Skip to main content
Log in

Steady Euler Flows on \({\mathbb {R}}^3\) with Wild and Universal Dynamics

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

Understanding complexity in fluid mechanics is a major problem that has attracted the attention of physicists and mathematicians during the last decades. Using the concept of renormalization in dynamics, we show the existence of a locally dense set \({\mathscr {G}}\) of stationary solutions to the Euler equations in \({\mathbb {R}}^3\) such that each vector field \(X\in {\mathscr {G}}\) is universal in the sense that any area preserving diffeomorphism of the disk can be approximated (with arbitrary precision) by the Poincaré map of X at some transverse section. We remark that this universality is approximate but occurs at all scales. In particular, our results establish that a steady Euler flow may exhibit any conservative finite codimensional dynamical phenomenon; this includes the existence of horseshoes accumulated by elliptic islands, increasing union of horseshoes of Hausdorff dimension 3 or homoclinic tangencies of arbitrarily high multiplicity. The steady solutions we construct are Beltrami fields with sharp decay at infinity. To prove these results we introduce new perturbation methods in the context of Beltrami fields that allow us to import deep techniques from bifurcation theory: the Gonchenko-Shilnikov-Turaev universality theory and the Newhouse and Duarte theorems on the geometry of wild hyperbolic sets. These perturbation methods rely on two tools from linear PDEs: global approximation and Cauchy–Kovalevskaya theorems. These results imply a strong version of V.I. Arnold’s vision on the complexity of Beltrami fields in Euclidean space.

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.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. See Sect. 4.1 for the definition of the involved topology.

  2. Theorem B is stronger than Arnold’s vision: there is no example of universal dynamics in celestial mechanics.

  3. A topologically generic subset is a set equal to a countable intersection of open and dense subsets.

  4. In particular, the closure \({\bar{\Gamma }}\) is diffeomorphic to \({\mathbb {R}}/{\mathbb {Z}}\times [-1,1]\).

  5. A hyperbolic horseshoe is a locally maximal, transitive hyperbolic Cantor set.

  6. More precisely, the map \(T_n\) associates to a point in \(R_n^0\), resp. in \(R_n^1\), its first iterate, resp. its 2n-th iterate.

  7. This is the case of Example 3.5, where U is the intersection of \({\mathbb {R}}^3\setminus \{(0,0)\}\times {\mathbb {R}}\) with a large ball.

  8. Equivalently, we can express the displacement function as the vector product \(\partial _t\ell (t)\wedge (\phi ^u_\epsilon (\ell (t))-\phi ^s_\epsilon (\ell (t)))\).

References

  1. Arnold, V. I., Khesin, B. A.: Topological methods in hydrodynamics, volume 125 of Applied Mathematical Sciences. Springer, Cham, (2021) \(\copyright \) 2021. Second edition [of 1612569]

  2. Arnold, V.: Small denominators and problems of stability of motion in classical and celestial mechanics. Russ. Math. Surv. 18, 85 (1963)

    MathSciNet  Google Scholar 

  3. Arnold, V.I.: Sur la topologie des écoulements stationnaires des fluides parfaits. C. R. Acad. Sci. Paris 261, 17–20 (1965)

    MathSciNet  MATH  Google Scholar 

  4. Arnold, V.: Sur la géométrie différentielle des groupes de Lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits. Ann. Inst. Fourier (Grenoble) 16(fasc.1), 319–361 (1966)

    MathSciNet  MATH  Google Scholar 

  5. Berger, P., Bochi, J.: On emergence and complexity of ergodic decompositions. Adv. Math. 390, 107904 (2021)

    MathSciNet  MATH  Google Scholar 

  6. Bonatti, C., Diaz, L.: On maximal transitive sets of generic diffeomorphisms. Publ. Math. Inst. Hautes Études Sci. 96, 171–197 (2003)

    MathSciNet  MATH  Google Scholar 

  7. Berger, P.: Generic family with robustly infinitely many sinks. Invent. Math. 205(1), 121–172 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  8. Berger, P.: Emergence and non-typicality of the finiteness of the attractors in many topologies. Proc. Steklov Inst. Math. 297(1), 1–27 (2017)

    MathSciNet  MATH  Google Scholar 

  9. Berger, P.: Coexistence of chaotic and elliptic behaviors among analytic, symplectic diffeomorphisms of any surface. (2021). arXiv:2105.08354

  10. Buzzard, G.T., Hruska, S.L., Ilyashenko, Y.: Kupka-Smale theorem for polynomial automorphisms of \({\mathbb{C} }^2\) and persistence of heteroclinic intersections. Invent. Math. 161(1), 45–89 (2005)

    ADS  MathSciNet  MATH  Google Scholar 

  11. Browder, F.E.: Approximation by solutions of partial differential equations. Amer. J. Math. 84, 134–160 (1962)

    MathSciNet  MATH  Google Scholar 

  12. Broer, H.W., Tangerman, F.M.: From a differentiable to a real analytic perturbation theory, applications to the Kupka Smale theorems. Ergod. Theory Dyn. Syst. 6(3), 345–362 (1986)

    MathSciNet  MATH  Google Scholar 

  13. Berger, P., Turaev, D.: On Herman’s positive entropy conjecture. Adv. Math. 349, 1234–1288 (2019)

    MathSciNet  MATH  Google Scholar 

  14. Buzzard, G.T.: Infinitely many periodic attractors for holomorphic maps of \(2\) variables. Ann. Math. 145(2), 389–417 (1997)

    MathSciNet  MATH  Google Scholar 

  15. Cardona, R.: Steady Euler flows and Beltrami fields in high dimensions. Ergodic Theor. & Dyn. Syst., pp. 1–24, (2020)

  16. Cardona, R., Miranda, E., Peralta-Salas, D.: Computability and Beltrami fields in Euclidean space. J. Math. Pures Appl. 169, 50–81 (2023)

    MathSciNet  MATH  Google Scholar 

  17. Cardona, R., Miranda, E., Peralta-Salas, D., Presas, F.: Universality of Euler flows and flexibility of Reeb embeddings. (2021). arXiv:1911.01963

  18. Cardona, R., Miranda, E., Peralta-Salas, D., Presas, F.: Constructing turing complete Euler flows in dimension. Proc. Natl. Acad. Sci. 118, e2026818118 (2021)

    MathSciNet  Google Scholar 

  19. Coullet, P., Tresser, C., Arneodo, A.: Transition to turbulence for doubly periodic flows. Phys. Lett. A 77(5), 327–331 (1980)

    ADS  MathSciNet  Google Scholar 

  20. Cieliebak, K., Volkov, E.: A note on the stationary Euler equations of hydrodynamics. Ergod. Theory Dyn. Syst. 37(2), 454–480 (2017)

    MathSciNet  MATH  Google Scholar 

  21. Delshams, A., Ramírez-Ros, R.: Poincaré-Melnikov-Arnold method for analytic planar maps. Nonlinearity 9(1), 1–26 (1996)

    ADS  MathSciNet  MATH  Google Scholar 

  22. Delshams, A., Ramírez-Ros, R.: Melnikov potential for exact symplectic maps. Comm. Math. Phys. 190(1), 213–245 (1997)

    ADS  MathSciNet  MATH  Google Scholar 

  23. Duarte, P.: Abundance of elliptic isles at conservative bifurcations. Dyn. Stab. Syst. 14(4), 339–356 (1999)

    MathSciNet  MATH  Google Scholar 

  24. Duarte, P.: Elliptic isles in families of area-preserving maps. Ergod. Theory Dyn. Syst. 28(6), 1781–1813 (2008)

    MathSciNet  MATH  Google Scholar 

  25. Enciso, A., García-Ferrero, M., Peralta-Salas, D.: Approximation theorems for parabolic equations and movement of local hot spots. Duke Math. J. 168(5), 897–939 (2019)

    MathSciNet  MATH  Google Scholar 

  26. Enciso, A., Luque, A., Peralta-Salas, D.: Beltrami fields with hyperbolic periodic orbits enclosed by knotted invariant tori. Adv. Math. 373, 107328 (2020)

    MathSciNet  MATH  Google Scholar 

  27. Enciso, A., Peralta-Salas, D.: Knots and links in steady solutions of the Euler equation. Ann. Math. 175(1), 345–367 (2012)

    MathSciNet  MATH  Google Scholar 

  28. Enciso, A., Peralta-Salas, D.: Existence of knotted vortex tubes in steady Euler flows. Acta Math. 214(1), 61–134 (2015)

    MathSciNet  MATH  Google Scholar 

  29. Enciso, A., Peralta-Salas, D.: Knotted vortex structures in incompressible fluid flows. Bull. London Math. Soc., in press (2023)

  30. Enciso, A., Peralta-Salas, D., Romaniega, A.: Beltrami fields exhibit knots and chaos almost surely. (2020). arXiv:2006.15033

  31. Enciso, A., Peralta-Salas, D., Torres de Lizaur, F.: Knotted structures in high-energy Beltrami fields on the torus and the sphere. Ann. Sci. Éc. Norm. Supér. 50(4), 995–1016 (2017)

