03.02.2015 Views

Representations and Cohomology of Finite Categories (DRAFT)

Representations and Cohomology of Finite Categories (DRAFT)

Representations and Cohomology of Finite Categories (DRAFT)

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Fei Xu<br />

<strong>Representations</strong> <strong>and</strong> <strong>Cohomology</strong> <strong>of</strong> <strong>Finite</strong><br />

<strong>Categories</strong> (<strong>DRAFT</strong>)<br />

Category Algebras & Simplicial Methods<br />

September 1, 2011


Preface<br />

These are exp<strong>and</strong>ed lecture notes that I used for a short course<br />

<strong>of</strong> the same title at Universitat Autònoma de Barcelona in<br />

2010-2011 academic year. The subtitle reveals the main tools<br />

we are using. It means that we shall study representations<br />

<strong>and</strong> cohomology <strong>of</strong> category algebras via simplicial modules.<br />

I assume the reader to have good background on homological<br />

algebra from classical books such as “A Course in Homological<br />

Algebra” by Hilton <strong>and</strong> Stammbach. The main theme <strong>of</strong> these<br />

notes is to answer the question: what we may do with an<br />

abstract finite category Here we regard a finite category as<br />

a generalization <strong>of</strong> an abstract finite group <strong>and</strong> <strong>of</strong> an abstract<br />

finite partially ordered set. Starting from a finite group <strong>and</strong><br />

a base ring, there are well developed group representation<br />

theory <strong>and</strong> group cohomology theory. Parallel theories have<br />

been introduced to finite partially ordered sets as well. This<br />

book presents a theory that extending all <strong>of</strong> the above.<br />

Category cohomology theory is a place where simplicial<br />

methods, homotopy theory <strong>and</strong> representation theory naturally<br />

meet. The objectives <strong>of</strong> our theory may be summarized<br />

in the following diagram.


vi<br />

C<br />

classifying space<br />

functor category<br />

nerve<br />

category algebra <strong>and</strong> modules<br />

BC<br />

geo.<br />

bar resolution<br />

<br />

NC<br />

RC−mod<br />

realiz.<br />

<strong>of</strong> the trivial RC−module R<br />

singular cohomology<br />

simplicial cohomology<br />

cohomology <strong>of</strong> modules<br />

H ∗ (C; N) ∼ ∗<br />

= lim ←−<br />

C N ∼ = Ext ∗ RC(R, N)<br />

Ext ∗ RC(M, N)<br />

Preface<br />

functor cohomology<br />

In the diagram, C is a finite category, NC is its nerve, a combinatorial<br />

construction, <strong>and</strong> BC is the geometric realization <strong>of</strong><br />

NC (or C) which is a CW-complex. Also R is a commutative<br />

ring with identity <strong>and</strong> M, N ∈ (R-mod) C are two covariant<br />

functors. (A functor in (R-mod) C should be considered as a<br />

diagram <strong>of</strong> R-modules.) In this picture, RC is the so-called<br />

category algebra. It was shown by B. Mitchell [56] that there<br />

exists an equivalence between (R-mod) C <strong>and</strong> RC-mod (the<br />

category <strong>of</strong> left RC-modules). Thus a functor from C to R-<br />

mod is always an RC-module. Notably when C is a group, the<br />

theory becomes the usual group cohomology theory. When C<br />

is a transporter category (see Chapter 6) defined over a group<br />

G, we recover the equivariant cohomology theory.<br />

We can similarly study category homology by replacing<br />

the last two rows in the picture by H∗(C; N) ∼ ∗<br />

= lim −→ N ∼ =<br />

C<br />

Tor RC<br />

∗ (R, N) <strong>and</strong> Tor RC<br />

∗ (M ′ , N) (M ′ is a right RC-module).<br />

In these notes we shall focus on cohomology theory because<br />

homology theory can be developed parallel to it.<br />

Much <strong>of</strong> the general theory is indeed established for all small<br />

categories. To make this book as useful as we can, we shall<br />

state results in their general forms, for small, not just finite,<br />

categories whenever it is the case. It is perhaps a good point<br />

to explain why small categories are <strong>of</strong> particular interests. In<br />

