We study the S3-orbifold of a rank three Heisenberg vertex algebra in terms of generators and relations. By using invariant theory, we prove that the orbifold algebra has a minimal strong generating set of vectors whose conformal weights are 1, 2, 3, 4, 5, 62 (two generators of degree 6). The structure of the cyclic Z3-orbifold is determined by similar methods. We also study characters of modules for the orbifold algebra.

For every vertex operator algebra V, the n-fold tensor product Vn has a natural vertex operator algebra structure. The symmetric group Sn acts on V(n) ≔ Vn by permuting tensor factors and thus Sn ⊂ Aut(V(n)). Denote by V(n)Sn the fixed point vertex operator subalgebra also called the Sn-orbifold of V(n). It is an open problem to classify irreducible modules of V(n)Sn although it is widely believed that every such module should come from a g-twisted Vn-module for some gSn. When it comes to the inner structure of V(n)Sn (e.g., a strong system of generators), very little is known. Even for the Heisenberg orbifold H(n)Sn, where H is the rank one Heisenberg vertex algebra, this problem seems quite difficult.

Finite and permutation orbifolds have been extensively studied both in physics and mathematics literature. According to Ref. 5, the earliest study of permutation orbifolds seems to be Ref. 17. The first systematic construction of cyclic orbifolds, including their twisted sectors appeared in Ref. 8. Explicit formulas for characters and modular transformation properties of permutation orbifolds of rational conformal field theories were given in Ref. 4.

In the literature on vertex algebras, the main focus has been on the classification and construction of twisted modules starting with Refs. 15 and 19. Further developments include “Quantum Galois” theory developed in Ref. 11, work on twisted sectors of permutation orbifolds6 (see also Ref. 7), and tensor category structure for general G-orbifolds of rational vertex algebras13,18. There are also numerous papers on orbifold vertex algebras for Abelian groups and groups of small order. Structure and representations of 2-permutation orbifolds were subjects of Refs. 1, 2, and 12; see also a more recent work on 3-permutation orbifolds of lattice vertex algebras.14 The ADE orbifolds of rank one lattice vertex algebras (e.g., Ref. 10) are important for classification of c = 1 rational vertex algebras. Orbifolds of irrational C2-cofinite vertex algebras have been investigated in Refs. 3, 9, and 2. Linshaw (see Refs. (9) and (20)–(22)) extensively studied orbifolds of “free field” vertex operator (super)algebras using the classical invariant theory26 and its (super) extensions. As a consequence of the main result in Ref. 22, every finite orbifold of an affine vertex algebra is finitely strongly generated. In particular, this implies that H(n)Sn has a finite strong set of generators. Characters of orbifolds of affine vertex algebras were investigated earlier in Ref. 16.

In this paper, we are concerned with the structure of one of the simplest non-Abelian orbifolds, H(3)S3, where H is the rank one Heisenberg vertex algebra [case H(2)S2 is well-understood12,1]. We prove three main results. The first result pertains to generators of H(3)S3. We show in Theorem 4.1 that this vertex algebra is isomorphic to a W-algebra of type (2, 3, 4, 5, 62) tensored with a rank one Heisenberg vertex algebra. Here, labels 2, 3, 4, 5, 62 indicate that our W-algebra is strongly generated by the Virasoro vector (of degree two) and five primary vectors of degrees 3, 4, 5 and two of degree 6. These five generators are explicitly given in Sec. IV where we denoted them by J1, J2, C1, C2, and C3. Our second result is about the cyclic Z3-orbifold of H(3) (see Theorem 5.1 for details).

In Secs. VI and VII, we discuss characters of certain H(3)S3-modules and their modular properties. It is expected that many irrational vertex algebras will enjoy modular invariance in a generalized sense involving iterated integrals instead of sums. For the rank n permutation orbifold H(n)Sn, the character of a module M is expected to transform as

(1.1)

where SM,MλiC and λiRi, 1 ≤ in, parametrize certain H(n)Sn-modules. Our third main result gives strong evidence for this conjecture for n = 3 (see Theorem 7.1).

Let H denote the rank one Heisenberg vertex operator algebra generated by α(1)1, with the usual conformal vector (and grading) given by ω=12α2(1)1. Let H(n)=Hn. For convenience, we suppress the tensor product symbol and let α1(1)1α(1)11H(n), and similarly, we define αi(1)1, i ≥ 2, such that H(n)=α1(1)1,,αn(1)1. We consider the natural action of Sn on H(n) given by

(2.1)

for 1 ≤ ijn, mj < 0, and σSn.