    MathSciNet  MATH  Google Scholar 

  32. Eilenberg, S., Steenrod, N.: Foundations of Algebraic Topology. Princeton University Press, Princeton, New Jersey (1952)

    MATH  Google Scholar 

  33. Feigenbaum, M.J.: The onset spectrum of turbulence. Phys. Lett. A 74(6), 375–378 (1979)

    ADS  MathSciNet  Google Scholar 

  34. Farge, M., Pellegrino, G., Schneider, K.: Coherent vortex extraction in 3d turbulent flows using orthogonal wavelets. Phys. Rev. Lett., 87, (2001)

  35. Gonchenko, M.S., Gonchenko, S.V.: On cascades of elliptic periodic points in two-dimensional symplectic maps with homoclinic tangencies. Regul. Chaotic Dyn. 14(1), 116–136 (2009)

    ADS  MathSciNet  MATH  Google Scholar 

  36. Gorodetski, A.: On stochastic sea of the standard map. Commun. Math. Phys. 309(1), 155–192 (2012)

    ADS  MathSciNet  MATH  Google Scholar 

  37. Glasser, M.L., Papageorgiou, V.G., Bountis, T.C.: Melnikov’s function for two-dimensional mappings. SIAM J. Appl. Math. 49(3), 692–703 (1989)

    MathSciNet  MATH  Google Scholar 

  38. Gonchenko, S. V., Shilnikov, L. P.: On two-dimensional analytic area-preserving diffeomorphisms with a countable set of elliptic stable periodic points. vol. 2, pp. 106–123. (1997). V. I. Arnold (on the occasion of his 60th birthday) (Russian)

  39. Gelfreich, V., Sauzin, D.: Borel summation and splitting of separatrices for the Hénon map. Ann. Inst. Fourier (Grenoble) 51(2), 513–567 (2001)

    MathSciNet  MATH  Google Scholar 

  40. Gonchenko, S.V., Shilnikov, L.P.: On two-dimensional area-preserving mappings with homoclinic tangencies. Dokl. Akad. Nauk 378(6), 727–732 (2001)

    MathSciNet  MATH  Google Scholar 

  41. Gonchenko, S. V., Shilnikov, L. P.: On two-dimensional area-preserving maps with homoclinic tangencies that have infinitely many generic elliptic periodic points. Zap. Nauchn. Sem. S.-Peterburg. Otdel. Mat. Inst. Steklov. (POMI), 300(Teor. Predst. Din. Sist. Spets. Vyp. 8):155–166, 288–289, (2003)

  42. Gonchenko, S. V., Shilnikov, L. P., Turaev, D. V.: Elliptic periodic orbits near a homoclinic tangency in four-dimensional symplectic maps and Hamiltonian systems with three degrees of freedom. vol. 3, pp. 3–26. (1998). J. Moser at 70 (Russian)

  43. Gonchenko, S., Shilnikov, L., Turaev, D.: Homoclinic tangencies of arbitrarily high orders in conservative and dissipative two-dimensional maps. Nonlinearity 20(2), 241–275 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  44. Gelfreich, V., Turaev, D.: Universal dynamics in a neighborhood of a generic elliptic periodic point. Regul. Chaotic Dyn. 15(2–3), 159–164 (2010)

    ADS  MathSciNet  MATH  Google Scholar 

  45. Gonchenko, S.V., Turaev, D.V., Shilnikov, L.P.: On the existence of Newhouse regions in a neighborhood of systems with a structurally unstable homoclinic Poincaré curve (the multidimensional case). Dokl. Akad. Nauk 329(4), 404–407 (1993)

    Google Scholar 

  46. Hénon, M.: Sur la topologie des lignes de courant dans un cas particulier. C. R. Acad. Sci. Paris 262, 312–314 (1966)

    Google Scholar 

  47. Hurewicz, W., Wallman, H.: Dimension Theory. Princeton Mathematical Series, vol. 4. Princeton University Press, Princeton, N. J., (1941)

  48. Irwin, M.C.: On the smoothness of the composition map. Quart. J. Math. Oxford Ser. 2(23), 113–133 (1972)

    ADS  MathSciNet  MATH  Google Scholar 

  49. Kupka, I.: Contribution à la théorie des champs génériques. Contrib. Differ. Equ. 2, 457–484 (1963)

    MATH  Google Scholar 

  50. Lomelí, H.E., Meiss, J.D., Ramírez-Ros, R.: Canonical Melnikov theory for diffeomorphisms. Nonlinearity 21(3), 485–508 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  51. Massey, W.S.: How to give an exposition of the Čech-Alexander-Spanier type homology theory. Amer. Math. Monthly 85(2), 75–83 (1978)

    MathSciNet  MATH  Google Scholar 

  52. Melnikov, V.K.: On the stability of a center for time-periodic perturbations. Trudy Moskov. Mat. Obšč. 12, 3–52 (1963)

    MathSciNet  Google Scholar 

  53. Miller, H.: Lecture notes in Algebraic Topology I. https://ocw.mit.edu/courses/mathematics/18-905-algebraic-topology-i-fall-2016/lecture-notes/, Fall (2016)

  54. Moise, E. E.: Geometric topology in dimensions\(2\)and\(3\). Graduate Texts in Mathematics, Vol. 47. Springer-Verlag, New York-Heidelberg, (1977)

  55. Mora, L., Romero, N.: Persistence of homoclinic tangencies for area-preserving maps. Ann. Fac. Sci. Toulouse Math. 6(4), 711–725 (1997)

    MathSciNet  MATH  Google Scholar 

  56. Monchaux, R., Ravelet, F., Dubrulle, B., Chiffaudel, A., Daviaud, F.: Properties of steady states in turbulent axisymmetric flows. Phys. Rev. Lett., 96, (2006)

  57. Morgulis, A., Yudovich, V.I., Zaslavsky, G.M.: Compressible helical flows. Comm. Pure Appl. Math. 48(5), 571–582 (1995)

    MathSciNet  MATH  Google Scholar 

  58. Newhouse, S. E.: Nondensity of axiom \({\rm A}({\rm a})\) on \(S^{2}\). In Global Analysis (Proc. Sympos. Pure Math., Vol. XIV, Berkeley, Calif., 1968), pp. 191–202. Amer. Math. Soc., Providence, R.I., (1970)

  59. Newhouse, S.E.: Diffeomorphisms with infinitely many sinks. Topology 13, 9–18 (1974)

    MathSciNet  MATH  Google Scholar 

  60. Newhouse, S., Ruelle, D., Takens, F.: Occurrence of strange Axiom A attractors near quasiperiodic flows on \(T^{m}\),\(\, m\ge 3\). Comm. Math. Phys. 64(1), 35–40 (1978)

    ADS  MathSciNet  MATH  Google Scholar 

  61. Poincaré, H.: Les méthodes nouvelles de la mécanique céleste. Gauthier-Villars, Paris (1899)

    MATH  Google Scholar 

  62. Ruelle, D., Takens, F.: On the nature of turbulence. Commun. Math. Phys. 20, 167–192 (1971)

    ADS  MathSciNet  MATH  Google Scholar 

  63. Smale, S.: Stable manifolds for differential equations and diffeomorphisms. Ann. Scuola Norm. Sup. Pisa Cl. Sci. 17, 97–116 (1963)

    MathSciNet  MATH  Google Scholar 

  64. Tao, T.: On the universality of the incompressible Euler equation on compact manifolds. Discrete Contin. Dyn. Syst. 38(3), 1553–1565 (2018)

    MathSciNet  MATH  Google Scholar 

  65. Taylor, M. E.: Partial differential equations I. Basic theory, volume 115 of Applied Mathematical Sciences. Springer, New York, second edition, (2011)

  66. Torres de Lizaur, F.: Chaos in the incompressible Euler equation on manifolds of high dimension. Invent. Math. 228, 687–715 (2022)

  67. Turaev, D.: Polynomial approximations of symplectic dynamics and richness of chaos in non-hyperbolic area-preserving maps. Nonlinearity 16(1), 123–135 (2003)

    ADS  MathSciNet  MATH  Google Scholar 

  68. Turaev, D.: Maps close to identity and universal maps in the Newhouse domain. Commun. Math. Phys. 335(3), 1235–1277 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  69. Vakulenko, S.: Strange attractors for Oberbeck-Boussinesq model. J. Dyn. Differ. Equ. 33(1), 303–343 (2021)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors are very grateful to Francisco Romero Ruiz del Portal for pointing us a reference for a proof that a countable union of totally disconnected compact sets is totally disconnected as well, cf. Fact 6.4.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre Berger.

Additional information

Communicated by C. Liverani.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Partially supported by the ERC project 818737 Emergence of wild differentiable dynamical systems. Partially supported by the project ANR CoSyDy (ANR-CE40-0014). Supported by the grants PID2019-106715GB GB-C21, CEX2019-000904-S and ‘Computational, dynamical and geometrical complexity in fluid dynamics”, Ayudas Fundacion BBVA a Proyectos de Investigacion Cientifica 2021.

