Costello
Costello
Costello
Theory
Volume 1 (8 May 2016)
Chapter 1. Introduction 1
1. The motivating example of quantum mechanics 2
2. A preliminary definition of prefactorization algebras 6
3. Prefactorization algebras in quantum field theory 7
4. Comparisons with other formalizations of quantum field theory 9
5. Overview of this volume 13
6. Acknowledgements 14
1. Introduction 289
2. Linear algebra and homological algebra in DVS 290
3. Spectral sequences 292
4. Differentiable pro-cochain complexes 294
5. Homotopy colimits 296
Appendix D. The Atiyah-Bott Lemma 307
Bibliography 309
Index 315
CHAPTER 1
Introduction
This two-volume book will provide the analog, in quantum field theory, of the
deformation quantization approach to quantum mechanics. In this introduction, we
will start by recalling how deformation quantization works in quantum mechanics.
The collection of observables in a quantum mechanical system forms an asso-
ciative algebra. The observables of a classical mechanical system form a Poisson
algebra. In the deformation quantization approach to quantum mechanics, one
starts with a Poisson algebra Acl and attempts to construct an associative algebra
Aq , which is an algebra flat over the ring C[[~]], together with an isomorphism of
associative algebras Aq /~ Acl . In addition, if a, b ∈ Acl , and e
a, e
b are any lifts of
q
a, b to A , then
1 e
lim [e a, b] = {a, b} ∈ Acl .
~→0 ~
1.1. A particle in a box. For the general framework we want to present, the
details of the physical system under study are not so important. However, for con-
creteness, we will focus attention on a very simple system: that of a single particle
confined to some region of space. We confine our particle inside some box and
occasionally take measurements of this system. The set of possible trajectories of
the particle around the box constitute all the imaginable behaviors of this particle;
we might describe this space of behaviors mathematically as Maps(I, Box), where
1. THE MOTIVATING EXAMPLE OF QUANTUM MECHANICS 3
I ⊂ R denotes the time interval over which we conduct the experiment. We say the
set of possible behaviors forms a space of fields on the timeline of the particle.
The behavior of our theory is governed by an action functional, which is a
function on Maps(I, Box). The simplest case typically studied is the massless free
field theory, whose value on a trajectory f : I → Box is
Z
S(f) = ( f (t), f¨(t))dt.
t∈I
Here we use (−, −) to denote the usual inner product on Rn , where we view the box
as a subspace of Rn , and f¨ to denote the second derivative of f in the time variable
t.
The aim of this section is to outline the structure one would expect the observ-
ables — that is, the possible measurements one can make of this system — should
satisfy.
1.2. Classical mechanics. Let us start by considering the simpler case where
our particle is treated as a classical system. In that case, the trajectory of the particle
is constrained to be in a solution to the Euler-Lagrange equations of our theory,
which is a differential equation determined by the action functional. For example,
if the action functional governing our theory is that of the massless free theory,
then a map f : I → Box satisfies the Euler-Lagrange equation if it is a straight line.
(Since we are just trying to provide a conceptual narrative here, we will assume
that Box becomes all of Rn so that we do not need to worry about what happens at
the boundary of the box.)
We are interested in the observables for this classical field theory. Since the
trajectory of our particle is constrained to be a solution to the Euler-Lagrange equa-
tion, the only measurements one can make are functions on the space of solutions
to the Euler-Lagrange equation.
If U ⊂ R is an open subset, we will let Fields(U) denote the space of fields on
U, that is, the space of maps f : U → Box. We will let
EL(U) ⊂ Fields(U)
denote the subspace consisting of those maps f : U → Box that are solutions to
the Euler-Lagrange equation. As U varies, EL(U) forms a sheaf of spaces on R.
We will let Obscl (U) denote the commutative algebra of functions on EL(U)
(the precise class of functions we will consider will be discussed later). We will
think of Obscl (U) as the collection of observables for our classical system that only
depend upon the behavior of the particle during the time period U. As U varies, the
algebras Obscl (U) vary and together constitute a cosheaf of commutative algebras
on R.
we couple the two physical systems so that they interact, then decouple them and
record how the measurement device has modified from its initial condition. (Of
course, there is a symmetry in this situation: both systems are affected by their
interaction, so a measurement inherently disturbs the system under study.)
The observables for a physical system are all the imaginable measurements we
could take of the system. Instead of considering all possible observables, we might
also consider those observables which occur within a specified time period. This
period can be specified by an open interval U ⊂ R.
Thus, we arrive at the following principle.
Principle 1. For every open subset U ⊂ R, we have a set
Obs(U) of observables one can make during U.
Our second principle is a minimal version of the linearity implied by, e.g., the
superposition principle.
Principle 2. The set Obs(U) is a complex vector space.
We think of Obs(U) as being the collection of ways of coupling a measure-
ment device to our system during the time period U. Thus, there is a natural map
Obs(U) → Obs(V) if U ⊂ V is a shorter time interval. This means that the space
Obs(U) forms a precosheaf.
1.4. Combining observables. Measurements (and so observables) differ qual-
itatively in the classical and quantum settings. If we study a classical particle, the
system is not noticeably disturbed by measurements, and so we can do multiple
measurements at the same time. (To be a little less sloppy, we suppose that by
refining our measuring devices, we can make the impact on the particle as small as
we would like.) Hence, on each interval J we have a commutative multiplication
map Obs(J) ⊗ Obs(J) → Obs(J). We also have maps Obs(I) ⊗ Obs(J) → Obs(K)
for every pair of disjoint intervals I, J contained in an interval K, as well as the
maps that let us combine observables on disjoint intervals.
For a quantum particle, however, a measurement typically disturbs the system
significantly. Taking two measurements simultaneously is incoherent, as the mea-
surement devices are coupled to each other and thus also affect each other, so that
we are no longer measuring just the particle. Quantum observables thus do not
form a cosheaf of commutative algebras on the interval. However, there are no
such problems with combining measurements occurring at different times. Thus,
we find the following.
Principle 3. If U, U 0 are disjoint open subsets of R, and
U, U 0 ⊂ V where V is also open, then there is a map
? : Obs(U) ⊗ Obs(U 0 ) → Obs(V).
If O ∈ Obs(U) and O0 ∈ Obs(U 0 ), then O ? O0 is defined
by coupling our system to measuring device O during the
period U and to device O0 during the period U 0 .
Further, there are maps for an finite collection of dis-
joint time intervals contained in a long time interval, and
these maps are compatible under composition of such maps.
1. THE MOTIVATING EXAMPLE OF QUANTUM MECHANICS 5
Weyl algebra, which one expects to find as the algebra of observables for quantum
mechanics of a particle moving in Rn .
This kind of geometric interpretation of algebra should be familiar to topolo-
gists: associative algebras are algebras over the operad of little intervals in R, and
this is precisely what we have described. As we explain in Section 4, this relation-
ship continues and so our quantization theorem produces many new examples of
algebras over the operad En of little n-discs.
An important point to take away from this discussion is that associative al-
gebras appear in quantum mechanics because associative algebras are connected
with the geometry of R. There is no fundamental connection between associative
algebras and any concept of “quantization”: associative algebras only appear when
one considers one-dimensional quantum field theories. As we will see later, when
one considers topological quantum field theories on n-dimensional space times,
one finds a structure reminiscent of an En -algebra instead of an E1 -algebra.
Remark: As a caveat to the strong assertion above (and jumping ahead of our story),
note that for a manifold of the form X → R, one can push forward a factorization
algebra Obs on X × R to a factorization algebra π∗ Obs on R along the projection
map π : X × R → R. In this case, π∗ Obs((a, b)) = Obs(X × (a, b)). Hence, a quan-
tization of a higher dimensional theory will produce, via such pushforwards to R,
deformations of associative algebras, but knowing only the pushforward is typi-
cally insufficient to reconstruct the factorization algebra on the higher dimensional
manifold. ^
factorization algebra. We do not know how to turn this idea into a precise theorem
in general. We do, however, have a precise theorem in one special case.
Beilinson-Drinfeld show that a factorization algebra in their sense on the affine
line A1 , which is also translation and rotation equivariant, is the same as a vertex
algebra. We have a similar theorem. We show in Chapter 5 that a factorization
algebra on C that is translation and rotation invariant, and also has a certain “holo-
morphic” property, gives rise to a vertex algebra. Therefore, in this special case,
we can show how a factorization algebra in our sense gives rise to one in the sense
used by Beilinson and Drinfeld.
4.2. Segal’s axioms for quantum field theory. Segal has developed and stud-
ied some very natural axioms for quantum field theory Segal (2010). These axioms
were first studied in the world of topological field theory by Atiyah, Segal, and
Witten, and in conformal field theory by Kontsevich and Segal (2004).
According to Segal’s philosophy, a d-dimensional quantum field theory (in Eu-
clidean signature) is a symmetric functor from the category CobRiem
d of d-dimensional
Riem
Riemannian cobordisms. An object of the category Cobd is a compact d − 1-
manifold together with a germ of a d-dimensional Riemannian structure. A mor-
phism is a d-dimensional Riemannian cobordism. The symmetric monoidal struc-
ture arises from disjoint union. As defined, this category does not have identity
morphisms, but they can be added in formally.
4.2.1 Definition. A Segal field theory is a symmetric monoidal functor from CobRiem
d
to the category of (topological) vector spaces.
We won’t get into details about what kind of topological vector spaces one
should consider, because our aim is just to sketch a heuristic relationship between
Segal’s picture and our picture.
In our approach to studying quantum field theory, the fundamental objects are
not the Hilbert spaces associated to codimension 1 manifolds, but rather the spaces
of observables. Any reasonable axiom system for quantum field theory should
be able to capture the notion of observable. In particular, we should be able to
understand observables in terms of Segal’s axioms.
Segal (in lectures and conversations) has explained how to do this. Suppose
we have a Riemannian manifold M and a point x ∈ M. Consider a ball B(x, r)
of radius r around x, whose boundary is a sphere S (x, r). Segal explains that the
Hilbert space Z(S (x, r)) should be thought of as the space of operators on the ball
B(x, r).
If r < r0 , there is a cobordism S (x, r) → S (x, r0 ) given by the complement of
B(x, r) in the closed ball B(x, r0 ). This gives rise to maps Z(S (x, r)) → Z(S (x, r0 )).
Segal defines the space of local operators at x to be the limit
lim Z(S (x, r))
r→0
of this inverse system.
One can understand from this idea of Segal’s how one should construct some-
thing like a prefactorization algebra on any Riemannian manifold M of dimension
4. COMPARISONS WITH OTHER FORMALIZATIONS OF QUANTUM FIELD THEORY 11
example, in dimension 2, the pair of pants with k legs provides the k-ary operations
for the E2 algebra structure on Z(S 1 ).
This story fits nicely with our approach. If we have a locally constant factoriza-
tion algebra F on Rn , then F (Dn ) is an En algebra. Further, we interpret F (Dn ) as
being the space of observables supported on an n-disc. Since F is locally constant,
this may as well be the observables supported on a point, because it is independent
of the radius of the n-disc.
The final chapters in this book develop the concept of factorization algebra,
by adding a certain local-to-global axiom to the definition of prefactorization alge-
bra. In Chapters 6, we provide the definition, discuss the relation between locally-
constant factorization algebras and En algebras, and explain how to construct sev-
eral large classes of examples. In 7, we develop some formal properties of the
theory of factorization algebras. Finally, in 8, we move beyond the formal and
analyze some interesting explicit examples. For instance, we compute the fac-
torization homology of the Kac-Moody enveloping factorization algebras, and we
explain how Abelian Chern-Simons theory produces a quantum group.
5.2. A comment on functional analysis and algebra. This book uses an un-
usual array of mathematical techniques, including both homological algebra and
functional analysis. The homological algebra appears because our factorization al-
gebras live in the world of cochain complexes (ultimately, because they come from
the BV formalism for field theory). The functional analysis appears because our
factorization algebras are built from vector spaces of an analytic nature, such as the
space of distributions on a manifold. We have included an expository introduction
to the techniques we use from homological algebra, operads, and sheaf theory in
Appendix A.
It is well-known that it is hard to make homological algebra and functional
analysis work well together. One reason is that, traditionally, the vector spaces
that arise in analysis are viewed as topological vector spaces, and the category of
topological vector spaces is not an Abelian category. In Appendix B, we introduce
the concept of differentiable vector spaces. Differentiable vector spaces are more
flexible than topological vector spaces, yet retain enough analytic structure for our
purposes. We show that the category of differentiable vector spaces is an Abelian
category, and indeed satisfies the strongest version of the axioms of an Abelian
category: it is a locally-presentable AB5 category. This means that homological al-
gebra in the category of differentiable vector spaces works very nicely. We develop
this in Appendix C.
A gentle introduction to differentiable vector spaces, containing more than
enough to follow everything in both volumes, is contained in Chapter 3, section 5.
6. Acknowledgements
The project of writing this book has stretched over many more years than we
ever anticipated. In that course of time, we have benefited from conversations with
many people, the chance to present this work at several workshops and confer-
ences, and the feedback of various readers. The book and our understanding of
this material is much better due to the interest and engagement of so many others.
Thank you.
We would like to thank directly the following people, although this list is un-
doubtedly incomplete: David Ayala, David Ben-Zvi, Dan Berwick-Evans, Damien
Calaque, Alberto Cattaneo, Ivan Contreras, Vivek Dhand, Chris Douglas, Chris
Elliott, John Francis, Dennis Gaitsgory, Sachin Gautam, Ezra Getzler, Greg Ginot,
6. ACKNOWLEDGEMENTS 15
Ryan Grady, Rune Haugseng, Theo Johnson-Freyd, David Kazhdan, Si Li, Ja-
cob Lurie, Takuo Matsuoka, Pavel Mnev, David Nadler, Thomas Nikolaus, Fred
Paugam, Dmitri Pavlov, Toly Preygel, Kasia Rejzner, Nick Rozenblyum, Clau-
dia Scheimbauer, Graeme Segal, Thel Seraphim, Yuan Shen, Jim Stasheff, Stephan
Stolz, Dennis Sullivan, Matt Szczesny, Hiro Tanaka, Peter Teichner, David Treumann,
Philsang Yoo, Brian Williams, and Eric Zaslow for helpful conversations. K.C. is
particularly grateful to Graeme Segal for many illuminating conversations about
quantum field theory over the years. O.G. would like to thank the participants in
the Spring 2014 Berkeley seminar and Fall 2014 MPIM seminar for their interest
and extensive feedback; he is also grateful to Rune Haugseng and Dmitri Pavlov
for help at the intersection of functional analysis and category theory. Peter Te-
ichner’s feedback after teaching a Berkeley course in Spring 2016 on this material
improved the book considerably. Finally, we are both grateful to John Francis and
Jacob Lurie for introducing us to factorization algebras, in their topological incar-
nation, in 2008.
Our work has taken place at many institutions — Northwestern University,
the University of California, Berkeley, the Max Planck Institute for Mathematics,
and the Perimeter Institute for Theoretical Physics — which provided a supportive
environment for writing and research. Research at the Perimeter Institute is sup-
ported by the Government of Canada through Industry Canada and by the Province
of Ontario through the Ministry of Economic Development and Innovation, and
K.C.’s research at the Perimeter Institute is supported by the Krembil Foundation.
We have also benefitted from the support of the NSF: K.C. and O.G. were par-
tially supported by NSF grants DMS 0706945 and DMS 1007168, and O.G. was
supported by NSF postdoctoral fellowship DMS-1204826. K.C. was also partially
supported by a Sloan fellowship.
Finally, during the period of this project, we have depended upon the unrelent-
ing support and love of our families. In fact, we have both gotten married and had
children over these years! Thank you Josie, Dara, and Laszlo for being models of
insatiable curiosity, ready sources of exuberant distraction, and endless fonts of joy.
Thank you Lauren and Sophie for joining and building our lives together. Amidst
all the tumult and the moves, you always encourage us, you always share your wit
and wisdom and warmth. And it makes all the difference.
Part 1
Prefactorization algebras
CHAPTER 2
19
20 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
scalar field theory on a Riemannian manifold (M, g), where the space of fields is
the space C ∞ (M) of smooth functions on M, and the action funtional is
Z
S (φ) = 12 φ(4g + m2 )φ dvolg .
M
Here 4g refers to the Laplacian with the convention that its eigenvalues are non-
negative, and dvolg denotes the Riemannian volume form associated to the metric.
The positive real number m is the mass of the theory. The main quantities of
interest in the free field theory are the correlation functions, defined by the heuristic
expression
Z
hφ(x1 ) · · · φ(xn )i = φ(x1 ) . . . φ(xn ) e−S (φ) dφ,
φ∈C ∞ (M)
where LV refers to the Lie derivative. Thus, the divergence of V measures the
infinitesimal change in volume that arises when one applies the infinitesimal dif-
feomorphism V.
In coordinates, the divergence is given by the formula
X ∂ ! X X ∂ fi
(†) DivωA fi =− fi x j Ai j + .
∂xi i, j i
∂x i
Cc∞ (U), where every element f ∈ Cc∞ (U) defines a linear functional on C ∞ (U) by
the formula Z
φ 7→ f φ dvolg .
U
ym (People sometimes call them “smeared” because these do not include beloved
functionals like delta functions, only smoothed-out approximations to them.)
As a first approximation to the algebra we wish to use, we define the space of
polynomial functions on Cc∞ (U) to be
e ∞ (U)) = Sym Cc∞ (U),
P(C
i.e., the symmetric algebra on Cc∞ (U). An element of P(C e c∞ (U)) that is homoge-
neous of degree n can be written as a finite sum of monomials f1 · · · fn where the
fi ∈ Cc∞ (U). Such a monomial defines a function on the space C ∞ (U) of fields by
the formula
Z
φ 7→ f1 (x1 )φ(x1 ) . . . fn (xn )φ(xn ) dvolg (x1 ) ∧ · · · ∧ dvolg (xn ).
(x1 ,...,xn )∈U n
Note that because Cc∞ (U) is a topological vector space, it is more natural to use
an appropriate completion of this purely algebraic symmetric power Symn Cc∞ (U).
Because this version of the algebra of polynomial functions is a little less natural
than the completed version, which we will introduce shortly, we use the notation
e The completed version is denoted P.
P.
We define the space of polynomial vector fields in a similar way. Recall that if
V is a finite-dimensional vector space, then the space of polynomial vector fields
on V is isomorphic toP(V) ⊗ V, where P(V) is the space of polynomial functions
on V. An element X = f ⊗ v, with f a polynomial, acts on a polynomial g by the
formula
∂g
X(g) = f .
∂v
In particular, if g is homogeneous of degree n and we pick a representative e g ∈
(V ) , then ∂g/∂v denotes the degree n − 1 polynomial
∗ ⊗n
w 7→ e
g(v ⊗ w ⊗ · · · ⊗ w).
In other words, for polynomials, differentiation is a version of contraction.
In the same way, we would expect to work with
g ∞ (U)) = P(C
Vect(C e ∞ (U)) ⊗ C ∞ (U).
We are interested, in fact, in a different class of vector fields. The space C ∞ (U)
has a foliation, coming from the linear subspace Cc∞ (U) ⊂ C ∞ (U). We are actually
interested in vector fields along this foliation, due to the role of variational calcu-
lus in field theory. This restriction along the foliation is clearest in terms of the
divergence operator we describe below, so we explain it after Definition 2.0.1.
Thus, let
g c (C ∞ (U)) = P(C
Vect e ∞ (U)) ⊗ Cc∞ (U).
Again, it is more natural to use a completion of this space that takes account of the
topology on Cc∞ (U). We will discuss such completions shortly.
24 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
Note that this formula is entirely parallel to the formula for divergence of a
Gaussian measure in finite dimensions, given in formula (†). Indeed, the formula
makes sense even when φ is not compactly supported; however, the term
f1 · · · fn (4 + m2 )φ
need not be compactly supported if φ is not compactly supported. To ensure that
the image of the divergence operator is in P(C e c∞ (U)), we only work with vector
fields with compact support, namely Vect g c (C ∞ (U)).
As we mentioned above, it is more natural to use a completion of the spaces
P(C ∞ (U)) and Vectc (C ∞ (U)) of polynomial functions and polynomial vector fields.
e
We now explain a geometric approach to such a completion.
Let dvolgn denote the Riemannian volume form on the product space U n arising
from the natural n-fold product metric induced by the metric g on U. Any element
F ∈ Cc∞ (U n ) then defines a polynomial function on C ∞ (U) by
Z
φ 7→ F(x1 , . . . , xn )φ(x1 ) · · · φ(xn ) dvolgn .
Un
This functional does not change if we permute the arguments of F by an element
of the symmetric group S n , so that this function only depends on the image of F
in the coinvariants of Cc∞ (U n ) by the symmetric group action. This quotient, of
course, is isomorphic to invariants for the symmetric group action.
2. DIVERGENCE IN INFINITE DIMENSIONS 25
Therefore we define
M
P(C ∞ (U)) = Cc∞ (U n )S n ,
n≥0
where the subscript indicates coinvariants. The (purely algebraic) symmetric power
Symn Cc∞ (U) provides a dense subspace of Cc∞ (U n )S n Cc∞ (U n )S n . Thus, P(C
e ∞ (U))
∞
is a dense subspace of P(C (U)).
In a similar way, we define Vectc (C ∞ (U)) by
M
Vectc (C ∞ (U)) = Cc∞ (U n+1 )S n ,
n≥0
where the symmetric group S n acts only on the first n factors. A dense subspace
of Cc∞ (U n+1 )S n is given by Symn Cc∞ (U) ⊗ Cc∞ (U) so that Vect
g c (C ∞ (U)) is a dense
∞
subspace of Vectc (C (U)).
2.0.2 Lemma. The divergence map
Div g c (C ∞ (U)) → P(C
g : Vect e ∞ (U))
extends continuously to a map
Div : Vectc (C ∞ (U)) → P(C ∞ (U)).
Proof. Suppose that
F(x1 , . . . , xn+1 ) ∈ Cc∞ (U n+1 )S n ⊂ Vectc (C ∞ (U)).
The divergence map in equation (‡) extends to a map that sends F to
n Z
X
−4 xn+1 F(x1 , . . . , xn+1 ) + F(x1 , . . . , xi , . . . , xn , xi ) dvolg .
i=1 xi ∈U
Here, 4 xn+1 denotes the Laplacian acting only on the n + 1st copy of U. Note that
the integral produces a function on U n−1 .
With these objects in hand, we are able to define the quantum observables of a
free field theory.
2.0.3 Definition. For an open subset U ⊂ M, let
H 0 (Obsq (U)) = P(C ∞ (U))/ Im Div .
In other words, H 0 (Obsq (U)) be the cokernel of the operator Div. Later we
will see that this linear map Div naturally extends to a cochain complex of quantum
observables, which we will denote Obsq , whose zeroth cohomology is what we just
defined. This extension is why we write H 0 .
Let us explain why we should interpret this space as the quantum observables.
We expect that an observable in a field theory is a function on the space of fields.
An observable on a field theory on an open subset U ⊂ M is a function on the fields
that only depends on the behaviour of the fields inside U. Speaking conceptually,
the expectation value of the observable is the integral of this function against the
“functional measure” on the space of fields.
26 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
Our approach is that we will not try to define the functional measure, but in-
stead we define the divergence operator. If we have some functional on C ∞ (U) that
is the divergence of a vector field, then the expectation value of the corresponding
observable is zero. Thus, we would expect that the observable given by a diver-
gence is not a physically interesting quantity, since its value is zero. Thus, we
might as well identify it with zero.
The appropriate vector fields on C ∞ (U) — the ones for which our divergence
operator makes sense — are vector fields along the foliation of C ∞ (U) by com-
pactly supported fields. Thus, the quotient of functions on C ∞ (U) by the subspace
of divergences of such vector fields gives a definition of observables.
where EωA denotes the expectation value map for this measure. But the image of
the divergence operator is not an ideal. (Indeed, usually an expectation value map
is not an algebra map!) This fact suggests that, in the BV formalism, the quantum
observables should not form a commutative algebra. One can check quickly that
for our definition above, H 0 (Obsq (U)) is not an algebra.
However, we will find that some shadow of this commutative algebra structure
exists, which allows us to combine observables on disjoint subsets. This residual
structure will give the spaces H 0 (Obsq (U)) of observables, viewed as a functor on
the category of open subsets U ⊂ M, the structure of a prefactorization algebra.
Let us make these statements precise. Note that P(C ∞ (U)) is a commutative
algebra, as it is a space of polynomial functions on C ∞ (U). Further, if U ⊂ V
there is a map of commutative algebras ext : P(C ∞ (U)) → P(C ∞ (V)), extending
a polynomial map F : C ∞ (U) → R to the polynomial map F ◦ res : C ∞ (V) → R
by precomposing with the restriction map res : C ∞ (V) → C ∞ (U). This map ext
is injective. We will sometimes refer to an element of the subspace P(C ∞ (U)) ⊂
P(C ∞ (V)) as an element of P(C ∞ (V)) with support in U.
3.0.1 Lemma. The product map
obtained by combining the inclusion maps P(Ui ) ,→ P(V) with the product map on
P(V). This map does descend to a map
H 0 (Obs(U1 )) ⊗ H 0 (Obs(U2 )) → H 0 (Obs(V)).
In other words, although the product of general observables does not make
sense, the product of observables with disjoint support does.
The same equation holds for the divergence operator we have defined in infinite
dimensions. If X ∈ Vectc (C ∞ (V)) and F ∈ P(Cc∞ (V)), then
Div(FX) − F DivX = X(F).
This computation tells us that the image of Div is not an ideal, as there exist X(F)
not in the image of Div.
When F is in P(C ∞ (U1 )) and X is in Vectc (C ∞ (U2 )), however, X(F) = 0, as
their supports are disjoint. Thus, we have precisely the desired relation F Div(X) =
Div(FX).
We now prove that X(F) = 0 when X and F have disjoint support.
We are working with polynomial functions and vector fields, so that all compu-
tations can be done in a purely algebraic fashion; in other words, we will work with
derivations as in algebraic geometry. Let ε satisfy ε2 = 0. For a polynomial vector
field X and a polynomial function F on a vector space V, we define the function
X(F) to assign to the vector v ∈ V, the ε component of F(v + εXv ) − F(v). Here Xv
denotes the tangent vector at v that X produces.
In our situation, we know that for any φ ∈ C ∞ (V), we have the following
properties:
• F(φ) only depends on the restriction of φ to U1 , and
• Xφ is a function with support in U2 and hence vanishes away from U2 .
Thus, F(φ + εX(φ)) = F(φ) as the restriction of φ + εX(φ) to U1 agrees with the
restriction of φ. We see then that
d
X(F)(φ) = F(φ + εXφ ) − F(φ) = 0,
dε
as asserted.
In a similar way, if U1 , . . . , Un are disjoint opens all contained in V, then there
is a map
H 0 (Obsq (U1 )) ⊗ · · · ⊗ H 0 (Obsq (Un )) → H 0 (Obsq (V))
descending from the map
P(U1 ) ⊗ · · · ⊗ P(Un ) → P(V)
given by inclusion followed by multiplication.
Thus, we see that the spaces H 0 (Obsq (U)) for open sets U ⊂ M are naturally
equipped with the structure maps necessary to define a prefactorization algebra.
(See Section 2 for a sketch of the definition of a factorization algebra, and Sec-
tion 1 for more details on the definition). It is straightforward to check, using the
arguments from the proof above, that these structure maps satisfy the necessary
compatibility conditions to define a prefactorization algebra.
of motion. We will now examine how our construction can be seen as just such a
deformation.
Let us see first why this holds for a class of measures on finite dimensional
vector spaces. Let S be a polynomial function on Rn . Let dn x denote the Lebesgue
measure on Rn , and consider the measure
ω = e−S /~ dn x,
where ~ is a small parameter. The divergence with respect to ω is given by the
formula
X ∂ ! 1 X ∂S X ∂ fi
Divω fi =− fi + .
∂xi ~ ∂xi ∂xi
As before, let P(Rn ) denote the space of polynomial functions on Rn and let Vect(Rn )
denote the space of polynomial vector fields. The divergence operator Divω is a lin-
ear map map Vect(Rn ) → P(Rn ). Note that the operators Divω and ~ Divω have the
same image so long as ~ , 0. When ~ = 0, the operator ~ Divω becomes the
operator
X ∂ X ∂S
fi 7→ − fi .
∂xi ∂xi
Therefore, the ~ → 0 limit of the image of Divω is the Jacobian ideal
∂S
!
Jac(S ) = ⊂ P(Rn ),
∂xi
which corresponds to the critical locus of S in Rn . Hence, the ~ → 0 limit of the ob-
servables P(Rn )/ Im Divω is the commutative algebra P(Rn )/ Jac(S ) that describes
functions on the critical locus of S .
Let us now check the analogous property for the observables of a free scalar
field theory on a manifold M. We will consider the divergence for the putative
Gaussian measure Z !
exp − 1~ φ(4 + m2 )φ dφ
M
on C ∞ (M). For any open subset U ⊂ M, this divergence operator gives us a map
Div~ : Vectc (C ∞ (U)) → P(C ∞ (U))
with
∂ X Z
f1 · · · fn 7 − ~1 f1 · · · fn (4 + m2 )φ +
→ f1 · · · b
fi f φ.
∂φ i M
Note how ~ appears in this formula; it is just the same modification of the operator
(‡) as Divω is of the divergence operator for the measure e−S dn x.
As in the finite dimensional case, the first term dominates in the ~ → 0 limit.
The ~ → 0 limit of the image of Div~ is the closed subspace of P(C ∞ (U)) spanned
by functionals of the form f1 · · · fn (4 + m2 )φ, where fi and φ are in Cc∞ (U)). This
subspace is the topological ideal in P(C ∞ (U)) generated by linear functionals of the
form (4 + m2 ) f , where f ∈ Cc∞ (U). If S (φ) = φ(4 + m2 )φ is the action functional
R
30 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
of our theory, then this subspace is precisely the topological ideal generated by all
the functional derivatives
∂S
( )
: φ ∈ Cc (U) .
∞
∂φ
In other words, it is the Jacobian ideal for S and hence is cut out by the Euler-
Lagrange equations. Let us call this ideal IEL (U).
A more precise statement of what we have just sketched is the following. De-
fine a prefactorization algebra H 0 (Obscl (U)) (the superscript cl stands for classi-
cal) that assigns to U the quotient algebra P(C ∞ (U))/IEL (U). Thus, H 0 (Obscl (U))
should be thought of as the polynomial functions on the space of solutions to the
Euler-Lagrange equations. Note that each constituent space H 0 (Obscl (U)) in this
prefactorization algebra has the structure of a commutative algebra, and the struc-
ture maps are all maps of commutative algebras. In short, H 0 (Obscl ) forms a com-
mutative prefactorization algebra. Heuristically, this terminology means that the
product map defining the factorization structure is defined for all pairs of opens
U1 , U2 ⊂ V, and not just disjoint pairs.
Our work in the section is summarized as follows.
q
4.0.1 Lemma. There is a prefactorization algebra H 0 (Obs~ ) over C[~] such that
when specialized to ~ = 1 is H 0 (Obsq ) and to ~ = 0 is H 0 (Obscl ).
This prefactorization algebra assigns to an open set the cokernel of the map
~ Div~ : Vectc (C ∞ (U))[~] → P(C ∞ (U))[~],
where Div~ is the map defined above.
q
We will see later that H 0 (Obs~ (U)) is free as an R[~]-module, although this is
a special property of free theories and is not always true for an interacting theory.
5. Correlation functions
We have seen that the observables of a free scalar field theory on a manifold M
give rise to a factorization algebra. In this section, we will explain how the structure
of a factorization algebra is enough to define correlation functions of observables.
We will calculate certain correlation functions explicitly and recover the standard
answers.
Suppose now that M is a compact Riemannian manifold, and, as before, let
us consider the observables of the free scalar field theory on M with mass m > 0.
Then we have the following result.
5.0.1 Lemma. If the mass m is positive, then H 0 (Obsq (M)) R.
Compare this result with the statement that for a Gaussian measure on Rn ,
the image of the divergence map is of codimension 1 in the space of polynomial
functions on Rn . The assumption here that the mass is positive is necessary to
ensure that the quadratic form M φ(4 + m2 )φ is non-degenerate.
R
This lemma will follow from our more detailed analysis of free theories in
Chapter 4, but we can sketch the idea here. The main point is that there is a family
5. CORRELATION FUNCTIONS 31
P(Rn ) → R
from polynomial functions to R. This map is the integral against the Gaussian
measure, normalized so that the integral of the function 1 is 1.
In our infinite dimensional situation, we are doing something very similar. Any
reasonable definition of the correlation function of the observables O1 , . . . , On ,
with Oi in P(C ∞ (Ui )), should only depend on the product function O1 · · · On ∈
P(C ∞ (M)). Thus, the correlation function map should be a linear map P(C ∞ (M)) →
R. Further, it should send divergences to zero. We have seen that there is only one
such map, up to an overall scale.
5.1. Comparison to physics. Next we will check explicitly that this correla-
tion function map really matches up with what physicists expect. Let fi ∈ Cc∞ (Ui )
be compactly-supported smooth functions on the open sets U1 , U2 ⊂ M. Let us
view each fi as a linear function on C ∞ (Ui ), and so as an element of the polyno-
mial functions P(C ∞ (Ui )).
Let G ∈ D(M × M) be the unique distribution on M × M with the property that
(4 x + m2 )G(x, y) = δDiag .
Here δDiag denotes the delta function supported on the diagonal copy of M inside
M × M. In other words, if we apply the operator 4 + m2 to the first factor of G,
we find the delta function on the diagonal. Thus, G is the kernel for the operator
(4 + m2 )−1 . In the physics literature, G is called the propagator; in the mathematics
literature, it is called the Green’s function for the operator 4 + m2 . Note that G is
smooth away from the diagonal.
32 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
5.1.1 Lemma. Given fi ∈ Cc∞ (Ui ), there are classes [ fi ] ∈ H 0 (Obsq (Ui )). Then
Z
h[ f1 ] [ f2 ]i = f1 (x)G(x, y) f2 (y).
x,y∈M
Since the observable 1 is chosen to be the basis element identifying H 0 (Obsq (M))
with R, the result follows.
With a little more work, the same arguments recover the usual Wick’s formula
for correlators of the form h f1 · · · fn i.
Remark: These kinds of formulas are standard knowledge in physics, but not in
mathematics. For a more extensive discussion in a mathematical style, see Chapter
6 of Glimm and Jaffe (1987) or Lecture 3 by Kazhdan in Deligne et al. (1999). ^
of observables, and that for linear observables we find the same formula that physi-
cists would write.
In Chapter 3 we will show that a certain class of factorization algebras on
the real line are equivalent to associative algebras, together with a derivation. As
Noether’s theorem suggest, the derivation arises from infinitesimal translation on
the real line, so that it encodes the Hamiltonian of the system.
In Chapter 4, we will analyze the factorization algebra of free field theories
in more detail. We will show that if we consider the free field theory on R, the
factorization algebra H 0 (Obsq ) corresponds (under the relationship between fac-
torization algebras on R and associative algebras) to the Weyl algebra. The Weyl
algebra is generated by observables p, q corresponding to position and momentum
satisfying [p, q] = 1. If we consider instead the family over R[~] of factoriza-
q
tion algebras H 0 (Obs~ ) discussed above, then we find the commutation relation
[p, q] = ~ of Heisenberg. This algebra, of course, is what is traditionally called the
algebra of observables of quantum mechanics. In this case, we will further see that
the derivation of this algebra (corresponding to infinitesimal time translation) is an
inner derivation, given by bracketing with the Hamiltonian
H = p2 − m2 q2 ,
which is the standard Hamiltonian for the quantum mechanics of the harmonic
oscillator.
More generally, we recover canonical quantization of the free scalar field the-
ory on higher dimensional manifolds as follows. Consider a free scalar theory
on the product Riemannian manifold N × R, where N is a compact Riemannian
manifold. This example gives rise to a factorization algebra on R that assigns to
an open subset U ⊂ R, the space H 0 (Obsq (N × U)). We will see that this fac-
torization algebra on R has a dense sub-factorization algebra corresponding to an
associative algebra. This associative algebra is the tensor product of Weyl alge-
bras, where each each eigenspace of the operator 4 + m2 on C ∞ (N) produces a
Weyl algebra. In other words, we find quantum mechanics on R with values in the
infinite-dimensional vector space C ∞ (N). Since that space has a natural spectral
decomposition for the operator 4 + m2 , it is natural to interpret as the algebra of
observables, the associative algebra given by tensoring together the Weyl algebra
for each eigenspace. This result is entirely consistent with standard arguments in
physics, being the factorization algebra analog of canonical quantization.
7. Interacting theories
In any approach to quantum field theory, free field theories are easy to con-
struct. The challenge is always to construct interacting theories. The core results
of this two-volume work show how to construct the factorization algebra corre-
sponding to interacting field theories, deforming the factorization algebra for free
field theories discussed above.
Let us explain a little bit about the challenges we need to overcome in order to
deal with interacting theories, and how we overcome these challenges.
34 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
always assume that the quadratic term in S is similar in form − 12 M φ(4 + m2 )φ,
i.e., we require an ellipticity condition. (Of course, the examples amenable to our
techniques apply to a very general class of interacting theories, including many
gauge theories.)
If U ⊂ M is an open subset, we can consider, as before, the spaces Vectc (C ∞ (M))
and P(C ∞ (M)) of polynomial functions and vector fields on M. By analogy with
the finite-dimensional situation, one can try to define the divergence for the putative
measure exp S (φ)/~dµ (where dµ refers to the “Lebesgue measure” on C ∞ (M))) by
the formula
∂ ∂S X
! Z
Div~ f1 . . . fn = ~ f1 . . . fn
1
+ f1 . . . fi . . . fn
b fi φ.
∂φ ∂φ M
This formula agrees with the formula we used when S was purely quadratic.
But now a problem arises. We defined P(C ∞ (U)) as the space of polynomial
functions whose Taylor terms are given by integration against a smooth function
on U n . That is, M
P(C ∞ (U)) = Cc∞ (U n )S n .
n
∂S
Unfortunately, if φ ∈ Cc (U)), then ∂φ is not necessarily in this space of
∞ functions.
For instance, if I(φ) = φ4 is the interaction term in the example of the φ4 theory,
R
then
∂I
Z
(ψ) = φψ3 dvol.
∂φ M
∂I
Thus ∂φ provides a cubic function on the space C ∞ (U), but it is not given by inte-
gration against an element in Cc∞ (U 3 ). Instead it is given by integrating against a
distribution on U 3 , namely, the delta-distribution on the diagonal.
We can try to solve this issue by using a larger class of polynomial functions.
Thus, we could let M
P(C ∞ (U)) = Dc (U n )S n
n
where Dc (U n ) is the space of compactly supported distributions on U. Similarly,
we could let M
Vectc (U) = Dc (U n+1 )S n .
n
The spaces P(C ∞ (U)) and Vectc (C ∞ (U)) are dense subspaces of these spaces.
7. INTERACTING THEORIES 35
More explicitly, this term involves pairing the distribution f φ on the diagonal in
M 2 with the delta-function on the diagonal.
If we consider the ~ → 0 limit of this putative operator ~ Div~ , we nonetheless
find a well-defined operator
where H 0 (Obscl (U)) is defined, as earlier, to be the quotient of P(C ∞ (U)) by the
ideal IEL (U) of the Euler-Lagrange equations.
This result is a version of elliptic regularity, and we prove it later, in Chapter 4.
Now the challenge we face should be clear. If S is the action functional for
the free field theory, then we have a factorization algebra of classical observables.
This factorization algebra deforms in two ways. First, we can deform it into the
factorization algebra of quantum observables for a free theory. Second, we can
deform it into the factorization algebra of classical observables for an interacting
field theory. The difficulty is to perform both of these deformations simultaneously.
To construct the observables of an interacting field theory, we use the renormal-
ization technique of Costello (2011b). In Costello (2011b), the first author gives a
definition of a quantum field theory and a cohomological method for constructing
field theories. A field theory as defined in Costello (2011b) gives us (essentially
36 2. FROM GAUSSIAN MEASURES TO FACTORIZATION ALGEBRAS
1. Prefactorization algebras
In this section we will give a formal definition of the notion of a prefactor-
ization algebra, starting concretely and then generalizing. In the first subsection,
using plain language, we describe a prefactorization algebra taking values in vec-
tor spaces. The reader is free to generalize by replacing “vector space” and “linear
map” with “object of a symmetric monoidal category C” and “morphism in C.”
(Our favorite target category is cochain complexes.) The next subsections give a
concise definition using the language of multicategories (also known as colored op-
erads) and allow an arbitrary multicategory as the target. In the final subsections,
we describe the category (and multicategory) of such prefactorization algebras.
1 ,...,U n
• There is a linear map mU V : F (U1 ) ⊗ · · · ⊗ F (Un ) → F (V) for
every finite collection of open sets where each Ui ⊂ V and the Ui are
pairwise disjoint. The following picture represents the situation.
V
U1
1 ,...,U n
mU
V : F (U1 ) ⊗ · · · ⊗ F (Un ) → F (V)
U2
... Un
• The maps are compatible in the obvious way, so that if Ui,1 t· · ·tUi,ni ⊆
Vi and V1 t · · · t Vk ⊆ W, the following diagram commutes.
Nk Nni Nk
i=1 j=1 F (U j ) i=1 F (Vi )
F (W)
Thus F resembles a precosheaf, except that we tensor the vector spaces rather than
take their direct sum.
For an explicit example of the associativity, consider the following picture.
V1 U1,2 W
F (U1,1 ) ⊗ F (U1,2 ) ⊗ F (U2,1 )
U1,1
V2
The case of k = n1 = 2, n2 = 1
These axioms imply that F (∅) is a commutative algebra. We say that F is
a unital prefactorization algebra if F (∅) is a unital commutative algebra. In this
case, F (U) is a pointed vector space by the image of the unit 1 ∈ F (∅) under the
structure map F (∅) → F (U). In practice, for our examples, F (∅) is C, R, C[[~]],
or R[[~]].
Example: The crucial example to bear in mind is an associative algebra. Every
associative algebra A defines a prefactorization algebra A f act on R, as follows.
` To
each open interval (a, b), we set A f act ((a, b)) = N
A. To any open set U = j I j,
where each I j is an open interval, we set F (U) = j A. The structure maps simply
arise from the multiplication map for A. Figure 1 displays the structure of A f act .
Notice the resemblance to the notion of an E1 or A∞ algebra. (One takes an infinite
1. PREFACTORIZATION ALGEBRAS 39
a⊗b⊗c ∈ A⊗A⊗A
ab ⊗ c ∈ A⊗A
abc ∈ A
tensor products of unital algebras, as follows. Given an infinite set I, consider the
Nby inclusion. For0each finite subset J J⊂ I, we
poset of finite subsets of I, ordered
J0
can take the tensor product A J = j∈J A. For J ,→ J , we define a map A → A
by tensoring with the identity 1 ∈ A for every j ∈ J 0 \J. Then AI is the colimit over
this poset.) ^
• For every (possibly empty) finite collection of open sets {Uα }α∈A and
open set V, there is a set of maps Disj M ({Uα }α∈A | V). If the Uα are
pairwise disjoint and all are contained in V, then the set of maps is a
single point. Otherwise, the set of maps is empty.
• The composition of maps is defined in the obvious way.
Remark: There is an important variation on this definition where one weakens the
requirement that the composition of structure maps holds “on the nose” and in-
stead requires homotopy coherence. For example, given disjoint opens U1 and U2
1 ,U 2
contained in V, which is then contained in W, we do not require that mU W =
V U1 ,U2
mW ◦ mV but that there is a “homotopy” between these maps. This kind of
situation arises naturally whenever the target category is best viewed as an ∞-
category, such as the category of cochain complexes. We will not develop here the
formalism necessary to treat homotopy-coherent prefactorization algebras because
our examples and constructions always satisfy the strictest version of composition.
The reader interested in seeing this variant developed should see the treatment in
Lurie (n.d.b). (We remark that one typically has “strictification” results that en-
sure that a homotopy-coherent algebra over a colored operad can be replaced by a
weakly-equivalent strict algebra over a colored operad, so that working with strict
algebras is sufficient for many purposes.) ^
1.3. Prefactorization algebras in the style of precosheaves. Any multicate-
gory C has an associated symmetric monoidal category SC, which is defined to be
the universal symmetric monoidal category equipped with a functor of multicate-
gories C → SC. We call it the symmetric monoidal envelope of C. Concretely, an
object of SC is a formal tensor product a1 ⊗ · · · ⊗ an of objects of C. Morphisms
in SC are characterized by the property that for any object b in C, the set of maps
1. PREFACTORIZATION ALGEBRAS 41
and we simply define the structure maps as the tensor product of the structure maps.
For instance, if U ⊂ V, then the structure map is
1.5.1 Definition. Let PreFAmc (X, C) denote the multicategory arising from the sym-
metric monoidal product on PreFA(X, SC). That is, if F1 , . . . Fn , G are prefactor-
ization algebras valued in C, we define the set of multi-morphisms to be
A
/ F ((a, b))
Id i(a,b)
(c,d)
A
/ F ((c, d))
commutes.
The product map m : A ⊗ A → A is defined as follows. Let a < b < c < d.
Then, the prefactorization structure on F gives a map
F ((a, b)) ⊗ F ((c, d)) → F ((a, d)),
and so, after identifying F ((a, b)), F ((c, d)) and F ((a, d)) with A, we get a map
A ⊗ A → A.
This is the multiplication in our algebra.
It remains to check the following.
((i)) This multiplication doesn’t depend on the intervals (a, b) t (c, d) ⊂
(a, d) we chose, as long as (a, b) < (c, d).
((ii)) This multiplication is associative and unital.
This is an easy (and instructive) exercise.
3. Modules as defects
We want to explain another simple but illuminating class of examples, and then
we apply this perspective in the context of quantum mechanics. We will work with
prefactorization algebras taking values in vector spaces with the tensor product as
symmetric monoidal structure.
a⊗m⊗b ∈ A⊗M⊗B
am ⊗ b ∈ M⊗B
amb ∈ M
interval I = (t0 , t1 ) such that t0 < p < t1 , we set F M (I) = M. The structure maps
are also determined by the bimodule structure. For example, given
s0 < s1 < t0 < p < t1 < u0 < u1 ,
we have
F M ((s0 , s1 ) t (t0 , t1 ) t (u0 , u1 )) = F M ((s0 , s1 )) ⊗ F M ((t0 , t1 )) ⊗ F M (u0 , u1 ))
= A⊗M⊗B
and the inclusion of these three intervals into (s0 , u1 ) is the map
F M ((s0 , s1 )) ⊗ F M ((t0 , t1 )) ⊗ F M )(u0 , u1 )) → F M ((s0 , u1 ))
.
a⊗m⊗b 7→ amb
The definition of a bimodule ensures that we have a prefactorization algebra. See
Figure 2.
