Skip to main content
Log in

A transfer principle: from periods to isoperiodic foliations

  • Published:
Geometric and Functional Analysis Aims and scope Submit manuscript

Abstract

We classify the possible closures of leaves of the isoperiodic foliation defined on the Hodge bundle over the moduli space of genus \(g\ge 2\) curves and prove that the foliation is ergodic on those sets. The results derive from the connectedness properties of the fibers of the period map defined on the Torelli cover of the moduli space. Some consequences on the topology of Hurwitz spaces of primitive branched coverings over elliptic curves are also obtained. To prove the results we develop the theory of augmented Torelli space, the branched Torelli cover of the Deligne–Mumford compactification of the moduli space of curves.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

Similar content being viewed by others

Notes

  1. In the literature, this foliation is also called the kernel foliation, or the absolute period foliation.

  2. Usually what is refered to as Birman’s exact sequence is for a surface punctured at only one point, but the idea of the proof is the same in the case of two points: given an element of the braid group \([\varphi ]\in B(\Sigma _{g,2})\), represented by an isotopy class of diffeomorphism \(\varphi \in \text {Diff} (\Sigma _g, q_1,q_2)\), there exists an isotopy \((\varphi _t)_{t\in [0,1]}\) in \(\text {Diff}(\Sigma _g)\) such that \(\varphi _0=\varphi \) and \(\varphi _1 = Id\). The loop \((\varphi _t (q_1),\varphi _t (q_2))\in \Sigma _g^2 {\setminus } \Delta \) defines an element \(\gamma \in \pi _1 (\Sigma _g^2 {\setminus } \Delta , (q_1,q_2))\) that only depends on \([\varphi ]\) and the map \( [\varphi ] \mapsto \gamma \) induces an isomorphism between \(B(\Sigma _{g,2})\) and \(\pi _1 (\Sigma _g^2 {\setminus } \Delta , (q_1,q_2))\).

References

  1. W. Abikoff. Degenerating families of Riemann surfaces, Ann. Math. (2), 105(1) (1977), 29–44.

    MathSciNet  MATH  Google Scholar 

  2. L.V. Ahlfors. The complex analytic structure of the space of closed Riemann surfaces, In: Analytic functions, Princeton Univ. Press, Princeton, N.J., (1960), pp. 45—66.

  3. E. Arbarello, M. Cornalba, and P. Griffiths. Geometry of Algebraic Curves, Vol. II, Springer Verlag, (2011).

    MATH  Google Scholar 

  4. V.I. Arnold. A remark on the branching of hyperelliptic integrals as functions of the parameters, Funkcional. Anal. i Prilozen., 2(3) (1968), 1–3.

    MathSciNet  Google Scholar 

  5. M. Bainbridge. Euler characteristics of Teichmüller curves in genus two, Geom. Topol., 11(4), 1887–2073.

  6. M. Bainbridge, D. Chen, Q. Gendron, S. Grushevsky, and M. Möller. Compactification of strata of Abelian differentials, Duke Math. J., 167(12) (2018), 2347–2416.

    MathSciNet  MATH  Google Scholar 

  7. M. Bainbridge, D. Chen, Q. Gendron, S. Grushevsky, and M. Möller. The moduli of multi-scale differentials, 1910.13492, pp. 1–122.

  8. M. Bainbridge, C. Johnson, C. Judge, and I. Park. Haupt’s theorem for strata of abelian differentials. 2002.12901.

  9. Benson Farb and Dan Margalit. A primer on Mapping Class groups, Pinceton University Press, (2001).

    MATH  Google Scholar 

  10. I. Berstein and A.L. Edmonds. On the classification of generic branched coverings of surfaces, Illinois Jr. Math., 24(Nr. 1) (1984), 64–82.

    MathSciNet  MATH  Google Scholar 

  11. L. Bers. Spaces of degenerating Riemann surfaces, Discontinuous groups and Riemann surfaces (Proc. Conf., Univ. Maryland, College Park, Md., 1973), Princeton Univ. Press, Princeton, N.J., (1974), pp. 43–55. Ann. of Math. Studies, no. 79.

  12. L. Bers. Deformations and moduli of Riemann surfaces with nodes and signatures, Math. Scand., 36 (1975), 12–16, collection of articles dedicated to Werner Fenchel on his 70th birthday.

  13. L. Bers. Finite dimensional Teichmüller spaces and generalizations, Bull. AMS, Vol. 5, Nr. 2, (1981).

  14. P. Buser. Geometry and Spectra of Compact Riemann Surfaces, Progress in Mathematics, Vol. 106, Birkhäuser Boston Inc., Boston, MA, (1992).

  15. G. Calsamiglia, B. Deroin, and S. Francaviglia. Branched projective structures with Fuchsian holonomy. Geom. Topol., 18 (2014), 379–446.

    MathSciNet  MATH  Google Scholar 

  16. A. Calderon and N. Salter. Relative homological representations of framed mapping class groups, to appear in B. Lond. Math. Soc. (2020).

  17. A. Calderon and N. Salter. Connected components of strata of Abelian differentials over Teichmüller space Comment, Math. Helv., 95(2) (2020), 361–420.

    MathSciNet  MATH  Google Scholar 

  18. K. Calta. Veech surfaces and complete periodicity in genus two, J. Am. Math. Soc., 17(4) (2004), 871–908.

    MathSciNet  MATH  Google Scholar 

  19. C.H. Clemens. A Scrapbook of Complex Curve Theory, Plenum Press, New York and London, (1980).

    MATH  Google Scholar 

  20. P. Deligne and D. Mumford. The irreducibility of the space of curves of given genus, Inst. Hautes Études Sci. Publ. Math., (36) (1969), 71–109.

  21. E.B. Dynkin. Maximal subgroups of the classical groups, Trudy Moskov. Mat. Obsc., 1 (1952), 39–166.

    MathSciNet  MATH  Google Scholar 

  22. A.L. Edmonds. Deformation of maps to branched coverings in dimension two, Ann. Math. (2), 110(1) (1979), 113–125.

    MathSciNet  MATH  Google Scholar 

  23. A. Eskin, H. Masur, and A. Zorich. Moduli spaces of abelian differentials: the principal boundary, counting problems and the Siegel-Veech constants, Publ. Math. Inst. Hautes Études Sci. No., 97 (2003), 61–179.

    MathSciNet  MATH  Google Scholar 

  24. A. Eskin and M. Mirzakhani. Invariant and stationary measures for the action on Moduli space. Publ. Math. Inst. Hautes Études Sci. No., 127 (2018), 95–324.

    MathSciNet  MATH  Google Scholar 

  25. B. Farb and D. Margalit. A Primer on Mapping Class Groups, Princeton University Press, (2012).

    MATH  Google Scholar 

  26. L. Gerritzen and F. Herrlich. The extended Schottky space. J. Reine Angew. Math., 389 (1988), 190–208.

    MathSciNet  MATH  Google Scholar 

  27. V. Gorbatsevich, A. Onishchik, and È. Vinberg. Lie Groups and Lie Algebras. III Springer-Verlag, Berlin, (1994). Structure of Lie groups and Lie algebras.

  28. M. Gromov. Rigid transformations groups, In: Gèometrie differentielle (Paris 1986) (D. Bernard, Y. Choquet-Bruhat ed.) Travaux en Cours, 33, Hermann, Paris (1988), pp. 65–139.