We clearly have a natural linear isomorphism

(2.2)

induced by αi(−m − 1) ↦ xi(m) for m ≥ 0. Using the terminology of Ref. 21, we say that C[xi(m)|1in,m0] is the associated graded algebra of the vertex algebra H(n). Further, we may endow the polynomial algebra C[xi(m)|1in,m0] with the structure of a -ring by defining the map

(2.3)

where the action of is extended to the whole space via the Leibniz rule. This definition of is compatible with the translation operator in H(n) given by T(v)=v21.

The following lemma is from Ref. 20.

Lemma 2.1.

LetAbe a vertex algebra with aZ0filtration, whereÃis the associated ∂-ring. If {ãi|iI} generatesÃ, then {ai|iI} strongly generatesA, where aiand ãiare related via the natural linear isomorphism described by theZ0filtration.

Further, we recall that the invariant ring C[x1,,xn]Sn has a variety of generating sets, including the power sum polynomials
In addition to this, it is common to study the invariant theory of the ring of infinitely many commuting copies of this polynomial algebra, where we denote by xi(m) the copy of xi from the mth copy of the polynomial algebra. A theorem of Weyl26 shows that C[xi(m)|1in,m0]Sn is generated by the polarizations of these polynomials
(2.4)
for 1 ≤ kn. Now, applying Lemma 2.1, we have an initial strong generating set for the orbifold H(n)Sn given by the vectors
(2.5)
for 1 ≤ kn and mj ≥ 0. It should be noted that the conformal vector of H(n) is 12ω2(0,0), making the orbifold a vertex operator algebra.

Here, we describe the structure of H(2)S2. This case is well-known, so we only provide a few details. Denote by M(1)+H(1)Z2 the fixed point subalgebra under the action α(1)1α(1)1, where α is the Heisenberg generator [in the physics literature, M(1)+ is often denoted by W(2, 4)]. This vertex algebra was thoroughly studied in Ref. 12 and elsewhere. Let

Observe that as vertex algebras

where H(1)h=h and H(1)h=h with conformal vector ω. The nontrivial element of the group S2 fixes the first tensor factor and

Thus, we immediately get

(3.1)

It is easy to see that

is contained in ⟨h⟩ and is primary of conformal weight 4 and thus a generator of M(1)+.

As described above, we know that the orbifold H(3)S3 will be strongly generated by the vectors

(4.1)

The following change of basis of the generating set will allow for an efficient reduction in the generating set of the corresponding orbifold

(4.2)

where η is a primitive third root of unity. Using this generating set, we have σβ1(1)1=β1(1)1 for all σS3. Further, examining the action of the generators of S3 on the generators of H(3), we have

(4.3)

and

(4.4)

From this action, we see that an initial generating set for the orbifold H(3)S3 may be taken to be

(4.5)

for 0 ≤ abc. In fact, we can explicitly write the relation between our original generators (4.1) and our new generators (4.5) as follows:

(4.6)

We may make a similar change of variables for the generating set of C[xi(m)|1i3,m0] by setting

(4.7)

for mi ≥ 0, and

(4.8)

for a, b, c ≥ 0.

Using the translation operator, T, restricted to the orbifold H(3)S3, the initial strong generating set (4.5) can be reduced per the following lemma:

Lemma 4.1.
The orbifoldH(3)S3is strongly generated by the vectors
(4.9)

Proof.
It is clear that since
(4.10)
the linear portion of the generating set (4.5) can be immediately minimized to contain the single weight one element ω1(0).
Moving on toward the quadratic terms in (4.5), we consider the vector spaces
(4.11)
and a natural family of subspaces
(4.12)
Observe that
(4.13)
where βi(z)=Y(βi(1)1,z). Also notice that for all a, b ≥ 0, we have ω20(a,b)=ω20(b,a). We may take an initial basis for A2(m) to be the set
(4.14)
which implies that
(4.15)
Further, the set
(4.16)
is clearly a basis for ∂A2(m). It follows that
(4.17)
and thus by induction, we may take a more convenient basis of A2(m) to be
(4.18)
Thus, a more efficient set of quadratic generators for the H(3)S3 may be taken to be ω20(0,2a) for a ≥ 0.
Finally, we consider the cubic generators. Analogous to (4.13), we have
(4.19)
which may be used inductively to write all cubic generators in terms of those of the form ω30(0,a,b) with 0 ≤ ab. Throughout, we use the fact that ω30(a,b,c) is invariant under any permutation of the entries a, b, c.

