Skip to main content
Log in

Complete Solution of the LSZ Model via Topological Recursion

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

Abstract

We prove that the Langmann–Szabo–Zarembo (LSZ) model with quartic potential, a toy model for a quantum field theory on noncommutative spaces grasped as a complex matrix model, obeys topological recursion of Chekhov, Eynard and Orantin. By introducing two families of correlation functions, one corresponding to the meromorphic differentials \(\omega _{g,n}\) of topological recursion, we obtain Dyson–Schwinger equations that eventually lead to the abstract loop equations being, together with their pole structure, the necessary condition for topological recursion. This strategy to show the exact solvability of the LSZ model establishes another approach towards the exceptional property of integrability in some quantum field theories. We compare differences in the loop equations for the LSZ model (with complex fields) and the Grosse–Wulkenhaar model (with hermitian fields) and their consequences for the resulting particular type of topological recursion that governs the models.

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

Similar content being viewed by others

Data Availability

The article has no associated data.

Notes

  1. The distinction between E and \(\tilde{E}\) is a technical assumption to uniquely define the boundary creation operator (see Definition 3.3). The limit \(\tilde{E}=E\) is well-defined in all formulas.

  2. This is true even for non-commuting matrices E and \(\tilde{E}\) since the measure is invariant under transformations \(\Phi \rightarrow U\Phi V\) with two independent unitary matrices U and V which can be chosen in a way to diagonalise E and \(\tilde{E}\) simultaneously; see proof of Proposition 2.7 for details.

  3. Introducing another unitary transformation \(\Phi ^{(U)} = \Phi U\) and thus \((\Phi ^\dagger )^{(U)} = U^\dagger \Phi ^\dagger \) guarantees the claimed diagonalisability of \(\tilde{E}\) in the same manner.

  4. “Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 427320536 – SFB 1442, as well as under Germany’s Excellence Strategy EXC 2044 390685587, Mathematics Münster: Dynamics – Geometry – Structure.”

  5. “Funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Project-ID 465029630.

  6. Their proof builds on properties of the inverse Cauchy matrix, however it is sufficient to work with the fundamental polynomials of the Lagrange interpolation formula.

References

  1. Aizenman, M., Duminil-Copin, H.: Marginal triviality of the scaling limits of critical 4D Ising and \(\phi _4^4\) models. Ann. Math. 194, 163–235 (2021). https://doi.org/10.4007/annals.2021.194.1.3. arXiv:1912.07973

    Article  MathSciNet  MATH  Google Scholar 

  2. Aizenman, M.: Proof of the triviality of \(\phi ^4_d\) field theory and some mean field features of Ising models for \(d > 4\). Phys. Rev. Lett. 47, 1–4 (1981). https://doi.org/10.1103/PhysRevLett.47.1

  3. Belliard, R., Charbonnier, S., Eynard, B., Garcia-Failde, E.: Topological recursion for generalised Kontsevich graphs and r-spin intersection numbers (2021). arXiv:2105.08035

  4. Borot, G., Charbonnier, S., Garcia-Failde, E.: Topological recursion for fully simple maps from ciliated maps (2021). arXiv:2106.09002

  5. Borot, G., Charbonnier, S., Garcia-Failde, E., Leid, F., Shadrin, S.: Analytic theory of higher order free cumulants (2021). arXiv:2112.12184

  6. Bouchard, V., Eynard, B.: Think globally, compute locally. JHEP 02, 143 (2013). https://doi.org/10.1007/JHEP02(2013)143. arXiv:1211.2302

    Article  ADS  MathSciNet  MATH  Google Scholar 

  7. Bouchard, V., Eynard, B.: Reconstructing WKB from topological recursion (2016). https://doi.org/10.5802/jep.58. arXiv:1606.04498

    Article  Google Scholar 

  8. Borot, G., Eynard, B., Mulase, M., Safnuk, B.: A Matrix model for simple Hurwitz numbers, and topological recursion. J. Geom. Phys. 61, 522–540 (2011). https://doi.org/10.1016/j.geomphys.2010.10.017. arXiv:0906.1206

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Borot, G., Eynard, B., Orantin, N.: Abstract loop equations, topological recursion and new applications. Commun. Number Theory Phys. 09, 51–187 (2015). https://doi.org/10.4310/CNTP.2015.v9.n1.a2. arXiv:1303.5808

    Article  MathSciNet  MATH  Google Scholar 

  10. Bernardi, O., Fusy, E.: Bijections for planar maps with boundaries. J. Comb. Theory Ser. A 158, 176–227 (2018). https://doi.org/10.1016/j.jcta.2018.03.001. arXiv:1510.05194

    Article  MathSciNet  MATH  Google Scholar 

  11. Borot, G., Garcia-Failde, E.: Simple maps, Hurwitz numbers, and topological recursion. Commun. Math. Phys. 380(2), 581–654 (2020). https://doi.org/10.1007/s00220-020-03867-1. arXiv:1710.07851

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Brezin, E., Itzykson, C., Parisi, G., Zuber, J.B.: Planar diagrams. Commun. Math. Phys. 59, 35–51 (1978). https://doi.org/10.1007/BF01614153

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Branahl, J., Grosse, H., Hock, A., Wulkenhaar, R.: From scalar fields on quantum spaces to blobbed topological recursion. J. Phys. A 55(42), 423001 (2022). https://doi.org/10.1088/1751-8121/ac9260. arXiv:2110.11789

    Article  MathSciNet  MATH  Google Scholar 

  14. Branahl, J., Hock, A.: A spectral curve for the generation of bipartite maps in topological recursion (2022). arXiv:2204.05181

  15. Branahl, J., Hock, A., Wulkenhaar, R.: Perturbative and geometric analysis of the quartic Kontsevich Model. SIGMA 17, 085 (2021). https://doi.org/10.3842/SIGMA.2021.085. arXiv:2012.02622

    Article  MathSciNet  MATH  Google Scholar 

  16. Branahl, J., Hock, A., Wulkenhaar, R.: Blobbed topological recursion of the quartic Kontsevich model I: Loop equations and conjectures. Commun. Math. Phys. 393(3), 1529–1582 (2022). https://doi.org/10.1007/s00220-022-04392-z. arXiv:2008.12201

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Bouchard, V., Mariño, M.: Hurwitz numbers, matrix models and enumerative geometry. In: From Hodge Theory to Integrability and TQFT: tt*-Geometry, vol. 78 of PProceedings of Symposia in Pure Mathematics, pp. 263–283. American Mathematical Society, Providence (2008). https://doi.org/10.1090/pspum/078/2483754. arXiv:0709.1458

  18. Borot, G.: Formal multidimensional integrals, stuffed maps, and topological recursion. Ann. Henri Poincaré 1, 225–264 (2014). https://doi.org/10.4171/AIHPD/7. arXiv:1307.4957

    Article  MathSciNet  MATH  Google Scholar 

  19. Borot, G., Shadrin, S.: Blobbed topological recursion: properties and applications. Math. Proc. Camb. Philos. Soc. 162(1), 39–87 (2017). https://doi.org/10.1017/S0305004116000323. arXiv:1502.00981

    Article  MathSciNet  MATH  Google Scholar 

  20. Chekhov, L., Eynard, B., Orantin, N.: Free energy topological expansion for the 2-matrix model. JHEP 12, 053 (2006). https://doi.org/10.1088/1126-6708/2006/12/053. arXiv:math-ph/0603003

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Connes, A.: Noncommutative Geometry. Academic Press, Inc., London (1994)

    MATH  Google Scholar 

  22. Doplicher, S., Fredenhagen, K., Roberts, J.E.: The quantum structure of space-time at the Planck scale and quantum fields. Commun. Math. Phys. 172, 187–220 (1995). https://doi.org/10.1007/BF02104515. arXiv:hep-th/0303037

