TC Quantum Field Intro
TC Quantum Field Intro
Kevin Walker
0 Introduction 1
2 Z2 Homology as a TQFT 9
2.1 The Basic Construction . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Gluing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Cylinder Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.4 Semisimplicity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3 Topological Fields 17
3.1 Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Topological Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
3.3 Extended Isotopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
iii
iv CONTENTS
7 1+1-dimensional Examples 67
7.1 Generalities on 1+1-dimensional Theories . . . . . . . . . . . . . . . 67
7.2 Finite Group Theories in 1 + 1 dimensions . . . . . . . . . . . . . . . 67
8 2+1-dimensional Examples 69
8.1 Temperley-Lieb Theories . . . . . . . . . . . . . . . . . . . . . . . . . 69
8.2 Spherical Categories in General . . . . . . . . . . . . . . . . . . . . . 69
8.3 A2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
8.4 G2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
8.5 Finite Group Theories in 2+1 dimensions . . . . . . . . . . . . . . . 74
8.6 Jones Planar Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 74
9 3+1-dimensional Examples 75
9.1 Theories From Ribbon Categories . . . . . . . . . . . . . . . . . . . . 75
9.2 Decategorification . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
9.3 More on Chern-Simons Theories . . . . . . . . . . . . . . . . . . . . 84
9.4 Contact Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
10 4+1-dimensional Examples 87
10.1 Theories From Khovanov Homology . . . . . . . . . . . . . . . . . . 87
11 n+1-dimensional Examples 89
11.1 Finite Group Theories . . . . . . . . . . . . . . . . . . . . . . . . . . 89
11.2 Finite Total Homotopy Theories . . . . . . . . . . . . . . . . . . . . 89
11.3 String Category . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
12 Reconstruction Results 91
13 [Other Chapters] 93
A Categories 95
A.1 Definitions and Notation . . . . . . . . . . . . . . . . . . . . . . . . . 95
A.2 [Still to do] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
B Semisimple Categories 97
B.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
B.2 Conditions Defining Semisimplicity . . . . . . . . . . . . . . . . . . . 98
B.3 The Density Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . 99
B.4 Semisimple Categories . . . . . . . . . . . . . . . . . . . . . . . . . . 101
B.5 To Do List . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Chapter 0
Introduction
• my notational preferences evolved as this was written, so there are some no-
tational inconsistencies
• key idea: Computing the path integral by actually integrating against some
measure on the space of fields is notoriously difficult, so to develop a mathe-
matically rigorous theory we need to take a detour around this obstacle. Most
previous approaches follow Segal and Atiyah [need refs] and take a rather
wide detour: a TQFT is defined (roughly) to be any choice of vector spaces
for n-manifolds and vectors for n+1-manifolds which satisfy algebraic prop-
erties suggested by the path integral. Here we make a tighter detour around
the path integral. We retain the notion of fields on manifolds and require
1
2 CHAPTER 0. INTRODUCTION
that the vector spaces for n-manifolds be constructed out of these fields by
local relations (or dually local projections). These local relations carry the
same information as the path integral of the n+1-dimensional ball B n+1 with
all possible boundary conditions. One advantage of this approach is that the
higher algebra / higher codimension / multi-tier structure [cite Freed (and also
xxxx?)] is automatically present all the way down to dimension zero. Another
advantage is that we can construct most (all?) known TQFTs in a uniform
framework.
• I’ve tried to give good citations, but there are bound to be inadvertant omis-
sions (comments welcome) [this is for the future — there are very few citations
at the moment]
• hard to decide what order to put things; some readers might want to skip
ahead to examples before reading all of the general theory
• I’ve gone part way toward making things general in a category-theoretic sense,
but not all the way. (e.g. I usually assume the target category for the TQFT is
complex vector spaces, but the proofs for the most part work more generally.)
T : C(M ; c) → T.
Here T ⊂ C denotes the the circle group. (In more traditional notation, T = eiS
for some S : C(M ; c) → R. S is called the action for the theory.) Assume that T is
local, in the sense that it is multiplicative (i.e. S is additive) with respect to disjoint
unions and it commutes with gluing. In other words, we can cut M into pieces and
3
4 CHAPTER 1. FROM PATH INTEGRALS TO LOCAL RELATIONS
a
d
d
c
c
M Mgl
Since (Y × I) ∪Y (Y × I) ∼
= Y × I, we have, by (1.1.2),
Z
KY (a, y)KY (y, b) = KY (a, b).
1.1.3 y∈C(Y ;d)
1.2. FUNCTORIAL PROPERTIES OF THE PATH INTEGRAL 5
πY2 = πY .
Z(M ) : C(∂M ) → C
c 7→ Z(M ; c).
Then clearly
Z(M ) ∈ Z(∂M ). 1.2.1
(Proof: Glue a collar of ∂M to M and apply (1.1.2). The result is still M .) It
is also easy to see that for any closed n-manifolds Y1 and Y2 , there is a natural
identification
Z(Y1 ⊔ Y2 ) = Z(Y1 ) ⊗ Z(Y2 ),
and that for any n+1-manifolds M1 and M2
Z(Y ) ⊗ Z(−Y ) → C
Z
f ⊗ g 7→ f (x)g(x).
x∈C(Y )
and
Z(Mgl ) = trY (Z(M )).
6 CHAPTER 1. FROM PATH INTEGRALS TO LOCAL RELATIONS
E
x x E
D
Y′
′
Y a′ D a′
∂Y ∂Y
a
If ∂M = Y ∪ −Y ∪ W (setup for gluing with corners [include figure?]), then for each
c ∈ C(∂Y ) there is a trace map
trY,d
Z(Y ; d) ⊗ Z(−Y ; d) ⊗ Z(W ; d, d) −→ Z(W ; d, d).
These maps (for each d) combine to determine a map
Y tr
Z(∂M ) −→ Z(∂Mgl ),
(see (6.1.4)) and again
Z(Mgl ) = trY (Z(M )).
The above constitutes a gluing (with corners) formula for n+1-manifolds.
D×I
D×I D×I D×I D×I D×I
(This is a product rather than a direct sum because non-zero components are allowed
for infinitely many x.) Define PD : F(Y ; a ∪ a′ ) → F(Y ; a ∪ a′ ) by
Y
PD = π(D;x) ⊗ id .
x∈C(E;∂a)
(Here we have suppressed a and a′ from the notation on the left hand side.) By the
same argument as before, PD2 = PD . PD is the local projection corresponding to
gluing a collar D × I to D ⊂ Y .
Now the key point: Y × I can be constructed by gluing Di × I to Y × {0}, where
Di runs through an open cover of Y . (See Figure (1.3.2).) It follows that πY can
be written as composition of the PDi ’s. This implies that
\
im(πY ) = im(PDi ).
Di
(Note that πY and the PDi ’s all commute, because the corresponding topological
gluings commute.)
Requiring that f ∈ im(PD ) is a local condition, in the sense that it only depends
on what’s going on inside a disk in Y . Thus we have obtained a local description of
im(πY ).
Dually, we can define y ∼ z if PD (y) = PD (z), where y and z are linear combina-
tions of fields on D. These are the local relations with which we will be concerned
for the rest of the book.
In practice, it turns out to be easier to define systems of local relations than it
is to define a path integral, so in the Chapter 4 we axiomatize these local relations.
Note that PD encodes the same information as Z(D × I) = Z(B n+1 ), so speci-
fying a local relation is tantamount to specifying the path integral of the n+1-ball.
8 CHAPTER 1. FROM PATH INTEGRALS TO LOCAL RELATIONS
Chapter 2
Z2 Homology as a TQFT
9
10 CHAPTER 2. Z2 HOMOLOGY AS A TQFT
(a) ∼ (b) ∼
(c) ∼
2.2 Gluing
Let Y be a 2-manifold with gluing data: ∂Y is decomposed into three disjoint
pieces, ∂Y = S+ ⊔ S− ⊔ S0 , and there is an identification S+ = −S− (i.e. S+
is identified with S− with reversed orientation). (See Figure (2.2.1).) Let Ygl
denote the 2-manifold obtained by gluing S+ to S− . Note that ∂Ygl = S0 . Given
(a, b, c) ∈ C(S+ )×C(S− )×C(S0 ) = C(∂Y ), we have the vector space A(Y ; a, b, c). We
would like to calculate A(Ygl ; c) in terms of the vector spaces A(Y ; a, b, c) for various
a and b. (c ∈ C(S0 ) = C(∂Ygl ) will be fixed throughout the following discussion.)
2.2. GLUING 11
S+ S−
gl
Ygl
Y
S0 S0
e ; x, x, c) → A(Y
glx : A(Y e gl ; c).
The latter map is compatible with the local relations, and so induces a linear map
If we can describe the kernel of this map, we will have proved the desired gluing
theorem.
Let C ∈ C(Y ; a, b, c) for some a, b ∈ C(S± ), and let e ∈ C(S± × I; a, b). (See
Figure (2.2.2).) Let Ce = C ∪S+ e ∈ C(Y ; b, b, c) and eC = C ∪S− e ∈ C(Y ; a, a, c).
Then clearly
glb (Ce) ∼ gla (eC) ∈ C(Ygl ; c)
(via an isotopy which shifts along the gluing locus in Ygl ), and so, moving from
C(Ygl ; c) to A(Ygl ; c),
M
Ce − eC ∈ ker(gl) ⊂ A(Y ; x, x, c).
x∈C(S± )
a b
e
C C
Ce b a eC
glb gla
(The quotient maps are denoted by p and q.) The top gluing mapL is injective and
the other three maps are surjective. Let K denote the subspace of x A(Y e ; x, x, c)
generated by ker(p) and elements of the form Ce − eC. It suffices to show that
e gl ; c) contains all elements of the form A − A′ , where A and A′ differ
gl(K) ⊂ A(Y
by a generating local relation supported in some disk D ⊂ Ygl . Arbitrary isotopies
on Ygl are generated by (a) isotopies supported away from the gluing locus in Ygl
and (b) a shift isotopy supported in a collar of the gluing locus which moves the
gluing locus off of itself. If E and E ′ ∈ C(Ygl ; c) differ by an isotopy of type (a),
then E − E ′ = gl(F ) where F ∈ ker(p). If E and E ′ ∈ C(Ygl ; c) differ by an isotopy
of type (b), then E = gl(eC) and E ′ = gl(Ce) for some C and e as above. It follows
that if E and E ′ ∈ C(Ygl ; c) differ by a general isotopy, then E − E ′ ∈ gl(K). Hence
we can assume that the disk D above is disjoint from the gluing locus, since D can
be isotoped off of the gluing locus. Clearly in this case A − A′ = gl(F ) for some
F ∈ ker(p).
2.3. CYLINDER CATEGORIES 13
Theorem. A(Ygl ; c) is the universal object with the above two properties. In other
words, if there is a vector space W and linear maps fa : Vaa → W for all x, such
that for all e : x → y the following diagram commutes
= Vaa BB
ida ×e {{{ BB fa
{{{ BB
BB
{{ !
Vab C = W
CC |
CC |||
C
e×idb CC! ||
|| fb
Vbb
then there exists a map g : A(Ygl ; c) → W such that fa = g · gla for all x.
This is just a restatement of the first version of the gluing theorem. It has the
advantage of generalizing to the case where target category is not a vector space.
For example the theorem remains true if we systematically, in the above discus-
sion, replace A(Y, a) with the set of all properly embedded 1-submanifolds in Y ,
modulo isotopy, with boundary a. In this case A × Aop acts on the collection of sets
{Vab }, and A(Ygl ; c) is the universal set having the above properties.
The above universal construction is usually called the coend (dual to end) of the
A × Aop action. (See for example [Mac98, p. 226].) From our point of view, it would
be better to call this the gluing of a representation V of A × Aop , but “coend” is
already well-established so we stick with that terminology.
