Abstract
The unitary rational orbifold conformal field theories in the algebraic quantum field theory and subfactor theory framework are formulated. Under general conditions, it is shown that the orbifold of a given unitary rational conformal field theory generates a unitary modular category. Many new unitary modular categories are obtained. It is also shown that the irreducible representations of orbifolds of rank one lattice vertex operator algebras give rise to unitary modular categories and determine the corresponding modular matrices, which has been conjectured for some time.
1. Introduction
Cosets and orbifolds are two methods of producing new two-dimensional conformal field theories from given ones (1). In refs. 2–5, unitary coset conformal field theories are formulated in the algebraic quantum field theory and subfactor theory (6) framework, and such a formulation is used to solve many questions beyond the reach of other approaches. The main purpose of this paper is to formulate unitary orbifold conformal field theories in the same framework and to give some applications of this formulation.
There is another approach to conformal field theories by using the theory of vertex operator algebras (cf. refs. 7 and 8). In the case of orbifolds, this has been studied, for example, in refs. 9 and 10. Although there are various advantages to these different approaches, our main results, Theorems 4.3 and 5.4, have not been obtained previously by other methods.
Under general conditions, as specified in Theorem 4.3, it is shown that the orbifold of given unitary rational conformal field theories generates a unitary modular category. Theorem 4.3 gives a large family of new unitary modular categories, which can be found in Sections 5 and 6. As an application of this general theory, it is shown in Theorem 5.4 that the irreducible representations of orbifolds of rank one lattice vertex operator algebras give rise to unitary modular categories [hence a unitary three-dimensional topological quantum field theory, cf. ref. 11] and the corresponding modular matrices are determined. More precisely, the simple objects of the modular categories are in one-to-one correspondence with the irreducible representations of these vertex operator algebras, which were classified in ref. 10. These simple objects and the modular matrices first appeared as examples in ref. 12 on the basis of certain heuristic arguments, and these examples can be clearly interpreted as a conjecture on the existence of certain unitary modular categories with the same modular matrices. Theorem 5.4 thus confirms this conjecture.
2. Preliminaries
First, the notion of irreducible conformal precosheaf and its covariant representations is recalled, as described in ref. 13.
By an interval, I shall always mean an open connected subset I of S1 such that I and the interior I′ of its complement are nonempty. I shall denote by ℐ the set of intervals in S1. I shall denote by PSL(2, R) the group of conformal transformations on the complex plane that preserve the orientation and leave the unit circle S1 globally invariant. Denote by G the universal covering group of PSL(2, R). Notice that G is a simple Lie group and has a natural action on the unit circle S1.
Denote by R(ϑ) the (lifting to G of the) rotation by an angle ϑ. This one-parameter subgroup of G will be referred to as rotation group (denoted by Rot) in the following. An irreducible conformal precosheaf 𝒜 of von Neumann algebras on the intervals of S1 is a map
from ℐ to the von Neumann algebras on a separable Hilbert space ℋ that satisfies the following properties:
- A. Isotony. If I1, I2 are intervals and I1 ⊂ I2, then
- B. Conformal invariance. There is a nontrivial unitary representation U of G on ℋ such that
C. Positivity of the energy. The generator of the rotation subgroup U(R(ϑ)) is positive.
D. Locality. If I0, I are disjoint intervals, then 𝒜(I0) and A(I) commute. The lattice symbol ∨ will denote “the von Neumann algebra generated by.”
E. Existence of the vacuum. There exists a unit vector Ω (vacuum vector), which is U(G) invariant and cyclic for ∨I∈ℐ𝒜(I).
F. Irreducibility. The only U(G)-invariant vectors are the scalar multiples of Ω.
The term irreducibility refers to the fact (cf. Proposition 1.2 of ref. 13) that under the assumption F ∨I∈ℐ𝒜(I) = B(ℋ).
A covariant representation π of 𝒜 is a family of representations πI of the von Neumann algebras 𝒜(I), I ∈ ℐ on a separable Hilbert space ℋπ and a unitary representation Uπ of the covering group G of PSL(2, R), such that the following properties hold:
A covariant representation π is called irreducible if ∨I∈ℐπ(𝒜(I)) = B(ℋπ). By our definition, the irreducible conformal precosheaf is in fact an irreducible representation of itself, and we will call this representation the vacuum representation.
