Feigin–Semikhatov conjecture and related topics
Abstract.
Feigin–Semikhatov conjecture, now established, states algebraic isomorphisms between the cosets of the subregular -algebras and the principal -superalgebras of type A by their full Heisenberg subalgebras. It can be seen as a variant of Feigin–Frenkel duality between the -algebras and also as a generalization of the connection between the superconformal algebra and the affine algebra .
We review the recent developments on the correspondence of the subregular -algebras and the principal -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 -superalgebras and their affine cosets, that is, subalgebras whose fields commute with affine subalgebras inside give a rich class of such vertex algebras. The -superalgebras, denoted by , are obtained from the affine vertex superalgebra through the quantum Drinfeld–Sokolov reduction parametrized by nilpotent elements inside . In this paper, we will write and for their simple quotients. When is a simply-laced Lie algebra and is called principal (or regular), 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 , the triality for , also known as Fateev–Lukyanov’s -algebra [36], is depicted in Fig. 1 below.
Another kind of this tirality was suggested first by Feigin and Semikhatov [38] for the -algebra associated with the subregular nilpotent element as in Fig. 2.
Here we take the coset of -superalgebras by the maximal affine subalgebra inside, which is rank one Heisenberg algebra . 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 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 Super Yang–Mills gauge theory. The above examples correspond to the following diagrams which describe the ranks of the gauge group and the half-BPS interfaces placed as three dimensional boundary conditions.
A large class of this triality is now proven by Creutzig and Linshaw [29, 30] mathematically and its application to the representation theory of -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 and
2.1. Coset construction of
The superconformal algebra with central charge has four generating fields: , of even parity and of odd parity satisfying the following operator product expansions (OPEs)
For clarification, let denote the (universal) vertex algebra freely generated by these fields and by its unique simple quotient. Kazama and Suzuki[62] gave a family of realizations of based on Hermitian symmetric spaces . We are interested in the case . More precisely, we consider the vertex superalgebra of the universal affine vertex algebra of at level and copies of -systems. Since we have homomorphisms and , we may take the diagonal coset , i.e., the subalgebra consisting of elements whose fields commute with the diagonal -action on . Then the coset contains with specific central charge whereas the whole coset is conjectured[58, 28] to be isomorphic to the principal -superalgebra at certain level:
(1) |
When , is exactly the principal -superalgebra and gives the honest coset construction of . More explicitly, the isomorphism is given by the following:
(6) |
with
(7) |
Here we have used the boson-fermion correspondence between the -system and the lattice vertex superalgebra associated with the lattice ; is the vertex operator for the element in the lattice ; is the Heisenberg field satisfying the OPE ; is the Heisenberg vertex algebra generated by . The construction (6) descends to the simple quotients
(8) |
When 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 and share some similarity. This means that one can not only go from the -side to the -side, but also in the opposite direction. Feigin–Semikhatov–Tipunin [40] found the crucial coset type realization of in terms of by using the negative-definite lattice vertex superalgebra . Namely, they established the following isomorphism of vertex algebras:
(13) |
with . Again, (13) descends to the simple quotients
(14) |
2.2. Equivalence of module categories
Let us recall the block-wise equivalence [40, 70] of representation categories for and by the coset constructions (6) and (13). From now on, we always assume as in (7). Since the coset constructions themselves use Heisenberg fields, it is reasonable to consider those modules of and which behave nicely for their Heisenberg fields inside them. This motivates to introduce the category of weight modules: let denote the category of -modules which decompose into direct sums of Fock modules with respect to . Note that the eigenvalues of on is . Then the category admits a block decomposition in terms of the spectrum of modulo :
Similarly, we introduce the category of weight -modules in terms of . Then it decomposes into
by using the spectrum of on , which is . 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 to , take an arbitrary weight module and decompose as a module over :
(15) |
with
and the Fock module of on which acts by . If lies in , then lies in with . Therefore, defines a functor
(16) |
It is non-zero if and only if . To get a functor in the opposite direction, take an arbitrary -module and decompose as a module over :
(17) |
Then we obtain a functor
(18) |
which is non-zero if and only if .
