Skip to main content
Log in

Reeb flows, pseudo-holomorphic curves and transverse foliations

  • Special issue commemorating the Golden Jubilee of the Institute of Mathematics and Statistics of the University of São Paulo
  • Published:
São Paulo Journal of Mathematical Sciences Aims and scope Submit manuscript

Abstract

Pseudo-holomorphic curves in symplectizations, as introduced by Hofer in 1993, and then developed by Hofer, Wysocki, and Zehnder, have brought new insights to Hamiltonian dynamics, providing new approaches to some classical questions in Celestial Mechanics. This short survey presents some recent developments in Reeb dynamics based on the theory of pseudo-holomorphic curves in symplectizations, focusing on transverse foliations near critical energy surfaces.

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
Fig. 6

Similar content being viewed by others

References

  1. Abbas, C.: An introduction to compactness results in symplectic field theory. Springer, Berlin (2014)

    Book  Google Scholar 

  2. Albers, P., Frauenfelder, U., van Koert, O., Paternain, G.P.: Contact geometry of the restricted three-body problem. Commun. Pure Appl. Math. 65(2), 229–263 (2012)

    Article  MathSciNet  Google Scholar 

  3. Birkhoff, G.: Dynamical Systems, vol. 9. American Mathematical Society, New York (1927)

    MATH  Google Scholar 

  4. Bourgeois, F., Eliashberg, Y., Hofer, H., Wysocki, K., Zehnder, E.: Compactness results in symplectic field theory. Geom. Topol. 7(2), 799–888 (2003)

    Article  MathSciNet  Google Scholar 

  5. Colin, V., Dehornoy, P., Rechtman, A.: On the existence of supporting broken book decompositions for contact forms in dimension 3, arXiv:2001.01448

  6. Conley, C., Zehnder, E.: The Birkhoff–Lewis fixed point theorem and a conjecture of VI Arnold. Invent. Math. 73(1), 33–49 (1983)

    Article  MathSciNet  Google Scholar 

  7. de Paulo, N.V., Salomão, P.A.S.: On the multiplicity of periodic orbits and homoclinics near critical energy levels of Hamiltonian systems in \({\mathbb{R}}^4\). Trans. Am. Math. Soc. 372(2), 859–887 (2019)

  8. de Paulo, N.V., Salomão, P.A.S.: Systems of transversal sections near critical energy levels of Hamiltonians systems in \({\mathbb{R}}^4\). Mem. Am. Math. Soc. 252, 1202 (2018) 1–105

  9. Fish, J.W., Siefring, R.: Connected sums and finite energy foliations I: Contact connected sums. J. Symplectic Geom. 16(6), 1639–1748 (2018)

    Article  MathSciNet  Google Scholar 

  10. Floer, A.: Morse theory for Lagrangian intersections. J. Differ. Geom. 28(3), 513–547 (1988)

    Article  MathSciNet  Google Scholar 

  11. Franks, J.: Area preserving homeomorphisms of open surfaces of genus zero. N. Y. J. Math. 2(1), 19 (1996)

    MathSciNet  MATH  Google Scholar 

  12. Gromov, M.: Pseudo holomorphic curves in symplectic manifolds. Invent. Math. 82(2), 307–347 (1985)

    Article  MathSciNet  Google Scholar 

  13. Grotta-Ragazzo, C., Salomão, P.A.S.: The Conley–Zehnder index and the saddle-center equilibrium. J. Differ. Equ. 220(1), 259–278 (2006)

    Article  MathSciNet  Google Scholar 

  14. Grotta-Ragazzo, C., Salomão, P.A.S.: Global surfaces of section in non-regular convex energy levels of Hamiltonian systems. Math. Z. 255(2), 323–334 (2007)

    Article  MathSciNet  Google Scholar 

  15. Hofer, H., Wysocki, K., Zehnder, E.: Properties of pseudoholomorphic curves in symplectisations I: Asymptotics. Ann. Inst. H. Poincaré Anal. NonLinéaire 13, 337–379 (1996)

  16. Hofer, H., Wysocki, K., Zehnder, E.: Properties of pseudoholomorphic curves in symplectizations III: Fredholm theory. Topics in nonlinear analysis. Birkhäuser, Basel, pp. 381–475 (1999)

  17. Hofer, H.: Pseudoholomorphic curves in symplectisations with applications to the Weinstein conjecture in dimension three. Invent. Math. 114, 515–563 (1993)

    Article  MathSciNet  Google Scholar 

  18. Hofer, H., Wysocki, K., Zehnder, E.: A characterization of the tight three sphere. Duke Math. J. 81(1), 159–226 (1995)

