Skip to main content
Log in

Complex Arnol’d – Liouville Maps

  • Published:
Regular and Chaotic Dynamics Aims and scope Submit manuscript

This article has been updated

Abstract

We discuss the holomorphic properties of the complex continuation of the classical Arnol’d – Liouville action-angle variables for real analytic 1 degree-of-freedom Hamiltonian systems depending on external parameters in suitable Generic Standard Form, with particular regard to the behaviour near separatrices. In particular, we show that near separatrices the actions, regarded as functions of the energy, have a special universal representation in terms of affine functions of the logarithm with coefficients analytic functions. Then, we study the analyticity radii of the action-angle variables in arbitrary neighborhoods of separatrices and describe their behaviour in terms of a (suitably rescaled) distance from separatrices. Finally, we investigate the convexity of the energy functions (defined as the inverse of the action functions) near separatrices, and prove that, in particular cases (in the outer regions outside the main separatrix, and in the case the potential is close to a cosine), the convexity is strictly defined, while in general it can be shown that inside separatrices there are inflection points.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Change history

  • 29 October 2023

    The numbering issue has been changed to 4-5.

Notes

  1. Compare [3] and references therein for general information.

  2. For references on Arnol’d diffusion, besides [1]; compare also [12, 24, 18, 23, 5, 13, 10, 16] among many other interesting results.

  3. By (2.4) and (2.8), \(\beta\leqslant|{\bar{\mathtt{G}}}(\theta_{i})-{\bar{\mathtt{G}}}(\theta_{j})|\leqslant 2\max_{\mathbb{T}}|{\bar{\mathtt{G}}}|\leqslant 2\varepsilon\).

  4. Explicitly: \(\partial_{\mathrm{p}_{1}}\mathrm{h}(\mathrm{p}^{0})=0\) and \(\partial^{2}_{\mathrm{p}_{1}^{2}}\mathrm{h}(\mathrm{p}^{0})\neq 0\).

  5. See, e. g., ([11], Appendix A).

  6. For brevity we write \(u\) and \(v\) instead of \(u(\hat{p})\) and \(v(\hat{p},q_{1})\), respectively.

  7. \(\kappa\) as in ( 2.10 ).

  8. For the external curves (\(i=0,2N\)) the orientation is to the right in \({\mathcal{M}}^{2N}(\hat{p})\), to the left in \({\mathcal{M}}^{0}(\hat{p})\).

  9. If \({\rm Re}(1+\nu)\geqslant 1/2\) (see (3.27)), then \(|(1+\nu)^{-1/2}-1|\leqslant|\nu|.\)

  10. This inclusion follows noting that, for every \(\theta,\) we have \(|(v_{0}+re^{{\rm i}\theta})^{2}-v_{0}^{2}|\geqslant r^{2}.\) The last inequality follows noting that it is equivalent to \(|re^{{\rm i}2\theta}+2v_{0}e^{{\rm i}\theta}|=|re^{{\rm i}\theta}+2v_{0}|\geqslant r\), which follows from \(v_{0}>r.\)

  11. Indeed, in one dimension, from a complex point of view, the Birkhoff normal form is the same both in the hyperbolic and in the elliptic case.

  12. The definition of \(\mathcal{P}\) is given in Lemma 2.

  13. Notice that there is no problem in \(\Theta_{{}_{\!\star}}(E)\), \(\Theta^{{}^{\!\star}}(E)\) where the square root vanishes. Actually, close to these points it is convenient to write \(\Omega(E)\) as a normal set with respect to \(q_{1}\) and not to \(p_{1}\).

  14. For real values of \(\hat{y}\) and \(E\) we are in the case \(E<E_{2j}(\hat{y})\), namely, \(z>0\).

  15. Omitting \(\hat{y}\).

  16. Recall (3.66).

  17. Uniquely fixing, e. g., \(\varphi^{(i)}_{1}(p,0)=0\).

  18. Recall (2.8)(2.10).

  19. Recall (3.3) and (3.5).

  20. Actually, a better estimate holds: it is smaller than some constant by \(\mu_{\rm o}{\mathtt{s}}\), where \(\mu_{\rm o}\) was defined in (3.13).

  21. We are considering the square root as a holomorphic function in the complex plane excluding the negative real axis.

  22. Close to a hyperbolic point the estimates for the action analyticity radius in (4.13) is \(\rho=\lambda|\log\lambda|\sqrt{\varepsilon}/C\) since \(\partial_{E}I\sim|\log\lambda|/\sqrt{\varepsilon}\) (see (4.27)), \(\lambda\varepsilon\) being the distance in energy from the critical energy of the hyperbolic point (see (4.30) below). Far away from the hyperbolic point the derivative is smaller (namely, \(\partial_{E}I\sim 1/\sqrt{\varepsilon}\)), but the distance in energy is bigger (being \(\sim\varepsilon\)).

  23. Observe that the action function \(E\to I_{1}^{i}(E,\hat{p})\) is strictly increasing and hence invertible.

  24. Compare [9].

  25. And the analogous formula for \(\partial^{2}_{I_{1}I_{1}}\tilde{\mathtt{E}}^{(i)}\big{(}\tilde{I}^{(i)}_{1}(E)\big{)}\).

  26. \(u\) is the solution of the fixed point equation \(u(y)=-b\big{(}y+u(y)\big{)}\) in the space of \(2\pi\)-periodic real-analytic function with holomorphic extension on the strip \(\{|{\rm Im}y|<1/5\}\) and \(|u|_{1/5}\leqslant 18\sqrt{\mathtt{g}}\).