2.4 Semisimplicity
Back to the Z2 -homology vector spaces A(Y ; a). The categories A(S) (where S is a
closed 1-manifold) have the additional property of being semisimple. (See Appendix
B for details.) This means, among other things, that there is a fixed list L(S) of
irreducible representations of A(S), and that any representation ρ of A(S) can be
written as M
ρ∼= ρα ⊗ α,
α∈L(S)
where the isomorphism is natural, and ρα = hom(α, ρ). It is also the case that
L(A(S ⊔ S ′ )) = L(A(S)) × L(A(S ′ )) (i.e. the irreps of A × A′ , where A and A′ are
semi-simple, are isomorphic to the products of irreps of A and A′ ).
More specifically, let e1 , e2 , e3 , e4 be the morphisms of A(S 1 ) shown in Figure
(2.4.1). The ei ’s are a complete set of minimal idempotents for A(S 1 ).PThat is,
ei ei = ei , ei ej = 0 if i 6= j, any morphism f of A(S 1 ) can be written as 41 gi ei hi
for some {gi , hi }, and no larger set of idempotents has this property.
Let αi = A(S 1 )ei as a left A(S 1 ) representation. Geometrically, αi is S 1 × I with
ei fixed on one end, arbitrary curves-mod-relations on the rest of S 1 × I, and A(S 1 )
acting on the other end (see Figure (2.4.2)).
2.4. SEMISIMPLICITY 15
1 1 1 1
e1 = + e3 = +
2 2 2 2
1 1 1 1
e2 = − e4 = −
2 2 2 2
ei
Topological Fields
3.1 Manifolds
Our manifolds will always be oriented and compact. They might have additional
structure (e.g. [need examples of additional structure]). They could be PL or
smooth. [for smooth, need to say more about corners; also could be topological?]
Morphisms between manifolds of the same dimension will always be orientation-
preserving homeomorphisms (or diffeomorphisms in the case of smooth manifolds)
unless specified otherwise.
17
18 CHAPTER 3. TOPOLOGICAL FIELDS
Y -Y
gl
Y'
Y
gl
-Y
Let Mi denote the category whose objects are compact, oriented manifolds of
dimension i, and whose morphisms are orientation-preserving homeomorphisms. In-
cluded in each Mi is ∅, the empty manifold of dimension i.
The collection of categories {Mi }i≥0 has the following familiar additional struc-
ture.
Y -Y Y'
Y'
gl
Y -Y
Y'
gl
Let S denote the category of sets. A topological field (of top dimension n)
consists of a collection of functors
Ci : Mi → S
(0 ≤ i ≤ n) which satisfy the following conditions. (We will usually omit the
subscript i from the notation.)
Boundary. There are boundary maps
∂ : C(X) → C(∂X).
∂ ∂
Ci−1 (∂f )
Ci−1 (∂X) / Ci−1 (∂Y )
In other words, ∂ is a natural transformation between the two functors Ci and Ci−1 ◦∂
from Mi to S.
Examples. For α ∈ MF(X), ∂α = α|∂X . For β ∈ S(X) (or β ∈ DF(X)), ∂β is
the (transverse) intersection of β with ∂X (preserving colors and orientations for
DF).
Orientation reversal. If X is closed, there is a bijection
C(X; c) ↔ C(−X, b
c), a↔b
a.
Examples. For MF, these are the obvious bijections resulting from ignoring the
orientation of X. For DF, the bijections could involve reversing orientations on
parts of the design; see discussion at (3.2.1).
Disjoint union. The fields preserve monoidal structure: there is an identification
and these identifications are compatible with morphisms, orientation reversal, and
boundaries in the obvious ways. Note that this implies that C(∅) consists of a single
element.
3.2.2 Gluing without corners. Given an identification
∂X = Y ⊔ −Y ⊔ W
Using the boundary and orientation reversal maps, we get two maps from C(X) to
C(Y ).
∂ / C(∂X) pr / C(Y )
EqY (C(X)) / C(X)
JJ O
JJ pr
JJ −
JJ
J$
C(−Y )
Let EqY (C(X)) denote the equalizer of these two maps. (That is, the set of all fields
in C(X) on which the two maps agree. In the case where Y and −Y lie in separate
components of X, EqY (C(X)) is a fibered product.) Then there is an injection
gl : EqY (C(X)) → C(glY (X)) such that the following diagram commutes.
/ C(X) ∂ / C(∂X)
EqY (C(X))
gl prW
∂
C(glY (X)) / C(W )
We will often identify EqY (C(X)) is its image gl(EqY (C(X))) ⊂ C(glY (X)). If
X = X1 ⊔ X2 , Y ⊂ ∂X1 and −Y ⊂ ∂X2 , then we will often use the notation α1 ∪ α2
instead of gl((α1 , α2 )). (Here αi ⊂ C(Xi ).)
Furthermore, we require that for any x ∈ C(glY (X)) and any neighborhood N
of the image of Y in glY (X) there exists a homeomorphism f of glY (X), isotopic
to the identity and supported in N , such that f∗ (x) ∈ EqY (C(X)). In other words,
any field on the glued manifold is close to a field obtained by gluing.
[Need to define restriction c|S for codim-0 submanifold S.]
Examples. For MF, EqY (MF(X)) is all of MF(glY (X)). For S, EqY (S(X))
consists of all submanifolds of glY (X) which are transverse to the image of the
gluing region Y in glY (X).
Gluing with corners. Assume an identification
∂X = Y ∪ −Y ∪ W,
3.3. EXTENDED ISOTOPY 23
∂
C(X) / C(∂X)
O O
gl
?
EqY (C ⋔(X))
/ C ⋔(X) ∂ / Eq (C(Y ⊔ −Y ⊔ W )) pr / C(Y )
S RRR O
RRR pr
RRR
RRR −
R(
C(−Y )
Let Eq∂Y (C(W )) ⊂ C(W ) be as in (3.2.2), for the gluing W → gl∂Y (W ) = ∂(glY (X)).
There is a map ∂ : EqY (C ⋔(X)) → Eq∂Y (C(W )) induced by ∂ : C ⋔(X) → C(∂X).
The gluing with corners axiom requires that there is an injection gl : EqY (C ⋔(X)) →
C(glY (X)) such that the following diagram commutes.
∂
EqY (C ⋔(X)) / Eq (C(W ))
∂Y
gl gl
∂
C(glY (X)) / C(gl (W ))
∂Y
Furthermore, we require that for any x ∈ C ⋔(X) and any neighborhood N of the
image of Y in glY (X) there exists a homeomorphism f of glY (X), isotopic (rel
boundary) to the identity and supported in N , such that f∗ (x) ∈ EqY (C ⋔(X)). In
other words, any field on the glued manifold whose boundary is obtained by gluing
is close to a field obtained by gluing.
As before, we will sometimes use the notation α1 ∪α2 instead of gl((α1 , α2 )) when
we are gluing distinct components of X together. The subspace EqY (C ⋔(X)) ⊂
C(glY (X)) is called the gluing subspace relative to this decomposition of glY (X).
Note that the gluing without corners axiom (3.2.2) is a special case of this one.
• note that this is a local relation, since both isotopy and collars can be localized
(need further field assumptions here? (for collars))
• purpose of this section: do some tedious proofs here so as not to interrupt the
flow (w?) in later chapters
to do:
25
26 CHAPTER 4. BASIC CONSTRUCTIONS AND GLUING I
• If a, b ∈ C(B; c) are related by an isotopy, f (a) = f (b) for all f ∈ Z(B; c).
Hence M
e )⊇
A(Y e
A(B; e ′; b
c) ⊗ A(Y c).
c∈C(∂B)
4.2. THE BASIC CONSTRUCTION IN CODIMENSION 1 27
P
c2
c1
c0 B Y
Let Y
′
ZB ⊆ e ′; b
Z(B; c) ⊗ Z(Y e ).
c) ⊂ Z(Y
c∈C(∂B)
Define \
′
Z(Y ) = ZB ,
B
where B ranges over all subballs of Y . [need to revise notation above; ⊗ is not the
right thing; fix below too]
More generally, let Y be an n-manifold with boundary, c ∈ C(∂Y ), and B ⊂ Y be
a subball, possibly intersecting ∂Y . Let S0 = ∂B ∩ ∂Y , ∂B = S0 ∪ S1 , ∂Y = S0 ∪ S2 ,
and P = ∂S0 = ∂S1 = ∂S2 . (See Figure (4.2.1).) Let c = c0 ∪ c2 with respect to
the decomposition ∂Y = S0 ∪ S2 , and let ∂c0 = ∂c2 = d ∈ C(P ). Let Y = Y ′ ∪ B,
so that ∂Y ′ = (−S1 ) ∪ S2 . Then we have decompositions
[
C(Y ; c) ⊇ C(B; c0 ∪ c1 ) × C(Y ′ ; cb1 ∪ c2 )
c1 ∈C(S1 ;d)
and M
e ; c) ⊇
A(Y e
A(B; e ′ ; cb1 ∪ c2 ).
c0 ∪ c1 ) ⊗ A(Y
c1 ∈C(S1 ;d)
Let Y
′
ZB ⊆ e ′ ; cb1 ∪ c2 ) ⊂ Z(Y
Z(B; c0 ∪ c1 ) ⊗ Z(Y e ; c).
c1 ∈C(S1 ;d)
As before, define \
′
Z(Y ; c) = ZB ,
B
More generally, if S has boundary and c ∈ C(∂S) we can define a category A(S; c)
with objects C(S; c) and morphisms A(S × I; b a, b), where a, b ∈ C(S; c). Here, as
usual, S × I denotes a cylinder with the vertical boundary ∂S × I pinched, so that
∂(S×I) is naturally identified with (−S)∪S (the double of S) and b a ∪b ∈ C(∂(S×I)).
If S is the empty n−1-manifold, then A(S) is the trivial category consisting of
a single object and (multiples of) its identity morphism.
There is a natural identification
denote this representation (collection of vector spaces plus left action of A(∂Y )) by
Z(Y ).
For any closed n−1-manifold S define Z(S) = Rep(A(S)), the category whose
objects are left representations of A(S) and whose morphisms are natural transfor-
mations (also called intertwiners in this context). Then we have, for all n-manifolds
Y,
4.3.2 Z(Y ) ∈ Z(∂Y ).
This is the categorified version of (1.2.1) and (6.1.1).
4.3.3 Assume now that A(∂Y ) is semisimple (see Appendix B) and let α be a right
irrep of A(∂Y ). Define
A(Y ; α) = mor(α, A(Y )),
the space of intertwiners from α to A(Y ). Let L be a complete set of irreps for
A(∂Y ). Then there is a canonical isomorphism
M
A(Y ) ∼
= A(Y ; α) ⊗ α.
α∈L
A(Y ; α) ∼
= A(Y ; b)eα .
(where now L denotes a complete set of left irreps of A(∂Y )). Note: In the old-
fashioned axiomatization of TQFTs (see, e.g., [Walk91]) these are the spaces asso-
ciated to surfaces with labeled boundary.
−S +S
R R
Y Ygl
Our goal is to describe A(Ygl ; c) in terms of the various A(Y ; ·, ·, c) and the action
of A(−S ⊔ S) = A(S)op × A(S) on these spaces.
We’ll start with the most abstract formulation of the codimension-1 gluing the-
orem, and then work our way toward more concrete statements.
Theorem. Let Y , S, Ygl , c be as above. 4.4.2
(a) For each object x of A(S) there is a map
b, x, c) → A(Ygl ; c).
glx : A(Y ; x
b, x, c)
A(Y ; x
e×1nnnn
nn7 NNN
NNNglx
n NNN
nn NN'
nnn
A(Y ; yb, x, c) A(Ygl ; c)
PPP 7
PPP
ppppp
PPP
1×e PPP' ppp
ppp gly
A(Y ; yb, y, c)
(c) A(Ygl ; c) is the universal object (vector space, set, or whatever flavor of A
we’re using) with properties (a) and (b). In other words, given a W and maps
gl′x : A(Y ; x
b, x, c) → W (for all x) such that the diagram analogous to the one in (b)
above commutes for all e, there is a unique θ : A(Ygl ; c) → W such that gl′x = θ ◦ glx
for all x.
b, x, c)
A(Y ; x ′
e×1nnnn
nn7 NNN
NNN
glx
NN
nnn glx NNN'
nnn '
A(Y ; yb, x, c) A(Ygl ; c) _θ _ _7/ W
PPP p 7
PPP glyppp
PPP pp
1×e PPP' pp
ppp gl′y
A(Y ; yb, y, c)
In other words, A(Ygl ; c) is the coend (see Appendix A [need more specific reference])
of the action of A(S)op × A(S) on A(Y ; ·, ·, c).
