Skip to main content
Log in

Classifying Relaxed Highest-Weight Modules for Admissible-Level Bershadsky–Polyakov Algebras

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

Abstract

The Bershadsky–Polyakov algebras are the minimal quantum hamiltonian reductions of the affine vertex algebras associated to \(\mathfrak {sl}_{3}\) and their simple quotients have a long history of applications in conformal field theory and string theory. Their representation theories are therefore quite interesting. Here, we classify the simple relaxed highest-weight modules, with finite-dimensional weight spaces, for all admissible but nonintegral levels, significantly generalising the known highest-weight classifications (Arakawa in Commun Math Phys 323:627–633, 2013, Adamović and Kontrec in Classification of irreducible modules for Bershadsky–Polyakov algebra at certain levels). In particular, we prove that the simple Bershadsky–Polyakov algebras with admissible nonintegral \(\mathsf {k}\) are always rational in category \(\mathscr {O}\), whilst they always admit nonsemisimple relaxed highest-weight modules unless \(\mathsf {k}+\frac{3}{2} \in \mathbb {Z}_{\geqslant 0}\).

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

References

  1. Arakawa, T.: Rationality of Bershadsky-Polyakov vertex algebras. Commun. Math. Phys. 323, 627–633 (2013). arXiv:1005.0185 [math.QA]

  2. Adamović, D., Kontrec, A.: Classification of irreducible modules for Bershadsky–Polyakov algebra at certain levels. arXiv:1910.13781 [math.QA]

  3. Adamović, D., Kontrec, A.: Bershadsky–Polyakov vertex algebras at positive integer levels and duality. arXiv:2011.10021 [math.QA]

  4. Polyakov, A.: Gauge transformations and diffeomorphisms. Int. J. Mod. Phys. A 5, 833–842 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  5. Bershadsky, M.: Conformal field theories via Hamiltonian reduction. Commun. Math. Phys. 139, 71–82 (1991)

    Article  ADS  MathSciNet  Google Scholar 

  6. Kac, V., Roan, S., Wakimoto, M.: Quantum reduction for affine superalgebras. Commun. Math. Phys. 241, 307–342 (2003). arXiv:math-ph/0302015

    Article  ADS  MathSciNet  Google Scholar 

  7. Arakawa, T.: Associated varieties of modules over Kac–Moody algebras and \(C_2\)-cofiniteness of W-algebras. Int. Math. Res. Not. 2015, 11605–11666 (2015). arXiv:1004.1554 [math.QA]

  8. Kac, V., Wakimoto, M.: Quantum reduction and representation theory of superconformal algebras. Adv. Math. 185, 400–458 (2004). arXiv:math-ph/0304011

    Article  MathSciNet  Google Scholar 

  9. Arakawa, T.: Representation theory of superconformal algebras and the Kac-Roan-Wakimoto conjecture. Duke Math. J. 130, 435–478 (2005). arXiv:math-ph/0405015

    Article  MathSciNet  Google Scholar 

  10. Adamović, D., Milas, A.: Vertex operator algebras associated to modular invariant representations of \(A_{1}^{\left(1\right)}\). Math. Res. Lett. 2, 563–575 (1995). arXiv:q-alg/9509025

  11. Feigin, B., Semikhatov, A., Yu Tipunin, I.: Equivalence between chain categories of representations of affine \(sl \left(2 \right)\) and \(N = 2\) superconformal algebras. J. Math. Phys. 39, 3865–3905 (1998). arXiv:hep-th/9701043

    Article  ADS  MathSciNet  Google Scholar 

  12. Ridout, D., Wood, S.: Relaxed singular vectors, Jack symmetric functions and fractional level \({\widehat{\mathfrak{sl}}} \left(2 \right)\) models. Nucl. Phys. B 894, 621–664 (2015). arXiv:1501.07318 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  13. Gaberdiel, M.: Fusion rules and logarithmic representations of a WZW model at fractional level. Nucl. Phys. B 618, 407–436 (2001). arXiv:hep-th/0105046

    Article  ADS  MathSciNet  Google Scholar 

  14. Ridout, D.: \({\widehat{\mathfrak{sl}}} \left(2 \right)_{-1/2}\) and the triplet model. Nucl. Phys. B 835, 314–342 (2010). arXiv:1001.3960 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  15. Creutzig, T., Ridout, D.: Modular data and Verlinde formulae for fractional level WZW models I. Nucl. Phys. B 865, 83–114 (2012). arXiv:1205.6513 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  16. Creutzig, T., Ridout, D.: Modular data and Verlinde formulae for fractional level WZW models II. Nucl. Phys. B 875, 423–458 (2013). arXiv:1306.4388 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  17. Auger, J., Creutzig, T., Ridout, D.: Modularity of logarithmic parafermion vertex algebras. Lett. Math. Phys. 108, 2543–2587 (2018). arXiv:1704.05168 [math.QA]

  18. Adamović, D.: Realizations of simple affine vertex algebras and their modules: the cases \(\widehat{sl(2)}\) and \(\widehat{osp(1,2)}\). Commun. Math. Phys. 366, 1025–1067 (2019). arXiv:1711.11342 [math.QA]

  19. Kawasetsu, K., Ridout, D.: Relaxed highest-weight modules I: rank \(1\) cases. Commun. Math. Phys. 368, 627–663 (2019). arXiv:1803.01989 [math.RT]

