartheta \) and \(\textrm{g}_2\), we assume that
do not vanish. Note that \({\nu }_1\) in (19) is the same as in (8). WeFootnote 14 let (under the same notations as in the previous section)
We now describe the rôle of the p-coordinates in the Hamiltonian (5). We denote as
where
the Hamiltonian (5) expressed in terms of \({{p}}\), and
the doubly averaged perturbing function. We look at the expansion
where \(\alpha \,{:}{=}\,\frac{a_1}{a_2}\) is the semi-major axes ratio. We focus on the function \(\textrm{P}\). Let \(\mathcal{L}\) as in (14); \(c\in (0,1)\), and put
and finally
Moreover, we let
Note that phase points in \(\mathcal{N}_{0}\) has the geometrical meaning of co-planar motions with the outer planet in retrograde motion.
Proposition 2.2
(Pinzari 2018, Section IV) The 4 degrees of freedom Hamiltonian \(\textrm{H}_{{p}}\) is real-analytic in \( \mathcal{N} \). It has an equilibrium on \(\mathcal{N}_{0}\). Such equilibrium turns to be hyperbolicFootnote 15 for \(\textrm{P}\).
2.3 Existence and Co-Existence of two Families of Tori
Theorems 1.1 and 1.2 can now be used to prove the existence of both full-dimensional and whiskered, co-dimension 2 tori in the three-body problem. Indeed,
-
Under conditions (1), by Theorem 1.1, an invariantFootnote 16 set \(\mathcal{F}\subset \mathcal{M}_{\varepsilon } \) for the Hamiltonian \(\textrm{H}_{{rps}_{\pi }}\) with 5-dimensional frequencies is found, whose measure satisfies
$$\begin{aligned} {\, \mathrm meas}\mathcal{M}_{\varepsilon }>{\, \mathrm meas}\mathcal{F}> \Big (1- C_* {\varepsilon }^{\frac{1}{2}+\overline{s}}\Big ) {\, \mathrm meas}\mathcal{M}_{\varepsilon } \end{aligned}$$(25)where \(\overline{s}=s-2\).
-
Under conditions (3) with \(a=\alpha _+\), by Theorem 1.2, for any \(\textrm{G}\in {{\mathbb {R}}}_+\), one finds an invariant set \(\mathcal{H}(\textrm{G})\subset \mathcal{N}_{0}(\textrm{G})\) with 3-dimensional frequencies for \(\textrm{H}_{{p}}\) and equipped with 4-dimensional stable and unstable manifoldsFootnote 17, whose measure satisfies
$$\begin{aligned} {\, \mathrm meas}\mathcal{N}_{0}(\textrm{G})>{\, \mathrm meas}\mathcal{H}(\textrm{G})> \Big (1- C_* \sqrt{ \alpha _+ }\Big ) {\, \mathrm meas}\mathcal{N}_{0}(\textrm{G})\ . \end{aligned}$$(26)
In the next, we show that the invariant sets \(\mathcal{F}\) and \(\mathcal{H}(\textrm{G})\) constructed above “have a common domain of existence”. We have to make this assertion more precise, mainly because \(\mathcal{F}\) and \(\mathcal{H}(\textrm{G})\) have been constructed with different formalisms.
Let
the canonical change of coordinates between \({rps}_{\pi }\) and \({p}\), well defined in a full measure set.
Let \({{\mathbb {G}}}_*\), \({{\mathbb {G}}}_0\) the respective images under the function (12):
of the sets \(\mathcal{M}_{\varepsilon }\), \(\mathcal{F}\). As \(\mathcal{F}\subset \mathcal{M}_{\varepsilon }\), then \({\mathbb G}_*\subset {{\mathbb {G}}}_0\). For any \(\textrm{G}_0\in {{\mathbb {G}}}_0\), \(\textrm{G}_*\in {{\mathbb {G}}}_*\), let
\(\mathcal{M}_{\varepsilon }(\textrm{G}_0)\) and \(\mathcal{F}(\textrm{G}_*)\) are invariant sets because \(\textrm{G}_{{rps}_{\pi }}\) is conserved along the motions of \(\textrm{H}_{{rps}}\).
Define:
At the cost of eliminating zero-measure sets from \({{\mathbb {G}}}_0\), \({{\mathbb {G}}}_*\), the sets \(\mathcal{F}'(\textrm{G}_*)\), \(\mathcal{M}'_{\varepsilon }(\textrm{G}_0) \) are well-defined, for all \(\textrm{G}_0\in {{\mathbb {G}}}_0\), \(\textrm{G}_*\in {{\mathbb {G}}}_*\). Then split
The volume-preserving property of \(\phi _{{rps}_{\pi }}^{{p}}\) in (27), the monotonicity of the Lebesgue integral and the bounds in (25) guarantee that
with some \(C_1>0\).
Recall now the definition of \(\mathcal{N}(\textrm{G})\), \(\mathcal{N}_{0}(\textrm{G})\) in (23) and \(\mathcal{H}(\textrm{G})\) in (26). The main result of the paper is the following
Theorem 2.1
Let \({\sigma }>0\) half-integer. There exist \(\varepsilon _*\), \(c_0\in (0, 1)\) such that, if \(\varepsilon <\varepsilon _*\), \(\textrm{G}_*\in {\mathbb G}_*\), \(\textrm{G}_*>c_0^{-1}\varepsilon ^2\), \(\alpha _+\le c_0 \varepsilon ^{12}\) and \(\mu \) verifies (1), (3) with \(a=\alpha _+\) and \(s=\sigma +\frac{7}{2}\), then there exists a non-empty set \(\mathcal{A}_{\star }(\textrm{G}_*) \) such that, letting
and denoting \(\widehat{\mathcal{F}}_*'(\textrm{G}_*)\), \(\widehat{\mathcal{H}}_*(\textrm{G}_*)\) the respective intersections of \(\widehat{\mathcal{F}}'(\textrm{G}_*)\), \(\widehat{\mathcal{H}}(\textrm{G}_*)\) with \(\mathcal{Q}(\textrm{G}_*) \), \(\mathcal{Q}_0(\textrm{G}_*) \) then \(\widehat{\mathcal{F}}_*'(\textrm{G}_*)\), \(\widehat{\mathcal{H}}_*(\textrm{G}_*)\) are non-empty and in fact verify
The proof of Theorem 2.1 relies on some technical result (Propositions 2.3, 2.4 and 2.5) which we now state and prove later.
Proposition 2.3
Let, for a suitable pure number \({\underline{k}}\in (1, 2)\), \(\Lambda _-<\textrm{G}\), \(k_-\le {\underline{k}}\) \(k_+\ge 2 \), \(\alpha _+\le \frac{c^2}{16}\). Choose \({\Lambda }_+\) as the unique value of \({\Lambda }_2> \textrm{G}\) such that \(\mathcal{C} \) and the straight line \({\Lambda }_1=2 {\Lambda }_2\) meet at \(({\Lambda }_1, {\Lambda }_2)=(2{\Lambda }_+, {\Lambda }_+)\). Let
Then, the set
is a subset of \(\mathcal{N}(\textrm{G})\).
Proposition 2.4
There exists \(c_1\in (0, 1)\) depending only on \({\Lambda }_+/\textrm{G}\), \({\Lambda }_-/\textrm{G}\) such that, letting, for any \({\gamma }< c_1^2\varepsilon ^2\),
then the set
is a subset of \( \widehat{\mathcal{M}}'_{\varepsilon }(\textrm{G})\).
Proposition 2.5
Assume \(\textrm{G}\ge 10c_1^2\varepsilon ^2\) and \({{\alpha }_+}<\frac{c^2}{16}\). Then, \(\mathcal{A}_0(\textrm{G})\) and \(\mathcal{A}_1(\textrm{G})\) have a non-empty intersection \(\mathcal{A}_\star (\textrm{G})\), verifying
We prove how Theorem 2.1 follows from the above propositions. \(\mathcal{Q}(\textrm{G}_*)\) is a subset of \(\widehat{\mathcal{M}}'_{\varepsilon }(\textrm{G}_*)\) and \(\mathcal{N}_{\varepsilon }(\textrm{G}_*)\), and
The bound in (28) guarantees that
On the other hand, if \(\widehat{\mathcal{F}}'_{\varepsilon }(\textrm{G}_*)\cap \mathcal{Q}(\textrm{G}_*) \) was empty, we would have
which contradicts the previous inequality if \(\overline{s}> \frac{3}{2}\) and \(\varepsilon \) is small. Finally, if \(\textrm{G}_*\in {{\mathbb {G}}}_*\),
and we have (29) with \({\sigma }=\overline{s}-\frac{3}{2}=s-\frac{7}{2}\), with \(s\ge 4\). The proof of (30) is similar.