Remark 4.1.
Using (4.6), we may take
(4.20)
as our strong generating set. We take advantage of this translation tool as generators of this form are somewhat more natural to the parent algebra, H(3).

Remark 4.2.
This simplification of the generating set also holds if we consider the -ring associated with H(3)S3, which has un-reduced generators given by (2.4). We may reduce these to a minimal generating set for C[xi(m)|1i3,m0]S3 given by
(4.21)

We now present the following relations among the generators of C[xi(m)|1i3,m0]S3.

Lemma 4.2.
Fora = (a1, a2, a3, a4, a5, a6) with a1, a2, a3, a4, a5, a6 ≥ 0, we have
(4.22)
(4.23)
and
(4.24)
fora = (a1, a2, a3, a4, a5) with a1, a2, a3, a4, a5 ≥ 0.

Remark 4.3.

These relations play the role of the determinant (and similar) relations in the invariant theory of classical Lie groups. We do not claim that these two families of expressions generate all relations in this case. In fact, there are such relations at every degree, and since the Frobenius number of 5 and 6 is 19, there are new relations at least up to this degree.

We identify the relations from Lemma 4.2 via the isomorphism (2.2) with expressions involving the generators of the orbifold H(3)S3. For a = (a1, a2, a3, a4, a5, a6) with a1, a2, a3, a4, a5, a6 ≥ 0, set
(4.25)
(4.26)
and
(4.27)
for a = (a1, a2, a3, a4, a5) with a1, a2, a3, a4, a5 ≥ 0.
Observe that for i ∈ {1, 2}, the weight of the expression D6i(a) is |a| + 6, while the weight of D5(a) is |a| + 5, where we take |(u1, …, un)| = u1 + ⋯ + un for any multi-index. By Lemma 4.2 and repeated applications of the weak associativity properties of vertex algebras, along with our linear isomorphism (2.2), we see that for i ∈ {1, 2}, we may rewrite
(4.28)
where
(4.29)
and
(4.30)
where μa,b,c,d(4) and μa,b(2) are appropriate constants. Importantly, this expansion has no terms that are “cubic” in the generators ω20(a,b) or “quadratic” in the generators ω30(a,b,c), each of which would contain a combination of six of the original βi vectors with i ∈ {2, 3}. We have a similar (and simpler) decomposition
(4.31)
where
(4.32)
where μa,b,c(3) are constants. Again, Lemma 4.2, weak associativity, and (2.2) imply that this expansion of will not contain terms that are “products” of the generators ω20(a,b) and ω30(a,b,c) or otherwise a combination of five of the original vectors β2 and β3.

Now we are poised to use the expressions D6i(a) and D5(a) to further reduce the generating set of H(3)S3 described in Lemma 4.1.

Lemma 4.3.

The full list of quadratic generators described in Lemma 4.1 can be replaced with the set{ω20(0,0),ω20(0,2),ω20(0,4)}.

Proof.
Theorem 4.7 of Ref. 21 implies that, using only combination of quadratic generators, we may immediately reduce our quadratic generators to the set
(4.33)
In order to remove the remaining two quadratic generators, we use decoupling relations that involve both quadratic and cubic generators. Using the expression D61(0,0,0,0,1,1), the decomposition described in (4.28)–(4.30) may be used to construct the following equation:
(4.34)
and thus, we may remove ω20(0,6) from (4.33). Using (4.34) and D61(0,0,0,0,2,2), we can similarly write ω20(0,8) as a vertex algebraic polynomial using the quadratic generators ω20(0,0), ω20(0,2), ω20(0,4), along with cubic generators ω30(0,0,0), ω30(0,0,1), ω30(0,1,1), ω30(0,0,2), and ω30(0,2,2).

Lemma 4.4.
The full list of cubic generators described in Lemma 4.1 can be replaced with the set
(4.35)
Moreover, these generators together withω10(0),ω20(0,0),ω20(0,2),ω20(0,4)form a strong generating set ofH(3)S3.