The functors are either zero or an equivalence. In particular, the block-wise equivalence of categories is stated in the following way:
Theorem 2.1.
The proof is established by giving natural isomorphisms connecting modules and their images in under the compositions and .
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 is irrational. On -side, the Weyl module of highest weight , the induced module whose top space is the -dimensional -module, has a two step resolution by the affine Verma modules:
(20) |
Applying the functor , we obtain the following resolution of the simple -module of highest weight :
Each component has the following character :
in terms of the -Pochhammer symbols . As indicated by the characters, is identified with a quotient of a Verma module, namely, , called a topological Verma module[1, 39]. The other one is obtained from by spectral flow twist defined in general by
which transforms the module characters in the following way:
Therefore, the counterpart of (20) is
To obtain resolutions of simple -modules by Verma modules (they are called massive Verma modules), we have to replace the affine Verma modules by thicker -modules. They are given by relaxed highest weight modules () induced by the weight -modules
Here is the Casimir element of and is the one dimensional module over with , . Then has the character
which gives the desired character on -side:
Indeed, we have an isomorphism . The relaxed highest weight modules and their spectral flow twists defined through
are the most natural class of -modules since all the simple weight -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 introduced by Adamović[3]:
(21) |
where is the half-lattice vertex algebra
and
Let us introduce the following modules[37] over
Here is the Verma module of highest weight over the Virasoro algebra of central charge with
and is the -module given by the sum of Fock modules
In particular, has the following character
and is isomorphic to the relaxed highest weight module
(22) |
under the assumption for all positive odd integers . The assumption says that there is no highest weight vector in otherwise these two modules are non-isomorphic indecomposable modules. The other modules are obtained from by spectral flow twists:
Let be generic so that (22) holds. The simple quotient () of has a two step resolution[45, 57] by Verma modules due to Feigin–Fuchs:
Except for finitely many values of , coincides with the simple quotient of , which we denote by . In this case, the above resolution induces a resolution of by relaxed highest Verma modules
Applying the functor , we obtain the following resolution of a simple -module by massive Verma modules:
with
The above approach should also work at admissible levels. Some resolutions of -modules and -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 -superalgebras, see Tab. 2 at the end of this article.
-side | -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 -mod and -mod when 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 -modules is given by
and that the list[1] of simple -modules is given by the highest weight modules
where runs through the set
They consist of all the unitary minimal representations[34] of . The module categories decompose into
and the lists of simple modules lying in each block are given by
In Fig. 4, we illustrate the block-wise equivalence for .
The fusion rules are given by
where runs through
-
•
and through[2] either of
-
•
,
, -
•
,
,
where we set . In particular, is a simple current and satisfies
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
Let be a rational -cofinite vertex operator algebra so that the module category -mod is a semisimple braided tensor category of finitely many simple objects. We assume that -mod has simple current modules which form an abelian group by fusion product
Note that the fusion ring of -modules is naturally a ring over by fusion product and a ring graded by the dual group by monodromy: with . By a simple current extension of by a lattice theory, we mean a vertex operator superalgebra extension of the form
for some non-degenerate integral lattice with a group homomorphism . Here is the dual lattice with which we may parametrize the simple -modules (). Take the sublattice so that . Then the fusion rings and are related by the formulas
In our setting, they are
(23) | |||
(24) |
We note that in Ref. [2] the fusion rules of 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 , it enjoys the level-rank duality[12] and is identified with a simple principal -algebra
(25) |
The -algebra is rational and -cofinite[8, 7] and known as one of the discrete series representation. The module category -mod is a semisimple braided tensor category with simple modules parametrized by those of , say , satisfying the same fusion rules.[13, 48, 20] On the other hand, the double commutant extending the Heisenberg vertex algebra turns out to be a lattice vertex algebra . Then decomposes as a module over into
and, similarly,
Therefore, and are simple current extensions by lattice theories, which imply the following description of fusion rings:
3. Feigin–Semikhatov conjecture
3.1. Main statements
To obtain a fruitful generalization of the relation between and , it is better to view as the principal -superalgebra associated with :
and then rewrite the relation (25) into
(26) |
Note that the numbers two and one in each factor are the dual Coxeter numbers of and . This implies that the relation between and is a variant of the celebrated Feigin–Frenkel duality[41] for the principal -algebras, which asserts an isomorphism
(27) |
in the case of type A.