    Article  ADS  MATH  Google Scholar 

  23. Disertori, M., Gurau, R., Magnen, J., Rivasseau, V.: Vanishing of beta function of non-commutative \(\phi ^4_4\) theory to all orders. Phys. Lett. B B649, 95–102 (2007). https://doi.org/10.1016/j.physletb.2007.04.007. arXiv:hep-th/0612251

    Article  ADS  MATH  Google Scholar 

  24. de Jong, J., Hock, A., Wulkenhaar, R.: Nested Catalan tables and a recurrence relation in noncommutative quantum field theory. Ann. Henri Poincaré D 9, 47–72 (2022). https://doi.org/10.4171/AIHPD/113. arXiv:1904.11231

    Article  MathSciNet  MATH  Google Scholar 

  25. Eynard, B., Mulase, M., Safnuk, B.: The Laplace transform of the cut-and-join equation and the Bouchard–Marino conjecture on Hurwitz numbers. Publ. Res. Inst. Math. Sci. 47, 629–670 (2011). https://doi.org/10.2977/PRIMS/47. arXiv:0907.5224

    Article  MathSciNet  MATH  Google Scholar 

  26. Eynard, B., Orantin, N.: Mixed correlation functions in the 2-matrix model, and the Bethe ansatz. JHEP 8, 28 (2005). https://doi.org/10.1088/1126-6708/2005/08/028. arXiv:hep-th/0504029

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Eynard, B., Orantin, N.: Invariants of algebraic curves and topological expansion. Commun. Number Theory Phys. 1, 347–452 (2007). https://doi.org/10.4310/CNTP.2007.v1.n2.a4. arXiv:math-ph/0702045

    Article  MathSciNet  MATH  Google Scholar 

  28. Eynard, B., Orantin, N.: Topological expansion and boundary conditions. JHEP 6, 37 (2008). https://doi.org/10.1088/1126-6708/2008/06/037. arXiv:0710.0223

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Eynard, B.: Invariants of spectral curves and intersection theory of moduli spaces of complex curves. Commun. Number Theory Phys. 8, 541–588 (2011). https://doi.org/10.4310/CNTP.2014.v8.n3.a4. arXiv:1110.2949

    Article  MathSciNet  MATH  Google Scholar 

  30. Eynard, B.: Counting Surfaces. Progress in Mathematical Physics, vol. 70. Birkhäuser, Basel (2016). https://doi.org/10.1007/978-3-7643-8797-6

    Book  MATH  Google Scholar 

  31. Eynard, B.: The Geometry of integrable systems. Tau functions and homology of spectral curves. Perturbative definition (2017). arXiv:1706.04938

  32. Fröhlich, J.: On the triviality of \(\lambda \phi ^4_d\) theories and the approach to the critical point in \(d \ge 4\) dimensions. Nucl. Phys. B 200, 281–296 (1982). https://doi.org/10.1016/0550-3213(82)90088-8

  33. Grosse, H., Hock, A., Wulkenhaar, R.: Solution of all quartic matrix models (2019). arXiv:1906.04600

  34. Grosse, H., Hock, A., Wulkenhaar, R.: Solution of the self-dual \(\Phi ^4\) QFT-model on four-dimensional Moyal space. JHEP 01, 081 (2020). https://doi.org/10.1007/JHEP01(2020)081. arXiv:1908.04543

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Grosse, H., Wulkenhaar, R.: Renormalisation of \(\phi ^4\)-theory on noncommutative \({\mathbb{R} }^4\) in the matrix base. Commun. Math. Phys. 256, 305–374 (2005). https://doi.org/10.1007/s00220-004-1285-2. arXiv:hep-th/0401128

    Article  ADS  MATH  Google Scholar 

  36. Grosse, H., Wulkenhaar, R.: Self-dual noncommutative \(\phi ^4\)-theory in four dimensions is a non-perturbatively solvable and non-trivial quantum field theory. Commun. Math. Phys. 329, 1069–1130 (2014). https://doi.org/10.1007/s00220-014-1906-3. arXiv:1205.0465

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Harish-Chandra: Differential operators on a semisimple lie algebra. Am. J. Math. 79, 87–120 (1957). https://doi.org/10.2307/2372387

    Article  MathSciNet  MATH  Google Scholar 

  38. Hirota, R.: Exact solution of the Korteweg–de Vries equation for multiple collisions of solitons. Phys. Rev. Lett. 27, 1192–1194 (1971). https://doi.org/10.1103/PhysRevLett.27.1192

    Article  ADS  MATH  Google Scholar 

  39. Hock, A.: Matrix field theory. Ph.D. thesis, WWU Münster (2020). arXiv:2005.07525

  40. Hock, A.: A simple formula for the \(x\)-\(y\) symplectic transformation in topological recursion (2022). arXiv:2211.08917

  41. Hock, A.: On the \(x\)-\(y\) symmetry of correlators in topological recursion via loop insertion operator (2022). arXiv:2201.05357

  42. Hock, A., Wulkenhaar, R.: Noncommutative 3-colour scalar quantum field theory model in 2D. Eur. Phys. J. C C78(7), 580 (2018). https://doi.org/10.1140/epjc/s10052-018-6042-3. arXiv:1804.06075

    Article  ADS  Google Scholar 

  43. Hock, A., Wulkenhaar, R.: Blobbed topological recursion of the quartic Kontsevich model II: Genus=0 (2021). arXiv:2103.13271

  44. Hock, A., Wulkenhaar, R.: Blobbed topological recursion from extended loop equations (2023). arXiv:2301.04068

  45. Itzykson, C., Zuber, J.B.: The planar approximation. II. J. Math. Phys. 21, 411–421 (1980). https://doi.org/10.1063/1.524438

    Article  ADS  MathSciNet  MATH  Google Scholar 

  46. Kontsevich, M.: Intersection theory on the moduli space of curves and the matrix Airy function. Commun. Math. Phys. 147, 1–23 (1992). https://doi.org/10.1007/BF02099526

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. Landau, L.D., Abrikosov, A.A., Khalatnikov, I.M.: On the removal of infinities in quantum electrodynamics. Dokl. Akad. Nauk SSSR 95, 497–500 (1954). (in Russian)

    MathSciNet  MATH  Google Scholar 

  48. Langmann, E., Szabo, R.J.: Duality in scalar field theory on noncommutative phase spaces. Phys. Lett. B B533, 168–177 (2002). https://doi.org/10.1016/S0370-2693(02)01650-7. arXiv:hep-th/0202039

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. Langmann, E., Szabo, R.J., Zarembo, K.: Exact solution of quantum field theory on noncommutative phase spaces. JHEP 01, 17 (2004). https://doi.org/10.1088/1126-6708/2004/01/017. hep-th/0308043

    Article  ADS  MathSciNet  MATH  Google Scholar 

  50. Morris, T.R.: Checkered surfaces and complex matrices. Nucl. Phys. B 356, 703–728 (1991). https://doi.org/10.1016/0550-3213(91)90383-9

    Article  ADS  Google Scholar 

  51. Panzer, E., Wulkenhaar, R.: Lambert-W solves the noncommutative \(\Phi ^4\)-model. Commun. Math. Phys. 374, 1935–1961 (2020). https://doi.org/10.1007/s00220-019-03592-4. arXiv:1807.02945

    Article  ADS  MathSciNet  MATH  Google Scholar 

  52. Rieffel, M.A.: Deformation quantization of Heisenberg manifolds. Commun. Math. Phys. 122(4), 531–562 (1989). https://doi.org/10.1007/BF01256492

