Skip to main content
Log in

Regularized Integrals on Elliptic Curves and Holomorphic Anomaly Equations

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

Abstract

We derive residue formulas for the regularized integrals (introduced by Li and Zhou in Commun Math Phys 388:1403–1474, 2021) on configuration spaces of elliptic curves. Based on these formulas, we prove that the regularized integrals satisfy holomorphic anomaly equations, providing a mathematical formulation of the so-called contact term singularities. We also discuss residue formulas for the ordered A-cycle integrals and establish their relations with those for the regularized integrals.

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

Access this article

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

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. The limit \(\lim \limits _{\varepsilon \rightarrow 0}\) exists and does not depend on the choice of local coordinate z [LZ21, Definition 2.11].

  2. Strictly speaking, there are further relations among the generators in the ring \({\mathcal {J}}_{n}\) such as the Weierstrass equations.

  3. In fact, \(F_{I}\) in this system of choices also admits a formal integral formula

    $$\begin{aligned} \int _{-{\widehat{Z}}_{i_{1},i_{m}}}^{{\widehat{Z}}_{i_{m}}+0} dx_{i_{m}}\cdots \int _{-{\widehat{Z}}_{i_{1},i_{k}}}^{{\widehat{Z}}_{i_{k},i_{k+1}}+ x_{i_{k+1}}}dx_{i_{k}}\cdots \int _{-{\widehat{Z}}_{i_{1},i_{2}}}^{{\widehat{Z}}_{i_{2},i_{3}}+ x_{i_{3}}}dx_{i_{2}} \int _{0}^{{\widehat{Z}}_{i_{1},i_{2}}+ x_{i_{2}}}dx_{i_{1}}. \end{aligned}$$
  4. Lemma 3.2 can also be proved based on (A.4) by using the residue calculations detailed in Lemma A.5 below. Together with the reasoning here that proves (3.4), the derivation of Theorem 3.1 actually does not rely on the explicit residue formulas (e.g., Lemma 2.6) for regularized integrals, but only on the relation (A.4) between regularized integrals and ordered A-cycle integrals.

  5. The combinatorial data of \(\varvec{r}\) as a labeled rooted forest F is discussed in [LZ21], with the root labeled by 0. The number \(C_{J,\mathfrak {r}_{J}}\) is the quantity p(F) in [LZ21, Lemma 3.36], called tree factorial in [Kre00].

  6. Namely, it is a homogeneous Lie element in the free Lie algebra generated by the set of operator variables \(R^{(a)}_0, a\in [n]\).

  7. This is essentially [LZ21, Lemma 3.40]. The following argument also gives a shorter proof of [LZ21, Lemma 3.42].

  8. In [LZ21, Lemma 3.40, Lemma 3.42], the vanishing is proved by using quite involved combinatorics on labeled rooted forests. Here the proof is simplified by using only the nested commutator structure and the resulting anti-symmetry.

References

  1. Böhm, J., Bringmann, K., Buchholz, A., Markwig, H.: Tropical mirror symmetry for elliptic curves. J. für die reine und angewandte Mathematik (Crelles Journal) 732, 211–246 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bershadsky, M., Cecotti, S., Ooguri, H., Vafa, C.: Kodaira-Spencer theory of gravity and exact results for quantum string amplitudes. Commun. Math. Phys. 165, 311–428 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Dijkgraaf, R.: Mirror Symmetry and Elliptic Curves, The Moduli Space of Curves (Texel Island,: Progr. Math., vol. 129. Birkhäuser Boston, Boston, MA, 1995, 149–163 (1994)

  4. Dijkgraaf, R.: Chiral deformations of conformal field theories. Nucl. Phys. B 493(3), 588–612 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Douglas, M.: Conformal field theory techniques in large N Yang-Mills theory. In: Baulieu, L., Dotsenko, V., Kazakov, V., Windey, P. (eds.) Quantum Field Theory and String Theory. NATO ASI Series (Series B: Physics), vol. 328. Springer, Boston, MA (1995)

  6. Dabholkar, A., Murthy, S., Zagier, D.: Quantum Black Holes, Wall Crossing, and Mock Modular Forms, arXiv:1208.4074 [hep-th]

