Skip to main content
Log in

Holomorphic anomaly equations and the Igusa cusp form conjecture

  • Published:
Inventiones mathematicae Aims and scope

Abstract

Let S be a K3 surface and let E be an elliptic curve. We solve the reduced Gromov–Witten theory of the Calabi–Yau threefold \(S \times E\) for all curve classes which are primitive in the K3 factor. In particular, we deduce the Igusa cusp form conjecture. The proof relies on new results in the Gromov–Witten theory of elliptic curves and K3 surfaces. We show the generating series of Gromov–Witten classes of an elliptic curve are cycle-valued quasimodular forms and satisfy a holomorphic anomaly equation. The quasimodularity generalizes a result by Okounkov and Pandharipande, and the holomorphic anomaly equation proves a conjecture of Milanov, Ruan and Shen. We further conjecture quasimodularity and holomorphic anomaly equations for the cycle-valued Gromov–Witten theory of every elliptic fibration with section. The conjecture generalizes the holomorphic anomaly equations for elliptic Calabi–Yau threefolds predicted by Bershadsky, Cecotti, Ooguri, and Vafa. We show a modified conjecture holds numerically for the reduced Gromov–Witten theory of K3 surfaces in primitive classes.

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

Similar content being viewed by others

Notes

  1. The Katz–Klemm–Vafa conjecture usually refers to the result proven in [44].

  2. Since S is holomorphic symplectic the (ordinary) virtual fundamental class vanishes. The theory is non-trivial only after reduction [35].

  3. See Gritsenko–Nikulin [16].

  4. We assume here that gn lie in the stable range i.e. take only those values for which the moduli spaces \({\overline{M}}_{g,n}\) and \({\overline{M}}_{g,n}(E,d)\) are Deligne–Mumford stacks. We follow the same convention throughout the paper. In all equations or diagrams or sums we assume (gn) to lie in the range where all moduli spaces are Deligne–Mumford stacks.

  5. Our argument is independent of [39, 40] and in fact yields a new proof.

  6. See [21, Equations (3.8) and (3.9)] and [1] for a discussion in the elliptic case.

  7. The examples in [1] suggest that the congruence subgroup should be \(\Gamma _1(N)\) in general. For elliptic orbifold \(\mathbb {P}^1\)s we have strictly \(\Gamma (N)\) modular forms; however this is not a counterexample since the target is an orbifold. We leave determining the exact congruence subgroup for elliptic fibrations without a section to a later date.

  8. A Calabi–Yau fibration is a flat connected morphism of non-singular projective varieties whose general fiber has trivial canonical class.

  9. We thank B. Poonen for discussions on this point.

  10. See [18, Section 0.3.2] for the definition of a stable graph.

  11. This corresponds to the following convention: Assume the dual graph of the target \(C_n\) is depicted in the plane with labels increasing in clockwise direction as in Fig. 1. Let \(e = \{ h,h' \}\) be an edge with \(\mathbf {w}(h) > 0\) and \(h \in v_i\) and \(h' \in v_j\). Then the chain corresponding to e ’travels’ clockwise from \(P_i\) to \(P_j\) around the cycle.

  12. The factor of 2 in the first term above cancels with the factor of 2 from the deleted loop’s contribution to \(\mathrm {Aut}(\Gamma )\).

  13. Polynomiality here follows from the polynomiality of the double ramification cycle [18, 45].

  14. The log canonical class of the pair (SE) is non-zero.

  15. The evaluation \(\langle \tau _1(F) \rangle _1 = \frac{2 C_2}{\Delta }\) follows also from the holomorphic anomaly equation.

  16. The variables \(z\in {\mathbb {C}}\) and \(\tau \in {\mathbb {H}}\) of [10] are related to (uq) by \(u = 2 \pi z\) and \(q = e^{2 \pi i \tau }\).

  17. The arguments of [3] carry over to the reduced virtual class. Alternatively, we may use a degeneration argument similar to [35, Proposition 5] to reduce to the standard case.

  18. There exist relations among the generators of \(\mathsf {E}\) but they do not involve \(C_2\). The ring \(\mathsf {E}\) is free over \({\mathbb {Q}}[ C_4, C_6, \wp ^{(k)}(z)|k\ge 0]\) and the derivative with respect to \(C_2\) is well-defined.

  19. The first part of Theorem 7 can also be found in work of Goujard and Möller [13]. Our argument gives a new proof of their result. We thank M. Raum for pointing out this connection.

  20. A sequence \(x_1, x_2, x_3, \ldots \) is non-recurring if \(x_i \ne x_j\) for all \(i \ne j\).

  21. Since F and \(\mathsf {A}\) are both 1-periodic we may assume there is no shift by an integer.

References

  1. Alim, M., Scheidegger, E.: Topological strings on elliptic fibrations. Commun. Number Theory Phys. 8(4), 729–800 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  2. Beauville, A.: Counting rational curves on \(K3\) surfaces. Duke Math. J. 97(1), 99–108 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  3. Behrend, K.: The product formula for Gromov–Witten invariants. J. Algebraic Geom. 8(3), 529–541 (1999)

    MathSciNet  MATH  Google Scholar 

  4. Behrend, K.: Donaldson–Thomas type invariants via microlocal geometry. Ann. Math. (2) 170(3), 1307–1338 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bryan, J.: The Donaldson–Thomas theory of \(K3 \times E\) via the topological vertex. arXiv:1504.02920

  6. Bryan, J., Leung, N.C.: The enumerative geometry of \(K3\) surfaces and modular forms. J. Am. Math. Soc. 13(2), 371–410 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  7. Bershadsky, M., Cecotti, S., Ooguri, H., Vafa, C.: Kodaira-Spencer theory of gravity and exact results for quantum string amplitudes. Comm. Math. Phys. 165(2), 311–427 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bryan, J., Oberdieck, G., Pandharipande, R., Yin, Q.: Curve counting on abelian surfaces and threefolds. Algebr. Geom. arXiv:1506.00841 (to appear)

