Skip to main content
Log in

Heavenly metrics, BPS indices and twistors

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

Abstract

Recently, T. Bridgeland defined a complex hyperkähler metric on the tangent bundle over the space of stability conditions of a triangulated category, based on a Riemann–Hilbert problem determined by the Donaldson–Thomas invariants. This metric is encoded in a function \(W(z,\theta )\) satisfying a heavenly equation, or a potential \(F(z,\theta )\) satisfying an isomonodromy equation. After recasting the RH problem into a system of TBA-type equations, we obtain integral expressions for both W and F in terms of solutions of that system. These expressions are recognized as conformal limits of the ‘instanton generating potential’ and ‘contact potential’ appearing in studies of D-instantons and BPS black holes. By solving the TBA equations iteratively, we reproduce Joyce’s original construction of F as a formal series in the rational DT invariants. Furthermore, we produce similar solutions to deformed versions of the heavenly and isomonodromy equations involving a non-commutative star product. In the case of a finite uncoupled BPS structure, we rederive the results previously obtained by Bridgeland and obtain the so-called \(\tau \) function for arbitrary values of the fiber coordinates \(\theta \), in terms of a suitable two-variable generalization of Barnes’ G function.

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

Access this article

Subscribe and save

Springer+ Basic
$34.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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. Since hyperkähler (HK) and quaternion-Kähler (QK) spaces are parametrized by solutions of nonlinear differential equations often known as heavenly equations, we refer to such quaternionic metrics as ‘heavenly,’ irrespective of their precise nature. See [1, 2] for the original descriptions in the four-dimensional HK and QK cases, and [3] for a recent discussion of the higher dimensional case.

  2. In the context of class S field theories, the full space of stability conditions \(\mathrm{Stab}({\mathcal {D}})\) [16] can be interpreted as the total space of the Coulomb branch fibration over the conformal manifold, but the significance of the complex HK metric on \({\mathcal {M}}\) is yet to be understood.

  3. The factor \(1/(2\pi )^2\) appearing in this and a few other equations is due to our normalization of the coordinates \(z^a\) and \(\theta ^a\) which coincides with normalization used in [18], but differs by the factor \(2\pi \mathrm {i}\) from the one in [12, 13].

  4. In [12, 13], W has been called the Joyce function, but we prefer to reserve this attribution for F.

  5. Here, \(\sigma _{\gamma }\) is a quadratic refinement of the skew-symmetric pairing on the charge lattice \(\Gamma \), a technical device which allows to trivialize the twisted group law \({\mathcal {X}}_\gamma \,{\mathcal {X}}_{\gamma '} =(-1)^{\langle \gamma ,\gamma '\rangle } {\mathcal {X}}_{\gamma +\gamma '}\).

  6. The index ‘sf’ stands for ‘semi-flat,’ since the HK metric encoded by the twistor lines (2.6) is flat along the fibers parametrized by \(\theta ^a\). In fact, in contrast to the HK metrics appearing in the physical context before taking the conformal limit, it is also flat along the base \({\mathcal {S}}\) so that the full complex HK space encoded by (2.6) is simply \(\mathbb {C}^{2m}\) endowed with a flat metric of signature (2m, 2m).

  7. Our convention is opposite to the one in [12, 18, 21], which can be reached by flipping the sign of the skew-symmetric pairing \(\langle \gamma ,\gamma '\rangle \).

  8. The zero mode of F does not contribute to the vector field \({\hat{f}}\) and may be set to 0.

  9. The quadratic refinement \(\sigma _{\gamma _i}\) was not included in [19], which relied on a slightly different version of rational DT invariants, which lacks the property of being deformation invariant, see, e.g., [35] for details.

  10. Note that this iteration procedure is distinct from the usual iteration of the TBA equations (2.10) which leads to a formal series in powers of \({\mathcal {X}}^{\mathrm{sf}}_{\gamma }\). This second expansion is the subject of the next subsection.

  11. Alternatively, one can sum over unrooted labeled tress, in which case the coefficient \(1/|\mathrm{Aut}({\mathcal {T}})|\) in \(K_{\mathcal {T}}\) is replaced by 1/n!.

  12. We remark that this change of variables can be done already in (3.17), i.e., it has a generalization to any tree \({\mathcal {T}}\). In that case one defines \(w_{ij}=z_{j} u_i-z_i u_{j}\) for any edge \(e:i\rightarrow j\). Then, the Jacobian is given by \(\left( \sum _i z_i\right) \frac{\prod _{e:i\rightarrow j}z_i z_j}{\prod _i z_i}\) and leads to the integral \(\int _{R_{\mathcal {T}}} \prod _{e:i\rightarrow j} \frac{\mathrm {d}w_{ij}}{w_{ij}}\). The results presented in (3.21) and (3.22) correspond to the particular case of the tree of the simplest topology \(\bullet \!\text{--- }\!\bullet \!\text{-- }\cdots \text{-- }\!\bullet \!\text{--- }\!\bullet \,\).

  13. The first formula corrects a sign misprint in [44, Eq.(25)]. The second is obtained from [44, Eq.(20)] by twice integrating by parts. Note that the second integration produces the boundary contribution \(-\frac{1}{12}\log (z)\).

References

  1. Plebański, J.F.: Some solutions of complex Einstein equations. J. Math. Phys. 16, 2395–2402 (1975)

    Article  ADS  MathSciNet  Google Scholar 

  2. Przanowski, M.: Locally Hermite Einstein, selfdual gravitational instantons. Acta Phys. Polon. B 14, 625–627 (1983)

    MathSciNet  Google Scholar 

  3. Adamo, T., Mason, L., Sharma, A.: Twistor sigma models for quaternionic geometry and graviton scattering, arXiv:2103.16984

  4. Seiberg, N., Witten, E.: Gauge dynamics and compactification to three-dimensions. In: Conference on the Mathematical Beauty of Physics (In Memory of C. Itzykson), vol. 6 (1996). arXiv:hep-th/9607163

  5. Gaiotto, D., Moore, G.W., Neitzke, A.: Four-dimensional wall-crossing via three-dimensional field theory. Commun. Math. Phys. 299, 163–224 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  6. Douglas, M.R.: D-branes, categories and N = 1 supersymmetry. J. Math. Phys. 42, 2818–2843 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  7. Alexandrov, S., Pioline, B., Saueressig, F., Vandoren, S.: D-instantons and twistors. JHEP 03, 044 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  8. Alexandrov, S.: D-instantons and twistors: some exact results. J. Phys. A 42, 335402 (2009)

  9. Alexandrov, S., Persson, D., Pioline, B.: Wall-crossing, Rogers dilogarithm, and the QK/HK correspondence. JHEP 1112, 027 (2011)

    Article  ADS  MathSciNet  Google Scholar 

  10. Alexandrov, S.: Twistor approach to string compactifications: a review. Phys. Rep. 522, 1–57 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  11. Alexandrov, S., Manschot, J., Persson, D., Pioline, B.: Quantum hypermultiplet moduli spaces in N=2 string vacua: a review. In: Proceedings, String-Math 2012, Bonn, Germany, July 16–21, 2012, pp. 181–212. (2013) arXiv:1304.0766

  12. Bridgeland, T.: Geometry from Donaldson–Thomas invariants, arXiv:1912.06504

  13. Bridgeland, T., Strachan, I.A.B.: Complex hyperKähler structures defined by Donaldson–Thomas invariants, arXiv:2006.13059