Recall that the -superalgebras in general are defined from the affine vertex superalgebras for (simple) basic classical Lie superalgebras with even supersymmetric invariant bilinear forms via the quantum Drinfeld–Sokolov reductions parametrized by the conjugacy classes of even nilpotent elements ,
When (), the bilinear forms are all proportional to the super trace and we identify with its ratio . The conjugacy classes are in one-to-one correspondence with the pair of partitions of and . For they are and correspond to the Virasoro vertex algebra and , respectively. For , they are and correspond to the superconformal algebra and , respectively.
The set of all the even nilpotent elements forms an algebraic subvariety inside and decomposes into conjugacy classes . There is a unique class, called regular or principal, characterized as maximizing the dimension . The partition is given by and the corresponding -superalgebra with is called principal. If , then there is a unique class, called subregular, of the second largest dimension. The partition is given by and the corresponding -algebra with is called subregular.
Feigin and Semikhatov[38] conjectured that the pair and generalizes to the series of pairs
(28) |
The -superalgebras (28) have rank one Heisenberg vertex subalgebras as the maximal affine subalgebras: we normalize the generators
(29) |
so that the eigenvalues of the zero mode on the -superalgebras are normalized to . Then the OPEs are
with
(30) |
The precise analogy of the Feigin–Frenkel duality for the pairs (28) is the duality between the Heisenberg cosets of these -superalgebras:
These pairs indeed enjoy the coset constructions:
Theorem 3.2.
Note that we can prove Theorem 3.2 for for by checking the OPEs directly. However, we cannot use this approach literally since we do not know the whole defining OPEs of and . 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 -superalgebras in general as proved by Creutzig and Linshaw [29, 30] and the other one is free field realizations of these -superalgebras.
3.2. Uniqueness property
The -algebra has a free strong generating set of the form
that is, a generating set whose conformal weights are 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 and the field of conformal weight two is the Virasoro field. The fields of conformal weight are those generators which have nontrivial -weights, namely .
Similarly, the -superalgebra has a free strong generating set of the form
Again, the field of conformal weight one is the Heisenberg field , the field of conformal weight two is the Virasoro field and the fields of conformal weight are those generators which have nontrivial -weights, namely and are the only odd generators.
The Heisenberg cosets
(32) |
have a special feature: they have a common strong generating set of the form
(33) |
which is the same as the principal -algebra , 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 , which is a consequence of a -character (35) below.
The important fact is that one-parameter families of vertex algebras of type are always obtained as quotients of the two-parameter universal -algebra 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 and has a strong free generating set of the form
By setting to be the generating field of conformal degree , parametrizes the central charge of the Virasoro field and parametrizes the undetermined scalar appearing in the OPE
The remaining OPEs are uniquely determined by the Jacobi identities under the assumption that there is an automorphism sending . Both of the coset algebras (32) and () are obtained as one-parameter families of vertex algebras from with different algebraic relations between and , called truncation curves in general. In the case , the parameters are
giving a coordinate of the truncation curve
The -algebra is obtained from the specialization of at these values via the quotient by the ideal generated by the null vector of conformal weight . On the other hand, (32) is obtained from by the specialization
which parametrizes the truncation curve
The above uniqueness ensures the Heisenberg cosets (32) coincide. The whole -superalgebra (resp. ) is characterized[29] by the vertex superalgebra containing the tensor product of the coset and (resp. ), extended by even (resp. odd) highest weight vectors of conformal weight (resp. ) and of Heisenberg weight . 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 and .
Theorem 3.3.
