This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Feigin–Semikhatov conjecture and related topics

Shigenori Nakatsuka Department of Mathematical and Statistical Sciences, University of Alberta, CAB 632, Edmonton, AB T6G 2G1, Canada [email protected]
Abstract.

Feigin–Semikhatov conjecture, now established, states algebraic isomorphisms between the cosets of the subregular 𝒲\mathcal{W}-algebras and the principal 𝒲\mathcal{W}-superalgebras of type A by their full Heisenberg subalgebras. It can be seen as a variant of Feigin–Frenkel duality between the 𝒲n\mathcal{W}_{n}-algebras and also as a generalization of the connection between the 𝒩=2\mathcal{N}=2 superconformal algebra and the affine algebra 𝔰𝔩^2,k\widehat{\mathfrak{sl}}_{2,k}.

We review the recent developments on the correspondence of the subregular 𝒲\mathcal{W}-algebras and the principal 𝒲\mathcal{W}-superalgebras of type A at the level of algebras, modules and intertwining operators, including fusion rules.

1. Introduction

Vertex (super)algebras are symmetry algebras (or chiral algebras) for two dimensional conformal field theories and have been introduced in the mathematical literature by Borcherds[18]. They also appear as algebras of local operators in several higher dimensional quantum field theories, which have attracted an intensive attention in the last decade, see Refs. 6, 16, 54, 22, 47, 42, 21, 19 for example.

The 𝒲\mathcal{W}-superalgebras and their affine cosets, that is, subalgebras whose fields commute with affine subalgebras inside give a rich class of such vertex algebras. The 𝒲\mathcal{W}-superalgebras, denoted by 𝒲k(𝔤,f)\mathcal{W}^{k}(\mathfrak{g},f), are obtained from the affine vertex superalgebra Vk(𝔤)V^{k}(\mathfrak{g}) through the quantum Drinfeld–Sokolov reduction parametrized by nilpotent elements ff inside 𝔤\mathfrak{g}. In this paper, we will write 𝒲k(𝔤,f)\mathcal{W}_{k}(\mathfrak{g},f) and Lk(𝔤)L_{k}(\mathfrak{g}) for their simple quotients. When 𝔤\mathfrak{g} is a simply-laced Lie algebra and f=fprinf=f_{\mathrm{prin}} is called principal (or regular), 𝒲k(𝔤)=𝒲k(𝔤,fprin)\mathcal{W}^{k}(\mathfrak{g})=\mathcal{W}^{k}(\mathfrak{g},f_{\mathrm{prin}}) enjoys the triality consisting of the Feigin–Frenkel duality [41] and the Goddard–Kent–Olive type construction [56, 10]— an example of affine cosets—, which asserts that all of them give the same algebra. In particular when 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}, the triality for 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}), also known as Fateev–Lukyanov’s 𝒲n\mathcal{W}_{n}-algebra [36], is depicted in Fig. 1 below.

Figure 1.
𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n})𝒲kˇ(𝔰𝔩n)\mathcal{W}^{\check{k}}(\mathfrak{sl}_{n})Com(V(𝔰𝔩n),V1(𝔰𝔩n)L1(𝔰𝔩n))\operatorname{Com}(V^{\ell}(\mathfrak{sl}_{n}),V^{\ell-1}(\mathfrak{sl}_{n})\otimes L_{1}(\mathfrak{sl}_{n}))FF dualityGKO1k+n+1+n=1\tfrac{1}{k+n}+\tfrac{1}{\ell+n}=1(k+n)(kˇ+n)=1(k+n)(\check{k}+n)=1Relation of levels

Another kind of this tirality was suggested first by Feigin and Semikhatov [38] for the 𝒲\mathcal{W}-algebra 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) associated with the subregular nilpotent element fsubf_{\mathrm{sub}} as in Fig. 2.

Figure 2.
Com(π,𝒲k(𝔰𝔩n,fsub))\operatorname{Com}(\pi,\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}))Com(π,𝒲kˇ(𝔰𝔩n|1,fprin))\operatorname{Com}(\pi,\mathcal{W}^{\check{k}}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}))Com(V(𝔤𝔩n),V(𝔰𝔩n|1))\operatorname{Com}(V^{\ell}(\mathfrak{gl}_{n}),V^{\ell}(\mathfrak{sl}_{n|1}))1k+n+1+n=1\tfrac{1}{k+n}+\tfrac{1}{\ell+n}=1(k+n)(kˇ+n1)=1(k+n)(\check{k}+n-1)=1Relation of levels

Here we take the coset of 𝒲\mathcal{W}-superalgebras by the maximal affine subalgebra inside, which is rank one Heisenberg algebra π\pi. The very beginning case of the Feigin–Frenkel type duality in this family is indeed derived from the so-called Kazam–Suzuki coset construction [62] of the 𝒩=2\mathcal{N}=2 superconformal algebra.

Gaiotto and Rapčák[54] found a vast generalization of these trialities for vertex algebras at the corner appearing as local operators at the junctions of supersymmetric interfaces in 𝒩=4\mathcal{N}=4 Super Yang–Mills gauge theory. The above examples correspond to the following diagrams which describe the ranks of the gauge group U(n)U(n) and the half-BPS interfaces placed as three dimensional boundary conditions.

Figure 3.
nnnnnnnn11nn11nn11

A large class of this triality is now proven by Creutzig and Linshaw [29, 30] mathematically and its application to the representation theory of 𝒲\mathcal{W}-superalgebras is also under work in progress [23, 24, 32].

In this article, we restrict ourselves to the case considered by Feigin and Semikhatov and review the results obtained at this time with an emphasis on how the Kazama–Suzuki coset construction is generalized and refined in this setting.

Acknowledgments

The author would like to express his deepest gratitude to Thomas Creutzig for collaboration and valuable comments during his preparation of this article. He also thanks Naoki Genra, Andrew. R. Linshaw and Ryo Sato for collaboration and for Masahito Yamazaki and Yuto Moriwaki for useful discussions. The author is supported by JSPS KAKENHI Grant Number 20J1014 and JSPS Overseas Research Fellowships Grant Number 202260077. The work is supported by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan.

2. The case of 𝔰𝔩^2,k\widehat{\mathfrak{sl}}_{2,k} and 𝔫𝔰2,c\mathfrak{ns}_{2,c}

2.1. Coset construction of 𝔫𝔰2,c\mathfrak{ns}_{2,c}

The 𝒩=2\mathcal{N}=2 superconformal algebra 𝔫𝔰2,c\mathfrak{ns}_{2,c} with central charge cc has four generating fields: L(z)L(z), J(z)J(z) of even parity and G±(z)G^{\pm}(z) of odd parity satisfying the following operator product expansions (OPEs)

L(z)L(w)12c(zw)4+2L(w)(zw)2+L(w)zw,J(z)J(w)13c(zw)2,\displaystyle L(z)L(w)\sim\frac{\tfrac{1}{2}c}{(z-w)^{4}}+\frac{2L(w)}{(z-w)^{2}}+\frac{\partial L(w)}{z-w},\quad J(z)J(w)\sim\frac{\tfrac{1}{3}c}{(z-w)^{2}},
L(z)J(w)J(w)(zw)2+J(w)(zw),L(z)G±(w)32G±(w)zw+wG±(w)zw,\displaystyle L(z)J(w)\sim\frac{J(w)}{(z-w)^{2}}+\frac{\partial J(w)}{(z-w)},\quad L(z)G^{\pm}(w)\sim\frac{\frac{3}{2}G^{\pm}(w)}{z-w}+\frac{\partial_{w}G^{\pm}(w)}{z-w},
J(z)G±(w)±G±(w)(zw),G±(z)G±(w)0,\displaystyle J(z)G^{\pm}(w)\sim\frac{\pm G^{\pm}(w)}{(z-w)},\quad G^{\pm}(z)G^{\pm}(w)\sim 0,
G±(z)G(w)23c(zw)3+2J(w)(zw)2+2L(w)+J(w)zw.\displaystyle G^{\pm}(z)G^{\mp}(w)\sim\frac{\tfrac{2}{3}c}{(z-w)^{3}}+\frac{2J(w)}{(z-w)^{2}}+\frac{2L(w)+\partial J(w)}{z-w}.

For clarification, let Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) denote the (universal) vertex algebra freely generated by these fields and by Lc(𝔫𝔰2)L_{c}(\mathfrak{ns}_{2}) its unique simple quotient. Kazama and Suzuki[62] gave a family of realizations of Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) based on Hermitian symmetric spaces G/HG/H. We are interested in the case G/H=SU(n+1)/SU(n)×U(1)G/H=SU(n+1)/SU(n)\times U(1). More precisely, we consider the vertex superalgebra Vk(𝔰𝔩n+1)bcnV^{k}(\mathfrak{sl}_{n+1})\otimes bc^{\otimes n} of the universal affine vertex algebra of 𝔰𝔩n+1\mathfrak{sl}_{n+1} at level kk and nn copies of bcbc-systems. Since we have homomorphisms Vk(𝔤𝔩n)Vk(𝔰𝔩n+1)V^{k}(\mathfrak{gl}_{n})\rightarrow V^{k}(\mathfrak{sl}_{n+1}) and V1(𝔤𝔩n)bcnV^{1}(\mathfrak{gl}_{n})\rightarrow bc^{\otimes n}, we may take the diagonal coset Com(Vk+1(𝔤𝔩n),Vk(𝔰𝔩n+1)bcn)\operatorname{Com}(V^{k+1}(\mathfrak{gl}_{n}),V^{k}(\mathfrak{sl}_{n+1})\otimes bc^{\otimes n}), i.e., the subalgebra consisting of elements whose fields commute with the diagonal 𝔤𝔩^n,k+1\widehat{\mathfrak{gl}}_{n,k+1}-action on Vk(𝔰𝔩n+1)bcnV^{k}(\mathfrak{sl}_{n+1})\otimes bc^{\otimes n}. Then the coset contains Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) with specific central charge cc whereas the whole coset is conjectured[58, 28] to be isomorphic to the principal 𝒲\mathcal{W}-superalgebra 𝒲(𝔰𝔩n+1|n)\mathcal{W}(\mathfrak{sl}_{n+1|n}) at certain level:

Vc(𝔫𝔰2)𝒲(𝔰𝔩n+1|n)Com(Vk+1(𝔤𝔩n),Vk(𝔰𝔩n+1)bcn).\displaystyle V^{c}(\mathfrak{ns}_{2})\hookrightarrow\mathcal{W}(\mathfrak{sl}_{n+1|n})\simeq\operatorname{Com}\left(V^{k+1}(\mathfrak{gl}_{n}),V^{k}(\mathfrak{sl}_{n+1})\otimes bc^{\otimes n}\right). (1)

When n=1n=1, Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) is exactly the principal 𝒲\mathcal{W}-superalgebra 𝒲(𝔰𝔩2|1)\mathcal{W}(\mathfrak{sl}_{2|1}) and gives the honest coset construction of Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}). More explicitly, the isomorphism is given by the following:

𝐊𝐒:Vc(𝔫𝔰2)Com(πH+Δ,Vk(𝔰𝔩2)V)G+(z)2k+2e(z)V1(z)G(z)2k+2f(z)V1(z)J(z)2k+2H+Δ(z)+1b1(z)\displaystyle\begin{array}[]{cccc}{\bf KS}\colon&V^{c}(\mathfrak{ns}_{2})&\xrightarrow{\simeq}&\operatorname{Com}\left(\pi^{H_{+}^{\Delta}},V^{k}(\mathfrak{sl}_{2})\otimes V_{\mathbb{Z}}\right)\\ &G^{+}(z)&\mapsto&\sqrt{\frac{2}{k+2}}e(z)\otimes V_{1}(z)\\ &G^{-}(z)&\mapsto&\sqrt{\frac{2}{k+2}}f(z)\otimes V_{-1}(z)\\ &J(z)&\mapsto&\frac{-2}{k+2}H_{+}^{\Delta}(z)+1\otimes b_{1}(z)\end{array} (6)

with

H+Δ(z)=12h(z)1+1b1(z),c=3kk+2.\displaystyle H_{+}^{\Delta}(z)=-\frac{1}{2}h(z)\otimes 1+1\otimes b_{1}(z),\quad c=\frac{3k}{k+2}. (7)

Here we have used the boson-fermion correspondence between the bcbc-system and the lattice vertex superalgebra VV_{\mathbb{Z}} associated with the lattice \mathbb{Z}; Vn(z)V_{n}(z) is the vertex operator for the element in the lattice nn\in\mathbb{Z}; bn(z)b_{n}(z) is the Heisenberg field satisfying the OPE bn(z)bn(z)n2/(zw)2b_{n}(z)b_{n}(z)\sim n^{2}/(z-w)^{2}; πH+Δ\pi^{H_{+}^{\Delta}} is the Heisenberg vertex algebra generated by H+Δ(z)H_{+}^{\Delta}(z). The construction (6) descends to the simple quotients

Lc(𝔫𝔰2)Com(πH+Δ,Lk(𝔰𝔩2)V).\displaystyle L_{c}(\mathfrak{ns}_{2})\xrightarrow{\simeq}\operatorname{Com}\left(\pi^{H_{+}^{\Delta}},L_{k}(\mathfrak{sl}_{2})\otimes V_{\mathbb{Z}}\right). (8)

When kk is a non-negative integer, this realization has been used to construct the unitary minimal representations [34] as a variant of the celebrated Goddard–Kent–Olive construction[56] of those representations for the Virasoro algebra. We will return back to this point later.

The coset construction (6) implies that the representation theories of Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) and Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) share some similarity. This means that one can not only go from the Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-side to the Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-side, but also in the opposite direction. Feigin–Semikhatov–Tipunin [40] found the crucial coset type realization of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) in terms of Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) by using the negative-definite lattice vertex superalgebra V1V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}. Namely, they established the following isomorphism of vertex algebras:

𝐅𝐒𝐓:Vk(𝔰𝔩2)Com(πHΔ,Vc(𝔫𝔰2)V1)e(z)k+22G+(z)V1(z)f(z)k+22G(z)V-1(z)12h(z)k+22HΔ(z)1b1(z)\displaystyle\begin{array}[]{cccc}{\bf FST}\colon&V^{k}(\mathfrak{sl}_{2})&\xrightarrow{\simeq}&\operatorname{Com}(\pi^{H_{-}^{\Delta}},V^{c}(\mathfrak{ns}_{2})\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}})\\ &e(z)&\mapsto&\sqrt{\frac{k+2}{2}}G^{+}(z)\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}}(z)\\ &f(z)&\mapsto&\sqrt{\frac{k+2}{2}}G^{-}(z)\otimes V_{\text{-}\operatorname{\sqrt{\smash[b]{-1}}}}(z)\\ &\frac{1}{2}h(z)&\mapsto&\frac{k+2}{2}H_{-}^{\Delta}(z)-1\otimes b_{\operatorname{\sqrt{\smash[b]{-1}}}}(z)\end{array} (13)

with HΔ(z)=J(z)1+1b1(z)H_{-}^{\Delta}(z)=J(z)\otimes 1+1\otimes b_{\operatorname{\sqrt{\smash[b]{-1}}}}(z). Again, (13) descends to the simple quotients

Lk(𝔰𝔩2)Com(πHΔ,Lc(𝔫𝔰2)V1).\displaystyle L_{k}(\mathfrak{sl}_{2})\xrightarrow{\simeq}\operatorname{Com}\left(\pi^{H_{-}^{\Delta}},L_{c}(\mathfrak{ns}_{2})\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\right). (14)

2.2. Equivalence of module categories

Let us recall the block-wise equivalence [40, 70] of representation categories for Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) and Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) by the coset constructions (6) and (13). From now on, we always assume c=3kk+2c=\frac{3k}{k+2} as in (7). Since the coset constructions themselves use Heisenberg fields, it is reasonable to consider those modules of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) and Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) which behave nicely for their Heisenberg fields inside them. This motivates to introduce the category of weight modules: let Vk(𝔰𝔩2)-modwtV^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}} denote the category of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-modules which decompose into direct sums of Fock modules with respect to π12h\pi^{\frac{1}{2}h}. Note that the eigenvalues of 12h0\frac{1}{2}h_{0} on Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) is \mathbb{Z}. Then the category Vk(𝔰𝔩2)-modwtV^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}} admits a block decomposition in terms of the spectrum of 12h0\frac{1}{2}h_{0} modulo \mathbb{Z}:

Vk(𝔰𝔩2)-modwt=[λ]/Vk(𝔰𝔩2)-modwt[λ].\displaystyle V^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}}=\bigoplus_{[\lambda]\in\mathbb{C}/\mathbb{Z}}V^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}}^{[\lambda]}.

Similarly, we introduce the category Vc(𝔫𝔰2)-modwtV^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}} of weight Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-modules in terms of πJ\pi^{J}. Then it decomposes into

Vc(𝔫𝔰2)-modwt=[μ]/Vc(𝔫𝔰2)-modwt[μ]\displaystyle V^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}}=\bigoplus_{[\mu]\in\mathbb{C}/\mathbb{Z}}V^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}}^{[\mu]}

by using the spectrum of J0J_{0} on Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}), which is \mathbb{Z}. To compare the categories of weight modules, it suffices to notice that the coset constructions for algebras (6) and (13) are generalized to modules by using the multiplicity spaces of suitable Fock modules. To obtain a functor from Vk(𝔰𝔩2)-modwtV^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}} to Vc(𝔫𝔰2)-modwtV^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}}, take an arbitrary weight module MM and decompose MVM\otimes V_{\mathbb{Z}} as a module over Vc(𝔫𝔰2)πH+ΔV^{c}(\mathfrak{ns}_{2})\otimes\pi^{H_{+}^{\Delta}}:

MVξΩξ+(M)πξH+Δ\displaystyle M\otimes V_{\mathbb{Z}}\simeq\bigoplus_{\xi\in\mathbb{C}}\Omega^{+}_{\xi}(M)\otimes\pi^{H_{+}^{\Delta}}_{\xi} (15)

with

Ωξ+(M):={aMVH+,nΔa=δn,0ξa,(n0)}\displaystyle\Omega_{\xi}^{+}(M):=\{a\in M\otimes V_{\mathbb{Z}}\mid H_{+,n}^{\Delta}a=\delta_{n,0}\xi\ a,\ (n\geq 0)\}

and πξH+Δ\pi^{H_{+}^{\Delta}}_{\xi} the Fock module of πH+Δ\pi^{H_{+}^{\Delta}} on which H+,0ΔH_{+,0}^{\Delta} acts by ξ\xi. If MM lies in Vk(𝔰𝔩2)-modwt[λ]V^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}}^{[\lambda]}, then Ωξ+(M)\Omega_{\xi}^{+}(M) lies in Vc(𝔫𝔰2)-modwt[εξ]V^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}}^{[-\varepsilon\xi]} with ε=2k+2\varepsilon=\frac{2}{k+2}. Therefore, Ωξ+\Omega_{\xi}^{+} defines a functor

Ωξ+:Vk(𝔰𝔩2)-modwt[λ]Vc(𝔫𝔰2)-modwt[εξ].\displaystyle\Omega_{\xi}^{+}\colon V^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}}^{[\lambda]}\rightarrow V^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}}^{[-\varepsilon\xi]}. (16)

It is non-zero if and only if ξλ+\xi\in-\lambda+\mathbb{Z}. To get a functor in the opposite direction, take an arbitrary Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-module NN and decompose NV1N\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}} as a module over Vk(𝔰𝔩2)πHΔV^{k}(\mathfrak{sl}_{2})\otimes\pi^{H_{-}^{\Delta}}:

NV1ξΩξ(N)πξHΔ.\displaystyle N\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\simeq\bigoplus_{\xi\in\mathbb{C}}\Omega^{-}_{\xi}(N)\otimes\pi^{H_{-}^{\Delta}}_{\xi}. (17)

Then we obtain a functor

Ωξ:Vc(𝔫𝔰2)-modwt[μ]Vk(𝔰𝔩2)-modwt[ε1ξ],\displaystyle\Omega^{-}_{\xi}\colon V^{c}(\mathfrak{ns}_{2})\text{-mod}_{\mathrm{wt}}^{[\mu]}\rightarrow V^{k}(\mathfrak{sl}_{2})\text{-mod}_{\mathrm{wt}}^{[\varepsilon^{-1}\xi]}, (18)

which is non-zero if and only if ξλ+\xi\in\lambda+\mathbb{Z}.