  14. Gaiotto, D.: Opers and TBA, arXiv:1403.6137

  15. Bridgeland, T.: Stability conditions on triangulated categories. Ann. Math. (2) 166(2), 317–345 (2007)

  16. Bridgeland, T., Smith, I.: Quadratic differentials as stability conditions. Publications mathématiques de l’IHÉS 121(1), 155–278 (2015)

  17. Dunajski, M.: Null Kähler geometry and isomonodromic deformations, arXiv:2010.11216

  18. Bridgeland, T.: Riemann–Hilbert problems for the resolved conifold and non-perturbative partition functions. J. Differ. Geom. 115(3), 395–435 (2020)

    Article  MathSciNet  Google Scholar 

  19. Joyce, D.: Holomorphic generating functions for invariants counting coherent sheaves on Calabi–Yau 3-folds. Geom. Topol. 11(2), 667–725 (2007)

    Article  MathSciNet  Google Scholar 

  20. Filippini, S.A., Garcia-Fernandez, M., Stoppa, J.: Stability data, irregular connections and tropical curves. Sel. Math. 23(3), 1355–1418 (2017)

    Article  MathSciNet  Google Scholar 

  21. Bridgeland, T.: Riemann–Hilbert problems from Donaldson–Thomas theory. Invent. Math. 216, 69–124 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  22. Alexandrov, S., Pioline, B.: Conformal TBA for resolved conifolds, arXiv:2106.12006

  23. Alexandrov, S., Pioline, B., Saueressig, F., Vandoren, S.: Linear perturbations of quaternionic metrics. Commun. Math. Phys. 296, 353–403 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  24. Alexandrov, S., Moore, G.W., Neitzke, A., Pioline, B.: \({\mathbb{R}}^3\) index for four-dimensional \(N=2\) field theories. Phys. Rev. Lett. 114, 121601 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  25. Alexandrov, S., Pioline, B.: Black holes and higher depth mock modular forms. Commun. Math. Phys. 374(2), 549–625 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  26. Alexandrov, S., Roche, P.: TBA for non-perturbative moduli spaces. JHEP 1006, 066 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  27. Bridgeland, T., Laredo, V.T.: Stokes factors and multilogarithms. Journal für die reine und angewandte Mathematik 2013(682), 89–128 (2013)

    Article  MathSciNet  Google Scholar 

  28. Barbieri, A.: A Riemann–Hilbert problem for uncoupled BPS structures. Manuscripta Math. 2, 1–21 (2019)

    MathSciNet  MATH  Google Scholar 

  29. Alexandrov, S., Manschot, J., Pioline, B.: S-duality and refined BPS indices. Commun. Math. Phys. 380(2), 755–810 (2020)

    Article  ADS  MathSciNet  Google Scholar 

  30. Cecotti, S., Neitzke, A., Vafa, C.: Twistorial topological strings and a \({\rm tt}^*\) geometry for \({\cal{N}} = 2\) theories in \(4d\). Adv. Theor. Math. Phys. 20, 193–312 (2016)

    Article  MathSciNet  Google Scholar 

  31. Barbieri, A., Bridgeland, T., Stoppa, J.: A quantized Riemann–Hilbert problem in Donaldson–Thomas theory, arXiv:1905.00748

  32. Kontsevich, M., Soibelman, Y.: Stability structures, motivic Donaldson–Thomas invariants and cluster transformations, arXiv:0811.2435

  33. Hitchin, N.J., Karlhede, A., Lindström, U., Roček, M.: Hyperkähler metrics and supersymmetry. Commun. Math. Phys. 108, 535 (1987)

    Article  ADS  Google Scholar 

  34. Alexandrov, S., Pioline, B., Saueressig, F., Vandoren, S.: Linear perturbations of Hyperkahler metrics. Lett. Math. Phys. 87, 225–265 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  35. Bridgeland, T.: An introduction to motivic Hall algebras. Adv. Math. 229(1), 102–138 (2012)

    Article  MathSciNet  Google Scholar 

  36. Chen, K.-T.: Iterated path integrals. Bull. Am. Math. Soc. 83(5), 831–879 (1977)

    Article  MathSciNet  Google Scholar 

  37. Goncharov, A.B.: Multiple polylogarithms, cyclotomy and modular complexes. Math. Res. Lett. 5, 497–516 (1998)

    Article  MathSciNet  Google Scholar 

  38. Alexandrov, S., Banerjee, S., Manschot, J., Pioline, B.: Multiple D3-instantons and mock modular forms II. Commun. Math. Phys. 359(1), 297–346 (2018)

    Article  MathSciNet  Google Scholar 

  39. Strachan, I.A.B.: The Moyal algebra and integrable deformations of the selfdual Einstein equations. Phys. Lett. B 283, 63–66 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  40. Takasaki, K.: Dressing operator approach to Moyal algebraic deformation of selfdual gravity. J. Geom. Phys. 14, 111–120 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  41. Whittaker, E.T., Watson, G.N.: A Course of Modern Analysis. University Press (1927)

  42. Nemes, G.: Generalization of Binet’s gamma function formulas. Integral Transforms Special Funct. 24(8), 597–606 (2013)

  43. Vignéras, M.-F.: L’équation fonctionnelle de la fonction zêta de Selberg du groupe modulaire PSL (\((2,Z)\). Astérisque. 61, 235–249 (1979)

  44. Adamchik, V.S.: Contributions to the theory of the Barnes function. Int. J. Math. Comput. Sci. 9(1), 11–30 (2014)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We are indebted to Tom Bridgeland for his nice set of lectures at the work of the Simons collaboration on Special Holonomy in Geometry, Analysis and Physics (Jan 11–13, 2021), which stimulated our interest in this topic, and for subsequent correspondence. We are also grateful to Andy Neitzke for discussions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Boris Pioline.

Additional information

Publisher's Note

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

Appendices

Proofs

In this appendix, we present several proofs which have been omitted in the main text.

1.1 Heavenly equation

Let us show that the function W (3.3) satisfies the heavenly equation (2.20). As in the proof of the section condition, we do this with help of an infinite series generated by iterating the relation (3.5). Namely, using (3.8), we get

(A.1)

where in the last line we inverted the labeling of first \(n'\) variables and set \(n=n'+n''\). \(\square \)

1.2 Isomonodromy equation

To prove that the isomonodromy equation (2.15), we first rewrite it in terms of the potentials F and W which can be achieved by multiplying by \(e^{2\pi \mathrm {i}\theta _\gamma }\) and summing over the lattice. This leads to the following equation

$$\begin{aligned} \mathrm {d}F=\frac{1}{(2\pi \mathrm {i})^2}\sum _{a, b,c}\omega ^{ab}\partial _{\theta ^a}F \partial _{\theta ^b}\partial _{\theta ^c}W\, \mathrm {d}z^a, \end{aligned}$$
(A.2)

where the differential is supposed to act only on the variables \(z^a\). Substituting the expression (3.4) for the Joyce potential and using (3.8) and equation (3.2), one arrives at the condition

(A.3)

Then, note that the first term on the l.h.s. can be rewritten as

(A.4)

Substituting this result into (A.3) and bringing all terms to the l.h.s., the resulting expression becomes

(A.5)

where in the last line we inverted the labeling of the variables labeled from \(n'+1\) to \(n'+n''\) on the previous line. \(\square \)

1.3 Conjecture 1 at low rank

Here, we prove Conjecture 1 presented in Sect. 3.2 in the simplest cases \(n=3\) and \(n=4\). Let us denote \(\gamma _{ij}=\langle \gamma _i,\gamma _j\rangle \) and \(k_{ij}=\frac{t_i t_j}{t_j-t_i}\) so that \(K_{ij}=\gamma _{ij} k_{ij}/(2\pi \mathrm {i})\). The functions \(k_{ij}\) satisfy

$$\begin{aligned} k_{ij}k_{jk}-k_{ik}k_{jk}-k_{ij}k_{ik}=0. \end{aligned}$$
(A.6)

The proof can be reduced essentially to a repetitive use of this identity.

