Skip to main content
Log in

Reconstructing GKZ via Topological Recursion

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

Abstract

In this article, a novel description of the hypergeometric differential equation found from Gel’fand–Kapranov–Zelevinsky’s system (referred to as GKZ equation) for Givental’s J-function in the Gromov–Witten theory will be proposed. The GKZ equation involves a parameter \(\hbar \), and we will reconstruct it as a quantum curve from the classical limit \(\hbar \rightarrow 0\) via the topological recursion. In this analysis, the spectral curve (referred to as GKZ curve) plays a central role, and it can be described by the critical point set of the mirror Landau–Ginzburg potential. Our novel description is derived via the duality relations of the string theories, and various physical interpretations suggest that the GKZ equation is identified with the quantum curve for the brane partition function in the cohomological limit. As an application of our novel picture for the GKZ equation, we will discuss the Stokes phenomenon for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model, and the wall-crossing formula for the total Stokes matrix will be examined. And as a byproduct of this analysis, we will study Dubrovin’s conjecture for this equivariant model.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. In general, we can identify these differential equations after some coordinate change through the mirror map. This implies that the J-functions are written in terms of the oscillatory integrals. To describe precise relation between these objects, we need the notion of the Gamma class introduced in [73, 80]; see [53, 54] for details.

  2. Since spectral curves discussed in this paper are of genus 0, we do not include the choice of \(\omega _2^{(0)}\) in the definition of spectral curve.

  3. A relationship between Gromov–Witten theory (or Frobenius structures) and the topological recursion has already been discussed in [42, 48] etc. However, our viewpoint is different from that of these works. We will focus on the reconstruction of GKZ equation as quantum curves.

  4. In view of the GKZ equation (1.1), this is regarded as the case where all equivariant parameters \(w_i\) and \(\lambda _a\) are equal to 0.

  5. Such similarities were also pointed out in [90].

  6. We remark that the operators \({\widehat{x}}\) and \({\widehat{y}}\) obey the commutation relation \( \left[ {\widehat{y}}, {\widehat{x}}\right] =\hbar . \)

  7. The commutation relation becomes \(\left[ {\widehat{y}}, {\widehat{x}}\right] =x \hbar \) which is sightly modified from the previous one.

  8. Using the operators \({\widehat{x}}\) and \({\widehat{y}}\) in (3.3) one can represent operators \(\widehat{{\mathsf {x}}}\) and \(\widehat{{\mathsf {y}}}\) as \(\widehat{{\mathsf {x}}}=\mathrm {e}^{{\widehat{x}}}\)and \(\widehat{{\mathsf {y}}}=\mathrm {e}^{{\widehat{y}}}\).

  9. In [89], such local coordinates are specified by the Lagrangian singularity of \(\Sigma _X\).

  10. The convergence in Remark 3.14 is confirmed by looking at the \(z\rightarrow \infty \) behavior of the building blocks of the topological recursion (3.12) such as Y(x(z)) and \(\int _{w_*}^{w} B(\cdot ,z)\).

  11. Precisely speaking, these labels are not well-defined since the labels exchange if x move around a ramification point (\(\infty \) is a ramification point). Here we consider a situation that x moves along a straight path to infinity.

  12. There are no mathematically rigorous definition for the brane partition function as the generating function of the open Gromov–Witten invariants in general. But switched to the type IIA superstring picture, we can find it as the enumeration of degeneracies of the open BPS states [91] which arise from D0-D2-D4 brane bound states for the case of local toric Calabi–Yau 3-fold. In this sense we have only the string theoretical definition of the brane partition function.

  13. This q-difference equation is defined simply as the annihilating equation of the brane partition function. Via mirror symmetry (see discussions in Sect. 5.3), the brane partition function is regarded as the wave function [3], and in this sense, we can identify this q-difference equation as the quantum curve. (See Remark 3.5.)

  14. We have directly checked this up to some orders.

  15. We have directly checked this up to some orders.

  16. Stokes graph is also known as an example of spectral networks [51].

  17. These conditions are sufficient conditions because, even if the image of Lefschetz thimbles intersects, the oscillatory integral is well-defined and the saddle point approximation is valid if the corresponding vanishing cycles never intersect.

  18. In this context, the GKZ curve appears as the chiral ring studied by N. Dorey [36].

  19. The name on-shell is inherited from the (“on-shell”) vortex partition function (5.2). Indeed for the choice \(p=w_0\), \({\mathcal {J}}_X(x)\) agrees with \(Z_{\text {vortex}}^{X}(x)\).

  20. In [59, 64] the similar analysis is discussed for the colored Jones polynomial of the knot in \({\mathbb {S}}^3\).

References

  1. Aganagic, M., Vafa, C.: Mirror symmetry, D-branes and counting holomorphic discs. arXiv:hep-th/0012041

  2. Aganagic, M., Cheng, M.C.N., Dijkgraaf, R., Krefl, D., Vafa, C.: Quantum geometry of refined topological strings. JHEP 1211, 019 (2012). arXiv:1105.0630 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  3. Aganagic, M., Dijkgraaf, D., Klemm, A., Marino, M., Vafa, C.: Topological strings and integrable hierarchies. Commun. Math. Phys. 261, 451 (2006). arXiv:hep-th/0312085

    ADS  MathSciNet  MATH  Google Scholar 

  4. Aganagic, M., Klemm, A., Marino, M., Vafa, C.: The topological vertex. Commun. Math. Phys. 254, 425 (2005). arXiv:hep-th/0305132

    ADS  MathSciNet  MATH  Google Scholar 

  5. Aganagic, M., Yamazaki, M.: Open BPS wall crossing and M-theory. Nucl. Phys. B 834, 258 (2010). arXiv:0911.5342 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  6. Alday, L.F., Tachikawa, Y.: Affine \(SL(2)\) conformal blocks from 4d gauge theories. Lett. Math. Phys. 94, 87 (2010). arXiv:1005.4469 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  7. Aoki, T., Iwaki, K., Takahashi, T.: Exact WKB analysis of Schrödinger equations with a Stokes curve of loop type. Funkcialaj Ekvacioj 62, 1–34 (2019)

    MathSciNet  MATH  Google Scholar 

  8. Awata, H., Fuji, H., Kanno, H., Manabe, H., Yamada, Y.: Localization with a surface operator, irregular conformal blocks and open topological string. Adv. Theor. Math. Phys 16(3), 725 (2012). arXiv:1008.0574 [hep-th]

    MathSciNet  MATH  Google Scholar 

  9. Benini, F., Cremonesi, S.: Partition functions of \({{\cal{N}}=(2,2)}\) gauge theories on \(S^2\) and vortices. Commun. Math. Phys 334(3), 1483 (2015). arXiv:1206.2356 [hep-th]

    ADS  MATH  Google Scholar 

  10. Benini, F., Peelaers, W.: Higgs branch localization in three dimensions. JHEP 1405, 030 (2014). arXiv:1312.6078 [hep-th]

    ADS  Google Scholar 

  11. Benini, F., Zaffaroni, A.: A topologically twisted index for three-dimensional supersymmetric theories. JHEP 1507, 127 (2015). arXiv:1504.03698 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  12. 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 (1994). arXiv:hep-th/9309140

    ADS  MathSciNet  MATH  Google Scholar 

  13. Bonelli, G., Sciarappa, A., Tanzini, A., Vasko, P.: Vortex partition functions, wall crossing and equivariant Gromov–Witten invariants. Commun. Math. Phys 333(2), 717 (2015). arXiv:1307.5997 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  14. Bonelli, G., Tanzini, A., Zhao, J.: Vertices, vortices and interacting surface operators. JHEP 1206, 178 (2012). arXiv:1102.0184 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  15. Bouchard, V., Eynard, B.: Think globally, compute locally. JHEP 1302, 143 (2013). arXiv:1211.2302 [math-ph]

    ADS  MathSciNet  MATH  Google Scholar 

  16. Bouchard, V., Eynard, B.: Reconstructing WKB from topological recursion. Journal de l’Ecole polytechnique - Mathematiques 4, 845–908 (2017). arXiv:1606.04498 [math-ph]

    MathSciNet  MATH  Google Scholar 

  17. Bouchard, V., Hutchinson, J., Loliencar, P., Meiers, M., Rupert, M.: A generalized topological recursion for arbitrary ramification. Annales Henri Poincare 15, 143 (2014). arXiv:1208.6035 [math-ph]

    ADS  MathSciNet  MATH  Google Scholar 

  18. Bouchard, V., Klemm, A., Marino, M., Pasquetti, S.: Remodeling the B-model. Commun. Math. Phys. 287, 117 (2009). arXiv:0709.1453 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  19. Bouchard, V., Klemm, A., Mariño, M., Pasquetti, S.: Topological open strings on orbifolds. Commun. Math. Phys. 296, 589 (2010). arXiv:0807.0597 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  20. Braverman, A.: Instanton counting via affine Lie algebras. 1. Equivariant \(J\) functions of (affine) flag manifolds and Whittaker vectors. CRM Proc. Lecture Notes 38, 113–132 (2004). arXiv:math/0401409 [math-ag]