The functors Ωξ±\Omega^{\pm}_{\xi} are either zero or an equivalence. In particular, the block-wise equivalence of categories is stated in the following way:

Theorem 2.1.

[40, 70] The functors

Ωλ+:Vk(𝔰𝔩2)-modwt[λ]Vc(𝔫𝔰2)-modwt[ελ]:Ωελ\displaystyle\Omega_{-\lambda}^{+}\colon V^{k}(\mathfrak{sl}_{2})\text{-}\mathrm{mod}_{\mathrm{wt}}^{[\lambda]}\ \rightleftarrows\ V^{c}(\mathfrak{ns}_{2})\text{-}\mathrm{mod}_{\mathrm{wt}}^{[\varepsilon\lambda]}\colon\Omega^{-}_{\varepsilon\lambda} (19)

are quasi-inverse to each other and thus give an equivalence of categories.

The proof is established by giving natural isomorphisms connecting modules MM and their images in MVV1M\otimes V_{\mathbb{Z}}\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}} under the compositions Ωλ+Ωελ\Omega_{-\lambda}^{+}\circ\Omega_{\varepsilon\lambda}^{-} and ΩελΩλ+\Omega_{\varepsilon\lambda}^{-}\circ\Omega_{-\lambda}^{+}.

2.3. Some basic modules and resolutions

To illustrate the block-wise equivalence, we compare resolutions of simple modules by some basic modules when the level kk is irrational. On Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-side, the Weyl module 𝕍nϖ1k\mathbb{V}^{k}_{n\varpi_{1}} of highest weight nϖ1n\varpi_{1}, the induced module whose top space is the n+1n+1-dimensional 𝔰𝔩2\mathfrak{sl}_{2}-module, has a two step resolution by the affine Verma modules:

0𝕄(n+2)ϖ1k𝕄nϖ1k𝕍nϖ1k0.\displaystyle 0\rightarrow\mathbb{M}^{k}_{-(n+2)\varpi_{1}}\rightarrow\mathbb{M}^{k}_{n\varpi_{1}}\rightarrow\mathbb{V}^{k}_{n\varpi_{1}}\rightarrow 0. (20)

Applying the functor Ωn/2+\Omega_{-n/2}^{+}, we obtain the following resolution of the simple Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-module Lc(14εn,12εn)L_{c}(\frac{1}{4}\varepsilon n,\frac{1}{2}\varepsilon n) of highest weight (L0,J0)=(14εn,12εn)(L_{0},J_{0})=(\frac{1}{4}\varepsilon n,\frac{1}{2}\varepsilon n):

0Ωn/2+(𝕄(n+2)ϖ1k)Ωn/2+(𝕄nϖ1k)Lc(14εn,12εn)0.\displaystyle 0\rightarrow\Omega^{+}_{-n/2}(\mathbb{M}^{k}_{-(n+2)\varpi_{1}})\rightarrow\Omega^{+}_{-n/2}(\mathbb{M}^{k}_{n\varpi_{1}})\rightarrow L_{c}(\tfrac{1}{4}\varepsilon n,\tfrac{1}{2}\varepsilon n)\rightarrow 0.

Each component has the following character tr(qL0zJ0)\mathrm{tr}_{\bullet}(q^{L_{0}}z^{J_{0}}):

chLc(14εn,12εn)=q14εnz12εn1qn+11+z1qn+12(zq32,z1q12;q)(q;q)2,\displaystyle\mathrm{ch}\ L_{c}(\tfrac{1}{4}\varepsilon n,\tfrac{1}{2}\varepsilon n)=q^{\frac{1}{4}\varepsilon n}z^{\frac{1}{2}\varepsilon n}\frac{1-q^{n+1}}{1+z^{-1}q^{n+\frac{1}{2}}}\frac{\left(-zq^{\frac{3}{2}},-z^{-1}q^{\frac{1}{2}};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{2}},
chΩn/2+(𝕄nϖ1k)=q14εnz12εn(zq32,z1q12;q)(q;q)2,\displaystyle\mathrm{ch}\ \Omega^{+}_{-n/2}(\mathbb{M}^{k}_{n\varpi_{1}})=q^{\frac{1}{4}\varepsilon n}z^{\frac{1}{2}\varepsilon n}\frac{\left(-zq^{\frac{3}{2}},-z^{-1}q^{\frac{1}{2}};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{2}},
chΩn/2+(𝕄(n+2)ϖ1k)=q14εn+12(n+1)2z12εn(n+1)(zq12n,z1q32+n;q)(q;q)2,\displaystyle\mathrm{ch}\ \Omega^{+}_{-n/2}(\mathbb{M}^{k}_{-(n+2)\varpi_{1}})=q^{\frac{1}{4}\varepsilon n+\frac{1}{2}(n+1)^{2}}z^{\frac{1}{2}\varepsilon n-(n+1)}\frac{\left(-zq^{\frac{1}{2}-n},-z^{-1}q^{\frac{3}{2}+n};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{2}},

in terms of the qq-Pochhammer symbols (a1,,am;q)=i=1mN=0(1aiqN)(a_{1},\cdots,a_{m};q)_{\infty}=\prod_{i=1}^{m}\prod_{N=0}^{\infty}(1-a_{i}q^{N}). As indicated by the characters, Ωn/2+(𝕄nϖ1k)\Omega^{+}_{-n/2}(\mathbb{M}^{k}_{n\varpi_{1}}) is identified with a quotient of a Verma module, namely, 𝐌c(14ϵn,12ϵn)=𝕄c(14εn,12εn)/(G1/2+|14εn,12εn)\mathbf{M}^{c}(\frac{1}{4}\epsilon n,\frac{1}{2}\epsilon n)=\mathbb{M}^{c}(\frac{1}{4}\varepsilon n,\frac{1}{2}\varepsilon n)/(G_{-1/2}^{+}|\frac{1}{4}\varepsilon n,\frac{1}{2}\varepsilon n\rangle), called a topological Verma module[1, 39]. The other one Ωn/2+(𝕄(n+2)ϖ1k)\Omega^{+}_{-n/2}(\mathbb{M}^{k}_{-(n+2)\varpi_{1}}) is obtained from 𝐌c(14ε(n+2),12ε(n+2))\mathbf{M}^{c}(\tfrac{-1}{4}\varepsilon(n+2),\tfrac{-1}{2}\varepsilon(n+2)) by spectral flow twist Sn1S_{-n-1} defined in general by

Sθ:Ga±Ga±θ±,JaJa+c3θδa,0,LaLa+θJa+c6θ2δn,0,(θ),\displaystyle S_{\theta}\colon G^{\pm}_{a}\mapsto G^{\pm}_{a\pm\theta},\quad J_{a}\mapsto J_{a}+\tfrac{c}{3}\theta\delta_{a,0},\quad L_{a}\mapsto L_{a}+\theta J_{a}+\tfrac{c}{6}\theta^{2}\delta_{n,0},\quad(\theta\in\mathbb{Z}),

which transforms the module characters in the following way:

chSpM(q,z)=qc6p2zc3pchM(q,zqp).\mathrm{ch}\ S_{p}M(q,z)=q^{\frac{c}{6}p^{2}}z^{\frac{c}{3}p}\mathrm{ch}\ M(q,zq^{p}).

Therefore, the counterpart of (20) is

0Sn1𝐌c(14ε(n+2),12ε(n+2))𝐌c(14ϵn,12ϵn)Lc(14εn,12εn)0.0\rightarrow S_{-n-1}\mathbf{M}^{c}(\tfrac{-1}{4}\varepsilon(n+2),\tfrac{-1}{2}\varepsilon(n+2))\rightarrow\mathbf{M}^{c}(\tfrac{1}{4}\epsilon n,\tfrac{1}{2}\epsilon n)\rightarrow L_{c}(\tfrac{1}{4}\varepsilon n,\tfrac{1}{2}\varepsilon n)\rightarrow 0.

To obtain resolutions of simple Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-modules by Verma modules 𝕄c(h,m)\mathbb{M}^{c}(h,m) (they are called massive Verma modules), we have to replace the affine Verma modules 𝕄λk\mathbb{M}^{k}_{\lambda} by thicker Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-modules. They are given by relaxed highest weight modules Ra,bkR_{a,b}^{k} (a,ba,b\in\mathbb{C}) induced by the weight 𝔰𝔩2\mathfrak{sl}_{2}-modules

Ra,b:=𝒰(𝔰𝔩2)[Ω,h]a,bn>0enva,bva,bn>0fnva,b.\displaystyle R_{a,b}:=\mathcal{U}(\mathfrak{sl}_{2})\otimes_{\mathbb{C}[\Omega,h]}\mathbb{C}_{a,b}\simeq\bigoplus_{n>0}\mathbb{C}e^{n}v_{a,b}\oplus v_{a,b}\oplus\bigoplus_{n>0}\mathbb{C}f^{n}v_{a,b}.

Here Ω=12h2+ef+fe\Omega=\frac{1}{2}h^{2}+ef+fe is the Casimir element of 𝔰𝔩2\mathfrak{sl}_{2} and a,b\mathbb{C}_{a,b} is the one dimensional module over [Ω,h]\mathbb{C}[\Omega,h] with Ω=a\Omega=a, h=bh=b. Then Ra,bkR_{a,b}^{k} has the character trqL0zh/2\mathrm{tr}_{\bullet}q^{L_{0}}z^{h/2}

chRa,bk=qa2(k+2)zb2nzn(q,zq,z1q;q)=qa2(k+2)zb2nzn(q;q)3,\mathrm{ch}R_{a,b}^{k}=q^{\frac{a}{2(k+2)}}z^{\frac{b}{2}}\frac{\sum_{n\in\mathbb{Z}}z^{n}}{\left(q,zq,z^{-1}q;q\right)_{\infty}}=q^{\frac{a}{2(k+2)}}z^{\frac{b}{2}}\frac{\sum_{n\in\mathbb{Z}}z^{n}}{\left(q;q\right)_{\infty}^{3}},

which gives the desired character on Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-side:

chΩn/2+(Rn(n+2)2,nk)=q14εnz12εn(zq12,z1q12;q)(q;q)2.\mathrm{ch}\ \Omega_{-n/2}^{+}\left(R_{\frac{n(n+2)}{2},n}^{k}\right)=q^{\frac{1}{4}\varepsilon n}z^{\frac{1}{2}\varepsilon n}\frac{\left(-zq^{\frac{1}{2}},-z^{-1}q^{\frac{1}{2}};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{2}}.

Indeed, we have an isomorphism Ωn/2+(Rn(n+2)/2,nk)𝕄c(14εn,12εn)\Omega_{-n/2}^{+}(R_{n(n+2)/2,n}^{k})\simeq\mathbb{M}^{c}(\frac{1}{4}\varepsilon n,\frac{1}{2}\varepsilon n). The relaxed highest weight modules and their spectral flow twists defined through

Sθ:eaea+θ,haha+kθδa,0,fafaθ,(θ),S_{\theta}\colon e_{a}\mapsto e_{a+\theta},\quad h_{a}\mapsto h_{a}+k\theta\delta_{a,0},\quad f_{a}\mapsto f_{a-\theta},\quad(\theta\in\mathbb{Z}),

are the most natural class of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-modules since all the simple weight Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-modules are obtained[52] as their simple quotients. A nice realization of relaxed highest weight modules is implemented by the inverse of the quantum Drinfeld–Sokolov reduction of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) introduced by Adamović[3]:

Vk(𝔰𝔩2)𝒲k(𝔰𝔩2)Πe(z)1e𝐚(z),12h(z)1𝐛+(z),f(z)(k+2)L(z)e𝐚(z),1:(𝐛(z)2+(k+1)(z𝐛(z))e𝐚(z):,\displaystyle\begin{split}&V^{k}(\mathfrak{sl}_{2})\longrightarrow\mathcal{W}^{k}(\mathfrak{sl}_{2})\otimes\Pi\\ &e(z)\mapsto 1\otimes\mathrm{e}^{\mathbf{a}}(z),\quad\tfrac{1}{2}h(z)\mapsto 1\otimes\mathbf{b}^{+}(z),\\ &f(z)\mapsto(k+2)L(z)\otimes\mathrm{e}^{-\mathbf{a}}(z),\\ &\hskip 56.9055pt-1\otimes:\left(\mathbf{b}^{-}(z)^{2}+(k+1)(\partial_{z}\mathbf{b}^{-}(z)\right)\mathrm{e}^{-\mathbf{a}}(z):,\end{split} (21)

where Π\Pi is the half-lattice vertex algebra

Π:=V(1+1)π(1+1)π1V1,\Pi:=V_{(1+\operatorname{\sqrt{\smash[b]{-1}}})\mathbb{Z}}\underset{\pi^{(1+\operatorname{\sqrt{\smash[b]{-1}}})\mathbb{Z}}}{\otimes}\pi^{\mathbb{Z}\oplus\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\subset V_{\mathbb{Z}\oplus\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}},

and

𝐚(z)=b1+1(z),𝐛±(z)=±k4b1+1(z)+12b11(z).\displaystyle\mathbf{a}(z)=b_{1+\operatorname{\sqrt{\smash[b]{-1}}}}(z),\quad\mathbf{b}^{\pm}(z)=\pm\tfrac{k}{4}b_{1+\operatorname{\sqrt{\smash[b]{-1}}}}(z)+\tfrac{1}{2}b_{1-\operatorname{\sqrt{\smash[b]{-1}}}}(z).

Let us introduce the following modules[37] over Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})

Mr,sk[θ,λ]:=Mr,sc(k)Πθ[λ],(r,s,θ,λ).\displaystyle\mathrm{M}^{k}_{r,s}[\theta,\lambda]:=\mathrm{M}^{c(k)}_{r,s}\otimes\Pi_{\theta}[\lambda],\quad(r,s,\theta\in\mathbb{Z},\ \lambda\in\mathbb{C}).

Here Mr,sc(k)\mathrm{M}^{c(k)}_{r,s} is the Verma module of highest weight L0=hr,sL_{0}=h_{r,s} over the Virasoro algebra of central charge c(k)c(k) with

c(k)=16(k+1)2(k+2),hr,s=(r(k+2)s)2(k+1)24(k+2),c(k)=1-6\tfrac{(k+1)^{2}}{(k+2)},\quad h_{r,s}=\tfrac{\left(r(k+2)-s\right)^{2}-(k+1)^{2}}{4(k+2)},

and Πθ[λ]\Pi_{\theta}[\lambda] is the Π\Pi-module given by the sum of Fock modules

Πθ[λ]=nπθ𝐛++(λ+n)𝐚1.\Pi_{\theta}[\lambda]=\bigoplus_{n\in\mathbb{Z}}\pi^{\mathbb{Z}\oplus\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}_{\theta\mathbf{b}^{+}+(\lambda+n)\mathbf{a}}.

In particular, Mr,sk[1,λ]\mathrm{M}_{r,s}^{k}[-1,\lambda] has the following character

chMr,sk[1,λ]=qhr,s+k4zλk2nzn(q;q)3\displaystyle\mathrm{ch}\ \mathrm{M}_{r,s}^{k}[-1,\lambda]=q^{h_{r,s}+\frac{k}{4}}z^{\lambda-\frac{k}{2}}\frac{\sum_{n\in\mathbb{Z}}z^{n}}{\left(q;q\right)_{\infty}^{3}}

and is isomorphic to the relaxed highest weight module

Mr,sk[1,λ]Ra,bk,(a=2(k+2)(hr,s+k4),b=2(λk2))\displaystyle M_{r,s}^{k}[-1,\lambda]\simeq R_{a,b}^{k},\quad(a=2(k+2)(h_{r,s}+\tfrac{k}{4}),\quad b=2(\lambda-\tfrac{k}{2})) (22)

under the assumption (bp)22a+1(b-p)^{2}\neq 2a+1 for all positive odd integers pp\in\mathbb{Z}. The assumption says that there is no highest weight vector in Ra,bkR_{a,b}^{k} otherwise these two modules are non-isomorphic indecomposable modules. The other modules Mr,sk[θ,λ]M_{r,s}^{k}[\theta,\lambda] are obtained from Mr,sk[1,λ]M_{r,s}^{k}[-1,\lambda] by spectral flow twists:

Mr,sk[θ,λ]Sθ+1Mr,sk[1,λ].M_{r,s}^{k}[\theta,\lambda]\simeq S_{\theta+1}M_{r,s}^{k}[-1,\lambda].

Let λ\lambda be generic so that (22) holds. The simple quotient Lr,sc(k)\mathrm{L}_{r,s}^{c(k)} (r,s1r,s\geq 1) of Mr,sc(k)\mathrm{M}_{r,s}^{c(k)} has a two step resolution[45, 57] by Verma modules due to Feigin–Fuchs:

0Mr,sc(k)Mr,sc(k)Lr,sc(k)0.0\rightarrow\mathrm{M}^{c(k)}_{-r,s}\rightarrow\mathrm{M}^{c(k)}_{r,s}\rightarrow\mathrm{L}_{r,s}^{c(k)}\rightarrow 0.

