Skip to main content
Log in

BV and BFV for the H-Twisted Poisson Sigma Model

  • Original Paper
  • Published:
Annales Henri Poincaré Aims and scope Submit manuscript

A Correction to this article was published on 25 February 2021

This article has been updated

Abstract

We present the BFV and the BV extension of the Poisson sigma model (PSM) twisted by a closed 3-form H. There exist superfield versions of these functionals such as for the PSM and, more generally, for the AKSZ sigma models. However, in contrast to those theories, they depend on the Euler vector field of the source manifold and contain terms mixing data from the source and the target manifold. Using an auxiliary connection \(\nabla \) on the target manifold M, we obtain alternative, purely geometrical expressions without the use of superfields, which are new also for the ordinary PSM and promise adaptations to other Lie algebroid-based gauge theories: The BV functional, in particular, is the sum of the classical action, the Hamiltonian lift of the (only on-shell nilpotent) BRST differential, and a term quadratic in the antifields which is essentially the basic curvature and measures the compatibility of \(\nabla \) with the Lie algebroid structure on \(T^*M\). We finally construct a \(\hbox {Diff}(M)\)-equivariant isomorphism between the two BV formulations.

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.

Similar content being viewed by others

Change history

Notes

  1. As recalled above, it was shown by Ševera–Weinstein that twisted Poisson structures correspond to particular Dirac structures. Dirac structures are Lie algebroids. Since these Dirac structures inside \(TM \oplus T^*M\) are projectable to \(T ^*M\), the Lie algebroid structure can be transferred to \(T ^*M\).

  2. Certainly, in the formula for the gauge transformations which contain the corresponding connection coefficients, one can replace them by connection coefficients of a torsionless connection plus a contorsion contribution and view the latter term as an independent one to the gauge transformations. However, not only this is also unnatural, but it renders the calculations more complicated.

  3. We are grateful to A. Hancharuk for making us aware of a mistake in a first version of this formula.

  4. In [20, 26], the focus was put more on a Lie derivative version of the gauge symmetries, mentioning the connection version only as a side remark, but the considerations are very similar and one of the two versions is needed for consistency, as recalled also in the paragraph below. The Dirac sigma model [21] provides an example of a Lie algebroid theory, where the gauge symmetries of this form are not sufficient, requiring a second connection on E and an additional term proportional to \(*F^i\) in the parametrization of the infinitesimal symmetries. See also the more recent development about Dirac sigma models as a universal gauge theory in two dimensions and geometrical interpretations of the two connections: [42,43,44,45,46,47].

  5. T.S. is grateful to S. Fischer for clarifying discussions in this context.—See also [21].

  6. Our conventions are such that a covariant derivative of a vector field v along \(\partial _i\) has components \(\nabla _i v^j \equiv v^j_{;i} = v^j_{,i} + \Gamma ^j_{ki} v^k\).

  7. Also (2.13) has covariance problems, but \(F^\nabla _i := F_i + \Gamma _{ij}^kA_k \wedge F^j\) results into \(F^\nabla = \mathrm {D} A - \tfrac{1}{2}T(A,A).\)

  8. \(\underline{\mathrm {d}x^i} \in X^* T^*M\) must not be confused with \(X^* \mathrm {d}x^i \equiv \mathrm {d}X^i \in \Omega ^1(S^1)\).

  9. Indices between parentheses and square brackets are antisymmetrized and symmetrized over, respectively. Thus, for example, \(v_{(ij)}= \tfrac{1}{2} (v_{ij} + v_{ji})\) and \(v_{[ij]}= \tfrac{1}{2} (v_{ij} - v_{ji})\).

  10. We suppress arguments in these formulas, but one may imagine that \(\phi \), \(\psi \), and \(\xi \) depend on \(\sigma \), \(\sigma ^{\prime }\), and \(\sigma ^{\prime \prime }\), respectively.

  11. The polynomial degree of the antifields, used here to split terms, must not be confused with the antighost number used in the literature (also called antifield number sometimes): Conventionally, the ghost number is understood as being the total degree of an additional bigrading, composed of the pure ghost number pgh and the antighost number agh. By definition, \(\mathrm {gh} = \mathrm {pgh} - \mathrm {agh}\). While \(\mathrm {agh}(X^+)=\mathrm {agh}(A^+)=1\), one agrees on \(\mathrm {agh}(c^+)=2\) and the last term in (4.27), now inside \(S_{\mathrm{BV}}^{(1)}\)—see (4.30)—has antighost number two. But while the BV bracket is of degree minus one in the degree of BV momenta (the polynomial degree of the antifields), it is not homogeneous with respect to the antighost number. Thus, the counting in the main text above seems more useful for our purposes.

  12. Here, we add the superscript \(\nabla \) since it depends on a connection and is a global result. The traditional expressions, resulting from this functional by the choice of connection coefficients with vanishing symmetric part within a local patch on M, will be simply denoted by \(S_{\mathrm{BV}}\). Somewhat miraculously, these expressions will turn out to be globally well defined as well, but with a different lift of the coordinate transformations to the field space.