    Article  MathSciNet  Google Scholar 

  19. Hofer, H., Wysocki, K., Zehnder, E.: Properties of pseudoholomorphic curves in symplectisations II: Embedding controls and algebraic invariants. Geom. Funct. Anal. 5(2), 270–328 (1995)

    Article  MathSciNet  Google Scholar 

  20. Hofer, H., Wysocki, K., Zehnder, E.: The dynamics of strictly convex energy surfaces in \({\mathbb{R}}^4\). Ann. Math. 148, 197–289 (1998)

    Article  MathSciNet  Google Scholar 

  21. Hofer, H., Wysocki, K., Zehnder, E.: A characterization of the tight three sphere II. Commun. Pure Appl. Math. 55(9), 1139–1177 (1999)

    Article  Google Scholar 

  22. Hofer, H., Wysocki, K., Zehnder, E.: Finite energy foliations of tight three-spheres and Hamiltonian dynamics. Ann. Math. 157, 125–255 (2003)

    Article  MathSciNet  Google Scholar 

  23. Hryniewicz, U., Salomão, P.A.S., Siefring, R.: Global surfaces of section with positive genus for dynamically convex Reeb flows. J. Fixed Point Theory Appl. (to appear)

  24. Hryniewicz, U., Salomão, P.A.S., Wysocki, K.: Genus zero global surfaces of section for Reeb flows and a result of Birkhoff. arXiv: 1912.01078

  25. Hryniewicz, U.: Systems of global surfaces of section for dynamically convex Reeb flows on the \(3\)-sphere. J. Symplectic Geom. 12, 4 (2014)

    Article  MathSciNet  Google Scholar 

  26. Hryniewicz, U., Salomão, P.A.S.: On the existence of disk-like global sections for Reeb flows on the tight 3-sphere. Duke Math. J. 160(3), 415–465 (2011)

    Article  MathSciNet  Google Scholar 

  27. Hryniewicz, U., Salomão, P.A.S.: Elliptic bindings for dynamically convex Reeb flows on the real projective three-space. Calc. Var. Part. Differ. Equ. 55(2), 43 (2016)

    Article  MathSciNet  Google Scholar 

  28. Hryniewicz, U., Licata, J., Salomão, P.A.S.: A dynamical characterization of universally tight lens spaces. Proc. Lond. Math. Soc. 110, 213–269 (2014)

    Article  MathSciNet  Google Scholar 

  29. Kim, J., Kim, Y., van Koert, O.: Reeb flows without simple global surfaces of section. arXiv:2104.03728

  30. Kim, S.: On a convex embedding of the Euler problem of two fixed centers. Regul. Chaotic Dyn. 23(3), 304–324 (2018)

    Article  MathSciNet  Google Scholar 

  31. Lemos de Oliveira, C.: 3-2-1 foliations for Reeb flows on \(S^3\). PhD thesis, University of São Paulo (2020)

  32. Lemos de Oliveira, C.: 3-2-1 foliations for Reeb flows on the tight 3-sphere. Preprint arXiv:2104.10295 (2021)

  33. Poincaré, H.: Sur un théorème de géométrie. Rend. Circ. Mat. Palermo 33, 375–407 (1912)

    Article  Google Scholar 

  34. Salomão, P.A.S.: Convex energy levels of Hamiltonian systems. Qual. Theory Dyn. Syst. 4(2), 439–454 (2004)

    Article  MathSciNet  Google Scholar 

  35. Schneider, A.: Global surfaces of section for dynamically convex Reeb flows on lens spaces. Trans. Am. Math. Soc. 373(4), 2775–2803 (2020)

    Article  MathSciNet  Google Scholar 

  36. Siefring, R.: Finite-energy pseudoholomorphic planes with multiple asymptotic limits. Math. Ann. 368(1), 367–390 (2017)

    Article  MathSciNet  Google Scholar 

  37. van Koert, O.: A Reeb flow on the three-sphere without a disk-like global surface of section. Qual. Theory Dyn. Syst. 19(1), 36 (2020)

    Article  MathSciNet  Google Scholar 

  38. Wendl, C.: Finite energy foliations on overtwisted contact manifolds. Geom. Topol. 12(1), 531–616 (2008)

    Article  MathSciNet  Google Scholar 

  39. Wendl, C.: Automatic transversality and orbifolds of punctured holomorphic curves in dimension four. Comment. Mathe. Helvet. 85(2), 347–407 (2010)

    Article  MathSciNet  Google Scholar 

  40. Wendl, C.: Open book decompositions and stable Hamiltonian structures. Expo. Math. 28(2), 187–199 (2010)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

P. Salomão acknowledges the support of NYU-ECNU Institute of Mathematical Sciences at NYU Shanghai. P. Salomão is partially supported by FAPESP 2016/25053-8 and CNPq 306106/2016-7.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pedro A. S. Salomão.