The condition comes from the construction of that does not vanish, which prohibits degenerations to the Virasoro vertex algebra. The case 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 itself is identified with the chiral algebras of Argyres–Douglas theories of type [4].
3.3. A digression on -characters
Let us briefly discuss the -character of the Heisenberg cosets (32). Since the character of is expressed as
the -character of the Heisenberg cosets is given by the residue
(34) |
As a -independent form, it can be rewritten into
(35) |
by the unary false theta functions
(36) |
From this, we find that there are two independent linear relations among the PBW basis for strong generators at conformal dimension 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 , which is . It is based on the following fact. When we extend by , the resulting vertex algebra is of type and the -character is given by the MacMahon function
It is known as the generating function counting plane partitions (i.e., three dimensional partitions located in the region ). In the same line, the function is the generating function of plane partitions with a pit[17] at as in Fig. 6. The first excluded plane partition is the one in Fig. 6, which has boxes. This corresponds to the fact that the field drops off from the strong generating set in (33).
3.4. Free field realization
Beside the uniqueness property of the -superalgebras based on , the free field realization of -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]
which embeds the -superalgebra into a tensor product of the affine vertex algebra at a suitable level and the symplectic boson associated with the symplectic vector superspace . Here we use the good grading used in the quantum Drinfeld–Sokolov reduction. In our cases,
(37) |
Here is the subspace of the standard Cartan subalgebra orthogonal to and respectively. When 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.
We compose them with the Wakimoto realization
(38) |
and then the bosonization[51] of the -system
(39) |
In the end, we arrive at the following free field realizations of the -superalgebras:
(40) |
At each step (38)-(39), the images are described by the kernel of a single screeing operator. Therefore, at generic levels , 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.
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):
(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 , when 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 , whose image is also characterized by an even screening operator. We stress that the screening operators for and have different form and that an automorphism of is used to identify them. Nontheless, it is generalized to our setting together with the -side in the following way:
(48) |
Here is the Heisenberg field corresponding to the matrix inside in terms of the elementary matrices . The embedding
has been introduced and used by Adamović–Kawasetsu–Ridout[5] for the Bershadsky–Polyakov algebra 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 -superalgebras are contained as
where they just stretch out in different lattices, namely and . This explains why the module categories of and look similar in general.
4. Correspondence of representation categories
Since we have the coset construction (Theorem 3.2) as in the case of and 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 . Motivated by this, we write
for the weight module categories for and , respectively. They decompose into blocks
in terms of the eigenvalues of as the spectrum of of the -superalgebras is . Then we have functors
which assign the multiplicity space of the modules inside to a -module and the multiplicity space of the modules inside to a -module , respectively. Note that the ratios are the norms of .
Theorem 4.1.
Thanks to the relation (31), the value coincides with the ratio of the Heisenberg fields , namely . Later, we will use the square root:
4.2. Fusion rules: rational case
Here we compare the fusion data of the full subcategories
when form an exceptional pair[14, 60], i.e.
(50) |
In this case, is rational [14] and -cofinite [7], and thus -mod is a semisimple braided tensor category with finitely many simple objects. These algebraic and categorical properties are inherited to -side [23, 24].
To describe the fusion data for the simple exceptional -algebra , it is better to use the level-rank duality (Theorem 3.3):
The module category -mod is a semisimple braided tensor category with simple modules satisfying the same fusion rules as -modules:
(51) |
Note that () are simple currents. They form a group since
Then the fusion ring is a ring over by fusion products
where permutes the fundamental weights , and also admits a grading by the dual group coming from the monodromy
with and . Then decomposes[29] into
Therefore, is a simple current extension of by a lattice theory, which implies that the fusion ring is described[14, 24] as
by the level-rank duality between the fusion rings of affine vertex algebras[49, 67, 24].