  27. See, in particular, Lemma 0 and Appendix A.3 in [12].

  28. According to Lie’s series method.

  29. As well as (3.52) and (3.56) by Lemma 8.

  30. Since \(b(x_{m}-2\pi)=b(x_{m})=\pi-x_{m}\).

  31. Namely, taking a cut in the negative real line.

  32. Using that \(\frac{1}{2}|z|^{2}-(\cosh|z|-1-\frac{1}{2}|z|^{2})\leqslant|1-\cos z|\leqslant\cosh|z|-1\).

  33. Using that for \(|z|<1\) we have \(\frac{4}{5}|z|\leqslant|\sin z|\leqslant\frac{6}{5}|z|\).

  34. Using that for \(|z|\leqslant 1/2\) we have \(|\arcsin z|\leqslant\frac{2}{\sqrt{3}}|z|.\)

  35. Using that for \(|z|\leqslant 1\) we have \(|\cos z|\leqslant\sqrt{3}.\)

References

  1. Arnold, V. I., On the Nonstability of Dynamical Systems with Many Degrees of Freedom, Soviet Math. Dokl., 1964, vol. 5, no. 3, pp. 581–585; see also: Dokl. Akad. Nauk SSSR, 1964, vol. 156, no. 1, pp. 9-12.

    MATH  Google Scholar 

  2. Arnol’d, V. I., Mathematical Methods of Classical Mechanics, 2nd ed., Grad. Texts in Math., vol. 60, New York: Springer, 1997.

    Google Scholar 

  3. Arnol’d, V. I., Kozlov, V. V., and Neishtadt, A. I., Mathematical Aspects of Classical and Celestial Mechanics, 3rd ed., Encyclopaedia Math. Sci., vol. 3, Berlin: Springer, 2006.

    Book  Google Scholar 

  4. Bambusi, D., Fusè, A., and Sansottera, M., Exponential Stability in the Perturbed Central Force Problem, Regul. Chaotic Dyn., 2018, vol. 23, no. 7–8, pp. 821–841.

    Article  MathSciNet  MATH  Google Scholar 

  5. Bernard, P., Kaloshin, V., and Zhang, K., Arnold Diffusion in Arbitrary Degrees of Freedom and Normally Hyperbolic Invariant Cylinders, Acta Math., 2016, vol. 217, no. 1, pp. 1–79.

    Article  MathSciNet  MATH  Google Scholar 

  6. Biasco, L. and Chierchia, L., On the Topology of Nearly-Integrable Hamiltonians at Simple Resonances, Nonlinearity, 2020, vol. 33, no. 7, pp. 3526–3567.

    Article  MathSciNet  MATH  Google Scholar 

  7. Biasco, L. and Chierchia, L., Quasi-Periodic Motions in Generic Nearly-Integrable Mechanical Systems, Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur., 2022, vol. 33, no. 3, pp. 575–580.

    Article  MathSciNet  MATH  Google Scholar 

  8. Biasco, L. and Chierchia, L., Global Properties of Generic Real-Analytic Nearly-Integrable Hamiltonian Systems, in preparation (2023).

  9. Biasco, L. and Chierchia, L., Singular KAM Theory, in preparation (2023).

  10. Chen, Q. and de la Llave, R., Analytic Genericity of Diffusing Orbits in a priori Unstable Hamiltonian Systems, Nonlinearity, 2022, vol. 35, no. 4, pp. 1986–2019.

    Article  MathSciNet  MATH  Google Scholar 

  11. Chierchia, L., Kolmogorov – Arnold – Moser (KAM) Theory, in Mathematics of Complexity and Dynamical Systems: Vol. 2, R. A. Meyers (Ed.), New York: Springer, 2012, pp. 810–836.

    Chapter  Google Scholar 

  12. Chierchia, L. and Gallavotti, G., Drift and Diffusion in Phase Space, Ann. Inst. H. Poincaré Phys. Théor., 1994, vol. 60, no. 1, 144 pp. (Erratum: Drift and Diffusion in Phase Space, Ann. Inst. H. Poincaré Phys.Théor., 1998, vol. 68, no. 1, pp. 135.

    MathSciNet  MATH  Google Scholar 

  13. Delshams, A., de la Llave, R., and Seara, T. M., Instability of High Dimensional Hamiltonian Systems: Multiple Resonances Do Not Impede Diffusion, Adv. Math., 2016, vol. 294, pp. 689–755.

    Article  MathSciNet  MATH  Google Scholar 

  14. Giorgilli, A., Unstable Equilibria of Hamiltonian Systems, Discrete Contin. Dynam. Systems, 2001, vol. 7, no. 4, pp. 855–871.

    Article  MathSciNet  MATH  Google Scholar 

  15. Hofer, H. and Zehnder, E., Symplectic Invariants and Hamiltonian Dynamics, Birkhäuser Advanced Texts: Basler Lehrbücher, Basel: Birkhäuser, 1994.

    Book  MATH  Google Scholar 

  16. Kaloshin, V. and Zhang, K., Arnold Diffusion for Smooth Systems of Two and a Half Degrees of Freedom, Ann. Math. Stud., vol. 208, Princeton, N.J.: Princeton Univ. Press, 2020.

    Book  MATH  Google Scholar 

  17. de la Llave, R. and Wayne, C. E., Whiskered and Low Dimensional Tori in Nearly Integrable Hamiltonian Systems, Math. Phys. Electron. J., 2004, vol. 10, Paper 5, 45 pp.

    MathSciNet  MATH  Google Scholar 

  18. Mather, J. N., Arnold Diffusion by Variational Methods, in Essays in Mathematics and Its Applications, Heidelberg: Springer, 2012, pp. 271–285.

    Chapter  Google Scholar 

  19. Medvedev, A. G., Neishtadt, A. I., and Treschev, D. V., Lagrangian Tori near Resonances of Near-Integrable Hamiltonian Systems, Nonlinearity, 2015, vol. 28, no. 7, pp. 2105–2130.

    Article  MathSciNet  MATH  Google Scholar 

  20. Neishtadt, A. I., Problems of Perturbation Theory for Nonlinear Resonant Systems, Doctoral Dissertation, Moscow State University, Moscow, 1988, 342pp. (Russian).

