vertex operator algebras modular forms and moonside
vertex operator algebras modular forms and moonside
Moonshine
Geoffrey Mason
University of California, Santa Cruz
Contents
Lecture 1.
1. The Monster simple group
2. J and V \
3. Monstrous Moonshine
4. Vertex algebras
5. Locality and quantum fields
6. CFT axioms
Lecture 2.
7. Lie algebras and local fields
8. Vertex operator algebras
9. Super vertex operator algebras
Lecture 3.
10. Modules over a VOA
11. Lattice theories
12. Partition functions and modular-invariance
Lecture 4.
13. The Lie algebra on V1
14. Automorphisms
15. Twisted sectors
16. The Moonshine module
17. AutV \
References.
1
Lecture 1
§1 The Monster simple group
Group theorists conceived the Monster sporadic simple group M in the early
1970s, although it was not officially born until 1982. Many features of M
were understood well before that time, however. In particular the complete
character table was already known. Here is a small part of it.
1 2A 2B
χ1 1 1 1
χ2 196883 4371 275
χ3 21296876 91884 − 2324
Let Vi be the M -module that affords the character χi . From the character
table one can compute branching rules Vi ⊗ Vj = ⊕k cijk Vk . In particular,
the tensor square V2⊗2 decomposes into the sum of symmetric and exterior
squares S 2 (V2 ) ⊕ Λ2 (V2 ), and the branching rules show that c222 = 1 with
V2 ⊆ S 2 (V2 ). So there is a canonical M -invariant surjection V2 ⊗ V2 → V2 ,
and it gives rise to a commutative, nonassociatve algebra structure on V1
whose automorphism group contains M . We can formally add an identity
element 1 to obtain a unital, commutative algebra
B = V1 ⊕ V2 (1)
§2 J and V \
Up to an undetermined constant, there is a unique modular function of weight
0 on the full modular group Γ := SL2 (Z) with a simple pole of residue 1 at
∞. Such functions can be represented as quotients of holomorphic modular
forms of equal weight. For example we have
θE8 (τ )3
J + 744 = = q −1 + 744 + 196884q + 21493760q 2 + . . . (2)
∆(τ )
θΛ (τ )
J + 24 = = q −1 + 24 + 196884q + 21493760q 2 + . . . . (3)
∆(τ )
2
Here,
X
θL (τ ) = q (α,α)/2 (6)
α∈L
is the theta function of an even lattice L, and E8 , Λ denote the E8 root lattice
and Leech lattice respectively.
John McKay noticed that the first few Fourier coefficients in (4) are sim-
ple linear combinations of dimensions of the irreducible M -modules Vi with
nonnegative coefficients. This suggests that we replace the coefficients by
the putative M -modules that correspond to them - a process that sometimes
goes by the abysmal name of ‘categorification’. From (3)-(6), the coefficients
of J are all nonnegative, so at least they correspond to linear spaces. Shifting
the grading by 1 for later convenience, we obtain a Z-graded linear space
with
V0\ = V1
V2\ = V1 ⊕ V2 (8)
V3\ = V1 ⊕ V2 ⊕ V3
...
and with dim Vn\ = coefficient of q n−1 in (4). McKay’s observation was pro-
moted to the conjecture that each Vn\ carries a ‘natural’ action of M . Note
that V2\ is identified with the algebra B.
3
§3 Monstrous Moonshine
With the conjectured Z-graded M -module V \ in hand, for each g ∈ M we
can take the graded trace of g and obtain another q-series
∞
X
−1
Zg = Zg (q) := q TrVn\ (g)q n . (9)
n=0
It was John Thompson who first asked what one can say about these
additional q-expansions. (There are 174 of them, one for each conjugacy
class of M .) We have Z1A (1, q) = J by construction, and from the character
table and (8) we see that
η(τ )24
Z2B (q) = + 24, (11)
η(2τ )24
where η(τ ) is the Dedekind eta-function
∞
Y
1/24
η(τ ) = q (1 − q n ). (12)
n=1
4
§4 Vertex algebras
The problem is now to define a natural action of the Monster M on V \ (59)
so that the graded traces Zg satisfy the Conway-Norton moonshine conjec-
tures (10). Borcherds’ radical proposed solution involved the idea of a vertex
algebra, which may be defined as follows. It is a pair (V, 1) consisting of a
nonzero C-linear space V and a distinguished vector 1 6= 0. Moreover, V is
equipped with bilinear products
µn : V ⊗ V → V (n ∈ Z),
u ⊗ v 7→ u(n)v (u, v ∈ V ),
satisfying the following axioms for all u, v, w ∈ V :
1) There is n0 = n0 (u, v) ∈ Z such that u(n)v = 0 for n ≥ n0 , (13)
2) v(n)1 = 0 (n ≥ 0) and v(−1)1 = v, (14)
3) For all p, q, r ∈ Z we have
∞
X p
{u(r + i)v}(p + q − i)w = (15)
i=0
i
∞
i r
X
(−1) {u(p + r − i)v(q + i)w − (−1)r v(q + r − i)u(p + i)w}.
