Skip to main content
Log in

Quantum Flag Manifold \(\sigma \)-Models and Hermitian Ricci Flow

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

Abstract

We show that flag manifold \(\sigma \)-models (including \(\mathbb{C}\mathbb{P}^{n-1}\), Grassmannian models as special cases) and their deformed versions may be cast in the form of gauged bosonic Thirring/Gross-Neveu-type systems. Quantum mechanically the gauging is violated by chiral anomalies, which may be cancelled by adding fermions. We conjecture that such models are integrable and check on some examples that the trigonometrically deformed geometries satisfy the generalized Ricci flow equations.

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

Similar content being viewed by others

Notes

  1. In [27] we presented a derivation of the sausage model from the perspective of Pohlmeyer reduction [31].

  2. In [27] we found that the flow is also linear in the case of generalized Kähler deformation of \(\mathbb{C}\mathbb{P}^{n-1}\)—this is the deformation that arises from the construction of [23]. Besides, as found in [36], the same type of flow holds for the so-called \(\lambda \)-deformations of symmetric space models [37, 38]. Finally, the elliptic deformation of the models studied here exhibits a similar behavior at least in the \({\mathfrak {s}}{\mathfrak {l}}_2\)-case [39].

  3. Here and below \(\mathbb {1}_n\) denotes the unit matrix.

  4. Another option is taking an \({\mathcal {R}}\)-matrix induced by a complex structure on the Lie group G with Lie algebra \(\mathfrak {g}\) [47]. In this case (2.8) is the condition of vanishing of the Nijenhuis tensor.

  5. The Boltzmann weight in the path integral is \(e^{-{1\over \hbar }\int \,d^2z\, \mathscr {L}}\), so that the kinetic term in \(\mathscr {L}\) is imaginary, and the potential term is real and positive (for \(\hbar >0\)), as it should be in the case of first order actions in Euclidean space, cf. [49].

  6. In fact, in the context of \(\beta \gamma \)-systems it is somewhat more convenient to talk about complex ‘phase space’, which in this case is \(T^*\mathbb {C}^n\). The coordinates \(U \in \textrm{Hom}(\mathbb {C}, \mathbb {C}^n)\) and \(V\in \textrm{Hom}(\mathbb {C}^n, \mathbb {C})\) used below are naturally thought of as the coordinates in the base and fiber of \(T^*\mathbb {C}^n=\mathbb {C}^n\oplus \mathbb {C}^n\). This abstraction is not strictly necessary for the purposes of the present paper but proves indispensable for the generalizations, in particular the ones involving fermions, cf. [51] and the brief review [39].

  7. In Minkowski signature this would have been the usual \(U(1)\times U(1)\) chiral symmetry. This difference in chiral transformations has been observed in [56, 57].

  8. Here we are talking about UV divergences.

  9. Here \(\mathbb {C}^n\otimes \mathbb {C}^m\) is symbolic notation, which is supposed to mean that the fields are matrix-valued, taking values in \(\mathbb {C}^n\otimes \mathbb {C}^m=\mathbb {C}^{mn}\simeq \textrm{Hom}(\mathbb {C}^m, \mathbb {C}^n)\simeq \textrm{Hom}(\mathbb {C}^n, \mathbb {C}^m)\).

  10. For the \(\beta \)-functions in the r.h.s. of these equations see [60, 61].

  11. In fact, there is no local function of the fields \({\mathcal {A}}, \overline{{\mathcal {A}}}\) and their derivatives that would be invariant w.r.t. this transformation.

  12. See [66] for a review of spinors in Euclidean spaces.

  13. From the point of view of \(\sigma \)-models this may also be seen in a direct calculation of the one-loop \(\beta \)-function, see [73] for a review.

  14. The verification was done numerically for low values of n.

  15. In the case of Hermitian symmetric spaces, and in the homogeneous/rational situation, the B-field is topological, but the Lax pair constructed according to Pohlmeyer’s procedure is different from the conventional Lax pair in the absence of the B-field [3].

  16. Ultralocal Lax pairs for the deformed \(\mathbb{C}\mathbb{P}^1\)-model were constructed in [21, 81].

  17. We introduce a factor of 2 in the definition of \(k_{ij}\) so that the metric satisfies the canonically normalized Ricci flow Eq. (3.23). One other way to fix this normalization is via the Einstein condition (5.1) for undeformed metrics.

References

  1. Bykov, D.: Integrable properties of sigma-models with non-symmetric target spaces. Nucl. Phys. B 894, 254–267 (2015). https://doi.org/10.1016/j.nuclphysb.2015.03.005. arXiv:1412.3746 [hep-th]

    Article  ADS  MATH  Google Scholar 

  2. Bykov, D.: Classical solutions of a flag manifold -model. Nucl. Phys. B 902, 292–301 (2016). https://doi.org/10.1016/j.nuclphysb.2015.11.015. arXiv:1506.08156 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Bykov, D.: Complex structures and zero-curvature equations for -models. Phys. Lett. B 760, 341 (2016). https://doi.org/10.1016/j.physletb.2016.06.071. arXiv:1605.01093 [hep-th]

    Article  ADS  MATH  Google Scholar 

  4. Costello, K., Yamazaki, M.: Gauge Theory and Integrability, III. arXiv:1908.02289 [hep-th]

  5. Affleck, I., Bykov, D., Wamer, K.: Flag manifold sigma models: spin chains and integrable theories. Phys. Rept. 953, 1–93 (2022). arXiv:2101.11638 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. D’Adda, A., Lüscher, M., Di Vecchia, P.: A \(1\over n\) expandable series of nonlinear sigma models with instantons. Nucl. Phys. B 146, 63–76 (1978). https://doi.org/10.1016/0550-3213(78)90432-7