  29. U. Hamenstädt. Dynamical properties of the absolute period foliation, Isr. J. Math., 225 (2018), 661—680.

    MATH  Google Scholar 

  30. U. Hamenstädt. Quotients of the orbifold fundamental group of strata of abelian differentials. Preprint, (2018).

  31. U. Hamenstädt. On the orbifold fundamental group of the odd component of the stratum H(2,..,2). Preprint, (2020).

  32. J. Harris and I. Morrison. Moduli of Curves, Graduate Texts in Mathematics 187, Springer, New York, (1998).

    Google Scholar 

  33. R. Hartshorne. Algebraic Geometry, Graduate Texts in Mathematics, Vol. 52, Springer-Verlag, N.Y, (1997).

    Google Scholar 

  34. O. Haupt. Ein Satz über die Abelschen integrale I, Gattung. Math. Z., 6(3-4) (1920), 219–237.

    MATH  Google Scholar 

  35. W.J. Harvey. Boundary Structure of the Modular Group, Riemann Surfaces and Related Topics. Proceedings of the 1978 Stony Brook Conference, (1981).

  36. F. Herrlich. The extended Teichmüller space, Math. Z., 203(2) (1990), 279–291.

    MathSciNet  MATH  Google Scholar 

  37. V. Hinich and A. Vaintrob. Augmented Teichmüller spaces and orbifolds, Sel. Math. New Ser., 16 (2010), 533–629.

    MATH  Google Scholar 

  38. W.P. Hooper and B. Weiss. Rel leaves of the Arnoux-Yoccoz surfaces, Selecta Mathematica, 24 (2018), 875–934.

    MathSciNet  MATH  Google Scholar 

  39. J.H. Hubbard and S. Koch. An analytic construction of the Deligne-Mumford compactification of the moduli space of curves, J. Diff. Geom., 98 (2014), 261–313.

    MathSciNet  MATH  Google Scholar 

  40. P. Hubert, H. Masur, T.A. Schmidt, and A. Zorich, Problems on billiards, flat surfaces and translation surfaces, In: Problems on mapping class groups and related topics, Proc. Sympos. Pure Math., vol. 74, Amer. Math. Soc, (2006), pp. 233–243.

  41. D. Johnson. Conjugacy relations in subgroups of the mappingclass group and a group-theoretic description of the Rochlin invariant, Math. Ann., 249(3) (1980), 243–263.

    MathSciNet  MATH  Google Scholar 

  42. M. Kapovich. Periods of abelian differentials and dynamics. “Dynamics: Topology and Numbers” (Proceedings of Kolyada Memorial Concerence), Contemporary Mathematics, AMS, Vol. 744, (2020), pp. 297–315.

  43. M. Kontsevich and A. Zorich. Connected components of the moduli spaces of Abelian differentials with prescribed singularities, Invent. Math., 153(3) (2003), 631–678.

    MathSciNet  MATH  Google Scholar 

  44. T. Le Fils. Periods of abelian differentials with prescribed singularities. 2003.02216.

  45. A.G. Maier. Trajectories on the closed orientable surfaces, (In Russian) Sb. Math., 12(54) (1943), 71–83.

    MathSciNet  Google Scholar 

  46. G.A. Margulis. Discrete Subgroups of Semi-simple Lie Groups, Springer-Verlag, (1991).

    MATH  Google Scholar 

  47. H. Masur. Closed trajectories for quadratic differentials with an application to billiards, Duke Math. J., 53(2) (1986), 307–314.

    MathSciNet  MATH  Google Scholar 

  48. H. Masur. Extension of the Weil–Petersson metric to the boundary of Teichmüller space, Duke Math. J., 43(3) (1976), 623–635.

    MathSciNet  MATH  Google Scholar 

  49. H. Masur. Interval exchange transformations and measured foliations, Ann. Math., 2/ 115 (1982), 169—200.

  50. C. McMullen. Moduli spaces of isoperiodic forms on Riemann surfaces. Duke Math. J., 163(12) (2014), 2271–2323.

    MathSciNet  MATH  Google Scholar 

  51. C. McMullen. Navigating moduli space with complex twists, J. Eur. Math. Soc., 15 (2013), 1223–1243.

    MathSciNet  MATH  Google Scholar 

  52. C. McMullen. Dynamics of \({\rm SL}_2(\mathbb{R})\) over moduli space in genus two, Ann. Math. (2), 165(2) (2007), 397–456.

    MathSciNet  Google Scholar 

  53. C. McMullen. Foliations of Hilbert Modular Surfaces, American Journal of Mathematics, 129(1) (2007), 183–215.

    MathSciNet  MATH  Google Scholar 

  54. G. Mess. The Torelli groups for genus 2 and 3 surfaces, Topology, 31(4) (1992), 775–790.

    MathSciNet  MATH  Google Scholar 

  55. C.C. Moore. Ergodicity of flows on homogeneous spaces, Am. J. Math., 88 (1966), 154–178.

    MathSciNet  MATH  Google Scholar 

  56. H. Movasati. On elliptic modular foliations, Indagationes Mathematicae, 19(2) (2008), 263–286.

    MathSciNet  MATH  Google Scholar 

  57. A. Putman. Cutting and pasting in the Torelli group, Geom. Topol., 11 (2007), 829–865.

    MathSciNet  MATH  Google Scholar 

  58. M.S. Raghunathan. Discrete subgroups of Lie groups. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 68, Springer-Verlag, New York-Heidelberg, (1972). ix+227.

  59. M. Ratner. Raghunathan’s topological conjecture and distributions of unipotent flows, Duke Math. J., 63(1) (1991), 235–280.

    MathSciNet  MATH  Google Scholar 

  60. M. Schiffer. A method of variation within the family of simple functions, Proc. Lond. Math. Soc., 44 (1938), 450–452.

    MathSciNet  MATH  Google Scholar 

  61. M. Schmoll. Spaces of Elliptic Differentials. Algebraic and Topological Dynamics. Contemp. Math. vol. 385, AMS, Providence, RI, (2006), pp. 193–206.

  62. C. Simpson. Lefschetz theorems for the integral leaves of a holomorphic one-form, Compositio Math., 87(1) (1993), 99–113.

    MathSciNet  MATH  Google Scholar 

  63. K. Strebel. Quadratic Differentials, Ergebnisse der Mathematik, Springer-Verlag, (1984).

    MATH  Google Scholar 

  64. R. Torelli. Sulle varieté di Jacobi, Rendiconti della Reale accademia nazionale dei Lincei., 22(5) (1913), 98–103.

    MATH  Google Scholar 

  65. W.A. Veech. Moduli spaces of quadratic differentials, J. Analyse Math., 55 (1990), 117—171.

    MathSciNet  MATH  Google Scholar 

  66. C. Voisin. Hodge Theory and Complex Algebraic Geometry II, Cambridge Studies in Advanced Mathematics. (2003).

  67. M. Wolf and S. Wolpert. Real analytic structures on the moduli space of curves, Am. J. Math., 114(5) (1992), 1079–1102.

    MathSciNet  MATH  Google Scholar 

  68. S. Wolpert. Geometry of the Weil-Petersson completion of Teichmüller space, Surveys in Differential Geometry, Vol. VIII (Boston, MA, 2002), Surv. Differ. Geom., VIII, Int. Press, Somerville, MA, (2003), pp. 357–393.