  21. Neishtadt, A. I., On the Change in the Adiabatic Invariant on Crossing a Separatrix in Systems with Two Degrees of Freedom, J. Appl. Math. Mech., 1987, vol. 51, no. 5, pp. 586–592; see also: Prikl. Mat. Mekh., 1987, vol. 51, no. 5, pp. 750-757.

    Article  MathSciNet  MATH  Google Scholar 

  22. Okunev, A., On the Fourier Coefficients of the Perturbation Written Using the Angle Variable Near a Separatrix Loop, Loughborough University, https://doi.org/10.17028/rd.lboro.12278741.v1, (2020).

  23. Treschev, D., Arnold Diffusion Far from Strong Resonances in Multidimensional a priori Unstable Hamiltonian Systems, Nonlinearity, 2012, vol. 25, no. 9, pp. 2717–2757.

    Article  MathSciNet  MATH  Google Scholar 

  24. Zhang, K., Speed of Arnold Diffusion for Analytic Hamiltonian Systems, Invent. Math., 2011, vol. 186, no. 2, pp. 255–290.

    Article  MathSciNet  MATH  Google Scholar 

Download references

ACKNOWLEDGMENTS

The authors are grateful to A. Neishtadt for providing parts of his thesis [20, in Russian] related to the present paper.

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Luca Biasco or Luigi Chierchia.

Ethics declarations

The authors declare that they have no conflicts of interest.

Additional information

MSC2010

37J05, 37J35, 37J40, 70H05, 70H08, 70H15

APPENDIX A. PROOFS OF PROPOSITION 2

First, recalling (3.45) and (3.46), we define the symplectic transformation

$$\Phi_{*}:({\mathtt{p}},{\mathtt{q}})\in D_{7{\mathtt{r}}/8,7{\mathtt{s}}/8}\ \longrightarrow\ \big{(}{\mathtt{p}},{\mathtt{q}}_{1}+\theta_{2j}(\hat{\mathtt{p}}),\hat{\mathtt{q}}+{\mathtt{p}}_{1}\partial_{\hat{\mathtt{p}}}\theta_{2j}(\hat{\mathtt{p}})\big{)}\in D_{{\mathtt{r}},{\mathtt{s}}},$$
(A.1)
transforming the Hamiltonian \({\mathtt{H}}\) in (2.6) into
$${\mathtt{H}}_{*}:={\mathtt{H}}\circ\Phi_{*}({\mathtt{p}},{\mathtt{q}})=:\big{(}1+\nu_{*}({\mathtt{p}},{\mathtt{q}}_{1})\big{)}{\mathtt{p}}_{1}^{2}+{\mathtt{G}}_{*}(\hat{\mathtt{p}},{\mathtt{q}}_{1}).$$
(A.2)
By Taylor expansion at \(({\mathtt{p}},{\mathtt{q}}_{1})=(0,\hat{\mathtt{p}},0)\), recalling (A.2), (3.47), (3.49) and (2.10), we get
$$\displaystyle{\mathtt{H}}_{*}=E_{2j}(\hat{\mathtt{p}})+\big{(}1+\nu_{*}(0,\hat{\mathtt{p}},0)\big{)}{\mathtt{p}}_{1}^{2}-\lambda^{2}(\hat{\mathtt{p}}){\mathtt{q}}_{1}^{2}+R_{*}({\mathtt{p}},{\mathtt{q}}_{1}),\qquad\mbox{with}$$
$$\displaystyle R_{*}({\mathtt{p}},{\mathtt{q}}_{1}):=\big{(}\nu_{*}({\mathtt{p}},{\mathtt{q}}_{1})-\nu_{*}(0,\hat{\mathtt{p}},0)\big{)}{\mathtt{p}}_{1}^{2}+{\mathtt{G}}_{*}(\hat{\mathtt{p}},{\mathtt{q}}_{1})-\frac{1}{2}\partial^{2}_{{\mathtt{q}}_{1}}{\mathtt{G}}_{*}(\hat{\mathtt{p}},0){\mathtt{q}}_{1}^{2}.$$
(A.3)
Then, the following trivial lemma holds.

Lemma 7

There exists a constant \(0<{\bf c_{{{}_{0}}}}<1/8\) , depending only on \(\kappa,n\) , such that, defining the symplectic transformation

$$\displaystyle\Phi_{0}:\{|Y_{1}|<{\bf c_{{{}_{0}}}}\varepsilon^{1/4}\}\times\hat{D}_{3{\mathtt{r}}/4}\times\{|X_{1}|<{\bf c_{{{}_{0}}}}\varepsilon^{1/4}\}\times\mathbb{T}^{n-1}_{3{\mathtt{s}}/4}\ \longrightarrow\ D_{7{\mathtt{r}}/8,7{\mathtt{s}}/8},$$
(A.4)
$$\displaystyle{\mathtt{p}}_{1}=\delta(\hat{Y})Y_{1},\qquad\hat{\mathtt{p}}=\hat{Y},\qquad{\mathtt{q}}_{1}=\frac{1}{\delta(\hat{Y})}X_{1},\qquad\hat{\mathtt{q}}=\hat{X}-\frac{\partial_{\hat{Y}}\delta(\hat{Y})}{\delta(\hat{Y})}Y_{1}X_{1},$$
we have that \({\mathtt{H}}_{0}:={\mathtt{H}}_{*}\circ\Phi_{0}\) has the form
$${\mathtt{H}}_{0}=E_{2j}(\hat{Y})+g(\hat{Y})(Y_{1}^{2}-X_{1}^{2})+\varepsilon R_{0}(\varepsilon^{-1/4}Y_{1},\hat{Y},\varepsilon^{-1/4}X_{1}),$$
(A.5)
where \(R_{0}(\tilde{Y}_{1},\hat{Y},\tilde{X}_{1})\) is holomorphic on
$$\{|\tilde{Y}_{1}|<{\bf c_{{{}_{0}}}}\}\times\hat{D}_{3{\mathtt{r}}/4}\times\{|\tilde{X}_{1}|<{\bf c_{{{}_{0}}}}\},$$
with \(|R_{0}|\lessdot 1\) and, finally, it is at least cubic in \(\tilde{Y}_{1},\tilde{X}_{1}\) .