32 CHAPTER 4. BASIC CONSTRUCTIONS AND GLUING I
Proof. Part (a) is obvious; fields which agree on ±S can be glued to yield a field on
Ygl , and all local relations on Y are also local relations on Ygl .
Part (b) follows from the fact that shifting a collar across the gluing submanifold
changes the field by an isotopy. The top arrows in the diagram correspond to gluing
e to −S, and then gluing −S to S to obtain a field on Ygl . The bottom arrows
correspond to gluing e to S first. The two fields thus obtained differ by an isotopy
of Ygl which shifts a collar across ±S, so they represent the same element of A(Ygl ; c)
and the diagram commutes.
Part (c) requires a little more work. Let g ∈ C(Ygl ; c) and ḡ be the equivalence
class of g in A(Ygl ; c). After an isotopy, we may assume that g is transverse to
the gluing submanifold (image of ±S) in Ygl . Then ḡ = glx (g♯ ) for some g♯ ∈
A(Y ; x b, x, c), where x is g restricted to ±S. Necessarily, we define θ(ḡ) = gl′x (g♯ ).
Thus θ is unique if it exists.
It remains to be shown that the above procedure for defining θ yields well-defined
results. We must show that if g and h are equal in A(Ygl ; c), then gl′x (g♯ ) = gl′y (h♯ )
(where y is h restricted to ±S). If g and h are locally related with respect to a ball
B ⊂ Ygl disjoint from ±S, then g♯ = h♯ , so a fortiori gl′x (g♯ ) = gl′x (h♯ ). If g and h
are related by a collar shift isotopy along ±S, then gl′x (g♯ ) = gl′y (h♯ ) by the assumed
commutativity of the above diagram for W and gl′ . Since collar shift isotopies plus
isotopies supported in balls disjoint from ±S generate arbitrary isotopies on Ygl ,
gl′x (g♯ ) = gl′y (h♯ ) if g and h are related by an arbitrary isotopy. Finally, suppose g
and h are locally related with respect to a ball B not disjoint from ±S. Let ϕ be
an isotopy of Ygl which moves B off of ±S. Note that ϕ(g)♯ is related to ϕ(h)♯ by a
local relation supported in ϕ(B). Then gl′x (g♯ ) = gl′w (ϕ(g)♯ ) = gl′z (ϕ(h)♯ ) = gl′y (h♯ ),
where w and z are the restrictions of ϕ(g) and ϕ(h) to ±S.
There is, of course, and dual version of the gluing theorem for Z:
rx : Z(Ygl ; c) → Z(Y ; x
b, x, c).
b, x, c)
Z(Y ; x
nn gNNN
e×1nnnn NNNrx
n NNN
nn NN
nw nn
Z(Y ; yb, x, c) Z(Ygl ; c)
gPPP
PPP ppp
PPP pppp
1×eop PPP p r
pw pp y
Z(Y ; yb, y, c)
4.4. GLUING CODIMENSION-1 MANIFOLDS 33
(c) Z(Ygl ; c) is the universal object with properties (a) and (b). In other words,
given a W and maps rx′ : W → Z(Y ; x b, x, c) (for all x) such that the diagram
analogous to the one in (b) above commutes for all e, there is a unique θ : W →
Z(Ygl ; c) such that rx′ = rx ◦ θ for all x.
b, x, c) l
Z(Y ; x
nn gNNN rx′
e×1nnnn NNN
N
nnn rx NNNN
wnnn
Z(Y ; yb, x, c) Z(Ygl ; c) o_θ _ _ W
gPPP p
PPP ry ppp
PPP ppp
1×e PPP p
pw pp r ry′
Z(Y ; yb, y, c)
In other words, Z(Ygl ; c) is the end (See Appendix A [need more specific reference])
of the action of A(S)op × A(S) on Z(Y ; ·, ·, c).
The proof is dual to the proof of (4.4.2). The map rx : Z(Ygl ; c) → Z(Y ; x b, x, c) is
b, x, c)) of C(Ygl ; c).
given by restricting a function in Z(Ygl ; c) to the subset glx (C(Y ; x
In the remainder of this section we state a number of special cases of the
codimension-1 gluing theorem.
Corollary. Let Y , S, Ygl , c be as above. If the target category of A (on n-manifolds) 4.4.4
is the category of vector spaces, then
M
A(Ygl ; c) = A(Y ; xb, x, c) /hev ∼ vei.
x∈C(S)
L
By hev ∼ vei we mean the subspace of x∈C(S) A(Y ; x b, x, c) generated by all ev − ve,
for all morphisms e : x → y of A(S) and all v ∈ A(Y ; y, x, c). Here we write the
action of A(S) as juxtaposition on the right and the action of A(S)op as juxtaposition
on the left.
L
Note that above means finite linear combinations.
Corollary. Let Y , S, Ygl , c be as above. If the target category of Z (on n-manifolds)
is vector spaces, then
Y
Z(Ygl ; c) = {(vx ) ∈ b, x, c) | vx e = evy for all e}.
Z(Y ; x
x∈C(S) 4.4.5
(See Appendix A for the definition of ⊗S .) In other words, we have a functor from
the category of n-dimensional bordisms to the category of bimodules.
(Some authors use the above functorial property as an axiom for the codimension-
1 gluing properties of TQFTs, but note that without auxiliary axioms it fails to
handle the case of gluing an n-manifold to itself, nor does it cover interrelationships
between different bordisms with the same underlying manifold (i.e. moving a part
of the incoming boundary to the outgoing boundary, or vice-versa).)
Let Y , S, R, Ygl , c be as before. Assume now that A(S) and A(R) are semisimple
categories. Let L be a complete set of irreps for A(S) and K be a complete set of
irreps for A(R). Then
M
A(Y ) = A(Y ; α∗ , β, γ) ⊗ α∗ ⊗ β ⊗ γ,
α,β,γ
where the sum is over α, β ∈ L and γ ∈ K. Note that the dual representation α∗ is
a right irrep of A(S)op = A(−S), and so α∗ ⊗ β ⊗ γ runs through a complete set of
irreps for A(∂Y ). (See (4.3.3) above.) Also
M
A(Ygl ) = A(Ygl ; γ) ⊗ γ.
γ∈K
−S
+S
P
−P
+P
R Rgl
R Y Ygl
Rgl
Theorem. A(Ygl ; c) is the coend of the action of A(S; z)op ×A(S; z) on A(Y ; ·, ·, c♯ ). 4.4.9
(a) For each object x of A(S; z) there is a map
b, x, c♯ ) → A(Ygl ; c).
glx : A(Y ; x
b, x, c♯ )
A(Y ; x
nn6
e×1nnnn
NNN
NNNglx
nnn NNN
nnn NN'
A(Y ; yb, x, c♯ ) A(Ygl ; c)
PPP p7
PPP
PPP ppppp
1×e PPP( pp
ppp gly
A(Y ; yb, y, c♯ )
(c) A(Ygl ; c) is the universal object (vector space, set, or whatever flavor of A
we’re using) with properties (a) and (b). In other words, given a W and maps
gl′x : A(Y ; x
b, x, c♯ ) → W (for all x) such that the diagram analogous to the one
in (b) above commutes for all e, there is a unique θ : A(Ygl ; c) → W such that
36 CHAPTER 4. BASIC CONSTRUCTIONS AND GLUING I
b, x, c♯ )
A(Y ; x
n6 NNN gl′x
e×1nnnnn NNN
nnn NN
nnn glx NNN'
'
A(Y ; yb, x, c♯ ) A(Ygl ; c) _θ _ _/7 W
PPP pp 7
PPP gly
PPP ppppp
1×e PPP( p
ppp gl′ y
A(Y ; yb, y, c♯ )
The proof does not differ in any interesting way from the proof of (4.4.2). All of
the other results in this section generalize easily to gluing n-manifolds with corners.
Note, however, that (4.4.9) is a somewhat unsatisfactory result. We want to
know not only A(Ygl ; c) for various c ∈ C(∂Ygl ), but also the action of A(∂Ygl ) on
A(Ygl ) (and also Z(Ygl )). For gluing without corners, ∂Ygl is a closed submanifold of
∂Y and knowledge of the action of A(∂Y ) on A(Y ) translates easily into knowledge
of the action of A(∂Ygl ) on A(Ygl ). For gluing with corners, this is not the case. In
order to state a more powerful result on gluing n-manifolds with corners, we first
need to understand gluing n−1-manifolds without corners. That is, we need to be
able to describe the category A(Rgl ) is terms of the categories A(R; zb, z) and the
action of the 2-category A(P ) on them. This is done in Chapter 5.
• YYYYY: insert examples into later sections (A and Z defs, cylinder cats, ...)
• ?? retain old manifolds → cat → n-vect space thing? put it in some other
chapter?
Chapter 5
This Chapter is analogous to Chapter 4, but one dimension lower and hence one
category level higher. We show that n−1-manifolds with boundary give rise to col-
lections of 1-categories; that these collections afford a representation of a 2-category
associated to the n−2-dimensional boundary of the n−1-manifold; and that these
2-category actions can be used to state and prove gluing theorems for A and Z of
1-manifolds.
[need to finish this intro once chapter is complete]
[WARNING: in order to be consistent with other chapters, need to switch right and
left actions, and modify most of the figures]
37
38 CHAPTER 5. BASIC CONSTRUCTIONS AND GLUING II
• [need to decide whether other type of composition of 2-mors is part of the data
(as opposed to derived)]
x y y
e u
x y
eu
y v
g
u
x
eg
e×I
denoted e) from A(S, y) to A(S, x). On objects this functor is given by gluing a
copy of e to S along ∂S (i.e. we add a collar to ∂S containing the field e). On
morphisms this functor is given by gluing a copy of e × I to S × I along ∂S × I.
(This involves unpinching parts of the boundaries; see [need ref for this? or is it
clear?].) It’s clear that composition is preserved, so this defines a functor. (See
Figure (5.1.1).)
For e ∈ Cxy1 and f ∈ C 1 we need an invertible natural transformation c
yz ef con-
necting the composition of the actions for e and f with the action of the composition
ef . This is given by the track of an isotopy, supported in a collar neighborhood of
∂S, connecting the corresponding (compositions of) collaring homeomorphisms.
For h ∈ Cef2 = A(∂S × I × I; e, f ) we need a natural transformation between the
functors for e and f . This is given, at u ∈ C(S; y), by gluing a copy of h to u × I (see
Figure (5.1.2)). The proof that this collection of morphisms actually does comprise
a natural transformation, that is, that this diagram
em
eu / ev
hu hv
fu / fv
fm
f
h
u
e
hu
h u×I
fv fv
fm hv
fu = ev
hu em
eu eu
the field defs; clearly works for our basic examples (pictures and maps); or, better,
do the proof in the fields section/chapter]
For e ∈ Cxy 1 , u ∈ C(S; x) and v ∈ C(S; y). we need a bijection (isomorphism)
ē u u
↔
v e v
−P x P P
u glx (u)
P ×I
m glx (m)
to be the 2-category of all right representations of A(P ) [need to say what sort of
2-cat this is and why], then we have, for all n−1-manifolds S,
5.2 Gluing
Next we use the above representations to prove gluing theorems for A and Z of
n−1-manifolds. Suppose we have the usual gluing (without corners) scenario: an
n−1-manifold S with an identification ∂S = (−P )⊔P ⊔Q. For notational simplicity
we will henceforth ignore Q. We have an action of the 2-category A(−P ⊔ P ) ∼ =
op
A(P ) × A(P ) on {A(S; ·, ·)} (equivalently, commuting right and left actions of
A(P )), and we would like to compute the 1-category A(Sgl ) from this action.