2.4 Proof of Propositions 2.3, 2.4 and 2.5
Proof of Proposition 2.3
We only need to prove that \(\mathcal{L}_0(\textrm{G})\subset \mathcal{L}_{{p}}(\textrm{G})\). We switch to the coordinates
We denote as \(\mathcal{X}_{{p}}\,{:}{=}\,\textrm{G}^{-1}\mathcal{L} _{{p}}\) the domain of (y, x), and as
\(\mathcal{X}_{{p}}\) can be written as the intersection of the three sets:
We prove \(\mathcal{X}_0\,{:}{=}\,\textrm{G}^{-1}\mathcal{L} _0\) is a subset of all of them. The curve
passes through \(P_0=(1, 2)\). We denote as \({\underline{k}}\) the slope of the straight line \(y=kx\) which is tangent at \(\mathcal{C}\) at \(P_0\). The slope of the straight line \(y=kx\) through \(P_0\) is obviously \({\overline{k}}=2\). We assume that
and choose \((x_+, y_+)\) as the only (x, y) with \(x>1\) such that \(\mathcal{C} \) meets \(y=2 x\) at (x, y). Under such assumptions, we have:
The straight line which is tangent at \(\mathcal{C} \) at \(P_0=(1, 2)\) has equation
Since we \({\alpha }_+<\frac{{c^2}}{4}\), \(x>1\) and \(\mathcal{C} \) is convex, we have
This shows that \(\mathcal{X}_2\supset \mathcal{X}_0\). As for \(\mathcal{X}_3\), we note that for
it is
with
As \(y_-<0\) and \((y-2)(y_+-y)\ge 0\) on \(\mathcal{X}_0\), we have that \(\mathcal{X}_3\supset \mathcal{X}_0\) (Figs. 1, 2). \(\square \)
Remark 2.1
The numbers \({\underline{k}}\), \({\Lambda }_+\) of Proposition 2.3 can be chosen as
\({\Lambda }_+\) is related to the number \(x_+\) computed along the proof via \({\Lambda }_+=x_+\textrm{G}\). \({\underline{k}}\) is defined as the slope of the straight line \(y= kx \) which is tangent at \(\mathcal{C} \). We can compute it eliminating y between the two equations; we obtain the cubic equation
The tangency condition is imposed identifying this equation with
where a is the abscissa of the tangency point. Equating the respective coefficients of (31) and (32), we obtain
Eliminating b through the second and the third equations, we obtain
which has the following three roots:
The only admissible value is then
In correspondence of this value for a, solving the system in (33), we find
\(\square \)
Proof of Proposition 2.4
From (11), we get
From the equality
and the definition of \(\mathcal{N}_1(\textrm{G})\), the assertion trivially follows. \(\square \)
Proof of Proposition 2.5
Let \({\Lambda }_2^\star \) be the abscissa, in the plane \((\Lambda _2, \Lambda _1)\), of the intersection point between the curves
Using the coordinate \(x\,{:}{=}\,\frac{{\Lambda }_2}{\textrm{G}}\). With \(x^\star \,{:}{=}\,\frac{{\Lambda }_2^\star }{\textrm{G}}\), \(\theta \,{:}{=}\,\frac{c_1^2\varepsilon ^2}{\textrm{G}}\), \(\zeta \,{:}{=}\,\frac{{\gamma }}{\textrm{G}}\), where \(\zeta <\theta \), the set \(\mathcal{A}_\star (\textrm{G})\,{:}{=}\,\mathcal{A}_0(\textrm{G})\cap \mathcal{A}_1(\textrm{G})\) has measure
where
and where, for short, we have let \(m\,{:}{=}\,1-\frac{2}{ c}\sqrt{{\alpha }_+}\). Then,
To go further, we need a quantitative bound on \(x^\star \). Indeed, we have
Claim 2.1
If \(0<\theta <\frac{1}{10}\), then \(1+4\theta<x^\star <1+6\theta \).
The proof of the claim is postponed below, in order not to interrupt the main proof.
Since we have assumed \(\textrm{G}\ge 10c_1^2\varepsilon ^2\) and \(\alpha _+\le \frac{c^2}{16}\), then \(\textrm{G}\ge \frac{\frac{12}{ c}\sqrt{{\alpha }_+}}{1-\frac{2}{ c}\sqrt{{\alpha }_+}}c_1^2\varepsilon ^2\). In the new variables, this is \(\theta \le \frac{m}{6(1-m)}\). But then
whence
Observe that the second inequality is well put, because \(x^\star >1+4\theta \), as said. The function \(F_1(x)\) splits in the same intervals:
Since \(\zeta <\theta \), a lower bound to the integral in (34) is given by
with
the function in the second line in (35). Since F is the difference of a linear function and a convex one, it is concave. Then, we have
since \(F(x^\star )=0\) and \(F(1)=\theta \), this inequality becomes
hence
having used \(1+4\theta<x^\star <1+6\theta \).
It remains to prove Claim 2.1. \(x^\star \) is defined as the zero of the function F in (36) in the range \((1, +\infty )\). Multiplying the left hand side of Equation
by \(x+1+\theta +(1+x)\sqrt{\frac{x+4}{5}}\), we obtain the algebraic equation of degree three
which, for \(x\ge -1\) is completely equivalent to the initial equation. We aim to apply a bisection argument to the function at left hand side, which we denote as G(x). We have
and it is immediate to check that
To prove uniqueness, just observe that the function \(x\in (0, +\infty )\rightarrow G(x)\) is increasing for all \(\theta >0\). This completes the proof. \(\square \)
3 Quantitative kam Theory
3.1 Proof of Theorem 1.1
The proof of Theorem 1.1 is based on an application of Chierchia and Pinzari (2010, Proposition 3). The method is completely analogous to the one used in the proof of Chierchia and Pinzari (2010, Theorem 1.3), so we shall only say what to change in the proof of Chierchia and Pinzari (2010, Theorem 1.3) in order to obtain the proof of Theorem 1.1. The polynomial N(I, r) in the first non-numbered formula in Chierchia and Pinzari (2010, Section 4) is to be changed as
Equations (60) and (61) in Chierchia and Pinzari (2010) can be modified, respectively, as
Analogously to Chierchia and Pinzari (2010), one next applies Lemma A.1 in Chierchia and Pinzari (2010), but modifying the choice of K as
and leaving the other quantities unvaried. A bound as in Equation (62) in Chierchia and Pinzari (2010) is so obtained, with \(H_0\) as in Chierchia and Pinzari (2010), \(N(\overline{I},\overline{r})\) as in (37), \({\mu }{\widetilde{P}}_\textrm{av}(\overline{p}, \overline{q}, \overline{I})=f_{{bnf}}(I,\overline{p},\overline{q})-N(\overline{I},\overline{r})\) uniformly bounded by \(C{\mu }\varepsilon ^{2s+1}\), by (A\({}_2\)). Due to the choice of K in (39) and the one for \(\overline{{\gamma }}\) in (38), a bound similar to the one in Equation (63) in Chierchia and Pinzari (2010) holds, with the right hand side replaced by \(\overline{C}{\mu }\varepsilon ^{2s+1}\). At this point, one follows the indications in Step 2 of the proof of Theorem 1.3 in Chierchia and Pinzari (2010). Namely, one has to repeat the procedure in Steps 5 and 6 of the proof Theorem 1.4 [previously proved in altchierchiaPi10], with the following modification. The annulus \(\mathcal{A} (\varepsilon )\) in Equation (47) in Chierchia and Pinzari (2010) is to be taken as
and the number \(\breve{\rho }\) in
Equation (48) in Chierchia and Pinzari (2010) is to be replaced with \(\breve{\rho }\,{:}{=}\,\min \{{\check{c}}_1\varepsilon ^{s+\frac{1}{2}}/2,\ \overline{{\rho }}/ {48}\}\). The other quantities remain unvaried. In the remaining Steps 5 and 6 of the proof of Theorem 1.4 in Chierchia and Pinzari (2010) replace the number “5” appearing in all the formulae with \((2s+1)\) and \(\varepsilon ^{n_2/2}\) in Equation (56) (and the formulae below) in Chierchia and Pinzari (2010) with \(\varepsilon ^{m(s-\frac{3}{2})}\). \(\square \)
3.2 Proof of Theorem 1.2
The proof of Theorem 1.2 proceeds along the same lines as the proof of Theorem 1.1, apart for being based on a generalization (Theorem 3.1 below) of Chierchia and Pinzari (2010, Proposition 3) which now we state.