Except for finitely many values of λ\lambda, Lr,sc(k)Π1[λ]\mathrm{L}_{r,s}^{c(k)}\otimes\Pi_{-1}[\lambda] coincides with the simple quotient of Mr,sk[1,λ]\mathrm{M}^{k}_{r,s}[-1,\lambda], which we denote by Lr,sk[1,λ]\mathrm{L}^{k}_{r,s}[-1,\lambda]. In this case, the above resolution induces a resolution of Lr,sk[1,λ]\mathrm{L}^{k}_{r,s}[-1,\lambda] by relaxed highest Verma modules

0Mr,sk[1,λ]Mr,sk[1,λ]Lr,sk[1,λ]0.\displaystyle 0\rightarrow\mathrm{M}_{-r,s}^{k}[-1,\lambda]\rightarrow\mathrm{M}_{r,s}^{k}[-1,\lambda]\rightarrow\mathrm{L}_{r,s}^{k}[-1,\lambda]\rightarrow 0.

Applying the functor Ωb/2+\Omega^{+}_{-b/2}, we obtain the following resolution of a simple Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-module by massive Verma modules:

0𝕄c(α,βλ)𝕄c(α+,βλ)Lc(α+,βλ)0\displaystyle 0\rightarrow\mathbb{M}^{c}(\alpha_{-},\beta_{\lambda})\rightarrow\mathbb{M}^{c}(\alpha_{+},\beta_{\lambda})\rightarrow L_{c}(\alpha_{+},\beta_{\lambda})\rightarrow 0

with

α±=h±r,s+12ε(β0βλ2),βλ=ε(λ+1)1.\alpha_{\pm}=h_{\pm r,s}+\tfrac{1}{2\varepsilon}\left(\beta_{0}-\beta_{\lambda}^{2}\right),\quad\beta_{\lambda}=\varepsilon(\lambda+1)-1.

The above approach should also work at admissible levels. Some resolutions of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-modules and Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-modules in this case has been obtained in Refs. 61, 39, 63 . It might be an interesting problem to compare these two.

We record the following dictionary[39] on the correspondence of some classes of basic modules up to spectral flow twists. One of the aims of this article is to present a generalization for other 𝒲\mathcal{W}-superalgebras, see Tab. 2 at the end of this article.

Table 1.
Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2})-side Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2})-side
affine Verma modules topological Verma modules
relaxed highest weight modules massive Verma modules

2.4. Fusion rules: rational case

Let us compare the fusion rings for Lk(𝔰𝔩2)L_{k}(\mathfrak{sl}_{2})-mod and Lc(𝔫𝔰2)L_{c}(\mathfrak{ns}_{2})-mod when kk is a non-negative integer, i.e. the rational case, when both categories are semisimple braided tensor categories with finitely many simple objects. Although the equivalence Theorem 2.1 is stated at the level of abelian categories, it is strong enough to deduce the relation of their fusion rules in this case by the theory of simple current extensions in the language of braided tensor category[26, 27, 71, 24].

Recall that the list of simple Lk(𝔰𝔩2)L_{k}(\mathfrak{sl}_{2})-modules is given by

Lk,0,Lk,ϖ1,,Lk,kϖ1L_{k,0},\ L_{k,\varpi_{1}},\ \cdots,\ L_{k,k\varpi_{1}}

and that the list[1] of simple Lc(𝔫𝔰2)L_{c}(\mathfrak{ns}_{2})-modules is given by the highest weight modules

Lc[u,v]:=Lc(hu,v,su,v),hu,v=1k+2(uv14),su,v=uvk+2,\displaystyle L_{c}[u,v]:=L_{c}(h_{u,v},s_{u,v}),\quad h_{u,v}=\tfrac{1}{k+2}(uv-\tfrac{1}{4}),\ s_{u,v}=\tfrac{u-v}{k+2},

where 𝐮=(u,v)\mathbf{u}=(u,v) runs through the set

u,v12+0, 0u,v,u+vk+2.u,v\in\tfrac{1}{2}+\mathbb{Z}_{\geq 0},\ 0\leq u,v,u+v\leq k+2.

They consist of all the unitary minimal representations[34] of 𝔫𝔰2,c\mathfrak{ns}_{2,c}. The module categories decompose into

Lk(𝔰𝔩2)-mod=i2Lk(𝔰𝔩2)-mod[i2],Lc(𝔫𝔰2)-mod=jk+2Lc(𝔫𝔰2)-mod[jk+2]\displaystyle L_{k}(\mathfrak{sl}_{2})\text{-mod}=\bigoplus_{i\in\mathbb{Z}_{2}}L_{k}(\mathfrak{sl}_{2})\text{-mod}^{[\frac{i}{2}]},\quad L_{c}(\mathfrak{ns}_{2})\text{-mod}=\bigoplus_{j\in\mathbb{Z}_{k+2}}L_{c}(\mathfrak{ns}_{2})\text{-mod}^{[\frac{j}{k+2}]}

and the lists of simple modules lying in each block are given by

Lk,pϖ1(pimod 2),Lc[𝐮](uvjmodk+2).\displaystyle L_{k,p\varpi_{1}}\ (p\equiv i\ \text{mod}\ 2),\quad L_{c}[\mathbf{u}]\ (u-v\equiv j\ \text{mod}\ k+2).

In Fig. 4, we illustrate the block-wise equivalence for k=2k=2.

Figure 4. Correspondence of simple modules for k=2k=2
(λ=0\lambda=0)(λ=12\lambda=\tfrac{1}{2})L2,0,L2,2ϖ1L_{2,0},\ L_{2,2\varpi_{1}}L2,ϖ1L_{2,\varpi_{1}}Lc[12,12],Lc[32,32L_{c}[\tfrac{1}{2},\tfrac{1}{2}],\ L_{c}[\tfrac{3}{2},\tfrac{3}{2}]Lc[32,12]L_{c}[\tfrac{3}{2},\tfrac{1}{2}]Lc[52,12],Lc[12,52]L_{c}[\tfrac{5}{2},\tfrac{1}{2}],\ L_{c}[\tfrac{1}{2},\tfrac{5}{2}]Lc[12,32]L_{c}[\tfrac{1}{2},\tfrac{3}{2}](μ=0\mu=0)(μ=14\mu=\tfrac{1}{4})(μ=24\mu=\tfrac{2}{4})(μ=34\mu=\tfrac{3}{4})Ω0+\Omega^{+}_{0}Ω1+\Omega^{+}_{-1}Ω1/2+\Omega^{+}_{-1/2}Ω1/2+\Omega^{+}_{1/2}

The fusion rules are given by

Lk,pϖ1Lk,qϖ1rLk,rϖ1,Lc[𝐮1]Lc[𝐮2]𝐮3Lc[𝐮3]L_{k,p\varpi_{1}}\boxtimes L_{k,q\varpi_{1}}\simeq\bigoplus_{r}L_{k,r\varpi_{1}},\quad L_{c}[\mathbf{u}_{1}]\boxtimes L_{c}[\mathbf{u}_{2}]\simeq\bigoplus_{\mathbf{u}_{3}}L_{c}[\mathbf{u}_{3}]

where rr runs through

  • |pq|rMin{p+q,2k(p+q)},|p-q|\leq r\leq\mathrm{Min}\{p+q,2k-(p+q)\},
    rp+q mod 2,r\equiv p+q\text{ mod }2,

and 𝐮3\mathbf{u}_{3} through[2] either of

  • |𝐮1+𝐮2+|<𝐮3+<Min{𝐮1++𝐮2+,2(k+2)(𝐮1++𝐮2+)}|\mathbf{u}_{1}^{+}-\mathbf{u}_{2}^{+}|<\mathbf{u}_{3}^{+}<\mathrm{Min}\{\mathbf{u}_{1}^{+}+\mathbf{u}_{2}^{+},2(k+2)-(\mathbf{u}_{1}^{+}+\mathbf{u}_{2}^{+})\},
    𝐮3=𝐮1+𝐮2\mathbf{u}_{3}^{-}=\mathbf{u}_{1}^{-}+\mathbf{u}_{2}^{-},

  • |𝐮1+𝐮2+|<(k+2)𝐮3+<Min{𝐮1++𝐮2+,2(k+2)(𝐮1++𝐮2+)}|\mathbf{u}_{1}^{+}-\mathbf{u}_{2}^{+}|<(k+2)-\mathbf{u}_{3}^{+}<\mathrm{Min}\{\mathbf{u}_{1}^{+}+\mathbf{u}_{2}^{+},2(k+2)-(\mathbf{u}_{1}^{+}+\mathbf{u}_{2}^{+})\},
    𝐮3=𝐮1+𝐮2±(k+2)\mathbf{u}_{3}^{-}=\mathbf{u}_{1}^{-}+\mathbf{u}_{2}^{-}\pm(k+2),

where we set 𝐮i±=ui±vi\mathbf{u}^{\pm}_{i}=u_{i}\pm v_{i}. In particular, Lk,kϖ1L_{k,k\varpi_{1}} is a simple current and satisfies

Lk,kϖ1Lk,pϖ1Lk,(kp)ϖ1.L_{k,k\varpi_{1}}\boxtimes L_{k,p\varpi_{1}}\simeq L_{k,(k-p)\varpi_{1}}.

For a more transparent comparison of the fusion rings, we use a general phenomenon of simple current extensions by lattice theories[71, 27, 24] applied for the decomposition

Lc(𝔫𝔰2)V1Lk,0V2(k+2)Lk,ϖ1V(k+2)2(k+2)+2(k+2).\displaystyle L_{c}(\mathfrak{ns}_{2})\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\simeq L_{k,0}\otimes V_{\operatorname{\sqrt{\smash[b]{-2(k+2)}}}\mathbb{Z}}\oplus L_{k,\varpi_{1}}\otimes V_{\frac{(k+2)}{\operatorname{\sqrt{\smash[b]{-2(k+2)}}}}+\operatorname{\sqrt{\smash[b]{-2(k+2)}}}\mathbb{Z}}.

Let VV be a rational C2C_{2}-cofinite vertex operator algebra so that the module category VV-mod is a semisimple braided tensor category of finitely many simple objects. We assume that VV-mod has simple current modules SaS_{a} which form an abelian group GG by fusion product

SaSbSa+b.S_{a}\boxtimes S_{b}\simeq S_{a+b}.

Note that the fusion ring 𝒦(V)\mathcal{K}(V) of VV-modules is naturally a ring over [G]\mathbb{Z}[G] by fusion product a.M=SaMa.M=S_{a}\boxtimes M and a ring graded by the dual group G=Hom(G,×)G^{\vee}=\mathrm{Hom}(G,\mathbb{C}^{\times}) by monodromy: Sa,M=ξ(a)idSaM\mathcal{M}_{S_{a},M}=\xi(a)\mathrm{id}_{S_{a}\boxtimes M} with ξG\xi\in G^{\vee}. By a simple current extension of VV by a lattice theory, we mean a vertex operator superalgebra extension of the form

=aGSaVF(a)+L\mathcal{E}=\bigoplus_{a\in G}S_{a}\otimes V_{F(a)+L}

for some non-degenerate integral lattice LL with a group homomorphism F:GL/LF\colon G\hookrightarrow L^{\prime}/L. Here LL^{\prime} is the dual lattice L={aL(a,L)}L^{\prime}=\{a\in\mathbb{Q}\otimes_{\mathbb{Z}}L\mid(a,L)\subset\mathbb{Z}\} with which we may parametrize the simple VLV_{L}-modules Vλ+LV_{\lambda+L} (λL/L\lambda\in L^{\prime}/L). Take the sublattice NLN\subset L^{\prime} so that F:GN/LL/LF\colon G\simeq N/L\subset L^{\prime}/L. Then the fusion rings 𝒦(V)\mathcal{K}(V) and 𝒦()\mathcal{K}(\mathcal{E}) are related by the formulas

𝒦()(𝒦(V)[N/L][L/L])N/L,𝒦(V)(𝒦()[N/L][L/L])N/L.\displaystyle\mathcal{K}(\mathcal{E})\simeq\left(\mathcal{K}(V)\underset{\mathbb{Z}[N/L]}{\otimes}\mathbb{Z}[L^{\prime}/L]\right)^{N/L},\quad\mathcal{K}(V)\simeq\left(\mathcal{K}(\mathcal{E})\underset{\mathbb{Z}[N^{\prime}/L]}{\otimes}\mathbb{Z}[L^{\prime}/L]\right)^{N^{\prime}/L}.

In our setting, they are

𝒦(Lc(𝔫𝔰2))(𝒦(Lk(𝔰𝔩2))[2][2(k+2)])2,\displaystyle\mathcal{K}(L_{c}(\mathfrak{ns}_{2}))\simeq\left(\mathcal{K}(L_{k}(\mathfrak{sl}_{2}))\underset{\mathbb{Z}[\mathbb{Z}_{2}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{2(k+2)}]\right)^{\mathbb{Z}_{2}}, (23)
𝒦(Lk(𝔰𝔩2))(𝒦(Lc(𝔫𝔰2))[k+2][2(k+2)])k+2.\displaystyle\mathcal{K}(L_{k}(\mathfrak{sl}_{2}))\simeq\left(\mathcal{K}(L_{c}(\mathfrak{ns}_{2}))\underset{\mathbb{Z}[\mathbb{Z}_{k+2}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{2(k+2)}]\right)^{\mathbb{Z}_{k+2}}. (24)

We note that in Ref. [2] the fusion rules of Lc(𝔫𝔰2)L_{c}(\mathfrak{ns}_{2}) were actually derived from (23).

Finally, we compare the fusion rings from the common ground, that is the fusion ring of their Heisenberg coset, known as the parafermion algebra. For k0k\in\mathbb{Z}_{\geq 0}, it enjoys the level-rank duality[12] and is identified with a simple principal 𝒲\mathcal{W}-algebra

Com(πh,Lk(𝔰𝔩2))𝒲(𝔰𝔩k),=k+k+2k+1.\displaystyle\operatorname{Com}\left(\pi^{h},L_{k}(\mathfrak{sl}_{2})\right)\simeq\mathcal{W}_{\ell}(\mathfrak{sl}_{k}),\quad\ell=-k+\frac{k+2}{k+1}. (25)

The 𝒲\mathcal{W}-algebra 𝒲(𝔰𝔩k)\mathcal{W}_{\ell}(\mathfrak{sl}_{k}) is rational and C2C_{2}-cofinite[8, 7] and known as one of the discrete series representation. The module category 𝒲(𝔰𝔩k)\mathcal{W}_{\ell}(\mathfrak{sl}_{k})-mod is a semisimple braided tensor category with simple modules parametrized by those of L2(𝔰𝔩k)L_{2}(\mathfrak{sl}_{k}), say 𝐋𝒲k(λ)\mathbf{L}_{\mathcal{W}}^{k}(\lambda), satisfying the same fusion rules.[13, 48, 20] On the other hand, the double commutant Com(𝒲(𝔰𝔩k),Lk(𝔰𝔩2))\operatorname{Com}(\mathcal{W}_{\ell}(\mathfrak{sl}_{k}),L_{k}(\mathfrak{sl}_{2})) extending the Heisenberg vertex algebra πh\pi^{h} turns out to be a lattice vertex algebra V2kV_{\sqrt{2k}\mathbb{Z}}. Then Lk(𝔰𝔩2)L_{k}(\mathfrak{sl}_{2}) decomposes as a module over 𝒲(𝔰𝔩k)V2k\mathcal{W}_{\ell}(\mathfrak{sl}_{k})\otimes V_{\sqrt{2k}\mathbb{Z}} into

Lk(𝔰𝔩2)ik𝐋𝒲k(2ϖi)V2i2k+2k\displaystyle L_{k}(\mathfrak{sl}_{2})\simeq\bigoplus_{i\in\mathbb{Z}_{k}}\mathbf{L}_{\mathcal{W}}^{k}(2\varpi_{i})\otimes V_{\frac{2i}{\sqrt{2k}}+\sqrt{2k}\mathbb{Z}}

and, similarly,

Lc(𝔫𝔰2)ik𝐋𝒲k(2ϖi)V(k+2)i(k+2)k+(k+2)k.\displaystyle L_{c}(\mathfrak{ns}_{2})\simeq\bigoplus_{i\in\mathbb{Z}_{k}}\mathbf{L}_{\mathcal{W}}^{k}(2\varpi_{i})\otimes V_{\frac{(k+2)i}{\sqrt{(k+2)k}}+\sqrt{(k+2)k}\mathbb{Z}}.

Therefore, Lk(𝔰𝔩2)L_{k}(\mathfrak{sl}_{2}) and Lc(𝔫𝔰2)L_{c}(\mathfrak{ns}_{2}) are simple current extensions by lattice theories, which imply the following description of fusion rings:

𝒦(Lk(𝔰𝔩2))(𝒦(L2(𝔰𝔩k))[k][2k])k,\displaystyle\mathcal{K}(L_{k}(\mathfrak{sl}_{2}))\simeq\left(\mathcal{K}(L_{2}(\mathfrak{sl}_{k}))\underset{\mathbb{Z}[\mathbb{Z}_{k}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{2k}]\right)^{\mathbb{Z}_{k}},
𝒦(Lc(𝔰𝔩2))(𝒦(L2(𝔰𝔩k))[(k+2)k][(k+2)k])k.\displaystyle\mathcal{K}(L_{c}(\mathfrak{sl}_{2}))\simeq\left(\mathcal{K}(L_{2}(\mathfrak{sl}_{k}))\underset{\mathbb{Z}[\mathbb{Z}_{(k+2)k}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{(k+2)k}]\right)^{\mathbb{Z}_{k}}.

3. Feigin–Semikhatov conjecture

3.1. Main statements

To obtain a fruitful generalization of the relation between Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) and Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}), it is better to view Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) as the principal 𝒲\mathcal{W}-superalgebra associated with 𝔰𝔩2|1\mathfrak{sl}_{2|1}:

Vc(𝔫𝔰2)𝒲(𝔰𝔩2|1),c=3(2+1),\displaystyle V^{c}(\mathfrak{ns}_{2})\simeq\mathcal{W}^{\ell}(\mathfrak{sl}_{2|1}),\quad c=-3(2\ell+1),

and then rewrite the relation (25) into

(k+2)(+1)=1.\displaystyle(k+2)(\ell+1)=1. (26)

Note that the numbers two and one in each factor are the dual Coxeter numbers of 𝔰𝔩2\mathfrak{sl}_{2} and 𝔰𝔩2|1\mathfrak{sl}_{2|1}. This implies that the relation between Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) and Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) is a variant of the celebrated Feigin–Frenkel duality[41] for the principal 𝒲\mathcal{W}-algebras, which asserts an isomorphism

𝒲k(𝔰𝔩n)𝒲kˇ(𝔰𝔩n),(k+n)(kˇ+n)=1.\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n})\simeq\mathcal{W}^{\check{k}}(\mathfrak{sl}_{n}),\quad(k+n)(\check{k}+n)=1. (27)

in the case of type A.