Additional information

Communicated by Claudio Gorodski.

Publisher's Note

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

On the occasion of the Golden Jubilee Celebration of IME-USP.

Appendices

Appendix A: The action functional

Let M be a smooth 3-manifold equipped with a contact form \(\lambda\). Denote by \(\varphi _t,t\in {\mathbb {R}},\) the Reeb flow of \(\lambda\).

The action functional associated with \(\lambda\) is defined on the space of smooth curves \(\gamma : {\mathbb {R}}/ {\mathbb {Z}}\rightarrow M\) as

$$\begin{aligned} {\mathcal {A}}(\gamma ) = \int _{{\mathbb {R}}/ {\mathbb {Z}}} \gamma ^*\lambda , \quad \forall \gamma \in C^\infty ({\mathbb {R}}/ {\mathbb {Z}},M). \end{aligned}$$

Let \(\eta \in \Gamma (\gamma ^*TM)\) be a vector field along \(\gamma\) and let \(u:(-\epsilon ,\epsilon )\times {\mathbb {R}}/ {\mathbb {Z}}\rightarrow M\) be a variation of \(\gamma\) so that \(u(0,t)=\gamma (t)\) for all \(t\in {\mathbb {R}}/ {\mathbb {Z}}\) and \(u_s(0,\cdot )=\eta\). We compute the first variation of \({\mathcal {A}}\) at \(\gamma\) in the direction of \(\eta\)

$$\begin{aligned} \begin{aligned} d{\mathcal {A}}(\gamma ) \cdot \eta&= \int _{0 \times {\mathbb {R}}/ {\mathbb {Z}}} {\mathcal {L}}_{\partial _s} (u^*\lambda )\\&= \int _{0 \times {\mathbb {R}}/ {\mathbb {Z}}} d i_{\partial _s} (u^*\lambda )+\int _{0 \times {\mathbb {R}}/ {\mathbb {Z}}} i_{\partial _s} (u^*d\lambda )\\&= \int _0^1 d\lambda |_{\gamma (t)}(\eta (t), {{\dot{\gamma }}}(t))dt. \end{aligned} \end{aligned}$$
(6)

Since \(d\lambda |_{\xi =\ker \lambda }\) is nondegenerate, this implies that \(\gamma\) is a critical point of \({\mathcal {A}}\) if and only if \({{\dot{\gamma }}} \subset \ker d\lambda |_\gamma .\) In particular, if \(\dot{\gamma }\) never vanishes, then \(\gamma\) is a critical point of \({\mathcal {A}}\) if and only if \(\gamma\) is a reparametrization of a closed Reeb orbit.

Now let \(P=(x,T)\) be a closed Reeb orbit of \(\lambda\). Then \(x_T = x(T\cdot ):{\mathbb {R}}/ {\mathbb {Z}}\rightarrow M\) is a critical point of \({\mathcal {A}}\). Let \(\eta \in \Gamma (x_T^*\xi )\) and let \(u:(-\epsilon ,\epsilon ) \times {\mathbb {R}}/ {\mathbb {Z}}\rightarrow M\) be a variation of \(x_T\) so that \(u_s(0,\cdot )=\eta\). We can assume that \(u_s(s,\cdot )\subset \xi , \forall s\). Using a finite covering of a tubular neighborhood of \(x_T({\mathbb {R}})\), we may assume that \(x_T\) is an embedding. Let \(\Xi\) be a vector field extending \(\eta\) in a neighborhood of \(x({\mathbb {R}})\).

The second variation of \({\mathcal {A}}\) at \(x_T\) in the direction of \(\eta\) is

$$\begin{aligned} \begin{aligned} d^2{\mathcal {A}}(x_T) \cdot (\eta ,\eta )&= \int _{0 \times {\mathbb {R}}/ {\mathbb {Z}}} {\mathcal {L}}_{\partial _s}({\mathcal {L}}_{\partial _s} (u^*\lambda ))\\&= \int _{0 \times {\mathbb {R}}/ {\mathbb {Z}}} i_{\partial _s}d (i_{\partial _s} (u^*d\lambda )))\\&= \int _{ {\mathbb {R}}/ {\mathbb {Z}}} x_T^* (i_\Xi d(i_\Xi d\lambda ))\\&= \int _0^1 d(i_\eta d\lambda )|_{x_T(t)}(\eta (t), \dot{x}_T(t))dt\\&= \int _0^1 d\lambda |_{x_T(t)} ( -({\mathcal {L}}_{\dot{x}_T} \eta )(t),\eta (t)) \, dt.\\ \end{aligned} \end{aligned}$$
(7)

We have used the usual relation \(d\alpha (X,Y)=X \cdot \alpha (Y) - Y \cdot \alpha (X) - \alpha ([X,Y])\), where XY are vector fields and \(\alpha\) is a 1-form.