i=0
i
Thanks to 1), both sums in 3) are finite so that the identity in question is
sensible.
At this point the reader may well be asking, where did these identities
come from, what are they good for, and what do they have to do with Mon-
strous Moonshine? The point of these lectures is to address these questions.
We begin by specializing (15) in various ways. It is convenient to consider
u(n) ∈ End(V ) to be the linear operator such that v 7→ u(n)v (v ∈ V ).
Taking r = 0, the binomial ri vanishes unless i = 0 and (15) reduces to the
operator identity
∞
X p
[u(p), v(q)] = {u(i)v}(p + q − i), (16)
i=0
i
called the commutator formula. Similarly, taking p = 0 yields the associativ-
ity formula
∞
i r
X
{u(r)v}(q) = (−1) {u(r − i)v(q + i) − (−1)r v(q + r − i)u(i)}. (17)
i=0
i
5
With n0 (u, v) as in (13), we obtain
∞
X
i r
(−1) {u(p + r − i)v(q + i) − (−1)r v(q + r − i)u(p + i)} = 0 (18)
i=0
i
6
We use obvious notation when manipulating fields, eg.,
X
Y (u, z)v := {u(n)v}z −n−1 ∈ V [[z]][z, z −1 ].
n
This means that the (operator) coefficients of each monomial z1p z2q in the
following identity coincide:
Indeed,
Therefore also
7
whence locality (21), (22) holds if, and only if, for all integers p, q, and some
nonnegative integer k we have
k
X k
i
(−1) ap+k−i bq+i − (−1)k bq+k−i ap+i = 0.
i=0
i
The last display is identical with the commutivity formula (18) if we take
r = k. Because (18) holds for all r ≥ n0 , it certainly holds for some positive
integer k in place of r. Combining this with (19), we have established
With this convention, identifying the operator coefficients for each monomial
z0p z1q z2r on the lhs and rhs of (24) yields exactly the identity (15).
§6 CFT axioms
(23) is the ‘main’ axiom for (2-dimensional) conformal field theory (CFT). We
now discuss the other axioms. Let (V, 1) be a vertex algebra, and introduce
the endomorphism
D : V → V, u 7→ u(−2)1. (25)
8
Using (14) and associativity (17) with q = 2, we have
∞
X n i
{u(n)v}(−2)1 = (−1) {u(n − i)v(−2 + i)1}
i=0
i
= u(n)v(−2)1 − nu(n − 1)v(−1)1
= u(n)v(−2)1 − nu(n − 1)v. (26)
Therefore,
X
[D, Y (u, z)]v = {Du(n)v − u(n)Dv}z −n−1
n
X
= {(u(n)v)(−2)1 − u(n)v(−2)1}z −n−1
n
X
= {−nu(n − 1)v}z −n−1
n
X
= {(−n − 1)u(n)v}z −n−2
n
d
= Y (u, z)v,
dz
where d/dz is the formal derivative. Hence, we obtain
d
[D, Y (u, z)] = Y (u, z).
dz
If we take u = v = w = 1 and p = q = r = −1 in (15) we find that
1(−2)1 = 1(−2)1 + 1(−2)1. Thus 1(−2)1 = 0, that is D1 = 0.