Appendices

Appendices

Proof of Proposition 3.7: Right Inverse of the Melnikov Operator

The aim of this appendix is to construct an inverse of the Melnikov operator for a vector field \(X\in \Gamma ^\infty _\textrm{Leb}(U)\) displaying a strong heteroclinic link in U, for some open subset U of \({\mathbb {R}}^3\).

1.1 Displacement and Melnikov operators

The aim of this section is to introduce the definitions of displacement operator, Melnikov operator and explain their relation.

Let \(\Sigma \subset {\mathbb {R}}^2\) and let \(\Omega \) be the standard symplectic, smooth 2-form on \(\Sigma \subset {\mathbb {R}}^2\). Let \(f\in C^\infty (\Sigma ,{\mathbb {R}}^2)\) be a smooth, symplectic diffeomorphisms from \(\Sigma \) onto its image. We assume that f displays two hyperbolic fixed point \(p^+\), \(p^-\) and a heteroclinic link L from \(p^-\) to \(p^+\). The displacement operator will describe the unfolding of L when we consider a perturbation of f.

Let \({\mathcal {N}}\subset C^\infty (\Sigma ,{\mathbb {R}}^2)\) be a neighborhood of 0. Assume that \({\mathcal {N}}\) is small enough such that, for every \(\epsilon \in {\mathcal {N}}\), the diffeomorphism \(f+\epsilon \) admits unique hyperbolic fixed points \(p^\pm _\epsilon \) nearby \(p^\pm \), called their hyperbolic continuations. Let \(\ell :{\mathbb {R}}\rightarrow L\) be a smooth parametrization of the heteroclinic connection L such that

$$\begin{aligned} \lim _{t\rightarrow \pm \infty }\ell (t)=p^\pm \quad \text {and}\quad f(\ell (t))=\ell (t+1). \end{aligned}$$

Fix local unstable and local stable manifolds \(W^u_{loc}(p^-;f)\) and \(W^s_{loc}(p^+;f)\), of respectively \(p^-\) and \(p^+\), which both contain \(\ell ([0,1])\).

For a fixed \(\epsilon \in C^\infty (\Sigma ,{\mathbb {R}}^2)\) small, consider the smooth immersed submanifold \(W^s_{loc}(p^+_\epsilon ; f+\epsilon )\). We can find a smooth diffeomorphism \(\phi ^s_\epsilon :W^s_{loc}(p^+; f)\rightarrow W^s_{loc}(p^+_\epsilon ; f+\epsilon )\). Similarly, we find a smooth diffeomorphism \(\phi ^u_\epsilon :W^u_{loc}(p^-; f)\rightarrow W^u_{loc}(p^-_\epsilon ; f+\epsilon )\). By Irwin’s proof [Irw72, Theorem 28] of the stable/unstable manifold theorem (which is based on the implicit function theorem), both \(\phi ^s_\epsilon \) and \(\phi ^u_\epsilon \) can be chosen so that they depend smoothly on \(\epsilon \). Up to a smooth reparametrization, we can assume that

$$\begin{aligned} (f+\epsilon )\phi ^u_\epsilon (\ell (0))=\phi ^u_\epsilon (\ell (1))\qquad \text {and}\qquad (f+\epsilon )\phi ^s_\epsilon (\ell (0))=\phi ^s_\epsilon (\ell (1))\, . \end{aligned}$$
(B.1)

We define, for \(t\in [0,1]\),Footnote 8

$$\begin{aligned} \textrm{displ}(\epsilon )(t)=\Omega (\partial _t\ell (t),\phi ^u_\epsilon (\ell (t))-\phi ^s_\epsilon (\ell (t))). \end{aligned}$$
(B.2)

Roughly speaking, \(\textrm{displ}(\epsilon )(t)\) is the algebraic distance, along an affine line passing through \(\ell (t)\) and normal to L, between the perturbed local stable and unstable manifolds. See Fig. 6.

Fig. 6
figure 6

The displacement function is the algebraic distance between perturbed local stable and unstable manifolds

By Eq. (B.1), \(\phi ^u_\epsilon (\ell ([0,1]))\) and \(\phi ^s_\epsilon (\ell ([0,1]))\) are fundamental domains. Then, the function \(\textrm{displ}(\epsilon )\) vanishes if and only if \(\phi ^s_\epsilon (\ell (t))= \phi ^u_\epsilon (\ell (t))\) for every \(t\in [0,1]\). Thus, the heteroclinic link persists if and only if \(\textrm{displ}(\epsilon )\equiv 0\).

Definition A.1

The displacement operator is the following map from a neighborhood \({\mathcal {N}}\) of 0 in \(C^\infty (\Sigma ,{\mathbb {R}}^2)\):

$$\begin{aligned}{} & {} \textrm{displ}:{\mathcal {N}}\rightarrow C^\infty ([0,1],{\mathbb {R}}) \\{} & {} \quad \epsilon \mapsto \textrm{displ}(\epsilon ), \end{aligned}$$

where the function \(\textrm{displ}(\epsilon )\) is defined in Eq. (B.2).

Remark A.2

The displacement operator is not uniquely defined: it depends on the parametrization \(\ell \) of L and on the families \((\phi ^u_\epsilon )_\epsilon \) and \((\phi ^s_\epsilon )_\epsilon \).

We recall that \(\phi ^s_\epsilon \) and \(\phi ^u_\epsilon \) depend smoothly on \(\epsilon \), so the following proposition holds.

Proposition A.3

The map \(\textrm{displ}:{\mathcal {N}}\rightarrow C^\infty ([0,1],{\mathbb {R}})\) is smooth.

The distance between the perturbed stable and unstable manifolds can be studied through the so-called Melnikov functions. We present here the main definitions and properties and refer for detailed proofs to [LMRR08, DRR97, DRR96, GPB89]; the historical references are [Poi99] and [Mel63].

Definition A.4

Let \(f\in C^\infty (\Sigma ,{\mathbb {R}}^2)\) be a symplectic diffeomorphism, which displays a heteroclinic link L, parametrized by \(\ell :{\mathbb {R}}\rightarrow L\). The Melnikov function associated to a smooth family \(r\mapsto \epsilon (r)\in C^\infty (\Sigma ,{\mathbb {R}}^2)\), with \(\epsilon (0)=0\), is:

$$\begin{aligned} \begin{aligned} \mathcal {M}(\epsilon (s)):&\ {\mathbb {R}}\rightarrow {\mathbb {R}}\\ t\mapsto {\mathcal {M}}(\epsilon (s))(t ):=&\,\sum _{k\in {\mathbb {Z}}}\Omega \left( \partial _t\ell (t), (f^*)^k(\partial _r\epsilon (r)|_{r=0}\circ f^{-1})\circ \ell (t)\right) \, , \end{aligned} \end{aligned}$$
(B.3)

where \(\partial _r\epsilon (r)|_{r=0}:=\lim _{r\rightarrow 0}\frac{\epsilon (r)}{r}\).

Remark A.5

In Eq. (B.3) the presence of \(\partial _r\epsilon (r)|_{r=0}\circ f^{-1}\) is due to the fact that, for the given family \((f+\epsilon (r))_r\), it is the vector field \(X=\partial _r\epsilon (r)|_{r=0}\circ f^{-1}\) which satisfies \(\frac{d}{dr}(f+\epsilon (r))|_{r=0}=X\circ f\). See [LMRR08, Section 2].

Recall that f is symplectic and, since \(f(\ell (t))=\ell (t+1)\), \(\partial _t\ell (t)= Df^{-k}(\ell (t+k))\partial _t\ell (t+k)\). Thus, it holds

$$\begin{aligned} {\mathcal {M}}(\epsilon (s))(t)=\sum _{k\in {\mathbb {Z}}}\Omega \left( \partial _t\ell (t+k), \partial _s\epsilon |_{s=0}(\ell (t+k-1))\right) . \end{aligned}$$
(B.4)

Note that the Melnikov function belongs to the space \(C^\infty ({\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}})\) of smooth 1-periodic functions.

Definition A.6

Let f be a symplectic diffeomorphism in \(C^\infty (\Sigma ,{\mathbb {R}}^2)\) displaying a heteroclinic link L, parametrized by \(\ell :{\mathbb {R}}\rightarrow L\). The Melnikov operator is a linear and continuous map

$$\begin{aligned} {\mathcal {M}}:C^\infty (\Sigma ,{\mathbb {R}}^2)\ni \epsilon \mapsto {\mathcal {M}}(\epsilon )\in C^\infty ({\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}})\, , \end{aligned}$$

where \({\mathcal {M}}(\epsilon )\) is the Melnikov function with respect to the smooth family \(r\mapsto r\epsilon \).

Remark A.7

Observe that \({\mathcal {M}}(\epsilon )\) depends only on \(\epsilon |_L\). Thus, we can define \({\mathcal {M}}(\epsilon )\) for \(\epsilon \in C^\infty (L,{\mathbb {R}}^2)\).