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 -superalgebras and as modules over the tensor product of the Heisenberg coset (32) and the Heisenberg vertex algebras inside them:
(52) |
Here and are the space of highest weight vector of the Fock modules. Then Theorem 3.1 implies that
as modules over the Heisenberg cosets. In other words, the difference between these two -superalgebras comes from the norms of the generators of the Heisenberg vertex subalgebras and . This difference can be seen most explicitly when their simple quotients and degenerate to free field algebras:
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 and 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 relative to the horizontal subalgebra . Let and be Heisenberg vertex algebras generated by the fields and satisfying the OPEs
Then the diagonal one
(53) |
is a commutative vertex algebra which acts on the tensor product of Fock modules of . In this setting, we have
(54) |
Therefore, to replace the Fock module inside with simultaneously in order to obtain the shape of , we introduce the following module
(55) |
over . Then (54) already implies that
(56) |
as modules over . It is crucial here to observe that (55) admits a vertex superalgebra structure
by changing basis of the two dimensional subspace of Heisenberg fields inside. Here again appears the lattice vertex superalgebra 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 is not a mere coincidence: indeed, when we manipulate the opposite direction in the same way, we find that appears as the main part of the gluing object
To summarize, we obtain the following reformulation of the coset construction:
Theorem 5.1.
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
(59) |
By Theorem 5.1, they give rise to functors
It turns out that they are equivalent to the coset functors:
Theorem 5.2.
[24] We have natural isomorphisms
In particular, the functors and give an equivalence of categories
(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:
for so that the values of highest weight vectors are given by
Then, given a logarithmic intertwining operator of -modules
we have a logarithmic intertwining operator
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
In this way, we obtain a linear map which assigns -intertwining operators to -intertwining operators. Obviously, we also have a linear map in the opposite direction, which we denote by .
Theorem 5.3.
[24] The linear maps are isomorphisms between the spaces of (logarithmic) intertwining operators
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 -superalgebras[29, 30] in Feigin–Frenkel type duality more generally.
Let the level be irrational. The -modules playing the role of the (contragradient of) affine Verma modules for are Wakimoto modules [55], which are defined through the composition of (37) and (38):
(61) |
This embedding is actually induced from the Wakimoto realization [44, 46] of the affine vertex algebra
(62) |
which can be extended to a resolution
(63) |
by Wakimoto modules of :
The sum runs over the elements of Weyl group of whose length are and denotes the dot action with the Weyl vector. For Weyl modules (), we have similar resolutions
(64) |
by replacing with . Introduce the following -modules:
The latter is called a Wakimoto module of and acts on it via (61). Then, by applying to (64), we obtain a resolution
(65) |
with . When , this resolution for is the contragradient dual of (20). Since (61) maps
with the Virasoro field by the Segal–Sugarawa construction, the character of the Wakimoto module is given by
with Note that agrees with the conformal weight of the free field module of as implicated by (48).
Introduce the following -modules:
Applying the functor to (65), we obtain a resolution of :
with
Here () are the spectral flow twists given[64] by
The character of is given by
Let us compare it with the characters of the Wakimoto modules over . These modules are defined as
through the free field realization (40). Since (40) sends
their characters are given by
(66) |
with Then we find that
Here is the image of under the orthogonal decomposition Indeed, is a submodule of (“thin Wakimoto module”). This is proven by introducing “thick” Wakimoto modules
whose image coincide with the Wakimoto modules . The -module structure on is given by the composition of (61) and the bosonization , that is, the free field realization (40). Replacing with its modules () below, we obtain the general Wakimoto module for .
To obtain Verma-type modules on -side as in the case of , we need to replace the Fock modules by Verma modules for and thus need the twist (40) and use (48), which satisfies
where is the Virasoro field of and are given by
Let () be the Verma module of equipped with the central character . It is generated by the one dimensional representation of the Zhu’s algebra . Under this notation, we have a natural homomorphism of -modules, which sends the highest weight vector to the other. Let be the -module defined by
Then we define the relaxed highest weight module of as
It is positively graded if and only if and we set , whose character is given by
Applying the functor , we find that the image has the following character :
When is even, the image is given by the spectral flow twist with of the Verma-type -module , which is defined in general as
where is the Verma module of with a natural homomorphism and acts on through . 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.