We have arrived at the following set-up: a quadruple (V, Y, 1, D) consist-
ing of a linear space V , a distinguished nonzero vector 1 ∈ V , an endomor-
phism D : V → V with D1 = 0, and a linear injection Y : V 7→ F(V ),
satisfying the following for all u, v ∈ V :
9
CFT. Conversely if (V, Y, 1, D) is a CFT then it can be shown that (V, 1) is
a vertex algebra. Basically, this means that the full strength of (15) can be
recovered (27).
The nomenclature in (27) is fairly standard in the physical literature, and
we use it in what follows. In addition, 1 is the vacuum vector, V is a Fock
space, elements in V are states, Y is the state-field correspondence, u(n) is
the nth mode of Y (u, z). Creativity is interpreted to mean that the state u
is created from the vacuum by the field Y (u, z) corresponding to u.
There are several other useful identities that follow without difficulty from
our axiomatic set-up. Among them we mention the following.
Y (1, z) = IdV , (28)
∞ n
X D u n
Y (u, z)1 = ezD u = z ,
n=0
n!
∞
X (−D)i
u(n)v = (−1)n+1 v(n + i)u. (29)
i=0
i!
Lecture 2
§7 Lie algebras and local fields
Certain infinite-dimensional Lie algebras naturally give rise to mutually
local fields. In this Section we discuss some important examples that illus-
trate some of the ideas developed so far.
1. Affine algebras.
Let L be a (complex) Lie algebra with bracket [a, b] (a, b ∈ L), equipped with
a symmetric, invariant, bilinear form h , i : L ⊗ L → C. (Invariant means
that h[a, b], ci = ha, [b, c]i for a, b, c ∈ L.) The associated affine Lie algebra is
b := L ⊗ C[t, t−1 ] ⊕ CK with central element K and bracket
L
[a ⊗ tm , b ⊗ tn ] = [a, b] ⊗ tm+n + mδm+n,0 ha, biK.
There is a triangular decomposition
L b− ⊕ L
b=L b0 ⊕ L
b+
10
with
b− := {a ⊗ tm | m < 0}, L
L b+ := {a ⊗ tm | m > 0}, L
b0 := {a ⊗ t0 } ⊕ CK.
is a left L-module
b affording the representation π, say. (The linear isomor-
phism in (30) comes from the Poincaré-Birkhoff-Witt theorem.) A typical
vector in V is a sum of vectors that look like
(b1 ⊗ tn1 ) . . . (bk ⊗ tnk ) ⊗ w (bi ∈ L, w ∈ W, n1 ≤ . . . ≤ nk ≤ −1),
and
π(a ⊗ tn ){(b1 ⊗ tn1 ) . . . (bk ⊗ tnk ) ⊗ w} =
(a ⊗ tn )(b1 ⊗ tn1 ) . . . (bk ⊗ tnk ) ⊗ w (31)
11
X X
= z2−p−2 (z1 − z2 )2 π([a, b] ⊗ tp ) z1−m−1 z2m+1 +
p∈Z m∈Z
X
z2−2 (z1 2
− z2 ) ha, biπ(K) mz1−m−1 z2m+1
m∈Z
2 X 2
z1 z1 z1 z1
= z2−p −1 p
π([a, b] ⊗ t )δ − − 1 ha, biπ(K)δ 0
.
z2 p∈Z
z2 z2 z2
(33)
Here
P δ(z)n−1is as in Section 4 (cf. comments preceding (24)), and δ 0 (z) :=
n∈Z nz . Now check that (z − 1)k δ(z) = 0 for k ≥ 1, (z − 1)k δ 0 (z) = 0
for k ≥ 2. In particular, (33) vanishes and (z1 − z2 )2 [Y (a, z1 ), Y (b, z2 )] = 0,
as asserted.
When W = Cv0 is the trivial 1-dimensional L-module we can go further,
and see the begginnings of a CFT. Here,
12
(Because Cv0 is the trivial L-module then (a ⊗ tn )v0 = 0 for n ≥ 0.) Y (a, z)
is also translation covariant (loc. cit.): if m ≥ 1 then
13
( )
d X
= − {a ⊗ tn−1 .b ⊗ t−m ⊗ v0 }z −n
dz n∈Z
d
= − Y (a, z)b ⊗ t−m .
dz
This shows that [D, Y (a, z)] = d/dzY (a, z) where D = −d/dt, and be-
cause 1 is independent of t then D1 = 0. It should come as no surprise that
in fact (V (l, Cv0 ), Y, 1 ⊗ v0 , −d/dt) is a vertex algebra/CFT. Indeed, based
on what we already know, the result follows from the following general result.