Proof

The fact that \(\Phi_{0}\) is well defined on its domain follows by the explicit expression in (A.4), by (3.50), (2.10) and (2.8) (in particular, \(\varepsilon\leqslant{\mathtt{r}}^{2}/2^{16}\)). Eq. (A.5) follows by (A.3), setting

$$R_{0}(\tilde{Y}_{1},\hat{Y},\tilde{X}_{1}):=\varepsilon^{-1}R_{*}\left(\delta(\hat{Y})\varepsilon^{1/4}\tilde{Y}_{1},\hat{Y},\frac{\varepsilon^{1/4}}{\delta(\hat{Y})}\tilde{X}_{1}\right).$$
(A.6)
Finally, the estimate \(|R_{0}|\lessdot 1\) follows from (3.47) and (3.50).

Next, we shall use the following well-known result, whose proof can be found, e.g., inFootnote

See, in particular, Lemma 0 and Appendix A.3 in [12].

[12] or in [14].

See, in particular, Lemma 0 and Appendix A.3 in [12].

Lemma 8

Given a Hamiltonian \({\mathtt{H}}_{0}\) as in (A.5) . For suitable constants \(0<{\bf c}_{{{}_{1}}}<{\bf c_{{{}_{0}}}}/8n{\bf c}_{{{}_{2}}}\) , depending only on \(\kappa,n\) , there exist a \((\) near-identity \()\) symplectic transformation

$$\displaystyle\begin{aligned} \Phi_{1}:&\{|y_{1}|<{\bf c}_{{{}_{1}}}\varepsilon^{1/4}\}\times\hat{D}_{{\mathtt{r}}/2}\times\{|x_{1}|<{\bf c}_{{{}_{1}}}\varepsilon^{1/4}\}\times\mathbb{T}^{n-1}_{{\mathtt{s}}/2}\ \longrightarrow\\ &\{|Y_{1}|<{\bf c_{{{}_{0}}}}\varepsilon^{1/4}\}\times\hat{D}_{3{\mathtt{r}}/4}\times\{|X_{1}|<{\bf c_{{{}_{0}}}}\varepsilon^{1/4}\}\times\mathbb{T}^{n-1}_{3{\mathtt{s}}/4}, \end{aligned}$$
(A.7)
and a function \(R_{\rm hp}(z,\hat{y})\) satisfying (3.52) and (3.56) , such that \({\mathtt{H}}_{\rm hp}(y,x_{1}):={\mathtt{H}}_{0}\circ\Phi_{1}(y,x)\) satisfies (3.53) . Moreover, \(\Phi_{1}\) has the form
$$\displaystyle Y_{1}=y_{1}+\varepsilon^{1/4}a_{1}(\varepsilon^{-1/4}y_{1},\hat{y},\varepsilon^{-1/4}x_{1}),\quad\hat{Y}=\hat{y},$$
(A.8)
$$\displaystyle X_{1}=x_{1}+\varepsilon^{1/4}a_{2}(\varepsilon^{-1/4}y_{1},\hat{y},\varepsilon^{-1/4}x_{1}),\quad\hat{X}=\hat{x}+\sqrt{\varepsilon}{\mathtt{r}}^{-1}a_{3}(\varepsilon^{-1/4}y_{1},\hat{y},\varepsilon^{-1/4}x_{1}),$$
for suitable functions \(a_{i}(\tilde{y}_{1},\hat{y},\tilde{x}_{1})\) , \(i=1,2,3\) , which are holomorphic and bounded by \({\bf c}_{{{}_{2}}}\) on
$$\{|\tilde{y}_{1}|<{\bf c_{{{}_{0}}}}/2\}\times\hat{D}_{{\mathtt{r}}/2}\times\{|\tilde{x}_{1}|<{\bf c_{{{}_{0}}}}/2\},$$
moreover, \(a_{1},a_{2}\) and \(a_{3}\) are, respectively, at least quadratic and cubic in \(\tilde{y}_{1},\tilde{x}_{1}\) .

Remark 6

\({\mathtt{H}}_{\rm hp}\) is simply the hyperbolic Birkhoff normal form of \({\mathtt{H}}_{0}\). Any canonical transformation of the form \(y_{1}=\alpha\tilde{y}_{1}+\beta\tilde{x}_{1},\) \(x_{1}=\beta\tilde{y}_{1}+\alpha\tilde{x}_{1},\) with \(\alpha^{2}-\beta^{2}=1\) and \(\hat{y}=\hat{\tilde{y}}\) leaves \({\mathtt{H}}_{\rm hp}\) invariant since \(y_{1}^{2}-x_{1}^{2}=\tilde{y}_{1}^{2}-\tilde{x}_{1}^{2}.\) Namely, the integrating transformation \(\Phi_{1}\) is not unique. However, as is well known, the form of the integrated Hamiltonian \({\mathtt{H}}_{\rm hp}\) in (3.53) is unique, in the sense that \(E_{2j}\), \(g\) and \(R\) are unique.

Note also that the map \(\Phi_{1}\) is close to the identity, for small \({\bf c}_{{{}_{1}}}\), since its Jacobian is the identity plus a matrix whose entries are (by the Cauchy estimates) uniformly bounded on its domain in (3.51) by \(2{\bf c}_{{{}_{2}}}{\bf c}_{{{}_{1}}}/{\bf c_{{{}_{0}}}}\leqslant 1/4n\).

Let us come back to the proof of Proposition 2 and let us prove (3.56).