We will start by noting some relationships between A(Sgl ) and A(S). We will
then prove that A(Sgl ) is the universal 1-category with these properties (what one
might (and we will) call a categorified coend or 2-coend construction).
To simplify notation (and also allow for its reuse), denote the 2-category A(P )
by A, denote the 1-category A(S; x, y) by Wxy (where x, y ∈ A0 = C(P )), and denote
the 1-category A(Sgl ) by C.
For each x ∈ A0 we have a functor
−P P
x y x y
e
u
P ×I
gly (ue)
sex
glx (eu)
5.2.4 se at x
For each e ∈ A1xy we can construct two functors from Wxy to C, gly ◦(e × 1) and
(1 × e) ◦ glx , and there is an invertible natural transformation se between these two
functors:
Wyy
y< BB
BBgly
e×1 yyy
BB
yy BB
yy
Wxy se =C
FF ||
FF |
FF |
1×e FF ||
" || glx
5.2.3 Wxx
(1 × h) • idglx
sf
=
se
(h × 1) • idgly
1 Wyy sf
I >>
f ×1 >>
>>
>> gl
>> y
h×1 >>
>>
>>
e×1 >>
Wxy se @C
1×e
1×h glx
1×f
.W 5.2.5
xx
f e fe
1 × cf e
sf e
= se
cf e × 1
sf
.W
(f e)×1 x < zz==
x ==
cef × 1 xx x ==
xxx f ×1 == sef
==glz
Wyz ==
y< EE
EE1×f sf ==
y yy EE ==
y E ==
y e×1 E"
yy gly
=
WxzE id Wyy /C
@
EE y <
EE1×f yyy
EE yy
E" yy e×1 se
Wxy
FF glx
FF1×e
1 × cef FF
FF
1×f e "
5.2.7 0 Wxx
Definition. Let A be an arbitrary disk-like 2-category, and let {Wxy } be a collection 5.2.9
of 1-categories affording an Aop × A action. A 1-category C, together with functors
{glx } and invertible natural transformations {se } satisfying (5.2.1), (5.2.3), (5.2.5)
and (5.2.7), is called the 2-coend of the Aop × A action if it is universal in the
following sense. If C ′ , {gl′x } and {s′e } also satisfy (5.2.1) through (5.2.7), then
there exists a functor θ : C → C ′ and, for all x ∈ A0 , a natural transformation
ηx : θ ◦ glx → gl′x , such that
gl′y
Wyy
y< BB s′e
e×1 yyy BB ηy
BB
yy gly BB
yy
θ
Wxy se =C
/ C′
FF H
FF glx ||
FF ||
1×e FF || ηx
" ||
Wxx
gl′x
Theorem. A(Sgl ) is the 2-coend of the A(P )op × A(P ) action on {A(S; x, y)}. 5.2.10
Proof. We will introduce two new categories, G (G for generators and relations) and
P (P for parallelogram). Both will have concrete algebraic descriptions. It will be
easy to show that G is (up to natural isomorphism, of course) the 2-coend of the
theorem, and that P is isomorphic to A(Sgl ). We then show that G and P are
isomorphic. (So G and P provide two alternative, more concrete descriptions of the
2-coend.)
S 0 . The morphisms will be
First we define G. The objects of G are x∈A0 Wxx
defined
S in terms of generators and relations. There are two types of generators,
1 0
x∈A0 Wxx (with the obvious choice of range and domain), and, for all x, y ∈ A ,
1 0
e ∈ Axy and u ∈ Wyx , morphisms
σeu : eu → ue
−1
σeu : ue → eu
5.2.13 • σe· is a natural transformation between the right and left actions of e ∈ A1xy :
c σev = σeu fce
for all f ∈ (Wyx )1uv , ef
Next we show that G is the 2-coend of the A × Aop action. There are obvious
functors glx : Wxx → G for all x (sending f to fb), and, with σe· playing the role of
se , the above relations for G guarantee that these data satisfy (5.2.1), (5.2.3), (5.2.5)
and (5.2.7). Let C ′ , {gl′x } and {s′e } be as in the definition of 2-coend. We must
define a functor θ : G → C ′ satisfying the conditions of (5.2.9). On the objects of G
define θ so that θ ◦ glx = gl′x for all x. On the generating morphisms coming from
Wxx again define θ so that θ ◦ glx = gl′x for all x. Finally, define θ(σeu ) = s′e,u and
θ(σeu−1 ) = (s′ )−1 .
e u
We must show that the above definition of θ on the generators of G 1 respects
the relations. Since the relations are just translations of the commutative diagrams
defining the 2-coend, this is easy to do. (5.2.11) follows from the fact that gl′x is
a functor. (5.2.12) follows from the fact that s′e and (s′e )−1 are mutually inverse.
(5.2.13) follows from the fact that s′e is a natural transformation. (5.2.14) follows
from (5.2.5). (5.2.15) follows from (5.2.7). Thus θ is a well-defined functor.
Finally, we must define, for all x, an invertible natural transformation ηx : gl′x →
θ ◦ glx . Since these two functors are equal, we can take ηx to be the identity natural
transformation. S
Next we define P. The objects of P are again x∈A0 Wxx 0 . The morphisms from
0
u ∈ Wxx to v ∈ Wyy are0
,
M
mor(eu, ve) p ◦ (h • idu ) ∼ (idv •h) ◦ p
5.2.16 e∈A1yx
Here mor(eu, ve) = (Wyx )1eu,ve and h ∈ A2ef . The equivalence class of a ∈ mor(eu, ve)
1 . We think of a morphism of P as a parallelogram with
will be denoted gle (a) ∈ Puv
short sides labeled by e and long sides labeled by u and v. See Figure (5.2.17).
(More literally, a morphism of P can be thought of as a field on Sgl × I restricting
to e on P± × I.)
[need to replace most (all?) occurrences of idblah • with simple juxtaposition]
To complete the definition of P we must specify how to compose morphisms. Let
p ∈ mor(eu, ve) and q ∈ mor(f v, wf ), where e ∈ A1xy , f ∈ A1zx , u ∈ Wxx
0 , v ∈ W0 ,
yy
0 . Then define
w ∈ Wzz
h
e
v v h
h e ∼ f
e f
u u
(There should be some associators inserted above, but they have been omitted for
simplicity.) See Figure (5.2.18).
The objects of P are essentially all of the objects of A(Sgl ). (More precisely,
they are all of the objects which are transverse to the gluing surface.) It follow from
(4.4.4) that the morphisms of the two categories are identical. So all that remains
is to show that the two composition rules coincide. This is illustrated in Figure
(5.2.19). The basic idea is that the equivalence relation on morphisms of P allows
us to eliminate ide and idf from the definition of composition in P, yielding the
result of composing in A(Sgl ).
Now we must show that G and P are isomorphic. Predictably, we define functors
α : G → P and β : P → G and show that α ◦ β is the identity functor on P and β ◦ α
is the identity functor on G. Note that G 0 = P 0 , so we can (and do) define both α
and β to be the identity on objects.
First some notation. For e ∈ A1xy , let Vee∗ ∈ A2 be the 2-morphism conjugate to
ide with domain idx and range ee∗ . Let Λee∗ ∈ A2 be the 2-morphism conjugate to
ide with domain ee∗ and range idx . See Figure (5.2.20). [put these defs earlier and
just recall them here??]
Keeping the topological interpretation of the algebra in mind, it’s easy to define
−1 ) = gl (id 1
α. Let α(σeu e eue ) and α(σeu ) = gle∗ (Λe∗ e • idu •Vee∗ ). For h ∈ (Wxx )uv ,
let α(b
h) = glidx (h̃), where h̃ is the morphism from idx •u to v • idx guaranteed by
the definition of 2-category action. [need to make sure we included this in the def]
See Figure (5.2.21). We must verify that the above assignments of generators of
G 1 obey the relations. This straightforward exercise is left to the reader.
Defining β : P → G is a little more complicated. The key topological idea is that
an arbitrary field on Sgl × I is isotopic to a composition of fields (morphisms) which
either: (a) are product fields near P± × I ⊂ Sgl × I, and so are gluings of fields on
S × I; or (b) are tracks of isotopies of Sgl which shift a collar neighborhood across
P± . The proof of this is illustrated in Figure (5.2.22). Accordingly, we define, for
48 CHAPTER 5. BASIC CONSTRUCTIONS AND GLUING II
w v
q f p e
f e
v u
e×I
qe fp
f ×I
w
f
f qe
e fp e
5.2.18 Composition in P
w w
f f
f q
f
= e
e e
e
u u
x y
e
x×I e ē
Λeē Veē
e ē x×I
e u
e u
−1 e
u α(σeu )
e e
α(σeu ) ē e
u
ē
e u
v
x x
h
v u
x×I
α(h)
x×I u
5.2.21 Definition of α : G → P
a ∈ (Wyx )1eu,ve ,
β(gle (a)) = (id\ b′
v •Λee∗ ) ◦ σe∗ ,ve ◦ a ,
where a′ is the adjoint of a with domain u and range e∗ ve. [here we ignore some
morphisms coming from the action of identity 1-morphisms of A; need to say this
more precisely; or maybe include those morphisms]
Now that α and β have been defined, we must verify that they are mutually
inverse. [relatively easy for P → G → P; harder for G → P → G; proofs illustrated
in figs xxxx, yyyy, zzz]
Wyy sf
>^ >
f ×1 >>
>>
>> r
>> y
h×1 >>
>>
>>
e×1 >>
Wxy nq se C
U
1×e
1×h rx
1×f
Wxx 5.2.25
Also, for all e ∈ A1xy and f ∈ A1yz ,
(f e)×1
Wzz=^
xx ==
cef × 1 xxx ==
==
x
|xx f ×1 == r sef
Wyz == z
bEE sf ==
yyy EE1×f ==
yye×1 EE ==
yy EE ==
|y ry
Wxz
P bEE id Wyy o C
EE1×f yyy
EE y
EE yy
y| y e×1 se
Wxy
Fb F rx
FF1×e
1 × cef FF
FF
1×f e
Wxx 5.2.26
h•id id •h
sf,a
f ay / ax f 5.2.27
52 CHAPTER 5. BASIC CONSTRUCTIONS AND GLUING II
It follows from (5.2.26) that for all e ∈ A1xy and f ∈ A1yz the following diagram
commutes
sef,a
(ef )az / ax (ef )
cL
ef cR
ef
esf,a se,a f
5.2.28 a(f az ) / eay f / (ax e)f
where cL R
ef [cef ] is the associator associated to the left [right] action of A on W .
def
For each morphism m : a → b of C ′ we have a collection of morphisms {mx =
′
rx (m) : ax → bx } (indexed by x ∈ A0 ). Because se is a natural transformation, the
following diagram commutes for all e ∈ A1
se,a
eay / ax e
emy mx e
se,b
5.2.29 eby / bx e
We are thus led to define a 1-category E whose objects are collections {ax ∈
0 } and {s
Wxx e,a : eay → ax e} satisfying (5.2.27) and (5.2.28). Morphisms of E are
collections {mx : ax → bx } satisfying (5.2.29). Composition is given by {mx } ◦
def
{nx } = {mx ◦ nx }. (It is easy to verify that {mx ◦ nx } satisfies (5.2.29).)
There are obvious functors rx : E → Wxx (for all x ∈ A0 ), and the individual
morphisms se,a fit together to give natural transformations se as in (5.2.24). [no-
tational problem here: se,a plays two roles, one from def of E and one from def of
2-end. should fix this.] The above discussion motivating the definition of E can also
be read as a proof that E, {rx } and {se } are universal in the appropriate sense.
Thus we have the desired concrete description of the 2-end.
One of the simplest examples of a 2-end is when A above is a 2-category with
only one object (i.e. a spherical tensor 1-category) and Aop × A acts on itself via left
and right tensor multiplication. The 2-end of this action has a more familiar name:
the Drinfeld center of the tensor 1-category A. Indeed, if we specialize the definition
of E above to this case we obtain the usual definition of the Drinfeld center. (See
for example [Kas95, p. 330]. Note however that we make no assumptions about
strict associativity.) [D. center works for general tensor 1-cat (don’t need spherical).
comment on this?]