In this situation, we say that S generates V . Thus (V (l, Cv0 ), Y, 1⊗v0 , −d/dt)
is a vertex algebra generated by L = L ⊗ t−1 . We will denote this vertex
algebra by V (L, l).
2. Virasoro algebra. (Several aspects of this case are similar to the previous
one, so we give less detail.)
The Virasoro algebra is the Lie algebra with underlying linear space
V ir := ⊕n∈Z CLn ⊕ CK with central element K and bracket
m3 − m
[Lm , Ln ] = (m − n)Lm+n + δm+n,0 K. (36)
12
(The denominator 12 is conventional here; it can be removed by rescaling.)
There is a triangular decomposition
with
Let W = Cv0 be the 1-dimensional V ir0 -module such that L0 v0 = hv0 , Kv0 =
cv0 , extend to a V ir+ ⊕ V ir0 -module by letting V ir+ annihilate v0 , and form
14
the induced module
V = V (c, h) = IndVV ir
ir+ ⊕V ir0 W
∼
= S(V ir− ) ⊗ Cv0
= S(⊕n<0 CLn ) ⊗ Cv0
∼
= C1 ⊕ CL−1 ⊕ . . .
where 1 := 1 ⊗ v0 . h and c are called the conformal weight and central charge
respectively. Introduce
X
Y (ω, z) := Ln z −n−2 . (37)
n∈Z
One sees easily that Y (ω, z) ∈ F(V ). Note the slight change in convention
regarding powers of z in (37), which is standard. The reader may enjoy
proving that Y (ω, z) is a (self-) local field. Indeed, we have
Note that
X
Y (ω, z)1 = {Ln 1}z −n−2 = h1z −2 + L−1 1z −1 + L−2 1 + . . . . (39)
n∈Z
15
that the following hold:
X
1) Y (ω, z) = L(n)z −n−2 and the modes L(n) generate an action of
n∈Z
the Virasoro algebra V ir (36) in which K acts on V as a scalar c,
called the central charge of V .
2) L(0) is a semisimple operator on V . Its eigenvalues lie in Z, are
bounded below, and have finite-dimensional eigenspaces.
3) D = L(−1).
Thanks to 3) and the last display, translation covariance may then be written
16
The vertex algebra V (L, l) can sometimes be given the structure of a VOA
- we just have to find the right conformal vector. We describe two important
cases where this can be achieved.
u(n) : V p → V p+|u| .
17
Finally, we require the super version of the basic identity (15), namely
∞
X p
{u(r + i)v}(p + q − i)w =
i=0
i
∞
X r
i
(−1) {u(p + r − i)v(q + i)w − (−1)r+|u||v| v(q + r − i)u(p + i)w}.
i=0
i
Lecture 3
18
YW (v, z) = n∈Z vW (n)z −n−1 such that YW (1, z) = IdW and the analog of
P
(15) holds, i.e. for all u, v ∈ V, w ∈ W we have
∞
X p
{u(r + i)v}W (p + q − i)w = (42)
i=0
i
∞
i r
X
(−1) {u(p + r − i)W v(q + i)W w − (−1)r v(q + r − i)W u(p + i)W w}.
i=0
i
19
the only ideals are the trivial ones V and 0. For example, any Heisenberg
VOA V (l, Cv0 ) is simple.
3. If (V, Y, 1, ω) is a SVOA (cf. Section 9) the odd part V 1 is a module over
the even part V 0 . In this case the conformal weight of V 1 is necessarily
a half-integer. (A VOA may have no modules with half-integral conformal
weight except for (direct sums of) the adjoint module. Thus the VOAs that
can occur as the even part of a SVOA are severely restricted.)
4. Recall the rank l Heisenberg VOA V (l, Cv0 ) (cf. Section 8) generated by a
rank l abelian Lie algebra L. For an L-module W we constructed (Section 7)
a space V (l, W ) and mutually local fields Y (a, z) ∈ F(V (l, W )) (a ∈ L). It is
not hard to see that V (l, W ) is a V (l, Cv0 )-module, and it is simple whenever
dim W = 1.