    Article  ADS  Google Scholar 

  7. D’Adda, A., Di Vecchia, P., Lüscher, M.: Confinement and chiral symmetry breaking in \({\mathbb{C}\mathbb{P} }^{n-1}\) models with quarks. Nucl. Phys. B 152, 125–144 (1979). https://doi.org/10.1016/0550-3213(79)90083-X

    Article  ADS  Google Scholar 

  8. Eichenherr, H., Forger, M.: On the dual symmetry of the nonlinear sigma models. Nucl. Phys. B 155, 381–393 (1979). https://doi.org/10.1016/0550-3213(79)90276-1

    Article  ADS  MathSciNet  Google Scholar 

  9. Eichenherr, H., Forger, M.: More about non-linear sigma models on symmetric spaces. Nucl. Phys. B 164, 528–535 (1980). https://doi.org/10.1016/0550-3213(80)90525-8

    Article  ADS  Google Scholar 

  10. Din, A., Zakrzewski, W.: General classical solutions in the \({\mathbb{C}\mathbb{P} }^{n-1}\) model. Nucl. Phys. B 174, 397–406 (1980). https://doi.org/10.1016/0550-3213(80)90291-6

    Article  ADS  Google Scholar 

  11. Din, A., Zakrzewski, W.: Classical solutions in Grassmannian models. Lett. Math. Phys. 5, 553 (1981). https://doi.org/10.1007/BF00408138

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Perelomov, A.: Solutions of the instanton type in chiral models. Sov. Phys. Usp. 24, 645–661 (1981). https://doi.org/10.1070/PU1981v024n08ABEH004829

    Article  ADS  MathSciNet  Google Scholar 

  13. Morozov, A., Perelomov, A., Shifman, M.A.: Exact Gell-Mann-low function of supersymmetric Kähler sigma models. Nucl. Phys. B 248, 279 (1984). https://doi.org/10.1016/0550-3213(84)90598-4

    Article  ADS  Google Scholar 

  14. Perelomov, A.: Chiral models: geometrical aspects. Phys. Rept. 146, 135–213 (1987). https://doi.org/10.1016/0370-1573(87)90044-5

    Article  Google Scholar 

  15. Wang, H.-C.: Closed manifolds with homogeneous complex structure. Am. J. Math. 76(1), 1–32 (1954). https://doi.org/10.2307/2372397

    Article  MathSciNet  MATH  Google Scholar 

  16. Cherednik, I.V.: Relativistically invariant quasiclassical limits of integrable two-dimensional quantum models. Theoret. Math. Phys. 47(2), 422–425 (1981). https://doi.org/10.1007/BF01086395

    Article  Google Scholar 

  17. Fateev, V.: The sigma model (dual) representation for a two-parameter family of integrable quantum field theories. Nucl. Phys. B 473, 509–538 (1996). https://doi.org/10.1016/0550-3213(96)00256-8

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. Klimčík, C.: Yang-Baxter sigma models and dS/AdS T duality. JHEP 0212, 051 (2002). https://doi.org/10.1088/1126-6708/2002/12/051. arXiv:hep-th/0210095

    Article  ADS  MathSciNet  Google Scholar 

  19. Klimčík, C.: On integrability of the Yang-Baxter sigma-model. J. Math. Phys. 50, 043508 (2009). https://doi.org/10.1063/1.3116242. arXiv:0802.3518 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Klimčík, C.: Integrability of the bi-Yang-Baxter sigma-model. Lett. Math. Phys. 104, 1095 (2014). https://doi.org/10.1007/s11005-014-0709-y. arXiv:1402.2105 [math-ph]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Lukyanov, S.L.: The integrable harmonic map problem versus Ricci flow. Nucl. Phys. B 865, 308–329 (2012). https://doi.org/10.1016/j.nuclphysb.2012.08.002. arXiv:1205.3201 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Delduc, F., Hoare, B., Kameyama, T., Magro, M.: Combining the bi-Yang-Baxter deformation, the Wess-Zumino term and TsT transformations in one integrable -model. JHEP 10, 212 (2017). https://doi.org/10.1007/JHEP10(2017)212. arXiv:1707.08371 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Delduc, F., Magro, M., Vicedo, B.: On classical \(q\)-deformations of integrable sigma-models. JHEP 1311, 192 (2013). https://doi.org/10.1007/JHEP11(2013)192. arXiv:1308.3581 [hep-th]

    Article  ADS  MATH  Google Scholar 

  24. Costello, K., Witten, E., Yamazaki, M.: Gauge Theory and Integrability, I. https://doi.org/10.4310/ICCM.2018.v6.n1.a6. arXiv:1709.09993 [hep-th]

  25. Costello, K., Witten, E., Yamazaki, M.: Gauge Theory and Integrability, II. https://doi.org/10.4310/ICCM.2018.v6.n1.a7. arXiv:1802.01579 [hep-th]

  26. Demulder, S., Hassler, F., Piccinini, G., Thompson, D.C.: Integrable deformation of \(\mathbb{C}\mathbb{P} ^n\) and generalised Kähler geometry. JHEP 10, 086 (2020). https://doi.org/10.1007/JHEP10(2020)086. arXiv:2002.11144 [hep-th]

    Article  ADS  MATH  Google Scholar 

  27. Bykov, D., Lüst, D.: Deformed -models, Ricci flow and Toda field theories. Lett. Math. Phys. 111, 150 (2021). arXiv:2005.01812 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Litvinov, A.V.: Integrable \({\mathfrak{g} }{\mathfrak{l} }(n|n)\) Toda field theory and its sigma-model dual. Pisma Zh. Eksp. Teor. Fiz. 110(11), 723 (2019). https://doi.org/10.1134/S0021364019230048. arXiv:1901.04799 [hep-th]