Theorem A.8

(Poincaré-Melnikov, [Poi99]–[Mel63]) Let f be a symplectic diffeomorphism in \(C^\infty (\Sigma ,{\mathbb {R}}^2)\), displaying a heteroclinic link L, parametrized by \(\ell :{\mathbb {R}}\rightarrow L\). Then, the partial differential of the displacement operator at 0 is equal to the Melnikov operator at \(\epsilon \), i.e.

$$\begin{aligned} D_0\textrm{displ}={\mathcal {M}}. \end{aligned}$$
(B.5)

Thus, for any \(r\ge 1\), in the \(C^r\)-uniform norm it holds

$$\begin{aligned} \textrm{displ}(\epsilon )= \textrm{displ}(0)+ D_0\textrm{displ}(\epsilon )+O(\epsilon ^2)={\mathcal {M}}(\epsilon )+O(\epsilon ^2). \end{aligned}$$

For a detailed proof, we refer for example to [DRR97] and [GPB89].

1.2 Inverse Melnikov operator for flows in \({\mathbb {R}}^3\)

Let U be an open subset of \({\mathbb {R}}^3\). Let \(X\in \Gamma ^\infty _\textrm{Leb}(U)\) display a strong heteroclinic link \(\Gamma \subset U\) between two saddle periodic orbits \(\gamma ^\pm \subset U\). In particular, the closure of the link \({\bar{\Gamma }}\) is diffeomorphic to \({\mathbb {R}}/{\mathbb {Z}}\times [-1,1]\).

Let \({\bar{L}}\) be the closure of the half strong stable manifold of a point in \(\gamma ^+\) in \(\Gamma \). Let \({\hat{\Sigma }}\subset {\mathbb {R}}^3\) be a disk which contains \({\bar{L}}\) and that intersects transversally \(\Gamma \). Up to reducing \({\hat{\Sigma }}\), it is transverse to the vector field X. Let \(\Sigma \Subset {\hat{\Sigma }}\) be a neighborhood of L in \({\hat{\Sigma }}\) such that for every point \(z \in \Sigma \), there exists \(\tau (z)>0\) minimal such that the orbit of z intersect transersally \({\hat{\Sigma }}\) at time \(\tau (z)\) at a point f(z). From the definition of strong heteroclinic link, the first return time \(\tau |_{ \Sigma \cap \Gamma }\) is constant, in particular \(\gamma ^+\) and \(\gamma ^-\) have the same period. Indeed, since \(\Sigma \cap \Gamma =L\) is a half strong stable manifold of a point in \(\gamma ^+\), it is \(T(\gamma ^+)\)-periodic, where \(T(\gamma ^+)\) denotes the period of \(\gamma ^+\) (similarly, for a half strong unstable manifold of a point if \(\gamma ^-\)). The disks \(\Sigma \subset {\hat{\Sigma }}\) are called Poincaré sections and f is called a Poincaré first return map. The maps f and \(\tau \) are in, respectively, \(C^\infty (\Sigma ,{\hat{\Sigma }})\) and \( C^\infty (\Sigma ,{\mathbb {R}}^+)\). Moreover, it is standard that there exists a smooth symplectic form \(\Omega \) in \({\hat{\Sigma }}\) which is invariant by the action of \(f|_\Sigma \). By Darboux’s Theorem, we can identify \(\Sigma \) with a disk of \({\mathbb {R}}^2\) so that \(\Omega \) is the standard symplectic form.

Let \(W\in \Gamma ^\infty _\textrm{Leb}(U)\) be a small perturbation of X: it induces a small perturbation \(\epsilon \in C^\infty (\Sigma ,{\mathbb {R}}^2)\) of the Poincaré map f. So we can study the unfolding of \(\Gamma \) using the tools of Melnikov operator for the heteroclinic link \(L={\hat{\Sigma }}\cap \Gamma \) presented above.

Fix a parametrization \(\ell :{\mathbb {R}}\rightarrow L\) of L, such that \(\lim _{t\rightarrow \pm \infty }\ell (t)=p^\pm \), where \(p^\pm :={\hat{\Sigma }}\cap \gamma ^\pm \), and \(f(\ell (t))=\ell (t+1)\). Fix also \((\phi ^s_\epsilon )_\epsilon \) and \((\phi ^u_\epsilon )_\epsilon \) as in A.1. Let \({\mathcal {M}}\) be the Melnikov operator associated to this setting.

Definition A.9

The Melnikov operator is the following linear and continuous map from a neighborhood \({\mathcal {N}}\) of 0 in \(\Gamma ^\infty _\textrm{Leb}(U)\)

$$\begin{aligned} {\mathcal {M}}: {\mathcal {N}}\ni W\mapsto {\mathcal {M}}(W):={\mathcal {M}}(\epsilon )\in C^\infty ({\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}})\, , \end{aligned}$$

where \(\epsilon \in C^\infty (\Sigma ,{\mathbb {R}}^2)\) is the perturbation of the first return map f associated to the vector field W.

Remark A.10

Observe that, since \({\mathcal {M}}(\epsilon )\) depends only on \(\epsilon |_L\) (see Remark A.7), and so on \(W|_\Gamma \), the operator \({\mathcal {M}}\) can be defined for \(W\in C^\infty (\Gamma ,{\mathbb {R}}^3)\).

The aim of this subsection is to give the proof of Proposition 3.7: such a proposition gives an inverse operator of the Melnikov one.

Proof of Proposition 3.7

We recall that \(\Gamma \) is an invariant cylinder smoothly embedded in \({\mathbb {R}}^3\), and that \(L\subset \Gamma \) is a line transverse to the flow whose return time is a constant \(T>0\). As the statement is invariant by reparametrizing the flow, we can assume that \(T=1\).

This defines a canonical diffeomorphism \(\Gamma \rightarrow L\times {\mathbb {R}}/{\mathbb {Z}}\) which maps L to \(L\times \{0\}\). Recall that \(\ell :{\mathbb {R}}\rightarrow L\) is a parametrization of L such that \(f\circ \ell (t)=\ell (t+1)\) for every \(t\in {\mathbb {R}}\), where f is the first return map to \( \Sigma \). Then using \(\ell \), we identify \(\Gamma \) to \({\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\) and L to \({\mathbb {R}}\times \{0\}\):

$$\begin{aligned} \Gamma \equiv {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\quad \text {and}\quad L\equiv {\mathbb {R}}\times \{0\} \; , \end{aligned}$$

such that, with \(p_2: \Gamma \rightarrow {\mathbb {R}}/{\mathbb {Z}}\) the second coordinate projection, the flow \((\Phi ^t_X)_{t\in {\mathbb {R}}}\) of X satisfies:

$$\begin{aligned} p_2\circ \Phi ^t_X(s,\theta )= \theta +t\ \mod 1\; , \quad \forall (s,\theta )\in {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\equiv \Gamma \, . \end{aligned}$$

Lemma A.11

There is a compactly supported function \(\Psi \in C^\infty ({\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}}^+)\) such that:

$$\begin{aligned} \int _{\mathbb {R}}\Psi \circ \Phi ^t_X(s,0)dt=1\;, \quad \forall s\in {\mathbb {R}}. \end{aligned}$$

Proof

Let \(\Psi _0\in C^\infty ({\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}}^+)\) be a smooth, compactly supported, non-negative function which is positive on \([\tfrac{1}{2}, \tfrac{3}{2}]\times \{0\}\). For every \((s,\theta )\in {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\equiv \Gamma \) define

$$\begin{aligned} C(s,\theta ):=\int _{\mathbb {R}}\Psi _0\circ \Phi ^t_X(s,\theta )dt\, . \end{aligned}$$

Note that every orbit intersects \([\tfrac{1}{2}, \tfrac{3}{2}]\times \{0\}\), so C is positive. Also the function C is \(\Phi _V^t\)-invariant and smooth. Actually, \(C(s,\theta )\) is equal to some constant C, because \(\Phi ^t_X|_{\Gamma }\) does not have any non trivial continuous first integral. Define then the function \(\Psi \in C^\infty ({\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}, {\mathbb {R}}^+)\) as

$$\begin{aligned} (s,\theta )\in {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\mapsto \Psi (s,\theta ):=\dfrac{\Psi _0(s,\theta )}{C}\in {\mathbb {R}}^+\, . \end{aligned}$$

It is obvious that the mean of \(\Psi \) along the orbits is one, as desired. \(\square \)

Note that for every \((s,\theta )\in {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\), the orbit of \((s,\theta )\) intersects L at a set \((\varphi (s,\theta )+{\mathbb {Z}})\times \{0\}\). This defines a \(\Phi _X^t\)-invariant map:

$$\begin{aligned} \varphi : {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\rightarrow {\mathbb {R}}/{\mathbb {Z}}. \end{aligned}$$

Note that for every \(s,t\in {\mathbb {R}}\) it holds

$$\begin{aligned} \varphi \circ \Phi ^t_X(s,0)=s \mod 1\, . \end{aligned}$$
(B.6)

We recall that \(\Sigma \) is a Poincaré section containing L and transverse to \(\Gamma \) and that \(\Omega \) denote an invariant symplectic form for the first return map f. Let \(N:{\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\equiv \Gamma \rightarrow {\mathbb {R}}^3\) be a vector field such that:

$$\begin{aligned} T\Gamma \oplus {\mathbb {R}}N= {\mathbb {R}}^3\quad \text {and}\quad \Omega (\partial _t \ell (s), N_\Sigma \circ \ell (s))= 1\quad \forall s\in {\mathbb {R}}\; , \end{aligned}$$
(B.7)

where \(N_\Sigma \) denotes the projection of \(N|_\Gamma \) into the section \(\Sigma \).