  20. Arakawa, T., Futorny, V., Ramirez, L.: Weight representations of admissible affine vertex algebras. Commun. Math. Phys. 353, 1151–1178 (2017). arXiv:1605.07580 [math.RT]

  21. Ridout, D., Snadden, J., Wood, S.: An admissible level \(\widehat{\mathfrak{osp}} \left(1 \vert 2 \right)\)-model: modular transformations and the Verlinde formula. Lett. Math. Phys. 108, 2363–2423 (2018). arXiv:1705.04006 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  22. Wood, S.: Admissible level \({\mathfrak{osp}} \left(1 \vert 2 \right)\) minimal models and their relaxed highest weight modules. Transf. Groups 25, 887–943 (2020)

    Article  Google Scholar 

  23. Creutzig, T., Kanade, S., Liu, T., Ridout, D.: Cosets, characters and fusion for admissible-level \({\mathfrak{osp}}(1\vert 2)\) minimal models. Nucl. Phys. B 938, 22–55 (2018). arXiv:1806.09146 [hep-th]

    Article  ADS  Google Scholar 

  24. Kawasetsu, K., Ridout, D.: Relaxed highest-weight modules II: classifications for affine vertex algebras. arXiv:1906.02935 [math.RT]

  25. Kawasetsu, K.: Relaxed highest-weight modules III: character formulae. arXiv:2003.10148 [math.RT]

  26. Futorny, V., Morales, O., Ramirez, L.: Simple modules for affine vertex algebras in the minimal nilpotent orbit. arXiv:2002.05568 [math.RT]

  27. Futorny, V., Křižka, L.: Positive energy representations of affine vertex algebras. arXiv:2002.05586 [math.RT]

  28. Adamović, D.: A realization of certain modules for the \(N=4\) superconformal algebra and the affine Lie algebra \(A_{2}^{(1)}\). Transform. Groups 21, 299–327 (2016). arXiv:1407.1527 [math.QA]

  29. Ridout, D., Wood, S.: Bosonic ghosts at \(c=2\) as a logarithmic CFT. Lett. Math. Phys. 105, 279–307 (2015). arXiv:1408.4185 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  30. Fehily, Z., Ridout, D.: (in preparation)

  31. Adamović, D., Kawasetsu, K., Ridout, D.: A realisation of the Bershadsky–Polyakov algebras and their relaxed modules. arXiv:2007.00396 [math.QA]

  32. Zamolodchikov, A.: Infinite additional symmetries in two-dimensional conformal quantum field theory. Theoret. Math. Phys. 65, 1205–1213 (1985)

    Article  Google Scholar 

  33. Semikhatov, A.: Inverting the Hamiltonian reduction in string theory. In: 28th International Symposium on Particle Theory, Wendisch-Rietz, Germany, pp. 156–167. arXiv:hep-th/9410109 (1994)

  34. Mathieu, O.: Classification of irreducible weight modules. Ann. Inst. Fourier (Grenoble) 50, 537–592 (2000)

    Article  MathSciNet  Google Scholar 

  35. Creutzig, T., Ridout, D.: Logarithmic conformal field theory: beyond an introduction. J. Phys. A 46, 494006 (2013). arXiv:1303.0847 [hep-th]

    Article  MathSciNet  Google Scholar 

  36. Ridout, D., Wood, S.: The Verlinde formula in logarithmic CFT. J. Phys: Conf. Ser. 597, 012065 (2015). arXiv:1409.0670 [hep-th]

    Google Scholar 

  37. Smith, S.: A class of algebras similar to the enveloping algebra of \(sl(2)\). Trans. Am. Math. Soc. 322, 285–314 (1990)

    MathSciNet  MATH  Google Scholar 

  38. Kac, V., Wakimoto, M.: Modular invariant representations of infinite-dimensional Lie algebras and superalgebras. Proc. Nat. Acad. Sci. USA 85, 4956–4960 (1988)

    Article  ADS  MathSciNet  Google Scholar 

  39. Arakawa, T.: Rationality of admissible affine vertex algebras in the category \({\cal O\it }\). Duke Math. J. 165, 67–93 (2016). arXiv:1207.4857 [math.QA]

  40. Adamović, D., Kac, V., Möseneder Frajria, P., Papi, P., Perše, O.: Conformal embeddings of affine vertex algebras in minimal W-algebras II: decompositions. Jpn. J. Math. 12, 261–315 (2017). arXiv:1604.00893 [math.RT]

  41. Gorelik, M., Kac, V.: On simplicity of vacuum modules. Adv. Math. 211, 621–677 (2007). arXiv:math-ph/0606002

    Article  MathSciNet  Google Scholar 

  42. Zhu, Y.: Modular invariance of characters of vertex operator algebras. J. Am. Math. Soc. 9, 237–302 (1996)

    Article  MathSciNet  Google Scholar 

  43. Feigin, B., Nakanishi, T., Ooguri, H.: The annihilating ideals of minimal models. Int. J. Mod. Phys. A 7, 217–238 (1992)

    Article  ADS  MathSciNet  Google Scholar 

  44. Li, H.: Representation theory and tensor product theory for vertex operator algebras. Ph.D. thesis, Rutgers University. arXiv:hep-th/9406211 (1994)