5.2.30 So a special case of the codimension-2 gluing theorem (5.2.23) is the following.
Suppose P is a closed n−2-manifold such that the 2-category A(P ) has only one
object, and thus can be thought of as a tensor 1-category. Then Z(P ), the repre-
sentations of A(P ), can also be thought of as a tensor 1-category, and Z(P × S 1 ) is
the Drinfeld center of Z(P ). [need to say more here; need additional assumptions
on A(P )]
5.2. GLUING 53
Note that for any closed P the category Z(P × S 1 ) has the structure of a braided
tensor category. Let X 2 be a twice-punctured disk. Then Z(P ×X) can be thought of
as a functor Z(P × S 1 ) × Z(P × S 1 ) → Z(P × S 1 ), which gives a tensor (monoidal)
structure to Z(P × S 1 ). The geometric braiding of X gives rise to an algebraic
braiding of Z(P × S 1 ). [need to show that in special case of D. center this def of
tensor and braiding agrees with the usual one.] [need to go into more detail on
above]
Still to do:
• annularization
• in earlier (fields) section, prove a few relevant isotopy identities carefully, using
field axioms (and refer to that from this chapter)
54 CHAPTER 5. BASIC CONSTRUCTIONS AND GLUING II
Chapter 6
First, the path integral Z(M n+1 ) is a function on fields on its boundary com- 6.1.1
patible with local relations, so
Z(M ) ∈ Z(∂M )
55
56 CHAPTER 6. FROM LOCAL RELATIONS TO PATH INTEGRALS
for all n+1-manifolds M . (If ∂M = ∅, this means that Z(M ) ∈ C, or whatever the
ground ring is.) Equivalently, Z(M ) is a function
Z(M ) : A(∂M ) → C.
6.1.2 Second, for all n-manifolds Y and boundary conditions c ∈ C(∂Y ), we have
nondegenerate pairings
c) → C
Z(Y ; c) ⊗ Z(−Y ; b
x ⊗ y 7→ hx, yi
and
c) → C
A(Y ; c) ⊗ A(−Y ; b
x ⊗ y 7→ hx, yi.
The pairings should be compatible with the actions of A(∂Y ): for all x ∈ A(Y ; c),
y ∈ A(−Y ; bb) and e ∈ A(∂Y )1cb = A(∂Y × I; b
c, b),
and similarly for Z(Y ; c). (Here we are using the identification A(−∂Y ) = A(∂Y )op ,
so A(∂Y ) has a left action on A(−Y ; ·).) More generally, we could replace ∂Y above
with a codimension-0 submanifold of ∂Y .
Recall that associated to reversing the orientation of Y we have an isomorphism
(conjugate linear if the ground ring is C) A(Y ; c) ∼ = A(−Y ; b c). Combining this
isomorphism with the above pairings we get [sesquilinear] inner products
A(Y ; c) ⊗ A(Y ; c) → C
def
x ⊗ y 7→ hx, yi = hx, ybi
and similarly
Z(Y ; c) ⊗ Z(Y ; c) → C
def
x ⊗ y 7→ hx, yi = hx, ybi.
(Note that we have “overloaded” the angle brackets h·, ·i to denote both the pairings
and the inner products. Which meaning is intended can be deduced from what’s
being plugged into the angle brackets.) We require these inner products to be [skew]
symmetric. The inner products induce isomorphisms between A(Y ; c) and Z(Y ; c)
(recall that these spaces are mutually dual), and these isomorphisms should preserve
the pairings.
(In all of the examples we will study, A(Y ; c) and Z(Y ; c) will be finite dimen-
sional, so we won’t worry about completeness.)
6.1. AXIOMS FOR PATH INTEGRALS 57
−Y +Y
W Wgl
W M Mgl
Wgl
cr ctr
Z(∂M ) −→ Z(Y ; c) ⊗ Z(−Y ; b
c) ⊗ Z(W ; b
c, c) −→ Z(W ; b
c, c),
where rc is a restriction map and trc comes from the pairing for Z(Y ; c). Call the
composite map ϕc : Z(∂M ) → Z(W ; b c, c). Recall from (4.4.5) that
Y
Z(Wgl ) = {(vx ) ∈ b, x) | vx e = evy for all e ∈ A(∂(Y ))1 }.
Z(Y ; x
x∈C(∂Y )
Then the gluing relation for n+1-manifolds states that (ϕx (Z(M ))) determines an
element of Z(Wgl ) (i.e. ϕx (Z(M ))e = eϕy (Z(M )) for all e ∈ A(∂(Y ))1xy ) and that
Z(Mgl ) is equal to this element of Z(Wgl ) = Z(∂M ). More succinctly,
where trY : Z(∂M ) → Z(∂Mgl ) is induced by the various ϕc (see (6.2.2) below).
Here’s a more concrete version of the gluing relation. Assume that A(Y ; c) is 6.1.7
finite dimensional. Fix c ∈ C(∂Y ) and let x ∈ A(W ; b c, c). We want to evaluate the
function Z(Mgl ) on the glued field glc (x) ∈ A(∂Mgl ) in terms of Z(M ) evaluated on
elements of A(∂M ) = A(Y ∪ −Y ∪ W ). Let {ei } be a basis of A(Y ; c) and let {ei }
def def
be the dual basis of Z(Y ; c). Let gij = hei , ej i and gij = hei , ej i. (Note that the
matrices (gij ) and (gij ) are mutually inverse.) Then straightforward linear algebra
shows that the above gluing relation is equivalent to
X
Z(Mgl )(glc (x)) = Z(M )(ei ∪ ebj ∪ x)gij .
i,j 6.1.8
58 CHAPTER 6. FROM LOCAL RELATIONS TO PATH INTEGRALS
which implies that Z(Y × I)(ebj ∪ ei ) gives the inverse of the matrix gij . In other
words,
hei , ej i = Z(Y × I)(ebi ∪ ej ),
which implies (since the basis {ei } was arbitrary) that
for all x, y ∈ A(Y ; c). Note that if we use (6.2.1) to define the inner product then it
is automatically compatible with the boundary category actions as in (6.1.3):
6.2.2 If (6.2.1) holds for Y then the gluing formulas of (6.1.4) automatically yield
functions on C(∂Mgl ) which lie in Z(∂Mgl ). Let b, c ∈ C(∂Y ), m ∈ A(∂Y × I; bb, c)
6.2. CONSEQUENCES OF THE AXIOMS 59
x x
b b m c c
ek fi
M Y ×I M
el fj
b b m c c
x x
and x ∈ A(W ; b, b c). Let {ei } be a basis of A(Y ; b) with dual inner product matrix
gij . Let {fi } be a basis of A(Y ; c) with dual inner product matrix hij . (See Figure
(6.2.3).) Then
X
Z(Mgl )(glc (xm)) = Z(M )(fi ∪ fbj ∪ xm)hij
i,j
X
= Z(M )(ek ∪ fbj ∪ x)gkl hel , mfi ihij
i,j,k,l
X
= Z(M )(ek ∪ fbj ∪ x)gkl hel m, fi ihij
i,j,k,l
X
= Z(M )(ek ∪ ebl ∪ mx)gkl
k,l
= Z(Mgl )(glb (mx)).
It is also easy to see that if e, f ∈ A(R; c)1xx are minimal idempotents for two non-
isomorphic irreps, then he, f i = 0.
Next we describe how gluing n-manifolds affects the pairings. Let ∂Y = R ∪ 6.2.5
S ∪ −S, and b ∈ C(R). (For simplicity we are suppressing from the notation labels
for ∂S and ∂R = ∂S ⊔ ∂(−S).) We want to compute the pairing for A(Ygl ; gl(b))
60 CHAPTER 6. FROM LOCAL RELATIONS TO PATH INTEGRALS
M1 W3
Y31 M3
W1 Y12
Y23
M2
W2
(We continue to suppress from the notation c and its various restrictions.) It now
follows from (6.2.5) that
X
Z(M1 ∪ M2 ∪ M3 )(c) = Z(M1 ∪ M2 )(fbik
31
∪ fil23 ) · Z(M3 )(fbil23 ∪ fik
31
) · hei , ei i−1
i,k,l
X
= Z(M1 )(fbik
31
∪ fij12 ) · Z(M2 )(fbij12 ∪ fil23 ) · Z(M3 )(fbil23 ∪ fik
31
) · hei , ei i−1 .
i,j,k,l
[need to adjust alignment above] Note that the above expression is symmetric in
1,2,3. In other words, if we applied the gluing formula in a different order (say
by first gluing M1 and M3 , then gluing M2 to M1 ∪ M3 ), the final answer for
Z(M1 ∪ M2 ∪ M3 ) would not change. This self-consistency property of the gluing
formula will be used in the proof of (6.3.1).
While the above expression might look complicated, the idea is simple: Orthog-
onally decompose the three spaces A(Yxy ; ·) according to the minimal idempotents
(irreps) of A(R; b), then do the obvious tensorial contractions, but adjust them by
a correction factor hei , ei i−1 . There is an obvious generalization for any number of
Mi ’s glued together around a “corner” R.
[need to say something about total finiteness in an n-cat]
1. there exists z ∈ Z(S n ) such that the induced inner product A(B n ; c)⊗A(B n ; c) →
C given by a ⊗ b 7→ z(b a ∪ b) is positive definite for all c ∈ C(S n−1 ); and
Then there exists a unique path integral Z(M n+1 ) ∈ Z(∂M ) (for all n+1-manifolds
M ) satisfying the axioms of (6.1) and such that Z(B n+1 ) = z.
Proof. We will define Z(M ) by decomposing M into handles (each of which is home-
omorphic to B n+1 ) and applying the gluing formula (6.1.7). The challenge is to show
that this computation is independent of the choice of handle decomposition. We will
do this by inducting on the index of the handles.
First, some terminology. By a (k, l)-body we mean a k-dimensional manifold
equipped with a handle decomposition with all handles of index ≤ l. We will use the
same letter to denote both a (k, l)-body and the underlying k-dimensional manifold;
which meaning is intended should be clear from context. If X is a (k, l)-body then
X × I will denote the (k + 1, l)-body given by thickening all the handles (increasing
their dimension by one while leaving the index unchanged).
Our inductive hypotheses are
2. For each (n, i)-body Y , A(Y ; c) is finite-dimensional for all c ∈ C(∂Y ) and we
have defined a positive definite inner product on A(Y ; c) which is unchanged
by handle slides and handle cancellations of index ≤ i.
3. For each (n − 1, i)-body R, A(R; b) is semisimple with finitely many irreps for
all b ∈ C(∂R).
The hypotheses in the statement of the theorem imply the inductive hypotheses
for i = 0. Note that a (k, 0)-body is homeomorphic to a disjoint union of copies of
B k , and there are no handle slides or cancellations of dimension ≤ 0 to consider.
A(B n−1 ; c) is semisimple because of the positive definite inner product. It has finitely
many irreps because otherwise A(B n−1 × S 1 ; c × S 1 ) would be infinite dimensional
[refer to relevant gluing theorem].
We’ll verify the inductive hypotheses for i = 1, then i = 2, then the general case.
Let M be an (n + 1, 1)-body. Choose an ordering of the 1-handles, then use this
ordering to define Z(M ) ∈ Z(∂M ) by attaching each 1-handle in turn and applying
(6.1.7). (Here we use the fact that A(B n × S 0 ; c) is finite dimensional for all c.)
Since the attaching targets of the 1-handles are disjoint, this calculation is clearly
independent of the ordering of the 1-handles. Consider a slide of a 1-handle α over
a 1-handle β. If α come before β in the ordering then it is clear that the calculation
of Z(M ) using (6.1.7) is not affected, since if α is already attached then sliding β
is simply an isotopy of the attaching target for β. Therefore Z(M ) is unaffected by
handle slides. Next consider a 0-handle α which is canceled by a 1-handle β. Let
α′ be the 0-handle at the other end of β. Instead of gluing β simultaneously to α
and α′ , we can first glue β to α′ , then glue α to β. Since the gluing targets are
disjoint, the latter order of gluing yields the same computation of Z(M ) as the first.
