The Landau–Ginzburg/Conformal Field Theory (LG/CFT) correspondence predicts tensor equivalences between categories of matrix factorisations of certain polynomials and categories associated to the N = 2 supersymmetric conformal field theories. We realise this correspondence for the potential xd for any d ≥ 2, where previous results were limited to odd d. Our proof first establishes the fact that both sides of the correspondence carry the structure of module tensor categories over the category of Zd-graded vector spaces equipped with a non-trivial braiding. This allows us to describe the CFT side as generated by a single object as a module tensor category, and use this to efficiently provide a functor realising the tensor equivalence.

We establish the Landau-Ginzburg/Conformal Field Theory (LG/CFT) correspondence for a family of self-defects of the potential xd for any integer d ≥ 2. This generalises the main theorem of Ref. 1 to even d. The LG/CFT correspondence appeared in the physics literature in the late 1980s and early 1990s, see e.g., Refs. 2–6. It comes from the observation that every N = 2-supersymmetric Landau-Ginzburg model characterized by a potential W at the infrared fixed point of the renormalisation group flow is a conformal field theory of central charge cW. At the level of boundary conditions, this correspondence predicts equivalences of C-linear categories between (on the CFT-side) categories of representations of vertex operator algebras (VOAs) and (on the LG-side) categories of matrix factorisations of W. At the level of self-defects, these equivalences should be tensor. There have been some recent efforts toward providing a mathematical framework for this correspondence, see e.g., Refs. 1, 7, and 8.

The only instance of LG/CFT realised as a tensor equivalence is that from Ref. 1. Here, the first author and collaborators provided a tensor equivalence between the Neveu-Schwarz part CNSd of the category E(d) of representations of the super-VOA of the unitary N = 2 minimal model with central charge c=3(d2)d and the permutation type subcategory Pd of the category of matrix factorisations of xdyd, for odd d. However, independent of d’s parity the fusion rules for Pd (first described in Ref. 9, fully established in Ref. 1) agree with those of CNS(d). We prove that this agreement can be promoted to a tensor functor for any d ≥ 2:

Theorem 1.

Let d ≥ 2. There is a tensor equivalence CNS(d)Pd.

Our proof strategy provides a fresh perspective on the structure of the categories involved. We regard them as Vd-module tensor categories (cf. Ref. 10), where Vd is the category of Zd-graded vector spaces, equipped with a non-trivial braiding and trivial associator (Definition 1). These are (in particular) tensor categories A equipped with a central functor VdZ(A). As VdCNS(d) is a braided subcategory, CNS(d) is naturally a Vd-module tensor category. This allows us to in effect quotient by Vd:

Theorem 2.

The category CNS(d) is equivalent as Vd-module tensor category to the pivotal free Vd-module fusion category generated by a single object that is self-dual of charge −1 and quantum dimension 2cos(πd).

This is Theorem 7 in the main text. Intuitively, this means that CNS(d) is generated by a single object that is self dual up to the action of 1Vd, and by freely acting by Vd. This is a novel generalisation of the well-known construction of fusion categories generated by a single self-dual object as a quotient of Temperley-Lieb categories. Functors out of a category generated by a single self-dual object are easy to describe: one just needs to identify a viable target for the object and its duality data.

On the LG-side, the Vd-module tensor category structure comes from the Zd-action as automorphisms of C[x] fixing xd. We provide a trivialisation of this action on Pd to get (Theorem 32 in the main text):

Theorem 3.

The category Pd can be given the structure of a spherical Vd-module tensor category.

To establish tensor equivalence between CNS(d) and Pd we provide a self-dual object of charge −1 in Pd, giving rise to a tensor functor CNS(d)Pd. Standard arguments show that this is an equivalence.

Theorem 4.

The categories CNS(d) and Pd are equivalent as spherical free Vd-module tensor categories.

This appears as Theorem 33. Theorem 1 is a direct corollary of this.

Our generalisation of the Temperley-Lieb construction is similar to another, more powerful, generalisation discussed in Ref. 11. While our use of charged duality data falls outside the direct applicability of those results, we expect that their framework can be adapted to encompass it.

In the physics context the Zd-action arises as a symmetry of the Lagrangian defining the LG-model, and the associated chiral ring. This symmetry is preserved under the renormalization group flow, so the corresponding CFT would also have this symmetry group. Mathematically, this translates to an action by the automorphisms G of the potential W on the category of matrix (bi-)factorisations for W. We expect that this gives rise to the structure of a Vect[G]-module category to the category of matrix factorisation for W, that in turn can be lifted to a module tensor category structure on a subcategory, after picking an appropriate quadratic form on G. Identifying the associated Vect[G]-module category structure gives another hand-hold for establishing the LG/CFT correspondence in other examples, we will explore this in future work.

The paper is organized as follows. In Sec. II we review the category CNS(d) and some basics on Temperley-Lieb categories. We then describe the structure of CNS(d) as a Vd-module tensor category. In Sec. III we provide a brief overview of matrix factorisations and introduce the category Pd of permutation type matrix factorisations. We then equip this category with the structure of a Vd-module tensor category. Finally, in Sec. IV we state and prove our main result.

1. Notation and conventions

We will work over C. Fix d ≥ 2 and the primitive d-th root of unity η=e2πid throughout.

When tensoring two objects in a category we will usually omit (especially in diagrams) the ⊗ symbol and for tensor products of morphisms we will use the notation ⊗ = ·. We will write 1c for the identity on an object c, and will often omit identity morphisms tensored with any other morphisms if clear from the context.

We will use string diagram calculus for braided fusion categories to streamline our arguments. Readers unfamiliar with this are referred to e.g., Refs. 12 and 13. As is customary, unitors and associators are suppressed in the string diagrams.

2. The category Vd

We remind the reader of the classical result14 that braidings on monoidal categories with a finite group as their set of objects are completely determined by a quadratic form on the group. This allows us to define:

Definition 1.

The pointed ribbon fusion category Vd is the spherical fusion category of Zd-graded vector spaces equipped with the braiding and associators induced by the quadratic form qd(k)=e2πidk2.

Lemma 2.

The fusion category Vd has trivial associators.

Proof.
Following the proof of Ref. 14, Theorem 12, we compute that the associators for the objects l,m,nZd on Vd determined by the quadratic form are given by 1l+m+n times
These are indeed trivial.□

In this section we set up the CFT side of the correspondence. We briefly recall the categorical structure of the N = 2 unitary minimal model with central charge c=3(d2)d. After this we discuss Temperley-Lieb categories, which will serve as a template for our description of CNS(d) as generated by a single object. This first part of this section is essentially a summary of Ref. 1, Secs. 2.1 and 2.3. We then move on to our result describing CNS(d) as a Vd-module tensor category, first discussing some basics of module tensor categories, and then proving that CNS(d) can be viewed as a Vd-module tensor category generated by a single object.