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. Sato, M.: Soliton equations as dynamical systems on infinite-dimensional Grassmann manifold. Nonlinear partial differential equations in applied science. Stud. Math. 81, 259–271 (1983). https://doi.org/10.1016/S0304-0208(08)72096-6

    Article  Google Scholar 

  54. Schwinger, J.: Euclidean quantum electrodynamics. Phys. Rev. 115, 721–731 (1959). https://doi.org/10.1103/PhysRev.115.721

    Article  ADS  MathSciNet  MATH  Google Scholar 

  55. Santilli, L., Tierz, M.: Complex (super)-matrix models with external sources and q-ensembles of Chern–Simons and ABJ(M) type. J. Phys. A Math. Theor. 53, 425201 (2020). https://doi.org/10.1088/1751-8121/abb6b0. arXiv:1805.10543

    Article  ADS  MathSciNet  MATH  Google Scholar 

  56. Schürmann, J., Wulkenhaar, R.: Towards integrability of a quartic analogue of the Kontsevich model (2019). arXiv:1912.03979

  57. ’t Hooft, G.: A Planar Diagram Theory for Strong Interactions. Nucl. Phys. B 72, 461 (1974). https://doi.org/10.1016/0550-3213(74)90154-0

    Article  ADS  Google Scholar 

  58. Tutte, W.: A census of planar maps. Can. J. Math. 15, 249–271 (1963). https://doi.org/10.4153/CJM-1963-029-x

    Article  MathSciNet  MATH  Google Scholar 

  59. Witten, E.: Two-dimensional gravity and intersection theory on moduli space. Surv. Differ. Geom. 1, 243–310 (1991). https://doi.org/10.4310/SDG.1990.v1.n1.a5

    Article  MathSciNet  MATH  Google Scholar 

  60. Wulkenhaar, R.: Quantum field theory on noncommutative spaces. In: Chamseddine, A., Consani, C., Higson, N., Khalkhali, M., Moscovici, H., Yu, G. (eds.) Advances in Noncommutative Geometry, pp. 607–690. Springer, Berlin (2019). https://doi.org/10.1007/978-3-030-29597-411

    Chapter  MATH  Google Scholar 

  61. Zinn-Justin, P., Zuber, J.B.: Matrix integrals and the generation and counting of virtual tangles and links. J. Knot Theory Ramif. 13, 325–356 (2004). https://doi.org/10.1142/S0218216504003172. arXiv:math-ph/0303049

Download references

Acknowledgements

JB is supported by the Cluster of Excellence Mathematics Münster. He would like to thank the University of Oxford for its hospitality. The work of JB at the University of Oxford was additionally financed by the RTG 2149 Strong and Weak Interactions—from Hadrons to Dark Matter. AH is supported by the Walter–Benjamin fellowship. The authors are grateful to Harald Grosse and Raimar Wulkenhaar, since many of the used techniques are results of our collaboration over the last years.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander Hock.

Ethics declarations

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Additional information

Communicated by A. Giulian.

Publisher's Note

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

Appendices

Appendix A. Solution of the 2-Point Function

Throughout the appendix, we will need the Lagrange interpolation formula that we want recall:

Lemma A.1

Let f be a polynomial of degree \(d\ge 0\) and \(x_1,\ldots ,x_{d+1}\) be pairwise distinct complex numbers. Then, for all \(x\in \mathbb {C}\),

$$\begin{aligned} f(x)=L(x)\sum _{j=1}^{d+1}\frac{f_j}{(x-x_j)L'(x_j)},\qquad \text {where } L(x)=\prod _{j=1}^d(x-x_j) \text { and } f_j=f(x_j). \end{aligned}$$

This section is dedicated to the proof of the spectral curve, encoded in the solution of the analytically continued 2-point function.

1.1 Proof of Theorem 4.3

Most methods build on the solution strategy in the QKM, [SW19].Footnote 6 We start with the complexified Dyson–Schwinger equation for the 2-point function. We assume that there are undetermined variable transforms x(z) and y(z) that can be uniquely specified by the upper ansatz. After transformation, we have for Eq. (3.1), \(g=0\):

$$\begin{aligned}&\bigg (x(z)+y(w)+ \frac{\lambda }{N} \sum _{k=1}^{{\tilde{d}}} \tilde{r}_k \mathcal {G}^{(0)}(z,\tilde{\varepsilon }_k) \bigg )\mathcal {G}^{(0)}(z,w) \end{aligned}$$
(6.9)
$$\begin{aligned}&= 1+ \frac{\lambda }{N} \sum _{k=1}^d r_k\frac{\mathcal {G}^{(0)}(\varepsilon _k,w)-\mathcal {G}^{(0)}(z,w)}{x(\varepsilon _k)-x(z)}. \end{aligned}$$
(A.1)

The ansatz of Theorem 4.3 turns this equation into

$$\begin{aligned} (y(w) -y(z))\mathcal {G}^{(0)}(z,w) = 1+ \frac{\lambda }{N} \sum _{k=1}^d r_k \frac{\mathcal {G}^{(0)}(\varepsilon _k,w)}{x(\varepsilon _k)-x(z)}. \end{aligned}$$
(A.2)

We set \(z= \hat{\tilde{w}}^l\) to get rid off the prefactor on the lhs under the finiteness assumption from Theorem 4.3 of \(\mathcal {G}^{(0)}(z= \hat{\tilde{w}}^l,w)\). This leads to d equations since y was assumed to be of rational degree d

$$\begin{aligned} \frac{\lambda }{N} \sum _{k=1}^d r_k \frac{\mathcal {G}^{(0)}(\varepsilon _k,w)}{x(\hat{\tilde{w}}^l)-x(\varepsilon _k)}=1. \end{aligned}$$

Define the fundamental Lagrange interpolation polynomials \(A(v) = \prod _{j=1}^d (v-x(\hat{\tilde{w}}^j))\) and \(B(v) =\prod _{j=1}^d (v-x(\varepsilon _j))\). Then we can read off

$$\begin{aligned} \frac{\lambda r_k}{N} \mathcal {G}^{(0)}(\varepsilon _k,w) =- \frac{\prod _{j=1}^d (x(\varepsilon _k)-x(\hat{\tilde{w}}^j)}{\prod _{j \ne k}^d (x(\varepsilon _k)-x(\varepsilon _j))} \end{aligned}$$
(A.3)

since the following sum over all residues at \(x(\varepsilon _j)\) for the poles in B(v) reads 1:

$$\begin{aligned} \sum _{j=1}^d {{\,\textrm{Res}\,}}_{v \rightarrow x(\varepsilon _k)} \frac{A(v)}{(x(\hat{\tilde{w}}^l)-v)B(v)} = \sum _{j=1}^d \frac{A(x(\varepsilon _k))}{(x(\hat{\tilde{w}}^l)-x(\varepsilon _j))B'(x(\varepsilon _k))}=1. \end{aligned}$$

With the same trick, we allow for \(d+1\) factors in the interpolation formula and include x(z). This gives the possibility to write

$$\begin{aligned} \frac{\prod _j (x(z)-x(\hat{\tilde{w}}^j))}{\prod _j (x(z)-x(\varepsilon _j))} + \sum _{k=1}^d \frac{\prod _j(x(\varepsilon _k)-x(\hat{\tilde{w}}^j))}{(x(\varepsilon _k)-x(z))\prod _{j \ne k}(x(\varepsilon _k)-x(\varepsilon _j))}=1. \end{aligned}$$