Evaluating (3.53) for \(\mu=0\), we get

$$\displaystyle{\bar{\mathtt{H}}}_{1}(y,x_{1}):={\bar{\mathtt{H}}}_{1}(y,x_{1})|_{\mu=0}={\bar{\mathtt{H}}}_{0}\circ\bar{\Phi}_{1}(y,x)$$
$$\displaystyle=\bar{E}_{2j}+\bar{g}(y_{1}^{2}-x_{1}^{2})+\varepsilon\bar{R}_{\rm hp}\left(\frac{y_{1}^{2}-x_{1}^{2}}{\sqrt{\varepsilon}}\right)=O(\varepsilon)$$
(A.9)
on the domain defined in (3.51). Let us denote \({\bar{\mathtt{H}}}_{0}:={\mathtt{H}}_{0}|_{\mu=0}\). Since by (A.6), (3.47) one has \({\mathtt{H}}_{0}-{\bar{\mathtt{H}}}_{0}=O(\varepsilon\mu)\),
$${\mathtt{H}}_{0}\circ\bar{\Phi}_{1}={\bar{\mathtt{H}}}_{0}\circ\bar{\Phi}_{1}+({\mathtt{H}}_{0}-{\bar{\mathtt{H}}}_{0})\circ\bar{\Phi}_{1}={\bar{\mathtt{H}}}_{1}+R_{1},\quad\mbox{with}\ \ \ R_{1}=O(\varepsilon\mu),$$
namely, the system is integrated up to a small term of order \(\varepsilon\mu\). Note also that, since \(\bar{\Phi}_{1}\) has the form in (A.8), it leaves invariant the terms of order \(\leqslant 2\) in \((y_{1},x_{1})\), namely,
$${\mathtt{H}}_{0}\circ\bar{\Phi}_{1}=E_{2j}+g(y_{1}^{2}-x_{1}^{2})+\varepsilon\bar{R}+Q,\quad\mbox{with}\quad Q=O(\varepsilon\mu).$$
(A.10)
Now we want to construct a symplectic transformation \(\Phi_{\mu}\) integrating \({\mathtt{H}}_{0}\circ\bar{\Phi}_{1}\). Since \({\bar{\mathtt{H}}}_{1}\) is already in normal form, we claim that the integrating transformation \(\Phi_{\mu}\) is \(O(\varepsilon^{1/4}\mu)\)-close to the identity and
$${\mathtt{H}}_{0}\circ\bar{\Phi}_{1}\circ\Phi_{\mu}=({\bar{\mathtt{H}}}_{1}+R_{1})\circ\Phi_{\mu}=:{\mathtt{H}}_{\rm hp}^{\prime}={\bar{\mathtt{H}}}_{1}+O(\varepsilon\mu),$$
(A.11)
where \({\mathtt{H}}_{\rm hp}^{\prime}\) is in normal form, namely, as in (3.53). By the unicity of the Birkhoff normal form, we deduce that \({\mathtt{H}}_{\rm hp}={\mathtt{H}}_{\rm hp}^{\prime}={\bar{\mathtt{H}}}_{1}+O(\varepsilon\mu)\). By (3.53), (A.9), (3.50) and (3.2), we get (3.56).

It remains to prove (A.11). The crucial point here is that the generating functionFootnote

According to Lie’s series method.

\(\chi\) of the integrating transformation \(\Phi_{\mu}\) is \(O(\sqrt{\varepsilon}\mu)\) and its gradient is, by the Cauchy estimates, \(O(\varepsilon^{1/4}\mu)\) in a domain \(\{|y_{1}|,|x_{1}|\lessdot\varepsilon^{1/4}\}\). The fact that \(\chi=O(\sqrt{\varepsilon}\mu)\) can be easily seen by passing, as is usual in Birkhoff’s normal form, to the coordinate \(\xi=(y_{1}-x_{1})/\sqrt{2}\), \(\eta=(y_{1}+x_{1})/\sqrt{2}\). In these coordinates, recalling (A.10), we get
$${\mathtt{H}}_{0}\circ\bar{\Phi}_{1}=E_{2j}+2g\xi\eta+\varepsilon\bar{R}^{\prime}(\xi,\eta)+Q^{\prime}(\xi,\eta)$$
with \(\bar{R}^{\prime}=\bar{R}_{\rm hp}(2\xi\eta/\sqrt{\varepsilon})=O(1)\) and \(Q^{\prime}=O(\varepsilon\mu)\). Note that the Taylor expansion of \(\bar{R}^{\prime}\) contains only a monomial of the form \(\bar{R}^{\prime}_{hh}\xi^{h}\eta^{h}\). At the first step, we have to cancel all the monomials of \(Q^{\prime}\) of the form \(Q^{\prime}_{hk}\xi^{h}\eta^{k}\) with \(h+k=3\). The generating function\(\chi^{(3)}\) of the first step is exactly
$$\chi^{(3)}=\sum_{h+k=3}\frac{Q^{\prime}_{hk}}{2g(h-k)}\xi^{h}\eta^{k}\stackrel{{\scriptstyle{(3.50)}}}{{=}}O(\sqrt{\varepsilon}\mu).$$
After this first step the Hamiltonian becomes \(E_{2j}+2g\xi\eta+\varepsilon\bar{R}^{\prime}(\xi,\eta)+Q^{\prime\prime}(\xi,\eta)\) with \(Q^{\prime\prime}=O(\varepsilon\mu)\). At the second step, we have to cancel all the monomials of \(Q^{\prime\prime}\) of the form \(Q^{\prime\prime}_{hk}\xi^{h}\eta^{k}\) with \(h+k=4\), \(h\neq k\). We proceed as in the first step with analogous estimates. Analogously for the other infinite steps, obtaining
$$E_{2j}+2g\xi\eta+\varepsilon\bar{R}^{\prime}(\xi,\eta)+\bar{Q}(\xi,\eta)$$
with \(\bar{Q}=O(\varepsilon\mu)\) and \(\bar{Q}_{hk}=0\) for \(h\neq k\), proving (A.11) (recall (3.50) and (3.2)).