    Article  Google Scholar 

  29. Fateev, V.: Classical and Quantum Integrable Sigma Models. Ricci Flow, “Nice Duality” and Perturbed Rational Conformal Field Theories. J. Exp. Theor. Phys. 129, 566 (2019). https://doi.org/10.1134/S1063776119100042. arXiv:1902.02811 [hep-th]

  30. Fateev, V.A., Onofri, E., Zamolodchikov, A.B.: Integrable deformations of the \(O(3)\) sigma model. The sausage model. Nucl. Phys. B 406, 521 (1993). https://doi.org/10.1016/0550-3213(93)90001-6

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Pohlmeyer, K.: Integrable Hamiltonian systems and interactions through quadratic constraints. Commun. Math. Phys. 46, 207 (1976). https://doi.org/10.1007/BF01609119

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Perelman, G.: The Entropy Formula for the Ricci Flow and its Geometric Applications. arXiv:math/0211159

  33. Hamilton, R.S.: Three-manifolds with positive Ricci curvature. J. Differ. Geom. 17(2), 255–306 (1982). https://doi.org/10.4310/jdg/1214436922

    Article  MathSciNet  MATH  Google Scholar 

  34. Hamilton, R.S.: The Ricci flow on surfaces. In: Mathematics and General Relativity (Santa Cruz, CA, 1986), Contemporary Mathematics, vol. 71, pp. 237–262. American Mathematical Society, Providence, RI (1988). https://doi.org/10.1090/conm/071/954419

  35. Bakas, I., Kong, S., Ni, L.: Ancient solutions of Ricci flow on spheres and generalized Hopf fibrations. J. Reine Angew. Math. 663, 209–248 (2012). https://doi.org/10.1515/CRELLE.2011.101. arXiv:0906.0589 [math.DG]

  36. Appadu, C., Hollowood, T.J.: Beta function of k deformed \(\text{ AdS}_{5}\,\times \text{ S}^{5}\) string theory. JHEP 11, 095 (2015). https://doi.org/10.1007/JHEP11(2015)095. arXiv:1507.05420 [hep-th]

    Article  ADS  Google Scholar 

  37. Sfetsos, K.: Integrable interpolations: from exact CFTs to non-Abelian T-duals. Nucl. Phys. B 880, 225–246 (2014). https://doi.org/10.1016/j.nuclphysb.2014.01.004. arXiv:1312.4560 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Hollowood, T.J., Miramontes, J.L., Schmidtt, D.M.: Integrable deformations of strings on symmetric spaces. JHEP 11, 009 (2014). https://doi.org/10.1007/JHEP11(2014)009. arXiv:1407.2840 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  39. Bykov, D.: Sigma models as Gross-Neveu models. Teor. Mat. Fiz. 208(2), 165–179 (2021). https://doi.org/10.1134/S0040577921080018. arXiv:2106.15598 [hep-th]

    Article  MathSciNet  MATH  Google Scholar 

  40. Nekrasov, N.A.: Lectures on curved beta-gamma systems, pure spinors, and anomalies. arXiv:hep-th/0511008

  41. Witten, E.: Two-dimensional models with (0,2) supersymmetry: Perturbative aspects. Adv. Theor. Math. Phys. 11(1), 1–63 (2007). https://doi.org/10.4310/ATMP.2007.v11.n1.a1. arXiv:hep-th/0504078

    Article  MathSciNet  MATH  Google Scholar 

  42. Lüscher, M.: Quantum nonlocal charges and absence of particle production in the two-dimensional nonlinear sigma model. Nucl. Phys. B 135, 1–19 (1978). https://doi.org/10.1016/0550-3213(78)90211-0

    Article  ADS  Google Scholar 

  43. Abdalla, E., Abdalla, M., Gomes, M.: Anomaly cancellations in the supersymmetric CP\(^{(N-1)}\) model. Phys. Rev. D 25, 452 (1982). https://doi.org/10.1103/PhysRevD.25.452

    Article  ADS  Google Scholar 

  44. Abdalla, E., Forger, M., Lima Santos, A.: Nonlocal charges for nonlinear models on Grassmann manifolds. Nucl. Phys. B 256, 145–180 (1985). https://doi.org/10.1016/0550-3213(85)90389-X

    Article  ADS  MathSciNet  Google Scholar 

  45. Abdalla, E., Abdalla, M.C.B., Rothe, K.D.: Non-perturbative Methods in Two-dimensional Quantum Field Theory. World Scientific, Singapore (2001). https://doi.org/10.1142/4678

  46. Belavin, A.A., Drinfeld, V.G.: Solutions of the classical Yang-Baxter equation for simple Lie algebras. Funktsional. Anal. i Prilozhen. 16(3), 1–29 (1982). https://doi.org/10.1007/BF01081585. ( Funct. Anal. Appl. 16(3), 159-180 (1982))

  47. Bykov, D.: Complex structure-induced deformations of -models. JHEP 03, 130 (2017). https://doi.org/10.1007/JHEP03(2017)130. arXiv:1611.07116 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. Bykov, D.: Flag manifold sigma-models and nilpotent orbits. Proc. Steklov Inst. Math. 309, 78–86 (2020). https://doi.org/10.1134/S0081543820030062. arXiv:1911.07768 [hep-th]

    Article  MathSciNet  MATH  Google Scholar 

  49. Zinn-Justin, J.: Quantum Field Theory and Critical Phenomena. Oxford University Press, Oxford (1996). https://doi.org/10.1093/acprof:oso/9780198509233.001.0001

  50. Green, M.B., Schwarz, J.H., Witten, E.: Superstring Theory. Vol. 1: Introduction. Cambridge University Press, Cambridge (2012). https://doi.org/10.1017/CBO9781139248563