fact this is a set-theoretic issue. In the above diagram, if we<br />

<br />

(R−mod


Preface<br />

consider a small category C then we can still construct a simplicial<br />

set NC, a topological space BC, an algebra RC <strong>and</strong><br />

a functor category (R-mod) C . In all these constructions, explicitly<br />

defined in the main text, the class <strong>of</strong> morphisms in C,<br />

Mor C, has to be a set, which is precisely the smallness condition<br />

on C. It lies in the applications <strong>of</strong> category cohomology<br />

<strong>and</strong> representations that we have in mind that C is frequently<br />

finite (Mor C is a finite set). The chief reason is that we underst<strong>and</strong><br />

quite well the representations <strong>of</strong> RC, which becomes<br />

an associative algebra with identity <strong>and</strong> is <strong>of</strong> finite R-rank.<br />

There exists another functor (co)homology theory [25], used<br />

in Steenrod algebras <strong>and</strong> cohomology <strong>of</strong> finite group schemes.<br />

In that (co)homology theory, one mainly investigates representations<br />

<strong>and</strong> cohomology <strong>of</strong> the concrete (essentially small<br />

but infinite) category <strong>of</strong> finite-dimensional vector spaces over<br />

a field. The relationship between these two theories is comparable<br />

to that between the cohomology theories <strong>of</strong> abstract<br />

finite groups <strong>and</strong> <strong>of</strong> general linear groups. They start <strong>of</strong>f the<br />

same foundation but are <strong>of</strong> different flavors <strong>and</strong> use quite different<br />

methods, even though they supply important ideas <strong>and</strong><br />

results to each another. Thus although there are something in<br />

common between [25] <strong>and</strong> this book, the main bulks <strong>of</strong> these<br />

two are different <strong>and</strong> to some extent complementary to each<br />

other.<br />

The ingredients shown in the previous diagram have been<br />

studied since the early stage <strong>of</strong> homological algebra in one<br />

way or another. In this book we collect many existing materials<br />

to form a source for self-learning as well as a reference<br />

for researchers. The major inputs from the author are firstly<br />

to provide a systematic introduction to the uses <strong>of</strong> simplicial<br />

methods, <strong>and</strong> secondly to develop tools for comparing<br />

cohomology <strong>of</strong> two small categories connected by a functor.<br />

vii


viii<br />

Preface<br />

Simplicial methods help us to construct some (complexes <strong>of</strong>)<br />

modules combinatorially while the tools we have mentioned<br />

tell us how to compare these (complexes <strong>of</strong>) modules (<strong>and</strong><br />

their cohomology). We shall explain these two features next.<br />

Simplicial methods should be considered as natural tools<br />

for algebraists. If A is a ring, then a simplicial A-module M<br />

is a collection <strong>of</strong> A-modules M n , n ≥ 0, equipped with certain<br />

maps among them. By Dold-Kan Correspondence, the<br />

category <strong>of</strong> non-negatively (or non-positively) graded chain<br />

complexes <strong>of</strong> A-modules is equivalent to the category <strong>of</strong> simplicial<br />

(or cosimplicial) A-modules. It also implies the category<br />

<strong>of</strong> first quadrant double complexes is equivalent to the<br />

category <strong>of</strong> bisimplicial A-modules. The complexes <strong>of</strong> modules<br />

have long been used in (at least the homological aspects<br />

<strong>of</strong>) the representation theory <strong>of</strong> algebras, so there is no reason<br />

why we do not consider their combinatorial predecessors.<br />

In this book we shall focus on the case for A = RC, where<br />

simplicial methods seem to be truly powerful.<br />

Let u : D → C be a functor between small categories. We<br />

construct a simplicial RC-module R[N(u/−)] (a bar construction).<br />

It has the properties that, for each n ≥ 0, R[N(u/−)] n<br />

is a left projective RC-module, <strong>and</strong> that, for each x ∈ Ob C,<br />

R[N(u/x)] is a simplicial R-module. Here u/x is the category<br />

over x <strong>and</strong> N(u/x) is its nerve, a simplicial set. Much <strong>of</strong> category<br />