According to Lie’s series method.

We can conclude the proof of Proposition 2:

The composition of the symplectic transformations defined in (A.1), (A.4), (3.51) integrates \({\mathtt{H}}\), namely, (3.53) holdsFootnote

As well as (3.52) and (3.56) by Lemma 8.

with \(\Phi_{\rm hp}:=\Phi_{*}\circ\Phi_{0}\circ\Phi_{1}\) satisfying (3.51), (3.54) and (3.55). The inclusion (3.57) follows by (3.54) and (3.50).

As well as (3.52) and (3.56) by Lemma 8.

APPENDIX B. PROOFS OF TWO SIMPLE LEMMATA

B.1. Proof of Lemma 1

We know that \(\partial_{\theta}{\bar{\mathtt{G}}}(\bar{\theta}_{i})=0\) and we want to solve the equation \(\partial_{\theta}{\mathtt{G}}\big{(}\hat{p},\theta_{i}(\hat{p})\big{)}=0\). Equivalently, for \(\mu\leqslant 2^{-8}\kappa^{-6}\), we want to find a real-analytic \(y=y(\hat{p}),\) \(\hat{p}\in\hat{D}_{\mathtt{r}}\), with

$$\textstyle\sup_{\hat{D}_{\mathtt{r}}}|y|\leqslant\rho:=\frac{2\varepsilon\mu}{\beta{\mathtt{s}}}\stackrel{{\scriptstyle{\rm(2.10)}}}{{\leqslant}}\frac{{\mathtt{s}}}{2},$$
(B.1)
by solving the equation
$$\partial_{\theta}{\mathtt{G}}\big{(}\hat{p},\bar{\theta}_{i}+y(\hat{p})\big{)}=0,$$
(B.2)
so that \(\theta_{i}(\hat{p})=\bar{\theta}_{i}+y(\hat{p})\). We have
$$\partial_{\theta}{\mathtt{G}}(\hat{p},\bar{\theta}_{i}+y)=\partial_{\theta}{\mathtt{G}}(\hat{p},\bar{\theta}_{i})+g(\hat{p},y)y,\quad\mbox{where}\quad g(\hat{p},y):=\int_{0}^{1}\partial_{\theta}^{2}{\mathtt{G}}(\hat{p},\bar{\theta}_{i}+ty)dt.$$
Then (B.2) can be written as the fixed point equation
$$y=\Psi(y),\quad\mbox{where}\quad\Psi(y):=-\frac{\partial_{\theta}{\mathtt{G}}(\hat{p},\bar{\theta}_{i})}{g(\hat{p},y)}$$
to be solved in the closed set of the real-analytic functions \(y=y(\hat{p})\) on \(\hat{D}_{\mathtt{r}}\) satisfying the bound (B.1). Note that, since \(\partial_{\theta}{\bar{\mathtt{G}}}(\bar{\theta}_{i})=0\), by (2.4) we have \(|\partial_{\theta}^{2}{\bar{\mathtt{G}}}(\bar{\theta}_{i})|\geqslant\beta\). Moreover, by (2.8) and the Cauchy estimates, we get for \(|y|\leqslant\rho\) and \(\hat{p}\in\hat{D}_{\mathtt{r}}\)
$$|g-\partial_{\theta}^{2}{\bar{\mathtt{G}}}(\bar{\theta}_{i})|\leqslant\frac{4\varepsilon\mu}{{\mathtt{s}}^{2}},\quad\mbox{which implies}\quad|g|\geqslant\beta-\frac{4\varepsilon\mu}{{\mathtt{s}}^{2}}\stackrel{{\scriptstyle{\rm(2.10)}}}{{\geqslant}}\frac{\beta}{2}.$$
(B.3)
Again by \(\partial_{\theta}{\bar{\mathtt{G}}}(\bar{\theta}_{i})=0\), (2.8) and the Cauchy estimates, we find uniformly on \(\hat{D}_{\mathtt{r}}\) that
$$|\partial_{\theta}{\mathtt{G}}(\hat{p},\bar{\theta}_{i})|\leqslant\varepsilon\mu/{\mathtt{s}}.$$
(B.4)
Then by (B.3) we obtain for \(|y|\leqslant\rho\) and \(\hat{p}\in\hat{D}_{\mathtt{r}}\)
$$\textstyle|\Psi|\leqslant\frac{2\varepsilon\mu}{\beta{\mathtt{s}}}=\rho,$$
(B.5)
by (2.10) and (B.1). Moreover,
$$\partial_{y}\Psi(y):=\frac{\partial_{\theta}{\mathtt{G}}(\hat{p},\bar{\theta}_{i})}{\big{(}g(\hat{p},y)\big{)}^{2}}\partial_{y}g(\hat{p},y).$$
Then for \(|y|\leqslant\rho\) and \(\hat{p}\in\hat{D}_{\mathtt{r}}\) we get
$$\textstyle|\partial_{y}\Psi|<2^{6}\frac{\varepsilon^{2}\mu}{\beta^{2}{\mathtt{s}}^{4}}\leqslant 2^{6}\kappa^{6}\mu\leqslant 1,$$
(B.6)
by (B.4), (B.3), (2.10) and since \(|\partial_{y}g(\hat{p},y)|<{16\varepsilon}/{{\mathtt{s}}^{3}}\) by (2.8) and the Cauchy estimates. In conclusion, by (B.5) and (B.6) we have that \(\Psi\) is a contraction and the fixed point theorem applies proving the first estimate in (3.2).

Let us now show the second estimate in (3.2).