  51. Bykov, D.: The \({{\sf CP}^{{\sf n-1}}}\)-model with fermions: a new look. To appear in Adv. Theor. Math. Phys. arXiv:2009.04608 [hep-th]

  52. Witten, E.: Topological sigma models. Commun. Math. Phys. 118, 411 (1988). https://doi.org/10.1007/BF01466725

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. Gawedzki, K.: Noncompact WZW conformal field theories. arXiv:hep-th/9110076

  54. de Boer, J., Shatashvili, S.L.: Two-dimensional conformal field theories on AdS(2d+1) backgrounds. JHEP 06, 013 (1999). https://doi.org/10.1088/1126-6708/1999/06/013. arXiv:hep-th/9905032

    Article  MathSciNet  Google Scholar 

  55. Witten, E.: Chiral symmetry, the 1/n expansion, and the SU(N) Thirring model. Nucl. Phys. B 145, 110–118 (1978). https://doi.org/10.1016/0550-3213(78)90416-9

    Article  ADS  Google Scholar 

  56. Zumino, B.: Euclidean supersymmetry and the many-instanton problem. Phys. Lett. B 69, 369 (1977). https://doi.org/10.1016/0370-2693(77)90568-8

    Article  ADS  Google Scholar 

  57. Mehta, M.R.: Euclidean continuation of the Dirac fermion. Phys. Rev. Lett. 65, 1983–1986 (1990). https://doi.org/10.1103/PhysRevLett.65.1983

    Article  ADS  MathSciNet  MATH  Google Scholar 

  58. Kac, V.G.: Automorphisms of finite order of semisimple Lie algebras. Funktsional. Anal. i Prilozhen. 3(3), 94–96 (1969). https://doi.org/10.1007/BF01676631. ( Funct. Anal. Appl., 3:3 (1969), 252-254)

  59. Bykov, D.: Flag manifold -models: the \(\frac{1}{N}\)-expansion and the anomaly two-form. Nucl. Phys. B 941, 316–360 (2019). https://doi.org/10.1016/j.nuclphysb.2019.02.006. arXiv:1901.02861 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  60. Curci, G., Paffuti, G.: Consistency between the string background field equation of motion and the vanishing of the conformal anomaly. Nucl. Phys. B 286, 399–408 (1987). https://doi.org/10.1016/0550-3213(87)90447-0

    Article  ADS  Google Scholar 

  61. Polchinski, J.: String Theory, vol. 1. Cambridge University Press, Cambridge (1998). https://doi.org/10.1017/CBO9780511816079

  62. Buscher, T.: Path integral derivation of quantum duality in nonlinear sigma models. Phys. Lett. B 201, 466–472 (1988). https://doi.org/10.1016/0370-2693(88)90602-8

    Article  ADS  MathSciNet  Google Scholar 

  63. Tseytlin, A.A.: Effective action of gauged WZW model and exact string solutions. Nucl. Phys. B 399, 601–622 (1993). https://doi.org/10.1016/0550-3213(93)90511-M. arXiv:hep-th/9301015

    Article  ADS  MathSciNet  Google Scholar 

  64. Schwarz, A.S., Tseytlin, A.A.: Dilaton shift under duality and torsion of elliptic complex. Nucl. Phys. B 399, 691–708 (1993). https://doi.org/10.1016/0550-3213(93)90514-P. arXiv:hep-th/9210015

    Article  ADS  Google Scholar 

  65. Alvarez-Gaume, L., Moore, G.W., Vafa, C.: Theta functions, modular invariance and strings. Commun. Math. Phys. 106, 1–40 (1986). https://doi.org/10.1007/BF01210925

    Article  ADS  MathSciNet  MATH  Google Scholar 

  66. van Nieuwenhuizen, P., Waldron, A.: On Euclidean spinors and Wick rotations. Phys. Lett. B 389, 29–36 (1996). https://doi.org/10.1016/S0370-2693(96)01251-8. arXiv:hep-th/9608174

    Article  ADS  MathSciNet  Google Scholar 

  67. Wulff, L., Tseytlin, A.: Kappa-symmetry of superstring sigma model and generalized 10d supergravity equations. JHEP 06, 174 (2016). https://doi.org/10.1007/JHEP06(2016)174. arXiv:1605.04884 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  68. Arutyunov, G., Frolov, S., Hoare, B., Roiban, R., Tseytlin, A.: Scale invariance of the \(\eta \)-deformed \(AdS_5\times S^5\) superstring, T-duality and modified type II equations. Nucl. Phys. B 903, 262–303 (2016). https://doi.org/10.1016/j.nuclphysb.2015.12.012. arXiv:1511.05795 [hep-th]

    Article  ADS  MATH  Google Scholar 

  69. Abdalla, E., Forger, M.: Integrable nonlinear models with fermions. Commun. Math. Phys. 104, 123 (1986). https://doi.org/10.1007/BF01210796

    Article  ADS  MathSciNet  MATH  Google Scholar 

  70. Bykov, D.: The worldsheet low-energy limit of the \(AdS_4 \times {\mathbb{C}\mathbb{P} }^3\) superstring. Nucl. Phys. B 838, 47–74 (2010). https://doi.org/10.1016/j.nuclphysb.2010.05.013. arXiv:1003.2199 [hep-th]

    Article  ADS  Google Scholar 

  71. Basso, B., Rej, A.: On the integrability of two-dimensional models with \(U(1)\times SU(N)\) symmetry. Nucl. Phys. B 866, 337–377 (2013). https://doi.org/10.1016/j.nuclphysb.2012.09.003. arXiv:1207.0413 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  72. Neveu, A., Papanicolaou, N.: Integrability of the classical scalar and symmetric scalar-pseudoscalar contact fermi interactions in two-dimensions. Commun. Math. Phys. 58, 31 (1978). https://doi.org/10.1007/BF01624787