Inserting Eq. (A.4) into Eq. (A.3) gives

$$\begin{aligned} \mathcal {G}^{(0)}(z,w) = \frac{1}{y(w)-y(z)} \prod _{j=1}^{d} \frac{x(z)-x\big (\widehat{\tilde{w}}^j\big )}{x(z)-x(\varepsilon _j)} \end{aligned}$$
(A.4)

and thus, we principally finished the proof of the first representation. However, it is still unclear why y(z) is necessarily given by \( y(z)=-z+\frac{\lambda }{N}\sum _{n=1}^{d}\frac{r_n}{x'(\varepsilon _n)(z-\varepsilon _n)}+C\) where the constant C must vanish in the end, and the same for x(z). This can be shown using Liouville’s theorem. Inserting \(z=\varepsilon _k\) into Eq. (A.5) and comparing with Eq. (A.4), y(z) must have d poles at \(z= \varepsilon _k\). Moreover, we have a simple pole at \(z=\infty \). Remembering the degree \(d+1\) of the assumption, we already have the complete set of poles. Thus, \(-z+\frac{\lambda }{N}\sum _{n=1}^{d}\frac{r_n}{x'(\varepsilon _n)(z-\varepsilon _n)}+C-y(z)\) is a bounded holomorphic function and by Liouville a constant. Analogously to [SW19], one can show with some technical effort, that the initial ansatz guarantees that \(C=0\).

Now we exchange the role of x and y, where we have to take the DSE of Remark 3.2. Carrying out the analytic continuation of Definition 4.4, we get a second complexified DSE for the 2-point function

$$\begin{aligned} (x(z) -x(w))\mathcal {G}^{(0)}(z,w) = 1+ \frac{\lambda }{N} \sum _{k=1}^{\tilde{d}} \tilde{r}_k \frac{\mathcal {G}^{(0)}(z,\tilde{\varepsilon }_k)}{y(\tilde{\varepsilon }_k)-y(w)}. \end{aligned}$$
(A.5)

Applying exactly the same steps yields the second representation, where x(z) is necessarily given by \(z-\frac{\lambda }{N}\sum _{k=1}^{\tilde{d}}\frac{\tilde{r}_k}{y'(\tilde{\varepsilon }_k)(z-\tilde{\varepsilon }_k)}\). \(\square \)

We remark that the form of the spectral cuarve is not surprising—it is the typical interwoven structure between x and y—as implicitly defined equations—that usually appears when solving a matrix model with external field (for instance the matrix model for simple Hurwitz numbers [BEMS11] or the Kontsevich model [EO07, Sec. 10.4]).

Appendix B. Proof of \(\Omega ^{(0)}_2\)

It remains the proof for the second part of the initial data, namely the Bergman kernel. By rational parametrisation of \(\omega _{0,1}\), the spectral curve is a planar algebraic curve and thus, the shape of the fundamental form of the second kind is automatically determined, which we will show to coincide with \(\Omega _{0,2}\).

Proposition B.1

The cylinder amplitude of the LSZ model reads

$$\begin{aligned} \Omega _{0,2}(z_1,z_2)=\frac{1}{x'(z_1)x'(z_2)(z_1-z_2)^2}. \end{aligned}$$
(A.6)

The proof is split into two lemmata that follow from techniques of [BHW22]. Taking Corollary 4.5 for \((g,n)=(0,2)\), setting \(v=y(z)\) (the first term vanishes since H is rational in v) and inserting the DSE of the 2-point function with exchanged x and y (Eq. A.6), we deduce that \(\Omega ^{(0)}_{2}\) satisfies the DSE

$$\begin{aligned}&x'(z)\mathfrak {G}_0(z)\Omega ^{(0)}_{2}(u,z) -\frac{\lambda }{N^2}\sum _{k,n=1}^{d,\tilde{d}} \frac{ r_k \tilde{r}_n \mathcal {T}^{(0)}(u\Vert \varepsilon _k,\tilde{\varepsilon }_n|)}{ (x(\varepsilon _k)-x(z))(y(\tilde{\varepsilon }_n)-y(z))}\\ =&-\frac{\partial }{\partial x(u)} \mathcal {G}^{(0)}(u,z) \;.\nonumber \end{aligned}$$
(B.1)

where \(\mathfrak {G}_0(z)=-H_{0,1}(y(z);z)\) and thus \(\mathfrak {G}_0(z)={{\,\textrm{Res}\,}}_{u\rightarrow z} \mathcal {G}^{(0)}(z,u)du\) by Theorem 4.3. To specify the pole structure of the 2-point function, we show the following decomposition:

Lemma B.2

The planar 2-point function can be rewritten in the following decomposition of its poles:

$$\begin{aligned}&\mathcal {G}^{(0)}(z,u) = \frac{1}{u-z} \bigg (1+ \frac{\lambda ^2}{N^2} \sum _{k,n=1}^d \sum _{l,m=1}^{\tilde{d}} \frac{C_{k,l}^{m,n}}{\Big (z-\widehat{\varepsilon }_k^{\,m}\Big )\Big (u-\widehat{\tilde{\varepsilon _l}}^n\Big )} \biggl ) \quad , \end{aligned}$$
(B.2)

where

$$\begin{aligned}&C_{k,l}^{m,n} = \frac{\Big (\widehat{\tilde{\varepsilon _l}}^n - \widehat{\varepsilon }_k^{\,m}\Big ) r_k \tilde{r}_l \mathcal {G}^{(0)}(\varepsilon _k,\tilde{\varepsilon }_l)}{x'\big (\widehat{\varepsilon }_k^{\,m} \big ) y' \big ( \widehat{\tilde{\varepsilon _l}}^n\Big ) \Big ( y\big (\widehat{\varepsilon }_k^{\,m} \big ) - y (\tilde{\varepsilon }_l) \Big ) \Big ( x\big (\widehat{\tilde{\varepsilon _l}}^n \big ) - x( \varepsilon _k) \Big ) }. \end{aligned}$$

Proof

To figure out all possible poles, we have to look at the two DSE’s (A.3) and (A.6). After dividing by \(y(w)-y(z)\) (or \(x(z)-x(w)\)) and the regularity assumption of Theorem 4.3, possible poles are located at \(z=u\) as well as at \(z=\hat{\varepsilon }_k^m\) and \(u=\widehat{\tilde{\varepsilon _l}}^n\). Then, we look at the function \((u-z) \mathcal {G}^{(0)}(u,z)\), which approaches 1 for \(z,u \rightarrow \infty \). Again, the basic relations Eqs. (A.3) and (A.6) give the chance to directly read off the residues at \(z=\widehat{\varepsilon _k}^m\) and \(u=\widehat{\tilde{\varepsilon _l}}^n\) as stated. For the regularity discussion at \(u={\hat{z}}^k\) and \(u=\widehat{\tilde{z}^k}\), one can look at the finite limit of the product representation of \(\mathcal {G}(z,u)\) given in Theorem 4.3. \(\square \)

Liouville’s theorem now proves:

Lemma B.3

Assume that (for generic u) the function \(\Omega ^{(0)}_2(u,z)\) is regular at any zero z of \(\mathfrak {G}_0(z)\). Then

$$\begin{aligned} x'(z)\Omega _{0,2}(z,u)-\frac{1}{x'(u)(z-u)^2}=0. \end{aligned}$$
(B.3)