  69. S. Wolpert. Families of Riemann surfaces and Weil-Petersson geometry, CBMS Regional Conference Series in Mathematics, Vol. 113, published for the Conference Board of the Mathematical Sciences, Washington, DC, (2010).

  70. A. Wright. Cylinder deformations in orbit closures of translation surfaces. Geom. Topol., 19(1) (2015), 414–438.

    MathSciNet  MATH  Google Scholar 

  71. S. Yamada. On the geometry of Weil-Petersson completion of Teichmüller spaces, Math. Res. Lett., 11(2-3) (2004), 327–344.

    MathSciNet  MATH  Google Scholar 

  72. F. Ygouf. A criterion for density of the isoperiodic leaves in rank 1 affine invariant suborbifolds. 2002.01186.

  73. R.J. Zimmer. Ergodic Theory and Semi-simple Groups. Monographs in Mathematics, Vol. 81. Birkhäuser, (1984).

Download references

Acknowledgements

We warmly thank the referees of this article for the careful reading and the many improvements that their comments made possible. We also hank U. Hamenstädt, P. Hubert, M. Kapovich, E. Lanneau, F. Loray, D. Margalit, M. Möller, G. Mondello, H. Movasati, A. Putman and A. Wright for useful conversations. This paper was partially supported by the France-Brazil agreement in Mathematics. G. Calsamiglia was partially supported by Faperj/CNPq/CAPES/Mathamsud/Cofecub, B. Deroin by ANR project LAMBDA ANR-13-BS01-0002, and S. Francaviglia by GNSAGA group of INdAM, and by PRIN 2017JZ2SW5. It was mainly developed at Universidade Federal Fluminense, IMPA, ENS/Paris, U. Cergy-Pontoise, UPMC, and Università de Bologna, to whom we thank the nice working conditions provided.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gabriel Calsamiglia.

Additional information

Publisher's Note

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

Appendices

Appendix I: Proof of Proposition 3.10

We first begin by providing a normal form for periods of positive volume and finite primitive degree, which implies the first item of Proposition 3.10 in the case where the subspace (18) is rational.

Lemma 9.1

Given a surjective homomorphism \(p:{\mathbb {Z}}^{2g}\rightarrow {\mathbb {Z}}+ i {\mathbb {Z}}\) of volume (and primitive degree) \(d\ge 2\), there exists \(M\in \text {Sp}(2g,{\mathbb {Z}})\) and a symplectic basis \(\{a_1,b_1,\ldots , a_g,b_g\}\) of \({\mathbb {Z}}^{2g}\) such that

$$\begin{aligned} p\circ M=a_1^*+i(db_1^*+a_2^*) \end{aligned}$$

where the star denotes the symplectic dual of the given element.

Proof

Denote by \(x=\Re (p)\) and \(y=\Im (p)\) the elements in \(({\mathbb {Z}}^{2g})^*\). They satisfy \(x\cdot y=d\). Choose \(a_1\in {\mathbb {Z}}^{2g}\) the symplectic dual of x. Choose some \(b_1\in {\mathbb {Z}}^{2g}\) such that \(a_1\cdot b_1=1\) and write

$$\begin{aligned} y=m_1a_1^*+db_1^{*}+m_2a_2^{*} \end{aligned}$$
(58)

where \(a_2\) is a primitive element satisfying \(a_1\cdot a_2=b_1\cdot a_2=0\) and \(m_1,d\) and \(m_2\) are co-prime. Complete those elements to a symplectic basis \(a_1,b_1, \ldots , a_g,b_g\) of \({\mathbb {Z}}^{2g}\). The image of an element \(u=\sum _{k\ge 1}\alpha _ka_k+\beta _kb_k\) under p is

$$\begin{aligned} p(u)=(a_{1}^{*}(u), (m_1a_1^*+db_1^{*}+m_2a_2^{*})(u))=(\alpha _1, m_1\alpha _1+d\beta _1+m_2\alpha _2). \end{aligned}$$

Therefore

$$\begin{aligned} p(u)= & {} 1\Leftrightarrow \alpha _1=1\text { and } m_1+d\beta _2+m_2\alpha _2=0, \\ p(u)= & {} i\Leftrightarrow \alpha _1=0\text { and } d\beta _1+m_2\alpha _2=1. \end{aligned}$$

Hence, p is surjective if and only if d and \(m_2\) are co-prime.

To conclude, we are going to show that under the hypothesis of d and \(m_2\) co-prime, there exists a choice of \(b_1\) such that the coefficients \(m_1\) and \(m_2\) in the splitting given by (58) are 0 and 1 respectively.

Let us analyze the effect of a change of the first given \(b_1\). Write \(\widetilde{a_1}^*=a_1^*\) and

$$\begin{aligned} \widetilde{b_1}^*=u_1a_1^*+b_1^*+\sum _{k\ge 2}u_ka_k^*+v_kb_k^* \end{aligned}$$

for some \(u_k,v_k\in {\mathbb {Z}}\). The new decomposition \(y=\widetilde{m_1}\widetilde{a_1}^*+d\widetilde{b_1}^*+\widetilde{m_2}\widetilde{a_2}^*\) has the properties

$$\begin{aligned} (y-\widetilde{m_1}\widetilde{a_1}^*-d\widetilde{b_1}^*)\cdot \widetilde{a_1}^*=0\text { and } (y-\widetilde{m_1}\widetilde{a_1}^*-d\widetilde{b_1}^*)\cdot \widetilde{b_1}^*=0. \end{aligned}$$

The first equation is automatically satisfied, and the second gives

$$\begin{aligned} \widetilde{m_1}=m_1-du_1+m_2v_2. \end{aligned}$$
(59)

Since d and \(m_2\) are co-prime we can already choose \(u_1\) and \(v_2\) to get \(\widetilde{m_1}=0\) and restart the argument by supposing \(m_1=0\).

For this choice, we have

$$\begin{aligned} \widetilde{m_2}^*\widetilde{a_2}^*=y-\widetilde{m_1}\widetilde{a_1}^*-d\widetilde{b_1}^*=m_2v_2 a_1^{*}+(m_2-du_2)a_2^*-dv_2b_2^*-d(\sum _{k\ge 3}u_ka_k^*+v_kb_k^*) \end{aligned}$$

and from this

$$\begin{aligned} \widetilde{m_2}=\text {gcd}(-m_2v_2, m_2-du_2, -dv_2, -du_3, -dv_3, \ldots , -d u_g, -dv_g). \end{aligned}$$

If we still want \(\widetilde{m_1}=0\) we need to impose \(m_2v_2=du_1\) by equation (59). The choice \(v_2=d\), \(u_1=m_2\) and all other coefficients equal to zero gives \(\widetilde{m_1}=0\) and \(\widetilde{m_2}=1\), as desired. \(\square \)

We continue the proof of Proposition 3.10, which is reminiscent of Ratner’s theory.