    Article  ADS  Google Scholar 

  73. Zarembo, K.: Integrability in Sigma-Models, Les Houches Lect. Notes 106 (2019) https://doi.org/10.1093/oso/9780198828150.003.0005. arXiv:1712.07725 [hep-th]

  74. Wang, M.Y., Ziller, W.: On normal homogeneous Einstein manifolds. Annales scientifiques de l’É.N.S. 4e série, tome 18(4), 563–633 (1985). https://doi.org/10.24033/asens.1497

  75. Alekseevsky, D.V.: Homogeneous Einstein metrics. In: Differential Geometry and its Applications (Proceedings of Brno Conference), pp. 1–21. University of J. E. Purkyne, Czechoslovakia (1987)

  76. Arvanitoyeorgos, A.: New invariant Einstein metrics on generalized flag manifolds. Trans. Am. Math. Soc. 337(2), 981–995 (1993). https://doi.org/10.1090/S0002-9947-1993-1097162-3

    Article  MathSciNet  MATH  Google Scholar 

  77. Alekseevsky, D.V.: Flag manifolds. In: 11th Yugoslav Geometrical Seminar (Divcibare, 1996), Zb. Rad. Mat. Inst. Beograd. (N.S.) 6(14), 3–35 (1997)

  78. Bytsko, A.: The zero curvature representation for nonlinear \(O(3)\) sigma model. J. Math. Sci. 85, 1619–1628 (1994). https://doi.org/10.1007/BF02355322. arXiv:hep-th/9403101

    Article  MathSciNet  MATH  Google Scholar 

  79. Brodbeck, O., Zagermann, M.: Dimensionally reduced gravity, Hermitian symmetric spaces and the Ashtekar variables. Class. Quant. Grav. 17, 2749–2764 (2000). https://doi.org/10.1088/0264-9381/17/14/310. arXiv:gr-qc/9911118

    Article  ADS  MathSciNet  MATH  Google Scholar 

  80. Delduc, F., Kameyama, T., Lacroix, S., Magro, M., Vicedo, B.: Ultralocal Lax connection for para-complex \({\mathbb{Z} }_T\)-cosets. Nucl. Phys. B 949, 114821 (2019). https://doi.org/10.1016/j.nuclphysb.2019.114821. arXiv:1909.00742 [hep-th]

    Article  MATH  Google Scholar 

  81. Bazhanov, V.V., Kotousov, G.A., Lukyanov, S.L.: Quantum transfer-matrices for the sausage model. JHEP 01, 021 (2018). https://doi.org/10.1007/JHEP01(2018)021. arXiv:1706.09941 [hep-th]

    Article  ADS  MathSciNet  MATH  Google Scholar 

  82. Blau, M.: Lecture notes on general relativity. Accessed on 21 June 2020. http://www.blau.itp.unibe.ch/GRLecturenotes.html

  83. Bykov, D.: Integrable sigma models on Riemann surfaces. https://doi.org/10.48550/arXiv.2202.12805 (2022)

Download references

Acknowledgements

I would like to thank A. A. Slavnov for support and G. Arutyunov, D. Lüst for discussions. I am especially grateful to K. Zarembo for reading the manuscript and many useful remarks and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dmitri Bykov.

Additional information

Communicated by V. Pestun

Publisher's Note

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

Appendices

Metric, B-Field and Dilaton of the Deformed \(\mathbb {C}^n\)-Model

Here we derive the formulas (3.24) from the main text. We start by writing out the quadratic form in the Lagrangian (3.1):

(A.1)

The constants \(\upalpha , \upbeta , \upgamma \) were defined in (3.9). Inverting the quadratic form Q, we obtain the Lagrangian of the \(\sigma \)-model:

$$\begin{aligned} \mathscr {L}=\sum \limits _{k, m=1}^n\,\partial {\overline{U}}_k \left( \delta _{km}{{\hat{\lambda }}_k}-\frac{{\hat{\lambda }}_k {\hat{\lambda }}_m\,U_k {\overline{U}}_m}{\upkappa +\sum \limits _p\,{\hat{\lambda }}_p\,|U_p|^2}\right) {\overline{\partial }}U_m,\quad \upkappa =(b-{\upgamma \over n})^{-1}. \end{aligned}$$
(A.2)

It is convenient to introduce the parametrization \(U_j=e^{t_j- i\,\phi _j}\) and the notation \(\lambda _j:={\hat{\lambda }}_j \,e^{2t_j}\) (we already used it in Sect. 3.4). In this case \(d U_j\wedge d{\overline{U}}_j=2i\,e^{2t_j}\,dt_j\wedge d\phi _j\) and \(\sum \limits _{j=1}^n\,{\hat{\lambda }}_j\,U_j \,d{\overline{U}}_j=\sum \limits _{j=1}^n\,\lambda _j\,(dt_j- i\,d\phi _j)\). ThereforeFootnote 17

$$\begin{aligned}{} & {} ds^2= \sum \limits _{i, j=1}^n\,k_{ij} (dt_i\,dt_j+d\phi _i d\phi _j). \end{aligned}$$
(A.3)
$$\begin{aligned}{} & {} B= \sum \limits _{i, j=1}^n\,k_{ij}\, dt_i\wedge d\phi _j,\quad \text {where}\quad {1\over 2}\,k_{ij}=\lambda _i\,\delta _{ij}-\frac{\lambda _i \lambda _j}{\upkappa +\sum \limits _k \lambda _k}. \end{aligned}$$
(A.4)