There is a structure map that we have not discussed yet, though. The inclusion
of the empty set into an interval I containing p means that we need to pick an
element mI of M for each interval. The simplest case is to fix one element m ∈
M and simply use it for every interval. If we are assigning the unit of A as the
distinguished element for F M on every interval to the left of p and the unit of B for
every interval to the right of p, then the distinguished elements, then the structure
maps we have given clearly respect these distinguished elements.
These distinguished elements, however, can change with the intervals, so long
as they are preserved by the structure maps. For an interesting example, see the
discussion of quantum mechanics below.
Let us examine one more interesting case. Suppose we have algebras A, B, and
C, and an A − B-bimodule M and a B − C-bimodule N. There is a prefactorization
algebra on R describing the natural algebra for this situation.
Fix points p < q. Let F M,N be the prefactorization algebra on R such that
• on {x ∈ R | x < p}, it agrees with FA ,
• on {x ∈ R | p < x < q}, it agrees with FB ,
• on {x ∈ R | q < x}, it agrees with FC ,
and
• to intervals (t0 , t1 ) with t0 < p < t1 < q, it assigns M, and
• to intervals (t0 , t1 ) with p < t0 < q < t1 , it assigns N.
3. MODULES AS DEFECTS 45
In other words, on {x ∈ R | x < q}, this prefactorization algebra F M,N behaves like
F M , and like FN on {x ∈ R | p < x}.
We still need to describe what it does on an interval I of the form (T 0 , T 1 ) with
T 0 < p < q < T 1 . There is a natural choice, dictated by the requirement that we
produce a prefactorization algebra.
We know that F M,N (I) must receive maps from M and N, by considering
smaller intervals that only overlap either p or q. It also receives maps from A,
B, and C from intervals not hitting these marked points, but these factor through
intervals containing one of the marked points. Finally, we must have a structure
map
µ : M ⊗ N → F M,N (I)
for each pair of disjoint intervals hitting both marked points.
Note, in particular, what the associativity condition requires in the situation
where we have three disjoint intervals given by
s0 < p < s1 < t0 < t1 < u0 < q < u1 ,
contained in I. We can factor the inclusion of these three intervals through the pair
of intervals (s0 , t1 ) t (u0 , u1 ) or the pair of intervals (s0 , s1 ) t (t0 , u1 ) Thus, our
structure map
M ⊗ B ⊗ N → F M,N (I)
must satisfy that µ((mb)⊗n) = µ(m⊗(bn)) for every m ∈ M, b ∈ B, and n ∈ N. This
condition means that F M,N (I) receives a canonical map from the tensor product
M ⊗B N.
Hence, the most natural choice is to set F M,N (I) = M ⊗B N. One can make
other choices for how to extend to these longer intervals, but such a prefactorization
algebra will receive a map from this one. The local-to-global principle satisfied by
a factorization algebra is motivated by this kind of reasoning.
Remark: We have shown how thinking about prefactorization algebras on a real
line “decorated” with points (i.e., with a kind of stratification) reflects familiar
algebraic objects like bimodules. By moving into higher dimensions and allow-
ing more interesting submanifolds and stratifications, one generalizes this familiar
algebra into new, largely-unexplored directions. See Ayala et al. (n.d.a) for an ex-
tensive development of these ideas in the setting of locally constant factorization
algebras. ^
3.2. Standard quantum mechanics as a prefactorization algebra. We will
now explain how to express the standard formalism of quantum mechanics in
the language of prefactorization algebras, using the kind of construction just de-
scribed. As our goal is to emphasize the formal structure, we will work with a
finite-dimensional complex Hilbert space and avoid discussions of functional anal-
ysis.
Remark: In a sense, this section is a digression from the central theme of the book.
Throughout this book we take the path integral formalism as fundamental, and
hence we do not focus on the Hamiltonian, or operator, approach to quantum
46 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
physics. Hopefully, juxtaposed with our work in Section 3, this example clarifies
how to connect our methods with others. ^
Let V denote a finite-dimensional complex Hilbert space V. That is, there is a
nondegenerate symmetric sesquilinear form (−, −) : V × V → C so that
(λv, λ0 v0 ) = λλ0 (v, v0 )
where λ, λ0 are complex numbers and v, v0 are vectors in V. Let A = End(V)
T
denote the algebra of endomorphisms, which has a ∗-structure via M ∗ = M , the
conjugate-transpose. The space V is a representation of A, and the ∗-structure is
characterized by the property that
(M ∗ v, v0 ) = (v, Mv0 ).
It should be clear that one could work more generally with a Hilbert space equipped
with the action of a ∗-algebra of operators, aka observables.
Now that we have fixed the kinematics of the situation, let’s turn to the dynam-
ics. Let (Ut )t∈R be a one-parameter group of unitary operators on V. Since we are
in the finite-dimensional setting, there is no problem identifying
Ut = eitH
for some Hermitian operator H that we call the Hamiltonian. We view V as a state
space for our system, A as where the observables live, and H as determining the
time evolution of our system.
We now rephrase this structure to make it easier to articulate via the factoriza-
tion picture. Let V denote V with the conjugate complex structure. We will denote
elements of V by “kets” | v0 i and elements of V by “bras” hv | , and we provide a
bilinear pairing between them by
v | v0 = (v, v0 ).
We equip V with the right A-module structure by
hv | M = hM ∗ v |
We will write hv | M | v0 i as can think of M acting on v0 from the left or on v from
the right and it will produce the same number.
Our goal is to describe a scattering-type experiment.
• At time t = 0, we prepare our system in the initial state hvin | .
• We modify the governing Hamiltonian over some finite time interval
(i.e., apply an operator, or equivalently, an observable).
• At time t = T , we measure whether our system is in the final state
| vout i.
If we run this experiment many times, with the same initial and final states and the
same operator, we should find a statistical pattern in our data. If an operator O acts
during a time interval (t, t0 ), then we are trying to compute the number
0
hvin | eitH Oei(T −t )H | vout i.
3. MODULES AS DEFECTS 47
Our formalism does not include the idealized situation of an operator O acting at
a single moment t0 in time, but in the appropriate limit of shorter and shorter time
intervals around t0 , we would compute
hvin | eit0 H Oei(T −t0 )H | vout i,
which agrees with the usual prescription.
Note that we use a bra hvin | for the “incoming” state and a ket | vout i for
the“outgoing” state so that the left-to-right ordering will agree with the left-to-
right ordering of the real line viewed as a time-line. The prefactorization algebra F
on the interval [0, T ] describing this situation has the following structure. Interior
open intervals describe moments when operators can act on our system. An interval
that contains 0 (but not the other end) should describe possible “incoming” states
of the system; dually, an interval containing the other endpoint should describe
“outgoing” states. Let us now spell things out explicitly.
To open subintervals, our prefactorization algebra F assigns the following vec-
tor spaces:
• [0, t) 7→ V
• (s, t) 7→ A
• (t, T ] 7→ V
• [0, T ] 7→ C.
In light of our discussion about modules in the preceding section, note that the
natural choice for the value F([0, T ]) is the vector space V ⊗A V. However, V ⊗End(V)
V is isomorphic to the ground field C due to the compatibility of the left and right
actions with the inner product. We must now describe the structure maps coming
from inclusion of intervals; we will describe enough so that the mechanism is clear.
The case that determines the rest is that the inclusion
[0, t0 ) t (t1 , t2 ) t (t3 , T ] ⊂ [0, T ]
corresponds to the structure map
V ⊗A⊗V → C
.
hv0 | ⊗ O ⊗ | v1 i 7→ hv0 | ei(t1 −t0 )H Oei(t3 −t2 )H | v1 i
In other words, the system evolves according the Hamiltonian during the closed
intervals in the complement of opens during which we specify the incoming and
outgoing states and the operator.
Note that if we set O = ei(t2 −t1 )H , then we obtain
hv0 | ei(t1 −t3 )H | v1 i
and so recover the expected value of being in state hv0 | at time t0 and going to
state | v1 i at time t3 . Setting t3 = t0 , we recover the inner product on V.
For another example of structure maps, the inclusion (t0 , t1 ) ⊂ (t00 , t10 ) goes to
0 0
O 7→ ei(t0 −t0 )H Oei(t1 −t1 )H . More generally, for k disjoint open intervals inside a big
open interval
(t0 , t1 ) t · · · t (t2k−2 , t2k−1 ) ⊂ (t00 , t10 ),
48 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
we again time-order the operators and then multiply with evolution operators in-
serted for the closed intervals between them: the structure map for F is
0 0
O1 ⊗ · · · ⊗ Ok 7→ ei(t0 −t0 )H O1 ei(t2 −t1 )H · · · Ok ei(t1 −t2k−1 )H .
Note that these structure maps reduce to those for an arbitrary associative algebra
if we set H = 0 and hence have the identity map as the evolution operator. (In
general, a one-parameter semigroup of algebra automorphisms can be used like
this to “twist” the prefactorization algebra associated to an associative algebra.)
As further examples, we have:
0
• the inclusion [0, t) ⊂ [0, t0 ) has structure map hv0 | 7→ hv0 | ei(t −t)H .
0
• the inclusion (t, T ] ⊂ (t0 , T ] has structure map | v1 i 7→ ei(t−t )H | v1 i.
• the inclusion [0, t0 ) t (t1 , t2 ) ⊂ [0, t ) has structure map hv0 | ⊗ O 7→
0
0
hv0 | ei(t1 −t0 )H Oei(t −t2 )H .
All these choices ensure that we are free to choose what happens during open inter-
vals but that the system evolves according to H during their closed complements.
In this way, the prefactorization algebra encodes the basic abstract structure of
quantum mechanics. The open interior (0, T ) encodes the algebra of observables,
and the boundaries encode the state spaces.
The attentive reader might notice that we have not discussed, e.g., the structure
map associated to (t0 , t1 ) ⊂ [0, t0 ). Here we need to use the distinguished elements
in F((t0 , t1 )) = A and F([0, t0 )) = V arising from the inclusion of the empty set
into these opens. If we fix hvin | as the “idealized” initial state at time 0, then we
0
set hvin | eit H to be the distinguished element in F([0, t0 )) = V. The distinguished
element of F((t0 , t1 )) is the evolution operator ei(t1 −t0 )H . Thus, the structure map for
(t0 , t1 ) ⊂ [0, t0 ) is naturally
0
O 7→ hvin | eit0 H Oei(t −t1 )H
We now describe the dual situation for intervals containing the other endpoint.
Here we specify an “idealized” final state | vout i so that the distinguished element
of F((t, T ]) is ei(T −t)H | vout i. We do not need these initial or final states to recover
the quantum mechanical formalism from the prefactorization algebra, so it is inter-
esting that the prefactorization perspective pushes towards fixing these boundary
states in the form of the distinguished elements.
Remark: Although we have explained here how to start with the standard ingredi-
ents of quantum mechanics and encode them as a prefactorization algebra, one can
also turn the situation around and motivate (or interpret), via the factorization per-
spective, aspects of the quantum mechanical formalism. For instance, time-reversal
amounts to reflection across a point in R. Requiring a locally constant prefactoriza-
tion algebra on R to be equivariant under time-reversal corresponds to equipping
the corresponding associative algebra A with an involutive algebra antiautomor-
phism. In other words, the prefactorization algebra corresponds to a ∗-algebra A.
Likewise, suppose we want a prefactorization algebra on a closed interval [0, T ]
such that it corresponds to A on the open interior. The right end point corresponds
4. A CONSTRUCTION OF THE UNIVERSAL ENVELOPING ALGEBRA 49
to a left A-module V, viewed as the “outgoing” states. It is natural to want the “in-
coming” states — given by some right A-module V — to be “of the same type” as
V (e.g., abstractly isomorphic) and to want the global sections to be C. This forces
us to have a map
V ⊗ A ⊗ V → C.
This map induces a pairing between V and V, which provides a pre-Hilbert struc-
ture. ^
Remark: Our construction above captures much of the standard formalism of quan-
tum mechanics, but there are a few loose ends to address.
First, in standard quantum mechanics, a state is not a vector in V but a line.
Above, however, we fixed vectors hvin | and | vout i, so there seems to be a dis-
crepancy. The observation that rescues us is a natural one, from the mathematical
viewpoint. Consider scaling the initial and final states by elements of C× . This de-
fines a new factorization algebra, but it is isomorphic to what we described above,
and the expected value “hv0 |Ov1 i” of an operator depends linearly in the rescaling
of the input and output vectors. More precisely, there is a natural equivalence rela-
tion we can place on the factorization algebras described above that corresponds to
the usual notion of state in quantum mechanics. In other words, we could make a
groupoid of factorization algebras where the underlying vector spaces and structure
maps are all the same, but the distinguished elements are allowed to change.
Another issue that might bother the reader is that our formalism only matches
nicely with experiments that resemble scattering experiments. It does not seem
well-suited to descriptions of systems like bound states (e.g., an atom sitting qui-
etly, minding its own business). For such systems, we might consider running
over the whole space of states, which is described as a groupoid in the previous
paragraph. Alternatively, we might drop the endpoints and simply work with the
factorization algebra on the open interval, which focuses on the algebra of opera-
tors. ^
algebra taking value in dg Lie algebras equipped with direct sum ⊕ as the symmet-
ric monoidal product (here direct sum means the sum of the underlying cochain
complexes, which inherits a natural Lie bracket).
Let C∗ h denote the Chevalley-Eilenberg complex for Lie algebra homology,
written as a cochain complex. In other words, C∗ h is the graded vector space
Sym(h[1]) with a differential determined by the bracket of h. See Appendix 3 for
the definition and further discussion of this construction.
Our main result shows how to construct the universal enveloping algebra Ug
using C∗ (gR ).
4.0.1 Proposition. Let H denote the cohomology prefactorization algebra of C∗ (gR ).
That is, we take the cohomology of every open and every structure map, so
H(U) = H ∗ (C∗ (gR (U)))
for any open U. Then H is locally constant, and the corresponding associative
algebra is isomorphic to Ug, the universal enveloping algebra of g.
Remark: Recall Lemma 2.0.2, which says that every locally constant prefactoriza-
tion algebra on R corresponds to an associative algebra. This proposition above
provides a homological mechanism for recovering the universal enveloping alge-
bra of a Lie algebra, but the reader has probably noticed that we could apply the
same construction with R replaced by any smooth manifold M. In Chapter 6, we
will explain how to understand what this general procedure means. ^
Proof. Local constancy of H is immediate from the fact that, if I ⊂ J is the
inclusion of one interval into another, the map of dg Lie algebras
Ω∗c (I) ⊗ g → Ω∗c (J) ⊗ g
is a quasi-isomorphism. We let Ag be the associative algebra constructed from H
by Lemma 2.0.2.
The underlying vector space of Ag is the space H(I) for any interval I. To be
concrete, we will use the interval R, so that we identify
Ag = H(R) = H ∗ (C∗ (Ω∗c (R) ⊗ g)).
We now identify that vector space.
The dg Lie algebra Ω∗c (R) ⊗ g is concentrated in degrees 0 and 1 and maps
quasi-isomorphically to its cohomology Hc∗ (R) ⊗ g = g[−1], which is concentrated
in degree 1 by the Poincaré lemma. This cohomology is an Abelian Lie algebra
because the cup product on Hc∗ (I) is zero. It follows that C∗ (Ω∗c (R) ⊗ g) is quasi-
isomorphic to the Chevalley-Eilenberg chains of the Abelian Lie algebra g[−1],
which is simply Sym g. Thus, as a vector space, Ag is isomorphic to the symmetric
algebra Sym g.
There is a map
Φ : g → Ag
that sends an element X ∈ g to ε ⊗ X where ε ∈ Hc1 (I) is a basis element for the
compactly supported cohomology of the interval I; we require the integral of ε to
be 1. We will show that Φ is a map of Lie algebras, where Ag is given the Lie
4. A CONSTRUCTION OF THE UNIVERSAL ENVELOPING ALGEBRA 51
bracket coming from the associative structure. This result immediately implies the
theorem, as we then have an map of associative algebra Φ : Ug → Ag that is an
isomorphism of vector spaces.
Let us check explicitly that Φ is a map of Lie algebras. Let δ > 0 be small
number, and let f0 ∈ Cc∞ (−δ, δ) be a compactly supported smooth function with
f0 dx = 1. Let ft (x) = f0 (x−t). Note that ft is supported on the interval (t−δ, t+δ).
R
5.1. Differentiable vector spaces. The most common way to encode the an-
alytic structure on a vector space such as the space of smooth functions on a man-
ifold is to endow it with a topology. Homological algebra with topological vector
spaces is not easy, however. (For instance, topological vector spaces do not form
an abelian category.) To get around this issue, we will work with differentiable
vector spaces. Let us first define the slightly weaker notion of a C ∞ -module.
5.1.1 Definition. Let Mfld be the site of smooth manifolds, i.e., the category of
smooth manifolds and smooth maps between them, where a cover is a surjective
local diffeomorphism.
Let C ∞ denote the sheaf of rings on Mfld that assigns to any manifold M the
commutative algebra C ∞ (M). A C ∞ -module is a module sheaf over C ∞ on Mfld.
Almost all of the differentiable vector spaces we will consider are concrete
in nature. Indeed, most satisfy the formal definition of a being a concrete sheaf,
which we now explain. Let Set denote the category of sets and let Set(S , T ) denote
the collection of functions from the set S to the set T . For any sheaf F on Mfld
taking values in Set, there is a natural map F (M) → Set(M, F (∗)): each element
of the set F (M) has an associated function from the underlying set M to the set
F (∗), the value of the sheaf on a point. We say a sheaf F is concrete if this map
F (M) → Set(M, F (∗)) is injective. Hence, for a concrete sheaf, one can think
of any section on M as just a particular function from M to F (∗); a section is
just a “smooth” function on M with values in the set F (∗). As an example of a
concrete sheaf, consider the sheaf X associated to a smooth manifold X, where
X(M) = C ∞ (M, X). In this case, a section of the sheaf X really is just a function to
X. This sheaf just identifies which set-theoretic functions are smooth.
We often work with the differentiable vector space arising from a topological
vector space V, which, just like the example X, simply records which set-theoretic
maps are smooth. For this reason, we will normally think of a differentiable vector
space F as being an ordinary vector space, given by its value on a point F (∗),
together with extra structure. We will often refer to the sections F (M) (i.e., the
value of the sheaf F on the manifold M) as the space of smooth maps to the value
on a point. If V is a differentiable vector space, we often write C ∞ (M, V) for the
54 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
space of smooth maps from a manifold M. Abusively, we will also often call a map
of differentiable vector spaces simply a smooth map.
5.2. Differentiable cochain complexes. Although we typically work with dif-
ferentiable vector spaces coming from topological vector spaces, the category DVS
is much easier to use — for our purposes — than that of topological vector spaces.
The key reason is that they are (essentially) just sheaves on a site, and homological
algebra for such objects is very well-developed, as we explain in Appendix C.
5.2.1 Definition. A differentiable cochain complex is a cochain complex in the
category of differentiable vector spaces.
A cochain map f : V → W of differentiable cochain complexes is a quasi-
isomorphism if the map C ∞ (M, V) → C ∞ (M, W) is a quasi-isomorphism for all
manifolds M. This condition is equivalent to asking that the map be a quasi-
isomorphism at the level of stalks.
respect to the connections on the Vi and W. For every manifold M, and for every
vi ∈ Vi (M), we require that
Xn
∇Φ(v1 , . . . , vn ) = Φ(v1 , . . . , ∇vi , . . . , vn ) ∈ Ω1 (M, W).
i=1
We let DVS(V1 , . . . , Vn | W) denote this space of smooth multilinear maps.
that a map out of the bigger space is determined by the subspace. In such cases,
we are working only with bornological vector spaces and depending upon the fact
that di fβ is a full and faithful functor. ^
So far, however, we have not discussed how completeness of topological vector
spaces appears in this context. We need a notion of completeness for a topological
vector space that only depends on smooth maps to that vector space. The relevant
concept was developed in Kriegl and Michor (1997). We will view the category
BVS as being a full subcategory of DVS.
5.4.1 Definition. A topological vector space V ∈ BVS is c∞ -complete, or conve-
nient, if every smooth map c : R → V has an antiderivative.
We denote the category of convenient vector spaces and bounded linear maps
by CVS.
This completeness condition is a little weaker than the one normally studied
for topological vector spaces. That is, every complete topological vector space is
c∞ -complete.
Proposition. The full subcategory di fc : CVS ⊂ DVS is closed under the forma-
tion of limits, countable coproducts, and sequential colimits of closed embeddings.
We give CVS the multicategory structure inherited from BVS. Since BVS is a
full sub multicategory of DVS, so is CVS.
Theorem. The multicategory structure on CVS is represented by a symmetric
⊗β .
monoidal structure b
This symmetric monoidal structure is called the completed bornological tensor
product. If E, F ∈ CVS, this completed bornological tensor product is written
as Eb⊗β F. The statement that it represents the multicategory structure means that
smooth (equivalently, bounded) bilinear maps f : E1 × E2 → F are the same as
⊗β E2 → F, for objects E1 , E2 , F of CVS.
bounded linear maps f 0 : E1b
When it should cause no confusion, we may use the symbol ⊗ instead of b ⊗β
for this tensor product.
5.5. Examples from differential geometry. Let us now give some examples
of differentiable vector spaces. These examples will include the basic building
blocks for most of the factorization algebras we will consider.
Let E be a vector bundle on a manifold X. We let E (X) denote the vector
space of smooth sections of E on X, and we let Ec (X) denote the vector space of
compactly supported sections of E on X.
Let us give these vector spaces the structure of differentiable vector spaces, as
follows. If M is a manifold, we say a smooth map from M to E is a smooth section
of the bundle π∗X E on M × X. We denote this set of smooth maps C ∞ (M, E (X)).
Sending M to C ∞ (M, E (X)) defines a sheaf of C ∞ -modules on the site of smooth
manifolds with a flat connection, and so a differentiable vector space.
5. SOME FUNCTIONAL ANALYSIS 57
5.6. Multilinear maps and enriched spaces of maps. The category of dif-
ferentiable vector spaces has a natural tensor product. In other words, it is a sym-
metric monoidal category. The tensor product in DVS is very simple: if V, W are
58 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
differentiable vector spaces, then they are C ∞ -modules (by forgetting the flat con-
nections) so V ⊗C ∞ W is another C ∞ -module, but it inherits a natural flat connection
∇V ⊗ IdW + IdV ⊗∇W , so we obtain a new differentiable vector space.
One must be careful with this tensor product, however. From the point of view
of analysis, this tensor product is not very meaningful: it is somewhat similar to an
un-completed tensor product of topological vector spaces. For example, if M and N
are manifolds, then C ∞ (M) and C ∞ (N) naturally have the structure of differentiable
vector spaces. It is not true that
(†) C ∞ (M) ⊗C ∞ C ∞ (N) = C ∞ (M × N).
This issue means that our examples will not assign to a disjoint union of opens the
tensor product of values on the components.
The multicategory structure on DVS that we use coincides with this symmetric
monoidal structure. The multicategory structure is better behaved than the sym-
metric monoidal structure: when restricted to the full subcategory CVS the multi-
category becomes the one associated to the symmetric monoidal structure on CVS,
which has good analytical properties (and in particular satisfies the equality in (†)).
Similarly, there is an internal hom in the category of C ∞ -modules, and this
sheaf hom likewise inherits a natural flat connection. We denote this sheaf hom by
HomDVS (V, W). This sheaf hom is not as well-behaved as one would hope, how-
ever, and does not capture the concept of “smooth families of maps”. In particular,
it is not true that the value of the sheaf HomDVS (V, W) on a point is the vector
space DVS(V, W) of smooth maps from V to W. For any reasonable definition of
the notion of smooth family of maps parametrized by a manifold, a smooth family
of maps parametrized by a point should be simply a map, and the self-enrichment
given by HomDVS (V, W) does not satisfy this.
There is, however, another way to enrich the category DVS over itself that
better captures the notion of smooth family of maps. (For a careful treatment, see
Section 6.) Before we define this enrichment, we need the following definition.
5.6.1 Definition. For V a differentiable vector space and M a manifold, let C∞ (M, V)
denote the differentiable vector space whose value on a manifold N is C ∞ (N ×
M, V). The flat connection map
∇N,C∞ (M,V) : C ∞ (N, C∞ (M, V)) → Ω1 (N, C∞ (M, V))
is the composition of the flat connection
∇N×M : C ∞ (M × N, V) → Ω1 (M × N, V) = Ω1 (M, C∞ (N, V)) ⊕ Ω1 (N, C∞ (M, V))
with the projection onto Ω1 (N, C∞ (M, V)).
a ring object in the category DVS, and for any differentiable vector space V, the
mapping space C∞ (M, V) is a C ∞ (M)-module.
This construction generalizes in a natural way. If E is a vector bundle on a
manifold M and V is a differentiable vector space, then we can define
C∞ (M, E ⊗ V) := E (M) ⊗C ∞ (M) C∞ (M, V)
where on the right hand side we are taking tensor products in the category DVS
over the differentiable ring C ∞ (M). Although, in general, tensor products in DVS
are not analytically well-behaved, in this case there are no problems because E (M)
is a projective C ∞ (M) module of finite rank. We will denote C∞ (M, T ∗ M ⊗ E) by
Ω1 (M, E).
5.6.2 Definition. Let V1 , . . . , Vn , W be differentiable vector spaces. Given a mani-
fold M, a smooth family of multi-linear maps V1 × · · · × Vn → W parametrized by
M is an element of
DVS(V1 , . . . , Vn | C∞ (M, W))
where we regard C∞ (M, W) as being a differentiable vector space using the defini-
tion above.
where E F denotes the external tensor product of vector bundles and Γ denotes
smooth sections.
Remark: An alternative approach to the one we’ve taken is to use the category of
nuclear topological vector spaces, with the completed projective tensor product,
instead of the category of convenient (or differentiable) vector spaces. Using nu-
clear spaces raises a number of technical issues, but one immediate issue is the
⊗πC ∞ (Y) = C ∞ (X × Y), it seems not to
following: although it is true that C ∞ (X)b
be true that the same statement holds if we use compactly supported smooth func-
tions. The problem stems from the fact that the projective tensor product does not
commute with colimits, whereas the bornological tensor product does. ^
We can define symmetric powers of convenient vector spaces using the sym-
metric monoidal structure we have described. If, as before, E is a vector bundle on
X and U is an open subset of X, this proposition allows us to identify
Symn (Ec (U)) = Cc∞ (U n , E n )S n .
The symmetric algebra Sym Ec (U) is defined as usual to be the direct sum of the
symmetric powers. It is an algebra in the symmetric monoidal category of conve-
nient vector spaces.
A related construction is the algebra of functions on a differentiable vector
space. If V is a differentiable vector space, we can define, as we have seen, the
space of linear functionals on V to be the space of maps HomDVS (V, R). Because
the category DVS is self-enriched, this is again a differentiable vector space. In a
similar way, we can define the space of polynomial functions on V homogeneous
of degree n to be the space
Pn (V) = HomDVS (V, . . . , V | C ∞ )S n .
| {z }
n times
In other words, we take smooth multilinar maps from n copies of V to R, and
then take the S n coinvariants. The self-enrichment of DVS gives this the structure
of differentiable vector space. Concretely, a smoooth map from a manifold M to
Pn (V) is
C ∞ (M, Pn (V)) = HomDVS (V, . . . , V | C ∞ (M))S n .
| {z }
n times
One can then define the algebra of functions on V by
Y
O(V) = Pn (V).
n
(In this formula, we take the product rather than the direct sum, so that O(V) should
be thought of as a space of formal power series on V. One can, of course, also
consider the version using the sum). The space O(V) is a commutative algebra
object of the category DVS in a natural way.
This construction is a very general one, of course: one can define the algebra
of functions on any object in any multicategory in the same way.
An important example is the following.
6. THE FACTORIZATION ENVELOPE OF A SHEAF OF LIE ALGEBRAS 61
Note that O(E (U)) is naturally the same as HomDVS (Sym E (U), R), i.e., it is
the dual of the symmetric algebra of E (U).
6.1. The key idea. Thus, let M be a manifold. Let L be a fine sheaf of dg
Lie algebras on M. Let Lc denote the cosheaf of compactly supported sections of
L. (We restrict to fine sheaves so that taking compactly supported sections is a
straightforward operation.)
Remark: Note that, although Lc is a cosheaf of cochain complexes, and a pre-
cosheaf of dg Lie algebras, it is not a cosheaf of dg Lie algebras. This is because
colimits of dg Lie algebras are not the same as colimits of cochain complexes. ^
We can view Lc as a prefactorization algebra valued in the category of dg Lie
algebras with symmetric monoidal structure given by direct sum. Indeed, if {Ui } is
a finite collection of disjoint opens in M contained in the open V, there is a natural
map of dg Lie algebras
M
Lc (Ui ) = Lc (ti Ui ) → Lc (V)
i
6.2. Local Lie algebras. In practice, we will need an elaboration of this con-
struction which involves a small amount of analysis.
6.2.1 Definition. Let M be a manifold. A local dg Lie algebra on M consists of the
following data:
((i)) a graded vector bundle L on M, whose sheaf of smooth sections will
be denoted L
((ii)) a differential operator d : L → L, of cohomological degree 1 and
square 0;
6. THE FACTORIZATION ENVELOPE OF A SHEAF OF LIE ALGEBRAS 63
Remark: This definition will play an important role in our approach to interacting
classical field theories, developed in Volume 2. ^
If U ⊂ M, then L(U) is a topological vector space, because it is the space of
smooth sections of a graded vector bundle on U. We would like to form, as above,
the Chevalley-Eilenberg chain complex C∗ (Lc (U)). The underlying vector space
of C∗ (Lc (U)) is the (graded) symmetric algebra on Lc (U)[1]. We need to take
account of the topological structure on Lc (U) when we take the tensor powers of
Lc (U).
We explained how to do this in the Section 5: we define (Lc (U))⊗n to be the
tensor power defined using the completed projective tensor product on the topo-
logical vector space Lc (U). Concretely, if Ln denotes the vector bundle on M n
obtained as the external tensor product, then
(Lc (U))⊗n = Γc (U n , Ln )
is the space of compactly supported smooth sections of Ln on U n . Symmetric
(or exterior) powers of Lc (U) are defined by taking coinvariants of Lc (U)⊗n with
respect to the action of the symmetric group S n . The completed symmetric algebra
on Lc (U)[−1] that is the underlying graded vector space of C∗ (Lc (U)) is defined
using these completed symmetric powers. The Chevalley-Eilenberg differential is
continuous, and therefore defines a differential on the completed symmetric algebra
of Lc (U)[−1], giving us the cochain complex C∗ (Lc (U)).
Example: Let g be a Lie algebra. Consider the sheaf Ω∗R ⊗ g of dg Lie algebras
on R, which assigns to the open U the dg Lie algebra Ω∗ (U) ⊗ g. Then, as we
saw in detail in Section 4, the factorization envelope of this sheaf of Lie algebras
encodes the universal enveloping algebra Ug of g. Indeed, factorization algebras on
R with an additional “locally constant” property give rise to associative algebras,
and the associative algebra associated to U(Ω∗R ⊗ g) recovers the ordinary universal
enveloping algebra.
In the same way, for any Lie algebra g we can construct a factorization algebra
on Rn as the factorization envelope of Ω∗Rn ⊗ g. The resulting factorization algebra
is locally constant: it has the property that the inclusion map from one disc to
another is a quasi-isomorphism. A theorem of Lurie Lurie (n.d.b) tells us that
locally constant factorization algebras on Rn are the same as En algebras. The En
algebra we have constructed is the En -enveloping algebra of g. (See the discussion
in Section 4 for more about locally constant factorization algebras, En algebras,
and the En enveloping functor.) ^
6.3. Shifted central extensions and the twisted envelope. Many interesting
factorization algebras — such as the Kac-Moody factorization algebra, and the
factorization algebra associated to a free field theory — can be constructed from a
64 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
variant of the factorization envelope construction, which we call the twisted factor-
ization envelope.
6.3.1 Definition. Let L be a sheaf of dg Lie algebras on M. A k-shifted central
extension of Lc is a precosheaf of dg Lie algebras L
ec fitting into an exact sequence
0 → C[k] → L
ec → Lc → 0
of precosheaves, where C[k] is the constant presheaf that assigns the one-dimensional
vector space C[k] in degree −k to every open.
If L is a local Lie algebra, we require in addition that locally there is a splitting
ec (U) = C[k] ⊕ Lc (U)
L
such that the differential and bracket maps from Lc (U) → C[k] and Lc (U)⊗2 →
C[k] are continuous.
6.3.2 Definition. In this situation, the twisted factorization envelope is the prefac-
torization algebra Ue L that sends an open set U to C∗ (L ec (U)). (In the case that L
is a local dg Lie algebra, we use the completed tensor product as above.)
The chain complex C∗ (L ec (U)) is a module over chains on the Abelian Lie al-
gebra C[k] for every k. Thus, we will view the twisted factorization envelope as a
prefactorization algebra in modules for C[c] where c has degree −k − 1.
Under the assumption that L is a homotopy sheaf, Theorem 6.0.1 shows that
the twisted factorization envelope UL e is a factorization algebra over the base ring
C[c]. Of particular interest is the case when k = −1, so that the central parameter c
is of degree 0.
Let us now introduce some important examples of this construction.
Example: Let g be a Lie algebra and consider the local Lie algebra Ω∗R ⊗ g on the
real line R, which we will denote gR . Given a skew-symmetric, invariant bilinear
form ω on g, there is a natural shifted extension of gRc where
Z
[α ⊗ X, β ⊗ Y]ω = α ∧ β ⊗ [X, Y] + α ∧ β ω(X, Y) c,
R
where we use c to denote the generator in degree 1 of the central extension. Let
Uω gR denote the twisted factorization envelope for this central extension. By mim-
icking the proof of Lemma 4.0.1, one can see that the cohomology of this twisted
factorization envelope recovers the enveloping algebra Ub g of the central extension
of g given by ω. ^
Example: Let g be a simple Lie algebra, and let h−, −ig denote a symmetric invari-
ant pairing on g. We define the Kac-Moody factorization algebra as follows.
Let Σ be a Riemann surface, and consider the local Lie algebra Ω0,∗ ⊗ g on Σ,
which sends an open subset U to the dg Lie algebra Ω0,∗ (U) ⊗ g. The differential
here is ∂,
¯ so we are describing the Dolbeault analog of the de Rham construction
in Section 4.
7. EQUIVARIANT PREFACTORIZATION ALGEBRAS 65
7.1. Discrete equivariance. We begin with the case where G is viewed sim-
ply as a group (i.e., we do not require any compatibility between the action and a
possible smooth structure on G). We give a concrete definition.
7.1.1 Definition. Let F be a prefactorization algebra on M. We say F is G-
equivariant if for each g ∈ G and each open subset U ⊂ M we are given iso-
morphisms
σg : F (U) −
→ F (gU)
satisfying the following conditions.
((i)) For all g, h ∈ G and all opens U, σg ◦ σh = σgh : F (U) → F (ghU).
((ii)) Every σg respects the factorization product. That is, for any finite tuple
U1 , . . . , Uk of disjoint opens contained in an open V, the diagram
F (U1 ) ⊗ · · · ⊗ F (Un ) / F (gU1 ) ⊗ · · · ⊗ F (gUn )
F (V) / F (gV)
commutes.
There is a map of colored operads Disj M → DisjGM , sending each open to itself
and sending a nonempty operation U1 , . . . , Un → V to the identity in Gn . Hence,
given an algebra A over DisjGM , there is an “underlying” prefactorization algebra
on M.
Next, observe that for an open set U, the set DisjGM (U | gU) is the coset g ·
Stab(U), where the stabilizer subgroup Stab(U) ⊂ G consists of elements in G
that preserve U. Hence for any algebra A over DisjGM , we see that there is an
isomorphism σg : A(U) → A(gU) for every g ∈ G and that they compose in the
natural way. Hence the “underlying” prefactorization algebra is G-equivariant.
Now observe that every operation in DisjGM factors as a collection of unary
operations σg arising from the G-action followed by a operation from the “under-
lying” prefactorization algebra. (If the input opens Ui already sit in the output open
V, we are done. Otherwise, pick elements gi of G that move the input opens inside
V and keep them pairwise disjoint.) Hence we obtain the following lemma.
7.2.2 Lemma. For G a group, every G-equivariant prefactorization algebra on M
produces a unique algebra over DisjGM , and conversely.
Some notation will make it easier to understand how the G-action intertwines
with the structure of the prefactorization algebra A. If (g1 , . . . , gn ) ∈ DisjGM (U1 , . . . , Un |V),
then we denote the associated operation by
m(g1 ,...,gk ) : A(U1 ) ⊗ · · · ⊗ A(Uk ) → A(V).
It can understood as the composition
O ⊗σgi O
A(Ui ) −−−→ A(gi Ui ) → A(V),
i i
where the second map is the structure map of the prefactorization algebra and the
first map is given by the unary operations arising from the action of G on M.
7.3. Smooth equivariance. From now on, we focus upon the situation where
G is a Lie group acting smoothly on a smooth manifold M. In this situation, an
algebra A over DisjGM has an underlying prefactorization algebra on M but now we
want the operations to vary smoothly as the input opens are moved by elements of
G. To accomplish this, we need to equip with sets of operations with a “smooth
structure” and we need the target category, in which an algebra takes values, to
admit a notion of “smoothly varying family of multilinear maps.”
Throughout this book, our prefactorization algebras take values in the category
DVS of differentiable vector spaces or in the category Ch(DVS) of cochain com-
plexes of differentiable vector spaces. We view these categories as enriched over
themselves. In the case of DVS, the self-enrichement is discussed in Section 5.6.
If V, W are differentiable vector spaces, we denote by HomDVS (V, W) the differen-
tiable vector space of maps from V to W. The key feature of HomDVS (V, W) is that
its value on a point is DVS(V, W) and, more generally, its value on a manifold X is
smooth families over X of maps of differentiable vector spaces from V to W. The
self-enrichment of DVS leads, in an obvious way, to a self-enrichment of Ch(DVS).
68 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
7.3.1. The colored operad. Equipping each set of operations DisjGM (U1 , . . . , Un |V)
with some kind of smooth structure is a little subtle. We might hope such a set,
which is a subset of Gn , inherits a manifold structure from Gn , but it is easy to pro-
duce examples of opens Ui and V where the set of operations is far from being a
submanifold of Gn . We take the following approach instead. Given a subset S ⊂ X
of a smooth manifold, there is a set of smooth maps Y → X that factor through S
for any smooth manifold Y. In other words, S provides a sheaf of sets on the site
Mfld of smooth manifolds, sometimes known as a “generalized manifold.” (See
Definition 2.0.1 for the site.)
Let Shv(Mfld) denote the category of sheaves (of sets). Then every collection
of operations
DisjGM (U1 , . . . , Un | V)
provides such a sheaf, so we naturally obtain a colored operad enriched in “gener-
alized manifolds,” i.e., Shv(Mfld).
g GM denote the colored operad where the set of colors is the
7.3.1 Definition. Let Disj
g GM (U1 , . . . , Un | V) are the sheaves
set of opens in M and where the operations Disj
on Mfld determined by the subset
DisjGM (U1 , . . . , Un | V) ⊂ Gn .
G
Let us unpack what this definition means. An algebra F over Disjg M in DVS
associates to each open U ⊂ M, a differentiable vector space F (U). To each finite
collection of opens U1 , . . . , Un , V, we have a map of sheaves
g GM (U1 , . . . , Un | V) → HomDVS (F (U1 ), . . . , F (Un ) | F (V)).
Disj
Thus, for each smooth manifold Y and for each smooth map Y → Gn factoring
through DisjGM (U1 , . . . , Un | V), we obtain a section in
C ∞ (Y, HomDVS (F (U1 ), . . . , F (Un ) | F (V)),
which encodes a Y-family of multilinear morphisms from the F (Ui ) to F (V).
Note that we can evaluate each sheaf Disjg GM (U1 , . . . , Un | V) at the point ∗ ∈
Mfld to obtain the underlying set DisjGM (U1 , . . . , Un | V). Thus, each algebra over
g GM has an underlying G-equivariant prefactorization algebra.
Disj
An algebra over this new colored operad is very close to what we need for our
purposes in this book. What’s missing so far is the ability to differentiate the action
of G on the prefactorization algebra to obtain an action of the Lie algebra g.
Ideally, this action of g would simply exist. Instead, we will put it in by hand,
as data, since that suffices for our purposes. After giving our definition, however,
we will indicate a condition on an algebra over DisjGM that provides the desired
action automatically.
7.3.2. Derivations. First, we need to introduce the notion of a derivation of a
prefactorization algebra on a manifold M. We will construct a differential graded
Lie algebra of derivations of any prefactorization algebra.
7. EQUIVARIANT PREFACTORIZATION ALGEBRAS 69
a relation known as the Leibniz rule. We simply write down the analog for an
algebra over the colored operad Disj M taking values in cochain complexes.
7.3.2 Definition. A degree k derivation of a prefactorization algebra F is a collec-
tion of maps DU : F (U) → F (U) of cohomological degree k for each open subset
U ⊂ M, with the property that for any finite collection of pairwise disjoint opens
U1 , . . . , Un , all contained in an open V, and an element αi ∈ F (Ui ) for each open,
the derivation acts by a Leibniz rule on the structure maps:
1 ,...,U n 1 ,...,U n
X
DV mU V (α1 , . . . , α n ) = (−1)k(|α1 |+···+|αi−1 |) mU
V (α1 , . . . , DUi αi , . . . , αn ),
i
Let Derk (F ) denote the derivations of degree k. It is easy to verify that the
derivations of all degrees Der∗ (F ) forms a differential graded Lie algebra. The
differential is defined by (dD)U = [dU , DU ], where dU is the differential on F (U).
The Lie bracket is defined by [D, D0 ]U = [DU , D0U ].
The concept of derivation allows us to talk about the action of a dg Lie algebra
g on a prefactorization algebra F . Such an action is simply a homomorphism of
differential graded Lie algebras from g to Der∗ (F ).
7.3.3. The main definition. We now provide the main definition. As our pref-
actorization algebra F takes values in the category of differentiable cochain com-
plexes, it makes sense to differentiate an element of F (U) for any open U.
7.3.3 Definition. A smoothly G-equivariant prefactorization algebra on M is an
algebra F over Disj g GM and an action of the Lie algebra g of G on the under-
lying prefactorization algebra of F such that for every X ∈ g, every operation
g GM (U1 , . . . , Un | V), and every 1 ≤ i ≤ n, we have
(g1 , . . . , gn ) ∈ Disj
∂
mg ,...,g (α1 , . . . , αn ) = mg1 ,...,gn (α1 , . . . , X(αi ), . . . , αn ).
∂Xi 1 n
∂
On the left hand side, ∂X i
indicates the action of the left-invariant vector field on
G associated to X in the ith factor of Gk and zero in the remaining factors.
k
Remark: In some cases, an algebra A over Disj g GM should possess a natural action of
g. We want to recover how g acts on each value A(U) from the action of G on A.
Suppose V is an open such that gU ⊂ V for every g in some neighborhood of the
identity in G. Then we can differentiate the structure maps mg : A(U) → A(V) to
obtain a map X : A(U) → A(V) for every X ∈ g. If A(U) = limV⊃U A(V), we ob-
tain via this limit, an action of g on A(U). Hence, if we have A(U) = limV⊃U A(V)
70 3. PREFACTORIZATION ALGEBRAS AND BASIC EXAMPLES
∨e f /~ ω0
PVi (M) / Ωn−i (M)
Div~ ddR
∨e f /~ ω0
PVi−1 (M) / Ωn−i+1 (M)
commutes. Here we use ∨e f /~ ω0 to denote the contraction map; it is the natural ex-
tension to polyvector fields of the contractions we defined for functions and vector
fields.
In summary, after contracting with the volume e f /~ ω0 , the divergence complex
becomes the de Rham complex. It is easy to check that, as maps from PVi (M) to
PVi−1 (M), we have
Div~ = ∨~−1 d f + Divω0
where ∨d f is the operator of contracting with the 1-form d f . In the ~ → 0 limit,
the dominant term is ∨~−1 d f .
More precisely, by multiplying the differential by ~, we see that there is a flat
family of cochain complexes over the algebra C[~] with the properties that
((i)) the family is isomorphic to the divergence complex when ~ , 0, and
((ii)) at ~ = 0, the cochain complex is
∨d f ∨d f
· · · → PV2 (M) −−−→ PV1 (M) −−−→ PV0 (M).
Note that this second cochain complex is a differential graded algebra, which is not
the case for the divergence complex.
The image of the map ∨d f : PV1 (M) → PV0 (M) is the ideal cutting out the
critical locus. Indeed, this whole complex is the Koszul complex for the equations
cutting out the critical locus. This observation leads to the following definition.
1.1.1 Definition. The derived critical locus of f is the locally dg ringed space
whose underlying manifold is M and whose dg commutative algebra of functions
is the complex PV∗ (M) with differential ∨d f .
where 4 is the Laplacian on M. (Normally we will reserve the symbol 4 for the
Batalin-Vilkovisky Laplacian, but that’s not necessary in this section.) This func-
tional is not well-defined on a field φ unless φ4φ is integrable, but it is helpful to
bear in mind that for classical field theory, what is crucial is the Euler-Lagrange
equations, which in this case is 4φ = 0 and which is thus well-defined on all
smooth functions. The action can be viewed here as a device for producing these
partial differential equations.
If U ⊂ M is an open subset, then the space of solutions to the equation of
motion on U is the space of harmonic functions on U. In this book, we will always
consider the derived space of solutions of the equation of motion. (For more details
about the derived philosophy, the reader should consult Volume 2.) In this simple
situation, the derived space of solutions to the free field equations, on an open
subset U ⊂ M, is the two-term complex
4
E (U) = C ∞ (U) −→ C ∞ (U)[−1] ,
We also want to keep track of where measurements are taking place on the
manifold M, so we will organize the observables by where they are supported. The
classical observables with support in U ⊂ M are then the symmetric algebra of
4
E (U)∨ = Dc (U)[1] −→ Dc (U) ,
for compactly supported sections of E ! . Note that there is a natural map of cochain
complexes Ec! (U) → E (U)∨ , given by viewing a compactly supported function as
a distribution.
2.1.1 Lemma. The inclusion map Ec! (U) → E (U)∨ is a cochain homotopy equiv-
alence of differentiable cochain complexes.