But each of the gluings in the latter order is equivalent to attaching a collar to a
copy of B n in the boundary of an n+1-maanifold, and thus has no effect on Z(M ).
6.3. COMPUTING THE PATH INTEGRAL 63
def
Yαγ = α∩γ ∼
= Bn × S0
def
Yαβ = α∩β ∼
= B n−1 × B 1
def
Yβγ = β∩γ ∼
= B n−1 × B 1 .
64 CHAPTER 6. FROM LOCAL RELATIONS TO PATH INTEGRALS
Yαβ
Yαγ
Yβγ
∂γ
(See Figure (6.3.2).) Note that if we attach β to γ along Yβγ this is equivalent to
attaching a boundary collar to Yβγ ⊂ ∂γ and so has no effect on Z(γ). [need ref
for this] Similarly, attaching α to γ ∪ β along Yαγ ∪ Yαβ is equivalent to attaching a
boundary collar to Yαγ ∪ Yαβ ⊂ ∂(γ ∪ β), and so again does not affect the calculation
of Z(γ). (Yαγ ∪ Yαβ here has an obvious (n, 1)-body structure, and we use this
structure to define the inner product.) So it suffices to show that attaching α then
β yields the same result for Z and attaching first β then α. This follows from (6.2.6).
The proof that Z(M ) is independent of slides and cancelations is now complete.
Let Y be an (n, 2)-body. As above, since Y × I is an (n + 1, 2)-body, we can
use (6.2.1) to define an inner product on A(Y ; c). Handle slides and cancelations
for Y are mirrored by handle slides and cancelations for Y × I, so the inner product
is invariant under these changes of handle decomposition. By hypothesis the inner
product is positive for the handles (copies of B n ) with any boundary conditions.
The argument of (6.2.5) now shows that the inner product on A(Y ; c) is positive
definite.
Let R be an (n−1, 2)-body. Since R×I is an (n, 2)-body, the morphism spaces of
A(R; b) have positive definite inner products. The same argument as before shows
that these inner products are compatible with the action of A(R; b) on itself. It
follows from (B.5.3) that A(R; b) is semisimple. Again, A(R; b) has finitely many
irreps since otherwise A(R × S 1 ; b × S 1 ) would be infinite dimensional.
We have now established the inductive hypotheses for i = 2. The proof for
arbitrary i is very similar. The only part worth commenting on is the cancelation
of i- and i−1-handles.
Let α be an i−1-handle which is canceled by an i-handle β. Assume that α
immediately precedes β in the ordering, and let γ denote the n+1-manifold resulting
from all gluings which precede α in the ordering. Attaching α and then β to γ results
6.3. COMPUTING THE PATH INTEGRAL 65
Still to do:
• do general state model? only 4d? will need to make assumptions about general
n-cat for general case. maybe just do 4-dim’l case as generally as possible and
say that it’s clear this could be generalized
Chapter 7
1+1-dimensional Examples
67
68 CHAPTER 7. 1+1-DIMENSIONAL EXAMPLES
Chapter 8
2+1-dimensional Examples
• idempotents
• quotients
• 6j?
69
70 CHAPTER 8. 2+1-DIMENSIONAL EXAMPLES
1. isotopy,
2. reversing the orientation of an arc and changing the label to its dual,
3. replacing “identity” coupons with parallel arcs (note that this includes
cups and caps because of Frobenius reciprocity),
4. erasing arcs labeled by the trivial object (and adjusting labels of coupons
if necessary),
5. combining two adjacent coupons into one (using composition of mor-
phisms in C), and
6. replacing a diagram containing a coupon labeled by a linear combination
of morphisms with the corresponding linear combination of fields.
• Note that any labeled graph in S 2 is equivalent, via the above local relations,
to some multiple of the empty graph. We call this multiple the “standard
evaluation” of a graph in S 2 .
8.2.2 • (Give refinement of description of A(Y ; c), including case when Y is a disk.
Observe that in this case we have an orthogonal basis.)
Assume now that C is s semisimple spherical category with finitely many irreps.
It follows from (8.2.1) that A(Y ; c) is finite dimensional for all 2-manifolds Y and
c ∈ C(∂Y ). Note that Z(S 2 ) is 1-dimensional, and the standard evaluation of triva-
lent graphs (which evaluates to 1 ∈ C on the empty graph) is a basis. It follows
8.2. SPHERICAL CATEGORIES IN GENERAL 71
from (8.2.2) that any non-zero element z ∈ Z(S 2 ) determines a nondegenerate in-
ner product on A(D 2 ; c) for all c. Assume that these inner products are positive
definite. [comment on this assumption.] We can now apply (6.3.1) to construct
a path integral for 3-manifolds. Following the proof of (6.3.1) we will construct a
state sum description of the path integral in terms of labelings of the cells of a cell
decomposition of a 3-manifold. We will see that this state model turns out to be
the Turaev-Viro state model [TV92, Tur94].
Choose z ∈ Z(S 2 ); z is λ times the standard evaluation of trivalent graphs for
some λ ∈ C. [for positive definiteness, need λ ∈ R+ ] (Recall from (6.3.1) that we
define Z(B 3 ) = z.) Our first task is the compute bases and inner products for
A(Y ; c), where (Y ; c) runs through all attaching targets for handles that we will
need below.
Recall that a field on 1-manifold R consists of a finite number of oriented points
in R each labeled by an irrep. If R is contained in the boundary of a 2-manifold
we will assume that all of the points are oriented inward. Fields will be denoted
as a sequence of irreps, so, for example, A(D 2 ; a, b, c) means a disk with boundary
conditions given by three inward pointing points labeled by a, b and c. [this remark
on notation should go earlier]
Let a and b be irreps (or equivalently minimal idempotents or simple objects) 8.2.3
of C. Then A(D 2 ; ba, b) is 1-dimensional if a and b are equivalent and 0-dimensional
otherwise. A basis for A(D 2 , ba, a) is given by a single oriented arc in D 2 connecting
the boundary points and labeled by a. Call this basis element ebaa . Then
where θabci is the value of the standard evaluation on a theta graph labeled by a, b,
c, eabci and ebabci .
Next we consider S 1 × I. We will see below that we will only need to know inner 8.2.5
products on A(S 1 × I; c) when c is the empty boundary condition ∅. For a an irrep
of C, consider the element ea ∈ A(S 1 ×I; ∅) represented by a loop S 1 ×{pt} ⊂ S 1 ×I
labeled by a. It follows from (4.4.6) that {ea }, where a runs through a complete set
of irreps of C, is a basis of A(S 1 × I; ∅). Recalling the definition of ebaa from above,
72 CHAPTER 8. 2+1-DIMENSIONAL EXAMPLES
(see Figure (8.2.6)). A similar argument shows that hea , eb i = 0 if a and b are not
equivalent. So {ea } is an orthonormal basis of A(S 1 × I; ∅).
8.2.7 Finally we consider S 2 . Let e∅ ∈ A(S 2 ) be represented by the empty trivalent
graph; e∅ spans A(S 2 ). Let L be a complete set of irreps of C. Decomposing
S 2 × I = (D 2 × I) ∪ (D 2 × I) and applying (6.1.9) and (8.2.5) we have
def P 2
where D = a da .
Armed with the above computations, we can now calculate the path integral
Z(M ) of a 3-manifold M in terms of a generic cell decomposition of M . By generic
we mean dual to a triangulation, so that each 1-cell is incident to three 2-cells, and
the 1- and 2-cells incident to a 0-cell form a tetrahedral pattern. By thickening the
cells we get a handle decomposition of M . For simplicity, we will at first assume
the M is closed. Later we will indicate how to extend the results to more general
handles decompositions and to 3-manifolds with boundary.
We will start by analyzing the effect of adding 3-handles and work our way down
to the 0-handles. Let Mi ⊂ M be the union of all handles of index less than or equal
to i. Let ni be the number of i-handles, and let Hi denote the set of i-handles.
Applying (6.1.9) and (8.2.7) to the decomposition M = M2 ∪ {3-handles}, we see
8.2. SPHERICAL CATEGORIES IN GENERAL 73
that
Y
Z(M ) = Z(M2 )(en∅ 3 ) z(e∅ )he∅ , e∅ i−1
H3
Y
= Z(M2 )(en∅ 3 ) λ−1 D −1 .
H3
X Y
Z(M2 )(en∅ 3 ) = Z(M1 )(gα ) λd(α, f ),
α∈L2 f ∈H2
where gα denotes is the graph on ∂M1 consisting of a loop for the attaching target of
each 2-handle labeled according to α, and d(α, f ) = dα(f ) , the standard evaluation
of a loop labeled by α(f ).
Next we consider the decomposition M1 = M0 ∪ {1-handles}. We want to eval-
uate Z(M1 ) on gα . In applying (6.1.9) and (8.2.4) we will place a trivalent vertex
(some eabci ) in each D 2 in ∂(M0 ⊔ {1-handles}) which results from cutting M1 . [need
figure(?)] The resulting field on the boundary of a 1-handle is a theta graph labeled
according to α and the eabci ’s on the two attaching disks. If these two basis elements
are not the same then the theta graph will evaluate to zero, so we may assume that
they are the same. In other words, given α we have fixed arc labels a, b, c for the
theta graph of each 1-handle, and we sum over i’s so that eabci runs through a ba-
sis of A(D 2 , a, b, c). Let L1,α denote the set of all such labelings of the 1-handles
consistent with α. For β ∈ L1,α and e ∈ H1 let Θ(α, β, e) denote the standard eval-
uation of the labeled theta graph on the boundary of e. For each 0-handle v ∈ H0
there is a labeled tetrahedral graph on the boundary of the 0-handle. (Labels of the
edges of the tetrahedron come from α and labels of the vertices of the tetrahedron
from from β.) Let Tet(α, β, v) denote the standard evaluation of this graph. Note
that the contribution of a 1-handle e to the he∗ , e∗ i−1 factor of (6.1.9) is, by (8.2.4),
(λΘ(α, β, e))−2 , since we glue along two pairs of disks for each 1-handle. Putting
this all together, we have
X Y Y
Z(M1 )(gα ) = λΘ(α, β, e)(λΘ(α, β, e))−2 λ Tet(α, β, v)
β∈L1,α e∈H1 v∈H0
X Y Y
= (λΘ(α, β, e))−1 λ Tet(α, β, v)
β∈L1,α e∈H1 v∈H0
74 CHAPTER 8. 2+1-DIMENSIONAL EXAMPLES
Here χ(M ) denotes the Euler characteristic of M , which is always zero for a closed 3-
manifold. We leave it in the formula because we want to indicate how to generalize
to non-closed 3-manifolds, and because we want to emphasize the parallels with
a similar expression for 4-manifolds. [need forward ref] The above expression for
Z(M ) is essentially the Turaev-Viro state sum [TV92, Tur94]. But note that we do
not need to show that it is invariant under Pachner moves on the triangulation —
invariance follows from the easier and more general (6.3.1).
For a general cell decomposition of M (one not necessarily dual to a triangu-
lation), The factors of Θ and above will be replaced with an evaluation of “multi-
barred theta graph” reflecting the number of 2-cells incident to a 1-cell. [need figure]
The factors of Tet will be replaced with an evaluation of the graph which is the link
of the 2-skeleton of the decomposition around the vertex. If M has boundary then
we first fix a graph g on ∂M , which we assume is in a neighborhood of the 1-skeleton
of the cell decomposition restricted to the boundary. Then there is a similar state
sum for Z(M )(g). [need to give more details]
8.3 A2
(follow Kuperberg then extend some; include lots of detail)
8.4 G2
(follow Kuperberg then extend some; include lots of detail)
3+1-dimensional Examples
• Note that this section is very similar to Section 8.2, so the reader might want
to compare the two, or treat Section 8.2 as a warm-up for this section.