Recall that the 𝒲\mathcal{W}-superalgebras in general are defined from the affine vertex superalgebras Vκ(𝔤)V^{\kappa}(\mathfrak{g}) for (simple) basic classical Lie superalgebras 𝔤\mathfrak{g} with even supersymmetric invariant bilinear forms κ\kappa via the quantum Drinfeld–Sokolov reductions parametrized by the conjugacy classes of even nilpotent elements ff,

𝒲κ(𝔤,f):=HDS,f0(Vκ(𝔤)).\displaystyle\mathcal{W}^{\kappa}(\mathfrak{g},f):=H^{0}_{\mathrm{DS},f}(V^{\kappa}(\mathfrak{g})).

When 𝔤=𝔰𝔩n|m\mathfrak{g}=\mathfrak{sl}_{n|m} (nmn\neq m), the bilinear forms κ\kappa are all proportional to the super trace str\mathrm{str} and we identify κ\kappa with its ratio kk\in\mathbb{C}. The conjugacy classes are in one-to-one correspondence with the pair of partitions of nn and mm. For 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} they are 2,122,1^{2} and correspond to the Virasoro vertex algebra and Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}), respectively. For 𝔤=𝔰𝔩2|1\mathfrak{g}=\mathfrak{sl}_{2|1}, they are (2|1),(12|1)(2|1),(1^{2}|1) and correspond to the 𝒩=2\mathcal{N}=2 superconformal algebra and Vk(𝔰𝔩2|1)V^{k}(\mathfrak{sl}_{2|1}), respectively.

The set of all the even nilpotent elements forms an algebraic subvariety 𝒩\mathcal{N} inside 𝔰𝔩n|m\mathfrak{sl}_{n|m} and decomposes into conjugacy classes 𝒩=λ𝒩λ\mathcal{N}=\sqcup_{\lambda}\mathcal{N}_{\lambda}. There is a unique class, called regular or principal, characterized as maximizing the dimension dim𝒩λ\mathrm{dim}\ \mathcal{N}_{\lambda}. The partition is given by (n|m)(n|m) and the corresponding 𝒲\mathcal{W}-superalgebra 𝒲k(𝔰𝔩n|m)=𝒲k(𝔰𝔩n|m,fprin)\mathcal{W}^{k}(\mathfrak{sl}_{n|m})=\mathcal{W}^{k}(\mathfrak{sl}_{n|m},f_{\mathrm{prin}}) with fprin=fn|mf_{\mathrm{prin}}=f_{n|m} is called principal. If 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}, then there is a unique class, called subregular, of the second largest dimension. The partition is given by (n1)111(n-1)^{1}1^{1} and the corresponding 𝒲\mathcal{W}-algebra 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) with fsub=f(n1)111f_{\mathrm{sub}}=f_{(n-1)^{1}1^{1}} is called subregular.

Feigin and Semikhatov[38] conjectured that the pair Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) and 𝒲(𝔰𝔩2|1)\mathcal{W}^{\ell}(\mathfrak{sl}_{2|1}) generalizes to the series of pairs

𝒲k(𝔰𝔩n,fsub),𝒲(𝔰𝔩n|1,fprin).\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}),\quad\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}). (28)

The 𝒲\mathcal{W}-superalgebras (28) have rank one Heisenberg vertex subalgebras as the maximal affine subalgebras: we normalize the generators

H+(z)=nH+,nzn1,H(z)=nH,nzn1\displaystyle H_{+}(z)=\sum_{n\in\mathbb{Z}}H_{+,n}z^{-n-1},\quad H_{-}(z)=\sum_{n\in\mathbb{Z}}H_{-,n}z^{-n-1} (29)

so that the eigenvalues of the zero mode H±,0H_{\pm,0} on the 𝒲\mathcal{W}-superalgebras are normalized to \mathbb{Z}. Then the OPEs are

H+(z)H+(w)ε+(zw)2,H(z)H(w)ε(zw)2.\displaystyle H_{+}(z)H_{+}(w)\sim\frac{\varepsilon_{+}}{(z-w)^{2}},\quad H_{-}(z)H_{-}(w)\sim\frac{\varepsilon_{-}}{(z-w)^{2}}.

with

ε+=n1n(k+n)1,ε=nn1(+n1)+1\displaystyle\varepsilon_{+}=\tfrac{n-1}{n}(k+n)-1,\quad\varepsilon_{-}=-\tfrac{n}{n-1}(\ell+n-1)+1 (30)

The precise analogy of the Feigin–Frenkel duality for the pairs (28) is the duality between the Heisenberg cosets of these 𝒲\mathcal{W}-superalgebras:

Theorem 3.1.

[29, 23] There is an isomorphism of vertex algebras

𝐅𝐒:Com(πH+,𝒲k(𝔰𝔩n,fsub))Com(πH,𝒲(𝔰𝔩n|1,fprin)){\bf FS}\colon\operatorname{Com}\left(\pi^{H_{+}},\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\right)\simeq\operatorname{Com}\left(\pi^{H_{-}},\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\right)

for (k,)(n+nn1,(n1)+n1n)(k,\ell)\neq(-n+\frac{n}{n-1},-(n-1)+\frac{n-1}{n}) satisfying the relation

(k+n)(+n1)=1.\displaystyle(k+n)(\ell+n-1)=1. (31)

These pairs indeed enjoy the coset constructions:

Theorem 3.2.

[23] There are isomorphisms of vertex superalgebras

𝐊𝐒:𝒲(𝔰𝔩n|1,fprin)Com(πH+Δ,𝒲k(𝔰𝔩n,fsub)V),\displaystyle{\bf KS}\colon\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\xrightarrow{\simeq}\operatorname{Com}(\pi^{H_{+}^{\Delta}},\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\otimes V_{\mathbb{Z}}),
𝐅𝐒𝐓:𝒲k(𝔰𝔩n,fsub)Com(πHΔ,𝒲(𝔰𝔩n|1,fprin)V1)\displaystyle{\bf FST}\colon\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\xrightarrow{\simeq}\operatorname{Com}(\pi^{H_{-}^{\Delta}},\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}})

for all (k,)(k,\ell) satisfying (31). Here we set

H+Δ(z)=H+(z)+b1(z),HΔ(z)=H(z)+b1(z).H_{+}^{\Delta}(z)=-H_{+}(z)+b_{1}(z),\quad H_{-}^{\Delta}(z)=H_{-}(z)+b_{\operatorname{\sqrt{\smash[b]{-1}}}}(z).

Note that we can prove Theorem 3.2 for n=2n=2 for Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) by checking the OPEs directly. However, we cannot use this approach literally since we do not know the whole defining OPEs of 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}). There are two approaches to overcome this difficulty: one is the uniqueness property of these algebras which is am important feature for the hook-type 𝒲\mathcal{W}-superalgebras in general as proved by Creutzig and Linshaw [29, 30] and the other one is free field realizations of these 𝒲\mathcal{W}-superalgebras.

3.2. Uniqueness property

The 𝒲\mathcal{W}-algebra 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) has a free strong generating set of the form

𝒲k(𝔰𝔩n,fsub)=𝒲(1,2,,n1,(n2)2),\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})=\mathcal{W}\left(1,2,\cdots,n-1,(\tfrac{n}{2})^{2}\right),

that is, a generating set whose conformal weights are 1,2,n21,2\cdots,\tfrac{n}{2} with described multiplicity such that the ordered monomials of their negative modes give a basis of the algebra. The field of conformal weight one is the Heisenberg field H+(z)H_{+}(z) and the field of conformal weight two is the Virasoro field. The fields of conformal weight n2\frac{n}{2} are those generators which have nontrivial H+,0H_{+,0}-weights, namely ±1\pm 1.

Similarly, the 𝒲\mathcal{W}-superalgebra 𝒲(𝔰𝔩n|1,frpin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{rpin}}) has a free strong generating set of the form

𝒲(𝔰𝔩n|1,fprin)=𝒲(1,2,,n,(n+12)2).\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})=\mathcal{W}\left(1,2,\cdots,n,(\tfrac{n+1}{2})^{2}\right).

Again, the field of conformal weight one is the Heisenberg field H(z)H_{-}(z), the field of conformal weight two is the Virasoro field and the fields of conformal weight n+12\frac{n+1}{2} are those generators which have nontrivial H,0H_{-,0}-weights, namely ±1\pm 1 and are the only odd generators.

The Heisenberg cosets

Com(πH+,𝒲k(𝔰𝔩n,fsub)),Com(πH,𝒲(𝔰𝔩n|1,fprin))\displaystyle\operatorname{Com}\left(\pi^{H_{+}},\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\right),\quad\operatorname{Com}\left(\pi^{H_{-}},\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\right) (32)

have a special feature: they have a common strong generating set of the form

𝒲(2,3,,2n+1)\displaystyle\mathcal{W}(2,3,\cdots,2n+1) (33)

which is the same as the principal 𝒲\mathcal{W}-algebra 𝒲k(𝔰𝔩2n+1)\mathcal{W}^{k}(\mathfrak{sl}_{2n+1}), in which case it is free. In our case, the generating set is not free and indeed has two linearly independent relation at conformal dimension 2n+42n+4, which is a consequence of a qq-character (35) below.

The important fact is that one-parameter families of vertex algebras of type 𝒲(2,3,,N)\mathcal{W}(2,3,\cdots,N) are always obtained as quotients of the two-parameter universal 𝒲\mathcal{W}_{\infty}-algebra 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] constructed mathematically by Linshaw [65], whose existence has been conjectured for decades, see the references therein. It is a vertex algebra over the commutative ring [c,λ]\mathbb{C}[c,\lambda] and has a strong free generating set of the form

𝒲[c,λ]=𝒲(2,3,4,).\mathcal{W}_{\infty}[c,\lambda]=\mathcal{W}(2,3,4,\cdots).

By setting Wm(z)W_{m}(z) to be the generating field of conformal degree mm, cc parametrizes the central charge of the Virasoro field W2(z)W_{2}(z) and λ\lambda parametrizes the undetermined scalar appearing in the OPE

W2(z)W5(w)5(16(c+2)λ37)W3(w)(zw)4\displaystyle W_{2}(z)W_{5}(w)\sim\frac{-5(16(c+2)\lambda-37)W_{3}(w)}{(z-w)^{4}}
+(16(c+2)λ55)wW3(w)(zw)3+5W5(w)(zw)2+wW5(w)(zw).\displaystyle\hskip 113.81102pt+\frac{-(16(c+2)\lambda-55)\partial_{w}W_{3}(w)}{(z-w)^{3}}+\frac{5W_{5}(w)}{(z-w)^{2}}+\frac{\partial_{w}W_{5}(w)}{(z-w)}.

The remaining OPEs are uniquely determined by the Jacobi identities under the assumption that there is an automorphism sending Wm(z)(1)mWm(z)W_{m}(z)\mapsto(-1)^{m}W_{m}(z). Both of the coset algebras (32) and 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}) (n3n\geq 3) are obtained as one-parameter families of vertex algebras from 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] with different algebraic relations between cc and λ\lambda, called truncation curves in general. In the case 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}), the parameters c,λc,\lambda are

c=(n1)(1n(n+1)(k+n1)2k+n),\displaystyle c=(n-1)\left(1-\frac{n(n+1)(k+n-1)^{2}}{k+n}\right),
λ=n+k(n2)(n2+nkn2)(n2+nk+n+2k),\displaystyle\lambda=-\frac{n+k}{(n-2)(n^{2}+nk-n-2)(n^{2}+nk+n+2k)},

giving a coordinate of the truncation curve 𝐂pr(n)\mathbf{C}_{\mathrm{pr}}(n)

λ(n2)(3n2n2+c(n+2))=(n21).\lambda(n-2)(3n^{2}-n-2+c(n+2))=(n^{2}-1).

The 𝒲\mathcal{W}-algebra 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}) is obtained from the specialization of 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] at these values via the quotient by the ideal generated by the null vector of conformal weight n+1n+1. On the other hand, (32) is obtained from 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] by the specialization

c=n(n(n3)+k(n2)+1)(n(n2)+k(n1)1)k+n,\displaystyle c=-\frac{n(n(n-3)+k(n-2)+1)(n(n-2)+k(n-1)-1)}{k+n},
λ=(k+n)(k+n1)(n(n4)+k(n3)+2)(n(n2)+k(n1)2)(n2+nk+k),\displaystyle\lambda=-\frac{(k+n)(k+n-1)}{(n(n-4)+k(n-3)+2)(n(n-2)+k(n-1)-2)(n^{2}+nk+k)},

which parametrizes the truncation curve 𝐂sub(n)\mathbf{C}_{\mathrm{sub}}(n)

c3λ2(n3)(n2)(n+1)(n+2)+c2λ(3(4λ+1)+λn(8n433n3+2n2+99n28)2(n1)n(n2n5))+cn(13λ+λ2(n1)(12n488n3+179n2+λn20)λn(10n337n2+18n+34)+n(n22n1)+2)n2(λ(n5)(3n+1)2(n+1))(5λ+(4λ+1)(n3)n+2)=0.\displaystyle\begin{split}&c^{3}\lambda^{2}(n-3)(n-2)(n+1)(n+2)\\ &+c^{2}\lambda(-3(4\lambda+1)+\lambda n(8n^{4}-33n^{3}+2n^{2}+99n-28)-2(n-1)n(n^{2}-n-5))\\ &+cn\binom{13\lambda+\lambda^{2}(n-1)(12n^{4}-88n^{3}+179n^{2}+\lambda n-20)\hskip 85.35826pt}{\hskip 85.35826pt-\lambda n(10n^{3}-37n^{2}+18n+34)+n(n^{2}-2n-1)+2}\\ &-n^{2}(\lambda(n-5)(3n+1)-2(n+1))(5\lambda+(4\lambda+1)(n-3)n+2)\\ &=0.\end{split}

The above uniqueness ensures the Heisenberg cosets (32) coincide. The whole 𝒲\mathcal{W}-superalgebra 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) (resp. 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})) is characterized[29] by the vertex superalgebra containing the tensor product of the coset and πH+\pi^{H_{+}} (resp. πH\pi^{H_{-}}), extended by even (resp. odd) highest weight vectors of conformal weight n2\frac{n}{2} (resp. n+12\frac{n+1}{2}) and of Heisenberg weight ±1\pm 1. This gives a proof of Theorem 3.2 at generic levels.

The level-rank duality for the parafermion algebra (25) can be systematically found[29] at the intersection points of the truncation curves 𝐂sub(n)\mathbf{C}_{\mathrm{sub}}(n) and 𝐂pr(m)\mathbf{C}_{\mathrm{pr}}(m).

Theorem 3.3.

[12, 9, 29] There is an isomorphism of vertex algebras

Com(πH+,𝒲k(𝔰𝔩n,fsub))𝒲r(𝔰𝔩m)\operatorname{Com}\left(\pi^{H_{+}},\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\right)\simeq\mathcal{W}_{r}(\mathfrak{sl}_{m})

for the following shifted levels (k+n,r+m)(k+n,r+m) with m3m\geq 3:

(n+mn1,m+1m+n),(nn+m1,m1m+n1),(nmnm1,nmnm1).\displaystyle\left(\frac{n+m}{n-1},\frac{m+1}{m+n}\right),\quad\left(\frac{n}{n+m-1},\frac{m-1}{m+n-1}\right),\quad\left(\frac{n-m}{n-m-1},\frac{n-m}{n-m-1}\right).

The condition m3m\geq 3 comes from the construction of 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] that W3(z)W_{3}(z) does not vanish, which prohibits degenerations to the Virasoro vertex algebra. The case m=2m=2 is subtle: the statement is true in the first case [24], but not in the second case [4]. The latter case is of particular interest since the Heisenberg coset is the singlet algebra, which is an simple current extension of the Virasoro vertex algebra of infinite order. The algebra 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) itself is identified with the chiral algebras of Argyres–Douglas theories of type (A1,A2n1)(A_{1},A_{2n-1})[4].

3.3. A digression on qq-characters

Let us briefly discuss the qq-character of the Heisenberg cosets (32). Since the character of 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) is expressed as

tr𝒲k(𝔰𝔩n,fsub)(qL0zH+,0)=1(q,qn1,zqn2z1qn2;q),\displaystyle\mathrm{tr}_{\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})}(q^{L_{0}}z^{H_{+,0}})=\frac{1}{(q,\cdots q^{n-1},zq^{\frac{n}{2}}z^{-1}q^{\frac{n}{2}};q)_{\infty}},

the qq-character of the Heisenberg cosets is given by the residue

F(q):=dzz1(q2,qn1,zqn2,z1qn2;q).\displaystyle F(q):=\int\frac{dz}{z}\frac{1}{(q^{2},\cdots q^{n-1},zq^{\frac{n}{2}},z^{-1}q^{\frac{n}{2}};q)_{\infty}}. (34)

As a zz-independent form, it can be rewritten into

1(q;q)2(q,,qn;q)m(1)mΦm(q)(qn+12+m,qn+12m;q)qm(m+1)2\displaystyle\frac{1}{(q;q)_{\infty}^{2}(q,\cdots,q^{n};q)_{\infty}}\sum_{m\in\mathbb{Z}}(-1)^{m}\Phi_{m}(q)\left(q^{\frac{n+1}{2}+m},q^{\frac{n+1}{2}-m};q\right)_{\infty}q^{\frac{m(m+1)}{2}} (35)

by the unary false theta functions

Φm(q)=N=0(1)NqN(N+1)2+Nm.\displaystyle\Phi_{m}(q)=\sum_{N=0}^{\infty}(-1)^{N}q^{\frac{N(N+1)}{2}+Nm}. (36)

From this, we find that there are two independent linear relations among the PBW basis for strong generators at conformal dimension 2n+42n+4 in (33).

There is a combinatorial meaning[54], called melting crystals, for (34) when multiplied by the character of the rank one Heisenberg vertex algebra π\pi, which is 1/(q;q)1/(q;q)_{\infty}. It is based on the following fact. When we extend 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] by π\pi, the resulting vertex algebra 𝒲1+=π𝒲[c,λ]\mathcal{W}_{1+\infty}=\pi\otimes\mathcal{W}_{\infty}[c,\lambda] is of type 𝒲(1,2,3,)\mathcal{W}(1,2,3,\cdots) and the qq-character is given by the MacMahon function

1n=1(1qn)n.\frac{1}{\prod_{n=1}^{\infty}(1-q^{n})^{n}}.

It is known as the generating function counting plane partitions (i.e., three dimensional partitions located in the region {x,y,z0}3\{x,y,z\geq 0\}\subset\mathbb{R}^{3}). In the same line, the function F(q)/(q;q)F(q)/(q;q)_{\infty} is the generating function of plane partitions with a pit[17] at (2,n+1)(2,n+1) as in Fig. 6. The first excluded plane partition is the one in Fig. 6, which has 2(n+1)2(n+1) boxes. This corresponds to the fact that the field W2n+2(z)W_{2n+2}(z) drops off from the strong generating set in (33).