The dilaton \(\Phi \) arises from the integration over the V-variables, namely from the determinant of the quadratic form in (A.1):

$$\begin{aligned} e^{-2\Phi }=\textrm{Det}(Q)=\prod \limits _{k=1}^n\,{1\over {\hat{\lambda }}_k}\times \left( 1+{1\over \upkappa }\,\sum \limits _{k=1}^n{\hat{\lambda }}_k |U_k|^2\right) . \end{aligned}$$
(A.5)

1.1 \(n=2\): the deformed \(\mathbb {C}^2\)

In computing the Ricci tensor and other geometric quantities it is much more convenient to deal with the inverse metric \(\kappa =k^{-1}\), which has the simple form

(A.6)

After the shift of \(t_i\)-variables described in Sect. 3.2 the functions \(\upchi _i\) are found to be

$$\begin{aligned} \upchi _1={1\over 2}\,{1+e^{2\tau }\over 1-e^{2\tau }}+{e^{\tau }\over 1-e^{2\tau }}\,e^{t_2-t_1},\quad \upchi _2={1\over 2}\,{1+e^{2\tau }\over 1-e^{2\tau }}+{e^{\tau }\over 1-e^{2\tau }}\,e^{t_1-t_2}. \end{aligned}$$
(A.7)

We start by computing the angular components \(R^{\phi _i \phi _j}\) (again, with upper indices) of the Ricci tensor. These may be obtained using the formula \(R^\mu _{\nu }\,K^\nu ={1\over \sqrt{g}}\,\partial _\rho \left( \sqrt{g}\,\nabla ^{\mu }K^\rho \right) \), valid for any Killing vector field K (cf. [82]). As it follows from (A.6), \({\partial \kappa ^{ij}\over \partial t_k}={1\over 2}{\partial \upchi _i\over \partial t_k}\,\delta ^{ij}\). Using this as well as the fact that \(\upchi _i=\upchi _i(t_1-t_2)\), we arrive at the formula for the \(\phi \)-components of the Ricci tensor:

$$\begin{aligned} R^{\phi _i \phi _j}=\sum \limits _{p=1}^n\,{1\over 8k}\,{\partial \over \partial t_p}\left( k\,\upchi _p\,{\partial \upchi _i\over \partial t_p}\right) \,\delta _{ij}-k_{ij}\,\sum \limits _{p=1}^n\,{1\over 8}\,\upchi _p\,{\partial \upchi _i\over \partial t_p} \,{\partial \upchi _j\over \partial t_p},\quad k=\textrm{det}(k_{ij}).\nonumber \\ \end{aligned}$$
(A.8)

For other ingredients of the r.h.s. of (3.23) we also obtain expressions in terms of \(\upchi _i\):

$$\begin{aligned} (H^2)^{\phi _i \phi _j}=-{1\over 2}{\partial \upchi _i\over \partial t_j}\,{\partial \upchi _j\over \partial t_i}+{1\over 2}\,k_{ij}\,\sum \limits _{p=1}^n\,\upchi _p\,{\partial \upchi _i\over \partial t_p} \,{\partial \upchi _j\over \partial t_p}\, \end{aligned}$$
(A.9)
$$\begin{aligned} 2\,\nabla ^{\phi _i} \nabla ^{\phi _j} \Phi ={1\over 8}\,\delta ^{ij}\,\sum \limits _{p=1}^n\,\upchi _p\,{\partial \upchi _i\over \partial t_p}\,\left( 2-{1\over k}\,{\partial k\over \partial t_p}\right) . \end{aligned}$$
(A.10)

Here we have used that the dilaton is \(\Phi =-t_1-t_2+{1\over 2}\log {k}\). Assembling the pieces, we arrive at the following expression for the r.h.s of the Ricci flow Eq. (3.23):

(A.11)

Using the explicit expressions (A.7), we find that

$$\begin{aligned} {d \kappa ^{ij}\over d\tau }=R^{\phi _i \phi _j}+{1\over 4}\,(H^2)^{\phi _i \phi _j}+2\,\nabla ^{\phi _i} \nabla ^{\phi _j} \Phi . \end{aligned}$$
(A.12)

The mixed \((t, \phi )\)-components of both sides of the Ricci flow equations are identically zero, so it remains to compute the (tt)-components. Here we do not find any significant simplifications, so we just outline the strategy of how this can be done. The (tt)-components of the Ricci tensor may be written as follows:

$$\begin{aligned} R_{t_i t_j}=R_{ij}^{(k)}-{1\over 2}\,\nabla _i \nabla _j \,\log {k}-{1\over 4}\,\textrm{Tr}\left( {\partial k^{-1}\over \partial t_i}\,k\, {\partial k^{-1}\over \partial t_j}\,k\right) \, \end{aligned}$$
(A.13)

where \(R_{ij}^{(k)}\) is the Ricci tensor of the two-dimensional metric \(k_{ij}\). As for any such metric, \(R_{ij}^{(k)}={R\over 2}\,k_{ij}\), where R is the Ricci scalar. The latter can be computed by noticing that the metric has additional isometry \(t_i\rightarrow t_i+\mathrm {const.}\) Introducing the variables \(x=t_1-t_2\) and \(y=t_1+t_2\) and shifting y to remove the cross-term, we bring the metric to diagonal form:

$$\begin{aligned} k_{ij}\,dt_i\,dt_j=\frac{2\,dx^2}{\upchi _1+\upchi _2}+{k\over 2}\,(\upchi _1+\upchi _2)\,dy^2,\quad k(x)={\upkappa \over \upchi _1+\upchi _2+\upkappa \,\upchi _1\upchi _2}. \end{aligned}$$
(A.14)

Denoting \(\upchi _1+\upchi _2:=c(x)\), we calculate the Ricci scalar \(2R=-c''-{k''\over k}\,c-{3\over 2}\,{k'\over k}\,c'+{c\over 2}\,{(k')^2\over k^2}\). One also has a simple formula for the Christoffel symbols with all upper indices: \(\Gamma ^{p|mn}=-{1\over 8} \upchi _n\,\partial _n\upchi _m\,\delta ^{pm}-{1\over 8} \upchi _m\,\partial _m\upchi _n\,\delta ^{pn}+{1\over 8}\,\upchi _p\,\partial _p\upchi _n\,\delta ^{mn}\). The rest is a matter of direct calculation, which shows that the Ricci flow Eq. (3.23) is satisfied. It would be interesting to find a proper geometrical framework that would facilitate such calculations.