  9. Buryak, A., Shadrin, S., Spitz, L., Zvonkine, D.: Integrals of \(\psi \)-classes over double ramification cycles. Am. J. Math. 137(3), 699–737 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  10. Eichler, M., Zagier, D.: The Theory of Jacobi Forms, Volume 55 of Progress in Mathematics. Birkhäuser Boston Inc., Boston, MA (1985)

    Book  MATH  Google Scholar 

  11. Faber, C., Pandharipande, R.: Relative maps and tautological classes. J. Eur. Math. Soc. 7(1), 13–49 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  12. Faber, C., Pandharipande, R.: Tautological and non-tautological cohomology of the moduli space of curves. In: Handbook of Moduli, vol. 1, pp. 293–330. Adv. Lect. Math. (ALM), 24, INT Press, Somerville (2013)

  13. Goujard, E., Möller, M.: Counting Feynman-like graphs: quasimodularity and Siegel–Veech weight. J. Eur. Math. Soc. arXiv:1609.01658 (to appear)

  14. Graber, T., Pandharipande, R.: Constructions of nontautological classes on moduli spaces of curves. Mich. Math. J. 51(1), 93–109 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  15. Greer, F. (in preparation)

  16. Gritsenko, V.A., Nikulin, V.V.: Siegel automorphic form corrections of some Lorentzian Kac–Moody Lie algebras. Am. J. Math. 119, 181–224 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  17. Janda, F.: Gromov–Witten theory of target curves and the tautological ring. Mich. Math. J. 66(4), 683–698 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  18. Janda, F., Pandharipande, R., Pixton, A., Zvonkine, D.: Double ramification cycles on the moduli spaces of curves. Publ. Math. Inst. Hautes Études Sci. 125, 221–266 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  19. Kaneko, M., Zagier, D.: A generalized Jacobi theta function and quasimodular forms in the moduli space of curves, (Texel Island, 1994), pp. 165–172, Prog. Math., 129, Birkhäuser Boston, Boston (1995)

  20. Katz, S., Klemm, A., Vafa, C.: M-theory, topological strings, and spinning black holes. Adv. Theor. Math. Phys. 3, 1445–1537 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  21. Klemm, A., Manschot, K., Wotschke, T.: Quantum geometry of elliptic Calabi–Yau manifolds. Commun. Number Theory Phys. 6(4), 849–917 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  22. Kool, M., Thomas, R.: Reduced classes and curve counting on surfaces I: theory. Algebr. Geom. 1(3), 334–383 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  23. Lee, Y.-P., Qu, F.: A product formula for log Gromov–Witten invariants. arXiv:1701.04527

  24. Lho, H., Pandharipande, R.: Stable quotients and the holomorphic anomaly equation. arXiv:1702.06096

  25. Li, J.: Stable morphisms to singular schemes and relative stable morphisms. J. Differ. Geom. 57(3), 509–578 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  26. Li, J.: A degeneration formula for Gromov–Witten invariants. J. Differ. Geom. 60(2), 199–293 (2002)

    Article  Google Scholar 

  27. Maulik, D.: Gromov–Witten theory of \(A_n\)-resolutions. Geom. Top. Pattern 13, 1729–1773 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  28. Maulik, D., Pandharipande, R.: A topological view of Gromov–Witten theory. Topology 45(5), 887–918 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Maulik, D., Pandharipande, R.: Gromov–Witten theory and Noether–Lefschetz theory. In: A Celebration of Algebraic Geometry, Clay Mathematics Proceedings, vol. 18, pp. 469–507. AMS (2010)

  30. Maulik, D., Pandharipande, R., Thomas, R.P.: Curves on \(K3\) surfaces and modular forms, with an appendix by A. Pixton. J. Topol. 3(4), 937–996 (2010)

  31. Milanov, T., Ruan, Y., Shen, Y.: Gromov–Witten theory and cycle-valued modular forms. Journal fur die Reine und Angewandte Mathematik, vol 2015 (2015)

  32. Oberdieck, G.: Gromov–Witten invariants of the Hilbert scheme of points of a K3 surface. Geom. Topol. 22(1), 323–437 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  33. Oberdieck, G.: Gromov–Witten theory of \(\text{K3} \times \mathbb{P}^1\) and quasi-Jacobi forms. Int. Math. Res. Not. arXiv:1605.05238 (to appear)

  34. Oberdieck, G.: On reduced stable pair invariants. Math. Z. arXiv:1605.04631 (to appear)

  35. Oberdieck, G., Pandharipande, R.: Curve counting on \(K3\times E\), the Igusa cusp form \(\chi _{10}\), and descendent integration. In: Faber, C., Farkas, G., van der Geer, G. (eds.) K3 Surfaces and Their Moduli. Birkhauser Prog. Math. 315, 245–278 (2016)

  36. Oberdieck, G., Shen, J.: Curve counting on elliptic Calabi–Yau threefolds via derived categories. JEMS. arXiv:1608.07073 (to appear)

  37. Oberdieck, G., Shen, J.: Reduced Donaldson–Thomas invariants and the ring of dual numbers. arXiv:1612.03102

  38. Oberdieck, G., Pixton, A.: Gromov–Witten theory of elliptic fibrations: Jacobi forms and holomorphic anomaly equations. arXiv:1709.01481

  39. Okounkov, A., Pandharipande, R.: Gromov–Witten theory, Hurwitz theory, and completed cycles. Ann. Math. (2) 163(2), 517–560 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  40. Okounkov, A., Pandharipande, R.: Virasoro constraints for target curves. Invent. Math. 163(1), 47–108 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  41. Pandharipande, R., Pixton, A.: Gromov–Witten/pairs descendent correspondence for toric 3-folds. Geom. Topol. 18(5), 2747–2821 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  42. Pandharipande, R., Pixton, A.: Gromov-Witten/pairs correspondence for the quintic 3-fold. J. Amer. Math. Soc. 30(2), 389–449 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  43. Pandharipande, R., Thomas, R.P.: Curve counting via stable pairs in the derived category. Invent. Math. 178(2), 407–447 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  44. Pandharipande, R., Thomas, R.P.: The Katz–Klemm–Vafa conjecture for K3 surfaces. In: Forum of Mathematics, Pi, vol. 4 (2016), 111 pp

  45. Pixton, A., Zagier, D. (in preparation)

  46. Toda, Y.: Stability conditions and curve counting invariants on Calabi–Yau 3-folds. Kyoto J. Math. 52(1), 1–50 (2012)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank J. Bryan, F. Janda, D. Maulik, R. Pandharipande, J. Shen and Q. Yin for useful discussions on curve counting on K3 surfaces and elliptic curves. We are also very grateful to T. Milanov, Y. Ruan and Y. Shen for discussions about their paper [31]. We would also like to thank the anonymous referees for their comments.

