Skip to main content
Log in

Quantum curves and q-deformed Painlevé equations

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

Abstract

We propose that the grand canonical topological string partition functions satisfy finite-difference equations in the closed string moduli. In the case of genus one mirror curve, these are conjectured to be the q-difference Painlevé equations as in Sakai’s classification. More precisely, we propose that the tau functions of q-Painlevé equations are related to the grand canonical topological string partition functions on the corresponding geometry. In the toric cases, we use topological string/spectral theory duality to give a Fredholm determinant representation for the above tau functions in terms of the underlying quantum mirror curve. As a consequence, the zeroes of the tau functions compute the exact spectrum of the associated quantum integrable systems. We provide details of this construction for the local \(\mathbb {P}^1\times \mathbb {P}^1\) case, which is related to q-difference Painlevé with affine \(A_1\) symmetry, to SU(2) Super Yang–Mills in five dimensions and to relativistic Toda system.

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

Similar content being viewed by others

Notes

  1. In [26], based on [27, 28], a different type of finite-dimensional determinant was considered to compute \(\tau \) functions.

  2. We would like to thank Marcos Mariño for a discussion on this point.

  3. In Sakai’s classification [31], this corresponds to surface type \(A_7^{(1)'} \) and symmetry type \(A_1^{(1)}\), where the superscript (1) stands for affine extension of the Dynkin algebra. Notice that this is not the unique q-P equation leading to differential Painlevé \({\mathrm{III}}_3\) in the continuous limit. Indeed, as pointed out for instance in [39], also the q-P equation with surface type \(A_7^{(1)} \) and symmetry type \(A_1^{(1)'}\) makes contact with Painlevé \({\mathrm{III}}_3\). This is perfectly consistent with the picture developed in this paper since the corresponding Newton polygon is identified with local \({\mathbb {F}}_1\) (see Fig. 1). By the geometric engineering construction [18, 19], we know that topological string theory on both \({\mathbb {F}}_1\) and \({\mathbb {P}}^1\times {\mathbb {P}}^1\) reduces to pure SU(2) theory in four dimensions. However, in this work we denote by q-P \({\mathrm{III}}_3\) only the one associated with surface type \(A_7^{(1)'} \) and symmetry type \(A_1^{(1)}\).

  4. For the higher genus generalization, see [22].

  5. Provided some positivity constraints are imposed on the mass parameters and \(\hbar \).

  6. This corresponds to ABJM theory with level \(k=2\) [57]. In the ABJM context, the self-dual point corresponds to an enhancement of the supersymmetry from \({\mathcal {N}}=6\) to \({\mathcal {N}}=8\).

  7. For the sake of notation, we will simply denote \(\Xi ^{\mathrm{TS}}(\kappa , \xi , \hbar )\) instead of \(\Xi ^{\mathrm{TS}}_{{\mathbb {P}}^1\times {\mathbb {P}}^1}(\kappa , \xi , \hbar )\).

  8. This corresponds to ABJ theory with level \(k=2\) and gauge group \(U(N)\times U(N+1)\) [57].

  9. After the preliminary version of this work was sent to Yasuhiko Yamada, he proved the equality between the two spectral problems arising in the quantization of the two polygons in Fig. 4.

  10. We thank Yasuhiko Yamada for discussions on this point.

  11. Alternatively one can in principle approach the autonomous limit starting from the result at generic C, but in this case one has to implement a suitable regularization scheme, see [86].

  12. For the sake of notation, we omit the subscript \({\mathbb {P}}^1\times {\mathbb {P}}^1\) in \( \Xi ^{\mathrm{TS}}\).

  13. We thank Yasuhiko Yamada for bringing our attention on this identity.

  14. We thank Jie Gu for discussions on this point.

  15. We follow the notation of [8].

  16. By a detailed numerical analysis of the spectrum of the operator (5.8).

  17. After the preliminary version of this paper appeared, it was shown in [86] that the equations found in [20] for the one-period phase of the 4d SU(N) gauge theory hold also for a more generic phase where all the SU(N) periods are non-vanishing. Moreover, in [86] the corresponding finite-difference equations were also presented. We expect the solution to such difference equations to be given by the generalized Fredholm determinants associated with the \(Y^{N,m} \) geometries computed in [20, 22].

References

  1. Seiberg, N., Witten, E.: Monopoles, duality and chiral symmetry breaking in \(\text{ N }=2\) supersymmetric QCD. Nucl. Phys. B 431, 484–550 (1994). arXiv:hep-th/9408099

    ADS  MathSciNet  MATH  Google Scholar 

  2. Seiberg, N., Witten, E.: Electric–magnetic duality, monopole condensation, and confinement in N = 2 supersymmetric Yang–Mills theory. Nucl. Phys. B 426, 19–52 (1994). arXiv:hep-th/9407087

    ADS  MathSciNet  MATH  Google Scholar 

  3. Martinec, E.J., Warner, N.P.: Integrable systems and supersymmetric gauge theory. Nucl. Phys. B 459, 97–112 (1996). arXiv:hep-th/9509161

    ADS  MathSciNet  MATH  Google Scholar 

  4. Gorsky, A., Krichever, I., Marshakov, A., Mironov, A., Morozov, A.: Integrability and Seiberg–Witten exact solution. Phys. Lett. B 355, 466–474 (1995). arXiv:hep-th/9505035

    ADS  MathSciNet  MATH  Google Scholar 

  5. Nekrasov, N., Okounkov, A.: Seiberg–Witten theory and random partitions. Prog. Math. 244, 525–596 (2006). arXiv:hep-th/0306238

    MathSciNet  MATH  Google Scholar 

  6. Nekrasov, N.A., Shatashvili, S.L.: Quantization of integrable systems and four dimensional gauge theories. arXiv:0908.4052

  7. Iorgov, N., Lisovyy, O., Tykhyy, Yu.: Painlevé VI connection problem and monodromy of \(c=1\) conformal blocks. JHEP 12, 029 (2013). arXiv:1308.4092

    ADS  MATH  Google Scholar 

  8. Its, A., Lisovyy, O., Tykhyy, Yu.: Connection problem for the sine-Gordon/Painlevé III tau function and irregular conformal blocks. Int. Math. Res. Not. 18, 8903–8924 (2015). arXiv:1403.1235

    MATH  Google Scholar 

  9. Gamayun, O., Iorgov, N., Lisovyy, O.: Conformal field theory of Painlevé VI. JHEP 10, 038 (2012). arXiv:1207.0787

    ADS  MATH  Google Scholar 

  10. Iorgov, N., Lisovyy, O., Teschner, J.: Isomonodromic tau-functions from Liouville conformal blocks. Commun. Math. Phys. 336, 671–694 (2015). arXiv:1401.6104

    ADS  MathSciNet  MATH  Google Scholar 

  11. Bershtein, M.A., Shchechkin, A.I.: Bilinear equations on Painlevé \(\tau \) functions from CFT. Commun. Math. Phys. 339, 1021–1061 (2015). arXiv:1406.3008

    ADS  MATH  Google Scholar 

  12. Gamayun, O., Iorgov, N., Lisovyy, O.: How instanton combinatorics solves Painlevé VI, V and IIIs. J. Phys. A 46, 335203 (2013). arXiv:1302.1832

    MathSciNet  MATH  Google Scholar 

  13. Bonelli, G., Lisovyy, O., Maruyoshi, K., Sciarappa, A., Tanzini, A.: On Painlevé/gauge theory correspondence. Lett. Math. Phys. 107, 2359–2413 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  14. Nagoya, H.: Conformal blocks and Painlevé functions. arXiv:1611.08971

  15. Gavrylenko, P., Lisovyy, O.: Fredholm determinant and Nekrasov sum representations of isomonodromic tau functions. Commun. Math. Phys. 363, 1–58 (2018). arXiv:1608.00958

    ADS  MathSciNet  MATH  Google Scholar 

  16. Gavrylenko, P., Lisovyy, O.: Pure \(SU(2)\) gauge theory partition function and generalized Bessel kernel. Proc. Symp. Pure Math. 18, 181–208 (2018). arXiv:1705.01869