  13. Note the analogy of (5.7) and (5.8) with its Hamiltonian counterpart, (3.11) and (3.12).

  14. We will clarify the relative signs in these formulas after explanations about the conventions. In this context, the new notation with underlining the fields will become clearer as well; it is in part due to different conventions about their commutativity properties.

  15. Note that fdeg now counts the polynomial degree of \(\theta \) for functions on \(T[1]\Sigma \), not to be confused with a form degree of graded differential forms on this supermanifold. In particular, on \(T[1]\Sigma \) one now has \(\mathrm {deg}(\theta ^\mu ) = \mathrm {fdeg}(\theta ^\mu ) = 1\) but \(\mathrm {deg}(\mathrm {d}\sigma ^\mu ) = \mathrm {fdeg}(\mathrm {d}\sigma ^\mu ) = 0\) where \(\mathrm {d}\sigma ^\mu \in \Omega ^1(T[1]\Sigma )\). It is thus also important that we change the notation consistently from the previous sections to the present one: What previously was a differential form on \(\Sigma \) like \(\alpha = \mathrm {d}\sigma ^\mu \, \alpha _\mu \) will now be underlined and written as the function \(\underline{\alpha } = \theta ^\mu \, \alpha _\mu (\sigma )\), to avoid any confusion of the above sort.

  16. As before, one may imagine that \({\underline{\phi }}{}\), \({\underline{\psi }}{}\), and \({\underline{\xi }}{}\) depend on \((\sigma , \theta )\), \((\sigma ^{\prime }, \theta ^{\prime })\), and \((\sigma ^{\prime \prime }, \theta ^{\prime \prime })\), respectively.

  17. By this notation, we mean that when replacing \({{\underline{A}}}{}\), \({{\underline{X}}^+}{}\), and \({{\underline{c}}^+}{}\) by the expressions inverse to Eqs. (5.40) and (5.42) in \(S_{\mathrm{BV}}\!\left[ X, {{\underline{A}}}{}, {{\underline{c}}}{}, {{\underline{X}}^+}{},{{\underline{A}}^+}{}, {{\underline{c}}^+}{}\right] \), we obtain the expression given in this formula and denoted by \(S_{\mathrm{BV}}\!\left[ X, {{\underline{A}}}{}^{{\nabla }}, {{\underline{c}}}{}\,, {{\underline{X}}^+}{}^{\!{\nabla }}\!,{{\underline{A}}^+}{}, {{\underline{c}}^+}{}^{{\nabla }} \right] \); in other words, it describes the same abstract functional \(S_{\mathrm{BV}}\), but written in different fields (coordinates), which are specified behind it in the squared brackets.

  18. Using the superfield expansions (5.9) and (5.10), one ends up with an apparently different sign for the last term. However, this then is still a function integrated over \(T[1]\Sigma \); transcribing this as an integral over a differential form on \(\Sigma \), this term receives a change in its sign, as explained in detail in Sect. 5.2 (see in particular Eq. (5.13)).

  19. This transformation is analogous to Eq. (5.5) in [55] and the proof of the statements below follow by the same steps as those provided there.

  20. The tensor S for a general Lie algebroid E appeared first in [29]. Its geometrical meaning is to measure the violation of the compatibility [28] of a connection \(\nabla \) on E with the Lie algebroid structure [31]—but see also [30], where this tensor was termed “basic curvature.”

References

  1. Ikeda, N., Izawa, K.I.: General form of Dilaton gravity and nonlinear gauge theory. Prog. Theor. Phys. 90(1):237–245 (1993). https://doi.org/:10.1143/PTP.90.237

  2. Ikeda, N.: Two-dimensional gravity and nonlinear gauge theory. Ann. Phys. 235, 435 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Schaller, P., Strobl, T.: Poisson structure induced (topological) field theories. Mod. Phys. Lett. A 9, 3129 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Schaller, P.,T. Strobl, T.: A brief introduction to Poisson sigma models. Lect. Notes Phys. 469, 321–333 (1996)

  5. Batalin, I.A., Vilkovisky, G.A.: Gauge algebra and quantization. Phys. Lett. B 102, 27 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  6. Batalin, I.A., Vilkovisky, G.A.: Quantization of gauge theories with linearly dependent generators. Phys. Rev. D 28, 2567 (1983). (Erratum-ibid. D 30 (1984) 508)

    Article  ADS  MathSciNet  Google Scholar 

  7. Henneaux, M., Teitelboim, C.: Quantization of Gauge Systems. Princeton University Press, Princeton (1991)

    MATH  Google Scholar 

  8. Alexandrov, M., Kontsevich, M., Schwartz, A., Zaboronsky, O.: The geometry of the master equation and topological quantum field theory. Int. J. Mod. Phys. A 12, 1405 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Cattaneo, A.S., Felder, G.: A path integral approach to the Kontsevich quantization formula. Commun. Math. Phys. 212, 591 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Cattaneo, A.S., Felder, G.: On the AKSZ formulation of the Poisson sigma model. Lett. Math. Phys. 56, 163 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  11. Roytenberg, D.: AKSZ-BV formalism and courant algebroid-induced topological field theories. Lett. Math. Phys. 79, 143 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  12. Ikeda, N.: Lectures on AKSZ sigma models for physicists. In: Noncommutative Geometry and Physics 4, Workshop on Strings, Membranes and Topological Field Theory, pp. 79–169. World Scientific, Singapore (2017)