Proof.
To get started, we notice that
(4.36)
and in other words, all cubic generators of conformal weight 6 or less may be written in terms of the generators {ω30(0,0,0),ω30(0,0,2),ω30(0,1,2)} using only the translation operator. Furthermore, we have
(4.37)
that is, all vectors of the form ω30(0,a,b) with 0 ≤ ab ≤ 4 can be written, using only the translation operator, in terms of our proposed generating set (4.35) with the addition of the vectors ω30(0,0,4), ω30(0,1,4), ω30(0,2,4), and ω30(0,c,d) with 0 ≤ cd where d ≥ 5. The remainder of our argument will be concerned with eliminating the need for these additional vectors.
Using D5(0, 0, 0, 1, 1) and the decomposition described in (4.31) and (4.32), we have
(4.38)
where again this calculation was performed using Ref. 25. Using D5(0, 0, 1, 1, 1), a similar decoupling equation for ω30(0,1,4) can be found. Furthermore, a linear combination of D5(0, 0, 0, 1, 3) and D5(0, 0, 1, 1, 2) leads to the decoupling equation
(4.39)
Next, for aN, each of the six expressions D5(0, 0, 0, 1, a − 3), D5(0, 0, 0, 2, a − 4), D5(0, 0, 1, 1, a − 4), D5(0, 0, 0, 3, a − 5), D5(0, 0, 1, 2, a − 5), and D5(0, 1, 1, 1, a − 5) can be expanded as linear combinations of the six vectors ω30(0,0,a), ω30(0,1,a1), ω30(0,2,a2), ω30(0,3,a3), ω30(0,4,a4), and ω30(0,5,a5) along with terms of the form ω3(0,m1,m2)1n1 where m1 + m2 + n = a with n ≥ 1. The linear independence of these can be determined by considering the determinant of the matrix, A, whose (i, j) entry is the coefficient of the ith vector in the jth expression as described above. A lengthy elementary calculation shows that, in the case that a is even
(4.40)
and a similar expression exists for odd a. In particular, for a ≥ 5, we can write each of the vectors ω30(0,0,a), ω30(0,1,a1), ω30(0,2,a2), ω30(0,3,a3), ω30(0,4,a4), and ω30(0,5,a5) as a vertex algebraic polynomial in terms of lower weight terms.
Next, for 5 ≤ a < b, D5(b + 2, a − 1, 0, 0, 0) can be used to construct
(4.41)
where Ψ is a vertex algebraic polynomial with terms of the form ω30(0,m1,m2)1n1 and ω20(r1,r2)1ω30(0,s1,s2) where m1 + m2 + n = a + b + 1 and r2 + r2 + s1 + s2 = a + b − 1 with n ≥ 1.

Finally, (4.38)–(4.41) provide us with a clear inductive path to eliminating all cubic generators other than ω30(0,0,0),ω30(0,0,2), and ω30(0,1,2). This proves the first assertion. To prove the second claim, we use Lemma 4.3, where we reduced all quadratic generators down to ω20(0,0),ω20(0,2),ω20(0,4). Although in this reduction, we additionally used cubic generators ω30(0,0,1), ω30(0,1,1), and ω30(0,2,2), we can remove them using the first two relations in (4.36), the second relation in (4.37), and relation (4.38).

Theorem 4.1.

  • The vertex operator algebraH(3)S3is simple of type (1, 2, 3, 4, 5, 62), i.e., it is strongly generated by seven vectors whose conformal weights are: 1, 2, 3, 4, 5, 6, 6. This generating set is minimal.

  • H(3)S3is isomorphic toH(1)W, where W is of type (2, 3, 4, 5, 62).

  • H(3)S3is not freely generated (by any set of generators).

Proof.
Clearly, by a result from Ref. 11, this vertex algebra is simple. By Lemma 4.4 and earlier discussion, we see that H(3)S3 is strongly generated by vectors of conformal weights: 1 (linear generator), 2, 4, 6 (quadratic generators) and 3, 5, 6 (cubic generators). From the character formula (see Proposition 6.1), we see that
where (a; q)i≥0(1 − aqi). An easy analysis shows that dropping one (or more) generators from this generating set would imply that certain graded dimensions of H(3)S3 are strictly bigger than the corresponding graded dimension for the smaller subalgebra. Thus, the proposed set of generators must be a minimal generating set. Clearly, this vertex algebra is not freely generated by these generators due to
which implies that there must be a nontrivial relation at degree 9. In fact, one can quickly argue that this is impossible simply from the fact that ch[H(3)S3] is modular, which is impossible to achieve with a free generating set.