    MathSciNet  MATH  Google Scholar 

  21. Braverman, A., Etingof, P.: Instanton counting via affine Lie algebras II: from Whittaker vectors to the Seiberg–Witten prepotential. Studies in Lie Theory, 61–78 (2006) arXiv:math/0409441 [math-ag]

  22. Braverman, A., Feigin, B., Finkelberg, M., Rybnikov, L.: A finite analog of the AGT relation I: F inite \(W\)-algebras and quasimaps’ spaces. Commun. Math. Phys. 308, 457 (2011). arXiv:1008.3655 [math.AG]

    ADS  MATH  Google Scholar 

  23. Cecotti, C., Vafa, C.: On classification of N = 2 supersymmetric theories. Commun. Math. Phys. 158, 596 (1993). arXiv:hep-th/9211097

    MathSciNet  MATH  Google Scholar 

  24. Chiang, T.M., Klemm, A., Yau, S.T., Zaslow, E.: Local mirror symmetry: calculations and interpretations. Adv. Theor. Math. Phys. 3, 495 (1999). arXiv:hep-th/9903053

    MathSciNet  MATH  Google Scholar 

  25. Chriss, N., Ginzburg, V.: Representation Theory and Complex Geometry. Birkhäuser Mathematics, p. 508 (1997)

  26. Closset, C., Cremonesi, S., Park, D.S.: The equivariant A-twist and gauged linear sigma models on the two-sphere. JHEP 1506, 076 (2015). arXiv:1504.06308 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  27. Coates, T., Corti, A., Iritani, H., Tseng, H.: Hodge-theoretic mirror symmetry for toric stacks. arXiv:1606.07254 [math.AG]

  28. Coates, T., Givental, A.: Quantum Riemann-Roch, Lefschetz and Serre. Ann. Math. 165, 15–53 (2007). arXiv:math/0110142 [math.AG]

    MathSciNet  MATH  Google Scholar 

  29. Coates, T., Iritani, H., Jiang, Y.: The Crepant transformation conjecture for toric complete intersections. arXiv:1410.0024 [math.AG]

  30. Costin, O.: Asymptotics and Borel Summability. Monographs and Surveys in Pure and Applied Mathematics, vol. 141. Chapmann and Hall/CRC, London (2008)

    MATH  Google Scholar 

  31. Delabaere, E., Dillinger, H., Pham, F.: Résurgence de Voros et périodes des courbes hyperelliptiques. Ann. Inst. Fourier (Grenoble) 43, 163–199 (1993)

    MathSciNet  MATH  Google Scholar 

  32. Delabaere, E., Howls, C.J.: Global asymptotics for multiple integrals with boundaries. Duke Math. J. 112, 199–264 (2002)

    MathSciNet  MATH  Google Scholar 

  33. Dijkgraaf, R., Hollands, L., Sulkowski, P.: Quantum curves and D-modules. JHEP 0911, 047 (2009). arXiv:0810.4157 [hep-th]

    ADS  MathSciNet  Google Scholar 

  34. Dijkgraaf, R., Hollands, L., Sulkowski, P., Vafa, C.: Supersymmetric gauge theories, intersecting branes and free fermions. JHEP 0802, 106 (2008). arXiv:0709.4446 [hep-th]

    ADS  MathSciNet  Google Scholar 

  35. Dimofte, T., Gukov, S., Hollands, L.: Vortex counting and Lagrangian 3-manifolds. Lett. Math. Phys. 98, 225 (2011). arXiv:1006.0977 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  36. Dorey, N.: The BPS spectra of two-dimensional supersymmetric gauge theories with twisted mass terms. JHEP 9811, 005 (1998). arXiv:hep-th/9806056

    ADS  MathSciNet  MATH  Google Scholar 

  37. Doroud, N., Gomis, J., Le Floch, B., Lee, S.: Exact results in \(D=2\) supersymmetric gauge theories. JHEP 1305, 093 (2013). arXiv:1206.2606 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  38. Dubrovin, B.: Geometry of 2D topological field theories. Springer Lecture Notes in Mathematics 1620, 120–348 (1996). arXiv:hep-th/9407018

  39. Dubrovin, B.: Geometry and analytic theory of Frobenius manifolds. In: Proceedings of the International Congress of Mathematicians, Vol. II (Berlin, 1998). Doc. Math. Extra Vol. II, 315–326 (1998)

  40. Dumitrescu, O., Mulase, M.: Quantum curves for Hitchin fibrations and the Eynard–Orantin theory. Lett. Math. Phys. 104, 635 (2014). arXiv:1310.6022 [math.AG]

    ADS  MathSciNet  MATH  Google Scholar 

  41. Dunin-Barkowski, P., Mulase, M., Norbury, P., Popolitov, A., Shadrin, S.: Quantum spectral curve for the Gromov–Witten theory of the complex projective line. arXiv:1312.5336 [math-ph]

  42. Dunin-Barkowski, P., Orantin, N., Shadrin, S., Spitz, L.: Identification of the Givental formula with the spectral curve topological recursion procedure. Commun. Math. Phys. 328(2), 669–700 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  43. Eynard, B., Orantin, N.: Invariants of algebraic curves and topological expansion. Commun. Numer. Theor. Phys. 1, 347 (2007). arXiv:math/9901001

    MathSciNet  MATH  Google Scholar 

  44. Eynard, B., Orantin, N.: Computation of open Gromov–Witten invariants for toric Calabi–Yau 3-folds by topological recursion, a proof of the BKMP conjecture. Commun. Math. Phys 337(2), 483 (2015). arXiv:1205.1103 [math-ph]

    ADS  MathSciNet  MATH  Google Scholar 

  45. Fang, B.: Central charges of T-dual branes for toric varieties. arXiv:1611.05153

  46. Fang, B., Liu, C.C.M., Treumann, D., Zaslow, E.: T-duality and homological mirror symmetry of toric varieties. Adv. Math. 229, 1873–1911 (2012). arXiv:0811.1228 [math.AG]

    MathSciNet  MATH  Google Scholar 

  47. Fang, B., Liu, C.C.M., Zong, Z.: All genus open-closed mirror symmetry for affine toric Calabi–Yau 3-orbifolds. Proc. Symp. Pure Math. 93, 1 (2015). arXiv:1310.4818 [math.AG]

    MATH  Google Scholar 

  48. Fang, B., Liu, C.C.M., Zong, Z.: The Eynard–Orantin recursion and equivariant mirror symmetry for the projective line. Geom. Topol. 21, 2049–2092 (2017). arXiv:1411.3557 [math.AG]

    MathSciNet  MATH  Google Scholar 

  49. Fang, B., Liu, C.C.M., Zong, Z.: On the remodeling conjecture for toric Calabi–Yau 3-orbifolds. arXiv:1604.07123 [math.AG]

  50. Fujitsuka, M., Honda, M., Yoshida, Y.: Higgs branch localization of 3d \({\cal{N}}=2\) theories. PTEP 2014(12), 123B02 (2014). arXiv:1312.3627 [hep-th]

    MATH  Google Scholar 

  51. Gaiotto, D., Moore, G.W., Neitzke, A.: Wall-crossing in coupled 2d–4d systems. JHEP 12, 082 (2012). arXiv:1103.2598 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  52. Gaiotto, D., Moore, G.W., Neitzke, A.: Spectral networks. Ann. Henri Poincaré 14, 1643–1731 (2013). arXiv:1204.4824 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  53. Galkin, S., Golyshev, V., Iritani, H.: Gamma classes and quantum cohomology of Fano manifolds: gamma conjectures. Duke Math. J 165(11), 2005–2077 (2016). arXiv:1404.6407 [math.AG]

    MathSciNet  MATH  Google Scholar 

  54. Galkin, S., Iritani, H.: Gamma conjecture via mirror symmetry. to appear in Adv. Stud. Pure Math. arXiv:1508.00719 [math.AG]

  55. Gelfand, I.M., Graev, M.I., Zelevinsky, A.V.: Holonomic systems of equations and series of hypergeometric type. Soviet Math. Doklady 36, 5–10 (1988)

    MathSciNet  Google Scholar 

  56. Gelfand, I.M., Kapranov, M.M., Zelevinsky, A.V.: Generalized Euler integrals and A-hypergeometric functions. Adv. Math. 84, 255–271 (1990)

    MathSciNet  MATH  Google Scholar 

  57. Givental, A.: Homological geometry I. Projective hypersurfaces. Selecta Math. (N.S.) 1, 325–345 (1995)

    MathSciNet  MATH  Google Scholar 

  58. Givental, A.: Equivariant Gromov–Witten invariants. Internat Math. Res. Notices, 613–663 (1996) arXiv:alg-geom/9603021

    Google Scholar 

  59. Gukov, S.: Three-dimensional quantum gravity, Chern–Simons theory, and the A polynomial. Commun. Math. Phys. 255, 577 (2005). arXiv:hep-th/0306165

    ADS  MathSciNet  MATH  Google Scholar 

  60. Gukov, S., Sulkowski, P.: A-polynomial, B-model, and quantization. JHEP 1202, 070 (2012). arXiv:1108.0002 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  61. Gukov, S., Witten, E.: Gauge theory, ramification, and the geometric langlands program. arXiv:hep-th/0612073

