Skip to main content
Log in

An \(\mathfrak {sl}_2\)-type tensor category for the Virasoro algebra at central charge 25 and applications

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

Let \({\mathcal {O}}_{25}\) be the vertex algebraic braided tensor category of finite-length modules for the Virasoro Lie algebra at central charge 25 whose composition factors are the irreducible quotients of reducible Verma modules. We show that \({\mathcal {O}}_{25}\) is rigid and that its simple objects generate a semisimple tensor subcategory that is braided tensor equivalent to an abelian 3-cocycle twist of the category of finite-dimensional \(\mathfrak {sl}_2\)-modules. We also show that this \(\mathfrak {sl}_2\)-type subcategory is braid-reversed tensor equivalent to a similar category for the Virasoro algebra at central charge 1. As an application, we construct a simple conformal vertex algebra which contains the Virasoro vertex operator algebra of central charge 25 as a \(PSL_2({\mathbb {C}})\)-orbifold. We also use our results to study Arakawa’s chiral universal centralizer algebra of \(SL_2\) at level \(-1\), showing that it has a symmetric tensor category of representations equivalent to \(\textrm{Rep}\,PSL_2({\mathbb {C}})\). This algebra is an extension of the tensor product of Virasoro vertex operator algebras of central charges 1 and 25, analogous to the modified regular representations of the Virasoro algebra constructed earlier for generic central charges by I. Frenkel–Styrkas and I. Frenkel–M. Zhu.

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

References

  1. Adamović, D., Lin, X., Milas, A.: \(ADE\) subalgebras of the triplet vertex algebra \({\cal{W} }(p)\): \(A\)-series. Commun. Contemp. Math. 15(6), 1350028 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  2. Arakawa, T.: Chiral algebras of class \({\cal{S}}\) and Moore–Tachikawa symplectic varieties, arXiv:1811.01577 (2022)

  3. Belavin, A., Polyakov, A., Zamolodchikov, A.: Infinite conformal symmetry in two-dimensional quantum field theory. Nucl. Phys. B 241(2), 333–380 (1984)

    Article  MathSciNet  MATH  Google Scholar 

  4. Creutzig, T., Jiang, C., Orosz Hunziker, F., Ridout, D., Yang, J.: Tensor categories arising from the Virasoro algebra. Adv. Math. 380, 107601 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  5. Creutzig, T., Kanade, S., McRae, R.: Tensor categories for vertex operator superalgebra extensions, to appear in Mem. American Mathematical Society. arXiv:1705.05017 (2022)

  6. Creutzig, T., Kanade, S., McRae, R.: Gluing vertex algebras. Adv. Math. 396, 108174 (2022)

    Article  MathSciNet  MATH  Google Scholar 

  7. Creutzig, T., McRae, R., Yang, J.: Direct limit completions of vertex tensor categories. Commun. Contemp. Math. 24(2), 2150033 (2022)

    Article  MathSciNet  MATH  Google Scholar 

  8. Creutzig, T., McRae, R., Yang, J.: On ribbon categories for singlet vertex algebras. Comm. Math. Phys. 387(2), 865–925 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  9. Creutzig, T., McRae, R., Yang, J.: Tensor structure on the Kazhdan-Lusztig category for affine \(\mathfrak{gl} (1|1)\). Int. Math. Res. Not. (2021). https://doi.org/10.1093/imrn/rnab080

    Article  MATH  Google Scholar 

  10. Dong, C., Griess, R.: Rank one lattice type vertex operator algebras and their automorphism groups. J. Algebra 208(1), 262–275 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  11. Dong, C., Lepowsky, J.: Generalized Vertex Algebras and Relative Vertex Operators, Progress in Mathematics, 112, p. 202. Birkhäuser, Boston (1993)

    Book  MATH  Google Scholar 

  12. Dong, C., Mason, G.: Rational vertex operator algebras and the effective central charge. Int. Math. Res. Not. 56, 2989–3008 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  13. Etingof, P., Gelaki, S., Nikshych, D., Ostrik, V.: Tensor Categories, Mathematical Surveys and Monographs, vol. 205, p. 343. American Mathematical Society, Providence (2015)

    MATH  Google Scholar 

  14. Feigin, B., Frenkel, E.: Affine Kac-Moody Algebras at the Critical Level and Gelfand-Dikii Algebras, Infinite Analysis, Part A, B (Kyoto, 1991). Advances in Mathematical Physics, vol. 16, pp. 197–215. World Science Publications, River Edge (1992)

    MATH  Google Scholar 

  15. Feigin, B., Fuchs, D.: Representations of the Virasoro algebra Representation of Lie groups and related topics Advanced Studies in Contemporary Mathematics, vol. 7, pp. 465–554. Gordon and Breach, New York (1990)

    Google Scholar 

  16. Frenkel, I., Huang, Y.-Z., Lepowsky, J.: On axiomatic approaches to vertex operator algebras and modules. Mem. Am. Math. Soc. 104(494), 8–64 (1993)

    MathSciNet  MATH  Google Scholar 

  17. Frenkel, I., Lepowsky, J., Meurman, A.: Vertex Operator Algebras and the Monster, Pure and Applied Mathematics, vol. 134, p. 508. Academic Press Inc, Boston (1988)

    MATH  Google Scholar 

  18. Frenkel, I., Styrkas, K.: Modified regular representations of affine and Virasoro algebras, VOA structure and semi-infinite cohomology. Adv. Math. 206(1), 57–111 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  19. Frenkel, I., Zhu, Y.: Vertex operator algebras associated to representations of affine and Virasoro algebras. Duke Math. J. 66(1), 123–168 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  20. Frenkel, I., Zhu, M.: Vertex algebras associated to modified regular representations of the Virasoro algebra. Adv. Math. 229(6), 3468–3507 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  21. Gannon, T., Negron, C.: Quantum \(SL(2)\) and logarithmic vertex operator algebras at \((p,1)\)-central charge. arXiv:2104.12821 (2022)

  22. Gorelik, M., Kac, V.: On complete reducibility for infinite-dimensional Lie algebras. Adv. Math. 226(2), 1911–1972 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  23. Huang, Y.-Z.: Virasoro vertex operator algebras, the (nonmeromorphic) operator product expansion and the tensor product theory. J. Algebra 182(1), 201–234 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  24. Huang, Y.-Z.: Rigidity and modularity of vertex tensor categories. Commun. Contemp. Math. 10(suppl. 1), 871–911 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  25. Huang, Y.-Z., Kirillov, A., Lepowsky, J.: Braided tensor categories and extensions of vertex operator algebras. Comm. Math. Phys. 337(3), 1143–1159 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  26. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra. I: Introduction and strongly graded algebras and their generalized modules. Conformal field theories and tensor categories. In: Lecture Notes in Mathematics, pp. 169–248. Peking University, Springer, Heidelberg (2014)

    Google Scholar 

  27. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, II: logarithmic formal calculus and properties of logarithmic intertwining operators. arXiv:1012.4196 (2022)

  28. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, III: intertwining maps and tensor product bifunctors. arXiv:1012.4197 (2022)

  29. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, IV: constructions of tensor product bifunctors and the compatibility conditions. arXiv:1012.4198 (2022)

  30. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, V: convergence condition for intertwining maps and the corresponding compatibility condition. arXiv:1012.4199 (2022)

  31. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, VI: expansion condition, associativity of logarithmic intertwining operators, and the associativity isomorphisms. arXiv:1012.4202 (2022)

  32. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, VII: Convergence and extension properties and applications to expansion for intertwining maps. arXiv:1110.1929 (2022)

  33. Huang, Y.-Z., Lepowsky, J., Zhang, L.: Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, VIII: Braided tensor category structure on categories of generalized modules for a conformal vertex algebra., arXiv:1110.1931 (2022)

  34. Iohara, K., Koga, Y.: Representation Theory of the Virasoro Algebra, Springer Monographs in Mathematics, pp. 18–474. Springer-Verlag, London Ltd, London (2011)

    Book  MATH  Google Scholar 

  35. Kassel, C.: Quantum Groups, Graduate Texts in Mathematics, vol. 155, pp. 12–531. Springer-Verlag, New York (1995)

    Google Scholar 

  36. Kazhdan, D., Lusztig, G.: Tensor structures arising from affine Lie algebras, IV. J. Am. Math. Soc. 7(2), 383–453 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  37. Kazhdan, D., Wenzl, H.: Reconstructing monoidal categories. In: Gelfand, I.M. (ed.) Seminar Advances in Soviet Mathematics, vol. 16, pp. 111–136. American Mathematical Society, Providence (1993)

    Google Scholar 

  38. Kirillov, A., Jr., Ostrik, V.: On a \(q\)-analogue of the McKay correspondence and the \(ADE\) classification of \(\mathfrak{sl} _2\) conformal field theories. Adv. Math. 171(2), 183–227 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  39. Lepowsky, J., Li, H.: Introduction to Vertex Operator Algebras and Their Representations, Progress in Mathematics, pp. 14–318. Birkhäuser, Boston (2004)

    Google Scholar 

  40. Li, H.: Symmetric invariant bilinear forms on vertex operator algebras. J. Pure Appl. Algebra 96(3), 279–297 (1994)

    Article  MathSciNet  MATH  Google Scholar 

  41. Li, H., Xu, X.: A characterization of vertex algebras associated to even lattices. J. Algebra 173(2), 253–270 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  42. Lusztig, G.: Introduction to Quantum Groups, Progress in Mathematics, vol. 110, pp. 12–341. Birkhäuser, Boston (1993)