Associated to this L-module is the simple V (l, Cv0 )-module V (l, Ceβ ). Note
that Ce0 is the trivial L-module, so that it can be identified with Cv0 . Also,
we have a linear isomorphism V (l, Ceβ ) ∼= S(H b − ) ⊗ Ceβ (cf. (30)). We form
the Fock space
M
VL := V (l, Ceβ )
β∈L
∼
M
b −) ⊗
= S(H Ceβ (45)
β∈L
b−
= S(H ) ⊗ C[L].
20
(It is convenient to identify the group algebra C[L] of L with ⊕β Ceβ .) We
discuss the following result:
VL carries the structure of a SVOA; if L is an even
lattice (i.e. hβ, βi ∈ 2Z for β ∈ L), then VL is a VOA.
S(H b − ) is naturally identified with the Heisenberg VOA itself, and in par-
ticular it is generated (cf. (35)) by the fields Y (α, z) (α ∈ H). Because each
V (l, Ceβ ) is a Heisenberg module, the Y (α, z) naturally extend to (mutually
local) fields on VL . To get a generating set of fields for VL (loc. cit.) we would
need to extend the set of Y (α, z) to a larger set of mutually (super) local
fields by defining fields Y (1 ⊗ eβ , z) (β ∈ L) directly. We will skip the details
here. RecallP (cf. Section 8) that the conformal vector for the Heisenberg VOA
is ω := 1/2 li=1 vi (−1)vi for an orthonormal basis {vi } of H. This state is
also taken as the conformal vector of VL . In Pparticular, the central charge of
−n−2
VL is the rank l of L. The field Y (ω, z) = n L(n)z determined by ω is
defined in the natural way, i.e. on V (l, Ceβ ) it acts as YV (l,Ceβ ) (ω, z). Since
each summand in (45) is a Heisenberg module, L(0) acts semisimply on each
of them, and therefore on VL . We consider the eigenvalues and eigenspaces
of L(0) in the next Section. Finally, we note that VL is a simple VOA if L is
even.
§12 Partition functions
Suppose that (V, Y, 1, ω) is a VOA of central charge c (cf. Section 8, axiom
1)), and spectral decomposition (40) into L(0)-eigenspaces. The partition
function of V is the formal q-series
∞
X
−c/24
Z(q) = ZV (q) := q dim Vn q n . (46)
n=n0
(This is the first place that c has played a rôle in the proceedings.) More
generally, for a simple V -module W with spectral decomposition (43), the
corresponding partition function is
∞
X
Z(q) = ZW (q) := q h−c/24 dim Vn q n . (47)
n=0
21
can often check the VOA axioms regarding the conformal vector (Section 8,
axiom 2)) by directly computing the corresponding partition function. We
will carry this out in the case of the Fock spaces for the Heisenberg VOA
and the lattice theory VL .
For the rank l Heisenberg theory V = V (l, Cv0 ) we saw (34) that V
has a tensor decomposition S(⊕∞ m=1 L ⊗ t
−m
) ⊗ Cv0 (L is the abelian Lie
algebra of rank l). It is not hard to see that the L(0)-grading respects this
decomposition, and that L ⊗ t−m is an eigenspace with eigenvalue m. Since
symmetric powers are multiplicative over direct sums, we obtain
∞
Y
−l/24
partition function of S(L ⊗ t−m )
ZV (l,Cv0 ) (q) = q
m=1
Y∞
= q −l/24 (1 + q m + q 2m + . . .)l
m=1
Y∞
= q −l/24 (1 − q m )−l = η(q)−l ,
m=1
(Here, we used that vi (j) moves across the tensor sign if j ≥ 0 and annihilates
eβ if j ≥ 1, as well as (44). The last equality holds because {vi } is an
22
orthonormal basis of H.) The upshot is that 1 ⊗ eβ is an eigenvector for L(0)
with eigenvalue 1/2hβ, βi. We therefore see that
X
partition function of C[L] = q 1/2hβ,βi = θL (q)
β∈L
23
One point that we will not pursue but that deserves mention is this:
the partition function of a VOA is a formal q-expansion, with no a priori
convergence properties. On the other hand, at least for a regular VOA, the
partition function turns out to be holomorphic in the complex upper half-
plane H when we think of it as a function ZV (τ ) with q = e2πiτ , τ ∈ H. For
this reason, we now write partition functions as functions of τ rather than q.