The second author was supported by a fellowship from the Clay Mathematics Institute.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Aaron Pixton.

Appendices

Appendix A. Elliptic functions and quasimodular forms

1.1 Overview

We prove that for certain multivariate elliptic functions F, the constant term of the Fourier expansion of F (in the elliptic parameter) is a quasimodular form. We also calculate the \(C_2\)-derivative of these quasimodular forms. In Sect. A.3 we treat the single variable case as a warm-up for the general case which appears in Sect. A.4. The main result of this appendix is Theorem 7.

1.2 Preliminaries

Let \(z \in {\mathbb {C}}\) and \(\tau \in {\mathbb {H}}\), where \({\mathbb {H}}= \{ z \in {\mathbb {C}}| \mathrm {Im}(z)>0 \}\) is the upper half plane. We will use the auxiliary variables

$$\begin{aligned} w = 2 \pi i z, \quad p = e^{2 \pi i z}, \quad q = e^{2 \pi i \tau }. \end{aligned}$$

The operator of differentiation with respect to z is denoted

$$\begin{aligned} \partial _z = \frac{1}{2 \pi i} \frac{d}{dz} = \frac{d}{dw} = p \frac{d}{dp} \end{aligned}$$

and for the kth derivative of a function f(z) we write

$$\begin{aligned} f^{(k)}(z) = \partial _z^k f(z). \end{aligned}$$

For any meromorphic function f(z) we let \([ f(z) ]_{(z-a)^{\ell }}\) denote the coefficient of \((z-a)^{\ell }\) in the Laurent expansion around a. The residue at a is

$$\begin{aligned} \mathrm {Res}_{z=a} f(z) \, = \, \big [ f(z) \big ]_{(z-a)^{-1}}. \end{aligned}$$

If \(f(z) = g(z) h(z)\) where h(z) is regular at a we have

$$\begin{aligned} \mathrm {Res}_{z=a} f(z) = \sum _{k \ge 1} \big [ g(z) \big ]_{(z-a)^{-k}} \frac{(2 \pi i)^{k-1} h^{(k-1)}(a)}{(k-1)!}. \end{aligned}$$
(63)

1.3 Elliptic functions

Consider the Eisenstein series \(C_{2k}(\tau )\) defined in (5) as functions on \({\mathbb {H}}\) under the change of variables \(q= e^{2 \pi i \tau }\). Consider also the Weierstraß function \(\wp (z)\) which has Laurent expansion

$$\begin{aligned} \wp (z) = \frac{1}{12} + \frac{p}{(1-p)^2} + \sum _{d \ge 1} \sum _{k | d} k (p^k - 2 + p^{-k}) q^{d} \end{aligned}$$
(64)

in the region \(0< |q|< |p| < 1\), and has Laurent expansion

$$\begin{aligned} \wp (z) = \frac{1}{w^2} + \sum _{k \ge 2} (2k-1) 2k C_{2k}(\tau ) w^{2k-2} \end{aligned}$$

at \(w=0\).

Let \(\mathsf {E}\) be the ring generated by quasimodular forms and derivatives of the Weierstraß function,

$$\begin{aligned} \mathsf {E}= {\mathbb {Q}}\left[ C_2(\tau ), C_4(\tau ), C_6(\tau ), \wp ^{(k)}(z) \big | \, k \ge 0 \, \right] . \end{aligned}$$

The ring is graded by weight:

$$\begin{aligned} \mathsf {E}= \bigoplus _{k \ge 0} \mathsf {E}_{k}, \end{aligned}$$

where \(C_{k}\) has weight k and \(\wp ^{(k)}(z)\) has weight \(2+k\). We also let

$$\begin{aligned} \frac{d}{dC_2} : \mathsf {E}\rightarrow \mathsf {E}\end{aligned}$$

be the formal differentiation with respect to the generator \(C_2\).Footnote 18

Every \(F(z) \in \mathsf {E}\) admits a Fourier expansion in the region \(0< |q|< |p|<1\),

$$\begin{aligned} F(z) = \sum _{n \in {\mathbb {Z}}} a_n(\tau ) p^n. \end{aligned}$$

The constant term in the expansion is denoted by

$$\begin{aligned} \big [ F(z) \big ]_{p^0} = a_0(\tau ). \end{aligned}$$

As a warm-up for the general case we prove the following proposition.

Proposition 8

For every \(F \in \mathsf {E}_k\) the series \(\big [ F \big ]_{p^0}\) is a quasimodular form of weight k and we have

$$\begin{aligned} \frac{d}{dC_2} \Big [ \, F \, \Big ]_{p^0} = \left[ \frac{d}{dC_2} F \right] _{p^0} - 2 \big [ \, F \, \big ]_{w^{-2}}. \end{aligned}$$

Consider the function

$$\begin{aligned} \mathsf {A}(z)&= - \frac{1}{2} - \sum _{m \ne 0} \frac{p^m}{1-q^m} \\&= \frac{1}{w} - \sum _{\ell \ge 1} 2\ell C_{2\ell }(q) w^{2 \ell - 1}, \end{aligned}$$

where the expansion in pq is taken in the region \(0< |q|< |p| < 1\). For the proof of the Proposition we require the following Lemma.

Lemma 18

\(\mathsf {A}(z + \lambda \tau + \mu ) = \mathsf {A}(z) - \lambda \) for every \(\lambda , \mu \in {\mathbb {Z}}\).