1.3.1 \(n=3\)

In this case, we need to show that

$$\begin{aligned} \,\mathrm{Sym}\, \Bigl [ \gamma _{12}\gamma _{23} \,k_{12} k_{23}\Bigr ]= & {} \, {1\over 2}\,\mathrm{Sym}\, \Bigl [ \left( \gamma _{12}\gamma _{23}+ \gamma _{12}\gamma _{13}+ \gamma _{13}\gamma _{23}\right) k_{12} k_{23}\Bigr ] \nonumber \\= & {} \, {1\over 2}\,\mathrm{Sym}\, \Bigl [ \gamma _{12}\gamma _{23}\left( k_{12} k_{23}+k_{12}k_{13}+k_{13}k_{23}\right) \Bigr ]. \end{aligned}$$
(A.7)

This is a direct consequence of (A.6) because it ensures that the expression in the round brackets on the r.h.s. is equal to \(2k_{12}k_{23}\).

1.3.2 \(n=4\)

In this case, we need to prove two relations

$$\begin{aligned}&\,\mathrm{Sym}\, \Bigl [ \gamma _{12}\gamma _{23}\gamma _{34} \,k_{12} k_{23}k_{34}\Bigr ] = \frac{1}{4}\,\mathrm{Sym}\, \Bigl [ \left( \gamma _{12}\gamma _{23}\gamma _{34} + \gamma _{12}\gamma _{13}\gamma _{34}+\gamma _{12}\gamma _{24}\gamma _{34}+\gamma _{12}\gamma _{14}\gamma _{34} \right. \nonumber \\&\qquad + \gamma _{13}\gamma _{23}\gamma _{24}+\gamma _{14}\gamma _{23}\gamma _{24}+\gamma _{12}\gamma _{23}\gamma _{14} +\gamma _{13}\gamma _{24}\gamma _{34}+\gamma _{12}\gamma _{13}\gamma _{24}+\gamma _{14}\gamma _{23}\gamma _{34} \nonumber \\&\left. \qquad +\gamma _{13}\gamma _{23}\gamma _{14}+\gamma _{13}\gamma _{14}\gamma _{24} \right) k_{12} k_{23}k_{34}\Bigr ], \end{aligned}$$
(A.8)
$$\begin{aligned}&\frac{1}{3} \,\mathrm{Sym}\, \Bigl [ \gamma _{12}\gamma _{13}\gamma _{14} \,k_{12} k_{13}k_{14}\Bigr ] \nonumber \\&\quad = \frac{1}{4}\,\mathrm{Sym}\, \Bigl [ \left( \gamma _{12}\gamma _{13}\gamma _{14}+\gamma _{12}\gamma _{23}\gamma _{24}+ \gamma _{13}\gamma _{23}\gamma _{34}+\gamma _{14}\gamma _{24}\gamma _{34}\right) k_{12} k_{23}k_{34}\Bigr ]. \end{aligned}$$
(A.9)

Proceeding as in (A.7), in each relation we rewrite all terms so that they have the same product of \(\gamma _{ij}\) factors.

The first relation is then equivalent to

$$\begin{aligned} \begin{aligned}&k_{12}k_{23}k_{34}+k_{12}k_{13}k_{34}+k_{12}k_{24}k_{34}+k_{12}k_{14}k_{34}+k_{13}k_{23}k_{24}+k_{13}k_{24}k_{34} +k_{13}k_{23}k_{14} \\&\quad +k_{14}k_{23}k_{34} +k_{12}k_{23}k_{14}+k_{14}k_{23}k_{24} +k_{12}k_{13}k_{24}-k_{13}k_{14}k_{24}=4 k_{12}k_{23}k_{34}. \end{aligned}\nonumber \\ \end{aligned}$$
(A.10)

The last 8 terms on the l.h.s. combined pairwise using the identity (A.6) give

$$\begin{aligned} 2k_{13}k_{23}k_{34}+k_{12}k_{23}k_{24}+k_{12}k_{13}k_{14}. \end{aligned}$$
(A.11)

The last two terms can in turn be combined with the third and fourth terms in (A.10) so that the full l.h.s. becomes

$$\begin{aligned} 2\left( k_{12}k_{23}k_{34}+k_{12}k_{13}k_{34}+k_{13}k_{23}k_{34}\right) . \end{aligned}$$
(A.12)

Due to (A.6), the sum of the last two terms here is equal to the first one so that one indeed obtains \(4 k_{12}k_{23}k_{34}\) as required.

The second relation (A.9) follows from

$$\begin{aligned} \begin{aligned}&k_{12}k_{23}k_{34}+k_{12}k_{14}k_{23}-k_{13}k_{14}k_{23}+k_{12}k_{13}k_{24}-k_{13}k_{14}k_{24} -k_{13}k_{23}k_{24}-k_{14}k_{23}k_{24} \\&\quad +k_{12}k_{13}k_{34}-k_{12}k_{14}k_{34}+k_{14}k_{23}k_{34} -k_{12}k_{24}k_{34}-k_{13}k_{24}k_{34}=4k_{12}k_{13}k_{14}. \end{aligned}\nonumber \\ \end{aligned}$$
(A.13)

To prove it, we note that combining the following pairs of terms on the l.h.s., (2,3), (4,5), (8,9), (1,11), (6,10), (5,12), one finds

$$\begin{aligned} 3k_{12}k_{13}k_{14}+k_{12}k_{23}k_{24}+k_{14}k_{24}k_{34}-k_{13}k_{23}k_{34}. \end{aligned}$$
(A.14)

This further can be rewritten as

$$\begin{aligned} \begin{aligned}&3k_{12}k_{13}k_{14}+(k_{12}k_{13}+k_{13}k_{23})k_{24}+k_{24}(k_{13}k_{34}-k_{13}k_{14}) -k_{13}(k_{24}k_{34}+k_{23}k_{24}) \\&\quad = 3k_{12}k_{13}k_{14}+k_{12}k_{13}k_{24}-k_{24}k_{13}k_{14} \end{aligned} \end{aligned}$$
(A.15)

Since the sum of the last two terms is combined into the first one, one indeed obtains \(4k_{12}k_{13}k_{14}\) as required.

1.4 Iterated integrals

Here, we prove that the functions \(J(z_1,\dots , z_n)\) defined in (3.19) satisfies the axioms (a)–(d) spelled out in (2.17).

(a) The homogeneity and holomorphicity are manifest from the representation on the second line of (3.20).