Proof. This assertion is a special case of a general result proved in Appen-
dix D. Note that by differentiable homotopy equivalence we mean that there is an
“inverse” map E (U)∨ → Ec! (U), and differentiable cochain homotopies between
the two composed maps and the identity maps. “Differentiable” here means that
all maps are in the category DVS of differentiable vector spaces. As these are
convenient cochain complexes, suffices to construct a continuous homotopy equiv-
alence.
This lemma says that, since we are working homotopically, we can replace
a distributional linear observable by a smooth linear observable. In other words,
any distributional observable that is closed in the cochain complex E (U)∨ is chain
homotopy equivalent to a closed smooth observable.We think of the smooth linear
observables as “smeared.” For example, we can replace a delta function δ x by some
bump function supported near the point x.
The observables we will work with is the space of “smeared observables”,
defined by
4
Obscl (U) = Sym(Ec! (U)) = Sym(Cc∞ (U)[1] −
→ Cc∞ (U)),
2. THE PREFACTORIZATION ALGEBRA OF A FREE FIELD THEORY 79
the action functional. Before making this idea precise, we recall the finite-dimensional
model of a quadratic function Q(x) = (x, Ax) on a vector space V. The contrac-
tion of the 1-form dQ with a tangent vector v0 ∈ T 0 V gives the linear functional
x 7→ (v0 , Ax). As S is quadratic in φ, we expect that the contraction of dS with a
tangent vector φ0 inC ∞ (U) is the linear functional
Z
φ 7→ 21
φ0 4φ.
But we run into an issue here: the functional S is not well-defined for all fields φ,
because the integral may not converge, and similarly the linear functional above is
not well-defined for arbitrary smooth functions φ0 and φ. However, the expression
∂S
Z
(φ) = 12 φ0 4φ
∂φ0
does make sense for any φ ∈ C ∞ (U) and φ0 ∈ Cc∞ (U). (Note that we now only
consider tangent vectors φ0 with compact support.) In other words, the desired
one-form dS does not make sense as a section of the cotangent bundle of C ∞ (U),
but it is well-defined as a section of the space T c∗C ∞ (U), the dual of the subbun-
dle T cC ∞ (U) ⊂ TC ∞ (U) describing vector fields along the leaves. This leafwise
one-form is closed. Such one-forms are the kinds of things we can contract with
elements of PVc (C ∞ (U)). The differential on PVc (C ∞ (U)) thus matches the differ-
ential on Obscl given by contracting with dS .
2.3. General free field theories. With this example in mind, we introduce a
general definition.
2.3.1 Definition. Let M be a manifold. A free field theory on M is the following
data:
((i)) A graded vector bundle E on M, whose sheaf of sections will be de-
noted E , and whose compactly supported sections will be denoted Ec .
((ii)) A differential operator d : E → E , of cohomological degree 1 and
square zero, making E into an elliptic complex.
((iii)) Let E ! = E ∨ ⊗ Dens M , and let E ! be the sections of E ! . Let d! be the
differential on E ! which is the formal adjoint to the differential on E .
Note that there is a natural pairing between Ec (U) and E ! (U), and this
pairing is compatible with differentials.
We require an isomorphism E → E ! [−1] compatible with differ-
entials, with the property that the induced pairing of cohomological
degree −1 on each Ec (U) is graded anti-symmetric.
2. THE PREFACTORIZATION ALGEBRA OF A FREE FIELD THEORY 81
The complex E (U) is the derived version of the solutions to the equations of
motion of the theory on an open subset U. More motivation for this definition is
presented in Volume 2.
Note that the equations of motion for a free theory are always linear, so that
the space of solutions is a vector space. Similarly, the derived space of solutions
of the equations of motion of a free field theory is a cochain complex, which is
a linear derived stack. The cochain complex E (U) should be thought of as the
derived space of solutions to the equations of motion on an open subset U. As we
explain in Volume 2, the pairing on Ec (U) arises from the fact that the equations
of motion of a field theory are not arbitrary differential equations, but describe the
critical locus of an action functional.
For example, for the free scalar field theory on a manifold M with mass m, we
have, as above,
4+m2
E = C ∞ (U) −−−−→ C ∞ (U)[−1].
∂ 2
Our convention is that 4 is a non-negative operator, so that on Rn , 4 = −
P
∂xi .
The pairing on Ec (U) is defined by
D E Z
φ0 , φ1 = φ0 φ1
M
for φk
in the graded piece C ∞ (U)[k].
As another example, let us describe Abelian Yang-Mills theory (with gauge
group R) in this language. Let M be a manifold of dimension 4. If A ∈ Ω1 (M) is
a connection on the trivial R-bundle on a manifold M, then the Yang-Mills action
functional applied to A is
Z Z
S Y M (A) = − 2
1
dA ∧ ∗dA = 2
1
A(d ∗ d)A.
M M
The equations of motion are that d ∗ dA = 0. There is also gauge symmetry, given
by X ∈ Ω0 (M), which acts on A by A → A + dX. The complex E describing this
theory is
d d∗d d
E = Ω0 (M)[1] →
− Ω1 (M) −−→ Ω3 (M)[−1] →
− Ω4 (M)[−2].
We explain how to derive this statement in Volume 2. For now, note that H 0 (E ) is
the space of those A ∈ Ω1 (M) which satisfy the Yang-Mills equation d ∗ dA = 0,
modulo gauge symmetry.
For any free field theory with cochain complex of fields E , we define the clas-
sical observables of the theory as
Obscl (U) = Sym(Ec! (U)) = Sym(Ec (U)[1]).
It is clear that classical observables form a prefactorization algebra (recall example
1.1). Indeed, Obscl (U) is a commutative differential graded algebra for every open
U. If U ⊂ V, there is a natural algebra homomorphism
iU cl cl
V : Obs (U) → Obs (V),
which on generators is the extension-by-zero map Cc∞ (U) → Cc∞ (V).
82 4. FREE FIELD THEORIES
i=1
which is the product of the value that each observable αi takes on φ restricted to
Ui .
2.4. The one-dimensional case, in detail. This space is particularly simple
in dimension 1. Indeed, we recover the usual answer at the level of cohomology.
2.4.1 Lemma. If U = (a, b) ⊂ R is an interval in R, then the algebra of classical
observables for the free field with mass m ≥ 0 has cohomology
H ∗ (Obscl ((a, b))) = R[p, q],
the polynomial algebra in two variables.
Proof. The idea is the following. The equations of motion for free classical
mechanics on the interval (a, b) are that the field φ satisfies (4 + m2 )φ = 0. This
space is two dimensional, spanned by the functions {e±mx }, for m > 0, and by the
functions {1, x}, for m = 0. Classical observables are functions on the space of so-
lutions to the equations of motion. We would this expect that classical observables
are a polynomial algebra in two generators.
We need to be careful, however, because we use the derived version of the
space of solutions to the equations of motion. We will show that the complex
2
!
−1 4+m
Ec ((a, b)) = Cc ((a, b)) −−−−→ Cc ((a, b))
! ∞ ∞ 0
φ p (x) = 2m
1
emx − e−mx .
For any value of m, the functions φ p and φq are annihilated by the operator −∂2x +m2 ,
and they form a basis for the kernel of this operator. Further, φ p (0) = 1 and φq (0) =
0, whereas φ0p (0) = 0 and φ0q (0) = 1. Finally, φq = φ0p .
Define a map
π : Ec! ((−1, 1)) → R{p, q}
2. THE PREFACTORIZATION ALGEBRA OF A FREE FIELD THEORY 83
by Z Z
g 7→ π(g) = q g(x)φq dx + p g(x)φ p dx.
The fact that G is the Green’s function implies that this operator is the inverse to
4 + m2 . It is clear that the operator of convolution with G is smooth (and even
continuous), so the result follows.
2.5. The Poisson bracket. We now return to the general case and construct
the P0 algebra structure on classical observables.
Suppose we have any free field theory on a manifold M, with complex of fields
E . Classical observables are the symmetric algebra Sym Ec (U)[1]. Recall that the
complex Ec (U) is equipped with an antisymmetric pairing of cohomological degree
−1. Thus, Ec (U)[1] is equipped with a symmetric pairing of degree 1.
2.5.1 Lemma. There is a unique smooth Poisson bracket on Obscl (U) of cohomo-
logical degree 1, with the property that
{α, β} = hα, βi
for any two linear observables α, β ∈ Ec (U)[1].
Recall that “smooth” means that the Poisson bracket is a smooth bilinear map
{−, −} : Obscl (U) × Obscl (U) → Obscl (U)
as defined in Section 5.
84 4. FREE FIELD THEORIES
Proof. The argument we will give is very general, and it applies in any reason-
able symmetric monoidal category. Recall that, as stated in Section 5, the category
of convenient vector spaces is a symmetric monoidal category with internal Homs
and a hom-tensor adjunction.
Let A be a commutative algebra object in the category Ch(CVS) of convenient
cochain complexes, and let M be a dg A-module. Then we define Der(A, M) to
be the space of algebra homomorphisms φ : A → A ⊕ M that are the identity on
A modulo the ideal M. (Here A ⊕ M is given the “square-zero” algebra structure,
where the product of any two elements in M is zero and the product of an element
in A with one in M is via the module structure.)
Since the category of convenient cochain complexes has internal Homs, this
cochain complex Der(A, M) is again an A-module in Ch(CVS).
The commutative algebra
Obscl (U) = Sym Ec! (U)
is the initial commutative algebra in the category Ch(CVS) of convenient cochain
complexes equipped with a smooth linear cochain map Ec! (U) → Obscl (U). (We
are simply stating the universal property characterizing Sym.) It follows that for
any dg module M in Ch(CVS) over the algebra Sym Ec! (U),
Der(Sym Ec! (U), M) = HomDVS (Ec! (U), M).
A Poisson bracket on Sym Ec! (U) is, in particular, a biderivation. A biderivation
is something that assigns to an element of Sym Ec! (U) a derivation of the algebra
Sym Ec! (U). Thus, the space of biderivations is the space
Der Sym Ec! (U), Der Sym Ec! (U), Sym Ec! (U) .
What we have said so far thus identifies the space of biderivations with
HomDVS (Ec! (U), HomDVS (Ec! (U), Sym Ec! (U))),
or equivalently with
HomDVS (Ec! (U) ⊗ Ec! (U), Sym Ec! (U))
via the hom-tensor adjunction in the category Ch(CVS).
The Poisson bracket we are constructing corresponds to the biderivation which
is the pairing on Ec! (U) viewed as a map
Ec! (U) ⊗ Ec! (U) → R = Sym0 Ec! (U).
This biderivation is antisymmetric and satisfies the Jacobi identity. Since Poisson
brackets are a subspace of biderivations, we have proved both the existence and
uniqueness clauses.
Note that for U1 , U2 disjoint open subsets of V and for observables αi ∈
Obscl (Ui ), we have
{iU
V α1 , iV α2 } = 0
1 U2
in Obscl (V). That is, observables coming from disjoint open subsets commute with
respect to the Poisson bracket. This property of Obscl (U) is the definition of a P0
2. THE PREFACTORIZATION ALGEBRA OF A FREE FIELD THEORY 85
where the differential arises from the Lie bracket and differential on Ebc (U). The
symbol [1] indicates a shift of degree down by one. (Recall that we always work
with cochain complexes, so our grading convention of C∗ is the negative of one
common convention.)
Remark: Those readers who are operadically inclined might notice that the Lie al-
gebra chain complex of a Lie algebra g is the E0 version of the universal enveloping
algebra of a Lie algebra. Thus, our construction is an E0 version of the familiar
86 4. FREE FIELD THEORIES
By PVc (C ∞ (U)) we mean polyvector fields along the foliation T cC ∞ (U) ⊂ TC ∞ (U)
of compactly-supported variants of a field. Concretely, the cochain complex Obscl (U)
is
∨dS ∨dS ∨dS
· · · −−−→ ∧2Cc∞ (U) ⊗ Sym Cc∞ (U) −−−→ Cc∞ (U) ⊗ Sym Cc∞ (U) −−−→ Sym Cc∞ (U),
⊗β .
where all tensor products appearing in this expression are b
In a similar way, the cochain complex of quantum observables Obsq (U) in-
volves just a modification of the differential. It is
∨dS +~ Div ∨dS +~ Div
. . . −−−−−−−−→ Cc∞ (U) ⊗ Sym Cc∞ (U) −−−−−−−−→ Sym Cc∞ (U).
The operator Div is the extension to all polyvector fields of the operator (see Defi-
nition 2.0.1) defined in Chapter 2 as a map from polynomial vector fields on C ∞ (U)
to polynomial functions. Thus, H 0 (Obsq (U)) is the same vector space (and same
prefactorization algebra) that we defined in Chapter 2.
As a graded vector space, there is an isomorphism
but it does not respect the differentials. In particular, the differential d on the quan-
tum observables satisfies
((i)) modulo ~, d coincides with the differential on Obscl (U), and
((ii)) the equation
Here, · indicates the commutative product on Obscl (U). As we will explain in Vol-
ume 2, these properties imply that Obsq defines a prefactorization algebra valued
in Beilinson-Drinfeld algebras and that Obsq quantizes the prefactorization algebra
Obscl valued in P0 algebras. (See Section 2.2 for the definition of these types of
algebras.) It is a characterizing feature of the Batalin-Vilkovisky formalism that
the Poisson bracket measures the failure of the differential d to be a derivation, and
the language of P0 and BD algebras is an operadic formalization of this concept.
We prove in Section 5 that Obscl is a factorization algebra, which implies Obsq is
a factorization algebra over R[~] .
3. QUANTUM MECHANICS AND THE WEYL ALGEBRA 87
preserves the cocycle defining the central extension. Therefore, it acts naturally
on the Chevalley-Eilenberg chains of the Heisenberg algebra. One can check that
∂
this operator is a derivation for the factorization product. That is, the operator ∂x
commutes with inclusions of one open subset into another, and if U, V are disjoint
and α ∈ Obsq (U), β ∈ Obsq (V), we have
∂ ∂ ∂
(α · β) = (α) · β + α · (β) ∈ Obsq (U t V).
∂x ∂x ∂x
In other words, the prefactorization algebra is, in a sense, translation-invariant!
(Derivations, and this interpretation, are discussed in more detail in Section 8.)
∂
It follows immediately that ∂x defines a derivation of the associative algebra
Am coming from the cohomology of Obsq ((0, 1)).
∂
Remark: These observations about ∂x apply to the classical observables, too. In-
deed, one discovers that infinitesimal translation induces a derivation of the Poisson
88 4. FREE FIELD THEORIES
algebra R[p, q]. As this derivation preserves the Poisson bracket, we know there is
an element H of R[p, q] such that {H, −} is the derivation. Here, H = p2 − m2 q2 .
(On a symplectic vector space, every symplectic vector field is Hamiltonian.) ^
3.0.2 Definition. The Hamiltonian H is the derivation of the associative algebra
∂
Am arising from the derivation − ∂x of the prefactorization algebra Obsq of observ-
ables of the free scalar field theory with mass m.
3.0.3 Proposition. The associative algebra Am coming from the free scalar field
theory with mass m is the Weyl algebra, generated by p, q, ~ with the relation
[p, q] = ~ and all other commutators being zero.
The Hamiltonian H is the derivation
H(a) = 1
2~ [p
2
− m2 q2 , a].
Note that it is an inner, or principal, derivation.
Remark: It might seem a priori that different action functionals, which encode
different theories, should have different algebras of observables. In this case, we
see that all the theories lead to the same associative algebra. The Hamiltonian is
what distinguishes these algebras of observables; that is, different actions differ by
“how to relate the same observable applied at different moments in time.”
There is an important feature of the Weyl algebra that illuminates the role of
the Hamiltonian. The Weyl algebra is rigid: the Hochschild cohomology of the
Weyl algebra A(n) for a symplectic vector space of dimension 2n vanishes, ex-
cept in degree 0, where the cohomology is one-dimensional. Hence, in particular,
HH 2 (A(n) ) = 0, so there are no nontrivial deformations of the Weyl algebra. In
consequence, any action functional whose underlying free theory is the free scalar
field (when the coupling constants are formal parameters) will have the same alge-
bra of observables, as an associative algebra. Since HH 1 (A(n) ) = 0, we know that
every derivation is inner and hence is represented by some element of A(n) . Thus,
the derivation arising from translation must have a Hamiltonian operator. ^
Proof. We will start by writing down elements of Obsq corresponding to the
position and momentum observables. Recall that
Obsq ((a, b)) = Sym Cc∞ ((a, b))−1 ⊕ Cc∞ ((a, b))0 [~]
φ p (x) = 2m
1
emx − e−mx .
Thus, φq , φ p are both in the null space of the operator −∂2x + m2 , with the properties
that ∂ x φ p = φq and that φq is symmetric under x 7→ −x, whereas φ p is antisymmet-
ric.
3. QUANTUM MECHANICS AND THE WEYL ALGEBRA 89
of the fact that H ∗ Obsq is locally constant. (The other commutators among [P0 ]
and [Q0 ] vanish automatically.)
We also need to calculate the Hamiltonian. We see that
∂
− Qt = − ft0 = Pt = dtd f (x − t) = dtd Qt ,
∂x
∂
− Pt = ft00 .
∂x
Hence, the Hamiltonian acting on [P0 ] gives the t-derivative of [Pt ] at t = 0 and
similarly for Q0 . In other words, the parameter t encodes translation of the real
line, and the infinitesimal action of translation acts correctly on the observables.
In cohomology, the image of −∂2x + m2 is zero, we see that
d
dt [Pt ] = 0 + m2 [ f (x − t)] = m2 [Qt ],
as we expect from usual classical mechanics. In particular, when m = 0, [Pt ] is
independent of t: in other words, momentum is conserved.
Thus, the Hamiltonian H satisfies
H([P0 ]) = m2 [Q0 ]
H([Q0 ]) = [P0 ].
If we assume that the commutation relation [[P0 ], [Q0 ]] = ~ holds, then
h i
H(a) = 2~
1
[P0 ]2 − m2 [Q0 ]2 , a ,
as desired.
90 4. FREE FIELD THEORIES
Thus, to complete the proof, it only remains to verify the Heisenberg commu-
tation relation.
If a(t) is a function t, let ȧ(t) denote the t-derivative. It follows from the equa-
tions above for the t-derivatives of [Pt ] and [Qt ] that for any function a(t) satisfying
ä(t) = m2 a(t), the observable
a(t)Pt − ȧ(t)Qt
is independent of t at the level of cohomology.
et is the linear observable of cohomological degree −1 given
More precisely, if Q
by the function ft in Cc ((t − 21 , t + 12 )[1], we have
∞
∂
(a(t)Pt − ȧ(t)Qt ) = −d(a(t)Q et )
∂t
where d is the differential on observables. Explicitly, we compute
∂ ∂ ∂ ∂ ∂
(a(t)Pt − ȧ(t)Qt ) = −ȧ(t) ft (x) − a(t) ft (x) − ä(t) ft (x) − ȧ(t) ft (x)
∂t ∂x ∂t ∂x ∂t
∂2
= a(t) 2 ft (x) − ä(t) ft (x)
∂x
∂2
= a(t) 2 ft (x) − m2 a(t) ft (x)
∂x
= −d(a(t)Q et ).
We will define modified observables Pt , Qt that are independent of t at the
cohomological level. We let
Pt = φq (t)Pt − φ0q (t)Qt
Qt = φ p (t)Pt − φ0p (t)Qt .
Since φ p and φq are in the null space of the operator −∂2x + m2 , the observables
Pt , Qt are independent of t at the level of cohomology. In the case m = 0, then
φq (t) = 1 so that Pt = Pt . The statement that Pt is independent of t corresponds,
in this case, to conservation of momentum.
In general, P0 = P0 and Q0 = Q0 . We also have
∂
Pt = −d(φq (t)Q
et ),
∂t
and similarly for Qt .
It follows that if we define a linear degree −1 observable h s,t by
Z t
h s,t = φq (u)Q
eu (x)du,
u=s
then
dh s,t = P s − Pt .
Note that if |t| > 1, the observables Pt and Qt have disjoint support from the
observables P0 and Q0 . Thus, we can use the prefactorization structure map
Obsq ((− 21 , 12 )) ⊗ Obsq ((t − 21 , t + 12 )) → Obsq (R)
3. QUANTUM MECHANICS AND THE WEYL ALGEBRA 91
Therefore Z
[Q0 Pt ] − [Q0 P−t ] = ~ f (x)h−t,t (x).
It remains to compute the integral. This integral can be rewritten as
Z t Z ∞
f (x) f (x − u)φq (u)du.
u=−t x=−∞
Note that the answer is automatically independent of t for t sufficiently large, be-
cause f (x) is supported near the origin so that f (x) f (x − u) = 0 for u sufficiently
large. Thus, we can sent t → ∞.
Since f is also symmetric under x → −x, we can replace f (x − u) by f (u − x).
We can perform the u-integral by changing coordinates u → u − x, leaving the
integrand as f (x) f (u)φq (u + x). Note that
φq (u + x) = 12 em(x+u) + e−m(x+u) .
Now, by assumption on f ,
Z Z
f (x)e dx =
mx
f (x)e−mx dx = 1.
92 4. FREE FIELD THEORIES
N×R
R
(−1, 1)
It follows that Z ∞
φq (u + x) f (u)du = φq (x).
u=−∞
We can then perform the remaining x integral φq (x) f (x)dx, which gives 1, as
R
desired.
Thus, we have proven
[[Q0 ], [P0 ]] = ~,
as desired.
to be the colimit of the finite tensor products of the Fi under the inclusion maps
coming from the unit 1 ∈ Fi (U) for any open subset. Equivalently, we can view
this tensor product as being associated to the Heisenberg central extension of the
complex (†) above.
Because the complex (†) is a dense subspace of the complex whose Heisenberg
extension defines π∗ Obsq , we see that there is a map of prefactorization algebras
with dense image
F → π∗ Obsq .
Passing to cohomology, we have a map
H ∗ F → H ∗ (π∗ Obsq ).
As the prefactorization algebra H ∗ (Fi ) corresponds to the Weyl algebra A √λi , we
see that H ∗ F corresponds to the algebra AN .
= (dg + α0 g) + (d f )g
= (d + α)g,
5. ABELIAN CHERN-SIMONS THEORY 95
In contrast to the free scalar field, where the cohomology of the linear observ-
ables was often infinite-dimensional, the cohomology of the linear observables is
finite-dimensional, for any open U. Thus, the cohomology prefactorization algebra
H ∗ Obscl is relatively easy to understand.
This pleasant situation continues with the quantum observables. Consider the
natural filtration on Obsq (U) by
F k Obsq (U) = Sym≤k Ω∗c (U)[2] [~].
This filtration induces a spectral sequence, and the first page is H ∗ Obscl (U)[~],
which we have already computed. The differential on the first page is zero, but the
differential on the second page arises from the pairing on linear observables (recall
that the pairing provides the Lie bracket of the Heisenberg Lie algebra, and this
bracket provides the differential in the Chevalley-Eilenberg chain complex). Note
that the pairing on linear observables descends to the cohomology of the linear
observables by
h[α], [β]iH ∗ = hα, βi.
Hence, the differential d2 on the second page of the spectral sequence vanishes on
constant and linear terms (i.e., from Sym0 and Sym1 ) and satisfies
d2 ([α][β]) = ~h[α], [β]iH ∗
for a pure quadratic term. The BD algebra axiom (recall equation (2.6.1)) then
determines d2 on terms that are cubic and higher.
96 4. FREE FIELD THEORIES
case, however, the solutions to the equations of motion – namely, de Rham co-
homology classes of Σ – are translation-invariant,Rso the Hamiltonian is trivial, as
follows. Fix a function f ∈ Cc∞ ((− 12 , 12 )) such that R f (x)dx = 1. Hence the 1-form
f (x)dx generates Hc1 (R). Let ft (x) = f (x − t). Then we get linear observables in
C∗b g((t − 12 , t + 12 )) by
At, j = ft (x)dx ⊗ α j ,
Bt,k = ft (x)dx ⊗ βk ,
νt = ft (x)dx ⊗ ν,
µt = ft (x)dx ⊗ µ.
These elements are all cohomologically nontrivial, and their cohomology classes
generate the Weyl algebra. Observe that, for instance,
∂
At, j = ft0 (x)dx ⊗ α j = d ft ⊗ α j ,
∂x
so that the infinitesimal translation is cohomologically trivial. This argument ap-
plies to all the generators, so we see that the derivation on the Weyl algebra induced
∂
from ∂x must be trivial.
98 4. FREE FIELD THEORIES
N Σ × [0, 2)
M
Σ
Hc2 (N 0 ) → Hc2 (M), the kernel Ker j! is a Lagrangian subspace. We also see that
j! : Hc3 (N 0 ) Hc3 (M) and j! : Hc1 (N 0 ) → Hc1 (M) is the zero map.
Thus, the graded subspace Ker j! ⊂ Hc∗ (N 0 ) is Lagrangian.
For the observables, the map j! describes how the linear observables on N 0 be-
have as linear observables on M. The observables in Ker j! thus vanish on M, and
so the ideal they generate in AΣ vanishes on the module H ∗ Obsq (M). We need to
show this ideal is the entire annihilator. Note that the action of AΣ is governed by
the integration pairing in Hc∗ (M), and the extension-by-zero map Hc∗ (N 0 ) → Hc∗ (M)
respects the integration pairing, so there are no nontrivial relations involving non-
linear observables.
5.4. Knots and links. Chern-Simons theory is famous for its relationship with
knot theory, notably the Vassiliev invariants. Here we will explain a classic exam-
ple: how the Gauss linking number appears in Abelian Chern-Simons theory. For
simplicity, we work in R3 .
Let K : S 1 ,→ R3 and K 0 : S 1 ,→ R3 be two disjoint knots. The linking
number `(K, K 0 ) of K and K 0 is the degree of the Gauss map G K,K 0 : S 1 × S 1 → S 2
that sends a pair of points (x, y) ∈ K × K 0 to the unit vector (x − y)/|x − y|. One
way to compute it – indeed, the way Gauss first introduced it – is to pull back the
standard volume form on S 2 and integrate over K × K 0 . Another way is to pick an
oriented disc whose boundary is K 0 and to count, with signs, the number of times
K intersects D (we may wiggle D gently to make it transverse to K).
To relate this notion to Chern-Simons theory, we need to find an observable
associated to a knot. The central object of Chern-Simons theory is a 1-form A,
which we view as describing a connection d + A for the trivial line bundle on R3 .
Thus, the natural observable for the knot K : S 1 → R3 is
Z
OK (A) = A,
K
where we have fixed an orientation on S 1. Note that this operator is linear in A.
Remark: The Wilson loop observable WK is exp(OK ), which is the holonomy of
the connection d + A around K. We do not work with it here because we want to
work with the symmetric algebra on linear observables, rather than the completed
symmetric algebra. In the non-Abelian setting, we do work with the completed
symmetric algebra, so the analogous Wilson loop observable lives in our quantum
observables for non-Abelian Chern-Simons theory. ^
There is a minor issue with this observable, however: it is distributional in na-
ture, as its support sits on the knot K. It is easy to find a smooth (or “smeared”) ob-
servable that captures the same information. (We know such a smooth observable
exists abstractly by Lemma 2.1.1, but here we will give an explicit representative.)
The key is again to use the pushforward.
Consider the 3-manifold T = S 1 × R2 . We view the core C = S 1 × (0, 0) ⊂ T
as the “distinguished” knot in T . Any embedding κ : T ,→ R3 induces a pushfor-
ward of observables so that we can understand the observables supported in tubular
neighborhood κ(T ) of the knot κ(C) by understanding the observables on T .
6. ANOTHER TAKE ON QUANTIZING CLASSICAL OBSERVABLES 101
Heisenberg dg Lie algebra. Our goal in this section is to provide another useful
description.
The prefactorization algebra of quantum observables Obsq , viewed as a graded
prefactorization algebra with no differential, coincides with Obscl [~]. In particular,
the structure maps are the same. The only difference between Obsq and Obscl [~] is
in the differential.
In deformation quantization, we deform, however, the product of observables.
We saw in Section 3 that taking cohomology, the structure maps for the quantum
observables do change and correspond to the Weyl algebra. The change in differen-
tial modifies the structure maps at the level of cohomology. However, it is possible
to approach quantization differently and deform the structure maps at the cochain
level, instead of the differential.
Hence, in this section we will give an alternative, but equivalent, description of
Obsq . We will construct an isomorphism of precosheaves Obsq Obscl [~] that is
compatible with differentials. This isomorphism is not compatible, however, with
the factorization product. Thus, this isomorphism induces a deformed factorization
product on Obscl [~] corresponding to the factorization product on Obsq .
In other words, instead of viewing Obsq as being obtained from Obscl by keep-
ing the factorization product fixed but deforming the differential, we will show that
it can be obtained from Obscl by keeping the differential fixed but deforming the
product.
One advantage of this alternative description is that it is easier to construct
correlation functions and vacua in this language. We will now develop these ideas
in the case of the free scalar field.
6.1. Green’s functions. The isomorphism we will construct between the cochain
complexes of quantum and classical observables relies on a Green’s function for
the Laplacian.
6.1.1 Definition. A Green’s function is a distribution G on M× M, preserved under
reflection across the diagonal (i.e., symmetric) and satisfying
(4 ⊗ 1)G = δ4 ,
where δ4 is the δ-distribution on the diagonal M ,→ M × M.
A Green’s function for the Laplacian with mass satisfies the equation
(4 ⊗ 1)G + m2G = δ4 .
∂ 2
(The convention is that 4 has positive eigenvalues, so that on Rn , 4 = −
P
∂xi .)
6.2. The isomorphism of graded vector spaces. Let us now turn to the con-
struction of the isomorphism of graded vector spaces between Obscl (U) and Obsq (U)
in the presence of a Green’s function.
The underlying graded vector space of Obsq (U) is
Sym(Cc∞ (U)[1] ⊕ Cc∞ (U))[~].
In general, for any vector space V, each element P ∈ (V ∨ )⊗2 defines a differential
operator ∂P of order two on Sym V, viewed as a commutative algebra, where ∂P
is uniquely characterized by the conditions that it is zero on Sym≤1 V and that on
Sym2 V it is given by contraction with P. The same holds when we define the
symmetric algebra using the completed tensor product.
In the same way, for every distribution P on U ×U, we can define a continuous,
second-order differential operator on Sym(Cc∞ (U)[1] ⊕ Cc∞ (U)) that is uniquely
characterized by the properties that it vanishes on Sym≤1 , that it vanishes elements
of negative cohomological degree in Sym2 (Cc∞ (U)[1] ⊕ Cc∞ (U)), and that
Z
∂P (φψ) = P(x, y)φ(x)ψ(y)
U×U
for any φ, ψ ∈ Cc∞ (U)
of cohomological degree zero. (On the left hand side, we
view φψ as an element of Sym2 .)
Choose a Green’s function G for the Laplacian on M. Thus, G restricts to a
Green’s function for the Laplacian on any open subset U of M.
Therefore, we can define a second-order differential operator ∂G on the algebra
Sym(Cc∞ (U)[1] ⊕ Cc∞ (U)). We extend this operator by R[~]-linearity to an operator
on the graded vector space Obsq (U).
Now, the differential on Obsq (U) can be written as d = d1 + d2 , where d1 is
a first-order differential operator and d2 is a second-order operator. The operator
4
d1 is the derivation arising from the differential on the complex Cc∞ (U) −
→ Cc∞ (U).
The operator d2 arises from the Lie bracket on the Heisenberg dg Lie algebra; it
is a continuous, ~-linear, second-order differential operator uniquely characterized
by the property that
Z
d2 (φ φ ) = ~
−1 0
φ−1 (x)φ0 (x)
M
for φk in the copy of Cc∞ (U) in degree k, sitting inside of Obsq (U).
104 4. FREE FIELD THEORIES
((4 + m2 ) ⊗ 1)G = δ4
[~∂G , d1 ] = d2 .
On the first line, the element φ0 φ−1 lives in Sym2 , as does its image under d1 .
On the fourth line, we have used the fact that 4y + m2 applied to G(x, y) is the
delta-distribution on the diagonal.
It is also immediate that ∂G commutes with d2 . Thus, if we define W(α) =
e~∂G (α), we have
(d1 + d2 )W(α) = W(d1 α).
In other words, there is an R[~]-linear cochain isomorphism
where the domain is the complex Obscl (U)[~] with differential d1 and the codomain
is the same graded vector space with differential d1 + d2 . This isomorphism holds
for every open U.
Note that W is not a map of prefactorization algebras. Thus, W induces a fac-
torization product (i.e., structure maps) on classical observables that quantizes the
original factorization product. Let us denote this quantum product by ?~ , whereas
the original product on classical observables will be denoted by ·. If U, V are dis-
joint open subsets of M with α ∈ Obscl (U), β ∈ Obscl (V), we have
α ?~ β = e−~∂G e~∂G α · e~∂G β .
This product is an analog of the Moyal formula for the product on the Weyl algebra.
6. ANOTHER TAKE ON QUANTIZING CLASSICAL OBSERVABLES 105
6.3. Beyond the free scalar field. We have described this construction for the
case of a free scalar field theory. This construction can be readily generalized to
the case of an arbitrary free theory. Suppose we have such a theory on a manifold
M, with space of fields E (M) and differential d. Instead of a Green’s function, we
require a symmetric and continuous linear operator G : (E (M)[1])⊗2 → R such
that
Gd(e1 ⊗ e2 ) = he1 , e2 i
where h−, −i is the pairing on E (M), which is part of the data of a free field theory.
(When M is compact and H ∗ (E (M)) = 0, the propagator of the theory satisfies
this property. For M = Rn , we can generally construct such a G from the Green’s
function for the Laplacian.) In this context, the operator e~∂G is, in the terminology
of Costello (2011b), the renormalization group flow operator from scale zero to
scale ∞.
This construction often makes it easier to analyze the structure of the quantum
observables. We will use it, for instance, to understand the correlation functions of
some chiral conformal field theories in the next chapter.
Example: We can also use it to provide another proof of Lemma 5.4.1, which as-
serts that Abelian Chern-Simons knows about linking number. Recall that for two
disjoint knots K and K 0 , we introduced smoothed versions, µκ and µκ0 , of the cur-
rents with support along the knots.
It is easy to check that, up to a constant, the 2-form
3
1 X
ω= (−1)i (xi − yi ) d(x1 − y1 ) ∧ d(x[
i − yi ) ∧ d(x3 − y3 )
|x − y|3 i=1
7. Correlation functions
Suppose we have a free field theory on a compact manifold M, with the prop-
erty that H ∗ (E (M)) = 0. As an example, consider the massive scalar field theory
on M where
4+m2
E = C∞
M −−−−→ C M .
∞
Our conventions are such that the eigenvalues of 4 are non-negative. Adding a
non-zero mass term m2 gives an operator with no zero eigenvalues, so that this
complex has no cohomology, by Hodge theory.
In this situation, the dg Lie algebra Eb(M) we constructed above has coho-
mology spanned by the central element ~ in degree 1. It follows that there is an
isomorphism of R[~]-modules
H ∗ (Obsq (M)) R[~].
Let us normalize this isomorphism by asking that the element 1 in
Sym0 (Eb[1]) = R ⊂ C∗ (Eb[1])
is sent to 1 ∈ R[~].
7.0.1 Definition. In this situation, we define the correlation functions of the free
theory as follows. For a finite collection U1 , . . . , Un ⊂ M of disjoint opens and a
closed element Oi ∈ Obsq (Ui ) for each open, we define
hO1 · · · On i = [O1 · · · On ] ∈ H ∗ (Obsq (M)) = R[~].
Here O1 · · · On ∈ Obsq (M) denotes the image under the structure map
O
Obsq (Ui ) → Obsq (M)
i
In other words, our normalization gives a value in R[~] for each cohomology
class, and we use this to read off the “expected value” of the product observable.
The map W constructed in the previous section allows us to calculate correla-
tion functions. Since we have a non-zero mass term and M is compact, there is a
unique Green’s function for the operator 4 + m2 .
7.0.2 Lemma (Wick’s lemma). Let U1 , . . . , Un be pairwise disjoint open sets. Let
αi ∈ Obscl (Ui ) = Sym(Cc∞ (Ui )[1] ⊕ Cc∞ (Ui ))
be classical observables, and let
W(αi ) = e~∂G αi ∈ Obsq (Ui )
be the corresponding quantum observables under the isomorphism W of cochain
complexes. Then
hW(α1 ) · · · W(αn )i = W −1 (W(α1 ) · · · W(αn )) (0) ∈ R[~].
7. CORRELATION FUNCTIONS 107
On the right hand side, W(α1 )·W(α2 ) indicates the product in the algebra Sym(C ∞ (M)[1]⊕
C ∞ (M))[~]. The symbol (0) indicates evaluating a function on C ∞ (M)[1]⊕C ∞ (M)[~]
at zero, i.e., taking its component in Sym0 .
Proof. The map W is an isomorphism of cochain complexes between Obsq (U)
and Obscl (U)[~] for every open subset U of M. As above, let ?~ denote the fac-
torization product on Obscl [~] corresponding, under the isomorphism W, to the
factorization product on Obsq . That is, we apply W to each input, multiply in
Obsq , and then apply W −1 to the output: explicitly, we have
α1 ?~ · · · ?~ αn = e−~∂G (e~∂G α1 ) · · · (e~∂G αn ) .
This statement does not resemble the usual statement in a physics text, so we
unpack it in an example to clarify the relationship with other versions of Wick’s
lemma.
As an example, let us suppose that we have two linear observables α1 , α2 of
cohomological degree 0, defined on open sets U, V. Thus, α1 ∈ Cc∞ (U) and α2 ∈
Cc∞ (V). As these are linear observables, the second-order operator ∂G does nothing,
so W(αk ) = αk . But the product is quadratic so we have
W −1 (α1 α2 ) = α1 α2 − ~∂G (α1 α2 ).
As Z
∂G (α1 α2 ) = G(x, y)α1 (x)α2 (y),
M×M
it follows that
Z
hW(α1 )W(α2 )i = −~ G(x, y)α1 (x)α2 (y),
M×M
which is what a physicist would write down as the expected value of two linear
observables.
Remark: We have set things up so that we are computing the functional integral
against the measure eS /~ dµ ,where S is the action functional and dµ is the (non-
existent) “Lebesgue measure” on the space of fields. Physicists often use the mea-
sures e−S /~ or eiS /~ . By changing ~ (e.g., ~ 7→ −~), one can move between these
different conventions. ^.
108 4. FREE FIELD THEORIES
τx
F (V) / F (τ x V)
commutes. (Here the vertical arrows are the structure maps of the prefactorization
algebra.)
Example: Let us consider, again, the prefactorization algebra of the free scalar field
on Rn . Observables on U are Chevalley-Eilenberg chains of the Heisenberg algebra
Ebc (U).
Let X denote a Killing vector field on Rn (i.e., X is an infinitesimal isometry).
For example, we could take X to be a translation vector field ∂x∂ j . Then the Heisen-
berg dg Lie algebra has a derivation where φk 7→ Xφk , for k = 0, 1, and the central
element ~ is annihilated. (Recall that φk is notation for an element of the copy of
Cc∞ (U) situated in degree k.) Because X is a Killing vector field on Rn , it commutes
with the differential on the Heisenberg algebra and satisfies the equality
Z Z
(Xφ )φ + φ0 (Xφ1 ) = 0.
0 1
disjoint and contained in V. It parametrizes the way we can move the open sets
without causing overlaps. Let us assume that Disj(U1 , . . . , Uk | V) has non-empty
interior, which happens when the closure of the Ui are disjoint and contained in V.
Let F be any discretely translation-invariant prefactorization algebra. For each
(x1 , . . . , xk ) ∈ Disj(U1 , . . . , Uk | V), we have a multilinear map obtained as a com-
position
m x1 ,...,xk : F (U1 ) × · · · × F (Uk ) → F (τ x1 U1 ) × · · · × F (τ xk Uk ) → F (V),
where the second map arises from the inclusion τ x1 U1 t · · · t τ xk Uk ,→ V.
8.1.3 Definition. A discretely translation-invariant prefactorization algebra F is
smoothly translation-invariant if the following conditions hold.
((i)) The map m x1 ,...,xk above depends smoothly on (x1 , . . . , xk ) ∈ Disj(U1 , . . . , Uk |
V).
((ii)) The prefactorization algebra F is equipped with an action of the Abelian
Lie algebra Rn of translations. If v ∈ Rn , we will denote the corre-
sponding action maps by
d
: F (U) → F (U).
dv
We view this Lie algebra action as an infinitesimal version of the global
translation invariance.
((iii)) The infinitesimal action is compatible with the global translation in-
variance in the following sense. For v ∈ Rn , let vi ∈ (Rn )k denote the
vector (0, . . . , v, . . . , 0), with v placed in the i-slot and 0 in the other
k − 1 slots. If αi ∈ F (Ui ), then we require that
!
d d
m x ,...,x (α1 , . . . , αk ) = m x1 ,...,xk α1 , . . . , αi , . . . , αk .
dvi 1 k dv
When we refer to a translation-invariant prefactorization algebra without fur-
ther qualification, we will always mean a smoothly translation-invariant prefactor-
ization algebra.
Example: We have already seen that the prefactorization algebra of the free scalar
field theory on Rn is discretely translation invariant and is also equipped with an
action of the Abelian Lie algebra Rn by derivations. It is easy to verify that this
prefactorization algebra is smoothly translation-invariant. ^
Example: This example is a special case of example 7.3.3.
Let F be the cohomology of the prefactorization algebra of observables of the
free scalar field theory on R with mass m. We have seen in Section 3 that the
corresponding associative algebra A is the Weyl algebra, generated by p, q, ~ with
commutation relation [p, q] = ~.
It is straightforward to verify that F is a locally-constant, smoothly translation
invariant prefactorization algebra on R, valued in vector spaces. As the action of R
on A is smooth, it differentiates to an infinitesimal action of the Lie algebra R on A
8. TRANSLATION-INVARIANT PREFACTORIZATION ALGEBRAS 111
∂
by derivations. The basis element ∂x of R becomes a derivation H of A, called the
Hamiltonian. The Hamiltonian hereis given by H(a) = 2~1
[p2 − m2 q2 , a]. ^
8.2. An operadic reformulation. Next, we will explain how to think of the
structure of a translation-invariant prefactorization algebra on Rn in more operadic
terms. This description has a lot in common with the En algebras familiar from
topology.
Let r1 , . . . , rk , s ∈ R>0 . Let
Discsn (r1 , . . . , rk | s) ⊂ (Rn )k
be the (possibly empty) open subset consisting of k-tuples x1 , . . . , xk ∈ Rn with the
property that the closures of the balls Bri (xi ) are all disjoint and contained in Bs (0).
Here, Br (x) denotes the open ball of radius r around x.
8.2.1 Definition. Let Discsn be the R>0 -colored operad in the category of smooth
manifolds whose space of k-ary morphisms Discsn (r1 , . . . , rk | s) between ri , s ∈
R>0 is described above.
The first property is that, for each configuration p ∈ Discsn (r1 , . . . , rk | s),
the map m[p] is a multilinear map of cohomological degree 0, compatible with
differentials.
Second, let N be a manifold and let fi : N → Frdi i be smooth maps into the
space Frdi i of elements of degree di . The smoothness properties of the map m[p]
mean that the map
N × Discsn (r1 , . . . , rk | s) → Fs
(x, p) 7 → m[p]( f1 (x), . . . , fk (x))
is smooth. Thus, we can work with smooth families of multiplications.
Next, note that a permutation σ ∈ S k gives an isomorphism
σ : Discsn (r1 , . . . , rk | s) → Discsn (rσ(1) , . . . , rσ(k) | s).
We require that for each p ∈ Discsn (r1 , . . . , rk | s) and collection of elements
αi ∈ Fri ,
m[σ(p)](ασ(1) , . . . , ασ(k) ) = m[p](α1 , . . . , αk ).
Finally, we require that the maps m[p] are compatible with composition, in
the following sense. For any configurations p ∈ Discsn (r1 , . . . , rk | ti ) and q ∈
Discsn (t1 , . . . , tl | s) and for collections αi ∈ Fri and β j ∈ Ft j , we require that
m[q](β1 , . . . , βi−1 , m[p](α1 , . . . , αk ), βi+1 , . . . , βl )
= m[q ◦i p](β1 , . . . , βi−1 , α1 , . . . , αk , βi+1 , . . . , βl ).
In addition, the action of Rn on each F r is compatible with these multiplication
maps, in the manner described above.
8.3. A co-operadic reformulation. Let us give one more equivalent way of
rewriting these axioms, which will be useful when we discuss the holomorphic
context. These alternative axioms will say that the spaces C ∞ (Discsn (r1 , . . . , rk |
s)) form an R>0 -colored co-operad when we use the appropriate completed tensor
product. Since we know how to tensor a differentiable vector space with the space
of smooth functions on a manifold, it makes sense to talk about an algebra over
this colored co-operad in the category of differentiable cochain complexes.
We can rephrase the smoothness axiom for the product map
m[p] : Fr1 ⊗ · · · ⊗ Frk → F s ,
as p varies in Discsn (r1 , . . . , rk | s), as follows. For any differentiable vector space
V and smooth manifold M, we use the notation V⊗C ∞ (M) interchangeably with the
notation C ∞ (M, V); both indicate the differentiable vector space of smooth maps
M → V. The smoothness axiom states that the map above extends to a smooth
map of differentiable spaces
µ(r1 , . . . , rk | s) : Fr1 × · · · × Frk → F s ⊗ C ∞ (Discsn (r1 , . . . , rk | s)).
In general, if V1 , . . . , Vk , W are differentiable vector spaces and if X is a smooth
manifold, let
C ∞ (X, HomDVS (V1 , . . . , Vk | W))
denote the space of smooth multilinear maps V1 × · · · × Vk → C ∞ (X, W).
8. TRANSLATION-INVARIANT PREFACTORIZATION ALGEBRAS 113
Recall that “smooth” means the following. The ring R itself is a differen-
tiable vector space: for instance, the real numbers R as a topological vector space
corresponds to the sheaf C ∞ on the site of smooth manifolds. As F (Rn ) is a dif-
ferentiable vector space, it is also a sheaf on the site of smooth manifolds, and
H ∗ (F (Rn )) denotes the cohomology sheaf on the site of smooth manifolds. Unrav-
eling the structures in the definition above, we find that for each smooth manifold
M and each element α ∈ H ∗ (C ∞ (M, F (Rn )), the element hαi is a smooth R-valued
function on M.