Let H be a simply connected simply-laced compact Lie group. By Theorem 3.2 of ref. 14, the vacuum positive energy representation of the loop group LH (cf. refs. 15 or 16) at level k gives rise to an irreducible conformal precosheaf denoted by 𝒜Hk. By Theorem 3.3 of ref. 14, every irreducible positive energy representation of the loop group LH at level k gives rise to an irreducible covariant representation of 𝒜Hk. When Hk ⊂ Gl is a connected subgroup of a simply connected Lie group, Proposition 2.2 in ref. 2 gives an irreducible conformal precosheaf, which will be denoted by 𝒜Gl/Hk, and this is referred to as the coset conformal precosheaf. We will see such examples in Section 4.
Next we will recall some of the results of ref. 17 and introduce notations.
Let {[ρi], i ∈ I} be a finite set of equivalence classes of irreducible covariant representations of an irreducible conformal precosheaf with finite index. For the definitions of the conjugation and composition of covariant representations, see Section 4 of ref. 14 or Section 2 of ref. 13.
Suppose this set is closed under conjugation and composition. We will denote the conjugate of [ρi] by [ρı̄] and the identity sector by [1] if no confusion arises and let Nijk = 〈[ρi][ρj], [ρk]〉. Here 〈x, y〉 denotes the dimension of the space of intertwinners from x to y [denoted by Hom(x, y)] for any representations x and y (by Theorem 2.3 of ref. 13, we do not have to distinguish between local and global intertwinners here). We will denote by {Te} a basis of isometries in Hom(ρk, ρiρj). The univalence of ρi and the statistical dimension of (cf. Section 2 of ref. 13) will be denoted by ωρi and dρi respectively.
Let φi be the unique minimal left inverse of ρi; define:
where ɛ(ρj, ρi) is the unitary braiding operator (cf. ref. 13).
Define σ̃ := Σidρi2ωρi−1. If the matrix (Yij) is invertible, by Proposition on p. 351 of ref. 17, σ̃ satisfies |σ̃|2 = Σidρi2. Suppose σ̃ = |σ̃|exp(ix), x ∈ ℝ. Define matrices
where C := exp(i(x/3)). Then these matrices satisfy the algebraic relations:
where Ĉij = δi is the conjugation matrix. Moreover,
1 |
Eq. 1 is known as the Verlinde formula, which determines the fusion rules Nijk from the S matrix.
3. The Orbifolds
Let 𝒜 be an irreducible conformal precosheaf on a Hilbert space ℋ, and let G be a finite group. Let V : G → U(ℋ) be a faithful§ unitary representation of G on ℋ.
Definition 3.1:
We say that G acts properly on 𝒜 if the following conditions are satisfied:
(i) For each fixed interval I and each g ∈ G, αg(a) := V(g)aV(g*) ∈ 𝒜(I), ∀a ∈ 𝒜(I);
(ii) For each g ∈ G, V(g)Ω = Ω, ∀g ∈ G.
Remark 3.1:
As pointed out to us by Roberto Longo, conditions i and ii above imply that for each g ∈ G and h ∈ G, [V(g), U(h)] = 0 (cf. ref. 18), which is a condition we initially added in Definition 3.1. However, in all known examples, all three conditions above are easily checked.
Suppose a finite group G acts properly on 𝒜 as above. For each interval I, define ℬ(I) := {a ∈ 𝒜(I)|V(g)aV(g*) = a, ∀g ∈ G}. Let ℋ0 = {x ∈ ℋ|V(g)x = x, ∀g ∈ G} and P0, the projection from ℋ to ℋ0. Notice that P0 commutes with every element of ℬ(I) and U(g), ∀g ∈ G.
Define 𝒜G(I) := ℬ(I)P0 on ℋ0. The unitary representation U of G on ℋ restricts to an unitary representation (still denoted by U) of G on ℋ0. Then by using the definitions, one can check the following:
Proposition 3.2. The map I ∈ ℐ → 𝒜G(I) on ℋ0 together with the unitary representation (still denoted by U) of G on ℋ0 is an irreducible conformal precosheaf.
The irreducible conformal precosheaf in Proposition 3.2 will be denoted by 𝒜G and will be called the orbifold of 𝒜 with respect to G.