  45. Kac, V., Wang, W.: Vertex operator superalgebras and their representations. In: Mathematical Aspects of Conformal and Topological Field Theories and Quantum Groups, Volume 175 of Contemporary Mathematics, pp. 161–191, Providence. American Mathematical Society. arXiv:hep-th/9312065 (1994)

  46. Dong, C., Li, H., Mason, G.: Twisted representations of vertex operator algebras. Math. Ann. 310, 571–600 (1998). arXiv:q-alg/9509005

  47. Blondeau-Fournier, O., Mathieu, P., Ridout, D., Wood, S.: Superconformal minimal models and admissible Jack polynomials. Adv. Math. 314, 71–123 (2017). arXiv:1606.04187 [hep-th]

    Article  MathSciNet  Google Scholar 

  48. De Sole, A., Kac, V.: Finite vs affine W-algebras. Jpn. J. Math. 1, 137–261 (2006). arXiv:math-ph/0511055

    Article  MathSciNet  Google Scholar 

  49. Tjin, T.: Finite W-algebras. Phys. Lett. B292, 60–66 (1992). arXiv:hep-th/9203077

    Article  ADS  MathSciNet  Google Scholar 

  50. Mazorchuk, V.: Lectures on \({\mathfrak{sl}}_{2} \left( {\mathbb{C}} \right)\)-Modules. Imperial College Press, London (2010)

    Google Scholar 

  51. Frenkel, I., Zhu, Y.: Vertex operator algebras associated to representations of affine and Virasoro algebras. Duke Math. J. 66, 123–168 (1992)

    Article  MathSciNet  Google Scholar 

  52. Di Francesco, P., Mathieu, P., Sénéchal, D.: Conformal Field Theory. Graduate Texts in Contemporary Physics. Springer, New York (1997)

    Book  Google Scholar 

  53. Kac, V., Wakimoto, M.: Classification of modular invariant representations of affine algebras. Infinite-Dimensional Lie Algebras and Groups (Luminy-Marseille. 1988), Volume 7 of Advanced Series in Mathematical Physics, pp. 138–177. World Scientific, New Jersey (1989)

  54. Arakawa, T.: Rationality of W-algebras: principal nilpotent cases. Ann. Math. 182, 565–604 (2015). arXiv:1211.7124 [math.QA]

  55. Adamović, D.: Some rational vertex algebras. Glas. Mat. Ser. III(29), 25–40 (1994). arXiv:q-alg/9502015

  56. Li, H.: The physics superselection principle in vertex operator algebra theory. J. Algebra 196, 436–457 (1997)

    Article  MathSciNet  Google Scholar 

  57. Creutzig, T., Ridout, D., Wood, S.: Coset constructions of logarithmic \(\left(1, p \right)\)-models. Lett. Math. Phys. 104, 553–583 (2014). arXiv:1305.2665 [math.QA]

  58. Adamović, D., Milas, A.: Lattice construction of logarithmic modules for certain vertex algebras. Selecta Math. New Ser. 15, 535–561 (2009). arXiv:0902.3417 [math.QA]

  59. Nagatomo, K., Tsuchiya, A.: The triplet vertex operator algebra \(W \left( p \right)\) and the restricted quantum group \({\overline{U}}_{q} \left( sl_2 \right)\) at \(q = e^{\frac{\pi i}{p}}\). Adv. Stud. Pure Math. 61, 1–49 (2011). arXiv:0902.4607 [math.QA]

  60. Ridout, D., Wood, S.: Modular transformations and Verlinde formulae for logarithmic \(\left( p_+, p_- \right)\)-models. Nucl. Phys. B 880, 175–202 (2014). arXiv:1310.6479 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  61. Ridout, D.: Fusion in fractional level \(\widehat{\mathfrak{sl}} \left(2 \right)\)-theories with \(k=-\tfrac{1}{2}\). Nucl. Phys. B 848, 216–250 (2011). arXiv:1012.2905 [hep-th]

    Article  ADS  MathSciNet  Google Scholar 

  62. Allen, R., Wood, S.: Bosonic ghostbusting—the bosonic ghost vertex algebra admits a logarithmic module category with rigid fusion. arXiv:2001.05986 [math.QA]

  63. Creutzig, T., Huang, Y.-Z., Yang, J.: Braided tensor categories of admissible modules for affine Lie algebras. Commun. Math. Phys. 362, 827–854 (2018). arXiv:1709.01865 [math.QA]

  64. Fiebig, P.: The combinatorics of category \(\mathscr {O}\) over symmetrizable Kac–Moody algebras. Transform. Groups 11, 29–49 (2006). arXiv:math.RT/0305378

  65. Kac, V.: Infinite-Dimensional Lie Algebras. Cambridge University Press, Cambridge (1990)

    Book  Google Scholar 

  66. Carter, R.: Lie Algebras of Finite and Affine Type. Cambridge Studies in Advanced Mathematics, vol. 96. Cambridge University Press, Cambridge (2005)

  67. Arakawa, T., van Ekeren, J.: Rationality and fusion rules of exceptional W-algebras. arXiv:1905.11473 [math.RT]

Download references

Acknowledgements

We thank Dražen Adamović, Tomoyuki Arakawa and Thomas Creutzig for interesting discussions related to the research reported here. ZF’s research is supported by an Australian Government Research Training Program (RTP) Scholarship. KK’s research is partially supported by MEXT Japan “Leading Initiative for Excellent Young Researchers (LEADER)”, JSPS Kakenhi Grant numbers 19KK0065 and 19J01093 and Australian Research Council Discovery Project DP160101520. DR’s research is supported by the Australian Research Council Discovery Project DP160101520 and the Australian Research Council Centre of Excellence for Mathematical and Statistical Frontiers CE140100049.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zachary Fehily.