Proof

We have \(\mathsf {A}(z) = \partial _z \log \Theta (z)\) where \(\Theta \) is the Jacobi theta function

$$\begin{aligned} \Theta (z) = (p^{1/2} - p^{-1/2}) \prod _{m \ge 1} \frac{ (1-pq^m) (1-p^{-1}q^m)}{ (1-q^m)^2 }. \end{aligned}$$

A direct check using this definition shows

$$\begin{aligned} \Theta (z + \lambda \tau + \mu ) = (-1)^{\lambda + \mu } p^{-\lambda } q^{-\lambda ^2/2} \Theta (z) \end{aligned}$$

for all \(\lambda , \mu \in {\mathbb {Z}}\) which implies the claim. \(\square \)

Proof of Proposition 8

We have

$$\begin{aligned} \big [ F(z) \big ]_{p^0} = \int _{\mathsf {C}_a} F(z) \mathop {}\!\mathrm {d}{z} \end{aligned}$$

where \(\mathsf {C}_a\) is the line segment from a to \(a+1\) for some \(a \in {\mathbb {C}}\) with \(0< \mathrm {Im}(a) < \mathrm {Im}(\tau )\). Since F(z) is periodic, i.e.

$$\begin{aligned} F(z + \lambda \tau + \mu ) = F(z) \end{aligned}$$

for every \(\lambda , \mu \in {\mathbb {Z}}\), we may instead assume \(-\mathrm {Im}(\tau )< \mathrm {Im}(a) < 0\).

By Lemma 18 the function \(f(z) = F(z) \cdot \mathsf {A}(z)\) satisfies

$$\begin{aligned} f(z+1) = f(z), \quad f(z+\tau ) = f(z) - F(z). \end{aligned}$$

Hence we may replace the integral of F over \(\mathsf {C}_a\) by the integral of \(F \cdot \mathsf {A}\) over the boundary of the fundamental domain \(B_a\) depicted in Fig. 2,

$$\begin{aligned} \big [ F(z) \big ]_{p^0} = \oint _{B_a} F(z) \cdot \mathsf {A}(z) \mathop {}\!\mathrm {d}{z}. \end{aligned}$$
Fig. 2
figure 2

The closed path \(B_{a}\)

Since both F and \(\mathsf {A}\) have poles inside \(B_a\) only at 0, an application of the residue theorem gives

$$\begin{aligned} \big [ F(z) \big ]_{p^0}&= \big [ F(z) \cdot \mathsf {A}(z) \big ]_{w^{-1}} \\&= [ F ]_{w^0} - \sum _{\ell \ge 1} 2 \ell C_{2\ell }(\tau ) [ F(z) ]_{w^{-2 \ell }}. \end{aligned}$$

An inspection of the Laurent series of F(z) yields now both claims. \(\square \)

1.4 Multiple variables

Let \(n \ge 2\) and \(z = (z_1, \ldots , z_n) \in {\mathbb {C}}^n\), and denote

$$\begin{aligned} w_a = 2 \pi i z_a, \quad p_a = e^{2 \pi i z_a}, \quad a \in \{ 1, \ldots , n \}. \end{aligned}$$

Every permutation \(\sigma \in S_n\) determines a region \(U_{\sigma } \subset {\mathbb {C}}^n\) by requiring

$$\begin{aligned} \mathrm {Im}(\tau )> \mathrm {Im}(z_a - z_b) > 0 \end{aligned}$$
(65)

whenever \(\sigma (a) > \sigma (b)\), or equivalently by

$$\begin{aligned} \mathrm {Im}(z_{\sigma ^{-1}(n)})> \cdots> \mathrm {Im}(z_{\sigma ^{-1}(1)}) > \mathrm {Im}(z_{\sigma ^{-1}(n)} - \tau ). \end{aligned}$$
(66)

Consider the ring of multivariate elliptic functions

$$\begin{aligned} \mathsf {ME}= {\mathbb {Q}}\left[ \, C_2(\tau ), C_4(\tau ), C_6(\tau ), \wp ^{(k)}( z_a - z_b )\ \big |\ k \ge 0,\, 1 \le a < b \le n \, \right] . \end{aligned}$$

We assign \(\wp ^{(k)}\) and \(C_k\) the weights \(2+k\) and k respectively and let

$$\begin{aligned} \mathsf {ME}= \bigoplus _{k \ge 0} \mathsf {ME}_k \end{aligned}$$

be the induced grading by weight k. Let

$$\begin{aligned} \frac{d}{dC_2} : \mathsf {ME}\rightarrow \mathsf {ME}\end{aligned}$$

be the formal differentiation with respect to the generator \(C_2\).

Every \(F \in \mathsf {ME}\) has a well-defined Fourier expansion in the region \(U_{\sigma }\),

$$\begin{aligned} F = \sum _{k_1, \ldots , k_n \in {\mathbb {Z}}} a_{k_1 \ldots k_n}(\tau ) p_1^{k_1} \ldots p_n^{k_n}, \quad (z_1, \ldots , z_n) \in U_{\sigma }. \end{aligned}$$

The constant coefficient in this expansion, i.e. the coefficient of \(\prod _i p_i^0\), is denoted

$$\begin{aligned} \left[ F \right] _{p^0, \sigma } = a_{0\ldots 0}(\tau ). \end{aligned}$$

Define the constant coefficient of F averaged over all permutation \(\sigma \),

$$\begin{aligned} \left[ F(z_1, \ldots , z_n) \right] _{p^0} = \frac{1}{n!} \sum _{\sigma \in S_n} \left[ F(z_1, \ldots , z_n) \right] _{p^0, \sigma }. \end{aligned}$$
(67)

The following is the main result of this appendix.Footnote 19

Theorem 7

Let \(F \in \mathsf {ME}_k\). Then the following holds.

  1. (1)

    \(\big [ F(z) \big ]_{p^0, \sigma } \in \mathsf {QMod}_{\le k}\) for every permutation \(\sigma \).

  2. (2)

    \(\big [ F(z) \big ]_{p^0} \in \mathsf {QMod}_k\).