1. Coset construction of CNS(d)

Let V be the vertex operator superalgebra corresponding to the N = 2 unitary minimal model with central charge c=3(d2)d with dZ2.15 The degree-zero part V0 of this superVOA can be identified with the coset su2̂d2u1̂4u1̂2d.16,17 The category of representations of this part of the VOA can be described as follows.18 Consider the category
where Rep(su(2)̂d2) is the category of integrable highest weight representations of the affine Lie algebra su2 at the level d − 2, and Repu1̂2d is the category of representations of the vertex operator algebra for u1 rationally extended by two fields of weight d. The latter is equivalent to the pointed fusion category with simple objects represented by Z2d and braiding and associators induced from the quadratic form on Z2d given by
(1)
The superscript op indicates taking the opposite braiding and ribbon twist.

Simples in Ed are labeled by triples l,r,s with l ∈ {0, …, d − 2}, rZ2d and sZ4. The category of representations of V0 can be realized as the category of local modules over the object A0,0,0d2,d,2 in Ed. Because d2,d,2 is invertible and acts freely, simple A-modules are free. That is, simple modules are of the form Al,r,s. One checks that they are local when l + r + s is even.

We are interested in the Neveu-Schwarz subcategory CNS(d), this consists of those objects with s even. In particular, we can write any simple of CNS(d) as l,rAl,r,0.

2. Categorical structure

We will work at the level of the braided fusion categorical structure of CNS(d). The fusion rules are
(2)
The associators on the Rep(su(2)̂d2)-factor are rather complicated, see Ref. 19, Appendix A.2, and we will not work with them directly. CNS(d) is ribbon, hence spherical, and the quantum dimensions are:
where ζ=eπid (so that ζ2 = η). From the fusion rules we see that there is a fusion subcategory spanned by [0, 2k] for kZd. This is a copy of Vd, its quadratic form is that from Eq. (1) pulled back to Zd:
(3)

3. Temperley-Lieb categories

We want to describe CNS(d) as category generated by a single object in the appropriate sense. To do this, we use a generalisation of the classical construction of fusion categories generated by a single self-dual object using Temperley-Lieb categories. In this section we recall this construction, closely following the treatment in Ref. 20. To start, fix a complex number κ.

Recall that a Kauffman diagram of type m,n is a planar curve connecting a line of m points to a parallel line of n points, confined to the strip between these lines. Note that there are no such curves when m and n have different parity. We denote the set of isotopy classes Kauffman diagrams by Km,n. Concatenating diagrams induces maps Km,n × Kn,rKm,r, juxtaposition gives maps Km,n × Km′,nKm+m′,n+n.

Definition 3.

The Temperley-Lieb category TLκ is the idempotent completion of the following tensor category Tκ. Its objects are spanned by the non-negative integers. For objects m and n, we set Tκm,n=CKm,n. Composition of morphisms is induced from concatenation of diagrams, subject to the relation that a circle is equivalent to the empty diagram times κ. The tensor product is addition on the objects and induced from juxtapostion on the morphisms.

The category Tκ is generated by the object corresponding to 1N. This object is self-dual, with evaluation and co-evaluation given by the cap- and cup-shaped Kauffman diagrams. Its quantum dimension is κ. This category is spherical by construction.

To find the idempotent completion, we need to examine the endomorphism algebras in Tκ. These algebras are the Temperley-Lieb algebras TLnκ. Adding a strand on the right of Kauffman diagrams in Kn,n gives a map Kn,nKn+1,n+1 which induces a map on the Temperley-Lieb algebras TLnκTLn+1κ. This allows one to inductively construct a sequence of idempotents called the Jones-Wenzl projectors pnTLnκ. Pick ζ such that κ = ζ + ζ−1. The induction terminates when the so-called quantum number
vanishes. When ζ=eiπd, this happens when n = d − 1, and we have that tr(pd−1) = [d]ζ = 0 where tr denotes the Markov trace. For further properties of the Wenzl-Jones idempotents, as well as of Temperley-Lieb categories, we refer to e.g., Refs. 20 and 21.

These idempotents give rise to a sequence of subobjects ⟨⟨n⟩⟩ ⊂ n, and for each n < d − 1 the rest of the direct summands of n can be identified with ⟨⟨i⟩⟩ ⊂ for i < n of the same parity as n. This gives the fusion rules for this category, the fusion morphisms correspond to elements of Kn,i sandwiched between pn and pi.

The final well-defined projector pd−1 is what is called a negligible morphism (see e.g., Ref. 21, Sec. 1.3). It in fact generates the tensor ideal of negligible morphisms [Ref. 21, Proposition 2.1]. The quotient by this ideal is a spherical fusion category. Mapping the object associated to 1 to the object [1]Rep(su(2)̂d2) induces a tensor functor TκRep(su(2)̂d2), which factors through this quotient as the image of pd−1 in Rep(su(2)̂d2) is zero. Using semi-simplicity one then sees that this functor is an equivalence identifying this quotient with the spherical fusion category underlying Rep(su(2)̂d2).

For d odd the category CNSd is equivalent as a braided fusion category to the Deligne tensor product TL2cos(πd)pd1Vd by Ref. 1, Proposition 2.1.

For even d, this does not hold, as there is no self-dual object to generate the Temperley-Lieb factor. In this section we will show this can be remedied by taking into account that Vd acts on CNS(d).

1. Module tensor categories

The following definition (see Ref. 11, Definition 3.1 and Ref. 10, Sec. 3.2 for more details) allows us to capture the action of Vd on CNS(d):

Definition 4.

Let D be a braided tensor category. A module tensor category over D is a tensor category M together with a braided tensor functor ΦZ:DZM from D to the Drinfeld center of M. We will call M a module fusion category if both D and M are fusion, and similarly pivotal (or spherical) if D, M and ΦZ are.

Write Φ for the composite of ΦZ with the forgetful functor Z(M)M. Then M is called free if for any simple mM and simples v,vD we have Φ(v)m ≅ Φ(v′)m if and only if vv′.

Equation (3) implies that CNS(d) is a Vd-module tensor category: it is a tensor category equipped with a braided functor Φ:VdZ(CNS(d)), which in this case factors through CNS(d)Z(CNS(d)).