9. STATES AND VACUA FOR TRANSLATION INVARIANT THEORIES 115
Example: This extended example examines vacua of the free scalar field with mass
m.
We have seen that the choice of a Green’s function G for the operator 4 + m2
leads to an isomorphism of cochain complexes Obscl (U)[~] Obsq (U) for every
open subsets U ⊂ Rn . This produces an isomorphism of prefactorization algebras
if we endow Obscl [~] with a deformed factorization product ?~ , defined using the
Green’s function.
If m > 0, there is a unique Green’s function G of the form G(x, y) = f (x − y),
where f is a distribution on Rn that is smooth away from the origin and tends
to zero exponentially fast at infinity. For example, if n = 1, the function f is
f (x) = 2m
1 −m|x|
e .
If m = 0, we already know
4π1d/2 Γ(d/2 − 1) |x − y|2−n if n , 2
G(x, y) =
− 1 log |x − y|
2π if n = 2
is the canonically-defined Green’s function.
It is standard to interpret the Green’s function as the two-point correlation func-
tion hδ x δ0 i. Hence, we see that this two-point function exhibits the necessary be-
havior: for m > 0, it decays exponentially fast, and it decays slowly in the massless
case. Strictly speaking, this observable is not in our prefactorization algebra, as
it is distributional, but it does exist at the cohomological level. Alternatively, we
could work with smeared versions.
Because Obscl (Rn ) is, as a cochain complex, the symmetric algebra on the
4+m2
complex Cc∞ (Rn )[1] −−−−→ Cc∞ (Rn ), there is a map from Obscl (Rn ) to R that is
116 4. FREE FIELD THEORIES
the identity on Sym0 and sends Sym>0 to 0. (In other words, we have a natural
augmentation map.) This map extends to an R[~]-linear cochain map
h−i : Obscl (Rn )[~] → R[~].
Clearly, this map is translation invariant and smooth. Thus, because we have a
cochain isomorphism between Obscl (Rn )[~] and Obsq (Rn ), we have produced a
translation-invariant state.
9.0.3 Lemma. For m > 0, this state is a massive vacuum. For m = 0, this state is a
vacuum if n > 2; for n ≤ 2, it does not satisfy the cluster decomposition principle.
Proof. Let F1 ∈ Cc∞ (D(0, r1 ))⊗k1 and F2 ∈ Cc∞ (D(0, r2 ))⊗k2 . We will view
F1 , F2 as observables in Obscl (Bri (0)) by using the natural map from Cc∞ (U)⊗k to
the coinvariants Symk Cc∞ (U). Let τc F2 be the translation of F2 by c, where c is
sufficiently large so that the discs D(0, r1 ) and D(c, r2 ) are disjoint. Explicitly, τc F2
is represented by the function F2 (y1 −c, . . . , yk2 −c). We are interested in computing
the expected value hF1 (τc F2 )i.
We gave above an explicit formula for the quantum factorization product ?~
on Obscl [~]. In this case, it states that F1 ?~ τc F2 is given by
min(k
X 1 ,k2 ) X Z Z r
Y
~ r
F1 (x)F2 (τc y) G(xik , y jk ),
r=0 1≤i1 <···<ir ≤k1 x∈Rnk1 y∈Rnk2 k=1
1≤ j1 <···< jr ≤k2
where x = (x1 , . . . , xk1 ) and τc y = (y1 − c, . . . , yk2 − c). Note that after performing
the integral on the right hand side, we are left with a function of the k1 + k2 − 2r
copies of Rn that we have not integrated over. This function is then viewed as an
observable in Symk1 +k2 −2r Cc∞ (Rn ).
If k1 , k2 > 0, then hF1 i = 0 and hF2 i = 0. Further, hF1 (τc F2 )i selects the
constant term in the expression for F1 ?~ τc F2 . There is only a non-zero constant
term if k1 = k2 = k. In that case, the constant term hF1 , τc F2 i is
Z
~k F1 (x1 , . . . , xk )F2 (y1 − c, . . . , yk − c)G(x1 , y1 ) . . . G(xk , yk ).
xi ,yi ∈Rn
This lemma simply reduces the full condition on a vacuum down to the Green’s
function, and hence to the usual perspective. ^
9. STATES AND VACUA FOR TRANSLATION INVARIANT THEORIES 117
Example: We can give a more abstract construction for the vacuum of associated
to a massive scalar field theory on Rn . It illuminates how boundary or asymptotic
conditions on fields relate to observables.
4+m2
Let E = C ∞ −−−−→ C ∞ be, as before, the complex of fields. Let Ebc be the
Heisenberg central extension Ebc = Ec ⊕ R · ~ where ~ has degree 1. We defined the
complex of observables on U as the Chevalley-Eilenberg chain complex of Ebc (U).
4+m2
Let ES (Rn ) be the complex S(Rn ) −−−−→ S(Rn ), where S(Rn ) is the space of
Schwartz functions on Rn . (Recall that a function is Schwartz if it tends to zero at
∞ faster than the reciprocal of any polynomial.)
A Heisenberg dg Lie algebra can be defined using ES (Rn ) instead R of Ec (R ).
n
This chapter serves two purposes. On the one hand, we develop several exam-
ples that exhibit how to understand the observables of a two-dimensional theory
from the point of view of prefactorization algebras and how this approach recovers
standard examples of vertex algebras. On the other hand, we provide a precise def-
inition of a prefactorization algebra on Cn whose structure maps vary holomorphi-
cally, much as we defined translation-invariant prefactorization algebras in Section
8. We then give a proof that when n = 1 and the prefactorization algebra possesses
a U(1)-action, we can extract a vertex algebra. For n > 1, the structure we find is
a higher-dimensional analog of a vertex algebra. (We leave a detailed analysis of
the structure present in the higher-dimensional case for future work). Such higher-
dimensional vertex algebras appear, for example, as the prefactorization algebra of
observables of partial twists of supersymmetric gauge theories.
Remark: The usage of the term “field” in the theory of vertex operators often pro-
vokes confusion. In this book, the term field is used to refer to a configuration in a
classical field theory: for example, in a scalar field theory on a manifold M, a field
is an element of C ∞ (M). The term “field” as used in the theory of vertex algebras
is not an example of this usage. Instead, as we will see below, a “field” in a vertex
algebra corresponds to an observable with support at a point of the surface; it can
be understand as a special kind of operator. ^
0.1.2 Definition. [Definition 1.3.1, Frenkel and Ben-Zvi (2004)] A vertex algebra
is the following data:
• a vector space V over C (the state space);
• a nonzero vector |0i ∈ V (the vacuum vector);
• a linear map T : V → V (the shift operator);
119
120 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
The vertex operation is best understood in terms of the following intuition. The
vector space V represents the set of pointwise measurements one can make of the
fields, and one should imagine labeling each point z ∈ C by a copy of V, which
we’ll denote Vz . Moreover, in a disc D containing the point z, the measurements
at z are a dense subspace of the measurements one can make in D (this is a feature
of chiral theories). We’ll denote the observables on D by VD . The vertex operation
is a way of combining pointwise measurements. Let D be a disc centered on the
origin. For z , 0, we can multiply observables to get a map
z
D Yz : V0 ⊗ Vz → VD .
0
algebra. Let Fr denote the cochain complex F (D(0, r)) that F assigns to a disc
of radius r. Recall, as we explained in Section 8 of Chapter 4, if F is smoothly
translation invariant, then we have an operator product map
µ : Fr1 ⊗ · · · ⊗ Frn → C ∞ (Discs(r1 , . . . , rk | s), F s ),
where Discs(r1 , . . . , rk | s) refers to the open subset of Ck consisting of points
z1 , . . . , zk such that the discs of radius ri around zi are all disjoint and contained in
the disc of radius s around the origin. If F is holomorphically translation invariant,
then µ lifts to a cochain map
µ : Fr1 ⊗ · · · ⊗ Frn → Ω0,∗ (Discs(r1 , . . . , rk | s), F s ),
where on the right hand side we have used the Dolbeault complex of the complex
manifold Discs(r1 , . . . , rk | s). We also require some compatibility of these lifts
with compositions, which we will detail later.
In other words, being holomorphically translation invariant means that the op-
erator product map is holomorphic (up to homotopy) in the location of the discs.
0.1.3 Theorem. Let F be a holomorphically translation invariant prefactorization
algebra on C. Let F be equivariant under the action of S 1 on C by rotation, and
let Frk denote the weight k eigenspace of the S 1 action on the complex Fr . Assume
that for every r < s, the extension map Frk → F sk associated to the inclusion
D(0, r) ⊂ D(0, s) is a quasi-isomorphism. Finally, we need to assume that the S 1
action on each Fr satisfies a certain technical “tameness” condition.
Then the vector space M
VF = H ∗ (Frk )
k∈Z
has the structure of a vertex algebra. The vertex algebra structure map
YF : VF ⊗ VF → VF [[z, z−1 ]]
is the Laurent expansion of operator product map
H ∗ µ : H ∗ (Frk11 ) ⊗ H ∗ (Frk22 ) → Hol(Discs(r1 , r2 | s), H ∗ (F s )).
On the right hand side, Hol denotes the space of holomorphic maps.
In other words, the intuition that the vertex algebra structure map is the opera-
tor product expansion is made precise in our formalism.
This result should be compared with the classic result of Huang (1997), who
relates vertex algebras with chiral conformal field theories at genus 0, in the sense
used by Segal. As we have seen in Section 4, the axioms for prefactorization
algebras are closely related to Segal’s axioms. Our axioms for a holomorphically
translation invariant field theory are similarly related to Segal’s axioms for a two-
dimensional chiral field theory. Although our result is closely related to Huang’s,
it is a little different because of the technical differences between a prefactorization
algebra and a Segal-style chiral conformal field theory.
One nice feature of our definition of holomorphically translation-invariant pref-
actorization algebra is that it makes sense in any complex dimension. The structure
122 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
We should understand this definition as saying that the vector fields ∂z∂ i act
homotopically trivially on F . In physics terminology, the operators ∂x∂ i , where xi
are real coordinates on Cn = R2n , are related to the energy-momentum tensor. We
are asking that the components of the energy momentum tensor in the zi directions
are exact for the differential on observables (in physics, this might be called “BRST
exact”). This is rather similar to a phenomenon that sometimes appears in the study
of topological field theory, in which a topological theory depends on a metric, but
the variation of the metric is exact for the BRST differential. See Witten (1988) for
discussion.
Explicitly, the colored cooperad structure is given as follows. The operad struc-
ture on the complex manifolds PDiscsn (r1 , . . . , rk | ti ) is given by maps
We let
arbitrary smooth maps between differentiable vector spaces. Since the cooperad
structure maps in our colored dg cooperad Ω0,∗ (PDiscsn ) arise from maps of com-
plex manifolds, it makes sense to ask for a coalgebra over this cooperad in the
category of differentiable vector spaces.
1.3.1 Proposition. Let F be a holomorphically translation-invariant
prefactoriza-
tion algebra on Cn . Then F defines an algebra over the cooperad Ω0,∗ (PDiscsn )
n∈N
in differentiable cochain complexes.
126 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
PDt (0)
PDr (0)
µ∂
0 PD s (z) Fr × F s Ω0,∗ (PDiscs(r, s | t), Ft )
that is closed with respect to the differential ∂ + dF , where dF refers to the natu-
ral differential on the differentiable cochain complex of smooth multilinear maps
HomDVS (Fr1 , . . . , Frk | F s ).
We will produce the desired element
Here dzi j li j denotes the operation of wedging with dzi j after applying the operator
li j . Note that these two operators graded-commute.
Next, we need to verify that (∂ + dF )µ∂ = 0, where ∂ + dF is the differential
on graded vector space (†) above that arises by combining the ∂ operator with the
differential arising from the complexes Fri , F s .
We use the following identities:
X X
[dF , dzi j li j ] = dzi j ri j ,
i, j i, j
∂ 0
ri j µ0 = µ ,
∂zi j
dF µ0 = 0.
2. A GENERAL METHOD FOR CONSTRUCTING VERTEX ALGEBRAS 129
The second and third identities are part of the axioms for a smoothly translation-
invariant prefactorization algebra.
These identities allows us to calculate that
(∂ + dF )µ∂ = (∂ + dF ) exp(−
X
dzi j li j )µ0
X X X
= − exp − dzi j li j dzi j ri j µ0 + exp − dzi j li j (∂ + dF )µ0
∂
X X !
= exp − dzi j li j − dzi j + ∂ µ0
∂zi j
= 0.
Thus, µ∂ is closed.
It is straightforward to verify that the elements µ∂ are compatible with compo-
sition and with the symmetric group actions.
words, every differentiable vector space E discussed here will assign a complex
vector space E(M) to a manifold M, the restriction maps are complex-linear, and
so on.
By their very definition, we know how to differentiate “smooth maps to a differ-
entiable vector space.” Thus, if M is a complex manifold and E is a differentiable
cochain complex, we have, for every open U ⊂ M, a cochain map
∂ : C ∞ (U, E) → Ω0,1 (U, E).
These maps form a map of sheaves of cochain complexes on M. Taking the coho-
mology sheaves of E, we obtain a graded differentiable vector space H ∗ (E) and we
have a map of sheaves of graded vector spaces
∂ : C ∞ (U, H ∗ (E)) → Ω0,1 (U, H ∗ (E)).
The kernel of this map defines a sheaf on M, whose sections we denote by Hol(U, H ∗ (E)).
We call these the holomorphic sections of H ∗ (E) on M.
2.1.1. The theorem on vertex algebras will require an extra hypothesis re-
garding the S 1 -action on the prefactorization algebra. Given any compact Lie
group G, we will formulate a concept of tameness for a G-action on a differen-
tiable vector space. We will require that the S 1 -action on the spaces Fr in our
factorization algebra is tame.
Note that for any compact Lie group G, the space D(G) of distributions on G
is an algebra under convolution. Since spaces of distributions are naturally differ-
entiable vector spaces, and the convolution product
∗ : D(G) × D(G) → D(G)
is smooth, D(G) forms an algebra in the category of differentiable vector spaces.
There is a map
δ : G → D(G)
sending an element g to the δ-distribution at g. It is a smooth map and a homomor-
phism of monoids.
2.1.2 Definition. Let E be a differentiable vector space and G be a compact Lie
group. A tame action of G on E is a smooth action of the algebra D(G) on E.
Note that this is, in particular, an action of G on E via the smooth map of groups
G → D(G)× sending g to δg .
If E is a differentiable cochain complex, a tame action is an action of D(G) on
E which commutes with the differential on E.
Let us now specialize to the case when G = S 1 , which is the case relevant
for the theorem on vertex algebras. For each integer k, there is an irreducible
representation ρk of S 1 given by the function ρk : λ 7→ λk . (Here we use λ to
denote a complex number by viewing S 1 as a subset of C. ) We can view ρk as a
map of sheaves on the site of smooth manifolds from the sheaf represented by the
manifold S 1 to the differentiable vector space C ∞ , which assigns complex-valued
ρk of the
smooth functions to each manifold. It lifts naturally to a representation e
2. A GENERAL METHOD FOR CONSTRUCTING VERTEX ALGEBRAS 131
each graded piece Gri F . This condition implies, by a spectral sequence, that they
hold on each F /F i F , allowing us to construct an inverse system of vertex algebras
associated to the prefactorization algebra F /F i F . The inverse limit of this system
of vertex algebras is the vertex algebra associated to F . ^
2.3. Using the theorem. The conditions of the theorem are always satisfied
in practice by prefactorization algebras arising from quantizing a holomorphically
translation invariant classical field theory. We describe in great detail the example
of the free βγ system in Section 3.
There is also the problem of recognizing the vertex algebra produced by such
a prefactorization algebra. Thankfully, there is a useful “reconstruction” theorem
that provides simple criteria to uniquely construct a vertex algebra given “genera-
tors and relations.” We will exploit this theorem to verify that we have recovered
standard vertex algebras in our examples later in this chapter.
2.3.1 Theorem (Reconstruction, Theorem 4.4.1, Frenkel and Ben-Zvi (2004)).
Let V be a complex vector space equipped with a nonzero vector |0i, an endomor-
phism T , a countable ordered set {aα }α∈S of vectors, and fields
n∈Z
such that
((i)) for all α, aα (z)|0i = aα + O(z);
((ii)) T |0i = 0 and [T, aα (z)] = ∂z aα (z) for all α;
((iii)) all fields aα (z) are mutually local;
((iv)) V is spanned by the vectors
where
X X
a(z)+ = an zn and a(z)− = an zn .
n≥0 n<0
2.4. The proof of the main theorem. The strategy of the proof is as follows.
We will analyze the structure on V given to us by the axioms of a translation-
invariant prefactorization algebra. The prefactorization product will become the
operator product expansion or state-field map. The locality axiom of a vertex alge-
bra will follow from the associativity axioms of the prefactorization algebra.
2.4.1. The vector space structure. Let us begin by describing important fea-
tures of V and its natural completion V.
Because our prefactorization algebra F is translation invariant, the cochain
complex F (D(z, r)) associated to a disc of radius r is independent of z. We there-
fore use the notation F (r) for F (D(z, r)) and F (∞) to denote F (C).
Let Fk (r) denote the weight k space of the S 1 action on F . The S 1 action on
each F (r) extends to an action of D(S 1 ). The projection map F (r) → Fk (r) onto
the weight k space induces a map at the level of cohomology that we also denote
πk : H ∗ (F (r)) → H ∗ (Fk (r)) = Vk .
It is a map of differentiable vector spaces, splitting the natural inclusion. Fur-
thermore, the extension map H ∗ (Fk (r)) → H ∗ (Fk (r0 )) associated to the inclusion
D(0, r) ,→ D(0, r0 ) is the identity on V.
By assumption, the differentiable vector space Vk = H ∗ (Fk (r)) is the colimit of
its finite-dimensional subspaces. This means that a section of Vk on the manifold
M is given, locally on M, by a smooth map (in the ordinary sense) to a finite
dimensional subspace of Vk .
By construction there is a canonical map
M
V= Vk → H ∗ (F (r))
k
for every r > 0. We emphasize that V need not be the limit limr→0 H ∗ (F (r)).
Instead, it is an algebraically tractable vector space mapping into that limit (so to
speak, the finite sums of modes).
We introduce now a natural partner to V into which all the H ∗ (F (r)) map. Let
Y
V= Vk ,
k
where the product is taken in the category of differentiable vector spaces. There is
a canonical map
Y
πk : H ∗ (F (r)) → V
k
given by the product of all the projection maps. We emphasize that V need not
be the colimit colimr→∞ H ∗ (F (r)). Instead, it is an algebraically tractable vector
space receiving a map from that colimit. Note that V is a completion of V in the
sense that we now allow arbitrary, infinite sums of the modes.
2.4.2. The operator product. Let us start to analyze the structure on V given
by the prefactorization algebra structure on F . Our approach will be to use instead
the structure as an algebra over the cooperad Ω0,∗ (PDiscs1 ). (Since a 1-dimensional
134 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
polydisc is simply a disc, we will simply write disc from hereon.) We will call the
structure maps for this algebra the operator product. Recall that we use
If α ∈ Ω0,∗ (X, F (∞)) is a cocycle, this map is defined by first extracting the com-
ponent α0 of α in Ω0,0 (X, F (∞)). The fact that α is closed means that α0 is closed
for the differential DF on F (∞) and that ∂α0 is exact. Thus, the cohomology class
[α0 ] is a section of F (∞) on X, and ∂[α0 ] = 0 so that [α0 ] is holomorphic.
Hence, at the level of cohomology, the operator product produces a smooth
multilinear map
mz1 ,...,zk : H ∗ (F (r1 )) × · · · × H ∗ (F (rk )) → Hol(Discs(r1 , . . . , rk | ∞), H ∗ (F (∞))).
Here zi indicate the positions of the centers of the discs in Discs(r1 , . . . , rk | ∞).
Consider what happens if we shrink the discs. If ri0 < ri , then
Discs(r1 , . . . , rk | ∞) ⊂ Discs(r10 , . . . , rk0 | ∞).
We also have the extension map F (r0 ) → F (r), from the inclusion D(0, r0 ) ,→
D(0, r). Note that the following diagram commutes:
H ∗ (F (r10 )) ⊗ · · · ⊗ H ∗ (F (rk0 )) / Hol(Discs(r0 , . . . , r0 | ∞), H ∗ (F (∞)))
1 k
H ∗ (F (r1 )) ⊗ · · · ⊗ H ∗ (F (rk )) / Hol(Discs(r1 , . . . , rk | ∞), H ∗ (F (∞))).
but
lim Hol(Discs(r, . . . , r | ∞), H ∗ (F (∞))) = Hol(Conf k (C), H ∗ (F (∞))),
r→0
where Conf k (C) is the configuration space of k ordered distinct points in C. Note
that if the zi lie in a disc D(0, r), this map provides a map
(r))
mzH∗(F
1 ,...,zk
: V ⊗k → Hol(Conf k (D(0, r)), H ∗ (F (r))).
2. A GENERAL METHOD FOR CONSTRUCTING VERTEX ALGEBRAS 135
∗
(F (∞))
Composing the prefactorization product map mzH1 ,...,z k
with the map
Y Y
πn : H ∗ (F (∞)) → V = Vn
n n
gives a map
Y
mz1 ,...,zk : V ⊗k → Hol(Conf k (C), V) = Hol(Conf k (C), Vn ).
n
This map does not involve the spaces F (r) anymore, only the space V and its
natural completion V. If there is potential confusion, we will refer to this version
of the operator product map by mVz1 ,...,zk instead of just mz1 ,...,zk .
The vertex operator will, of course, be constructed from the maps mz1 ,...,zk .
Consider the map
Y
mz,0 : V ⊗ V → Hol(C× , Vn ),
n
where we have restricted the map mVz,w to the locus where w = 0. Since each space
Vn is a discrete vector space (i.e., a colimit of finite dimensional vector spaces), we
can form the ordinary Laurent expansion of an element in Hol(C× , Vn ) to get a map
Lz mVz,0 : V ⊗ V → V[[z, z−1 ]].
It has the following important property.
2.4.1 Lemma. The image of Lz mVz,0 is in the subspace V((z)).
∂ ∂
becomes a map ∂z : Vk → Vk−1 . Let T : V → V be the map that is ∂z
on Vk .
((iii)) The state-field map Y : V → End(V)[[z, z−1 ]] : We let
Y(v, z)(v0 ) = Lz mz,0 (v, v0 ) ∈ V((z)).
Note that Y(v, z) is a field in the sense used in the axioms of a vertex
algebra, because Y(v, z)(v0 ) has only finitely many negative powers of
z.
It remains to check the axioms of a vertex algebra. We need to verify:
((i)) Vacuum axiom: Y(|0i , z)(v) = v.
((ii)) Translation axiom:
∂
Y(T v1 , z)(v2 ) = Y(v1 , z)(v2 ).
∂z
and T |0i = 0.
((iii)) Locality axiom:
(z1 − z2 )N [Y(v1 , z1 ), Y(v2 , z2 )] = 0
for N sufficiently large.
The vacuum axiom follows from the fact that the unit |0i ∈ F (∅), viewed as an
element of F (D(0, r)), is a unit for the prefactorization product.
The translation axiom follows immediately from the corresponding axiom of
prefactorization algebras, which is built into our definition of holomorphic transla-
tion invariance: namely,
∂
mz,0 (v1 , v2 ) = mz,0 (∂z v1 , v2 ).
∂z
The fact that T |0i = 0 follows from the fact that the derivation ∂z of F gives a
derivation of the commutative algebra F (∅), which must therefore send the unit to
zero.
It remains to prove the locality axiom. Essentially, it follows from the associa-
tivity property of prefactorization algebras.
2.5. Proof of the locality axiom. Let us observe some useful properties of the
operator product maps
mz1 ,...,zk : V ⊗k → Hol(Conf k (C), V).
Firstly, the map mz1 ,...,zk is S 1 -equivariant, where we use the diagonal S 1 action on
V ⊗k on the left hand side and on the right hand side we use the rotation action of S 1
on Conf k (C) coupled to the S 1 action on V coming from F ’s S 1 -equivariance. The
operator product is also S k -equivariant, where S k acts on V ⊗k and on Conf k (C) in
the evident way. It is also invariant under translation, in the sense that for arbitrary
vi ∈ V,
mz1 +λ,...,zk +λ (v1 , . . . , vk ) = τλ mz1 ,...,zk (v1 , . . . , vk ) ∈ V
We write X
Lz mz,0 (v1 , v2 ) = zk mk (v1 , v2 )
k
where mk (v1 , v2 ) ∈ V. As we have seen, mk (v1 , v2 ) is zero for k 0.
The key proposition in proving the locality axiom is the following bit of com-
plex analysis.
2.5.1 Proposition. Let Ui j ⊂ Conf k (C) be the open subset of configurations for
which z j − zi < z j − zl for l , i, j.
Then, for (z1 , . . . , zk ) ∈ Ui j , we have the following identity in Hol(Ui j , V):
X
mz1 ···zk (v1 , . . . , vk ) = z j ···zk (v1 , . . . , vi−1 , m (vi , v j ), . . . , v
(zi − z j )n mz1 ···b n
bj , . . . , vk ).
In particular, the sum on the right hand side converges. (In this expression, the
hats zbj and vbj indicate that we skip these entries.)
When k = 3, this identity allows us to expand mz1 ,z2 ,z3 in two different ways.
We find that
P
(z2 − z3 )k mz1 ,z3 (mk (v1 , v2 ), v3 ) if |z2 − z3 | < |z1 − z3 |
mz1 ,z2 ,z3 (v1 , v2 , v3 ) =
(z − z )k m (v , mk (v , v )) if |z − z | > |z − z | .
P
z2 ,z3
1 3 2 1 3 2 3 1 3
That is, 1 , v2 ) is the projection of mz,0 (v1 , v2 ) onto the weight |v1 + v2 | − k
zk mk (v
eigenspace of V.
We want to show that
X
(‡) (z1 − z2 )n mz2 ,z3 ,...,zk (mn (v1 , v2 ), . . . , vk ) = mz1 ,...,zk (v1 , . . . , vk )
n
where the zi lie in our open U 0 . This identity is equivalent to showing that
X
mz2 ,z3 ,...,zk (πn mz1 ,z2 (v1 , v2 ), . . . , vk ) = mz1 ,...,zk (v1 , . . . , vk ).
n
Indeed, the Laurent expansion of mz1 ,z2 (v1 , v2 ) and the expansion in terms of eigenspaces
of the S 1 action on V differ only by a reordering of the sum.
Fix v1 , . . . , vk ∈ U 0 . Define a map
Φ : D(S 1 ) → Hol(U 0 , V)
α 7→ mz2 ,z3 ,...,zk (α ∗ mz1 ,z2 (v1 , v2 ), v3 , . . . , vk ).
Here α ∗ − refers to the action of D(S 1 ) on H ∗ (F (δ)). Note that Φ is a smooth map.
Let δ1 denote the delta-function on S 1 with support at the identity 1. The
associativity identity (†) of prefactorization algebras implies that
Φ(δ1 ) = mz1 ,...,zk (v1 , . . . , vk ).
The point here is that δ1 ∗ is the identity on H ∗ (F (δ)).
To prove the identity (‡), it now suffices to prove that
X n
Φ λ = Φ(δ1 ).
n∈Z
The spaces D(S 1 ) and Hol(Uiδj , V) both lie in the essential image of the func-
tor from locally-convex topological vector spaces to differentiable vector spaces.
A result of Kriegl and Michor (1997) tells us that smooth linear maps between
topological vector spaces are bounded. Therefore, the map Φ is a bounded linear
map of topological vector spaces.
In Lemma 5.1.1 in Appendix B we show, using results of Kriegl and Michor
(1997), that the space of compactly supported distributions on any manifold has
the bornological property, meaning that a bounded linear map from it to any topo-
logical vector space is the same as a continuous linear map. It follows that Φ is
continuous, thus completing the proof.
As a corollary, we find the following.
2.5.2 Corollary. For v1 , . . . , vk ∈ V, mz1 ,...,zk (v1 , . . . , vk ) has finite order poles on
every diagonal in Conf k (C). That is, for some N sufficiently large, the function
Y
( (zi − z j )N )mz1 ,...,zk (v1 , . . . , vk )
i< j
1 , z2 ]]
Lz2 Lz1 F ∈ C[[z±1 ±1
by first taking the Laurent expansion with respect to z1 , yielding a series in z1 whose
coefficients are holomorphic functions of z2 ∈ C× , and then applying the Laurent
expansion with respect to z2 to each of the coefficient functions of the expansion
with respect to z1 .
Recall that we define
Y(v1 , z)(v2 ) = Lz mz,0 (v1 , v2 ) ∈ V((z)).
We define mk (v1 , v2 ) so that
X
Lz mz,0 (v1 , v2 ) = zk mk (v1 , v2 ).
k
140 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
k
Proposition 2.5.1 tells us that
X
mz1 ,z2 ,0 (v1 , v2 , v3 ) = zn2 mz1 ,0 v1 , mn (v2 , v3 )
n
as long as |z2 | < |z1 |. Thus,
Y(v1 , z1 )Y(v2 , z2 )(v3 ) = Lz1 Lz2 mz1 ,z2 ,0 (v1 , v2 , v3 ).
Similarly,
Y(v2 , z2 )Y(v1 , z1 )(v3 ) = Lz2 Lz1 mz2 ,z2 ,0 (v2 , v1 , v3 )
= Lz2 Lz1 mz1 ,z2 ,0 (v1 , v2 , v3 ).
Therefore,
[Y(v1 , z1 ), Y(v2 , z2 )](v3 ) = Lz1 Lz2 − Lz2 Lz1 mz1 ,z2 ,0 (v1 , v2 , v3 ).
Finally, we know that for N sufficiently large, (z1 − z2 )N z1N z2N mz1 ,z2 ,0 has no poles,
and so extends to a function on C2 . It follows from the fact that partial derivatives
commute that
Lz1 Lz2 − Lz2 Lz1 (z1 − z2 )N z1N z2N mz1 ,z2 ,0 (v1 , v2 , v3 ) = 0.
Since Laurent expansion is a map of C[z1 , z2 ]-modules, and since z1 , z2 act invert-
1 , z2 ]], the result follows.
ibly on C[[z±1 ±1
Remark: In the vertex algebra literature, a heuristic justification of the locality ax-
iom is frequently given by unpacking the consequences of pretending that Y(v1 , z1 )Y(v2 , z2 )(v3 )
and Y(v2 , z2 )Y(v1 , z1 )(v3 ) arise as expansions of a holomorphic function of z1 , z2 ∈
C× in the regions when |z1 | < |z2 | and |z2 | < |z1 |. Our approach makes this idea
rigorous. ^
The differential has a component ∂ arising from the underlying cochain complex
of H and a component arising from the Lie bracket, which we’ll denote ∆. It is the
BV Laplacian for this theory.
Below we will unpack the information Obsq that encodes by examining some
simple open sets and the cohomology H ∗ Obsq on those open sets. As usual, the
142 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
meaning of a complex is easiest to garner through its cohomology. The main theo-
rem of this section is that the vertex algebra extracted from Obsq is the well-known
βγ vertex algebra.
ai 7→ −iai−1 ,
a∗i 7→ −(i − 1)a∗i−1 ,
n<0 n≥0
∂a −n
and
X X ∂
Y(a∗0 , z) = a∗n z−n − z−n .
n≤0 n>0
∂a−n
V~=2πi W.
3.4. The criteria to obtain a vertex algebra. The βγ field theory is mani-
∂
festly translation-invariant, so it remains to verify that the action of ∂z is homotopi-
cally trivial. Consider the operator
d
η= ,
d(dz)
which acts on the space of fields E . The operator η maps Ω1,1 to Ω1,0 and Ω0,1 to
Ω0,0 . Then we see that
d
[∂ + ~∆, η] = ,
dz
so that the action of d/dz is homotopically trivial, as desired.
We would like to apply the result of Theorem 2.2.1, which shows that a holo-
morphically translation invariant prefactorization algebra with certain additional
conditions gives rise to a vertex algebra. We need to check the following condi-
tions.
((i)) The prefactorization algebra must have an action of S 1 covering the
action on C by rotation.
((ii)) For every disc D(0, r) ⊂ C, including r = ∞, the S 1 action on Obsq (D(0, r))
must extend to an action of the algebra D(S 1 ) of distributions on S 1 .
q
((iii)) If Obsk (D(0, r)) denotes the weight k eigenspace of the S 1 action, then
we require that the extension map
q q
H ∗ (Obsk (D(0, r))) → H ∗ (Obsk (D(0, s)))
is an isomorphism for r < s.
q
((iv)) Finally, we require that the space H ∗ (Obsk (D(0, r))) is a discrete vector
space, that is, a colimit of finite dimensional vector spaces.
The first condition is obvious in our example: the S 1 action arises from the natural
action of S 1 on Ω0,∗
c (C) and Ωc (C). The second condition is also easy to check: if
1,∗
and ∂ preserves the symmetric powers. Hence, we can analyze each symmetric
power
c (U)[1] ⊕ Ωc (U)[1] , ∂)
(Symn Ω1,∗ 0,∗
separately. The nth symmetric power can then be studied as an elliptic complex on
the complex manifold U n .
3.5.1. Recollections. We remind the reader of some facts from the theory of
several complex variables (references for this material are Gunning and Rossi
(1965), Forster (1991), and Serre (1953)). We then use these facts to describe
the cohomology of the observables.
3.5.1 Proposition. Every open set U ⊂ C is Stein Forster (1991). As the product
of Stein manifolds is Stein, every product U n ⊂ Cn is Stein.
Remark: Behnke and Stein (1949) proved that every noncompact Riemann surface
is Stein, so the arguments we develop here extend farther than we exploit them. ^
We need a particular instance of Cartan’s theorem B about coherent analytic
sheaves. See Gunning and Rossi (1965).
3.5.2 Theorem (Cartan’s Theorem B). Let X be a Stein manifold and let E be a
holomorphic vector bundle on X. Then,
(
0, k,0
H (Ω (X, E), ∂) =
k 0,∗
Hol(X, E), k = 0,
where Hol(X, E) denotes the holomorphic sections of E on X.
We now use a corollary noted by Serre (1953); it is a special case of the Serre
duality theorem. (Nowadays, people normally talk about the Serre duality theorem
for compact complex manifolds, but in Serre’s original paper he proved it for non-
compact manifolds too, under some additional hypothesis that will be satisfied on
Stein manifolds).
Note that we use the Fréchet topology on Hol(X, E), obtained as a closed sub-
space of C ∞ (X, E). We let E ! be the holomorphic vector bundle E ∨ ⊗ KX where KX
is the canonical bundle of X.
3.5.3 Corollary. For X a Stein manifold of complex dimension n, the compactly-
supported Dolbeault cohomology is
(
0, k,n
H k (Ω0,∗ (X, E), ∂) =
c (Hol(X, E ! ))∨ , k = n,
3. THE βγ SYSTEM AND VERTEX ALGEBRAS 145
Proof. The first statement follows from the second statement. Locally on X,
the Dolbeault-Grothendieck lemma tells us that the sheaf Ω0,∗ (X, E) has no higher
cohomology, and the sheaves C ∞ (M, Ω0,i (X, E)) are certainly fine.
To prove the second statement, consider the exact sequence
0 → Hol(X, E) → Ω0,0 (X, E) → Ω0,1 (X, E) · · · → Ω0,n (X, E) → 0.
It is an exact sequence of Fréchet spaces. Proposition 3 of section I.1.2 in Grothendieck
(1955) tells us that the completed projective tensor product of nuclear Fréchet
spaces is an exact functor, that is, takes exact sequences to exact sequences. An-
other result of Grothendieck (1952) tells us that for any complete locally con-
vex topological vector space F, C ∞ (M, F) is naturally isomorphic to C ∞ (M)b ⊗π F,
where b ⊗π denotes the completed projective tensor product. The result then fol-
lows.
Proof of the theorem. We need to produce an isomorphism of differentiable
vector spaces, i.e., an isomorphism of sheaves. In other words, we need to find an
isomorphism between their sections on each manifold M,
c (Dr1 ,...,rn ))) C (M, Hol((P \ Dr1 ) × . . . (P × Drn ), O(−1) )),
C ∞ (M, H n (Ω0,∗ ∞ 1 1 n
and this isomorphism must be natural in M. We also need to show that the sections
c (Dr1 ,...,rn ))) of the other cohomology sheaves, with i < n, are zero.
C ∞ (M, H i (Ω0,∗
The first thing we prove is that, for any complex manifold X and holomorphic
vector bundle E on X, and any open subset U ⊂ X, there is an exact sequence
(‡) 0 → C ∞ (M, Ω0,∗
c (U, E)) → C (M, Ω (X, E)) → C (M, Ω (X \ U, E)) → 0
∞ 0,∗ ∞ 0,∗
as sheaves on M. Now, we claim that this sequence is exact on the right, as sheaves.
The point is that, locally on M, every smooth function on M × (X \ U) extends to
a smooth function on some M × (X \ Ki ) and by applying a bump function that is 1
148 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
Taking the colimit as ε → 0, combined with our previous calculations, gives the
desired result. Note that this colimit must be taken in the category of sheaves on
M. We are also using the fact that sequential colimits commute with formation of
cohomology.
Let us introduce some notation to deal with products of these. For a multi-
index K = (k1 , . . . , kn ) or L = (l1 , . . . , lm ), we let
bK (x) = bk1 (x) · · · bkn (x),
cL (x) = cl1 (x) · · · clm (x).
Of course these products make sense only if ki > 0 for all i. Note that under the
natural S 1 action on H 0 (Obscl (D(x, r))), the elements bK (x) and cL (x) are of weight
− |K| and − |L|, where |K| = i ki and similarly for L.
P
3.6.2 Lemma. These observables generate the cohomology of the classical ob-
servables in the following sense.
((i)) For i , 0 the cohomology groups H i (Obscl (D(x, r))) vanish as differ-
entiable vector spaces.
((ii)) The monomials {bK (x)cL (x)}|K|+|L|=n form a basis for the weight n space
H 0 (Obscl (D(x, r)))n . Further, they form a basis in the sense of differ-
entiable vector spaces, meaning that any smooth section
f : M → Hn0 (Obscl (D(x, r)))
can be expressed uniquely as a sum
X
f = fKL bK (x)cL (x),
|K|+|L|=n
where fKL ∈ C ∞ (M) and locally on M all but finitely many of the fKL
are zero.
((iii)) More generally, any smooth section
f : M → H 0 (Obscl (D(x, r)))
150 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
K,L
Under this isomorphism, the observable bK (0)cL (0) goes to the function
1
z−k1 dz1 · · · z−kn −l1 −1
n dzn w1 · · · w−l
m
m −1
.
(2πi)n+m 1
Everything in the statement is now immediate. On H 0 (Obscl (D(0, r))) we use the
topology that is the colimit of the topologies on holomorphic functions on (P1 \
D(0, r − ε))n as ε → 0.
3.6.3 Lemma. We have
H ∗ (Obsq (U)) = H ∗ (Obscl (U))[~]
3. THE βγ SYSTEM AND VERTEX ALGEBRAS 151
right hand side. Explicitly, bi (x) and c j (y) are derivatives of delta-functions and we
apply them to the Green’s function.
We can identify Obsq (U) as a graded vector space with Obscl (U)[~]. The
quantum differential is d = d1 + ~d2 , a sum of two terms, where d1 is the dif-
ferential on Obscl (U) and d2 is the differential arising from the Lie bracket in the
shifted Heisenberg Lie algebra whose Chevalley-Eilenberg chain complex defines
Obsq (U).
It is easy to check that
[~∂P , d1 ] = d2 .
This identity follows immediately from the fact that (1, 1)-current ∂P on C2 is the
delta-current on the diagonal.
As we explained in Section 6, we get an isomorphism of cochain complexes
W : Obscl (U)[~] 7→ Obsq (U)
.
α 7→ e~∂P α
Further, for U1 , U2 disjoint opens in V, the prefactorization product map
?~ : Obscl (U1 )[~] × Obscl (U2 )[~] → Obscl (V)[~],
arising from that on Obsq under the identification W, is given by the formula
α ?~ β = e−~∂P e~∂P α · e~∂P β .
Here · refers to the commutative product on classical observables.
Let us apply this formula to α = b1 (z) and β in the algebra generated by
bi (0) and c j (0). First, note that since b1 (z) is linear, ~∂P b1 (z) = 0. Note also
that [∂P , b1 (z)] commutes with ∂P . Thus, we find that
b1 (z) ?~ β = e−~∂P b1 (z)e~∂P β
= b1 (z)β − [~∂P , b1 (z)]β.
Note that [∂P , b1 (z)] is an order one operator, and so a derivation. So it suffices to
calculate what it does on generators. We find that
1 1 ∂j 1
!
[∂P , b1 (z)]c j (0) =
2πi j! ∂ j w z − w w=0
1 − j−1
= z ,
2πi
and
[∂P , b1 (z)]b j (0) = 0.
In other words,
∞
1 X − j−1 ∂
[∂P , b1 (z)] = z .
2πi j=0 ∂c j (0)
Note as well that for |z| < r, we can expand the cohomology class b1 (z) in H 0 (Obscl (D(0, r))
as a sum
∞
X
b1 (z) = bn+1 (0)zn .
n=0
154 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
Indeed, for a classical field γ ∈ Ω1hol (D(0, r)) satisfying the equations of motion,
we have
b1 (z)(γ) = γ(z)
∞
X 1
= zn γ(n) (0)
n=0
n!
∞
X
= zn bn1 (0)(γ).
n=0
Putting all these computations together, we find that, for β in the algebra generated
by c j (0), bi (0), we have
∞ ∞
X ~ X ∂
b1 (z) ?~ β = bn+1 (0)zn + z−m−1 β ∈ H 0 (Obscl (D(0, r))[~].
n=0
2πi m=0 ∂cm (0)
Thus, if we set ~ = 2πi, we see that the operator product on the space V~=2πi
matches the one on W if we sent bn (0) to a−n and cn (0) to a∗−n . A similar calculation
of the operator product with c0 (z) completes the proof.
we should recover κ-twisted differential operators. When the twist is integral, the
twist corresponds to a line bundle on BunG (Σ) and the twisted differential operators
are precisely differential operators for that line bundle.
It is nontrivial to properly define differential operators on the stack BunG (Σ),
and we will not attempt it here. At the formal neighbourhood of a point, however,
there are no difficulties and our statements are rigorous. ^
4.2. The main result. Note that if we take our Riemann surface to be C, the
prefactorization algebra F κ is holomorphically translation invariant because the
∂ ∂
derivation ∂z is homotopically trivial, via the homotopy given by ∂dz . It follows
that we are in a situation where we might be able to apply Theorem 2.2.1. The
main result of this section is the following.
Theorem. The holomorphically translation invariant prefactorization algebra F κ
on C satisfies the conditions of Theorem 2.2.1, and so defines a vertex algebra.
This vertex algebra is isomorphic to the affine Kac-Moody vertex algebra.
Before we can prove this statement, we of course need to describe the affine
Kac-Moody vertex aigebra.
Recall that the Kac-Moody Lie algebra is the central extension of the loop
algebra Lg = g[t, t−1 ],
0 → C · c →b gκ → Lg → 0.
As vector spaces, we have gκ = g[t, t ] ⊕ C · c, and the Lie bracket is given by the
b −1
formula
I !
[ f (t) ⊗ X, g(t) ⊗ Y]κ := f (t)g(t) ⊗ [X, Y] + f ∂g κ(X, Y)
Observe that g[t] is a sub Lie algebra of the Kac-Moody algebra. The vacuum
module W for the Kac-Moody algebra is the induced representation from the trivial
rank one representation of g[t]. The induction-restriction adjunction provides a
natural map C → W of g[t]-modules, where C is the trivial representation of g[t].
Let |0i ∈ W denote the image of 1 ∈ C.
There is a useful, explicit description of W as a vector space. Consider the sub
Lie algebra C · c ⊕ t−1 g[t−1 ] of b
gκ , which is complementary to g[t]. The action of
this Lie algebra on the vacuum element |0i ∈ W gives a canonical isomorphism
U(C · c ⊕ t−1 g[t−1 ]) W
as vector spaces.
The vacuum module W is also a C[c] module in a natural way, because C[c] is
inside the universal enveloping algebra of b
gκ .
4.2.1 Definition. The Kac-Moody vertex algebra is defined as follows. It is a vertex
algebra structure over the base ring C[c] on the vector space W. (Working over
4. KAC-MOODY ALGEBRAS AND FACTORIZATION ENVELOPES 157
the base ring C[c] simply means all maps are C[c]-linear.) By the reconstruction
theorem 2.3.1, to specify the vertex algebra structure it suffices to specify the state-
field map on a subset of elements of W that generate all of W (in the sense of
the reconstruction theorem). The following state-field operations define the vertex
algebra structure on W.
((i)) The vacuum element |0i ∈ W is the unit for the vertex algebra, that is,
Y(|0i , z) is the identity.
((ii)) If X ∈ t−1 g ⊂ bgκ , we have an element X |0i ∈ W. We declare that
X
Y(X |0i , z) = Xn z−1−n
n∈Z
where Xn = tn X ∈ b
gκ , and we are viewing elements of b
gκ as endomor-
phisms of W.
4.4. Proof of the theorem. Let us now prove that the vertex algebra associ-
ated to Fκ is isomorphic to the Kac-Moody vertex algebra. The proof will be a
little different than the proof of the corresponding result for the βγ system.
158 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
is, in fact, a subcomplex. Applying this fact to U = ρ−1 (I) for an interval I ⊂ R>0
and taking cohomology, we obtain a linear map
−1 κ
H 1 (Ω0,∗ 0
c (ρ (I))) ⊗ g) ⊕ C · c → H (F (I)).
Note that we have a natural identification
c (ρ (I))) = Ωhol (ρ (I)) ,
−1 −1 ∨
H 1 (Ω0,∗ 1
wise, for each integer n, performing a contour integral against zn around this circle
c (ρ (I))) that we call φ(z ).
defines an element of H 1 (Ω0,∗ −1 n
The map
gκ → H ∗ (F κ (ρ−1 (I)))
i :b
constructed by the theorem factors through the map
i: gκ → H 1 (Ω0,∗
b −1
c (ρ (I))) ⊗ g ⊕ C · c
n
Xz 7→ φ(zn ) ⊗ X .
c 7→ c
4. KAC-MOODY ALGEBRAS AND FACTORIZATION ENVELOPES 159
Note that Cauchy’s integral theorem ensures that, in a sense, it does not matter
what circle we use for the contour integrals that define φ. Hence, this map should
produce a map from the universal enveloping algebra of b gκ , viewed as a locally
∗ κ
constant prefactorization algebra, to H F .