As in Chierchia and Pinzari (2010) \(\mathcal{D}_{\gamma _1, \gamma _2,{\tau }}\subset {{\mathbb {R}}}^{n}\) denotes the set of vectors \({\omega }=({\omega }_1, {\omega }_2)\in {{\mathbb {R}}}^{n_1}\times {{\mathbb {R}}}^{n_2}\) satisfying for any \(k=(k_1,k_2)\in {{\mathbb {Z}}}^{n_1}\times {{\mathbb {Z}}}^{n_2}\setminus \{0\}\), inequality
Theorem 3.1
Let \({n_1}\), \({n_2}\in {{\mathbb {N}}}\), \({n}\,{:}{=}\,{n_1}+{n_2}\), \(\tau >n\), \({{\gamma }_1}\ge {{\gamma }_2}>0\), \(0<s\le \frac{\varepsilon }{\overline{\varepsilon }+\varepsilon }\), \({\rho }>0\), \(A\,{:}{=}\,{D}_{\rho }\times B^2_{\overline{\varepsilon }+\varepsilon }\), and let
be real-analytic on \(A\times {{\mathbb {T}}}_{\overline{s}+s}^{n}\). Let
with \(\omega _k(I_1, I_2, pq)\,{:}{=}\,\partial _{I_k} \textrm{h}(I_1, I_2, pq)\), and assume that the map \(I\in D_{\rho }\rightarrow {\omega }(I,J)\) is a diffeomorphism of \(D_{\rho }\) for all \(J=pq\), with \((p,q)\in B^2_{\varepsilon }\), with non-singular Hessian matrix \(U(I,J)\,{:}{=}\,\partial _{I}^2\textrm{h}(I,J)\). LetFootnote 18
Assume, forFootnote 19 simplicity,
Define
Finally, let \(\overline{M}_1\), \(\overline{M}_2\) upper bounds on the norms of the sub-matrices \(n_1\times n\), \(n_2\times n\) of \(U^{-1}\) of the first \(n_1\), last \(n_2\) rowsFootnote 20. Assume the perturbation \(\textrm{f}\) so small that the following “KAM conditions” hold
Then, for any \(({\pi }, {\kappa })\in B^2_{\overline{\varepsilon }}\) and any \({\omega }_*\in {\Omega }_*({\pi }{\kappa })\,{:}{=}\,{\omega }({D}, {\pi }{\kappa })\cap \mathcal{D} _{\gamma _1, \gamma _2,\tau }\), one can find a unique real-analytic embedding
such that \(\mathcal{M} _{{\omega }_*}\,{:}{=}\,\phi _{{{\omega }_*}}({{\mathbb {T}}}^n\times B^2_{\overline{\varepsilon }})\) is a real-analytic \((n+2)\)-dimensional manifold, on which the \(\textrm{H}\)-flow is analytically conjugated to
In particular, the manifolds
are real-analytic n-dimensional \(\textrm{H}\)-invariant tori embedded in \({\, \mathrm Re\,}({D}_r)\times {{\mathbb {T}}}^{n}\times B^2_{\overline{\varepsilon }}\), equipped with \((n+1)\)-dimensional manifolds
on which the motions leave, approach \(\textrm{T}_{{{\omega }_*}}\) at an exponential rate. Let \( \textrm{T}_{\omega _*, 0}\) denote the projection of \(\textrm{T}_{\omega _*}\) on the \((I, \varphi )\)-variables, and \(\displaystyle \textrm{K}_0\,{:}{=}\,\bigcup _{{\omega }_*\in {\Omega }_*}\textrm{T}_{{\omega }_*, 0}\). Then \( \textrm{K}_0\) satisfies the following measureFootnote 21 estimate:
where \({D}_{\gamma _1, \gamma _2,\tau }\) denotes the \({\omega }_0(\cdot , 0)\)-preimage of \(\mathcal{D} _{\gamma _1, \gamma _2,\tau }\) and \(c_n\) can be taken to be \(\displaystyle c_n=(1+(1+2^8nE)^{2n})^2\).
Finally, the following uniform estimates hold for the embedding \(\phi _{\omega _*}\):
where \(v(\vartheta , {\pi }, {\kappa }; {\omega }_*)=(v_1(\vartheta , {\pi }, {\kappa }; {\omega }_*), v_2(\vartheta ,{\pi }, {\kappa }; {\omega }_*))\) and \(I^0({\pi }{\kappa }; {\omega }_*)=(I^0_1({\pi }{\kappa }; {\omega }_*),I^0_2({\pi }{\kappa }; {\omega }_*))\in D\) is the \({\omega }(\cdot , {\pi }{\kappa })\)—pre-image of \({\omega }_*\in {\Omega }_*({\pi }{\kappa })\). where \(r\,{:}{=}\,8 {n} {\widehat{E}} {\widetilde{{\rho }}}\), \( r'=2{\widehat{E}}\varepsilon \)
The proof of Theorem 3.1 is deferred to the next Sect. 3.3. Here, we prove how Theorem 1.2 follows from it.
As said, we follow the same ideas of the proof of Theorem 3.1, which in turn follows (Chierchia and Pinzari 2010, Theorem 1.3). By \((A'_{2})\),
At this point, proceeding as in Chierchia and Pinzari (2010, Proof of Theorem 1.3, Step 1) but with \(\epsilon ^5\) replaced by \(\epsilon \), under condition
by an application of Chierchia and Pinzari (2010, Lemma A.1), with \({\overline{K}}=\frac{6}{s_0}\log {\epsilon ^{-1}}\), \(r_p=r_q={\epsilon }_0\), \(r=4\rho ={\overline{{\rho }}}\,{:}{=}\,\min \left\{ \frac{{\overline{{\gamma }}}}{2\overline{M}{\overline{K}}^{{\tau }+1}},\ {\rho }_0\right\} \) (with \(\overline{M}\,{:}{=}\,\sup |\partial ^2_{I_1}H_0|\)), \({\rho }_p={\rho }_q={\epsilon }_0/4\), \({\sigma }=s_0/4\), \(\ell _1=n_1\), \(\ell _2=0\), \(m=n_2\) \(h=H_0\), \(g\equiv 0\), \(f={\mu }P\), \(A={\overline{D}}\,{:}{=}\,\omega _0^{-1}\mathcal{D} _{\gamma , \tau }\) (where \(\omega _0\) is as in \(A_1\) and \(\mathcal{D} _{\gamma , \tau }\) is the usual Diophantine set in \({\mathbb {R}}^n\), namely the set (40) with \(\gamma _1=\gamma _2\)), \({B}={B'}=\{0\}\), \(s=s_0\), \({\alpha }_1={\alpha }_2={\overline{{\alpha }}}=\frac{{\overline{{\gamma }}}}{2{\overline{K}}^{\tau }}\), and \({\Lambda }=\{0\}\), on the domain \(W_{\overline{v},\overline{s}}\) where \(\overline{v}=({\overline{{\rho }}}/2,{\epsilon }_0/2)\) and \(\overline{s}=s_0/2\), one finds a real-analytic and symplectic transformation \(\overline{\phi }\) which carries \(\textrm{H}\) to
where
whence (by (47)) also \({\overline{P}}={\mu }{\widetilde{P}}_\textrm{av}+{\widetilde{P}}\) is bounded by \({C}{\mu }a \Vert P_0\Vert \) on \(W_{{\overline{v}},{\overline{s}} }\).
The next step is to apply Theorem 3.1 to the Hamiltonian \(\overline{H}\). Since we can take
the numbers L, K, \({\widehat{\rho }}\) and \({\widetilde{\rho }}\) can be bounded, respectively, as
and
having let \(\gamma _2\,{:}{=}\,\mu \Vert P_0\Vert \overline{\gamma }_2\). Condition (41) is trivially satisfied for any \(\overline{\gamma }<1\), \(s\le 6\), while, from the bounds
one sees that conditions (42) hold taking
with a suitable \({\widehat{C}}>1\). By the thesis of Theorem 3.1, we can find a set of n-dimensional invariant tori \(\mathcal{K}\subset \mathcal{P} \) whose projection \(\mathcal{K} _0\) on \(\mathcal{P} _0\) satisfies the measure estimate
\(\square \)
3.3 Proof of Theorem 3.1
We fix the following notations.
-
in \({{\mathbb {R}}}^{n}\) we fix the 1-norm: \( |I|\,{:}{=}\,|I|_1\,{:}{=}\,\sum _{1\le i\le n_1}|I_i|\);
-
in \({{\mathbb {T}}}^{n}\) we fix the “sup-metric”: \( |\varphi |\,{:}{=}\,|\varphi |_{\infty }\,{:}{=}\,\max _{1\le i\le n}|\varphi _i|\) (mod \(2{\pi }\));
-
in \({{\mathbb {R}}}\) we fix the sup norm: \( |(p, q)|\,{:}{=}\,|(p, q)|_{\infty }\,{:}{=}\,\max \{|p|, |q|\}\);
-
for matrices we use the “sup-norm”: \( |{\beta }|\,{:}{=}\,|{\beta }|_{\infty }\,{:}{=}\,\max _{i,j}|{\beta }_{ij}|\);
-
we denote as \(B^n_{\varepsilon }(z_0)\) the complex ball having radius \(\varepsilon \) centered at \(z_0\in {{\mathbb {C}}}^n\). If \(z_0=0\), we simply write \(B^n_{\varepsilon }\).