  13. Klimcik, C., Strobl, T.: WZW-Poisson manifolds. J. Geom. Phys. 43, 341 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Park, J.-S.: Topological open p-branes. In: Symplectic Geometry and Mirror Symmetry (Seoul, 2000), pp. 311–384. World Scientific Publishing, River Edge (2001)

  15. Ševera, P., Weinstein, A.: Poisson geometry with a 3-form background. Prog. Theor. Phys. Suppl. 144, 145 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Ikeda, N., Xu, X.: Canonical functions, differential graded symplectic pairs in supergeometry, and Alexandrov–Kontsevich–Schwartz–Zaboronsky sigma models with boundaries. J. Math. Phys. 55, 113505 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Batalin, I.A., Vilkovisky, G.A.: Relativistic S matrix of dynamical systems with boson and fermion constraints. Phys. Lett. 69B, 309 (1977)

    Article  ADS  Google Scholar 

  18. Batalin, I.A., Fradkin, E.S.: A generalized canonical formalism and quantization of reducible gauge theories. Phys. Lett. 122B, 157 (1983)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Ikeda, N., Strobl, T.: On the relation of Lie algebroids to constrained systems and their BV/BFV formulation. Ann. H. Poincare 20(2), 527 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  20. Bojowald, M., Kotov, A., Strobl, T.: Lie algebroid morphisms, Poisson sigma models, and off-shell closed gauge symmetries. J. Geom. Phys. 54, 400 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Kotov, A., Schaller, P., Strobl, T.: Dirac sigma models. Commun. Math. Phys. 260, 455 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Baulieu, L., Losev, A.S., Nekrasov, N.A.: Target space symmetries in topological theories I. JHEP 2002, 021 (2002)

    Article  ADS  MathSciNet  Google Scholar 

  23. Cattaneo, A.S., Felder, G.: On the globalization of Kontsevich’s star product and the perturbative Poisson sigma model. Prog. Theor. Phys. Suppl. 144, 38 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  24. Cattaneo, A.S., Moshayedi, N., Wernli, K.: Globalization for Perturbative quantization of nonlinear split AKSZ sigma models on manifolds with boundary. Commun. Math. Phys. 372(1), 213 (2019)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Cattaneo, A.S., Moshayedi, N., Wernli, K.: On the globalization of the Poisson sigma model in the BV–BFV formalism. arXiv:1808.01832 [math-ph]

  26. Strobl, T.: Algebroid Yang–Mills theories. Phys. Rev. Lett. 93, 211601 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  27. Zucchini, R.: Algebraic formulation of higher gauge theory. J. Math. Phys. 58(6), 062301 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Blaom, A.D.: Geometric structures as deformed infinitesimal symmetries. Trans. Am. Math. Soc. 358, 3651 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  29. Mayer, C., Strobl, T.: Lie algebroid Yang–Mills with matter fields. J. Geom. Phys. 59, 1613 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. Abad, C.A., Crainic, M.: Representations up to homotopy of Lie algebroids. J. Reine Angew. Math. (Crelles J.) 663, 91 (2011)

    MathSciNet  MATH  Google Scholar 

  31. Kotov, A., Strobl, T.: Lie algebroids, gauge theories, and compatible geometrical structures. Rev. Math. Phys. 31(04), 1950015 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  32. Batalin, I.A., Bering, K., Damgaard, P.H.: Superfield quantization. Nucl. Phys. B 515, 455 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. Batalin, I.A., Bering, K., Damgaard, P.H.: Superfield formulation of the phase space path integral. Phys. Lett. B 446, 175 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  34. Grigoriev, M.A., Damgaard, P.H.: Superfield BRST charge and the master action. Phys. Lett. B 474, 323 (2000)

    Article  ADS  Google Scholar 

  35. Grigoriev, M.A.: Parent formulations, frame-like Lagrangians, and generalized auxiliary fields. JHEP 1212, 048 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. Kotov, A., Salnikov, V., Strobl, T.: 2D gauge theories and generalized geometry. JHEP 1408, 021 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Koszul, J.-L.: Crochet de Schouten-Nijenhuis et cohomologie. Astérisque Hors Série 137, 257–271 (1985)

    MathSciNet  MATH  Google Scholar 

  38. Grützmann, M., Strobl, T.: General Yang–Mills type gauge theories for \(p\)-form gauge fields: From physics-based ideas to a mathematical framework or From Bianchi identities to twisted Courant algebroids. Int. J. Geom. Methods Mod. Phys. 12, 1550009 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  39. Kotov, A., Strobl, T.: Characteristic classes associated to Q-bundles. Int. J. Geom. Methods Mod. Phys. 12(01), 1550006 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  40. Carow-Watamura, U., Heller, M.A., Ikeda, N., Kaneko, Y., Watamura, S.: Higher gauge theories from Lie n-algebras and off-shell covariantization. JHEP 1607, 125 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. Carow-Watamura, U., Heller, M.A., Ikeda, N., Kaneko, T., Watamura, S.: Off-shell covariantization of algebroid gauge theories. PTEP 2017(8), 083B01 (2017)