Computing the \(\beta \)-Function in the Inhomogeneous Gauge

In this appendix we consider the renormalization of the theory (4.14) at one loop. We wish to show that all divergences may be reabsorbed in a redefinition of the deformation parameter s entering \(\upalpha , \upbeta , \upgamma \). The quartic vertices in (4.14) look exactly as in (2.9), but with the summation restricted to \(1 \ldots n-1\) and the truncated r-matrix of the form

$$\begin{aligned} {\widetilde{r}}_{ij}^{kl}=a_{ij}\delta _i^k \delta _j^l+\upgamma \,\delta _{ij}\delta ^{kl},\quad i, j, k, l=1\ldots n-1. \end{aligned}$$
(B.1)

One contribution to the \(\beta \)-function comes from exactly the same diagrams as before, shown in Fig. 4, where in the vertices one replaces \(r\rightarrow {\widetilde{r}}\):

$$\begin{aligned} \left( \beta _{ij}^{kl}\right) _{1}= & {} \left[ \frac{(n-1)s}{(1-s)^2}+(i-j)a_{ij}\right] \,\left( \delta _i^k \delta _j^l-{1\over n-1}\delta _{ij}\delta ^{kl}\right) \nonumber \\{} & {} =\left[ (n-1)\,\upalpha \upbeta +(i-j)a_{ij}\right] \,\delta _i^k \delta _j^l-\upalpha \upbeta \,\delta _{ij}\delta ^{kl}. \end{aligned}$$
(B.2)
Fig. 4
figure 4

Diagrams contributing to \(\left( \beta _{ij}^{kl}\right) _{1}\). Here \(i, j, k, l=1, \ldots , n-1\)

Fig. 5
figure 5

Bubbles in the sextic vertex, contributing to \(\left( \beta _{ij}^{kl}\right) _{2}\)

An additional contribution comes from bubbles in the sextic vertex (see Fig. 5):

$$\begin{aligned} \left( \beta _{ij}^{kl}\right) _{2}=\upalpha \upbeta (n+1)\,\delta _{ij}\delta ^{kl}+\upalpha \upbeta \delta _i^k \delta _j^l. \end{aligned}$$
(B.3)

Summing the two contributions, we obtain

$$\begin{aligned} \beta _{ij}^{kl}=\left[ n\,\upalpha \upbeta +(i-j)a_{ij}\right] \,\delta _i^k \delta _j^l+n\upalpha \upbeta \,\delta _{ij}\delta ^{kl} \end{aligned}$$
(B.4)

As explained in Sect. 3.2, the part proportional to \((i-j)a_{ij}\) can be removed by a redefinition of coordinates, and the remaining terms give rise to the Ricci flow equation \({\dot{\upgamma }}=n\upalpha \upbeta \) that is solved by \(s=e^{n\tau }\).

Fig. 6
figure 6

A bubble in the quartic vertex, contributing to \(\upbeta \,|V_k|^2\) terms

As another illustration let us consider the renormalization of the term \(\upbeta \,|V_k|^2\) in (4.14). The divergent graphs that contribute to the \(\beta \)-function of this ‘vertex’ come from bubbles in the quartic vertex, see Fig. 6:

$$\begin{aligned} \beta _{i}^j=\delta _{i}^j\,\upbeta \,\sum \limits _{k=1}^{n-1}\,{\widetilde{r}}_{\,ki}^{\,kj}=\delta _{i}^j\,\upbeta \,\left( \sum \limits _{k=1}^{n-1}\,a_{kj}+\upgamma \right) =\delta _{i}^j\,\upbeta \,((n-j)+\upalpha \, n) \end{aligned}$$
(B.5)

The first term in brackets is again cancelled by a redefinition of variables. We showed in Sect. 3.2 that the U-variables evolve with the Ricci flow as \(|U_j|^2=e^{j\,\tau }\). However, since we have chosen a gauge \(U_n=1\), we have to perform a compensating gauge transformation, and as a result the evolution of the variables is given by \(|U_j|^2=e^{(j-n)\,\tau }\). The V-variables evolve as inverses of these, \(|V_j|^2=e^{-(j-n)\,\tau }\), and this eliminates the unwanted term in (B.5). The remaining flow equation is \({\dot{\upbeta }}=n\upalpha \upbeta \), which is again solved by \(s=e^{n\tau }\). A similar analysis could be performed for other vertices in the Lagrangian (4.14).