  7. Eicher, M., Zagier, D.: The theory of Jacobi forms, Progress in Math. 55, Birkhäuser Boston, Boston, MA, (1985)

  8. Getzler, E.: Resolving mixed hodge modules on configuration spaces. Duke Math. J. 96(1), 175–203 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  9. Goujard, E., Möller, M.: Counting Feynman-like graphs: Quasimodularity and Siegel-Veech weight. J. Eur. Math. Soc. 22(2), 365–412 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  10. Gui, Z., Li, S.: Elliptic Trace Map on Chiral Algebras, arXiv:2112.14572 [math.QA]

  11. Kreimer, D.: Chen’s Iterated Integral represents the operator product expansion. Adv. Theor. Math. Phys. 3, 627–670 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  12. Kaneko, M., Zagier, D.: A generalized Jacobi theta function and quasimodular forms, The moduli space of curves (Texel Island,: Progr. Math., vol. 129. Birkhäuser Boston, Boston, MA, vol. 1995, 165–172 (1994)

  13. Li, S.: Feynman graph integrals and almost modular forms. Commun. Number Theory Phys. 6, 129–157 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  14. Li, S., Zhou, J.: Regularized integrals on Riemann surfaces and modular forms. Commun. Math. Phys. 388, 1403–1474 (2021)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Libgober, A.: Elliptic genera, real algebraic varieties and quasi-Jacobi forms, arXiv:0904.1026 [math.AG]

  16. Oberdieck, G., Pixton, A.: Holomorphic anomaly equations and the Igusa cusp form conjecture. Invent. Math. 213, 507–587 (2018)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Rudd, R.: The string partition function for QCD on the torus, arXiv: 9407176 [hep-th]

  18. Roth, M., Yui, N.: Mirror symmetry for elliptic curves: the B-model (bosonic) counting, preprint

  19. Ruan, Y., Zhang, Y., Zhou, J.: Genus Two Quasi-Siegel Modular Forms and Gromov-Witten Theory of Toric Calabi-Yau Threefolds, arxiv:1911.07204 [math.AG]

  20. Silverman, J.H.: The Arithmetic of Elliptic Curves, Graduate Texts in Mathematics, vol. 106. Springer, New York, NY (2009)

    Book  Google Scholar 

  21. Totaro, B.: Configuration spaces of algebraic varieties. Topology 35, 1057–1067 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  22. Urban, E.: Nearly overconvergent modular forms. In: Bouganis, T., Venjakob, O. (eds.) Iwasawa Theory. Contributions in Mathematical and Computational Sciences, vol. 7, pp. 401–441. Springer, Berlin, Heidelberg (2012)

    Chapter  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Robbert Dijkgraaf, Zhengping Gui, Xinxing Tang and Zijun Zhou for helpful communications and discussions.

Funding

Both authors are supported by the national key research and development program of China (NO. 2020YFA0713000). J. Z. is also partially supported by the Young overseas high-level talents introduction plan of China.

Author information

Authors and Affiliations

Authors

Contributions

SL and JZ designed research, performed research and wrote the paper. All authors gave final approval for publication.

Corresponding author

Correspondence to Si Li.

Ethics declarations

Conflict of interest

The authors have no competing interests.

Additional information

Communicated by C. Schweigert.

Publisher's Note

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

A Some Combinatorial Properties of Residue Formulas

A Some Combinatorial Properties of Residue Formulas

We discuss some combinatorial properties of residue formulas and their applications in this Appendix.

1.1 A.1 Further commutation relations of residue operators

We first give an extension of Lemma 3.3.

Lemma A.1

Introduce the notation \(R^{(a)}_{0}=\int _{A} dz_{a}\). Then one has the following identities of operators acting on quasi-elliptic functions in z

  1. (i).
    $$\begin{aligned} R^{(a)}_{c}R^{(b)}_{c}-R^{(b)}_{c}R^{(a)}_{c}=R^{(b)}_{c}R^{(a)}_{b} =-R^{(a)}_{c}R^{(b)}_{a},\quad a,b\in [n],~ c\in [n] \cup \{0\}.\nonumber \\ \end{aligned}$$
    (A.1)
  2. (ii).
    $$\begin{aligned} R^{(e)}_{c}R^{(a)}_{b}=R^{(a)}_{b}R^{(e)}_{c}, \quad \textrm{if}\, \quad a,b,e\in [n], c\in [n]\cup \{0\},\quad \{a,b\}\cap \{c,e\}=\emptyset .\nonumber \\ \end{aligned}$$
    (A.2)
  3. (iii).