    MathSciNet  Google Scholar 

  17. Bonelli, G., Grassi, A., Tanzini, A.: Seiberg–Witten theory as a Fermi gas. Lett. Math. Phys. 107, 1–30 (2017). arXiv:1603.01174

    ADS  MathSciNet  MATH  Google Scholar 

  18. Katz, S.H., Klemm, A., Vafa, C.: Geometric engineering of quantum field theories. Nucl. Phys. B 497, 173–195 (1997). arXiv:hep-th/9609239

    ADS  MathSciNet  MATH  Google Scholar 

  19. Klemm, A., Lerche, W., Mayr, P., Vafa, C., Warner, N.P.: Selfdual strings and N = 2 supersymmetric field theory. Nucl. Phys. B 477, 746–766 (1996). arXiv:hep-th/9604034

    ADS  MATH  Google Scholar 

  20. Bonelli, G., Grassi, A., Tanzini, A.: New results in \(\cal{N}=2\) theories from non-perturbative string. Ann. Henri Poincare 19, 743–774 (2018). arXiv:1704.01517

    MathSciNet  MATH  Google Scholar 

  21. Grassi, A., Hatsuda, Y., Marino, M.: Topological strings from quantum mechanics. Ann. Henri Poincaré (2016). arXiv:1410.3382

  22. Codesido, S., Grassi, A., Mariño, M.: Spectral theory and mirror curves of higher genus. Ann. Henri Poincaré 18, 559–622 (2017). arXiv:1507.02096

    ADS  MathSciNet  MATH  Google Scholar 

  23. Gavrylenko, P.: Isomonodromic \(\tau \)-functions and \(\text{ W }_{N}\) conformal blocks. JHEP 09, 167 (2015). arXiv:1505.00259

    ADS  MathSciNet  MATH  Google Scholar 

  24. Bershtein, M.A., Shchechkin, A.I.: q-Deformed Painlevé \(\tau \) function and q-deformed conformal blocks. J. Phys. A 50, 085202 (2017). arXiv:1608.02566

    ADS  MathSciNet  MATH  Google Scholar 

  25. Jimbo, M., Nagoya, H., Sakai, H.: CFT approach to the \(q\)-Painlevé vi equation. J. Integr. Syst. (2017). https://doi.org/10.1093/integr/xyx009

  26. Mironov, A., Morozov, A.: q-Painlev equation from Virasoro constraints. Phys. Lett. B 785, 207–210 (2018). arXiv:1708.07479

    ADS  MATH  Google Scholar 

  27. Mironov, A., Morozov, A.: Check-operators and quantum spectral curves. SIGMA 13, 047 (2017). arXiv:1701.03057

    MathSciNet  MATH  Google Scholar 

  28. Mironov, A., Morozov, A.: On determinant representation and integrability of Nekrasov functions. Phys. Lett. B 773, 34–46 (2017). arXiv:1707.02443

    ADS  MathSciNet  MATH  Google Scholar 

  29. Kajiwara, K., Noumi, M., Yamada, Y.: Geometric aspects of Painlevé equations. J. Phys. A: Math. Theor. 50, 073001 (2017). arXiv:1509.08186

    ADS  MATH  Google Scholar 

  30. Grammaticos, B., Ramani, A.: Discrete Painlevé Equations: A Review, pp. 245–321. Springer, Berlin (2004)

    MATH  Google Scholar 

  31. Sakai, H.: Rational surfaces associated with a ne root systems and geometry of the Painlevé equations. Commun. Math. Phys. 220, 165–229 (2001)

    ADS  MATH  Google Scholar 

  32. Minahan, J.A., Nemeschansky, D.: Superconformal fixed points with E(n) global symmetry. Nucl. Phys. B 489, 24–46 (1997). arXiv:hep-th/9610076

    ADS  MathSciNet  MATH  Google Scholar 

  33. Mizoguchi, S., Yamada, Y.: W(E(10)) symmetry, M theory and Painleve equations. Phys. Lett. B 537, 130–140 (2002). arXiv:hep-th/0202152

    ADS  MathSciNet  MATH  Google Scholar 

  34. Yamada, Y.: Amoebae of type e. http://www.math.kobe-u.ac.jp/~yamaday/aE.pdf (2017). Accessed 1 Oct 2017

  35. Ormerod, C.M., Yamada, Y.: From polygons to ultradiscrete Painlevé equations. arXiv:1408.5643

  36. Goncharov, A.B., Kenyon, R.: Dimers and cluster integrable systems. arXiv:1107.5588

  37. Fock, V.V., Marshakov, A.: Loop groups, clusters, dimers and integrable systems. arXiv:1401.1606

  38. Witten, E.: Quantum background independence in string theory. arXiv:hep-th/9306122

  39. Grammaticos, B., Ramani, A.: Parameterless discrete Painlevé equations and their Miura relations. J. Nonlinear Math. Phys. 23, 141 (2016)

    MathSciNet  Google Scholar 

  40. Zamolodchikov, A.B.: Painleve III and 2-d polymers. Nucl. Phys. B 432, 427–456 (1994). arXiv:hep-th/9409108

    ADS  MathSciNet  MATH  Google Scholar 

  41. McCoy, B.M., Tracy, C.A., Wu, T.T.: Painleve functions of the third kind. J. Math. Phys. 18, 1058 (1977)

    ADS  MathSciNet  MATH  Google Scholar 

  42. Grassi, A., Hatsuda, Y., Marino, M.: Quantization conditions and functional equations in ABJ(M) theories. J. Phys. A 49, 115401 (2016). arXiv:1410.7658

    ADS  MathSciNet  MATH  Google Scholar 

  43. Huang, M.-X., Klemm, A., Poretschkin, M.: Refined stable pair invariants for E-, M- and \([p, q]\)-strings. JHEP 1311, 112 (2013). arXiv:1308.0619

    ADS  MATH  Google Scholar 

  44. Huang, M.-X., Klemm, A., Reuter, J., Schiereck, M.: Quantum geometry of del Pezzo surfaces in the Nekrasov–Shatashvili limit. JHEP 1502, 031 (2015). arXiv:1401.4723

    ADS  MathSciNet  MATH  Google Scholar 

  45. Kashaev, R., Mariño, M., Zakany, S.: Matrix models from operators and topological strings, 2. Ann. Henri Poincaré 17, 2741–2781 (2016). arXiv:1505.02243

    ADS  MathSciNet  MATH  Google Scholar 

  46. Batyrev, V.V.: Dual polyhedra and mirror symmetry for Calabi–Yau hypersurfaces in toric varieties. J. Alg. Geom. 3, 493–545 (1994). arXiv:alg-geom/9310003

    MathSciNet  MATH  Google Scholar 

  47. Chiang, T., Klemm, A., Yau, S.-T., Zaslow, E.: Local mirror symmetry: calculations and interpretations. Adv. Theor. Math. Phys. 3, 495–565 (1999). arXiv:hep-th/9903053

    MathSciNet  MATH  Google Scholar 

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

  49. Kashaev, R., Marino, M.: Operators from mirror curves and the quantum dilogarithm. Commun. Math. Phys. 346, 967 (2016). arXiv:1501.01014

    ADS  MathSciNet  MATH  Google Scholar 

  50. Laptev, A., Schimmer, L., Takhtajan, L.A.: Weyl type asymptotics and bounds for the eigenvalues of functional-difference operators for mirror curves. Geom. Funct. Anal. 26, 288–305 (2016). arXiv:1510.00045

    MathSciNet  MATH  Google Scholar 

  51. Marino, M., Putrov, P.: ABJM theory as a Fermi gas. J. Stat. Mech. 1203, P03001 (2012). arXiv:1110.4066

    MathSciNet  Google Scholar 

  52. Hatsuda, Y., Moriyama, S., Okuyama, K.: Instanton effects in ABJM theory from Fermi gas approach. JHEP 1301, 158 (2013). arXiv:1211.1251

    ADS  MathSciNet  MATH  Google Scholar 

  53. Calvo, F., Marino, M.: Membrane instantons from a semiclassical TBA. JHEP 1305, 006 (2013). arXiv:1212.5118

    ADS  Google Scholar 

  54. Hatsuda, Y., Moriyama, S., Okuyama, K.: Instanton bound states in ABJM theory. JHEP 1305, 054 (2013). arXiv:1301.5184