Equipp \(\mathbb R^{2g}\) with its canonical symplectic form \(\omega (x,y) = \sum _{1\le k\le g} x_{2k}y_{2k+1}-x_{2k+1}y_{2k}\). The volume of a period \(p\in \mathbb C^{2g}\) is the symplectic product \( V(p)= \omega (\Re p , \Im p ) \). Since the action of \(\Gamma \) is linear, and that the volume is multiplicative, namely \(V(\lambda p ) = |\lambda |^2 V(p)\) for every \(\lambda \in \mathbb C\) and \(p\in \mathbb C^{2g}\), we can restrict our attention to the action of \(\Gamma \) on the subset \(X\subset \mathbb C^{2g}\) whose elements have volume 1. In real and imaginary coordinates the set of periods of volume \(1\) is then the set of pairs \((x,y)\in {\mathbb {R}}^{2g}\times {\mathbb {R}}^{2g}\) such that \(\omega (x,y)=1\).

The simple real Lie group \(G = \text {Sp}(2g, \mathbb R)\) acts transitively on the set of couples \((x,y)\in (\mathbb R^{2g})^2\) such that \(\omega (x,y)=1\), and that the stabilizer of the couple

$$\begin{aligned} \big ( (1,0,\ldots ,0), (0,1,0,\ldots ,0) \big )\text { is the group } \left( \begin{array}{ccc} 1 &{} &{} \\ &{} 1 &{} \\ &{} &{} \text {Sp}(2g-2,\mathbb R)\end{array} \right) \end{aligned}$$

that we will denote by U in the sequel. Our set X is isomorphic to the homogeneous space G/U. The linear action of \(\Gamma \) on X is under the isomorphism \(X \simeq G/U\) given by left multiplication on G/U.

Since the group G is simple, that U is generated by unipotent elements, and that \(\Gamma \) is a lattice in G, Ratner’s theorem [Rat91] tells us that the closure of the \(\Gamma \)-orbits on X are homogeneous in the following sense

Theorem 9.2

(Ratner). For every \(p\in X\) of the form \(p = gU\), there exists a closed subgroup H of G containing \(U^g= gUg^{-1}\), such that \(\Gamma \cap H\) is a lattice in H, and such that \(\overline{\Gamma \cdot p} = \Gamma H p\).

Notice that in our situation, we have \(U^g = I_{|W} \oplus \text {Sp}(W^\perp ) \simeq \text {Sp} (2g-2,\mathbb R)\) where \(W={\mathbb {R}}\Re p+{\mathbb {R}}\Im p \subset {\mathbb {R}}^{2g}\) is the symplectic subspace associated to the volume one \(p\in {\mathbb {C}}^2\).

Let \(H_0\) be the connected component of H containing the identity: then \(\Gamma \cap H_0\) is still a lattice in \(H_0\), and \(U^g \) is contained in \(H_0\).

If \(H_0=G\) then \(\overline{\Gamma \cdot p}=G\) and we deduce that the orbit closure is dense in X. Since the closure of \(\Lambda (p)\) contains all the \(\Lambda (q)\) of elements \(q\in \overline{\Gamma \cdot p}\), we have \(\overline{\Lambda (p)}={\mathbb {C}}\).

If \(H_0\) is a proper subgroup of G, Kapovich observes that it falls into two categories

  • (Semi-simple case) \(H_0\) is of the form \(S \oplus \text {Sp} (W^\perp )\), where S is a Lie subgroup of \(\text {Sp}(W)\).

  • (Non semi-simple case) \(H_0\) is not semi-simple and preserves a line \(L \subset W\).

The proof of this dichotomy can be found in [Kap20, p. 12], and is based on Dynkin’s classification of maximal connected complex Lie subgroups of \(\text {Sp}(2g,\mathbb C)\), see [Dyn52]. Let L be a maximal complex Lie subgroup of \(\text {Sp}(2g,\mathbb C)\) which contains \(H_0\). If \(H_0\ne \text {Sp} (2g,\mathbb R)\), its Zariski closure in the complex domain is a strict subgroup of \(\text {Sp}(2g,\mathbb C)\), so it is contained in a maximal complex Lie (strict) subgroup of \(\text {Sp}(2g,\mathbb C)\). It satisfies one of the following properties (see [GOV94, Ch. 6, Thm 3.1, 3.2]):

  1. (1)

    \(L= \text {Sp} (V) \oplus \text {Sp} (V^\perp )\) for some complex symplectic subspace \(V\subset \mathbb C^{2g}\),

  2. (2)

    L is conjugated to \(\text {Sp} (s,\mathbb C) \otimes \text {SO} (t,\mathbb C)\) where \(2g = st\), \(s\ge 2\), \(t\ge 3\), \(t\ne 4\) or \(t=4 \) and \(s=2\),

  3. (3)

    L preserves a line of \(\mathbb C^{2g}\).

Since \(H_0\) contains \(U^g\), L contains the complexification of \(U^g\), which is nothing but \(\text {Id}_{W_ \mathbb C}\oplus \text {Sp} (W_ \mathbb C ^\perp )\), where \(W_\mathbb C\) denotes the complexification \(W \otimes _\mathbb R \mathbb C\) of W. In case (1), the only possibility is that up to permutation of V and \(V^\perp \), we have \(W_\mathbb C= V\). In particular, \(H_0\) is a subgroup of \(\text {Sp} (W) \oplus \text {Sp}(W^\perp )\). Since it contains \(\text {Id}_{|W} \oplus \text {Sp}(W^\perp )\), it must be of the form \(S \oplus \text {Sp} (W^\perp )\), where S is a Lie subgroup of \(\text {Sp}(W)\). Case (2) cannot occur. In case (3), observe that the line L needs to be in \(W_\mathbb C\), since the group \(\text {Id}_{|W_\mathbb C} \oplus \text {Sp} (W_\mathbb C ^\perp )\) preserves this line. If L is defined over the reals, we are done. If not, both L and \({\overline{L}}\) (the image of L by the complex conjugation) are preserved by \(H_0\), and thus \(H_0\) is a subgroup of \(\text {Sp} (W) \oplus \text {Sp} (W^\perp )\). As before, because it contains \(\text {Id}_W \oplus \text {Sp} (W^\perp )\), it must be of the form \(S \oplus \text {Sp}(W^\perp )\), where S is a Lie subgroup of \(\text {Sp}(W)\).

Remark 9.3

If \(\Gamma _W:= \Gamma \cap (\text {Id}_W \oplus \text {Sp}(W^\perp ))\) is a lattice in \(\text {Id}_W \oplus \text {Sp}(W^\perp )\) then W is defined over \({\mathbb {Q}}\) and the conclusion of the first item of Proposition 3.10 follows easily. Indeed, \(\Gamma _W \) acts by the identity on \(W^\sigma \) for every Galois automorphism \(\sigma \). The Zariski closure of \(\Gamma _W\) being \(\text {Id}_W \oplus \text {Sp} (W^\perp )\) (by Borel density theorem, see [Zim84]), \(\text {Id}_W \oplus \text {Sp}(W^\perp )\) acts by the identity on \(W^\sigma \) as well. This implies that \(W^\sigma = W\) for every \(\sigma \) (otherwise \(\text {Id}\oplus \text {Sp}(W^{\perp })\) acts by the identity on the non-trivial subspace \((W+W^{\sigma })\cap W^{\perp }\)), and so W is rational.

Let us proceed to analyze the different cases.