Observe that the last expression obtained in (7) depends only on the linearized flow along P

$$\begin{aligned} \begin{aligned} ({\mathcal {L}}_{\dot{x}_T} \eta )(t)&: = \frac{d}{ds}\big |_{s=0} \left\{ D\varphi ^{-1}_{sT}(x_T(t+s)) \cdot \eta (t+s) \right\} . \end{aligned} \end{aligned}$$

A similar analysis of the action functional can be found in [32].

Appendix B: The asymptotic operator

Let \(P=(x,T)\) be a closed orbit of the Reeb flow of \(\lambda\) and let \(J:\xi \rightarrow \xi\) be a \(d\lambda\)-compatible complex structure. The asymptotic operator associated with P and J is the linear operator

$$\begin{aligned} \begin{aligned} A_{P,J}(\eta ) := -J|_{x_T} \cdot {\mathcal {L}}_{\dot{x}_T}\eta \in L^2(x_T^* \xi ), \quad \forall \eta \in W^{1,2}(x_T^*\xi ), \\ \end{aligned} \end{aligned}$$

where \({\mathcal {L}}_{\dot{x}_T} \eta\) is defined in the previous section.

It is an exercise to check that \(A_{P,J}\) admits 0 as an eigenvalue if and only if P is degenerate.

Since J preserves \(d\lambda\) we see from (7) that

$$\begin{aligned} d^2{\mathcal {A}}(x_T) \cdot (\eta ,\eta ) = \int _0^1 d\lambda |_{x_T(t)} (A_{P,J} \cdot \eta (t), J_t\cdot \eta (t))\, dt, \end{aligned}$$

where \(J_t = J|_{x_T(t)}\).

If \(\nu \in {\mathbb {R}}\) is an eigenvalue of \(A_{P,J}\) and \(\eta _\nu\) is a non-trivial \(\nu\)-eigenfunction, then

$$\begin{aligned} \begin{aligned} d^2{\mathcal {A}}(x_T) \cdot (\eta _\nu ,\eta _\nu )&= \int _0^1 d\lambda _{x_T(t)}(A_{P,J} \cdot \eta _\nu (t), J_t\cdot \eta _\nu (t))\, dt,\\&= \nu \int _0^1 g_{J_t}(\eta _\nu (t),\eta _{\nu }(t))\, dt, \end{aligned} \end{aligned}$$

where \(g_J(\cdot , \cdot ) = d\lambda (\cdot , J \cdot )\) is the positive-definite inner product on \(\xi\) induced by \(d\lambda\) and J. Hence \(d^2{\mathcal {A}}(x_T) \cdot (\eta _\nu ,\eta _\nu )\) vanishes if and only if \(\nu =0\) and, otherwise, its sign coincides with the sign of \(\nu\).

In order to better describe the spectrum of the operator \(A_{P,J}\) we need some normalization. Choose a unitary trivialization \(\Psi : x_T^* \xi \rightarrow {\mathbb {R}}/ {\mathbb {Z}}\times {\mathbb {R}}^2\). This means that

$$\begin{aligned} \Psi ^*(dx \wedge dy) = d\lambda |_\xi \quad {\text{ and }} \quad \Psi _* J = J_0:=\left( \begin{array}{cc}0 &{} -1 \\ 1 &{} 0 \end{array}\right) . \end{aligned}$$

Here (xy) are coordinates in \({\mathbb {R}}^2\).

We claim that the operator

$$\begin{aligned} {\mathcal {L}}_S:= \Psi \circ A_{P,J} \circ \Psi ^{-1} \end{aligned}$$

has the form

$$\begin{aligned} {\mathcal {L}}_S= -J_0\frac{d}{dt} - S, \end{aligned}$$
(8)

for some smooth loop \(t \mapsto S(t)\) of \(2\times 2\) symmetric matrices.