Proof

We deduce the representation for \(\mathfrak {G}_0(z)={{\,\textrm{Res}\,}}_{u\rightarrow z} \mathcal {G}^{(0)}(z,u)du\):

$$\begin{aligned}&\mathfrak {G}_0(z)= 1- \frac{\lambda ^2}{N^2} \sum _{k,n}^d \sum _{l,m}^{\tilde{d}} \frac{C_{k,l}^{m,n}}{\Big (z-\widehat{\varepsilon }_k^{\,m}\Big )\Big (z-\widehat{\tilde{\varepsilon _l}}^n\Big )} \end{aligned}$$

and therefore the partical fraction decomposition

$$\begin{aligned}&\mathcal {G}^{(0)}(u,z) = \frac{\mathfrak {G}_0(z)}{u-z} + \frac{\lambda ^2}{N^2} \sum _{k,n}^d \sum _{l,m}^{\tilde{d}} \frac{C_{k,l}^{m,n}}{\Big (z-\widehat{\varepsilon }_k^{\,m}\Big )\Big ( z-\widehat{\tilde{\varepsilon _l}}^n\Big )\Big (u-\widehat{\tilde{\varepsilon _l}}^n\Big )}. \end{aligned}$$

Comparing with Eq. (B.2), the right hand side of Eq. (B.4) reads with the upper partial fraction decomposition

$$\begin{aligned}&\frac{1}{\mathfrak {G}_0(z)}\bigg [ \frac{\lambda }{N^2}\sum _{k=1}^d \sum _{n=1}^{\tilde{d}} r_k \tilde{r}_n \frac{\mathcal {T}^{(0)}(u\Vert \varepsilon _k,\tilde{\varepsilon }_n|)}{ (x(\varepsilon _k)-x(z))(y(\tilde{\varepsilon }_n)-y(z))} \\&\quad + \frac{\lambda ^2}{N^2 x'(u)} \sum _{k,n}^d \sum _{l,m}^{\tilde{d}} \frac{C_{k,l}^{m,n}}{\Big (z-\widehat{\varepsilon }_k^{\,m}\Big )\Big (z-\widehat{\tilde{\varepsilon _l}}^n\Big )\Big (u-\widehat{\tilde{\varepsilon _l}}^n\Big )^2} \bigg ]. \end{aligned}$$

The simple poles at \(z= \widehat{\varepsilon }_k^{\,m},\widehat{\tilde{\varepsilon _l}}^n\) cancel due to the prefactor \( \frac{1}{\mathfrak {G}_0(z)}\). By assumption, we can exclude further poles on the rhs at any zero z of \(\mathfrak {G}_0(z)\). Thus, the lhs and rhs of Eq. (B.4) must be a constant, and sending \(z \rightarrow \infty \), the constant has to be zero as claimed. \(\square \)

Appendix C. Proofs of Sect. 5

Section 5 is completely free of any proofs. This has two reasons: First, the proofs are very technical without greater learning effects. Second, the proofs are carried out in quite similar manner to established results for hermitian fields. Proposition 5.1 and Theorem 5.8 show up in a slightly more general form for hermitian fields in [Hoc20]. The other two results are generalisations of results found in [BHW22, Chapter 4].

Proof of Proposition 5.1

Define the differential operator

$$\begin{aligned} {\hat{D}} = \frac{\partial ^{2b-2}}{\partial J_{p^2q^2}\partial J^\dagger _{q^2p^2}\ldots \partial J_{p^bq^b}\partial J^\dagger _{q^bp^b}}. \end{aligned}$$

The same steps as for Proposition 3.1, namely using Proposition 2.8, give:

$$\begin{aligned} G_{|pq|\mathcal {J}}&=N^{b-2} {\hat{D}} \frac{\partial ^2}{\partial J_{pq} \partial J^\dagger _{qp}} \log [\mathcal {Z}(J,J^\dagger )]_{J,J^\dagger =0}\\&=- \frac{\lambda N^{b-4} {\hat{D}}}{E_p+\tilde{E}_q} \frac{\partial }{\partial J_{qp}^\dagger } \frac{1}{ \mathcal {Z}(J,J^\dagger )}\sum _{m,n=1}^N\frac{\partial ^3}{\partial J_{pm}\partial J_{mn}^\dagger \partial J_{nq}} \mathcal {Z}(J,J^\dagger )_{J,J^\dagger =0}\\&= - \frac{\lambda N^{b-4} {\hat{D}}}{E_p+\tilde{E}_q} \frac{\partial }{\partial J_{qp}^\dagger } \frac{1}{\mathcal {Z}(J,J^\dagger )} \biggl [ \frac{\partial }{\partial J_{pq} } W_p[J,J^\dagger ] \mathcal {Z}(J,J^\dagger )\\&\quad + \sum _{m,n=1}^N \frac{N}{E_n-E_p} \frac{\partial }{\partial J_{nq}} \bigg ( J^\dagger _{mp} \frac{\partial }{\partial J^\dagger _{mn}}-J_{nm} \frac{\partial }{\partial J_{pm}} \bigg ) \mathcal {Z}(J,J^\dagger )\bigg ] _{J,J^\dagger =0}. \end{aligned}$$

Note that we generalised the steps in Example 2.6 to the general operator \({\hat{D}}\). For the second line, write

$$\begin{aligned} \frac{1}{\mathcal {Z}(J,J^\dagger )} \frac{\partial }{\partial J_{pq} } \big ( W_p[J,J^\dagger ] \mathcal {Z}(J,J^\dagger ) \big ) = \frac{\partial }{\partial J_{pq} } W_p[J,J^\dagger ] + W_p[J,J^\dagger ] \frac{\partial }{\partial J_{pq} } \log [\mathcal {Z}(J,J^\dagger )]\;, \end{aligned}$$

with \(W_p[J,J^\dagger ]\) given by Proposition 2.8. Applying all derivatives on \(W_p\) gives by definition at \(J=J^\dagger =0\):

$$\begin{aligned} N^{b-2} {\hat{D}} \frac{\partial ^2 }{\partial J_{pq}\partial J_{qp}^\dagger } W_p[J,J^\dagger ] = \frac{1}{N^2} T_{p\Vert pq|\mathcal {J}|} . \end{aligned}$$

The second term, however, gives rise to a partitioned sum of two subsets, dependent on which subset of indices in \({\hat{D}}\) hit \( \log [\mathcal {Z}(J,J^\dagger )]\) itself:

$$\begin{aligned} N^{b-4} {\hat{D}} \bigg ( W_p[J,J^\dagger ] \frac{\partial ^2 }{\partial J_{pq}\partial J_{qp}^\dagger } \log [\mathcal {Z}(J,J^\dagger )]\bigg )\bigg \vert _{J=0}= \sum _{\mathcal {J}'\uplus \mathcal {J}''=\mathcal {J}} T_{p\Vert \mathcal {J}'|} G_{|pq|\mathcal {J}''|}. \end{aligned}$$

Demanding \(\mathcal {J}'\ne \emptyset \) gives \(\Omega _pG_{|pq|\mathcal {J}}\) in Proposition 5.1. The other terms arise from the part of the Ward identity. Excluding the term \(l=p\), we can write \( \frac{1}{N}\sum _{ l=1, l\ne p}^N \frac{G_{|lq|\mathcal {J}|}-G_{|pq|\mathcal {J}|}}{E_l-E_p}\) with the same argument as in Proposition 3.1. The excluded term can be included in \(\frac{1}{N^2} T_{p\Vert pq|\mathcal {J}|}\) to write it as \(\frac{1}{N}\frac{\partial G_{|pq|\mathcal {J}|}}{\partial E_p}\). The other realisation of fixing the indices nm by \(p^s\) and \(q^s\) (when \(\frac{\partial }{\partial J_{p^sq^s}} \in \hat{D}\) hits \(J_{mn}\) and vice versa) gives the very last term in Proposition 5.1 where two boundaries merge. \(\square \)

Proof of Proposition 5.5

The second term of the lhs of the DSE of Corollary 5.6 is rewritten into polynomial \(f(.;w|I|\mathcal {J})\) of degree \(d-1\) and a denominator also appearing in \(\mathcal {G}^{(0)}(z,w)\).