    ADS  MATH  Google Scholar 

  55. Hatsuda, Y., Marino, M., Moriyama, S., Okuyama, K.: Non-perturbative effects and the refined topological string. JHEP 1409, 168 (2014). arXiv:1306.1734

    ADS  MathSciNet  MATH  Google Scholar 

  56. Marino, M.: Spectral theory and mirror symmetry. Proc. Symp. Pure Math. 98, 259 (2018). arXiv:1506.07757

    MathSciNet  Google Scholar 

  57. Codesido, S., Grassi, A., Marino, M.: Exact results in N = 8 Chern–Simons-matter theories and quantum geometry. JHEP 1507, 011 (2015). arXiv:1409.1799

    ADS  MathSciNet  MATH  Google Scholar 

  58. Sun, K., Wang, X., Huang, M.-X.: Exact quantization conditions, toric Calabi–Yau and nonperturbative topological string. JHEP 01, 061 (2017). arXiv:1606.07330

    ADS  Google Scholar 

  59. Marino, M., Zakany, S.: Matrix models from operators and topological strings. Ann. Henri Poincare 17, 1075–1108 (2016). arXiv:1502.02958

    ADS  MathSciNet  MATH  Google Scholar 

  60. Gu, J., Klemm, A., Marino, M., Reuter, J.: Exact solutions to quantum spectral curves by topological string theory. JHEP 10, 025 (2015). arXiv:1506.09176

    ADS  MathSciNet  MATH  Google Scholar 

  61. Okuyama, K., Zakany, S.: TBA-like integral equations from quantized mirror curves. JHEP 03, 101 (2016). arXiv:1512.06904

    ADS  MATH  Google Scholar 

  62. Wang, X., Zhang, G., Huang, M.-X.: New exact quantization condition for toric Calabi–Yau geometries. Phys. Rev. Lett. 115, 121601 (2015). arXiv:1505.05360

    ADS  Google Scholar 

  63. Hatsuda, Y., Marino, M.: Exact quantization conditions for the relativistic Toda lattice. JHEP 05, 133 (2016). arXiv:1511.02860

    ADS  MathSciNet  MATH  Google Scholar 

  64. Huang, M.-X., Wang, X.-F.: Topological strings and quantum spectral problems. JHEP 1409, 150 (2014). arXiv:1406.6178

    ADS  MathSciNet  MATH  Google Scholar 

  65. Franco, S., Hatsuda, Y., Marino, M.: Exact quantization conditions for cluster integrable systems. J. Stat. Mech. 1606, 063107 (2016). arXiv:1512.03061

    MathSciNet  Google Scholar 

  66. Grassi, A.: Spectral determinants and quantum theta functions. J. Phys. A 49, 505401 (2016). arXiv:1604.06786

    MathSciNet  MATH  Google Scholar 

  67. Hatsuda, Y., Katsura, H., Tachikawa, Y.: Hofstadter’s butterfly in quantum geometry. New J. Phys. 18, 103023 (2016). arXiv:1606.01894

    ADS  MathSciNet  Google Scholar 

  68. Mariño, M., Zakany, S.: Exact eigenfunctions and the open topological string. J. Phys. A 50, 325401 (2017). arXiv:1606.05297

    MathSciNet  MATH  Google Scholar 

  69. Codesido, S., Gu, J., Mariño, M.: Operators and higher genus mirror curves. JHEP 02, 092 (2017). arXiv:1609.00708

    ADS  MathSciNet  MATH  Google Scholar 

  70. Grassi, A., Gu, J.: BPS relations from spectral problems and blowup equations. arXiv:1609.05914

  71. Sciarappa, A.: Exact relativistic Toda chain eigenfunctions from separation of variables and gauge theory. JHEP 10, 116 (2017). arXiv:1706.05142

    ADS  MathSciNet  MATH  Google Scholar 

  72. Couso-Santamaría, R., Marino, M., Schiappa, R.: Resurgence matches quantization. J. Phys. A 50, 145402 (2017). arXiv:1610.06782

    ADS  MathSciNet  MATH  Google Scholar 

  73. Sugimoto, Y.: Geometric transition in the nonperturbative topological string. Phys. Rev. D 94, 055010 (2016). arXiv:1607.01534

    ADS  MathSciNet  Google Scholar 

  74. Marino, M., Zakany, S.: Wavefunctions, integrability, and open strings. arXiv:1706.07402

  75. Hatsuda, Y., Sugimoto, Y., Xu, Z.: Calabi–Yau geometry and electrons on 2d lattices. Phys. Rev. D 95, 086004 (2017). arXiv:1701.01561

    ADS  MathSciNet  Google Scholar 

  76. Gu, J., Huang, M.-X., Kashani-Poor, A.-K., Klemm, A.: Refined BPS invariants of 6d SCFTs from anomalies and modularity. JHEP 05, 130 (2017). arXiv:1701.00764

    ADS  MathSciNet  MATH  Google Scholar 

  77. Grassi, A., Marino, M.: The complex side of the TS/ST correspondence. J. Phys. A 52, 055402 (2019). arXiv:1708.08642

    ADS  Google Scholar 

  78. Kashaev, R.M., Sergeev, S.M.: Spectral equations for the modular oscillator. arXiv:1703.06016

  79. Aganagic, M., Bouchard, V., Klemm, A.: Topological strings and (almost) modular forms. Commun. Math. Phys. 277, 771–819 (2008). arXiv:hep-th/0607100

    ADS  MathSciNet  MATH  Google Scholar 

  80. Marino, M., Putrov, P.: Exact results in ABJM theory from topological strings. JHEP 1006, 011 (2010). arXiv:0912.3074

    ADS  MathSciNet  MATH  Google Scholar 

  81. Drukker, N., Marino, M., Putrov, P.: From weak to strong coupling in ABJM theory. Commun. Math. Phys. 306, 511–563 (2011). arXiv:1007.3837

    ADS  MathSciNet  MATH  Google Scholar 

  82. Jimbo, M., Sakai, H.: A q-analog of the sixth Painlevé equation. Lett. Math. Phys. 38, 145–154 (1996)

    ADS  MathSciNet  MATH  Google Scholar 

  83. Quispel, G.R.W., Roberts, J.A.G., Thompson, C.J.: Integrable mappings and soliton equations. Phys. Lett. A 127, 419–421 (1988)

    ADS  MathSciNet  MATH  Google Scholar 

  84. Quispel, G.R.W., Roberts, J.A.G., Thompson, C.J.: Integrable mappings and soliton equations II. Physica D 34, 183–192 (1989)

    ADS  MathSciNet  MATH  Google Scholar 

  85. Tsuda, T.: Integrable mappings via rational elliptic surfaces. J. Phys. A: Math. Gen. 37, 2721 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

  86. Bershtein, M., Gavrylenko, P., Marshakov, A.: Cluster Toda chains and Nekrasov functions. Teor. Mat. Fiz. 198, 179 (2019). arXiv:1804.10145

    Google Scholar 

  87. Aganagic, M., Cheng, M.C., Dijkgraaf, R., Krefl, D., Vafa, C.: Quantum geometry of refined topological strings. JHEP 1211, 019 (2012). arXiv:1105.0630

    ADS  MathSciNet  MATH  Google Scholar 

  88. Hatsuda, Y., Okuyama, K.: Resummations and non-perturbative corrections. JHEP 09, 051 (2015). arXiv:1505.07460

    MathSciNet  MATH  Google Scholar 

  89. Gromak, V.: Reducibility of the Painlevé equations. Differ. Equ. 20, 1191–1198 (1984)

    MathSciNet  MATH  Google Scholar 

  90. Bershtein, M.A., Shchechkin, A.I.: Backlund transformation of Painleve III(\(D_8\)) tau function. J. Phys. A 50, 115205 (2017). arXiv:1608.02568

    ADS  MathSciNet  MATH  Google Scholar 

  91. Bridgeland, T.: Riemann–Hilbert problems for the resolved conifold. arXiv:1703.02776

  92. Scalise, J., Stoppa, J.: Variations of BPS structure and a large rank limit. J. Inst. Math. Jussieu (2019). https://doi.org/10.1017/S1474748019000136