Indeed, recall that the Reeb flow of \(\lambda\) preserves \(d\lambda |_\xi\) and thus we find a smooth path of \(2\times 2\) symplectic matrices \(\Phi (t)\), satisfying \(\Phi (0)=I\) and

$$\begin{aligned} \Phi (1+t) = \Phi (t) \Phi (1), \quad \forall t\in {\mathbb {R}}/ {\mathbb {Z}}. \end{aligned}$$
(9)

It represents the linearized flow of \(\frac{1}{T}\lambda\) restricted to \((\xi ,d\lambda |_\xi )\) along \(x_T\) in coordinates \(\Psi\). In particular, a solution \({\bar{\zeta }}(t)\in {\mathbb {R}}^2 \simeq \xi |_{x_T(t)}\) to the linearized flow with initial condition \(\bar{\zeta }(t+s)=\zeta (t+s) \in {\mathbb {R}}^2 \simeq \xi _{x_T(t+s)}\) satisfies

$$\begin{aligned} {\bar{\zeta }}(t) = \Phi (t)\Phi (t+s)^{-1}\zeta (t+s), \quad \forall s,t. \end{aligned}$$

Note that \({\bar{\zeta }}(t)\) depends on s.

Denoting

$$\begin{aligned} \zeta (t) = \Psi \circ \eta (t)\in {\mathbb {R}}^2 \quad \forall t, \end{aligned}$$

let us compute \({\mathcal {L}}_{\dot{x}_T} \eta\) in coordinates \(\Psi\)

$$\begin{aligned} \begin{aligned} \Psi \circ {\mathcal {L}}_{\dot{x}_T}\eta (t)&= \lim _{s\rightarrow 0} \frac{{\bar{\zeta }}(t) - \zeta (t)}{s}\\&= \lim _{s\rightarrow 0} \frac{\Phi (t)\Phi (t+s)^{-1}\zeta (t+s) - \zeta (t)}{s}\\&= \lim _{s\rightarrow 0} \left\{ \Phi (t)\Phi (t+s)^{-1}\frac{\zeta (t+s)-\zeta (t)}{s}+\Phi (t)\frac{\Phi (t+s)^{-1}-\Phi (t)^{-1}}{s}\zeta (t)\right\} \\&= {{\dot{\zeta }}}(t) +\Phi (t)\dot{[\Phi (t)^{-1}]}\zeta (t)\\&= {{\dot{\zeta }}}(t) - {{\dot{\Phi }}}(t) \Phi (t)^{-1}\zeta (t) \end{aligned} \end{aligned}$$
(10)

Now let

$$\begin{aligned} S(t):= -J_0 {{\dot{\Phi }}}(t) \Phi (t)^{-1}. \end{aligned}$$
(11)

Since \(\Phi (t)\) is symplectic we have

$$\begin{aligned} \Phi (t)^T J_0 \Phi (t) = J_0 \Rightarrow {{\dot{\Phi }}}(t)^T J_0 \Phi (t) + \Phi (t)^T J_0 {{\dot{\Phi }}}(t) = 0, \quad \forall t. \end{aligned}$$

Hence

$$\begin{aligned} {{\dot{\Phi }}}(t)^TJ_0 = \Phi (t)^T S(t). \end{aligned}$$
(12)

Using the definition of S we compute

$$\begin{aligned} S(t)^T = (\Phi (t)^T)^{-1}{{\dot{\Phi }}}(t)^T J_0 \Rightarrow \dot{\Phi }(t)^T J_0 = \Phi (t)^T S(t)^T. \end{aligned}$$

It follows from (12) that \(S(t)^T=S(t), \forall t\). The reader can easily check that (9) implies \(S(t+1)=S(t) \forall t.\) Finally, (8) follows from (10), (11) and the identity

$$\begin{aligned} {\mathcal {L}}_S = -J_0 \cdot \Psi \circ {\mathcal {L}}_{\dot{x}_T} \circ \Psi ^{-1}. \end{aligned}$$

We check the symmetry of \(A_{P,J}\) using coordinates \(\Psi\). The \(L^2\)-product on \(\Gamma (x_T^*\xi )\) induced by \(d\lambda\) and J takes the form

$$\begin{aligned} \langle \zeta _1,\zeta _2\rangle = \int _0^1 \langle \zeta _1(t),\zeta _2(t) \rangle \, dt, \quad \forall \zeta _1,\zeta _2\in L^2({\mathbb {R}}/ {\mathbb {Z}},{\mathbb {R}}^2). \end{aligned}$$

Integrating by parts, we obtain

$$\begin{aligned} \begin{aligned} \langle \zeta _1,{\mathcal {L}}_S\cdot \zeta _2\rangle&= \int _0^1 \langle \zeta _1(t),-J_0{{\dot{\zeta }}}_2(t) - S(t)\zeta _2(t)\rangle \, dt\\&= \int _0^1 \langle -J_0{{\dot{\zeta }}}_1(t) - S(t)\zeta _1(t),\zeta _2(t)\rangle \, dt\\&= \langle {\mathcal {L}}_S \cdot \zeta _1,\zeta _2\rangle . \end{aligned} \end{aligned}$$