Remark 3.3:
The net ℬ(I) ⊂ 𝒜(I) is a standard net of inclusions (cf. ref. 19) with conditional expectation ɛ defined by
Note that ɛ has finite index. We can therefore apply the theory in ref. 20 (also cf. refs. 21 and 22) to this setting. It follows, for example, that if G acts properly on 𝒜 as in Definition 3.1, then for each I, the action of G on 𝒜(I) is outer, i.e., ℬ(I)′ ∩ 𝒜(I) ≃ ℂ, where ℬ(I) is the fixed point subalgebra of 𝒜(I) under the action of G.
4. Complete Rationality
As in ref. 13, by an interval of the circle, we mean an open connected proper subset of the circle. If I is such an interval, then I′ will denote the interior of the complement of I in the circle. We will denote by ℐ the set of such intervals. Let I1, I2 ∈ ℐ. We say that I1, I2 are disjoint if Ī1 ∩ Ī2 = ∅, where Ī is the closure of I in S1. When a I1, I2 are disjoint, I1 ∪ I2 is called a 1-disconnected interval in ref. 23. Denote by ℐ2 the set of unions of disjoint 2 elements in ℐ. Let 𝒜 be an irreducible conformal precosheaf, as in Section 2.1. For E = I1 ∪ I2 ∈ ℐ2, let I3 ∪ I4 be the interior of the complement of I1 ∪ I2 in S1, where I3, I4 are disjoint intervals. Let
Note that 𝒜(E) ⊂ 𝒜̂(E). Recall that a net 𝒜 is split if 𝒜(I1) ∨ 𝒜(I2) is naturally isomorphic to the tensor product of von Neumann algebras 𝒜(I1) ⊗ 𝒜(I2) for any disjoint intervals I1, I2 ∈ ℐ. 𝒜 is strongly additive if 𝒜(I1) ∨ 𝒜(I2) = 𝒜(I), where I1 ∪ I2 is obtained by removing an interior point from I.
Definition 4.1:
(complete rationality of ref. 24). 𝒜 is said to be completely rational, or μ rational, if 𝒜 is split, strongly additive, and the index [𝒜̂(E) : 𝒜(E)] is finite for some E ∈ ℐ2. The value of the index [𝒜̂(E) : 𝒜(E)] (it is independent of E by Proposition 5 of ref. 24) is denoted by μ𝒜 and is called the μ index of 𝒜.
Proposition 4.2. Let 𝒜 be an irreducible conformal precosheaf, and let G be a finite group acting properly on 𝒜. Suppose that 𝒜 is split and strongly additive. Then 𝒜G is also split and strongly additive.
Proof:
It follows from the definitions that 𝒜G is split.
Let I be an interval and I1, I2 be the connected components of a set obtained from I by removing an interior point of I. To show that 𝒜G is strongly additive, it is sufficient to show that 𝒜(I1) ∨ ℬ(I2) = 𝒜(I). Let us show that 𝒜(I1) ∨ ℬ(I2) = 𝒜(I). First note that [𝒜(I) : 𝒜(I1) ∨ ℬ(I2)] < ∞. In fact, let I2(n) ⊂ I2 be an increasing sequence of intervals such that I2(n) have one boundary point in common with I2, Ī1 ∩ Ī2(n) = ∅, and ∪nI2(n) = I2. By the additivity of the conformal net 𝒜 (cf. ref. 13), we have that 𝒜(I1) ∨ ℬ(I2(n)) [respectively 𝒜(I1) ∨ 𝒜(I2(n))] are increasing sequences of von Neumann algebras such that
where we have used the assumption that 𝒜 is strongly additive. By the splitting property, we have
It follows (cf. Proposition 3 of ref. 24) that [𝒜(I1) : 𝒜(I1) ∨ ℬ(I2)] ≤ |G|.
So there exists a faithful normal conditional expectation ɛ̃ : 𝒜(I) → 𝒜(I1) ∨ ℬ(I2). Note that
and so ɛ̃(𝒜(I2)) in an intermediate von Neumann algebra between ℬ(I2) and 𝒜(I2). So (cf. ref. 25 or refs. therein) there exists a subgroup K of G such that ɛ̃(𝒜(I2)) is the pointwise fixed subalgebra of 𝒜(I2) under the action of K. Because ℬ(I2) ⊂ 𝒜(I2) is irreducible by Remark 3.3, ɛ̃(𝒜(I2)) ⊂ 𝒜(I2) is also irreducible, and it follows that there exists a unique conditional expectation from 𝒜(I2) to ɛ̃(𝒜(I2)), given by
Let us show that K is the trivial subgroup, i.e., if k ∈ K, then k is the identity element of G.