Figure 5. Plane partiton with a pit at (2,n+1)(2,n+1)
7524231101\vdots\vdots\vdots\vdots\vdots\vdots\cdots\cdots\cdots\cdots}\left.\begin{array}[]{c}\\ \\ \\ \end{array}\right\}n+1n+1
Figure 6. First excluded partition
111111\vdots\vdots}\left.\begin{array}[]{c}\\ \\ \\ \end{array}\right\}n+1n+1

3.4. Free field realization

Beside the uniqueness property of the 𝒲\mathcal{W}-superalgebras based on 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda], the free field realization of 𝒲\mathcal{W}-superalgebras is also a powerful tool as in the original proof[41] of the Feigin–Frenkel duality (27). In general, we have the Miura map [59]

𝒲k(𝔤,f)Vk(𝔤0)SB(𝔤1/2),\mathcal{W}^{k}(\mathfrak{g},f)\hookrightarrow V^{k^{\prime}}(\mathfrak{g}_{0})\otimes SB(\mathfrak{g}_{1/2}),

which embeds the 𝒲\mathcal{W}-superalgebra into a tensor product of the affine vertex algebra Vk(𝔤0)V^{k^{\prime}}(\mathfrak{g}_{0}) at a suitable level and the symplectic boson associated with the symplectic vector superspace 𝔤1/2\mathfrak{g}_{1/2}. Here we use the good grading 𝔤=j12𝔤j\mathfrak{g}=\bigoplus_{j\in\frac{1}{2}\mathbb{Z}}\mathfrak{g}_{j} used in the quantum Drinfeld–Sokolov reduction. In our cases,

𝒲k(𝔰𝔩n,fsub)Vk(𝔰𝔩2)π𝔥k+n,𝒲(𝔰𝔩n|1,fprin)V(𝔤𝔩1|1)π𝔥+n1.\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\hookrightarrow V^{k^{\prime}}(\mathfrak{sl}_{2})\otimes\pi^{k+n}_{\mathfrak{h}^{\perp}},\quad\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\hookrightarrow V^{\ell^{\prime}}(\mathfrak{gl}_{1|1})\otimes\pi^{\ell+n-1}_{\mathfrak{h}^{\perp}}. (37)

Here 𝔥𝔤\mathfrak{h}^{\perp}\subset\mathfrak{g} is the subspace of the standard Cartan subalgebra 𝔥\mathfrak{h} orthogonal to 𝔰𝔩2\mathfrak{sl}_{2} and 𝔤𝔩1|1\mathfrak{gl}_{1|1} respectively. When k,k,\ell are generic, the images coincide with the joint kernels of screening operators associated with highest weight vectors of Fock modules for the Heisenberg vertex algebras. In the weighted Dynkin diagrams below, they are canonically attached to the nodes of weight one.

Figure 7. 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}
11α1\alpha_{1}11α2\alpha_{2}\cdot\cdot\cdot11αn2\alpha_{n-2}0αn1\alpha_{n-1}
Figure 8. 𝔤=𝔰𝔩n|1\mathfrak{g}=\mathfrak{sl}_{n|1}
11α1\alpha_{1}11α2\alpha_{2}\cdot\cdot\cdot11αn1\alpha_{n-1}×\times0αn\alpha_{n}

We compose them with the Wakimoto realization

Vk(𝔰𝔩2)βγπ𝔥k+n,V(𝔤𝔩1|1)bcπ𝔥+n1Vπ𝔥+n1\displaystyle V^{k^{\prime}}(\mathfrak{sl}_{2})\hookrightarrow\beta\gamma\otimes\pi^{k+n}_{\mathfrak{h}},\quad V^{\ell^{\prime}}(\mathfrak{gl}_{1|1})\hookrightarrow bc\otimes\pi_{\mathfrak{h}}^{\ell+n-1}\simeq V_{\mathbb{Z}}\otimes\pi_{\mathfrak{h}}^{\ell+n-1} (38)

and then the bosonization[51] of the βγ\beta\gamma-system

βγΠ:=V(1+1)π(1+1)π1.\displaystyle\beta\gamma\hookrightarrow\Pi:=V_{(1+\operatorname{\sqrt{\smash[b]{-1}}})\mathbb{Z}}\underset{\pi^{(1+\operatorname{\sqrt{\smash[b]{-1}}})\mathbb{Z}}}{\otimes}\pi^{\mathbb{Z}\oplus\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}. (39)

In the end, we arrive at the following free field realizations of the 𝒲\mathcal{W}-superalgebras:

𝒲k(𝔰𝔩n,fsub)Ππ𝔥k+n,𝒲(𝔰𝔩n|1,fprin)Vπ𝔥+n1.\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\hookrightarrow\Pi\otimes\pi^{k+n}_{\mathfrak{h}},\quad\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\hookrightarrow V_{\mathbb{Z}}\otimes\pi^{\ell+n-1}_{\mathfrak{h}}. (40)

At each step (38)-(39), the images are described by the kernel of a single screeing operator. Therefore, at generic levels k,k,\ell, the images of (40) are described by the joint kernels of screening operators. The screening operators for (38) are attached to the nodes of weight zero in Fig. 8-8, respectively. The screening operator for (39) adds one odd node to Fig. 8 as below.

Figure 9. Screening operators for 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})
11α1\alpha_{1}11α2\alpha_{2}\cdot\cdot\cdot11αn2\alpha_{n-2}0αn1\alpha_{n-1}×\times

Now, we may apply the Feigin–Frenkel duality for the Virasoro vertex algebras for each (even) node in Fig. 8-9 since the Heisenberg cosets concentrate on the Heisenberg vertex algebras by (40):

Com(πH+,𝒲k(𝔰𝔩n,fsub))Com(πH+,π1π𝔥k+n),Com(πH,𝒲(𝔰𝔩n|1,fprin))Com(πH,ππ𝔥+n1).\displaystyle\begin{split}&\operatorname{Com}\left(\pi^{H_{+}},\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\right)\hookrightarrow\operatorname{Com}\left(\pi^{H_{+}},\pi^{\mathbb{Z}\oplus\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\otimes\pi^{k+n}_{\mathfrak{h}}\right),\\ &\operatorname{Com}\left(\pi^{H_{-}},\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\right)\hookrightarrow\operatorname{Com}\left(\pi^{H_{-}},\pi^{\mathbb{Z}}\otimes\pi^{\ell+n-1}_{\mathfrak{h}}\right).\end{split} (41)

This proves Theorem 3.1 for generic levels. Note that these cosets form continuous families of subspaces inside the common Heisenberg vertex algebra and that the Feigin–Frenkel duality for the Virasoro vertex algebras asserts that they are the same as vector spaces. Then it follows that the isomorphism in Theorem 3.1 remains true for all levels except for the level (k,)=(n+nn1,(n1)+n1n)(k,\ell)=(-n+\frac{n}{n-1},-(n-1)+\frac{n-1}{n}), when πH±\pi^{H_{\pm}} degenerate to commutative vertex algebras. Theorem 3.2 can be obtained in the same way by using (40) itself.

Remark 3.4.

The screening realizations of the Heisenberg cosets (41) coincide with those proposed in Refs. 68, 53 based on Miura operators. Therefore, they coincide with the vertex algebras introduced by Bershtein–Feigin–Merzon [17], which act on the cohomology of the moduli spaces of spiked instantons of Nekrasov.[69]

From the perspective of the free field realization (40), the inverse Drinfeld–Sokolov reduction (21) is naturally explained: the appearance of the Virasoro vertex algebra comes from its free free field realization inside π𝔥k+2\pi^{k+2}_{\mathfrak{h}}, whose image is also characterized by an even screening operator. We stress that the screening operators for 𝒲k(𝔰𝔩2)\mathcal{W}^{k}(\mathfrak{sl}_{2}) and Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) have different form and that an automorphism of π𝔥k+2Π\pi^{k+2}_{\mathfrak{h}}\otimes\Pi is used to identify them. Nontheless, it is generalized to our setting together with the 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-side in the following way:

(48)

Here h(z)h_{\perp}(z) is the Heisenberg field corresponding to the matrix i=1nEi,i+nEn+1,n+1\sum_{i=1}^{n}E_{i,i}+nE_{n+1,n+1} inside 𝔰𝔩n|1\mathfrak{sl}_{n|1} in terms of the elementary matrices Ei,jE_{i,j}. The embedding

𝒲k(𝔰𝔩n,fsub)𝒲k(𝔰𝔩n)Π\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\hookrightarrow\mathcal{W}^{k}(\mathfrak{sl}_{n})\otimes\Pi

has been introduced and used by Adamović–Kawasetsu–Ridout[5] for the Bershadsky–Polyakov algebra 𝒲k(𝔰𝔩3,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{3},f_{\mathrm{sub}}) and by Fehily[37] in general, to construct relaxed highest weight modules which are a source of logarithmic conformal field theories. It is obvious that we may convert the horizontal arrow in (48) by using the other coset construction (Theorem 3.2) by adding necessary lattice vertex superalgebras. In the end, we find that these two 𝒲\mathcal{W}-superalgebras are contained as

𝒲k(𝔰𝔩n,fsub)𝒲k(𝔰𝔩n)V1𝒲(𝔰𝔩n|1,fprin),\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\subset\mathcal{W}^{k}(\mathfrak{sl}_{n})\otimes V_{\mathbb{Z}\oplus\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\supset\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}),

where they just stretch out in different lattices, namely (1+1)(1+\operatorname{\sqrt{\smash[b]{-1}}})\mathbb{Z} and \mathbb{Z}. This explains why the module categories of 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}) look similar in general.

4. Correspondence of representation categories

Since we have the coset construction (Theorem 3.2) as in the case of Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) and Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}) in Section 2.1, we can generalize the equivalence of weight module categories and compare the fusion data in the rational setting.

4.1. Equivalence of categories

By using the normalized Heisenberg fields in (30), we introduce the category of weight modules consisting of all the modules which decompose into direct sums of Fock modules when restricted to the Heisenberg vertex subalgebras, that is Kazhdan–Lusztig objects for 𝔤𝔩^1\widehat{\mathfrak{gl}}_{1}. Motivated by this, we write

𝐊𝐋kn1,1,𝐊𝐋n|1\mathbf{KL}_{k}^{n-1,1},\quad\mathbf{KL}_{\ell}^{n|1}

for the weight module categories for 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}), respectively. They decompose into blocks

𝐊𝐋kn1,1=[λ]/𝐊𝐋k,[λ]n1,1,𝐊𝐋n|1=[μ]/𝐊𝐋,[μ]n|1\displaystyle\mathbf{KL}_{k}^{n-1,1}=\bigoplus_{[\lambda]\in\mathbb{C}/\mathbb{Z}}\mathbf{KL}_{k,[\lambda]}^{n-1,1},\quad\mathbf{KL}_{\ell}^{n|1}=\bigoplus_{[\mu]\in\mathbb{C}/\mathbb{Z}}\mathbf{KL}_{\ell,[\mu]}^{n|1}

in terms of the eigenvalues of H±,0H_{\pm,0} as the spectrum of H±,0H_{\pm,0} of the 𝒲\mathcal{W}-superalgebras is \mathbb{Z}. Then we have functors

Ωξ+:𝐊𝐋k,[λ]n1,1𝐊𝐋,[(ε1)ξ]n|1,Ωξ:𝐊𝐋,[μ]n|1𝐊𝐋k,[(ε++1)ξ]n1,1,\displaystyle\Omega^{+}_{\xi}\colon\mathbf{KL}_{k,[\lambda]}^{n-1,1}\rightarrow\mathbf{KL}_{\ell,[(\varepsilon_{-}-1)\xi]}^{n|1},\quad\Omega^{-}_{\xi^{\prime}}\colon\mathbf{KL}_{\ell,[\mu]}^{n|1}\rightarrow\mathbf{KL}_{k,[(\varepsilon_{+}+1)\xi^{\prime}]}^{n-1,1},

which assign the multiplicity space Ωξ+(M)\Omega^{+}_{\xi}(M) of the modules πξH+Δ\pi^{H_{+}^{\Delta}}_{\xi} inside MVM\otimes V_{\mathbb{Z}} to a 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-module MM and the multiplicity space Ωξ(N)\Omega^{-}_{\xi^{\prime}}(N) of the modules πξHΔ\pi^{H_{-}^{\Delta}}_{\xi^{\prime}} inside NV1N\otimes V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}} to a 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-module NN, respectively. Note that the ratios ε±±1\varepsilon_{\pm}\pm 1 are the norms of H±Δ(z)H_{\pm}^{\Delta}(z).

Theorem 4.1.

[24] The functors

Ωλ+:𝐊𝐋k,[λ]n1,1𝐊𝐋,[(1ε)λ]n|1:Ω(1ε)λ\displaystyle\Omega_{-\lambda}^{+}\colon\mathbf{KL}_{k,[\lambda]}^{n-1,1}\ \rightleftarrows\ \mathbf{KL}_{\ell,[(1-\varepsilon_{-})\lambda]}^{n|1}\colon\Omega^{-}_{(1-\varepsilon_{-})\lambda} (49)

are quasi-inverse to each other and thus give an equivalence of categories.

Thanks to the relation (31), the value 1ε1-\varepsilon_{-} coincides with the ratio of the Heisenberg fields H±(z)H_{\pm}(z), namely ε/ε+\varepsilon_{-}/\varepsilon_{+}. Later, we will use the square root:

ϵ=1ε=ε/ε+.\epsilon=\sqrt{1-\varepsilon_{-}}=\sqrt{\varepsilon_{-}/\varepsilon_{+}}.

4.2. Fusion rules: rational case

Here we compare the fusion data of the full subcategories

𝒲k(𝔰𝔩n,fsub)-mod𝐊𝐋kn1,1,𝒲(𝔰𝔩n|1,fprin)-mod𝐊𝐋n|1\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\text{-mod}\subset\mathbf{KL}_{k}^{n-1,1},\quad\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\text{-mod}\subset\mathbf{KL}_{\ell}^{n|1}

when (k,fsub)(k,f_{\mathrm{sub}}) form an exceptional pair[14, 60], i.e.

k=n+n+rn1,(r2,gcd(n+r,n1)=1).\displaystyle k=-n+\tfrac{n+r}{n-1},\quad(r\geq 2,\ \mathrm{gcd}(n+r,n-1)=1). (50)

In this case, 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) is rational [14] and C2C_{2}-cofinite [7], and thus 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-mod is a semisimple braided tensor category with finitely many simple objects. These algebraic and categorical properties are inherited to 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-side [23, 24].

To describe the fusion data for the simple exceptional 𝒲\mathcal{W}-algebra 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}), it is better to use the level-rank duality (Theorem 3.3):

Com(πH+,𝒲n+n+rn1(𝔰𝔩n,fsub))𝒲r+r+nr+1(𝔰𝔩r).\operatorname{Com}\left(\pi^{H_{+}},\mathcal{W}_{-n+\frac{n+r}{n-1}}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\right)\simeq\mathcal{W}_{-r+\frac{r+n}{r+1}}(\mathfrak{sl}_{r}).

The module category 𝒲r+r+nr+1(𝔰𝔩r)\mathcal{W}_{-r+\frac{r+n}{r+1}}(\mathfrak{sl}_{r})-mod is a semisimple braided tensor category with simple modules 𝐋𝒲r(λ)\mathbf{L}_{\mathcal{W}}^{r}(\lambda) satisfying the same fusion rules as Ln(𝔰𝔩r)L_{n}(\mathfrak{sl}_{r})-modules:

𝒦(𝒲r+r+nr+1(𝔰𝔩r))𝒦(Ln(𝔰𝔩r)),𝐋𝒲r(λ)Ln,λ.\displaystyle\mathcal{K}\left(\mathcal{W}_{-r+\frac{r+n}{r+1}}(\mathfrak{sl}_{r})\right)\xrightarrow{\simeq}\mathcal{K}(L_{n}(\mathfrak{sl}_{r})),\quad\mathbf{L}_{\mathcal{W}}^{r}(\lambda)\mapsto L_{n,\lambda}. (51)

Note that 𝐋𝒲r(nϖi)\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i}) (0ir10\leq i\leq r-1) are simple currents. They form a group r\mathbb{Z}_{r} since

𝐋𝒲r(nϖi)𝐋𝒲r(nϖj)𝐋𝒲r(nϖi+j).\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i})\boxtimes\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{j})\simeq\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i+j}).

Then the fusion ring is a ring over [r]\mathbb{Z}[\mathbb{Z}_{r}] by fusion products

𝐋𝒲r(nϖi)𝐋𝒲r(λ)𝐋𝒲r(σiλ)\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i})\boxtimes\mathbf{L}_{\mathcal{W}}^{r}(\lambda)\simeq\mathbf{L}_{\mathcal{W}}^{r}(\sigma^{i}\lambda)

where σ\sigma permutes the fundamental weights σϖi=ϖi+1\sigma\varpi_{i}=\varpi_{i+1}, and also admits a grading by the dual group r=Hom(r,×)\mathbb{Z}_{r}^{\vee}=\mathrm{Hom}(\mathbb{Z}_{r},\mathbb{C}^{\times}) coming from the monodromy

𝐋𝒲r(nϖi),𝐋𝒲r(λ)=ζriπP/Q(λ)\mathcal{M}_{\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i}),\mathbf{L}_{\mathcal{W}}^{r}(\lambda)}=\zeta_{r}^{-i\pi_{P/Q}(\lambda)}

with ζr=exp(2π1/r)\zeta_{r}=\mathrm{exp}(2\pi\operatorname{\sqrt{\smash[b]{-1}}}/r) and πP/Q(ϖi)=i\pi_{P/Q}(\varpi_{i})=i. Then 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) decomposes[29] into

𝒲k(𝔰𝔩n,fsub)ir𝐋𝒲r(nϖi)Vninr+nr.\displaystyle\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\simeq\bigoplus_{i\in\mathbb{Z}_{r}}\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i})\otimes V_{\frac{ni}{\sqrt{nr}}+\sqrt{nr}\mathbb{Z}}.

Therefore, 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) is a simple current extension of 𝒲r+r+nr+1(𝔰𝔩r)\mathcal{W}_{-r+\frac{r+n}{r+1}}(\mathfrak{sl}_{r}) by a lattice theory, which implies that the fusion ring is described[14, 24] as

𝒦(𝒲k(𝔰𝔩n,fsub))(𝒦(𝒲r+r+nr+1(𝔰𝔩r))[r][nr])r𝒦(Lr(𝔰𝔩n))\displaystyle\mathcal{K}(\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}))\simeq\left(\mathcal{K}\left(\mathcal{W}_{-r+\frac{r+n}{r+1}}(\mathfrak{sl}_{r})\right)\underset{\mathbb{Z}[\mathbb{Z}_{r}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{nr}]\right)^{\mathbb{Z}_{r}}\simeq\mathcal{K}(L_{r}(\mathfrak{sl}_{n}))

by the level-rank duality between the fusion rings of affine vertex algebras[49, 67, 24].