-
if \(A\subset {{\mathbb {R}}}^{n}\), and \(r>0\), we denote by \(A_r\,{:}{=}\,\bigcup _{x_0\in A} B^n_r(x_0)\) the complex r-neighborhood of A (according to the prefixed norms/metrics above);
-
given \(A\subset {{\mathbb {R}}}^n\) and positive numbers r, \(\varepsilon \), s, we let
$$\begin{aligned} v\,{:}{=}\,(r, \varepsilon )\ ,\quad U_{v}\,{:}{=}\,{A}_r \times B^2_{ \varepsilon }\ ,\quad W_{v, s}\,{:}{=}\,{U}_v \times {\mathbb T}^{n}_{ s}\end{aligned}$$ -
if f is real-analytic on a complex domain of the form \(W_{v_0, s_0}\), with \(v_0=(r_0, \varepsilon _0)\), \(r_0>r\), \(\varepsilon _0>\varepsilon \), \(s_0>s\), we denote by \(\Vert f\Vert _{v,s}\) its “sup-Taylor–Fourier norm”:
$$\begin{aligned} \Vert f\Vert _{v, s}\,{:}{=}\,\sum _{k,{\alpha },{\beta }}\sup _{U_v}|f_{{\alpha },{\beta }, k}|e^{|k|s}\varepsilon ^{|({\alpha },{\beta })|}\end{aligned}$$(48)with \(|k|\,{:}{=}\,|k|_1\), \(|({\alpha },{\beta })|\,{:}{=}\,|{\alpha }|_1+|{\beta }|_1\), where \(f_{k, {\alpha }, {\beta }}(I)\) denotes the coefficients in the expansion
$$\begin{aligned} f=\sum _{\begin{array}{c} (k,{\alpha },{\beta })\in \textbf{Z}^{n}\times {{\mathbb {N}}}^\ell \times {{\mathbb {N}}}^\ell \\ {{\alpha }_i\ne {\beta }_i\forall i} \end{array}}f_{k, {\alpha },{\beta }}(I)e^{ik\cdot \varphi } p^{\alpha }q^{\beta }; \end{aligned}$$ -
if f is as in the previous item, \(K>0\) and \({{\mathbb {L}}}\) is a sub-lattice of \(\textbf{Z}^n\), \(T_Kf\) and \(\P _{{\mathbb {L}}} f\) denote, respectively, the K-truncation and the \({{\mathbb {L}}}\)-projection of f:
$$\begin{aligned} T_Kf{} & {} :=\sum _{\begin{array}{c} (k,{\alpha },{\beta })\in \textbf{Z}^{n}\times {{\mathbb {N}}}^\ell \times {{\mathbb {N}}}^\ell \\ {{\alpha }_i\ne {\beta }_i\forall i} |k|_1\le K \end{array}}f_{k, {\alpha },{\beta }}(I)e^{ik\cdot \varphi }p^{\alpha }q^{\beta }\ ,\quad \P _{{\mathbb {L}}} f\\{} & {} :=\sum _{\begin{array}{c} (k,{\alpha },{\beta })\in \textbf{Z}^{n}\times {{\mathbb {N}}}^\ell \times {\mathbb N}^\ell \\ {{\alpha }_i\ne {\beta }_i\forall i}, k\in {{\mathbb {L}}} \end{array}}f_{k, {\alpha },{\beta }}(I)e^{ik\cdot \varphi }p^{\alpha }q^{\beta }\end{aligned}$$with \(f_{k, {\alpha },{\beta }}(I)\,{:}{=}\,f_{k, {\alpha },{\beta }}(I, 0, 0)\). We say that f is \((K, {{\mathbb {L}}})\) in normal form if \(f=\P _{{\mathbb {L}}} T_{K}f\). If \({{\mathbb {L}}}\) is strictly larger than \(\{0\}\), we say that f is resonant normal form.
Proposition 3.1
(Partially hyperbolic averaging theory) Let \(H=h(I_1, I_2, pq)+ f(I,{\varphi },p,q) \) be a real-analytic function on \(W_{v_0, s_0}\), with \(v_0=(r_0, \varepsilon _0)\). Let K, r, s, \(\varepsilon \), \({\hat{r}}\), \({\hat{s}}\), positive numbers, with \({\hat{r}}<r/4\), \({\hat{s}}<s/4\) and \(\hat{\varepsilon }<\varepsilon /4\). Put \({\hat{\sigma }}\,{:}{=}\,\min \left\{ {\hat{s}}, \frac{{\hat{\varepsilon }}}{\varepsilon }\right\} \). Assume there exist positive numbers \({\alpha }_1\), \({\alpha }_2>0\), with \(\alpha _1\ge \alpha _2\), such that, for all \(k=(k_1, k_2, k_3)\in {{\mathbb {Z}}}^{n+1}\), \(0<|k|\le K\) and for all \((I, p, q)\in U_{r, \varepsilon }\),
and
with a suitable number \(c_1\). Then, one can find a real-analytic and symplectic transformation
with \(r_*=r-4{\hat{r}}\), \(s_*=s-4{\hat{s}}\), \(\varepsilon _*=\varepsilon -4{\hat{\varepsilon }}\), which conjugates H to
where g is \((K, \{0\})\) in normal form, and g, f verify
Finally, \(\Phi _*\) verifies
Proposition 3.1 is an extension of the Normal Form Lemma by Pöschel (1993). The extension pertains at introducing the (p, q) coordinates in the integrable part and leaving the amounts of analyticity \({\hat{r}}\), \({\hat{s}}\) and \({\hat{\varepsilon }}\) as independent. This is needed in order to construct the motions (44), where the coordinates \(({\pi }, {\kappa })\) are not set to (0, 0), but take value in a small neighborhood of it. A more complete statement implying Proposition 3.1 is quoted and proved in Sect. 3.4.
Below, we let \(B\,{:}{=}\,B^2_{\overline{\varepsilon }}(0)\); therefore, \(B_\varepsilon \) will stand for \(B^2_{\overline{\varepsilon }+\varepsilon }(0)\).
Lemma 3.1
(kam Step Lemma) Under the same assumptions and notations as in Theorem 3.1, there exists a sequence of numbers \({\rho }_j\), \(\varepsilon _j\), \(s_j\); of domains
and a real-analytic and symplectic transformations
such that
and such that the following holds. Letting \(E_0\,{:}{=}\,E\), \((M_{0}, \overline{M}_{0}, \widehat{M}_{0}, L_{0})=(M, \overline{M}, {\widehat{M}}, L)\), \(s_0\,{:}{=}\,s\), \(\rho _0\,{:}{=}\,\rho \), \(\varepsilon _0\,{:}{=}\,\varepsilon _0\), \(\lambda _0\,{:}{=}\,\lambda \) and, given, for \(0\le j\in {{\mathbb {Z}}}\), \(E_j\), \((M_{j}, \overline{M}_{j}, {\widehat{M}}_{j}, L_{j})\), \(s_j\), \(\rho _j\), \(\varepsilon _j\), \(\lambda _j\), define
Then, for all \((p_{j+1}, q_{j+1})\in B_{\varepsilon _{j+1}}\),
-
(i)
\(D_{j+1}(p_{j+1}q_{j+1})\subseteq {(D_{j}}(p_{j+1}q_{j+1}))_{{\widehat{{\rho }}}_{j}/ {4}}\). Letting
$$\begin{aligned}{} & {} \varpi _{j+1}\,{:}{=}\,\partial _{(I_{j+1}, p_{j+1}q_{j+1})}\textrm{h}_{j+1}(I_{j+1}, p_{j+1}q_{j+1}))\\{} & {} \quad =(\omega _{j+1}(I_{j+1}, p_{j+1}q_{j+1}), \nu _{j+1}(I_{j+1}, p_{j+1}q_{j+1})) \end{aligned}$$the map \(I_{j+1}\rightarrow \omega _{j+1}(I_{j+1}, p_{j+1}q_{j+1})\) is a diffeomorphism of \(\big (D_{j+1}(p_{j+1}q_{j+1})\big )_{{\rho }_j}\) verifying
$$\begin{aligned} \omega _{j+1}(D_{j+1}(p_{j+1}q_{j+1})), p_{j+1}q_{j+1})=\omega _{j}(D_{j}(p_{j+1}q_{j+1})), p_{j+1}q_{j+1}) . \end{aligned}$$The map
$$\begin{aligned} {\widehat{\iota }}_{j{+1}}(p_{j+1}q_{j+1})= & {} ({\widehat{\iota }}_{j{+1,} 1}(p_{j+1}q_{j+1}),{\widehat{\iota }}_{j{+1,}2}(p_{j+1}q_{j+1})):\\ D_{j}(p_{j+1}q_{j+1})\rightarrow & {} D_{j+1}(p_{j+1}q_{j+1})\\ I_j(p_{j+1}q_{j+1})\rightarrow & {} I_{j+1}(p_{j+1}q_{j+1})\,{:}{=}\,\omega _{j+1}^{-1}\big (\omega _{j}(I_j, p_{j+1}q_{j+1}), p_{j+1}q_{j+1}\big ) \end{aligned}$$verifies