Note that CNS(d) is a spherical fusion category, and for any simple cCNS(d) we have [0, 2k]c ≅ [0, 2l]c if and only if kl mod d. This means that:

Proposition 5.

CNS(d) is a spherical free Vd-module fusion category.

2. Charged morphisms

We see from the fusion rules in Eq. (2) that any object of CNS(d) is a subobject of a tensor product of tensor powers of [1,1] and [0,2]. This means that [1,1] generates CNS(d) as a Vd-module category. Additionally
meaning that [1,1] is self-dual up to tensoring by [0, −2], which corresponds to -1̲Vd. To capture this, we use the notion (see e.g., Ref. 22) of enriching a module category to an enriched category: a category whose hom-sets have been replaced by objects in the monoidal category that acts, see e.g., Refs. 23 and 24 for further details.
We enrich CNS(d) to a Vd-enriched category CNS(d) by setting, for any c,cCNS(d) and k̲Vd:

The inverse to this enrichment is applying the lax monoidal functor Vd(0̲,) to the hom-objects. We will use this to pass between CNS(d) and CNS(d) as Vd-module tensor category freely.

The enrichment is characterised by the defining adjunction relation:

Notation 6.

We will denote an element of the k̲ summand of CNS(d)(c,c) by f: ck c′, and the corresponding element of CNS(d)([0,2k]c,c) by f̄.

Note that f: ck c′ and f̄:[0,2k]cc determine each other. We will refer to a morphism f: ck c′ as a morphism of charge k, and to the pair f and f̄ as each other’s mates. The composition of f: ck c′ with f′: c′ →k c″ is f′◦f: ck+k c″, determined by:
The category CNS(d) is monoidal in the sense of Ref. 22, Definition 2.1. On objects the monoidal structure is that of CNS(d). On morphisms, the monoidal product of two morphisms f1:c1k1c1 and f1:c2k2c2 is the mate f1f2:c1c2k1+k2c1c2 to
where β is the braiding in CNS(d).

3. Charged duality

In this language, [1,1] is self-dual of charge −1. That is, we have an evaluation ev: [1,1]2−1 [0, 0] exposing [1,1] as its own left dual. It is defined by
where e is an evaluation map witnessing [1, −1] as the left dual of [1,1]. The co-evaluation has charge 1 and is the morphism coev: [0, 0] →1 [1,1]2 defined through
where ce is the co-evaluation associated to e, and β[0,2],[1,1] is the braiding. The isomorphism [0, 2][1, −1] ≅ [1, 1] is chosen to be the dual element under the composition pairing to the isomorphism [1, 1] ≅ [0, 2][1, −1] used in the definition of the evaluation. A quick computation shows that these morphisms satisfy the usual snake relations. For the right duality, we have eṽ:[1,1]21[0,0] defined by
where p(e) denotes e composed with the pivotal structure. The corresponding co-evaluation is defined by
One computes that evcoeṽ and eṽcoev agree and are equal to
showing that this number deserves to be called the quantum dimension of [1,1] also in this context.

4. Generating CNS(d)

We can now prove:

Theorem 7.

The category CNS(d) is equivalent as Vd-module tensor category to the pivotal free Vd-module fusion category generated by a single object that is self-dual of charge −1 and quantum dimension 2cos(πd).

Proof.

We need to show that the data of a self-dual object of charge −1 and quantum dimension κ=2cos(π2) is sufficient to recover CNS(d) as a Vd-module tensor category. To do this, we will generate a spherical free Vd-module tensor category K from this data, and provide an equivalence between the maximal fusion quotient of its idempotent completion and CNS(d).

We will closely follow the classical treatment of Temperley-Lieb categories as outlined in Sec. II A 3. The difference is that we will generate a Vd-enriched and tensored category to accommodate the charge. The objects of K are pairs (k,n)Zd×N (which will eventually be mapped to [0, 2k] ⊗ [1,1]n). On objects, the tensor product on K is just:
The morphisms in this category are generated from the duality data ev and coev and by enforcing that the resulting category is a free Vd-module category. That is, morphisms from (k, n) to (l, m) are a tensor product of lk̲ with the span of Kauffman diagrams Km,n between m and n up to isotopy:
Here, we make C[Km,n] into a Zd-graded vector space by setting the degree of a diagram to be the number of co-evaluations minus the number of evaluations in the diagram. This encodes the charge of the duality data. Observe that C[Km,n] is trivial when mn is odd, and has degree mn2 otherwise.

Composition of morphisms from (k, n) to (l, m) with morphisms from (l, m) to (j, r) is induced from the map Km,n × Kr,mKr,n that glues diagrams, together with the obvious isomorphism lk̲rl̲rk̲. The tensor product of morphisms is induced from juxtaposition of diagrams.

We now want to compute the idempotent completion of K. To do this, we need to examine the endomorphism algebras in K. The endomorpisms of (k, n) are of charge 0, meaning that finding all the idempotents in K reduces to the classical (uncharged) case. The hom-object K((k,n),(k,n)) is for each k the classical Temperley-Lieb algebra TLn(κ). Finding the idempotents then proceeds like in Sec. II A 3.

This gives for each k a sequence {pk,n ∈ End((k, n))}, giving rise to objects ⟨⟨k, n⟩⟩ ⊂ (k, n) with 0 ≤ nd − 2 in the idempotent completion of K.

Observe that for any k and l we have (k, n) ≅ (l, n) along an isomorphism of charge lk. These isomorphisms descend to isomorphisms
(4)
giving for each n the subcategory spanned by the ⟨⟨k, n⟩⟩ the structure of a free rank one module category over Vd.
In the classical charge 0 case, one uses the Kauffman diagrams to find the complementary summands to ⟨⟨k, n⟩⟩ in (k, n). In our charge −1 case, this analysis stays almost the same, (This should not come as a surprise. Idempotent completion only sees the underlying linear category, and does not heed the charges) the only difference being that the inclusions can now carry charge. Taking this into account and using Eq. (4) to obtain a morphism of charge zero, we find
where the projection onto the first summand is constructed from pn−1 and ev together with a charge 1 isomorphism.

Just like in the classical case Ref. 21, Proposition 2.1 the final well-defined projectors pk,d−1 generate the tensor ideal of negligible morphisms. This means we just need to provide a self-dual object in CNS(d) of charge −1, quantum dimension κ and vanishing Jones-Wenzl projector. By the discussion preceding the Theorem, [1,1]CNS(d) is such an object.□

Remark 8.