$$\begin{aligned} -\frac{\lambda }{N}\sum _{k=1}^d r_k \frac{\mathcal {T}^{(g)}(I\Vert \varepsilon _k,w|\mathcal {J}|)}{x(\varepsilon _k)-x(z)}&=\frac{\frac{\lambda }{N}\sum _{k=1}^d r_k \mathcal {T}^{(g)}(I\Vert \varepsilon _k,w|\mathcal {J}|) \prod _{i\ne k}^d (x(z)-x(\varepsilon _i))}{\prod _{j=1}^d(x(z)-x(\varepsilon _j))}\\&=:\frac{f(x(z);w|I|\mathcal {J})}{\prod _{j=1}^d (x(z)-x(\varepsilon _j))}, \end{aligned}$$

with one of the product representations of the 2-point function, Theorem 4.3, application of the interpolation formula (see (A.1)) with \(L_w(x(z)):=\prod _{j=1}^d (x(z)-x(\hat{\tilde{w}}^j)\), yields:

$$\begin{aligned} \frac{f(x(z);w|I|\mathcal {J})}{\prod _{j=1}^d (x(z)-x(\varepsilon _j))}&=\frac{ L_w(x(z))}{\prod _{j=1}^d (x(z)-x(\varepsilon _j))}\sum _{j=1}^d\frac{f(x(\hat{\tilde{w}}^j);w|I|\mathcal {J}|)}{(x(z)-x(\hat{\tilde{w}}^j) L_w'(x(\hat{\tilde{w}}^j)}\\&=-\lambda (y(w){-}y(z))\mathcal {G}^{(0)}(z,w) \sum _{j=1}^d{{\,\textrm{Res}\,}}_{t\rightarrow \hat{\tilde{w}}^j} \frac{f(x(t);w|I|\mathcal {J}|)x'(t)dt}{(x(z)-x(t)) L_w(x(t))}, \end{aligned}$$

where the analyticity of \(f(x(z);w|I|\mathcal {J})\) at \(z=\hat{\tilde{w}}^j\) was used. Next, insert the rhs of Corollary 5.6 for \(z\mapsto t\) near \(t=\hat{\tilde{w}}^j\) at which the first term of the lhs vanishes (here it is important that the integrand has only a simple pole at \(t=\hat{\tilde{w}}^j\)). Inserting it for \(f(x(t);w|I|\mathcal {J})\) leads to

$$\begin{aligned}&-\frac{\lambda }{N}\sum _{k=1}^dr_k \frac{\mathcal {T}^{(g)}(I\Vert \varepsilon _k,w|)}{x(\varepsilon _k)-x(z)} \\&=-\lambda (y(w){-}y(z))\mathcal {G}^{(0)}(z,w) \sum _{j=1}^d {{\,\textrm{Res}\,}}_{t\rightarrow \hat{\tilde{w}}^j} \frac{x'(t)dt}{(x(z){-}x(t))(y(w){-}y(t))\mathcal {G}^{(0)}(t,w)} \nonumber \\&\quad \times \bigg [ \sum _{\begin{array}{c} I_1\uplus I_2=I\\ g_1+g_2=g,~(g_1,I_1)\ne (0,\emptyset ) \end{array}} \Omega ^{(g_1)}_{|I_1|+1}(I_1,t) \mathcal {T}^{(g_2)}(I_2\Vert t,w|\mathcal {J}|) \nonumber \\&\quad +\mathcal {T}^{(g-1)}(I,t\Vert t,w|\mathcal {J}|) + \sum _{\begin{array}{c} I_1\uplus I_2=I\\ \mathcal {J}_1\uplus \mathcal {J}_2=\mathcal {J},~ \mathcal {J}_1\ne \emptyset \\ g_1+g_2=g \end{array}} \mathcal {T}^{(g_1)}(I_1,t\Vert \mathcal {J}_1|) \mathcal {T}^{(g_2)}(I_2\Vert t,w|\mathcal {J}_2|) \nonumber \\&\quad + \sum _{i=1}^m \frac{\partial }{\partial x(u_i)} \Big (\frac{\mathcal {T}^{(g)}(I{\setminus } u_i\Vert u_i,w|\mathcal {J}|)}{ x(u_i)-x(t)}\Big ) \nonumber \\&\quad - \sum _{s=2}^b\frac{ \mathcal {T}^{(g)}(I\Vert z^s,w^s,z^s,w|\mathcal {J}{\setminus } J^s|) - \mathcal {T}^{(g)}(I\Vert t,w^s,z^s,w|\mathcal {J}{\setminus } J^s|) }{x(z^s)-x(t)}. \nonumber \end{aligned}$$
(B.4)

Next, compute for the same integrand the residue at \(t= z\) (for arbitrary z):

$$\begin{aligned}&\lambda (y(w)-y(z))\mathcal {G}^{(0)}(z,w){{\,\textrm{Res}\,}}_{t\rightarrow z}\frac{x'(t)dt}{(x(z)-x(t))(y(w)-y(t))\mathcal {G}^{(0)}(t,w)} \\&\quad \times \bigg [ \sum _{\begin{array}{c} I_1\uplus I_2=I\\ g_1+g_2=g,~(g_1,I_1)\ne (0,\emptyset ) \end{array}} \Omega ^{(g_1)}_{|I_1|+1}(I_1,t) \mathcal {T}^{(g_2)}(I_2\Vert z,w|\mathcal {J}|) \\&\quad +\mathcal {T}^{(g-1)}(I,t\Vert t,w|\mathcal {J}|) + \sum _{\begin{array}{c} I_1\uplus I_2=I\\ \mathcal {J}_1\uplus \mathcal {J}_2=\mathcal {J},~ \mathcal {J}_1\ne \emptyset \\ g_1+g_2=g \end{array}} \mathcal {T}^{(g_1)}(I_1,t\Vert \mathcal {J}_1|) \mathcal {T}^{(g_2)}(I_2\Vert t,w|\mathcal {J}_2|)\\&\quad + \sum _{i=1}^m \frac{\partial }{\partial x(u_i)} \Big (\frac{\mathcal {T}^{(g)}(I{\setminus } u_i\Vert u_i,w|\mathcal {J}|)}{ x(u_i)-x(t)}\Big )\\&\quad - \sum _{s=2}^b\frac{ \mathcal {T}^{(g)}(I\Vert z^s,w^s,z^s,w|\mathcal {J}{\setminus } J^s|) - \mathcal {T}^{(g)}(I\Vert t,w^s,z^s,w|\mathcal {J}{\setminus } J^s|) }{x(z^s)-x(t)}\\&=(y(w)-y(z))\mathcal {T}^{(g)}(I\Vert z,w|\mathcal {J}|) -\frac{\lambda }{N}\sum _{k=1}^d r_k \frac{\mathcal {T}^{(g)}(I\Vert \varepsilon _k,w|\mathcal {J}|)}{x(\varepsilon _k)-x(z)}\;. \end{aligned}$$

Summing both expressions finishes the proof. \(\square \)

Proof of Corollary 5.6

First of all, we need to proof analyticity at certain points: \(\square \)

Lemma C.1

Let \(\mathcal {J}=\{J^2,\ldots ,J^b\}\) for \(J^s=[z^s,w^s]\) and \(I=\{u_1,\ldots ,u_m\}\). The generalised correlation functions \(\mathcal {T}^{(g)}(I\Vert z,w|\mathcal {J})\) are analytic at \(z=u_i\) and \(z=z^s\).