$$\begin{aligned}{} & {} \sup _{{ D}_{j}}|{\widehat{\iota }}_{j+1, 1}(p_{j+1}q_{j+1})-{\, \mathrm id \,}|\le 3{n} \frac{{\overline{M}}_1}{{\overline{M}}}{\widehat{E}} _{j}{\widetilde{{\rho }}}_{j}\le 3{n} {\widehat{E}} _{j}{\widetilde{{\rho }}}_{j}\ ,\nonumber \\{} & {} \sup _{{ D}_{j}}|{\widehat{\iota }}_{j+1, 2}(p_{j+1}q_{j+1})-{\, \mathrm id \,}|\le 3{n} \frac{{\overline{M}}_2}{{\overline{M}}}{\widehat{E}} _{j}{\widetilde{{\rho }}}_{j}\le 3{n}{\widehat{E}} _{j}{\widetilde{{\rho }}}_{j} \end{aligned}$$(57)$$\begin{aligned}{} & {} \mathcal{L}(\widehat{\iota }_{j+1}(p_{j+1}q_{j+1})-{\, \mathrm id \,})\le 2^9{n} {\widehat{E}} _{j} \end{aligned}$$(58) -
(ii)
the perturbation \({f}_j\) has sup-Fourier norm
$$\begin{aligned}\Vert {f}_j\Vert _{(W_j)_{\rho _j, \varepsilon _j, s_j}}\le E_j \end{aligned}$$ -
(iii)
the real-analytic symplectomorphisms \(\Psi _{j+1}\) in (53) verify
$$\begin{aligned}{} & {} \sup _{(W_{j+1})_{{\rho }_{j+1}, \varepsilon _{j+1},s_{j+1}}}|{I_{j, 1}}(I_{j+1},\varphi _{j+1}, p_{j+1}, q_{j+1})-{I_{j+1,1}}|\le \frac{3}{ {4 }} \frac{{\widehat{M}}_j }{ {M_j}} {\widehat{E}}_j{{\widetilde{{\rho }}}_j}\nonumber \\{} & {} \sup _{(W_{j+1})_{{\rho }_{j+1}, \varepsilon _{j+1},s_{j+1}}}|{I_{j, 2}}(I_{j+1},\varphi _{j+1}, p_{j+1}, q_{j+1})-{I_{j+1,2}}|\le \frac{3}{ {4 }}{\widehat{E}}_j {{\widetilde{{\rho }}}_j}\nonumber \\{} & {} \sup _{(W_{j+1})_{{\rho }_{j+1}, \varepsilon _{j+1},s_{j+1}}}|\varphi _{j}(I_{j+1},\varphi _{j+1}, p_{j+1}, q_{j+1})-\varphi _{j+1}|\le \frac{3}{ {4 }}{\widehat{E}}_j s_j\nonumber \\{} & {} {\sup _{(W_{j+1})_{{\rho }_{j+1}, \varepsilon _{j+1},s_{j+1}}}|p_{j}(I_{j+1},\varphi _{j+1}, p_{j+1}, q_{j+1})-p_{j+1}|\le \frac{3}{ {4 }}{\widehat{E}}_j \varepsilon _j}\nonumber \\{} & {} {\sup _{(W_{j+1})_{{\rho }_{j+1}, \varepsilon _{j+1},s_{j+1}}}|q_{j}(I_{j+1},\varphi _{j+1}, p_{j+1}, q_{j+1})-q_{j+1}|\le \frac{3}{ {4 }}{\widehat{E}}_j \varepsilon _j} . \end{aligned}$$(59)The rescaled dimensionless map \({\check{\Phi }}_{j+1}\,{:}{=}\,{\, \mathrm id \,}+{1}_{{\widehat{{\rho }}}_0^{-1},s_0^{-1}, \varepsilon _0^{-1}}\left( \Phi _{j+1}-{\, \mathrm id \,}\right) \circ {1}_{{\widehat{{\rho }}}_{0},s_{0}, \varepsilon _{0}}\) has Lipschitz constant on \((W_{j+1})_{{\rho }_{j+1}/{\widehat{\rho }}_0, \varepsilon _{j+1}/\varepsilon _0,s_{j+1}/s_0}\)
$$\begin{aligned} \mathcal{L}({\check{\Phi }}_{j+1}-{\, \mathrm id \,})\le 6(n+1)\big (12\cdot (24)^\tau \big )^{j}{\widehat{E}} _{j}\,;\end{aligned}$$(60) -
(iv)
for any \(j\ge 0\), \({\widehat{E}} _{j+1}<{\widehat{E}} _{j}^2\), \(\lambda _j\ge \frac{\lambda _0}{2}\).
Proof
The proof of this proposition is obtained generalizing (Chierchia and Pinzari 2010, Lemma B.1). We shall limit ourselves to describe only the different points, leaving to the interested reader the easy work of completing details.
We construct the transformations (53) by recursion, based on Proposition 3.1. For simplicity of notations, we shall systematically eliminate the sub-fix “j” and replace “\(j+1\)” with a “+”. As an example, instead of (53), we shall write
When needed, the base step will be labeled as “0” (e.g., (76) below). Let us assume (inductively) that
Condition (61) is verified at the base step provided one takes \(D_0=\omega _0^{-1}(\mathcal{D}_{\gamma _1, \gamma _2, \tau }, p_0q_0)\); (62) is so by assumption, while (63) follows from (41):
We aim to apply Proposition 3.1 with \(\varepsilon \), s of Proposition 3.1 corresponding now to \(\overline{\varepsilon }+\varepsilon \), \(\overline{s}+s\), and//
We check that that (61) and (62) imply conditions (49) and (50). We start with (49). If \((I, p, q)=(I_{1}, I_{2}, p_, q)\in A_{{\widehat{\rho }}, \varepsilon }\) and \(k\in \textbf{Z}^3\setminus \{0\}\), with \(|k|_1\le K\), then there exists some \(I_0(pq)=(I_{01}(pq), I_{02}(pq))\) such that \(|I-I_0(pq)|<{\widehat{\rho }}\) and \(\omega (I_0(pq), pq)=(\omega _{01}, \omega _{02})\in \mathcal{D}_{\gamma _1, \gamma _2, \tau }\). We have
having used (63). The bounds above have been obtained considering separately the cases \(k_3\ne 0\) and \(k_3= 0\), and:
–if \(k_3\ne 0\), taking the infimum of the modulus of the imaginary part of the expression between the |’s; observing that \(\overline{\omega }_0=(\overline{\omega }_{01}, \overline{\omega }_{02})\) are real and bounding the differences \(|{\, \mathrm Im\,}\big ({\omega }_i(I, pq)-{\omega }_i(I(pq), pq)\big )|\) with \(MK{\widehat{\rho }}\) (when \(i=1\)), \(\widehat{M}K{\widehat{\rho }}\) (when \(i=2\)) and using the definition of \({\widehat{\rho }}\) in (55).
–if \(k_3=0\), using the Diophantine inequality and again bounding the differences \(|{\, \mathrm Im\,}\big ({\omega }_i(I, pq)-{\omega }_i(I(pq), pq)\big )|\) as in the previous case and using the definition \({\widehat{\rho }}\).
We now check condition (50). The inequality \(Ks> 8\log 2\) is trivial by definition of K (see (54)), and also, the smallness condition (50) is easily met, since \({\hat{\sigma }}=\min \{\frac{1}{8}\frac{\varepsilon }{\overline{\varepsilon }+\varepsilon } ,\ \frac{s}{8} \}=\frac{s}{8}\), \(\delta =2^{-6}\min \{{\hat{\rho }} s ,\ \varepsilon ^2\}=2^{-6}{\widetilde{{\rho }}} s\) (by the definition of \({\widetilde{\rho }}\) in (56)), whence
having used \(L\ge {\widehat{M}}^{-1}\), \(M^{-1}\), so \({\alpha }_2\ge K L^{-1}{\widehat{{\rho }}}\), \(2^{6}c_1<{\widehat{c}}\), and (62). Thus, by Proposition 3.1, \(\textrm{H}\) may be conjugated to
where
while, by (51) and the choice of K,
The conjugation is realized by an analytic transformation
Using (52), \({\widetilde{\rho }}\le \varepsilon ^2/s\), \({\alpha }_1\ge {MK{\widehat{{\rho }}}}\), \({\alpha }_2=\frac{\gamma _2}{2K^\tau }\ge {L^{-1} K{\widehat{{\rho }}}}\), \(Ks\ge {6}\) and the definition of \({\widehat{E}}\), we obtain the bound (59) with, at the left hand side, the set \(W_{{\widehat{{\rho }}}/2, \varepsilon /2, s/2}\). Below we shall prove that \({W_{+}}_{{\rho }_{+}, \varepsilon _{+},s_{+}}\subset W_{{\widehat{{\rho }}}/2, \varepsilon /2, s/2}\), so we shall have (59).
We now evaluate the generalized frequency
with
(the “new frequency map”) and
(the “new Lyapunov exponent”).