    When acting on almost-elliptic functions in z, one has

    $$\begin{aligned} R^{(b)}_{c}R^{(a)}_{b}+R^{(c)}_{a}R^{(b)}_{c}+R^{(a)}_{b}R^{(c)}_{a}=0, \quad \textrm{if}\, \quad a,b,c\in [n]\quad \text {are distinct}.\nonumber \\ \end{aligned}$$
    (A.3)

Here the various operators are understood to be composed in the natural operator ordering.

Proof

The first two identities with \(c\ne 0\) are covered in Lemma 3.3. The first identity (A.1) with \(c=0\) is proved by using the standard integral contour deformation argument on the complex plane. For the identity (A.2), by the global residue theorem and Lemma 3.3 one has

$$\begin{aligned} R^{(e)}_{0}R^{(a)}_{b}= & {} -(\sum _{\begin{array}{c} i:\, e\prec i\\ i\ne b \end{array}} R^{(e)}_{i} +R^{(e)}_{b}) R^{(a)}_{b}\nonumber \\= & {} -R^{(a)}_{b} \sum _{\begin{array}{c} i:\, e\prec i\\ i\ne b \end{array}} R^{(e)}_{i} -R^{(a)}_{b}R^{(e)}_{b}-R^{(a)}_{b}R^{(e)}_{a}+R^{(a)}_{b}R^{(e)}_{b}+R^{(a)}_{b}R^{(e)}_{a}-R^{(e)}_{b} R^{(a)}_{b}\nonumber \\= & {} R^{(a)}_{b} R^{(e)}_{0}. \end{aligned}$$

Here in the first equality we have used the fact that \(R^{(a)}_{b}\) preserves meromorphicity, so that the global residue theorem can be applied to \(R^{(e)}_{0}R^{(a)}_{b}\). The identity (A.3) is proved by a local calculation in a similar way as in Lemma 3.3. \(\square \)

The special case with \(c=0\) in (A.2) reads that

$$\begin{aligned} R^{(a)}_{b}\int _{A_{i}}=\int _{A_{i}}R^{(a)}_{b} \,\quad \text {for distinct}~a,b,i,~\text {when acting on quasi-elliptic functions}. \end{aligned}$$

We also record the following interesting result that is parallel to the above commutation relation.

Corollary A.2

One has the following identity of operators

Proof

One proof is already implicitly given when proving (3.4) in Theorem 3.1 using the residue formula for regularized integrals. An alternative proof using the relation between regularized integrals and ordered A-cycle integrals is given as follows.Footnote 4 For any almost-elliptic function \(\Psi \), we write

$$\begin{aligned} \Psi =\sum _{k} \Psi _{k} ({{\overline{z}}_i-{z}_i\over {\overline{\tau }}-\tau })^k, \end{aligned}$$

where \(\Psi _{k}\) is a quasi-elliptic function in \(z_{i}\). Then using the expression of \(\textrm{vol}\) given in Remark 2.7 and the Stokes theorem (see [LZ21, Lemma 3.26]), we have

(A.4)

By (A.4), linearity and Lemma A.1 with \(c=0\), it suffices to prove the following identity

$$\begin{aligned} R^{(a)}_{b}\sum _{r:\,r\ne i}R^{(i)}_{r} \left( ({\overline{z}}_{i}-z_{i})^{k+1} \Psi _k\right) =\sum _{r:\,r\ne i,a}R^{(i)}_{r} \left( ({\overline{z}}_{i}-z_{i})^{k+1} R^{(a)}_{b} \Psi _k\right) , \end{aligned}$$

where \(\Psi _k\) is quasi-elliptic in \(z_{i}\). This follows from Lemma 3.3 that is derived by local computations and is valid when acting on \(({\overline{z}}_{i}-z_{i})^{k+1} \Psi _k\). \(\square \)

1.2 A.2 Summation over planar rooted labeled forests

The sequence r in Propositions  2.11, 4.4 corresponds to rooted, labeled forest, as has been discussed in [LZ21]. We now elaborate on this. For each sequence r we associate a map

$$\begin{aligned} \varvec{r}:[n-1]\rightarrow [n],\quad \quad \varvec{r}(j)=r_{j}, ~j\in [n-1]. \end{aligned}$$
(A.5)

satisfying

$$\begin{aligned} j<\varvec{r}(j),\quad j\in [n-1]. \end{aligned}$$
(A.6)

The data (A.5) gives rise to a rooted, labeled forest, with the vertices labeled by \(\varvec{r}^{-1}(n)\) designated as the roots. Define an equivalence relation \(\sim \) on \([n-1]\) by declaring

$$\begin{aligned} j\sim \varvec{r}(j),\quad j\in [n-1]. \end{aligned}$$

Then \([n-1]\) is decomposed into several disjoint subsets

$$\begin{aligned}{}[n-1]=\bigsqcup _{k\in \varvec{r}^{-1}(n)} \Gamma _{k}. \end{aligned}$$
(A.7)

We call \(\varvec{r}|_{\Gamma _{k}}\) a tree in the forest \(\varvec{r}\).