We worked with the object [1,1] for concreteness, but any [1, m] with m odd generates CNS(d) as a Vd-module category, and is self dual of charge −m with quantum dimension 2cos(πd). This gives a presentation of CNS(d) for each odd number in between 0 and 2d. If d is itself odd, this means that the object [1, d] is self-dual of charge d ≡ 0 mod d and generates CNS(d). Running through the Proof of Theorem 7 for this object, we recover the observation from Ref. 1, Proposition 2.1 that CNS(d) is, for odd d, the Deligne product of the fusion quotient of a Temperley-Lieb category with Vd.

We now turn our attention to the Landau-Ginzburg side of the correspondence. This side is described using matrix factorisations, and we briefly recall some details about these. After that, we define Pd, and prove that it can be given the structure of a Vd-module tensor category.

Here we introduce some basics of matrix factorisations. Denote by R a commutative C-algebra, and let us pick WR.

Definition 9.

A matrix factorisation of W consists of a pair M,dM, where:

  • M is a finite-rank, free, Z2-graded R-module M = M0M1; and

  • dM: MM is a degree 1 R-linear endomorphism dM=0d1Md0M0 called the twisted differential, satisfying dMdM = W.idM.

If clear from the context, we will simply use M to denote a matrix factorisation M,dM. If (R′, W′) is another ring with chosen element, a matrix factorisation for RCR and potential W ⊗ 1 − 1 ⊗ W′ is called a matrix bifactorisation following Ref. 19. We will be interested in the case where R = R′ and W = W′: bifactorisations of (R, W). The underlying module M for such a bifactorisation of W is an RR-bimodule, and dMdM = W.idM − idM.W. Note here that a R-bimodule is free if it is free as an RCR-module.

Definition 10.

Let M,dM, N,dN be two matrix bifactorisations of W. We define a morphism of matrix factorisations to be an R-linear morphism f: MN. Note that these morphisms inherit the Z2-grading of the base module, so we can write them in components as f=f0,f1.

The category MFbiW with matrix bifactorisations as objects and morphisms as defined above has the structure of a differential Z2-graded category,25 with differential in the morphism space defined as follows: let fHomMFbi(W)M,N, then

We will work in the associated homotopy category:

Definition 11.

The category HMFbiW is the category with the same objects as MFbiW, and as hom-objects the zero degree homology of the hom-objects of MFbiW.

Assuming that R=Cx1,,xn and W is a polynomial satisfying that
is finite (i.e., a potential) the category HMFbiW is tensor:19,26

Definition 12.

Given M,dM, N,dNHMFbiW, their tensor product MN is defined as follows: the base module is MR N and the twisted differential is dMN = dMidN + idMdN.

This tensor product defines a monoidal structure with trivial associators, we will discuss what is needed about the unitors below. In terms of the Z2-grading, the underlying module is
To see that the differential on the tensor product squares to the desired potential, one needs to use that the composition of tensor products of graded morphisms follows the Koszul sign rule.

Notation 13.

In this paper, we will often need to consider a morphism f in HMFbiW between a matrix factorisation M′ and a tensor product MN of two matrix factorisations. Both when MN is the source and when it is the target such a morphism has four components: the maps MiNjMi+j with i,jZ2. In the rest of this paper, we will denote these components by fij.

For further details on the (higher) categorical structure of HMFbiW, we refer to Refs. 19, 26, and 27.

Let us fix from now on R=Cx and W = xd for dZ2. We are interested in a particular subcategory Pd of HMFbiW consisting of so-called permutation type matrix factorisations. After explaining what these are we examine the structure of this category as fusion category and prove that it admits a Vd-module tensor category structure.

1. Permutation type matrix factorisations

Consider the matrix bifactorisation in HMFbiW with base module P=P0P1=Cx,y2 and twisted differential given by the matrix
where S ⊂ {0, …, d − 1}. Here, η can be chosen to be any primitive d-th root of unity, for us this is η=e2πid, cf. Sec. I A 1. We will denote such a matrix factorisation as PS and we will call it a permutation type matrix factorisation. Note that the tensor unit of HMFbiW is the permutation type matrix factorisation I = P{0}.

Notation 14.

We set S = {m, …, m + l} and denote the associated permutation type matrix factorisations as Pm;lPS. We will use PS and Pm;l, and S and ml interchangeably.

In what follows, we will only be interested in the full subcategory Pd of HMFbiW spanned by the simples Pm;l. This is a semi-simple category.1 Many of the results listed here are specialised to Pd from results for HMFbiW for any W. For the general theory see Ref. 19.

2. The monoidal structure of Pd

The subcategory Pd is closed under tensor product (see Refs. 1 and 9). The fusion rules are
(5)
for l, l′ ∈ {0, …, d − 2}, and m,mZ2d. The objects Pm;0 span a copy of Vect[Zd] with trivial associator, this is the tensor category underlying Vd (see Lemma 2). This gives an embedding of tensor categories
(6)
sending the one dimensional degree-m vector space CmVd to m̲Pm;0. We will eventually lift this embedding to a central functor VdZ(Pd), this is Theorem 32.
The left and right unitors λSλPS:P0;0PSPS and ρSρPS:PSP0;0PS are given by (recall Notation 13)
(7)
Here the maps LP and RP are defined for a C[x]-bimodule P as

3. Duals in Pd

Any matrix factorisation M in HMFbiW has a left dual M (see Refs. 26 and 27) with the same underlying module M and twisted differential dM=dM. For a permutation type matrix factorisation PSPd, there is an isomorphism
given by multiplication by (1)|S|+1jSηj on the odd summand and the identity on the even summand [Ref. 1, Eq. (3.4)]. Combining this isomorphism with the evaluation and coevaluation maps from Ref. 1, Page 12 gives maps
exhibiting PS as the left dual PS* of PS. This defines a functor ()*:PdopPd. The explicit form of the coevaluation map is in components (see Notation 13):
(8)
The evaluation map has components that are maps C[x,y,z]C[x,y] given by
(9)
The map GS is defined using a contour integral over y and is C[x,z]-linear, see Ref. 1. For our purposes it suffices to know that
(10)
as one can see from a brief computation. For further details we refer to Refs. 1 and 27. Note that Pd is a semi-simple rigid tensor category with simple unit: a fusion category.28 

4. A spherical structure for Pd

Choosing the left dual PS* to a simple PSPd to be PS induces a pivotal structure by identifying PS**=(PS)*=PS along the identity map. With this pivotal structure we can compute the quantum dimensions of simple objects PS by computing the composite
We only need to evaluate either the even or the odd part of this map. Furthermore, note [Eq. (9)] that (evS)11=0. So it suffices to compute [using that GS is C[x,z]-linear and Eq. (8)]
We will only need
where we used Eq. (10). As qdimR(Pm;l) is the complex conjugate of qdimL(Pm;l), this pivotal structure is not spherical. This is undesirable: CNS(d) is spherical. Note however that Pd is a pseudo-unitary fusion category [its fusion matrices are those of Rep(su(2)̂d2) extended by zeros], and therefore admits a unique spherical structure [Ref. 29, Proposition 9.5.1]. By [Ref. 30, Proposition 2.4] this spherical structure is such that
where qdim denotes the dimension in the spherical structure. In particular we have that
(11)
in this spherical structure. Henceforth we will use this spherical structure.