-side | -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 -algebras. On the other hand, since the positively graded modules are controlled by the finite -superalgebras[33], it is natural to compare the finite -superalgebras in our case. In the literature[35, 66], the finite -algebras for have a nice realization as infinitesimal Cherednik algebras.
6. Conclusions
The pair of -superalgebras and enjoy the Feigin–Frenkel type duality for their Heisenberg cosets. It can be proven either by using the uniqueness property of the universal -algebra or by the technique of free field realization. Furthermore, the duality is refined to a reconstruction type theorem of one of the -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, at , known also as the chiral algebras of Argyres–Douglas theories of type [4] have suitable module categories which have a non-semisimple braided tensor category structure [15]. It is an interesting problem to understand the -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 -superalgebras, namely hook-type -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 -characters are work in progress.[31]
References
- [1] Dražen Adamović. Representations of the superconformal vertex algebra. Internat. Math. Res. Notices, (2):61–79, 1999.
- [2] Dražen Adamović. Vertex algebra approach to fusion rules for superconformal minimal models. J. Algebra, 239(2):549–572, 2001.
- [3] Dražen Adamović. Realizations of simple affine vertex algebras and their modules: the cases and . Comm. Math. Phys., 366(3):1025–1067, 2019.
- [4] Dražen Adamović, Thomas Creutzig, Naoki Genra, and Jinwei Yang. The vertex algebras and . 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 -cofiniteness of -algebras. Int. Math. Res. Not. IMRN, (22):11605–11666, 2015.
- [8] Tomoyuki Arakawa. Rationality of -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 -algebras of type . Selecta Math. (N.S.), 23(4):2369–2395, 2017.
- [10] Tomoyuki Arakawa, Thomas Creutzig, and Andrew R. Linshaw. -algebras as coset vertex algebras. Invent. Math., 218(1):145–195, 2019.
- [11] Tomoyuki Arakawa and Edward Frenkel. Quantum Langlands duality of representations of -algebras. Compos. Math., 155(12):2235–2262, 2019.
- [12] Tomoyuki Arakawa, Ching Hung Lam, and Hiromichi Yamada. Parafermion vertex operator algebras and -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 -algebras. Comm. Math. Phys., 370(1):205–247, 2019.
- [14] Tomoyuki Arakawa and Jethro van Ekeren. Rationality and fusion rules of exceptional -algebras, 2019.
- [15] Jean Auger, Thomas Creutzig, Shashank Kanade, and Matthew Rupert. Braided tensor categories related to vertex algebras. Comm. Math. Phys., 378(1):219–260, 2020.
- [16] Christopher Beem, Leonardo Rastelli, and Balt C. van Rees. 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 -algebras and principal -superalgebras. Adv. Math., 383:Paper No. 107685, 52, 2021.
- [24] Thomas Creutzig, Naoki Genra, Shigenori Nakatsuka, and Ryo Sato. Correspondences of categories for subregular -algebras and principal -superalgebras, 2021.
- [25] Thomas Creutzig, Yasuaki Hikida, and Devon Stockal. Correlator correspondences for subregular -algebras and principal -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 -algebras, 2020.
- [30] Thomas Creutzig and Andrew R. Linshaw. Trialities of orthosymplectic -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 -algebras, 2022.
- [33] Alberto De Sole and Victor G. Kac. Finite vs affine -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 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 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. 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 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 and superconformal algebras. J. Math. Phys., 39(7):3865–3905, 1998.
- [41] Boris Feigin and Edward Frenkel. Duality in -algebras. Internat. Math. Res. Notices, (6):75–82, 1991.
- [42] Boris Feigin and Sergei Gukov. . 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, -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 -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 -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 -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. superconformal 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 -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 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 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 -algebra and vertex algebras of type . Compos. Math., 157(1):12–82, 2021.
- [66] I. Losev and A. Tsymbaliuk. Infinitesimal Cherednik algebras as -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. -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 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.