(b) Let us denote \(z_{k,l}=\sum _{i=k}^l z_i\) and \(z=z_{1,n}\). Then, bringing all terms in the recursion relation (2.17b) to one side, one finds

$$\begin{aligned}&\mathrm {d}J_n(z_1,\dots , z_n) - \sum _{k=1}^{n-1}\, J_k(z_1,\dots z_k)\,J_{n-k}(z_{k+1},\dots z_n) \left[ \frac{\mathrm {d}z_{k+1,n}}{z_{k+1,n}} -\frac{\mathrm {d}z_{1,k}}{z_{1,k}}\right] \nonumber \\&\quad = \frac{1}{(2\pi \mathrm {i})^{n-1}}\prod _{i=1}^n\left( \int _{\ell _{\gamma _i}}\frac{\mathrm {d}t_i}{t_i^2}\, e^{-2\pi \mathrm {i}z_i/t_i}\right) \prod _{i=1}^{n-1} \frac{t_i t_{i+1}}{t_{i+1}-t_i} \left[ \mathrm {d}z- 2\pi \mathrm {i}z\sum _{k=1}^{n}\frac{\mathrm {d}z_k}{t_k} \right. \nonumber \\&\left. \qquad -2\pi \mathrm {i}\sum _{k=1}^{n-1}\left( \frac{1}{t_k}-\frac{1}{t_{k+1}}\right) \left( z_{1,k}\mathrm {d}z_{k+1,n} -z_{k+1,n}\mathrm {d}z_{1,k}\right) \right] . \end{aligned}$$
(A.16)

Since the last line sums up to \(-2\pi \mathrm {i}\sum _{k=1}^{n}\frac{1}{t_k}\left( z_k\mathrm {d}z-z\mathrm {d}z_k\right) \), the total expression reduces to

$$\begin{aligned} \frac{\mathrm {d}z}{(2\pi \mathrm {i})^{n-1}}\prod _{i=1}^n\left( \int _{\ell _{\gamma _i}}\frac{\mathrm {d}t_i}{t_i^2}\, e^{-2\pi \mathrm {i}z_i/t_i}\right) \prod _{i=1}^{n-1} \frac{t_i t_{i+1}}{t_{i+1}-t_i} \left( 1- 2\pi \mathrm {i}\sum _{k=1}^{n}\frac{z_k}{t_k} \right) . \end{aligned}$$
(A.17)

Changing variables to \(t_i=\mathrm {i}z_i/(s u_i)\) with \(s, u_i\in \mathbb {R}^+\) and \(\sum _{i=1}^n u_i=1\), the integral factorizes exactly as in (3.20) into

$$\begin{aligned}&- \frac{\mathrm {i}\mathrm {d}z}{(2\pi \mathrm {i})^{n-1} \prod _{i=1}^n z_i} \int _0^\infty \mathrm {d}s\left( 1-2\pi s\right) e^{-2\pi s}\, \int _{\begin{array}{c} 0\le u_1\le 1\\ \sum _{i=1}^n u_1=1 \end{array}} \mathrm {d}u_1\cdots \mathrm {d}u_{n-1}\nonumber \\&\quad \prod _{i=1}^{n-1} \frac{z_i z_{i+1}}{z_{i+1} u_i-z_i u_{i+1}}\, . \end{aligned}$$
(A.18)

Since the integral over s vanishes, this proves the recursion relation (2.17b).

(c) The vanishing property (2.17c) is a consequence of the identity

$$\begin{aligned} \sum _{s\in \Sigma _n}\prod _{i=1}^{n-1}\frac{t_{s(i)} t_{s(i+1)}}{t_{s(i+1)}-t_{s(i)}}=0. \end{aligned}$$
(A.19)

To show that this holds, let \(u_i=1/t_i\), and consider

$$\begin{aligned} S_n(u_1,\dots , u_n) = \sum _{s\in \Sigma _n}\prod _{i=1}^{n-1} \frac{1}{u_{s(i)}- u_{s(i+1)}}\, . \end{aligned}$$
(A.20)

We can show that \(S_n=0\) inductively on n. For \(n=1\) and \(n=2\), this is clear. For \(n>2\), the residue of \(S_n\) at \(u_{n-1}=u_n\) is equal to \(S_{n-2}(u_1,\dots , u_{n-2})\), which vanishes by the induction hypothesis. By symmetry, the same holds for the residue at \(u_i=u_j\). Since \(S_n\) is a rational function with no poles, it must be a constant. And since it vanishes at large \(u_i\), it vanishes identically.

(d) The growth condition is evident from the representation (3.22) since singularities can only arise when the integration region \(R_n(z_1,\dots , z_n)\) touches one of the divisors \(w_i=0\), leading at most to logarithmic singularities. The behavior as \(z_k\rightarrow 0\) is easily read off from (3.20):

$$\begin{aligned} J_n(z_1,\dots , z_n)&{\mathop {\sim }\limits ^{z_1\rightarrow 0}}&-\frac{1}{(2\pi \mathrm {i})^2}\, J_{n-1}(z_2,z_3,\dots , z_n)\log (z_1), \nonumber \\ J_n(z_1,\dots , z_n)&{\mathop {\sim }\limits ^{z_n\rightarrow 0}}&\frac{1}{(2\pi \mathrm {i})^2}\, J_{n-1}(z_1,\dots , z_{n-1})\log (z_n), \end{aligned}$$
(A.21)

where we used the explicit form of \(J_2\) (2.18), whereas \(J_n(z_1,\dots , z_n)\) stays finite as \(z_k\rightarrow 0\) for \(1<k<n\), as the logarithmic divergences at \(u_k=z_k u_{k\pm 1}/z_{k\pm 1}\) cancel. \(\square \)

1.5 Refined potentials

Here, we prove that the refined versions of the Plebański and Joyce potentials, (5.8) and (5.9), satisfy the deformed heavenly equation (5.4) and isomonodromy equation (5.6), respectively. As in the unrefined case, the proof relies on an asymptotic expansion. But the difference is that now this will be the expansion in powers of \({\mathcal {X}}^{\mathrm{sf}}_\gamma \).

Thus, the starting point to establish (5.4) is the formal series (5.10) which implies

(A.22)
(A.23)

where in the last line we, as usual, set \(n=n'+n''\). Substituting these results into the deformed heavenly equation, one obtains

(A.24)

where

$$\begin{aligned} S_{ab}=\sum _{k,l=1}^N\frac{1}{t_k} \left( q_{k,a}q_{l,b}-q_{k,b}q_{l,a}\right) -\sum _{m=1}^{n-1} \left( \frac{1}{t_m}-\frac{1}{t_{m+1}}\right) \sum _{k=1}^m\sum _{l=m+1}^n \left( q_{k,a}q_{l,b}-q_{k,b}q_{l,a}\right) .\nonumber \\ \end{aligned}$$
(A.25)

It is straightforward to verify that \(S_{ab}\) actually vanishes. Indeed, denoting \(Q_{m,a}=\sum _{k=1}^m q_{k,a}\), we have

$$\begin{aligned} S_{ab}= & {} \sum _{m=1}^n\frac{1}{t_m}\left( q_{m,a}Q_{n,b}-q_{m,b}Q_{n,a}\right) -\sum _{m=1}^{n-1} \left( \frac{1}{t_m}-\frac{1}{t_{m+1}}\right) \left( Q_{m,a}Q_{n,b}-Q_{m,b}Q_{n,a}\right) \nonumber \\= & {} \frac{1}{t_n}\left( q_{n,a}Q_{n,b}-q_{n,b}Q_{n,a}\right) +\frac{1}{t_n}\left( Q_{n-1,a}Q_{n,b}-Q_{n-1,b}Q_{n,a}\right) =0. \end{aligned}$$
(A.26)

Thus, the deformed heavenly equation (5.4) holds.