The eigenvalues of \({\mathcal {L}}_S\) are real. A non-trivial \(\nu\)-eigenfunction \(\zeta _\nu\) of \({\mathcal {L}}_S\) satisfies the smooth linear ODE

$$\begin{aligned} {{\dot{\zeta }}}_\nu (t)= J_0 (S(t)+\nu I) \zeta _\nu (t) , \quad \forall t, \end{aligned}$$

and thus \(\zeta _\nu\) is smooth and never vanishes. In particular, \(\zeta _\nu\) has a well-defined winding number

$$\begin{aligned} \mathrm{wind}(\nu ) = \frac{\Theta (1) - \Theta (0)}{2\pi } \end{aligned}$$

where \(\zeta _\nu (t) = (r(t)\cos (\Theta (t)),r(t) \sin (\Theta (t))) \forall t,\) for continuous functions \(r>0,\Theta\). It does not depend on the \(\nu\)-eigenfunction.

If the linearized first return map \(D\varphi _T(x_T(0)):\xi |_{x_T(0)} \rightarrow \xi _{x_T(0)}\) is the identity map then it is always possible to choose \(J\in {\mathcal {J}}_+(\xi )\) and a trivialization \(\Psi\) so that \(S \equiv 0\). In this case, the eigenvalues of \({\mathcal {L}}_0=-J_0\frac{d}{dt}\) are

$$\begin{aligned} \nu _k = 2\pi k, k \in {\mathbb {Z}}. \end{aligned}$$

Each \(\nu _k\) admits a 2-dimensional eigenspace

$$\begin{aligned} \{t\mapsto A(\cos (2 \pi kt),\sin (2\pi kt)) + B(\sin (2\pi kt),-\cos (2\pi kt)), t\in {\mathbb {R}}/ {\mathbb {Z}},A,B\in {\mathbb {R}}\}, \end{aligned}$$

whose eigenfunctions have winding number k.

The next theorem asserts that the general case is similar.

Theorem B.1

([19]) Let \(t\mapsto S(t)\) be a smooth loop of \(2\times 2\) symmetric matrices and let \({\mathcal {L}}_S\) be defined as in (8). Then

  1. (i)

    The spectrum \(\sigma ({\mathcal {L}}_S)\) consists of real eigenvalues which accumulate precisely at \(\pm \infty\).

  2. (ii)

    The winding number \(\mathrm{wind}(\nu )\in {\mathbb {Z}},\nu \in \sigma ({\mathcal {L}}_S),\) is independent of the \(\nu\)-eigenfunction.

  3. (iii)

    The map

    $$\begin{aligned} \nu \mapsto \mathrm{wind}(\nu )\in {\mathbb {Z}}, \end{aligned}$$

    is a surjective increasing map. For every \(k\in {\mathbb {Z}}\), there exist precisely two eigenvalues \(\nu _k^1,\nu _k^2\), counting multiplicities, so that

    $$\begin{aligned} \mathrm{wind}(\nu _k^1)=\mathrm{wind}(\nu _k^2)=k. \end{aligned}$$
  4. (iv)

    If \(\nu _1\ne \nu _2\) satisfy \(\mathrm{wind}(\nu _1)=\mathrm{wind}(\nu _2)\), then any two \(\nu _1,\nu _2\)-eigenfunctions are pointwise linearly independent.

Appendix C: The generalized Conley–Zehnder index

Let \(P=(x,T)\) be a closed Reeb orbit of \(\lambda\) and let \(\tau\) be a unitary trivialization of \(x_T^*\xi\). Let \({\mathcal {L}}_S\) be the operator defined in (8). Let

$$\begin{aligned} \begin{aligned} \nu ^\tau _-(P):= \max \{ \nu : \nu \in \sigma ({\mathcal {L}}_S)\cap (-\infty ,0)\},\\ \nu ^\tau _+(P):= \min \{ \nu : \nu \in \sigma ({\mathcal {L}}_S)\cap [0,+\infty )\}, \end{aligned} \end{aligned}$$

and let

$$\begin{aligned} \mathrm{wind}_+^\tau (P):= \mathrm{wind}(\nu ^\tau _+(P)) \quad {\text{ and }} \quad \mathrm{wind}^\tau _-(P):= \mathrm{wind}(\nu ^\tau _-(P)) . \end{aligned}$$

These winding numbers satisfy \(0\le \mathrm{wind}^\tau _+(P)-\mathrm{wind}^\tau _-(P)\le 1\) and they do not depend on J.