Additional information

Communicated by H. -T. Yau

Publisher's Note

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

Appendix A: Proof of Theorem 4.8

Appendix A: Proof of Theorem 4.8

In this appendix, we adopt the notation of Sect. 4.1 and assume throughout that \(\lambda _0 \notin \mathbb {Z}_{\geqslant 0}\) so that \(H^0(\mathcal {L}_{\lambda }) \ne 0\) (and that the level \(\mathsf {k}\) is as in Assumption 1). With these assumptions, the aim is to prove the following assertion:

$$\begin{aligned} \mathsf {I}^{\mathsf {k}}\cdot \mathcal {L}_{\lambda } \ne 0 \quad \Rightarrow \quad H^0(\mathsf {I}^{\mathsf {k}}) \cdot H^0(\mathcal {L}_{\lambda }) \ne 0. \end{aligned}$$
(A.1)

\(H^0({-})\) is a cohomological functor which involves tensoring with a ghost vertex operator superalgebra whose vacuum element will be denoted by \({|}{0}{\rangle }\). With this, we shall prove (A.1) by exhibiting elements \(\chi \in \mathsf {I}^{\mathsf {k}}\) and \(v \in \mathcal {L}_{\lambda }\) for which \(\chi \otimes {|}{0}{\rangle }\) and \(v\otimes {|}{0}{\rangle }\) are (degree-0) closed elements of the appropriate BRST complexes and the (clearly closed) element \(\chi _n v\otimes {|}{0}{\rangle }\) is not exact, for some \(n \in \mathbb {Z}\). Using brackets to denote cohomology classes, \([\chi _n v\otimes {|}{0}{\rangle }]\) then gives a nonzero element of \(H^0(\mathsf {I}^{\mathsf {k}}) \cdot H^0(\mathcal {L}_{\lambda })\):

$$\begin{aligned} {[}\chi \otimes {|}{0}{\rangle }] \cdot [v \otimes {|}{0}{\rangle }] \equiv [\chi \otimes {|}{0}{\rangle }](z) [v \otimes {|}{0}{\rangle }] = [\chi (z) v \otimes {|}{0}{\rangle }] \ne 0. \end{aligned}$$
(A.2)

As noted at the end of Sect. 4.1, this amounts to a proof of Theorem 4.8. To prove (A.1) however, we need to delve a little deeper into the details of minimal quantum hamiltonian reduction for .

1.1 A.1. Minimal quantum hamiltonian reduction

Recall from [6] that the minimal quantum hamiltonian reduction functor \(H^0({-})\) computes the cohomology of the tensor product of a given -module with certain ghost vertex operator superalgebras. Specifically, we need a fermionic ghost system \(\mathsf {F}^{\alpha }\) for each positive root \(\alpha \in \Delta _+\) of \(\mathfrak {sl}_{3}\) and one bosonic ghost system \(\mathsf {B}\) corresponding to the two simple roots \(\alpha _{1}\) and \(\alpha _{2}\). Denoting the fermionic ghosts by \(b^{\alpha }\) and \(c^{\alpha }\), \(\alpha \in \Delta _+\), and the bosonic ghosts by \(\beta \) and \(\gamma \), we take the defining operator product expansions to be

(A.3)

understanding that the remaining operator product expansions between ghost generating fields are regular. The tensor product of these ghost vertex operator superalgebras will be denoted by \(\mathsf {G}= \mathsf {F}^{\alpha _{1}} \otimes \mathsf {F}^{\alpha _{2}} \otimes \mathsf {F}^{\theta } \otimes \mathsf {B}\), for convenience.

We fix a basis of \(\mathfrak {sl}_{3}\) for the computations to follow. Let \(E_{ij}\) denote the \(3 \times 3\) matrix with 1 in the (ij)-th position and zeroes elsewhere. Then, we set

$$\begin{aligned} e^{\theta } = E_{13}, \quad \begin{aligned} e^{\alpha _{1}}&= E_{12},&h^{\alpha _{1}}&= E_{11}-E_{22},&f^{\alpha _{1}}&= E_{21}, \\ e^{\alpha _{2}}&= E_{23},&h^{\alpha _{2}}&= E_{22}-E_{33},&f^{\alpha _{2}}&= E_{32}, \end{aligned} \quad f^{\theta } = E_{31}. \end{aligned}$$
(A.4)

Here, \(\theta = \alpha _{1} + \alpha _{2}\) is the highest root of \(\mathfrak {sl}_{3}\) and we shall also set \(h^{\theta } = h^{\alpha _{1}} + h^{\alpha _{2}} = E_{11}-E_{33}\).