    MathSciNet  MATH  Google Scholar 

  42. Wright, K.: Lie algebroid gauging of non-linear sigma models. J. Geom. Phys. 146, 103490 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  43. Chatzistavrakidis, A., Deser, A., Jonke, L., Strobl, T.: Beyond the standard gauging: gauge symmetries of Dirac sigma models. JHEP 1608, 172 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Chatzistavrakidis, A., Deser, A., Jonke, L., Strobl, T.: Strings in singular space-times and their universal gauge theory. Ann. H. Poincare 18(8), 2641 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  45. Chatzistavrakidis, A., Deser, A., Jonke, L., Strobl, T.: Gauging as constraining: the universal generalised geometry action in two dimensions. PoS CORFU 2016, 087 (2017)

    Google Scholar 

  46. Ševera, P., Strobl, T.: Transverse generalized metrics and 2d sigma models. J. Geom. Phys. 146, 103509 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  47. Ikeda, N.: Momentum sections in Hamiltonian mechanics and sigma models. SIGMA 15, 076 (2019)

    MathSciNet  MATH  Google Scholar 

  48. Dirac, P.A.M.: Generalized Hamiltonian dynamics. Can. J. Math. 2, 129 (1950)

    Article  MathSciNet  MATH  Google Scholar 

  49. Faddeev, L.D., Jackiw, R.: Hamiltonian reduction of unconstrained and constrained systems. Phys. Rev. Lett. 60, 1692 (1988)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  50. Alekseev, A., Strobl, T.: Current algebras and differential geometry. JHEP 0503, 035 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  51. Kochan, D., Ševera, P.: Differential gorms, differential worms. In: Mathematical Physics, vol. 128 (2005), Proceedings of the XI Regional Conference, Tehran, Iran, 3–6 May (2004)

  52. Vaintrob, A.: Lie algebroids and homological vector fields. Uspekhi Mat. Nauk 52(2), 161–162 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  53. Cattaneo, A.S., Cotta-Ramusino, P., Rossi, C.A.: Loop observables for BF theories in any dimension and the cohomology of knots. Lett. Math. Phys. 51, 301 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  54. Ikeda, N.: Deformation of BF theories, topological open membrane and a generalization of the star deformation. JHEP 0107, 037 (2001)

    Article  ADS  MathSciNet  Google Scholar 

  55. Ikeda, N., Strobl, T.: From BFV to BV and spacetime covariance. arXiv:2007.15912 [hep-th]

  56. Heller, M.A., Ikeda, N., Watamura, S.: Unified picture of non-geometric fluxes and T-duality in double field theory via graded symplectic manifolds. JHEP 1702, 078 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

N.I. is grateful to the Max Plank institute for mathematics, the university of Geneva, and the Prague Charles university, where part of this work was carried out during 2017 and 2018, for their hospitality. T.S. is grateful to Satoshi Watamura for his invitation to Sendai in 2016, where this collaboration all began. We are grateful to the Erwin Schrödinger International Institute for Mathematics and Physics for support within the program “Higher Structure and Field Theory”. This work was supported by the LABEX MILYON (ANR-10-LABX-0070) of Université de Lyon, within the program “Investissements d’Avenir” (ANR-11-IDEX-0007) operated by the French National Research Agency (ANR).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas Strobl.

Additional information

Communicated by Boris Pioline.

Publisher's Note

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

Appendices

BV and BFV for the Poisson Sigma Model

The classical action of the Poisson sigma model [2, 3] results from putting to zero the 3-form H in the twisted generalization (2.1) studied in the present paper:

$$\begin{aligned} S = \int _{\Sigma } \left( A_i \wedge \mathrm {d}X{}^i + \tfrac{1}{2} \pi ^{ij}(X{}) A_i \wedge A_j \right) . \end{aligned}$$
(A.1)

It agrees with (the classical part of) the AKSZ model [8] in two dimensions [10]. In this appendix, we recall the BV formulation [9] of the action (A.1)—which is the starting point in the AKSZ approach—as well as the BFV formulation of the PSM.

1.1 Recollection of BV for the PSM

The BV symplectic form is

$$\begin{aligned} \omega _{\mathrm{BV}}= & {} \int _{T[1]\Sigma } d^2r\sigma d^2\theta \, \delta \varvec{X}^i \delta \varvec{A}_i, \end{aligned}$$
(A.2)

where \(\varvec{X}^i\) and \(\varvec{A}_i\) can be viewed as unconstrained superfields on \(T[1]\Sigma \) of total degree zero and one, respectively, see (5.9) and (5.10) in the main text.