Let v ∈ 𝒜(I2) be the isometry (cf. Section 2 of ref. 19) such that
Define a map γ̃ : 𝒜(I) → 𝒜(I1) ∨ ℬ(I2) by:
One checks easily that
It follows that γ̃(xy) = γ̃(x)γ̃(xy) for any x, y ∈ 𝒜(I), because 𝒜(I) is generated by two commuting subalgebras A(I1) and A(I2).
For any k ∈ K, define vk = αk(v) and
Then one checks that
It follows that
for any x, y ∈ 𝒜(I), because 𝒜(I) is generated by two commuting subalgebras A(I1) and A(I2). One can also check similarly α̃k(α̃k−1(x)) = x, ∀x ∈ 𝒜(I). So α̃k is an automorphism of 𝒜(I). Because 𝒜(I) is a type III factor, there exists a unitary operator Uk ∈ B(H0) such that α̃k(x) = UkxU*k, ∀x ∈ 𝒜(I). Because α̃k(x1) = x1, ∀x1 ∈ 𝒜(I1), we have Uk ∈ 𝒜(I1)′ = 𝒜(I′1) by Haag duality (cf. Section 2 of ref. 13).
If for all unitary U′ ∈ 𝒜(I′), (UkU′Ω, Ω) = 0, then (U′Ω, U*kΩ) = 0, and it follows that (𝒜(I′)Ω, U*kΩ) = 0. Because 𝒜(I′)Ω is dense in ℋ by the Reeh–Schlieder theorem (cf. ref. 13), it follows that U*kΩ = 0, which implies U*k = 0 by using the Reeh–Schlieder theorem again because U*k ∈ 𝒜(I′1). Hence there exists a unitary U′ ∈ 𝒜(I′) such that (UkU′Ω, Ω) ≠ 0. Note that 𝒜(I′) ⊂ 𝒜(I′1). Replacing Uk by UkU′ if necessary, we may assume that (UkΩ, Ω) ≠ 0.
Let gn ∈ G be a sequence of elements, such that gnI′1 = I′1, and gnI2 is an increasing sequence of intervals containing I2, i.e., I2 ⊂ gnI2 ⊂ gn+1I2, and ∪ngnI2 = I′1 (one may take gn to be a sequence of dilations). By applying Ad(U(gn)) to the equation
and by using αk(Ad(U(gn))x2) = Ad(U(gn))(αk(x2)), we get
It follows that
Let U be a weak limit of Ad(U(gn))(Uk). Note that
because Ad(U(gn))(Uk) ∈ 𝒜(I′1) by our choice of gn. Because (Ad(U(gn))(Uk)Ω, Ω) = (UkΩ, Ω) ≠ 0, where we use the fact that Ω is invariant under the action of U(gn), it follows that
so U ≠ 0, and we have
Because ∪ngnI2 = I′1, ∨n𝒜(gnI2) = A(I′1) by the additivity of the conformal net 𝒜 (cf. ref. 13), it follows that
Recalling that αk is an automorphism of 𝒜(I′1) and 𝒜(I′1) is a factor, it follows that
Changing U into (1/)U if necessary, we may assume that U is unitary, and so we have
So U ∈ B(I′1)′ ∩ 𝒜(I′1), and by Remark 3.3, AdU(x) = x = αk(x), ∀x ∈ 𝒜(I′1). It follows that VkxΩ = xΩ, ∀x ∈ 𝒜(I′1), and by the Reeh–Schlieder theorem, Vk = id, so k is the identity element in G. Because k ∈ K is arbitary, we have shown that K is the trivial group.
So 𝒜(I1) ∨ ℬ(I2) = 𝒜(I), and
Theorem 4.3. Let 𝒜 be an irreducible conformal precosheaf, and let G be a finite group acting properly on 𝒜. Suppose 𝒜 is completely rational or μ rational as in Definition 4.1. Then there are only a finite number of irreducible covariant representations of 𝒜G and they give rise to a unitary modular category as defined in II.5 of ref. 11 by the construction as given in Section 1.7 of ref. 5.
Proof:
This follows immediately from Proposition 4.2, Correlation 32 of ref. 24 and (2) of Correlation 1.7.3 of ref. 5.