Now for every \(g\in C^\infty ({\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}})\) we define the smooth vector field \({\mathcal {I}} (g)\in C^\infty ({\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}},{\mathbb {R}}^3)\) as:

$$\begin{aligned} {\mathcal {I}} (g):= (s , \theta )\in \Gamma \mapsto g\circ \varphi (s,\theta )\cdot \Psi (s,\theta )\cdot N(s, \theta )\in {\mathbb {R}}^3. \end{aligned}$$
(B.8)

We shall prove that the operator which associates to g the Melnikov function of \({\mathcal {I}} (g)\) is the identity. In order to do this, let us go back to the two dimensional setting. Let \(\epsilon (g)\in C^\infty (\Sigma , {\mathbb {R}}^2)\) be the perturbation induced by the vector field \(X+{\mathcal {I}}(g)\) on the first return map f. Consider the smooth family \(r\mapsto \epsilon (rg)\). Then the partial derivative \(\partial _r\epsilon |_{r=0}\) is:

$$\begin{aligned} \partial _r\epsilon (rg)|_{r=0}=\lim _{r\rightarrow 0} \frac{1}{r} \epsilon (r\cdot g) \, . \end{aligned}$$

Lemma A.12

For every \(s\in {\mathbb {R}}\) it holds:

$$\begin{aligned} \Omega (\partial _t\ell (s+1), \partial _r\epsilon (rg)|_{r=0}\circ \ell (s))=g(s)\cdot \left( \int _0^{1}\Psi \circ \Phi ^t_X(s,0)dt\right) \, , \end{aligned}$$

where g is identified with a 1-periodic real function.

Proof

A flow-box coordinate neighborhood V of \(\Gamma \) is diffeomorphic to \(\Gamma \times (-1,1)\). This diffeomorphism induces an identification of V with \({\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\times (-1,1)\) and can be chosen so that N is constantly equal to (0, 0, 1), \(\Sigma \cap V\) is identified with \({\mathbb {R}}\times \{0\}\times {\mathbb {R}}\) and X has zero third component.

Let \(X_g:= X+{{\mathcal {I}}}(g)\) and denote by \((\Phi _{X_g}^t)_{t\in {\mathbb {R}}}\) its flow. Let \(\tau _g:\Sigma \rightarrow {\mathbb {R}}\) be the first return time on \(\Sigma \) associated to \(X_g\). In the coordinates given by the identification, we have:

$$\begin{aligned} X_g(s,t)= X(s,\theta )+(0,0, g\circ \varphi (s,\theta )\cdot \Psi (s,\theta ))\; ,\quad \forall (s,\theta )\in {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\; . \end{aligned}$$

Let \(z=\ell (s)\) and express the perturbation \(\epsilon (g)\) associated to \({\mathcal {I}} (g)\) as

$$\begin{aligned} \epsilon (g)(z)=\left( (f+\epsilon (g))(z)-z\right) -(f(z)-z)\,. \end{aligned}$$

In the coordinates of the tubular neighborhood of \(\Gamma \) we have:

$$\begin{aligned} (0,\epsilon (g)(z))=\int _0^{\tau _g(z)}(X+{\mathcal {I}}(g))\circ \Phi ^t_{X+{\mathcal {I}}(g)}(z)dt-\int _0^{1}X\circ \Phi ^t_X(z)dt\, , \end{aligned}$$

where the coordinate 0 is the coordinate in \({\mathbb {R}}/{\mathbb {Z}}\) (we have changed the ordering of the coordinates for the ease of notation).

Hence for every \(r\ge 1\):

$$\begin{aligned} (0,\epsilon (g)(z))={} & {} \int _1^{\tau _g(z)}X\circ \Phi ^t_{X}(z)dt+ \int _0^{1} {\mathcal {I}}(g) \circ \Phi ^t_{X}(z)dt+ \\ {}{} & {} \int _0^{1}\left( X\circ \Phi ^t_{X+{\mathcal {I}}(g)}(z)-X\circ \Phi ^t_X(z)\right) dt+ O(\Vert g\Vert ^2_{C^r})\, , \end{aligned}$$

where we have used that \(\int _1^{\tau _g(z)}X\circ \Phi ^t_{X+{\mathcal {I}}(g)}(z) dt = \int _1^{\tau _g(z)} X\circ \Phi ^t_X(z)dt+ O(\Vert g\Vert ^2_{C^r})\). Since, with \(p_3: {\mathbb {R}}\times {\mathbb {R}}/{\mathbb {Z}}\times (-1,1)\rightarrow (-1,1)\) the third coordinate projection, it holds \(p_3(X)=0\) on the neighborhood V, the first and third terms have their third components null while the second term has its first and second components null. Thus, it holds:

$$\begin{aligned} p_3(0,\epsilon (g)(z))= p_3\left( \int _0^{1} {\mathcal {I}}(g) \circ \Phi ^t_{X}(z)dt\right) + O(\Vert g\Vert ^2_{C^r})\; . \end{aligned}$$

Now we plug Eq. (B.8) to obtain:

$$\begin{aligned} p_3(0,\epsilon (g)(z))= p_3\left( \int _0^{1} g\circ \varphi \circ \Phi ^t_{X}(z)\cdot \Psi \circ \Phi ^t_{X}(z)\cdot N \circ \Phi ^t_{X}(z)dt\right) + O(\Vert g\Vert ^2_{C^r})\; . \end{aligned}$$

Now we use that \(p_3\circ N\equiv 1\) and that, from (B.7), \(\Omega (\partial _t \ell (s+1) , \epsilon (g)\circ \ell (s))= p_3(0,\epsilon (g)(z))\), to obtain:

$$\begin{aligned} \Omega (\partial _t \ell (s+1), \epsilon (g) \circ \ell (s))= \int _0^{1} g\circ \varphi \circ \Phi ^t_{X}(z)\cdot \Psi \circ \Phi ^t_{X}(z)dt + O(\Vert g\Vert ^2_{C^r})\; . \end{aligned}$$
(B.9)

Using the invariance of \(\varphi \) by the flow, we have \( g\circ \varphi \circ \Phi ^t_{X}(z)= g\circ \ell (s)\) for every t. Using Eq.(B.9) with the perturbation \(\epsilon (rg)\), taking its derivative with respect to r and evaluating at \(r=0\) yield the sought equality. \(\square \)

The Melnikov function associated to \({\mathcal {I}}(g)\) is

$$\begin{aligned} {\mathcal {M}}({\mathcal {I}}(g))(s)={\mathcal {M}}(\epsilon (g))(s)=\sum _{k\in {\mathbb {Z}}}\Omega (\partial _t\ell (s+k),\partial _r\epsilon (rg)|_{r=0}(\ell (s+k-1))). \end{aligned}$$

By Lemma A.12, since the function g is 1-periodic and by Proposition A.11, we deduce that

$$\begin{aligned} {\mathcal {M}}({\mathcal {I}}(g))(s)=g(s)\cdot \sum _{k\in {\mathbb {Z}}} \left( \int _0^{1}\Psi \circ \Phi ^t_X(s+k,0)\,dt\right) \end{aligned}$$

and, since \(\Phi ^t_X(s+k,0)=\Phi ^{t+k}_X(s,0)\) for every \(k\in {\mathbb {Z}}\), we conclude that

$$\begin{aligned} {\mathcal {M}}({\mathcal {I}}(g))(s)= g(s)\cdot \int _{\mathbb {R}}\Psi \circ \Phi ^t_X(s,0) dt = g(s)\, . \end{aligned}$$

\(\square \)

Proof of Theorem 3.4: Cauchy–Kovalevskaya’s Theorem for Curl

1.1 Application of Cauchy–Kovalevskaya’s Theorem for curl

Let \(\Sigma \subset {\mathbb {R}}^3\) be a bounded, oriented, analytic surface that can be analytically extended in a neighborhood. Given a vector field W on \(\Sigma \), we can decompose it as follows:

$$\begin{aligned} W=W_T+(W\cdot N)N, \end{aligned}$$

where N is a unit normal to \(\Sigma \), while \(W_T\) is the component of W tangent to \(\Sigma \). We refer to W as the Cauchy datum for the Cauchy–Kovalevskaya theorem. Denote as \(g:=W\cdot N\) the normal component of the Cauchy datum.