Semi-simple case. Since the group \(H_0 = S\oplus \text {Sp} (W^\perp )\) contains a lattice, it must be unimodular. In particular, either S is the trivial group, or a 1-parameter subgroup, or the whole \(\text {Sp} (W)\). If S is trivial, then Ratner’s Theorem tells us \(\Gamma _W:= \Gamma \cap (\text {Id}_W \oplus \text {Sp}(W^\perp ))\) is a lattice in \(\text {Id}_W \oplus \text {Sp}(W^\perp )\) and we fall in case (1) by Remark 9.3.

If S is 1-dimensional, \(S\oplus \text {Id}_{W^\perp }\) would be the radical of \(H_0\), and a theorem of Wolf and Raghunathan, see [Rag72], shows that it would intersect \(\Gamma \) in a lattice. This implies that the intersection of \(\Gamma \) with \(\text {Id}_W \oplus \text {Sp}(W^\perp )\) is also a lattice. Indeed, let \({\overline{\Gamma }}\) denote the natural projection of \(\Gamma \) on \(\text {Sp}(W^{\perp })\). Since the sequence

$$\begin{aligned} 0\rightarrow S\cap \Gamma \rightarrow \Gamma \rightarrow {\overline{\Gamma }}\rightarrow 1 \end{aligned}$$

is exact, the map \(\Gamma \rightarrow (S\cap \Gamma )\backslash S\) induced by the projection of \(S\oplus \text {Sp}(W^{\perp })\) to S induces a morphism \({\overline{\Gamma }}\rightarrow (\Gamma \cap S)\backslash S\). The group \((\Gamma \cap S)\backslash S\) is abelian, and since \({\overline{\Gamma }}\) has Kazdhan property (T), the image group in \((\Gamma \cap S)\backslash S\) is finite. We thus conclude that the image of \(\Gamma \) on the factor S is a lattice and therefore \(\Gamma \cap (\text {Id}_W \oplus \text {Sp}(W^\perp ))\) as well. We conclude as in Remark 9.3.

Finally, it remains to treat the case where \(S= \text {Sp}(W)\). This case splits into two subcases, depending on the lattice \(\Gamma \cap \text {Sp} (W) \oplus \text {Sp} (W^\perp )\) being reducible or irreducible. If it is reducible, this implies that \(\Gamma \cap (\text {Id}_W \oplus \text {Sp} (W^\perp ))\) is a lattice, and then W must be rational by the above considerations. Assume now that we are in the irreducible case. Then \(g=2\), by a theorem of Margulis [Mar91]. Assume W is not rational, otherwise we are done. Let \(\sigma \) be a Galois automorphism such that \(W^\sigma \ne W\). The group \(\Gamma \cap \text {Sp} (W) \oplus \text {Sp} (W^\perp )\) preserves the splitting \(W^\sigma \oplus (W^\sigma )^\perp \), since \(\Gamma \) and the symplectic form are defined over the rationals. Borel density theorem applied to the lattice \(\Gamma \cap \text {Sp} (W) \oplus \text {Sp} (W^\perp )\) shows that \(\text {Sp}(W) \oplus \text {Sp} (W^\perp )\) preserves the splitting \(W^\sigma \oplus (W^\sigma )^\perp \). This implies that \(W^\sigma = W^\perp \) and \((W^\sigma )^\sigma = W\). This being true for every Galois automorphism, this means that W is defined over a totally real quadratic field K, and we have \(W^\sigma = W^\perp \) where \(\sigma \) is the Galois automorphism of K. This is the only situation where we fall in the last case of Proposition 3.10.

Non semi-simple case. We suppose that \(H_0\) is not semi-simple and prove in this case that the periods p satisfy the second case of Proposition 3.10. We already know that there is a line \(L\subset W\) that is invariant by the action.

For this, we will first need to understand in detail the subgroup B of \(\text {Sp}(2g,\mathbb R) \) formed by all elements that stabilize the line L, see [Kap20, p. 10]. To unscrew the structure of B, notice that any element of B stabilizes both L and \(L^\perp \) so that we have an exact sequence

$$\begin{aligned} CH_{2g}\rightarrow B \rightarrow \text {Sp} (L^\perp /L)\simeq \text {Sp}(2g-2). \end{aligned}$$

The group \( CH_{2g}\) is then the set of elements \(M\in \text {Sp} (2g)\) which induce the identity map on \(L^\perp /L\).

We now have another exact sequence

$$\begin{aligned} H_{2g-1} \rightarrow CH_{2g} \rightarrow \text {GL}(L) \simeq \mathbb R^*, \end{aligned}$$
(60)

the last arrow being given by the restriction of an element \(M\in CH_{2g}\) to the line L. Hence the subgroup \(H_{2g-1}\subset CH_{2g}\) is the group of elements \(M\in \text {Sp} (2g)\) which act as the identity on L and on \(L^\perp / L\). Such M are easily seen to be of the form \(M_{\varphi , \alpha }\), for some \(\varphi \in (L^\perp /L)^*\) and \(\alpha \in \mathbb R\), where

  • the restriction of \(M_{\varphi , \alpha }\) to \(L^\perp \) equals \(id_{|L^\perp } + \varphi a_1\),

  • \( M_{\varphi , \alpha } (b_1) = \alpha a_1 + b_1 + \sum _{k\ge 2} \varphi (b_k) a_k -\varphi (a_k) b_k\),

where \(a_1,b_1,\ldots , a_g,b_g\) is a symplectic basis such that \(L= \mathbb R a_1\). The group structure on \(H_{2g-1}\) is then given by the following relation

$$\begin{aligned} M_{\varphi , \alpha } M_{\varphi ' , \alpha '} = M_{\varphi + \varphi ', \alpha + \alpha ' + \omega (\varphi , \varphi ')}, \end{aligned}$$
(61)

where \(\omega (\varphi , \varphi ')\) is the natural symplectic product induced by \(\omega \) on \((L^\perp /L)^*\), namely

$$\begin{aligned} \omega (\varphi , \varphi ') = \sum _{k\ge 2} \varphi (a_k)\varphi '(b_k)-\varphi '(a_k) \varphi (b_k). \end{aligned}$$

Equation (61) is a straightforward computation. An equivalent formulation is that \(H_{2g-1}\) is the central extension

$$\begin{aligned} \mathbb R \rightarrow H_{2g-1} \rightarrow (L^\perp / L)^*, \end{aligned}$$

defined by the 2-cocycle \( (\varphi , \varphi ') \mapsto \omega (\varphi , \varphi ')\). The group \(H_3\) is isomorphic to the classical Heisenberg group of upper triangular real matrices of size \(3\times 3\) with 1’s on the diagonal.

Now \(CH_{2g} \) is a semi-direct product of \(\mathbb R^*\) by \(H_{2g-1}\), see (6073). To understand its structure, we introduce for every \(\lambda \), one of its lift \(S_\lambda \in CH_{2g}\) defined by

$$\begin{aligned} S_\lambda (a_1) = \lambda a_1, \ S_\lambda (b_1) = \frac{1}{\lambda } b_1, \ S_\lambda (a_k) = a_k,\ S_\lambda (b_k)= b_k\text { for } k\ge 2.\end{aligned}$$

A trivial computation shows that for any \(\lambda \in \mathbb R^*\), every \(\varphi \in (L^\perp /L)^*\) and every \(\alpha \in \mathbb R\), we have

$$\begin{aligned} S_\lambda M_{\varphi , \alpha } S_\lambda ^{-1} = M_{\lambda \varphi , \lambda ^2 \alpha }. \end{aligned}$$
(62)

This shows that \(CH_{2g}\) is not unimodular, and consequently does not contain any lattice.