  62. Guzzetti, D.: Stokes matrices and monodromy of the quantum cohomology of projective spaces. Commun. Math. Phys. 207, 341–383 (1999). arXiv:math/9904099 [math.AG]

    ADS  MathSciNet  MATH  Google Scholar 

  63. Harvey, R., Lawson, H.B.: Calibrated geometries. Acta Math. 148, 47 (1982)

    MathSciNet  MATH  Google Scholar 

  64. Hikami, K.: Generalized volume conjecture and the A-polynomials: the Neumann–Zagier potential function as a classical limit of quantum invariant. J. Geom. Phys. 57, 1895 (2007). arXiv:math/0604094 [math.QA]

    ADS  MathSciNet  MATH  Google Scholar 

  65. Honda, D., Okuda, T.: Exact results for boundaries and domain walls in 2d supersymmetric theories. JHEP 1509, 140 (2015). arXiv:1308.2217 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  66. Hori, K., Iqbal, A., Vafa, C.: D-branes and mirror symmetry. arXiv:hep-th/0005247

  67. Hori, K., Katz, S., Klemm, A., Pandharipande, R., Thomas, R., Vafa, C., Vakil, R., Zaslow, E.: Mirror Symmetry. Clay Mathematics Monographs, p. 929. American Mathematical Society, Providence (2003)

    MATH  Google Scholar 

  68. Hori, K., Romo, M.: Exact results in two-dimensional \((2,2)\) supersymmetric gauge theories with boundary. arXiv:1308.2438 [hep-th]

  69. Hori, K., Vafa, C.: Mirror symmetry. arXiv:hep-th/0002222

  70. Iqbal, A., Kashani-Poor, A.K.: The vertex on a strip. Adv. Theor. Math. Phys 10(3), 317 (2006). arXiv:hep-th/0410174

    MathSciNet  MATH  Google Scholar 

  71. Iqbal, A., Nekrasov, N., Okounkov, A., Vafa, C.: Quantum foam and topological strings. JHEP 0804, 011 (2008). arXiv:hep-th/0312022

    ADS  MathSciNet  MATH  Google Scholar 

  72. Iritani, H.: Quantum D-modules and equivariant Floer theory for free loop spaces. Math. Z. 252(3), 577–622 (2006). arXiv:math/0410487 [math.DG]

    MathSciNet  MATH  Google Scholar 

  73. Iritani, H.: An integral structure in quantum cohomology and mirror symmetry for toric orbifolds. Adv. Math. 222(3), 1016–1079 (2009). arXiv:0903.1463 [math.AG]

    MathSciNet  MATH  Google Scholar 

  74. Iritani, H.: A mirror construction for the big equivariant quantum cohomology of toric manifolds. Math. Ann. 368(1), 279–316 (2017). arXiv:1503.02919 [math.AG]

    MathSciNet  MATH  Google Scholar 

  75. Iwaki, K., Nakanishi, T.: Exact WKB analysis and cluster algebras. J. Phys. A: Math. Theor. 47, 474009 (2014). arXiv:1401.7094 [math.CA]

    MathSciNet  MATH  Google Scholar 

  76. Iwaki, K., Takahash, A.: Stokes matrices for the quantum cohomologies of orbifold projective lines. Math. Phys. A54, 101701 (2013). arXiv:1305.5775 [math.AG]

    ADS  MathSciNet  MATH  Google Scholar 

  77. Kanno, H., Tachikawa, Y.: Instanton counting with a surface operator and the chain-saw quiver. JHEP 1106, 119 (2011). arXiv:1105.0357 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  78. Katz, S.H., Klemm, A., Vafa, C.: Geometric engineering of quantum field theories. Nucl. Phys. B 497, 173 (1997). arXiv:hep-th/9609239

    ADS  MathSciNet  MATH  Google Scholar 

  79. Katz, S., Mayr, P., Vafa, C.: Mirror symmetry and exact solution of 4D \(N=2\) gauge theories: 1. Adv. Theor. Math. Phys. 1, 53 (1998). arXiv:hep-th/9706110

    MATH  Google Scholar 

  80. Katzarkov, L., Kontsevich, M., Pantev, T.: Hodge theoretic aspects of mirror symmetry. In: From Hodge theory to integrability and TQFT \({{\rm tt}}^{\ast }\)-geometry. Proceedings of Symposia on Pure Mathematics. vol. 78, pp. 87–174. American Mathematical Society, Providence, RI (2008) arXiv:0806.0107 [math.AG]

  81. Kawai, T., Takei, Y.: Algebraic Analysis of Singular Perturbation Theory. Translations of Mathematical Monographs 227, AMS, pp 129 (2005) (Japanese ver. 1998) (2005)

  82. Koike, T., Schäfke, R.: On the Borel summability of WKB solutions of Schrödinger equations with rational potentials and its application. in preparation; also Talk given by Koike, T. in the RIMS workshop “Exact WKB analysis — Borel summability of WKB solutions” September, (2010)

  83. Kozcaz, C., Pasquetti, S., Passerini, F., Wyllard, N.: Affine \(sl(N)\) conformal blocks from \({\cal{N}}=2\) \(SU(N)\) gauge theories. JHEP 1101, 045 (2011). arXiv:1008.1412 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  84. Kronheimer, P.B., Mrowka, T.S.: Gauge theory for embedded surfaces: I. Topology 32(4), 773–826 (1993)

    MathSciNet  MATH  Google Scholar 

  85. Kronheimer, P.B., Mrowka, T.S.: Knot homology groups from instantons. J. Topol. 4(4), 835–918 (2011). arXiv:0806.1053 [math.GT]

    MathSciNet  MATH  Google Scholar 

  86. Lerche, W., Mayr, P.: On \(N=1\) mirror symmetry for open type II strings. arXiv:hep-th/0111113

  87. Mariño, M.: Open string amplitudes and large order behavior in topological string theory. JHEP 0803, 060 (2008). arXiv:hep-th/0612127

    ADS  MathSciNet  Google Scholar 

  88. Mariño, M.: Chern–Simons Theory, Matrix Models, and Topological Strings, p. 197. Oxford University Press, Oxford (2015)

    MATH  Google Scholar 

  89. Mulase, M., Sulkowski, P.: Spectral curves and the Schrödinger equations for the Eynard–Orantin recursion. Adv. Theor. Math. Phys. 19, 955 (2015). arXiv:1210.3006 [math-ph]

    MathSciNet  MATH  Google Scholar 

  90. Nawata, S.: Givental J-functions, quantum integrable systems, AGT relation with surface operator. Adv. Theor. Math. Phys. 19, 1277 (2015). arXiv:1408.4132 [hep-th]

    MathSciNet  MATH  Google Scholar 

  91. Ooguri, H., Vafa, C.: Knot invariants and topological strings. Nucl. Phys. B 577, 419 (2000). https://doi.org/10.1016/S0550-3213(00)00118-8. arXiv:hep-th/9912123

    Article  ADS  MathSciNet  MATH  Google Scholar 

  92. Peelaers, W.: Higgs branch localization of \({\cal{N}}=1\) theories on \(S^3\times S^1\). JHEP 1408, 060 (2014). arXiv:1403.2711 [hep-th]

    ADS  Google Scholar 

  93. Saito, K.: Period mapping associated to a primitive form. Publ. RIMS 19, 1231–1264 (1983)

    MathSciNet  MATH  Google Scholar 

  94. Sanda, F., Shamoto, Y.: An analogue of Dubrovin’s conjecture. arXiv:1705.05989 [math.AG]

  95. Shadchin, S.: On F-term contribution to effective action. JHEP 0708, 052 (2007). arXiv:hep-th/0611278

    ADS  MathSciNet  MATH  Google Scholar 

  96. Sugishita, S., Terashima, S.: Exact results in supersymmetric field theories on manifolds with boundaries. JHEP 1311, 021 (2013). arXiv:1308.1973 [hep-th]

    ADS  MathSciNet  MATH  Google Scholar 

  97. Takei, Y.: WKB analysis and Stokes geometry of differential equations. RIMS preprint 1848, March (2016)

  98. Ueda, K.: Stokes matrices for the quantum cohomologies of Grassmannians. Int. Math. Res. Notices 34, 2075–2086 (2005). arXiv:math/0503355 [math.AG]

    MathSciNet  MATH  Google Scholar 

  99. Ueda, K.: Stokes matrix for the quantum cohomology of cubic surfaces. arXiv:math.AG/0505350

  100. Ueda, K., Yoshida, Y.: Equivariant A-twisted GLSM and Gromov–Witten invariants of CY 3-folds in Grassmannians. arXiv:1602.02487 [hep-th]

  101. Voros, A.: The return of the quartic oscillator. The complex WKB method. Ann. Inst. Henri Poincaré 39, 211–338 (1983)

    MathSciNet  MATH  Google Scholar 

  102. Witten, E.: Phases of \(N=2\) theories in two-dimensions. Nucl. Phys. B 403, 159 (1993) AMS/IP Stud. Adv. Math. 1, 143 (1996). arxiv:hep-th/9301042

  103. Yoshida, Y.: Localization of vortex partition functions in \({\cal{N}}=(2,2)\) super Yang–Mills theory. arXiv:1101.0872 [hep-th]