By (2.8), the first estimate in (3.2), (2.10) and the Cauchy estimates, we get

$$\displaystyle|E_{i}(\hat{p})-\bar{E}_{i}|\leqslant|{\mathtt{G}}\big{(}\hat{p},\theta_{j}(\hat{p})\big{)}-{\bar{\mathtt{G}}}\big{(}\theta_{j}(\hat{p})\big{)}|+|{\bar{\mathtt{G}}}\big{(}\theta_{j}(\hat{p})\big{)}-{\bar{\mathtt{G}}}(\bar{\theta}_{j})|$$
$$\displaystyle\leqslant\textstyle\varepsilon\mu+\frac{2\varepsilon^{2}\mu}{\beta{\mathtt{s}}^{2}}\leqslant 3\kappa^{3}\varepsilon\mu,$$
proving the second estimate in (3.2).

Let us prove the final claim. By (2.11) (applied to \({\bar{\mathtt{G}}}\)) and by the Cauchy estimates, it follows that the minimal distance between two critical points of \({\bar{\mathtt{G}}}\) can be estimated from below by \(2\beta{\mathtt{s}}^{2}/\varepsilon\). Thus, by the first estimates in (3.2), it follows that the relative order of the critical points of \({\bar{\mathtt{G}}}\) is preserved, provided \(8\varepsilon^{3}\mu^{2}<\beta^{3}{\mathtt{s}}^{4}\), which, using (2.10), is implied by \(2^{3}\kappa^{7}\mu^{2}<1\), which, in turn, is implied by the hypothesis \(\mu\leqslant 1/(2\kappa)^{6}\).

As for critical energies, since \({\bar{\mathtt{G}}}\) is \(\beta\)-Morse, they are at least \(\beta\) apart; hence, from the second estimate in (3.2) the claim follows provided \(3\kappa^{3}\varepsilon\mu<\beta\), which, by (2.10), is implied by \(\mu<1/(3\kappa^{4})\), which, again, is implied by the hypothesis.

B.2. Proof of Lemma 6

First denote \(R(z):={w}(z)-\cos z,\) so that \(|R|_{1}\leqslant{{\mathtt{g}}_{\rm o}}.\) We note that, on the real line, \({w}\) has exactly two critical points: a maximum \(x_{M}\) (with \({w}(x_{M})=1\)) and a minimum \(x_{m}\) (with \({w}(x_{m})=-1\)) in the interval \([-\pi/2,3\pi/2).\) Indeed, since by the Cauchy estimates \(\sup_{R}|{w}^{\prime}|\leqslant{{\mathtt{g}}_{\rm o}},\) the equation \({w}^{\prime}(x)=-\sin x+R^{\prime}(x)=0\) in the interval \([-\pi/2,3\pi/2)\) has only two solutions \(x_{M},x_{m}\) with \(|x_{M}|,|x_{m}-\pi|\leqslant 1.0001{{\mathtt{g}}_{\rm o}}\leqslant 0.001.\) Obviously, \(x_{M}+b(x_{M})=0\) and \(x_{m}+b(x_{m})=\pi.\)

On the real line the function \(b\) is given by the \(2\pi\)-periodic continuousFootnote

Since \(b(x_{m}-2\pi)=b(x_{m})=\pi-x_{m}\).

function defined in the interval \([x_{m}-2\pi,x_{m}]\) by the expression
$$b(x):={\rm sign}(x-x_{M})\arccos\big{(}{w}(x)\big{)}-x.$$
Let us consider first the complex domain \(\Omega_{0}:=\{0.4<{\rm Re}z<\pi-0.4,\ |{\rm Im}z|<1/4\}\) where \(b(z)\) is clearly extendible to a holomorphic function. Here we have \(\sup_{\Omega_{0}}|\cos z|\leqslant 0.913\) and, therefore, \(\sup_{\Omega_{0}}|\cos z|+|R(z)|\leqslant 0.914.\) Then for \(z\in\Omega_{0}\) we get
$$|b(z)|=|\arccos\big{(}\cos z+R(z)\big{)}-z|\leqslant\int_{0}^{1}\left|\frac{R(z)}{\sqrt{1-\big{(}\cos z+tR(z)\big{)}^{2}}}\right|dt\leqslant 6.1{{\mathtt{g}}_{\rm o}}.$$

Since \(b(x_{m}-2\pi)=b(x_{m})=\pi-x_{m}\).

We now prove that \(b(z)\) is extendible to a holomorphic function for \(|z|<1/2\). First we prove that there exists a real-analytic positive function \(d\) with holomorphic extension on \(|z|<1/2\) such that \({w}(z)=1-\frac{1}{2}\big{(}(z-x_{M})d(z)\big{)}^{2}\). By Taylor’s expansion at \(z=x_{M}\) we have that \(d^{2}(z)=-2\int_{0}^{1}(1-t){w}^{\prime\prime}\big{(}x_{M}+t(z-x_{M})\big{)}dt\) and, therefore, for \(|z|<1/2\)

$$|d^{2}(z)-1|\leqslant 1-\cos x_{M}+\sup_{|z|<1/2}|\sin z||z-x_{M}|+2{{\mathtt{g}}_{\rm o}}\leqslant 0.55.$$
Then we can take the principle square rootFootnote

Namely, taking a cut in the negative real line.

of \(d^{2}(z)\), obtaining the function \(d(z)\). Now consider the holomorphic function \(a(z)\) defined for \(|z|<2\) such that \(a^{\prime}(z)=1/\sqrt{1-(z/2)^{2}}\) and \(a(0)=0\). Then for real \(x\) we get \(a(x)={\rm sign}(x)\arccos(1-x^{2}/2)\) and also (with \(d(x)>0\))
$$b(x):={\rm sign}(x-x_{M})\arccos\left(1-\frac{1}{2}\big{(}(x-x_{M})d(x)\big{)}^{2}\right)-x=a\big{(}(x-x_{M})d(x)\big{)}-x.$$
Then \(a\big{(}(z-x_{M})d(z)\big{)}-z\) is a holomorphic extension of \(b\) for \(|z|<2\). An analogous argument holds for \(|z-\pi|<2.\)

Namely, taking a cut in the negative real line.