In a similar way, to establish (5.6), we compute

(A.27)

where \(Z_k=Z_{\gamma _k}\), \(\zeta _m=\sum _{k=1}^m Z_k\) and the function \(\kappa (x)\) was defined at the beginning of Sect. 5. Further noting that

(A.28)

we see that (A.27) can be rewritten as

(A.29)

where

$$\begin{aligned} S'= & {} \sum _{k=1}^n \frac{1}{t_k}\left( Z_k\mathrm {d}\zeta _n -\zeta _n\mathrm {d}Z_k\right) -\sum _{m=1}^{n-1} \left( \frac{1}{t_m}-\frac{1}{t_{m+1}}\right) \left( \zeta _m\mathrm {d}\zeta _n-\zeta _n\mathrm {d}\zeta _m\right) \nonumber \\= & {} \frac{1}{t_n}\left( Z_n\mathrm {d}\zeta _n -\zeta _n\mathrm {d}Z_n\right) +\frac{1}{t_n}\left( \zeta _{n-1}\mathrm {d}\zeta _n-\zeta _n \mathrm {d}\zeta _{n-1}\right) =0. \end{aligned}$$
(A.30)

This proves the deformed isomonodromy equation (5.6).

Special functions and useful identities

In this appendix, we collect the definitions and some properties of several special functions as well as various useful identities used in Sect. 4.

1.1 Miscellaneous

1.1.1 Bernoulli polynomials

The Bernoulli polynomials have the following generating function

$$\begin{aligned} \sum _{m=0}^\infty \frac{x^m}{m!}\, B_m(\xi )= \frac{x\, e^{\xi x}}{e^x-1}. \end{aligned}$$
(B.1)

The first few polynomials are given by

$$\begin{aligned} \begin{aligned} B_0(x)=&\, 1, \\ B_1(x)=&\, x-{1\over 2}\, , \\ B_2(x)=&\, x^2-x+\frac{1}{6}\, , \\ B_3(x)=&\, x^3-\frac{3}{2}\, x^2+\frac{1}{2}\, x. \end{aligned} \end{aligned}$$
(B.2)

At \(x=0\) they reduce to Bernoulli numbers \(B_n\) and have the following symmetry property

$$\begin{aligned} B_n(1-x)=(-1)^n B_n(x). \end{aligned}$$
(B.3)

Importantly, the Bernoulli polynomials arise in the inversion formula for polylogarithms, namely,

$$\begin{aligned} \mathrm{Li}_n(e^{2\pi \mathrm {i}x})+(-1)^n\mathrm{Li}_n(e^{-2\pi \mathrm {i}x})=-\frac{(2\pi \mathrm {i})^n}{n!}\, B_n([x]), \end{aligned}$$
(B.4)

where

$$\begin{aligned}{}[x]= {\left\{ \begin{array}{ll} x-\lfloor \,\mathrm{Re}\,x\rfloor , &{} \quad \text{ if } \,\mathrm{Im}\,x\ge 0, \\ x+\lfloor -\,\mathrm{Re}\,x\rfloor +1, &{} \quad \text{ if } \,\mathrm{Im}\,x< 0. \end{array}\right. } \end{aligned}$$
(B.5)

Note that for \(\,\mathrm{Im}\,x\ne 0\) or \(x\notin \mathbb {Z}\), the bracket satisfies \([-x]=1-[x]\).

1.1.2 Useful identities

For \(\,\mathrm{Re}\,a,\,\mathrm{Re}\,b>0\), one has

$$\begin{aligned} \int _0^\infty \frac{\mathrm {d}x}{x}\left[ e^{-ax} -e^{-bx}\right]= & {} -\log \frac{a}{b}\, , \end{aligned}$$
(B.6a)
$$\begin{aligned} \int _0^\infty \frac{\mathrm {d}x}{x^2}\left[ e^{-ax} -(1-(a-b)x)\,e^{-bx}\right]= & {} b-a+a\log \frac{a}{b}\, , \end{aligned}$$
(B.6b)
$$\begin{aligned} \int _0^\infty \frac{\mathrm {d}x}{x^3}\left[ e^{-ax} -\left( 1-(a-b)x+{1\over 2}\,(a-b)^2 x^2\right) \,e^{-bx}\right]= & {} \frac{3a^2+b^2}{4}-ab-\frac{a^2}{2}\,\log \frac{a}{b}\qquad \nonumber \\ \end{aligned}$$
(B.6c)

1.2 Integral representations of Gamma and Barnes functions

1.2.1 Gamma function

The logarithm of the Gamma function has two integral representations due to Binet which are both valid for \(\,\mathrm{Re}\,z>0\) [41, p248-250]:

  • first Binet formula

    $$\begin{aligned} \log \Gamma (z) = \left( z-\frac{1}{2}\right) \log z - z + \frac{1}{2}\,\log (2\pi ) + \int _0^\infty \frac{\mathrm {d}s}{s} \left( {1\over 2}-\frac{1}{s}+\frac{1}{e^s-1} \right) e^{-z s}, \end{aligned}$$
    (B.7)
  • second Binet formula

    $$\begin{aligned} \log \Gamma (z) = \left( z-\frac{1}{2}\right) \log z - z + \frac{1}{2}\,\log (2\pi ) - \frac{z}{\pi } \int _0^{\infty } \frac{\mathrm {d}s}{s^2+z^2}\, \log \left( 1- e^{-2\pi s} \right) . \end{aligned}$$
    (B.8)

The first formula has the following useful generalization due to Hermite, valid for \(\,\mathrm{Re}\,z>0\) and \(\,\mathrm{Re}\,(z+\eta )>0\) [42]

$$\begin{aligned} \log \Gamma (z+\eta ) = \left( z+\eta -\frac{1}{2}\right) \log z - z + \frac{1}{2}\,\log (2\pi ) + \int _0^\infty \frac{\mathrm {d}s}{s} \left( \eta -{1\over 2}-\frac{1}{s}+\frac{e^{(1-\eta )s}}{e^s-1} \right) e^{-z s},\nonumber \\ \end{aligned}$$
(B.9)

which follows from (B.7) by using the identities (B.6).

1.2.2 Barnes function

The Barnes function (see, e.g., [43]) is the unique meromorphic function G(z) which satisfies the functional equation

$$\begin{aligned} G(z+1)=\Gamma (z)\, G(z)\, , \end{aligned}$$
(B.10)

the convexity condition \(\frac{d^3}{dz^3}\log G(z)\ge 0\) for all \(z>1\), and the normalization condition \(G(1)=1\). It has the following useful properties:

  • its logarithmic derivative is

    $$\begin{aligned} \frac{\mathrm {d}}{\mathrm {d}z} \log G(z+1) = \frac{1}{2}\log (2\pi ) -z + \frac{1}{2} + z \frac{\mathrm {d}}{\mathrm {d}z} \log \Gamma (z), \end{aligned}$$
    (B.11)
  • Kinkelin’s reflection formula

    $$\begin{aligned} \log \frac{G(1-z)}{G(1+z)} = z \log \left( \frac{\sin (\pi z)}{2\pi z}\right) - \int _0^z \pi x\, \log \sin (\pi x)\, \mathrm {d}x, \end{aligned}$$
    (B.12)
  • two integral representations of Binet type, both valid for valid for \(\,\mathrm{Re}\,z>0\) [44],Footnote 13

    $$\begin{aligned} \begin{aligned} \log G(z+1)=&\,z\log \Gamma (z)+ \frac{z^2}{4} -\frac{B_2(z)}{2}\, \log z +\zeta '(-1) -\frac{1}{12} \\&\,- \int _0^\infty \frac{\mathrm {d}s}{s^2} \left( \frac{1}{2}+ \frac{1}{s} +\frac{s}{12} - \frac{1}{1-e^{-s}} \right) e^{-zs}, \end{aligned} \end{aligned}$$
    (B.13)

    and

    $$\begin{aligned} \begin{aligned} \log G(z+1) =&\, \frac{1}{2}\left( z^2-\frac{1}{6}\right) \log z-\frac{3z^2}{4} + \zeta '(-1) + \frac{z}{2}\, \log (2\pi )\\&\, +\frac{1}{2\pi ^2}\int _0^{\infty } \frac{s \mathrm {d}s }{s^2+z^2}\, \Bigl [ 2\pi s \log \left( 1-e^{-2\pi s} \right) - \mathrm{Li}_2\left( e^{-2\pi s} \right) \Bigr ]. \end{aligned}\nonumber \\ \end{aligned}$$
    (B.14)

    where \(B_2(z)\) is the Bernoulli polynomial and \(\zeta '(-1)\) is related to Glaisher’s constant \(\gamma _E\) via \(\zeta '(-1)=\frac{1}{12}-\log \gamma _E\).

We will need a generalization of the first integral representation similar to Hermite’s generalization (B.9) of the first Binet formula. It is given by the following

Proposition 1

For \(\,\mathrm{Re}\,z>0\) and \(\,\mathrm{Re}\,(z+\eta )>0\), one has

$$\begin{aligned} \begin{aligned} \log G(z+\eta +1) =&\, \zeta '(-1) -\frac{3}{4} \,z^2-z\eta +\frac{1}{2}\, (z+\eta )\log (2\pi ) +\frac{1}{2}\left( (z+\eta )^2-\frac{1}{6}\right) \log z\\&\, + \int _0^\infty \frac{\mathrm {d}s}{s} \left( \frac{1}{s^2}-\frac{\eta }{s}+\frac{\eta ^2}{2} -\frac{1}{12} - \frac{e^{(1-\eta )s}}{(e^s-1 )^2} \right) e^{-zs}. \end{aligned} \end{aligned}$$
(B.15)

Proof

First, let us note the following identity

$$\begin{aligned}&\int _0^\infty \frac{\mathrm {d}s}{s} \left( \frac{1}{s^2} -\frac{1}{12} - \frac{e^{s}}{(e^s-1 )^2}\right) e^{-zs}\nonumber \\&\quad = z\log \Gamma (z)-z\left( z-{1\over 2}\right) \log (z)+z^2-\frac{z}{2}\log (2\pi )-\frac{1}{12} \nonumber \\&\qquad -\int _0^\infty \frac{\mathrm {d}s}{s^2} \left( \frac{1}{2}+ \frac{1}{s} +\frac{s}{12} - \frac{e^s}{e^s-1} \right) e^{-z s}, \end{aligned}$$
(B.16)

which can be established by integration by parts and the use of the first Binet formula (B.7). Therefore, the integral representation (B.13) can be rewritten as

$$\begin{aligned} \begin{aligned} \log G(z+1)=&\,\left( \frac{z^2}{2}-\frac{1}{12}\right) \log z -\frac{3z^2}{4}+\frac{z}{2}\log (2\pi ) +\zeta '(-1)\\&\,+\int _0^\infty \frac{\mathrm {d}s}{s} \left( \frac{1}{s^2} -\frac{1}{12} - \frac{e^{s}}{(e^s-1 )^2}\right) e^{-zs}. \end{aligned} \end{aligned}$$
(B.17)

Let us now substitute \(z\rightarrow z+\eta \). Then, the last integral term differs from the one in (B.15) by the following contribution

$$\begin{aligned} \begin{aligned} \int _0^\infty \frac{\mathrm {d}s}{s}&\left[ \left( \frac{1}{s^2} -\frac{1}{12}\right) e^{-(z+\eta ) s} -\left( \frac{1}{s^2}-\frac{\eta }{s}+\frac{\eta ^2}{2} -\frac{1}{12}\right) e^{-zs}\right] \\&\quad = \frac{3\eta ^2}{4}+\frac{z\eta }{2}-{1\over 2}\left( (z+\eta )^2-\frac{1}{6}\right) \log \frac{z+\eta }{z}\,, \end{aligned} \end{aligned}$$
(B.18)

where we used (B.6). Combining this result with terms in the first line of (B.17) after the replacement \(z\rightarrow z+\eta \), one reproduces the first line in (B.15), which completes the proof. \(\square \)

1.3 Generalized Gamma function

The generalized Gamma function \(\Lambda (z,\eta )\) is a function on \(\mathbb {C}^\times \times \mathbb {C}\) defined by [28]

$$\begin{aligned} \Lambda (z,\eta ) =\frac{e^z\,\Gamma (z+\eta )}{\sqrt{2\pi } z^{z+\eta -1/2}}\, . \end{aligned}$$
(B.19)

This function has the following properties:

  • twisted periodicity

    $$\begin{aligned} \Lambda (z,\eta +1) = \frac{z+\eta }{z}\, \Lambda (z,\eta ), \end{aligned}$$
    (B.20)
  • reflection property

    $$\begin{aligned} \Lambda (-z,\eta ) \, \Lambda (z,1-\eta ) = {\left\{ \begin{array}{ll} \left( 1-e^{2\pi \mathrm {i}( z-\eta )} \right) ^{-1}, \quad &{} \,\mathrm{Im}\,z>0,\\ \left( 1-e^{2\pi \mathrm {i}(\eta -z)} \right) ^{-1}, \quad &{} \,\mathrm{Im}\,z<0, \end{array}\right. } \end{aligned}$$
    (B.21)
  • two integral representations directly following from the generalized version of the first Binet formula (B.9) and the second Binet formula (B.8), respectively,

    $$\begin{aligned} \log \Lambda (z,\eta )= & {} \int _0^\infty \frac{\mathrm {d}s}{s} \left( \eta -{1\over 2}-\frac{1}{s}+\frac{e^{(1-\eta )s}}{e^s-1} \right) e^{-z s} \end{aligned}$$
    (B.22)
    $$\begin{aligned}= & {} \, \left( z+\eta -{1\over 2}\right) \log \frac{z+\eta }{z}\nonumber \\&-\eta -\frac{z+\eta }{\pi }\int _0^\infty \mathrm {d}s\,\frac{ \log \left( 1-e^{-2\pi s}\right) }{s^2+(z+\eta )^2}, \end{aligned}$$
    (B.23)

    where the first representation is valid for \(\,\mathrm{Re}\,z>0\), \(\,\mathrm{Re}\,(z+\eta )>0\), whereas the second requires only the second condition.

  • asymptotic series at large z [12, §8.3]

    $$\begin{aligned} \log \Lambda (z,\eta ) = \sum _{k= 2}^\infty \frac{(-1)^k B_k(\eta )}{k(k-1)}\, z^{1-k}. \end{aligned}$$
    (B.24)

    It follows from (B.22) by using \(\frac{s e^{x s}}{e^s-1}=\sum _{k=0}^{\infty } \frac{s^k}{k!} B_k(x)\) and integrating term by term using \(\int _0^\infty s^n e^{-zs}\mathrm {d}s= n!/ z^{n+1}\). Note that this asymptotic expansion is consistent with the periodicity relation (B.20), since \(B_k(\eta +1)-B_k(\eta )=k \eta ^{k-1}\).