Consider the blackboard orientation in which vertices that have valency zero or one are ordered from right to left, vertices within a path in a tree are ordered from bottom to the root in the top. Due to the constraint (A.6), each sequence \(\varvec{r}\) above gives a planar/embdedded, rooted, labelled forest (called forest for short), whose labelings are compatible with the blackboard orientation in the sense that with respect to <

  • labelings for the roots \(\varvec{r}^{-1}(n)\) increases from right to left;

  • within a rooted tree, labelings increases from from bottom to top (towards the root);

  • within a rooted tree, labelings of children of the same vertex increase from right to left.

By Lemma 3.3, when acting on almost-elliptic functions the operator \(R^{[n-1]}_{\varvec{r}}= R^{(n-1)}_{\varvec{r}(n-1)} \cdots R^{(2)}_{\varvec{r}(2)} R^{(1)}_{\varvec{r}(1)}\) can be factored into a product of residue operators associated to the trees

$$\begin{aligned} R^{[n-1]}_{\varvec{r}}=\prod _{k}R_{\Gamma _{k}}. \end{aligned}$$
(A.8)

Similarly, the integration operator (2.10) factor into a product of commuting integration operators. The condition (A.6) or equivalently the blackboard orientation records the operator orderings in the residue and integration operators within each \(\Gamma _{k}\) (permutations of different trees and of the leaves within the trees are not allowed by the embedding data). In particular, when acting on an elliptic function \(\Phi \) one has

(A.9)

where the summation \(\sum ^{<}\) on the right hand side is over forests whose labelings are compatible with the blackboard orientation, with respect to <.

Remark A.3

A rooted labeled forest \(\varvec{r}\) above gives rise to a rooted labeled tree obtained by joining all the roots in the rooted forest to the vertex n now designated as the new root. The joint tree corresponds to an element in the basis of the cohomology group \(H^{n-1}(\textrm{Conf}_{n}(\mathbb {C}),\mathbb {C})\) whose rank is the signed first Stirling number \((-1)^{n-1}s(n,1)=(n-1)!\). The relation \(R^{(j)}_{n}R^{(i)}_{j}=-R^{(i)}_{n}R^{(j)}_{i}\) and the Jacobi identity

$$\begin{aligned}{}[[R^{(a)}_{n},R^{(b)}_{n} ],R^{(c)}_{n}]+[[R^{(b)}_{n},R^{(c)}_{n} ],R^{(a)}_{n}]+[[R^{(c)}_{n},R^{(a)}_{n} ],R^{(b)}_{n}]=0. \end{aligned}$$

from Lemma A.1 then correspond to the anti-symmetry and Arnold relation (assembling the same form as Lemma A.1 (iii)) in the cohomology ring \(H^{*}(\textrm{Conf}_{n}(\mathbb {C}),\mathbb {C})\). See [Tot96, Get99] for details. It seems that our decomposition here is related to the mixed Hodge structures on \(H^{*}(\textrm{Conf}_{n}(E),\mathbb {C})\). We hope to discuss the algebraic formulation of regularized integrals using the language of mixed Hodge structures in a future investigation, following the lines in [Tot96, Get99].

There is in fact a bijection between the set of planar, rooted, labeled forests that are compatible with blackboard orientation and \(\mathfrak {S}_{[n-1]}\) given as follows.

  • For each \(\phi \in \mathfrak {S}_{[n-1]}\), consider its one-line notation \(\phi =\phi _1 \phi _2\cdots \phi _{n-1}\). Read from left to right, the maxima correspond to roots of the trees. From the remaining entries of \(\phi \), \(\phi _{a}\) is a child of \(\phi _{b}\) if \(\phi _{b}\) is the rightmost entry that precedes \(\phi _{a}\) and is larger than \(\phi _{a}\). This gives a rooted, labeled forest and thus a unique planar one that is compatible with backboard orientation.