  104. Yoshida, Y.: Factorization of 4d \({\cal{N}}=1\) superconformal index. arXiv:1403.0891 [hep-th]

  105. Zhou, J.: Local mirror symmetry for the topological vertex. arXiv:0911.2343 [math.AG]

Download references

Acknowledgements

The authors thank Prof.  Hiroshi Iritani who suggests his idea on the equivariant version of the Dubrovin’s conjecture. KI also thanks Dr.  Fumihiko Sanda for fruitful discussion. HF and MM thank Prof. Piotr Sułkowski for stimulating discussions and useful comments. The research of HF and IS is supported by the Grant-in-Aid for Challenging Research (Exploratory) [# 17K18781]. The research of HF is also supported by the Grant-in-Aid for Scientific Research(C) [# 17K05239], and Grant-in-Aid for Scientific Research(B) [# 16H03927] from the Japan Ministry of Education, Culture, Sports, Science and Technology, and Fund for Promotion of Academic Research from Department of Education in Kagawa University. The research of KI is supported by the Grant-in-Aid for JSPS KAKENHI KIBAN(S) [# 16H06337], Young Scientists Grant-in-Aid for (B) [# 16K17613] from the Japan Ministry of Education, Culture, Sports, Science and Technology. The work of MM is supported by the ERC Starting Grant no. 335739 “Quantum fields and knot homologies” funded by the European Research Council under the European Union’s Seventh Framework Programme. The work of MM is also supported by Max-Planck-Institut für Mathematik in Bonn.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hiroyuki Fuji.

Additional information

Communicated by X. Yin

Publisher's Note

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

Appendices

Appendix A. GKZ curve from the J-function

In this appendix we will discuss a heuristic derivation of the GKZ curves from the J-functions for the projective space and complete intersections. For this purpose, we will introduce the on-shell equivariant J-function.Footnote 19 Let X be the projective space \({\mathbb {C}}\mathbf{P }_{\varvec{w}}^{N-1}\) or the complete intersection \(X=X_{\varvec{l};\varvec{w},\varvec{\lambda }}\) in \({\mathbb {C}}\mathbf{P }^{N-1}\). Let \(\ell :H^*_T({\mathbb {C}}\mathbf{P }^{N-1}) \rightarrow {\mathbb {C}}\) be a \({\mathbb {C}}\)-linear map. For the equivariant J-function \(J_{X}(x)\), the composite map \((\ell \circ J_{X})(x)\) is a \({\mathbb {C}}\)-valued function satisfying GKZ equation. We assume that \(\ell \) is a \({\mathbb {C}}\)-algebra homomorphism. Then by

$$\begin{aligned} H^*_T({\mathbb {C}}\mathbf{P }^{N-1}) \cong {\mathbb {C}}[p]/(\prod _{i=0}^{N-1}(p-w_i)), \end{aligned}$$
(A.1)

\(\ell (p)\) must be one of equivariant parameters \(w_i\)\((i=0,\ldots ,N-1)\).

Definition A.1

The function \({\mathcal {J}}_X(x)\) is named on-shell equivariant J-function, if the second equivariant cohomology element \(p\in H_T^{2}(X)\) in Proposition 2.3 is replaced with one of equivariant parameters \(w_i\)\((i=0,\ldots ,N-1)\):

$$\begin{aligned} {\mathcal {J}}_X(x)=J_{X}(x)|_{p=w_i}. \end{aligned}$$
(A.2)

By the construction, we have the following lemma:

Lemma A.2

The on-shell equivariant J-function obeys the GKZ equation.

$$\begin{aligned} {\widehat{A}}_X({\widehat{x}},{\widehat{y}}){\mathcal {J}}_X(x)=0. \end{aligned}$$
(A.3)

We remark that if \(w_i \ne w_j\) for \(i\ne j\), then \({\mathbb {C}}[p]/(\prod _{i=0}^{N-1}(p-w_i)) \cong ({\mathbb {C}}[p]/(p-w_0)) \times \cdots \times ({\mathbb {C}}[p]/(p-w_{N-1}))\) is a product of N-copies of the \({\mathbb {C}}\)-algebra \({\mathbb {C}}\). Then the \({\mathbb {C}}\)-algebra homomorphisms \(H^*_T({\mathbb {C}}\mathbf{P }^{N-1}) \rightarrow {\mathbb {C}}, p \mapsto w_i\) give a \({\mathbb {C}}\)-basis of the space of \({\mathbb {C}}\)-linear maps: \(H^*_T({\mathbb {C}}\mathbf{P }^{N-1}) \rightarrow {\mathbb {C}}\). Thus the on-shell J-functions give basis of the solution of GKZ equation.

Now we will consider the asymptotic expansion of the on-shell equivariant J-function.

$$\begin{aligned} {\mathcal {J}}_X(x)\sim \mathrm {exp}\left( \sum _{m=0}^{\infty } \hbar ^{m-1}S_m(x)\right) . \end{aligned}$$
(A.4)

Using (2.28) and this asymptotic expansion we find the defining equation of the GKZ equation \(A_X(x,y)=0\) in \((x,y)\in {\mathbb {C}}^*\times {\mathbb {C}}\) from the relation:

$$\begin{aligned} y=x\frac{dS_0(x)}{dx}\in {\mathbb {C}}. \end{aligned}$$
(A.5)

In the following we will evaluate the saddle point value \(S_0(x)\) of the on-shell equivariant J-function \({\mathcal {J}}_X(x)\) for the projective space \(X={\mathbb {C}}\mathbf{P }_{\varvec{w}}^{N-1}\) and the smooth Fano complete intersection \(X_{\varvec{l};\varvec{w},\varvec{\lambda }}\) in a heuristic way. As a consequence we will show that the defining equation \(A_X(x,y)=0\) of the GKZ curve is obtained for these two cases of X.

(1) Projective space \({\mathbb {C}}\mathbf{P }_{\varvec{w}}^{N-1}\):

Let p denote one of equivariant parameters \(w_i\)\((i=0,\ldots ,N-1)\). We focus on the factor \(\prod _{m=1}^{d}(p-w+m\hbar )\) to find the saddle point value of the on-shell equivariant J-function \({\mathcal {J}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\) in (2.7). In the \(\hbar \rightarrow 0\) limit while keeping \(d\hbar =z\) finite, we use the Riemann integral as follows:

$$\begin{aligned}&\prod _{m=1}^{d}(p-w+m\hbar )= \mathrm {exp}\left[ \sum _{m=1}^d\log (p-w+m\hbar )\right] \nonumber \\&\underset{\begin{array}{c} \hbar \rightarrow 0\\ d\hbar =u:\mathrm {finite} \end{array}}{\sim } \mathrm {exp}\left[ \frac{1}{\hbar }\int ^u_0 du'\,\log (p-w+u')\right] \nonumber \\&\quad =\mathrm {exp}\left[ \frac{1}{\hbar } \left( (u+p-w)\log (u+p-w)-u-(p-w)\log (p-w)\right) \right] . \end{aligned}$$
(A.6)

Adopting this factor we can approximate \({\mathcal {J}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\) by the integral on z as

$$\begin{aligned}&{\mathcal {J}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x) \underset{\begin{array}{c} \hbar \rightarrow 0 \end{array}}{\sim } \int _{\gamma } du \,\exp \left[ \frac{1}{\hbar }{\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u;x) \right] , \nonumber \\&\quad {\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u;x) =(u+p)\log x\nonumber \\&\quad -\sum _{i=0}^{N-1}\bigl ((u+p-w_i)\log (u+p-w_i)-u-(p-w_i)\log (p-w_i)\bigr ), \end{aligned}$$
(A.7)

where we interpret the term \((p-w_i)\log (p-w_i)\) as 0 when \(p=w_i\). Here we call \({\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u;x)\)effective superpotential, and an analytical continuation can be performed by deforming the integration path \(\gamma \) on the complex u plane.Footnote 20 Assuming such analytical continuation, we can approximate the integral (A.7) by the saddle point value in \(\hbar \rightarrow 0\) limit:

$$\begin{aligned} S_0(x)={\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u_c;x),\qquad \frac{\partial {\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u;x)}{\partial u}\Bigg |_{u=u_c} =0. \end{aligned}$$

The saddle point condition is then given by

$$\begin{aligned} \frac{\partial {\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u;x)}{\partial u}\Bigg |_{u=u_c} =\log x-\sum _{i=0}^{N-1}\log (u_c+p-w_i) =0, \end{aligned}$$
(A.8)

and by (A.5) one has

$$\begin{aligned} y=x\frac{\partial {\mathcal {W}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(u;x)}{\partial x}\Bigg |_{u=u_c} =u_c+p. \end{aligned}$$
(A.9)

By eliminating the variable \(u_c\) from these relations (A.8) and (A.9), we find a constraint equation on \((x,y)\in {\mathbb {C}}^*\times {\mathbb {C}}\):

$$\begin{aligned} \prod _{i=0}^{N-1}(y-w_i)-x=0, \end{aligned}$$
(A.10)

which agrees with the defining equation \(A_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x,y) = 0\) of the GKZ curve (2.22).