The BV action takes the form of the classical action (A.1), written in terms of the superfields:

$$\begin{aligned} S_{\mathrm{BV}} = \int _{T[1]\Sigma } d^2\sigma d^2\theta \, \left( \varvec{A}_i \varvec{\mathrm {d}}\varvec{X}^i + \tfrac{1}{2} \pi ^{ij}(\varvec{X}) \varvec{A}_i \varvec{A}_j \right) . \end{aligned}$$
(A.3)

Expanded in terms of the original fields, the ghost fields, and their antifields, this becomes:Footnote 18

$$\begin{aligned} S_{\mathrm{BV}}= & {} \!\!\! \int _{\Sigma } \biggl (A_i \wedge \mathrm {d}X{}^i + \tfrac{1}{2} \pi ^{ij}(X{}) A_i \wedge A_j - \pi ^{ij}(X{}) X{}^{+}_i c_j + A^{+ i} \wedge \left( \mathrm {d}c_{i} + \pi ^{jk},_i(X{}) A_j c_k \right) \nonumber \\&+ \tfrac{1}{2} \pi ^{ij},_k(X{}) c^{+k} c_i c_j - \tfrac{1}{4} \pi ^{ij},_{kl}(X{}) A^{+k} \wedge A^{+l} c_i c_j \biggr ). \end{aligned}$$
(A.4)

In the AKSZ formalism, it is evident—by its very construction—that this action satisfies the classical master equation \((S_{\mathrm{BV}},S_{\mathrm{BV}})=0\): Indeed, the AKSZ sigma model can be seen as the Hamiltonian lift of the sum of two commuting differentials to the mapping space \(\underline{\mathrm {Hom}}(T[1] \Sigma ,T^*[1]M)\), one being the de Rham differential on the source space \(T[1]\Sigma \) and the other one the Lie algebroid differential on target space \(T^*[1]M\). Note that this construction cannot be applied as such to the H-twisted Lie algebroid structure on \(T^*[1]M\) since, for \(H\ne 0\), the corresponding nilpotent vector field Q is not compatible with the symplectic form on \(T^*[1]M\); this, however, is needed for the Hamiltonian lift in the AKSZ formalism, see [8, 10,11,12].

1.2 BFV for the PSM

The BFV formulation [3] of the PSM permits a superfield description similar to the one above, too. The BFV symplectic form can be written as

$$\begin{aligned} \omega _{\mathrm{BFV}}= & {} \int _{T[1]S^1} d \sigma d \theta \, \delta {\widetilde{X}}{}^i \wedge \delta {\widetilde{A}}{}_i. \end{aligned}$$
(A.5)

In terms of the expansion (3.8), this gives rise to the following Poisson brackets for the fields and ghosts:

$$\begin{aligned} \{X{}^i(\sigma ) p_{j}(\sigma ^{\prime })\}= & {} \delta ^i{}_j\delta (\sigma - \sigma ^{\prime }), \end{aligned}$$
(A.6)
$$\begin{aligned} \{c_j(\sigma ) b^i(\sigma ^{\prime })\}= & {} \delta ^i{}_j\delta (\sigma - \sigma ^{\prime }). \end{aligned}$$
(A.7)

Likewise, also the Hamiltonian BFV or BRST charge permits a compact description in terms of these superfields:

$$\begin{aligned} S_{\mathrm{BFV}} = \int _{T[1]S^1} d \sigma d \theta \, \left( {\widetilde{A}}{}_i {\mathrm {d}}{\widetilde{X}}{}^i + \tfrac{1}{2} \pi ^{ij}({\widetilde{X}}{}) {\widetilde{A}}{}_i {\widetilde{A}}{}_j \right) , \end{aligned}$$
(A.8)

which, when decomposed into its components takes the form:

$$\begin{aligned} S_{\mathrm{BFV}}= & {} \int _{S^1} d \sigma \left( c_i \partial X{}^i + \pi ^{ij}(X{}) c_i p_{j} + \tfrac{1}{2} \pi ^{ij},_k(X{}) b^k c_i c_j \right) . \end{aligned}$$
(A.9)

It satisfies \(\{S_{\mathrm{BFV}}, S_{\mathrm{BFV}}\} =0\). The BFV symplectic form and the BFV-BRST charge are just a one-dimensional superextension of the classical symplectic form (on the cotangent space of loop space) and the classical action, respectively.

Naive Generalization of AKSZ Does not Work for the HPSM

We already explained at the end of Sect. A.1, why—at least in its direct, unmodified form—the AKSZ procedure cannot be applied to the HPSM for non-vanishing H; in fact, the PSM is the most general AKSZ sigma model for a two-dimensional choice of \(\Sigma \), see, e.g., [11]. One may still hope that there may be some elegant superfield formalism that yields the BV theory of the HPSM without much explicit work. We want to show here that, at least in the most direct way, this is not the case.

Rewriting the classical symplectic form (2.27) of the HPSM and its classical action (2.1) in terms of superfields on \(T[1]\Sigma \) as before, one would arrive at a BV symplectic form

$$\begin{aligned} \omega _{\mathrm{BV}}= & {} \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \, \left( \delta \varvec{X}^i \delta \varvec{A}_i + \tfrac{1}{2} H_{ijk}(\varvec{X}) \varvec{\mathrm {d}}\varvec{X}^k \delta \varvec{X}^i \delta \varvec{X}^j \right) , \end{aligned}$$
(B.1)

and a BV action

$$\begin{aligned} S_{\mathrm{BV}}= & {} \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \ \left( \varvec{A}_i \varvec{\mathrm {d}}\varvec{X}^i + \tfrac{1}{2} \pi ^{ij}(\varvec{X}) \varvec{A}_i \varvec{A}_j \right) \nonumber \\&+ \int _{T[1]N} d^3 \sigma d^3 \theta \ \tfrac{1}{3!} H_{ijk}(\varvec{X}) \varvec{\mathrm {d}}\varvec{X}^i \varvec{\mathrm {d}}\varvec{X}^j \varvec{\mathrm {d}}\varvec{X}^k. \end{aligned}$$
(B.2)