  3. (3)

    We have

    $$\begin{aligned} \frac{d}{dC_2} \Big [ F(z) \Big ]_{p^0} = \left[ \frac{d}{dC_2} F \right] _{p^0} - \sum _{\begin{array}{c} a,b = 1 \\ a \ne b \end{array}}^{n} \left[ (2 \pi i)^2 \mathrm {Res}_{z_a = z_b}\Big ( (z_a - z_b) \cdot F \Big ) \right] _{p^0}. \end{aligned}$$

1.5 Preparations for the proof

We prove a series of results leading up to the proof of Theorem 7 in Sect. A.6.

Lemma 19

Let \(F \in \mathsf {ME}\) and \(\sigma \in S_n\). Then

$$\begin{aligned}{}[ F ]_{p^0, \sigma } = [ F ]_{p^0, \widetilde{\sigma }} \end{aligned}$$

for every cyclic permutation \(\widetilde{\sigma }\) of \(\sigma \).

Proof

Let \((a_1, \ldots , a_n) \in U_{\sigma }\) and let \(\mathsf {C}_{a_i}\) be the line segment from \(a_i\) to \(a_i + 1\) in the \(z_i\)-plane. Then

$$\begin{aligned}{}[ F ]_{p^0, \sigma } = \int _{\mathsf {C}_{a_1}} \cdots \int _{\mathsf {C}_{a_n}} F(z_1, \ldots , z_n) \mathop {}\!\mathrm {d}{z_n} \cdots \mathop {}\!\mathrm {d}{z_{1}}. \end{aligned}$$

Since F is periodic, i.e. \(F(z + \lambda \tau + \mu ) = F(z)\) for every \(\lambda , \mu \in {\mathbb {Z}}^n\), we may replace the integral over \(\mathsf {C}_{a_{\sigma ^{-1}(n)}}\) by the integral over \(\mathsf {C}_{a_{\sigma ^{-1}(n)}-\tau }\). But comparing with (66) this corresponds to taking the constant coefficient of F with respect to a cyclic permutation of \(\sigma \). \(\square \)

For every \(a \ne b\) let \(R_{ab}\) denote the operation of taking the residue in \(z_a = z_b\) written as a right operator,

$$\begin{aligned} f(z_1, \ldots , z_n) R_{ab} := 2 \pi i \cdot \mathrm {Res}_{z_a = z_b} f(z_1, \ldots , z_n). \end{aligned}$$

We also write

$$\begin{aligned} \mathsf {A}_{ab} = \mathsf {A}(z_a - z_b). \end{aligned}$$

Lemma 20

Let \(F(z) \in \mathsf {ME}_k\), and let \(i_1, \ldots , i_m \in \{ 1, \ldots , n \}\) be pairwise distinct. Then for any \(r \ge 0\) we have

$$\begin{aligned} \big ( F(z) \mathsf {A}_{i_1 i_{m}}^{r} \big ) R_{i_1 i_2} R_{i_2 i_3} \cdots R_{i_{m-1} i_{m}} \in \mathsf {ME}_{k+r-(m-1)}. \end{aligned}$$

Proof

Let \(F(z) \in \mathsf {ME}_k\) be a monomial in the generators and consider the splitting

$$\begin{aligned} F(z) = F_{ab}(z_a - z_b) \cdot \widetilde{F}_{ab}(z), \end{aligned}$$

where \(F_{ab}\) is the product of all factors in F of the form \(\wp ^{(s)}(z_a - a_b)\) for some s. In particular, \(\widetilde{F}_{ab}(z)\) is regular at \(z_a = z_b\).

Consider the action of \(R_{ab}\) on \(F(z) \mathsf {A}_{ac}^r\). If \(b \ne c\) we have

$$\begin{aligned} (F \mathsf {A}_{ac}^r) R_{a b} = \sum _{\ell \ge 1} \left[ F_{ab} \right] _{(w_a - w_b)^{-\ell }} \frac{1}{(\ell - 1)!} \partial _{z_a}^{\ell -1} \left( \widetilde{F}_{ab} \mathsf {A}^r_{ac} \right) \Big |_{z_a = z_b}, \end{aligned}$$
(68)

where we have used (63). Since

$$\begin{aligned} \partial _z \mathsf {A}(z) = - \wp (z) - 2 C_2(\tau ) \end{aligned}$$

the right hand side of (68) can be written as a sum of terms

$$\begin{aligned} F'(z) \cdot \mathsf {A}^{r'}_{bc}, \end{aligned}$$

where \(F' \in \mathsf {ME}_{k'}\) with \(k' + r' = k+r-1\). Similarly, if \(b = c\) we have

$$\begin{aligned} ( F(z) \mathsf {A}_{ac}^r ) R_{a c} \in \mathsf {ME}_{k+r-1}. \end{aligned}$$

The claim follows from the steps above and an induction argument. \(\square \)

Let \(\sigma \in S_n\) be a permutation, let

$$\begin{aligned} g_{ab} = {\left\{ \begin{array}{ll} 1 &{} \text { if } \sigma (a) > \sigma (b), \\ 0 &{} \text { otherwise}, \end{array}\right. } \end{aligned}$$

and for all \(x \in {\mathbb {C}}\) and non-negative integers a define

$$\begin{aligned} \left( {\begin{array}{c}x\\ a\end{array}}\right) = \frac{x \cdot (x-1) \cdots (x-a+1)}{a!}. \end{aligned}$$

Proposition 9

Let \(F(z) \in \mathsf {ME}\). Then

$$\begin{aligned}&\big [ F(z) \big ]_{p^0, \sigma } = \sum _{\ell \ge 1} \sum _{i_1, i_2, \ldots , i_{\ell }}\\&\quad \times \left[ F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + \ell - 2 - g_{i_1 i_2} - \ldots - g_{i_{\ell -1} i_{\ell }} \\ \ell -1\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{\ell -1} i_{\ell }} \right] _{p^0, \sigma }, \end{aligned}$$

where the inner sum is over all non-recurringFootnote 20 sequences \(i_1, \ldots , i_{\ell } \in \{ 1, \ldots , n \}\) with endpoints \(i_1=1\) and \(i_{\ell }=n\).