    Google Scholar 

  43. McRae, R.: On the tensor structure of modules for compact orbifold vertex operator algebras. Math. Z. 296(1–2), 409–452 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  44. McRae, R.: A general mirror equivalence theorem for coset vertex operator algebras. arXiv:2107.06577 (2022)

  45. McRae, R., Yang, J.: Structure of Virasoro tensor categories at central charge \(13-6p-6p^{-1}\) for integers \(p>1\). arXiv:2011.02170 (2022)

  46. Milas, A.: Fusion rings for degenerate minimal models. J. Algebra 254(2), 300–335 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  47. Orosz Hunziker, F.: Fusion rules for the Virasoro algebra of central charge \(25\). Algebr. Represent. Theory 23(5), 2013–2031 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  48. Tsuchiya, A., Wood, S.: The tensor structure on the representation category of the \({\cal{W} }_{p}\) triplet algebra. J. Phys. A 46(44), 445203 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  49. Wang, W.: Rationality of Virasoro vertex operator algebras. Internat. Math. Res. Notices 7, 197–211 (1993)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

We thank Tomoyuki Arakawa for pointing out to us his construction of the chiral universal centralizer algebra and for suggesting that we study its representations. We also thank Drazen Adamović, Naoki Genra, Andrew Linshaw, and the referee for comments. R. McRae is supported by a startup grant from Tsinghua University. J. Yang is supported by a startup grant from Shanghai Jiao Tong University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Robert McRae.

Additional information

Publisher's Note

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

Appendices

Appendix A: Algebras in equivalent tensor categories

The tensor-categorical results in this appendix are straightforward and certainly known, but we include some details from their proofs to make this paper more self-contained. We also use this appendix to recall notation from, for example, [5, 38] for algebras in tensor categories and their modules.

Let \({\mathcal {C}}\) be a braided tensor category with tensor product bifunctor \(\boxtimes \) and unit object \({\textbf{1}}\). We recall (from [5, 38], for example) that a (commutative) \({\mathcal {C}}\)-algebra \((A,\mu _A,\iota _A)\) is an object A of \({\mathcal {C}}\) equipped with multiplication and unit morphisms

$$\begin{aligned} \mu _A: A\boxtimes A\longrightarrow A,\qquad \iota _A: {\textbf{1}}\longrightarrow A \end{aligned}$$

which satisfy natural unit and associativity (and commutativity) axioms. An isomorphism between two \({\mathcal {C}}\)-algebras \((A,\mu _A,\iota _A)\) and \((B,\mu _B,\iota _B)\) is a \({\mathcal {C}}\)-morphism \(f: A\rightarrow B\) such that \(f\circ \mu _A =\mu _B\circ (f\boxtimes f)\) and \(f\circ \iota _A=\iota _B\).

Given a commutative algebra A, the tensor category \({{\,\textrm{Rep}\,}}A\) of (left) A-modules has objects \((X,\mu _X)\) where X is an object of \({\mathcal {C}}\) and \(\mu _X: A\boxtimes X\rightarrow X\) is a morphism satisfying natural left unit and associativity axioms. The category \({{\,\textrm{Rep}\,}}A\) has a braided tensor subcategory \({{\,\textrm{Rep}\,}}^0A\) consisting of all objects \((X,\mu _X)\) such that

$$\begin{aligned} \mu _X\circ {\mathcal {M}}_{A,X}=\mu _X, \end{aligned}$$

where \({\mathcal {M}}_{A,X}:={\mathcal {R}}_{X,A}\circ {\mathcal {R}}_{A,X}\) is the natural double braiding, or monodromy, isomorphism in \({\mathcal {C}}\). Given an object \((X,\mu _X)\) of \({{\,\textrm{Rep}\,}}A\), an A-submodule of X is an object \((W,\mu _W)\) of \({{\,\textrm{Rep}\,}}A\) equipped with a \({{\,\textrm{Rep}\,}}A\)-injection \(i: W\rightarrow X\). We say that \((X,\mu _X)\) is simple if every submodule \(i: W\rightarrow X\) is either 0 or an isomorphism. If A is commutative, we can identify ideals of A with A-submodules of A, so that A is simple as a \({\mathcal {C}}\)-algebra if and only if it simple as an A-module.

Let \({\mathcal {F}}:{\mathcal {C}}\rightarrow {\mathcal {D}}\) be a tensor functor, so that there is an isomorphism \(\varphi :{\textbf{1}}_{\mathcal {D}}\rightarrow {\mathcal {F}}({\textbf{1}}_{\mathcal {C}})\) and a natural isomorphism

$$\begin{aligned} F: \boxtimes _{\mathcal {D}}\circ ({\mathcal {F}}\times {\mathcal {F}})\longrightarrow {\mathcal {F}}\circ \boxtimes _{\mathcal {C}}\end{aligned}$$
(A.1)

which are suitably compatible with the unit and associativity isomorphisms of \({\mathcal {C}}\) and \({\mathcal {D}}\). If \((A,\mu _A,\iota _A)\) is a \({\mathcal {C}}\)-algebra, then \(({\mathcal {F}}(A),\mu _{{\mathcal {F}}(A)},\iota _{{\mathcal {F}}(A)})\) is a \({\mathcal {D}}\)-algebra, where

$$\begin{aligned} \mu _{{\mathcal {F}}(A)}: {\mathcal {F}}(A)\boxtimes _{\mathcal {D}}{\mathcal {F}}(A)\xrightarrow {F_{A,A}} {\mathcal {F}}(A\boxtimes _{\mathcal {C}}A) \xrightarrow {{\mathcal {F}}(\mu _A)} {\mathcal {F}}(A) \end{aligned}$$

and

$$\begin{aligned} \iota _{{\mathcal {F}}(A)}: {\textbf{1}}_{\mathcal {D}}\xrightarrow {\varphi }{\mathcal {F}}({\textbf{1}}_{\mathcal {C}})\xrightarrow {{\mathcal {F}}(\iota _A)} {\mathcal {F}}(A). \end{aligned}$$

If A is commutative and \({\mathcal {F}}\) is braided or braid-reversing, then \({\mathcal {F}}(A)\) is also commutative.

If \((X,\mu _X)\) is an object of \({{\,\textrm{Rep}\,}}A\), then \(({\mathcal {F}}(X),\mu _{{\mathcal {F}}(X)})\) is an object of \({{\,\textrm{Rep}\,}}{\mathcal {F}}(A)\), where

$$\begin{aligned} \mu _{{\mathcal {F}}(X)}: {\mathcal {F}}(A)\boxtimes _{\mathcal {D}}{\mathcal {F}}(X)\xrightarrow {F_{A,X}} {\mathcal {F}}(A\boxtimes _{\mathcal {C}}X)\xrightarrow {{\mathcal {F}}(\mu _X)} {\mathcal {F}}(X). \end{aligned}$$

Moreover, if \(f: X_1\rightarrow X_2\) is a morphism in \({{\,\textrm{Rep}\,}}A\), then \({\mathcal {F}}(f): {\mathcal {F}}(X_1)\rightarrow {\mathcal {F}}(X_2)\) is a morphism in \({{\,\textrm{Rep}\,}}{\mathcal {F}}(A)\). If A is commutative, \({\mathcal {F}}\) is braided or braid-reversing, and \((X,\mu _X)\) is an object of \({{\,\textrm{Rep}\,}}^0A\), then \(({\mathcal {F}}(X),\mu _{{\mathcal {F}}(X)})\) is an object of \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\). Thus if \({\mathcal {F}}\) is braided or braid-reversing, it restricts to a functor from \({{\,\textrm{Rep}\,}}^0 A\) to \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\). In the case that \({\mathcal {F}}\) is an equivalence of categories, we get:

Proposition A.1

Let \({\mathcal {F}}:{\mathcal {C}}\rightarrow {\mathcal {D}}\) be a braided or braid-reversing tensor equivalence. If A is a commutative \({\mathcal {C}}\)-algebra, then \({\mathcal {F}}:{{\,\textrm{Rep}\,}}^0 A\rightarrow {{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\) is an equivalence of categories.

Proof

To show that \({\mathcal {F}}\) is essentially surjective onto \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\), suppose \((W,\mu _W)\) is an object of \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\). Since \({\mathcal {F}}\) is essentially surjective onto \({\mathcal {D}}\), there is an isomorphism \(f: W\rightarrow {\mathcal {F}}(X)\) in \({\mathcal {D}}\) for some object X in \({\mathcal {C}}\), and then \(f: (W,\mu _W)\rightarrow ({\mathcal {F}}(X),\mu _{{\mathcal {F}}(X)})\) is an isomorphism in \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\) where

$$\begin{aligned} \mu _{{\mathcal {F}}(X)}=f\circ \mu _W\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}f^{-1}). \end{aligned}$$

Using the full faithfulness of \({\mathcal {F}}\), we define \(\mu _X: A\boxtimes _{\mathcal {C}}X\rightarrow X\) to be the unique \({\mathcal {C}}\)-morphism such that

$$\begin{aligned} {\mathcal {F}}(\mu _X)=\mu _{{\mathcal {F}}(X)}\circ F_{A,X}^{-1}. \end{aligned}$$

It is then straightforward to check that \((X,\mu _X)\) is an object of \({{\,\textrm{Rep}\,}}A\) such that \({\mathcal {F}}(X,\mu _X)=({\mathcal {F}}(X),\mu _{{\mathcal {F}}(X)})\cong (W,\mu _W)\). For example, to prove the associativity of \(\mu _X\), the faithfulness of \({\mathcal {F}}\) implies it is enough to show

$$\begin{aligned} {\mathcal {F}}(\mu _X\circ (\textrm{Id}_A\boxtimes _{\mathcal {C}}\mu _X)) = {\mathcal {F}}(\mu _X\circ (\mu _A\boxtimes _{\mathcal {C}}\textrm{Id}_X)\circ {\mathcal {A}}_{A,A,X}), \end{aligned}$$

which follows from the definitions and the associativity of \(({\mathcal {F}}(X),\mu _{{\mathcal {F}}(X)})\). Similarly, since W, equivalently \({\mathcal {F}}(X)\), is an object of \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\), then

$$\begin{aligned} {\mathcal {F}}(\mu _X)={\mathcal {F}}(\mu _X\circ {\mathcal {M}}_{A,X}^{\pm 1}) \end{aligned}$$

since \({\mathcal {F}}\) is braided or braid-reversing, and it follows that X is an object of \({{\,\textrm{Rep}\,}}^0A\).