(2) Complete intersection \(X_{\varvec{l};\varvec{w},\varvec{\lambda }}\) in \({\mathbb {C}}\mathbf{P }_{\varvec{w}}^{N-1}\):

Adopting the similar approximation for the factor (A.6) in \({\mathcal {J}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\), we obtain the effective superpotential \({\mathcal {W}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(u;x)\) for the on-shell equivariant J-function \({\mathcal {J}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x)\) in (2.8) as

$$\begin{aligned}&{\mathcal {J}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x) \underset{\begin{array}{c} \hbar \rightarrow 0 \end{array}}{\sim } \int _{\gamma }du\,\exp \left[ {\mathcal {W}}_{X_{\varvec{l}; \varvec{w},\varvec{\lambda }}}(u;x)\right] ,\nonumber \\&{\mathcal {W}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(u;x)=(u+p)\log x\nonumber \\&\quad -\sum _{i=0}^{N-1}\bigl ((u+p-w_i)\log (u+p-w_i)-u-(p-w_i)\log (p-w_i)\bigr ) \nonumber \\&\quad +\sum _{a=1}^n\bigl ((l_a u+l_a p-\lambda _a) \log (l_a u+l_a p-\lambda _a)-l_a u - (l_ap-\lambda _a)\log (l_ap-\lambda _a)\bigr ), \end{aligned}$$
(A.11)

where p denotes one of equivariant parameters \(w_i\)\((i=0,\ldots ,N-1)\) and we interpret the term \((p-w_i)\log (p-w_i)\) as 0 when \(p=w_i\). The saddle point condition is then given by

$$\begin{aligned} \frac{\partial {\mathcal {W}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(u;x)}{\partial u}\Bigg |_{u=u_c}&=\log x-\sum _{i=0}^{N-1}\log (u_c+p-w_i) \nonumber \\&\quad -\sum _{a=1}^{n}\log (l_a u_c+l_ap-\lambda _a)^{-l_a} =0, \end{aligned}$$
(A.12)

and by (A.5) one has

$$\begin{aligned} y=x\frac{\partial {\mathcal {W}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(u;x)}{\partial x}\Bigg |_{u=u_c} =u_c+p. \end{aligned}$$
(A.13)

As a result of the elimination of the variable \(u_c\) from these relations (A.12) and (A.13), we find a constraint equation on \((x,y)\in {\mathbb {C}}^*\times {\mathbb {C}}\):

$$\begin{aligned} \prod _{i=0}^{N-1}(y-w_i)-x\prod _{a=1}^n(l_ay-\lambda _a)^{l_a}=0, \end{aligned}$$
(A.14)

which agrees with the defining equation \(A_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x,y) = 0\) of the GKZ curve (2.24).

Appendix B. GKZ equations for oscillatory integrals

In this appendix we will give a proof of Propositions 2.9 and 2.12.

1.1 B.1 Proof of Propositions 2.9

We will show that the oscillatory integral \({\mathcal {I}}_X(x)\) satisfies the GKZ equation for the projective space \(X={\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}\) and the Fano complete intersection \(X_{\varvec{l};\varvec{w},\varvec{\lambda }}\) separately.

Proposition B.1

The oscillatory integral \({\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\) in (2.17) satisfies the GKZ equation (2.9).

Proof

Act a differential operator \(\left( \hbar x\frac{d}{dx}-w_i\right) \left( \hbar x\frac{d}{dx}-w_0\right) \) on the oscillatory integral \({\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\),

$$\begin{aligned}&\left( \hbar x\frac{d}{dx}-w_i\right) \left( \hbar x\frac{d}{dx}-w_0\right) {\mathcal {I}}_{{\mathbb {C}} \mathbf{P }^{N-1}_{\varvec{w}}}(x)\\&\quad =\left( \hbar x\frac{d}{dx}-w_i\right) \int _{\Gamma }\prod _{i=1}^{N-1} du_i\,\frac{x}{(u_1\cdots u_{N-1})^2} \,\mathrm {e}^{\frac{1}{\hbar }W_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}} (u_1,\ldots ,u_{N-1};x)}\\&\quad =\int _{\Gamma }\prod _{i=1}^{N-1}du_i\,\frac{x}{(u_1\cdots u_{N-1})^2} \left( \frac{x}{u_1\cdots u_{N-1}}+w_0-w_i+\hbar \right) \mathrm {e}^{\frac{1}{\hbar }W_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}} (u_1,\ldots ,u_{N-1};x)}, \end{aligned}$$

where \(W_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}\) is the Landau–Ginzburg potential given in (2.14). To manipulate further we will use the following integration by parts:

$$\begin{aligned} 0&=\int _{\Gamma }\prod _{i=1}^{N-1}du_i\, \hbar \frac{d}{du_i}\left( \frac{1}{u_i} \,\mathrm {e}^{\frac{1}{\hbar }W_{{\mathbb {C}} \mathbf{P }^{N-1}_{\varvec{w}}}(u_1,\ldots ,u_{N-1};x)}\right) \\&=\int _{\Gamma }\prod _{i=1}^{N-1}du_i\,\frac{1}{u_i^2}\Biggl [ \left( u_i+w_i-w_0-\hbar -\frac{x}{u_1\cdots u_{N-1}}\right) \Biggr ] \mathrm {e}^{\frac{1}{\hbar }W_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}} (u_1,\ldots ,u_{N-1};x)}. \end{aligned}$$

In this computation, any boundary contributions do not appear, because the image of the Lefschetz thimble \(\Gamma \) is a relative cycle starting from a non-degenerate critical point \(p_{\mathrm {crit}}\) of the Landau–Ginzburg potential \(W_X\) to the infinity. Then one finds that

$$\begin{aligned}&\left( \hbar x\frac{d}{dx}-w_i\right) \left( \hbar x\frac{d}{dx}-w_0\right) {\mathcal {I}}_{{\mathbb {C}} \mathbf{P }^{N-1}_{\varvec{w}}}(x)\\&\quad =\int _{\Gamma }\prod _{i=1}^{N-1}du_i\,\frac{xu_i}{(u_1\cdots u_{N-1})^2} \,\mathrm {e}^{\frac{1}{\hbar }W_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}} (u_1,\ldots ,u_{N-1};x)}. \end{aligned}$$

Repeating the above manipulations for \(\left[ \prod _{i=1}^{N-1}\left( \hbar x\frac{d}{dx}-w_i\right) \right] \left( \hbar x\frac{d}{dx}-w_0\right) \), the following relation is obtained:

$$\begin{aligned} \begin{aligned}&\left[ \prod _{i=1}^{N-1}\left( \hbar x\frac{d}{dx}-w_i\right) \right] \left( \hbar x\frac{d}{dx}-w_0\right) {\mathcal {I}}_{{\mathbb {C}} \mathbf{P }^{N-1}_{\varvec{w}}}(x)\\&\quad =\int _{\Gamma }\prod _{i=1}^{N-1}du_i\,\frac{xu_1\cdots u_{N-1}}{(u_1\cdots u_{N-1})^2} \,\mathrm {e}^{\frac{1}{\hbar }W_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}} (u_1,\ldots ,u_{N-1};x)} =x{\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x). \end{aligned} \end{aligned}$$
(B.1)

This differential equation is the same as the GKZ equation (2.9) for the J-function \(J_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}\). \(\quad \square \)

Proposition B.2

The oscillatory integral \({\mathcal {I}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x)\) in (2.18) satisfies the GKZ equation (2.9).

Proof

Consider the GKZ equation (B.1) for the oscillatory integral \({\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\) for the mirror Landau–Ginzburg model of the projective space \(X={\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}\) denoted by

$$\begin{aligned} 0&={\widehat{A}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}} \left( {\widehat{x}},{\widehat{y}}\right) {\mathcal {I}}_{{\mathbb {C}} \mathbf{P }^{N-1}_{\varvec{w}}}(x) =\prod _{i=0}^{N-1}\left( {\widehat{y}}-w_i\right) {\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x) -{\widehat{x}}{\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x), \end{aligned}$$

where \({\widehat{x}}\) (resp. \({\widehat{y}}\)) acts on \({\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(x)\) as x (resp. \(\hbar x d/dx\)). Perform the Laplace transformation of this differential equation:

$$\begin{aligned} 0&=\int _0^{\infty }dv_1\cdots \int _0^{\infty }dv_n\, \mathrm {e}^{-\frac{\sum _{i=1}^n(v_i+\lambda _i\log v_i)}{\hbar }} \\ {}&\quad {\widehat{A}}_{{\mathbb {C}}\mathbf {P }^{N-1}_{\varvec{w}}}\left( v_1^{l_1}\cdots v_n^{l_n}{\widehat{x}},{\widehat{y}}\right) {\mathcal {I}}_{{\mathbb {C}} \mathbf {P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x)\\ {}&=\int _0^{\infty }dv_1\cdots \int _0^{\infty }dv_n\, \mathrm {e}^{-\frac{\sum _{a=1}^n(v_a+\lambda _a\log v_a)}{\hbar }}\\ {}&\quad \prod _{i=0}^{N-1}\left( \hbar x\frac{d}{dx}-w_i\right) {\mathcal {I}}_{{\mathbb {C}}\mathbf {P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x)\\ {}&\quad -\int _0^{\infty }dv_1\cdots \int _0^{\infty }dv_n\, \mathrm {e}^{-\frac{\sum _{a=1}^n(v_a+\lambda _a\log v_a)}{\hbar }} v_1^{l_1}\cdots v_n^{l_n}x{\mathcal {I}}_{{\mathbb {C}}\mathbf {P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x)\\&=\prod _{i=0}^{N-1}\left( \hbar x\frac{d}{dx}-w_i\right) {\mathcal {I}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x)\\ {}&\quad -x\int _0^{\infty }dv_1\cdots \int _0^{\infty }dv_n\, \mathrm {e}^{-\frac{\sum _{a=1}^n(v_a+\lambda _a\log v_a)}{\hbar }} v_1^{l_1}\cdots v_n^{l_n} {\mathcal {I}}_{{\mathbb {C}}\mathbf {P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x). \end{aligned}$$