Proof. In Chapter 3, Section 4, we showed how the universal enveloping alge-
bra of any Lie algebra a arises as a prefactorization envelope. Let U f act (a) denote
the prefactorization algebra on R that assigns to an interval I, the cochain complex
U(a) f act (I) = C∗ (Ω∗c (I, a)).
We showed that the cohomology of U(a) f act is locally constant and corresponds to
the ordinary universal enveloping algebra Ua.
Let us apply this construction to a = b
gκ . To prove the theorem, we will produce
a map of prefactorization algebras on R>0
gκ ) f act → ρ∗ F κ .
i : U(b
Since both sides are defined as the Chevalley-Eilenberg chains of local Lie alge-
bras, it suffices to produce such a map at the level of dg Lie algebras.
We will produce such a map in a homotopical sense. To be explicit, we will
introduce several precosheaves of Lie algebras and quasi-isomorphisms
' '
(†) Ω∗c ⊗b
gκ −→ L1 L01 −→ L2 .
In the middle, between L1 and L01 , there will be a cochain homotopy equivalence.
At the level of cohomology, we then obtain an isomorphism
gκ = H ∗ (Ω∗c ⊗b
b gκ ) −→ H ∗ L2
that proves the theorem.
Let us remark on a small but important point. These precosheaves will satisfy
a stronger condition that ensures the functor C∗ of Chevalley-Eilenberg chains pro-
duces a prefactorization algebra. Consider the symmetric monoidal category whose
underlying category is dg Lie algebras but whose symmetric monoidal structure is
given by direct sum as cochain complexes. (The coproduct in dg Lie algebras is not
this direct sum.) All four precosheaves from the sequence (†) are prefactorization
algebras valued in this symmetric monoidal category. In general, a precosheaf L
of dg Lie algebras is such a prefactorization algebra if it has the property that that
for any two disjoint opens U, V in W, the elements in L(W) coming from L(U)
commute with those coming from L(V).
We now describe these prefactorization dg Lie algebras.
Let L1 be the prefactorization dg Lie algebras on R that assigns to an interval
I the dg Lie algebra
L1 (I) = Ω∗c (I) ⊗ g[z, z−1 ] ⊕ C · c[−1].
Thus, L1 is a central extension of Ω∗c ⊗ g[z, z−1 ]. The cocycle defining the central
extension on the interval I is
Z ! I !
αXzn ⊗ βYzm 7→ α ∧ β κ(X, Y) zn ∂z zm ,
I
160 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
Finally, we check that this map has the properties discussed just after the state-
ment of the theorem. Let I ⊂ R>0 be an interval. Then, under the isomorphism
→ H ∗ (C∗ (L01 (I))),
gκ ) −
i : U(b
the element Xzn is represented by the element
f (r)zn Xdr ∈ L01 (U)[1],
where f has compact support and is chosen so that f (r)dr = 1.
R
F κ (D(0, r0 )) ⊗ F κ (A(s, s0 )) / F κ (D(0, s0 ))
commutes.
We need to show that this action identifies V with the vacuum representation
W. The putative map from W to V sends the vacuum element |0i to the unit ob-
servable 1 ∈ V. To show that this map is well-defined, we need to show that 1 ∈ V
is annihilated by the elements zn X ∈ b
gκ for n ≥ 0.
The unit axiom for prefactorization algebras implies that the following diagram
commutes:
U(bgκ ) / H ∗ (F κ (A(s, s0 )) .
V / H ∗ (F κ (D(0, s0 ))
The left vertical arrow is given by the action of U(b gκ ) on the unit element 1 ∈ V, and
the right vertical arrow is the map arising from the inclusion A(r, r0 ) ⊂ D(0, r0 ). The
bottom right arrow is the inclusion into the direct sum of S 1 -eigenspaces, which is
injective.
The proof of Theorem 4.4.1 gives an explicit representative for the element
Xzn ∈ b gκ in F κ (A(s, s0 )). Namely, let f (r) be a function supported in the interval
(s, s ) and with f (r)dr = 1. Then Xzn is represented by
R
0
1 iθ n
2 e f (r)dzz X ∈ Ω0,1 0
c (A(s, s )) ⊗ g.
We will show that this element is exact when it is viewed as an element of Ω0,1 0
c (D(0, s ))⊗
n
g, so that Xz vanishes when going right and then down in the square above.
Set Z r
h(r) = f (t)dt,
∞
∂
so h(r) = 0 for r and h(r) = 1 for r < s. Also,
s0 ∂r h = f . The polar-coordinate
∂
representation of ∂z tells us that
Thus, we have shown that the elements Xzn ∈ b gκ , where n ≥ 0, act by zero on the
element 1 ∈ V. We therefore have a unique map of U(b gκ -modules from W to V
sending |0i to 1.
It remains to show that this map is an isomorphism. Note that every object
we are discussing is filtered. The universal enveloping algebra U(b gκ ) is filtered, as
usual, by setting F i U(b gκ ) to be the subspace spanned by products of ≤ i elements
of b
gκ . (This filtration is inherited from the tensor algebra.) Similarly, the space
F κ (U) = C∗ (Ω0,∗
c (U, g) ⊕ C · c[−1])
is filtered by saying that F i F κ (U) is the subcomplex C≤i . All the maps we have
been discussing are compatible with these increasing filtrations.
The associated graded of U(b gκ ) is the symmetric algebra of b
gκ , and the associ-
ated graded of F κ (U) is the appropriate symmetric algebra on Ω0,∗ c (U, g)[1] ⊕ C · c
in differentiable cochain complexes. Upon taking associated graded, the maps
gκ ) → Gr H ∗ F κ (A(s, s0 )),
Gr U(b
Gr H ∗ (F κ (A(s, s0 )) → Gr H ∗ F κ (D(0, s0 ))
are maps of commutative algebras. It follows that the map
gκ ) → Gr V ⊂ H ∗ Sym Ω0,∗
Gr U(b c (D(0, s), g)[1] ⊕ C · c
given by the product of the projection maps to the S 1 -eigenspaces, we get a map
mz,0 : V ⊗ V → V.
This map depends holomorphically on z. The operator product map is obtained as
the Laurent expansion of mz,0 .
Our aim is to calculate the operator product map and identify it with the vertex
operator map in the Kac-Moody vertex algebra. We use the following notation. If
X ∈ g, let Xi = zi X ∈ b gκ on V by the symbol ·. Then we
gκ . We denote the action of b
have the following.
4.4.3 Proposition. For all v ∈ V,
X
mz,0 (X−1 · 1, v) = z−i−1 (Xi · v) ∈ V
i∈Z
where the sum on the right hand side converges.
Before we prove this proposition, let us observe that it proves our main result.
4.4.4 Corollary. This isomorphism of U(b gκ )-modules from the vacuum represen-
tation W to V is an isomorphism of vertex algebras, where V is given the vertex
algebra structure arising from the prefactorization algebra F κ , and W is given the
Kac-Moody vertex algebra structure defined in 4.2.1.
Proof. This corollary follows immediately from the reconstruction theorem
2.3.1.
Proof of the proposition. The inclusion of a disc into an annulus induces a
structure map of the prefactorization algebra. Each element v of V thus determines
an element of the prefactorization algebra on that annulus, and hence a map from
V to V. The formula will follow by applying our earlier analysis of the structure
maps.
As before, we let Xi ∈ b gκ denote Xzi for X ∈ g. As we explained in the
discussion following Theorem 4.4.1, we can view the element
X−1 · 1 ∈ V
as being represented at the cochain level by X times the linear functional
Ω1hol (D(0, s)) → R C
α 7→ |z|=r z−1 α.
Cauchy’s theorem tells us that this linear functional sends h(z)dz, when h is holo-
morphic, to 2πih(0).
Now fix z0 ∈ A(s, s0 ), and let
ιz0 : V → H ∗ (F κ (A(s, s0 )))
166 5. HOLOMORPHIC FIELD THEORIES AND VERTEX ALGEBRAS
denote the map arising from the restriction to V of the structure map
H ∗ (F κ (D(0, ε))) = H ∗ (F κ (D(z0 , ε))) → H ∗ (F κ (A(s, s0 )))
arising from the inclusion of the disc D(z0 , ε) into the annulus A(s, s0 ). It is clear
from the definition of mz0 ,0 and the axioms of a prefactorization algebra that the
following diagram commutes:
mz0 ,0
V ⊗ V ⊗ Id /V .
O
ιz0
H ∗ (F κ (A(s, s0 )) ⊗ V / H ∗ (F κ (D(0, ∞)))
Here the bottom right arrow is the restriction to V of the prefactorization structure
map
H ∗ (F κ (A(s, s0 )) ⊗ H ∗ (F κ (D(0, ε)) → H ∗ (F κ (D(0, ∞)).
It therefore suffices to show that
∈ H ∗ (F κ (A(s, s0 )))
X
ιz0 (X−1 · 1) = Xi z−i−i
0
i∈Z
Expanding (z−z0 )−1 in the regions when |z| < |z0 | (relevant for the first integral) and
when |z| > |z0 | (relevant for the second integral) gives the desired expression.
Part 3
Factorization algebras
CHAPTER 6
1. Factorization algebras
A factorization algebra is a prefactorization algebra that satisfies a local-to-
global axiom. This axiom is the analog of the gluing axiom for sheaves; it ex-
presses how the values on big open sets are determined by the values on small
open sets. We thus begin by reviewing the notion of a (co)sheaf, with more back-
ground and references available in Appendix 4. Next, we introduce Weiss covers,
which are the type of covers appropriate for factorization algebras. Finally, we give
the definition of factorization algebras,
given by taking the product of the structure maps A(U) → A(Ui ). A presheaf A
is a sheaf if for every U and for every open cover of U, this map equalizes the pair
of maps Y Y
A(Ui ) ⇒ A(Ui ∩ U j )
i i, j
given by the two inclusions Ui ∩ U j → Ui and Ui ∩ U j → U j . That is, A(U) is
the limit of that diagram. In words, this condition means that a section s of A on
U can be described as sections si on each element of the cover that agree on the
overlaps, i.e., si |Ui ∩U j = s j |Ui ∩U j .
We now dualize this definition. A precosheaf Φ on M with values in a category
C is a functor from Opens(M) to C. It is a cosheaf if for every open cover of U, the
canonical map given by taking the coproduct of the structure map Φ(Uk ) → Φ(U),
a
Φ(Uk ) → Φ(U),
k
coequalizes the pair of maps
a a
Φ(Ui ∩ U j ) ⇒ Φ(Uk )
i, j k
is exact, where the first map is the difference of the structure maps for the two
inclusions. In the remainder of this book, we will phrase our gluing axioms in this
style, as we will always restrict to linear target categories.
1.2. Weiss covers. For factorization algebras, we require our covers to be fine
enough that they capture all the “multiplicative structure” — the structure maps —
of the underlying prefactorization algebra. In fact, a factorization algebra will be a
cosheaf with respect to this modified notion of cover.
1.2.1 Definition. Let U be an open set. A collection of open sets U = {Ui | i ∈ I}
is a Weiss cover of U if for any finite collection of points {x1 , . . . , xk } in U, there is
an open set Ui ∈ U such that {x1 , . . . , xk } ⊂ Ui .
The Weiss covers define a Grothendieck topology on Opens(M), the poset cate-
gory of open subsets of a space M. We call it the Weiss topology of M. (This notion
1. FACTORIZATION ALGEBRAS 171
was introduced in Weiss (1999) and it is explained nicely and further developed in
Boavida de Brito and Weiss (2013).)
Remark: A Weiss cover is certainly a cover in the usual sense, but a Weiss cover
typically contains an enormous number of opens. It is a kind of “exponentiation”
of the usual notion of cover, because a Weiss cover is well-suited to studying all
configuration spaces of finitely many points in U. For instance, given a Weiss cover
U of U, the collection
{Uin | i ∈ I}
provides a cover in the usual sense of U n ⊂ M n for every positive integer n. ^
Example: For a smooth n-manifold M, there are several simple ways to construct
a Weiss cover for M. One construction is simply to take the collection of open sets
in M diffeomorphic to a disjoint union of finitely many copies of the open n-disc.
Alternatively, the collection of opens {M \ q | q ∈ M} is a Weiss cover of M. One
can also produce a Weiss cover of a manifold with countably many elements. Fix
a Riemannian metric on M. Pick a collection of points {qn ∈ M | n ∈ N} such that
the union of the unit “discs” D1 (qn ) = {q ∈ M | d(qn , q) < 1} covers M. (Such
an open may not be homeomorphic to a disc if the injectivity radius of qn is less
than 1.) Consider the collection of all “discs” D = {D1/m (qn ) | m ∈ N, n ∈ N}.
The collection of all finite, pairwise disjoint union of elements of D is a countable
Weiss cover. ^
The examples above suggest the following definition.
1.2.2 Definition. A cover U = {Uα } of M generates the Weiss cover V if every open
V ∈ V is given by a finite disjoint union of opens Uα from U.
is an isomorphism.
172 6. FACTORIZATION ALGEBRAS: DEFINITIONS AND CONSTRUCTIONS
term is M
Φ(U j0 ∩ · · · ∩ U jk ),
j0 ,..., jk ∈I
ji all distinct
and note that this term has an “internal” differential arising from Φ. The “external”
differential is the alternating sum of the structure maps arising from inclusion of
k + 1-fold intersections into k-fold intersections. The Čech complex of U with
coefficients in Φ is the totalization of this double complex:
∞
M M
Č(U, Φ) =
Φ(U j0 ∩ · · · ∩ U jk )[k] .
k=0 j0 ,..., jk ∈I
ji all distinct
There is a natural map from the 0th term (of the double complex) to Φ(U) given by
the sum of the structure maps Φ(Ui ) → Φ(U). Thus, the Čech complex possesses
a natural map to Φ(U). We say that Φ is a homotopy cosheaf if the natural map
from the Čech complex to Φ(U) is a quasi-isomorphism for every open U ⊂ M and
every open cover of U.
Remark: This Čech complex is simply the dual of the Čech complex for sheaves
of cochain complexes. Note that this Čech complex is also the normalized cochain
complex arising from a simplicial cochain complex, where we evaluate Φ on the
simplicial space U• associated to the cover U. In more sophisticated language, the
Čech complex describes the homotopy colimit of Φ on this simplicial diagram. ^.
We can now define our main object of interest. Let C be a multicategory whose
underlying category is a Grothendieck abelian category. Then there is a natural
multicategory whose underlying category is Ch(C), the category of cochain com-
plexes in C in which the weak equivalences are quasi-isomorphisms.
1.4.1 Definition. A homotopy factorization algebra is a prefactorization algebra
F on X valued in Ch(C), with the property that for every open set U ⊂ X and Weiss
cover U of U, the natural map
Č(U, F ) → F (U)
is a quasi-isomorphism. That is, F is a homotopy cosheaf with respect to the Weiss
topology.
Remark: In light of the weak equivalences, the strict notion of multiplicative fac-
torization algebra is not appropriate for the world of cochain complexes. Whenever
174 6. FACTORIZATION ALGEBRAS: DEFINITIONS AND CONSTRUCTIONS
Remark: Although the multiplicativity axiom does not directly extend to the setting
of multicategories, we can examine the bilinear structure map
mU,V
UtV ∈ Ch(DVS)(F (U), F (V) | F (U t V))
for two disjoint opens U and V, and we can ask about the image of mU,V UtV . For
almost all the examples constructed in this book, the differentiable vector spaces
that appear are convenient vector spaces, which have a natural topology. In our
constructions, one can typically see that the image of the structure map is dense
with respect to this natural topology on F (U t V), which is a cousin of the mul-
tiplicativity axiom. In practice, though, it is the descent axiom that is especially
important for our purposes. ^
1.5. Weak equivalences. The notion of weak equivalence, or quasi-isomorphism,
on cochain complexes induces such a notion on homotopy factorization algebras.
1.5.1 Definition. Let F, G be homotopy factorization algebras valued in Ch(C), as
above. A map φ : F → G of homotopy factorization algebras is a weak equivalence
if, for all open subsets U ⊂ M, the map F(U) → G(U) is a quasi-isomorphism of
cochain complexes.
The answer reveals a key limitation of our axioms: the concept of factorization
algebra is only appropriate for perturbative quantum field theories.
Indeed, in a perturbative gauge theory, the gauge field (i.e., the connection) is
taken to be an infinitesimally small perturbation A0 + δA of a fixed connection A0 ,
which is a solution to the equations of motion. There is a well-known formula (the
time-ordered exponential) expressing the holonomy of A0 + δA as a power series in
δA, where the coefficients of the power series are given as integrals over Lk , where
L is the loop which we are considering. This expression shows that the holonomy
of A0 +δA can be built up from observables supported at points (which happen to lie
on the loop L). Thus, the holonomy observable will form part of our factorization
algebra.
However, if we are not working in a perturbative setting, this formula does not
apply, and we would not expect (in general) that the prefactorization algebra of
observables satisfies the local-to-global axiom.
3.1. Varying in smooth families. It would be nice, for instance, to let the
diffeomorphism group act smoothly, in some sense, and not just as a discrete group.
Within the framework we’ve developed, there is a natural approach, parallel to
our definition of a smoothly equivariant prefactorization algebra (see Section 7.3).
Note that embeddings define a natural sheaf of sets on the site Mfld of smooth
manifolds: given n-manifolds M, N, the sheaf Emb(M, N) assigns to the manifold
X the subset
{ f ∈ C ∞ (X × M, N) | ∀x ∈ X, f (x, −) is an embedding of M in N}.
In other words, it gives smooth families over X of embeddings of M into N.
3.1.1 Definition. Let Embn denote the category enriched over Sh(Mfld) whose
objects are smooth n-manifolds and in which Embn (M, N) is the sheaf Emb(M, N).
It possesses a symmetric monoidal structure under disjoint union.
3.2. Other kinds of geometry. This kind of construction works very gener-
ally. For instance, if we are interested in complex manifolds, we could work in the
following setting.
3.2.1 Definition. Let Holn denote the category whose objects are complex n-manifolds
and whose morphisms are open holomorphic embeddings. It possesses a symmetric
monoidal structure under disjoint union.
0,∗
monoidal structure. Thus, the factorization envelope Uhol
n g : M 7→ C ∗ (Ωc (M) ⊗ g)
provides a prefactorization algebra on complex n-manifolds.
The Kac-Moody factorization algebras (see example 6.3) provide an example
on Hol1 (i.e., for Riemann surfaces). In Williams (n.d.), Williams constructs the
Virasoro factorization algebras on Hol1 . These examples extend naturally to Holn
for any n, although the choices of central extension become more interesting and
complicated. (Indeed, one finds extensions as L∞ algebras and not just dg Lie
algebras.)
It should be clear that one can enrich this category Holn over Sh(Mfld) and
hence to talk about smooth families of such holomorphic embeddings. (Or, indeed,
one can enrich it over a site of all complex manifolds.) Thus, one can talk, for
instance, about smooth prefactorization algebras on complex n-manifolds.
In general, let G denote some kind of local structure for n-manifolds, such
a Riemannian metric or complex structure or orientation. In other words, G is
a sheaf on Embn . A G-structure on an n-manifold M is a section G ∈ G(M).
There is a category EmbG whose objects are n-manifolds with G-structure (M, G M )
and whose morphisms are G-structure-preserving embeddings, i.e., embeddings
f : M ,→ N such that f ∗G N = G M . This category is fibered over EmbG . One can
then talk about prefactorization algebras on G-manifolds.
3.3. Application to field theory. This notion is quite useful in the context of
field theory. One often studies a field theory that makes sense on a large class of
manifolds. Indeed, physicists often search for action functionals that depend only
on a particular geometric structure. For instance, a conformal field theory only
depends on a conformal class of metric, and so the solutions to the equations of
motion form a sheaf on the site of conformal manifolds. For example, the free
βγ system makes sense on all conformal 2-manifolds, indeed, its solutions form
a sheaf on Hol1 . Similarly, the physicist’s examples of a topological field theory
only depend on the underlying smooth manifold (at least at the classical level). The
solutions to the equations of motion for Abelian Chern-Simons theory (see Section
5) make sense on the site of oriented 3-manifolds. (More generally, Chern-Simons
theory for gauge group G makes sense on the site of oriented 3-manifolds with a
principal G-bundle.) Finally, the free scalar field theory with mass m makes sense
on the site of Riemannian n-manifolds.
Since most of the constructions with field theories in this book have exploited
the enveloping factorization algebra construction, the classical and quantum ob-
servables we have constructed are typically prefactorization algebras on G-manifolds
for some G. This universality of these quantizations — that these prefactorization
algebras of quantum observables simultaneously work for all manifolds with some
geometric structure — illuminates why our constructions typically recover stan-
dard answers: these are the answers that work generally.
For interacting theories, the focus of Volume 2, the quantizations are rarely so
easily obtained or described. On a fixed manifold, a quantization of the prefactor-
ization algebra of classical observables might exist, but it might not exist for all
manifolds with that structure. More explicitly, quantization proceeds via Feynman
3. VARIANT DEFINITIONS OF FACTORIZATION ALGEBRAS 179
3.4. Descent axioms. So far we have only discussed a variant of the notion of
prefactorization algebra. We now turn to the local-to-global axiom in this context.
3.4.1 Definition. A Weiss cover of a G-manifold M is a collection of G-embeddings
{φi : Ui → M}i∈I such that for any finite set of points x1 , . . . , xn ∈ M, there is some
i such that {x1 , . . . , xn } ⊂ φi (Ui ).
With this definition in hand, we can formulate the natural generalization of our
earlier definition.
3.4.2 Definition. A factorization algebra on G-manifolds is a symmetric monoidal
functor F : EmbG → Ch(DVS) that is a homotopy cosheaf in the Weiss topology.
Let’s return to our running example. Our arguments in Section 6 show that
the factorization envelope of a dg Lie algebra using the de Rham complex is a
factorization algebra on Embn , and similarly the factorization enveloped using the
Dolbeault complex is a factorization algebra on Holn .
Remark: These definitions, indeed the spirit of this variation on factorization alge-
bras, are inspired by conversations with and the work of John Francis and David
Ayala. See Ayala and Francis (2015) for a wonderful overview of factorization
algebras in the setting of topological manifolds, along with deep theorems about
them. We note that they use the term “factorization homology,” rather than fac-
torization algebra, as they view the construction as a symmetric monoidal analog,
for n-manifolds, of ordinary homology. In this picture, the value on a disc defines
the coefficients for a homology theory on n-manifolds. (We discuss factorization
homology in the next section.) In further work Ayala et al. (n.d.b), they have devel-
oped a generalization of factorization techniques where the coefficients are (∞, n)-
categories. One payoff of their results is a proof of the Cobordism Hypothesis,
and hence an explanation of functorial topological field theories via factorization-
theoretic thinking. (See Lurie (2009b) for the canonical reference on the Cobor-
dism Hypothesis.) ^
Remark: The quantizations mentioned above produce factorization algebras and
not just prefactorization algebras. Hence, their values on big manifolds can be
computed from their values on smaller manifolds, via the local-to-global axiom. It
is interesting to compare this approach to understanding the global behavior of a
quantum field theory — by a kind of open cover — with the approach in functo-
rial field theory, in the style of Atiyah-Segal-Lurie, where a manifold is soldered
together from manifolds with corners. The Ayala-Francis work shows that the
180 6. FACTORIZATION ALGEBRAS: DEFINITIONS AND CONSTRUCTIONS
These maps correspond to the standard En structure on the n-fold loop space.
For a particularly nice example, consider the case where the source manifold
is M = R. The structure maps of F then describe the standard product on the space
Ω p X of based loops in X. At the level of components, we recover the standard
product on π0 Ω p X = π1 (X, p).
This prefactorization algebra F does not always satisfy the gluing axiom. How-
ever, Salvatore Salvatore (2001) and Lurie Lurie (n.d.b) have shown that if X is
sufficiently connected, this prefactorization algebra is in fact a factorization alge-
bra.
4.2. Relationship with Hochschild homology. The direct relationship be-
tween En algebras and locally constant factorization algebras on Rn raises the
question of what the local-to-global axiom means from an algebraic point of view.
Evaluating a locally constant factorization algebra on a manifold M is known as
factorization homology or topological chiral homology.
The following result, for n = 1, is striking and helpful.
4.2.1 Theorem. For A an E1 algebra (e.g., an associative algebra), there is a weak
equivalence
A f act (S 1 ) ' HH∗ (A),
where A f act denotes the locally constant factorization algebra on R associated to
A and HH∗ (A) denotes the Hochschild homology of A.
Here HH∗ (A) ' A ⊗LA⊗A A means any cochain complex quasi-isomorphic to
the usual bar complex (i.e., we are interested in more than the mere cohomology
groups). In particular, this theorem says that the cohomology of the global sections
on S 1 of A f act equals the Hochschild homology of the E1 algebra A.
This result is one of the primary motivations for the higher dimensional gener-
alizations of factorization homology. It has several proofs in the literature, depend-
ing on choice of gluing axiom and level of generality (for instance, one can work
with algebra objects in more general ∞-categories). See Theorem 3.9 of Ayala and
Francis (2015) or Theorem 5.5.3.11 of Lurie (n.d.b); a barehanded proof for dg
algebras can be found in Section 4.3 of Gwilliam (2012).
There is a generalization of this result even in dimension 1. Note that there
are more general ways to extend A f act from the real line to the circle, by allowing
“monodromy.” Let σ denote an automorphism of A. Pick an orientation of S 1 and
fix a point p in S 1 . Let (A f act )σ denote the prefactorization algebra on S 1 such
that on S 1 − {p} it agrees with A f act but where the structure maps across p use the
automorphism σ. For instance, if L is a small interval to the left of p, R is a small
interval to the right of p, and M is an interval containing both L and R, then the
structure map is
A⊗A → A
a ⊗ b 7→ a ⊗ σ(b)
where the leftmost copy of A corresponds to L and so on. It is natural to view the
copy of A associated to an interval containing p as the A − A bimodule Aσ where
A acts as the left by multiplication and the right by σ-twisted multiplication.
4. LOCALLY CONSTANT FACTORIZATION ALGEBRAS 183
E c (U) is the continuous dual to E ! (U) and that Ec (U) is the continuous dual to
!
E (U).
5.1.2 Lemma. A local cochain complex (E , d) is a sheaf of differentiable cochain
complexes. In fact, it is also a homotopy sheaf. Similarly, compactly supported
sections (Ec , d) is a cosheaf of differentiable cochain complexes, as well as a ho-
motopy cosheaf.
Proof. We describe the cosheaf case as the sheaf case is parallel.
In Section 7.2, we showed that precosheaf Eck given the degree k vector space
is a cosheaf with values in DVS. Thus we know that Ec , without the differential, is
a cosheaf of graded differentiable vector spaces. As colimits of cochain complexes
(note: not homotopy hocolimits) are computed degreewise, we see that (Ec , Q) is a
cosheaf of differentiable cochain complexes. Identical arguments apply to E c .
The homotopy cosheaf condition is only marginally more difficult. Suppose we
are interested in an open U and a cover U for U. Recall that the Čech complex is the
totalization of a double complex; we will view d as the vertical differential and the
Čech differential as horizontal, so that the double complex is concentrated in the
left half plane. We will work with the “augmented” double complex in which Ec (U)
is added as the degree 1 column. The contracting cochain homotopy constructed in
Section 7.2 applies to the cosheaf Eck of degree k elements. Hence this augmented
double complex is acyclic in the horizontal direction. The usual “staircase” or
“zigzag” then ensures that the whole double complex is quasi-isomorphic to the
rightmost column.
The factorization algebras we will discuss are constructed from the symmetric
! !
algebra on the vector spaces Ec (U) and E c (U). Note that, since E c (U) is dual to
d E !c (U) as the algebra of formal power series on E (U).
E (U), we can view Sym
Thus, we often write
d E !c (U) = O(E (U)),
Sym
because we view this algebra as the space of “functions on E (U).”
Note that if U → V is an inclusion of open sets in M, then there are natural
maps of commutative dg algebras
Sym Ec (U) → Sym Ec (V),
d Ec (U) → Sym
Sym d Ec (V),
Sym E c (U) → Sym E c (V),
d E c (U) → Sym
Sym d E c (V).
5.2. The main theorem. The main result of this section is the following.
186 6. FACTORIZATION ALGEBRAS: DEFINITIONS AND CONSTRUCTIONS
5.2.1 Theorem. We have the following parallel results for vector bundles and local
cochain complexes.
((i)) Let E be a vector bundle on M. Then
((a)) Sym Ec and Sym E c are strict (non-homotopical) factorization
algebras valued in the category of differentiable vector spaces,
and
d Ec and Sym
((b)) Sym d E c are strict (non-homotopical) factorization
algebras valued in the category of differentiable pro-vector spaces.
((ii)) Let E be a local cochain complex on M. Then
((a)) Sym Ec and Sym E c are homotopy factorization algebras valued
in the category of differentiable cochain complexes, and
d Ec and Sym
((b)) Sym d E c are homotopy factorization algebras valued
in the category of differentiable pro-cochain complexes.
Proof. Let us first prove the strict version of the result. To start with, consider
the case of Sym Ec . We need to verify the local-to-global axiom.
Let U be an open set in M and let U = {Ui | i ∈ I} be a Weiss cover of U. We
need to prove that Sym∗ Ec (U) is the cokernel of the map
M M
Sym(Ec (Ui ∩ U j )) → Sym(Ec (Ui ))
i, j∈I i∈I
Our cover U is a Weiss cover. This means that, for every finite set of points
x1 , . . . , xm ∈ U, we can find an open Ui in the cover U containing every x j . This
implies that the subsets of U m of the form (Ui )m , where i ∈ I, cover U m . Further,
(Ui )m ∩ (U j )m = (Ui ∩ U j )m .
The desired isomorphism now follows from the fact that Ecm is a cosheaf on M m .
The same argument applies to show that Sym Ec! is a factorization algebra.
In the completed case, essentially the same argument applies, with the subtlety
6. FACTORIZATION ALGEBRAS FROM LOCAL LIE ALGEBRAS 187
(see Section 4) that, when working with pro-cochain complexes, the direct sum is
completed.
For the homotopy case, the argument is similar. Let U = {Ui | i ∈ I} be a
Weiss cover of an open subset U of M. Let F = Sym Ec denote the prefactoriza-
tion algebras we are considering (the argument below will apply when we use the
completed symmetric product or use E c instead of Ec ). We need to show that the
map
Č(U, F ) → F (U)
is an equivalence, where the left hand side is equipped with the standard Čech
differential.
Let F m (U) = Symm Ec . Both sides above split as a direct sum over m, and the
map is compatible with this splitting. (If we use the completed symmetric product,
this decomposition is as a product rather than a sum.)
We thus need to show that the map
M
Symm Ec Ui1 ∩ · · · ∩ Uin [n − 1] → Symm (Ec (U))
i1 ,...,in
is a weak equivalence.
For i ∈ I, we get an open subset Uim ⊂ U m . Since U is a Weiss cover of U,
these open subsets form a cover of U m . Note that
Uim1 ∩ · · · ∩ Uimn = Ui1 ∩ · · · ∩ Uin m .
Note that Ec (U)⊗m can be naturally identified with Γc (U m , E m ) (where the tensor
product is the completed projective tensor product).
Thus, to show that the Čech descent axiom holds, we need to verify that the
map M
Γc Uim1 ∩ · · · ∩ Uimn , E m [n − 1] → Γc (V m , E m )
i1 ,...,in ∈I
is a quasi-isomorphism. The left hand side above is the Čech complex for the
cosheaf of compactly supported sections of E m on V m . Standard partition of unity
arguments show that this map is a weak equivalence. (See Section 4.4 for discus-
sion and references.)
Consider the Čech complex of OL with respect to some Weiss cover U for an
open U in M. Applying the filtration above to each side of the map
Č(U, OL) → OL(U),
we get a map of spectral sequences. On the first page, this map is a quasi-isomorphism
by Theorem 5.2.1. Hence the original map of filtered complexes is a quasi-isomorphism.
CHAPTER 7
Note that for the map to a point f : M → pt, the pushforward factorization
algebra f∗ F is simply the global sections of F . We also call this the factorization
homology of F on M.
2.2. The statement. In this section we will show that any U-factorization al-
gebra on X extends to a factorization algebra on X. This extension is unique up to
quasi-isomorphism.
For an open V ⊂ X, we use UV to denote the collection of all the opens in U
contained inside V.
Let F be a U-factorization algebra. Let us define a prefactorization algebra
ext(F ) on X by
ext(F )(V) = Č(UV , F ),
2. EXTENSION FROM A BASIS 193
complex obtained by taking the unnormalized cochains. Let C N (A• ) denote the
totalization of the double complex by taking the normalized cochains. Finally, let
shAB : C N (A• ) ⊗ C N (B• ) → C N (A• ⊗ B• )
denote the Eilenberg-Zilber shuffle map, which is a lax symmetric monoidal func-
tor from simplicial cochain complexes to cochain complexes. See Theorem 1.2.3
for some discussion.
2.3.1. Extending values. For an open V ⊂ X, recall UV denotes the Weiss
cover of V given by all the opens in U contained inside V. We then define
ext(F )(V) := Č(UV , F ).
This construction provides a precosheaf on X.
2.3.2. Extending structure maps. The Čech complex for the factorization glu-
ing axiom arises as the normalized cochain complex of a simplicial cochain com-
plex, as should be clear from its construction. We use Č(V, F )• to denote this
simplicial cochain complex, so that
Č(V, F ) = C N (Č(V, F )• ),
with F a factorization algebra and V a Weiss cover.
We construct the structure maps by using the simplicial cochain complexes
Č(V, F )• , as many properties are manifest at that level. For instance, the unary
maps ext(F )(V) → ext(F )(W) arising from inclusions V ,→ W are easy to un-
derstand: UV is a subset of UW and thus we get a map between every piece of the
simplicial cochain complex.
We now explain in detail the map
mVV 0 : ext(F )(V) ⊗ ext(F )(V 0 ) → ext(F )(V ∪ V 0 ),
where V ∩ V 0 = ∅. Note that knowing this map, we recover every other multiplica-
tion map
ext(F )(V) ⊗ ext(F )(V 0 ) / ext(F )(W)
6
mVV 0
)
ext(F )(V ∪ V 0 )
by postcomposing with a unary map.
The n-simplices Č(UV , F )n is the direct sum of the complexes
F (Ui0 ∩ · · · ∩ Uin ),
with each Uik in V. The n-simplices of the tensor product of simplicial cochain
complexes Č(UV , F )• ⊗Č(UV 0 , F )• are precisely the levelwise tensor product Č(UV , F )n ⊗
Č(UV 0 , F )n , which breaks down into a direct sum of terms
F (Ui0 ∩ · · · ∩ Uin ) ⊗ F (U j0 ∩ · · · ∩ U jn )
with the Uik ’s in V and the U jk ’s in V 0 . We need to define a map
mVV 0 ,n : Č(UV , F )n ⊗ Č(UV 0 , F )n → Č(UV∪V 0 , F )n
for every n, and we will express it in terms of these direct summands.
2. EXTENSION FROM A BASIS 195
Now
because F is defined on a factorizing basis. The right hand term is one of the direct
summands for Č(UV∪V 0 , F )n . Summing over the direct summands on the left, we
obtain the desired levelwise map.
Finally, the composition
C N (Č(U ∩ V, F )• ) ⊗ C N (Č(UV 0 , F )• )
sh
C N Č(U ∩ V, F )• ⊗ Č(UV 0 , F )•
C N (mVV 0 ,• )
C N (Č(UV∪V 0 , F )• )
gives us mVV 0 .
A parallel argument works to construct the multiplication maps from n disjoint
opens to a bigger open.
The desired associativity and equivariance are clear at the level of the simplicial
cochain complexes Č(UV , F ), since they are inherited from F itself.
2.3.3. Verifying gluing. We have constructed ext(F ) as a prefactorization al-
gebra, but it remains to verify that it is a factorization algebra. Thus, our goal is
the following.
2.3.1 Proposition. The extension ext(F ) is a cosheaf with respect to the Weiss
topology. In particular, for every open subset W and every Weiss cover W, the
complex Č(W, ext(F )) is quasi-isomorphic to ext(F )(W).
As our gluing axiom is simply the axiom for cosheaf — but using a funny class
of covers — the standard refinement arguments about Čech homology apply. We
now spell this out.
For W an open subset of X and W = {W j } j∈J a Weiss cover of W, there are two
associated covers that we will use:
(a) UW = {Ui ⊂ W | i ∈ I} and
(b) UW = {Ui | ∃ j such that Ui ⊂ W j }.
The first is just the factorizing basis of W induced by U, but the second consists of
the opens in U subordinate to the cover W. Both are Weiss covers of W.
To prove the proposition, we break the argument into two steps and exploit the
intermediary Weiss cover UW .
196 7. FORMAL ASPECTS OF FACTORIZATION ALGEBRAS
Note that each term F (U~i ) appearing in this source complex appears only once
in the target complex Č(UW , F ). Let f send each such term F (U~i ) to its unique
image in the target complex via the identity. This map f is a cochain map: it clearly
respects the internal differential of each term F (U~i ), and it is compatible with the
Čech differential by construction.
Second, we need to show f is a quasi-isomorphism. We will show this by
imposing a filtration on the map and showing the induced spectral sequence is a
quasi-isomorphism on the first page.
We filter the target complex Č(UW , F ) by
MM
F nČ(UW , F ) := F (U~i )[k].
k≤n ~i∈Id
k+1
Equip the source complex Č(W, ext(F )) with a filtration by pulling this filtration
back along f . In particular, for any U~i = Ui0 ∩ · · · ∩ Uin , the preimage under f
consists of a direct sum over all tuples W~j = W j0 ∩ · · · ∩ W jm such that every Uik
is a subset of W~j . Here, m can be any nonnegative integer (in particular, it can be
bigger than n).
Consider the associated graded complexes with respect to these filtrations. The
source complex has
MM M
Gr Č(UW , F ) = C[m] .
F (U~i )[n] ⊗
n≥0 ~i∈Id ~j∈ Jd
n+1 m+1 such that U ⊂W ∀k
ik ~j
The rightmost term (after the tensor product) corresponds to the chain complex for
a simplex — here, the simplex is infinite-dimensional — and hence is contractible.
In consequence, the map of spectral sequences is a quasi-isomorphism.
We now wish to relate the Čech complex on the intermediary UW to that on UW .
3. PULLING BACK ALONG AN OPEN IMMERSION 197
opens is another strongly convex open. Let C denote the Weiss cover generated by
C, which is thus a factorizing basis for X.
There is a closely associated basis on Y. Let C0 denote the collection of all
connected opens U in Y such that f (U) is in C. Let C0 be the factorizing basis of Y
generated by C0 .
Given F a multiplicative factorization algebra on X, let FC denote the asso-
ciated multiplicative factorization algebra on the basis C. We now define a multi-
plicative factorization algebra f ∗ FC0 on the basis C0 as follows. To each U ∈ C0 ,
let f ∗ FC0 (U) = F ( f (U)). To a finite collection of disjoint opens V1 , . . . , Vn in C0 ,
we assign
f ∗ FC0 (V1 t · · · t Vn ) = f ∗ FC0 (V1 ) ⊗ · · · ⊗ f ∗ FC0 (Vn ),
the tensor product of the values on those components, by multiplicativity. For U in
C0 (and thus convex) and V in C0 such that V ⊂ U, equip f ∗ FC0 with the obvious
structure map
f ∗ FC0 (V) = F (V) → F (U) = f ∗ FC0 (U).
By multiplicativity, this determines the structure map for any inclusion in the fac-
torizing basis C0
t j V j ⊂ ti Ui ,
where each Ui ∈ C0 . This inclusion is the disjoint union over i of the inclusions
Ui ∩ (t j V j ) ⊂ Ui ,
and we have already specified the structure map for each i. Let f ∗ F denote the
factorization algebra on Y obtained by extending f ∗ FC0 from the basis C0 .
Note that although we used a choice of Riemannian metric in making the con-
struction, the factorization algebra f ∗ F is unique up to isomorphism. It is enough
to exhibit a map locally and the pullbacks are locally isomorphic on Y.
1.1. Enveloping algebras. Let g be a Lie algebra and let gR denote the local
Lie algebra Ω∗R ⊗ g on the real line R. Recall Proposition 4.0.1 which showed that
the factorization envelope UgR recovers the universal enveloping algebra Ug.
This factorization algebra UgR is also defined on the circle.
1.1.1 Proposition. There is a weak equivalence
Proof. The wonderful fact here is that we do not need to pick a Weiss cover
and work with the Čech complex. Instead, we simply need to examine the cochain
complex C∗ (Ω∗ (S 1 ) ⊗ g).
201
202 8. FACTORIZATION ALGEBRAS: EXAMPLES
One approach is to use the natural spectral sequence arising from the filtration
F k = Sym≤k . The E1 page is C∗ (g, Ug), and one must verify that there are no
further pages.
Alternatively, recall that the circle is formal, so Ω∗ (S 1 ) is quasi-isomorphic to
its cohomology H ∗ (S 1 ) as a dg algebra. Thus, we get a homotopy equivalence of
dg Lie algebras
Ω∗ (S 1 ) ⊗ g ' g ⊕ g[−1],
where, on the right, g acts by the (shifted) adjoint action on g[−1]. Now,
C∗ (g ⊕ g[−1]) C∗ (g, Sym g),
where Sym g obtains a g-module structure from the Chevalley-Eilenberg homology
complex. Direct computation verifies this action is precisely the adjoint action of g
on Ug (we use, of course, that Ug and Sym g are isomorphic as vector spaces).
1.2. The free scalar field in dimension 1. Recall from Section 3 that the
Weyl algebra is recovered from the factorization algebra of quantum observables
Obsq for the free scalar field on R. We know that the global sections of Obsq on
a circle should thus have some relationship with the Hochschild homology of the
Weyl algebra, but we will see that the relationship depends on the ratio of the mass
to the radius of the circle.
For simplicity, we restrict attention to the case where S 1 has a rotation-invariant
metric, radius r, and total length L = 2πr. Let S 1L denote this circle R/ZL. We also
work with C-linear observables, rather than R-linear.
1.2.1 Proposition. If the mass m = 0, then H k Obsq (S 1L ) is C[~] for k = 0, −1 and
vanishes for all other k.
If the mass m satisfies mL = in for some integer n, then H k Obsq (S 1L ) is C[~]
for k = 0, −2 and vanishes for all other k.
For all other values of mass m, H 0 Obsq (S 1L ) C[~] and all other cohomology
groups vanish.
Hochschild homology with monodromy (recall Theorem 4.2.2) provides an
explanation for this result. The equations of motion for the free scalar field locally
have a two-dimensional space of solutions, but on a circle the space of solutions
depends on the relationship between the mass and the length of the circle. In the
massless case, a constant function is always a solution, no matter the length. In
the massive case, there is either a two-dimensional space of solutions (for certain
imaginary masses, because our conventions) or a zero-dimensional space. Viewing
the space of local solutions as a local system, we have monodromy around the
circle, determined by the Hamiltonian.
When the monodromy is trivial (so mL ∈ iZ), we are simply computing the
Hochschild homology of the Weyl algebra. Otherwise, we have a nontrivial auto-
morphism of the Weyl algebra.
Note that this proposition demonstrates, in a very primitive example, that the
factorization homology can detect spectral properties of the Hamiltonian. The
proof will exhibit how the existence of solutions to the equations of motion (in
1. SOME EXAMPLES OF COMPUTATIONS 203
this case, a very simple ordinary differential equation) affects the factorization ho-
mology. It would be interesting to develop further this kind of relationship.
Proof. We directly compute the global sections, in terms of the analysis of the
local Lie algebra
2
!
∞ 1 4+m
L = C (S L ) −−−−→ C (S L )[−1] ⊕ C~
∞ 1
Let Uω gΣ denote the twisted factorization envelope for this extended local Lie al-
gebra.
1.3.1 Proposition. Let g denote the genus of a closed Riemann surface Σ. Then
whose differential has the form ∂ + dLie . (Note that in the Lie algebra, c has coho-
mological degree 1.)
Consider the filtration F k = Sym≤k by polynomial degree. The purely an-
alytic piece ∂ preserves polynomial degree while the Lie part dLie lowers poly-
nomial degree by 1. Thus the E1 page of the associated spectral sequence has
Sym(H ∗ (Σ, O) ⊗ g ⊕ Cc) as its underlying graded vector space. The differential
now depends purely on the Lie-theoretic aspects. Moreover, the differential on
further pages of the spectral sequence are zero.
In the untwisted case (where there is no extension), we find that the E1 page is
precisely H ∗ (g, Ug⊗g ), by computations directly analogous to those for Proposition
1.1.1.
In the twisted case, the only subtlety is to understand what happens to the
extension. Note that H 0 (Ω0,∗ (Σ) is spanned by the constant functions. The pairing
ω vanishes if a constant function is an input, so we know that the central extension
does not contribute to the differential on the E1 page.
An alternative proof is to use the fact the Ω0,∗ (Σ) is homotopy equivalent to
its cohomology as a dg algebra. Hence we can compute C∗ (H ∗ (Σ, O) ⊗ g ⊕ Cc)
instead.
We can use the ideas of sections 6 and 7 to understand the “correlation func-
tions” of this factorization algebra Uω gΣ . More precisely, given a structure map
Then we have
−1 −1
−1
(∂ + dLie ) (∂ a1 )a2 · · · ak = ∂(∂ a1 )a2 · · · ak + dLie (∂ a1 )a2 · · · ak
k
−1 X −1
= ∂(∂ a1 )a2 · · · ak + [∂ a1 , a j ]a2 · · · abj · · · ak
j=2
k Z
X −1
+ ∂ a1 , ∂a j c a2 · · · abj · · · ak .
j=2 Σ g
−1
When a1 is in the complementary space to the harmonic forms, we know ∂∂ a1 =
a1 so we have the relation
k
X −1
0 = ba1 · · · ak c + b[∂ a1 , a j ]a2 · · · abj · · · ak c
j=2
k Z
X −1
+ ∂ a1 , ∂a j bc a2 · · · abj · · · ak c,
j=2 Σ g
which allows us to iteratively reduce the “polynomial” degree of the original term
ba1 · · · ak c (from k to k − 1 or less, in this case).
This relation looks somewhat complicated but it is easy to understand for k
small. For instance, in k = 1, we see that
0 = ba1 c
whenever a1 is in the space orthogonal to the harmonic forms. We get
Z
−1 −1
0 = ba1 a2 c + b[∂ a1 , a2 ]c + ∂ a1 , ∂a2 bcc.
Σ g
for k = 2.