Lemma 3.2
Let \((p_+, q_+)\in B_{\varepsilon /2}\). The new frequency map \( {{\omega }}_+\) is injective on \( D(p_+q_+)_{{\widehat{{\rho }}}/2}\) and maps \( D(p_+q_+)_{{\widehat{{\rho }}}/4}\) over \( {{\omega }}(D, p-+q_+)\). The map \({\widehat{\iota }}_+(p_{+}q_{+})=(\widehat{\iota }_{+1}(p_{+}q_{+}),{\widehat{\iota }}_{+2}(p_{+}q_{+}))\,{:}{=}\, {{\omega }}_+^{-1}\circ {{\omega }}|_{D(p_+q_+)}\) which assigns to a point \(I_0\in D(p_+q_+)\) the \( {{\omega }}_+(\cdot , p_+q_+)\)-preimage of \( {{\omega }}(I_0, p_+q_+)\) in \({ D}(p_+q_+)_{{\widehat{{\rho }}}/4}\) satisfies
The Jacobian matrix \(U_+\,{:}{=}\,\partial ^2_{ {I_+}}\textrm{h}_+ {(I_+, p_+q_+)}\) is non-singular on \( D_{{\widehat{{\rho }}}/ {4}} {\times B^2_{\varepsilon /2}}\) and the following bounds hold
where \( U_+^{-1}{=}{:}\left( \begin{array}{lrr} T_{+1}\\ T_{+2} \end{array} \right) \). Finally, the new Lyapunov exponent \(\nu _+(I_+, p_+q_+)\) satisfies
Postponing for the moment the proof of this lemma, we let \({\rho }_+\,{:}{=}\,{\widehat{{\rho }}}/ 2\), \(s_+\,{:}{=}\,s/2\), \(\varepsilon _+=\varepsilon /2\) and \( D_+(p_{+}q_{+})\,{:}{=}\,\widehat{\iota }_+(p_{+}q_{+})(D(p_{+}q_{+}))\). By Lemma 3.2, \(D_+\) is a subset of \(D_{{\widehat{{\rho }}}/4}\) and hence
We prove that \( {\widehat{E}} _+=\frac{E_+L_+}{\widehat{\rho }_+^2}\le {\widehat{E}} ^2 \). Since
we have
Finally, (42), (70) and the definition of \({\tilde{E}}\) imply \(\lambda _+\ge \frac{\lambda }{2}\). Collecting all bounds, we get
and
Now, using, in the last inequality, the bound
(since \({\hat{\rho \le }} \frac{{\gamma }_1}{ 2M K^{{\tau }+1}}\) and \(K\ge 6/s\)) we find
(having used \(s\le 1/2\)). We now prove that \(\lambda _{+}\ge \frac{\lambda _0}{2}\). Iterating (70) and using \({\widehat{\rho }}_k\le {\widehat{\rho }}_{k-1}/4\), \({\widetilde{\rho }}_k\le {\widetilde{\rho }}_{k-1}/4\), \(\varepsilon _k= \varepsilon _{k-1}/4\), \(L_k=2L_{k-1}\), (75) and the second condition in (42) with \({\tilde{c}}=2^6\), we get
This allows to check (63) at the next step: using (64) and (73), we have
Finally, (57) and (58) follow from (68), while the estimate in (60) is a consequence of (59), (71), (72), (74), inequality \(LM\ge 1\) and Cauchy estimates:
Proof of Lemma 3.2
The proof of this proposition is obtained generalizing (Chierchia and Pinzari 2010, Lemma B.2). As above, we limit to discuss only the different parts.
By (51),
(where \({\overline{f}}\) denotes the average of f). Therefore we may bound
This shows that the function (66) has a Jacobian matrix
which is invertible for all \((p_+, q_+)\in B^2_{\varepsilon /2}\) and satisfies
In a similar way one proves (69). Next, for any fixed \((p_+, q_+)\in B^2_{\varepsilon /2}\) and \(\overline{\omega }=\omega (I(p_+q_+), p_+q_+)\in \omega (D, p_+q_+) \) with \(I(p_+q_+)\in D\), we want to find \(I_+=I_+(p_+q_+)\in D_+\) such that
To this end, we consider the function
As F differs from \(\omega _+\) by a constant, we have
Similarly, we bound the quantities
and
Putting everything together, we get
By the implicit function theorem (e.g., (Celletti and Chierchia 1998, Theorem 1 and Remark 1)), Equation (77) has a unique solution
with
so we can take
This ensures that (61) holds also for \(D_+\).
Finally, the real part of the function (67) satisfies the lower bound
The proof of (68) proceeds as in Chierchia and Pinzari (2010, proof of Lemma B.2). \(\square \)
Proof of Theorem 3.1.
Step 1 Construction of the “generalized limit actions”
Let \(({\pi }, {\kappa })\in B_0=B^2_{\overline{\varepsilon }}=\bigcap _{j\ge 0}B_{\varepsilon _j}\). Define, on \(D_0({\pi }{\kappa })=\omega _0^{-1}(\mathcal{D} _{\gamma _1, \gamma _2, \tau }, {\pi }{\kappa })\cap D\),
Then \({\check{\iota }}_j({\pi }{\kappa })\) converge uniformly to a \({\check{\iota }} ({\pi }, {\kappa })=({\check{\iota }}_1({\pi }, {\kappa }), {\check{\iota }}_2({\pi }, {\kappa }))\) verifying
Moreover, as
we have
In particular, taking \(j=0\),
Moreover,
So \({\check{\iota }}({\pi }{\kappa })\) is bi-Lipschitz, with
Step 2 Construction of \(\phi _{\omega _*}\). For each \(j\ge 1\), the transformation
is defined on \((W_j)_{\rho _j, s_j, \varepsilon _j}\). If
then, by (79), \(W_*\subset \bigcap _j (W_j)_{\rho _j, s_j, \varepsilon _j}\). The sequence \(\Phi _j\) converges uniformly on \(W_*\) to a map \(\Phi \). We then let
with \(v(\vartheta , {\pi }, {\kappa }; \omega _*)\,{:}{=}\,\big (v_1(\vartheta , {\pi }, {\kappa }; \omega _*), v_2(\vartheta , {\pi }, {\kappa }; \omega _*)\big )\). Since (59) imply, on \(W_*\),Footnote 22
and similarly,
then, in view of (78), (81), (82), the definition of \(W_*\) and the triangular inequality, we have (46). Equations (80), (81), (82) also imply
where
Finally, with similar arguments as in Step 1, by (84), the rescaled map
has Lipschitz constant
In particular, \({\check{\Phi }}\), hence, \(\Phi \), and, finally, the map \((\vartheta ,{\pi }, {\kappa }; {\omega })\rightarrow \phi _{\omega }(\vartheta , {\pi }, {\kappa })\) are bi-Lipschitz, hence, injective.
Step 3 For any \({\omega }_*\in \mathcal{D} _{\gamma _1, \gamma _2,{\tau }}\cap {\omega }_0(D, 0)\), \(\textrm{T}_{{\omega }_*}\) in (83) is a n-dimensional \(\mathrm H\)-invariant torus with frequency \({\omega }_*\). This assertion is a trivial generalization of its analogue one in Chierchia and Pinzari (2010, Proof of Proposition 3, Step 3); therefore, its proof is omitted.
Step 4 Measure Estimates (proof of (45)) The proof of (45) proceeds as in Chierchia and Pinzari (2010, Proof of Proposition 3, Step 4), just replacing the quantities that in Chierchia and Pinzari (2010, Proof of Proposition 3, Step 4) are called
with the quantities here denoted as
\(\square \)
3.4 Normal Form Theory
Proposition 3.1 can be obtained from the more general Proposition 3.2, taking \(m=1\), \({\mathbb L}=\{0\}\) and changing coordinates as follows:
We define \(c_m\) to be the smallest number such that, for any two functions, real-analytic in \(W_{r, s, \varepsilon }\) and any choice of \({\hat{r}}<r\), \({\hat{s}}<s\), \({\hat{\varepsilon }}<\varepsilon \),
Proposition 3.2
Let \(\{0\}\subset {{\mathbb {L}}}\subset {{\mathbb {Z}}}\). Proposition 3.1 holds true taking
replacing \(c_1\) with \(c_m\), \(\P _0\) with \(\P _{{\mathbb {L}}}\) and condition (49) with
where
Lemma 3.3
Let \({\hat{r}}<r/2\) \({\hat{s}}<s/2\), \({\hat{\varepsilon }}<\varepsilon /2\) and \({\delta }\,{:}{=}\,\min \{{\hat{r}}{\hat{s}},\ {\hat{\varepsilon }}^2\}\). Let
be real-analytic on \(W_{v,s,\varepsilon }\). Assume that inequality (85) and
are satisfied. Then, one can find a real-analytic and symplectic transformation
defined by the time-one flowFootnote 23\(X_\phi ^1f\,{:}{=}\,f\circ \Phi \) of a suitable \(\phi \) verifying
such that
and, moreover, the following bounds hold
Finally, for any real-analytic function F on \(W _{v, s, \varepsilon }\),
Sketch of proof Lemma 3.3 is a straightforward generalization of Pöschel (1993, Iterative Lemma). To obtain such generalization, just replace the norm defined in Pöschel (1993, Section 1) with the norm (48), where
and bound the “ultraviolet remainders”, namely the norm of the functions whose expansion (87) includes only terms with \(|(k, \alpha -\beta )|_1>K\), as follows. Observe that, if \(|(k, \alpha -\beta )|_1>K\), then either \(|k|_1>K/2\) or \(|\alpha -\beta |_1>K/2\). In the latter case, a fortiori, \(|\alpha |_1+|\beta |_1\ge |\alpha -\beta |_1>K/2\). Then we have, for such functions, \(\Vert f\Vert _{r, s-{\hat{s}}, \varepsilon -{\hat{\varepsilon }}}\le \max \left\{ e^{-K{\hat{s}}/2} ,\ \left( \frac{\varepsilon -{\hat{\varepsilon }}}{\varepsilon }\right) ^{K/2}\right\} \Vert f\Vert _{r, s, \varepsilon }\). Other details are omitted.