    Google Scholar 

  93. Faddeev, L., Kashaev, R.: Quantum dilogarithm. Mod. Phys. Lett. A 9, 427–434 (1994). arXiv:hep-th/9310070

    ADS  MathSciNet  MATH  Google Scholar 

  94. Faddeev, L.: Discrete Heisenberg–Weyl group and modular group. Lett. Math. Phys. 34, 249–254 (1995). arXiv:hep-th/9504111

    ADS  MathSciNet  MATH  Google Scholar 

  95. Tracy, C.A., Widom, H.: Proofs of two conjectures related to the thermodynamic Bethe ansatz. Commun. Math. Phys. 179, 667–680 (1996). arXiv:solv-int/9509003

    ADS  MathSciNet  MATH  Google Scholar 

  96. Honda, M., Okuyama, K.: Exact results on ABJ theory and the refined topological string. JHEP 1408, 148 (2014). arXiv:1405.3653

    ADS  MathSciNet  MATH  Google Scholar 

  97. Awata, H., Hirano, S., Shigemori, M.: The partition function of ABJ theory. Prog. Theor. Exp. Phys. 2013, 053B04 (2013). arXiv:1212.2966

    Google Scholar 

  98. Honda, M.: Direct derivation of “mirror” ABJ partition function. JHEP 1312, 046 (2013). arXiv:1310.3126

    ADS  Google Scholar 

  99. Matsumoto, S., Moriyama, S.: ABJ fractional brane from ABJM Wilson loop. JHEP 1403, 079 (2014). arXiv:1310.8051

    ADS  Google Scholar 

  100. Grassi, A., Marino, M.: M-theoretic matrix models. JHEP 1502, 115 (2015). arXiv:1403.4276

    ADS  MathSciNet  MATH  Google Scholar 

  101. Bazhanov, V.V., Lukyanov, S.L., Zamolodchikov, A.B.: Integrable structure of conformal field theory. 2. Q operator and DDV equation. Commun. Math. Phys. 190, 247–278 (1997). arXiv:hep-th/9604044

    ADS  MathSciNet  MATH  Google Scholar 

  102. Putrov, P., Yamazaki, M.: Exact ABJM partition function from TBA. Mod. Phys. Lett. A 27, 1250200 (2012). arXiv:1207.5066

    ADS  MathSciNet  MATH  Google Scholar 

  103. Hatsuda, Y., Okuyama, K.: Probing non-perturbative effects in M-theory. JHEP 1410, 158 (2014). arXiv:1407.3786

    ADS  MathSciNet  MATH  Google Scholar 

  104. Aharony, O., Bergman, O., Jafferis, D.L.: Fractional M2-branes. JHEP 0811, 043 (2008). arXiv:0807.4924

    ADS  MathSciNet  Google Scholar 

  105. Moriyama, S., Nosaka, T., Yano, K.: Superconformal Chern–Simons theories from del Pezzo geometries. JHEP 11, 089 (2017). arXiv:1707.02420

    ADS  MathSciNet  MATH  Google Scholar 

  106. Moriyama, S., Nakayama, S., Nosaka, T.: Instanton effects in rank deformed superconformal Chern–Simons theories from topological strings. JHEP 08, 003 (2017). arXiv:1704.04358

    ADS  MathSciNet  MATH  Google Scholar 

  107. Cecotti, S., Vafa, C.: Topological antitopological fusion. Nucl. Phys. B 367, 359–461 (1991)

    ADS  MATH  Google Scholar 

  108. Cecotti, S., Gaiotto, D., Vafa, C.: \(tt^*\) geometry in 3 and 4 dimensions. JHEP 05, 055 (2014). arXiv:1312.1008

    ADS  MathSciNet  MATH  Google Scholar 

  109. Bullimore, M., Kim, H.-C., Koroteev, P.: Defects and quantum Seiberg–Witten geometry. JHEP 05, 095 (2015). arXiv:1412.6081

    ADS  MathSciNet  MATH  Google Scholar 

  110. Bullimore, M., Kim, H.-C.: The superconformal index of the (2,0) theory with defects. JHEP 05, 048 (2015). arXiv:1412.3872

    ADS  MathSciNet  MATH  Google Scholar 

  111. Klemm, A., Zaslow, E.: Local mirror symmetry at higher genus. arXiv:hep-th/9906046

  112. Bershadsky, M., Cecotti, S., Ooguri, H., Vafa, C.: Holomorphic anomalies in topological field theories. Nucl. Phys. B 405, 279–304 (1993). arXiv:hep-th/9302103

    ADS  MathSciNet  MATH  Google Scholar 

  113. Hatsuda, Y.: unpublished

  114. Mkrtchyan, R.L.: Nonperturbative universal Chern–Simons theory. JHEP 09, 054 (2013). arXiv:1302.1507

    ADS  MathSciNet  MATH  Google Scholar 

  115. Eguchi, T., Kanno, H.: Topological strings and Nekrasov’s formulas. JHEP 12, 006 (2003). arXiv:hep-th/0310235

    ADS  MathSciNet  Google Scholar 

  116. Taki, M.: Refined topological vertex and instanton counting. JHEP 03, 048 (2008). arXiv:0710.1776

    ADS  MathSciNet  Google Scholar 

  117. Iqbal, A., Kashani-Poor, A.-K.: SU(N) geometries and topological string amplitudes. Adv. Theor. Math. Phys. 10, 1–32 (2006). arXiv:hep-th/0306032

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We would like to thank Mikhail Bershtein, Jie Gu, Oleg Lisovyy, Marcos Mariño, Andrei Mironov, Alexei Morozov and especially Yasuhiko Yamada for many useful discussions and correspondence. We are also thankful to Jie Gu for a careful reading of the manuscript. We thank Bernard Julia and the participants of the workshop “Exceptional and ubiquitous Painlevé equations for Physics” for useful discussions and the stimulating atmosphere. The work of GB is supported by the PRIN project “Non-perturbative Aspects Of Gauge Theories And Strings”. GB and AG acknowledge support by INFN Iniziativa Specifica ST&FI. AT acknowledges support by INFN Iniziativa Specifica GAST.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alba Grassi.

Additional information

Publisher's Note

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

Appendices

The grand potential: definitions

In this section we review the definition of the topological string grand potential \({\mathsf {J}} _X\) associated with a toric CY X with genus one mirror curve. We mainly follow the notation of [22, 56]. As in Sect. 2.1, we denote

$$\begin{aligned} \kappa =\mathrm{e}^{\mu } \end{aligned}$$
(A.1)

the “true” modulus of X, \(\mathbf m_X\) the set of mass parameters and \({\varvec{\xi }}\) the rescaled mass parameters. The Kähler parameters of X are denoted by

$$\begin{aligned} t_i , \quad i=1, \ldots s \end{aligned}$$
(A.2)

and can be expressed in terms of the complex moduli through the mirror map. When \(\kappa \) is large we have (see [22, 56] and references therein)

$$\begin{aligned} t_i=c_i \mu +\sum _{j=1}^{r_X}a_{ij} \log m_X^{(j)}+\Pi (\kappa ^{-1}, \mathbf{m}_X) \end{aligned}$$
(A.3)

where \(\Pi \) is a convergent series in \(\kappa ^{-1}\) while \(c_i, a_{ij}\) are constants which can be read from the toric data of the CY [43, 44]. For instance for local \( {\mathbb {P}}^1\times {\mathbb {P}}^1\) we have \(s=2\), \(r_X=1\), \(c_1=c_2=2\), \(a_{11}=0\) and \(a_{21}=-1\). Hence,

$$\begin{aligned} t_2=t_1-\log m_{\mathrm{{\mathbb {P}}^1 \times {\mathbb {P}}^1}}. \end{aligned}$$
(A.4)