In order to prove that Pd admits the structure of a Vd-module tensor category, we will need concrete choices for the isomorphisms in Eq. (5). We further need to know how these choices interact under composition.

There is a Zd-action on Pd that helps choosing the fusion isomorphisms. We will refer to this action as twisting, and describe it in Sec. III C 1.

1. Twisting and untwisting matrix factorisations

The algebra automorphism σ:CxCx defined by xηx leaves the potential W = xd invariant. This induces a group isomorphism ZdAutCx,xd, given by kσk. These automorphisms then act on the bifactorisations of xd, we write PSba to denote a permutation type matrix factorisation whose underlying Cx-bimodule is equal to its original one as a Z2-graded C-graded vector space, but has its left and right actions twisted by σ in the following way. For pCx, and mCx,y we send
where · denotes the left or right action on the original bimodule.
It will often be convenient to keep track of which variable we are acting on by setting

Twisting PS can be untwisted to give PS with S′ a Zd-translate of S:

Definition 15.
sa,b:PSabaPSb is the isomorphism of matrix factorisations given in components as:

We will need a couple of facts about this twisting and untwisting.

Facts 16.

  1. (See Lemma 3.5, Ref. 1) Note that a,bZd, ab defines an autoequivalence of Pd. If b = −a, this auto-equivalence is tensor. When either a or b is zero, then we will abbreviate the notation by simply adding a subscript at the non-zero side, e.g., a10=a1. Note that we can relate any a̲(=Pa;0) and -a1 via s: we have s-a,0a̲=-a1.

  2. A direct consequence of this which we will use later is that (a+b)̲ is isomorphic to -aI-b along s-a,-b = -a(s0,-b) ◦ s-a,0, and similarly -a-bPS-aPS+b along -a(s-b;0)1=s-a,0s-a-b,01.

  3. Note that for any two underlying modules of two permutation type matrix factorisations respectively M and N, twisting the intermediate variable can be transferred through the tensor product:
  4. The maps s respect the addition in Zd in the sense that
    (12)

In what follows we will need the following relationship between the unitors and s:

Lemma 17.
For aZd and any PSPd, we have:

Proof.
Note that for the degree i summand of PS the left hand side acts on f(x,y)pC[x,y]C[y]Pi as
where . denotes the untwisted action of C[x] on the C[x]-bimodule Pi, and we have used that the twisted action .a of C[x] on aPi is given by x.a(−) = ηax.(−). This agrees with the definition of the unitor from Eq. (7). The proof of the second equality is analogous.□

2. Explicit fusion isomorphisms

The fusion rules from Eq. (5) are established in Ref. 1, Appendix B by finding explicit morphisms exhibiting the direct sum decomposition of the tensor product of two simples into summands. This gives a basis for the hom-spaces Pd(PSPS,PS) and the goal of this section is to establish how this interacts with the associators. The associators in Pd are trivial, meaning that they induce the identity map between Pd((PSPS)PS,PS) and Pd(PS(PSPS),PS). Our aim is to examine what this identity map looks like in the bases induced by composition of the bases from Ref. 1, Appendix B. This sets us up to give a coherent set of isomorphisms a̲Pm;la̲Pm;l in Sec. III D.

Establishing the results we need requires examining the underlying modules and the maps between them in detail, leading to a couple of lengthy arguments. Fortunately, the results can be summarised in simple string diagrams, and we work with those in the rest of the paper.

Following Ref. 1, Appendix B we take
[also denoted as ψa,(m;l)L if we choose to make the permutation set explicit] as the basis for Pd(a̲PS,PS+a). By naturality of the unitors, we can swap the last two morphisms in this composite. In string diagrams, we will depict this morphism as:

To aid readability, we will use blue strands to indicate objects in VdPd. This choice of basis behaves well under composition:

Lemma 18.
Let a,bZd and Pm;lPd. Then
or in string diagrams:

Proof.
Using the interchange law for the tensor product and naturality of the unitors, we can re-arrange both sides so that they start with
and end with λPS+a+b. As both these maps are isomorphisms, we can cancel them, leaving us with
for the left hand side, and
for the right hand side. Both sides are an image under the invertible functor I, so we can cancel the leftmost I. Fact 16.2 allows us to form commutative squares on the outermost two terms in both composites. This leaves to prove that
commutes, where we used naturality of λ to swap the middle maps in the top row, and inserted the inverse unitor. The rightmost square in the diagram clearly commutes, so we just need to show the leftmost square commutes. Using that s0;-a-bs-a,-b1=-a(sa,-a), we see this comes down to showing λ-a(sa,-a) = -aλ. This is Lemma 17, so we are done.□

This implies that the associators are given by the identity matrix in this basis. In the definition of ψa;SL we used the left unitor. There is a mirrored version
(also denoted ψ(m;l),aR) represented in string diagrams as
providing a basis vector for Pd(PSa̲,PS+a) using the right unitor. Similarly to before, with analogous proof, we have:

Lemma 19.
Let a,bZd and PSPd. Then
in string diagrams:

These morphisms interact well with the associators in the sense that:

Lemma 20.
Let a,bZd and PSPd. Then
which in string diagrams is

Proof.

This is immediate from the naturality of the unitors and the additivity of s Eq. (12).□

Using Lemma 17, we have an alternative form for these maps
The underlying modules for a̲PS (and PSa̲) are sums of C[x,y,z], that we can index by i, j ∈ {0, 1} corresponding to a̲iPm;lj. Recall from Eq. (7) that the unitors are zero on the one-summands of the unit. Tracing through the maps we find that the only nonzero components (in the sense of Notation 13) are
(13)
The maps ψL and ψR give us a priori two different choices of basis for Pd(a̲b̲,a+b̲), we will show they actually agree. This is in a sense a generalization to the result that λI=ρI.19 In HMFbiW and consequently also in Pd, morphisms are taken up to homology (see Definition 11), and so to prove this equality we need to find (à la [Ref. 19, Appendix A]) a degree 1 morphism h such that
(14)
with components hij, where i, j ∈ {0, 1} refers to the summand IiIj in II, a slight variation on Notation 13 for odd morphisms.