However, this naive candidate of the BV action does not satisfy the classical master equation. In fact, with the induced BV brackets

$$\begin{aligned} \left( \varvec{X}^i(\sigma , \theta ),\varvec{A}_j(\sigma ^{\prime }, \theta ^{\prime })\right)= & {} \delta ^i{}_j \delta ^2(\sigma - \sigma ^{\prime }) \delta ^2(\theta - \theta ^{\prime }), \end{aligned}$$
(B.3)
$$\begin{aligned} \left( \varvec{A}_i(\sigma , \theta ),\varvec{A}_j(\sigma ^{\prime }, \theta ^{\prime })\right)= & {} - H_{ijk}(\varvec{X}) \varvec{\mathrm {d}}\varvec{X}^k \delta ^2(\sigma - \sigma ^{\prime }) \delta ^2(\theta - \theta ^{\prime }), \end{aligned}$$
(B.4)

one obtains

$$\begin{aligned} \left( S_{\mathrm{BV}},S_{\mathrm{BV}}\right)= & {} \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \ \left( \pi ^{kl} H_{ijl} \varvec{\mathrm {d}}\varvec{X}^i \varvec{\mathrm {d}}\varvec{X}^j \varvec{A}_k - \pi ^{jm} \pi ^{kn} H_{imn} \varvec{\mathrm {d}}\varvec{X}^i \varvec{A}_j \varvec{A}_k \right. \nonumber \\&\left. + \tfrac{1}{3} \pi ^{il} \pi ^{jm} \pi ^{kn} H_{lmn} \varvec{A}_i \varvec{A}_j \varvec{A}_k \right) \end{aligned}$$
(B.5)
$$\begin{aligned}= & {} - \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \ \tfrac{1}{3} H_{ijk} ( \varvec{\mathrm {d}}\varvec{X}^i + \pi ^{il} \varvec{A}_l)( \varvec{\mathrm {d}}\varvec{X}^j + \pi ^{jm} \varvec{A}_m) ( \varvec{\mathrm {d}}\varvec{X}^k + \pi ^{kn} \varvec{A}_n).\nonumber \\ \end{aligned}$$
(B.6)

This is zero only for \(H = 0\). Note that the second expression shows that the master equation holds true on-shell (in some sense), i.e., upon usage of the obvious superfield extension of the field Eq. (2.2). A similar statement also follows if one drops the H contribution to (B.1), taking \(\omega _{\mathrm{BV}}\) to have Darboux form, the choice made in [14]. The on-shell validity of the master equation is, however, not sufficient for the BV formulation. In fact, this is precisely the reason imposing the use of the more elaborate BV formalism over its simpler BRST version (see also the discussion around (4.10)).

Let us try a more general ansatz. Suppose that one is not sure of how to distribute the 3-form H between the BV symplectic form and the BV action (see also “Appendix C” for a related question). We thus introduce two parameters \(\alpha ,\beta \in {{\mathbb {R}}}\) and attempt to use

$$\begin{aligned} \omega _{\mathrm{BV}}= & {} \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \, \left( \delta \varvec{X}^i \delta \varvec{A}_i + \tfrac{\alpha }{2} H_{ijk}(\varvec{X}) \varvec{\mathrm {d}}\varvec{X}^k \delta \varvec{X}^i \delta \varvec{X}^j \right) , \end{aligned}$$
(B.7)

as the BV symplectic form and

$$\begin{aligned} S_{\mathrm{BV}}&= \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \ \left( \varvec{A}_i \varvec{\mathrm {d}}\varvec{X}^i + \tfrac{1}{2} \pi ^{ij}(\varvec{X}) \varvec{A}_i \varvec{A}_j + \tfrac{\beta }{2} B_{ij}(\varvec{X}) \varvec{\mathrm {d}}\varvec{X}^i \varvec{\mathrm {d}}\varvec{X}^j \right) , \end{aligned}$$
(B.8)

as the BV action. Here, we assumed, for simplicity, that \(H = \mathrm {d}B\). The previous ansatz is reproduced for \(\alpha =\beta =1\), the considerations of [14] correspond to \(\alpha =0\) and \(\beta =1\), but one might still hope that for some other choice of parameters there is some decisive cancellation. However, a direct calculation yields

$$\begin{aligned} \left( S_{\mathrm{BV}},S_{BV}\right)= & {} \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \ \left( \pi ^{ij},_l \pi ^{lk} \varvec{A}_i \varvec{A}_j \varvec{A}_k \right. \nonumber \\&\left. - \alpha \pi ^{jm} \pi ^{kn} H_{imn} \varvec{\mathrm {d}}\varvec{X}^i \varvec{A}_j \varvec{A}_k + (2 \alpha - \beta ) \pi ^{kl} H_{ijl} \varvec{\mathrm {d}}\varvec{X}^i \varvec{\mathrm {d}}\varvec{X}^j \varvec{A}_k \right) .\nonumber \\ \end{aligned}$$
(B.9)

This calculation holds true for every choice of the bivector field \(\pi \) and the 3-form \(H=\mathrm {d}B\), not necessarily restricted by means of the twisted Poisson condition (1.1)—which was used to obtain (B.5B.6), to which it then reduces for the initial choice of constants.