Geometrically the mirror map can be expressed as the A-period of the underlying mirror curve (2.4). Since this is a genus one curve we denote by \({\mathcal {A}}\) the corresponding A-cycle. Then we have

$$\begin{aligned} t_1\propto \oint _{{\mathcal {A}}} p(x) \mathrm{d}x, \end{aligned}$$
(A.5)

where p(x) is determined by Eq. (2.4). As explained in Sect. 2.1 one can promote the mirror curve to an operator \({{\mathsf {O}}}_{ X}\) [see Eq. (2.8)]. As a consequence, the differential \(p(x) \mathrm{d}x\) is promoted to a quantum differential

$$\begin{aligned} p(x,\hbar ) \mathrm{d}x \end{aligned}$$
(A.6)

fulfilling

$$\begin{aligned} {{\mathsf {O}}}_{ X} \left( {1\over \sqrt{ p(x, \hbar )}}\exp \left[ {\mathrm{i}\over \hbar } \int ^x p(y, \hbar )\mathrm{d}y\right] \right) =0. \end{aligned}$$
(A.7)

The quantum mirror map is then defined as [87]

$$\begin{aligned} t_1(\hbar ) \propto \oint _{{\mathcal {A}}} p(x, \hbar ) \mathrm{d}x. \end{aligned}$$
(A.8)

As explained in [87], such map can be computed efficiently at large \(\kappa \) and one has

$$\begin{aligned} t_i ( \hbar )=c_i \mu +\sum _{j=1}^{r_X}a_{ij} \log m_X^{(j)}+\Pi (\kappa ^{-1}, \mathbf{m}_X, \hbar ), \end{aligned}$$
(A.9)

where \(\Pi \) is a series expansion in \(\kappa ^{-1}\) but it is exact in \(\hbar \). For instance, when X is the canonical bundle over \({\mathbb {P}}^1\times {\mathbb {P}}^1\), we have

$$\begin{aligned} t_1 ( \hbar )=t(\mu , \xi , \hbar ), \quad t_2 ( \hbar )= t(\mu , \xi , \hbar )-{\hbar \over 2 \pi } \xi , \end{aligned}$$
(A.10)

where the first few terms in the expansion are

$$\begin{aligned} \begin{aligned} t(\mu , \xi ,\hbar )&= 2 \mu -2 (m_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}+1) \kappa ^{-2}\\&\quad +\kappa ^{-4} \left( -3 m_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}^2-\frac{2 m_{{\mathbb {P}}^1 \times {\mathbb {P}}^1} \left( {\mathrm{e}}^{2\mathrm{i}\hbar }+4 {\mathrm{e}}^{\mathrm{i}\hbar }+1\right) }{{\mathrm{e}}^{\mathrm{i}\hbar }}-3\right) +O\left( \kappa ^{-6}\right) , \\&\quad \kappa =\mathrm{e}^{\mu }, \quad m_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}=\mathrm{e}^{{\hbar \over 2 \pi } \xi } . \end{aligned} \end{aligned}$$
(A.11)

Note that in this example we have a series in \(\kappa ^{-2}\). For a generic genus one geometry, we would have a series in \(\kappa ^{-r}\) where r is determined by the canonical class/intersection data of the geometry. A list of r-parameters for others genus one geometries can be found for instance in Table 1 of [21].

When \(\hbar \) is real, it has been checked numerically for many geometries that the series \(\Pi \) in (A.9) at large \(\kappa \) has a finite radius of convergence (see for instance [55]). For generic complex values of \(\hbar \), instead one has to perform a partial resummation, but it is still possible to organize (A.9) into a convergent series by using instanton calculus in 5d gauge theory. Indeed, from a 5d gauge theory viewpoint the quantum mirror map can be computed from the vev of the Wilson loop in the fundamental representation, see for instance [109, 110]. For the example of local \({\mathbb {P}}^1\times {\mathbb {P}}^1\), this is illustrated for instance in [77].

We introduce the topological string free energy

$$\begin{aligned} F^{\mathrm{top}}_{\mathrm{X}}\left( \mathbf{t}, g_s\right) ={1\over 6 g_s^2} a_{ijk} t_i t_j t_k +b_i t_i +F^{\mathrm{GV}}_X\left( \mathbf{t} , g_s\right) \end{aligned}$$
(A.12)

with

$$\begin{aligned} F^{\mathrm{GV}}_X\left( \mathbf{t}, g_s\right) =\sum _{g\ge 0} \sum _\mathbf{d} \sum _{w=1}^\infty {1\over w} n_g^{ \mathbf{d}} \left( 2 \sin { w g_s \over 2} \right) ^{2g-2} \mathrm{e}^{-w \mathbf{d} \cdot \mathbf{t}}, \end{aligned}$$
(A.13)

where \(n_g^{ \mathbf{d}}\) are the Gopakumar–Vafa invariants of X and \(g_s\) is the string coupling constant. Moreover,

$$\begin{aligned} \mathbf{d}=\{d_1, \ldots ,d_s\} \end{aligned}$$
(A.14)

and the sum over \( \mathbf{d}\) in (A.13) runs over all integers

$$\begin{aligned} d_i \in {\mathbb {Z}}, \quad d_i\ge 0 \end{aligned}$$
(A.15)

such that

$$\begin{aligned} \sum _{i=1}^s d_i^2 \ne 0. \end{aligned}$$
(A.16)

The coefficients \(a_{ijk}, b_i \) are determined by the classical data of X. In the limit \(g_s \rightarrow 0\), we have

$$\begin{aligned} F^{\mathrm{top}}_{\mathrm{X}}\left( \mathbf{t}, g_s\right) \sim \sum _{g\ge 0} F_g(\mathbf{t}) g_s^{2g-2}, \end{aligned}$$
(A.17)

where \(F_g(\mathbf{t}) \) are called the genus g free energies of topological string. For instance, we have

$$\begin{aligned} \begin{aligned} F_0(\mathbf{t})&={1\over 6} a_{ijk} t_i t_j t_k + \sum _{\mathbf{d}} N_0^{ \mathbf{d}} \mathrm{e}^{-\mathbf{d} \cdot \mathbf{t}},\\ F_1(\mathbf{t})&=b_i t_i + \sum _{\mathbf{d}} N_1^{ \mathbf{d}} \mathrm{e}^{-\mathbf{d} \cdot \mathbf{t}}, \end{aligned} \end{aligned}$$
(A.18)

where \(N_g^\mathbf{d}\) are the Gromov–Witten invariants. When X is a toric CY, one has explicit expressions for (A.18) in terms of hypergeometric and standard functions (see for instance [55, 80, 111]).

Let us discuss briefly the convergence properties of (A.13). We first note that (A.13) has poles for \(\pi ^{-1}g_s \in {\mathbb {Q}}\) which makes it ill defined on the real \(g_s\) axis. If instead \(g_s \in {\mathbb {C}}/{\mathbb {R}}\), then (A.13) diverges as a series in \(\mathrm{e}^{- \mathbf{t}}\). Nevertheless, by using instanton calculus it is possible to partially resummate it and organize it into a convergent series [24] at least if \(g_s \in {\mathbb {C}}/{\mathbb {R}}\).

Similarly, we define the Nekrasov–Shatashvili free energy as

$$\begin{aligned} F^{\mathrm{NS}}(\mathbf{t}, \hbar ) ={1\over 6 \hbar } a_{ijk} t_i t_j t_k +b^{\mathrm{NS}}_i t_i \hbar +F^{\mathrm{NS}}_{\mathrm{inst}}(\mathbf{t}, \hbar ), \end{aligned}$$
(A.19)

where

$$\begin{aligned} F^{\mathrm{NS}}_{\mathrm{inst}}(\mathbf{t}, \hbar ) =\sum _{j_L, j_R} \sum _{w\ge 1 } \sum _{\mathbf{d} } N^{\mathbf{d}}_{j_L, j_R} \frac{\sin \frac{\hbar w}{2}(2j_L+1)\sin \frac{\hbar w}{2}(2j_R+1)}{2 w^2 \sin ^3\frac{\hbar w}{2}} \mathrm{e}^{-w \mathbf{d}\cdot \mathbf{t}} \end{aligned}$$
(A.20)

and \(N^{\mathbf{d}}_{j_L, j_R}\) are the refined BPS invariants of X. The sum over \(\mathbf{d}\) is as in (A.13). Moreover, there exists a constant vector \(\mathbf{B} \), called the B-field, such that

$$\begin{aligned} N^{\mathbf{d}}_{j_L, j_R}\ne 0 \quad \leftrightarrow \quad (-1)^{2j_L + 2 j_R+1}= (-1)^{\mathbf{B} \cdot \mathbf{d}}. \end{aligned}$$
(A.21)