• A disklike 3-category with only one object (0-morphism) and only one 1-
morphism is a ribbon category. [need to give details] Thus we expect that
a from a ribbon category we can construct a 3+1-dimensional TQFT. We
will see in the next section that (with some additional assumptions on the
ribbon category) this 3+1-dimensional TQFT can be decategorified to yield
the Witten-Reshtikhin-Turaev type 2+1-dimensional TQFT that can be con-
structed directly (but less cleanly) from the same ribbon category.
1. isotopy,
2. reversing the orientation of a ribbon and changing the label to its dual,
75
76 CHAPTER 9. 3+1-DIMENSIONAL EXAMPLES
3. replacing “identity” coupons with parallel ribbons (note that this includes
cups and caps because of Frobenius reciprocity),
4. erasing ribbons labeled by the trivial object (and adjusting adjacent labels
of coupons as necessary),
5. combining two adjacent coupons into one (using composition of mor-
phisms in C), and
6. replacing a diagram containing a coupon labeled by a linear combination
of morphisms with the corresponding linear combination of fields.
• Note that any labeled ribbon graph in (B 3 , ∅) is equivalent, via the above
local relations, to some multiple of the empty graph. We call this multiple the
“standard evaluation” of a graph in (B 3 , ∅).
9.1.2 • (Give refinement of description of A(M ; c), including case when M is a disk.
Observe that in this case we have an orthogonal basis.)
Assume now that C is s semisimple ribbon category with finitely many irreps.
It follows from (9.1.1) that A(M ; c) is finite dimensional for all 3-manifolds M
and c ∈ C(∂M ). Note that Z(S 3 ) is 1-dimensional, and the standard evaluation
of trivalent ribbon graphs (which evaluates to 1 ∈ C on the empty graph) is a
basis. It follows from (9.1.2) that any non-zero element z ∈ Z(S 3 ) determines a
nondegenerate inner product on A(B 3 ; c) for all c. Assume that these inner products
are positive definite. [comment on this assumption.] We can now apply (6.3.1) to
construct a path integral for 4-manifolds.
9.1. THEORIES FROM RIBBON CATEGORIES 77
Following the proof of (6.3.1) we will construct a state sum description of the
path integral in terms of labelings of the handles of a handle decomposition of a 4-
manifold. [need to be consistent about handle decomposition vs cell decomposition]
We will see below that this state sum specializes to: (a) the well-known Witten-
Reshtikhin-Turaev framed link surgery formula, for handle decompositions consist-
ing of a single 0-handle and several 2-handles; (b) the Crane-Yetter state sum, for
handle decompositions which are dual to triangulations; and (c) the Turaev shadow
state sum, for handle decompositions which have tetrahedral singularities in their
2-skeleton. [need to be more precise here] [need refs for above]
Choose z ∈ Z(S 3 ); z is λ times the standard evaluation of trivalent graphs for
some λ ∈ C. [for positive definiteness, need λ ∈ R+ ] (Recall from (6.3.1) that we
define Z(B 4 ) = z.) Our first task is the compute bases and inner products for
A(M ; c), where (M ; c) runs through all attaching targets for handles that we will
need below.
Recall that a field on 2-manifold Y consists of a finite number of oriented framed
points in Y each labeled by an irrep. If Y is contained in the boundary of a 3-
manifold we will assume that all of the points are oriented inward. Fields will be
denoted as a sequence of irreps, so, for example, A(B 3 ; a, b, c) means a 3-ball with
boundary conditions given by three inward pointing points labeled by a, b and c.
[remark that the ordering of the irreps (points) in a connected component doesn’t
matter] [this remark on notation should go earlier]
Let a and b be irreps (or equivalently minimal idempotents or simple objects) of 9.1.3
C. Then A(B 3 ; ba, b) is 1-dimensional if a and b are equivalent and 0-dimensional oth-
erwise. A basis for A(B 3 , ba, a) is given by a single oriented ribbon in B 3 connecting
the boundary points and labeled by a. Call this basis element ebaa . Then
where θabci is the value of the standard evaluation on a theta graph labeled by a, b,
c, eabci and ebabci .
More generally, if c = (a1 , . . . , an ) (n ≥ 3), we can choose a trivalent ribbon 9.1.5
tree with boundary c, and labelings of the tree give an orthogonal basis of A(B 3 ; c)
(see (9.1.2)). The n − 3 internal edges of the tree are labeled by irreps and the n − 2
78 CHAPTER 9. 3+1-DIMENSIONAL EXAMPLES
trivalent vertices of the tree are labeled by some orthogonal bases of the appropriate
trivalent vertex spaces. Call these (multi) labels b and v respectively. Then
Y Y
hebv , ebv i = λ Θ(b, v, i) d(b, j)−1 ,
i∈V j∈E
where V denotes the set of vertices of the tree and E denotes the set of internal
edges. [need to give more detail and improve notation]
9.1.6 Next we consider S 1 × D 2 . We will see below that we will only need to know
inner products on A(S 1 × D 2 ; c) when c is the empty boundary condition ∅. For
a an irrep of C, consider the element ea ∈ A(S 1 × D 2 ; ∅) represented by a loop
S 1 × {pt} ⊂ S 1 × D 2 labeled by a. It follows from (4.4.6) that {ea }, where a runs
through a complete set of irreps of C, is a basis of A(S 1 × D 2 ; ∅). Recalling the
definition of ebaa from above, and applying (6.1.9), we see that
A similar argument shows that hea , eb i = 0 if a and b are not equivalent. So {ea } is
an orthonormal basis of A(S 1 × D 2 ; ∅).
9.1.7 Next we consider S 2 × I. In applications we will only need to consider empty
boundary conditions. A(S 2 × I; ∅) is 1-dimensional and spanned by e∅ , the empty
field. Let L be a complete set of irreps of C. Decomposing S 2 × I × I = (D 2 × I ×
I) ∪ (D 2 × I × I) and applying (6.1.9) and (9.1.6) we have
This completes the inner product calculations we will need for handle attachments.
Before doing general calculations we look at S 2 × I and S 2 × S 1 , and also show
that frequently we can choose λ so that Z(W 4 ) depends only on the bordism class
of W .
Consider A(S 2 ×I; b a, b), where the two ribbon ends lie in different components of
∂(S × I) and a and b are irreps (simple objects). It is easy to see that A(S 2 × I; b
2 a, b)
is 0 if a ≇ b, and that A(S 2 × I; b a, a) is spanned by xa , the obvious unknotted
ribbon connecting the two ribbon ends. So we must determine whether xa = 0 in
(S 2 × I; b
a, a).
Let Hab denote the standard evaluation of the (framed) Hopf link with labels a 9.1.9
and b. Note that H1a = da . Let xab denote xa plus a small linking circle labeled by
b (see Figure (9.1.10)). Then xab = (Hab /da )xa . But in S 2 × I the linking circle is
isotopic to an unlinked circle, so we also have xab = db xa . So (Hab /da − db )xa = 0
for all b ∈ L, which means that xa = 0 unless Hab = da db for all b ∈ L. By (9.1.1)
this is also a sufficient condition for xa 6= 0. Call such an a ∈ L degenerate. [find a
better name for this?] Thus A(S 2 × I; b a, a) is 1-dimensional when a is degenerate
and 0-dimensional otherwise. Note that 1 ∈ L is always degenerate.
Similar arguments show that a basis of A(S 2 × S 1 ) consists of {fa }, where a is
degenerate and fa ∈ A(S 2 × S 1 ) denotes the ribbon graph pt × S 1 ⊂ S 2 × S 1 labeled
by a ∈ L.
Suppose that dim A(S 2 ×S 1 ) = 1 (or equivalently 1 is the only degenerate irrep).
[in what follows, need to make clearer when this assumption is needed.] We will
show that in this case setting λ2 = 1/D yields a path integral Z(W 4 ) which depends
only on the bordism class of W . 4-dimensional bordism is generated by three types
of surgery (and their inverses):
∅ ↔ S4
B4 × S0 ↔ S3 × B1
B 3 × S 1 ↔ S 2 × B 2.
We have
Z(S 4 ) = Z(B 4 )(e′∅ ) · Z(B 4 )(e′∅ ) · he′∅ , e′∅ i−1 = λ2 D,
so the first type of surgery does not affect Z(W ) if λ2 = 1/D. We have
Z(B 4 × S 0 )(∅) = λ2 ,
80 CHAPTER 9. 3+1-DIMENSIONAL EXAMPLES
while
Z(S 3 × B 1 ) = he′∅ , e′∅ i = D −1 ,
so the second type of surgery does not affect Z(W ) if λ2 = 1/D.
Finally, we consider B 3 × S 1 ↔ S 2 × B 2 , with boundary S 2 × S 1 . With a
eye toward future applications, we will temporarily abandon our assumption that
dim A(S 2 × S 1 ) = 1. Let fa ∈ A(S 2 × S 1 ) be as defined above. Note that f1 is
equivalent to the empty field. We have
[remark that this is the well known Witten-Resheitkin-Turaev surgery formula] Note
that we do not need to show that this expression is invariant under handle slides; this
follows from (6.3.1). More generally, let K be a labeled ribbon graph (e.g. a framed
link) in ∂W . By general position, we may assume that K lies in the complement of
L in S 3 . Let J(L ∪ K, m) denote the standard evaluation of L ∪ K, with L labeled
by m. Then X
Z(W )(K) = (λJ(L ∪ K, m))(λk d(m)).
m∈L(L)
9.1.11 Next consider a 4-manifold W equipped with a generic cell decomposition (i.e. a
cell decomposition dual to a triangulation). This means that a neighborhood of each
9.1. THEORIES FROM RIBBON CATEGORIES 81
0-cell looks like an open cone on the boundary of a 4-simplex, and a neighborhood
of a point in a k-cell looks like a neighborhood of a point of a k-cell in this cone.
(In other words, the open cone on the boundary of a 4-simplex is the local model
for the cell decomposition.) We will use the gluing formula to derive a state sum
description of Z(W ) in terms of labelings of the cells. [refer to 2+1 dim’l case? need
to say that this is very similar to 2+1/TV case, and so we’ll use fewer words]
For simplicity, assume that W is closed.
[say something about the fact that there is a closely related handle decomposition
and we will treat the two as equivalent]
Let ni be the number of i-handles, Hi be the set of i-handles, and let Wi denote
the union of all handles of index less than or equal to i.
Applying the gluing formula (6.1.9) and the inner product calculation (9.1.8) to
the decomposition W = W3 ∪ {4-handles}, we have
Y
Z(W ) = Z(W3 )(∅) z(e′∅ )he′∅ , e′∅ i−1
H4
= Z(W3 )(∅)λn4 D n4
Applying the gluing formula and the inner product calculation (9.1.7) to the decom-
position W3 = W2 ∪ {3-handles}, we have
Y
Z(W3 )(∅) = Z(W2 )(∅) z(∅)he∅ , e∅ i−1
H3
−n3
= Z(W3 )(∅)λ D −n3
by β ∈ L2 to the 1-handle e, and the inner product is computed using the standard
evaluation (λ = 1). For each 1-handle e we have a factor of λE(α, β, e) coming
from the 1-handle and (λE(α, β, e))−2 coming from the gluing correction factor, so
the total contribution of a 1-handle is (λE(α, β, e))−1 . Each 0-handle v contributes
λΦ(α, β, v), where Φ(α, β, v) denotes the standard evaluation of the 1-skeleton of
the boundary of a 4-simplex. The labels of the edges of this graph come from α,
and the labels of the (4-valent) vertices come from β. Combining all of the above
and applying the gluing formula we have
X Y Y
Z(W1 )(Lα ) = λ−1 E(α, β, e)−1 λΦ(α, β, v)
β∈L2 e∈H1 v∈H0
If we choose for each 1-handle e a trivalent ribbon tree with boundary c(α, e) (equiv-
alently, choose a partition of the four labeled points of c(α, e) into two groups of two),
then labelings of the internal edge and two vertices of this tree give an orthogonal
basis of A(B 3 ; c(α, e)) (see (9.1.5)). The above expression becomes
X Y Y
Z(W1 )(Lα ) = λ−1 Θ1 (α, β, e)−1 Θ2 (α, β, e)−1 d(β, e) λΦ(α, β, v).