By construction, our group \(H_0\) is contained in B. We have an exact sequence \(CH_{2g} \rightarrow B \rightarrow \text {Sp}(L^\perp /L, \omega ) \). The image of \(H_0\) by the right arrow is onto since \(H_0\) contains \(U^g\), so that \(H_0\) itself splits as an exact sequence \(CH_{2g} \cap H_0 \rightarrow H_0 \rightarrow \text {Sp} (L^\perp /L, \omega )\). The group \(CH_{2g}\cap H_0\) is invariant under the action by conjugation of \(\text {Sp} (L^\perp /L, \omega )\simeq \text {Sp}(W^{\perp })\). The restriction of this action on \(H_{2g-1}\) can be described explicitly: for \(U\in \text {Sp}(L^{\perp }/L)\simeq \text {Sp}(W^{\perp })\) denote \(s(U)=Id_W\oplus U\in H_0\subset B\). Then

$$\begin{aligned} s(U)M_{\varphi ,\alpha }s(U)^{-1}=M_{\varphi \circ s(U)^{-1},\alpha }. \end{aligned}$$

Lemma 9.4

The closed non-trivial connected subgroups of \(CH_{2g}\) invariant by \(\text {Sp}(W^{\perp })\) are

(1):

\(Z(H_{2g-1}),\)

(2):

lifts of \(\text {GL}^+(L)\) in \(CH_{2g},\)

(3):

\(H_{2g-1},\)

(4):

\(\text {GL}^{+}(L) < imes Z(H_{2g-1}),\)

(5):

\(CH_{2g}.\)

The proof of Lemma 9.4 is an easy consequence of the previous exact sequences and calculations.

Now \(H_0\cap CH_{2g}\) cannot fall in cases (1) and (2) of Lemma 9.4 since in either of those, \(H_0\) would be semi-simple, contrary to hypothesis. It can neither fall in cases (4) or (5), since in those cases equation (62) does not allow \(H_0\) to be unimodular. Hence we are left with the possibility \(H_0\cap CH_{2g}=H_{2g-1}\) (and \(H_0\simeq \text {Sp}(W^{\perp }) < imes H_{2g-1}\)). In this case we will show that the invariant line L is rational, and thus we fall in the second possibility of Proposition 3.10.

The theorem of Raghunathan and Wolf cited above tells us that \(\Gamma \cap H_{2g-1}\) is a lattice in \( H_{2g-1}\). By using Borel’s density Theorem in [Gro88, p.91] we deduce that its Zariski closure is \(H_{2g-1}\). We have \(L\subset K:=\bigcap _{\gamma \in \Gamma \cap H_{2g-1} } \text {Ker} (\gamma - I)\) which is an intersection of rational spaces. If the inclusion is proper, then the Zariski closure of \(\Gamma \cap H_{2g-1}\) would not be the whole of \(H_{2g-1}\). This shows that \(L=K\) and it is a rational one-dimensional subspace of \(\mathbb R^{2g}\).

Up to a real affine change of coordinates on \(\mathbb C\), we can assume that the imaginary part of p generates L, and that it is a primitive element of \(\mathbb Z^{2g}\). Since the group \(H_{2g-1}\) acts transitively on the set of vectors \(v\in \mathbb R^{2g}\) such that \(v\cdot \Im p= 1\), while keeping the period \(\Im p\) fixed, we see that \(H\cdot p\) already contains all the periods q such that \(\Im q = \Im p \) and such that \(V(q)= V(p)= 1\). Since, \(\Gamma \) acts transitively on the set of primitive elements of \(\mathbb Z^{2g}\), we infer that \(\Gamma H p = \overline{\Gamma \cdot p}\) contains all the periods q with volume \(V(p) = 1\) and with a primitive integer imaginary part. Since any periods of \( \overline{\Gamma \cdot p}\) is of this form, we deduce that this situation is exactly the second case of the proposition. The proof of this latter is now complete.

Applying Moore’s ergodic theorem in [Moo66] to each case H above we deduce the final ergodicity part of Proposition 3.10.

Appendix II: Proof of Lemma 3.9

Up to composing \( p\) by an element of \(\text {GL}_2^+ (\mathbb R)\) we can assume that the image of \(p\) is the set of Gaussian integers. By Lemma 9.1, we can assume that there exists a symplectic basis \(a_1, b_1, \ldots , a_g, b_g\) in which the period \(p\) has the following form:

$$\begin{aligned} p(a_1)=p(a_2)=1, \ p(b_1)=p(b_2)=i \text { and } p(a_k)= p(b_k)=0 \text { for } k\ge 3. \end{aligned}$$

This basis permits to identify \( H_1(\Sigma _g,\mathbb Z) \) with \( \mathbb Z^{2g} \) equipped with the symplectic form

$$\begin{aligned} u\cdot v = u_1 v_2 - u_2 v_1 +\cdots + u_{2g-1} v_{2g} - u_{2g} v_{2g-1}, \end{aligned}$$

and the group \(\text {Aut}(H_1(\Sigma _g, \mathbb Z))\) with \(\text {Sp}(2g,\mathbb Z)\). In these coordinates we have

$$\begin{aligned} p (u_1, \ldots , u_{2g}) = (u_1+u_3) + (u_2+u_4)i . \end{aligned}$$

The form \( u_1+u_3 \in (\mathbb Z^{2g})^*\) is dual to the vector \(P_1= -(b_1+b_2)\) (meaning that \(u_1+u_3= - (b_1+b_2)\cdot u\)), whereas the form \( u_2+u_4\in (\mathbb Z^{2g})^*\) is dual to \(P_2= a_1+a_2\) (meaning \(u_2+u_4= (a_1+a_2)\cdot u\)). We then have that an edomorphism of \({\mathbb {Z}}^{2g}\) given by a matrix \(M\) stabilizes \(p\) if and only if \(M (P_k)=P_k\) for \(k=1,2\), and similarly an endomorphism of \((\mathbb Z/2\mathbb Z)^{2g}\) given by a matrix \(M[2]\) stabilizes p[2] if and only if \(M[2] (P_k[2])=P_k[2]\) for \(k=1,2\). The condition can be read in the columns of the matrices. If \(C_k\) (resp. \(C_k[2]\)) denotes the kth columun of M (resp. M[2]) then

$$\begin{aligned}{} & {} C_1 +C_3= a_1+a_2 \text { and } C_2 +C_4=b_1+b_2 \end{aligned}$$
(63)
$$\begin{aligned}{} & {} (\text {resp. } C_1[2]+C_3[2]=a_1[2]+a_2[2] \text { and } C_2[2]+C_4[2]=b_1[2]+b_2[2]).\end{aligned}$$
(64)

Let \(M[2]\in \text {Stab} _{\text {Sp}(2g,\mathbb Z / 2\mathbb Z)} (p[2])\) a symplectic isomorphism whose columns \( C_1[2] , \ldots , C_{2g}[2]\) satisfy (64). Let \( C_k' \in \mathbb Z^{2g}\) be representatives of the classes \(C_k[2]\in (\mathbb Z / 2\mathbb Z)^{2g}\), for \(k=1,\dots , 2g\), and \(M' \) be the square matrix whose columns are the \(C_k'\). Up to changing the representative in each class, we can suppose (63) is valid for the columns of \(M'\). We will assume in the sequel that these equations are always satisfied.