For (ii), we first observe that ⟨ω1(0)⟩ is isomorphic to H(1). By taking the commutant WComm(H(1),H(3)S3), we get H(1)WH(3)S3. Each generator of H(3)S3, except ω1(0), can be written as a linear combination of elements in H(1)W with a nonzero component in W [otherwise, it would imply that some of the generators are contained in ⟨ω1(0)⟩, contradicting the minimality in (i)]. These nonzero components form a generating set of W.

Using (4.6), we may take the original invariants ω1(0), ω2(0, 0), ω2(0, 2), ω2(0, 4), ω3(0, 0, 0), ω3(0, 0, 2), and ω3(0, 1, 2) as our minimal strong generating set. Furthermore, the following change of variables allows us to express the orbifold in terms of primary generators
(4.42)

We now consider the orbifold of H(3) under the action of the cyclic subgroup Z3123S3. Using a similar strategy to Subsection IV A, we may take

(5.1)

as our initial generating set. Analogous to Lemma 4.1, we have the following initial reduction of the generating set:

Lemma 5.1.
The orbifoldH(3)Z3is strongly generated by vectors
(5.2)

Proof.

The linear generator is the same vector found in Lemma 4.1 which may be used to produce all vectors of the form ω10(a) for a ≥ 0 as before.

The quadratic generators will be reduced using an argument similar to the proof of Lemma 4.1, where the main difference is due to the fact that for ab, ω2,30(a,b)ω2,30(b,a). We begin by setting
(5.3)
and notice that for all m ≥ 0, ∂A(m) is a subspace of A(m + 1) of co-dimension 1. In fact, we have
(5.4)
from which it follows that the set
(5.5)
is a basis of A(m). Our result follows.
Again, the reduction of the cubic generators follows from an analogue of (4.13). In particular,
(5.6)
for i ∈ {1, 2}, may be used inductively to achieve the desired set, where we use the fact that ωi,i,i0(a,b,c) is fixed under any permutation of the entries a, b, c.

We now present analogues of Lemmas 4.3 and 4.4 in this setting.

Lemma 5.2.
The full list of quadratic generators described in (5.2) can be replaced with
(5.7)

Proof.
Our main tool for this argument will be the expression
(5.8)
which may be expanded to
(5.9)
An elementary calculation allows us to write
(5.10)
and more generally,
(5.11)
Now combining (5.9)–(5.11), we have
(5.12)
which, for a = 0 and a ≥ 2, may be used to solve for ω2,30(0,a+4) in terms of vectors of lower conformal weight.
Inductively, together with the special case
(5.13)
this allows us to remove all but the necessary generators.

Lemma 5.3.
The full list of cubic generators described in (5.2) can be replaced with
(5.14)

Proof.
We follow a strategy similar to the proof of Lemma 4.4. In this case, our argument is greatly simplified due to the fact that our decoupling relations are simpler and occur at a lower initial conformal weight. We focus on the ω2,2,20(0,a,b) terms, as the ω3,3,30(0,a,b) are similar, starting with the observation that
(5.15)
meaning that, by using the translation operator, all cubic generators of weight five or less may be written in terms of (5.14).
The expression
(5.16)
will be our main tool moving forward. A special case of (5.16) can be used to construct the equation
(5.17)
which can be used inductively to remove all generators of the form ω2,2,23(0,0,a) for a ≥ 3 from the generating set. Furthermore, a more general version yields
(5.18)
which can be used to remove all generators of the form ω2,2,23(0,a,b), where 2 ≤ ab. This leaves only the generators of the form ω2,2,20(0,1,a) which can be removed using
(5.19)

Theorem 5.1.

  • The vertex operator algebraH(3)Z3is simple of type (1, 2, 33, 4, 53), i.e., it is strongly generated by seven vectors whose conformal weights are 1, 2, 3, 3, 3, 4, 5, 5, 5. This generating set is minimal.

  • H(3)Z3is isomorphic toH(1)W, where W is of type (2, 33, 4, 53).

  • H(3)Z3is not freely generated (by any set of generators).

Proof.