  • For each planar, rooted, labeled forest compatible with backboard orientation, one starts at the vertex n of the forest and then traverses the forest clockwise. Reading off each label as it is reached, this yields an element in \(\mathfrak {S}_{[n-1]}\) in its one-line notation.

Using the above correspondence, (A.9) can be alternatively written as

(A.10)

where \(\varvec{r}(\phi )\) and equivalently \(\Gamma (\phi )\) is the forest compatible with blackboard orientation that is determined by \(\phi \).

A different ordering \(\prec \) is determined by \(\sigma \in \mathfrak {S}_{[n-1]}\) via the integration region \(A_{\sigma ([n-1])}\) or \(E_{\sigma ([n-1])}\). The ordering \(\prec \) is related to the magnitude ordering \(<:1<2<\cdots < n-1\) by the action by \(\sigma \)

$$\begin{aligned} \sigma _{1}\prec \sigma _{2}\prec \cdots \prec \sigma _{n-1}. \end{aligned}$$
(A.11)

Correspondingly, the map \(\varvec{r}\) is changed to \(\varvec{r}':\sigma ([n-1])=[n-1]\rightarrow [n]\) with the convention \(\sigma (n)=n\). It satisfies

$$\begin{aligned} \sigma _{i}\prec \varvec{r}'(\sigma _{i}) \end{aligned}$$
(A.12)

instead of (A.6). The set of such functions \(\varvec{r}'\) is in one-to-one correspondence with the set of the functions \(\sigma (\varvec{r}):=\sigma \circ \varvec{r}\circ \sigma ^{-1}\), with the convention \(\sigma (n)=n\). Correspondingly, we have

(A.13)

with \(\varvec{r}\) satisfying (A.5), (A.6).

Similarly to (A.9), (A.13) is a sum over the collection of embedded, rooted, labeled forests in (A.9), with the new labelings obtained by applying \(\sigma \) on the original ones. Now the new labelings are compatible with the blackboard orientation not with respect to < but with respect to \(\prec \) in (A.11). See Fig. 1 below for an illustration.

Fig. 1
figure 1

From left to right: picking representatives compatible with blackboard orientation. From top to bottom: applying permutation \(\sigma \) to obtain labeled rooted forests in the residue formulas of regularized integrals, with the map \(\varvec{r}'\) given by \(\sigma \circ \varvec{r}\circ \sigma ^{-1}\)

By the independence on the ordering of the regularized integral, we have

(A.14)

Now summing over the orderings \(\prec \) in (A.11) is equivalent to summing over the relabelings of the embedded, labeled rooted forests in (A.9), with the operator orderings for the residue and integration operators given by the blackboard orientation. It follows that

(A.15)

where \(\sigma \circ \Gamma _{k}\) is the tree whose labelings are obtained from the \(\sigma \)-action on those in \(\Gamma _{k}\). Denote the set of labelings of \(\Gamma _{k}\) by \(\rho _{k}\subseteq [n-1]\), then the summation can be further expressed as a sum

$$\begin{aligned} \sum _{\Gamma =(\Gamma _{k})}^{<} \sum _{\sigma \in \mathfrak {S}_{[n-1]}} = \sum _{[\Gamma ]=([\Gamma _{k}])} \sum _{\rho =(\rho _{k})\in \Gamma _{[n-1]}}\sum _{\tau \in \prod _{k}\mathfrak {S}_{\rho _{k}}}, \end{aligned}$$

where \([\Gamma ]\) is the forest without the labelings and \(\Gamma _{[n-1]}\) is the set of compositions of \([n-1]\). Therefore, we obtain

(A.16)

One can also exchange the order of the summation by summing over \(\rho \) in the outmost one. Then the summation is related to summation over different types of forests with fixed number of vertices in each of the trees.

Example A.4

Consider the \(n=3\) case. The two forests in (A.9) are given by the top left two in Fig. 2 below. In this case, \(\mathfrak {S}_{n-1}\) consists of two elements \(\sigma =(1)(2), (12)\). Hence the forests in (A.15) are obtained by including two more forests indicated in the top right two in Fig. 2. The resummation in (A.16) is illustrated in the data in the bottom.

Fig. 2
figure 2

The left two forests correspond to summands in \(\int _{E_3}\int _{E_2}\int _{E_1}\), the right two correspond to the summands in \(\int _{E_3}\int _{E_1}\int _{E_2}\)

Similar results hold for ordered A-cycle integrals, except now one does not have (A.14) for a general elliptic function \(\Phi \).