β∈L2 e∈H1 v∈H0
Here Θ1 and Θ2 are the θ factors for the two trivalent vertices corresponding to e,
with labels coming from α and β. d(β, e) is the loop value for the internal edge of
the tree for e. Φ(α, β, v) can now be interpreted as a “15j” symbol: it is the standard
evaluation of a trivalent ribbon graph with 15 edges and 10 trivalent vertices.
Combining the expressions for all handles we have
X X
Z(W ) =λχ(W ) Dn4 −n3
α∈L2 β∈L2
Y Y Y
d(α, f ) Θ1 (α, β, e)−1 Θ2 (α, β, e)−1 d(β, e) Φ(α, β, v).
f ∈H2 e∈H1 v∈H0
where v runs over 0-cells of G and e runs over 1-cells of G. If the disk bundle over
Y is twisted with respect to the boundary trivialization determined by embedding
of ea , or equivalently the embedding of ea is changed by a twist, this becomes
where ta is the twist factor for a and k is the number of twists (Euler number of the
disk bundle). [need to say this better; also refer back to def of ta ]
Now we proceed as before. Gluing 3- and 4-handles presents nothing new:
where t(α, f ) is the twist factor for the label that α assigns to f , and kf is the euler
number for f . [comment on Turaev “gleams”] Gluing 1-handles to W0 gives
X Y Y
Z(W1 )(Lα ) = λ−1 Θ(α, β, e)−1 λ Tet(α, β, v).
β∈L2 e∈H1 v∈H0
Here L2 is the set of all labelings of the 1-handles by orthogonal basis vectors
associated to the three labels coming from α and the three 2-cells adjacent to the
84 CHAPTER 9. 3+1-DIMENSIONAL EXAMPLES
If we set λ2 = 1/D this is essentially the Turaev shadow state sum for the cell
decomposition of W . [need to check normalization] Note that we do not need to
show that this expression is independent of the choice of cell decomposition; this
follow from (6.3.1).
[need to remark that this can be generalized to manifolds with boundary]
[...]
9.2 Decategorification
Plan:
• choose λ to make Z a bordism invariant
• gluing
• A(S 1 × I)
86 CHAPTER 9. 3+1-DIMENSIONAL EXAMPLES
Chapter 10
4+1-dimensional Examples
• The main fact about Khovanov homology that we will use is that surface
bordisms in S 3 × I give maps on Khovanov homology. (See [MW06] and [need
ref for Jacobsson].)
87
88 CHAPTER 10. 4+1-DIMENSIONAL EXAMPLES
the tangle. Let α be the corresponding arc in W which joins the two
inner boundary components and is disjoint from the surface. Let W ′ be
W with a neighborhood of α removed. W ′ is a bordism in S 3 × I from
AB ∪ BC (a link in S 3 consisting of distant copies of AB and BC) to
AC. Thus W ′ gives a map from Kh(AB) ⊗ Kh(BC) → Kh(AC). This
is the definition of composition of 4-morphisms. It is not hard to show
that this is independent of the choice of p.
– The various conjugations needed to complete the definition of the 4-
category have obvious geometric definitions.
– [Need to also say something about (s)pin structures. Does this entail a
modification of the definition of 4-category?]
• In particular, A(B 4 ; L) ∼
= Kh(L) is this theory.
Chapter 11
n+1-dimensional Examples
89
90 CHAPTER 11. N +1-DIMENSIONAL EXAMPLES
Chapter 12
Reconstruction Results
In this Chapter we show how to construct various parts of a full TQFT from various
sorts of combinatorial data (link invariants with certain properties, Moore-Seiberg
data, modular tensor category, ...)
Maybe the point of this chapter should be to make contact with prior TQFT
literature.
Maybe get rid of this chapter and instead put the various reconstruction results
at the ends of the 1+1, 2+1, 3+1 chapters (?)
Plan:
• moore-seiberg equations
• TV model
91
92 CHAPTER 12. RECONSTRUCTION RESULTS
Chapter 13
[Other Chapters]
• ?? CS-U(1) theories
93
94 CHAPTER 13. [OTHER CHAPTERS]
Appendix A
Categories
• duality/conjugate stuff
95
96 APPENDIX A. CATEGORIES
Semisimple Categories
[need to complete transition from v-space to modules (seach for all ’vector’, ’space’,
’k’)]
[need to switch from left reps to right reps (in order to be consistent with he rest of
the book)]
In this appendix we prove structure theorems (see (B.4.5) and (B.4.6)) for
semisimple k-categories. We then use this structure theorem to prove various things
about these categories. When k is a field, some authors call k-categories “algebroids”
(in analogy to groups and groupoids), and indeed most facts about semisimple alge-
bras (or more generally semisimple rings) generalize easily to categories. The proofs
given here follow Chapter XVII of [Lang xxxx] closely; in many cases they have
been adapted almost verbatim. In what follows, the reader should keep in mind the
special case of algebras, thought of as categories with only one object.
B.1 Definitions
In this appendix category will always mean k-category (see (A.1.1)), where k is a
ring. An important special case is when k is a field. As in the rest of this book,
we will write the composition of f ∈ Cab1 and g ∈ C 1 as f g, not gf (think sticking
bc
arrows together, rather than composing functions).
For the remainder of this section let C be a k-category.
A left ideal L ⊂ C is a subset of C 1 which is closed under left composition by
morphsisms of C, and such that L ∩ Cab 1 is a submodule of C 1 for all a, b ∈ C 0 . In
ab
1
other words, for all f ∈ C and g ∈ L, f g ∈ L whenever the composition is defined.
Right ideals of C are defined similarly. Two-sided ideals are both left and right
ideals.
97
98 APPENDIX B. SEMISIMPLE CATEGORIES
Ca → Cv.
L
(i.e. {Faj } are independent subspaces of Ea for all a ∈ C 0 ). Let E ′ = j∈J F j . I
claim that E ′ = E. It suffices to show that for all i ∈ I, F i ⊂ E ′ . Consider the
subrepresentation (F i ∩ E ′ ) ⊂ F i . F i ∩ E ′ = 0 would contradict the maximality of
J. Therefore, by theL simplicity of F i , F i ∩ E ′ = F i and F i ⊂ E ′ .
2 ⇒ 3. Let E = i∈I F i with each F i simple and let G ⊂ E. Choose a maximal
P P
J ⊂ I such that G + ( j∈J F j ) is direct (i.e. Ga ∩ ( j∈J Faj ) = 0 for all a ∈ C 0 ).
P
Let E ′ = G + ( j∈J F j ). Then for all i ∈ I, F i ∩ E ′ = F i (because F i ∩ E ′ = 0
would contradict the maximality of J), so E ′ = E.
3 ⇒ 1. First we show that any subrepresentation G of E contains a simple sub-
representation. Choose a non-zero v ∈ Ga , a ∈ C 0 . Then Cv is a subrepresentation
of G and ker(Ca → Cv) is a left ideal L 6= C. Consider the partially ordered set of
left ideals strictly smaller than Ca and containing L. By Zorn’s Lemma there is a
maximal ideal M in this set. It follows that M v is a maximal subrepresentation of
Cv. By condition 3, E = M v ⊕ M ′ for some M ′ . Then (Cv ∩ M ′ ) ⊂ G is simple by
the maximality of M v. This is the desired simple subrepresentation of G.
Let E ′ be the sum of all the simple subrepresentations of E. If E ′ 6= E then
E = E ′ ⊕ F for some F 6= 0. But by the argument above F contains a simple
subrepresentation, which is a contradiction.
that αv = f v.
Proof. Let E = Cv ⊕ E ′ and let π : E → Cv be the projection. Then π ∈ R =
EndC (E), so αv = απb v = πa αv. Hence αv ∈ Cv, which implies that there is an
1 such that αv = f v.
f ∈ Cab
Next we use a diagonal trick to enhance the above result to work for a finite
number of vi ∈ Ea .
100 APPENDIX B. SEMISIMPLE CATEGORIES
B.3.2 Theorem (Density Theorem). Let C, E and R = EndC (E) be as above and
assume E is semisimple. Choose a, b ∈ C 0 and an R-map α : Eb → Ea . Choose
1 such that αv = f v for all i.
v1 , . . . , vm ∈ Eb . Then there exists f ∈ Cab i i
Proof. Let E (m) be the direct sum of m copies of E. Then EndC (E (m) ) is isomorphic
B.3.3 to m × m matrices with entries in R. (Proof: Clearly Matm (R) ⊂ End(E (m) ). Let
(m) (m)
h ∈ End(E (m) ). Then ha : Ea → Ea decomposes as a matrix (hij a ), where each
hij
a is a linear map from Ea to itself. It is easy to see that for fixed i and j and
varying a the collection {hij (m) : E (m) → E (m)
a } is in EndC (E) = R.) Note that α b a
(m) (m)
is an EndC (E )-map, and (vi ) can be thought of as an element of Ea . It now
follows from the previous lemma that there exists f ∈ Cab 1 such that αv = f v for
i i
all i.
e= L j j
Proof. Let E i,j E , where i, j runs through the index set for {vi }. (In general
there are multiple copies of each E j in the direct sum.) E e is a C-representation,
e consists of matrices whose (i, j; i′ , j ′ ) entry is a natural transformation
and EndC (E)
′
from E to E j . (This is similar to (B.3.3).) Note that α determines an EndC (E)-
j e
map from E eb to E
ea , and (v ) can be thought of as an element of E
j ea . It now follows
i
1 j j
from (B.3.1) that there exists f ∈ Cab such that αj vi = f vi for all i and j.
Proof. Consider the natural map C → EndR (E∗ ). By assumption this map is in-
jective. By (B.3.4) it is surjective: Fix b ∈ C 0 and choose {vij } of (B.3.4) to be a
collection of generators of the finitely many, finitely generated, non-zero Ebj .
B.4. SEMISIMPLE CATEGORIES 101
X X X
x = 1a x1b = eaj x ebk = eai xebi .
j∈Ia k∈Ib i∈Ia ∩Ib
102 APPENDIX B. SEMISIMPLE CATEGORIES
P
Furthermore, if x = i xi , with xi ∈ Ci , then necessarily xi = eai xebi . It follows
that C is a direct sum M
C= Ci .
i∈I
A category is called simple if it has only one isomorphism class of non-zero simple
left ideal. Clearly Ci above i simple for all i.
We have proved the following structure theorem for semisimple categories.
B.4.3 Theorem. Let C be a semisimple category. Let {Li }i∈I be a complete set of rep-
resentatives for the simple leftLideals of C, and let Ci be the sum of all left ideals
isomorphic to Li . Then C = i∈I Ci and each Ci is a simple category. For each
a, b ∈ C 0 , Ciab
1 is non-zero for only fintely many i.
b. We have proved
B.5. TO DO LIST 103
Theorem. Let C be a semisimple category over an algebraically closed field k. Let B.4.6
{Li }i∈I be a complete set of representatives for the simple left ideals of C. Then C
is naturally isomorphic to the complete category of graded vector spaces with each
object a ∈ C 0 corresponding to the graded vector space {Lia }i∈I .
B.5 To Do List
To do:
• representations
• idempotents
[NOTE] [Note: Bibliography is not complete. At the moment I’m just testing
different styles for entries.]
[TO DO] [To do: need to add Segal refs, Freed, FQ, Quinn, Barrett, Sawin,
Frohman, JKB, Bullock, more Turaev, BHMV, Przytycki, Kirby-
Melvin, Roberts, ...]
[CY93]
[Mac98] S. Mac Lane, Categories for the Working Mathematician, second edi-
tion, Springer, New York, 1998.
[TV92] V. Turaev and O. Viro, State sum invariants of 3-manifolds and quan-
tum 6-j symbols, Topology 31 (1992), 862–902.
105
106 BIBLIOGRAPHY
[Wit89a] E. Witten, Quantum field theory and the Jones polynomial, Comm.
Math. Phys. 121 (1989), 351–399.
[Wit89b] E. Witten, Gauge theories and integrable lattice models, Nucl. Phys. B
322 (1989), 629–697.