Although there will be no time to develop the general theory of regular
VOAs in these lectures, we can illustrate some of the ideas using the lattice
theory VL . If V is an arbitrary VOA, the set of modules over V are the objects
of a category V -Mod. A morphism f : W1 → W2 between two V -modules
W1 , W2 is a linear map such that
In terms of fields, this reads f YW1 (u, z) = YW2 (u, z)f . Roughly speaking, V
is called rational if V -Mod is semisimple, i.e. every V -module is a direct sum
of simple V -modules. (In fact, one has to include additional types of modules
that we did not discuss in Section 10.) It can be shown that a rational VOA
has only finitely many (isomorphism classes of) simple V -modules. A VOA
is regular if it is both rational in the above sense and satisfies an additional
condition that we will not discuss here.
If L is an even lattice as before then VL is indeed a regular VOA. It
therefore has only finitely many inequivalent simple modules, and in fact
they are enumerated by the quotient group L0 /L where L0 is the dual lattice
of L. If we set E := R ⊗Z L then the dual lattice is
L0 := {α ∈ E | hα, βi ∈ Z (β ∈ L)}.
∼
M
b −) ⊗
= S(H Ceβ (50)
β∈L+γ
b − ) ⊗ C[L + γ],
= S(H
24
(compare with (45)), where L + γ ∈ L0 /L. The partition function is
θL+γ (τ )
ZVL+γ (τ ) = ,
η(τ )l
Lecture 4
25
V \ , also of central charge 24, whose partition function is J (4), which has
constant term 0. This is the1 Moonshine module.
Although the VOAs VΛ and V \ have partition functions differing only in
their constant term, many of their algebraic properties are quite different.
Indeed, these properties are to a large extent governed by the constant term.
For this reason, we begin with a general discussion of this point. We restrict
attention to VOAs of CFT-type, which means that in the spectral decom-
position (40) the pieces Vn vanish for n < 0 and V0 = C1. (Recall that we
always have 1 ∈ V0 .) There are many interesting VOAs that are not of CFT-
type, nevertheless CFT-type theories are natural from a physical standpoint
because they arise from ’unitarity’ assumptions. Be that as it may, our basic
assumption here is that the spectral decomposition of V has the shape
V = C1 ⊕ V1 ⊕ V2 ⊕ . . .
26
root system embedded in the ambient Euclidean space E = R ⊗ L.) For
example, if L = 3E8 then the Lie algebra on (VL )1 is semisimple, being the
sum of three copies of the E8 Lie algebra. (Note that dim E8 = 248, so that
dim(VL )1 = 744, in agreement with (2).) Similarly, the Leech lattice Λ has
no roots, whence (VΛ )1 is abelian of rank l = 24.
Because J has no constant term, a VOA V \ with partition function J
and central charge c = 24 necessarily has no corresponding Lie algebra. In
particular, V \ cannot be a lattice theory, because the weight one piece never
vanishes for a lattice theory (cf. (51)).
§14 Automorphisms
Let V be a (S)VOA. An automorphism of V is an invertible linear map
g : V → V such that g(ω) = ω and gv(q)g −1 = g(v)(q) for all v, q, i.e.
gY (v, z)g −1 = Y (g(v), z) (v ∈ V ). (52)
We give some basic examples of automorphisms.
1. One checks (use induction and (16) or (17)) that for n ≥ 0,
n
i n
X
n
(u(0) v)(q) = (−1) u(0)n−i v(q)u(0)i (u, v ∈ V, q ∈ Z).
i=0
i
Therefore,
∞
u(0)
X 1
(u(0)n v)(q)
e .v (q) =
n=0
n!
∞ X
n
X (−1)i
= u(0)n−i v(q)u(0)i
n=0 i=0
i!(n − i)!
= e v(q)e−u(0) ,
u(0)
27
= − {(L(−1)u)(−1) − (L(0)u)(−2) + (L(1)u)(−3)} 1
= 0.