For local \( {\mathbb {P}}^1 \times {\mathbb {P}}^1 \), this can be set to zero [21, 55]. When \(\hbar \rightarrow 0\), we recover the following genus expansion

$$\begin{aligned} F^{\mathrm{NS}}(\mathbf{t}, \hbar ) =\sum _{g\ge 0} F_g^{\mathrm{NS}}(\mathbf{t})\hbar ^{2g-2}. \end{aligned}$$
(A.22)

The convergent properties for the NS free energy are analogous to the ones of (A.13). The topological string grand potential is defined as

$$\begin{aligned} \mathsf {J}_{X}({\mu }, \varvec{\xi },\hbar ) = \mathsf {J}^{\mathrm{WKB}}_X ({\mu }, \varvec{\xi },\hbar )+ \mathsf {J}^{\mathrm{WS}}_X ({\mu }, \varvec{\xi } , \hbar ), \end{aligned}$$
(A.23)

where

$$\begin{aligned} \mathsf {J}^{\mathrm{WS}}_X({\mu }, \varvec{\xi }, \hbar )=F^{\mathrm{GV}}_X\left( {2 \pi \over \hbar }{} \mathbf{t}(\hbar )+ \pi \mathrm{i}\mathbf{B} , {4 \pi ^2 \over \hbar } \right) . \end{aligned}$$
(A.24)

Moreover,

$$\begin{aligned} \mathsf {J}^{\mathrm{WKB}}_X({\mu }, \varvec{\xi }, \hbar )= & {} {t_i(\hbar ) \over 2 \pi } {\partial F^{\mathrm{NS}}(\mathbf{t}(\hbar ), \hbar ) \over \partial t_i} +{\hbar ^2 \over 2 \pi } {\partial \over \partial \hbar } \left( {F^{\mathrm{NS}}(\mathbf{t}(\hbar ), \hbar ) \over \hbar } \right) \nonumber \\&+ {2 \pi \over \hbar } b_i t_i(\hbar ) + A({\varvec{\xi }}, \hbar ), \end{aligned}$$
(A.25)

where \( A({\varvec{\xi }}, \hbar )\) denotes the so-called constant map contribution [112]. It is important to notice that, even tough both \(\mathsf {J}^{\mathrm{WS}}_X\) and \(\mathsf {J}^{\mathrm{WKB}}_X\) have a dense set of poles on the real \(\hbar \) axis, their sum (A.23) is free of poles. We usually refer to this mechanism as the HMO cancellation mechanism since it was first discovered in [52] in the context of ABJM theory. Moreover, we have

$$\begin{aligned} \mathsf {J}^{\mathrm{WKB}}_X({\mu }, \varvec{\xi }, \hbar )= & {} {1\over 12 \pi \hbar } a_{ijk} t_i(\hbar ) t_j(\hbar ) t_k(\hbar ) + \left( {2 \pi b_i \over \hbar } + {\hbar b_i^{\mathrm{NS}} \over 2 \pi } \right) t_i(\hbar ) \nonumber \\&+\, {\mathcal {O}}\left( \mathrm{e}^{-t_i(\hbar )} \right) + A({\varvec{\xi }}, \hbar ). \end{aligned}$$
(A.26)

Hence, it is convenient to split

$$\begin{aligned} \mathsf {J}^{\mathrm{WKB}}_X({\mu }, \varvec{\xi }, \hbar )= \mathrm{P}_X({{\mu }, \varvec{\xi }, \hbar })+ \mathsf {J}^{\mathrm{WKB, inst}}_X({\mu }, \varvec{\xi }, \hbar )+ A({\varvec{\xi }}, \hbar ) \end{aligned}$$
(A.27)

where \({\mathrm{P}}_X\) encodes the polynomial part in \(t_i\) of \( \mathsf {J}^{\mathrm{WKB}}_X\) and

$$\begin{aligned} \mathsf {J}^{{\mathrm{WKB, inst}}}_X({\mu }, \varvec{\xi }, \hbar )= {t_i(\hbar ) \over 2 \pi } {\partial F^{\mathrm{NS}}_{\mathrm{inst}}(\mathbf{t}(\hbar ), \hbar ) \over \partial t_i} +{\hbar ^2 \over 2 \pi } {\partial \over \partial \hbar } \left( {F^{\mathrm{NS}}_{\mathrm{inst}}(\mathbf{t}(\hbar ), \hbar ) \over \hbar } \right) . \end{aligned}$$
(A.28)

For local \( {\mathbb {P}}^1 \times {\mathbb {P}}^1 \), we have

$$\begin{aligned} {\mathrm{P}}_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}(\mu , \xi ,\hbar ) = -\frac{\xi t(\mu ,\xi ,\hbar )^2}{16 \pi ^2}+\frac{t(\mu ,\xi ,\hbar )^3}{12 \pi \hbar }-\frac{\hbar t(\mu ,\xi ,\hbar )}{24 \pi }+\frac{\pi t(\mu ,\xi ,\hbar )}{6 \hbar }-\frac{\xi }{24}. \end{aligned}$$
(A.29)

The constant map contribution for local \( {\mathbb {P}}^1 \times {\mathbb {P}}^1 \) reads [113]

$$\begin{aligned} A( \xi ,\hbar )= A_p(\xi ,\hbar )-J_{\mathrm{CS}}\left( {2\pi ^2 \over \hbar },\mathrm{i}\pi +{1\over 2} \xi \right) , \end{aligned}$$
(A.30)

where

$$\begin{aligned} A_p(\xi ,\hbar )={\hbar ^2 \over (4 \pi ^2)^2}\left[ \frac{\xi ^3}{24}+\frac{\pi ^2 \xi }{6 }\right] +A_c\left( {\hbar \over \pi }\right) , \end{aligned}$$
(A.31)

with

$$\begin{aligned} A_{\mathrm{c}}(k)= \frac{2\zeta (3)}{\pi ^2 k}\left( 1-\frac{k^3}{16}\right) +\frac{k^2}{\pi ^2} \int _0^\infty \frac{x}{\mathrm{e}^{k x}-1}\log (1-\mathrm{e}^{-2x})\mathrm{d}x. \end{aligned}$$
(A.32)

Moreover, \(J_{\mathrm{CS}}(g_s,t)\) is the non-perturbative Chern–Simons free energy [88, 114]. As explained in [88], this also coincides with the grand potential of the resolved conifold as defined in [55]. More precisely,

$$\begin{aligned} \begin{aligned} J_{\mathrm{CS}}(g_s,T+\mathrm{i}\pi )= J^{\mathrm{pert}}_{{ \mathrm CS}}\left( g_s,T+\mathrm{i}\pi \right) + J^{\mathrm{np}}_{{ \mathrm CS}}\left( g_s,T+\mathrm{i}\pi \right) \end{aligned} \end{aligned}$$
(A.33)

where

$$\begin{aligned} J^{\mathrm{pert}}_{{ \mathrm CS}}\left( g_s,T+\mathrm{i}\pi \right)= & {} -g_s^{-2}\frac{T^3}{12}-g_s^{-2}\frac{\pi ^2 T}{12}-{1\over 24}T+{1\over 2}A_c(4 \pi /g_s)\nonumber \\&+\sum _{n \ge 1}{1\over n}\left( 2 \sin {n g_s \over 2}\right) ^{-2}(-1)^n\mathrm{e}^{-n T}, \end{aligned}$$
(A.34)
$$\begin{aligned} J^{\mathrm{np}}_{{ \mathrm CS}}\left( g_s,T+\mathrm{i}\pi \right)= & {} -\sum _{n\ge 1} {1\over 4 \pi n^2}\csc \left( {2 \pi ^2 n \over g_s}\right) \nonumber \\&\times \left( { 2 \pi n \over g_s} T + {2 \pi ^2 n \over g_s}\cot \left( {2 \pi ^2 n \over g_s}\right) +1\right) \mathrm{e}^{-{2 \pi n \over g_s}T}. \end{aligned}$$
(A.35)

For instance, when \(g_s=\pi \) and \(T>0\), we have (see also [88], equation (5.11))

$$\begin{aligned} J_{\mathrm{CS}}(\pi ,T+\mathrm{i}\pi )= & {} -(\pi )^{-2}\frac{T^3}{12}-\frac{T}{12}-{1\over 24}T+{1\over 2}A_c(4)\nonumber \\&+{1\over 8 \pi ^2}\mathrm{Li}_3(\mathrm{e}^{-2T})+{T\over 4 \pi ^2}\mathrm{Li}_2(\mathrm{e}^{-2T})\nonumber \\&-\left( {T^2\over 4 \pi ^2}+{1\over 8}\right) \log \left( 1-\mathrm{e}^{-2T}\right) -{1\over 4}{{\mathrm{arctanh}}}(\mathrm{e}^{-T}). \end{aligned}$$
(A.36)

Notice that, even thought \(J^{\mathrm{np}}_{{ \mathrm CS}}\) and \(J^{\mathrm{pert}}_{{ \mathrm CS}}\) are ill defined at \(g_s=\pi \), their sum is perfectly well defined. This is yet another example of the HMO cancellation mechanism.