To manipulate further we will use the following integration by parts repeatedly for each \(v_i\)’s:

$$\begin{aligned} 0&=\int _0^{\infty }dv_a\,\hbar \frac{d}{dv_a}\left( v_a^{m} \mathrm {e}^{-\frac{v_a+\lambda _a\log v_a}{\hbar }} {\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x) \right) \\&=\int _0^{\infty }dv_a\left( -v_a-\lambda _a+m\hbar \right) v_a^{m-1}\mathrm {e}^{-\frac{v_a+\lambda _a\log v_a}{\hbar }} {\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x) \\&\quad +\int _0^{\infty }dv_av_a^{m}\mathrm {e}^{-\frac{v_a+\lambda _a\log v_a}{\hbar }} \hbar l_a\frac{x}{v_a}\frac{d}{dx}{\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x) \\&=-\int _0^{\infty }dv_a\,v_a^{m}\mathrm {e}^{-\frac{v_a+\lambda _a\log v_a}{\hbar }} {\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x) \\&\quad +\left( l_a\hbar x\frac{d}{dx}-\lambda _a+m\hbar \right) \int _0^{\infty } dv_a\,v_a^{m-1}\mathrm {e}^{-\frac{v_a+\lambda _a\log v_a}{\hbar }} {\mathcal {I}}_{{\mathbb {C}}\mathbf{P }^{N-1}_{\varvec{w}}}(v_1^{l_1}\cdots v_n^{l_n}x). \end{aligned}$$

Then one finds the GKZ equation (2.11) for the J-function \(J_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x)\):

$$\begin{aligned} \left[ \prod _{i=0}^{N-1}\left( \hbar x\frac{d}{dx}-w_i\right) -x\prod _{a=1}^{n}\prod _{m=1}^{l_a}\left( l_a\hbar x\frac{d}{dx}-\lambda _a+m\hbar \right) \right] {\mathcal {I}}_{X_{\varvec{l};\varvec{w},\varvec{\lambda }}}(x)=0. \end{aligned}$$
(B.2)

\(\square \)

1.2 B.2 Proof of Proposition 2.12

Here we investigate the behavior of coefficients when \(x \rightarrow \infty \) in the saddle point approximation (2.34) of the oscillatory integral for

$$\begin{aligned} W_X= & {} \sum _{i=1}^{N-1} (u_i + w_i \log u_i) - \sum _{a=1}^{n} (v_a + \lambda _a \log v_a) + \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \\&+w_0 \log \left( \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \right) , \end{aligned}$$

which is mirror to \(X = X_{\varvec{w},\varvec{\lambda }}\). In this subsection we write \(W = W_X\) for simplicity.

To prove (i) in Proposition 2.12, it is enough to find an asymptotic behavior of second derivatives of W:

$$\begin{aligned} \frac{\partial ^{2} W}{\partial u_{i}^2}= & {} - \frac{w_i - w_0}{u_i^2} + \frac{2}{u_i^2} \, \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \\ \frac{\partial ^{2} W}{\partial u_{i} \partial u_j}= & {} \frac{1}{u_i u_j} \, \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \quad (i \ne j)\\ \frac{\partial ^{2} W}{\partial u_{i} \partial v_a}= & {} - \frac{1}{u_i v_a} \, \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \\ \frac{\partial ^{2} W}{\partial v_{a} \partial v_b}= & {} \frac{1}{v_a v_b} \, \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \quad (a \ne b)\\ \frac{\partial ^{2} W}{\partial v_{a}^2}= & {} \frac{\lambda _a - w_0}{v_a^2}. \end{aligned}$$

At a critical point \(({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}})\), we can use

$$\begin{aligned} \left( \frac{v_1 \cdots v_n}{u_1 \cdots u_{N-1}} x \right) \biggl |_{({\varvec{u}},{\varvec{v}}) = ({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}})} = u_i^{\mathrm{(c)}} + w_i - w_0 = v_a^{\mathrm{(c)}} + \lambda _a - w_0. \end{aligned}$$

This behaves as \(O(x^{\frac{1}{N-n}})\) in the case of (2.32), and as \(\lambda _b - w_0 + O(x^{-1})\) in the case of (2.33). Therefore, we can find the behavior

$$\begin{aligned} \mathrm{Hess}({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}}) = {\left\{ \begin{array}{ll} O(x^{-\frac{N+n-1}{N-n}}) &{} \text {in the case of } (2.32), \\ O(x^{2}) &{} \text {in the case of }(2.33) \end{array}\right. } \end{aligned}$$

of the Hessian when \(x \rightarrow \infty \). The claim of (i) in Proposition 2.12 follows immediately from this computation.

Let us prove (ii) in Proposition 2.12. We take a coordinate at a critical point \(({\varvec{u}}^{\mathrm{(c)}}, {\varvec{v}}^\mathrm{(c)})\) of W:

$$\begin{aligned} {\varvec{\xi }} = (\xi _1, \dots , \xi _{N+n-1}) = (u_1 - u_1^\mathrm{(c)},\dots ,u_{N-1} - u_{N-1}^{\mathrm{(c)}}, v_1 - v_1^{\mathrm{(c)}}, \dots , v_n - v_n^{\mathrm{(c)}}), \end{aligned}$$

and consider the Taylor expansion of W at the critical point \({\varvec{\xi }}^{\mathrm{(c)}} = {\varvec{0}}\):

$$\begin{aligned} W({\varvec{\xi }};x)= & {} W({\varvec{\xi }}^{\mathrm{(c)}};x) + \frac{1}{2!}\sum _{i,j} W_{ij}({\varvec{\xi }}^{\mathrm{(c)}};x) \, \xi _i \xi _j \nonumber \\&+ \sum _{m \ge 3} \frac{1}{m!} \sum _{i_1, \dots , i_m} W_{i_1 \dots i_m}({\varvec{\xi }}^\mathrm{(c)};x) \, \xi _{i_1} \cdots \xi _{i_m}, \end{aligned}$$
(B.3)

where \(W_{i_1 \dots i_m} = (\partial ^m W)/(\partial \xi _{i_1} \cdots \partial \xi _{i_m})\). Since we have chosen generic \(w_i\) and \(\lambda _a\) so that the critical points are non-degenerate, the Hesse matrix H has non-zero determinant at \({\varvec{\xi }}^{\mathrm{(c)}}\). We further take a linear transformation \(\xi _i = (- \hbar )^{1/2}\sum _j C_{i}^j(x) s_j\) of the coordinates which transforms the quadratic part of W as

$$\begin{aligned} \frac{1}{2 \hbar }\sum _{i,j} W_{ij}({\varvec{\xi }}^{\mathrm{(c)}};x)\, \xi _i \xi _j = - \frac{1}{2} \sum _{i=1}^{N+n-1} s_i^2. \end{aligned}$$
(B.4)

We can find these coefficients \(C_{i}^{j}(x)\) by applying the simultaneous completing the square to the quadratic form in the left hand side of (B.4). Eventually we can find the behavior of the coefficients \(C_{i}^{j}(x)\) for large x as follows:

  • Let us consider the case when the critical point \(({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}})\) behaves as (2.32) in Lemma 2.11. Since \(W_{ij}({\varvec{\xi }}^\mathrm{(c)};x)\) behaves as \(O(x^{- \frac{1}{N-n}})\) in this case, we can show that

    $$\begin{aligned} C_{i}^{j}(x) = O(x^{\frac{1}{2(N-n)}}) \end{aligned}$$
    (B.5)

    holds for all \(i,j=1,\dots , N+n-1\), when \(x \rightarrow \infty \).