(For the last two equalities use translation covariance, L(0)u = u (because
u ∈ V1 ), L(1)u ∈ V0 = C1, L(n)u ∈ V1−n = 0 for n ≥ 2, and 1(q) = δq+1,0 IdV
(cf. (28).)
It follows from this calculation that if V is a VOA of CFT-type then
u(0)
{e | u ∈ V1 } is a set of automorphisms of V . In the previous Section we
learned that V1 carries the structure of a Lie algebra with bracket [uv] =
u(0)v. Now we see that the usual action of the associated Lie group G
generated by exponentials eadu extends to an action of G as automorphisms
of V .
2. Suppose that V is a SVOA. Then there is a canonical involutorial au-
tomorphism which acts as +1 on the even part of V and −1 on the odd
part.
3. A related example (and the one we will need later) is an involutorial
automorphism t of a lattice VOA VL , defined to be a lifting of the −1 au-
tomorphism of the lattice L. t also acts as −1 on the abelian Lie algebra
C ⊗ L and then acts as naturally on the associated Heisenberg VOA (cf. (34)
- where L is the Lie algebra, not the lattice!) and on VL , where
(cf. (45)).
28
b − ). Therefore by (34),
from states in the Heisenberg VOA Fock space S(H
m>0
Y
−l/24
= q (1 + q m )−l
m>0
l
η(τ )
= . (54)
η(2τ )
whenever g(u) = e−2πir/R (r ∈ Z), and Yg (1, z) = IdWg . The twisted vertex
operators Yg (u, z) are required to satisfy twisted analogs of the basic identity
(15). In the delta-function formulation (cf. (24)) this reads
−1 z1 − z2 −1 z2 − z1
z0 δ Yg (u, z1 )Yg (v, z2 ) − z0 δ Yg (v, z2 )Yg (u, z1 )
z0 −z0
−r/R
−1 z1 − z0 z1 − z2
= z2 δ Yg (Y (u, z0 )v, z2 ). (55)
z2 z0
Finally, the operator Lg (0) (the zero mode of Yg (ω, z)) is required to be
semisimple with finite-dimensional eigenspaces. The eigenvalues satisfy a
truncation condition analogous to that for V -modules (cf. the discussion in
Section 10 preceding display (43)). There is an obvious notion of irreducible
29
(or simple) g-twisted module, and as in the untwisted case (cf. (43)) the
spectral decomposition of a simple g-twisted module takes the form
∞
M
Wg = (Wg )hg +n/R (56)
n=0
for a scalar hg (the conformal weight). Needless to say, the twisted sector
has an associated partition function
∞
X
−c/24+hg
ZWg (τ ) := q dim(Wg )n q n/R .
n=0
Let us specialize to the case of the (even, self-dual) Leech lattice Λ with
its associated VOA VΛ and canonical involution t (cf. Section 14). In this
case there is (up to isomorphism) a unique simple t-twisted module, denoted
by VΛ (t, τ ). The following transformation law can be proved:
30
§16 The Moonshine Module
Retaining the notation of the previous Section, consider
VΛ ⊕ VΛ (t). (58)
The involution t acts naturally on the twisted sector: in the ‘usual way’
on S(⊕n>0 H ⊗ t−n/2 ) and as −1 on the 212 constant part. The Moonshine
Module is then defined to be the space of t-invariants
V \ := VΛ+ ⊕ VΛ (t)+ . (59)
Now every state u ⊗ eβ ∈ VΛ (β 6= 0) produces a t-invariant u ⊗ eβ +
t(u) ⊗ e−β . On the other hand, the partition function of the Heisenberg VOA
(consisting of states u⊗e0 ) is 1/∆(τ ) and the graded trace of t is ∆(τ )/∆(2τ )
(the case l = 24 of (54)). It follows that
ZV + (τ )
Λ
= (q −1 + 98580q + . . .) + (98304q + . . .)
= q −1 + 196884q + . . . (60)
It is clear from the above that ZV \ (τ ) is a modular function of weight 0 and
level at most 2, and it is easy to check that in fact it is invariant under the
full modular group. Thus from the q-expansion we arrive at the identity
ZV \ (τ ) = J.