Lemma 21.
For a,bZd we have, in Pd(a̲b̲,a+b̲),
so

Proof.
We will show that the morphism h̃ given by h̃00=ηah00, h̃10=ηah10, h̃01=h01, and h̃11=h11 for h as in Eq. (14) defines a chain homotopy between ψa,(b;0)L and ψ(a;0),bR. That is, we want to show that
Recall that we have
as the differentials with domain a̲1 and a̲0, respectively. To unclutter the notation, we will write di for diI in the rest of this proof.
On a̲0b̲0 we need, for fC[x,y,z],
Without loss of generality, we can send xηax, zηbz and replace f by f(ηa·, ·, ηb·). Note that this acts on the differentials as
This means we get, using that h is C-linear,
and this is exactly the 00 component of Eq. (14). Using the same transformation, one similarly shows that the other components hold.□

The maps ψL (or equivalently ψR) give maps
that we can use to expose -a̲ as the left dual of a̲. Suppressing as is customary unit strands from the string diagrams, this means we set

We claim that this evaluation agrees with the evaluation from Eq. (9). Because a̲ is invertible any two choices of evaluation are related by a scalar multiplication and hence it suffices to show that eva̲coeva̲=1I:

Lemma 22.
Let aVd and take eva̲ as defined above, and coev-a̲ from Eq. (8). Then

Proof.
We work degree by degree, using Notation 13. On the degree 1 summand I1=C[x,z] we need to compute
Now (eva)10(coev-a)10=0 as (eva)10=0, and using that (coev-a)01=η-a we see
where the factor ηa comes from the evaluation, see Eq. (13). So the degree 1 part of this map is indeed the identity.
For the degree 0 component, we have (eva)11=0, leaving us with
This finishes the proof.□

We will denote the coevaluation in string diagrams as
We will now move on to considering the interaction of these basis vectors with fusion of arbitrary simples. In order to do this, we need some information about the fusion isomorphisms from Ref. 1, Appendix B. These are inductively defined basis vectors for those Pd(Pm;l,Pm;lPm;l) that are nontrivial, starting from
(15)
We will not need the explicit form of these maps. It will suffice to know that as maps (Pm+m+12(1ϵ);l+ϵ)i+jC[x,y](Pm;1)i(Pm;l)jC[x,y,z] they are determined by multiplication by homogeneous gijϵC[x,y,z]. Writing ϵ for ±1, the degrees of these polynomials are
(16)
for i, j ∈ {0, 1}. In what follows we will depict these inductively defined morphisms φSS,SPd(PS,PSPS) simply by a trivalent vertex,
for |S′|, |S″| > 1. Note that
The reader will notice that these φSS,S map into the tensor product PSPS rather than out of it as ψL and ψR do. We will therefore work with (ψL)1 and (ψR)1 instead, (The reader may wonder why we introduced ψL rather than (ψL)1 in the first place. This is because the unitors are easier to describe than their inverses, but the basis vectors for Pd(Pm;l,Pm;lPm;l) can be expressed as multiplication by a polynomial while their inverses are more complicated.) and we will depict them by
Observe that, because they only involve isomorphisms, Lemmas 18, 19, 20, and 21 have clear analogues for these inverses that are obtained by reading the string diagrams downwards.

The result we are after is:

Proposition 23.
Let aVd and Pm;l,Pm;l,Pm;lPd. Then
where we have represented eva:a̲a̲I by a cap.

Whenever Pm;l does not appear as a summand of Pm′;lPm″;l, this proposition is trivially true. We will proceed by induction on l′, mirroring the argument from Ref. 1, Appendix B. Note that picking m′, l′, m″, l″ determines the possible choices for m, l. The base case for the induction (l′ = 1) is:

Lemma 24.
For aVd and the gm,(m;l)ϵ from Eq. (15), we have

Proof.
For brevity, write lϵ = l″ + ϵ and mϵ=m+m+12(1ϵ). We will proceed in two steps. To begin, we will show
where the left and right non-vertex dots in the right hand diagram represent s0,−a and sa,0, respectively. The non-trivial step is showing that the upper and lower composites in
coming from the left and right hand diagrams, respectively, agree up to ηa2(1ϵ). Because the polynomials (gm,(m;l)ϵ)ij are homogeneous, we have that σxaσyaσzagm,(m;l)ϵ=ηadeg(gm,(m;l)ϵ)gm,(m;l)ϵ, where deggm,(m;l)ϵ denotes the degree of the maps gm,(m;l)ϵ, see Eq. (16). Now, consider an element fx,z in Pmϵ;lϵ of degree k, where k = 0 if i = j and k = 1 otherwise. One can rewrite this as k=1211i+j. Tracing this element through the composition of maps we obtain
for the upper composite, and
for the bottom route. These indeed differ by a factor of ηa2(ϵ1).
What remains to be checked is the equality:
The inverses of ψL and ψR on the right hand side start with s0,−a and sa,0, just like appear in the left hand diagram. The next maps are then inverse unitors ρ(Pm;1)a1λ-aPm,l1. By the triangle equations for the unitors, we can replace this by λI11-aPm;lλ-aPm,l1. This is then followed by
Using Lemma 22, we see that this is the same as
where the last equality is Lemma 17. Composing this with the inverse unitors we get
where we have omitted some identity morphisms for brevity. Using Lemma 17 on s-a,a1 and λI1, we see this is just a composition of unitors, and we are done.□

The induction step is then:

Proof of Proposition 23.
Assume that Proposition 23 holds for all a,m,m,mZd, l, l″ ∈ {0, 1, …, d − 2} and l′ < L, we want to show that it holds for l′ = L + 1. For our induction we will use that P0;1Pm;L includes along g0,(m;L)+ onto Pm;L+1, and this allows us to use that the proposition holds for P0;1 and Pm;L. Using the associators we have for any m, m′, m″ and l′, l″ that
for some αC*. This allows us to compute

The first equality corresponds to applying this associator, the second and third use the induction hypothesis, the fourth is the inverse of the associator, and the last equality again uses the induction hypothesis. Composing with the projection morphism π: P0,1Pm,LPm;L+1 associated to the inclusion g0,(m;L)+ and using that πg0,(m;L)+=1Pm,L+1 now yields the result.□

We will now use the results from the previous sections to equip Pd with the structure of a Vd-module tensor category. That is, we will provide what is called a central lift VdZ(Pd) for the embedding VdPd from Eq. (6). Rather than doing this directly, it is convenient to instead trivialise the adjoint action of Vd on Pd, as we explain in Sec. III D 1.