Accordingly, we have[24]

𝒲(𝔰𝔩n|1,fprin)ir𝐋𝒲r(nϖi)V(n+r)i(n+r)r+(n+r)r\displaystyle\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\simeq\bigoplus_{i\in\mathbb{Z}_{r}}\mathbf{L}_{\mathcal{W}}^{r}(n\varpi_{i})\otimes V_{\frac{(n+r)i}{\sqrt{(n+r)r}}+\sqrt{(n+r)r}\mathbb{Z}}

and then

𝒦(𝒲(𝔰𝔩n|1,fprin))(𝒦(Ln(𝔰𝔩r))[r][(n+r)r])r.\displaystyle\mathcal{K}(\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}))\simeq\left(\mathcal{K}(L_{n}(\mathfrak{sl}_{r}))\underset{\mathbb{Z}[\mathbb{Z}_{r}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{(n+r)r}]\right)^{\mathbb{Z}_{r}}.

The fusion data is compared in the following way:

Theorem 4.2.

[24] For k=n+n+rn1k=-n+\frac{n+r}{n-1} and =(n1)+n1n+r\ell=-(n-1)+\frac{n-1}{n+r} as in (50), we have isomorphisms of rings

𝒦(𝒲(𝔰𝔩n|1,fprin))(𝒦(𝒲k(𝔰𝔩n,fsub))[n][n(n+r)])n,\displaystyle\mathcal{K}(\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}))\simeq\left(\mathcal{K}(\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}))\underset{\mathbb{Z}[\mathbb{Z}_{n}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{n(n+r)}]\right)^{\mathbb{Z}_{n}},
𝒦(𝒲k(𝔰𝔩n,fsub))(𝒦(𝒲(𝔰𝔩n|1,fprin))[n+r][n(n+r)])n+r,\displaystyle\mathcal{K}(\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}))\simeq\left(\mathcal{K}(\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}))\underset{\mathbb{Z}[\mathbb{Z}_{n+r}]}{\otimes}\mathbb{Z}[\mathbb{Z}_{n(n+r)}]\right)^{\mathbb{Z}_{n+r}},

and 𝒦(𝒲k(𝔰𝔩n,fsub))𝒦(Lr(𝔰𝔩n))\mathcal{K}(\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}))\simeq\mathcal{K}(L_{r}(\mathfrak{sl}_{n})).

5. Correspondence of intertwining operators

5.1. Gluing approach

Let us discuss the monoidal correspondence beyond the rational case. In this generality, we do not know whether there is a suitable tensor structure on the category of weight modules and, even if it exists, the technique of tensor category at hand seems not to be sufficient. Thus, let us compare the bare spaces of logarithmic intertwining operators. For this purpose, it is meaningful to take a closer look at the coset construction (Theorem 3.1). Decompose the 𝒲\mathcal{W}-superalgebras 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}) as modules over the tensor product of the Heisenberg coset (32) and the Heisenberg vertex algebras πH±\pi^{H_{\pm}} inside them:

𝒲k(𝔰𝔩n,fsub)a𝒞+,akπaH+,𝒲(𝔰𝔩n|1,fprin)a𝒞,aπaH.\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\simeq\bigoplus_{a\in\mathbb{Z}}\mathscr{C}^{k}_{+,a}\otimes\pi^{H_{+}}_{a},\quad\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\simeq\bigoplus_{a\in\mathbb{Z}}\mathscr{C}^{\ell}_{-,a}\otimes\pi^{H_{-}}_{a}. (52)

Here 𝒞+,ak\mathscr{C}^{k}_{+,a} and 𝒞,a\mathscr{C}^{\ell}_{-,a} are the space of highest weight vector of the Fock modules. Then Theorem 3.1 implies that

𝒞+,ak𝒞,a\mathscr{C}^{k}_{+,a}\simeq\mathscr{C}^{\ell}_{-,a}

as modules over the Heisenberg cosets. In other words, the difference between these two 𝒲\mathcal{W}-superalgebras comes from the norms of the generators of the Heisenberg vertex subalgebras H+(z)H_{+}(z) and H(z)H_{-}(z). This difference can be seen most explicitly when their simple quotients 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}) degenerate to free field algebras:

𝒲2m+2m+12m1(𝔰𝔩2m,fsub)V2m,𝒲(2m1)+2m12m+1(𝔰𝔩2m|1,fprin)V2m+1.\displaystyle\mathcal{W}_{-2m+\frac{2m+1}{2m-1}}(\mathfrak{sl}_{2m},f_{\mathrm{sub}})\simeq V_{\sqrt{2m}\mathbb{Z}},\quad\mathcal{W}_{-(2m-1)+\frac{2m-1}{2m+1}}(\mathfrak{sl}_{2m|1},f_{\mathrm{prin}})\simeq V_{\sqrt{2m+1}\mathbb{Z}}.

The difference coming from the normalizations of Heisenberg fields is indeed responsible for the difference of module categories: the block-wise equivalence of categories is a formulation to get rid of this effect and the difference of the periodicity of simple currents (equivalently, spectral flows) in the rational case manifests this difference. Now, it is natural to seek for a way to change the normalizations of Heisenberg fields more directly. Note that it is equivalent to swap the Fock modules πaH+\pi^{H_{+}}_{a} and πaH\pi^{H_{-}}_{a} simultaneously.

The very role is played by the relative semi-infinite cohomology functor founded by Feigin[43] and Frenkel–Garland–Zuckerman [50] as conjectured by Creutzig–Linsahw [29, 30]. In our case, we use the the semi-infinite cohomology of 𝔤𝔩^1\widehat{\mathfrak{gl}}_{1} relative to the horizontal subalgebra 𝔤𝔩1\mathfrak{gl}_{1}. Let πA+\pi^{A^{+}} and πA\pi^{A^{-}} be Heisenberg vertex algebras generated by the fields A+(z)A^{+}(z) and A(z)A^{-}(z) satisfying the OPEs

A+(z)A(w)+1(zw)2,A(z)A(w)1(zw)2.A^{+}(z)A(w)^{+}\sim\frac{1}{(z-w)^{2}},\quad A^{-}(z)A^{-}(w)\sim\frac{-1}{(z-w)^{2}}.

Then the diagonal one

πdiag:=A+(z)1+1A(z)πA+πA\displaystyle\pi^{\mathrm{diag}}:=\langle A^{+}(z)\otimes 1+1\otimes A^{-}(z)\rangle\subset\pi^{A^{+}}\otimes\pi^{A^{-}} (53)

is a commutative vertex algebra which acts on the tensor product of Fock modules παA+πβA\pi_{\alpha}^{A^{+}}\otimes\pi^{A^{-}}_{\beta} of πA+πA\pi^{A^{+}}\otimes\pi^{A^{-}}. In this setting, we have

Hrel,n(𝔤𝔩1;παA+πβA)δα+β,0δn,0[eαeβ].\displaystyle H^{\mathrm{rel},n}_{\infty}\left(\mathfrak{gl}_{1};\pi^{A^{+}}_{\alpha}\otimes\pi^{A^{-}}_{\beta}\right)\simeq\delta_{\alpha+\beta,0}\delta_{n,0}\mathbb{C}[\mathrm{e}^{\alpha}\otimes\mathrm{e}^{\beta}]. (54)

Therefore, to replace the Fock module πaH+\pi^{H_{+}}_{a} inside 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) with πaH\pi^{H_{-}}_{a} simultaneously in order to obtain the shape of 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}), we introduce the following module

K+:=aπa1H+πaH\displaystyle K_{+\rightarrow-}:=\bigoplus_{a\in\mathbb{Z}}\pi_{-a}^{\operatorname{\sqrt{\smash[b]{-1}}}H_{+}}\otimes\pi^{H_{-}}_{a} (55)

over π1H+πH\pi^{\operatorname{\sqrt{\smash[b]{-1}}}H_{+}}\otimes\pi^{H_{-}}. Then (54) already implies that

Hrel,0(𝔤𝔩1;𝒲k(𝔰𝔩n,fsub)K+)a𝒞+,akπaH𝒲(𝔰𝔩n|1,fprin)\displaystyle H_{\infty}^{\mathrm{rel},0}\left(\mathfrak{gl}_{1};\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\otimes K_{+\rightarrow-}\right)\simeq\bigoplus_{a\in\mathbb{Z}}\mathscr{C}_{+,a}^{k}\otimes\pi_{a}^{H_{-}}\simeq\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}) (56)

as modules over Com(πH,𝒲(𝔰𝔩n|1,fprin))πH\operatorname{Com}(\pi^{H_{-}},\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}))\otimes\pi^{H_{-}}. It is crucial here to observe that (55) admits a vertex superalgebra structure

K+VπAK_{+\rightarrow-}\simeq V_{\mathbb{Z}}\otimes\pi^{A^{-}}

by changing basis of the two dimensional subspace of Heisenberg fields inside. Here again appears the lattice vertex superalgebra VV_{\mathbb{Z}} used in the coset construction (Theorem 3.2). Thanks to this, on can show that (56) is an isomorphism of vertex superalgebras. The appearance of VV_{\mathbb{Z}} is not a mere coincidence: indeed, when we manipulate the opposite direction in the same way, we find that V1V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}} appears as the main part of the gluing object

𝒲k(𝔰𝔩n,fsub)Hrel,0(𝔤𝔩1;𝒲(𝔰𝔩n|1,fprin)K+),\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\simeq H_{\infty}^{\mathrm{rel},0}\left(\mathfrak{gl}_{1};\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\otimes K_{-\rightarrow+}\right),
K+V1πA+.\displaystyle K_{-\rightarrow+}\simeq V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\otimes\pi^{A^{+}}.

To summarize, we obtain the following reformulation of the coset construction:

Theorem 5.1.

[24] There exist isomorphisms of vertex superalgebras

𝒲(𝔰𝔩n|1,fprin)Hrel,0(𝔤𝔩1;𝒲k(𝔰𝔩n,fsub)K+),𝒲k(𝔰𝔩n,fsub)Hrel,0(𝔤𝔩1;𝒲(𝔰𝔩n|1,fprin)K+),\displaystyle\begin{split}&\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\simeq H_{\infty}^{\mathrm{rel},0}\left(\mathfrak{gl}_{1};\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\otimes K_{+\rightarrow-}\right),\\ &\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\simeq H_{\infty}^{\mathrm{rel},0}\left(\mathfrak{gl}_{1};\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\otimes K_{-\rightarrow+}\right),\end{split} (57)

with

K+=VπA,K+=V1πA+.\displaystyle K_{+\rightarrow-}=V_{\mathbb{Z}}\otimes\pi^{A^{-}},\quad K_{-\rightarrow+}=V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\otimes\pi^{A^{+}}. (58)

5.2. Correspondence

To study the correspondence of intertwining operators, let us start with reformulating the equivalence of weight module categories (Theorem 4.1) through the gluing approach. For the coset functor, we have used the multiplicity space of Fock modules with non-trivial highest weights. There is a similar room for the gluing objects (58): we may replace the rank one Heisenberg vertex algebra by the Fock modules. We set

K+λ:=VπλAμπμϵ1λ1H+πμ+ϵλH,K+λ:=V1πλA+μπμϵλ1Hπμ+ϵ1λH+.\displaystyle\begin{split}&K_{+\rightarrow-}^{\lambda}:=V_{\mathbb{Z}}\otimes\pi^{A^{-}}_{\lambda}\simeq\bigoplus_{\mu\in\mathbb{Z}}\pi^{\operatorname{\sqrt{\smash[b]{-1}}}H_{+}}_{-\mu-\epsilon^{-1}\lambda}\otimes\pi^{H_{-}}_{\mu+\epsilon\lambda},\\ &K_{-\rightarrow+}^{\lambda}:=V_{\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}\otimes\pi^{A^{+}}_{\lambda}\simeq\bigoplus_{\mu\in\mathbb{Z}}\pi^{\operatorname{\sqrt{\smash[b]{-1}}}H_{-}}_{-\mu-\epsilon\lambda}\otimes\pi^{H_{+}}_{\mu+\epsilon^{-1}\lambda}.\end{split} (59)

By Theorem 5.1, they give rise to functors

H+,λrel():=Hrel,0(𝔤𝔩1;K+λ):𝐊𝐋kn1,1𝐊𝐋n|1,\displaystyle H_{+,\lambda}^{\mathrm{rel}}(\bullet):=H_{\infty}^{\mathrm{rel},0}(\mathfrak{gl}_{1};\bullet\otimes K_{+\rightarrow-}^{\lambda})\colon\mathbf{KL}_{k}^{n-1,1}\rightarrow\mathbf{KL}_{\ell}^{n|1},
H,λrel():=Hrel,0(𝔤𝔩1;K+λ):𝐊𝐋n|1𝐊𝐋kn1,1.\displaystyle H_{-,\lambda}^{\mathrm{rel}}(\bullet):=H_{\infty}^{\mathrm{rel},0}(\mathfrak{gl}_{1};\bullet\otimes K_{-\rightarrow+}^{\lambda})\colon\mathbf{KL}_{\ell}^{n|1}\rightarrow\mathbf{KL}_{k}^{n-1,1}.

It turns out that they are equivalent to the coset functors:

Theorem 5.2.

[24] We have natural isomorphisms

H+,ϵλrelΩλ+:𝐊𝐋k,[λ]n1,1𝐊𝐋,[ϵ2λ]n|1,\displaystyle H_{+,\epsilon\lambda}^{\mathrm{rel}}\simeq\Omega_{-\lambda}^{+}\colon\mathbf{KL}_{k,[\lambda]}^{n-1,1}\rightarrow\mathbf{KL}_{\ell,[\epsilon^{2}\lambda]}^{n|1},
H,ϵλrelΩϵ2λ:𝐊𝐋,[ϵ2λ]n|1𝐊𝐋k,[λ]n1,1.\displaystyle H_{-,\epsilon\lambda}^{\mathrm{rel}}\simeq\Omega_{\epsilon^{2}\lambda}^{-}\colon\mathbf{KL}_{\ell,[\epsilon^{2}\lambda]}^{n|1}\rightarrow\mathbf{KL}_{k,[\lambda]}^{n-1,1}.

In particular, the functors H+,ϵλrelH_{+,\epsilon\lambda}^{\mathrm{rel}} and H,ϵλrelH_{-,\epsilon\lambda}^{\mathrm{rel}} give an equivalence of categories

H+,ϵλrel:𝐊𝐋k,[λ]n1,1𝐊𝐋,[ϵ2λ]n|1:H,ϵλrel.\displaystyle H_{+,\epsilon\lambda}^{\mathrm{rel}}\colon\mathbf{KL}_{k,[\lambda]}^{n-1,1}\rightleftarrows\mathbf{KL}_{\ell,[\epsilon^{2}\lambda]}^{n|1}\colon H_{-,\epsilon\lambda}^{\mathrm{rel}}. (60)

Now, it is clear how to incorporate with (60) the spaces of intertwining operators. Returning back to (59), we just use the intertwining operators of Fock modules.

More concretely, we fix the base of the space of intertwining operators as follows:

IπA+(πλ3A+πλ1A+πλ2A+)=Y+(,z),IπA(πλ3Aπλ1Aπλ2A)=Y(,z)\displaystyle I_{\pi^{A^{+}}}\binom{\pi^{A^{+}}_{\lambda_{3}}}{\pi^{A^{+}}_{\lambda_{1}}\ \pi^{A^{+}}_{\lambda_{2}}}=\mathbb{C}Y^{+}(\cdot,z),\quad I_{\pi^{A^{-}}}\binom{\pi^{A^{-}}_{\lambda_{3}}}{\pi^{A^{-}}_{\lambda_{1}}\ \pi^{A^{-}}_{\lambda_{2}}}=\mathbb{C}Y^{-}(\cdot,z)

for λ3=λ1+λ2\lambda_{3}=\lambda_{1}+\lambda_{2} so that the values of highest weight vectors are given by

Y±(eλ1,z)eλ2=z±λ1λ2exp(λ1m0Am±mzm)eλ3.\displaystyle Y^{\pm}(\mathrm{e}^{\lambda_{1}},z)\mathrm{e}^{\lambda_{2}}=z^{\pm\lambda_{1}\lambda_{2}}\mathrm{exp}\left(\lambda_{1}\sum_{m\geq 0}\frac{A^{\pm}_{-m}}{m}z^{m}\right)\mathrm{e}^{\lambda_{3}}.

Then, given a logarithmic intertwining operator of 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-modules

𝒴(,z):M1M2M3{z}[logz],\displaystyle\mathcal{Y}(\bullet,z)\colon M_{1}\otimes M_{2}\rightarrow M_{3}\{z\}[\log z],

we have a logarithmic intertwining operator

𝒴(,z)Y(,z):(M1πϵλ1A)(M2πϵλ2A)(M3πϵλ3A){z}[logz],\displaystyle\mathcal{Y}(\bullet,z)\otimes Y^{-}(\bullet,z)\colon(M_{1}\otimes\pi^{A^{-}}_{\epsilon\lambda_{1}})\otimes(M_{2}\otimes\pi^{A^{-}}_{\epsilon\lambda_{2}})\rightarrow(M_{3}\otimes\pi^{A^{-}}_{\epsilon\lambda_{3}})\{z\}[\log z],

which naturally extends to the whole complex defining the relative semi-infinite cohomology. It is straightforward to show that it descends to an intertwining operator at the level of cohomology which we denote by

+(𝒴):H+,ϵλ1rel(M1)H+,ϵλ2rel(M2)H+,ϵλ3rel(M3){z}[logz].\displaystyle\mathbb{H}_{+}(\mathcal{Y})\colon H_{+,\epsilon\lambda_{1}}^{\mathrm{rel}}(M_{1})\otimes H_{+,\epsilon\lambda_{2}}^{\mathrm{rel}}(M_{2})\rightarrow H_{+,\epsilon\lambda_{3}}^{\mathrm{rel}}(M_{3})\{z\}[\log z].

In this way, we obtain a linear map +\mathbb{H}_{+} which assigns 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-intertwining operators to 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-intertwining operators. Obviously, we also have a linear map in the opposite direction, which we denote by \mathbb{H}_{-}.

Theorem 5.3.