The Deformed Flag Manifold \(\sigma \)-Model in Geometric Form

We start this appendix by showing that there is a GLSM formulation of the flag manifold \(\sigma \)-model based on \(n\times n\)-matrices rather than on \(m\times n\) matrices as in Sect. 4.2. The quotient (4.20) suggests the following GLSM representation for a deformed flag manifold \(\sigma \)-model:

$$\begin{aligned} \mathscr {L}=\textrm{Tr}\left( V {\overline{D}} U\right) -\textrm{Tr}\left( V {\overline{D}} U\right) ^\dagger +\textrm{Tr}\left( r_s^{\textrm{T}}(U V) (U V)^\dagger \right) , \end{aligned}$$
(C.1)

where, as opposed to (4.21), \(U, V\in \textrm{End}(\mathbb {C}^n)\) are square matrices. The covariant derivatives are defined formally as before: \({\overline{D}}U={\overline{\partial }}U+i \,U \overline{{\mathcal {A}}}\), but now one should keep in mind that the gauge field takes values in \(\mathfrak {p}=\textrm{Lie}(P_{m_1, \ldots , m_{s}})\subset {\mathfrak {g}}{\mathfrak {l}}_n\). As we now explain, one can simplify this gauged theory to arrive at the Lagrangian (4.21) studied before, if one assumes that U is non-degenerate, i.e. \(U\in GL(n, \mathbb {C})\). Taking the variation of (C.1) w.r.t. \(\overline{{\mathcal {A}}}\), we find \(VU|_{\mathfrak {p}^\vee }=0\) (here we view \(VU\in {\mathfrak {g}}{\mathfrak {l}}_n^\vee \)). In general this is a rather non-trivial condition, however already from the requirement of vanishing of the first \(n_s\) rows of the matrix VU we find that the first \(n_s\) rows of V are orthogonal to all columns of U. Since U is non-generate, it follows that \(V_{ik}=0\) for \(i=1, \ldots , n_s\). Therefore we can safely truncate the matrix V by erasing its first \(n_s\) rows, forming a reduced matrix \({\widetilde{V}}\in \textrm{Hom}(\mathbb {C}^n, \mathbb {C}^m)\) with \(m=\sum \limits _{i=1}^{s-1}\,n_i\). It is then easy to see that the first \(n_s\) columns of the matrix U do not enter the Lagrangian (C.1) either, so we may analogously truncate the matrix U, forming a new matrix \({\widetilde{U}}\in \textrm{Hom}(\mathbb {C}^m, \mathbb {C}^n)\). We may then replace the fields U and V in the Lagrangian (C.1) by \({\widetilde{U}}, {\widetilde{V}}\) and truncate the gauge fields in the covariant derivatives accordingly. If one drops the tildes, one recognizes the formula (4.21).

Our next goal is to derive the metric form of the deformed flag manifold model (4.21), where U and V are \(m\times n\)-matrices. If necessary, one can always return from (4.21) to (C.1) by setting \(m=n\), so in a sense the \(m\ne n\) notation is more general. Our goal in this section is to eliminate the variables \(V, {\overline{V}}\), which enter the Lagrangian quadratically. The relevant equations to be solved are (2.12)–(2.13) with \({\hat{r}}=r\). To write down the solution, we introduce the projector \(\Pi _U=U({\overline{U}} U)^{-1}{\overline{U}}\) on \(\textrm{Im}(U)\) (indeed \(\Pi _U(UY)=UY\)). Then

$$\begin{aligned} V={1\over {\overline{U}} U} \,{\overline{U}} \left[ \frac{1}{\frac{1}{2}\,\frac{1+s}{1-s}\,\textrm{Id}+\Pi _U\,{i\over 2}\,{\mathcal {R}}}\right] \, \left( U {1\over {\overline{U}} U} D{\overline{U}} \right) . \end{aligned}$$
(C.2)

Using \(\Pi _U U=U\) and \(\overline{U} \Pi _U= \overline{U}\), it is easy to prove that in the denominator one can freely commute \(\Pi _U\) with \({\mathcal {R}}\). One shows that (2.13) is satisfied by noting that \({\overline{U}}\,r_s\,\Pi _U={\overline{U}} \left( \frac{1}{2}\,\frac{1+s}{1-s}\,\textrm{Id}+{i\over 2}\,{\mathcal {R}} \,\Pi _U\, \right) \) and using this commutation property.

The formula (C.2) for V can be considerably simplified by using part of the complex gauge symmetry. We will call this procedure ‘passing to the unitary frame’, for the following reason. We may write \(U={\hat{U}}{\hat{b}}\), where \(\overline{{\hat{U}}} {\hat{U}}=\mathbb {1}_m\) and \({\hat{b}}\in B\subset P_{m_1, \ldots , m_s}\) is a lower-triangular matrix (B is the Borel subgroup). This decomposition is simply the Gram–Schmidt orthogonalization of the m vectors in U. The remaining gauge group is then \(H=U(n_1)\times \cdots \times U(n_s)\), which is of course the denominator of the quotient (4.17). The solution for V clearly simplifies in the unitary gauge \({\overline{U}} U=\mathbb {1}_m\). Using this gauge and inserting V back into the Lagrangian, we find

$$\begin{aligned} \mathscr {L}\sim \textrm{Tr}\left( \left( U D{\overline{U}} \right) ^\dagger \left[ \frac{1}{\frac{1}{2}\,\frac{1+s}{1-s}\,\textrm{Id}+U{\overline{U}}\,{i\over 2}\,{\mathcal {R}}}\right] \, \left( U D{\overline{U}} \right) \right) \end{aligned}$$
(C.3)

An additional simplification occurs, if we use the formulation of the model where U is an \(n\times n\)-matrix. In that case the unitary gauge implies that U is unitary, so that, apart from \({\overline{U}} U=\mathbb {1}_n\) we have \(U {\overline{U}}=\mathbb {1}_n\), and as a result one can forget the projector in the denominator in (C.3), arriving at the simple formula

figure k

Of course in this case the complications are absorbed in the enlarged gauge field \({\mathcal {A}}\).

Rights and permissions

Springer Nature or its licensor 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

Bykov, D. Quantum Flag Manifold \(\sigma \)-Models and Hermitian Ricci Flow. Commun. Math. Phys. 401, 1–32 (2023). https://doi.org/10.1007/s00220-022-04532-5

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-022-04532-5

Navigation