1. Central functors and trivialisations of adjoint actions

In this section we recall some basics about central lifts, and record the observation that under suitable conditions central lifts correspond to trivialisations of adjoint actions. This allows us to use the results from the previous sections to give a central lift for the inclusion VdPd through specifying a trivialisation of the adjoint action of Vd.

Definition 25.

Let B be a braided monoidal category and C monoidal category.

  • A central functor (in the sense of Ref. 10) from a B to C is a braided monoidal functor BZ(C) into the Drinfeld center of C.

  • A central lift of a functor BC to a central functor is a factorisation BZ(C)C of that functor through the forgetful functor Z(C)C.

Concretely, a lift to a central functor means that we have to pick, for every bB, a natural isomorphism βb (the half-braiding) with components
natural in b, and satisfying for all b,bB and c,cC the hexagon equations
(17)
(18)
where we have suppressed the associators. The lift is a functor between braided monoidal categories, and we require this functor to be braided in the usual sense.

Even if B does not carry a braiding, it still makes sense to talk about lifts of functors BC to functors BZ(C). In what follows, this will be useful as an intermediate step, and we will refer to this as an unbraided central lift (or unbraided central functor).

Definition 26.
Suppose that B is a pointed fusion category, then the adjoint action associated to the inclusion ι:BC is the monoidal functor
from the Picard groupoid of invertible objects Pic(B) of B to the tensor automorphisms of C. Clearly Adb◦Adb ≅ Adbb. The monoidal coherence isomorphism μ for Adb is

We can rephrase the data of an unbraided central lifts in terms of trivialisations of the adjoint action:

Lemma 27.
Let ι:BC be an embedding of a spherical pointed fusion category into a fusion category. Then a lift of this embedding to an unbraided central functor is equivalently an isomorphism
in the groupoid of tensor automorphisms of C.

Proof.
Given a central lift BZ(C), define the components of τ to be the natural isomorphisms τb with components
or in string diagrams
This is natural in c by naturality of βb. These components are indeed monoidal in c, use Eq. (17) to see that the two compositions in the diagram
are
Here we used that coevb=evb1, as B is spherical and pointed.
This τ is natural in b as β is. To see that τ is monoidal, observe that Eq. (18) gives
This shows that a central lift gives rise to a trivialisation of the adjoint action.
For the converse we define βb for b a simple object by
(19)
that is
This determines βb,c for all bB. Similar arguments to the ones above show that this indeed satisfies the conditions for an unbraided central lift.□

2. VdPd and a central lift for it

In this subsection we will show that VdPd admits a central lift. We will do this through Lemma 27, using the results from Sec. III C 2 to produce the trivialisation for the adjoint action Ad of Vd on Pd. That is:

Proposition 28.
The natural isomorphism τ:AdIdPd with components for aZd̲ and Pm;lPd given by
is monoidal in both a and Pm;l.

We remark that we want τ to be a natural isomorphism in the category of monoidal functors Fun(Zd̲,Aut(Pd)) with values in the tensor automorphisms of Pd. As Zd̲ is a discrete category and Pd is semi-simple, showing naturality is trivial. We need to show that it is monoidal on Zd̲, and monoidal on Pd. Recall that Ada◦Adb ≅ Ada+b by fusion of a̲ and b̲ along ψa;(b;0)L and simultaneously fusing -a̲ and -b̲ to -a-b̲, so the monoidality on Zd̲ is:

Lemma 29.
Let a,bZd̲ and Pm;lPd. Then

Proof.

Apply Lemma 20, then Lemmas 18 and 19, followed by Lemma 21.□

Recall that the isomorphism Ada(PP′) ≅ Ada(P) ⊗ Ada(P′) is given by 1a̲Pcoeva1-a̲P. The monoidality in Pd is then a consequence of Proposition 23:

Corollary 30.
Let Pm;l,Pm;l,Pm;lPd and aZd̲. Then

Proof.

From the fusion rules Eq. (5) we see that m=m+m+12(l+lν), with ν the summation variable, this gives the exponent of ηa on the left hand side. Using Proposition 23, we see that the exponent of ηa on the right hand side is equal to this.□

This concludes the Proof of Proposition 28. Using Lemma 27 now gives an unbraided central lift for VdPd. We will now show that this unbraided lift is in fact braided, for the braiding on Vd as induced from the quadratic form qd (Definition 1). It suffices to show that the self-braiding on a̲ from the central lift [Eq. (19)] agrees with the self-braiding on Vd, which is given by qd(a)1a̲a̲=ηa21a̲a̲.

Lemma 31.
For aZd̲ we have

Proof.
By Lemma 20, we can swap the order of the fusion vertices on the left hand side. The first fusion vertex is then
by Lemmas 21 and 22. Now applying the snake relation for ev-a and coeva gives the result.□

In summary:

Theorem 32.

The category Pd can be given the structure of a free spherical Vd-module fusion category.

We are now in a position to prove our main result. Let us summarise what we have found so far. We have established that both CNS(d) (Proposition 5) and Pd (Theorem 32) are spherical free Vd-module fusion categories. We have further established that CNS(d) is generated by the charge −1 self-dual object [1,1] of quantum dimension 2cos(πd), Theorem 7. Recall that in the spherical pivotal structure P0;1 also has quantum dimension 2cos(πd) Eq. (11). Furthermore, its dual is P−1,1, making it self-dual of charge −1. This gives a tensor functor CNS(d)Pd mapping [1,1] to P0;1. Using semi-simplicity one sees that this functor is an equivalence. That is, we have:

Theorem 33.

The categories CNS(d) and Pd are equivalent Vd-module tensor categories for any d, with equivalence that sends [l, 2m + l] ↦ Pm;l.

In particular, CNS(d) and Pd are equivalent as spherical fusion categories.

Remark 34.

Following on from Remark 8, it is educative to compare our methods with Ref. 1. Remark 8 addresses how our view on CNS(d) changes. What it does not address is the absence of equivariant duality data for P0;1. To recover this, note that we have provided something stronger: we have a trivialisation of the adjoint action of Vd on all of Pd. Specialising this trivialisation to P0;1 gives rise to the data used in Ref. 1.