1.4 Generalized Barnes function

We define the generalized Barnes function \(\Upsilon (z,\eta )\) on \(\mathbb {C}^\times \times \mathbb {C}\) by

$$\begin{aligned} \Upsilon (z,\eta ) = \frac{e^{\frac{3}{4}z^2- \zeta '(-1)}\, G(z+\eta +1) }{(2\pi )^{z/2} \,z^{\frac{1}{2} z^2} \bigl [\Gamma (z+\eta ) \bigr ]^\eta }\, , \end{aligned}$$
(B.25)

where G(z) is the Barnes function. This definition generalizes the function \(\Upsilon (z)\) introduced in [21, Eq.(17)] which is obtained from (B.25) setting \(\eta =0\). It has the following properties:

  • relation to the generalized Gamma function (B.19)

    $$\begin{aligned} \frac{\partial }{\partial z} \log \Upsilon (z,\eta ) = z \frac{\partial }{\partial z} \log \Lambda (z,\eta ), \end{aligned}$$
    (B.26)

    which follows from the property of the Barnes function (B.11),

  • twisted periodicity

    $$\begin{aligned} \Upsilon (z,\eta +1) = (z+\eta )^{-\eta } \Upsilon (z,\eta ), \end{aligned}$$
    (B.27)
  • reflection property

    $$\begin{aligned}&\log \frac{\Upsilon (-z,\eta )}{\Upsilon (z,1-\eta )} \nonumber \\&\quad = -\frac{\pi \mathrm {i}}{12}\,B_2(\eta )+ {\left\{ \begin{array}{ll} z\log \left( 1-e^{2\pi \mathrm {i}( z-\eta )} \right) +\frac{1}{2\pi \mathrm {i}}\, \mathrm{Li}_2\left( e^{2\pi \mathrm {i}( z-\eta )} \right) , &{} \,\mathrm{Im}\,z>0\\ z\log \left( 1-e^{2\pi \mathrm {i}(\eta -z)} \right) -\frac{1}{2\pi \mathrm {i}}\, \mathrm{Li}_2\left( e^{2\pi \mathrm {i}(\eta -z)}\right) , &{} \,\mathrm{Im}\,z<0, \end{array}\right. }\nonumber \\ \end{aligned}$$
    (B.28)

    which is a consequence of (B.12),

  • two integral representations of Binet type

    $$\begin{aligned} \log \Upsilon (z,\eta )= & {} \int _0^\infty \frac{\mathrm {d}s}{s} \left( \frac{1}{s^2} - \frac{1}{2} \,B_2(\eta ) - \frac{\eta (e^s-1)+1}{(e^s-1)^2}\,e^{(1-\eta )s} \right) e^{-z s}\qquad \quad \quad \end{aligned}$$
    (B.29)
    $$\begin{aligned}&-\frac{1}{2} \,B_2(\eta ) \log z \nonumber \\= & {} \frac{z^2}{2}\, \log \frac{z+\eta }{z}-{1\over 2}\,B_2(\eta )\, \log (z+\eta )-\frac{z\eta }{2}+\frac{\eta ^2}{4} \nonumber \\&+\int _0^{\infty } \frac{\mathrm {d}s }{s^2+(z+\eta )^2} \left[ \frac{1}{\pi }\left( s^2+\eta (z+\eta )\right) \log \left( 1-e^{-2\pi s} \right) \right. \nonumber \\&\left. - \frac{s}{2\pi ^2}\, \mathrm{Li}_2\left( e^{-2\pi s} \right) \right] , \end{aligned}$$
    (B.30)

    directly following from (B.15), (B.9) and (B.14), (B.8), respectively,

  • asymptotic series at large z

    $$\begin{aligned} \log \Upsilon (z,\eta ) = -\frac{1}{2}\, B_2(\eta ) \log z + \sum _{k= 3}^\infty \frac{ (-1)^k B_k(\eta )}{k(k-2)}\, z^{2-k}, \end{aligned}$$
    (B.31)

    which follows from (B.29) taking into account

    $$\begin{aligned} \frac{1}{s^2} - \frac{1}{2}\, B_2(\eta ) - \frac{\eta (e^s-1)+1}{(e^s-1)^2}\,e^{(1-\eta )s} = \sum _{k=3}^{\infty } \frac{(k-1)}{k!}\, B_k(1-\eta )\, s^{k-2}\nonumber \\ \end{aligned}$$
    (B.32)

    and integrating term by term. For \(\eta =0\), this reduces to the result given in [21, §5.4].

Solution in the uncoupled finite case

Here, we provide some intermediate steps in deriving (4.6) and (4.7).

1.1 Darboux coordinates

In the uncoupled case, \({\mathcal {X}}_{\gamma '}\) appearing in the r.h.s. of the integral equation (1.4) can be replaced by \({\mathcal {X}}^{\mathrm{sf}}_{\gamma '}\). Then, after combining contributions of opposite charges, this equation takes the form

$$\begin{aligned} {\mathcal {X}}_\gamma (t) = {\mathcal {X}}^{\mathrm{sf}}_{\gamma }(t) \,\exp \left[ \sum _{\begin{array}{c} \gamma '\in \Gamma \\ \,\mathrm{Re}\,(Z_{\gamma '}/t)>0 \end{array}} \Omega (\gamma ') \langle \gamma ,\gamma '\rangle R_{\gamma '}(t)\right] , \end{aligned}$$
(C.1)

where

$$\begin{aligned} R_{\gamma }(t) =\frac{1}{2\pi \mathrm {i}}\int _{-\mathrm {i}\vartheta _\gamma }^{\infty } \mathrm {d}s \frac{\log \left( 1-e^{-2\pi s}\right) }{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )} -\frac{1}{2\pi \mathrm {i}}\int _{\mathrm {i}\vartheta _\gamma }^{\infty }\frac{\log \left( 1-e^{-2\pi s}\right) }{s+\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )}\qquad \end{aligned}$$
(C.2)

and we changed the integration variable \(t'=\mathrm {i}Z_{\gamma }/(s+\mathrm {i}\vartheta _\gamma )\). Splitting each integral into two parts, one obtains

$$\begin{aligned} \begin{aligned} R_{\gamma }(t) =&\, \frac{Z_\gamma /t-\vartheta _\gamma }{\pi }\int _0^{\infty }\mathrm {d}s\, \frac{\log \left( 1-e^{-2\pi s}\right) }{s^2+(Z_\gamma /t-\vartheta _\gamma )^2} \\&\, -\frac{1}{2\pi \mathrm {i}}\int _0^{-\mathrm {i}\vartheta _\gamma } \mathrm {d}s\,\frac{\log \left( 1-e^{-2\pi s}\right) }{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )} +\frac{1}{2\pi \mathrm {i}}\int _0^{-\mathrm {i}\vartheta _\gamma } \mathrm {d}s\, \frac{\log \left( 1-e^{2\pi s}\right) }{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )}\, , \end{aligned} \nonumber \\ \end{aligned}$$
(C.3)

where in the last term we flipped the sign of s. The last two terms can be combined using (B.4) at \(n=1\), leading to the contribution

$$\begin{aligned} -\int _0^{-\mathrm {i}\vartheta _\gamma } \mathrm {d}s\,\frac{[\mathrm {i}s]-{1\over 2}}{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )}\, . \end{aligned}$$
(C.4)