Now since \({\mathcal {F}}: {\mathcal {C}}\rightarrow {\mathcal {D}}\) is faithful, \({\mathcal {F}}\) remains faithful when restricted to \({{\,\textrm{Rep}\,}}^0A\). Then to show that the restriction to \({{\,\textrm{Rep}\,}}^0 A\) is full, let \(g: {\mathcal {F}}(X_1)\rightarrow {\mathcal {F}}(X_2)\) be a morphism in \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\), where \(X_1\) and \(X_2\) are objects of \({{\,\textrm{Rep}\,}}^0 A\). Then \(g={\mathcal {F}}(f)\) for some morphism \(f: X_1\rightarrow X_2\) in \({\mathcal {C}}\), and

$$\begin{aligned} {\mathcal {F}}(f\circ \mu _{X_1})&=g\circ \mu _{{\mathcal {F}}(X_1)}\circ F_{A,X_1}^{-1} =\mu _{{\mathcal {F}}(X_2)}\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}g)\circ F_{A,X_1}^{-1} \\&= \mu _{{\mathcal {F}}(X_2)}\circ F_{A,X_2}^{-1}\circ {\mathcal {F}}(\textrm{Id}_A\boxtimes _{\mathcal {C}}f) ={\mathcal {F}}(\mu _{X_2}\circ (\textrm{Id}_A\boxtimes _{\mathcal {C}}f)). \end{aligned}$$

Since \({\mathcal {F}}\) is faithful, it follows that f is a morphism in \({{\,\textrm{Rep}\,}}^0 A\). \(\square \)

Since equivalences of abelian categories preserve zero objects and thus also zero morphisms, if \({\mathcal {F}}:{\mathcal {C}}\rightarrow {\mathcal {D}}\) is an equivalence, then a morphism i in \({\mathcal {C}}\) is injective if and only if \({\mathcal {F}}(i)\) is injective in \({\mathcal {D}}\). Consequently, \({\mathcal {F}}\) also preserves simple objects. Thus:

Corollary A.2

Let \({\mathcal {F}}:{\mathcal {C}}\rightarrow {\mathcal {D}}\) be a braided or braid-reversing tensor functor. If A is a commutative \({\mathcal {C}}\)-algebra, then an object X of \({{\,\textrm{Rep}\,}}^0 A\) is simple if and only if \({\mathcal {F}}(X)\) is a simple object of \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\). In particular, A is a simple \({\mathcal {C}}\)-algebra if and only if \({\mathcal {F}}(A)\) is a simple \({\mathcal {D}}\)-algebra.

The next result is that if \({\mathcal {F}}: {\mathcal {C}}\rightarrow {\mathcal {D}}\) is a braided or braid-reversing tensor functor and A is a commutative \({\mathcal {C}}\)-algebra, then \({\mathcal {F}}\) is also braided or braid-reversing on \({{\,\textrm{Rep}\,}}^0A\). The proof is straightforward but somewhat long, so we only include the more non-trivial details:

Theorem A.3

Let \({\mathcal {F}}: {\mathcal {C}}\rightarrow {\mathcal {D}}\) be a right exact braided, respectively braid-reversing, tensor functor. If A is a commutative \({\mathcal {C}}\)-algebra, then \({\mathcal {F}}:{{\,\textrm{Rep}\,}}^0 A\rightarrow {{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\) has the structure of a braided, respectively braid-reversing, tensor functor.

Proof

Letting \(\boxtimes _A\) and \(\boxtimes _{{\mathcal {F}}(A)}\) denote the tensor products in \({{\,\textrm{Rep}\,}}^0 A\) and \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\), respectively, we need to obtain a natural isomorphism

$$\begin{aligned} F^A: \boxtimes _{{\mathcal {F}}(A)}\circ ({\mathcal {F}}\times {\mathcal {F}})\longrightarrow {\mathcal {F}}\circ \boxtimes _A. \end{aligned}$$

from the original natural isomorphism F of (A.1). To do so, we recall (from [5, Section 2.3] for example) that for objects \(X_1\) and \(X_2\) in \({{\,\textrm{Rep}\,}}^0 A\), \(X_1\boxtimes _A X_2\) is the cokernel of \(\mu ^{(1)}-\mu ^{(2)}\), where \(\mu ^{(1)}=\mu _{X_1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2}\) and \(\mu ^{(2)}\) is the composition

$$\begin{aligned} (A\boxtimes _{\mathcal {C}}X_1)\boxtimes _{{\mathcal {C}}} X_2&\xrightarrow {{\mathcal {R}}_{A,X_1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2}} (X_1\boxtimes _{\mathcal {C}}A)\boxtimes _{{\mathcal {C}}} X_2 \\&\xrightarrow {{\mathcal {A}}_{X_1,A,X_2}^{-1}} X_1\boxtimes _{\mathcal {C}}(A\boxtimes _{\mathcal {C}}X_2)\xrightarrow {\textrm{Id}_{X_1}\boxtimes _{\mathcal {C}}\mu _{X_2}} X_1\boxtimes _{\mathcal {C}}X_2. \end{aligned}$$

We use \(\eta _{X_1,X_2}: X_1\boxtimes _{\mathcal {C}}X_2\rightarrow X_1\boxtimes _A X_2\) to denote the cokernel morphism. Similar definitions and notation apply to objects in \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\). We would like to define morphisms \(F^A_{X_1,X_2}\) and \(G^A_{X_1,X_2}\) in \({\mathcal {D}}\) such that the diagrams

and

commute.

From the cokernel definitions of \({\mathcal {F}}(X_1)\boxtimes _A{\mathcal {F}}(X_2)\) and \(X_1\boxtimes _A X_2\) and by the right exactness of \({\mathcal {F}}\) (so that \({\mathcal {F}}(X_1\boxtimes _A X_2)\) is still a cokernel), the existence and uniqueness of \(F^A_{X_1,X_2}\) and \(G^A_{X_1,X_2}\) is equivalent to the identities

$$\begin{aligned} {\mathcal {F}}( \eta _{X_1,X_2})\circ F_{X_1,X_2}\circ \mu ^{(1)}&={\mathcal {F}}(\eta _{X_1,X_2})\circ F_{X_1,X_2}\circ \mu ^{(2)},\\ \eta _{{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\circ F^{-1}_{X_1,X_2}\circ {\mathcal {F}}(\mu ^{(1)})&=\eta _{{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\circ F^{-1}_{X_1,X_2}\circ {\mathcal {F}}(\mu ^{(2)}). \end{aligned}$$

When \({\mathcal {F}}\) is braid-reversing (the more interesting case), the first identity is proved as follows:

$$\begin{aligned}&{\mathcal {F}}( \eta _{X_1,X_2}) \circ F_{X_1,X_2}\circ (\textrm{Id}_{{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}\mu _{{\mathcal {F}}(X_2)})\circ {\mathcal {A}}_{{\mathcal {F}}(X_1),{\mathcal {F}}(A),{\mathcal {F}}(X_2)}^{-1}\circ ({\mathcal {R}}_{{\mathcal {F}}(A),{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)}) \\&\quad = {\mathcal {F}}(\eta _{X_1,X_2}) \circ F_{X_1,X_2}\circ (\textrm{Id}_{{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}{\mathcal {F}}(\mu _{X_2}))\circ (\textrm{Id}_{{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}F_{A,X_2})\circ \\&\quad \quad \circ {\mathcal {A}}_{{\mathcal {F}}(X_1),{\mathcal {F}}(A),{\mathcal {F}}(X_2)}^{-1}\circ ({\mathcal {R}}_{{\mathcal {F}}(A),{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)}) \\&\quad ={\mathcal {F}}(\eta _{X_1,X_2}\circ (\textrm{Id}_{X_1}\boxtimes _{\mathcal {C}}\mu _{X_2}))\circ F_{X_1,A\boxtimes _{\mathcal {C}}X_2}\circ (\textrm{Id}_{{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}F_{A,X_2})\circ \\&\quad \quad \circ {\mathcal {A}}_{{\mathcal {F}}(X_1),{\mathcal {F}}(A),{\mathcal {F}}(X_2)}^{-1}\circ ({\mathcal {R}}_{{\mathcal {F}}(A),{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)}) \\&\quad ={\mathcal {F}}(\eta _{X_1,X_2}\circ (\textrm{Id}_{X_1}\boxtimes _{\mathcal {C}}\mu _{X_2})\circ {\mathcal {A}}_{X_1,A,X_2}^{-1}\circ ({\mathcal {R}}^{-1}_{X_1,A}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2}))\circ \\&\quad \quad \circ F_{A\boxtimes _{\mathcal {C}}X_1,X_2}\circ (F_{A,X_1}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2}\circ (\textrm{Id}_{X_1}\boxtimes _{\mathcal {C}}\mu _{X_2})\circ {\mathcal {A}}_{X_1,A,X_2}^{-1}\circ ({\mathcal {R}}_{A,X_1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2}))\circ \\&\quad \quad \circ {\mathcal {F}}({\mathcal {M}}_{A,X_1}^{-1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2})\circ F_{A\boxtimes _{\mathcal {C}}X_1,X_2}\circ (F_{A,X_1}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2}\circ (\mu _{X_1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2})\circ ({\mathcal {M}}_{A,X_1}^{-1}\boxtimes \textrm{Id}_{X_2}))\circ F_{A\boxtimes _{\mathcal {C}}X_1,X_2}\circ (F_{A,X_1}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2})\circ F_{X_1,X_2}\circ ({\mathcal {F}}(\mu _{X_1})\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2})\circ (F_{A,X_1}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2})\circ F_{X_1,X_2}\circ \mu ^{(1)}, \end{aligned}$$