Definition C.1

( [20]) The generalized Conley–Zehnder index of \(P=(x,T)\) with respect to \(\tau\) is defined as

$$\begin{aligned} \mathrm{CZ}^\tau (P):=\mathrm{wind}_+^\tau (P)+\mathrm{wind}_-^\tau (P). \end{aligned}$$

This definition immediately implies that if \(\nu \in \sigma (A_{P,J})\), then

$$\begin{aligned} \begin{aligned} \nu < 0 \ \Rightarrow \mathrm{wind}^\tau (\nu ) \le \frac{\mathrm{CZ}^\tau (P)}{2},\\ \nu \ge 0 \ \Rightarrow \mathrm{wind}^\tau (\nu ) \ge \frac{\mathrm{CZ}^\tau (P)}{2}. \end{aligned} \end{aligned}$$

A more geometric definition of \(\mathrm{CZ}^\tau\) is as follows. Consider \(\Phi _t,t\in [0,1]\), the family of \(2\times 2\) symplectic matrices representing the linearized Reeb flow on \(\xi\) along P in coordinates induced by \(\tau\) as in the previous section.

For each initial condition \(0\ne {\bar{\zeta }}\) let \(\Theta _{\bar{\zeta }}(t)\) be a continuous argument of \(\Phi _t{\bar{\zeta }}\) with \(t\in [0,1]\). Let

$$\begin{aligned} \Delta \Theta ({\bar{\zeta }}) = \frac{\Theta _{{\bar{\zeta }}}(1) - \Theta _{{\bar{\zeta }}}(0)}{2\pi }, \end{aligned}$$

and

$$\begin{aligned} I^\tau (P):= \{ \Delta \Theta ({\bar{\zeta }}):0\ne {\bar{\zeta }} \in {\mathbb {R}}^2\}, \end{aligned}$$

be the interval containing the argument variations of all initial conditions.

The length of \(I^\tau (P)\) is less than \(\frac{1}{2}\) and thus, for each \(\epsilon >0\) small, either \(I^\tau (P)-\epsilon\) contains an integer k or is contained in between two consecutive integers k and \(k+1\). In the first case we define \({{\widetilde{\mu }}}^\tau (P) := 2k\) and in the second case, \({{\widetilde{\mu }}}^\tau (P)= 2k+1\).

It is a simple exercise to check that

$$\begin{aligned} \mathrm{CZ}^\tau (P) = {{\widetilde{\mu }}}^\tau (P). \end{aligned}$$

Moreover, P is nondegenerate if and only if the boundary of \(I^\tau (P)\) does not contain an integer. For a proof to these facts, see [22].

Appendix D: The quaternionic trivialization

Let \(S \subset {\mathbb {R}}^4\) be a regular energy level of a Hamiltonian function H, that is \(S=H^{-1}(c), c \in {\mathbb {R}},\) and \(\nabla H|_S\) never vanishes. The quaternion group induces an orthonormal frame

$$\begin{aligned} TS = \mathrm{span}\{X_1,X_2,X_3\}|_S, \end{aligned}$$
(13)

spanned by the vector fields

$$\begin{aligned} X_i=A_i \dfrac{\nabla H}{|\nabla H|} \subset TS, \,\,\, i=1,2,3. \end{aligned}$$
(14)

Here, the \(4\times 4\) matrices \(A_i,i=1,2,3,\) are

$$\begin{aligned} A_1 = \left( \begin{array}{cc} 0 &{} J \\ J &{} 0 \end{array} \right) \quad A_2 =\left( \begin{array}{cc} J &{} 0 \\ 0 &{} -J \end{array} \right) \quad A_3 = \left( \begin{array}{cc} 0 &{} I \\ -I &{} 0 \end{array} \right) , \end{aligned}$$

where 0, I and J are the \(2 \times 2\) matrices

$$\begin{aligned} 0= \left( \begin{array}{cc} 0 &{} 0 \\ 0 &{} 0 \end{array} \right) \quad I = \left( \begin{array}{cc} 1 &{} 0 \\ 0 &{} 1 \end{array} \right) \quad J = \left( \begin{array}{cc} 0 &{} 1 \\ -1 &{} 0 \end{array} \right) . \end{aligned}$$
(15)

Observe that \(X_3\) is parallel to the Hamiltonian vector field \(X_H=A_3 \nabla H\).

Denote by \(\phi _t\) the Hamiltonian flow of \(X_H\) restricted to S. Since \(d\phi _t: TS \rightarrow TS\) preserves the line bundle \({\mathbb {R}}X_3\), one may restrict the study of the linearized flow to

$$\begin{aligned} TS/{\mathbb {R}}X_3 \simeq \mathrm{span} \{X_1,X_2\}. \end{aligned}$$

As discussed in Sect. 1.1, if S has contact-type then \(X_H|_S\) is parallel to a Reeb vector field R of a contact form \(\lambda\) on S. In this case, the contact structure \(\xi =\ker \lambda\) is transverse to R and as a result we have \(\xi \simeq \mathrm{span} \{X_1,X_2\}\). This makes the quaternionic trivialization a useful tool for estimating Conley–Zehnder indices.