We will be interested in W satisfying the following condition:

where \(W^\flat _T\) is the 1-form dual to \(W_T\), push-forwarded on \(\Sigma \), and \(\sigma \) is the area form on \(\Sigma \) induced from the ambient space \({\mathbb {R}}^3\). The following generalizes Theorem 3.1 in [EPS12].

Theorem B.1

(Cauchy–Kovalevskaya’s Theorem for curl). Let \(\Sigma \subset {\mathbb {R}}^3\) be a surface as above and let W be the analytic Cauchy datum in \(C^\omega ({\bar{\Sigma }},{\mathbb {R}}^3)\). The equation

$$\begin{aligned} \textrm{curl }\, X&= X, \end{aligned}$$
(B.1)
$$\begin{aligned} X|_{\Sigma }&=W \end{aligned}$$
(B.2)

has a unique, analytic solution in a neighborhood \(\Omega \) of \(\Sigma \) if and only if W fulfills condition (\(*\)).

This theorem implies Theorem 3.4

Proof of Theorem 3.4

Let \(\Gamma \) be an analytic surface in \({\mathbb {R}}^3\) whose closure \({\bar{\Gamma }}\) is diffeomorphic to \({\mathbb {I}}\times {\mathbb {R}}/{\mathbb {Z}}\). We are going to show that for every \(g\in C^\omega ({\bar{\Gamma }},{\mathbb {R}})\), there exists a neighborhood \(\Omega \) of \({\bar{\Gamma }}\) in \({\mathbb {R}}^3\) and a Beltrami field \(X\in C^\omega (\Omega ,{\mathbb {R}}^3)\) such that the normal component to \(\Gamma \) of \(X|_\Gamma \) equals g. To be precise, with N a unit normal vector field on \(\Gamma \), this means that \((X|_\Gamma )\cdot N= g\).

In order to show this, it suffices to observe that, given the normal component \(g\in C^\omega ({\bar{\Gamma }},{\mathbb {R}})\) of a Cauchy datum on \(\Gamma \), there are many Cauchy data W satisfying condition (\(*\)) and such that \(W\cdot N=g\). Endow \({\bar{\Gamma }}\simeq {\mathbb {I}}\times {\mathbb {R}}/{\mathbb {Z}}\) with coordinates \((z,\theta )\in {\mathbb {I}}\times {\mathbb {R}}/{\mathbb {Z}}\) so that the area form \(\sigma \) is \(dz\wedge d\theta \). We are looking for a Cauchy datum W such that

$$\begin{aligned} d(W^\flat _T)=gdz\wedge d\theta . \end{aligned}$$

Defining

$$\begin{aligned} {\tilde{g}}(z,\theta ):=\int _{-1}^zg(s,\theta )ds, \end{aligned}$$

we set

$$\begin{aligned} W^\flat _T:={\tilde{g}}(z,\theta )d\theta . \end{aligned}$$

Thus, it clearly holds that \(d(W^\flat _T)=g(z,\theta )dz\wedge d\theta \), that is W satisfies condition (\(*\)). By Theorem B.1 applied at \(\Gamma \), we conclude that there exists a neighborhood \(\Omega \) of \(\Gamma \) in \({\mathbb {R}}^3\) and an Euclidean Beltrami field \(X(g)\equiv X\in C^\omega (\Omega ,{\mathbb {R}}^3)\), defined as

$$\begin{aligned} X|_{\Gamma }:= gN+({{\tilde{g}}}(z,\theta )d\theta )^\#, \end{aligned}$$

whose normal component is g. \(\square \)

Remark B.2

Any other \(W^\flat _T\) satisfying condition (\(*\)) is of the form \({\tilde{g}}(z,\theta )d\theta +\alpha \), where \(\alpha \) is a closed 1-form on \(\Sigma \). In the following, we will always choose \(\alpha =0\), which implies the uniqueness of the Beltrami field with normal datum g.

Remark B.3

Let \(g\in C^\omega ({\bar{\Sigma }},{\mathbb {R}})\mapsto X(g)\in C^\omega (\Omega ,{\mathbb {R}}^3)\) be the map that associates to every normal datum g the corresponding Beltrami field X(g) given by Corollary 3.4. Such a map is linear. Indeed, let \(g_1,g_2\) be normal data and \(c_1,c_2\in {\mathbb {R}}\). Then, for \(i=1,2\), there exists a neighborhood \(\Omega _i\) of \(\Sigma \) and a Beltrami field \(X^i:=X(g_i)\in C^\omega (\Omega _i,{\mathbb {R}}^3)\) such that

$$\begin{aligned} X^i|_{\Sigma }=W^i_T+g_iN \end{aligned}$$

where \(W^i_T:=({\tilde{g}}_i(z,\theta )d\theta )^\#\). Thus, the Beltrami field \(X(c_1 g_1+c_2 g_2)\in C^\omega (\Omega _1\cap \Omega _2,{\mathbb {R}}^3)\) satisfies

$$\begin{aligned}{} & {} X(c_1 g_1+c_2 g_2)|_{\Sigma }=:X=\\{} & {} \quad (c_1{\tilde{g}}_1(z,\theta )d\theta )^\#+(c_2{\tilde{g}}_2(z,\theta )d\theta )^\#+(c_1 g_1+c_2 g_2)N=(c_1 W^1_T+c_2 W^2_T) \\{} & {} \quad +(c_1 g_1+c_2 g_2)N= c_1(W^1_T+g_1N)+c_2(W^2_T+g_2N)=c_1 X^1+c_2 X^2. \end{aligned}$$

Since the equation \(\textrm{curl }\, X=X\) is linear, by uniqueness (see Remark B.2) we have that X associated to the normal datum \(c_1g_1+c_2g_2\) is

$$\begin{aligned} X=c_1 X^1|_{\Omega _1\cap \Omega _2}+c_2 X^2|_{\Omega _1\cap \Omega _2}. \end{aligned}$$

1.2 Proof of Theorem B.1

Let us first consider a toy model to understand Theorem B.1 and Condition (\(*\)), following [EPS20, Subsection 3.2]. Let us consider \(\Sigma =\{x_3=0\}\) with coordinates \((x_1,x_2,x_3)\) in \({\mathbb {R}}^3\). Fix an analytic Cauchy datum

$$\begin{aligned} W=W_1(x_1,x_2)\partial _{x_1}+W_2(x_1,x_2)\partial _{x_2}+W_3(x_1,x_2)\partial _{x_3}. \end{aligned}$$

In coordinates, the Beltrami equation \(\textrm{curl}\ X= X\) reads as

figure a

By the standard Cauchy-Kovaleskaya’s Theorem applied at \(\Sigma =\{x_3=0\}\), there exists a unique analytic solution to this system in a neighborhood \(\Omega \) of \(\Sigma \), with Cauchy datum W.

Considering \(\frac{\partial }{\partial x_1}\)(B.4)\(-\frac{\partial }{\partial x_2}\)(B.3) and using (B.5), we obtain

$$\begin{aligned} \frac{\partial }{\partial x_3}\left( \frac{\partial X_2}{\partial x_1}-\frac{\partial X_1}{\partial x_2}- X_3\right) =0. \end{aligned}$$

Thus

$$\begin{aligned} \frac{\partial X_2}{\partial x_1}-\frac{\partial X_1}{\partial x_2}= X_3+f(x_1,x_2), \end{aligned}$$
(B.6)

in \(\Omega \), for some analytic function f. Evaluating (B.6) at \(\Sigma =\{x_3=0\}\), we obtain

$$\begin{aligned} \frac{\partial W_2}{\partial x_1}-\frac{\partial W_1}{\partial x_2}=W_3+f(x_1,x_2). \end{aligned}$$

The vector field \(X=(X_1,X_2,X_3)\) is Beltrami in \(\Omega \) if and only if the Cauchy datum satisfies the constraint

$$\begin{aligned} \frac{\partial W_2}{\partial x_1}-\frac{\partial W_1}{\partial x_2}=W_3. \end{aligned}$$
(B.7)

In terms of the dual 1-form \(\beta :=W_1dx_1+W_2dx_2\), this is equivalent to

$$\begin{aligned} d\beta =W_3dx_1\wedge dx_2. \end{aligned}$$
(B.8)

We have to choose the tangent part \(\beta \) of the Cauchy datum so that (B.8) is fullfilled. If \(\Sigma \) is open, given \(W_3\), this is always possible.

We now prove Theorem B.1, whose proof follows closely that of Theorem 3.1 in [EPS12]. Observe that Condition (\(*\)) reduces to the condition of [EPS12, Theorem 3.1] when \(g=0\).