where the next to last equality uses the assumption that \(X_1\) is an object of \({{\,\textrm{Rep}\,}}^0A\). The second identity is proved similarly, so we get the desired morphism \(F^A_{X_1,X_2}\) in \({\mathcal {D}}\) with inverse \(G^A_{X_1,X_2}\).

To prove that \(F^A_{X_1,X_2}\) is an isomorphism in \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\), we use the definitions of \(\mu _{X_1\boxtimes _A X_2}\) and \(\mu _{{\mathcal {F}}(X_1)\boxtimes _{{\mathcal {F}}(A)}{\mathcal {F}}(X_2)}\) from [5, Section 2.3], along with compatibility of F with the associativity isomorphisms, to calculate

$$\begin{aligned}&\mu _{{\mathcal {F}}(X_1\boxtimes _A X_2)} \circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}F^A_{X_1,X_2})\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}\eta _{{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)})\\&\quad ={\mathcal {F}}(\mu _{X_1\boxtimes _A X_2})\circ F_{A,X_1\boxtimes _A X_2}\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}{\mathcal {F}}(\eta _{X_1,X_2}))\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}F_{X_1,X_2})\\&\quad ={\mathcal {F}}(\mu _{X_1\boxtimes _A X_2}\circ (\textrm{Id}_A\boxtimes _{\mathcal {C}}\eta _{X_1,X_2}))\circ F_{A,X_1\boxtimes _{\mathcal {C}}X_2}\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}F_{X_1,X_2})\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2})\circ {\mathcal {F}}(\mu _{X_1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2})\circ {\mathcal {F}}({\mathcal {A}}_{A,X_1,X_2})\circ F_{A,X_1\boxtimes _{\mathcal {C}}X_2}\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}F_{X_1,X_2})\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2})\circ {\mathcal {F}}(\mu _{X_1}\boxtimes _{\mathcal {C}}\textrm{Id}_{X_2})\circ F_{A\boxtimes _{\mathcal {C}}X_1,X_2}\circ (F_{A,X_1}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\circ {\mathcal {A}}_{{\mathcal {F}}(A),{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\\&\quad ={\mathcal {F}}(\eta _{X_1,X_2})\circ F_{X_1,X_2}\circ ({\mathcal {F}}(\mu _{X_1})\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\circ (F_{A,X_1}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\circ {\mathcal {A}}_{{\mathcal {F}}(A),{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\\&\quad = F^A_{X_1,X_2}\circ \eta _{{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\circ (\mu _{{\mathcal {F}}(X_1)}\boxtimes _{\mathcal {D}}\textrm{Id}_{{\mathcal {F}}(X_2)})\circ {\mathcal {A}}_{{\mathcal {F}}(A),{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\\&\quad = F^A_{X_1,X_2}\circ \mu _{{\mathcal {F}}(X_1)\boxtimes _{{\mathcal {F}}(A)}{\mathcal {F}}(X_2)}\circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}\eta _{{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}). \end{aligned}$$

Because \(\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}\eta _{{\mathcal {F}}(X_1),{\mathcal {F}}(X_2)}\) is surjective, it follows that

$$\begin{aligned} \mu _{{\mathcal {F}}(X_1\boxtimes _A X_2)} \circ (\textrm{Id}_{{\mathcal {F}}(A)}\boxtimes _{\mathcal {D}}F^A_{X_1,X_2}) =F^A_{X_1,X_2}\circ \mu _{{\mathcal {F}}(X_1)\boxtimes _{{\mathcal {F}}(A)}{\mathcal {F}}(X_2)} \end{aligned}$$

as required. The proof that \(F^A\) defines a natural transformation is also straightforward from the definitions in [5, Section 2.3] and the fact that F is a natural transformation.

To show that the natural isomorphism \(F^A\) gives \({\mathcal {F}}:{{\,\textrm{Rep}\,}}^0 A\rightarrow {{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\) the structure of a braided (or braid-reversing) tensor functor, we also need \(F^A\) to be compatible with the unit, associativity and braiding isomorphisms. Since the compatibility proofs are straightforward, we only discuss compatibility with the right unit and braiding isomorphisms in the case that \({\mathcal {F}}\) is braid-reversing. For the right unit isomorphisms, we need to show

$$\begin{aligned} {\mathcal {F}}(r^A_X)\circ F^A_{X,A}=r_{{\mathcal {F}}(X)}^{{\mathcal {F}}(A)} \end{aligned}$$

for any object X of \({{\,\textrm{Rep}\,}}^0 A\), where \(r^A\) and \(r^{{\mathcal {F}}(A)}\) are the right unit isomorphisms in \({{\,\textrm{Rep}\,}}^0 A\) and \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\), respectively. Indeed, from the definitions in [5, Section 2.3] and the assumption that X is an object of \({{\,\textrm{Rep}\,}}^0A\), we get

$$\begin{aligned} {\mathcal {F}}(r^A_X)\circ F^A_{X,A}\circ \eta _{{\mathcal {F}}(X),{\mathcal {F}}(A)}&= {\mathcal {F}}(r^A_X)\circ {\mathcal {F}}(\eta _{X,A})\circ F_{X,A} \\&={\mathcal {F}}(\mu _X)\circ {\mathcal {F}}({\mathcal {R}}_{A,X}^{-1})\circ F_{X,A} \\&={\mathcal {F}}(\mu _X)\circ {\mathcal {F}}({\mathcal {R}}_{X,A})\circ F_{X,A} \\&={\mathcal {F}}(\mu _X)\circ F_{A,X}\circ {\mathcal {R}}_{{\mathcal {F}}(A),{\mathcal {F}}(X)}^{-1} \\&=\mu _{{\mathcal {F}}(X)}\circ {\mathcal {R}}_{{\mathcal {F}}(A),{\mathcal {F}}(X)}^{-1} = r^{{\mathcal {F}}(A)}_{{\mathcal {F}}(X)}\circ \eta _{{\mathcal {F}}(X),{\mathcal {F}}(A)}. \end{aligned}$$

So the desired equality follows from surjectivity of \(\eta _{{\mathcal {F}}(X),{\mathcal {F}}(A)}\). Similarly, the braid-reversing properties of F and the definitions in [6, Section 2.6] show that

$$\begin{aligned} {\mathcal {F}}({\mathcal {R}}^A_{X_1,X_2})\circ F^A_{X_1,X_2}= F^A_{X_2,X_1}\circ {\mathcal {R}}^{-1}_{{\mathcal {F}}(X_2),{\mathcal {F}}(X_1)} \end{aligned}$$

for objects \(X_1\) and \(X_2\) of \({{\,\textrm{Rep}\,}}^0 A\), so that \({\mathcal {F}}\) defines a braid-reversed tensor functor from \({{\,\textrm{Rep}\,}}^0 A\) to \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\). \(\square \)

Since equivalences between abelian categories are automatically exact, we get the following corollary of Proposition A.1 and Theorem A.3:

Corollary A.4

Let \({\mathcal {F}}:{\mathcal {C}}\rightarrow {\mathcal {D}}\) be a braided, respectively braid-reversing, tensor equivalence. If A is a commutative \({\mathcal {C}}\)-algebra, then \({\mathcal {F}}\) induces a braided, respectively braid-reversing, tensor equivalence between \({{\,\textrm{Rep}\,}}^0 A\) and \({{\,\textrm{Rep}\,}}^0{\mathcal {F}}(A)\).

Appendix B: Uniqueness of \(L_1(\mathfrak {sl}_2)\)

In this appendix, we present some uniqueness results for the simple affine vertex operator algebra \(L_1(\mathfrak {sl}_2)\) associated to \(\mathfrak {sl}_2\) at level 1, and also for its non-trivial irreducible module. These results are perhaps known to experts, but we provide proofs here for completeness. The automorphism group of \(L_1(\mathfrak {sl}_2)\) is \(PSL_2({\mathbb {C}})\), with fixed-point subalgebra the Virasoro vertex operator algebra \(V_1\) [10, 46]. As a \(PSL_2({\mathbb {C}})\times V_1\)-module,

$$\begin{aligned} L_1(\mathfrak {sl}_2)\cong \bigoplus _{n=0}^\infty V(2n)\otimes {\mathcal {L}}_{2n+1,1}. \end{aligned}$$
(B.1)

We first establish the uniqueness of a simple vertex operator algebra with this decomposition as a \(V_1\)-module, while ignoring \(PSL_2({\mathbb {C}})\)-actions:

Theorem B.1

If V is a simple vertex operator algebra extension of \(V_1\) such that \(V\cong \bigoplus _{n=0}^\infty V(2n)\otimes {\mathcal {L}}_{2n+1,1}\) as \(V_1\)-modules (with V(2n) considered just as a \((2n+1)\)-dimensional vector space), then \(V\cong L_1(\mathfrak {sl}_2)\) as vertex operator algebras.