1.4. The free βγ system. In Section 3 we studied the local structure of the
free βγ theory; in other words, we carefully examined the simplest structure maps
for the factorization algebra Obsq of quantum observables on the plane C. It is
natural to ask about the global sections on a closed Riemann surface.
There are many ways, however, to extend this theory to a Riemann surface Σ.
Let L be a holomorphic line bundle on Σ. Then the L-twisted free βγ system has
fields
E = Ω0,∗ (Σ, L) ⊕ Ω1,∗ (Σ, L∨ ),
where L∨ denotes the dual line bundle. The −1-symplectic pairing on fields is to
apply pointwise the evaluation pairing between L and L∨ and then to integrate
the resulting density. The differential is the ∂ operator for these holomorphic line
bundles.
q
Let ObsL denote the factorization algebra of quantum observables for the L-
twisted βγ system. Locally on Σ, we know how to understand the structure maps:
pick a trivialization of L and employ our work from Section 3.
206 8. FACTORIZATION ALGEBRAS: EXAMPLES
q
1.4.1 Proposition. Let bk = dim H k (Σ, L). Then H ∗ ObsL (Σ) is a rank-one free
module over C[~] and concentrated in degree −b0 − b1 .
Proof. As usual, we use the spectral sequence arising from the filtration F k =
Sym≤k on Obsq . The first page just depends on the cohomology with respect to ∂,
so the underlying graded vector space is
Sym H ∗ (Σ, L)[1] ⊕ H ∗ (Σ, L∨ ⊗ Ω1hol )[1] [~],
relevant literature, but we will not give a complete proof. The interested reader is
encouraged to consult Costello (n.d.c) for more.
2.0.1. Overview of this section. Before we can explain the argument, we need
to formulate the statement. Thus, we begin by explaining how every En algebra —
q
and in particular ObsCS , which is an E3 algebra — has an associated ∞-category of
modules with a natural En−1 -monoidal structure. This terminology means that in
this category of modules, there is a space of ways to tensor together modules, and
this space is homotopy equivalent to the configuration space of points in Rn−1 . In
q
our situation, n = 3 so the ∞-category of modules for ObsCS will be E2 -monoidal,
which is the ∞-categorical version of being braided monoidal.
With these ideas about modules in hand, we can give a precise version of our
claim and provide a strategy of proof. In particular, we explain how these modules
q
for ObsCS are equivalent to the category of representations of a quantum group.
We finish this section by connecting the concept of module for an E3 algebra
to that of a line operator defined in a more physical sense. This will justify our
q
interpretation of the E2 -monoidal category of left modules of ObsCS as being the
category of line operators for Abelian Chern-Simons theory.
Remark: The kinds of assertions and constructions we want to make here rely on
methods and terminology from higher category theory and homotopical algebra.
Formalizing our ideas here would be a nontrivial endeavor, and orthogonal to our
primary goals in this book. Hence we will use the language of higher categories
here quite loosely, although our expressions can be made precise. In particular, in
this section only, we will sometimes simply use “category” in place of ∞-category,
except where the distinction becomes important. We will also use terminology like
“functor” as if working with ordinary categories. ^
2.1. Modules for En algebras. Even in classical algebra, there are different
notions of module for an associative algebra: left modules, right modules, and
bimodules. If we view associative algebras as 1-dimensional in nature, it makes
us suspect that for an n-dimensional algebra — such as an En algebra — there
might be a proliferation of distinct notions reasonably called “modules,” and that
suspicion is true. (See Ayala et al. (n.d.a) for examples.) Nevertheless, in the
context here, we will only need the notion of left (equivalently, right) modules.
The reason is that every En algebra A has an “underlying” E1 algebra A, which
is unique up to equivalence (we explain this “forgetful functor” momentarily).
Thus we can associate to A its category of left modules, which we will denote
LModA to emphasize that we mean left modules (and not bimodules). As we will
explain, this category LModA is naturally equipped with an En−1 -monoidal struc-
ture.
More generally, every En algebra has an underlying Ek algebra for all 0 ≤ k ≤
n. This fact can be seen in two distinct ways: via the operadic definition or via the
factorization algebra picture. They are both geometric in nature.
We start with the operadic approach. Fix a linear inclusion of ι : Rk ,→ Rn . We
will use ι : (x1 , . . . , xk ) 7→ (x1 , . . . , xk , 0, . . . , 0). Then if Dkr (p) denotes the k-disc
of radius r in Rk centered at p, then Dnr (ι(p)) ∩ ι(Rk ) = ι(Dkr (p)), where Dnr (ι(p))
208 8. FACTORIZATION ALGEBRAS: EXAMPLES
where the limit is taken over those opens V in X which contain U. The factorization
product on F defines one on i∗X F .
If F is locally constant, this construction is particularly nice: in that case, the
limit is eventually constant so that i∗X F (U) is F (V) for some sufficiently small open
V containing U.
If R ⊂ Rn is a line, we can restrict a locally constant factorization algebra on
n
R to one on R in this way, to produce an associative algebra. This restriction
procedure is how one describes the E1 algebra underlying an Ek algebra in the
language of factorization algebras.
2.1.1. Modules via geometry. We now want to examine the type of monoidal
structure on LModA arising from an En algebra structure on A. Again, one can use
results about operads and categories or one can use factorization algebras. Here we
begin with the factorization approach.
Let
Hn = {(x1 , . . . , xn ) ∈ Rn | xn ≥ 0}
denote the “upper half space.” Let ∂Hn denote the wall {xn = 0}. By “disc in H” we
will mean the intersection D ∩ Hn for any open disc D ⊂ Rn . Since we are always
working in Hn for the moment, we will use an open U to denote its intersection
with Hn .
Let A be a locally constant factorization algebra on Rn associated to the En
algebra A. Given M ∈ LModA and a point p ∈ ∂Hn living on the wall {xn = 0},
we construct a factorization algebra M p on Hn by extending from the following
factorization algebra on a basis. For an open U not containing p, we set M p (U) =
A(U), and for any disc D in Hn containing p, we set M p (D) = M. For any disjoint
union U t D with p ∈ D, we set M p (U t D) = A(U) ⊗ M. The structure maps are
easy to specify. If U t D ⊂ D0 , then U ⊂ D0 \ D, and on this hemispherical shell,
A(D0 \ D) ' A. Since we know how A acts on M, we have the composition
U2
A⊗A⊗M
U1
M
V2 V1
p
V3 A ⊗ M p ⊗ Mq
U
M p ⊗A Mq
V1 V2
p q
From such an M ~p , we can recover a left A-module in a very simple way. Con-
sider the projection map π : Hn → R≥0 sending (x1 , . . . , xn ) to xn . Then π∗ M ~p
returns A on any interval not containing the boundary point 0, but it has interesting
value on an interval containing 0.
In this way, we produce a left A-module for any configuration of distinct
points {p1 , . . . , pk } in ∂Hn Rn−1 and any corresponding k-tuple of left A-modules
{M1 , . . . , Mk }. In other words, we have sketched a process that can lead to making
LModA an En−1 algebra in some category of categories.
2.1.2. Modules via operads. Developing the picture sketched above into a for-
mal theorem would be somewhat non-trivial. Fortunately, Lurie has proved the
theorem we need in Lurie (n.d.b), using the operadic approach to thinking about
En algebras. We begin by stating the relevant result from his work and then gloss
the essential argument (although we hide a huge amount of technical machinery
developed to make such an argument possible).
We need some notation to state the result. If C is a symmetric monoidal ∞-
category, then C is itself an E∞ algebra in the ∞-category Cat∞ of ∞-categories,
equipped with the Cartesian symmetric monoidal structure (i.e., the product of cat-
egories). Hence, it makes sense to talk about left module categories over C, which
we denote LModC (Cat∞ ). Let Cat∞ (∆) denote the subcategory of ∞-categories
possessing geometric realizations and whose functors preserve geometric realiza-
tions.
2.1.1 Theorem (Theorem 4.8.5.5 and Corollary 5.1.2.6, Lurie (n.d.b)). Let C
denote a symmetric monoidal ∞-category possessing geometric realizations and
whose tensor product preserves colimits separately in each variable. Then there is
a fully faithful symmetric monoidal functor
LMod− : AlgE1 (C) → LModC (Cat∞ (∆))
sending an E1 algebra A to LModA . This functor determines a fully faithful functor
AlgEn (C) → AlgEn−1 (LModC (Cat∞ (∆))).
This result might seem quite technical but it is a natural generalization of more
familiar constructions from ordinary algebra. One might interpret the first part as
saying that
((i)) there is a big category whose objects are categories with an action of
the symmetric monoidal category C,
((ii)) the category of left modules over an algebra in C gives an object in this
big category, and
((iii)) there is a way of “tensoring” categories of modules such that
LModA LModB ' LModA⊗B .
Tensoring categories might not be familiar, but it is suggested by generalizing the
construction of the tensor product of ordinary vector spaces: given a product of
categories D × D0 , we might hope there is a category D D0 receiving a bifunctor
from D × D0 such that all other bifunctors from D × D0 factor through it. Note that
we have a natural bifunctor LModA × LModB → LModA⊗B .
2. ABELIAN CHERN-SIMONS THEORY AND QUANTUM GROUPS 211
To understand the second part of the theorem, it helps to know about Dunn
additivity: it is equivalent to give an object A ∈ C the structure of an En algebra as
it is to give A first the structure of an E1 algebra and then the structure of an En−1
algebra in the category of E1 algebras. In other words, there is an equivalence of
categories
AlgEn−1 (AlgE1 (C)) ' AlgEn (C).
(See Theorem 5.1.2.2 of Lurie (n.d.b).) By induction, to give A an En algebra is to
specify n compatible E1 structures on A.
Geometrically, this assertion should seem plausible. If one has an En structure
on A, then each way of stacking discs along a coordinate axes in Rn provides an
E1 multiplications. Hence, we obtain n different E1 multiplications on A. In the
other direction, knowing how to multiply along these axes, one should be able to
reconstruct a full En structure.
The theorem then says that we get a natural functor
LMod
AlgEn (C) ' AlgEn−1 (AlgE1 (C)) −−−−→ AlgEn−1 (LModC (Cat∞ (∆))),
where the first equivalence is Dunn additivity and the second functor exists because
LMod− is symmetric monoidal. This composed arrow is what allows one to turn
an En algebra into an En−1 monoidal category.
2.2. The main statement and the strategy of proof. Let ACS denote the E3
q
algebra associated to the locally constant factorization algebra ObsCS of quantum
observables for Abelian Chern-Simons theory with the rank 1 Abelian Lie algebra.
Since we will do everything over the field of complex numbers, we will identify
this Lie algebra with C. By the theory described above, the category LModACS is
an E2 -monoidal category. Recall that E2 -monoidal categories are the ∞-categorical
version of braided monoidal categories. There is another braided monoidal cate-
gory intimately connected with Chern-Simons theory: representations of the quan-
tum group. We want to relate these two categories.
In the case of interest, the quantum group is a quantization of the universal
enveloping algebra of the Abelian Lie algebra C. Typically, a quantum group de-
pends on a parameter often denoted q or ~. The size of this parameter is related to
the choice of an invariant pairing on the Lie algebra. In the Abelian case, multi-
plication by scalars is a symmetry of the Lie algebra that also scales the invariant
pairing. Thus, all values of the quantum parameter are equivalent, and we find only
a single quantum group.
It will be convenient to consider a completed version of the universal envelop-
ing algebra. Thus, our Abelian quantum group will be U(C) b = C[[t]]. The Hopf
algebra structure will be that on the universal enveloping algebra. Thus, it is com-
mutative and co-commutative and the variable t is primitive. The only interesting
structure we find is the R-matrix, given by the formula
R = et⊗t ∈ U(C)
b b ⊗U(C).
b
More formally, since V is a module for the Lie algebra C, we can take the Chevalley-
Eilenberg cochains C ∗ (C, V) of C with coefficients in V. This complex is a dg mod-
ule for C ∗ (C) = C[ε]. Let LMod0 (C[ε]) be the full subcategory of the ∞-category
of C[ε]-modules given by the essential image of the functor Φ.
2. ABELIAN CHERN-SIMONS THEORY AND QUANTUM GROUPS 213
2.2.4 Lemma. The equivalence of the previous lemma transfers the braided monoidal
category structure on Ho(LMod0 (ACS )) to a braided monoidal category on Rep(U(C)).
b
This braided monoidal category structure is realized by some R-matrix
R ∈ U(C)
b b ⊗U(C),
b
giving U(C)
b the structure of quasi-triangular Hopf algebra.
Proof. Let us sketch how one can see this using the Tannakian formalism,
before giving a more concrete approach to constructing the quasi-triangular Hopf
algebra structure.
The category Ho(LMod0 (ACS ) is a braided monoidal category. There is a
functor
Aug : Ho(LMod0 (ACS )) → Vect
sending a module M to M ⊗C[ε] C, using the equivalence of E2 algebras between
ACS and C[ε]. The augmentation map C[ε] → C is a map of E2 algebras, so this
augmentation functor is monoidal. Under the equivalence
Rep(U(C))
b ' Ho(LMod0 (ACS ))
the functor Aug becomes the forgetful functor, which takes a representation of
U(C)
b and forgets the U(C)
b action.
Let us recall how the Tannakian formalism works for braided monoidal cat-
egories. (See, for example, Etingof et al. (2015).) Suppose we have a braided
monoidal category C with a monoidal (but not necessarily braided monoidal) func-
tor F : C → Vect. Then, by considering the automorphisms of the fibre functor
F, one can construct a quasi-triangular Hopf algebra whose representations will be
our original braided monoidal category C, and where the functor F is the forget-
ful functor. (One needs some additional technical hypothesis on C and F for this
construction to work.)
We are in precisely the situation where this formalism applies. The braided
monoidal category is Ho(LMod0 (ACS ), and the fibre monoidal functor is the func-
tor Aug given by the augmentation of ACS as an E2 algebra. Since we already
know that Ho(LMod0 (ACS )) is equivalent to representations of the Hopf algebra
U(C),
b in such a way that the fibre functor becomes the forgetful functor, the Hopf
algebra produced by the Tannakian story is U(C).
b We therefore find that the braid-
ing is described by some R-matrix on this Hopf algebra.
Since this approach is so abstract, let us describe how to see concretely the
R-matrix on U(C).
b
The equivalence with Ho(LMod0 (ACS )) gives us a braided monoidal struc-
ture on representations of U(C).
b For any two representations V, W of U(C),
b this
braiding gives an isomorphism V ⊗ W W ⊗ V. There is already such an iso-
morphism, just of vector spaces, since vector spaces are a symmetric monoidal
category. Composing these two isomorphisms gives, for all representations V, W
of U(C),
b an isomorphism RV,W : V ⊗ W − → V ⊗ W. It is an isomorphism of vector
spaces and not necessarily of U(C)-modules,
b and it is natural in V and W.
2. ABELIAN CHERN-SIMONS THEORY AND QUANTUM GROUPS 215
The functor
Rep(U(C))
b × Rep(U(C))
b → Vect
(V, W) 7→ V ⊗ W
b b⊗2 module U(C)
is represented by the right U(C) b b ⊗U(C),
b i.e., the functor is given by
tensoring with this right module. The natural automorphism RV,W of this functor is
thus represented by an automorphism of this right module, which must be given by
left multiplication with some element
R ∈ U(C)
b b ⊗U(C).
b
This element is the universal R-matrix.
Now, the axioms of a braided monoidal category translate directly into the
properties that R must satisfy to define a quasi-triangular Hopf algebra structure on
U(C).
b
So far, we have seen that we can identify the braided monoidal category of rep-
resentations of U(C)
b with the braided monoidal category Ho LMod0 (ACS ), where
we equip U(C)
b with some R-matrix. It remains to show that there is essentially
only one possible R-matrix.
2.2.5 Proposition. Every R-matrix for U(C)
b satisfying the axioms defining a quasi-
triangular Hopf algebra is of the form
R = exp(c(t ⊗ t)) ∈ U(C)
b b ⊗U(C)
b
where c is a complex number.
Proof. Recall that an R-matrix of a Hopf algebra A is an invertible element R
of A ⊗ A such that
∆op (x) = R∆(x)R−1 for all x ∈ A,
∆ ⊗ idA (R) = R13 R23 ,
idA ⊗∆(R) = R13 R12 ,
where ∆ denotes the coproduct on A, ∆op denotes ∆ postcomposed with the flip
operation, and R13 , for example, denotes
X
ai ⊗ 1 ⊗ bi ∈ A ⊗ A ⊗ A
i
where R = i ai ⊗ bi ∈ A ⊗ A.
P
For us, due to the completion of the tensor product, an element R will live
⊗C[[t2 ]]. As C[[t]] is cocommutative, we see that the first equation will
in C[[t1 ]]b
be trivially satisfied by any invertible element. The only conditions are then the
remaining equations. As we are working with a deformation of the usual symmetric
braiding on C[[t]]-modules, an R-matrix is of the form
1 ⊗ 1 + c1 t1 ⊗ 1 + c2 1 ⊗ t2 + · · · .
216 8. FACTORIZATION ALGEBRAS: EXAMPLES
Writing R = i x(i) ⊗ y(i) with x(i) ∈ C[[t1 ]] and y(i) ∈ C[[t2 ]], we see that the first
P
equation becomes
X X
∆(x(i) ) ⊗ y(i) = (x(i) ⊗ 1 + 1 ⊗ x(i) ) ⊗ y(i) ,
i i
which forces each x(i) to be primitive. The other equation likewise forces each y(i)
to be primitive. Hence R = ct1 ⊗ t2 for some constant c and so R = exp(ct1 ⊗ t2 ).
The constant c is the only freedom available in choosing an R-matrix.
Thus, we have shown the desired equivalence of braided monoidal categories,
where U(C)
b is equipped with an R-matrix of the form exp(c(t ⊗ t)) for some value
of the constant c. This constant cannot be zero, as this would force ACS to be a
commutative E3 algebra, and ACS is not commutative. Any two non-zero values
of the constant c are related by scaling t, and so all the braided monoidal categories
are equivalent. This completes the proof of Theorem 2.2.2.
q
so we can pushforward ObsCS ,K from this neighborhood to S 1 . The underlying
E1 algebra AK should look something like ACS tensored with a Clifford algebra
(the observables for the uncoupled free fermion). Thus the line defect for Abelian
q q
Chern-Simons theory produces an interesting modification ObsCS ,K of ObsCS near
K; we will view this factorization algebra as a model for what a line defect should
be in the setting of factorization algebras.
By picking different field theories supported on K and coupling them to Chern-
Simons theory in the bulk, we obtain other factorization algebras. One can do this,
for instance, by varying the coupling constants in the example above or adding
more interactions terms. One can similarly imagine picking a 2-dimensional sub-
manifold and coupling a field theory on it to the 3-dimensional bulk.
2.3.2. Defects via factorization algebras. Let us now explain in the language
of factorization algebras how to couple a field theory to a quantum-mechanical
system living on a line. Suppose we have a factorization algebra F on Rn , which we
assume here is locally constant. Let us fix a copy of R ⊂ Rn . Let us place a locally-
constant factorization algebra G on R, which we think of as the observables of a
quantum-mechanical system. We will view G as a homotopy-associative algebra or
E1 algebra or A∞ algebra, but we will tend to write formulas as if everything were
strictly associative. (One can always replace an A∞ algebra by a weakly equivalent
but strictly associative dg algebra.)
Let i∗ G denote the pushforward of G to a factorization algebra on Rn . Note
that i∗ G takes value C on any open that does not intersect R, and it takes value
G(i−1 (U)) on any open which does intersect R.
Before we define a coupled factorization algebra, we need to understand the
situation where the field theory on Rn and the field theory on R are uncoupled. In
this case, the factorization algebra of observables for the system is simply F ⊗ i∗ G.
We are interested in deforming this uncoupled situation.
2.3.1 Definition. A coupling of our factorization algebra F to the factorization
algebra G living on R is an element α ∈ F |R ⊗ G of cohomological degree 1
satisfying the Maurer-Cartan equation
dα + 12 [α, α] = dα + α ? α = 0,
where ? is the product in F |R ⊗ G viewed as an associative algebra.
dα (x) = dx + [α, x] = dx + α ? x + x ? α.
π : Rn \ R → R × R>0
(x, v) 7→ (x, kvk).
Next, let us explain how to interpret our concept of “coupling” from the point
of view of physics. To do this, we need to say a little bit about local operators (that
is, operators supported at a point) from the point of view of factorization algebras.
This topic will be treated in more detail in volume 2. Here, we will be brief.
In any field theory on Rn whose observables are given by a factorization algebra
F , there is a co-stalk F x of F for each point x ∈ Rn . This co-stalk is the limit
lim x∈U F (U) of the spaces of observables on neighbourhoods U of x. The co-stalk
F x is a way to define local observables at x for the field theory: by definition an
element of F x is an observable contained in F (U) for every neighbourhood U of
x. For a general field theory, this limit might produce something hard to describe
(and one might also worry about the difference between the limit and the homotopy
limit). For a topological field theory, however, the limit is essentially constant, so
that we can identify F x with F (D) for any disc D around x.
As x varies in Rn , the space F x forms a vector bundle, typically of infinite rank.
This vector bundle has a flat connection, reflecting the fact that we can differentiate
local operators. Because F x may be infinite-dimensional, one can not in general
solve the parallel transport equation from the flat connection. Thus, for a general
factorization algebra, F x should be thought of as a D-module. For a topological
field theory, however, the bundle F x is a local system, meaning one can solve the
parallel transport equation. The equivalence F x ' Fy given by parallel transport
' '
arises by the quasi-isomorphisms F x −→ F (D) ←− Fy for any disc D containing
both x and y.
Let us consider, as before, a line R ⊂ Rn , with a field theory on Rn whose
observables are described by a factorization algebra F and with a field theory on
R whose observables are described by a factorization algebra G. Let us assume
that both factorization algebras are locally constant. For x ∈ R, we can consider
the space F x ⊗ G x , which is the tensor product of local operators of our two field
theories at x ∈ R ⊂ Rn . As we have seen, the space F x ⊗ G x forms, as x varies,
a vector bundle with flat connection on R. Let us denote this vector bundle by
Floc ⊗ Gloc . Since Floc ⊗ Gloc is a local system, there is a quasi-isomorphism
F |R (D) ⊗ G(D) ' Ω∗ (R, Floc ⊗ Gloc )
for any disc D ⊂ R.
From a physics point of view, we might expect that to couple the system spec-
ified by F to that specified by G, we need to give a Lagrangian density on the real
line R. In this case, such a Lagrangian would be a one-form on R with values in
the bundle of local observables Floc ⊗ Gloc , viewed as observables on the combined
fields for the two field theories.
Such a Lagrangian density does appear naturally in our earlier mathematical
definition. Under the isomorphism
F |R (D) ⊗ G(D) ' Ω∗ (R, Floc ⊗ Gloc ),
a Maurer-Cartan element α on the left hand side corresponds to a sum
S α = S α(0) + S α(1) ,
2. ABELIAN CHERN-SIMONS THEORY AND QUANTUM GROUPS 221
where S α(0) is a function along R with values in the degree 1 component of the
vector bundle of local observables and S α(1) is a 1-form with values in the degree 0
component of the local observables. The term S α(1) is thus a Lagrangian density. To
better understand what it means, we need to go a bit further.
Under the correspondence above, the Maurer-Cartan equation for α becomes
the equation
h i
ddR S α(0) + dF ⊗G S α(0) + dF ⊗G S α(1) + S α(0) , S α(1) = 0,
where dF ⊗G is the differential on Floc ⊗ Gloc . This equation is essentially a version
of the master equation in the Batalin-Vilkovisky formalism, but this relationship is
manifest only with factorization algebras produced by BV quantization, which is
the subject of Volume 2. Solving the equation perturbatively, however, we can get
the flavor of the relationship.
Consider the situation where we try to solve for S α iteratively in some formal
variable c: (0) (0)
(1) (1)
S α = c S α,1 + S α,1 + c2 S α,2 + S α,2 + ··· .
At the first stage, we work modulo c2 . We then want to solve the equation
(0) (1)
ddR S α,1 + dF ⊗G S α,1 = 0.
(Note that for grading reasons, we always need dF ⊗G S α(0) = 0.) And we are in-
terested in solutions modulo the image of the operator ddR + dF ⊗G on elements in
Ω∗ (R, Floc ⊗Gloc ) of cohomological degree 0; this space is the first-order version of
gauge equivalence between solutions of the Maurer-Cartan equation. Unraveling
these conditions, we see that we care about a Lagrangian density in Floc ⊗ Gloc up
to a total derivative. This data is what one expects from physics. Continuing onto
higher powers of c, one finds that our algebraic definition for coupling the quan-
tum mechanical system G to the field theory F matches with what one expects
from physics.
In sum, for any topological field theory whose observables are given by a fac-
torization algebra F on Rn , this discussion indicates that the category of projec-
tive left modules for F captures the category of Wilson line operators for the n-
dimensional field theory described by F . These categories both admit interesting
ways to “combine” or “fuse” objects: the left F -modules have an En−1 -monoidal
structure, by Lurie’s work, and the line defects can be fused, equipping them with
an En−1 -monoidal structure as well. (For a physically-oriented introduction to the
relationship between defects and higher categories, see Kapustin (2010).) We as-
sert that these agree: when one examines the physical constructions, they behave as
the homotopical algebra suggests. For some discussion of these issues, see Costello
(2014).
2.3.4. The take-away message. Let us apply this discussion to Abelian Chern-
Simons theory, where we have an E3 algebra ACS . The left modules that we find
in our discussion above — projective modules of rank n — match exactly with
the category LMod0 (ACS ) we introduced earlier. If we define the E2 -monoidal
category of line operators of Abelian Chern-Simons theory to be the E2 -monoidal
222 8. FACTORIZATION ALGEBRAS: EXAMPLES
category LMod0 (ACS ), we find that our theorem relating Abelian Chern-Simons
theory to the quantum group takes the following satisfying form.
Theorem. The following are equivalent as braided monoidal categories:
((i)) the homotopy category of the category of Wilson line operators of
Abelian Chern-Simons theory
((ii)) the category of representations of the quantum group for the Abelian
Lie algebra C.
In this case, the identification between braided monoidal structures can be seen
explicitly. We have already shown that the factorization algebra encodes the Wil-
son loop observables and recovers the Gauss linking number, but it is these rela-
tions that are used in providing the braiding on the line operators. (See, e.g., the
discussion in Kapustin and Saulina (2011).)
2.4. Proof of a technical proposition. Let us now give the proof of the fol-
lowing proposition, whose proof we skipped earlier.
2.4.1 Proposition. View ACS as an E2 algebra via the map of operads from E2 to
E3 . There is a canonical equivalence of E2 algebras from ACS to the commutative
algebra C[ε], where the parameter ε is of degree 1. We view C[ε] as an E2 algebra
via the map of operads from E2 to E∞ .
Proof. Recall that ACS is a twisted factorization envelope. Explicitly, it as-
signs to an open subset U ⊂ R3 , the cochain complex
C∗ (Ω∗c (U)[1] ⊕ C · ~[−1]) ⊗C[~] C~=1 .
The Lie bracket on the dg Lie algebra Ω∗c (U)[1] ⊕ C · ~[−1] is given by the shifted
central extension of the Abelian Lie algebra Ω∗c (U)[1] for the cocycle
Z
α ⊗ β 7→ α ∧ β.
where the limit is taken over those opens V in R3 that contain our chosen open
U ⊂ R2 . Note that since ACS is locally constant, this limit is eventually constant.
If we identify R3 = R2 × R, we can identify
i∗R2 ACS (U) = ACS (U × R).
In other words, we can identify i∗R2 ACS with the push-forward π∗ ACS where π :
R3 → R2 is the projection. This identification is as factorization algebras.
There is a quasi-isomorphism, for all U ⊂ R2 ,
Ω∗c (U)[−1] ' Ω∗c (U × R).
2. ABELIAN CHERN-SIMONS THEORY AND QUANTUM GROUPS 223
We can restrict this cocycle to one on the sub-cosheaf Ω∗c (U), where we use the
map Ω∗c (U) → Ω∗c (U × R)[1] discussed above. We find that it is zero, since for
ω1 , ω2 ∈ Ω∗c (U)[1], we have
Z
ω1 f (t)dt ∧ ω2 f (t)dt = 0.
U×R
Remark: There is another approach to this lemma that is less direct but applicable
more broadly, by using the machinery of deformation theory.
There is a Hochschild complex for En algebras (in cochain complexes over a
field of characteristic zero) akin to the Hochschild cochain complex for associative
algebras. Furthermore, given an En algebra A, there is an exact sequence of the
form
Def(A) → HHE∗ n (A) → A
relating the En Hochschild complex of A and the complex describing En algebra
deformations of A (see Francis (2013)). First-order deformations of the En algebra
structure are given by the n + 1st cohomology group of the complex Def(A). The
funny shift is due to the fact that the En deformation complex is an En+1 algebra
and the underlying dg Lie algebra is obtained by shifting to turn the Poisson bracket
into an unshifted Lie bracket. Compare, for instance, the fact that HH 2 (A) classifies
first-order deformations of associative algebras.
Moreover, there is an analog of the Hochschild-Kostant-Rosenberg theorem
for En algebras: for A a commutative dg algebra A, it identifies the En -deformation
complex for A with the shifted polyvector fields SymA (TA [−n]). Here TA denotes
the tangent complex of A. (See Calaque and Willwacher (2015) for the full state-
ment and a proof.)
For us, the classical observables of Abelian Chern-Simons are equivalent to
the commutative dg algebra C[ε], where ε has cohomological degree 1, because it
is the factorization envelope of an Abelian Lie algebra. The E2 HKR theorem then
tells us that
HHE∗ 2 (C[ε]) ' C [ε, ∂/∂ε] ,
where ε and ∂/∂ε have cohomological degree 1. There are no degree 3 elements
of our deformation complex, so the E2 deformation must be trivial. Hence ACS
is a trivial deformation as an E2 algebra. Note, by contrast, that there is a one-
dimensional space of first-order deformations of C[ε] as an E3 algebra because
HHE∗ 3 (C[ε]) is C [ε, ∂/∂ε], with ∂/∂ε now in degree 2. ^
APPENDIX A
Background
along with pointers to more thorough treatments. By no means does the reader
need to be expert in all these areas to use our results or follow our arguments. She
just needs a working knowledge of this background machinery, and this appendix
aims to provide the basic definitions, to state the results relevant for us, and to
explain the essential intuition.
We do assume that the reader is familiar with basic homological algebra and ba-
sic category theory. For homological algebra, there are numerous excellent sources,
in books and online, among which we recommend the complementary texts by
Weibel (1994) and Gelfand and Manin (2003). For category theory, the standard
reference Mac Lane (1998) is adequate for our needs; we also recommend the
series Borceux (1994a).
Remark: Our references are not meant to be complete, and we apologize in advance
for the omission of many important works. We simply point out sources that we
found pedagogically oriented or particularly accessible. ^
0.1. Reminders and Notation. We overview some terminology and notations
before embarking on our expositions.
For C a category, we often use x ∈ C to indicate that x is an object of C. We
typically write C(x, y) to the denote the set of morphisms between the objects x
and y, although occasionally we use HomC (x, y). The opposite category Cop has
the same objects but Cop (x, y) = C(y, x).
Given a collection of morphisms S in C, a localization of C with respect to
S is a category C[S −1 ] and a functor q : C → C[S −1 ] satisfying the following
conditions.
((ii)) For any functor F : C → D such that F(s) is an isomorphism for every
s ∈ S , there is a functor F : C[S −1 ] → C and a natural isomorphism
τ : F ◦ F ⇒ F.
((iii)) For every category D, the functor q∗ : Fun(C[S −1 ], D) → Fun(C, D)
sending G to G ◦ q is full and faithful.
These ensure that the localization is unique up to equivalence of categories, if it
exists. (We will not discuss size issues in this book.) The localization is the cate-
gory in which every s ∈ S becomes invertible. A morphism f : C[S −1 ](x, y) can
be concretely understood as a “zig zag”
' '
x ← z1 → z2 ← · · · → y
where the arrows going to the left (the “wrong way”) are all in S and the arrows
going to the right are arbitrary morphisms from C.
We typically apply localization in the setting of a category C with a class W
of weak equivalences. We call W a class of weak equivalences if W contains all
isomorphisms in C and satisfies the 2-out-of-3 property, which states that given
morphisms f ∈ C(x, y) and g ∈ C(y, z) such that two morphisms in the set { f, g, g ◦
f } are in W, then all three are in W. As examples, consider
((i)) any category with isomorphisms as weak equivalences,
((ii)) the category of cochain complexes with cochain homotopy equiva-
lences as weak equivalences,
((iii)) the category of cochain complexes with quasi-isomorphisms as weak
equivalences, or
((iv)) the category of topological spaces with weak homotopy equivalences
as weak equivalences.
We often denote the localization C[W −1 ] by Ho(C, W), or just Ho(C), and call it
the “homotopy category.”
1. Simplicial techniques
Simplicial sets are a combinatorial substitute for topological spaces, so it should
be no surprise that they can be quite useful. On the one hand, we can borrow intu-
ition for them from algebraic topology; on the other, simplicial sets are extremely
concrete to work with because of their combinatorial nature. In fact, many con-
structions in homological algebra are best understood via their simplicial origins.
Analogs of simplicial sets (i.e., simplicial objects in other categories) are useful as
well.
In this book, we use simplicial sets in two ways:
• when we want to talk about a family of QFTs or space of parametrices,
we will usually construct a simplicial set of such objects instead; and
• we accomplish some homological constructions by passing through
simplicial sets (e.g., in constructing the extension of a factorization
algebra from a basis).
After giving the essential definitions, we state the main theorems we use.
1. SIMPLICIAL TECHNIQUES 227
1.0.1 Definition. Let ∆ denote the category whose objects are totally ordered finite
sets and whose morphisms are non-decreasing maps between ordered sets. We
usually work with the skeletal subcategory whose objects are
[n] = {0 < 1 < · · · < n}.
A morphism f : [m] → [n] then satisfies f (i) ≤ f ( j) if i < j.
We will relate these objects to geometry below, but it helps to bear in mind the
following picture. The set [n] corresponds to the n-simplex 4n equipped with an
ordering of its vertices, as follows. View the n-simplex 4n as living in Rn+1 as the
set
X
, , . . . , = .
(x x x ) x 1 and 0 ≤ x ≤ 1 ∀ j
0 1 n j j
j
Identify the element 0 ∈ [n] with the 0th basis vector e0 = (1, 0, . . . , 0), 1 ∈ [n]
with e1 = (0, 1, 0, . . . , 0), and k ∈ [n] with the kth basis vector ek . The ordering
on [n] then prescribes a path along the edges of 4n , starting at e0 , then going to e1 ,
and continuing till the path ends at en .
Every map f : [m] → [n] induces a linear map f∗ : Rm+1 → Rn+1 by setting
f∗ (ek ) = e f (k) . This linear map induces a map of simplices f∗ : 4m → 4n .
There are particularly simple maps that play an important role throughout the
subject. Note that every map f factors into a surjection followed by an injection.
We can then factor every injection into a sequence of coface maps, namely maps
of the form
fk : [n] → ( [n + 1]
i, i≤k .
i 7→
i + 1, i > k
Similarly, we can factor every surjection into a sequence of codegeneracy maps,
namely maps of the form
dk : [n] → ( [n − 1]
i, i≤k .
i 7→
i − 1, i > k
The names face and degeneracy fit nicely with the picture of the geometric sim-
plices: a coface map corresponds to a choice of n-simplex in the boundary of the
n + 1-simplex, and a codegeneracy map corresponds to “collapsing” an edge of the
n-simplex to project the n-simplex onto an n − 1-simplex.
We now introduce the main character.
1.0.2 Definition. A simplicial set is a functor X : ∆op → S et, often denoted X• .
The set X([n]), often denoted Xn , is called the “set of n-simplices of X.” A map of
simplicial sets F : X → Y is a natural transformation of functors. We denote this
category of simplicial sets by sSet.
Let’s quickly examine what the coface maps tell us about a simplicial set X.
For example, by definition, a map fk : [n] → [n+1] in ∆ goes to X( fk ) : Xn+1 → Xn .
228 A. BACKGROUND
We interpret this map X( fk ) as describing the kth n-simplex sitting as a “face on the
boundary” of an n + 1-simplex of X. A similar interpretation applies to the dk .
must be taken, since there are topological spaces very different in nature from sim-
plicial or cell complexes and for which no simplicial set could provide an accurate
description. The key is only to think about topological spaces and simplicial sets
up to the appropriate notion of equivalence.
Remark: It is more satisfactory to define these homotopy notions directly in terms
of simplicial sets and then to verify that these match up with the topological no-
tions. We direct the reader to the references for this story, as the details are not
relevant to our work in the book. ^
We say a continuous map f : X → Y of topological spaces is a weak equiva-
lence if it induces a bijection between connected components and an isomorphism
πn ( f, x) : πn (X, x) → πn (Y, f (x)) for every n > 0 and every x ∈ X. Let Ho(Top),
the homotopy category of Top, denote the category of topological spaces where
we localize at the weak equivalences. There is a concrete way to think about
this homotopy category. For every topological space, there is some CW complex
weakly equivalent to it, under a zig zag of weak equivalences; and by the White-
head theorem, a weak equivalence between CW complexes is in fact a homotopy
equivalence. Thus, Ho(Top) is equivalent to the category of CW complexes with
morphisms given by continuous maps modulo homotopy equivalence.
Likewise, let Ho(sSet) denote the homotopy category of simplicial sets, where
we localize at the appropriate notion of simplicial homotopy. Then Quillen proved
the following wonderful theorem.
1.1.3 Theorem. The adjunction induces an equivalence
| − | : Ho(sSet) Ho(Top) : Sing
between the homotopy categories. (In particular, they provide a Quillen equiva-
lence between the standard model category structures on these categories.)
This theorem justifies the assertion that, from the perspective of homotopy
type, we are free to work with simplicial sets in place of topological spaces. In
addition, it helpful to know that algebraic topologists typically work with a better
behaved categories of topological spaces, such as compactly generated spaces.
Among simplicial sets, those that behave like topological spaces are known as
Kan complexes or fibrant simplicial sets. Their defining property is a simplicial
analogue of the homotopy lifting property, which we now describe explicitly. The
horn for the kth face of the n-simplex, denoted Λk [n], is the subsimplicial set of
∆[n] given by the union of the all the faces ∆[n − 1] ,→ ∆[n] except the kth. (As a
functor on ∆, the horn takes the [m] to monotonic maps [m] → [n] that do not have
k in the image.) A simplicial set X is a Kan complex if for every map of a horn
Λk [n] into X, we can extend the map to the n-simplex ∆[n]. Diagrammatically, we
can fill in the dotted arrow
Λk [n] /
=X
∆[n]
230 A. BACKGROUND
to get a commuting diagram. In general, one can always find a “fibrant replace-
ment” of a simplicial set X• (e.g., by taking Sing |X• | or via Kan’s Ex∞ functor)
that is weakly equivalent and a Kan complex.
1.2. Simplicial sets and homological algebra. Our other use for simplicial
sets relates to homological algebra. We always work with cochain complexes, so
our conventions will differ from those who prefer chain complexes. For instance,
the chain complex computing the homology of a topological space is concentrated
in non-negative degrees. We work instead with the cochain complex concentrated
in non-positive degrees. (To convert, simply swap the sign on the indices.)
1.2.1 Definition. A simplicial abelian group is a simplicial object A• in the cate-
gory of abelian groups, i.e., a functor A : ∆op → Ab.
where the fk run over the coface maps from [|m|−1] to [|m|]. The normalized chains
NA• is the cochain complex
|m|−1
\
(NA• )m = ker A• ( fk ),
k=0
where m ≤ 0 and where the fk run over the coface maps from [|m| − 1] to [|m|]. The
differential is A( f|m| ), the remaining coface map. One can check that the inclusion
NA• ,→ CA• is a quasi-isomorphism (in fact, a cochain homotopy equivalence).
Example: Given a topological space X, its singular chain complex C∗ (X) arises as a
composition of three functors in this simplicial world. (Because we prefer cochain
complexes, the singular chain complex is, in fact, a cochain complex concentrated
in nonpositive degrees.) First, we make the simplicial set Sing X, which knows
about all the ways of mapping a simplex into X. Then we apply the free abelian
group functor Z− : S ets → AbG ps levelwise to obtain the simplicial abelian group
Z Sing X : [n] 7→ Z(Top(4n , X)).
1. SIMPLICIAL TECHNIQUES 231
Then we apply the unnormalized chains functor to obtain the singular chain com-
plex
C∗ (X) = CZ Sing X.
In other words, the simplicial language lets us decompose the usual construction
into its atomic components. ^
It is clear from the constructions that we only ever obtain cochain complexes
concentrated in nonpositive degrees from simplicial abelian groups. In fact, the
Dold-Kan correspondence tells us that we are free to work with either kind of
object — simplicial abelian group or such a cochain complex — as we prefer.
Let Ch≤0 (Ab) denote the category of cochain complexes concentrated in non-
positive degrees, and let sAbG ps denote the category of simplicial abelian groups.
1.2.2 Theorem (Dold-Kan correspondence). The normalized chains functor
N : sAb → Ch≤0 (Ab)
is an equivalence of categories. Under this correspondence,
πn (A• ) H −n (NA• )
for all n ≥ 0, and simplicial homotopies go to chain homotopies.
Throughout the book, we will often work with cochain complexes equipped
with algebraic structures (e.g., commutative dg algebras or dg Lie algebras). Thank-
fully, it is well-understood how the Dold-Kan correspondence intertwines with the
tensor structures on sAb and Ch≤0 (Ab).
Let A• and B• be simplicial abelian groups. Then their tensor product (A ⊗ B)•
is the simplicial abelian group with n-simplices
(A ⊗ B)n = An ⊗Z Bn .
There is a natural transformation
∇A,B : CA ⊗ CB → C(A ⊗ B),
known as the Eilenberg-Zilber map or shuffle map, which relates the usual tensor
product of complexes with the tensor product of simplicial abelian groups. As
CA• ⊗ CB• is not isomorphic to C(A ⊗ B)• , it is not a strong monoidal functor but
instead a lax monoidal functor. The Eilenberg-Zilber map is, however, always a
quasi-isomorphism, and so it preserves weak equivalences.
1.2.3 Theorem. The unnormalized chains functor and the normalized chains func-
tor are both lax monoidal functors via the Eilenberg-Zilber map.
Thus, with a little care, we can relate algebra in the setting of simplicial mod-
ules with algebra in the setting of cochain complexes.
1.3. References. Friedman (2012) is a very accessible and concrete introduc-
tion to simplicial sets, with lots of intuition and pictures. Weibel (1994) explains
clearly how simplicial sets appear in homological algebra, notably for us, the Dold-
Kan correspondence and the Eilenberg-Zilber map. As usual, Gelfand and Manin
232 A. BACKGROUND
(2003) provides a nice complement to Weibel. The expository article Goerss and
Schemmerhorn (2007) provides a lucid and quick discussion of how simplicial
methods relate to model categories and related issues. For the standard, modern
reference on simplicial sets and homotopy theory, see the thorough and clear Go-
erss and Jardine (2009).
But we recognize that there should be more elaborate notions involving many dif-
ferent n-ary operations required to satisfy complicated relations. As a basic exam-
ple, a Poisson algebra has two binary operations.
Before we give the general definition of an operad in (dg) vector spaces, we
explain how to visualize such algebraic structures. An n-ary operation τ : A⊗n → A
is pictured as a rooted tree with n labeled leaves and one root.
a1 a2 ... an
µ(a1 , . . . , an )
For us, operations move down the page. To compose operations, we need to spec-
ify where to insert the output of each operation. We picture this as stacking rooted
trees. For example, given a binary operation µ, the composition µ ◦ (µ ⊗ 1) corre-
sponds to the tree
a1 a2 a3
µ 1
µ(µ(a1 , a2 ), a3 )
1 µ
µ(a1 , µ(a2 , a3 ))
σ⊗σ−1
/
O(n) ⊗ (O(m1 ) ⊗ · · · ⊗ O(mn )) O(n) ⊗ (O(mσ(1) ) ⊗ · · · ⊗ O(mσ(n) ))
◦ ◦
P σ(mσ(1) ,...,mσ(n) ) P
/O
n n
O j=1 m j j=1 m j
id ⊗(τ1 ⊗···⊗τn ))
(
O(n) ⊗ (O(m1 ) ⊗ · · · ⊗ O(mn )) O(n) ⊗ (O(mσ(1) ) ⊗ · · · ⊗ O(mσ(n) ))
◦ ◦
P τ1 ⊕···⊕τn P
/O
n n
O j=1 m j j=1 m j
n
X X
N= Lj = ` j,k .
i=1 ( j,k)∈M
2. OPERADS AND ALGEBRAS 235
n
N mj
N
!
O(n) ⊗ O(m j ) ⊗ O(` j,k )
j=1 k=1
shuffle
id ⊗(⊗ j ◦) n
N N mj
n N
O(n) ⊗ O(m j ) ⊗ O(` j,k )
j=1 j=1 k=1
n
N
O(n) ⊗ O(L j ) ◦⊗id
j=1
mj
n N
N
O(M) ⊗ O(` j,k )
◦ j=1 k=1
O(N)
commutes.
(4) The unit diagrams commute:
'
O(n) ⊗ K⊗n / O(n) ⊗ O(1)⊗n / O(n)
id ⊗η ⊗n ◦
and
&
K ⊗ O(n) / O(1) ⊗ O(n) / O(n) .
η⊗id ◦
interwining in the natural way with all the structure of the operads.
Example: The operad Com describing commutative algebras has Com(n) K, with
the trivial S n action, for all n. This is because there is only one way to multiply n
elements: even if we permute the inputs, we have the same output. ^
236 A. BACKGROUND
Example: The operad Ass describing associative algebras has Ass(n) = K[S n ], the
regular representation of S n , for all n. This is because the product of n elements
only depends on their left-to-right ordering, not on a choice of parantheses. We
should have exactly one n-ary product for each ordering of n elements. ^
Remark: One can describe operads via generators and relations. The two examples
above are generated by a single binary operation. In Ass, there is a relation between
the 3-ary operations generated by that binary operation — the associativity relation
— as already discussed. For a careful treatment of this style of description, we
direct the reader to the references. ^
We now explain the notion of an algebra over an operad. Our approach is
modeled on defining a representation of a group G on a vector space V as a group
homomorphism ρ : G → GL(V). Given a vector space V, there is an operad EndV
that contains all imaginable multilinear operations on V and how they compose,
just like GL(V) contains all linear automorphisms of V.