The orthonormal frame (13) induces a trivialization \(\Psi : TS \rightarrow S \times {\mathbb {R}}^3\)

$$\begin{aligned} \Psi : TS \ni a_1X_1 + a_2X_2 +a_3X_3 \mapsto (a_1,a_2,a_3)\in {\mathbb {R}}^3 \end{aligned}$$
(16)

which provides a simple form to analyze the transverse linearized flow along Hamiltonian trajectories.

Proposition D.1

In coordinates \((a_1,a_2)\in {\mathbb {R}}^2\) induced by the trivialization (16), a solution to the linearized flow along a non-constant trajectory of \(X_H\), projected to \(TS/{\mathbb {R}}X_3\), satisfies the equation

$$\begin{aligned} \left( \begin{array}{c}\dot{a}_1 \\ \dot{a}_2 \\ \end{array} \right) =-J M\left( \begin{array}{c} a_1 \\ a_2 \end{array} \right) , \end{aligned}$$
(17)

where J is given in (15), M is the symmetric matrix

$$\begin{aligned} M = \left( \begin{array}{cc} \kappa _{11} + \kappa _{33} &{} \kappa _{12} \\ \kappa _{12} &{} \kappa _{22}+\kappa _{33} \end{array} \right) , \end{aligned}$$

and

$$\begin{aligned} \kappa _{ij} = \langle \mathbf{H} X_i,X_j \rangle . \end{aligned}$$

The matrix \(\mathbf{H}=\mathbf{H}(x)\) is the Hessian of H at \(x\in S\), \(X_i, i=1,2,3\), is given by (14), and \(\langle \cdot ,\cdot \rangle\) is the standard inner product on \({\mathbb {R}}^4\).

Proof

Let \(x(t)\in S\) be a Hamiltonian trajectory, that is a solution to

$$\begin{aligned} \dot{x} = A_3 \nabla H(x). \end{aligned}$$
(18)

A solution \(y(t)\in T_{x(t)}S\) to the linearized flow \(d\phi _t:TS\rightarrow TS\) along x(t) satisfies the linear differential equation

$$\begin{aligned} \dot{y} = A_3 \mathbf{H}(x) y. \end{aligned}$$
(19)

Substituting \(y=a_1X_1 + a_2X_2+a_3X_3\) in (19), we obtain

$$\begin{aligned} \sum _{i=1}^3 (\dot{a}_iX_i+a_i\dot{X}_i) = \sum _{i=1}^3 a_i A_3 \mathbf{H} X_i. \end{aligned}$$
(20)

We may assume for simplicity that \(|\nabla H|=1\). In particular, it follows from (14) and (18) that

$$\begin{aligned} \dot{X}_i = A_i \mathbf{H} {\dot{x}}=A_i \mathbf{H}A_3 \nabla H=A_i\mathbf{H}X_3. \end{aligned}$$
(21)

Taking the inner product of the expression (20) with \(X_1\), using (21) and the relations

$$\begin{aligned} \begin{aligned} A_i^T&=-A_i , \quad A_1A_2 = A_3, \quad A_2A_3 = A_1, \quad A_3A_1=A_2,\\ \langle X_i,X_j\rangle&= \delta _{ij} \Rightarrow \langle \dot{X}_i, X_i\rangle =0, \quad \langle \dot{X}_i, X_j\rangle = - \langle \dot{X}_j,X_i\rangle \quad \forall i,j, \end{aligned} \end{aligned}$$

we obtain

$$\begin{aligned} \dot{a}_1 =- a_2\langle \mathbf{H}X_3,X_3\rangle -a_1\langle \mathbf{H}X_1,X_2\rangle - a_2\langle \mathbf{H}X_2,X_2 \rangle . \end{aligned}$$

This is precisely the expression for \(\dot{a}_1\) in (17). Analogously one obtains

$$\begin{aligned} \dot{a}_2 =a_1\langle \mathbf{H}X_1,X_1\rangle +a_1\langle \mathbf{H}X_3,X_3\rangle + a_2\langle \mathbf{H}X_1,X_2 \rangle . \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

de Paulo, N.V., Salomão, P.A.S. Reeb flows, pseudo-holomorphic curves and transverse foliations. São Paulo J. Math. Sci. 16, 314–339 (2022). https://doi.org/10.1007/s40863-022-00285-0

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40863-022-00285-0

Keywords

Navigation