Proof

In principal, this is clear when returning to the matrix base with a finite limit of coinciding indices. The result can be shown inductively in the Euler characteristic in DSE (5.4), with analogous considerations to [BHW22, Lemma 4.1]. \(\square \)

The strategy of the proof consists of rewriting a term in Proposition 5.5 in terms of all the others - but taking the residues at different points:

$$\begin{aligned}&\frac{\partial }{\partial x(u_i)}\sum _{j=1}^d {{\,\textrm{Res}\,}}_{t\rightarrow z,\hat{\tilde{w}}^j} \frac{x'(t)dt\prod _{k=1}^d(x(t)-x(\varepsilon _k)) }{(x(z)-x(t)) L_w(x(t))} \frac{\mathcal {T}^{(g)}(I\backslash u_i\Vert u_i,w|\mathcal {J})}{x(u_i)-x(t)}\\&=\frac{\partial }{\partial x(u_i)} \sum _{j=1}^d {{\,\textrm{Res}\,}}_{v\rightarrow x(z),x(\hat{\tilde{w}}^j)} \frac{dv\prod _{k=1}^d(v-x(\varepsilon _k)) }{(x(z)-v) L_w(v)} \frac{\mathcal {T}^{(g)}(I\backslash u_i\Vert u_i,w|\mathcal {J})}{x(u_i)-v}\\&=-\frac{\partial }{\partial x(u_i)}{{\,\textrm{Res}\,}}_{v\rightarrow x(u_i)} \frac{dv\prod _{k=1}^d(v-x(\varepsilon _k)) }{(x(z)-v) L_w(v)} \frac{\mathcal {T}^{(g)}(I\backslash u_i\Vert u_i,w|\mathcal {J})}{x(u_i)-v}\\&=\frac{1}{x'(u_i)}\frac{\partial }{\partial u_i} \frac{\prod _{k=1}^d(x(u_i)-x(\varepsilon _k)) }{(x(z)-x(u_i)) L_w(x(u_i))}\mathcal {T}^{(g)}(I\backslash u_i\Vert u_i,w|\mathcal {J})\\&={{\,\textrm{Res}\,}}_{t\rightarrow u_i}\frac{x'(t)dt \prod _{k=1}^d(x(t)-x(\varepsilon _k)) }{(x(z)-x(t)) L_w(x(t))}\Big \{\frac{1}{x'(u_i)x'(t)(t-u_i)^2 } \mathcal {T}^{(g)}(I\backslash u_i\Vert t,w|\mathcal {J})\Big \}\;, \end{aligned}$$

where we substituted \(t\mapsto v=x(t)\), then moved the integration contour and finally represented the result in form of a residue formula. \(\Omega ^{(0)}_2(z,w)\) is the only correlation function divergent on the diagonal so that the terms in \(\{~ \}\) extend to \(\sum _{I_1,I_2,g_1,g_2} \Omega ^{(g_1)}_{|I_1|+1}(I_1;t)\mathcal {T}^{(g_2)}(I_2\Vert t,w|\mathcal {J})\) and finally to

$$\begin{aligned}&\sum _{\begin{array}{c} I_1\uplus I_2=I\\ g_1+g_2=g,~(g_1,I_1)\ne (0,\emptyset ) \end{array}} \Omega ^{(g_1)}_{|I_1|+1}(I_1,t) \mathcal {T}^{(g_2)}(I_2\Vert t,w|\mathcal {J}|) \nonumber \\&\quad +\mathcal {T}^{(g-1)}(I,t\Vert t,w|\mathcal {J}|) + \sum _{\begin{array}{c} I_1\uplus I_2=I\\ \mathcal {J}_1\uplus \mathcal {J}_2=\mathcal {J},~ \mathcal {J}_1\ne \emptyset \\ g_1+g_2=g \end{array}} \mathcal {T}^{(g_1)}(I_1,t\Vert \mathcal {J}_1|) \mathcal {T}^{(g_2)}(I_2\Vert t,w|\mathcal {J}_2|) \nonumber \\&\quad + \sum _{s=2}^b\frac{ \mathcal {T}^{(g)}(I\Vert t,w^s,z^s,w|\mathcal {J}{\setminus } J^s|) }{x(z^s)-x(t)}, \end{aligned}$$
(C.1)

where the regularity at \(z=u_i\) was exploited—under the residue operation, we added an effective zero. In the same manner, we rewrite the term

$$\begin{aligned}&\sum _{j=1}^d {{\,\textrm{Res}\,}}_{t\rightarrow z,\hat{\tilde{w}}^j} \frac{x'(t)dt\prod _{k=1}^d(x(t)-x(\varepsilon _k)) }{(x(z)-x(t)) L_w(x(t))} \frac{\mathcal {T}^{(g)}(I\Vert z^s,w^s,z^s,w|\mathcal {J}{\setminus } J^s|)}{x(z^s)-x(t)} \end{aligned}$$

for \(s \in \{2,\ldots ,b\}\) into a residue formula where we take the residue at \(t=z^s\) by deforming the contour. Since we have just a pole at order one at \(t=z^s\), we set an in the integrand the function

$$\begin{aligned} \mathcal {T}^{(g)}(I\Vert z^s,w^s,z^s,w|\mathcal {J}{\setminus } J^s|)\quad \rightarrow \quad \mathcal {T}^{(g)}(I\Vert t,w^s,z^s,w|\mathcal {J}{\setminus } J^s|). \end{aligned}$$

Again, the regularity argument of the all the other terms for the integrand in Eq. (C.2) at \(t=z^s\) allows for adding another zero. \(\square \)

Proof of Theorem 5.8

Assume \(p^i_j\) and \(q^i_j\) such that all \(E_{p^i_j}\) and \(\tilde{E}_{q^i_j}\) are pairwise different. Set \(a=p^1_1\), \(d=q^1_1\) and \(c=q^1_{N_1}\) to have a clear distinction between these and the remaining \(p^i_j\) and \(q^i_j\). Again, we define a differential operator \({\hat{D}}_{dc}\):

$$\begin{aligned} \hat{D}_{dc}=\frac{\partial ^{2N_1+..+2N_b-2}}{\partial J^\dagger _{dp^1_2}\partial J_{p^1_2q^1_2}.. \partial J_{p^1_{N_1}c} \partial J_{p^2_1q^2_1}\partial J^\dagger _{q^2_1p^2_2}.. \partial J^\dagger _{q^2_{N_2}p^2_1} ..\partial J_{p^b_1q^b_1}.. \partial J^\dagger _{q^b_{N_b}p^b_1}}. \end{aligned}$$

Bringing the global denominator of the theorem to the other side yields by definition of the correlation function

$$\begin{aligned}&(\tilde{E}_{d}-\tilde{E}_{c})G_{|p_1^1 q^1_1..q^1_{N_1}|\mathcal {J}|}\\ =&(\tilde{E}_{d}-\tilde{E}_{c})N^{b-2} \hat{D}_{dc}\frac{\partial ^2}{\partial J^\dagger _{ca}\partial J_{ad}} \log \mathcal {Z}(J,J^\dagger )\big \vert _{J,J^\dagger =0}\\ =&N^{b-2} \hat{D}_{dc}\frac{\partial ^2}{\partial J^\dagger _{ca}\partial J_{ad}}[E_a+\tilde{E}_{d}-(E_a+\tilde{E}_{c})]\log \mathcal {Z}(J,J^\dagger )\big \vert _{J,J^\dagger =0}. \end{aligned}$$