Proof

We argue by induction on L that for every \(L \ge 1\) we have

$$\begin{aligned}&\big [ F(z) \big ]_{p^0, \sigma } =\sum _{\ell = 1}^{L} \sum _{i_1, i_2, \ldots , i_{\ell }}\nonumber \\&\quad \times \left[ F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + \ell - 2 - g_{i_1 i_2} - \ldots - g_{i_{\ell -1} i_{\ell }} \\ \ell -1\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{\ell -1} i_{\ell }} \right] _{p^0, \sigma },\nonumber \\ \end{aligned}$$
(69)

where the inner sum runs over all non-recurring sequences \((i_1, \ldots , i_{\ell })\) such that \(i_1 = 1\) and the following holds:

  • if \(\ell < L\) then \(i_{\ell } = n\),

  • if \(\ell = L\) and \(i_r = n\) then \(r=\ell \).

If \(L = 1\) equality (69) holds by definition. Hence we may assume the claim holds for \(L \ge 1\) and we show the case \(L+1\). Every summand on the right hand side of (69) with \(i_{\ell } \ne n\) is equal to the \(p^0\)-coefficient (in \(U_{\sigma }\)) of

$$\begin{aligned} \int _{\mathsf {C}_{a}} F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + \ell - 2 - g_{i_1 i_2} - \cdots - g_{i_{\ell -1} i_{\ell }} \\ \ell - 1\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{\ell -1} i_{\ell }} \mathop {}\!\mathrm {d}{z_{i_{\ell }}} \end{aligned}$$
(70)

for some \(a \in {\mathbb {C}}\) such that

$$\begin{aligned} (z_1, \ldots , z_{i_{\ell }-1}, a, z_{i_{\ell }+1}, \ldots , z_n) \in U_{\sigma } \end{aligned}$$

and \(\mathsf {C}_a\) is the line segment from a to \(a+1\) in the \(z_{i_{\ell }}\)-plane. Define the function

$$\begin{aligned} \mathsf {H}(z) = F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + \ell - 1 - g_{i_1 i_2} - \cdots - g_{i_{\ell -1} i_{\ell }} \\ \ell \end{array}}\right) R_{i_1 i_2} \cdots R_{i_{\ell -1} i_{\ell }} \end{aligned}$$

Using Lemma 18 and \(\mathrm {Res}_{z=r+s} f(z) = \mathrm {Res}_{z=r}f(z+s)\) repeatedly we find

$$\begin{aligned}&\mathsf {H}(z_1, \ldots , z_{i_{\ell }} + \tau , \ldots , z_n) = \mathsf {H}(z_1, \ldots , z_n) \\&\quad - F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + \ell - 2 - g_{i_1 i_2} - \cdots - g_{i_{\ell -1} i_{\ell }} \\ \ell - 1\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{\ell -1} i_{\ell }}. \end{aligned}$$

Hence arguing as in the proof of Proposition 8 we may replace (70) by an integral of \(\mathsf {H}(z)\) over the box \(B_{a}\) depicted in Fig. 2. The function \(\mathsf {H}\) has possible poles inside \(B_{a}\) only at the pointsFootnote 21

$$\begin{aligned} z_{i_{\ell }} = z_{i_{\ell +1}} + g_{i_{\ell } i_{\ell +1}} \tau \end{aligned}$$

for some \(i_{\ell +1} \notin \{ i_1, \ldots , i_{\ell } \}\). By the residue theorem (70) is therefore

$$\begin{aligned} 2 \pi i \sum _{i_{\ell +1} \notin \{ i_1, \ldots , i_{\ell } \}} \mathrm {Res}_{z_{i_{\ell }} = z_{i_{\ell +1}} + g_{i_{\ell } i_{\ell +1}} \tau } \mathsf {H}(z), \end{aligned}$$

which after moving the shift by \(g_{i_{\ell } i_{\ell +1}} \tau \) inside simplifies to

$$\begin{aligned} \sum _{i_{\ell +1} \notin \{ i_1, \ldots , i_{\ell } \}} F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + \ell - 1 - \sum _{a=1}^{\ell } g_{i_a i_{a+1}} \\ \ell \end{array}}\right) R_{i_1 i_2} \cdots R_{i_{\ell -1} i_{\ell }} R_{i_{\ell } i_{\ell +1}}. \end{aligned}$$

Plugging back into (69) we obtain the case \(L+1\). The induction is complete. \(\square \)

Averaging Proposition 9 over all permutations \(\sigma \) yields the following.

Proposition 10

Let \(F(z) \in \mathsf {ME}\). Then

$$\begin{aligned} \big [ F(z) \big ]_{p^0} = \sum _{m \ge 1} \sum _{i_1=1, i_2, \ldots , i_{m+1}=n} \left[ \left( F \cdot \frac{\mathsf {A}_{1n}^{m}}{m!}\right) R_{i_1 i_2} R_{i_2 i_3} \cdots R_{i_m i_{m+1}} \right] _{p^0}, \end{aligned}$$

where the inner sum runs over all non-recurring sequences \(i_1, \ldots , i_{\ell +1} \in \{ 1, \ldots , n \}\) with endpoints \(i_1=1\) and \(i_{\ell +1}=n\).

Proof

Setting \(\ell = m+1\) in Proposition 9 yields

$$\begin{aligned}&\big [ F(z) \big ]_{p^0, \sigma } = \sum _{m \ge 1} \sum _{i_1, i_2, \ldots , i_{m+1}} \\&\quad \times \left[ F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + m - 1 - g_{i_1 i_2} - \cdots - g_{i_{m} i_{m+1}} \\ m\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{m} i_{m+1}} \right] _{p^0, \sigma }, \end{aligned}$$

where the non-recurring sequence \((i_1, \ldots , i_{m+1})\) satisfies \(i_1 = 1\), \(i_{m+1} = n\).