In the following we will estimate \(b(z)\) for a strip \(|z|<1/2\), analogous arguments holds for \(|z-\pi|<1/2\). We will often use thatFootnote

Using that \(\frac{1}{2}|z|^{2}-(\cosh|z|-1-\frac{1}{2}|z|^{2})\leqslant|1-\cos z|\leqslant\cosh|z|-1\).

$$|z|\leqslant 1\quad\Longrightarrow\quad 0.45|z|^{2}\leqslant|1-\cos z|\leqslant 0.55|z|^{2}.$$
(B.7)
Now we prove that there exists a unique function \(b(z)\) defined for
$$\Omega_{1}:=\{3\sqrt{{\mathtt{g}}_{\rm o}}<|z|<1/2\}$$
satisfying \(\sup_{\Omega_{1}}|b|\leqslant\frac{3}{2}\sqrt{{\mathtt{g}}_{\rm o}},\) such that \({w}(z)=\cos\big{(}z+b(z)\big{)}\), as a fixed point of the equation
$$b(z)=\Psi(b)(z):=2\arcsin\left(\frac{-R(z)}{2\sin(z+b(z)/2)}\right).$$
Indeed,
$$\textstyle\cos\big{(}z+b(z)\big{)}-\cos z=-2\sin\left(z+b(z)/2\right)\sin\left(b(z)/2\right)=R(z).$$
For \(z\in\) we have \(|z+b(z)/2|\geqslant\frac{3}{2}\sqrt{{\mathtt{g}}_{\rm o}}\), which impliesFootnote

Using that for \(|z|<1\) we have \(\frac{4}{5}|z|\leqslant|\sin z|\leqslant\frac{6}{5}|z|\).

\(|\sin(z+b(z)/2)|\geqslant\frac{6}{5}\sqrt{{\mathtt{g}}_{\rm o}}\) andFootnote

Using that for \(|z|\leqslant 1/2\) we have \(|\arcsin z|\leqslant\frac{2}{\sqrt{3}}|z|.\)

\(\sup_{\Omega_{1}}|\Psi(b)(z)|<\sqrt{{\mathtt{g}}_{\rm o}}.\) Finally, \(\Psi\) is a contraction sinceFootnote

Using that for \(|z|\leqslant 1\) we have \(|\cos z|\leqslant\sqrt{3}.\)

$$\displaystyle\textstyle\sup_{\Omega_{1}}|\Psi(b)-\Psi(b^{\prime})|\leqslant\frac{2}{\sqrt{3}}{{\mathtt{g}}_{\rm o}}\sup_{\Omega_{1}}\left|\frac{1}{\sin(z+b(z)/2)}-\frac{1}{\sin(z+b^{\prime}(z)/2)}\right|$$
$$\displaystyle\textstyle\leqslant\frac{2}{\sqrt{3}}\frac{5^{2}}{6^{2}}2\left|\sin\left(\frac{b^{\prime}(z)-b(z)}{4}\right)\cos\left(z+\frac{b^{\prime}(z)+b(z)}{4}\right)\right|\leqslant\frac{5}{6}\sup_{\Omega_{1}}|b-b^{\prime}|.$$
In conclusion, we get \(\sup_{\Omega_{1}}|b|\leqslant\frac{3}{2}\sqrt{{\mathtt{g}}_{\rm o}}.\)

Using that \(\frac{1}{2}|z|^{2}-(\cosh|z|-1-\frac{1}{2}|z|^{2})\leqslant|1-\cos z|\leqslant\cosh|z|-1\).

Using that for \(|z|<1\) we have \(\frac{4}{5}|z|\leqslant|\sin z|\leqslant\frac{6}{5}|z|\).

Using that for \(|z|\leqslant 1/2\) we have \(|\arcsin z|\leqslant\frac{2}{\sqrt{3}}|z|.\)

Using that for \(|z|\leqslant 1\) we have \(|\cos z|\leqslant\sqrt{3}.\)

Next, we claim that in the domain \(\Omega_{2}:=\{|z|\leqslant 3\sqrt{{\mathtt{g}}_{\rm o}}\}\) we have that \(|b(z)|<9\sqrt{{\mathtt{g}}_{\rm o}}\). Indeed, by contradiction, assume that there exists \(z_{0}\in\Omega_{2}\) such that for every \(|z|<|z_{0}|\) we have \(|b(z)|<9\sqrt{{\mathtt{g}}_{\rm o}}\), but \(|b(z_{0})|=9\sqrt{{\mathtt{g}}_{\rm o}}\). Then \(|z_{0}+b(z_{0})|\leqslant 12\sqrt{{\mathtt{g}}_{\rm o}}\) and by (B.7) and since \(\cos\big{(}z_{0}+b(z_{0})\big{)}-1=\cos z_{0}-1+R(z_{0})\) we get

$$\displaystyle 16{{\mathtt{g}}_{\rm o}}\leqslant 0.45(|b(z_{0})|-|z_{0}|)^{2}\leqslant 0.45|z_{0}+b(z_{0})|^{2}\leqslant|\cos\big{(}z_{0}+b(z_{0})\big{)}-1|$$
$$\displaystyle\leqslant|\cos z_{0}-1|+|R(z_{0})|\leqslant 0.55|z_{0}|^{2}+{{\mathtt{g}}_{\rm o}}\leqslant 6{{\mathtt{g}}_{\rm o}},$$
which is a contradiction. Thus, \(\sup_{\Omega_{2}}|b(z)|\leqslant 9\sqrt{{\mathtt{g}}_{\rm o}}.\)     \(\square\)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Biasco, L., Chierchia, L. Complex Arnol’d – Liouville Maps. Regul. Chaot. Dyn. 28, 395–424 (2023). https://doi.org/10.1134/S1560354723520064

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1560354723520064

Keywords

Navigation