  • Let us consider the case when the critical point \(({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}})\) behaves as (2.33) in Lemma 2.11. Let \(i_b \in \{N,\dots ,N+n-1 \}\) be the label of \(v_b\); that is, \(\xi _{i_b} = v_b - v_b^{\mathrm{(c)}} \). Since we can arrange the quadratic part of W as

    $$\begin{aligned}&W_{i_b i_b}({\varvec{\xi }}^{\mathrm{(c)}};x) \xi _{i_b}^2 + 2\sum _{i \ne i_b} W_{i \, i_b}({\varvec{\xi }}^{\mathrm{(c)}};x) \xi _{i_b} \xi _i + \sum _{i,j \ne i_b} W_{ij}({\varvec{\xi }}^{\mathrm{(c)}};x) \xi _{i} \xi _j \\&\quad =W_{i_b i_b}({\varvec{\xi }}^{\mathrm{(c)}};x) \left( \xi _{i_b} + \sum _{i \ne i_b} \frac{W_{i\, i_b}({\varvec{\xi }}^{\mathrm{(c)}};x)}{W_{i_b i_b}({\varvec{\xi }}^{\mathrm{(c)}};x)} \xi _i \right) ^2 \\&\qquad - \frac{1}{W_{i_b i_b}({\varvec{\xi }}^{\mathrm{(c)}};x)} \left( \sum _{i \ne i_b} W_{i\, i_b}({\varvec{\xi }}^{\mathrm{(c)}};x) \xi _i\right) ^2 + \sum _{i,j \ne i_b} W_{ij}({\varvec{\xi }}^{\mathrm{(c)}};x) \xi _{i} \xi _j. \end{aligned}$$

    Thus we can choose

    $$\begin{aligned} s_{i_b} = W_{i_b i_b}^{1/2}({\varvec{\xi }}^{\mathrm{(c)}};x) \left( \xi _{i_b} + \sum _{i \ne i_b} \frac{W_{i\, i_b}({\varvec{\xi }}^{\mathrm{(c)}};x)}{W_{i_b i_b}({\varvec{\xi }}^{\mathrm{(c)}};x)} \xi _i \right) \end{aligned}$$

    as one of new coordinates, and other \(s_i\)’s are written in terms of \(\xi _i\)’s except for \(\xi _{i_b}\). Since \(W_{i_b i_b}({\varvec{\xi }}^\mathrm{(c)};x) = O(x^2)\), \( W_{i\, i_b}({\varvec{\xi }}^{\mathrm{(c)}};x) = O(x)\) if \(i \ne i_b\) and \(W_{ij}({\varvec{\xi }}^{\mathrm{(c)}};x)=O(1)\) for \(i,j \ne i_b\), we can conclude

    $$\begin{aligned} C_{i}^{j}(x) = {\left\{ \begin{array}{ll} O(x^{-1}) &{} \text {if }i = i_b \\ O(1) &{} \text {otherwise} \end{array}\right. } \end{aligned}$$
    (B.6)

    hold when \(x \rightarrow \infty \).

Let us proceed the computation of saddle point expansion. The above change of the coordinate yields

$$\begin{aligned} d\xi _1 \cdots d\xi _{N+n-1} = \frac{(-\hbar )^{\frac{N+n-1}{2}}}{\sqrt{\mathrm{Hess}({\varvec{u}}^\mathrm{(c)},{\varvec{v}}^{\mathrm{(c)}})}} \, ds_1\cdots ds_{N+n-1}. \end{aligned}$$

Then, by the standard argument of the saddle point method, the asymptotic expansion of the oscillatory integral is computed by term-wise integration:

$$\begin{aligned} {{\mathcal {I}}}_{X}(x)\sim & {} {} \exp \left( \frac{1}{\hbar } W({\varvec{\xi }}^\mathrm {(c)};x) \right) \, \frac{(-\hbar )^{\frac{N+n-1}{2}}}{u^{\mathrm {(c)}}_1 \cdots u^{\mathrm {(c)}}_{N-1} \,\sqrt{\mathrm {Hess}({\varvec{\xi }}^{\mathrm {(c)}})}}\nonumber \\&\times \left( 1 + \sum _{m = 1}^{\infty } \hbar ^{\frac{m}{2}} \sum _{i_1, \dots , i_{m}} f_{i_1 \dots i_m} \int _{{{\mathbb {R}}}^{N+n-1}} ds_1 \cdots ds_{N+n-1} \right. \nonumber \\&\left. \qquad \exp \left( - \frac{1}{2} \sum _{i=1}^{N+n-1} s_i^2 \right) s_{i_1} \cdots s_{i_m} \right) , \end{aligned}$$
(B.7)

where \(f_{i_1 \dots i_m}\) is the Taylor coefficient given by

$$\begin{aligned}&\frac{u^{\mathrm{(c)}}_1 \cdots u^{\mathrm{(c)}}_{N-1}}{u_1 \cdots u_{N-1}}\exp \left( \frac{1}{\hbar } \sum _{m \ge 3} \frac{1}{m!} \sum _{i_1, \dots , i_m} W_{i_1 \dots i_m}({\varvec{\xi }}^{\mathrm{(c)}};x) \xi _{i_1} \cdots \xi _{i_m} \right) \Biggl |_{\xi _i = (- \hbar )^{1/2}\sum _j C_{i}^j(x) s_j} \\&\quad =1 + \sum _{m = 1}^{\infty } \hbar ^{\frac{m}{2}} \sum _{i_1, \dots , i_{m}} f_{i_1 \dots i_m} s_{i_1} \cdots s_{i_m}. \end{aligned}$$
  • If \(({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}})\) behaves as (2.32), then \(W_{i_1 \dots i_m}({\varvec{\xi }}^{\mathrm{(c)}};x) = O(x^{- \frac{m-1}{N-n}})\), and hence,

    $$\begin{aligned} W_{i_1 \dots i_m}({\varvec{\xi }}^{\mathrm{(c)}};x) \, C_{i_1}^{j_1} \cdots C_{i_m}^{j_m} = O(x^{- \frac{m-2}{2(N-n)}}) \quad \text {for any }j_1, \dots , j_m. \end{aligned}$$

    For \(m \ge 3\), this tends to 0 when \(x \rightarrow \infty \).

  • If \(({\varvec{u}}^{\mathrm{(c)}},{\varvec{v}}^{\mathrm{(c)}})\) behaves as (2.33), then \(W_{i_1 \dots i_m}({\varvec{\xi }}^{\mathrm{(c)}};x) = O(x^{\ell _b})\), where \(\ell _b\) is the number of \(i_b\) in the indices \(i_1, \dots , i_m\). Therefore,

    $$\begin{aligned} W_{i_1 \dots i_m}({\varvec{\xi }}^{\mathrm{(c)}};x) \, C_{i_1}^{j_1} \cdots C_{i_m}^{j_m} = O(1) \quad \text {for any }j_1, \dots , j_m. \end{aligned}$$

We can also verify that

$$\begin{aligned} \frac{u_i^{\mathrm{(c)}}}{u_i} = \left( 1 + (-\hbar )^{1/2} \sum _{i} \frac{C_i^{j}}{u_i^{\mathrm{(c)}}} s_j \right) ^{-1}, \end{aligned}$$

and the coefficient satisfies

$$\begin{aligned} \frac{C_i^{j}}{u_i^{\mathrm{(c)}}} = {\left\{ \begin{array}{ll} O(x^{- \frac{1}{2(N-n)}}) &{} \text {for the case of }(2.32) \\ O(1) &{} \text {for the case of }(2.33) \end{array}\right. } \end{aligned}$$

when \(x \rightarrow \infty \). Therefore, we can conclude that

$$\begin{aligned} f_{i_1 \dots i_m} = {\left\{ \begin{array}{ll} O(x^{- \frac{1}{2(N-n)}}) &{} \text {for the case of } (2.32) \\ O(1) &{} \text {for the case of }(2.33) \end{array}\right. } \end{aligned}$$
(B.8)

for \(m \ge 1\). After evaluating the Gaussian integrals in (B.7) by using

$$\begin{aligned} \int _{{\mathbb {R}}} ds_i\, e^{- \frac{1}{2} s_i^2} \, s_i^{k} = {\left\{ \begin{array}{ll} 0 &{} \text {if }k\text { is odd}, \\ \displaystyle \sqrt{2\pi } \, {(k-1)!!} &{} \text {if }k\text { is even}, \end{array}\right. } \end{aligned}$$

we obtain the saddle point approximation (2.34). In particular, (B.8) proves the claim (ii) in Proposition 2.12.

Appendix C. Computational results by iteration and topological recursion

In this appendix we will firstly give some explicit computational results of the WKB solutions to the GKZ equations. In Appendix C.2 we will explicitly perform the WKB reconstruction (3.13) for the equivariant \({{\mathbb {C}}}\mathbf{P }^1\) model, and see agreements with the results in Appendix C.1.

1.1 C.1. Some iterative computations for the GKZ equation

Assume the saddle point approximation of the oscillatory integral

$$\begin{aligned} {\mathcal {I}}_X(x)\sim \exp \left( \sum _{m=0}^{\infty }\hbar ^{m-1}S_m(x)\right) , \end{aligned}$$

one finds a set of the first order differential equations for \(S_m\)’s by expanding the GKZ equation around \(\hbar =0\).

\(\underline{{\mathbb {C}}\mathbf{P }^{N-1}\hbox { model}}\)

The GKZ equation for the (non-equivariant) \({\mathbb {C}}\mathbf{P }^{N-1}\) model is

$$\begin{aligned} \left[ \left( \hbar x\frac{d}{dx}\right) ^{N}-x\right] {\mathcal {I}}_{{\mathbb {C}} \mathbf{P }^{N-1}}(x)=0. \end{aligned}$$
(C.1)

Some computational results of \(S_m\)’s for \(N=2, 3, 4\) are listed in Table 2.