We sum the previous equation over all permutations \(\sigma \in S_n\). By Lemmas 19 and 20 it is enough to sum over all \(\sigma \) with \(\sigma (n) = n\). It follows \(g_{i_{m} i_{m+1}} = 0\) above. We then split the sum over all such \(\sigma \) into a sum over orderings \(\rho \) of the variables \(z_{i}, i \notin \{ i_1, \ldots , i_{m}, n \}\), a sum over orderings \(\tau \in S_{m}\) of the variables \(z_{i_1}, \ldots , z_{i_{m}}\) and the \(\left( {\begin{array}{c}n-1\\ m\end{array}}\right) \) refinements of both orderings. Since

$$\begin{aligned} F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + m - 1 - g_{i_1 i_2} - \cdots - g_{i_{m-1} i_{m}} \\ m\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{m} i_{m+1}} \end{aligned}$$

depends only on the variables \(z_i\) with \(i \notin \{ i_1, \ldots , i_{m} \}\) we find

$$\begin{aligned}&\big [ F(z) \big ]_{p^0} = \sum _{m \ge 1} \sum _{i_1, i_2, \ldots , i_{m+1}} \sum _{\rho \in S_{n-m-1}} \frac{1}{(n-1)!} \cdot \left( {\begin{array}{c}n-1\\ m\end{array}}\right) \\&\quad \cdot \left[ \sum _{\tau \in S_{m}} F \cdot \left( {\begin{array}{c} \mathsf {A}_{1n} + m - 1 - g_{i_1 i_2} - \cdots - g_{i_{m-1} i_{m}} \\ m\end{array}}\right) R_{i_1 i_2} \cdots R_{i_{m} i_{m+1}} \right] _{p^0, \tilde{\rho }} \end{aligned}$$

where \(\tilde{\rho }\) is any fixed refinement of the ordering \(\rho \). The proposition follows now by an application of Worpitzky’s identity

$$\begin{aligned} \sum _{\tau \in S_{m}} \left( {\begin{array}{c}x+m-1-a_{\tau }\\ m\end{array}}\right) = x^m, \end{aligned}$$

where \(a_{\tau }\) is the number of ascents of \(\tau \), i.e. the number of \(i \in \{ 1, \ldots , \ell -1 \}\) with \(\tau (i+1) > \tau (i)\). \(\square \)

Lemma 21

The action of the residue operators \(R_{ab}\) on meromorphic functions of variables \(z_1,\ldots ,z_n\) with poles only along \(z_i-z_j = 0\) for \(i<j\) satisfy

$$\begin{aligned} R_{ab} R_{cb} = R_{cb} R_{ab} + R_{ca} R_{ab}, \ \quad \ R_{ab} R_{bc} = -R_{ba} R_{ac} \end{aligned}$$

for all pairwise distinct abc.

Proof

We may assume that

$$\begin{aligned} f(z) = \prod _{1 \le i< j < n} (z_i - z_j)^{m_{ij}} \end{aligned}$$

for some \(m_{ij} \in {\mathbb {Z}}\). The claim follows then from a direct calculation. \(\square \)

1.6 Proof of Theorem 7

We prove the quasimodularity of \([F]_{p^0, \sigma }\), the homogeneity of \([F]_{p^0}\), and the formula

$$\begin{aligned} \begin{aligned} \frac{d}{dC_2} \Big [ F(z) \Big ]_{p^0} =&\, \left[ \frac{d}{dC_2} F \right] _{p^0} - 2\sum _{a< b = n} \left[ (w_a - w_b) F R_{ab} \right] _{p^0} \\&- 2 \sum _{a< b < n} \left[ (w_b - w_a) F R_{ba} \right] _{p^0}, \end{aligned} \end{aligned}$$
(71)

which implies the formula in the Theorem by symmetrization over \(S_n\). We argue by induction on n, the number of variables \(z_i\) on which F depends.

If \(n = 1\), then F is a quasimodular form and all three statements hold by inspection. Assume the statement is known for all functions which depend on a smaller number of variables. By Proposition 10, we have

$$\begin{aligned} \Big [ F(z) \Big ]_{p^0} = \sum _{m \ge 1} \sum _{i_1=1, i_2, \ldots , i_{m+1}=n} \left[ \left( F \frac{\mathsf {A}_{1n}^{m}}{m!}\right) R_{i_1 i_2} R_{i_2 i_3} \cdots R_{i_m i_{m+1}} \right] _{p^0}. \end{aligned}$$

Each summand on the right side depends on fewer variables than F and is therefore a quasi-modular form of weight k by Lemma 20 and induction. To obtain (71) we apply the \(\frac{d}{dC_2}\) operator, use induction on the right side, and use Lemma 21 to commute the resulting \(R_{ab}\) operators past the \(R_{i_k i_{k+1}}\) operators. This yields (71) also for F. The quasimodularity of \([F]_{p^0, \sigma }\) (and the weight bound) follows similarly from Lemma 20 and Proposition 9. \(\square \)

Appendix B. Elliptic fibrations

1.1 Overview

We present a refinement of Conjecture B by weight, and give evidence in the case of elliptic Calabi–Yau threefolds in fiber classes.

1.2 Weight refinement

Let \(\pi : X \rightarrow B\) be an elliptic fibration with a section and integral fibers. The holomorphic anomaly equation of Conjecture B and the argument used in the proof of Corollary 1 yield a refinement of Conjecture A by weight as follows.

Recall the divisor class W defined in Sect. 0.5. The endomorphisms of \(H^{*}(X)\) defined by

$$\begin{aligned} T_+(\alpha ) = ( \pi ^{*} \pi _{*} \alpha ) \cup W, \quad T_-(\alpha ) = \pi ^{*} \pi _{*} ( \alpha \cup W ) \end{aligned}$$

satisfy \(T_+^2 = T_+\) and \(T_-^2 = T_-\) as well as \(T_+ T_- = T_- T_+ = 0\). Hence the cohomology of X splits as

$$\begin{aligned} H^{*}(X) = \mathrm {Im}(T_+) \oplus \mathrm {Im}(T_-) \oplus \big ( \mathrm {Ker}( T_+ ) \cap \mathrm {Ker}(T_-) \big ) . \end{aligned}$$

Define a modified degree function \(\underline{\deg }(\gamma )\) by the assignment

$$\begin{aligned} \underline{\deg }(\gamma ) = {\left\{ \begin{array}{ll} 2 &{} \text {if } \gamma \in \mathrm {Im}(T_+) \\ 1 &{} \text {if } \gamma \in \mathrm {Ker}( T_+ ) \cap \mathrm {Ker}(T_-) \\ 0 &{} \text {if } \gamma \in \mathrm {Im}(T_-). \\ \end{array}\right. } \end{aligned}$$

If X is an elliptic curve and B is a point then \(\underline{\deg }\) specializes to the real cohomological degree \(\deg _{{\mathbb {R}}}\).