The authors gratefully acknowledge the Topology and Geometry conference (2015) and the Blow Up in Bonn for making sure our paths crossed. A.R.C. is supported by Cardiff University. We also gratefully acknowledge support from Cardiff University for hospitality for Thomas during his visit. T.A.W. is supported by Ric Wade’s Royal Society Grant Nos. RF∖ERE∖210416 and RF∖ERE∖231013. T.A.W. is grateful to André Henriques, Márton Habliscek, and Christoph Weiss for helpful discussions during the preparation of this paper. T.A.W. also wants to thank the wardens of Skokholm Island, the Oxford Zoology Knowles Lab, and autumnal storms for a long and beautiful stay on an isolated island that inspired this paper.

The authors have no conflicts to disclose.

Ana Ros Camacho: Conceptualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Thomas A. Wasserman: Conceptualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
A.
Davydov
,
A.
Ros Camacho
, and
I.
Runkel
, “
N = 2 minimal conformal field theories and matrix bifactorisations of xd
,”
Commun. Math. Phys.
357
(
2
),
597
629
(
2018
).
2.
I.
Brunner
,
M.
Herbst
,
W.
Lerche
, and
B.
Scheuner
, “
Landau–Ginzburg realization of open string TFT
,”
J. High Energy Phys.
2006
,
043
.
3.
P. S.
Howe
and
P. C.
West
, “
N = 2 superconformal models, Landau–Ginsburg Hamiltonians and the ϵ expansion
,”
Phys. Lett. B
223
(
3–4
),
377
385
(
1989
).
4.
D. A.
Kastor
,
E. J.
Martinec
, and
S. H.
Shenker
, “
RG flow in N = 1 discrete series
,”
Nucl. Phys. B
316
(
3
),
590
608
(
1989
).
5.
E. J.
Martinec
, “
Algebraic geometry and effective Lagrangians
,”
Phys. Lett. B
217
(
4
),
431
437
(
1989
).
6.
C.
Vafa
and
N.
Warner
, “
Catastrophes and the classification of conformal theories
,”
Phys. Lett. B
218
(
1
),
51
58
(
1989
).
7.
N.
Carqueville
,
A.
Ros Camacho
, and
I.
Runkel
, “
Orbifold equivalent potentials
,”
J. Pure Appl. Algebra
220
(
2
),
759
781
(
2016
).
8.
A.
Ros Camacho
, “
On the Landau–Ginzburg/conformal field theory correspondence
,” in
Vertex Oper. Algebr. Number Theory Relat. Top.
,
Contemporary Mathematics
(
AMS
,
2020
), Vol.
753
.
9.
I.
Brunner
and
D.
Roggenkamp
, “
B-type defects in Landau–Ginzburg models
,”
J. High Energy Phys.
2007
(
08
),
093
.
10.
A. G.
Henriques
,
D.
Penneys
, and
J.
Tener
, “
Categorified trace for module tensor categories over braided tensor categories
,”
Doc. Math.
21
,
1089
1149
(
2016
).
11.
A.
Henriques
,
D.
Penneys
, and
J.
Tener
, “
Planar algebras in braided tensor categories
,”
Mem. Am. Math. Soc.
282
,
1392
(
2023
).
12.
A.
Joyal
and
R.
Street
, “
Braided tensor categories
,”
Adv. Math.
102
,
20
78
(
1993
).
13.
B.
Bartlett
, “
Fusion categories via string diagrams
,”
Commun. Contemp. Math.
18
(
05
),
1550080
(
2016
).
14.
A.
Joyal
and
R.
Street
, “
Braided monoidal categories
,”
Macquarie Math. Report No. 860081
,
1986
.
15.
D.
Adamović
, “
Representations of the N = 2 superconformal vertex algebra
,”
Int. Math. Res. Not.
,
1999
(
2
),
61
79
.
16.
S.
Carpi
,
R.
Hillier
,
Y.
Kawahigashi
,
R.
Longo
, and
F.
Xu
, “
N = 2 superconformal nets
,”
Commun. Math. Phys.
336
(
3
),
1285
1328
(
2015
).
17.
P.
Di Vecchia
,
J. L.
Petersen
,
M.
Yu
, and
H. B.
Zheng
, “
Explicit construction of unitary representations of the N = 2 superconformal algebra
,”
Phys. Lett. B
174
(
3
),
280
284
(
1986
).
18.
J.
Fröhlich
,
J.
Fuchs
,
I.
Runkel
, and
C.
Schweigert
, “
Algebras in tensor categories and coset conformal field theories
,”
Fortschr. Phys.
52
(
6–7
),
672
677
(
2004
).
19.
N.
Carqueville
and
I.
Runkel
, “
On the monoidal structure of matrix bi-factorizations
,”
J. Phys. A: Math. Theor.
43
(
27
),
275401
(
2010
).
20.
S.
Yamagami
, “
A categorical and diagrammatical approach to Temperley–Lieb algebras
,” arXiv:math/0405267 [math.QA].
21.
M. H.
Freedman
,
H.
Wenzl
, and
M. H.
Freedman
, “
A magnetic model with a possible Chern–Simons phase
,”
Commun. Math. Phys.
234
(
1
),
129
183
(
2003
).
22.
S.
Morrison
and
D.
Penneys
, “
Monoidal categories enriched in braided monoidal categories
,”
Int. Math. Res. Not.
2019
(
11
),
3527
3579
.
23.
G. M.
Kelly
,
Basic Concepts of Enriched Category Theory
(
Cambridge University Press
,
1982
).
24.
S. M.
Lane
,
Categories for the Working Mathematician
, 2nd ed. (
Springer-Verlag
,
New York, NY
,
1971
).
25.
C. A.
Weibel
,
An Introduction to Homological Algebra
, 1st ed. (
Cambridge University Press
,
1994
).
26.
N.
Carqueville
and
D.
Murfet
, “
Adjunctions and defects in Landau–Ginzburg models
,”
Adv. Math.
289
,
480
566
(
2016
).
27.
N.
Carqueville
and
I.
Runkel
, “
Rigidity and defect actions in Landau–Ginzburg models
,”
Commun. Math. Phys.
310
(
1
),
135
179
(
2012
).
28.
P.
Etingof
,
D.
Nikshych
, and
V.
Ostrik
, “
On fusion categories
,”
Ann. Math.
162
(
2
),
581
642
(
2005
).
29.
P.
Etingof
,
S.
Gelaki
,
D.
Nikshych
, and
V.
Ostrik
,
Tensor Categories
, 1st ed. (
American Mathematical Society
,
2015
), ISBN: 978-1-4704-2349-0.
30.
M.
Müger
, “
From subfactors to categories and topology I: Frobenius algebras in and Morita equivalence of tensor categories
,”
J. Pure Appl. Algebra
180
(
1–2
),
81
157
(
2003
).
Published open access through an agreement with JISC Collections