31
The space (58) has the structure of an abelian intertwining algebra, a
generalization of VOA and SVOA. The main missing ingredient, which we
cannot go into here, is the definition of fields YVΛ ⊕VΛ (t) (u, z) for states u in
the twisted sector satisfying an appropriate variation of the basic identity
(15), (24). Once this is done, t is seen to be an automorphism of this larger
structure. Then it is easy to see that the t-invariant subspace V \ , together
with the restriction of the fields to this subspace, defines the structure of a
VOA on V \ with central charge 24. Furthermore, VΛ+ ⊕ VΛ (t)− is a SVOA
with even part VΛ+ . (Indeed, it is an N = 1 superconformal field theory, a
term we alluded to but did not define in Section 9.)
Consider a VOA V of CFT-type (cf. Section 13) with trivial Lie algebra
V1 :
V = V0 ⊕ V2 ⊕ . . .
§17 AutV \
Consider the CFT
VΛ = C1 ⊕ (VΛ )1 ⊕ . . .
where Λ is, as before, the Leech lattice. Because Λ has no roots, it follows
from (51) that dim(VΛ )1 = 24, and the Lie algebra on (VΛ )1 is abelian. So the
automorphisms eu(0) (u ∈ (VΛ )1 ) generate a 24-dimensional complex torus
T . Additional automorphisms of VΛ arise from the automorphism group
Co0 := Aut(Λ) of the Leech lattice, and there is a (nonsplit) short exact
sequence
1 → T → AutVΛ → Co0 → 1.
32
The automorphism t of Λ (or of VΛ ) is a central involution of Co0 , and the
quotient Co1 := Co0 /hti is the largest sporadic (simple) Conway group of
order 221 . . ..
Because t acts as −1 on T , its only fixed elements are those of order at
most 2. So the centralizer C(t) of t in AutVΛ (the elements that commute
with t) is described by another short exact sequence (also nonsplit)
1 → 21+24 → C
b → Co1 → 1,
b ⊆ AutV \
C
|M | = 246 . . . 47.59.71
These results are not easily obtained, and we say no more about them here.
The graded traces ZV \ (g, τ ) for g ∈ M turn out to be hauptmoduln as
described in Section 1. This result is also difficult. We end these Notes with
the computation for a single automorphism g of order 2 that acts trivially
33
on VΛ+ and as −1 on VΛ (t)+ . A previous calculation shows that its graded
trace is a modular function of weight 0 and level 2. Specifically,
ZV \ (g, τ ) = ZV + (τ ) − ZVΛ (t)+ (τ )
Λ
−1
= (q + 98556q + . . .) − (98304q + . . .)
= q −1 + 276q + . . .
is the hauptmodul for the Monster element 2B (11).
References
The following textbooks and monographs cover the material in these Notes
and much more.
J. Conway et al, ATLAS of Finite Groups, Clarendon Press, OUP, 1985.
C. Dong and J. Lepowsky, Generalized Vertex Algebras and Relative Vertex
Operators, Birkhäuser, 1993.
I. Frenkel, J. Lepowsky and A. Meurman, Vertex Operator Algebras and the
Monster, Academic Press, 1988.
T. Gannon, Moonshine Beyond the Monster: The Bridge Connecting Alge-
bra, Modular Forms, and Physics, CUP, 2006.
V. Kac, Vertex Algebras for Beginners, 2nd ed., Univ. Lect. Ser. Vol. 10,
AMS, 1998.
J. Lepowsky and H. Li, Introduction to Vertex Operator Algebras and Their
Representations, Birkäuser, 2004.
A. Matsuo and K. Nagatomo, Axioms for a Vertex Algebra and the Locality
of Quantum Fields, Math. Soc. of Japan Memoirs, Vol. 4, 1999.
G. Mason and M. Tuite, Vertex operators and modular forms, in A Window
into Zeta and Modular Physics, MSRI Publ. No. 57, CUP, 2010.
Moonshine, The First Quarter Century and Beyond, J. Lepowsky, J. McKay
and M. Tuite eds., LMS Lect. Note Series No. 372, CUP, 2010.
Moonshine, the Monster, and Related Topics, C. Dong and G. Mason eds.,
Contemp. Math. No. 193, AMS, 1996.
34