Our goal is to modify the vectors \( C_k '\) by defining

$$\begin{aligned} C_k := C_k ' + 2E_k, \text { with } E_k\in \mathbb Z^{2g} \end{aligned}$$

in such a way that the matrix \( M:= (C_1,\ldots , C_{2g})\) not only stabilizes p (satisfying (63)) but also belongs to \( \text {Sp} (2g,\mathbb Z)\). It is therefore necessary to impose

$$\begin{aligned}{} & {} E_1+E_3=0 \text { and } E_2+E_4=0,\end{aligned}$$
(65)
$$\begin{aligned}{} & {} C_1,\ldots , C_{2g}\text { forms a symplectic basis of } {\mathbb {Z}}^{2g}, \end{aligned}$$
(66)

Main step: construction of \(C_1,C_2,C_3,C_4\). The conditions we need to satisfy are

  1. (1)

    \( C_1\cdot C_2= 1\),

  2. (2)

    \( C_1\cdot C_3=0 \), or equivalently \( C_1 \cdot (a_1+a_2) =0\),

  3. (3)

    \( C_1\cdot C_4= 0\), or equivalently, knowing (1): \( C_1\cdot (b_1+b_2)=C_1\cdot C_2 =1\),

  4. (4)

    \(C_2\cdot C_3=0\), or equivalently, knowing (1): \( C_2 \cdot (a_1+a_2)=C_2\cdot C_1=-1\),

  5. (5)

    \(C_2\cdot C_4=0\), or equivalently, \( C_2 \cdot (b_1+b_2) =0\),

  6. (6)

    \(C_3\cdot C_4=1\), or equivalently, \( ((a_1+a_2) -C_1)\cdot ((b_1+b_2) - C_2) =1\).

If (63) and (1)–(5) are true, we have that (6) is automatically satisfied

$$\begin{aligned}{} & {} ((a_1+a_2) -C_1)\cdot ((b_1+b_2) - C_2) \\{} & {} \quad =(a_1+a_2)\cdot (b_1+b_2) -(a_1+a_2) \cdot C_2 -C_1 \cdot (b_1+b_2) +C_1\cdot C_2= 2-1-1+1=1. \end{aligned}$$

Let us first find \(C_1, C_2, C_3, C_4\) so that conditions (2)–(5) are satisfied. In real and imaginary coordinates this is equivalent to

$$\begin{aligned} p (E_1) = \left( \frac{1-C_1'\cdot (b_1+b_2) }{2}, \frac{C_1' \cdot (a_1+a_2)}{2} \right) , \end{aligned}$$

and

$$\begin{aligned} p(E_2) = \left( \frac{(b_1+b_2)\cdot C_2' }{2} , \frac{1+C_2'\cdot (a_1+a_2)}{2} \right) . \end{aligned}$$

Observe that the right hand sides of the last two equations are integers because \(C_k'\)’s are lifts of - the elements of the symplectic basis- \(C_k[2]\)’s and satisfy (63). By surjectivity of \(p\) there exist solutions to the equations (2)–(5), and so we can assume that the \(C_k'\)’s satisfy (2)–(5). We can now replace \(C_k'\) by \( C_k = C_k'+2E_k\) for \(k=1,2\) with

$$\begin{aligned} E_k \in (a_1+a_2)^{\perp } \cap (b_1+b_2) ^\perp \text { for } k=1,2. \end{aligned}$$

to preserve conditions (2)–(5). Condition (1) is equivalent to

$$\begin{aligned} 1= C_1 \cdot (C_2' +2 E_2) = C_1\cdot C_2' + 2 C_1\cdot E_2, \end{aligned}$$

and we know that \( C_1 \cdot C_2'\) is odd. So we are done if we can choose \(C_1\) so that the map

$$\begin{aligned} E_2 \in (a_1+a_2)^{\perp } \cap (b_1+b_2) ^\perp \mapsto C_1\cdot E_2\in \mathbb Z \end{aligned}$$
(67)

is onto. We have \( (a_1+a_2)^{\perp } \cap (b_1+b_2) ^\perp = \mathbb Z (a_1-a_2) +\mathbb Z (b_1-b_2) +\sum _{k\ge 3} \mathbb Z a_k+\mathbb Z b_k\) So the map (67) is onto if and only if

$$\begin{aligned} \text {gcd} ( C_1 \cdot (a_1-a_2) , C_1 \cdot (b_1-b_2) , C_1 \cdot a_3 , C_1\cdot b_3, \ldots , C_1 \cdot a_g , C_1\cdot b_g) =1. \end{aligned}$$
(68)

However, observe the following:

  • \(C_1 \cdot (b_1-b_2)= C_1 \cdot (b_1+b_2) [2]= 1[2] \) is odd, and

  • If \(g\ge 3\), choosing \(E_1\) appropriately, we can assume that the value of \( C_1 \cdot a_3 \) is either \( 1\text { or } 2\).

So (68) is satisfied with these choices of \(E_1\). Hence the main step is achieved.

To conclude the proof of Lemma 3.9, we need to construct the columns \( C_5,\ldots , C_{2g}\). We will inductively find the pair \( C_5, C_6\), then the pair \( C_7, C_8\), etc.. Let us construct the first one. We need to find \(E_5, E_6\) so that

$$\begin{aligned} C_5, C_6 \in \left( C_1, C_2, C_3, C_4 \right) ^\perp \text { and } C_5\cdot C_6 = 1. \end{aligned}$$

This means that the equations

$$\begin{aligned} (E_k\cdot C_1, \ldots , E_k \cdot C_4) =-\frac{1}{2} (C_k'\cdot C_1,\ldots ,C_k'\cdot C_4) \text { for } k=5,6 \end{aligned}$$
(69)

and

$$\begin{aligned} C_5\cdot (C_6'+ 2 E_6) = 1 \end{aligned}$$
(70)

hold. We can find \(E_5\) and \(E_6\) so that (69) is satisfied, since \(C_1, C_2, C_3, C_4\) is a symplectic family (i.e. \(C_1\cdot C_2=C_3\cdot C_4=1\) and other products are zero). So we can assume that \(C_5'\) and \(C_6'\) belong to the orthogonal of \( C_1, C_2, C_3, C_4\), which is a symplectic submodule of \(\mathbb Z^{2g}\) isomorphic to \(\mathbb Z^{2g-4}\) with the canonical symplectic product. In these coordinates, one of the coordinates of \( C_5'\) is odd (since \(C_5[2]\), the reduction of \(C_5'\) modulo 2, is non zero) so by adding to \(C_5\) an even vector of \(\mathbb Z^{2g-4}\) one can assume that one of the coordinates of \( C_5\) is equal to \(1\). Hence, equation (70) being equivalent to \( C_5\cdot E_6= \frac{-C_5\cdot C_6'}{2}\) and the product \( C_5\cdot C_6\) being even, equation (70) can be solved for some suitable choice of \(E_6\). We get \(C_5\) and \(C_6\) in this way. For the construction of the other pairs by induction, the argument is similar.