The structures in (A.9) and (A.16) elucidate many properties of regularized integrals and ordered A-cycle integrals. For example, Corollary 2.13 follows directly from this and the choice \(F_{I}={\widehat{Z}}_{i_{1}}^{m}/m!\) given in (2.12) for a non-recurring sequence \(I=(i_{1},\cdots , i_{m}),i_{m+1}=n\).

1.3 A.3 Holomorphic limit of regularized integrals

In Proposition 4.5 (ii), firstly established by [LZ21, Theorem 3.4 (2)], we used Proposition 4.4 (iii) proved by Oberdieck-Pixton in [OP18, Proposition 10]. We now give a proof of Proposition 4.5 (ii) that is simpler than the one in [LZ21], following the residue operator approach. According to Remark 4.6, this also provides a different proof of Proposition 4.4 (iii).

Lemma A.5

[LZ21, Lemma 3.26, Lemma 3.37]. As operators acting on elliptic functions in \({\mathcal {J}}_{n}\), one has

(A.17)

where

  • \(\prec \) is the prescribed ordering < by magnitude as in \(\int _{A_{[n]}}\).

  • \(J=(j_1,\cdots , j_\ell )\) is an non-recurring sequence: \(j_1\prec j_2\prec \cdots \prec j_\ell \) with cardinality \(|J|=\ell \).

  • \(\mathfrak {r}_{J}=(\mathfrak {r}_{j_1},\cdots , \mathfrak {r}_{j_\ell })\) satisfies \(j_a\prec \mathfrak {r}_{j_{a}}, \mathfrak {r}_{j_{a}}\in [n]\) for \(a=1,2,\cdots ,\ell \).

  • \(C_{J,\mathfrak {r}_{J}}=\int _{0}^{1}dx_{\mathfrak {r}_{j_{\ell }}}\int _{0}^{x_{\mathfrak {r}_{j_{\ell }}}}dx_{j_{\ell }}\cdots \int _{0}^{ x_{\mathfrak {r}_{j_{1}}}}dx_{j_{1}} \).

  • \(R^{J}_{\mathfrak {r}_{J}}=R^{(j_\ell )}_{\mathfrak {r}_{j_\ell }}\circ \cdots \circ R^{(j_1)}_{\mathfrak {r}_{j_1}}\) and the notation \( \left( \int _{A_{[n]-J}} R^{J}_{\mathfrak {r}_{J}}\right) ^{\prec }\) stands for the operator arranged according the ordering \(\prec \).

Proof

We only sketch the proof. For any almost-elliptic function \(\Psi \), \( \Psi =\sum _{k} \Psi _{k} ({{\overline{z}}_i-{z}_i\over {\overline{\tau }}-\tau })^k, \) continuing with (A.4), one has

$$\begin{aligned}= & {} \sum _{k} {1\over k+1}\int _{A} \Psi _{k} +2\pi i \sum _{k}{1\over k+1}\sum _{r} R^{(i)}_{r} \left( ({{\overline{z}}_i-{\overline{z}}_r+{\overline{z}}_r-z_{r}+z_{r}-z_{i}\over {\overline{\tau }}-\tau })^{k+1} \Psi _{k}\right) \nonumber \\= & {} \sum _{k} {1\over k+1}\int _{A} \Psi _{k}\\{} & {} +2\pi i \sum _{k}{1\over k+1}\sum _{\ell :\, 0\le \ell \le k+1}{k+1\atopwithdelims ()\ell } ({{\overline{z}}_r-z_{r}\over {\overline{\tau }}-\tau })^{k+1-\ell }({1\over {\overline{\tau }}-\tau })^{\ell } \sum _{r}R^{(i)}_{r} \left( (z_{r}-z_{i})^{\ell } \Psi _{k}\right) . \end{aligned}$$

Here in the last equality we have used (2.8). It follows that as operators on the almost-elliptic function \(\Psi \) one has

One then expands iterated regularized integrals of almost-elliptic functions in terms of iterated A-cycle integrals and residues (see [LZ21, Lemma 3.26]). Applying the above reasoning to get rid of terms that do not contribute in the holomorphic limit, the desired claim then follows, similar to the proof of Proposition 2.11. Note that there is no term with \(J=[n]\) by the above constraint on \(J,\mathfrak {r}_{J}\), as it should be the case according to the behavior of the pole structure under residue operations. \(\square \)

Remark A.6

When performing the operations in Lemma A.5, we do not use \(z_{n}\) as the reference as in Proposition 2.11, so that the \(\mathfrak {S}_{n}\)-symmetry among all the indices in [n] is retained. Of course the same relation holds if we do apply the referencing.