If \(\,\mathrm{Re}\,\vartheta _\gamma \in (0,1)\), the integral is easily evaluated to

$$\begin{aligned} -\left( Z_\gamma /t-\vartheta _\gamma +{1\over 2}\right) \log \left( 1-\frac{t\vartheta _\gamma }{Z_\gamma }\right) -\vartheta _\gamma \, . \end{aligned}$$
(C.5)

Combining this with the first term in (C.3), one recognizes the integral representation (B.23) of the generalized Gamma function shifted by a logarithmic term

$$\begin{aligned} -\log \Lambda \left( \frac{Z_\gamma }{t}, -\vartheta _\gamma \right) -\log \left( 1-\frac{t\vartheta _\gamma }{Z_\gamma }\right) =-\log \Lambda \left( \frac{Z_\gamma }{t},1-\vartheta _\gamma \right) , \end{aligned}$$
(C.6)

where in the last step we used the twisted periodicity property (B.20). If however \(\,\mathrm{Re}\,\vartheta _\gamma \notin (0,1)\), there is an additional contribution given by

$$\begin{aligned} \epsilon _\gamma \sum _{n\in S_{\epsilon _\gamma }}\int _{-\mathrm {i}n}^{-\mathrm {i}\vartheta _\gamma }\frac{\mathrm {d}s}{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )} =-\epsilon _\gamma \sum _{n\in S_{\epsilon _\gamma }}\log \left( 1-\frac{t(\vartheta _\gamma -n)}{Z_\gamma }\right) , \end{aligned}$$
(C.7)

where \(\epsilon _\gamma =\text{ sgn }(\,\mathrm{Re}\,\vartheta _\gamma )\) and we introduced two finite sets \(S_+=\{1,\dots , \lfloor \,\mathrm{Re}\,\vartheta _\gamma \rfloor \}\) and \(S_-=\{0,\dots , 1+\lfloor \,\mathrm{Re}\,\vartheta _\gamma \rfloor \}\). Adding this contribution to (C.6) and again using repeatedly the twisted periodicity (B.20), one finally obtains

$$\begin{aligned} R(t) =-\log \Lambda \left( \frac{Z_\gamma }{t},1-[\vartheta _\gamma ]\right) . \end{aligned}$$
(C.8)

Substituting this result into (C.1), one gets formula (4.6) given in the main text.

1.2 \(\tau \) function

The starting point for deriving the \(\tau \) function is formula (3.29). In terms of the integer BPS indices (and after discarding an irrelevant function of \(\theta ^a\) coming from \(\log (\pm n)\) in the last term where n counts the reducibility of the charge), it takes the form

$$\begin{aligned} \log \tau= & {} \frac{1}{4\pi ^2}\sum _{\gamma \in \Gamma _{\!\star }} \biggl [\Omega (\gamma )\int _{\ell _\gamma } \frac{\mathrm {d}t'}{t'}\, \frac{t}{t'-t}\, \Bigl ( \mathrm{Li}_2\left( e^{2\pi \mathrm {i}(\vartheta _\gamma - Z_\gamma /t')}\right) \nonumber \\&-2\pi \mathrm {i}\, \frac{Z_\gamma }{t'}\,\log \left( 1-e^{2\pi \mathrm {i}(\vartheta _\gamma - Z_\gamma /t')}\right) \Bigl ) -\mathrm{Li}_2\left( e^{2\pi \mathrm {i}\vartheta _\gamma }\right) \log (Z_\gamma /t)\biggr ]. \end{aligned}$$
(C.9)

The next steps are similar to the one in the previous appendix. Setting \(t'=\mathrm {i}Z_{\gamma }/(s+\mathrm {i}\vartheta _\gamma )\) and combining the contributions of opposite charges, one finds

$$\begin{aligned} \log \tau= & {} -\frac{1}{4\pi ^2}\sum _{\gamma \in \Gamma _{\!\star }} \sigma _\gamma \Omega (\gamma )\biggl [ \int _{-\mathrm {i}\vartheta _\gamma }^\infty \mathrm {d}s\, \frac{\mathrm{Li}_2\left( e^{-2\pi s}\right) -2\pi (s+\mathrm {i}\vartheta _\gamma )\,\log \left( 1-e^{-2\pi s}\right) }{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )} \nonumber \\&\quad +\mathrm{Li}_2\left( e^{2\pi \mathrm {i}\vartheta _\gamma }\right) \log (Z_\gamma /t)\biggr ] \nonumber \\&=\sum _{\begin{array}{c} \gamma \in \Gamma \\ \,\mathrm{Re}\,(Z_{\gamma }/t)>0 \end{array}} \sigma _\gamma \Omega (\gamma )\nonumber \\&\qquad \left[ \int _{0}^\infty \mathrm {d}s \frac{-\frac{s}{2\pi ^2 }\, \mathrm{Li}_2\left( e^{-2\pi s}\right) +\frac{1}{\pi }\left( s^2+\vartheta _\gamma (\vartheta _\gamma -Z_\gamma /t)\right) \log \left( 1-e^{-2\pi s}\right) }{s^2+(Z_\gamma /t-\vartheta _\gamma )^2} \right. \nonumber \\&\left. +{1\over 2}\int _0^{-\mathrm {i}\vartheta _\gamma } \mathrm {d}s\, \frac{B_2([\mathrm {i}s]) -(\mathrm {i}s-\vartheta _\gamma )(2[\mathrm {i}s]-1)}{s-\mathrm {i}(Z_\gamma /t-\vartheta _\gamma )} -{1\over 2}\, B_2([\vartheta _\gamma ])\,\log (Z_\gamma /t)\right] .\nonumber \\ \end{aligned}$$
(C.10)

For \(\,\mathrm{Re}\,\vartheta _\gamma \in (0,1)\), the second integral is evaluated to

$$\begin{aligned} \frac{1}{4}\, \vartheta _\gamma ^2 +\frac{\vartheta _\gamma Z_\gamma }{2t } +{1\over 2}\left( \frac{Z_\gamma ^2}{t^2}-B_2(\vartheta _\gamma )\right) \log \left( 1-\frac{t\vartheta _\gamma }{Z_\gamma }\right) . \end{aligned}$$
(C.11)

Substituting this back into (C.10), one recognizes the square bracket as the integral representation (B.30) of the generalized Barnes function \(\Upsilon (z,\eta )\) shifted by a logarithmic term

$$\begin{aligned} \log \Upsilon \left( \frac{Z_\gamma }{t}, -\vartheta _\gamma \right) +\vartheta _\gamma \log \left( Z_\gamma /t-\vartheta _\gamma \right) =\log \Upsilon \left( \frac{Z_\gamma }{t},1-\vartheta _\gamma \right) , \qquad \end{aligned}$$
(C.12)

where in the last step we used the twisted periodicity property (B.27). For \(\,\mathrm{Re}\,\vartheta _\gamma \notin (0,1)\), there are additional contributions which can be computed as in (C.7) and amount to replacing \(\vartheta _\gamma \) by \([\vartheta _\gamma ]\), consistently with the periodicity of the original integral representation. As a result, one arrives at formula (4.7).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Alexandrov, S., Pioline, B. Heavenly metrics, BPS indices and twistors. Lett Math Phys 111, 116 (2021). https://doi.org/10.1007/s11005-021-01455-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11005-021-01455-5

Keywords

Mathematics Subject Classification

Navigation