Proof of Proposition 3.2
Let
By Lemma 3.3, we find a canonical transformation \(\Phi _1=X_{\phi _1}\) which is real-analytic on \(W_{ r_1, s_1, \varepsilon _1}\) and conjugates \(H=H_0\) to \(H_1=H_0\circ \Phi _1=h+g_1+f_1\), where \(g_1=\P _{{{\mathbb {L}}}}T_K f_0\) and
having used
We now focus on the case
otherwise the lemma isFootnote 24 proved. Then, we have
Note that
Assume now that, for some \(j\ge 1\), it is \(H_j=H_{j-1}\circ \Phi _j=h+g_j+f_j\), where
We have just proved this is true when \(j=1\). Let \(L\,{:}{=}\,\left[ \frac{K{\hat{\sigma }}_0}{8\log 2}\right] \). We prove that (88) is true for \(j+1\), for all \(1\le j\le L\). Let
Note that, for all \(1\le j\le L\), it is \({\hat{r}}_j<\frac{r_j}{2}\):
Similarly, \({\hat{s}}_j<\frac{s_j}{2}\), \({\hat{\varepsilon }}_j<\frac{\varepsilon _j}{2}\). Let then
so that \(r_j=r_1-2(j-1)\frac{{\hat{r}}_0}{L}\), etc., for all \(1\le j\le L\). Then
and Lemma 3.3 applies again, and \(H_j\) can be conjugated to \(H_{j+1}=H_j\circ \Phi _{j+1}=h+g_{j+1}+f_{j+1}\), with
To bound the right hand side of the latter expression, we use (89) and observe that
having used \(e^{-\frac{K\widehat{s}_0}{2L}}\le e^{-\frac{K{\hat{\sigma }}_0}{2L}}\), \(e^{-\frac{K{{\widehat{\varepsilon }}_0}}{2\varepsilon _1L}}\le e^{-\frac{K{{\widehat{\varepsilon }}_0}}{2\varepsilon _0L}}\le e^{-\frac{K{\hat{\sigma }}_0}{2L}}\) and \(L\le \frac{K{\hat{\sigma }}_0}{8\log 2}\). Moreover, writing
with \(f_{0}^{{{\mathbb {L}}}, K}\) real-analytic on \(W_{r_0, s_0, \varepsilon _0}\), while \(f_{j-1}^{{{\mathbb {L}}}, K}\) real-analytic on \(W_{r_{j-1}, s_{j-1}, \varepsilon _{j-1}}\) and verifying
we get
Collecting all such bounds we get
The inductive claim \(j\rightarrow j+1\) is thus proved, for all \(1\le j\le L\). Letting now \(\Phi _*\,{:}{=}\,\Phi _1\circ \cdots \circ \Phi _{L+1}\) and
and using \(L+1> \frac{K{\widehat{{\sigma }}}_0}{8\log 2}\), we get
as claimed. The bounds (52) are obtained from (86), by usual telescopic arguments. \(\square \)
Notes
Arnold (1963), Chierchia and Pinzari (2010), Pinzari (2018) deal with the “minimal” case \(s=2\). The case \(s=2\) is called here “minimal” as we work in the framework of generalizations of the Kolmogorov condition \((A_1)\) above. In Rüssmann (2001), Féjoz (2004), using different techniques, the case \(s=1\) has been considered.
Namely verifying
$$\begin{aligned} \Omega =d{\Lambda }\wedge d\ell +d\textrm{y}\wedge d\textrm{x}=\sum _{i=1}^2 d{\Lambda }_i\wedge d\ell _i+\sum _{i=1}^4d\textrm{y}_i\wedge d\textrm{x}_i . \end{aligned}$$With reference to the three-body Hamiltonian (5), the \(i^\textrm{th}\) instantaneous ellipse is the orbit generated by \(\textrm{h}_i:=\frac{|y^{\mathrm{(i)}}|^2}{2\textrm{m}_i}-\frac{\textrm{m}_i\textrm{M}_i}{|x^{\mathrm{(i)}}|}\) in a region of phase space where \(\textrm{h}_i\) is negative.
The reason of this is that the Hamiltonian (5) has first integrals, as recalled in the next section.
As pointed out in Pinzari (2018), the only note-worthing difference with the case studied in Chierchia and Pinzari (2011) (which deals with prograde motions of the planets, namely, revolving all in the same verse) is that here the elliptic character of the equilibrium does not follow for free from the symmetry of the Hamiltonian, but is checked manually.
Circular configurations correspond to \(\textrm{G}_i=\Lambda _i\); co-planar configurations correspond to \(\textrm{G}={\sigma }_1\textrm{G}_1+{\sigma }_2 \textrm{G}_2\), with \(({\sigma }_1, {\sigma }_2)\in \{\pm 1\}^2\setminus \{(-1, -1)\}\).
There is an inessential difference between the Definition (11) and the one in Pinzari (2018, Eqs. (25), (26))). Denoting as \(\overline{{rps}}_{\pi }\,{:}{=}\,\Big (({\Lambda }_1,{\Lambda }_2,\overline{{\lambda }}_1,\overline{{\lambda }}_2, \overline{\eta }_1,\overline{\eta }_2, \overline{\xi }_1,\overline{\xi }_2, \overline{p}, \overline{q}), (P, Q) \Big )\) the coordinates defined in Pinzari (2018), we have the following relations:
$$\begin{aligned} \left\{ \begin{array}{l} \displaystyle \overline{\lambda }_1=\lambda _1+\zeta \\ \displaystyle \overline{\lambda }_2=\lambda _2-\zeta \\ \displaystyle \overline{\eta }_1+\textrm{i}\overline{\xi }_1=(\eta _1+\textrm{i}\xi _1)e^{-\textrm{i}\zeta }\\ \displaystyle \overline{\eta }_2+\textrm{i}\overline{\xi }_2=(\eta _2+\textrm{i}\xi _2)e^{\textrm{i}\zeta }\\ \displaystyle \overline{p}+\textrm{i}\overline{q}=(p+\textrm{i}q) e^{\textrm{i}\zeta }\\ \displaystyle P+\textrm{i}Q=\sqrt{2(\textrm{G}-\textrm{Z})}\,e^{-\textrm{i}\zeta } \end{array}\right. \end{aligned}$$But as \((Z, \zeta )\), (P, Q) and \(\zeta \) do not appear in the Hamiltonian, its expression does not change.
Observe that \({\alpha }_-\) and \({\alpha }_+\) have the meaning of lower and upper bound for the semi-major axes ratio \(\alpha =a_1/a_2\), namely,
$$\begin{aligned} {\alpha }_-\le {\alpha }\le {\alpha }_+\qquad \forall \ ({\Lambda }_1, {\Lambda }_2)\in \mathcal{L} \ . \end{aligned}$$Indeed, from the formula
$$\begin{aligned} \frac{\Lambda _1}{\Lambda _2}=\frac{m_1}{m_2}\sqrt{\frac{m_0+{\mu }m_2}{m_0+{\mu }m_1}\alpha }\end{aligned}$$we find
$$\begin{aligned} {\alpha }\le \left( \frac{m_2}{m_1}\right) ^2\frac{m_0+{\mu }m_1}{m_0+{\mu }m_2} k_+^2={\alpha }_+ \end{aligned}$$and, similarly, \({\alpha }\ge {\alpha }_-\).
The reader might ask the reason of inequalities in (16). This is related to the fact that we want to investigate a region of phase space where the inner planet, labeled as “1”, has a larger angular momentum, namely, \(\mathrm G_1>\mathrm G_2\), and, simultaneously, the masses of the planets, as well as their eccentricities and mutual inclination are small. As, when eccentricities and mutual inclination go to zero, the \(\mathrm G_i\) reduce to \(\Lambda _i\), by (14), the number \(k_-\) in (15) needs to be strictly larger than 1. Conditions (16) are apt to ensure this, as in fact they immediately imply
$$\begin{aligned} 1-{\delta }\le \frac{m_0+{\mu }m_2}{m_0+{\mu }m_1}\le 1+{\delta }\end{aligned}$$hence, by (15),
$$\begin{aligned} k_-\ge \frac{m_1}{m_2}\sqrt{(1-{\delta }){\alpha }_-}>1 . \end{aligned}$$The proof in Pinzari (2018, Appendix A) is given with \(\delta = 1-\frac{1}{4\chi ^2}\ge \frac{3}{4}\), but works well also for any \(\delta \in (0, 1)\). Indeed, for \(({\Lambda }_1, {\Lambda }_2)\in \mathcal{L}\),
$$\begin{aligned} {\Lambda }_1-{\Lambda }_2={\Lambda }_2\left( \frac{{\Lambda }_1}{{\Lambda }_2}-1\right) \ge {\Lambda }_- (k_-1)\ge {\Lambda }_- \left( \frac{m_1}{m_2}\sqrt{(1-{\delta }){\alpha }_-}-1\right) \ . \end{aligned}$$Therefore, for \(({\Lambda }_1, {\Lambda }_2)\) on a complex neighborhood of \(\mathcal{L}\) depending on \({\Lambda }_-\), \(m_1\), \(m_2\), \({\alpha }_-\) and \({\delta }\) we shall have \(|{\Lambda }_1-{\Lambda }_2|\ge \frac{{\Lambda }_-}{2} \left( \frac{m_1}{m_2}\sqrt{(1-{\delta }){\alpha }_-}-1\right) \) and, as in the proof of Pinzari (Pinzari (2018), Proposition III.2), one can take \(\varepsilon _0<\frac{{\Lambda }_-}{2} \left( \frac{m_1}{m_2}\sqrt{(1-{\delta }){\alpha }_-}-1\right) \) in order that the denominators of the functions \(\textrm{c}_1^*\), \(\textrm{c}_2\), \(\textrm{c}^*_2\) in (Pinzari 2018, Appendix A) do not vanish, and so small that collisions are excluded.
Namely, analytic on a complex neighborhood of \(\mathcal{M}_{{rps}_\pi , \varepsilon _0}\) and real-valued on \(\mathcal{M}_{{rps}_\pi , \varepsilon _0}\).