To define \(H^0(\mathcal {M})\) for a -module \(\mathcal {M}\), one first grades \(\mathcal {M} \otimes \mathsf {G}\) by the fermionic ghost number, that is by the total number of c-modes minus the total number of b-modes. Equivalently, the ghost number is the eigenvalue of the zero mode of the field \(\sum _{\alpha \in \Delta _+} \mathopen {:} b^{\alpha }(z) c^{\alpha }(z) \mathclose {:}\). Next, one introduces [5, 6] the following fermionic field of ghost number 1:

(A.5)

A straightforward computation verifies that \(d(z) d(w) \sim 0\). We then form a differential complex by requiring that d(z) is homogeneous of conformal weight 1 and equipping \(\mathcal {M} \otimes \mathsf {G}\) with the differential \(d = d_0\) (which obviously squares to 0).

With (A.5), this requirement on d(z) requires that the conformal weight of \(c^{\theta }\) is also 1, whilst that of \(e^{\theta }\) is 0. The latter may be achieved by adding \(\frac{1}{2} \partial h^{\theta }\) to the standard Sugawara energy-momentum tensor \(T^{\text {Sug.}}\) of . When this is done, homogeneity and (A.5) now fix the conformal weight \(\widetilde{\Delta }\) of all the generating fields as in Table 1. The energy-momentum tensor of is thus

$$\begin{aligned} \begin{aligned}&\widetilde{L} = T^{\text {Sug.}} + \frac{1}{2} \partial h^{\theta } + \sum _{\alpha \in \Delta _+} T^{\mathsf {F}^{\alpha }} + T^{\mathsf {B}}, \\&\quad \text {where} \quad T^{\mathsf {F}^{\alpha _{i}}} = \frac{1}{2} \mathopen {:} \partial b^{\alpha _{i}} c^{\alpha _{i}} + \partial c^{\alpha _{i}} b^{\alpha _{i}} \mathclose {:},\\&\quad T^{\mathsf {F}^{\theta }} = \mathopen {:} \partial b^{\theta } c^{\theta } \mathclose {:} \quad \text {and} \quad T^{\mathsf {B}} = \frac{1}{2} \mathopen {:} \partial \gamma \beta - \partial \beta \gamma \mathclose {:}. \end{aligned} \end{aligned}$$
(A.6)

The central charge matches that of \(\mathsf {BP}^{\mathsf {k}}\), see (2.2):

$$\begin{aligned} \frac{8\mathsf {k}}{\mathsf {k}+3} - 6\mathsf {k}+ 1 + 1 - 2 - 1 = -\frac{(2\mathsf {k}+3)(3\mathsf {k}+1)}{\mathsf {k}+3}. \end{aligned}$$
(A.7)
Table 1 The ghost numbers \(\#\), charges \(\widetilde{j}\) and conformal weights \(\widetilde{\Delta }\) of the generating fields of the vertex operator superalgebra

As the notation suggests, \(\widetilde{L}\) is closed and its image in cohomology (that is, in ) is L. Note that the “symmetric” deformation of adding \(\frac{1}{2} \partial h^{\theta }\) to \(T^{\text {Sug.}}\) ensures this result. There are other deformations consistent with d being a differential—they correspond to adding a multiple of \(\partial J\) to L. Speaking of which, the element

$$\begin{aligned} \widetilde{J} = \frac{1}{3} (h^{\alpha _{1}} - h^{\alpha _{2}}) + \mathopen {:} b^{\alpha _{1}} c^{\alpha _{1}} \mathclose {:} - \mathopen {:} b^{\alpha _{2}} c^{\alpha _{2}} \mathclose {:} - \mathopen {:} \beta \gamma \mathclose {:} \end{aligned}$$
(A.8)

is likewise closed and its image in cohomology is J [6]. We give the charge (\(\widetilde{J}_0\)-eigenvalue) of the generating fields of in Table 1 for completeness. We also note that

$$\begin{aligned} \begin{aligned} \widetilde{G}^+&= f^{\alpha _{2}} + \mathopen {:} h^{\alpha _{2}} \beta \mathclose {:} - \mathopen {:} b^{\alpha _{1}} c^{\theta } \mathclose {:} - \mathopen {:} b^{\alpha _{1}} c^{\alpha _{1}} \beta \mathclose {:} + 2 \mathopen {:} b^{\alpha _{2}} c^{\alpha _{2}} \beta \mathclose {:} + \mathopen {:} b^{\theta } c^{\theta } \beta \mathclose {:} + \mathopen {:} \beta \beta \gamma \mathclose {:} + (\mathsf {k}+1) \partial \beta \\ \text {and} \quad \widetilde{G}^-&= f^{\alpha _{1}} - \mathopen {:} h^{\alpha _{1}} \gamma \mathclose {:} + \mathopen {:} b^{\alpha _{2}} c^{\theta } \mathclose {:} - 2 \mathopen {:} b^{\alpha _{1}} c^{\alpha _{1}} \gamma \mathclose {:} + \mathopen {:} b^{\alpha _{2}} c^{\alpha _{2}} \gamma \mathclose {:} - \mathopen {:} b^{\theta } c^{\theta } \gamma \mathclose {:} + \mathopen {:} \gamma \gamma \beta \mathclose {:} - (\mathsf {k}+1) \partial \gamma \end{aligned} \end{aligned}$$
(A.9)

are both closed. Their images in cohomology are \(G^+\) and \(G^-\), respectively [8].