Before proceeding to give another proof of Proposition 4.5 (ii), we introduce some notations. Similar to the discussions in Appendix 4.2, for any pair \((J,\mathfrak {r}_{J})\) in Lemma A.5, we define a map

$$\begin{aligned} \varvec{r}:[n]\rightarrow [n]\cup \{0\},\quad \quad \varvec{r}(j)=\mathfrak {r}_{j}, ~j\in J\subseteq [n],\quad \varvec{r}(k)=0,~ k\in [n]-J. \end{aligned}$$
(A.18)

It is clear that the data \((J,\mathfrak {r}_{J})\) is equivalent to a map \(\varvec{r}\) in (A.18) satisfying \(j\prec \varvec{r}(j)\) for any \(j\in \varvec{r}^{-1}([n])\). In what follows we shall not distinguish them. Denote

$$\begin{aligned} \mathcal {R}_{\varvec{r}}:= \left( \int _{A_{[n]-J}} R^{J}_{\mathfrak {r}_{J}}\right) ^{\prec }= R^{(n)}_{\varvec{r}(n)} \cdots R^{(2)}_{\varvec{r}(2)} R^{(1)}_{\varvec{r}(1)}. \end{aligned}$$
(A.19)

Again we call the data \(\varvec{r}\) in (A.18) associated to \((J,\mathfrak {r}_{J})\) in Lemma A.5 a forest, and any element in [n] a vertex. By defining an equivalence relation \(\sim \) on [n] by declaring

$$\begin{aligned} j\sim \varvec{r}(j),\quad j\in J, \end{aligned}$$

the set [n] is decomposed into several disjoint subsets called treesFootnote 5

$$\begin{aligned}{}[n]=\bigsqcup _{k\in \varvec{r}^{-1}(0)} \Gamma _{k}. \end{aligned}$$
(A.20)

It is clear that the index set \(\varvec{r}^{-1}(0)\) satisfies \(|\varvec{r}^{-1}(0)|=1\) if and only if \(|\mathfrak {r}_{J}-J|=1\), in which case the forest itself is a tree.

The following result Lemma A.7 relates the residue operators \(\mathcal {R}_{\varvec{r}} \) associated to trees and forests to elements in the free algebra generated by the operator variables \(R^{(i)}_{0},i\in [n]\).

Lemma A.7

Let the notations be as above.

  1. (i).

    (Residue operator associated to a tree) Assume \(2\le m\le n\). Let \(\mathfrak {r}=(\mathfrak {r}_{1},\cdots , \mathfrak {r}_{m-1})\) be a sequence satisfying

    $$\begin{aligned} \mathfrak {r}_{a}\in [m],~ \mathfrak {r}_{m-1}=m,\quad \quad a<\mathfrak {r}_{a},~\text {for}~a=1,2,\cdots , m-1. \end{aligned}$$

    Then for any \(c\in [n]\cup \{0\}-[m]\), the operator on quasi-elliptic functions

    $$\begin{aligned} R^{(m)}_{c}\circ R^{[m-1]}_{\mathfrak {r}} :=R^{(m)}_{c}\circ R^{(m-1)}_{m}\circ \cdots R^{(2)}_{\mathfrak {r}_{2}}\circ R^{(1)}_{\mathfrak {r}_{1}} \end{aligned}$$

    is sum of nested commutators of m residue operators of the form \(R^{(a)}_{c}, a\in [m] \).

  2. (ii).

    (Residue operator associated to a forest) The operator \(\mathcal {R}_{\varvec{r}}\) in (A.19) is a product \(L_{v}\cdots L_{1}\) of \(v=|\varvec{r}^{-1}(0)|\) factors \(L_{k},k=1,2,\cdots , v\), with the factors being the operators associated to the trees in the decomposition (A.7). In particular, each \(L_{k}\) is either of the form \(R^{(i)}_{0},i\in [n]\) or a sum of nested commutators of \(m_k\) such operators, where \(m_k\) is the number of vertices of the tree corresponding to \(L_{k}\).Footnote 6

The constructive proof of Lemma A.7 is essentially contained in [LZ21]. To prove Lemma A.7 (i), we first show the statement for operators of the form \(R^{(b)}_{c}\prod _{i}R^{(i)}_{b}\) using Lemma A.1, then use induction on the height (i.e., the smallest number h such that \(\varvec{r}^{h}(j)=0\)) of the vertices \(j\in [n]\). After that, one applies Lemma A.1 to move the residue operators \(R^{(a)}_{b},b\ne 0\) across the \(R^{(e)}_{0}\) operators and reduce the discussions on forests in Lemma A.7 (ii) to those on trees in Lemma A.7 (i). The details are elementary but tedious, and are therefore omitted.