Proof

For \(n\in {\mathbb {N}}\), we use \(v_{2n+1,1}\) to denote a generating vector of \({\mathcal {L}}_{2n+1,1}\) of minimal conformal weight, and we use \(\pi _{2n+1}: V\rightarrow V(2n)\otimes {\mathcal {L}}_{2n+1,1}\) to denote the \({\mathcal {V}}ir\)-module projection.

By assumption, the conformal-weight-1 space \(V_{(1)}=V(2)\otimes v_{3,1}\) is three-dimensional, and as in [17, Remark 8.9.1] it is a Lie algebra with bracket \([u,v]=u_0 v\) and invariant symmetric bilinear form \(\langle \cdot ,\cdot \rangle \) defined by \(u_1 v=\langle u,v\rangle {\textbf{1}}\) for \(u,v\in V_{(1)}\). We claim that \(\langle \cdot ,\cdot \rangle \) is non-degenerate. Indeed, consider \(a\otimes v_{3,1}, b\otimes v_{3,1}\in V_{(1)}\) for non-zero \(a,b\in V(2)\). Then we have an identification of \(V_1\)-module intertwining operators

$$\begin{aligned} \pi _1\circ Y_V\vert _{(a\otimes {\mathcal {L}}_{3,1})\otimes (b\otimes {\mathcal {L}}_{3,1})} =\langle a\otimes v_{3,1},b\otimes v_{3,1}\rangle {\mathcal {Y}}_{33}^1 \end{aligned}$$

where \({\mathcal {Y}}_{33}^1\) is the unique \(V_1\)-module intertwining operator of type \(\left( {\begin{array}{c}{\mathcal {L}}_{1,1}\\ {\mathcal {L}}_{3,1}\,{\mathcal {L}}_{3,1}\end{array}}\right) \) such that

$$\begin{aligned} {\mathcal {Y}}_{33}^1(v_{3,1},x)v_{3,1}\in x^{-2}({\textbf{1}}+x{\mathcal {L}}_{1,1}[[x]]). \end{aligned}$$

Since there is a non-zero intertwining operator of type \(\left( {\begin{array}{c}{\mathcal {L}}_{1,1}\\ {\mathcal {L}}_{r,1}\,{\mathcal {L}}_{3,1}\end{array}}\right) \) only for \(r=3\), it follows that if \(\langle a\otimes v_{3,1}, b\otimes v_{3,1}\rangle =0\) for all \(a\in V(2)\), then the V-submodule generated by \(b\otimes v_{3,1}\) is contained in \(\bigoplus _{n= 1}^\infty V(2n)\otimes {\mathcal {L}}_{2n+1,1}\). Since V is simple, this is impossible for \(b\ne 0\), and thus for such b there exists \(a\in V(2)\) such that \(\langle a\otimes v_{3,1}, b\otimes v_{3,1}\rangle \ne 0\). This proves that \(\langle \cdot ,\cdot \rangle \) is non-degenerate.

Since \(\langle \cdot ,\cdot \rangle \) is non-degenerate, we can take \(h\in V_{(1)}\) such that \(\langle h,h\rangle =1\). We would like to show that \(\omega =L_{-2}{\textbf{1}}\) is a multiple of \(h_{-1}^2{\textbf{1}}\). Because \(h_{2n+1,1}=n^2>2\) if \(n>1\), we have

$$\begin{aligned} h_{-1}^2{\textbf{1}}=c\cdot \omega +L_{-1}u \end{aligned}$$

for some \(c\in {\mathbb {C}}\) and \(u\in V_{(1)}\). We first claim that \(u=0\). To show this, observe that vertex operator modes acting on V satisfy

$$\begin{aligned} 2\sum _{m=0}^\infty h_{-m}h_{m+1}&=\textrm{Res}_x\,x^2 Y_V(h_{-1}^2{\textbf{1}},x) \\&=c\cdot \textrm{Res}_x\,x^2 Y_V(\omega ,x)+\textrm{Res}_x\,x^2\frac{d}{dx}Y_V(u,x) =c\cdot L_1-2 u_1. \end{aligned}$$

The left side of this equation annihilates \(V_{(1)}\) because

$$\begin{aligned} h_0h_1 v =\langle h,v\rangle h_0{\textbf{1}}=0 \end{aligned}$$

for all \(v\in V_{(1)}\), and \(L_1\) annihilates \(V_{(1)}\) as well since \(V_{(1)}\) consists of Virasoro primary vectors. Thus \(u_1\) also annihilates \(V_{(1)}\), that is, \(\langle u,v\rangle =0\) for all \(v\in V_{(1)}\). Since \(\langle \cdot ,\cdot \rangle \) is non-degenerate, this proves the claim.

We now have \(h_{-1}^2{\textbf{1}}=c\cdot \omega \), and it remains to calculate c. Let \((\cdot ,\cdot )\) be the unique non-degenerate invariant bilinear form on \({\mathcal {L}}_{1,1}\subseteq V\) such that \(({\textbf{1}},{\textbf{1}})=1\). Thus

$$\begin{aligned} (L_{-2}{\textbf{1}},L_{-2}{\textbf{1}})=({\textbf{1}},L_2 L_{-2}{\textbf{1}})=\frac{1}{2}, \end{aligned}$$

so that using (2.3) and the \(L_{-1}\)-derivative property,

$$\begin{aligned} \frac{c}{2}&= (L_{-2}{{\textbf {1}}},c\cdot {L}_{-2}{{\textbf {1}}}) =(L_{-2}{{\textbf {1}}},h_{-1}^2{{\textbf {1}}})\\ {}&=\text {Res}_x\,x^{-1}(L_{-2}{{\textbf {1}}},(\pi _1\circ Y_V)(h,x)h) =\text {Res}_x\,x^{-1}({{\textbf {1}}},L_2(\pi _1\circ Y_V)(h,x)h\rangle \\ {}&=\text {Res}\,x\left( {{\textbf {1}}}, x\frac{d}{dx} (\pi _1\circ Y_V)(h,x)h+3(\pi _1\circ Y_V)(L_0 h,x)h\right) \\ {}&=({{\textbf {1}}}, -2h_1 h+3h_1 h) =\langle h,h\rangle ({{\textbf {1}}},{{\textbf {1}}})=1. \end{aligned}$$

Thus \(c=2\) and we get \(\omega =\frac{1}{2}h_{-1}^2{\textbf{1}}\).

We will now use [41, Corollary 3.15] to show that V is isomorphic to the lattice vertex operator algebra \(V_L\) where L consists of all \(\alpha \in {\mathbb {C}}\) such that for some non-zero \(v\in V\), \(h_0 v=\alpha v\). To apply this result, we need to verify Conditions (1) – (3) of [41]: Condition (1) is satisfied because

$$\begin{aligned} L_n h=\delta _{n,0} h,\qquad h_n h=\delta _{n,1}{\textbf{1}}\end{aligned}$$

for \(n\ge 0\). Condition (2) is that the algebra generated by the operators \(h_n\), \(n\ge 0\), acts locally finitely on V. This is clear because V is a vertex operator algebra. Condition (3) is that \(G_0 ={\mathbb {C}}{\textbf{1}}\), where \(G_0\) is the intersection of the generalized \(h_0\)-eigenspace with generalized eigenvalue 0 with the Heisenberg vacuum space

$$\begin{aligned} G=\lbrace v\in V\,\vert \,h_nv =0\,\,\text {for}\,\,n>0\rbrace . \end{aligned}$$

Indeed, if \(v\in G\), then \(L_0 v=\frac{1}{2} h_0^2 v\), so if v is additionally a generalized eigenvector for h with generalized eigenvalue 0, then \(L_0^N v=0\) for some \(N\in {\mathbb {Z}}_+\). Thus v has conformal weight 0 and it follows that \(G_0={\mathbb {C}}{\textbf{1}}\).

It now follows from [41, Corollary 3.15] that V is isomorphic to a rank-one lattice vertex operator algebra \(V_L\). Since V has the same character as the \(\mathfrak {sl}_2\)-root lattice vertex operator algebra, which in turn is isomorphic to \(L_1(\mathfrak {sl}_2)\), it follows that V is isomorphic to \(L_1(\mathfrak {sl}_2)\). \(\square \)

We now consider what happens when a simple vertex operator algebra with decomposition (B.1) does have a \(PSL_2({\mathbb {C}})\)-action:

Theorem B.2

Suppose V is a simple vertex operator algebra extension of \(V_1\) with an action of \(PSL_2({\mathbb {C}})\) by vertex operator algebra automorphisms such that \(V\cong \bigoplus _{n=0}^\infty V(2n)\otimes {\mathcal {L}}_{2n+1,1}\) as a \(PSL_2({\mathbb {C}})\times V_1\)-module. Then the vertex operator algebra isomorphism \(L_1(\mathfrak {sl}_2)\rightarrow V\) guaranteed by Theorem B.1 can be chosen to be an isomorphism of \(PSL_2({\mathbb {C}})\)-modules.