We remark that deforming the energy-momentum tensor of means that we now have two distinct mode conventions for affine fields. Our convention will be that mode indices with respect to the deformed conformal weight will be denoted with parentheses. Thus, for an affine generator a with deformed conformal weight \(\widetilde{\Delta }\) as in Table 1, we shall write

$$\begin{aligned} a(z) = \sum _{n \in \mathbb {Z}} a_n z^{-n-1} = \sum _{n \in \mathbb {Z}- \widetilde{\Delta }} a_{(n)} z^{-n-\widetilde{\Delta }}. \end{aligned}$$
(A.10)

We shall not bother to so distinguish mode indices for ghost fields: their expansions will always be taken with respect to the conformal weights in Table 1.

1.2 A.2. The proof

We start with a well known fundamental result for the highest-weight vector v of \(\mathcal {L}_{\lambda }\), recalling that we are assuming throughout that \(\lambda _0 \notin \mathbb {Z}_{\geqslant 0}\) and that \(\mathsf {k}\) satisfies Assumption 1. Let \({|}{0}{\rangle }\) denote the vacuum vector of \(\mathsf {G}\). By [9, Lem. 4.6.1 and Prop. 4.7.1], we have the following lemma.

Lemma A.1

For all \(n \in \mathbb {Z}_{\geqslant 0}\), \((e^{\theta }_{-1})^n v \otimes {|}{0}{\rangle }\) is closed and inexact. In particular, \([v \otimes {|}{0}{\rangle }] \ne 0\).

We next consider the maximal ideal \(\mathsf {I}^{\mathsf {k}}\) of .

Lemma A.2

\(\mathsf {I}^{\mathsf {k}}\) is generated by a single singular vector \(\chi \) whose \(\mathfrak {sl}_{3}\)-weight and conformal weight are \((\mathsf {u}-2) \theta \) and \((\mathsf {u}-2) \mathsf {v}\), respectively. Moreover, \(\chi \otimes {|}{0}{\rangle }\) is closed.

Proof

This follows easily from [38, Cor. 1], which says that the maximal submodule of a Verma module whose highest weight is admissible is generated by singular vectors of known weight. In our case, the highest weight is \(\mathsf {k}\omega _{0}\) (which is admissible because \(\mathsf {k}\) is) and the only generating singular vector that is nonzero in the quotient of this Verma module has weight \(w_{} \cdot (\mathsf {k}\omega _{0})\), where \(w_{}\) is the Weyl reflection corresponding to the root \(-\theta + \mathsf {v}\delta \). Here, \(\delta \) denotes the standard imaginary root of \(\widehat{\mathfrak {sl}}_{3}\). This singular vector is \(\chi \) and its \(\mathfrak {sl}_{3}\)- and conformal weights are now easily computed. The fact that \(\chi \otimes {|}{0}{\rangle }\) is closed follows from \(\chi \) being a highest-weight vector. \(\square \)

In fact, \(\chi \otimes {|}{0}{\rangle }\) is also inexact, though we will not need to a priori establish this fact for what follows.

We remark that a nice conceptual proof of [38, Cor. 1] starts from the celebrated fact that the submodule structure of a Verma module only depends on the corresponding integral Weyl group [64]. This structure is therefore the same for all admissible highest-weight \(\widehat{\mathfrak {sl}}_{3}\)-modules, irrespective of their level. In particular, this structure matches that of a Verma module whose simple quotient is integrable, integrability being equivalent to admissibility for simple highest-weight modules with \(\mathsf {v}=1\). However, the fact that the maximal submodule is generated by singular vectors is well known in the integrable case, see [65] or [66].

Suppose now that \(\chi (z) v = 0\). Because \(\chi \) generates \(\mathsf {I}^{\mathsf {k}}\), it follows that \(\mathsf {I}^{\mathsf {k}}\cdot v = 0\). Since v generates \(\mathcal {L}_{\lambda }\), as a -module, and \(\mathsf {I}^{\mathsf {k}}\) is a two-sided ideal of , we get \(\mathsf {I}^{\mathsf {k}}\cdot \mathcal {L}_{\lambda } = 0\). Thus, the hypothesis of (A.1), that \(\mathcal {L}_{\lambda }\) is not an -module, requires that \(\chi _n v \ne 0\) for some \(n \in \mathbb {Z}\). As \(\chi \) has \(\mathfrak {sl}_{3}\)-weight \((\mathsf {u}-2) \theta \), our knowledge of the weights of \(\mathcal {L}_{\lambda }\) lets us refine this requirement to \(\chi _{-(\mathsf {u}-2)-i} v \ne 0\) for some \(i \in \mathbb {Z}_{\geqslant 0}\). There is therefore a minimal \(N \in \mathbb {Z}_{\geqslant 0}\) such that \(\chi _{-(\mathsf {u}-2)-N} v \ne 0\).