Remark 4.3:
An irreducible covariant representation of 𝒜G is called an untwisted representation if it appears as a summand in the restriction to 𝒜G of a covariant representation of 𝒜. A representation is called twisted if it is not untwisted. By computing the μ indices, we can show that the set of twisted representations of 𝒜G is not empty. This fact is first noticed in ref. 24 under the assumption that 𝒜G is strongly additive. Note that this is very different from the case of cosets (cf. ref. 4, Correlation 3.2, where it was shown that under certain conditions there are no twisted representations for the coset).
5. A Class of Orbifolds
The irreducible conformal precosheaf 𝒜U(1)2l associated with LU(1) at level 2l is studied in Section 3.5 of ref. 5. We have μ𝒜U(1)2l = 2l, and there are exactly 2l irreducible representations of 𝒜U(1)2l which is labeled by integer k, 0 ≤ k ≤ 2l − 1. By identifying ℝ2M = (x, y) → x + iy ∈ ℂM, where x, y are column vectors with M real entries, we have the following natural inclusion LSU(M)1 × LU(1)M ⊂ LSpin(2M)1, where U(1) acts on ℂM as a complex scalar, and the subscripts are the levels of the representations. Also note that we have natural inclusions LSpin(M)2 ⊂ LSU(M)1 ⊂ LSpin(2M)1. Define J := (IdM, −IdM) ∈ SO(2M) and lift it to Spin(2M). Note that for A ∈ SU(M), JAJ = Ā, and JAJ = A if A ∈ Spin(M). From the definition, one can check that AdJ generates a proper ℤ2 action on 𝒜SU(M)1, and by using the branching rules for the inclusion LSpin(M)2 ⊂ LSU(M)1 in Section 4 of ref. 26, we can prove the following:
Lemma 5.1.
Similarly, by using branching rules, we can prove the following:
Lemma 5.2.
One can use Lemmas 5.1, 5.2, Theorem 4.3, and the ideas of Section 4 of ref. 4 to determine the modular matrices as defined in Section 2 for 𝒜Spin(M)2 and the net
associated to the diagonal coset LSpin(M)2 ⊂ LSU(M)1 × LSpin(M)1. Details will appear elsewhere.
To state the results for the net 𝒜Spin(M)1×Spin(M)1/Spin(M)2, recall the set of integrable weights of irreducible positive energy representations of LSpin(2l) at level k is given by:
When l is even, this set admits a ℤ2 × ℤ2 automorphism generated by As, Av, where As, Av are given by:
When l is odd, this set admits a ℤ4 automorphism generated by As, where As is given by
These automorphisms will be called diagram automorphisms. Note that the set of diagram automorphisms is ℤ2 × ℤ2 when l is even and ℤ4 when l is odd.
Theorem 4.6 of ref. 2, which concerns the ring structure of general diagonal cosets of type A, now holds for our diagonal coset, with the action of ℤN there replaced by ℤ2 × ℤ2 when l is even and by ℤ4 when l is odd, because the proof of ref. 2 applies verbatim. As in Section 4.3 of ref. 2, I denote by [Λ̇, Λ̈; Λ] the orbit of (Λ̇, Λ̈; Λ) under the diagonal action of the diagram automorphisms, where (Λ̇, Λ̈) are the integrable weights (both at level 1) of LSpin(M) × LSpin(M). The following is a complete list of irreducible representations of 𝒜Spin(M)1×Spin(M)1/Spin(M)2.
2 |
I have also chosen the notations to make the comparisons with the notations of ref. 12 easy (l corresponds to N on pp. 517 and 518 of ref. 12).
The univalences of the above representations are given by:
By using the remark after Proposition 3.1 of ref. 4, the T matrix can be chosen to be
3 |
where ωx is given as above. One can then determine the S matrix [cf. (2) of Lemma 2.2 in ref. 3] for 𝒜Spin(M)1×Spin(M)1/Spin(M)2. The notations in (7) have been chosen so that the S matrix is given by Table 1.
Table 1.
× S | 1 | j | φlj | φk′ | σj | τj | ||
---|---|---|---|---|---|---|---|---|
1 | 1 | 1 | 1 | 2 | ||||
j | 1 | 1 | 1 | 2 | − | − | ||
φli | 1 | 1 | (−1)l | 2(−1)k′ | bij | bij | ||
φk | 2 | 2 | 2(−1)k | 4 cos | 0 | 0 | ||
σi | − | bij | 0 | aij | −aij | |||
τi | − | bij | 0 | −aij | aij |
Here aij = (1 + (2δi,j − 1)exp(−(πil/2))), bij = (−1)l+δi,j exp (πil/2), δi,j is the usual Delta function, and 1 ≤ i, j ≤ 2.