2.1.2 Definition. The endomorphism operad EndV of a vector space V has n-ary
operations EndV (n) = Hom(V ⊗n , V) and compositions
P
◦n;m1 ,...,mn : O(n) ⊗ (O(m1 ) ⊗ · · · ⊗ O(mn )) → O nj=1 m j
µn ⊗ (µm1 ⊗ · · · ⊗ µmn ) 7→ µn ◦ (µm1 ⊗ · · · ⊗ µmn )
are simply composition of multilinear maps.
2.1.1. References. For a very brief introduction to operads, see the “What is”
column by Stasheff (2004). The article Vallette (2014) provides a nice motivation
and overview for linear operads and their relation to homotopical algebra. For a
systematic treatment with an emphasis on Koszul duality, see Loday and Vallette
(2012). In Costello (n.d.a), there is a description of the basics emphasizing a dia-
grammatic approach: an operad is a functor on a category of rooted trees. Finally,
the book Fresse (n.d.) is wonderful.
2.2. The P0 and BD operads. Two linear operads play a central role in this
book and in the Batalin-Vilkovisky formalism. In physics, there is a basic division
into classical and quantum, and the BV formalism also breaks down along these
lines. The operad controlling the classical BV formalism is the P0 operad, encod-
ing Poisson-zero algebras, equivalently commutative dg algebras with a compatible
Poisson bracket of cohomological degree 1. The operad controlling the quantum
BV formalism is the BD operad, encoding Beilinson-Drinfeld algebras.
2.2.1. P0 . We will define the shifted Poisson operads Pk in terms of their al-
gebras. These operads live in the category of graded vector spaces over any field
of characteristic zero.
2.2.1 Definition. A Pk algebra is a cochain complex A that possesses a commu-
tative product · : A ⊗ A → A of cohomological degree 0 and a Lie bracket
{−, −} : A[k − 1] ⊗ A[k − 1] → A[k − 1] on the shifted complex A[k − 1] sat-
isfying the Poisson relation: for every a ∈ A, the graded linear map {a, −} is a
derivation for the commutative product.
Let’s unpack what that condition means for a BD algebra. Let (A, dA ) be a
cochain complex. To equip it with a BD algebra structure is to give a map ρ of dg
operads from BD to EndA . Forgetting the differentials, we see that the underlying
graded vector space A] is thus a P0 algebra. Now let’s examine what it means that
the map ρ(2) : BD(2) → Hom(A⊗A, A) is a cochain map. Let • denote the commu-
tative product on A, which is the element ρ(2)(·), and let {−, −} denote the Poisson
bracket on A, which is the element ρ(2)({−, −}). Then, since ρ(2) commutes with
the differentials, we must have that
dHom (ρ(2)(·)) = dA ◦ • − • ◦ (dA ⊗ 1 + 1 ⊗ dA )
equals
ρ(2)(d(·)) = ~{−, −}.
In other words, we need
dA (x • y) − (dA x) • y − x • (dA y) = ~{x, y}
for any x, y ∈ A. This equation is precisely the standard relation in the Batalin-
Vilkovisky formalism.
We should remark that this operad is quite similar to an operad known as the
Batalin-Vilkovisky operad, except that cohomological degrees are different. In the
mathematics literature, the BV operad typically refers to an operad related to the
little two-dimensional discs operad E2 . There is the potential for confusion here,
because the BV quantization procedure is related not to the E2 operad, but to the
E0 operad. (In the original physics literature, where these relations were first intro-
duced, everything was Z/2-graded and so it is impossible to distinguish between
these structures.) The role of this BD operad was first recognized in Beilinson and
Drinfeld (2004), and so we named it after them.
2.3. Colored operads aka multicategories. We now introduce a natural si-
multaneous generalization of the notions of a category and of an operad. The es-
sential idea is to have a collection of objects that we can “combine” or “multiply.”
If there is only one object, we recover the notion of an operad. If there are no ways
to combine multiple objects, then we recover the notion of a category.
We will give the symmetric version of this concept, just as we did with operads.
Let (C, ) be a symmetric monoidal category, such as (S ets, ×) or (VectK , ⊗K ).
We require C to have all reasonable colimits.
2.3.1 Definition. A multicategory (or colored operad) M over C consists of
(i) a collection of objects (or colors) ObM,
(ii) for every n + 1-tuple of objects (x1 , . . . , xn | y), an object
M(x1 , . . . , xn | y)
in C called the maps from the x j to y,
2. OPERADS AND ALGEBRAS 239
Usually, we will suppress the notation ρ and simply write x·m or [x, m]. Continuing
with the examples from above, the matrices gln acts on Kn by left multiplication,
so Kn is naturally a gln -module. Analogously, vector fields Vect(M) act on smooth
functions C ∞ (M) as derivations, and so C ∞ (M) is a Vect(M)-module.
There is a category g−mod whose objects are g-modules and whose morphisms
are the natural structure-preserving maps. To be explicit, a map f ∈ g − mod(M, N)
of g-modules is a linear map f : M → N such that [x, f (m)] = f ([x, m]) for every
x ∈ g and every m ∈ M.
Lie algebra homology and cohomology arise as the derived functors of two
natural functors on the category of g-modules. We define the invariants as the
functor
(−)g : g − mod → VectK
M 7→ Mg
where M g = {m | [x, m] = 0 ∀x ∈ g}. (A nonlinear analog is taking the fixed points
of a group action on a set. )The coinvariants is the functor
(−)g : g − mod → VectK
M 7→ Mg
where Mg = M/gM = M/{[x, m] | x ∈ g, m ∈ M}. (A nonlinear analog is taking
the quotient, or orbit space, of a group action on a set, i.e., the collection of orbits.)
To define the derived functors, we rework our constructions into the setting of
modules over associative algebras so that we can borrow the Tor and Ext functors.
The universal enveloping algebra of a Lie algebra g is
Ug = Tens(g)/(x ⊗ y − y ⊗ x − [x, y])
where Tens(g) = ⊕n≥0 g⊗n denotes the tensor algebra of g. Note that the ideal by
which we quotient ensures that the commutator in Ug agrees with the bracket in g:
for all x, y ∈ g,
x · y − y · x = [x, y],
where · denotes multiplication in Ug. It is straightforward to verify that there is an
adjunction
U : LieAlgK AssAlgK : Forget,
where Forget(A) views an associative algebra A over K as a vector space with
bracket given by the commutator of its product. As a consequence, we can view
g − mod as the category of left Ug-modules, which we’ll denote Ug − mod, without
harm.
Now observe that K is a trivial g-module for any Lie algebra g: x · k = 0 for all
x ∈ g and k ∈ K. Moreover, K is the quotient of Ug by the ideal (g) generated by g
itself, so that K is a bimodule over Ug. It is then straightforward to verify that
Mg = K ⊗Ug M and M g = HomUg (K, M)
for every module M.
3.1.1 Definition. For M a g-module, the Lie algebra homology of M is
H∗ (g, M) = TorUg
∗ (K, M),
242 A. BACKGROUND
where the hat ybk indicates removal. As this is a free resolution of K, we use it to
compute the relevant Tor and Ext groups: for coinvariants we have
K ⊗LUg M ' (· · · → ∧n g ⊗K M → · · · → g ⊗K M → M)
and for invariants we have
R HomUg (K, M) ' (M → g∨ ⊗K M → · · · → ∧n g∨ ⊗K M → · · · ),
where g∨ = HomK (g, K) is the linear dual. These resolutions were introduced by
Chevalley and Eilenberg and so their names are attached.
3.1.2 Definition. The Chevalley-Eilenberg complex for Lie algebra homology of
the g-module M is
C∗ (g, M) = (SymK (g[1]) ⊗K M, d)
where the differential d encodes the bracket of g on itself and on M. Explicitly, we
have
X
d(x1 ∧ · · · ∧ xn ⊗ m) = (−1) j+k [x j , xk ] ∧ x1 ∧ · · · xbj · · · xbk · · · ∧ xn ⊗ m
1≤ j<k≤n
n
X
+ (−1)n− j x1 ∧ · · · xbj · · · ∧ xn ⊗ [x j , m].
j=1
where the differential d encodes the linear dual to the bracket of g on itself and on
M. Fixing a linear basis {ek } for g and hence a dual basis {ek } for g∨ , we have
X X
d(ek ⊗ m) = − ek ([ei , e j ])ei ∧ e j ⊗ m + ek ∧ el ⊗ [el , m]
i< j l
When M is the trivial module K, we simply write C∗ (g) and C ∗ (g). It is impor-
tant for us that C ∗ (g) is a commutative dg algebra and that C∗ (g) is a cocommutative
dg coalgebra. This property suggests a geometric interpretation of the Chevalley-
Eilenberg complexes: under the philosophy that every commutative algebra should
be interpreted as the functions on some “space,” we view C ∗ (g) as “functions on
a space Bg” and C∗ (g) as ”distributions on Bg.” Here we interpret the natural pair-
ing between the two complexes as providing the pairing between functions and
distributions. This geometric perspective on the Chevalley-Eilenberg complexes
motivates the role of Lie algebras in deformation theory, as we explain in the fol-
lowing section.
3.1.1. References. Weibel (1994) contains a chapter on the homological alge-
bra of ordinary Lie algebras, of which we have given a gloss. In Lurie (n.d.a), Lurie
gives an efficient treatment of this homological algebra in the language of model
and infinity categories.
In other words, a dg Lie algebra is an algebra over the operad Lie in the cate-
gory of dg R-modules. In practice — and for the rest of the section — we require
the graded pieces gk to be projective R-modules so that we do not need to worry
about the tensor product or taking duals.
244 A. BACKGROUND
4.1. Sheaves.
op
4.1.1 Definition. A presheaf of vector spaces on a space X is a functor F : OpensX →
VectK where OpensX is the category encoding the partially ordered set of open sets
in X (i.e., the objects are open sets in X and there is a map from U to V exactly
when U ⊂ V).
4. SHEAVES, COSHEAVES, AND THEIR HOMOTOPICAL GENERALIZATIONS 245
In other words, a presheaf F assigns a vector space F (U) to each open U and
a restriction map resV⊃U : F (V) → F (U) whenever U ⊂ V. Bear in mind the
following two standard examples. The constant presheaf F = K has K(U) = K
(hence it assigns the same vector space to every open) and its restriction map is
always the identity. The presheaf of continuous functions C X0 assigns the vector
space C X0 (U) of continuous functions from U to R (or C, as one prefers) and the
restriction map resV⊃U consists precisely of restricting a continuous function from
V to a smaller open U.
A sheaf is a presheaf whose value on an open is determined by its behavior on
smaller opens.
4.1.2 Definition. A sheaf of vector spaces on a space X is a presheaf F such that
for every open U and every cover U = {Vi }i∈I of U, we have
Y Y
F (U) → lim F (Vi ) ⇒ F (Vi ∩ V j ) ,
i∈I i, j∈I
where the map out of F (U) is the product of the restriction maps for the inclusion
of each Vi into U and where, in the limit diagram, the top arrow is restriction for
the inclusion of Vi ∩ V j into Vi and the bottom arrow is restriction for the inclusion
of Vi ∩ V j into V j .
where the map into G(U) is the coproduct of the extension maps from the Vi to U
and where, in the colimit diagram, the top arrow is extension from Vi ∩ V j to Vi
and the bottom arrow is extension from Vi ∩ V j to V j .
The crucial example of a cosheaf (for us) is the functor Ec that assigns to the
open U, the vector space Ec (U) of compactly-supported smooth sections of E on
246 A. BACKGROUND
which has an internal differential inherited from G, and whose “external” differen-
tial is the alternating sum of the structure maps for the inclusion of an n + 1-fold
intersection into an n-fold intersection. We write
∞ M
M
Č(U, G) = Tot⊕ G(∩nj=0 Vi j )[n] ,
n=0 ~i∈I n+1
induces a map
M M
(†) extI : colim Vc (Ui ∩ U j ) ⇒ Vc (Ui ) → Vc (U)
i, j∈I i∈I
sending v to ρi v. (Note that here we use the fact that if the support of a section on
U is contained in Ui , it is in the image of exti .) The map
X M
ρ̃i : Vc (U) → Vc (Ui )
i i
has the property that
X X
exti ◦ ρ̃i = idVc (U)
i i
by the definition of a partition of unity. P
We now want to show that this map i ρ̃i induces an inverse to extI . Note that
the colimit appearing in (†) is a quotient of ⊕i Vc (Ui ). If we denote this quotient
space by Q, then we have an exact sequence
M f M
Vc (Ui ∩ U j ) → Vc (Ui ) → Q → 0,
i, j i
where f denotes the difference P of the two maps in the colimit appearing in (†).
Consider the composition of i ρ̃i followed by the projection to A. To show it is a
left inverse to extI , we
P will P show that every element v in ⊕i Vc (Ui ) is in the same
equivalence class as i ρ̃i ◦ i exti (v), i.e., they map to the same element in Q.
It suffices to verify this for a particularly simple class of elements that span
⊕i Vc (Ui ). Fix an index k ∈ I P v = (vi ) ∈ ⊕i Vc (Ui ) have the property that
and let P
vi = 0 if i , k. Let ṽ = (ṽi ) be i ρ̃i ◦ i exti (v). Then ṽi = ρi vk . We want to show
that v and ṽ map to the same element in Q, so we need their difference to live in
image of f . Consider the element w = (wi j ) in ⊕i, j Vc (Ui ∩U j ) where w Pik = ρi vk for
all i and wi j = 0 for j , k. Then f (w) = ( f (w)i ) is given by f (w)i = j (wi j − w ji )
and so f (w)i = ρi vk for i , k and
X
f (w)k = ρk vk − ρ j v k = ρk v k − v k .
j
Hence f (w) = ṽ − v. As this holds for any k, we know it holds for all vectors v.
The argument above is interesting because it shows not only that Vc is a
cosheaf but exhibits an explicit decomposition of a section into sections on the
cover, by using a partition of unity. In fact, the argument can be carried further.
4.4.3 Lemma. Let U = {Ui }i∈I be an open cover of an open U ⊂ M, and let {ρi }i∈I
be a partition of unity subordinate to the cover. There is a cochain homotopy
equivalence between Č(U, Vc ) and Vc (U) where the cochain map
X
σ= extUi ⊂U : Č(U, Vc ) → Vc (U)
i
is determined by the extension maps. Explicitly, σ vanishes on an element of
Vc (Ui0 ∩ · · · ∩ Uin ) for n > 0 and it sends an element of Vc (Ui ) to its extension by
zero.
This lemma immediately implies the following.
4. SHEAVES, COSHEAVES, AND THEIR HOMOTOPICAL GENERALIZATIONS 249
5. Elliptic complexes
Classical field theory involves the study of systems of partial differential equa-
tions (or generalizations), which is an enormously rich and sophisticated subject.
We focus in this book on a tractable and well-understood class of PDE that appear
throughout differential geometry: elliptic complexes. Here we will spell out the
basic definitions and some examples that suffice for our work in this book.
obtained by summing over the indices of order k and replacing the partial deriv-
ative ∂/∂x j by variables iξ j , where i is the usual square root of one. (It is the
standard convention to include the factor of i, due to the role of Fourier transforms
in motivating many of these constructions and definitions.) It is natural to view the
principal symbol as a function on the cotangent bundle T ∗ U that is a homogeneous
polynomial of degree k along the cotangent fibers, where ξ j is the linear functional
dual to dx j . The principal symbol controls the qualitative behavior of L. It also be-
haves nicely under changes of coordinates, transforming as a section of the bundle
Symk (T U ) → U.
Remark: Elsewhere in the text, we talk about differential operators in a more ab-
stract, homological setting. For instance, we say the BV Laplacian is a second-
order differential operator. Our use of the differential operators in this homologi-
cal setting is inspired by their use in analysis, but the definitions are modified, of
course. For instance, we view the odd directions as geometric in the BV formalism,
so the BV Laplacian is second-order. ^
5. ELLIPTIC COMPLEXES 251
This definition says that the principal symbol is an invertible linear operator
after evaluating at any nonzeroP covector. As an example, consider the Laplacian
on Rn : its principal symbol is ξ2j , which only vanishes when all the ξ j are zero.
5.2. Functional analytic consequences. Ellipticity is purely local and is an
easy property to check in practice. Globally, it has powerful consequences, of
which the following is the most famous.
5.2.1 Theorem. For X a closed manifold (i.e., compact and boundaryless), an
elliptic operator Q : E → F is Fredholm, so that its kernel and cokernel are
finite-dimensional vector spaces.
Thus, an elliptic operator on a closed manifold is invertible up to a finite-
dimensional “error.” Moreover, there is a rich body of techniques for constructing
these partial inverses, especially for classical operators such as the Laplacian.
The notions from above extend naturally to cochain complexes. A differential
complex on a manifold X is a Z-graded vector bundle ⊕n E n → X (with finite total
rank) and a differential operator Qn : E n → E n+1 for each integer n such that
Qn+1 ◦ Qn = 0. We typically denote this by (E , Q). There is an associated principal
symbol complex (π∗ E, σQ ) on T ∗ X by taking the principal symbol of each operator
Qn .
5.2.2 Definition. An elliptic complex is a differential complex whose principal
symbol complex is exact on T ∗ X \ X (i.e., the cohomology of the symbol complex
vanishes away from the zero section of the cotangent bundle).
252 A. BACKGROUND
5.4. Formal Hodge theory. We want to explain how to generalize the parametrix
approach to elliptic complexes. The essential idea to make a retraction of the ellip-
tic complex E onto its cohomology H ∗ E , viewed as a complex with zero differen-
tial:
π /
η
8E o ι
H∗E .
Second, let G denote the parametrix for D (where G is for “Green’s function”).
Set η = Q∗G. We need to show it a contracting homotopy with the desired proper-
ties. Let
id −DG = S and id −GD = T,
where S and T are smoothing endomorphisms of cohomological degree 0. Note
that
QD = QQ∗ Q = DQ,
so
G(QD)G = G(DQ)G =⇒ (GQ)(id −S ) = (id −T )(QG).
Hence we find GQ = QG − U, where U = QS − T QG is smoothing because Q is
a differential operator. In consequence,
[Q, η] = QQ∗G + Q∗GQ
= QQ∗G + Q∗ QG − Q∗ U
= DG − Q∗ U
= id −(S + Q∗ U).
The final term in parantheses is smoothing. We have thus verified that η has the
desired properties.
5.5. References. As usual, Atiyah and Bott (1967) explain beautifully the es-
sential ideas of elliptic complexes and pseudodifferential techniques and show how
to use them efficiently. For an accessible development of the analytic methods in
the geometric setting, we recommend Wells (2008). The full story and much more
is available in the classic works of Hörmander (2003).
APPENDIX B
Functional analysis
1. Introduction
The goal of this appendix is to introduce several types of vector spaces and
explain how they are related. In the end, most of the vector spaces we work with
— which are built out of smooth or distributional sections of vector bundles —
behave nicely in whichever framework one chooses to use, but it is important to
have a setting where abstract constructions behave well. In particular, we will
do homological algebra with infinite-dimensional vector spaces, and that requires
care. Below, we introduce the underlying “functional analysis” that we need (i.e.,
we describe here just the vector spaces and discuss the homological issues in a
separate appendix). For a briefer overview, see Section 5.
There are four main categories of vector spaces that we care about:
• LCTVS, the category of locally convex Hausdorff topological vector
spaces,
• BVS, the category of bornological vector spaces,
• CVS, the category of convenient vector spaces, and
• DVS, the category of differentiable vector spaces.
The first three categories are vector spaces equipped with some extra structure,
like a topology or bornology, satisfying some list of properties. The category DVS
consists of sheaves of vector spaces on the site of smooth manifolds, equipped with
some extra structure that allows us to differentiate sections (hence the name).
The main idea is that DVS provides a natural place to compare and relate vec-
tor spaces that arise in differential geometry and physics. As a summary of the
relationships between these categories, we have the following diagram of functors:
incβ
BVS / LCTVS .
c∞ di fβ di ft
%
CVS / DVS
di fc
All the functors into DVS preserve limits. The functors incβ and c∞ out of BVS
are left adjoints. The functors into DVS all factor as a right adjoint followed by the
functor di fβ . For example, di ft : LCTVS → DVS is the composition di fβ ◦ born,
where the bornologification functor born : LCTVS → BVS is the right adjoint to
the inclusion incβ : BVS → LCTVS.
255
256 B. FUNCTIONAL ANALYSIS
For our work, it is also important to understand multilinear maps between vec-
tor spaces (in the categories mentioned above). Of course, it is pleasant to have
a tensor product that represents bilinear maps. In the infinite-dimensional setting,
however, the tensor product is far more complicated than in the finite-dimensional
setting, with many different versions of the tensor product, each possessing various
virtues and defects. We will discuss certain natural tensor products that appear on
the categories above and how the functors intertwine with them.
As suggestive notation, we will typically denote the vector space V(X) of sec-
tions on X of V by C ∞ (X, V), to emphasize that we think of these as the smooth
functions from X to the smooth vector space V. A map φ : V → W then con-
sists of a map φX : C ∞ (X, V) → C ∞ (X, W) for every manifold X such that for
every smooth map f : X → Y, we have fW∗ ◦ φY = φX ◦ fV∗ , where we use
fV∗ : C ∞ (Y, V) → C ∞ (X, V) to denote the natural pullback map.
Remark: Given a presheaf F on Mfld that is a module over C ∞ (in the category of
presheaves of vector spaces), the sheafification is also naturally a C ∞ -module, and
this sheafification can be computed as simply a presheaf of sets or vector spaces.
Moreover, the forgetful functor from C ∞ -modules to smooth vector spaces is right
adjoint to the tensor product C ∞ ⊗−, i.e., base-change behaves as expected from or-
dinary algebra in this sheaf-theoretic context. See Kashiwara and Schapira (2006)
or Stacks Project Authors (2016) for more. ^
All the natural vector bundles in differential geometry provide examples of
C ∞ -modules. For instance, consider differential forms. Let Ωk denote the sheaf
that assigns Ωk (X) to each smooth manifold X. Given a C ∞ -module V, we define
the k-forms with values in V as
Ωk (−, V) := Ωk ⊗C ∞ V,
so that on a manifold X, we have
Ωk (X, V) = Ωk (X) ⊗C ∞ (X) C ∞ (X, V).
With this definition in hand, we can define our main object of interest.
2.0.3 Definition. A differentiable vector space is a C ∞ -module V equipped with a
flat connection
∇X,V : C ∞ (X, V) → Ω1 (X, V)
for every smooth manifold X such that pullback commutes with the connections,
f ∗ ◦ ∇Y,V = ∇X,V ◦ f ∗ ,
for every smooth map f : X → Y.
To say that ∇X,V is a connection means that it satisfies the Leibniz rule,
∇X,V ( f · v) = (d f )v + f ∇X,V v,
where f ∈ C ∞ (X) and v ∈ C ∞ (X, V). To say that it is flat means that the curvature
F(∇X,V ) = (∇X,V )2 : C ∞ (X, V) → Ω2 (X, V)
vanishes.
The flat connection ∇X,V thus allows us to differentiate sections of V on X. If
X is a vector field on X, we define
X(v) := X, ∇X,V v ∈ C ∞ (X, V),
258 B. FUNCTIONAL ANALYSIS
where h−, −i denotes the C ∞ (X)-pairing between vector fields and 1-forms
h−, −i : T (X) × Ω1 (X, V) → C ∞ (X, V).
Moreover, because the curvature vanishes, the Lie bracket of vector fields goes to
the commutator:
[X, Y](v) = X(Y(v)) − Y(X(v)).
Thus we can take as many derivatives as we would like.
This ability to do calculus with differentiable vector spaces is crucial to many
of our constructions. See, for instance, the sections on translation-invariant factor-
ization algebras, holomorphically translation-invariant factorization algebras, and
equivariant factorization algebras.
2.0.4 Definition. A map of differentiable vector spaces φ : V → W is a map of
C ∞ -modules such that
∇X,W ◦ φX = (idΩ1 (X) ⊗φX ) ◦ ∇X,V
for every smooth manifold X.
We denote the category of differentiable vector spaces by DVS. We denote the
vector space of morphisms from V to W by DVS(V, W)
We still need to equip these sheaves with flat connections to make them differ-
entiable vector spaces. There is a natural choice, due to the fact that the pullback
bundle π∗M E is trivial in the X-direction.
2. DIFFERENTIABLE VECTOR SPACES 259
2.1.2 Definition. Let X be a smooth manifold. Equip the pullback bundle π∗M E on
X × M with the natural flat connection along the fibers of the projection map π M
(i.e., we only differentiate in X-directions). We thus obtain a map
or equivalently a map
This map defines a flat connection on E (M) and hence gives it the structure of a
differentiable vector space.
As this flat connection does not increase support, it preserves the subspace
Ec (M) of sections with proper support over X, and it gives Ec (M) the structure of
a differentiable vector space.
Remark: Typically, one works with Γ(M, E) as a complete locally convex topolog-
ical vector space, via the Fréchet topology. Below, in Section 3, we will explain a
result of Kriegl and Michor that shows that every locally convex topological vector
space V naturally produces a differentiable vector space. Moreover, Γ(M, E) goes
to E under this functor. Hence, the differentiable vector space E (M) arises from
the standard topology on Γ(M, E). (An identical comment applies to Γc (M, E).) ^
2.1.1. We are also interested in distributional sections of a vector bundle p :
E → M. We will show that these also form a differentiable vector space, after
setting up the preliminaries about distributions.
Let D(M) denote the vector space of distributions on the smooth manifold
M. That is, D(M) is the continuous linear dual to the vector space Cc∞ (M). Let
Dc (M) denote the vector space of compactly supported distributions on M, i.e., the
continuous linear dual of C ∞ (M). This space is also a C ∞ (M)-module.
2.1.3 Definition. Let D(M) denote the C ∞ -module whose smooth sections C ∞ (X, D(M))
on the manifold X are the continuous linear maps from Cc∞ (M) to C ∞ (X).
Similarly, let Dc (M) denote the C ∞ -module whose smooth sections on the man-
ifold X, denoted C ∞ (X, Dc (M)), are the continuous linear maps from C ∞ (M) to
C ∞ (X).
Note that when X is a point, we recover the usual notion of a distribution. The
definition above arises by asking that a smooth map φ from X to D(M) correspond
to a smooth family of distributions {φ x } x∈X on M. If one evaluates on f ∈ Cc∞ (M),
then {φ x ( f )} x∈X should be a smooth function X.
We now equip these C ∞ -modules with a natural flat connection.
2.1.4 Definition. The vector fields TX on X act in a natural way on the vector
space C ∞ (X, D(M)) by
X · φ : f 7→ X(φ( f )),
260 B. FUNCTIONAL ANALYSIS
We call this sheaf the distributional sections of the vector bundle E. Similarly, the
compactly-supported distributional sections of E is the C ∞ -module E¯c (M) where
Note that setting X to be a point, we recover the original notion of the distribu-
tional sections.
There is a natural flat connection on E (M) and E c (M) arising from a natural
action of the vector fields TX on these spaces, just as in the case of D(M) and
Dc (M). Hence, these are differentiable vector spaces.
2.1.3. Holomorphic vector bundles and sections. Let M be a complex mani-
fold and p : E → M a holomorphic vector bundle. The holomorphic sections also
form a differentiable vector space; we will show this in the simplest case, as the
general case is completely parallel.
2.1.6 Definition. For M a complex manifold and X a smooth manifold, let f ∈
C ∞ (X, O(M)) denote a smooth function on the product manifold X × M such that
f x (m) := f (x, m) is holomorphic on M for every x ∈ X.
Note that C ∞ (−, O(M)) is a subsheaf of C ∞ (−, C ∞ (M)), using the definition
of the differentiable vector space C ∞ (M) from above. It is straightforward to see
that the flat connection on C ∞ (X, C ∞ (M)) preserves the subspace C ∞ (X, O(M)) for
every X. Hence C ∞ (−, O(M)) is a differentiable vector space that we will denote
O(M).
2. DIFFERENTIABLE VECTOR SPACES 261
2.2. Stalks and local properties. Let V be a differentiable vector space. Re-
stricting to Rn , we obtain a sheaf V|Rn on Rn . We can thus define the n-dimensional
stalk of V:
Stalkn (V) = colim0∈U⊂Rn V(U).
The colimit above is taken over open subsets of Rn containing the origin and is
computed in the category of vector spaces.
Note that the stalk of V at a point in any manifold can be defined in the same
way, but the stalk at a point in a n-dimensional manifold is the same as the stalk
at the origin in Rn . Because of this “uniformity” of differentiable vector spaces,
a local property (i.e., a property verified stalkwise) only needs to be checked in
countably many cases. As an example, consider how convenient the standard cri-
terion for the exactness of a sequence becomes in our setting.
2.2.1 Lemma. A sequence of differentiable vector spaces
0→A→B→C→0
is exact if and only if, for all n, the sequence
0 → Stalkn (A) → Stalkn (B) → Stalkn (C) → 0
of vector spaces is exact.
More generally, it is a standard fact that the functor of taking stalks preserves
colimits and finite limits. The following corollary will be useful for us.
2.2.2 Corollary. A map f : A → B of differentiable vector spaces is an isomor-
phism if and only if, for all n, the map Stalkn ( f ) : Stalkn (A) → Stalkn (B) is an
isomorphism.
2.3. Categorical properties. The category DVS is well-behaved in the sense
that all the usual constructions make sense. For example, we have the following.
2.3.1 Lemma. DVS contains all finite products.
for A a finite set. Hence the product of the connections for each Vα provides the
desired connection
Y Y
∞
∇X, Vα : C X,
Q Vα → Ω X,
1
Vα .
α∈A α∈A
It is straightforward to check that this construction satisfies the universal property
of a product.
Note in the proof that we used crucially the tensoring with Ω1 commutes with
finite products of C ∞ -modules. There is no reason to expect tensoring commutes
with infinite products. Below we give an indirect proof that DVS contains infinite
products as well.
We now provide kernels.
2.3.2 Lemma. For any map φ : V → W of differentiable vector spaces, the kernel,
given by
ker φ : X 7→ ker φ(X) : C ∞ (X, V) → C ∞ (X, W) ,
The connection
M M
∇X,L Vα : C ∞ (X, Vα ) → Ω1 (X, Vα )
α∈A α∈A
is the unique connection that restricts to ∇X,Vα for each Vα . It is manifest that any
cocone in DVS for this collection {Vα } factors through the differentiable vector
space just constructed.
Remark: For finite coproducts, the presheaf formula is already a sheaf and provides
the coproduct in C ∞ -modules. The natural connection makes it the coproduct of
differentiable vector spaces. ^
2.3.5 Lemma. For any map φ : V → W of differentiable vector spaces, the coker-
nel coker φ is given by sheafifying the presheaf
] φ : X 7→ coker φ(X) : C ∞ (X, V) → C ∞ (X, W) ,
coker
This C ∞ module is a differentiable vector space by a natural connection induced
from the connection ∇X,W .
] φ. (Recall Re-
Proof. Let Cφ denote the sheafification of the presheaf coker
mark 2: the sheafification can be as a presheaf of sets or vector spaces or C ∞ -
modules, as they all provide the same object, after suitable forgetting of module
structure.) This sheaf Cφ provides the cokernel in C ∞ -modules.
We need to produce the flat connection on Cφ . As Ω1 is a projective C ∞ -
module, it is a flat C ∞ -module so that
1⊗φ 1⊗qφ
Ω1 ⊗C ∞ V −→ Ω1 ⊗C ∞ W −→ Ω1 ⊗C ∞ Cφ −→ 0
is exact. As flat connections are C-linear but not C ∞ -linear (i.e., they are maps of
smooth vector spaces but not of C ∞ -modules), the commutative diagram
1⊗φ 1⊗qφ
Ω1 ⊗OC ∞ V / Ω1 ⊗C ∞ W / Ω1 ⊗C ∞ Cφ /0
O
∇V ∇W
φ qφ
V /W / Cφ /0
lives in the category of smooth vector spaces. Since smooth vector spaces form
an abelian category, there is naturally a C-linear map fCφ from Cφ to Ω1 ⊗C ∞ Cφ
inherited from the connection on W. We need to show that fCφ is a flat connection,
but this assertion follows from the fact that φ intertwines the flat connections on V
and W.
2.3.6 Corollary. DVS contains all colimits. Hence it is a cocomplete category.
Proof. Every colimit can be constructed out of coproducts and coequalizers.
Since we are in a linear setting, we can construct an coequalizer as a cokernel.
In fact, DVS contains all limits and is hence complete, but the argument is
indirect. We first prove that DVS is locally presentable, and by Corollary 1.28 of
264 B. FUNCTIONAL ANALYSIS
Adámek and Rosický (1994), we know that every locally presentable category is
complete.
2.3.7 Proposition. DVS is locally presentable.
2.3.8 Corollary. DVS contains all limits, and so it is a complete category.
The proof of this proposition depends on some technical results in category
theory, rather orthogonal from the rest of this book, so we will not give expository
background. Our main reference are Adámek and Rosický (1994) and Kashiwara
and Schapira (2006).
Proof of Proposition 2.3.7. As we know that DVS is cocomplete, it suffices to
prove that DVS is accessible (see Corollary 2.47 of Adámek and Rosický (1994)).
Our strategy is to present DVS as constructed out of known accessible categories
by operations that produce accessible categories. To be more precise, there is a 2-
category ACC of accessible categories, and it possesses certain “limits” understood
in the correct 2-categorical sense. In particular, ACC is closed under “inserters”
and “equifiers.” Below, we will define and use precisely these constructions.
First, we must show that our initial inputs are accessible. Recall that the mod-
ule sheaves for a sheaf of rings on a site is presentable and hence accessible. (See,
for instance, Theorem 18.1.6 of Kashiwara and Schapira (2006).) Thus the cat-
egory of smooth vector spaces Vect sm – i.e., sheaves of vector spaces on the site
Mfld – is accessible. Likewise, the category ModC ∞ of C ∞ -modules is accessible.
Consider the following two functors:
F : ModC ∞ → Vect sm ,
the forgetful functor, and
G := G ◦ Ω1 ⊗C ∞ : ModC ∞ → Vect sm ,
which sends a C ∞ -module V to the 1-forms valued in V, Ω1 ⊗C ∞ , and then views
it as a smooth vector space. The forgetful functor F preserves filtered colimits by
Theorem 18.1.6 of Kashiwara and Schapira (2006); hence F is accessible. The
functor Ω1 ⊗C ∞ − is a left adjoint from ModC ∞ to itself and hence preserves all col-
imits, so the composite G also preserves filtered colimits and is hence accessible.
The inserter of F and G is the following category, denoted Ins(F, G). An
object of Ins(F, G) is a morphism f : F(V) → G(V) with V an object in ModC ∞ .
A morphism in Ins(F, G) is a commuting square
f
F(V) / G(V)
F(φ) G(φ)
f0
F(V 0 ) / G(V 0 )
Remark: In the remainder of this section, we will use topological vector space to
mean locally convex Hausdorff topological vector space, for simplicity. ^
The key idea of Kriegl and Michor is simple and compelling: the notion of a
smooth curve is well-behaved, so build up everything from that notion.
In particular, it is possible to define smooth curves γ : R → V in a topological
vector space in a very simple way, in the style of elementary calculus. One then
defines a smooth map f : V → W between topological vector spaces to be a
map that sends smooth curves in V to smooth curves in W. This notion extends
naturally to define smooth maps from a finite-dimensional smooth manifold X into
a topological vector space V. Hence, we obtain a smooth vector space (i.e., a sheaf
on Mfld) from V: to the manifold X, we assign the set of smooth maps C ∞ (X, V).
Remark: In fact, a map f : M → V is smooth in this sense if and only if it is
smooth in the usual sense. See Lemma 3.0.7 below. ^
3.0.2 Definition. Let V ∈ LCTVS. A curve γ : R → V is differentiable if its
derivative
γ(t + s) − γ(t)
γ0 (t) := lim
s→0 s
exists at t ∈ R for all t.
A curve γ is smooth if all iterated derivatives exist.
With this definition in hand, we introduce the notion of a smooth map between
topological vector spaces, following Kriegl and Michor.
268 B. FUNCTIONAL ANALYSIS
As Kriegl and Michor explain, this notion arises by taking seriously the per-
spective of variational calculus.
Remark: We note that not every smooth function between topological vector spaces,
in this sense, is necessarily continuous. As Kriegl and Michor point out, this a
priori unappealing mismatch is not so strange: smoothness is about nonlinear phe-
nomena, and the usual topology on a topological vector space is focused on linear
phenomena. (For many nice topological vector spaces, however, a smooth function
is continuous.) ^
In particular, for an open U ⊂ Rn and V a topological vector space, we say that
f : U → V is smooth if any smooth curve γ : R → U goes to a smooth curve
f ◦ γ : R → V. For X a smooth manifold, a function f : X → V is smooth if
its restriction to every chart is smooth. Hence the notion is local on X, so that we
obtain the following definition.
3.0.4 Definition. Let V ∈ LCTVS. We define a C ∞ -module by
X ∈ Mfldop 7→ { f : X → V | f smooth}.
We denote this functor by C ∞ (−, V).
Let modt : LCTVS → ModC ∞ denote the functor that sends V to C ∞ (−, V). It
has the following useful property.
3.0.5 Lemma. The functor modt preserves limits.
Proof. The forgetful functor from sheaves of C ∞ -modules to sheaves of sets
preserves limits, because it is a right adjoint. Likewise, the forgetful functor from
sheaves of sets to presheaves of sets preserves limits, since it is a right adjoint. But
a limit of presheaves is computed objectwise, since it is simply a limit in a functor
category. We thus merely need to show that
C ∞ (X, modt (lim V)) C ∞ (X, lim modt (V))
I I
But this functor C ∞ (−, V) is even better. By construction, these maps C ∞ (X, V)
are differentiable, so we find that C ∞ (−, V) is naturally a differentiable vector
3. LOCALLY CONVEX TOPOLOGICAL VECTOR SPACES 269
Kernels are similar. Being the kernel of a continuous linear map is a closed
condition, and a closed linear subspace of a locally convex vector space is a locally
convex vector space.
3.1.2 Lemma. LCTVS admits all colimits.
Proof. We show the existence of cokernels and coproducts.
Cokernels are straightforward. For a continuous linear map φ : V → W, the
cokernel is W/φ(V), the quotient of W by the closure of the image of φ. We are
forced to use the closure as we work with Hausdorff spaces.
Coproducts are a little trickier. The underlying vector space is simply the direct
sum, but we equip it with the “diamond” topology. (We learned this term from the
helpful online lecture notes of Paul Garrett.) We construct a collection of convex
neighborhoods of the origin out of convex neighborhoods of the constituents as
follows.
Let iα : Vα ,→
L
α∈A Vα denote the canonical inclusion. Given a choice of
convex neighborhood Uα of Vα for every α ∈ A, take the convex hull of the union
S
α∈A iα (U α ). Declare such a convex hull to be an open. We take the vector space
topology induced by all such opens.
4.0.2 Definition. A bornological vector space (V, B) is a vector space whose bornol-
ogy is obtained from some locally convex Hausdorff topology on V.
A linear map f : V → W is bounded if the image of every bounded set in V is
a bounded set in W.
We denote by BVS the category where an object is a bornological vector space
and a morphism is a bounded linear map.
Note that a continuous linear map is always bounded, but the converse is false
in general. (Consider an infinite-dimensional Banach space V, and let Vw denote
the underlying vector space equipped with the weak topology. The identity map
id : Vw → V is bounded but not continuous.)
Consider the functor incβ : BVS → LCTVS that assigns to a bornological vec-
tor space the finest locally convex topology with the same underlying bounded sets.
By Lemma 4.2 of Kriegl and Michor (1997), this functor embeds BVS as a full sub-
category of LCTVS. In particular, a linear map f : V → W between bornological
vector spaces is bounded if and only if f : incβ V → incβ W is continuous.
4.0.3 Corollary. The functor incβ : BVS ,→ LCTVS preserves all colimits.
Proof. LCTVS contains all colimits. Given a diagram δ : D → BVS, let
colim incβ ◦ δ denote its colimit in LCTVS. If there exists a colimit colim δ in
BVS, then we know there is a canonical map
colim incβ ◦ δ → incβ colim δ.
Hence, it suffices to show that colim incβ ◦ δ is in the image of incβ to obtain the
claim.
For each object d ∈ D, we have a continuous linear map
incβ δ(d) → colim incβ ◦ δ,
which is thus a bounded linear map. Hence, we know that this map factors through
the underlying vector space of colim incβ ◦ δ, but now equipped with the finest lo-
cally convex topology with the same bounded sets. Hence, the colimit colim incβ ◦δ
is in the image of incβ .
By the very definition of incβ , we see that there is a natural right adjoint.
4.0.4 Corollary. Consider the functor
born : LCTVS → BVS
sending a locally convex Hausdorff topological vector space to its underlying bornolog-
ical vector space. This functor born is right adjoint to incβ .
It is important to know that it is the bornology of a topological vector space
that matters for smoothness, not the topology itself. The following two results from
Kriegl and Michor (1997) make this assertion precise.
4.0.5 Lemma (Kriegl and Michor (1997), Corollary 1.8). For V ∈ LCTVS, a
curve γ : R → V is smooth if and only if γ : R → incβ (born(V)) is smooth.
272 B. FUNCTIONAL ANALYSIS
4.0.6 Lemma (Kriegl and Michor (1997), Corollary 2.11). A linear map L :
V → W of locally convex vector spaces is bounded if and only if it maps smooth
curves in V to smooth curves in W.
In consequence, we have the following result.
4.0.7 Proposition. The functor
di fβ : BVS → DVS
V 7→ di ft (incβ (V))
embeds BVS as a full subcategory of DVS. It preserves all limits.
Note that BVS is therefore equivalent to the essential image of LCTVS in DVS.
Proof. Only the assertion about limits remains to be proved. Note that born
preserves limits, so that since BVS is a full subcategory of LCTVS, we can com-
pute a limit in LCTVS and then apply born. As di ft preserves limits and only cares
about the underlying bornology, we are finished.
4.1. Categorical properties. We have seen that incβ realizes BVS as a reflec-
tive subcategory of LCTVS. Thus, we know the following.
4.1.1 Lemma. BVS admits all limits and colimits.
Proof. LCTVS contains all limits, and born : LCTVS → BVS is a right
adjoint and hence preserves limits. A colimit of BVS is computed by applying
born to the colimit computed in LCTVS.
4.2. Multilinear maps and the bornological tensor product. As in any cat-
egory of linear objects, we can discuss multilinear maps. We can also ask if mul-
tilinear maps out of a tuple of vector spaces is co-represented by a vector space,
called the tensor product. For bornological vector spaces, there is such a natural
tensor product with all the properties we love about the tensor product of finite-
dimensional vector spaces.
4.2.1 Definition. Let BVS(V1 , . . . , Vn | W) denote the set of all bounded multilin-
ear maps from the set-theoretic product of bornological vector spaces V1 × · · · × Vn
to the bornological vector space W. Equip it with the bornology of uniform con-
vergence on bounded sets. This bornology makes BVS(V1 , . . . , Vn | W) into a
bornological vector space.
The following result is the first step to showing that BVS forms a closed sym-
metric monoidal category.
4.2.2 Proposition (Kriegl and Michor (1997), Proposition 5.2). There are natu-
ral bornological isomorphisms
BVS(V1 , . . . , Vn+m | W) BVS (V1 , . . . , Vn | BVS(Vn+1 , . . . , Vn+m | W)) .
Hence, BVS forms a multicategory where multimorphisms are themselves bornolog-
ical vector spaces and composition is bounded.
5. CONVENIENT VECTOR SPACES 273
di fc : CVS → DVS
V 7→ di fβ ◦ incc (V)
also preserves limits. Moreover, because CVS is a full subcategory of BVS and
BVS is a full subcategory of DVS, we have the following.
5.0.3 Lemma. The functor di fc : CVS → DVS embeds CVS as a full subcategory
that is closed under limits.
We note that while di fc does not preserve all colimits, it does preserve some.
5.0.4 Proposition. The functor di fc : CVS → DVS preserves countable coprod-
ucts and sequential colimits of closed embeddings.
Proof. A countable coproduct is a special case of a sequential colimit of closed
embeddings, so we will prove only the latter property.
Given a sequence
V1 → V2 → · · ·
of closed embeddings, let V denote colim Vi . We need to show that
5.1. Examples from differential geometry. Let us explain some further ex-
amples. Let E be a vector bundle on a manifold M, and, as before, let E , Ec , E , E c
refer to sections of E which are smooth, compactly supported, distributional, or
compactly-supported and distributional. All of these spaces have natural topologies
and so can be viewed as bornological vector spaces, and they are all c∞ -complete.
We explained in Section 2.1 how to view these vector spaces as differentiable.
It is easy to check that the smooth structure discussed there is the same as the
one that arises from the topology. Indeed, the Serre-Swan theorem tells us that any
vector bundle is a direct summand of a trivial vector bundle, so we can reduce to the
case that E is trivial. Then results of Grothendieck, summarized in Grothendieck
(1952), allow one to describe smooth maps to these various vector spaces using
the theory of nuclear vector spaces; in this way we arrive at the description given
earlier.
5.1.1 Lemma. For any vector bundle E on a manifold M, the standard nuclear
topologies on E c , E and Ec are bornological. It follows that these spaces are
convenient (because completeness in the locally-convex sense is stronger than c∞ -
completeness).
Proof. Because every vector bundle is a summand of a trivial one, it suffices to
prove the statement for the trivial vector bundle. According to Kriegl and Michor
(1997), 52.29, the strong dual of a Fréchet Montel space is bornological. The space
C ∞ (M) of smooth functions on a manifold is Fréchet Montel, because every nu-
clear space is Schwartz (see pages 579-581 of Kriegl and Michor (1997)) and every
Fréchet Schwarz space is Montel. Thus the strong dual of C ∞ (M) is bornological,
as desired.
Next, we will see that any Fréchet space is bornological. This property follows
immediately from proposition 14.8 of Trèves (1967) (see also the corollary on the
following page). It follows that C ∞ (M) is bornological, and that the same holds for
C K∞ (M) for any compact subset K ⊂ M.
Since bornological spaces are closed under formation of colimits, the same
holds for Cc∞ (M).
5.2. Multilinear maps and the completed tensor product. We can take the
bornological tensor product of any two convenient vector spaces V and W, but
V ⊗β W is rarely convenient itself. The completion of a tensor product c∞ (V ⊗β W) is
convenient, however, and equips CVS with a natural symmetric monoidal structure.
We denote this completed tensor product by b ⊗β . That is,
⊗β W := c∞ (V ⊗β W).
Vb
(Strictly speaking, we should write c∞ (incc (V) ⊗β incc (W)) to make clear where
everything lives.)