We see that, for \(H \ne 0\), (B.9) vanishes if and only if \(\alpha = 0\), \(\beta = 0\), and \(\pi \) Poisson. But in this case, H drops out from both (B.7) and (B.8) and we only reproduce the BV formulation of the untwisted Poisson sigma model (see “Appendix A.1”).

In terms of superfields as above and for the case that one writes the BV symplectic form in Darboux coordinates as in (A.2), the correct BV extension of the HPSM is found to be rather of the involved form (5.20) in the main text. A qualitatively new feature of this functional, which is at least not easy to guess, is the appearance of the Euler vector field \(\varvec{\varepsilon }\) inside \(S_{\mathrm{BV}}\)—a fact that holds true also for the simpler BFV functional \(S_{\mathrm{BFV}}\) of the HPSM, see (3.10).

BV Symplectic form with H

The graded symplectic form \(\omega _{\mathrm{BFV}}\) of the HPSM necessarily contains the 3-form H, see (3.4) or (3.9). This is not the case for the BV symplectic form \(\omega _{\mathrm{BV}}\), see (4.21) or (5.26) (in fact, for degree reasons, \(\omega _{\mathrm{BV}}\) is always exact [11]). One may be curious, however, how the BV functional of the HPSM would look like if one adds an H contribution to the BV symplectic form similar to the one in \(\omega _{\mathrm{BFV}}\). Indeed, in terms of the new fieldFootnote 19

$$\begin{aligned} \varvec{A}_i= & {} \varvec{\check{A}}_{i} + H_{ijk}(\varvec{X}) (\varepsilon \varvec{X}^j) {\mathrm {d}}\varvec{X}^k, \end{aligned}$$
(C.1)

the BV symplectic form (5.26) changes to

$$\begin{aligned} \omega _{\mathrm{BV}}= & {} \int _{T[1]\Sigma } d^2 \sigma d^2 \theta \, \left( \delta \varvec{X}^{i} \delta \varvec{\check{A}}_i - \tfrac{1}{2} H_{ijk} {\mathrm {d}}\varvec{X}^i \delta \varvec{X}^j \delta \varvec{X}^k \right) . \end{aligned}$$
(C.2)

and the BV action (5.20), becomes

$$\begin{aligned} S_{\mathrm{BV}}= & {} \int _{T[1]\Sigma } \!\!\!\! d^2 \sigma d^2 \theta \, \left[ \varvec{\check{A}}_i \, \varvec{\mathrm {d}}\varvec{X}^i + \tfrac{1}{2} \pi ^{ij}(\varvec{X})\, \varvec{\check{A}}_i \varvec{\check{A}}_{j} \right] + \int _{T[1]N} \!\!\!\! d^3 \sigma d^3 \theta \, H(\varvec{X}) \nonumber \\&+ \int _{T[1]\Sigma } \!\!\!\! d^2 \sigma d^2 \theta \, \left[ \tfrac{1}{4} (\pi ^{il} \pi ^{jm} H_{lmk})(\varvec{X}) \, \varvec{\check{A}}_i \varvec{\check{A}}_j \varvec{\varepsilon } \varvec{X}^k - \tfrac{1}{2} (\pi ^{il} H_{jkl}) (\varvec{X})\, \varvec{\check{A}}_i (\varvec{\mathrm {d}}\varvec{X}^j) \varvec{\varepsilon } \varvec{X}^k \right] \nonumber \\&+ \int _{T[1]\Sigma } \!\!\!\! d^2 \sigma d^2 \theta \, \left[ \tfrac{1}{8} (\pi ^{im} \pi ^{jn} \pi ^{pq} H_{mql}H_{npk})(\varvec{X})\, \varvec{\check{A}}_i \varvec{\check{A}}_j (\varvec{\varepsilon } \varvec{X}^k) \varvec{\varepsilon } \varvec{X}^l \right] . \nonumber \\ \end{aligned}$$
(C.3)

This answers the question posed in this appendix.

Bianchi Identity for the Basic Curvature S

In this appendix, we want to provide a non-sophisticated derivation of the Bianchi identity (4.35) for the tensor S, defined in (4.13).Footnote 20 We do this in terms of a component calculation. In a local coordinate system, the curvature R and the torsion \(\Theta \) of the affine TM connection \(\Gamma _{ij}^k\) as well as the E torsion T and the basic curvature S take the following form:

$$\begin{aligned} R_{lij}^k= & {} \partial _i \Gamma _{lj}^k - \partial _j \Gamma _{li}^k + \Gamma _{lj}^m \Gamma _{mi}^k - \Gamma _{li}^m \Gamma _{mj}^k, \end{aligned}$$
(D.1)
$$\begin{aligned} \Theta ^k_{ij}= & {} - \Gamma _{ij}^k + \Gamma _{ji}^k = \pi ^{kl} H_{ijl}, \end{aligned}$$
(D.2)
$$\begin{aligned} T_k^{ij}= & {} - f_k^{ij} - \pi ^{il} \Gamma _{kl}^j - \pi ^{lj} {\Gamma _{kl}^i}, \end{aligned}$$
(D.3)
$$\begin{aligned} S_{ij}{}^{kl}= & {} \nabla _j T_i^{kl} - \pi ^{mk} R^l_{ijm} + \pi ^{ml} R^k_{ijm} \nonumber \\&\equiv - f_{(i}^{kl}{},_{j)} + \Gamma _{(ij)}^m f_m^{kl} + 2 \Gamma _{m(j}^{[k} f_{i)}^{l]m} + 2 \pi ^{m[k} \Gamma ^{l]}_{ij},_m + 2 \pi ^{m[k},_j \Gamma ^{l]}_{im} - 2 \Gamma ^{[k}_{im}\Gamma ^{l]}_{nj} \pi ^{mn}. \nonumber \\ \end{aligned}$$
(D.4)

The Bianchi identities for \(R_{ijk}^l\) and \(\Theta _{ij}^k\) are

$$\begin{aligned}&R_{lij}^k = - R_{lji}^k, \end{aligned}$$
(D.5)
$$\begin{aligned}&R_{[ijk]}^l = \nabla _{[i} \Theta ^l_{jk]} - \Theta _{[ij}^m \Theta _{k]m}^l, \end{aligned}$$
(D.6)
$$\begin{aligned}&\nabla _{[i} R_{l|jk]}^m + \Theta _{[ij}^n R_{k]ln}^m =0. \end{aligned}$$
(D.7)

A twisted Poisson structure on M gives rise to a Lie algebroid structure on \(T^*M\), which, by means of a connection with the prescribed torsion as in (2.21), can be expressed as follows:

$$\begin{aligned}&[\pi ^{\sharp }(\alpha ), \pi ^{\sharp }(\beta )]^{\nabla } = \pi ^{\sharp }([\alpha , \beta ]_{\pi }^{\nabla }), \end{aligned}$$
(D.8)
$$\begin{aligned}&[[\alpha , \beta ]_{\pi }^{\nabla }, \gamma ]_{\pi }^{\nabla } + \text{ cyclic }(\alpha \beta \gamma ) = 0, \end{aligned}$$
(D.9)

where \(\alpha , \beta , \gamma \in \Omega ^1(M)\). Here, \([X, Y]^{\nabla }\) is the covariantized Lie bracket for vector fields \(X, Y \in \mathfrak {X}(M)\), which is defined by replacing the derivative \(\partial _i\) by the covariant derivative \(\nabla _i\). Similarly, \([\alpha , \beta ]_{\pi }^{\nabla }\) is the covariantized Koszul bracket [56]

$$\begin{aligned}&[\alpha , \beta ]_{\pi }^{\nabla } = L^{\nabla }_{\pi ^{\sharp } (\alpha )}\beta - L^{\nabla }_{\pi ^{\sharp } (\beta )} \alpha - \nabla (\pi (\alpha , \beta )), \end{aligned}$$
(D.10)

where \(L_X = \mathrm{D}\circ \iota _X + \iota _X \circ \mathrm{D}\) is the covariantized Lie derivative (as before, \(\mathrm{D}\) denotes the exterior covariant derivative induced by the connection \(\nabla \)). In local coordinates, (D.8) and (D.9) take the form

$$\begin{aligned}&\pi ^{l[i} \nabla _l \pi ^{jk]} =0, \end{aligned}$$
(D.11)
$$\begin{aligned}&\pi ^{m[i} \nabla _m T_l^{jk]} - T_{l}^{m[i} T_m^{jk]} - \pi ^{[i|m} \pi ^{|j|n} R_{lmn}^{k]} = 0. \end{aligned}$$
(D.12)

In addition, one may verify the following formulas:

$$\begin{aligned}&[\nabla _m, \nabla _k]T_n^{ij} = \Theta _{km}^l \nabla _l T_n^{ij} - R_{nmk}^l T_l^{ij} + R_{lmk}^i T_n^{lj} + R_{lmk}^j T_m^{il}, \end{aligned}$$
(D.13)

and (cf. Eq. (2.26))

$$\begin{aligned}&- \nabla _k \pi ^{ij} + \pi ^{il} \Theta _{kl}^j = T_k^{ij}. \end{aligned}$$
(D.14)

Using (D.11), (D.12), (D.13), and (D.14), one now verifies

$$\begin{aligned}&\pi ^{m[i} \nabla _m S_{nk}{}^{jl]} + T_m^{[ij} S_{nk}{}^{l]m} - T_n^{m[i} S_{mk}{}^{jl]} - T_k^{m[i} S_{nm}{}^{jl]} = 0, \end{aligned}$$
(D.15)

which is the searched-for Bianchi identity of the basic curvature \(S_{nk}{}^{ij}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ikeda, N., Strobl, T. BV and BFV for the H-Twisted Poisson Sigma Model. Ann. Henri Poincaré 22, 1267–1316 (2021). https://doi.org/10.1007/s00023-020-00988-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00023-020-00988-0

Navigation