The second equality in the first equation in (20) is implied by \(\textrm{C}=\textrm{C}^{(1)}+\textrm{C}^{(2)}\) and \(\textrm{C}^{(1)}\cdot \textrm{P}^{(1)}=0 \).
In Pinzari (2018) a slightly more general result is proved: the equilibrium is hyperbolic when \(\mathcal{L}_{{p}}\) in (21) is defined without the inequality
and \(\mathcal{G}_{{p}}\) in (22) is taken to be \(\Big \{\textrm{G}_2:\ \max \{\frac{2}{ c}\sqrt{{\alpha }_+}{\Lambda }_2, \textrm{G}\}<\textrm{G}_2<\min \{ {\Lambda }_2, \textrm{G}^{\star }\}\Big \}\), with \(\textrm{G}^\star \) the unique root of the polynomial \(\textrm{G}_2\rightarrow 5\Lambda _1^2\textrm{G}-(\textrm{G}+\textrm{G}_2)(4\textrm{G}+\textrm{G}_2)\). But as (24) ensures \(\Lambda _2<\textrm{G}^\star \), under such restriction, \(\mathcal{G}_{{p}}\) can be taken as in (22).
More precisely, Theorem 1.1 is applied to the Hamiltonian \(\textrm{H}_{{bnf}}\) in (17), hence with
$$\begin{aligned} n_1=2 ,\ n_2=3 ,\ V=\mathcal{L} ,\ \varepsilon =\varepsilon _1 ,\ H_0=\textrm{h}_\textrm{k} ,\ P=f_{{bnf}} \end{aligned}$$corresponding to the image under \({\overline{\phi }}\) of the invariant set obtained through the thesis of Theorem 1.1.
Theorem 1.2 is applied to the Hamiltonian \(\textrm{H}_{{p}}\) of Proposition 2.2, hence with
$$\begin{aligned} n_1=2 ,\ n_2=1 ,\ V=\mathcal{A} _{{p}}(\textrm{G}) \\ H_0=\textrm{h}_\textrm{k}-\frac{m_1 m_2}{a_2} ,\ P_0=-\frac{m_1 m_2}{a_2}\alpha ^2\textrm{P},\ P_1=-\frac{m_1 m_2}{a_2}\textrm{O}(\alpha ^3) ,\ a=\alpha _+ \end{aligned}$$The norms will be specified in the next Sect. 3.3.
(41) is a simplifying assumption. It may be relaxed.
That is, \(\displaystyle \overline{M}_i\ge \sup _{D_{\rho }}\Vert T_i\Vert \ ,\quad i=1,\ 2\ ,\quad \text {if}\quad U^{-1}=\left( \begin{array}{lrr} T_1\\ T_2 \end{array} \right) \ .\)
\({\, \mathrm meas}_n\) denotes the n-dimensional Lebesgue measure.
\(\P _z\) denotes the projection on the z-variables.
The time-one flow generated by \(\phi \) is defined as the differential operator
$$\begin{aligned} X_\phi ^1\,{:}{=}\,\sum _{k=0}^\infty \frac{\mathcal{L}_\phi ^k}{k!} \end{aligned}$$where \(\mathcal{L}^0_\phi f\,{:}{=}\,f\) and \(\mathcal{L}^k_\phi f\,{:}{=}\,\big \{\phi , \mathcal{L}^{k-1}_\phi f\big \}\), with \(k=1,\ 2,\ \cdots \).
Indeed, in such case,
$$\begin{aligned} \Vert f_1\Vert _{v_1,s_1,\varepsilon _1}\le 4 e^{-K{\hat{{\sigma }}}_0/2}\le e^{-K{\hat{{\sigma }}}_0/4} \end{aligned}$$because \(e^{-K{\hat{{\sigma }}}_0/4}\le \frac{1}{4}\) having chosen \(K\hat{\sigma }_0\ge 8\log 2\).
References
Arnold, V.I.: Small denominators and problems of stability of motion in classical and celestial mechanics. Russ. Math. Surv. 18(6), 85–191 (1963)
Arnold, V.I.: Instability of dynamical systems with many degrees of freedom. Dokl. Akad. Nauk SSSR 156, 9–12 (1964)
Bonetto, F., Gallavotti, G., Gentile, G., Mastropietro, V.: Lindstedt series, ultraviolet divergences and Moser’s theorem. Ann. Sc. Norm. Super Pisa. Cl. Sci. IV Ser. 26(3), 545–593 (1998)
Celletti, A., Chierchia, L.: Construction of stable periodic orbits for the spin-orbit problem of celestial mechanics. Regul. Chaotic Dyn. 3(3), 107–121 (1998). (J. Moser at 70 (Russian))
Chierchia, L., Gallavotti, G.: Drift and diffusion in phase space. Ann. Inst. H. Poincaré Phys. Théor. 60(1), 144 (1994)
Chierchia, L., Pinzari, G.: Properly–degenerate KAM theory (following V.I. Arnold). Discrete Contin. Dyn. Syst. Ser. S 3(4), 545–578 (2010)
Chierchia, L., Pinzari, G.: The planetary \(N\)-body problem: symplectic foliation, reductions and invariant tori. Invent. Math. 186(1), 1–77 (2011)
Chierchia, L., Procesi, M.: Kolmogorov–Arnold–Moser (KAM) Theory for Finite and Infinite Dimensional Systems, pp. 1–45. Springer, Berlin (2019)
Clarke, A., Féjoz, J., Guardia, M..: Why are inner planets not inclided? arXiv: 2210.11311v1 (2022)
Delshams, A., Kaloshin, V., de la Rosa, A., Seara, T.M.: Global instability in the restricted planar elliptic three body problem. Commun. Math. Phys. 366(3), 1173–1228 (2019)
Deprit, A.: Elimination of the nodes in problems of \(n\) bodies. Celestial Mech. 30(2), 181–195 (1983)
Féjoz, J.: Démonstration du ‘théorème d’Arnold’ sur la stabilité du système planétaire (d’après Herman). Ergod. Theory Dyn. Syst. 24(5), 1521–1582 (2004)
Féjoz, J., Guardia, M., Kaloshin, V., Roldan, P.: Kirkwood gaps and diffusion along mean motion resonances in the restricted planar three body problem. J. Eur. Math. Soc. 18, 2315 (2014)
Gallavotti, G.: Quasi-integrable mechanical systems. In: Phénomènes Critiques. Systèmes aléatoires, Théories de Jauge, Part I, II (Les Houches, 1984), pp. 539–624. North-Holland, Amsterdam (1986)
Gallavotti, G.: Twistless KAM tori. Commun. Math. Phys. 164(1), 145–156 (1994)
Gentile, G., Gallavotti, G.: Majorant convergence for twistless KAM tori. Ergod. Theory Dyn. Syst. 15(5), 587–869 (1995)
Guzzo, M., Efthymiopoulos, C., Paez, R.I.: Semi-analytic computations of the speed of Arnold diffusion along single resonances in a priori stable Hamiltonian systems. J. Nonlinear Sci. 30(3), 851–901 (2020)
Jacobi, C.G.J.: Sur l’élimination des noeuds dans le problème des trois corps. Astron Nachr 1, 81–102 (1842)
Laskar, J., Robutel, P.: Stability of the planetary three-body problem. I .Expansion of the planetary Hamiltonian. Celestial Mech. Dyn. Astronom. 62(3), 193–217 (1995)
Pinzari, G.: Perihelia reduction and global Kolmogorov tori in the planetary problem. Mem. Amer. Math. Soc., vol. 255, No. 1218 (2018)
Pinzari, G.: On the co-existence of maximal and whiskered tori in the planetary three-body problem. J. Math. Phys. 59(5), 052701 (2018)
Pöschel, J.: Nekhoroshev estimates for quasi-convex Hamiltonian systems. Math. Z. 213(2), 187–216 (1993)
Radau, R.: Sur une transformation des équations différentielles de la dynamique. Ann. Sci. Ec. Norm. Sup. 5, 311–375 (1868)
Rüssmann, H.: Invariant tori in non-degenerate nearly integrable Hamiltonian systems. Regul. Chaotic Dyn. 6(2), 119–204 (2001)
Acknowledgements
G.P. is indebted to G. Gallavotti for many highlighting discussions. Theorem 1.1 is part of the MSc thesis of X.L. This research is funded by the PRIN project “New frontiers of Celestial Mechanics: theory and applications” and has been developed under the auspices of INdAM and GNFM. The authors confirm that the manuscript has no associated data.
Funding
Open access funding provided by Università degli Studi di Padova within the CRUI-CARE Agreement.
Author information
Authors and Affiliations
Contributions
GP conceptualized and developed the work and wrote the main manuscript. XL wrote a draft of Theorem 1.1, which is part of his MSc thesis at the University of Padova. All authors reviewed the manuscript.
Corresponding author
Ethics declarations
Conflict of interest
The authors declare no conflict of interest.
Additional information
Communicated by Alain Goriely.
Publisher's Note
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
About this article
Cite this article
Pinzari, G., Liu, X. Quantitative kam Theory, with an Application to the Three-Body Problem. J Nonlinear Sci 33, 90 (2023). https://doi.org/10.1007/s00332-023-09948-4
Received:
Accepted:
Published:
DOI: https://doi.org/10.1007/s00332-023-09948-4