Corollary* 3

Assume Conjectures A and B hold. Then for any \(\underline{\deg }\)-homogeneous classes \(\gamma _1, \ldots , \gamma _n \in H^{*}(X)\) we have

$$\begin{aligned} {\mathcal C}^{\pi }_{g, \mathsf {k}}( \gamma _1, \ldots , \gamma _n ) \in H_{*}({\overline{M}}_{g,n}(B, \mathsf {k})) \otimes \frac{1}{\Delta (q)^{m}} \mathsf {QMod}_{\ell }, \end{aligned}$$

where \(m = -\frac{1}{2} c_1(N_{\iota }) \cdot \mathsf {k}\) and \(\ell = 2g - 2 + 12m + \sum _i \underline{\deg }(\gamma _i)\).

1.3 An example

Let X be a Calabi–Yau threefold and let \(\pi : X \rightarrow B\) be an elliptic fibration with section and integral fibers over a Fano surface B. We consider the genus g Gromov–Witten potentials in fiber classes

$$\begin{aligned} F_g(q) = \sum _{d = 0}^{\infty } q^d \int _{[ {\overline{M}}_{g,0}(X,dF) ]^{\text {vir}}} 1 \end{aligned}$$

with the convention that the summation starts at \(d=1\) if \(g \in \{ 0, 1 \}\). By Toda’s calculation [46, Thm 6.9], the Pandharipande–Thomas invariants \(\mathsf {P}_{n,\beta }\) of X in fiber classes form the generating series

$$\begin{aligned} \sum _{d = 0}^{\infty } \sum _{n \in {\mathbb {Z}}} \mathsf {P}_{n, dF} y^n q^d = \prod _{\ell , m \ge 1} (1- (-y)^{\ell } q^{m} )^{-\ell \cdot e(X)} \cdot \prod _{m \ge 1} (1-q^m)^{-e(B)} . \end{aligned}$$

Assuming X satisfies the Gromov–Witten/Pairs correspondence [41, 42], we therefore obtain

$$\begin{aligned} F_0(q)&= - e(X) \sum _{m, a \ge 1} \frac{1}{a^3} q^{ma} \\ F_1(q)&= \left( e(B) - \frac{1}{12} e(X) \right) \sum _{m,a \ge 1} \frac{1}{a} q^{ma} \\ F_g(q)&= e(X) \frac{(-1)^g B_{2g}}{4g} C_{2g-2}(q), \quad g \ge 2. \end{aligned}$$

If \(g \ge 2\) the series

$$\begin{aligned} \int {\mathcal C}^\pi _{g,0}() = F_g(q) \end{aligned}$$

is quasimodular of weight \(2g-2\) in agreement with Corollary* 3.

In genus \(g \le 1\) the series \(F_0\) and \(F_1\) are not quasimodular forms. However, this does not contradict Corollary* 3 since the moduli spaces \({\overline{M}}_{g,0}(\mathbb {P}^1,0)\) are unstable here and \({\mathcal C}^\pi _g()\) is not defined. Instead, we need to add additional insertions to stabilize the moduli space. In genus 0 we obtain

$$\begin{aligned} \int {\mathcal C}^\pi _{0,0}(W,W,W)&= \int _X W^3 + \left( q \frac{d}{dq} \right) ^3 F_0(q) = -12 e(X) C_4(q), \\ \int {\mathcal C}^\pi _{0,0}(\pi ^{*}D,W,W)&= \int _X \pi ^{*} D \cup W^2 = 0,\\ \int {\mathcal C}^\pi _{0,0}(\pi ^{*}D, \pi ^{*} D',W)&= \int _X \pi ^{*} D \cup \pi ^{*} D' \cup W = \int _B D \cdot D', \end{aligned}$$

for any \(D, D' \in H^{2}(B)\), where in the first equality we used

$$\begin{aligned} e(X) = -60 \int _{B} K_B^2. \end{aligned}$$

All three evaluations are in perfect agreement with Corollary* 3.

In genus 1 we obtain agreement with Corollary* 3 by

$$\begin{aligned} \int {\mathcal C}^{\pi }_{1,0}(W)&= \int _{{\overline{M}}_{1,1}(X,0)} {\text {ev}}_1^{*}(W) + \left( q \frac{d}{dq} \right) F_1(q) \\&= \left( e(B) -\frac{1}{12} e(X) \right) C_2(q), \end{aligned}$$

where we used

$$\begin{aligned} c_2(X) = \pi ^{*} c_2(B) + 11 \pi ^{*} c_1(B)^2 + 12 \iota _{*} c_1(B). \end{aligned}$$

A direct check shows that all evaluations above are also compatible with the conjectured holomorphic anomaly equation. For example, in genus 1 Conjecture B predicts correctly

$$\begin{aligned} \frac{d}{dC_2} \int {\mathcal C}^{\pi }_{1,0}(W)&= \int {\mathcal C}_{0,0}^{\pi }(W, \Delta _B) - 2 \int {\mathcal C}_{1,0}(\mathsf {1}) \psi _1 \\&= e(B) - 2 \int _{{\overline{M}}_{1,1}} \psi _1 \int _X c_3(X) \\&= e(B) - \frac{1}{12} e(X). \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Oberdieck, G., Pixton, A. Holomorphic anomaly equations and the Igusa cusp form conjecture. Invent. math. 213, 507–587 (2018). https://doi.org/10.1007/s00222-018-0794-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-018-0794-0

Navigation