5.2.1 Lemma. The multicategory structure on CVS induced by the functor di fc :
CVS → DVS is represented by b
⊗β .
276 B. FUNCTIONAL ANALYSIS
Proof. We have
Proof. Recall that C ∞ (V, W) denotes the set of functions that sends smooth
curves in V to smooth curves in W. We equip it with the initial topology such that
for any smooth γ : R → V, the pullback γ∗ : C ∞ (V, W) → C ∞ (R, W) is continuous.
This topology makes C ∞ (V, W) into a locally convex topological vector space.
Then BVS(V, W) ⊂ C ∞ (V, W) is a closed set because f ∈ C ∞ (V, W) is linear if
and only if
for any manifold X. In particular, the value on a point C ∞ (∗, HomDVS (V, W))
equals DVS(V, W).
We use this enriched category in our definition of equivariant prefactorization
algebras (see section 7) and of holomorphically translation-invariant prefactoriza-
tion algebras (see 1).
Another compelling feature of this category is that it matches nicely with the
internal hom in CVS, the category of convenient vector spaces, which is the source
of most of the vector spaces that we actually work with.
6.0.2 Lemma. For any manifold M and convenient vector spaces V, W, there is an
isomorphism
C ∞ (M, CVS(V, W)) = CVS(V, C ∞ (M, W)),
and it is natural in M. This identification is compatible with the flat connection
possessed by the C ∞ (M) modules on each side. Hence
CVS(V, W) HomDVS (V, W)
as differentiable vector spaces.
The proof appears in the third subsection.
6.1. Constructions with families. To start, we want to construct, for every
manifold X, a sheaf on Mfld that returns, intuitively speaking, “families over X of
horizontal sections of the differentiable vector space V.” To do this, we need to
specify what we mean by such families.
6.1.1 Definition. For a manifold X and a differentiable vector space V, let C∞ (X, V)
denote the presheaf sending a manifold Σ to C ∞ (Σ × X, V).
In other words, C∞ (X, V) is the composite of the product functor −×X followed
by evaluating the sheaf V.
6.1.2 Lemma. For any manifold X and any V ∈ DVS, the presheaf C∞ (X, V) is
canonically a differentiable vector space.
Proof. It is straightforward to see that C∞ (X, V) is a sheaf. Any cover of Σ
pulls back to a cover over Σ × X along the projection πΣ : Σ × X → Σ, and since V
is a sheaf, we can then reconstruct the value of V on Σ × X using this cover.
Likewise, as C ∞ (Σ, C∞ (X, V)) = C ∞ (Σ × X, V) is a C ∞ (Σ × V)-module, it is a
C (Σ)-module via the pullback map π∗Σ : C ∞ (Σ) → C ∞ (Σ × X).
∞
maps. In consequence,
Ω1 (Σ × X) Ω1 (X) ⊗C ∞ (X) C ∞ (Σ × X) ⊕ Ω1 (Σ) ⊗C ∞ (Σ) C ∞ (Σ × X),
where, for example, C ∞ (Σ × X) is a module over C ∞ (X) via the projection πX :
Σ × X → X. Hence we have a decomposition
Ω1 (Σ × X, V) Ω1 (Σ × X) ⊗C ∞ (Σ×X) C ∞ (Σ × X, V)
Ω1 (X) ⊗C ∞ (X) C ∞ (Σ × X, V) ⊕ Ω1 (Σ) ⊗C ∞ (Σ) C ∞ (Σ × X, V)
Ω1 (X, C∞ (Σ, V)) ⊕ Ω1 (Σ, C∞ (X, V)).
Define ∇Σ,C∞ (X,V) as the composition of ∇Σ×X,V followed by projection to Ω1 (Σ, C∞ (X, V)).
On Σ × X, we know that ∇Σ×X,V is a connection on Σ × X and hence satisfies the
Leibniz rule, so the projection onto Ω1 (Σ, C∞ (X, V)) also satisfies the Leibniz rule.
Likewise, the curvature of ∇Σ,C∞ (X,V) over Σ is the projection of the curvature of
∇Σ×X,V to Ω2 (Σ, C∞ (X, V)), and hence zero.
We also need a slight generalization of the mapping space C∞ (X, V). Note that
C ∞ (X) provides an algebra object in DVS and that C∞ (X, V) is a module sheaf over
C ∞ (X).
6.1.3 Definition. For E → X a vector bundle on a smooth manifold and V a
differentiable vector space, let C∞ (X, E ⊗ V) denote the differentiable vector space
E (X) ⊗C ∞ (X) C∞ (X, V).
The value of this sheaf on a manifold Σ is
C ∞ (Σ, C∞ (X, E ⊗ V)) = C ∞ (Σ, E (X)) ⊗C ∞ (Σ,C ∞ (X)) C ∞ (Σ, C∞ (X, V))
= Γ(Σ × X, π∗X E) ⊗C ∞ (Σ×X) C ∞ (Σ × X, V).
We use Ω1 (X, V) to denote C∞ (X, T X∗ ⊗ V).
Tensoring over C ∞ (X) is well-behaved here as E (X) is finite rank and projec-
tive over C ∞ (X).
We now prove a key result for constructing the enriched hom HomDVS .
6.1.4 Lemma. For any differentiable vector spaces V and for any manifold X,
there is a natural map of sheaves
∇C∞ (X,V) : C∞ (X, V) → Ω1 (X, V)
making C∞ (X, V) a C ∞ (X)-module sheaf with a flat connection.
Proof. The flat connection on C∞ (X, V) arises from the composition
∇X×Σ,V
C ∞ (Σ, C∞ (X, V)) = C ∞ (Σ × X, V) −−−−−→ Ω1 (Σ × X, V) → Ω1 (Σ, C∞ (X, V)),
∗
where the final projection map arises from the splitting T Σ×X π∗X T X∗ ⊕ π∗Σ T Σ∗ . This
∞
flat connection ∇Σ,C (X,V)) is C (X)-linear, since it only takes derivatives in the
∞
We now show that the flat connection on V induces a natural map of sheaves
∇C∞ (X,V) : C∞ (X, V) → Ω1 (X, V).
This map is a flat connection for C∞ (X, V) viewed as a C ∞ (X)-module in ModC∞ . It
is not a map in DVS because it anticommutes with the flat connections on C∞ (X, V)
and Ω1 (X, V). Nonetheless, as we will see, this property is enough to obtain the
lemma.
Let Σ be an arbitrary manifold. The splitting T Σ×X
∗ π∗X T X∗ ⊕ π∗Σ T Σ∗ produces
a connection in the X-direction by projection. More explicitly, the composition of
the connection over Σ × X followed by projection onto the first summand,
∇X×Σ,V
C ∞ (Σ × X, V) −−−−−→ Ω1 (Σ × X, V) → Γ(Σ × X, π∗X T X∗ ) ⊗C ∞ (Σ×X) C ∞ (Σ × X, V),
gives a map
∇C∞ (X,V) : C ∞ (Σ, C∞ (X, V)) → C ∞ (Σ, Ω1 (X, V)) = C ∞ (Σ, C∞ (X, T X∗ ⊗ V)).
Note that this map is C ∞ (Σ)-linear since it only takes derivatives in the X-direction.
Moreover, this construction is clearly natural in Σ and thus defines a map as C ∞ -
modules.
This map ∇C∞ (X,V) is a flat connection with respect to C ∞ (X), viewed as a
C -module. On Σ × X, we know that ∇Σ×X,V is a connection on Σ × X and hence
∞
satisfies the Leibniz rule, so the projection onto C ∞ (Σ, C∞ (X, T X∗ ⊗V)) also satisfies
the Leibniz rule. Likewise, the curvature of ∇C∞ (X,V) over Σ is the projection of the
curvature of ∇Σ×X,V to C ∞ (Σ, Ω2 (X, V)), and hence zero. In short, we know that
∇Σ×X,V = ∇Σ,C∞ (X,V) + ∇C∞ (X,V) ,
because they both arise as ∇Σ×X,V followed by projections onto the two summands
of Ω1 (Σ × X). As ∇Σ×X,V is flat, we obtain the for each summand separately.
Note that this decomposition also implies that
0 = ∇Σ,Ω1 (X,V) ◦ ∇C∞ (X,V) + ∇C∞ (X,V) ◦ ∇Σ,C∞ (X,V) ,
justifying our claim that ∇C∞ (X,V) anticommutes with the flat connections and hence
is not a map of differentiable vector spaces.
6.2. The enriched hom. We now introduce the appropriate notion of “fami-
lies of maps from V to W.”
6.2.1 Definition. For V and W differentiable vector spaces, let HomDVS (V, W) de-
note the following presheaf. The sections of the presheaf HomDVS (V, W) on a man-
ifold X are
HomDVS (V, W)(X) = DVS(V, C∞ (X, W)).
Note that, by definition, a section F assigns to each manifold Σ a map
FΣ : C ∞ (Σ, V) → C ∞ (Σ, C∞ (X, W)) = C ∞ (Σ × X, W),
and these maps FΣ intertwine with pullbacks along maps φ : Σ → Σ0 . To a smooth
map f : Y → X, the presheaf HomDVS (V, W) assigns the natural pullback
f ∗ : DVS(V, C∞ (X, W)) → DVS(V, C∞ (Y, W)),
6. THE RELEVANT ENRICHMENT OF DVS OVER ITSELF 281
Note that by definition, HomDVS (V, W)(∗) = DVS(V, W), so our definition has
one property necessary to provide an enrichment of DVS. We now see that it
naturally lifts to an element DVS.
6.2.2 Lemma. For any V, W ∈ DVS, the presheaf HomDVS (V, W) is a differentiable
vector space.
Proof. First, we explain why HomDVS (V, W) is a sheaf. Let U = {Ui } be a
cover of the manifold X in the site Mfld. Then, for any manifold Σ, there is a
natural product cover {Σ × Ui } of Σ × X, and
Y Y
C ∞ (Σ × X, W) lim C ∞ (Σ × Ui , W) ⇒ C ∞ (Σ × (Ui ∩ U j ), W) ,
i i, j
We want to show that HomDVS provides differentiable vector spaces with the
structure of a category. Thus, we need to explain how to compose such “families
of maps;” the definition is natural but somewhat elaborate upon first encounter.
Let U, V, W be differentiable vector spaces, and let X be a manifold. There is a
natural composition
•X : DVS(V, C∞ (X, W)) × DVS(U, C∞ (X, V)) → DVS(U, C∞ (X, W))
constructed as follows. Let G ∈ DVS(V, C∞ (X, W)) and F ∈ DVS(U, C∞ (X, V)),
and recall that FΣ : C ∞ (Σ, U) → C ∞ (Σ × X, V) denote how F acts on sections for
the manifold Σ. For each input manifold Σ, set
(G •X F)Σ = (idΣ ×∆X )∗ ◦ GΣ×X ◦ FΣ : C ∞ (Σ, V) → C ∞ (Σ × X, W),
where ∆X : X → X × X is the diagonal map. Observe that for any map φ : Σ → Σ0 ,
we have
(G •X F)Σ ◦ φ∗U = φ∗W ◦ (G •X F)Σ0 ,
as each constituent of (G •X F)Σ commutes with pullbacks. Hence, •X is well-
defined. Moreover, by construction, •X is C ∞ (X)-bilinear, and •X respects the flat
connections, since its constituents do.
6. THE RELEVANT ENRICHMENT OF DVS OVER ITSELF 283
To see that this operation • is well-defined, observe that for any map f : Y → X
of manifolds, we find
( f ∗G) •Y ( f ∗ F) = f ∗ ◦ (G •X F)
as each constituent plays naturally with pullbacks such as f ∗ : DVS(U, C∞ (X, W)) →
DVS(U, C∞ (Y, W)). Hence • defines a map of presheaves, in fact, a bilinear map
of differentiable vector spaces.
6.2.4 Lemma. This composition operation is associative:
− • (− • −) = (− • −) • −.
Proof. We need to show that for a manifold X and morphisms F ∈ DVS(T, C∞ (X, U)),
G ∈ DVS(U, C∞ (X, V)), and H ∈ DVS(V, C∞ (X, W)),
H •X (G •X F) = (H •X G) •X F.
Fix an input manifold Σ. Observe that (H •X (G •X F))Σ is given by
(idΣ ×∆X )∗ ◦ HΣ×X ◦ ((idΣ ×∆X )∗ ◦ GΣ×X ◦ FΣ )
and that ((H •X G) •X F)Σ is given by
(idΣ ×∆X )∗ ◦ ((idΣ×X ×∆X )∗ ◦ H(Σ×X)×X ◦ GΣ×X ) ◦ FΣ .
Using associativity of (ordinary) composition, we see that it suffices to show
HΣ×X ◦ (idΣ ×∆X )∗ = (idΣ×X ×∆X )∗ ◦ H(Σ×X)×X ,
but this equality holds because H is a map of sheaves on manifolds, and hence
intertwines the pullback of sections.
In consequence, we obtain the desired result.
6.2.5 Theorem. There is a category DVS whose objects are differentiable vector
spaces and in which the differentiable vector space of morphisms from V to W is
HomDVS (V, W). As
C ∞ (∗, HomDVS (V, W)) = DVS(V, W),
we recover the category DVS via the global sections functor.
It is also possible to enrich the multicategory structure of DVS. We define
HomDVS (V1 , . . . , Vn |W) as the differentiable vector space whose sections on X are
C ∞ (X, HomDVS (V1 , . . . , Vn |W)) = DVS(V1 , . . . , Vn |C∞ (X, W)),
which is analogous to our definition of the enriched hom.
284 B. FUNCTIONAL ANALYSIS
Let us now analyze the smooth maps from a manifold M to C ∞ (V, W). We
will repeatedly use the fact that the functor sending a convenient vector space E to
C ∞ (M, E) preserves limits. (This claim follows from combining Lemma 3.8 with
Lemma 3.11 of Kriegl and Michor (1997) and using the fact that limits commute.)
We compute
C ∞ (M, C ∞ (V, W)) = C ∞ (M, lim C ∞ (R, W))
γ∈C ∞ (R,V)
= lim C ∞ (M, C ∞ (R, W))
γ∈C ∞ (R,V)
= lim C ∞ (M × R, W)
γ∈C ∞ (R,V)
= lim C ∞ (R, C ∞ (M, W))
γ∈C ∞ (R,V)
= C (V, C (M, W)).
∞ ∞
(Alternatively, one can simply cite the exponential law, Theorem 3.12 of Kriegl
and Michor (1997).) Now the linear maps are a closed subset of the smooth maps
CVS(V, W) ⊂ C ∞ (V, W)
cut out by linear equations that enforce linearity of the smooth maps. (See the proof
of lemma 5.2.2.) It follows that a smooth map from M to CVS(V, W) is a smooth
map from M to C ∞ (V, W) that is pointwise in CVS(V, W). Similarly an element of
C ∞ (V, C ∞ (M, W)) is linear if for each m ∈ M the corresponding map from V to W
is linear. It follows that
C ∞ (M, CVS(V, W)) = CVS(V, C ∞ (M, W))
as desired.
Let us now explain an application of this result that we use repeatedly in this
book. Viewing E as a vector space, it is natural to try to define the algebra of
functions on E as something like the symmetric algebra Sym E ∗ on the dual to
E . We need to be careful about what we mean by taking the symmetric algebra,
however, so we give our preferred definition.
7.1.4 Definition. For E the convenient vector space of smooth sections of a vector
bundle E, the uncompleted algebra of functions on E is the convenient vector space
∞
M
CVS(E b⊗β n , R)S n .
n=0
The subscript indicates that one takes the coinvariants with respect to the natural
action of the symmetric group S n on the n-fold tensor product.
This vector space is convenient. “Uncompleted” here refers to using the direct
sum rather than product; it is the uncompleted symmetric algebra — namely, poly-
nomials — rather than the completed symmetric algebra — namely, formal power
series.
We will often work with the completed algebra of functions
∞
Y
O(E ) := CVS(E b⊗β n , R)S n ,
n=0
1. Introduction
In the study of field theories, one works with vector spaces of an analytical
nature, like the space of smooth functions or distributions on a manifold. To use
the Batalin-Vilkovisky formalism, we need to perform homological algebra in this
setting. The standard approach to working with objects of this nature is to treat
them as topological vector spaces, but it is not obvious how to set up a well-behaved
version of homological algebra with topological vector spaces.
Our approach here breaks the problem into two steps. First, our cochain com-
plexes are constructed out of very nice topological vector spaces that are already
convenient vector spaces. Hence, we view them as cochain complexes of conve-
nient vector spaces, since CVS is a better-behaved category (for our purposes) than
LCTVS. Second, we apply the functor di fc to view them as cochain complexes of
differentiable vector spaces; as CVS is a full subcategory of DVS, nothing dras-
tic has happened. The benefit, however, is that DVS is a Grothendieck abelian
category, so that standard homological algebra applies immediately.
1.1. Motivation. There are a few important observations to make about this
approach.
Recall that di fc does not preserve cokernels, so that di fc need not preserve
cohomology. Given a complex C ∗ in CVS, the cohomology group H k C ∗ is a cok-
ernel computed in CVS. Hence H k (di fcC ∗ ) could be different from di fc (H k C ∗ ).
In consequence, di fc need not preserve quasi-isomorphisms. We will view quasi-
isomorphisms as differentiable cochain complexes as the correct notion and avoid
discussing quasi-isomorphisms as convenient cochain complexes.
The functor di fc does preserve cochain homotopy equivalences, however. Thus,
certain classical results – such as the Atiyah-Bott lemma (see Section D) or the use
of partitions of unity (see Section 4.4) – play a crucial role for us. They establish
explicit cochain homotopy equivalences for convenient cochain complexes, which
go to cochain homotopy equivalences of differentiable cochain complexes. Later
constructions, such as the observables of BV theories, involve deforming the differ-
entials on these differentiable cochain complexes. In almost every situation, these
deformed cochain complexes are filtered in such a way that the first page of the
spectral sequence is the original, undeformed differential. Thus we can leverage
the classical result in the new deformed situation.
289
290 C. HOMOLOGICAL ALGEBRA IN DIFFERENTIABLE VECTOR SPACES
In short, many constructions in geometry and analysis are already done in CVS
(or cochain complexes thereof), but the homological algebra is better done in DVS.
We should explain why it’s better to do the algebra in DVS.
As discussed in Appendix B, convenient vector spaces can be understood via
their associated sheaves on the site of smooth manifolds. (In particular, morphisms
between convenient vector spaces are precisely the maps as such sheaves of vector
spaces.) In other words, doing convenient linear algebra is equivalent to thinking
about smooth families of points in the convenient spaces.
For instance, in thinking about a differential operator D : E → F, one can think
about smoothly varying the inputs and studying the associated family of outputs.
It is natural then to understand the kernel by thinking about smooth families of
solutions to Dφ = 0. As di fc preserves kernels, this way of thinking agrees with
simply picking out the vector subspace annihilated by D.
These ways of thinking differ when it comes to the cokernel. In DVS, one
studies the cokernel of D by thinking about when two smooth families in F differ
by a family of outputs from D. The cokernel in DVS just amounts to studying such
equivalence relations locally, i.e., in very small smooth families. In other words,
we patch together equivalence classes; we need to sheafify the cokernel as pre-
cosheaves. This approach reflects our initial impulse to study something by vary-
ing inputs and outputs. From this perspective, the cokernel in CVS is constructed
by a more opaque procedure. The cokernel in DVS maps to the cokernel in CVS,
viewed as a differentiable vector space, so the two approaches communicate.
Remark: For another approach to these issues, see Wallbridge (n.d.), where a monoidal
combinatorial model category structure is put on the category Ch(CVS) of un-
bounded chain complexes of convenient vector spaces. ^
1.2. Outline of the appendix. After establishing foundational properties of
DVS, we develop some homological techniques that we will use repeatedly through-
out the text. Most are simple reworkings of standard results of homological alge-
bra with sheaves. First, we will discuss spectral sequences, which we often use to
verify some map of differentiable cochain complexes is a quasi-isomorphism. Sec-
ond, we discuss the category of differentiable pro-cochain complexes, which is a
somewhat technical definition but which plays a crucial role for observables of in-
teracting theories. Finally, we will discuss homotopy colimits and explicit methods
for constructing them, which we use in our definition of factorization algebras.
3. Spectral sequences
We will often use versions of spectral sequence arguments in the category of
differentiable complexes.
3.0.1 Definition. A filtered differentiable cochain complex is a sequence of differ-
entiable cochain complexes
· · · → Fn−1C → FnC → Fn+1C → · · ·
where each map Fn−1C → FnC is a monomorphism (explicitly, a monomorphism
in each cohomological degree). A map of filtered differentiable cochain complexes
is a sequence of cochain maps { fn : FnC → Fn D}n∈Z that commute with the maps
of each sequence.
Recall the following useful theorem. (For a proof and further discussion, see
Eilenberg and Moore (1962).)
3.0.3 Theorem (Eilenberg-Moore Comparison Theorem). Let f : C → D be a
map of complete filtered cochain complexes in an AB4 abelian category. Suppose
for each integer n, there exists an integer p such that F pC n = 0 = F p Dn (i.e., in
each cohomological degree n, the filtration is bounded below).
If there is a natural number r such that the induced map between the rth pages
of the spectral sequences
pq pq pq
fr : E r C → E r D
3. SPECTRAL SEQUENCES 293
Proof. Under the hypothesis that the towers are eventually constant, we know
that the associated filtrations F• V and F• W are bounded below. Under the hypothe-
sis that they are bounded, the associated filtrations are exhaustive: V = colimn Fn V
and W = colimn Fn W.
The map f• then induces a map of filtered differentiable cochain complexes b f :
V → W. As Fn+1 V/Fn V Ker vn and likewise for the filtration on W, the spectral
sequence argument above implies that b f is a quasi-isomorphism. But lim f = bf , so
we are finished.
4.1. Colimits. Colimits are not just given by colimits as differentiable cochain
complexes: one must complete the “naive” colimit. To be explicit, there is an
inclusion functor of differentiable pro-cochain complexes into negatively-filtered
differentiable cochain complexes (i.e., filtered objects such that Fn V = F0 V for all
positive n). This inclusion functor is right adjoint to a “completion functor” that
sends a filtration
· · · ,→ Fn V ,→ Fn+1 V ,→ · · · ,→ F0 V = V
V = limn V/Fn V equipped with the filtration Fn b
to its completion b V = ker(b V →
V/Fn V). Note that completion is “idempotent” in the sense that a complete filtered
cochain complex is isomorphic to its completion.
We define the completed colimit of a diagram of differentiable pro-cochain
complexes as the completion of the colimit in the category of filtered differentiable
cochain complexes. To emphasize the role of completion, we will sometimes write
[ For example, we have the following.
colim.
4. DIFFERENTIABLE PRO-COCHAIN COMPLEXES 295
where on the right hand side we use the direct sum of differentiable cochain com-
plexes.
5. Homotopy colimits
A factorization algebra is, in particular, a cosheaf with respect to the Weiss
topology. When working with precosheaves in cochain complexes, the conceptu-
ally correct version of the cosheaf gluing axiom uses the homotopy colimit rather
than the usual colimit. Explicitly, a precosheaf F of cochain complexes is a cosheaf
if for any open set V and any Weiss cover U = {Ui }i∈I of V, the canonical map
hocolimČU F → F (V)
is a quasi-isomorphism, where the left hand side denotes the homotopy colimit
over the Čech cover of U.
In our situation, the homotopy colimit of a diagram F : I → Ch(DVS) can
be described by a familiar cochain complex, as we explain below. In the case of
the cosheaf gluing axiom, we recover precisely the formula that appears in the
definition of a factorization algebra (see Section 1).
We begin by discussing a general situation that includes Ch(DVS) as a special
case. In particular, we explain the homotopical universal property characterizing a
homotopy colimit and then state a theorem providing an explicit construction of a
homotopy colimit in this situation. In the next subsection, we give the proof of this
theorem. Finally, we treat the case of differentiable pro-cochain complexes.
Remark: The Čech complex we use is long-established in homological algebra and
easy to motivate. The reader satisfied with it should probably look no further, as the
arguments below are highly technical. The goal of this section, though, is to ensure
compatibility between our work in this book and the perspective emphasized by
higher category theory. It bridges an odd gap in the literature between homological
and homotopical algebra. ^
5.1. Reminder on homotopy colimits. Let A denote an abelian category, and
let A = Ch(A) denote the category of unbounded cochain complexes in A. Let I
be a category and let F : I → A be a functor, which we will call a diagram of
cochain complexes (or I-diagram to emphasize the shape of the diagram).
5. HOMOTOPY COLIMITS 297
In other words, the homotopy colimit of a diagram satisfies, at the level of ho-
motopy categories, a version of adjunction in the global definition of colimit. It
is the initial cochain complex, up to quasi-isomorphism, among all cochain com-
plexes, up to quasi-isomorphism, that receive a map from the diagram, up to quasi-
isomorphism of diagrams.
5.1.2. In Section 5.3 below, we describe the ∞-category of cochain com-
plexes in several ways, as a model category and as a quasi-category, following
Lurie. Given a diagram F : I → C in an ∞-category, there is an ∞-category of
cocones for F, which Lurie denotes as CF/ . A cocone is essentially a homotopy co-
herent version of the notion of cocone given in ordinary categories. A colimit for F
is then an initial object in CF/ , which means that any cocone admits a contractible
space of maps from the colimit. (Having a contractible space of morphisms is the
∞-categorical replacement of the notion of unique map.) This ∞-categorical col-
imit is often called the homotopy colimit to distinguish it from the 1-categorical
colimit, when there is the possibility of confusion. See section 1.2.13 of Lurie
(2009a) for the beginning of a systematic discussion.
where [p] denotes the category of the totally ordered set {0 < · · · < p} and σ varies
over all functors from [p] to I, the vertical differential (in the q-grading) is simply
the internal differentials of the complexes F(i), and the horizontal differential is
dhor : C I (F)−p,q → C I (F)1−p,q
Pp
i=0 (−1) (F(σ ◦ δi (0)))σ◦δi ,
(x)σ 7→ i
Remark: When I = ∆op , ConeI (F) is precisely the direct sum totalization of the
double complex that the Dold-Kan correspondence produces out of the simplicial
cochain complex F. ^
Our goal is to prove the following result (indeed, its ∞-categorical strengthen-
ing).
5. HOMOTOPY COLIMITS 299
5.4. The ∞-categorical version of the theorem. Our goal is to find a way of
describing colimits in the ∞-category D(A) in a concrete way. We will see that
the construction ConeI from above provides this colimit.
5.4.1. As an example for the kind of result we want, we review the following
construction of the homotopy limit of a diagram of simplicial sets, since we will
use it in our proof.
302 C. HOMOLOGICAL ALGEBRA IN DIFFERENTIABLE VECTOR SPACES
The twisted arrow category Tw(C) of a category C has objects the morphisms
in C but a morphism from f ∈ C(x, y) to f 0 ∈ C(x0 , y0 ) is a commuting diagram
X / X0 .
f f0
Yo Y0
Note that there is a natural functor from Tw(C) to C × Cop sending an object to its
source and target. Given a functor F : C × Cop → D, the end of F is the limit of
the composition Tw(C) → C × Cop → D. (The coend is the colimit.)
A cosimplicial simplicial set is a functor X • : ∆ → sSet. A key example is
∆ whose n-cosimplex is precisely the simplicial set ∆[n]. It is a kind of fattened
•
model of a point.
Recall that the totalization of a cosimplicial object X • : ∆ → sSet is the end of
sSet∆ (∆• , X • ), where sSet∆ will denote the internal hom of sSet. This construction
is dual to the more familiar geometric realization. (Geometric realization assem-
bles a space by attaching cells, so it is a colimit. Totalization assembles a space as
a tower of extensions, so it is a limit.)
5.4.1 Definition. Let F : I → sSet be a diagram of simplicial sets. The cosimpli-
cial cobar construction B̌• F is the cosimplicial simplicial set whose n-cosimplices
are Y
B̌n F = F(σ(n)),
σ:[n]→I
where σ runs over all functors from the poset [n] into I. The cobar construction
B̌(F) is the totalization of B̌• F.
Remark: The cosimplicial cobar construction keeps track, in a very explicit sense,
of all the ways that a given object F(i) of the diagram is mapped into. It remem-
bers every finite sequence of arrows whose terminus is F(i). The cobar construction
then looks at ways of coherently mapping a “fat point” into this enormous arrange-
ment. ^
Here is the statement at the level of homotopy categories.
5.4.2 Proposition (Bousfield and Kan (1972), XI.8.1). Let F : I → sSet be a
diagram such that F(i) is a Kan complex for every i ∈ I (i.e., F is objectwise
fibrant). Then B̌(F) is a representative for the homotopy limit holimI F in Ho(sSet).
There is a corresponding ∞-categorical statement. Let S denote the ∞-category
of spaces, which is N(sSet◦ ), i.e., the homotopy coherent nerve of sSet◦ , the cate-
gory of Kan complexes, which is the subcategory of fibrant objects in sSet. As an
immediate corollary of the proceeding proposition and Theorem 4.2.4.1 of Lurie
(2009a), we have the following.
5.4.3 Proposition. Let F : I → sSet◦ be a diagram. Then B̌(F) is a limit of the
associated diagram in spaces S.
5. HOMOTOPY COLIMITS 303
5.4.2. We will parallel this development to show that our cone construction
ConeI is a kind of bar construction and hence a colimit in ∞-categories.
Above, we used the fact that a diagram of simplicial sets naturally produced a
cosimplicial simplicial set, and we could then talk about mapping the “fat point”
into this cosimplicial cobar construction. Here, we will start with a diagram of
cochain complexes in A and then produce a simplicial cochain complex in A. To
do a bar construction, we need to find a replacement for the “fat point” ∆• in the
setting of cochain complexes.
We will use 4n to denote the cochain complex N∗ ∆[n] ∈ Ch, the natural
cochain complex associated to the simplicial set ∆[n] by the Dold-Kan correspon-
dence. There is thus a natural cosimplicial cochain complex 4• that provides our
“fat point.” Given a simplicial cochain complex X• : ∆op → A = Ch(A), we thus
have a functor
X• ⊗ 4• : ∆op × ∆ → A,
using the fact that A is tensored over Ch. The realization of X• is the colimit of the
composite functor Tw(∆op ) → ∆op × ∆ → A.
5.4.4 Definition. Let F : I → A be a diagram in A. The simplicial bar construction
B• F is the simplicial object in A whose n-simplices are
M
Bn F = F(σ(0)),
σ:[n]→I
where σ runs over all functors from the poset [n] into I. The bar construction B(F)
is the realization of B• F.
Indeed, this property characterizes the homotopy colimit of F. (See Remark A.3.3.13
of Lurie (2009a).)
Observe that for any simplicial set K, we have natural isomorphisms
for every m and n. Applying Dold-Kan thus produces a simplicial homotopy equiv-
alence
A∆ (∆[n] ⊗ Bm F, Z) sSet∆ (∆[n], A∆ (Bm F, Z))
for every m and n. Taking the limit over the twisted arrow category, we obtain a
simplicial homotopy equivalence
and hence a weak equivalence of simplicial sets. The right hand side is isomorphic
to the cobar construction,
5.6.1. Sequences, filtrations, and completions. Let Z≤0 denote the category
whose objects are the nonpositive integers {n ≤ 0} with a single morphism n → m if
n ≤ m and no morphism if n > m. Let Seq denote the functor category Fun(Z≤0 , A).
We call an object
· · · X(n) → X(n + 1) → · · · → X(0)
in Seq a sequence. Equip Seq with the projective model structure: a map of se-
quences f : X → Y is
• a weak equivalence if f (n) : X(n) → Y(n) is a quasi-isomorphism for
every n, and
• a fibration if f (n) : X(n) → Y(n) is a fibration for every n.
Note that we are working with levelwise weak equivalences, which we denote WL .
The following observation gives a nice explanation for the role of filtered
cochain complexes: they are the cofibrant sequences.
5.6.1 Lemma. A sequence X is cofibrant if and only if every map X(n) → X(n + 1)
is a cofibration (i.e., a monomorphism in every cohomological degree).
Hence, we introduce the following definition.
5.6.2 Definition. Let A f il denote the category Seqc of cofibrant objects, i.e., the
filtered cochain complexes.
Proof. We can view this diagram as living in A f il (and hence also the category
Seq), and by Lemma 5.6.4, it suffices to compute the colimit in A f il [WF−1 ]. We
will thus verify the (uncompleted) cone construction provides a model for this col-
imit. Completing this cone will provide a complete filtered cochain complex that
is filtered weak equivalent, and hence also a model for the desired colimit.
We compute this colimit in filtered cochain complexes by using model cate-
gories. Let SeqL denote Seq equipped with the projective model structure, so the
weak equivalences are WL . Observe that WL ⊂ WF as collections of morphisms
inside Seq. Then the left Bousfield localization at WF , which we will denote SeqF ,
has the same cofibrations, by definition. Hence, if one wants to compute homotopy
colimits in SeqF , they agree with the homotopy colimits in SeqL . (More explicitly,
the projective model structures on Fun(I, SeqF ) and Fun(I, SeqL ) have the same
cofibrant replacement functor, so homotopy colimits can be computed in either
category.) In SeqL , our computation becomes simple: a homotopy colimit of di-
agram in such a functor category can be computed objectwise, so we can apply
Theorem 5.4.6 objectwise. This construction naturally produces ConeI with the
natural filtration.
APPENDIX D
Atiyah and Bott (1967) shows that for an elliptic complex (E , Q) on a compact
closed manifold M, with E the smooth sections of a Z-graded vector bundle, there
is a homotopy equivalence (E , Q) ,→ (E , Q) into the elliptic complex of distribu-
tional sections. The argument follows from the existence of parametrices for el-
liptic operators. This result was generalized to the non-compact case in Tarkhanov
(1987).
Let M be a smooth manifold, not necessarily compact. Let (E , Q) be an elliptic
complex on M. Let E denote the complex of distributional sections of E . We will
endow both E and E with their natural topologies.
0.0.1 Lemma (Lemma 1.7, Tarkhanov (1987)). The natural inclusion (E , Q) ,→
(E , Q) admits a continuous homotopy inverse.
The continuous homotopy inverse
Φ:E →E
is given by a kernel KΦ ∈ E ! ⊗ E with proper support. The homotopy S : E → E
is a continuous linear map with
[Q, S ] = Φ − Id .
!
The kernel KS for S is a distribution, that is, an element of E ⊗ E with proper
support.
Adámek, Jiřı́, and Rosický, Jiřı́. 1994. Locally presentable and accessible cate-
gories. London Mathematical Society Lecture Note Series, vol. 189. Cambridge
University Press, Cambridge.
Atiyah, M., and Bott, R. 1967. A Lefschetz Fixed Point Formula for Elliptic Com-
plexes: I. Ann. of Math. (2), 86(2), 374–407.
Ayala, David, and Francis, John. 2015. Factorization homology of topological
manifolds. J. Topol., 8(4), 1045–1084.
Ayala, David, Francis, John, and Tanaka, Hiro Lee. Factorization homology for
stratified manifolds. Available at http://arxiv.org/abs/1409.0848.
Ayala, David, Francis, John, and Rozenblyum, Nick. Factorization homology from
higher categories. Available at http://front.math.ucdavis.edu/1504.
04007.
Behnke, Heinrich, and Stein, Karl. 1949. Entwicklung analytischer Funktionen auf
Riemannschen Flächen. Math. Ann., 120, 430–461.
Beilinson, Alexander, and Drinfeld, Vladimir. 2004. Chiral algebras. American
Mathematical Society Colloquium Publications, vol. 51. Providence, RI: Amer-
ican Mathematical Society.
Ben Zvi, David, Brochier, Adrien, and Jordan, David. Integrating quantum groups
over surfaces: quantum character varieties and topological field theory. Avail-
able at http://arxiv.org/abs/1501.04652.
Boardman, J. M., and Vogt, R. M. 1973. Homotopy invariant algebraic structures
on topological spaces. Lecture Notes in Mathematics, Vol. 347. Springer-Verlag,
Berlin-New York.
Boavida de Brito, Pedro, and Weiss, Michael. 2013. Manifold calculus and homo-
topy sheaves. Homology Homotopy Appl., 15(2), 361–383.
Bödigheimer, C.-F. 1987. Stable splittings of mapping spaces. Pages 174–187 of:
Algebraic topology (Seattle, Wash., 1985). Lecture Notes in Math., vol. 1286.
Springer, Berlin.
Borceux, Francis. 1994a. Handbook of categorical algebra. 1. Encyclopedia of
Mathematics and its Applications, vol. 50. Cambridge: Cambridge University
Press. Basic category theory.
Borceux, Francis. 1994b. Handbook of categorical algebra. 2. Encyclopedia of
Mathematics and its Applications, vol. 51. Cambridge University Press, Cam-
bridge. Categories and structures.
Bott, Raoul, and Tu, Loring W. 1982. Differential forms in algebraic topology.
Graduate Texts in Mathematics, vol. 82. Springer-Verlag, New York-Berlin.
309
310 Bibliography
Bousfield, A. K., and Kan, D. M. 1972. Homotopy limits, completions and local-
izations. Lecture Notes in Mathematics, Vol. 304. Springer-Verlag, Berlin-New
York.
Calaque, Damien, and Willwacher, Thomas. 2015. Triviality of the higher formal-
ity theorem. Proc. Amer. Math. Soc., 143(12), 5181–5193.
Costello, Kevin. The A∞ operad and the moduli space of curves. Available at
http://arxiv.org/abs/math/0402015.
Costello, Kevin. Renormalisation and the Batalin-Vilkovisky formalism. Available
at http://arxiv.org/abs/0706.1533.
Costello, Kevin. Supersymmetric gauge theory and the Yangian. Available at
http://arxiv.org/abs/1303.2632.
Costello, Kevin. 2007. Topological conformal field theories and Calabi-Yau cate-
gories. Adv. Math., 210(1), 165–214.
Costello, Kevin. 2010. A geometric construction of the Witten genus, I. In: Pro-
ceedings of the International Congress of Mathematicians, (Hyderabad, 2010).
Costello, Kevin. 2011a. A geometric construction of the Witten genus, II. Available
at http://arxiv.org/abs/1112.0816.
Costello, Kevin. 2011b. Renormalization and effective field theory. Mathematical
Surveys and Monographs, vol. 170. American Mathematical Society, Provi-
dence, RI.
Costello, Kevin. 2013. Notes on supersymmetric and holomorphic field theories in
dimensions 2 and 4. Pure Appl. Math. Q., 9(1), 73–165.
Costello, Kevin. 2014. Integrable lattice models from four-dimensional field the-
ories. Pages 3–23 of: String-Math 2013. Proc. Sympos. Pure Math., vol. 88.
Amer. Math. Soc., Providence, RI.
Costello, Kevin, and Li, Si. 2011. Quantum BCOV theory on Calabi-Yau man-
ifolds and the higher genus B-model. Available at http://arxiv.org/abs/
1201.4501.
Costello, Kevin, and Scheimbauer, Claudia. 2015. Lectures on mathematical as-
pects of (twisted) supersymmetric gauge theories. Pages 57–87 of: Mathemati-
cal aspects of quantum field theories. Math. Phys. Stud. Springer, Cham.
Curry, Justin. Sheaves, Cosheaves and Applications. Available at http://arxiv.
org/abs/1303.3255.
Deligne, Pierre, Etingof, Pavel, Freed, Daniel S., Jeffrey, Lisa C., Kazhdan, David,
Morgan, John W., Morrison, David R., and Witten, Edward (eds). 1999. Quan-
tum fields and strings: a course for mathematicians. Vol. 1, 2. American Mathe-
matical Society, Providence, RI; Institute for Advanced Study (IAS), Princeton,
NJ. Material from the Special Year on Quantum Field Theory held at the Insti-
tute for Advanced Study, Princeton, NJ, 1996–1997.
Dugger, Daniel. A primer on homotopy colimits. Available at http://pages.
uoregon.edu/ddugger/.
Dwyer, W. G., and Kan, D. M. 1980a. Calculating simplicial localizations. J. Pure
Appl. Algebra, 18(1), 17–35.
Dwyer, W. G., and Kan, D. M. 1980b. Simplicial localizations of categories. J.
Pure Appl. Algebra, 17(3), 267–284.
Bibliography 311
Eilenberg, Samuel, and Moore, John C. 1962. Limits and spectral sequences.
Topology, 1, 1–23.
Etingof, Pavel, Gelaki, Shlomo, Nikshych, Dmitri, and Ostrik, Victor. 2015. Ten-
sor categories. Mathematical Surveys and Monographs, vol. 205. American
Mathematical Society, Providence, RI.
Félix, Yves, Halperin, Stephen, and Thomas, Jean-Claude. 2001. Rational homo-
topy theory. Graduate Texts in Mathematics, vol. 205. Springer-Verlag, New
York.
Forster, Otto. 1991. Lectures on Riemann surfaces. Graduate Texts in Mathematics,
vol. 81. New York: Springer-Verlag. Translated from the 1977 German original
by Bruce Gilligan, Reprint of the 1981 English translation.
Francis, John. 2013. The tangent complex and Hochschild cohomology of En -
rings. Compos. Math., 149(3), 430–480.
Freed, Daniel S. 1994. Higher algebraic structures and quantization. Comm. Math.
Phys., 159(2), 343–398.
Frenkel, Edward, and Ben-Zvi, David. 2004. Vertex algebras and algebraic curves.
Second edn. Mathematical Surveys and Monographs, vol. 88. Providence, RI:
American Mathematical Society.
Fresse, Benoit. Homotopy of Operads and the Grothendieck-Teichmüller
group. Available at http://math.univ-lille1.fr/˜fresse/
OperadHomotopyBook/.
Friedman, Greg. 2012. Survey article: an elementary illustrated introduction to
simplicial sets. Rocky Mountain J. Math., 42(2), 353–423.
Gelfand, Sergei I., and Manin, Yuri I. 2003. Methods of homological algebra.
Second edn. Springer Monographs in Mathematics. Berlin: Springer-Verlag.
Getzler, E. 1994. Batalin-Vilkovisky algebras and two-dimensional topological
field theories. Comm. Math. Phys., 159(2), 265–285.
Ginot, Grégory. 2015. Notes on factorization algebras, factorization homology and
applications. Pages 429–552 of: Mathematical aspects of quantum field theories.
Math. Phys. Stud. Springer, Cham.
Ginot, Grégory, Tradler, Thomas, and Zeinalian, Mahmoud. Derived Higher
Hochschild cohomology, Brane topology and centralizers of En -algebra maps.
Available at http://arxiv.org/abs/1205.7056.
Ginot, Grégory, Tradler, Thomas, and Zeinalian, Mahmoud. Derived Higher
Hochschild Homology, Topological Chiral Homology and Factorization Alge-
bras. Available at http://arxiv.org/abs/1011.6483.
Glimm, James, and Jaffe, Arthur. 1987. Quantum physics. Second edn. Springer-
Verlag, New York. A functional integral point of view.
Goerss, Paul, and Schemmerhorn, Kristen. 2007. Model categories and simplicial
methods. Pages 3–49 of: Interactions between homotopy theory and algebra.
Contemp. Math., vol. 436. Providence, RI: Amer. Math. Soc.
Goerss, Paul G., and Jardine, John F. 2009. Simplicial homotopy theory. Modern
Birkhäuser Classics. Basel: Birkhäuser Verlag. Reprint of the 1999 edition
[MR1711612].
312 Bibliography
Grady, Ryan, Li, Qin, and Li, Si. BV quantization and the algebraic index. Avail-
able at http://arxiv.org/abs/1507.01812.
Griffiths, Phillip, and Harris, Joseph. 1994. Principles of algebraic geometry. Wi-
ley Classics Library. New York: John Wiley & Sons Inc. Reprint of the 1978
original.
Grothendieck, A. 1952. Résumé des résultats essentiels dans la théorie des produits
tensoriels topologiques et des espaces nucléaires. Ann. Inst. Fourier Grenoble,
4, 73–112 (1954).
Grothendieck, Alexander. 1957. Sur quelques points d’algèbre homologique.
Tôhoku Math. J. (2), 9, 119–221.
Grothendieck, Alexandre. 1955. Produits tensoriels topologiques et espaces
nucléaires. Mem. Amer. Math. Soc., 1955(16), 140.
Gunning, Robert C., and Rossi, Hugo. 1965. Analytic functions of several complex
variables. Englewood Cliffs, N.J.: Prentice-Hall Inc.
Gwilliam, Owen. 2012. Factorization Algebras and Free Field Theories. ProQuest
LLC, Ann Arbor, MI. Thesis (Ph.D.)–Northwestern University.
Gwilliam, Owen, and Grady, Ryan. 2014. One-dimensional Chern-Simons theory
and the  genus. Algebr. Geom. Topol., 14(4), 2299–2377.
Gwilliam, Owen, and Johnson-Freyd, Theo. How to derive Feynman diagrams
for finite-dimensional integrals directly from the BV formalism. Available at
http://arxiv.org/abs/1202.1554.
Hatcher, Allen. Notes on Basic 3-Manifold Topology. Available at http://www.
math.cornell.edu/˜hatcher/.
Haugseng, Rune. Weakly simplicial model categories and homotopy colimits.
Available upon request.
Horel, Geoffroy. Factorization homology and calculus à la Kontsevich Soibelman.
Available at http://arxiv.org/abs/1307.0322.
Hörmander, Lars. 2003. The analysis of linear partial differential operators.
I. Classics in Mathematics. Berlin: Springer-Verlag. Distribution theory
and Fourier analysis, Reprint of the second (1990) edition [Springer, Berlin;
MR1065993 (91m:35001a)].
Hörmann, Fritz. Homotopy Limits and Colimits in Nature: A Motivation
for Derivators. Available at http://home.mathematik.uni-freiburg.de/
hoermann/holimnature.pdf.
Huang, Yi-Zhi. 1997. Two-dimensional conformal geometry and vertex operator
algebras. Progress in Mathematics, vol. 148. Boston, MA: Birkhäuser Boston
Inc.
Johansen, A. 1995. Twisting of N = 1 SUSY gauge theories and heterotic topo-
logical theories. Internat. J. Modern Phys. A, 10(30), 4325–4357.
Kapustin, Anton. 2010. Topological field theory, higher categories, and their ap-
plications. Pages 2021–2043 of: Proceedings of the International Congress of
Mathematicians. Volume III. Hindustan Book Agency, New Delhi.
Kapustin, Anton, and Saulina, Natalia. 2011. Topological boundary conditions in
abelian Chern-Simons theory. Nuclear Phys. B, 845(3), 393–435.
Bibliography 313
315
316 INDEX