Remark A.8

We can in fact show that the operator \(R^{(m)}_{c}\circ R^{[m-1]}_{\mathfrak {r}}\) in Lemma A.7 (i) and each \(L_{k}\) in Lemma A.7 (ii), provided that the total number \(m_{k}\) of vertices in the corresponding tree is at least 2, can be expressed as a nested commutator instead of a sum of nested commutators. A graphic rule, which is essentially given by the blackboard orientation, can be found in [LZ21, Section 3.4].

Furthermore, using Lemma A.1 one can show that the compositions of residue operators \(\mathcal {R}_{\varvec{r}}\) can be conveniently described using the gluing of labeled rooted forests in the indicated orderings. Lemma A.1 can then be interpreted as saying that the algebra generated by the residue operators \(R^{(a)}_{b},a\in [n],b\in [n]\cup \{0\}\) is acted on by the underlying operad of labeled rooted forests.

Another proof of Proposition 4.5 (ii)

Consider the average of (A.17) over \(\sigma \in \mathfrak {S}_{n}\). Observe that \(C_{J,\mathfrak {r}_{J}}\) is independent of the ordering of elements in the underlying set of \(J\cup \mathfrak {r}_{J}\), as can be seen from its expression. Thanks to this property, to prove the desired statement Proposition 4.5 (ii) it is enough to show thatFootnote 7

$$\begin{aligned} \sum _{\sigma \in \mathfrak {S}_{n}}\mathcal {R}_{\sigma (\varvec{r})}=0,\quad ~\text {if}\quad \varvec{r}^{-1}([n])\ne \emptyset . \end{aligned}$$
(A.21)

Here as in (A.13) and (A.15), we have \(\mathcal {R}_{\sigma (\varvec{r})}= R^{(\sigma ([n-1]))}_{\sigma (\varvec{r})(\sigma ([n-1]))},\sigma (\varvec{r})=\sigma \circ \varvec{r}\circ \sigma ^{-1}\) with the convention \(\sigma (0)=0\).

For any such \(\varvec{r}\), applying Lemma A.7 (ii) we obtain a decomposition \(\mathcal {R}_{\varvec{r}}=L_{v}\cdots L_1\). The constraint on \(\varvec{r}\) in Lemma A.5 implies that the number \(v=|\varvec{r}^{-1}(0)|\) of trees in the decomposition (A.20) satisfies \(v\le n-1\). This tells that there is at least a factor above, say \(L_{*}\), such that the number m of vertices of the corresponding tree T is at least 2. Therefore, \(L_{*}\) is a sum of m-nested commutators. Denote the set of vertices of the tree T by V. The summation over \(\mathfrak {S}_{n}\) can be organized into a multi-summation similar to (A.16): over the choices for the underlying set of V, over the orderings of elements in the underlying set of V, and over the orderings of the rest of the elements in \(\{1,2,\cdots , n\}\). Now to prove (A.21) it suffices to show that

$$\begin{aligned} \sum _{\sigma \in \mathfrak {S}(V)}\sigma (L_{*})=0, \end{aligned}$$
(A.22)

where \(\sigma (L_{*})\) is the operator obtained by permuting the indices of the m-nested residue operators in \(L_{*}\) using \(\sigma \) from the permutation group \(\mathfrak {S}(V)\).

Let K be any m-nested commutator summand of \(L_{*}\). Consider the upper-indices, say \(a_1,a_2\), of an inner-most commutator of K. The summation \(\sum _{\sigma \in \mathfrak {S}(V)}\sigma (K)\) above is a multi-summation, with one of them being the sum over the permutation group of \(\{a_1,a_2\}\). The result is therefore zero by the anti-symmetryFootnote 8 of the nested commutator K in \(a_1,a_2\). By linearity, this then implies (A.22). The proof is now complete. \(\square \)

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

Li, S., Zhou, J. Regularized Integrals on Elliptic Curves and Holomorphic Anomaly Equations. Commun. Math. Phys. 401, 613–645 (2023). https://doi.org/10.1007/s00220-023-04647-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-023-04647-3

Navigation