Proof of Proposition B.1

Let \(\Sigma \subset {\mathbb {R}}^3\) be a bounded, oriented, analytic surface. Denote by

$$\begin{aligned} j_\Sigma :\Sigma \rightarrow {\mathbb {R}}^3 \end{aligned}$$

the analytic embedding of \(\Sigma \) into \({\mathbb {R}}^3\). Let W be an analytic vector field defined on \(\Sigma \). We will refer to it as Cauchy datum. We write W as

$$\begin{aligned} W=W_T+(W\cdot N) N, \end{aligned}$$

where N is the unit normal vector to \(\Sigma \), while \(W_T\) is the component of W tangent to \(\Sigma \). Let \(W^\flat \) be the dual 1-form of W, while \(W_T^\flat \) denotes the dual 1-form of \(W_T\). Let \(g:=W\cdot N\). Recall that the vector field W satisfies Condition (\(*\)) if

$$\begin{aligned} d W_T^\flat =g\sigma , \end{aligned}$$

where here d is the exterior derivative on \(\Sigma \) and \(\sigma \) is the induced area-form on \(\Sigma \).

Take local analytic coordinates in a neighborhood \(\Omega \) of \(\Sigma \):

$$\begin{aligned} (\rho ,\xi _1,\xi _2), \end{aligned}$$

where \(\rho \) is the signed distance function from \(\Sigma \): if \(\Omega \) is narrow enough, then this function is analytic. In these coordinates, the Euclidean metric reads as:

$$\begin{aligned} ds^2=d\rho ^2+h_{ij}(\rho ,\xi _1,\xi _2)d\xi _id\xi _j, \end{aligned}$$

where the last expression has to be understood as sum for \(i,j=1,2\). In the sequel, we will use such Einstein convention, i.e. repeated indices have to be understood as summed. Denote as \(h^{ij}\) the inverse matrix of \(h_{ij}\) and \(|h|:=\textrm{det}(h_{ij})\) its determinant.

The Beltrami field X reads in these coordinates as

$$\begin{aligned} X=a(\rho ,\xi _1,\xi _2)\partial _{\rho }+b^i(\rho ,\xi _1,\xi _2)\partial _{\xi _i} \end{aligned}$$

for some functions \(a(\rho ,\xi _1,\xi _2), b^1(\rho ,\xi _1,\xi _2),b^2(\rho ,\xi _1,\xi _2)\). The 1-form associated to X is so

$$\begin{aligned} \beta :=a(\rho ,\xi _1,\xi _2)d\rho +c_i(\rho ,\xi _1,\xi _2)d\xi _i, \end{aligned}$$

with \(c_i:=h_{ij}b^j\). In terms of the 1-form \(\beta \), the equation \(\textrm{curl}\,X=X\) reads as

$$\begin{aligned} \star d\beta = \beta , \end{aligned}$$
(B.9)

where \(\star \) denotes the Hodge star operator. In the coordinates \((\rho ,\xi _1,\xi _2)\), the term \(\star d\beta \) has the form

$$\begin{aligned}{} & {} \star d\beta = \dfrac{1}{|h|^{1/2}}\left( \dfrac{\partial c_2}{\partial \xi _1}-\dfrac{\partial c_1}{\partial \xi _2} \right) d\rho +|h|^{1/2}h^{2i}\left( \dfrac{\partial a}{\partial \xi _i} -\dfrac{\partial c_i}{\partial \rho }\right) d\xi _1+|h|^{1/2}h^{1i}\left( \dfrac{\partial c_i}{\partial \rho }-\dfrac{\partial a}{\partial \xi _i} \right) d\xi _2. \end{aligned}$$

The equation \(\star d\beta =\beta \) is then equivalent, in \((\rho ,\xi _1,\xi _2)\)-coordinates, to:

figure b

We start by showing the necessity of Condition (\(*\)). The tangential component of X on \(\Sigma \) is

$$\begin{aligned} X_T=b^i(0,\xi _1,\xi _2)\partial _{\xi _i}. \end{aligned}$$

Its dual 1-form is then \(X_T^\flat =c_i(0,\xi _1,\xi _2)d\xi _i\).

Thus

$$\begin{aligned} d(c_i(0,\xi _1,\xi _2)d\xi _i)=\left( \dfrac{\partial c_2(0,\xi _1,\xi _2)}{\partial \xi _1}-\dfrac{\partial c_1(0,\xi _1,\xi _2)}{\partial \xi _2} \right) d\xi _1\wedge d\xi _2. \end{aligned}$$

On the other hand the area-form \(\sigma \) on \(\Sigma \) is, since \(N=\partial _\rho \),

$$\begin{aligned} \sigma =|h|^{1/2}(0,\xi _1,\xi _2)d\xi _1\wedge d\xi _2. \end{aligned}$$

Evaluating then (B.10) at \(\rho =0\), i.e. at \(\Sigma \), we obtain

$$\begin{aligned} \dfrac{\partial c_2(0,\xi _1,\xi _2)}{\partial \xi _1}-\dfrac{\partial c_1(0,\xi _1,\xi _2)}{\partial \xi _2}=a(0,\xi _1,\xi _2)|h|^{1/2}(0,\xi _1,\xi _2), \end{aligned}$$

and, denoting \(g:=a(0,\xi _1,\xi _2)\), we obtain Condition (\(*\)):

$$\begin{aligned} d(X_T^\flat )=g\sigma . \end{aligned}$$

We now show the sufficiency of Condition (\(*\)). Consider the auxiliary problem:

$$\begin{aligned} {\left\{ \begin{array}{ll} (d+\delta )\psi =\star \psi \quad \text {in }\Omega \\ \psi |_\Sigma =W^\flat , \end{array}\right. } \end{aligned}$$
(A)

where \(W^\flat \) is the dual 1-form of W, \(\delta \) is the codifferential (i.e. for \(\omega \in \Omega ^k\) it holds \(\delta \omega =(-1)^{k}\star d\star \omega \)), and, denoting as \(\Omega ^k\) the space of k-forms, \(\psi \in \Omega ^0\oplus \Omega ^1\oplus \Omega ^2\oplus \Omega ^3\). In particular, \(\psi \) can be decomposed as \(\psi ^0\oplus \psi ^1\oplus \psi ^2\oplus \psi ^3\), with \(\psi ^k\) a k-form. The operator \(d+\delta \) is a first order elliptic operator (notice that \((d+\delta )^2=d\delta +\delta d=\Delta \) is the Hodge Laplacian). So, by the standard Cauchy–Kovalevskaya’s Theorem (see [Tay11, Page 502, Proposition 4.2]), the Cauchy problem (A) has a unique analytic solution in \(\Omega \), for some \(\Omega \) narrow enough.

Lemma B.4

If W satisfies Condition (\(*\)), then \(\delta \psi =0\) in \(\Omega \).

Proof of Lemma B.4

In [EPS12, Section 4] the authors define a parity operator

$$\begin{aligned} Q:\Omega ^0\oplus \Omega ^1\oplus \Omega ^2\oplus \Omega ^3\rightarrow \Omega ^0\oplus \Omega ^1\oplus \Omega ^2\oplus \Omega ^3 \end{aligned}$$

as

$$\begin{aligned} Q\psi :=\psi ^0\oplus -\psi ^1\oplus \psi ^2\oplus -\psi ^3, \end{aligned}$$

where \(\psi =\psi ^0\oplus \psi ^1\oplus \psi ^2\oplus \psi ^3\). First, since \(\star d\star =-Q\delta \), applying \(\star d\) at the equation (A) and since \(\star \star =\textrm{Id}\), we notice that \(\delta \psi \) satisfies the elliptic equation

$$\begin{aligned} (d+\delta )\delta \psi =-\star Q\delta \psi . \end{aligned}$$

Therefore, if we show that \(\delta \psi |_\Sigma =0\), then \(\delta \psi =0\) in \(\Omega \) by uniqueness of the solution by the Cauchy–Kovalevskaya’s Theorem, since the operator \(d+\delta \) is elliptic.