The first part of the lemma follows, namely the surjectivity of the map \(\text {Stab}_{\text {Aut}(H_1(\Sigma _g, \mathbb Z))}(p) \rightarrow \text {Stab}_{\text {Aut}(H_1(\Sigma _g, \mathbb Z / 2\mathbb Z))}(p[2])\).

For the second part, it suffices to remark that the stabilizer of \(p[2]\) in \( \text {Aut} (H_1(\Sigma _g, \mathbb Z / 2\mathbb Z))\) is the stabilizer of a pair of non colinear elements of \(H^1(\Sigma _g, \mathbb Z / 2\mathbb Z)\) that do not intersect each other. The number of elements in this group is given by the announced formula, by elementary considerations.

Appendix III: A result in Picard–Lefschetz theory

We prove here a result which is a well-known consequence of Picard–Lefschetz theory, but which we cannot find as such in the literature.

Let \(h : S\rightarrow C\) be a holomorphic map from a compact complex surface \(S\) to a compact complex curve \(C\). We assume both \(S\) and \(C\) are non singular, that the critical points of \(h\) are non degenerate, and that the fibers of \(h\) are connected.

Lemma 11.1

For every \(c\in C\), the sequence

$$\begin{aligned} \pi _1 (h^{-1} (c) ) \rightarrow \pi _1 (S) \rightarrow \pi _1 (C) \end{aligned}$$
(71)

where the map on the left is induced by inclusion and the one on the right by \(h_*\), is exact.

Proof

Let us first prove the claim for \(c\) a regular value of \(h\). To do so, given a lift \(w\in h^{-1} (c)\), we will use some useful loops \(\tilde{\mu _i}: [0,1] \rightarrow S\) starting and ending at \(w\) that do not intersect the critical fibers of \(h\).

Denote by \( c_i \), \(i= 1,\ldots , r\) the critical values of \(h\). At any regular point \(p_i\) of the curve \( h^{-1} (c_i )\), the map \(h\) induces a biholomorphism between a compact disc \(D_i\) transversal to \( h^{-1} (c_i \) at \(p_i\) and a neighborhood \(\delta _i\) of \(c_i \) in \(C\). We can assume that the discs \(\delta _i := h(D_i)\) are disjoint. Given a point \( w_i \in \partial D_i\), let \(\gamma _i:[0,1] \rightarrow E{\setminus } \cup _i \text {Int} (\delta _i) \) be a family of paths with origin \(c\) and end point \( h(w_i) \), having the property of being disjoint appart from their origin. Since \(h\) induces a \(C^\infty \)-fibration with connected fibers from \( S {\setminus } \cup _i h^{-1} (c_i ) \) to \( C {\setminus } \{c_i \} \), given a lift \(w\in h^{-1} (c)\), we can lift the path \( \gamma _i\) to a path \(\tilde{\gamma _i} \) starting at \( w \) and ending at \( w_i\). We denote by \(\mu _i\) the concatenation \(\gamma _i \star \partial \delta _i\star \gamma _i^{-1} \) and by \(\tilde{\mu _i}\) its lift \( \tilde{\gamma _i} \star \partial D_i \star \tilde{\gamma _i}^{-1}\). Notice the following two facts:

  1. (1)

    \(\tilde{\mu _i}\) is homotopically trivial in \(S \),

  2. (2)

    the representatives of the loops \(\mu _i\) in \(\pi _1 (C{\setminus } \{c_i \}, c)\) generate the kernel of the map

    $$\begin{aligned} \pi _1 (C{\setminus } \{z_i \}, c)\rightarrow \pi _1 (C, c) \end{aligned}$$
    (72)

    induced by inclusion.

(The first is clear, and the second comes from the fact that the paths \(\gamma _i\) are disjoint appart from their origin.)

We are now in a position to prove the claim in the case of a regular fiber. Consider an element in the kernel of the map \(h_\star : \pi _1 (S,w) \rightarrow \pi _1(C, c)\) represented by a loop \( \varepsilon : [0,1] \rightarrow S \) starting and ending at \(w\). Up to homotopy, we can assume that \(\varepsilon \) intersects none of the critical fibers of \(h\). The image \( h \circ \varepsilon \) is then a loop starting and ending at \(c\) which defines an element of the fundamental group of \( \pi _1 (C {\setminus } \{ z_i \} , c ) \) belonging to the kernel of the map (72). Hence, \(h\circ \varepsilon \) is homotopic to a word \(W (\mu _1, \ldots , \mu _r) \). Consider the path \(\varepsilon ' = \varepsilon \star W (\tilde{\mu _1}, \ldots , \tilde{\mu _r}) ^{-1} \); by construction,

  1. (i)

    \(\varepsilon ' \) is homotopic to \(\varepsilon \) is \(S\),

  2. (ii)

    it intersects none of the critical fibers of \(h\),

  3. (iii)

    its image \( h \circ \varepsilon '\) with fixed extremities \(c\) is homotopic in \( C{\setminus } \{c_i \} \) to the constant path \( c\).

Since the restriction of \(h\) induces a locally trivial \(C^\infty \) fibration from \(S {\setminus } \cup _i h^{-1} (c_i ) \) to \(C {\setminus } \{c_i \}\)—this is a proper submersion, so this is Ehresmann’s fibration theorem—we can lift the homotopy found in (iii) to a homotopy in \( S {\setminus } \cup _i h^{-1} (c_i ) \) between \(\varepsilon '\) and a loop contained in the fiber \( h^{-1} ( c)\). This establishes that \(\varepsilon \) is homotopic with fixed extremities to a loop contained in the fiber \( h^{-1} (c)\), and ends the proof of the claim in the case of a regular fiber.

It remains to prove the claim for the singular fibers of \(h\). For each \(i\), there exists a neighborhood \(U\) of \(h^{-1} (c_i)\) that retracts by deformation on \( h^{-1} (c_i)\) (see [ACG11, Chapter X, Section 9]); in particular, the images of \(\pi _1 (h^{-1} (c_i))\) and \(\pi _1 (U)\) in \(\pi _1 (S)\) coincide. Since \(U\) contains a regular fiber, we know from the prerceeding discussion that the image of \(\pi _1 (U)\) in \(\pi _1 (S)\) contains \( \text {Ker} (h_\star )\). On the other hand, the image of \( \pi _1 (h^{-1}(c_i))\) in \(\pi _1 (S)\) is contained in \( \text {Ker} (h_\star )\) since \(h\) is constant on \(h^{-1} (c_i)\), so the sequence (71) is exact for the critical value \(c=c_i\), and the Lemma follows. \(\square \)

Corollary 11.2

For every \(c\in C\), the sequence

$$\begin{aligned} H_1 (h^{-1} (c),\mathbb Z ) \rightarrow H _1 (S,\mathbb Z) \rightarrow H_1 (C,\mathbb Z) \end{aligned}$$
(73)

where the map on the left is induced by inclusion and the one on the right by \(h_*\), is exact.

Proof

This is a consequence of Lemma 11.1 and Hurewicz’ theorem. \(\square \)

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

Calsamiglia, G., Deroin, B. & Francaviglia, S. A transfer principle: from periods to isoperiodic foliations. Geom. Funct. Anal. 33, 57–169 (2023). https://doi.org/10.1007/s00039-023-00627-w

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00039-023-00627-w

Keywords and phrases

Mathematics Subject Classification

Navigation