Proof

By assumption, the conformal-weight-1 space \(V_{(1)}=V(2)\otimes v_{3,1}\) is a \(PSL_2({\mathbb {C}})\)-module isomorphic to \(\mathfrak {sl}_2\), so we may fix a standard basis \(\lbrace e, f, h\rbrace \) for \(V_{(1)}\) as a \(PSL_2({\mathbb {C}})\)-module. Moreover, because \(PSL_2({\mathbb {C}})\) acts on V by vertex operator algebra automorphisms, the map \(V_{(1)}\otimes V_{(1)} \rightarrow V_{(1)}\) given by \(u\otimes v\mapsto u_0 v\) is a \(PSL_2({\mathbb {C}})\)-module homomorphism, and thus it must be a multiple of the Lie bracket on \(\mathfrak {sl}_2\), that is,

$$\begin{aligned} e_0 f= c\cdot h,\qquad f_0 h=2c\cdot f,\qquad h_0 e=2c\cdot e \end{aligned}$$
(B.2)

for some \(c\in {\mathbb {C}}\). Since \(V_{(1)}\) is also a Lie algebra with bracket \([u,v]=u_0 v\), and since this Lie algebra is isomorphic to \(\mathfrak {sl}_2\) by Theorem B.1, we have \(c\ne 0\) (but note that we cannot immediately identify the \(\lbrace e, f, h\rbrace \) basis for \(V_{(1)}\) considered as a \(PSL_2({\mathbb {C}})\)-module with the corresponding basis for \(V_{(1)}\) as a Lie algebra).

We can now replace the vertex operator \(Y_{V}\) with \(\varphi ^{-1}\circ Y_{V}\circ (\varphi \otimes \varphi )\), where \(\varphi \) is the linear isomorphism that acts as the identity on \(V(2n)\otimes {\mathcal {L}}_{2n+1,1}\) for \(n\ne 1\) and by the scalar c on \(V(2)\otimes {\mathcal {L}}_{3,1}\). Note that

$$\begin{aligned} \varphi : (V,\varphi ^{-1}\circ Y_{V}\circ (\varphi \otimes \varphi ),{\textbf{1}},\omega ) \longrightarrow (V, Y_{V}, {\textbf{1}},\omega ) \end{aligned}$$

is a vertex operator algebra isomorphism that is also a \(PSL_2({\mathbb {C}})\)-module isomorphism. Thus we may assume that the simple vertex operator algebra structure on V satisfies (B.2) with \(c=1\). As in the proof of Theorem B.1, \((u,v)\rightarrow u_1 v\) defines a non-degenerate invariant bilinear form on \(V_{(1)}\cong \mathfrak {sl}_2\):

$$\begin{aligned} u_1 v=k\langle u,v\rangle {\textbf{1}}\end{aligned}$$

for some non-zero \(k\in {\mathbb {C}}\), where \(\langle \cdot ,\cdot \rangle \) is the invariant bilinear form on \(\mathfrak {sl}_2\) such that \(\langle h,h\rangle =2\). Thus \(h_1 h=2k{\textbf{1}}\), so by the proof of Theorem B.1, \(\omega =\frac{1}{4k} h_{-1}^2{\textbf{1}}\), and then

$$\begin{aligned} L_0 =\frac{1}{4k}h_0^2 +\frac{1}{2k}\sum _{n=1}^\infty h_{-n} h_n. \end{aligned}$$

Using \(h_0 e=2e\), \(h_1 e=k\langle h,e\rangle {\textbf{1}}=0\), and \(h_n e=0\) for \(n\ge 2\), we get \(e=L_0 e=\frac{1}{k}e\), and thus \(k=1\).

We can now conclude that there is a vertex algebra homomorphism from the universal affine vertex algebra \(V^1(\mathfrak {sl}_2)\) at level 1 to V which sends the generators \(e(-1){\textbf{1}}\), \(f(-1){\textbf{1}}\), and \(h(-1){\textbf{1}}\) of \(V^1(\mathfrak {sl}_2)\) to e, f, and h, respectively. Since V is simple and generated by e, f, and h (because V is abstractly isomorphic to \(L_1(\mathfrak {sl}_2)\)), this homomorphism descends to a vertex operator algebra isomorphism \(L_1(\mathfrak {sl}_2)\rightarrow V\). By construction, this isomorphism commutes with the \(PSL_2({\mathbb {C}})\)-actions on the weight-one subspaces of \(L_1(\mathfrak {sl}_2)\) and V. Then because both algebras are generated by their weight-one subspaces, and because \(PSL_2({\mathbb {C}})\) acts by automorphisms on both algebras, the isomorphism \(L_1(\mathfrak {sl}_2)\rightarrow V\) is an isomorphism of \(PSL_2({\mathbb {C}})\)-modules, as desired. \(\square \)

The vertex operator algebra \(L_1(\mathfrak {sl}_2)\), has a unique (up to isomorphism) irreducible module which is not isomorphic to \(L_1(\mathfrak {sl}_2)\). We denote this module by X. The action of \(\mathfrak {sl}_2\) on X by zero-modes of the vertex operators for elements of \(L_1(\mathfrak {sl}_2)_{(1)}\cong \mathfrak {sl}_2\) exponentiates to an action of \(SL_2({\mathbb {C}})\) on X such that

$$\begin{aligned} g\cdot Y_{X}(v,x)w = Y_{X}(g\cdot v,x)g\cdot w \end{aligned}$$
(B.3)

for \(g\in SL_2({\mathbb {C}})\), \(v\in L_1(\mathfrak {sl}_2)\), and \(w\in X\). In particular, the \(SL_2({\mathbb {C}})\)-action and \(V_1\)-action on X commute; we have

$$\begin{aligned} X=\bigoplus _{n=1}^{\infty } V(2n-1)\otimes {\mathcal {L}}_{2n,1} \end{aligned}$$

as an \(SL_2({\mathbb {C}})\times V_1\)-module. We now have a uniqueness result for the \(SL_2({\mathbb {C}})\)-action on X:

Theorem B.3

Suppose W is a simple \(L_1(\mathfrak {sl}_2)\)-module with an \(SL_2({\mathbb {C}})\)-action such that

$$\begin{aligned} g\cdot Y_W(v,x)w= Y_W(g\cdot v,x)g\cdot w \end{aligned}$$
(B.4)

for all \(g\in SL_2({\mathbb {C}})\), \(v\in L_1(\mathfrak {sl}_2)\), and \(w\in W\), and such that \(W\cong \bigoplus _{n=1}^\infty V(2n-1)\otimes {\mathcal {L}}_{2n,1}\) as an \(SL_2({\mathbb {C}})\times V_1\)-module. Then there is an \(L_1(\mathfrak {sl}_2)\)-module isomorphism \(X\rightarrow W\) which is also an \(SL_2({\mathbb {C}})\)-module isomorphism.

Proof

Let T(W) denote the top level of W, that is, the lowest conformal weight space \(V(1)\otimes v_{2,1}\). The assumed compatibility of the \(SL_2({\mathbb {C}})\)- and \(L_1(\mathfrak {sl}_2)\)-actions on W implies that the linear map \(L_1(\mathfrak {sl}_2)_{(1)}\otimes T(W)\rightarrow T(W)\) give by \(v\otimes w\mapsto v_0 w\) is an \(SL_2({\mathbb {C}})\)-module homomorphism, and thus must be a multiple of the \(\mathfrak {sl}_2\)-action on V(1). That is, if \(\lbrace e, f, h\rbrace \) is the standard basis of \(L_1(\mathfrak {sl}_2)_{(1)}\cong \mathfrak {sl}_2\), then there is a basis \(\lbrace v_1, v_{-1}\rbrace \) of T(W) such that

$$\begin{aligned} \begin{array}{lll} e_0 v_1 =0, &{} h_0 v_1 =c\cdot v_1, &{} f_0 v_1=c\cdot v_{-1},\\ e_0 v_{-1} = c\cdot v_1, &{} h_0 v_{-1} =-c\cdot v_{-1}, &{} f_0 v_{-1} =0\\ \end{array} \end{aligned}$$
(B.5)

for some \(c\in {\mathbb {C}}\).

On the other hand, the axioms for vertex operator algebras and modules show that the action of \(L_1(\mathfrak {sl}_2)_{(1)}\) on T(W) by zero-modes gives T(W) the structure of an \(\mathfrak {sl}_2\)-module. The formulas (B.5) describe an \(\mathfrak {sl}_2\)-module structure only when \(c=0\) or \(c=1\), since if \(c\ne 0\), then \(v_1\) is the unique (up to scale) highest-weight vector in T(W) and thus must have h-eigenvalue 1 as T(W) is two-dimensional. In fact, we may conclude \(c\ne 0\) because the classification of simple \(L_1(\mathfrak {sl}_2)\)-modules shows that W must be isomorphic to X as an \(L_1(\mathfrak {sl}_2)\)-module, and we know that \(L_1(\mathfrak {sl}_2)_{(1)}\) acts non-trivially on the top level of X. Thus (B.5) holds with \(c=1\), showing that the assumed \(SL_2({\mathbb {C}})\)-action on T(W) is the same as that obtained by exponentiating the zero-mode action of \(L_1(\mathfrak {sl}_2)_{(1)}\cong \mathfrak {sl}_2\) on T(W).