Our aim is then proving that \(\delta \psi |_\Sigma =0\). Write \(\psi \) as follows:

$$\begin{aligned} \psi =\Psi +d\rho \wedge {{\tilde{\Psi }}}, \end{aligned}$$

where \(\Psi ,{{\tilde{\Psi }}}\) are forms in \(\Omega ^0\oplus \Omega ^1\oplus \Omega ^2\oplus \Omega ^3\) such that \(i_{\partial _\rho }\Psi =i_{\partial _\rho }{{\tilde{\Psi }}}=0\). Since \(\psi |_\Sigma =W^\flat \), then

$$\begin{aligned} \Psi |_\Sigma =c_id\xi _i\quad \text {and}\quad {{\tilde{\Psi }}}|_\Sigma =g. \end{aligned}$$

From (A), it holds that

$$\begin{aligned} \delta \psi |_\Sigma =(\star \psi -d\psi )|_\Sigma . \end{aligned}$$
(B.13)

By the Cauchy condition, it holds that \(\star \psi |_\Sigma =\star W^\flat |_\Sigma \). Using that \(\star d\rho =|h|^{1/2}d\xi _1\wedge d\xi _2\), we deduce that

$$\begin{aligned} \star \psi |_\Sigma =\star W^\flat |_\Sigma =\star (gd\rho +c_1(0,\xi _1,\xi _2)d\xi _1+c_2(0,\xi _1,\xi _2)d\xi _2)= \end{aligned}$$
$$\begin{aligned} g|h|^{1/2}d\xi _1\wedge d\xi _2 +d\rho \wedge {{\tilde{\alpha }}}, \end{aligned}$$
(B.14)

for some 1-form \({{\tilde{\alpha }}}\). Denoting as \({\bar{d}}\) the exterior derivative only in the variables \(\xi _1,\xi _2\), we have

$$\begin{aligned} d\psi ={\bar{d}}\Psi +d\rho \wedge \dfrac{\partial \Psi }{\partial \rho }-d\rho \wedge {\bar{d}}{{\tilde{\Psi }}}={\bar{d}}\Psi +d\rho \wedge \left( \dfrac{\partial \Psi }{\partial \rho }-{\bar{d}}{{\tilde{\Psi }}} \right) . \end{aligned}$$

Since \(\Psi |_\Sigma =c_1(0,\xi _1,\xi _2)d\xi _1+c_2(0,\xi _1,\xi _2)d\xi _2\), it holds, using Condition (\(*\)),

$$\begin{aligned} {\bar{d}}\Psi |_\Sigma =g|h|^{1/2}d\xi _1\wedge d\xi _2. \end{aligned}$$

Consequently

$$\begin{aligned} d\psi |_\Sigma =g|h|^{1/2}d\xi _1\wedge d\xi _2+d\rho \wedge {\hat{\alpha }}, \end{aligned}$$
(B.15)

for some \({\hat{\alpha }}={\hat{\alpha ^{0}}}\oplus {\hat{\alpha ^{1}}}\oplus {\hat{\alpha ^{2}}}\) on \(\Sigma \). Putting together (B.13), (B.14) and (B.15), we deduce that

$$\begin{aligned} \delta \psi |_\Sigma =d\rho \wedge {{\check{\alpha }}}, \end{aligned}$$

for some \({{\check{\alpha }}}={{\tilde{\alpha }}}-{\hat{\alpha }}={{\check{\alpha }}}^0\oplus {{\check{\alpha }}}^1\oplus {{\check{\alpha }}}^2\) on \(\Sigma \). Noticing that \(\psi |_\Sigma =W^\flat \) is a 1-form, we obtain:

  1. (i)

    Since \(\delta \psi |_\Sigma =d\rho \wedge {{\check{\alpha }}}\) does not contain 0-forms, the components \(\delta \psi ^0|_\Sigma \) and \(\delta \psi ^1|_\Sigma \) are both zero;

  2. (ii)

    For \(p=2,3\) we have

    $$\begin{aligned} \delta \psi ^p|_\Sigma =(-1)^p(\star d\star \psi ^p)|_\Sigma =(-1)^p\star (d\rho \wedge \beta )\end{aligned}$$

    for some form \(\beta \). This comes from the fact that

    $$\begin{aligned} d(\star \psi ^2)|_\Sigma =d\rho \wedge \partial _\rho (\star \psi ^2)|_{\rho =0}+{\bar{d}}(\star \psi ^2)|_{\rho =0}, \end{aligned}$$

    but since \(\psi ^2|_{\rho =0}=0\), then \({\bar{d}}(\star \psi ^2)|_{\rho =0}={\bar{d}}(\star \psi ^2|_{\rho =0})=0\). Similarly, since \(\psi ^3|_\Sigma =0\), it holds that

    $$\begin{aligned} d(\star \psi ^3)|_\Sigma =d\rho \wedge \partial _\rho (\star \psi ^3)|_{\rho =0}. \end{aligned}$$

    Therefore, for \(p=2,3\), \(\delta \psi ^p|_\Sigma \) only contains elements with \(d\xi _1\) and \(d\xi _2\), for \(p=2\), and only \(d\xi _1\wedge d\xi _2\) for \(p=3\).

The fact that \(\delta \psi |_\Sigma \) does not contain the element \(d\rho \) contradicts the expression \(\delta \psi |_\Sigma =d\rho \wedge {{\check{\alpha }}}\) previously found, unless

$$\begin{aligned} \delta \psi |_\Sigma =0, \end{aligned}$$

as we wanted to prove. This complete the proof of Lemma B.4. \(\square \)

Going back to the problem (A), since \(\delta \psi =0\) in \(\Omega \), we have that

$$\begin{aligned} {\left\{ \begin{array}{ll} d\psi =\star \psi \quad \text {in }\Omega \\ \psi |_\Sigma =W^\flat \end{array}\right. } \end{aligned}$$

and, using the identity \(\star (\star \psi )=\psi \), it holds

$$\begin{aligned} {\left\{ \begin{array}{ll} \star d\psi =\psi \quad \text {in }\Omega \\ \psi |_\Sigma =W^\flat . \end{array}\right. } \end{aligned}$$

Since \(\star d\) maps 1-forms into 1-forms and \(W^\flat \) is a 1-form, it holds \(\psi ^0=\psi ^2=\psi ^3=0\). Hence

$$\begin{aligned} {\left\{ \begin{array}{ll} \star d\psi ^1=\psi ^1\quad \text {in }\Omega \\ \psi ^1|_\Sigma =W^\flat . \end{array}\right. } \end{aligned}$$

Taking the vector field X dual to \(\psi ^1\), it satisfies then \(\textrm{curl}\,X=X\) and the sufficiency follows. This concludes the proof of the sufficiency and so of Theorem B.1. \(\square \)

Disjoint Union of Compact Subsets with Connected Complement

The following result is used in the proof of Corollary 3.2. It is probably standard, but we include a proof for the sake of completeness.

Proposition C.1

Let \(K_+\) and \(K_-\) be two disjoint compact subsets of \({\mathbb {R}}^n \) with connected complement. Then \(K=K_+\sqcup K_-\) has connected complement.

Proof

Since K is compact, it is clear that \({\mathbb {R}}^n\backslash K\) is connected if and only if \({\mathbb {S}}^n\backslash K\) is connected, where the n-sphere \({\mathbb {S}}^n\) is understood as the one-point compactification of \({\mathbb {R}}^n\). Alexander’s duality then implies [Mas78] that

$$\begin{aligned} {{\tilde{H}}}_0({\mathbb {S}}^n\backslash K;{\mathbb {Z}})={{\tilde{H}}}^{n-1}_{CE}(K;{\mathbb {Z}})\,, \end{aligned}$$

where \({{\tilde{H}}}_0({\mathbb {S}}^n\backslash K;{\mathbb {Z}})\) stands for the reduced singular homology of \({\mathbb {S}}^n\backslash K\) with integer coefficients (well defined because it is a manifold) and \({{\tilde{H}}}^{n-1}_{CE}(K;{\mathbb {Z}})\) denotes the reduced Čech cohomology, which is defined for any compact subset of \({\mathbb {R}}^n\). Now, since the reduced k-th cohomology group coincides with the cohomology group for \(k\ge 1\) (this also holds for Čech cohomology, for which one can also write a long exact sequence [ES52]) and \(K_+\) and \(K_-\) are disjoint, we infer that

$$\begin{aligned} {{\tilde{H}}}^{n-1}_{CE}(K;{\mathbb {Z}})=H^{n-1}_{CE}(K;{\mathbb {Z}})=H^{n-1}_{CE}(K_+;{\mathbb {Z}})\oplus H^{n-1}_{CE}(K_-;{\mathbb {Z}})\,, \end{aligned}$$

where to write the last isomorphism we have used the standard Mayer-Vietoris sequence for Čech cohomology [Mil16]. Then, using again Alexander’s duality and that \(H^{n-1}_{CE}(K_{\pm };{\mathbb {Z}})={{\tilde{H}}}^{n-1}_{CE}(K_{\pm };{\mathbb {Z}})\), we obtain that

$$\begin{aligned} H^{n-1}_{CE}(K_+;{\mathbb {Z}})\oplus H^{n-1}_{CE}(K_-;{\mathbb {Z}})={{\tilde{H}}}_0({\mathbb {S}}^n\backslash K_+;{\mathbb {Z}})\oplus {{\tilde{H}}}_0({\mathbb {S}}^n\backslash K_-;{\mathbb {Z}})=0 \end{aligned}$$

because \({\mathbb {S}}^n\backslash K_{\pm }\) are connected by assumption, and the 0-th reduced singular homology group of a manifold is trivial if and only if the manifold is connected. Putting all these computations together we finally conclude that

$$\begin{aligned} {{\tilde{H}}}_0({\mathbb {S}}^n\backslash K;{\mathbb {Z}})=0\,, \end{aligned}$$

so \({\mathbb {S}}^n\backslash K\) is connected, as we wanted to prove. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Berger, P., Florio, A. & Peralta-Salas, D. Steady Euler Flows on \({\mathbb {R}}^3\) with Wild and Universal Dynamics. Commun. Math. Phys. 401, 937–983 (2023). https://doi.org/10.1007/s00220-023-04660-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-023-04660-6

Navigation