[24] The linear maps ±\mathbb{H}_{\pm} are isomorphisms between the spaces of (logarithmic) intertwining operators

+:I(M3M1M2)I(H+,ϵλ3rel(M3)H+,ϵλ1rel(M1)H+,ϵλ2rel(M2)),\displaystyle\mathbb{H}_{+}\colon I\binom{M_{3}}{M_{1}\ M_{2}}\overset{\sim}{\rightarrow}I\binom{H_{+,\epsilon\lambda_{3}}^{\mathrm{rel}}(M_{3})}{H_{+,\epsilon\lambda_{1}}^{\mathrm{rel}}(M_{1})\ H_{+,\epsilon\lambda_{2}}^{\mathrm{rel}}(M_{2})},
:I(N3N1N2)I(H,ϵμ3rel(N3)H,ϵμ1rel(N1)H+,ϵμ2rel(N2)).\displaystyle\mathbb{H}_{-}\colon I\binom{N_{3}}{N_{1}\ N_{2}}\overset{\sim}{\rightarrow}I\binom{H_{-,\epsilon\mu_{3}}^{\mathrm{rel}}(N_{3})}{H_{-,\epsilon\mu_{1}}^{\mathrm{rel}}(N_{1})\ H_{+,\epsilon\mu_{2}}^{\mathrm{rel}}(N_{2})}.

5.3. Some basic modules and resolutions

We present some basic modules and resolutions generalizing §2.3. The results here will be contained in Ref. [31] which treats the case of hook-type 𝒲\mathcal{W}-superalgebras[29, 30] in Feigin–Frenkel type duality more generally.

Let the level kk be irrational. The 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-modules playing the role of the (contragradient of) affine Verma modules for Vk(𝔰𝔩2)V^{k}(\mathfrak{sl}_{2}) are Wakimoto modules [55], which are defined through the composition of (37) and (38):

𝒲k(𝔰𝔩n,fsub)βγπ𝔥k+n.\displaystyle\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})\hookrightarrow\beta\gamma\otimes\pi^{k+n}_{\mathfrak{h}}. (61)

This embedding is actually induced from the Wakimoto realization [44, 46] of the affine vertex algebra Vk(𝔰𝔩n)V^{k}(\mathfrak{sl}_{n})

Vk(𝔰𝔩n)𝒜𝔫+π𝔥k+n,𝒜𝔫+:=βγdim𝔫+,\displaystyle V^{k}(\mathfrak{sl}_{n})\hookrightarrow\mathcal{A}_{\mathfrak{n}_{+}}\otimes\pi^{k+n}_{\mathfrak{h}},\quad\mathcal{A}_{\mathfrak{n}_{+}}:=\beta\gamma^{\otimes\dim\mathfrak{n}_{+}}, (62)

which can be extended to a resolution

0Vk(𝔰𝔩n)D0kD1k0\displaystyle 0\rightarrow V^{k}(\mathfrak{sl}_{n})\rightarrow D^{k}_{0}\rightarrow D^{k}_{1}\rightarrow\cdots\rightarrow 0 (63)

by Wakimoto modules of Vk(𝔰𝔩n)V^{k}(\mathfrak{sl}_{n}):

Dpk=l(w)=p𝕎w10k,𝕎λk:=𝒜𝔫+π𝔥,λk+n,(λ𝔥).\displaystyle D^{k}_{p}=\bigoplus_{l(w)=p}\mathbb{W}^{k}_{w^{-1}*0},\quad\mathbb{W}^{k}_{\lambda}:=\mathcal{A}_{\mathfrak{n}^{+}}\otimes\pi^{k+n}_{\mathfrak{h},\lambda},\quad(\lambda\in\mathfrak{h}^{*}).

The sum runs over the elements ww of Weyl group of 𝔰𝔩n\mathfrak{sl}_{n} whose length are pp and * denotes the dot action wλ=w(λ+ρ)ρw*\lambda=w(\lambda+\rho)-\rho with ρ=iϖi\rho=\sum_{i}\varpi_{i} the Weyl vector. For Weyl modules 𝕍λk\mathbb{V}_{\lambda}^{k} (λP+\lambda\in P_{+}), we have similar resolutions

0𝕍λkDλ,0kDλ,1k0\displaystyle 0\rightarrow\mathbb{V}^{k}_{\lambda}\rightarrow D^{k}_{\lambda,0}\rightarrow D^{k}_{\lambda,1}\rightarrow\cdots\rightarrow 0 (64)

by replacing DpkD^{k}_{p} with Dλ,pk=l(w)=p𝕎w1λkD^{k}_{\lambda,p}=\bigoplus_{l(w)=p}\mathbb{W}^{k}_{w^{-1}*\lambda}. Introduce the following 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-modules:

Tλk,+=HDS,fsub0(𝕍λk),𝕎λk,+:=βγπ𝔥,λk+n.T_{\lambda}^{k,+}=H_{\mathrm{DS},f_{\mathrm{sub}}}^{0}(\mathbb{V}_{\lambda}^{k}),\quad\mathbb{W}^{k,+}_{\lambda}:=\beta\gamma\otimes\pi^{k+n}_{\mathfrak{h},\lambda}.

The latter is called a Wakimoto module of 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) acts on it via (61). Then, by applying HDS,fsub0H_{\mathrm{DS},f_{\mathrm{sub}}}^{0} to (64), we obtain a resolution

0Tλk,+Eλ,0kEλ,1k0.\displaystyle 0\rightarrow T_{\lambda}^{k,+}\rightarrow E^{k}_{\lambda,0}\rightarrow E^{k}_{\lambda,1}\rightarrow\cdots\rightarrow 0. (65)

with Eλ,pkl(w)=p𝕎w1λk,+E^{k}_{\lambda,p}\simeq\bigoplus_{l(w)=p}\mathbb{W}^{k,+}_{w^{-1}*\lambda}. When 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2}, this resolution for λ=nϖ1\lambda=n\varpi_{1} is the contragradient dual of (20). Since (61) maps

L(z)(n2:β(z)γ(z)::γ(z)β(z):)1\displaystyle L(z)\mapsto(\tfrac{n}{2}\partial:\beta(z)\gamma(z):-:\gamma(z)\partial\beta(z):)\otimes 1
+1(Lsug(z)+(11k+n)ρ(z)n2ϖn1(z))\displaystyle\hskip 56.9055pt+1\otimes(L_{\mathrm{sug}}(z)+(1-\tfrac{1}{k+n})\partial\rho(z)-\tfrac{n}{2}\partial\varpi_{n-1}(z))
H+(z):β(z)γ(z):1+1ϖn1(z),\displaystyle H_{+}(z)\mapsto-:\beta(z)\gamma(z):\otimes 1+1\otimes\varpi_{n-1}(z),

with Lsug(z)L_{\mathrm{sug}}(z) the Virasoro field by the Segal–Sugarawa construction, the character trzH+,0qL0\mathrm{tr}_{\bullet}z^{H_{+,0}}q^{L_{0}} of the Wakimoto module 𝕎λk,+\mathbb{W}^{k,+}_{\lambda} is given by

ch𝕎λk,+=qΔλk(qn2z)(λ,ϖn1)(q;q)n1(zqn2,z1q1n2;q)\displaystyle\mathrm{ch}\ \mathbb{W}^{k,+}_{\lambda}=\frac{q^{\Delta_{\lambda}^{k}}(q^{\frac{n}{2}}z)^{(\lambda,\varpi_{n-1})}}{\left(q;q\right)_{\infty}^{n-1}\left(zq^{\frac{n}{2}},z^{-1}q^{1-\frac{n}{2}};q\right)_{\infty}}

with Δλk=12(k+n)(λ,λ)(11k+n)(ρ,λ).\Delta_{\lambda}^{k}=\tfrac{1}{2(k+n)}(\lambda,\lambda)-(1-\tfrac{1}{k+n})(\rho,\lambda). Note that Δλk\Delta_{\lambda}^{k} agrees with the conformal weight of the free field module π𝔥,λk+n\pi^{k+n}_{\mathfrak{h},\lambda} of 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}) as implicated by (48).

Introduce the following 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-modules:

Tλ,=H+,ϵ(λ,ϖn1)rel(Tλk,+),Wλ=H+,ϵ(λ,ϖn1)rel(𝕎λk,+).T^{\ell,-}_{\lambda}=H_{+,\epsilon(\lambda,\varpi_{n-1})}^{\mathrm{rel}}(T^{k,+}_{\lambda}),\quad W^{\ell}_{\lambda}=H_{+,\epsilon(\lambda,\varpi_{n-1})}^{\mathrm{rel}}(\mathbb{W}^{k,+}_{\lambda}).

Applying the functor H+,ϵ(λ,ϖn1)relH_{+,\epsilon(\lambda,\varpi_{n-1})}^{\mathrm{rel}} to (65), we obtain a resolution of Tλ,T^{\ell,-}_{\lambda}:

0Tλ,Fλ,0kFλ,1k0.\displaystyle 0\rightarrow T_{\lambda}^{\ell,-}\rightarrow F^{k}_{\lambda,0}\rightarrow F^{k}_{\lambda,1}\rightarrow\cdots\rightarrow 0.

with

Fλ,pkl(w)=pS(w1λλ,ϖn1)Ww1λ.F^{k}_{\lambda,p}\simeq\bigoplus_{l(w)=p}S_{(w^{-1}*\lambda-\lambda,\varpi_{n-1})}W_{w^{-1}*\lambda}^{\ell}.

Here SθS_{\theta} (θ\theta\in\mathbb{Z}) are the spectral flow twists given[64] by

Sθ:X(z)Y(Δ(θH,z)X,z),(X𝒲(𝔰𝔩n|1,fprin)),\displaystyle S_{\theta}\colon X(z)\mapsto Y(\Delta(\theta H_{-},z)X,z),\quad(X\in\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})),
Δ(θH,z):=zθH,0exp(θm=1H,mm(z)m).\displaystyle\Delta(\theta H_{-},z):=z^{\theta H_{-,0}}\mathrm{exp}\left(-\theta\sum_{m=1}^{\infty}\tfrac{H_{-,m}}{m}(-z)^{-m}\right).

The character trqL0zH,0\mathrm{tr}_{\bullet}q^{L_{0}}z^{H_{-,0}} of WλW_{\lambda}^{\ell} is given by

chWλ,=qΔλk+n2(λ,ϖn1)12ϵ2(λ,ϖn1)2zϵ2(λ,ϖn1)(zqn+12,z1qn+32;q)(q;q)n.\displaystyle\mathrm{ch}\ W^{\ell,-}_{\lambda}=q^{\Delta_{\lambda}^{k}+\frac{n}{2}(\lambda,\varpi_{n-1})-\frac{1}{2}\epsilon^{2}(\lambda,\varpi_{n-1})^{2}}z^{\epsilon^{2}(\lambda,\varpi_{n-1})}\frac{\left(-zq^{\frac{n+1}{2}},-z^{-1}q^{\frac{-n+3}{2}};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{n}}.

Let us compare it with the characters of the Wakimoto modules over 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}). These modules are defined as

𝕎μ,=Vπ𝔥,μ+n1,(μ𝔥),\mathbb{W}^{\ell,-}_{\mu}=V_{\mathbb{Z}}\otimes\pi^{\ell+n-1}_{\mathfrak{h},\mu},\quad(\mu\in\mathfrak{h}^{*}),

through the free field realization (40). Since (40) sends

L(z)(:c(z)b(z):+n12:c(z)b(z):)1\displaystyle L(z)\mapsto(:\partial c(z)\ b(z):+\tfrac{n-1}{2}\partial:c(z)b(z):)\otimes 1
+1(Lsug(z)+(11+n1)ρ(z)+1n2ϖn(z)),\displaystyle\hskip 56.9055pt+1\otimes(L_{\mathrm{sug}}(z)+(1-\tfrac{1}{\ell+n-1})\partial\rho(z)+\tfrac{1-n}{2}\partial\varpi_{n}(z)),
H(z):b(z)c(z):1+1ϖn(z),\displaystyle H_{-}(z)\mapsto:b(z)c(z):\otimes 1+1\otimes\varpi_{n}(z),

their characters tr(qL0zH,0)\mathrm{tr}_{\bullet}(q^{L_{0}}z^{H_{-},0}) are given by

ch𝕎μ,=qΔμ+n12(μ,ϖn)z(μ,ϖn)(zqn+12,z1qn+12;q)(q;q)n\displaystyle\mathrm{ch}\ \mathbb{W}^{\ell,-}_{\mu}=q^{\Delta_{\mu}^{\ell}+\frac{n-1}{2}(\mu,\varpi_{n})}z^{(\mu,\varpi_{n})}\frac{\left(-zq^{\frac{n+1}{2}},-z^{-1}q^{\frac{-n+1}{2}};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{n}} (66)

with Δμ=12(+n1)(μ,μ)(11+n1)(μ,ρ).\Delta_{\mu}^{\ell}=\tfrac{1}{2(\ell+n-1)}(\mu,\mu)-(1-\tfrac{1}{\ell+n-1})(\mu,\rho). Then we find that

chWλ=11+z1qn+12ch𝕎μ,,μ=(+n1)(λ~+(λ,ϖn1)ϖn).\displaystyle\mathrm{ch}\ W^{\ell}_{\lambda}=\frac{1}{1+z^{-1}q^{\frac{-n+1}{2}}}\mathrm{ch}\ \mathbb{W}^{\ell,-}_{\mu},\quad\mu=-(\ell+n-1)(\tilde{\lambda}+(\lambda,\varpi_{n-1})\varpi_{n}).

Here λ~\tilde{\lambda} is the image of λ\lambda under the orthogonal decomposition 𝔥(𝔰𝔩n|1)=𝔥(𝔰𝔩n)ϖn.\mathfrak{h}(\mathfrak{sl}_{n|1})=\mathfrak{h}(\mathfrak{sl}_{n})\oplus\mathbb{C}\varpi_{n}. Indeed, WλW^{\ell}_{\lambda} is a submodule of 𝕎μ,\mathbb{W}^{\ell,-}_{\mu} (“thin Wakimoto module”). This is proven by introducing “thick” Wakimoto modules

𝕎^λk,+:=π𝔥,λk+nΠ,(λ𝔥),\widehat{\mathbb{W}}^{k,+}_{\lambda}:=\pi^{k+n}_{\mathfrak{h},\lambda}\otimes\Pi,\quad(\lambda\in\mathfrak{h}^{*}),

whose image coincide with the Wakimoto modules 𝕎μ,\mathbb{W}^{\ell,-}_{\mu}. The 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-module structure on 𝕎^λk,+\widehat{\mathbb{W}}^{k,+}_{\lambda} is given by the composition of (61) and the bosonization βγΠ\beta\gamma\rightarrow\Pi, that is, the free field realization (40). Replacing Π\Pi with its modules Π0[a]\Pi_{0}[a] (aa\in\mathbb{C}) below, we obtain the general Wakimoto module for 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}).

To obtain Verma-type modules on 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-side as in the case of Vc(𝔫𝔰2)V^{c}(\mathfrak{ns}_{2}), we need to replace the Fock modules π𝔥,λk+n\pi^{k+n}_{\mathfrak{h},\lambda} by Verma modules for 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}) and thus need the twist (40) and use (48), which satisfies

𝒲k(𝔰𝔩n,fsub)𝒲k(𝔰𝔩n)ΠL(z)W2(z)1+1(12:c(z)d(z):n2𝐛(z))H(z)1𝐛+(z)\displaystyle\begin{array}[]{cll}\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})&\rightarrow&\hskip 28.45274pt\mathcal{W}^{k}(\mathfrak{sl}_{n})\otimes\Pi\\ L(z)&\mapsto&W_{2}(z)\otimes 1+1\otimes(\tfrac{1}{2}:c(z)d(z):-\tfrac{n}{2}\partial\mathbf{b}^{-}(z))\\ H_{-}(z)&\mapsto&\hskip 28.45274pt1\otimes\mathbf{b}^{+}(z)\end{array}

where W2(z)W_{2}(z) is the Virasoro field of 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}) and 𝐛±(z)\mathbf{b}^{\pm}(z) are given by

𝐛±(z)=±12(ε+1)c(z)+12d(z).\mathbf{b}^{\pm}(z)=\pm\tfrac{1}{2}(\varepsilon_{+}-1)c(z)+\tfrac{1}{2}d(z).

Let 𝕄χλk\mathbb{M}^{k}_{\chi_{\lambda}} (λ𝔥\lambda\in\mathfrak{h}^{*}) be the Verma module of 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n}) equipped with the central character χλ:𝒵(𝔰𝔩n)\chi_{\lambda}\colon\mathcal{Z}(\mathfrak{sl}_{n})\rightarrow\mathbb{C}. It is generated by the one dimensional representation χλ\chi_{\lambda} of the Zhu’s algebra Zhu(𝒲k(𝔰𝔩n))𝒵(𝔰𝔩n)\mathrm{Zhu}(\mathcal{W}^{k}(\mathfrak{sl}_{n}))\simeq\mathcal{Z}(\mathfrak{sl}_{n}). Under this notation, we have a natural homomorphism 𝕄χλkπ𝔥,λk+n\mathbb{M}^{k}_{\chi_{\lambda}}\rightarrow\pi^{k+n}_{\mathfrak{h},\lambda} of 𝒲k(𝔰𝔩n)\mathcal{W}^{k}(\mathfrak{sl}_{n})-modules, which sends the highest weight vector to the other. Let Πθ[a]\Pi_{\theta}[a] (θ,a)(\theta,a\in\mathbb{C}) be the Π\Pi-module defined by

Πθ[a]=mπθ𝐛++(m+a)c+1.\Pi_{\theta}[a]=\bigoplus_{m\in\mathbb{Z}}\pi^{\mathbb{Z}+\operatorname{\sqrt{\smash[b]{-1}}}\mathbb{Z}}_{\theta\mathbf{b}^{+}+(m+a)c}.

Then we define the relaxed highest weight module Rθk(χλ,a)R_{\theta}^{k}(\chi_{\lambda},a) of 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) as

Rθk,+(χλ,a):=𝕄χλkΠθ[a].R_{\theta}^{k,+}(\chi_{\lambda},a):=\mathbb{M}^{k}_{\chi_{\lambda}}\otimes\Pi_{\theta}[a].

It is positively graded if and only if θ=n/2\theta=-n/2 and we set Rk,+(χλ,a)=Rn/2k,+(χλ,a)R^{k,+}(\chi_{\lambda},a)=R^{k,+}_{-n/2}(\chi_{\lambda},a), whose character is given by

chRk,+(χλ,a)=qΔλk+18n2ε+zan2ε+mzm(q;q)n+1.\displaystyle\mathrm{ch}\ R^{k,+}(\chi_{\lambda},a)=q^{\Delta^{k}_{\lambda}+\frac{1}{8}n^{2}\varepsilon_{+}}z^{a-\frac{n}{2}\varepsilon_{+}}\frac{\sum_{m\in\mathbb{Z}}z^{m}}{\left(q;q\right)_{\infty}^{n+1}}.