As \(\mathcal {L}_{\lambda }\) is simple, there therefore exists a Poincaré–Birkhoff–Witt monomial such that

$$\begin{aligned} U \chi _{-(\mathsf {u}-2)-N} v = v \end{aligned}$$
(A.11)

(rescaling \(\chi \) if necessary). We choose an ordering for U so that

$$\begin{aligned} f^{\alpha }_{n\leqslant 0}< h^{\alpha }_{n<0}< e^{\alpha }_{n<0}< f^{\alpha }_{n>0}< h^{\alpha }_{n>0} < e^{\alpha }_{n\geqslant 0} \end{aligned}$$
(A.12)

(obviously we may omit the \(h^{\alpha }_0\) and K). This means, for example, that the \(f^{\alpha }_n\) with \(n\leqslant 0\) are ordered to the left while the \(e^{\alpha }_n\) with \(n\geqslant 0\) are ordered to the right. For \(n\geqslant 0\), we have \(e^{\alpha }_0 \chi = 0\) and \(e^{\alpha }_n v = 0\), hence

$$\begin{aligned} e^{\alpha }_n \chi _{-(\mathsf {u}-2)-N} v = (e^{\alpha }_0 \chi )_{-(\mathsf {u}-2)-N+n} v = 0. \end{aligned}$$
(A.13)

We may therefore assume that U contains no \(e^{\alpha }_n\)-modes with \(n\geqslant 0\). Similarly,

$$\begin{aligned} h^{\alpha }_n \chi _{-(\mathsf {u}-2)-N} v = (\mathsf {u}-2) \theta (h^{\alpha }) \chi _{-(\mathsf {u}-2)-(N-n)} v = 0 \end{aligned}$$
(A.14)

for \(n>0\), by the minimality of N. Thus, we may assume that U contains no \(h^{\alpha }_n\)-modes with \(n>0\) either. Finally, v is not in the image of any \(f^{\alpha }_n\), with \(n\leqslant 0\), \(h^{\alpha }_n\), with \(n<0\), or \(e^{\alpha }_n\), with \(n<0\). All these modes may therefore also be excluded from U. We conclude that U may be taken to be a monomial in the modes \(f^{\alpha }_n\) with \(n>0\).

Given a partition \(\xi = [\xi _1 \geqslant \xi _2 \geqslant \cdots ]\), let \(\ell (\xi )\) denote its length and \({|}{\xi }{|}\) denote its weight. We write \(f^{\alpha }_{\xi } = f^{\alpha }_{\xi _1} f^{\alpha }_{\xi _2} \cdots \). Then, there exist partitions \(\xi \), \(\pi \) and \(\rho \) such that \(U = f^{\theta }_{\xi } f^{\alpha _{2}}_{\pi } f^{\alpha _{1}}_{\rho }\) and so

$$\begin{aligned} f^{\theta }_{\xi } f^{\alpha _{2}}_{\pi } f^{\alpha _{1}}_{\rho } \chi _{-(\mathsf {u}-2)-N} v = v. \end{aligned}$$
(A.15)

Moreover, considering \(\mathfrak {sl}_{3}\)- and conformal weights gives

$$\begin{aligned} \ell (\pi ) = \ell (\rho ), \quad \ell (\xi ) + \ell (\pi ) = \mathsf {u}-2 \quad \text {and} \quad {|}{\xi }{|} + {|}{\pi }{|} + {|}{\rho }{|} = \mathsf {u}-2+N. \end{aligned}$$
(A.16)

Lemma A.3

Let F(z), \(F \in \mathfrak {sl}_{3}\), be an affine field and let \(U_0\) be a monomial in the negative root vectors \(f^{\alpha }_0\) of \(\widehat{\mathfrak {sl}}_{3}\). Then, the modes of the field \((U_0 \chi )(w)\) satisfy

(A.17)

Proof

Observe that \(U_0 \chi \) is annihilated by the \(F_m\) with \(m>0\). Consequently, the assertion follows easily from the operator product expansion

$$\begin{aligned} F(z) (U_0 \chi )(w) \sim \frac{(F_0 U_0 \chi )(w)}{z-w}. \end{aligned}$$
(A.18)

\(\square \)

We apply Lemma A.3 to the left-hand side of (A.15), noting that the \(f^{}\)-modes all annihilate v. The result is

(A.19)

using (A.16). This looks complicated, but it allows us to determine the partitions \(\xi \), \(\pi \) and \(\rho \).

Lemma A.4

If any of the parts of \(\xi \), \(\pi \) or \(\rho \) are greater than 1, then \(f^{\theta }_{\xi } f^{\alpha _{2}}_{\pi } f^{\alpha _{1}}_{\rho } \chi _{-(\mathsf {u}-2)-N} v = 0\).

Proof

Suppose that \(\xi \) has a part \(\xi _i > 1\) (the argument is identical if \(\pi \) or \(\rho \) has a part greater than 1). Then, we can form a new partition \(\xi '\) from \(\xi \) by subtracting 1 from \(\xi _i\) and reordering parts if necessary. Note that \(\ell (\xi ') = \ell (\xi )\) and \({|}{\xi '}{|} = {|}{\xi }{|} - 1\). Then, Lemma A.3 and N being minimal give

(A.20)

But, this is the right-hand side of (A.19). \(\quad \square \)

Combining (A.15), which is manifestly nonzero, with Lemma A.4 now forces all parts of \(\xi \), \(\pi \) and \(\rho \) to be 1. As partition lengths and weights are now equal, the relations of (A.16) are easily solved to give \({|}{\xi }{|} = \mathsf {u}-2-N\) and \({|}{\pi }{|} = {|}{\rho }{|} = N\). In particular, (A.15) now becomes

$$\begin{aligned} (f^{\theta }_1)^{\mathsf {u}-2-N} (f^{\alpha _{2}}_1)^N (f^{\alpha _{1}}_1)^N \chi _{-(\mathsf {u}-2)-N} v = v. \end{aligned}$$
(A.21)

By rescaling \(\chi \) again, if necessary, we arrive at following key result.