Nekrasov partition function and topological strings on local \({\mathbb {P}}^1\times {\mathbb {P}}^1\)

As explained in [77], when the topological string free energy has a gauge theory interpretation, it is sometimes better to replace the topological string expressions (A.13) and (A.19) with their Nekrasov form since the latter has better convergence properties when \(\hbar \) is complex, i.e.  \(|q| \ne 1\). We follow the convention of [24] and write the 5d Nekrasov partition function, or q-conformal blocks, as

$$\begin{aligned} \begin{aligned}&{\mathcal {Z}}(u,Z,q_1,q_2)=\sum _{\lambda ,\mu } \left( { Z\over q_1 q_2}\right) ^{|\lambda |+|\mu |}\\&\quad \frac{1}{N_{\lambda ,\lambda }(1,q_1,q_2)N_{\mu ,\mu }(1,q_1,q_2)N_{\lambda ,\mu }(u,q_1,q_2)N_{\mu ,\lambda }(u^{-1},q_1,q_2)} \end{aligned} \end{aligned}$$
(B.1)

where the sum runs over all pairs \((\mu , \lambda )\) of Young diagrams. Moreover, we use

$$\begin{aligned} \begin{aligned} N_{\lambda ,\mu }(u,q_1,q_2) =&\prod _{s\in \lambda } \left( 1-u q_2^{-a_\mu (s)-1}q_1^{\ell _\lambda (s)}\right) \cdot \prod _{s \in \mu } \left( 1-u q_2^{a_\lambda (s)}q_1^{-\ell _\mu (s)-1}\right) \end{aligned} \end{aligned}$$
(B.2)

where \(a_{\lambda }(s), l_{\lambda }(s)\) are the arm length and the leg length of the box \(s \in \lambda \). For instance, we have

$$\begin{aligned} \begin{aligned}&\log {\mathcal {Z}}(u,Z,q,q^{-1})= \frac{2 q u Z}{(q-1)^2 (u-1)^2} + \mathcal {O}\left( {Z^2}\right) ,\\&\mathrm{i}\lim _{q_2\rightarrow 1} \log \left( q_2\right) \log \left( {\mathcal {Z}}(u,Z,q,q_2)\right) = \frac{i q (q+1) u Z}{(q-1) (q-u) (q u-1)}+ \mathcal {O}\left( {Z^2}\right) . \end{aligned} \end{aligned}$$
(B.3)

As explained in [115,116,117], if we expand \( {\mathcal {Z}}(u,Z,q,q^{-1})\) at small u, we recover the partition function (without the one loop contribution) of topological string on local \({\mathbb {P}}^1\times {\mathbb {P}}^1\). Therefore, we identify

$$\begin{aligned} \exp \left[ \mathsf {J}^{\mathrm{WS}}_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}({\mu }, {\xi }, \hbar )\right] \end{aligned}$$
(B.4)

with

$$\begin{aligned} {{1 \over (Q_f q,q,q)_{\infty }^2}} {\mathcal {Z}}\left( Q_f ,{Q_b\over Q_f}, \mathrm{e}^{4\pi ^2 \mathrm{i}/\hbar },\mathrm{e}^{-4\pi ^2 \mathrm{i}/\hbar }\right) . \end{aligned}$$
(B.5)

where

$$\begin{aligned} Q_b=\mathrm{e}^{-{2 \pi \over \hbar }t (\mu , \xi , \hbar )}, \quad Q_f=\mathrm{e}^{\xi } Q_b \end{aligned}$$
(B.6)

and \(t (\mu , \xi , \hbar )\) is defined in (A.11).

Likewise for local \({\mathbb {P}}^1 \times {\mathbb {P}}^1\), we identify \( F^{\mathrm{NS}}_{\mathrm{inst}}(\mathbf{t}, \hbar ) \) given in (A.20) with the Nekrasov–Shatashvili limit of (B.1), see for instance [116]. More precisely, we have

$$\begin{aligned} {- \sum _{w\ge 1}{1\over w^2} \cot \left( {\hbar w\over 2}\right) Q_F^w}+ \mathrm{i}\lim _{q_2\rightarrow 1} \log \left( q_2\right) \log \left( {\mathcal {Z}}\left( Q_F,{Q_B \over Q_F},q_1,q_2\right) \right) , \end{aligned}$$
(B.7)

where

$$\begin{aligned} Q_B=\mathrm{e}^{-t (\mu , \xi , \hbar )}, \quad Q_F=m_{{\mathbb {P}}^1 \times {\mathbb {P}}^1} Q_B, \quad q_1=\mathrm{e}^{\mathrm{i}\hbar }. \end{aligned}$$
(B.8)

As pointed out in the previous section, both quantities in (B.3) are ill defined and have poles at \(|q|=1\). However, when we combine them in the grand potential (A.23), these poles cancel.

Some relevant shifts

We define

$$\begin{aligned} V_1(\mu ,\xi ,\hbar )=\mathrm{e}^{ J_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}^{\mathrm{WKB, inst}}(\mu , { \xi },\hbar )}. \end{aligned}$$
(C.1)

Then we have

$$\begin{aligned} \begin{aligned}&{V_1( \mu +\mathrm{i}\pi ,\xi ,\hbar ) V_1( \mu -\mathrm{i}\pi ,\xi ,\hbar )\over V_1( \mu ,\xi ,\hbar )^2}= 1, \\&{V_1( \mu -2 \pi \mathrm{i}, \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_1( \mu +2 \pi \mathrm{i},\xi + {4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_1( \mu ,m,\hbar )^2}= 1,\\&{V_1( \mu -\mathrm{i}\pi , \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_1( \mu +\mathrm{i}\pi ,\xi +{4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_1( \mu ,\xi ,\hbar )^2}= 1 . \end{aligned} \end{aligned}$$
(C.2)

Similarly, we define

$$\begin{aligned} V_2({\mu }, {\xi }, \hbar )= {(Q_f ^{-1}q,q,q)_{\infty } \over (Q_f q,q,q)_{\infty }}. \end{aligned}$$
(C.3)

Then we have

$$\begin{aligned}&{V_2( \mu +\mathrm{i}\pi ,\xi ,\hbar ) V_2( \mu -\mathrm{i}\pi ,\xi ,\hbar )\over V_2( \mu ,\xi ,\hbar )^2}= -{1 \over Q_f } , \nonumber \\&{V_2( \mu -2 \pi \mathrm{i}, \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_2( \mu +2 \pi \mathrm{i},\xi +{4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_2( \mu ,\xi ,\hbar )^2}= -{1 \over Q_f } , \end{aligned}$$
(C.4)
$$\begin{aligned}&{V_2( \mu -\mathrm{i}\pi , \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_2( \mu +\mathrm{i}\pi , \xi +{4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_2(\mu , \hbar ,\hbar )^2}=1, \end{aligned}$$
(C.5)

where we used

$$\begin{aligned} \left( u q;q,q\right) _{\infty } =\prod _{i,j\ge 0}\left( 1-u q q^{i+j}\right) =\exp \left[ - \sum _{s\ge 1}\frac{ u^s}{s \left( q^{s\over 2}-q^{-\frac{s}{2}}\right) ^2} \right] . \end{aligned}$$
(C.6)