This proof is similar to the proof of Theorem 4.1. In particular, we use the fact that the free W-algebra on a generating set containing the same weight elements with any one removed has certain graded dimensions strictly less than those described by the character of H(3)Z3, (6.1).

Finally, the orbifold H(3)Z3 is described in terms of primary generators with the following vectors:
(5.20)

We first derive a general formula for the character of the Sn-orbifold of a vertex algebra. Although formulas of this form have appeared in the literature (e.g., Ref. 4 or Ref. 16), we include the proof for completeness.

Theorem 6.1.
Let V be a VOA andfSn(q)be the character of the Sn-fixed point subalgebraV(n)SnVn (under the usual action). Then
where g acts by permutation of tensor factors.

Proof.
Consider the group algebra C[Sn] and the idempotent
where |Sn| = n!. The space Im(e) is precisely V(n)Sn. Since the eigenvalues of e are 1 and 0 (and 1 for the Sn-fixed subalgebra), the character can be computed simply by taking the trace of e

Now we specialize to n = 3 and V=H(1).

Proposition 6.1.

Proof.
It suffices to consider the following elements: g = 1, g = (12), and g = (123). By the previous theorem, we have
Clearly,
For tr(12), we compute the trace on the following basis of H(3)
where P is the set of partitions, and for pλi, we take obvious monomials of Heisenberg elements. The matrix representation of (12)End(H(3)) in this basis has a non-zero entry on the diagonal (=1) if and only if the corresponding basis element is
Since all pλ are generated by α(−i), the vectors contributing to the trace are generated by α(−i) ⊗ α(−i) ⊗ α(−j). Thus,
One similarly argues for 3-cycles, so we get

Corollary 6.1.

By a result from Ref. 11 (especially Sec. VII devoted to dihedral groups), we have a decomposition of H(3) in terms of H(3)S3-modules

(7.1)

where V(2) is the 2-dimensional standard representation of S3, and H(3)S3,sgn and H(3)S3,st are irreducible H(3)S3-modules.

Lemma 7.1.
We have

Proof.
The character of the sign module11 
is computed along the lines of Proposition 6.1 with the difference that in the sign representation 2-cycles act as −1. This sign contributes with a negative sign to the middle terms of the formula.

The second relation follows directly from the first formula, Proposition 6.1, and decomposition (7.1).

Next, we construct several H(3)S3-modules coming from ordinary and g-twisted H(3)-modules, gS3.

We first look at H(3)S3-modules coming from irreducible H(3)-modules (Fock representations). Let w=(w1,w2,w3)C3. We denote by Fw1,w2,w3 the irreducible H(3)-module with highest weight w such that αi(0) acts as multiplication by wi. Similarly, Fw denotes the Fock representation for H(1) with highest weight w.

Interestingly, Fw1,w2,w3 is generically irreducible as an H(3)S3-module.

Lemma 7.2.

Letw1,w2,w3C3such that wiwj, ij. Then theH(3)-moduleFw1,w2,w3is irreducible as anH(3)S3-module.

Proof.
Let V be a simple VOA, g be an automorphism of V of prime order p, and M be a simple V-module such that gM is not isomorphic to M as V-modules. Then according to Theorem 6.1,11M is a simple Vg-module. To apply this result, we first observe embeddings
Because of w1w2, (12)Fw1,w2,w3Fw1,w2,w3 as H(3)-modules. Indeed, for (12)Fw1,w2,w3, actions of α1(0) and α2(0) are switched, so (12)Fw1,w2,w3Fw2,w1,w3. Therefore, Fw1,w2,w3 is irreducible as a G(3)H(3)(<12>)-module. Similarly, we consider (123)Fw1,w2,w3 and Fw1,w2,w3 as G(3)-modules. Because of w1w3 and w3w2, they are not isomorphic as G(3)-modules, and thus Fw1,w2,w3 is irreducible as a G(3)<(123)> -module. The proof follows.
In the study of characters, we require modular variables so we let q = e2πiτ, where τH (the upper half-plane). Let
the Dedekind η-function. Clearly, for every w,
(7.2)

Next we consider g-twisted H(3)-modules, so g is either a 2- or 3-cycle. We may assume, without loss of generality, that g is either (12) or (123).