We have now shown that any \(L_1(\mathfrak {sl}_2)\)-module isomorphism \(X\rightarrow W\) restricts to an \(SL_2({\mathbb {C}})\)-module isomorphism on top levels, since the \(SL_2({\mathbb {C}})\)-actions on both top levels are obtained by exponentiating the zero-mode actions of \(L_1(\mathfrak {sl}_2)_{(1)}\). Then because both X and W are generated by their top levels, and because (B.3) and (B.4) both hold, we conclude that any \(L_1(\mathfrak {sl}_2)\)-module isomorphism \(X\rightarrow W\) is also an \(SL_2({\mathbb {C}})\)-module isomorphism. \(\square \)

Appendix C: Uniqueness of the chiral universal centralizer

In this appendix, we prove the uniqueness assertion in Theorem 8.1 using Theorems B.2 and B.3. Thus let V be any simple conformal vertex algebra containing \(V_1\otimes V_{25}\) as a vertex operator subalgebra and such that

$$\begin{aligned} V \cong \bigoplus _{r\in {\mathbb {Z}}_+} {\mathcal {L}}_{r,1}^{(1)}\otimes {\mathcal {L}}_{r,1}^{(25)} \end{aligned}$$

as a \(V_1\otimes V_{25}\)-module. The \(\mathfrak {sl}_2\)-type fusion rules for \(V_1\)- and \(V_{25}\)-modules imply that V has a conformal vertex algebra involution which acts as \((-1)^{r-1}\) on \({\mathcal {L}}_{r,1}^{(1)}\otimes {\mathcal {L}}_{r,1}^{(25)}\). Let \(V^0\) be the fixed-point subalgebra of this involution and \(V^1\) the eigenspace with eigenvalue \(-1\), so that

$$\begin{aligned} V^0 \cong \bigoplus _{n=0}^\infty {\mathcal {L}}_{2n+1,1}^{(1)}\otimes {\mathcal {L}}_{2n+1,1}^{(25)},\qquad V^1\cong \bigoplus _{n = 1}^\infty {\mathcal {L}}_{2n,1}^{(1)}\otimes {\mathcal {L}}_{2n,1}^{(25)} \end{aligned}$$

as \(V_1\otimes V_{25}\)-modules. It is easy to see that \(V^0\) is a simple conformal vertex algebra as follows: For any non-zero \(v\in V^0\),

$$\begin{aligned} V =\textrm{span}\lbrace u_n v\,\vert \,u\in V, n\in {\mathbb {Z}}\rbrace = \textrm{span}\lbrace u_n v\,\vert \,u\in V^0, n\in {\mathbb {Z}}\rbrace +\textrm{span}\lbrace u_n v\,\vert \,u\in V^1, n\in {\mathbb {Z}}\rbrace \end{aligned}$$

by [39, Proposition 4.5.6] since V is simple. But since \(V=V^0\oplus V^1\) and since

$$\begin{aligned} \textrm{span}\lbrace u_n v\,\vert \,u\in V^i, n\in {\mathbb {Z}}\rbrace \subseteq V^i \end{aligned}$$

for \(i=0,1\), it follows that

$$\begin{aligned} V^0=\textrm{span}\lbrace u_n v\,\vert \,u\in V^0, n\in {\mathbb {Z}}\rbrace \end{aligned}$$

for any non-zero \(v\in V^0\), and thus \(V^0\) is simple. A similar argument shows that \(V^1\) is a simple \(V^0\)-module.

Proposition C.1

Any simple conformal vertex algebra extension of \(V_1\otimes V_{25}\) which is isomorphic to \(V^0\) as a \(V_1\otimes V_{25}\)-module is isomorphic to \(V^0\) as a conformal vertex algebra. Moreover, any simple \(V^0\)-module which is isomorphic to \(V^1\) as a \(V_1\otimes V_{25}\)-module is isomorphic to \(V^1\) as a \(V^0\)-module.

Proof

From Proposition 6.1, \({\mathcal {O}}_{25}^0\) is braided tensor equivalent to \(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \) with a suitable braiding. Thus using [7, Theorem 7.5], uniqueness of the simple conformal vertex algebra structure on \(V^0=\bigoplus _{n=0}^\infty {\mathcal {L}}_{2n+1,1}^{(1)}\otimes {\mathcal {L}}_{2n+1,1}^{(25)}\) extending \(V_1\otimes V_{25}\) is equivalent to uniqueness of the simple commutative algebra structure on the object

$$\begin{aligned} A^0=\bigoplus _{n=0}^\infty V(2n)\otimes {\mathcal {L}}_{2n+1,1}^{(1)} \end{aligned}$$

in the Ind-category \(\textrm{Ind}(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \otimes {\mathcal {O}}_1^0)\). Moreover, uniqueness of the simple \(V^0\)-module structure on \(V^1\) is equivalent to uniqueness of the simple \(A^0\)-module structure on the object

$$\begin{aligned} A^1 =\bigoplus _{n=1}^\infty V(2n-1)\otimes {\mathcal {L}}_{2n,1}^{(1)} \end{aligned}$$

in \(\textrm{Ind}(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \otimes {\mathcal {O}}_{1}^0)\).

Since the \(\mathfrak {sl}_2\)-modules V(2n) for \(n\in {\mathbb {N}}\) are objects of the maximal symmetric tensor subcategory \({{\,\textrm{Rep}\,}}PSL_2({\mathbb {C}})\subseteq ({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \), we can use [7, Theorem 7.5] again to see that a simple commutative algebra structure on \(A^0\) in \(\textrm{Ind}(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \otimes {\mathcal {O}}_1^0)\) amounts to a vertex operator algebra structure \((A^0, Y_{A^0},{\textbf{1}},\omega )\) such that:

  1. (1)

    The \(PSL_2({\mathbb {C}})\)-action on the V(2n) factors in \(A^0\) gives a \(PSL_2({\mathbb {C}})\)-action by vertex operator algebra automorphisms on \(A^0\).

  2. (2)

    The only \(PSL_2({\mathbb {C}})\)-invariant ideals of \(A^0\) are 0 and \(A^0\).

Moreover, uniqueness of the simple commutative algebra structure on \(A^0\) in \(\textrm{Ind}(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \otimes {\mathcal {O}}_1^0)\) amounts to uniqueness of the vertex operator algebra structure on \(A^0\) up to an isomorphism that is also a \(PSL_2({\mathbb {C}})\)-module isomorphism.

Similarly, because \(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \) is equivalent to \({{\,\textrm{Rep}\,}}\mathfrak {sl}_2\) when considered as a module category for \({{\,\textrm{Rep}\,}}PSL_2({\mathbb {C}})\), a simple \(A^0\)-module structure on \(A^1\) (where \(A^0\) is considered as a commutative algebra in \(\textrm{Ind}(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \otimes {\mathcal {O}}_1^0)\)) is equivalent to an \(A^0\)-module structure \((A^1, Y_{A^1})\) (where \(A^0\) is considered as a vertex operator algebra) such that:

  1. (3)

    The \(SL_2({\mathbb {C}})\)-action on the \(V(2n-1)\) factors in \(A^1\) and the \(PSL_2({\mathbb {C}})\)-action on \(A^0\) are compatible in the sense that

    $$\begin{aligned} g\cdot Y_{A^1}(v,x)w =Y_{A^1}(g\cdot v,x)g\cdot w \end{aligned}$$

    for all \(g\in SL_2({\mathbb {C}})\), \(v\in A^0\), and \(w\in A^1\).

  2. (4)

    The only \(SL_2({\mathbb {C}})\)-invariant \(A^0\)-submodules of \(A^1\) are 0 and \(A^1\).

Moreover, uniqueness of the \(A^0\)-module structure on \(A^1\) (where \(A^1\) is considered as a commutative algebra in \(\textrm{Ind}(({{\,\textrm{Rep}\,}}\mathfrak {sl}_2)^\tau \otimes {\mathcal {O}}_1^0)\)) is equivalent to uniqueness of the \(A^0\)-module structure on \(A^1\) (where \(A^0\) is considered as a vertex operator algebra) up to an \(A^0\)-module isomorphism that is also an \(SL_2({\mathbb {C}})\)-module isomorphism.

We claim that Conditions (1) and (2) imply that \(A^0\) is a simple vertex operator algebra, and we claim that \(A^1\) is a simple \(A^0\)-module. Then the desired uniqueness assertions will follow from Theorems B.2 and B.3: \(A^0\) will be isomoprhic to the simple affine vertex operator algebra \(L_1(\mathfrak {sl}_2)\) with its standard \(PSL_2({\mathbb {C}})\)-action by automorphisms, and \(A^1\) will be isomorphic to the unique non-trivial simple \(L_1(\mathfrak {sl}_2)\)-module with its standard \(SL_2({\mathbb {C}})\)-action.