We generalise the steps in Example 2.6 to the general operator \({\hat{D}}_{dc}\):

$$\begin{aligned}&C \hat{D}_{dc}N^{b-1}\bigg (\frac{\partial }{\partial J^\dagger _{ca}} \frac{\exp (-N S_{int}(\frac{1}{N}\partial _J,\frac{1}{N}\partial _{J^\dagger }))J_{da}}{\mathcal {Z}(J,J^\dagger )}\\&\quad - \frac{\partial }{\partial J_{ad}} \frac{\exp (-N S_{int}(\frac{1}{N}\partial _J,\frac{1}{N}\partial _{J^\dagger }))J^\dagger _{ac}}{\mathcal {Z}(J,J^\dagger )}\bigg )\mathcal {Z}_{free}(J,J^\dagger )\big \vert _{J,J^\dagger =0}\\&=-N^{b-4}\lambda \hat{D}_{dc}\sum _{n,m=1}^N\left( \frac{\partial }{\partial J^\dagger _{ca}} \frac{\frac{\partial ^3}{\partial J_{an}\partial J^\dagger _{nm}\partial J_{md}}}{\mathcal {Z}(J,J^\dagger )}-\frac{\partial }{\partial J_{ad}} \frac{ \frac{\partial ^3}{\partial J^\dagger _{cm}\partial J_{mn}\partial J^\dagger _{na}}}{\mathcal {Z}(J,J^\dagger )}\right) \mathcal {Z}(J,J^\dagger )\big \vert _{J,J^\dagger =0}. \end{aligned}$$

For \(E_m=E_a\), the bracket vanishes for regular and non-regular terms. This comes from the fact that \(\frac{\partial }{\partial J^\dagger _{ca}}\) and \(\frac{\partial }{\partial J_{ad}}\) do not act on \(\frac{1}{\mathcal {Z}(J,J^\dagger )}\) because it gives 0 after taking \(J=0\) (no cycle in a), then setting \(m=a\) yields the same four derivatives.

Therefore, we can assume \(E_m\ne E_a\) and apply the simpler Ward identity of Proposition 2.7.

$$\begin{aligned}&=-\lambda \hat{D}_{dc}N^{b-3}\sum _{\begin{array}{c} n,m \\ m\ne a \end{array} }\bigg (\frac{\partial }{\partial J^\dagger _{ca}} \frac{\frac{\partial }{\partial J_{md}}}{\mathcal {Z}(J,J^\dagger )(E_m-E_a)} \big (J_{mn}\frac{\partial }{\partial J_{an}}- J^\dagger _{na}\frac{\partial }{\partial J^\dagger _{nm}}\big )\\&\quad +\frac{\partial }{\partial J_{ad}}\frac{ \frac{\partial }{\partial J^\dagger _{cm}}}{\mathcal {Z}(J,J^\dagger )(E_m-E_a)}\big (J_{an}\frac{\partial }{\partial J_{mn}}- J^\dagger _{nm}\frac{\partial }{\partial J^\dagger _{na}}\big )\bigg )\mathcal {Z}(J,J^\dagger )\big \vert _{J,J^\dagger =0}. \end{aligned}$$

The following derivatives lead to cancellations:

  • in the first line \(\frac{\partial }{\partial J^\dagger _{ca}}\) on \(J^\dagger _{na}\) is cancelled by \(\frac{\partial }{\partial J_{ad}}\) on \(J_{an}\) in the second line

  • if \(\frac{\partial }{\partial J_{md}}\) acts on \(J_{mn}\) is cancelled by \(\frac{\partial }{\partial J^\dagger _{cm}}\) on \(J^\dagger _{nm}\) in the second line

We end up with the surviving terms

$$\begin{aligned}&-\lambda \hat{D}_{dc}N^{b-3}\sum _{\begin{array}{c} n,m \\ m\ne a \end{array} }\frac{1}{E_m-E_a} \biggl (\frac{\partial }{\partial J^\dagger _{ca}}J_{mn}\frac{ \frac{\partial ^2}{\partial J_{an}\partial J_{md}}}{\mathcal {Z}(J,J^\dagger )}\\&\quad -\frac{\partial }{\partial J_{ad}}J^\dagger _{nm}\frac{ \frac{\partial ^2}{\partial J^\dagger _{cm}\partial J^\dagger _{na}}}{\mathcal {Z}(J,J^\dagger )}\biggl )\mathcal {Z}(J,J^\dagger )\big \vert _{J,J^\dagger =0}. \end{aligned}$$

We manipulate this equation using the general identity

$$\begin{aligned} \frac{\partial _x \partial _yf(x,y)}{f(x,y)}=\partial _x\partial _y\log [f(x,y)]+\partial _x\log [f(x,y)] \partial _y\log [f(x,y)] \end{aligned}$$

yields

$$\begin{aligned}&-\lambda \hat{D}_{dc}N^{b-3}\sum _{\begin{array}{c} n,m \\ m\ne a \end{array} }\frac{1}{E_m-E_a} \bigg \{\frac{\partial }{\partial J^\dagger _{ca}}J_{mn} \frac{\partial ^2}{\partial J_{an}\partial J_{md}}\log [\mathcal {Z}(J,J^\dagger )]\\&\quad -\frac{\partial }{\partial J_{ad}}J^\dagger _{nm} \frac{\partial ^2}{\partial J^\dagger _{cm}\partial J^\dagger _{na}}\log [\mathcal {Z}(J,J^\dagger )]\\&\quad + \frac{\partial }{\partial J^\dagger _{ca}}J_{mn} \bigg (\frac{\partial }{\partial J_{an}}\log [\mathcal {Z}(J,J^\dagger )]\bigg )\bigg (\frac{\partial }{\partial J_{md}}\log [\mathcal {Z}(J,J^\dagger )]\bigg )\\&\quad -\frac{\partial }{\partial J_{ad}}J^\dagger _{nm} \bigg (\frac{\partial }{\partial J^\dagger _{cm}}\log [\mathcal {Z}(J,J^\dagger )]\bigg ) \bigg (\frac{\partial }{\partial J^\dagger _{na}}\log [\mathcal {Z}(J,J^\dagger )]\bigg )\bigg \}\bigg \vert _{J,J^\dagger =0}. \end{aligned}$$

The nm are fixed by a derivative acting on \(J_{mn}\) (or \(J^\dagger _{nm}\)). We collect the following terms:

  • either a derivative of the form \(\frac{\partial }{\partial J^\dagger _{q^1_k p^1_{k}}}\) and \(\frac{\partial }{\partial J_{p^1_k q^1_{k}}}\), respectively, fixes the nm in the first two lines, which produces separated cycles

  • or a derivative of the form \(\frac{\partial }{\partial J^\dagger _{q^\beta _k p^\beta _{k}}}\) and \(\frac{\partial }{\partial J_{p^\beta _k q^\beta _{k}}}\), respectively, with \(\beta >1\) fixes the nm, which merges the first cycle with the \(\beta ^{\text {th}}\)-cycle.

  • in the last two lines it is only possible that a derivative of the form\(\frac{\partial }{\partial J^\dagger _{q^1_k p^1_{k}}}\) and \(\frac{\partial }{\partial J_{p^1_k q^1_{k}}}\), respectively, fixes the nm, otherwise setting \(J=0\) leads to vanishing contributions. Acting with the remaining derivatives of \(\hat{D}_{dc}\) on the product of the logarithms by considering the Leibniz rule leads to the assertion for pairwise different \(E_{p^i_j}\).

The expression stays true for coinciding \(E_{p^i_j}\) since the lhs is regular which induces a well-defined limit of the rhs by continuation to differentiable functions. A genus expansion and application of the boundary creation \(-N\frac{\partial }{\partial E_{p_i}}\) operator yields the assertion. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Branahl, J., Hock, A. Complete Solution of the LSZ Model via Topological Recursion. Commun. Math. Phys. 401, 2845–2899 (2023). https://doi.org/10.1007/s00220-023-04702-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-023-04702-z

Navigation