We first construct a family of θ-twisted H(3)-modules, where θ ≔ (12) switches the first two “coordinates.” By using the isomorphism in Sec. III, H(2)<(12)>H(1)M(1)+, for every w1,w3C, we get a θ-twisted H(3)-module
where M(1)(θ)C[α(1/2),α(3/2),.] is a Z2-twisted H(1)-module studied in Ref. 12. This module is of conformal weight 116 and its character is easily computed. We quickly get
(7.3)
Further, we construct a family of (123)-twisted H(3)-modules. Let σ = (123). We first construct a σ-twisted Heisenberg algebra h^σ as in Ref. 19. It is isomorphic to C[t,t1]t1/3C[t,t1]t2/3C[t,t1]Ck as a graded Lie algebra. Then for every wC, we have an associated σ-twisted H(3)-representation Fw(σ)U(h<0σ). Moreover,
where h=14p2i=1p1i(pi)ri, and p is the order of σ and ri (1 ≤ i ≤ 3) are dimensions of the eigenspaces of σ. Plugging in p = 3 and ri = 1 into the last formula, we obtain h=19 and thus
(7.4)

Here, we do not pursue decomposition and irreducibility of Fw1,w3(σ) and Fw(σ) as H(3)S3-modules. This will be a subject of Ref. 24.

In this part, we obtain a modularity result for the character of H(3)S3 under the S-transformation, τ1τ. This was briefly discussed in the Introduction.

We make use of a well-known modular relation for the Dedekind η-function

(7.5)

Let us also recall a higher dimensional Gauss’ integral formula

From the Gauss formula and (7.5), we immediately get the following relations.

Lemma 7.3.

Notice that only the numerator of the integrand depends on w. The denominator is placed inside the integral so that it matches the shape of (1.1).

In Proposition 6.1, the character formulas for Fw1,w2,w3, Fw1,w3(θ), and Fw(σ), that is, formulas (7.2)–(7.4), together with Lemma 7.3 imply the following theorem:

Theorem 7.1.

The character ofH(3)S3has a modular invariance property, in the sense of (1.1).

Let V be a vertex operator algebra. For a V-module M, we define its quantum dimension

In a situation where V is rational (in the strongest sense), this limit always exists and is nonzero (see Ref. 13). But for general V, this limit can be zero or even +.

As our characters can be expressed in terms of η(τ), we will make use of an asymptotic formula

(7.6)

Proposition 7.1.

AsH(3)S3-modules, we have

  • For any Fock representationFw1,w2,w3,

  • (b)

    For representations appearing insideH(3),

  • (c)

    Fw1,w3(θ)and Fw(σ) have quantum dimension +∞.

Proof.

We first observe that in the formulas for ch[H(3)S3](τ), ch[H(3)S3,sgn](τ), and ch[H(3)S3,st](τ), the infinite product q1/8(q;q)3=1η(τ)3, multiplied with a constant, dominates the asymptotics t → 0+. Therefore in order to compute quantum dimensions for these modules, we only have to compute the ratio of these constants. This implies assertions (a) and (b).

For (c), we need more precise asymptotic behaviors. We have
From these formulas, we get
so both ratios have growing terms at t = 0. The proof follows.

We finish with a conjecture which will be addressed in Ref. 24.

Conjecture 1.
Every irreducible (ordinary) H(3)S3-module M appears in the decomposition of a g-twistedH(3)-module, for gS3. Moreover,

Remark 8.1.

  1. In order to confirm Conjecture 1, we first have to describe the Zhu algebra of H(3)S3. This will be addressed in Ref. 24.

  2. In our very recent work (Ref. 23), we determine strong minimal generating sets for the permutation orbifold F(3)S3, where F is the free fermion vertex algebra, and for SF(3)S3, where SF is the rank one symplectic fermion vertex superalgebra.9 We prove that these vertex algebras are of type 12,2,4,92 and 12, 2, 33, 43, 55, 64, respectively.

We thank Drazen Adamovic, Thomas Creutzig, and especially Andrew Linshaw for useful discussions. We also thank the referee for constructive criticism and other valuable comments. Several computations in this paper are performed by the OPE package (Ref. 25) for Mathematica.

The first author was partially supported by the NSF Grant No. DMS-1601070.