Note that the ℤ2 action on 𝒜U(1)2l given by AdJ as defined before Lemma 5.1 is a proper action on 𝒜U(1)2l. The reader familiar with ref. 8 may notice that the action given by AdJ corresponds to −1 isometry of rank one lattice vertex operator algebras (cf. refs. 8 and 10). By using the branching rules in Section 4 of ref. 26, we can show:
Lemma 5.3.
By Lemma 5.3 and Theorem 4.3, we have proved the following theorem:
Theorem 5.4. All irreducible representations of 𝒜U(1)2lℤ2 are given by (2) for l ≥ 3. These irreducible representations give rise to a unitary modular category whose S and T matrices are given by Table 1 and (3).
When l = 2, Spin(4) = SU(2) × SU(2), one checks that Theorem 5.4 still holds in this case, where the integrable weights of LSpin(4) should be replaced by the integrable weights of LSU(2) × LSU(2). When l = 1, by using the fact that 𝒜U(1)2 ≃ 𝒜SU(2)1, one can check that
and 𝒜U(1)8 has already been studied in Section 3.5 of ref. 5.
Remark 5.5:
The reader may wonder why we identify the orbifold of 𝒜U(1)2l with respect to the natural ℤ2 action as a coset in Lemma 5.3 instead of considering such an orbifold directly as in ref. 10. The reason is that the “twisted representations” in ref. 10 are defined only algebraically. To show that these “twisted representations” in ref. 10 give rise to covariant representations of our orbifold, one needs to study the analytical properties of the twisted vertex operators in ref. 10, which are not trivial if one tries this directly. On the other hand, there are no such problems for cosets (cf. Remark 4.3). As noted before in Lemma 5.3, the ℤ2 action on 𝒜U(1)2l given by AdJ corresponds to the −1 isometry of rank one lattice vertex operator algebras (cf. refs. 8 and 10). The classification of irreducible representations of the orbifold rank one lattice vertex operator algebras is given in ref. 10, which corresponds to the first part of Theorem 5.4. We note that the S matrix can be identified with the S matrix on Proofs 517 and 518 of ref. 12. However, there are mistakes in the S matrix on Proofs 517 and 518 of ref. 12 corresponding to the entries of aij, bij in Table 1. Table 1 gives the correct S matrix.
6. More Examples and Questions
The lattice vertex operator algebras and their automorphism groups provide a rich source of examples of orbifolds (cf. refs. 8, 10, and 27). I have determined the S, T matrices and hence the fusion rules for the orbifold of rank 1 lattice Vertex Operator Algebras in Theorem 5.4. It would be interesting to generalize this theorem to higher-rank cases.
Let 𝒜SU(N)k be the irreducible conformal precosheaf, and let G be a finite subgroup of SU(N). Then there is a natural action of G on 𝒜SU(N)k, and it is easy to check that this action is proper (if the action of G is not faithful, one can replace G by a quotient G′ as explained in the footnote of Definition 3.1). By the results of refs. 28 and 23, 𝒜SU(N)k is completely rational. Hence Theorem 4.3 applies in this case, and we have a family of unitary modular categories. It would be interesting to study these modular categories in general.
Finally, let me mention that permutation orbifolds (cf. ref. 29 and refs. therein) provide another interesting class of orbifolds. Let us formulate these orbifolds in our setting. Let 𝒜 be an irreducible conformal precosheaf. Then the tensor product of 𝒜 with itself n times 𝒜⊗n := 𝒜 ⊗ 𝒜 ⊗ … ⊗ 𝒜 is also an irreducible conformal precosheaf. Let G ⊂ Sn be a finite subgroup of Sn, the permutation group on n letters. Note that any finite group is embedded in a permutation group by Cayley's theorem. There is an obvious action of G on 𝒜⊗n by permuting the n tensors, and one checks directly by definitions that this action of G on 𝒜⊗n is proper as defined in Section 3. Note that if 𝒜 is μ rational, so is 𝒜⊗n by definition. If 𝒜 is μ rational, by Theorem 2.6, we obtain a large family of unitary modular categories from the orbifold (𝒜⊗n)G. The modular matrices of the unitary modular categories associated with the permutation orbifolds above have been written down on the basis of heuristic physics arguments in ref. 30. It would be interesting to do the computations in our framework as in Section 4 and compare them with the results of ref. 30.