Similarly, we define

$$\begin{aligned} V_3( \xi , \hbar )=\exp [A_p( \xi ,\hbar ) ] \end{aligned}$$
(C.7)

and we have

$$\begin{aligned} \begin{aligned}&{V_3( \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_3( \xi +{4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_3( \xi ,\hbar )^2}= \mathrm{e}^{-\xi /4}. \end{aligned} \end{aligned}$$
(C.8)

Also for

$$\begin{aligned} V_4( \mu , \xi , \hbar )=\exp [{\mathrm{P}}_{{\mathbb {P}}^1\times {\mathbb {P}}^1}(\mu , \xi ,\hbar ) ] \end{aligned}$$
(C.9)

we have

$$\begin{aligned} \begin{aligned}&{V_4( \mu +\mathrm{i}\pi ,\xi ,\hbar ) V_4( \mu -\mathrm{i}\pi ,\xi ,\hbar )\over V_4( \mu ,\xi ,\hbar )^2}=\mathrm{e}^{\xi /2} Q_b , \\&{V_4( \mu -2 \pi \mathrm{i}, \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_4( \mu +2 \pi \mathrm{i},\xi +{4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_4( \mu ,\xi ,\hbar )^2}= \mathrm{e}^{2 \xi } Q_b^2,\\&{V_4( \mu -\mathrm{i}\pi , \xi {-4 \mathrm{i}\pi ^2/\hbar },\hbar ) V_4( \mu +\mathrm{i}\pi ,\xi +{4 \mathrm{i}\pi ^2/\hbar },\hbar )\over V_4( \mu ,\xi ,\hbar )^2}=\mathrm{e}^{\xi /2}. \end{aligned} \end{aligned}$$
(C.10)

Moreover, we have

$$\begin{aligned}&J^{\mathrm{np}}_{{ \mathrm CS}} \left( {2\pi ^2 \over \hbar },{1\over 2}\xi + {2 \pi ^2 \mathrm{i}\over \hbar }+\mathrm{i}\pi \right) +J^{\mathrm{np}}_{{ \mathrm CS}} \left( {2\pi ^2 \over \hbar },{1\over 2}\xi - {2 \pi ^2 \mathrm{i}\over \hbar }+\mathrm{i}\pi \right) \nonumber \\&\quad = 2J^{\mathrm{np}}_{{ \mathrm CS}} \left( {2\pi ^2 \over \hbar },{1\over 2}\xi +\mathrm{i}\pi \right) . \end{aligned}$$
(C.11)

Similarly,

$$\begin{aligned} \begin{aligned}&J^{\mathrm{pert}}_{{ \mathrm CS}} \left( {2\pi ^2 \over \hbar },{1\over 2}\xi + {2 \pi ^2 \mathrm{i}\over \hbar }+\mathrm{i}\pi \right) +J^{\mathrm{pert}}_{{ \mathrm CS}} \left( {2\pi ^2 \over \hbar },{1\over 2}\xi - {2 \pi ^2 \mathrm{i}\over \hbar }+\mathrm{i}\pi \right) \\&\qquad - 2J^{\mathrm{pert}}_{{ \mathrm CS}} \left( {2\pi ^2 \over \hbar },{1\over 2}\xi +\mathrm{i}\pi \right) ={1\over 4}\xi +\log \left( \mathrm{e}^{-\xi /2}+1\right) . \end{aligned} \end{aligned}$$
(C.12)

Some identities for \(\eta \) function

We denote

$$\begin{aligned} \eta (\tau )=\mathrm{e}^{\mathrm{i}\pi \tau /12}\prod _{n=1}^{\infty } \left( 1-\mathrm{e}^{2 \pi \mathrm{i}n \tau }\right) \end{aligned}$$
(D.1)

the Dedekind \(\eta \) function. The Weber modular functions are defined as

$$\begin{aligned} \begin{aligned}&f(\tau )={\eta ^2(\tau )\over \eta (\tau /2)\eta (2\tau )},\\&f_1(\tau )={\eta (\tau /2)\over \eta (\tau )},\\&f_2(\tau )=\sqrt{2}{\eta (2\tau )\over \eta (\tau )}.\\ \end{aligned} \end{aligned}$$
(D.2)

Standard identities of Weber modular functions are

$$\begin{aligned}&f_1(\tau )^8+f_2(\tau )^8=f(\tau )^8, \end{aligned}$$
(D.3)
$$\begin{aligned}&8 j(\tau )=(f_1(\tau )^{16}+f_2(\tau )^{16}+f^{16}(\tau ))^3, \end{aligned}$$
(D.4)

where j is the j-invariant:

$$\begin{aligned} j(\tau )={1\over {{\bar{q}}}}+ 744+196884{{\bar{q}}}+ \mathcal {O}({{{\bar{q}}}^2} ), \quad {{\bar{q}}}=\mathrm{e}^{2 \pi \mathrm{i}\tau } \end{aligned}$$
(D.5)

Summary of conventions

The paper connects different branches of mathematical physics and for this reason has a lot of notations which we summarize here. The complex moduli of local \({\mathbb {P}}_1 \times {\mathbb {P}}_1\) are denoted by

$$\begin{aligned} \kappa , \quad m_{{\mathbb {P}}_1\times {\mathbb {P}}_1}. \end{aligned}$$
(E.1)

The parameter \(\hbar \) is such that

$$\begin{aligned}{}[{\mathsf {x}}, {\mathsf {p}}]=\mathrm{i}\hbar , \end{aligned}$$
(E.2)

where \({\mathsf {x}}, {\mathsf {p}} \) are the operators appearing in the quantum mirror curve (3.1). It is sometimes useful to use \((\mu , \xi )\) which are related to (E.1) as

$$\begin{aligned} \kappa =\mathrm{e}^{\mu }, \quad \xi = {2 \pi \over \hbar }\log {m_{{\mathbb {P}}^1\times {\mathbb {P}}^1}}. \end{aligned}$$
(E.3)

The quantum mirror map of local \({\mathbb {P}}_1 \times {\mathbb {P}}_1\) is denoted by

$$\begin{aligned} t (\mu , \xi , \hbar ) . \end{aligned}$$
(E.4)

We also use

$$\begin{aligned} t_1 ( \hbar )=t(\mu , \xi , \hbar ), \quad t_2 ( \hbar )= t(\mu , \xi , \hbar )-{\hbar \over 2 \pi } \xi . \end{aligned}$$
(E.5)

To connect with the topological string/spectral theory correspondence, it is useful to introduce

$$\begin{aligned} Q_b=\mathrm{e}^{-{2 \pi \over \hbar }t (\mu , \xi , \hbar )}, \quad Q_f=\mathrm{e}^{\xi } Q_b , \quad Q_B=\mathrm{e}^{-t (\mu , \xi , \hbar )}, \quad Q_F=m_{{\mathbb {P}}_1\times {\mathbb {P}}_1} Q_B. \end{aligned}$$
(E.6)

Notice that

$$\begin{aligned} Q_b=Q_B^{2\pi /\hbar }, \quad Q_f=Q_F^{2\pi /\hbar }. \end{aligned}$$
(E.7)

In the q-Painlevé literature, one typically uses the variables

$$\begin{aligned} Z, u, q . \end{aligned}$$
(E.8)

These are related to the variables appearing in the topological string/spectral theory correspondence via

$$\begin{aligned} Z^{-1}=\mathrm{e}^{\xi }, \quad q=\mathrm{e}^{\mathrm{i}4 \pi ^2/\hbar }, \quad \quad u=\mathrm{e}^{\xi }Q_b= Q_f. \end{aligned}$$
(E.9)

In the ABJ language, the natural variables are the rank of the gauge group M and the Chern–Simons level k. These are related to the topological string variables via

$$\begin{aligned} \log m_{{\mathbb {P}}^1 \times {\mathbb {P}}^1}=\mathrm{i}\hbar -2 \pi \mathrm{i}M, \quad \hbar = \pi k. \end{aligned}$$
(E.10)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bonelli, G., Grassi, A. & Tanzini, A. Quantum curves and q-deformed Painlevé equations. Lett Math Phys 109, 1961–2001 (2019). https://doi.org/10.1007/s11005-019-01174-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11005-019-01174-y

Keywords

Mathematics Subject Classification

Navigation