1.
T.
Abe
, “
Fusion rules for the free bosonic orbifold vertex operator algebra
,”
J. Algebra
229
,
333
374
(
2000
).
2.
T.
Abe
, “
C2-cofiniteness of 2-cyclic permutation orbifold models
,”
Commun. Math. Phys.
317
,
425
445
(
2013
).
3.
D.
Adamovic
,
X.
Lin
, and
A.
Milas
, “
ADE subalgebras of the triplet vertex algebra: A-type
,”
Commun. Contemp. Math.
15
,
1350028
(
2013
).
4.
P.
Bantay
, “
Characters and modular properties of permutation orbifolds
,”
Phys. Lett. B
419
,
175
178
(
1998
).
5.
P.
Bantay
, “
Permutation orbifolds
,”
Nucl. Phys. B
633
,
365
378
(
2002
).
6.
K.
Barron
,
C.
Dong
, and
G.
Mason
, “
Twisted sectors for tensor product vertex operator algebras associated to permutation groups
,”
Commun. Math. Phys.
227
,
349
384
(
2002
).
7.
K.
Barron
,
Y.-Z.
Huang
, and
J.
Lepowsky
, “
An equivalence of two constructions of permutation-twisted modules for lattice vertex operator algebras
,”
J. Pure Appl. Algebra
210
,
797
826
(
2007
).
8.
L.
Borisov
,
M. B.
Halpern
, and
C.
Schweigert
, “
Systematic approach to cyclic orbifolds
,”
Int. J. Mod. Phys. A
13
,
125
168
(
1998
).
9.
T.
Creutzig
and
A.
Linshaw
, “
Orbifolds of symplectic fermion algebras
,”
Trans. Am. Math. Soc.
369
,
467
494
(
2017
).
10.
C.
Dong
and
C.
Jiang
, “
Representations of the vertex operator algebra VL2A4.
,”
J. Algebra
377
,
76
96
(
2013
).
11.
C.
Dong
and
G.
Mason
, “
On quantum Galois theory
,”
Duke Math. J.
86
,
305
321
(
1997
).
12.
C.
Dong
and
K.
Nagatomo
, “
Classification of irreducible modules for the vertex operator algebra M(1)+
,”
J. Algebra
216
,
384
404
(
1999
).
13.
C.
Dong
,
L.
Ren
, and
F.
Xu
, “
On orbifold theory
,”
Adv. Math.
321
,
1
30
(
2017
).
14.
C.
Dong
,
F.
Xu
, and
N.
Yu
, “
The 3-permutation orbifold of a lattice vertex operator algebra
,”
J. Pure Appl. Algebra
222
,
1316
1336
(
2018
).
15.
I.
Frenkel
,
J.
Lepowsky
, and
A.
Meurman
,
Vertex Operator Algebras and the Monster
, Pure and Applied Mathematics Vol. 134 (
Academic Press
,
1988
).
16.
V.
Kac
and
I.
Todorov
, “
Affine orbifolds and rational conformal field theory extensions of W1+∞
,”
Commun. Math. Phys.
190
,
57
111
(
1997
).
17.
A.
Klemm
and
M.
Schmidt
, “
Orbifolds by cyclic permutations of tensor product conformal field theories
,”
Phys. Lett. B
245
,
53
58
(
1990
).
18.
A.
Kirillov
, Jr.
, “
On G–equivariant modular categories
,” e-print arXiv:math/0401119 (
2004
).
19.
J.
Lepowsky
, “
Calculus of twisted vertex operators
,”
Proc. Natl. Acad. Sci. U. S. A.
82
,
8295
8299
(
1985
).
20.
A.
Linshaw
, “
A Hilbert theorem for vertex algebras
,”
Transform. Groups
15
,
427
448
(
2010
).
21.
A.
Linshaw
, “
Invariant theory and the Heisenberg vertex algebra
,”
Int. Math. Res. Not.
17
,
4014
4050
(
2012
).
22.
A.
Linshaw
, “
Invariant subalgebras of affine vertex algebras
,”
Adv. Math.
234
,
61
84
(
2013
).
23.
A.
Milas
,
M.
Penn
, and
J.
Wauchope
, “
S3-permutation orbifolds of fermionic vertex algebras
,” IndAM Proceedings Series (to appear).
24.
A.
Milas
,
M.
Penn
, and
J.
Wauchope
(unpublished).
25.
K.
Theilmans
, “
A mathematica package for computing operator product expansions
,”
Int. J. Mod. Phys.
C2
,
787
798
(
1991
).
26.
H.
Weyl
,
The Classical Groups: Their Invariants and Representations
(
Princeton University Press
,
1946
).