To prove the claims, first let \(I\subseteq A^0\) be a non-zero ideal. Since \(A^0\) is a semisimple \({\mathcal {V}}ir\)-module, so is I, and thus I contains \(v\otimes {\mathcal {L}}_{2n+1,1}^{(1)}\) for some \(n\in {\mathbb {N}}\) and non-zero \(v\in V(2n)\). Now consider the ideal generated by \(V(2n)\otimes {\mathcal {L}}_{2n+1,1}^{(1)}\); it is \(PSL_2({\mathbb {C}})\)-invariant by Condition (1), and thus it equals \(A^0\) by Condition (2). In particular, the ideal generated by \(V(2n)\otimes {\mathcal {L}}_{2n+1,1}^{(1)}\) contains \(V(0)\otimes {\mathcal {L}}_{1,1}^{(1)}\). Since there is a non-zero \(PSL_2({\mathbb {C}})\)-module homomorphism \(V(2m)\otimes V(2n)\rightarrow V(0)\) if and only if \(m=n\), it follows that the Virasoro intertwining operator

$$\begin{aligned} \pi _1\circ Y_{A^0}\vert _{(V(2n)\otimes {\mathcal {L}}_{2n+1,1}^{(1)})\otimes (V(2n)\otimes {\mathcal {L}}_{2n+1,1}^{(1)})}, \end{aligned}$$

where \(\pi _1: A^0\rightarrow V(0)\otimes {\mathcal {L}}_{1,1}^{(1)}\) is the projection, is non-zero. Thus this intertwining operator has the form \(b\otimes {\mathcal {Y}}\) where \(b: V(2n)\otimes V(2n)\rightarrow V(0)\) is a non-zero (and thus non-degenerate) invariant bilinear form and \({\mathcal {Y}}\) is a non-zero intertwining operator of type \(\left( {\begin{array}{c}{\mathcal {L}}_{1,1}^{(1)}\\ {\mathcal {L}}_{2n+1,1}^{(1)}\,{\mathcal {L}}_{2n+1,1}^{(1)}\end{array}}\right) \).

Now recall that our non-zero ideal \(I\subseteq A^0\) contains \(v\otimes {\mathcal {L}}_{2n+1,1}^{(1)}\) for some non-zero \(v\in V(2n)\). Thus letting \(v_{2n+1,1}\) denote a non-zero lowest-conformal-weight vector in \({\mathcal {L}}_{2n+1,1}^{(1)}\), we have now shown that there is some \(v'\in V(2n)\) such that

$$\begin{aligned} Y_{A^0}(v'\otimes v_{2n+1,1},x)(v\otimes v_{2n+1,1})\in I((x)) \end{aligned}$$

has non-zero projection to \(V(0)\otimes {\mathcal {L}}_{1,1}^{(1)}\). Then because I is a semisimple \({\mathcal {V}}ir\)-module, this shows that I contains \(V(0)\otimes {\mathcal {L}}_{1,1}^{(1)}\) and thus contains the vacuum vector \({\textbf{1}}\). Since \({\textbf{1}}\) generates \(A^0\) as an \(A^0\)-module, we conclude that \(I=A^0\), proving \(A^0\) is simple.

To show that \(A^1\) is a simple \(A^0\)-module, we can now use Theorem B.1, which shows that \(A^0\) is isomorphic to \(L_1(\mathfrak {sl}_2)\) as a vertex operator algebra. Then \(A^1\) must be simple because the only \(L_1(\mathfrak {sl}_2)\)-module with the same \(V_1\)-module decomposition as \(A^1\) is the non-trivial simple \(L_1(\mathfrak {sl}_2)\)-module. \(\square \)

We can now prove uniqueness of the simple conformal vertex algebra structure on \(V^0\oplus V^1\), and thus the uniqueness assertion in Theorem 8.1:

Proof

Let \((V,Y_V,{\textbf{1}},\omega )\) and \((V,{\widetilde{Y}}_V,{\textbf{1}},\omega )\) be two simple conformal vertex algebra structures on V such that

$$\begin{aligned} Y_V\vert _{({\mathcal {L}}^{(1)}_{1,1}\otimes {\mathcal {L}}_{1,1}^{(25)})\otimes V} ={\widetilde{Y}}_V\vert _{({\mathcal {L}}^{(1)}_{1,1}\otimes {\mathcal {L}}_{1,1}^{(25)})\otimes V}. \end{aligned}$$

Since \(V=V^0\oplus V^1\) where \(V^0\) is a simple subalgebra and \(V^1\) is a simple \(V^0\)-module with respect to either conformal vertex algebra structure on V, Proposition C.1 shows that we may assume

$$\begin{aligned} Y_V\vert _{V^0\otimes V} = {\widetilde{Y}}_V\vert _{V^0\otimes V}. \end{aligned}$$

Then skew-symmetry dictates that

$$\begin{aligned} Y_V\vert _{V^1\otimes V^0}=e^{xL_{-1}}Y_V(\cdot ,-x)\cdot \vert _{V^0\otimes V^1} =e^{xL_{-1}}{\widetilde{Y}}_V(\cdot ,-x)\cdot \vert _{V^0\otimes V^1}={\widetilde{Y}}_V\vert _{V^1\otimes V^0}, \end{aligned}$$

so it remains to consider \({\widetilde{Y}}_V\vert _{V^1\otimes V^1}\).

The \(\mathfrak {sl}_2\)-type fusion rules of \(V_1\)- and \(V_{25}\)-modules show that \(\textrm{Im}\,{\widetilde{Y}}_V\vert _{V^1\otimes V^1}\subseteq V^0\). We claim the space of \(V^0\)-module intertwining operators of type \(\left( {\begin{array}{c}V^0\\ V^1\,V^1\end{array}}\right) \) is one dimensional, so that

$$\begin{aligned} {\widetilde{Y}}_V\vert _{V^1\otimes V^1}=\alpha \cdot Y_V\vert _{V^1\otimes V^1} \end{aligned}$$

for some \(\alpha \in {\mathbb {C}}\) (note that \(Y_V\vert _{V^1\otimes V^1}\ne 0\) since otherwise \(V^1\) would be a proper ideal of \((V,Y_V,{\textbf{1}},\omega )\)). Assuming the claim, we have \(\alpha \ne 0\) since \((V,{\widetilde{Y}}_V,{\textbf{1}},\omega )\) is a simple conformal vertex algebra, so then

$$\begin{aligned} \sigma :=\textrm{Id}_{V^0}\oplus \sqrt{\alpha }\cdot \textrm{Id}_{V^1} \end{aligned}$$

is a linear isomorphism of V such that \(\sigma \circ {\widetilde{Y}}_V=Y_V\circ (\sigma \otimes \sigma )\), showing that the two simple conformal vertex algebra structures on V are isomorphic (this is a special case of the uniqueness proof for simple current extensions in [12, Proposition 5.3]).

To prove the claim, let \(q_r: {\mathcal {L}}_{r,1}^{(1)}\otimes {\mathcal {L}}_{r,1}^{(25)}\rightarrow V\) and \(\pi _r: V\rightarrow {\mathcal {L}}_{r,1}^{(1)}\otimes {\mathcal {L}}_{r,1}^{(25)}\) for \(r\in {\mathbb {Z}}_+\) denote the natural inclusion and projection, respectively. It is enough to show that the linear map

$$\begin{aligned} {\mathcal {Y}}\mapsto \pi _1\circ {\mathcal {Y}}\circ (q_{2}\otimes q_{2}) \end{aligned}$$

from the space of \(V^0\)-module intertwining operators of type \(\left( {\begin{array}{c}V^0\\ V^1\,V^1\end{array}}\right) \) to the one-dimensional space of \(V_1\otimes V_{25}\)-module intertwining operators of type \(\left( {\begin{array}{c}{\mathcal {L}}_{1,1}^{(1)}\otimes {\mathcal {L}}_{1,1}^{(25)}\\ {\mathcal {L}}_{2,1}^{(1)}\otimes {\mathcal {L}}_{2,1}^{(25)}\,{\mathcal {L}}_{2,1}^{(1)}\otimes {\mathcal {L}}_{2,1}^{(25)}\end{array}}\right) \) is injective. Indeed, let

$$\begin{aligned} {\mathcal {Y}}: V^1\otimes V^1&\rightarrow V^0((x))\\ u_1\otimes v_1&\mapsto {\mathcal {Y}}(u_1,x)v_1=\sum _{n\in {\mathbb {Z}}} (u_1)_n v_1\,x^{-n-1} \end{aligned}$$

be a non-zero intertwining operator, and fix a non-zero \(v_1\in {\mathcal {L}}_{2,1}^{(1)}\otimes {\mathcal {L}}_{2,1}^{(25)}\). Then

$$\begin{aligned} \textrm{span}\lbrace (u_1)_n v_1\,\vert \,u_1\in V^1, n\in {\mathbb {Z}}\rbrace \end{aligned}$$

is a \(V^0\)-submodule of \(V^0\) by the easy intertwining operator generalization of [39, Proposition 4.5.7]. Using the commutator formula for intertwining operators, we see that this submodule is non-zero, since \(v_1\) generates \(V^1\) as a \(V^0\)-module and \({\mathcal {Y}}\) is non-zero (see [11, Proposition 11.9]). So in fact

$$\begin{aligned} {\mathcal {L}}_{1,1}^{(1)}\otimes {\mathcal {L}}_{1,1}^{(25)} =\pi _1(V^0) =\pi _1(\textrm{span}\lbrace (u_1)_n v_1\,\vert \,u_1\in V^1, n\in {\mathbb {Z}}\rbrace ) \end{aligned}$$

since \(V^0\) is simple. That is, there exists \(m\in {\mathbb {Z}}_+\) such that

$$\begin{aligned} \pi _1\circ {\mathcal {Y}}\circ (q_{2m}\otimes q_2)\ne 0. \end{aligned}$$

Finally, the \(\mathfrak {sl}_2\)-type fusion rules of \(V_1\)- and \(V_{25}\)-modules force \(m=1\), completing the proof of the claim and also of the theorem. \(\square \)

Rights and permissions

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

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

McRae, R., Yang, J. An \(\mathfrak {sl}_2\)-type tensor category for the Virasoro algebra at central charge 25 and applications. Math. Z. 303, 32 (2023). https://doi.org/10.1007/s00209-022-03197-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00209-022-03197-z

Mathematics Subject Classification

Navigation