Proposition A.5

If N is the minimal integer such that \(\chi _{-(\mathsf {u}-2)-N} v \ne 0\), then

$$\begin{aligned} (f^{\alpha _{2}}_1)^N (f^{\alpha _{1}}_1)^N \chi _{-(\mathsf {u}-2)-N} v = (e^{\theta }_{-1})^{\mathsf {u}-2-N} v. \end{aligned}$$
(A.22)

The idea now is to use the fact that the right-hand side of (A.22) is inexact when tensored with \({|}{0}{\rangle }\) (Lemma A.1) to prove that the same is true for \(\chi _{-(\mathsf {u}-2)-N} v\). For this, we need to replace the action of \(f^{\alpha _{2}}_1\) and \(f^{\alpha _{1}}_1\) with elements that preserve exactness, for example any closed elements.

Lemma A.6

For all \(i,j \in \mathbb {Z}_{\geqslant 0}\), we have

(A.23)

Proof

We start with (A.9), which gives

$$\begin{aligned} \widetilde{G}^-_{(1/2)} = f^{\alpha _{1}}_{(1/2)} - \sum _{m \in \mathbb {Z}} h^{\alpha _{1}}_{(m)} \gamma _{-m+1/2} + \cdots = f^{\alpha _{1}}_1 - \sum _{m \in \mathbb {Z}} h^{\alpha _{1}}_m \gamma _{-m+1/2} + \cdots , \end{aligned}$$
(A.24)

where the \(\cdots \) stands for pure ghost terms. As these ghost terms annihilate \({|}{0}{\rangle }\), we have

(A.25)

for any \(j \in \mathbb {Z}_{\geqslant 0}\). Now, \(m\geqslant 1\) implies that \(h^{\alpha _{2}}_m v = 0\), hence that

(A.26)

The first commutator on the right-hand side is a sum of terms, each obtained from \((f^{\alpha _{1}}_1)^j\) by replacing one of the \(f^{\alpha _{1}}_1\) by \(-2 f^{\alpha _{1}}_{m+1}\). However, each of these terms is 0 by Lemma A.4. On the other hand, the second commutator is proportional to \(\chi _{-(\mathsf {u}-2)-(N-m)}\), so it annihilates v by minimality of N. We therefore obtain

(A.27)

from which we conclude inductively that , for all \(i\in \mathbb {Z}_{\geqslant 0}\).

To deduce (A.23), we now repeat the argument by acting with \(\widetilde{G}^+_{(1/2)}\) on \((f^{\alpha _{2}}_1)^i (f^{\alpha _{1}}_1)^j \chi _{-(\mathsf {u}-2)-N} v \otimes {|}{0}{\rangle }\). There are no essential differences between this case and that described above, so we omit the details.

Corollary A.7

\(\chi _{-(\mathsf {u}-2)-N} v \otimes {|}{0}{\rangle }\) is closed and inexact.

Proof

We have already seen that \(\chi _{-(\mathsf {u}-2)-N} v \otimes {|}{0}{\rangle }\) is closed. Suppose therefore that it is exact. As , since \(\widetilde{G}^{\pm }\) is closed, it now follows from Proposition A.5 and Lemma A.6 that

(A.28)

is also exact. But, this contradicts Lemma A.1. \(\square \)

This corollary completes the proof of Theorem 4.8.

We conclude with a few remarks about this proof. First, proving that quantum hamiltonian reduction indeed realises all simple highest-weight modules is obviously desirable and has been studied in several settings. However, Arakawa’s general results [9, 54] in this direction for universal minimal and regular W-algebras do not immediately imply the desired results for their simple quotients. Indeed, the cases where this completeness result for simple W-algebras is known seem to be cases in which the simple quotient is rational and \(C_2\)-cofinite, see for example [1, 54, 67]. Our proof, applying as it does to the nonrational and non-\(C_2\)-cofinite simple Bershadsky–Polyakov algebras, is therefore quite novel and seems to be very different from the rational proofs in the literature.

Second, this proof relies on certain key facts that might be regarded as special to the Bershadsky–Polyakov algebras. In particular, we use the explicit realisation (A.9) of the charged generators of \(\mathsf {BP}^{\mathsf {k}}\). However, the pure ghost terms played no role in the proof, so it may be possible to generalise this part of the argument to other minimal, or perhaps even subregular, W-algebras. On the other hand, the proof also exploits the fact that the maximal ideal of is generated by a single singular vector, which does not normally hold when generalising to nonadmissible levels. It is therefore not clear that this proof can be adapted for the nonadmissible case, but it would of course be interesting to try.

Alternatively, it may be that one can prove more general completeness results of this type by further developing the inverse quantum hamiltonian reduction methods introduced in [18, 33] and extended to the Bershadsky–Polyakov algebras in [31]. These methods have the advantage of building up the representation theory iteratively from that of the so-called exceptional W-algebras [67], in particular from the regular ones. This may then lead to uniform methods for all W-algebras, at least when the level is admissible and (sufficiently) nondegenerate. We hope to have the opportunity to report on this promising direction in the future.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Fehily, Z., Kawasetsu, K. & Ridout, D. Classifying Relaxed Highest-Weight Modules for Admissible-Level Bershadsky–Polyakov Algebras. Commun. Math. Phys. 385, 859–904 (2021). https://doi.org/10.1007/s00220-021-04008-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-021-04008-y

Navigation