Applying the functor H+,ϵ(an2ε+)relH^{\mathrm{rel}}_{+,\epsilon(a-\frac{n}{2}\varepsilon_{+})}, we find that the image has the following character :

chH+,ϵ(an2ε+)rel(Rk,+(χλ,a))\displaystyle\mathrm{ch}\ H^{\mathrm{rel}}_{+,\epsilon(a-\frac{n}{2}\varepsilon_{+})}(R^{k,+}(\chi_{\lambda},a))
=qΔλk+18n2ε+212ϵ2(an2ε+)2zϵ2(an2ε+)(zq12,z1q12;q)(q;q)n.\displaystyle\hskip 28.45274pt=q^{\Delta_{\lambda}^{k}+\frac{1}{8}n^{2}\varepsilon_{+}^{2}-\frac{1}{2}\epsilon^{2}(a-\frac{n}{2}\varepsilon_{+})^{2}}z^{\epsilon^{2}(a-\frac{n}{2}\varepsilon_{+})}\frac{\left(-zq^{\frac{1}{2}},-z^{-1}q^{\frac{1}{2}};q\right)_{\infty}}{\left(q;q\right)_{\infty}^{n}}.

When nn is even, the image is given by the spectral flow twist Sn/2(Rχμ,)S_{-n/2}(R^{\ell,-}_{\chi_{\mu}}) with μ=(+n1)(λ~+aϖn)\mu=-(\ell+n-1)(\tilde{\lambda}+a\varpi_{n}) of the Verma-type 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-module R,(χμ)R^{\ell,-}(\chi_{\mu}), which is defined in general as

R,(χμ)=𝕄χμV,(μ𝔥),R^{\ell,-}(\chi_{\mu})=\mathbb{M}^{\ell}_{\chi_{\mu}}\otimes V_{\mathbb{Z}},\quad(\mu\in\mathfrak{h}^{*}),

where 𝕄χμ\mathbb{M}^{\ell}_{\chi_{\mu}} is the Verma module of 𝒲(𝔤𝔩n)\mathcal{W}^{\ell}(\mathfrak{gl}_{n}) with a natural homomorphism 𝕄χμπ𝔥,μ+n1\mathbb{M}^{\ell}_{\chi_{\mu}}\rightarrow\pi^{\ell+n-1}_{\mathfrak{h},\mu} and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}) acts on R,(χμ,a)R^{\ell,-}(\chi_{\mu},a) through 𝒲(𝔰𝔩n|1,fprin)𝒲(𝔤𝔩n)V\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})\rightarrow\mathcal{W}^{\ell}(\mathfrak{gl}_{n})\otimes V_{\mathbb{Z}}. The spectral flow twist makes the above asymmetric Verma-type module into a “(symmetric) Verma module”. To summarize, we have obtained the following dictionary generalizing Tab. 2 below.

Table 2.
𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}})-side 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-side
Wakimoto modules “thin” Wakimoto modules
“thick” Wakimoto modules Wakimoto modules
relaxed highest weight modules “Verma modules”

It is an interesting problem to extend the correspondence of classes of modules in the spirit of the quantum Langlands duality[11] of Arakawa and Frenkel for modules over the princial 𝒲\mathcal{W}-algebras. On the other hand, since the positively graded modules are controlled by the finite 𝒲\mathcal{W}-superalgebras[33], it is natural to compare the finite 𝒲\mathcal{W}-superalgebras in our case. In the literature[35, 66], the finite 𝒲\mathcal{W}-algebras for 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) have a nice realization as infinitesimal Cherednik algebras.

6. Conclusions

The pair of 𝒲\mathcal{W}-superalgebras 𝒲k(𝔰𝔩n,fsub)\mathcal{W}^{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) and 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}^{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}}) enjoy the Feigin–Frenkel type duality for their Heisenberg cosets. It can be proven either by using the uniqueness property of the universal 𝒲\mathcal{W}_{\infty}-algebra 𝒲[c,λ]\mathcal{W}_{\infty}[c,\lambda] or by the technique of free field realization. Furthermore, the duality is refined to a reconstruction type theorem of one of the 𝒲\mathcal{W}-superalgebras from the other by the coset functor or the relative semi-infinite cohomology functor. Either of them leads to the block-wise equivalence of module categories consisting of Kazhdan–Lusztig objects with respect to the Heisenberg subalgebra. This equivalence is incorporated with isomorphisms of all the logarithmic intertwining operators. Moreover, we have obtained a dictionary of the correspondence of some fundamental class of modules. There are several interesting problems still left open.

  • The correspondence for correlation functions on the Riemannian sphere has been established[25] by path integral method, which is worth generalization to correlation functions on arbitrary Riemannian surfaces.

  • Apart from the rational cases, 𝒲k(𝔰𝔩n,fsub)\mathcal{W}_{k}(\mathfrak{sl}_{n},f_{\mathrm{sub}}) at k=n+nn+1k=-n+\frac{n}{n+1}, known also as the chiral algebras of Argyres–Douglas theories of type (A1,A2n1)(A_{1},A_{2n-1})[4] have suitable module categories which have a non-semisimple braided tensor category structure [15]. It is an interesting problem to understand the 𝒲(𝔰𝔩n|1,fprin)\mathcal{W}_{\ell}(\mathfrak{sl}_{n|1},f_{\mathrm{prin}})-side and compare the braided tensor category structure.

Obviously, it is interesting to pursue the correspondence of modules and intertwining operators, or correlation functions in general, for other pairs of 𝒲\mathcal{W}-superalgebras, namely hook-type 𝒲\mathcal{W}-superalgebras appearing in Refs. 54, 29, 30. The Feigin–Frenkel type duality is already established[29, 30] mathematically, the reconstruction type theorem is also established for a large class[32] and the correspondence of module categories and the behavior of qq-characters are work in progress.[31]

References

  • [1] Dražen Adamović. Representations of the N=2N=2 superconformal vertex algebra. Internat. Math. Res. Notices, (2):61–79, 1999.
  • [2] Dražen Adamović. Vertex algebra approach to fusion rules for N=2N=2 superconformal minimal models. J. Algebra, 239(2):549–572, 2001.
  • [3] Dražen Adamović. Realizations of simple affine vertex algebras and their modules: the cases sl(2)^\widehat{sl(2)} and osp(1,2)^\widehat{osp(1,2)}. Comm. Math. Phys., 366(3):1025–1067, 2019.
  • [4] Dražen Adamović, Thomas Creutzig, Naoki Genra, and Jinwei Yang. The vertex algebras mathcakR(p)mathcak{R}^{(p)} and 𝒱(p)\mathcal{V}^{(p)}. Comm. Math. Phys., 383(2):1207–1241, 2021.
  • [5] Dražen Adamović, Kazuya Kawasetsu, and David Ridout. A realisation of the Bershadsky-Polyakov algebras and their relaxed modules. Lett. Math. Phys., 111(2):Paper No. 38, 30, 2021.
  • [6] Luis F. Alday, Davide Gaiotto, and Yuji Tachikawa. Liouville correlation functions from four-dimensional gauge theories. Lett. Math. Phys., 91(2):167–197, 2010.
  • [7] Tomoyuki Arakawa. Associated varieties of modules over Kac-Moody algebras and C2C_{2}-cofiniteness of WW-algebras. Int. Math. Res. Not. IMRN, (22):11605–11666, 2015.
  • [8] Tomoyuki Arakawa. Rationality of WW-algebras: principal nilpotent cases. Ann. of Math. (2), 182(2):565–604, 2015.
  • [9] Tomoyuki Arakawa, Thomas Creutzig, and Andrew R. Linshaw. Cosets of Bershadsky-Polyakov algebras and rational 𝒲\mathcal{W}-algebras of type AA. Selecta Math. (N.S.), 23(4):2369–2395, 2017.
  • [10] Tomoyuki Arakawa, Thomas Creutzig, and Andrew R. Linshaw. WW-algebras as coset vertex algebras. Invent. Math., 218(1):145–195, 2019.
  • [11] Tomoyuki Arakawa and Edward Frenkel. Quantum Langlands duality of representations of 𝒲\mathcal{W}-algebras. Compos. Math., 155(12):2235–2262, 2019.
  • [12] Tomoyuki Arakawa, Ching Hung Lam, and Hiromichi Yamada. Parafermion vertex operator algebras and WW-algebras. Trans. Amer. Math. Soc., 371(6):4277–4301, 2019.
  • [13] Tomoyuki Arakawa and Jethro van Ekeren. Modularity of relatively rational vertex algebras and fusion rules of principal affine WW-algebras. Comm. Math. Phys., 370(1):205–247, 2019.
  • [14] Tomoyuki Arakawa and Jethro van Ekeren. Rationality and fusion rules of exceptional WW-algebras, 2019.
  • [15] Jean Auger, Thomas Creutzig, Shashank Kanade, and Matthew Rupert. Braided tensor categories related to p\mathcal{B}_{p} vertex algebras. Comm. Math. Phys., 378(1):219–260, 2020.
  • [16] Christopher Beem, Leonardo Rastelli, and Balt C. van Rees. 𝒲\mathcal{W} symmetry in six dimensions. J. High Energy Phys., (5):017, front matter+37, 2015.
  • [17] Mikhail Bershtein, Boris Feigin, and Grigory Merzon. Plane partitions with a “pit”: generating functions and representation theory. Selecta Math. (N.S.), 24(1):21–62, 2018.
  • [18] Richard E. Borcherds. Vertex algebras, Kac-Moody algebras, and the Monster. Proc. Nat. Acad. Sci. U.S.A., 83(10):3068–3071, 1986.
  • [19] Miranda C. N. Cheng, Sungbong Chun, Boris Feigin, Francesca Ferrari, Sergei Gukov, Sarah M. Harrison, and Davide Passaro. 3-manifolds and VOA characters, 2022.
  • [20] Thomas Creutzig. Fusion categories for affine vertex algebras at admissible levels. Selecta Math. (N.S.), 25(2):Paper No. 27, 21, 2019.
  • [21] Thomas Creutzig, Tudor Dimofte, Niklas Garner, and Nathan Geer. A QFT for non-semisimple TQFT, 2021.
  • [22] Thomas Creutzig and Davide Gaiotto. Vertex algebras for S-duality. Comm. Math. Phys., 379(3):785–845, 2020.
  • [23] Thomas Creutzig, Naoki Genra, and Shigenori Nakatsuka. Duality of subregular 𝒲\mathcal{W}-algebras and principal 𝒲\mathcal{W}-superalgebras. Adv. Math., 383:Paper No. 107685, 52, 2021.
  • [24] Thomas Creutzig, Naoki Genra, Shigenori Nakatsuka, and Ryo Sato. Correspondences of categories for subregular WW-algebras and principal WW-superalgebras, 2021.
  • [25] Thomas Creutzig, Yasuaki Hikida, and Devon Stockal. Correlator correspondences for subregular 𝒲\mathcal{W}-algebras and principal 𝒲\mathcal{W}-superalgebras. J. High Energy Phys., (10):Paper No. 032, 28, 2021.
  • [26] Thomas Creutzig, Shashank Kanade, and Andrew R. Linshaw. Simple current extensions beyond semi-simplicity. Commun. Contemp. Math., 22(1):1950001, 49, 2020.
  • [27] Thomas Creutzig, Shashank Kanade, and Robert McRae. Tensor categories for vertex operator superalgebra extensions, 2017.
  • [28] Thomas Creutzig and Andrew R. Linshaw. Cosets of affine vertex algebras inside larger structures. J. Algebra, 517:396–438, 2019.
  • [29] Thomas Creutzig and Andrew R. Linshaw. Trialities of 𝒲\mathcal{W}-algebras, 2020.
  • [30] Thomas Creutzig and Andrew R. Linshaw. Trialities of orthosymplectic 𝒲\mathcal{W}-algebras, 2021.
  • [31] Thomas Creutzig, Andrew R. Linshaw, and Shigenori Nakatsuka, work in progress.
  • [32] Thomas Creutzig, Andrew R. Linshaw, Shigenori Nakatsuka, and Ryo Sato. Duality via convolution of WW-algebras, 2022.
  • [33] Alberto De Sole and Victor G. Kac. Finite vs affine WW-algebras. Jpn. J. Math., 1(1):137–261, 2006.
  • [34] P. Di Vecchia, J. L. Petersen, M. Yu, and H. B. Zheng. Explicit construction of unitary representations of the N=2N=2 superconformal algebra. Phys. Lett. B, 174(3):280–284, 1986.
  • [35] Pavel Etingof, Wee Liang Gan, and Victor Ginzburg. Continuous Hecke algebras. Transform. Groups, 10(3-4):423–447, 2005.
  • [36] V. A. Fateev and S. L. Lykyanov. The models of two-dimensional conformal quantum field theory with ZnZ_{n} symmetry. Internat. J. Modern Phys. A, 3(2):507–520, 1988.
  • [37] Zachary Fehily. Subregular W-algebras of type A, 2021.
  • [38] B. L. Feigin and A. M. Semikhatov. 𝒲n(2)\mathscr{W}^{(2)}_{n} algebras. Nuclear Phys. B, 698(3):409–449, 2004.
  • [39] B. L. Feigin, A. M. Semikhatov, V. A. Sirota, and I. Yu. Tipunin. Resolutions and characters of irreducible representations of the N=2N=2 superconformal algebra. Nuclear Phys. B, 536(3):617–656, 1999.
  • [40] B. L. Feigin, A. M. Semikhatov, and I. Yu. Tipunin. Equivalence between chain categories of representations of affine sl(2){\rm sl}(2) and N=2N=2 superconformal algebras. J. Math. Phys., 39(7):3865–3905, 1998.
  • [41] Boris Feigin and Edward Frenkel. Duality in WW-algebras. Internat. Math. Res. Notices, (6):75–82, 1991.
  • [42] Boris Feigin and Sergei Gukov. VOA[M4]{\rm VOA}[M_{4}]. J. Math. Phys., 61(1):012302, 27, 2020.
  • [43] B. L. Feĭgin. Semi-infinite homology of Lie, Kac-Moody and Virasoro algebras. Uspekhi Mat. Nauk, 39(2(236)):195–196, 1984.
  • [44] B. L. Feĭgin and È. V. Frenkel. A family of representations of affine Lie algebras. Uspekhi Mat. Nauk, 43(5(263)):227–228, 1988.
  • [45] B. L. Feĭgin and D. B. Fuchs. Representations of the Virasoro algebra. In Representation of Lie groups and related topics, volume 7 of Adv. Stud. Contemp. Math., pages 465–554. Gordon and Breach, New York, 1990.
  • [46] Edward Frenkel. Langlands correspondence for loop groups, volume 103 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2007.
  • [47] Edward Frenkel and Davide Gaiotto. Quantum Langlands dualities of boundary conditions, DD-modules, and conformal blocks. Commun. Number Theory Phys., 14(2):199–313, 2020.
  • [48] Edward Frenkel, Victor Kac, and Minoru Wakimoto. Characters and fusion rules for WW-algebras via quantized Drinfeld-Sokolov reduction. Comm. Math. Phys., 147(2):295–328, 1992.
  • [49] I. B. Frenkel. Representations of affine Lie algebras, Hecke modular forms and Korteweg-de Vries type equations. In Lie algebras and related topics (New Brunswick, N.J., 1981), volume 933 of Lecture Notes in Math., pages 71–110. Springer, Berlin-New York, 1982.
  • [50] I. B. Frenkel, H. Garland, and G. J. Zuckerman. Semi-infinite cohomology and string theory. Proc. Nat. Acad. Sci. U.S.A., 83(22):8442–8446, 1986.
  • [51] Daniel Friedan, Emil Martinec, and Stephen Shenker. Conformal invariance, supersymmetry and string theory. Nuclear Phys. B, 271(1):93–165, 1986.
  • [52] V. M. Futorny. Irreducible non-dense A1(1)A^{(1)}_{1}-modules. Pacific J. Math., 172(1):83–99, 1996.
  • [53] Davide Gaiotto and Miroslav Rapcak. Miura operators, degenerate fields and the M2-M5 intersection, 2020.
  • [54] Davide Gaiotto and Miroslav Rapčák. Vertex algebras at the corner. J. High Energy Phys., (1):160, front matter+85, 2019.
  • [55] Naoki Genra. Screening operators and parabolic inductions for affine 𝒲\mathcal{W}-algebras. Adv. Math., 369:107179, 62, 2020.
  • [56] P. Goddard, A. Kent, and D. Olive. Virasoro algebras and coset space models. Phys. Lett. B, 152(1-2):88–92, 1985.
  • [57] Kenji Iohara and Yoshiyuki Koga. Representation theory of the Virasoro algebra. Springer Monographs in Mathematics. Springer-Verlag London, Ltd., London, 2011.
  • [58] Katsushi Ito. N=2N=2 superconformal 𝐂Pn{\bf C}{\rm P}_{n} model. Nuclear Phys. B, 370(1):123–142, 1992.
  • [59] Victor Kac, Shi-Shyr Roan, and Minoru Wakimoto. Quantum reduction for affine superalgebras. Comm. Math. Phys., 241(2-3):307–342, 2003.
  • [60] Victor G. Kac and Minoru Wakimoto. On rationality of WW-algebras. Transform. Groups, 13(3-4):671–713, 2008.
  • [61] Kazuya Kawasetsu and David Ridout. Relaxed highest-weight modules I: Rank 1 cases. Comm. Math. Phys., 368(2):627–663, 2019.
  • [62] Yoichi Kazama and Hisao Suzuki. New n=2n=2 superconformal field theories and superstring compactification. Nuclear Phys. B, 321(1):232–268, 1989.
  • [63] Shinji Koshida and Ryo Sato. On resolution of highest weight modules over the 𝒩=2\mathcal{N}=2 superconformal algebra, 2018.
  • [64] Haisheng Li. The physics superselection principle in vertex operator algebra theory. J. Algebra, 196(2):436–457, 1997.
  • [65] Andrew R. Linshaw. Universal two-parameter 𝒲\mathcal{W}_{\infty}-algebra and vertex algebras of type 𝒲(2,3,,N)\mathcal{W}(2,3,\ldots,N). Compos. Math., 157(1):12–82, 2021.
  • [66] I. Losev and A. Tsymbaliuk. Infinitesimal Cherednik algebras as WW-algebras. Transform. Groups, 19(2):495–526, 2014.
  • [67] Victor Ostrik and Michael Sun. Level-rank duality via tensor categories. Comm. Math. Phys., 326(1):49–61, 2014.
  • [68] Tomáš Procházka and Miroslav Rapčák. 𝒲\mathcal{W}-algebra modules, free fields, and Gukov-Witten defects. J. High Energy Phys., (5):159, 70, 2019.
  • [69] Miroslav Rapčák, Yan Soibelman, Yaping Yang, and Gufang Zhao. Cohomological Hall algebras, vertex algebras and instantons. Comm. Math. Phys., 376(3):1803–1873, 2020.
  • [70] Ryo Sato. Equivalences between weight modules via 𝒩=2\mathcal{N}=2 coset constructions, 2016.
  • [71] Hiromichi Yamada and Hiroshi Yamauchi. Simple current extensions of tensor products of vertex operator algebras. Int. Math. Res. Not. IMRN, (16):12778–12807, 2021.