Table 2 \(S_m(x)\) for the \({\mathbb {C}}\mathbf{P }^{N-1}\) models (\(N=2,3,4\)) obtained from the GKZ equation (C.1)

\(\underline{\hbox {Equivariant }{\mathbb {C}}\mathbf{P }^{1}\hbox { model}}\)

The GKZ equation for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model is

$$\begin{aligned} \left[ \left( \hbar x\frac{d}{dx}-w_0\right) \left( \hbar x\frac{d}{dx}-w_1\right) -x\right] {\mathcal {I}}_{{\mathbb {C}} \mathbf{P }^{1}_{\varvec{w}}}(x)=0. \end{aligned}$$
(C.2)

For this model there are two solutions which have the formal power series expansion:

$$\begin{aligned} {\mathcal {I}}^{(\pm )}_{{\mathbb {C}}\mathbf{P }^{1}_{\varvec{w}}}(x)\sim \exp \left( \sum _{m=0}^{\infty }\hbar ^{m-1}S_m^{(\pm )}(x) \right) . \end{aligned}$$

Computational results of \(S^{(\pm )}_m\) for \(m=0,1,2,3,4\) are listed in Table 3 modulo constant shifts.

Table 3 \(S_m^{(\pm )}(x)\) for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model obtained from the GKZ equation (C.2)

\(\underline{\hbox {Degree 1 hypersurface in }{\mathbb {C}}\mathbf{P }^{1}}\)

The GKZ equation for the degree 1 hypersurface \(X_{\varvec{w},\lambda }=X_{l=1;\varvec{w},\lambda }\) in \({\mathbb {C}}\mathbf{P }^{1}\) is

$$\begin{aligned} \left[ \left( \hbar x\frac{d}{dx}-w_0\right) \left( \hbar x\frac{d}{dx}-w_1\right) -x\left( \hbar x\frac{d}{dx}-\lambda +\hbar \right) \right] {\mathcal {I}}_{X_{\varvec{w}, \lambda }}(x)=0. \end{aligned}$$
(C.3)

For this model we also find two solutions which have the formal power series expansion:

$$\begin{aligned} {\mathcal {I}}^{(\pm )}_{X_{\varvec{w},\lambda }}(x) \sim \exp \left( \sum _{m=0}^{\infty }\hbar ^{m-1}S_m^{(\pm )}(x) \right) . \end{aligned}$$

Computational results of \(S^{(\pm )}_m\) for \(m=0,1,2,3\) are listed in Table 4 modulo constant shifts.

Table 4 \(S_m^{(\pm )}(x)\) for degree 1 hypersurface in \({\mathbb {C}}\mathbf{P }^{1}\) obtained from the GKZ equation (C.3)

Especially, focusing on the \(S^{(\pm )}_1(x)\) we find the following expansion around \(x=\infty \):

$$\begin{aligned} \begin{aligned} \mathrm {e}^{S^{(+)}_1(x)}&=1+\frac{(w_0-\lambda )(w_1-\lambda )}{2x^2}+O(x^{-3}),\\ \mathrm {e}^{S^{(-)}_1(x)}&=\frac{1}{x}-\frac{w_0+w_1 -2\lambda }{x^2}+O(x^{-3}), \end{aligned} \end{aligned}$$

and these asymptotic expansions are consistent with Proposition 2.12 (i).

1.2 C.2 Topological recursion for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model

In the following, for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model, we will explicitly recover the computational result in Table 3 by applying the topological recursion (3.9) in [43]. The GKZ curve

$$\begin{aligned} \Sigma _{{\mathbb {C}}\mathbf{P }^{1}_{\varvec{w}}}=\Big \{\; (x,y)\in {\mathbb {C}}^*\times {\mathbb {C}}\; \Big |\; (y-w_0)(y-w_1)-x=0\; \Big \} \end{aligned}$$

is parametrized by a local coordinate z as follows:

$$\begin{aligned} x(z)=z^2-\Lambda ,\qquad y(z)=z+\frac{1}{2}(w_0+w_1),\qquad \Lambda =\frac{1}{4}(w_0-w_1)^2. \end{aligned}$$

The spectral curve \(\Sigma _{{\mathbb {C}}\mathbf{P }^{1}_{\varvec{w}}}\) has only one simple ramification point at \(z=0\) in this local coordinate. Starting from

$$\begin{aligned} \omega _{1}^{(0)}(z)=0,\qquad \omega _{2}^{(0)}(z_1,z_2)=B(z_1,z_2)=\frac{dz_1dz_2}{(z_1-z_2)^2}, \end{aligned}$$

the differentials \(\omega _{n}^{(g)}\) for \((g,n) \ne (0,1), (0,2)\) are defined by the topological recursion (3.9)

$$\begin{aligned} \begin{aligned} \omega _{n+1}^{(g)}(z,\varvec{z}_N)&= \mathop {\mathrm {Res}}_{w=0}\, \frac{\int _{-w}^{w}B(\cdot , z)}{2\left( y(w)-y(-w)\right) dx(w)/x(w)}\bigg (\omega _{n+2}^{(g-1)} (w,-w,\varvec{z}_N)\\&\quad +\sum _{\ell =0}^{g}\sum _{\emptyset =J\subseteq N}\omega _{|J|+1}^{(g-\ell )}(w,\varvec{z}_J)\omega _{|N|-|J|+1}^{(\ell )} (-w,\varvec{z}_{N \backslash J}) \bigg ), \end{aligned} \end{aligned}$$

where \(N=\{1,2,\ldots ,n\}\supset J=\{i_1,i_2,\ldots ,i_j\}\), and \(N\backslash J=\{i_{j+1},i_{j+2},\ldots ,i_n\}\). Integrating these multi-differentials, one finds the free energies

$$\begin{aligned} \begin{aligned} F_{1}^{(0)}(x)&=\int ^z_{z_*} y(z')\frac{dx(z')}{x(z')},\qquad F_{2}^{(0)}(x)=\int ^z_{z_*}\int ^z_{z_*}\left( B(z_1',z_2') -\frac{dx(z_1')dx(z_2')}{(x(z_1')-x(z_2'))^2}\right) ,\\ F_{n}^{(g)}(x)&=\int ^z_{z_*} \cdots \int ^z_{z_*} \omega ^{(g)}_{n} (z_1',\ldots ,z_n'),\quad (g,n) \ne (0,1), (0,2), \end{aligned} \end{aligned}$$
(C.4)

where \(z_*\) denotes a reference point.

The WKB reconstruction (3.13) of wave function is defined by \(F_{n}^{(g)}(x)\)’s as

$$\begin{aligned} \psi _{{\mathbb {C}}\mathbf{P }^{1}_{\varvec{w}}}(x)= \exp \left( \sum _{g=0,n=1}^{\infty }\frac{\hbar ^{2g-2+n}}{n!} F^{(g)}_{n}(x)\right) . \end{aligned}$$
(C.5)

We fix the reference point by \(z_*=\infty \) which corresponds to \(x(z_*)=\infty \). Here note that \(F_{1}^{(0)}\) and \(F_{2}^{(0)}\) need to be regularized by certain constant shifts so as to depend on \(z_*\). Some explicit computational results of the free energies \(F_n^{(g)}(x)\) are listed in Table 5.

Table 5 Free energies for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model. Here we have two types free energies corresponding to the branches \(z=\pm \sqrt{x+\Lambda }\)

The wave function (C.5) is reorganized by

$$\begin{aligned} \psi _{{\mathbb {C}}\mathbf{P }^{1}_{\varvec{w}}}(x)=\exp \left( \sum _{m=0}^{\infty }\hbar ^{m-1}F_m(x)\right) ,\qquad F_m(x)=\sum _{\begin{array}{c} g\ge 0,\;n\ge 1,\\ 2g+n-1=m \end{array}}\frac{1}{n!}F_n^{(g)}(x), \end{aligned}$$

where corresponding to two branches \(z=\pm \sqrt{x+\Lambda }\) we find two types of free energies \(F_m(x)=F_m^{(\pm )}(x)\). Since the leading term \(S_0(x)\) of the asymptotic expansion of the J-function obeys

$$\begin{aligned} x\frac{dS_0(x)}{dx}=y(x), \end{aligned}$$

the free energy \(F_1^{(0)}(x)\) in (C.4) agrees with \(S_0^{(\pm )}(x)\) up to a constant shift. Using the computational results in Table 5, \(F_m^{(\pm )}(x)\)’s (\(m\ge 1\)) are computed immediately and summarized in Table 6.

Table 6 \(F_m(x)=F_m^{(\pm )}(x)\) for the equivariant \({\mathbb {C}}\mathbf{P }^{1}\) model

Comparing the computational results in Tables 3 and 6, one finds the agreement

$$\begin{aligned} F_m^{(\pm )}(x)=S_m^{(\pm )}(x), \end{aligned}$$

for \(m=1,2,3,4\) up to a constant shift of \(F_1^{(\pm )}(x)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fuji, H., Iwaki, K., Manabe, M. et al. Reconstructing GKZ via Topological Recursion. Commun. Math. Phys. 371, 839–920 (2019). https://doi.org/10.1007/s00220-019-03590-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03590-6

Navigation