Acknowledgments
This work is partially supported by National Science Foundation Grant DMS-9820935.
Footnotes
This paper was submitted directly (Track II) to the PNAS office.
Article published online before print: Proc. Natl. Acad. Sci. USA, 10.1073/pnas.260375597.
Article and publication date are at www.pnas.org/cgi/doi/10.1073/pnas.260375597
If V : G → U(ℋ) is not faithful, we can take G′ := G/kerV and consider G′ instead.
References
- 1.Moore G, Seiberg N. Lett Phys B. 1989;220:422–430. [Google Scholar]
- 2.Xu F. Commun Math Phys. 2000;211:1–43. [Google Scholar]
- 3.Xu F. Publ Res Inst Math Sci (Kyoto Univ) 1999;35:795–824. [Google Scholar]
- 4.Xu F. On a Conjecture of Kac-Wakimoto. 1999. , preprint math.RT/9904098. [Google Scholar]
- 5.Xu F. 3-Manifold Invariants from Cosets. 1999. , preprint math.GT/9907077. [Google Scholar]
- 6.Jones V F R. Inv Math. 1983;72:1–25. [Google Scholar]
- 7.Borcherds R. Proc Natl Acad Sci USA. 1986;83:3068–3071. doi: 10.1073/pnas.83.10.3068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Frenkel I B, Lepowsky J, Ries J. Vertex Operator Algebras and the Monster. New York: Academic; 1988. [Google Scholar]
- 9.Dong C, Mason G. Duke Math J. 1997;86:305–321. [Google Scholar]
- 10.Dong C, Nagatomo K. Commun Math Phys. 1999;202:169–195. [Google Scholar]
- 11.Turaev V G. Quantum Invariants of Knots and 3-Manifolds. Berlin: de Gruyter; 1994. [Google Scholar]
- 12.Dijkgraaf R, Vafa C, Verlinde E, Verlinde H. Commun Math Phys. 1989;123:429–487. [Google Scholar]
- 13.Guido D, Longo R. Commun Math Phys. 1996;181:11–35. [Google Scholar]
- 14.Fröhlich J, Gabbiani F. Commun Math Phys. 1993;155:569–640. [Google Scholar]
- 15.Pressley A, Segal G. Loop Groups. Oxford: Oxford Univ. Press; 1986. [Google Scholar]
- 16.Kac V G. Infinite Dimensional Lie Algebras. 3rd Ed. Cambridge, U.K.: Cambridge Univ. Press; 1990. [Google Scholar]
- 17.Rehren K-H. The Algebraic Theory of Superselection Sectors. Teaneck, NJ: World Scientific; 1990. [Google Scholar]
- 18.Guido D, Longo R. Commun Math Phys. 1992;148:521–551. [Google Scholar]
- 19.Longo R, Rehren K-H. Rev Math Phys. 1995;7:567–597. [Google Scholar]
- 20.Xu F. Commun Math Phys. 1998;192:349–403. [Google Scholar]
- 21.Böckenhauer J, Evans D E. Commun Math Phys. 1998;197:361–386. [Google Scholar]
- 22.Böckenhauer J, Evans D E. Commun Math Phys. 1999;200:57–103. [Google Scholar]
- 23.Xu F. Commun Contemp Math. 2000;2:307–347. [Google Scholar]
- 24.Kawahigashi Y, Longo R, Müger M. Multi-Interval Subfactors and Modularity of Representations in Conformal Field Theory. 1999. , preprint math.OA/9903104. [Google Scholar]
- 25.Izumi M, Longo R, Popa S. J Funct Anal. 1998;155:25–63. [Google Scholar]
- 26.Kac V G, Wakimoto M. Adv Math. 1988;70:156–234. [Google Scholar]
- 27.Dong C, Griess R, Ryba A. J Algebra. 1998;217:701–710. [Google Scholar]
- 28.Wassermann A. Inv Math. 1998;133:467–539. [Google Scholar]
- 29.Barron K, Dong C, Mason G. Twisted Sectors for Tensor Product Vertex Operator Algebras Associated to Permutation Groups. 1999. , preprint math. QA/9803118. [Google Scholar]
- 30.Bantay P. Permutation Orbifolds. 1998. , preprint hep-th/9910079. [Google Scholar]