(co)homology is developed based on this construction.<br />

For instance, R[N(Id C /−)] gives rise to the bar resolution<br />

<strong>of</strong> the trivial RC-module R such that R[N(Id C /−)] = RC.<br />

This is the reason why simplicial methods are indispensable<br />

to us. Due to the existence <strong>of</strong> these constructions, the Kan<br />

extensions are used, as generalizations <strong>of</strong> the usual induction<br />

<strong>and</strong> co-induction on which one relies to investigate group<br />

(co)homology, to compare (co)homology <strong>of</strong> C <strong>and</strong> D. In fact,


Preface<br />

let LK u : RD-mod → RC-mod be the left Kan extension<br />

along u. Then LK u R[N(Id D /−)] ∼ = R[N(u/−)]. Accordingly<br />

underst<strong>and</strong>ing various overcategories becomes an important<br />

issue as they show up in the constructions <strong>of</strong> simplicial modules<br />

as well as in the definitions <strong>of</strong> Kan extensions. An illuminating<br />

example is as follows. Suppose G is a finite group<br />

<strong>and</strong> H is a subgroup. Consider them as categories with one<br />

object •, equipped with the inclusion i : H → G. Then the<br />

only overcategory i/• is equivalent to the set <strong>of</strong> left cosets<br />

G/H. The left Kan extension <strong>of</strong> the bar resolution <strong>of</strong> the<br />

trivial RH-module is a projective resolution <strong>of</strong> the induced<br />

module R(G/H) = R ↑ G H . As in algebraic topology where<br />

they originate <strong>and</strong> their applications confine, we shall show<br />

that simplicial methods now become crucial theoretical <strong>and</strong><br />

computational tools in algebra too.<br />

What we do here in this book is really to take up the algebraic<br />

<strong>and</strong> combinatorial tools invented by topologists <strong>and</strong><br />

apply them to algebra. We shall see that we get satisfactory results.<br />

Simplicial methods were introduced by algebraic topologists<br />

<strong>and</strong> lie in the center <strong>of</strong> homotopy theory. In homotopy<br />

theory, one <strong>of</strong>ten considers diagrams <strong>of</strong> spaces, which are simply<br />

functors from an index category C to T op, the category<br />

<strong>of</strong> topological spaces. When u : D → C <strong>and</strong> F : C → T op are<br />

two functors, one can naturally obtain a new functor F ◦ u.<br />

This is called the restriction along u, written as Res u F. Indeed<br />

Res u is a functor from the category T op C to T op D . On the<br />

other h<strong>and</strong> if G : D → T op is a functor, we can construct<br />

two functors C → T op. These two functors are called the left<br />

<strong>and</strong> right Kan extensions <strong>of</strong> G along u. It is intuitive if we<br />

record topologists’ notations<br />

C ⊗ D G <strong>and</strong> Hom D (C, G).<br />

ix


x<br />

Preface<br />

These two functors C ⊗ D −, Hom D (C, −) are the left <strong>and</strong> right<br />

adjoints <strong>of</strong> Res u . In Hollender-Vogt [37], one can see that<br />

these two functors enjoy many properties that an algebraist<br />

would expect for similar functors, induced by a ring homomorphism<br />

R 1 → R 2 , between R 1 -mod <strong>and</strong> R 2 -mod. In fact<br />

in [37], a covariant functor in T op C is called a left C-module<br />

while a contravariant functor is called a right C-module. I<br />

would like to point out here that since a functor u : D → C<br />

does not lead to an algebra homomorphism from RD to RC,<br />

one has to be very careful when constructing RC ⊗ RD − <strong>and</strong><br />

Hom RD (RC, −). This is the reason why I refrain from using<br />

topologists’ notations <strong>of</strong> the Kan extensions. Plus the intuitive<br />

notations do not seems to be convenient in computations<br />

as they are in other places such as group (co)homology.<br />

However the Kan extensions are truly generalizations <strong>of</strong> well<br />

known functors. When i : H → G is the inclusion functor at<br />

the end <strong>of</strong> last paragraph, the left <strong>and</strong> right Kan extensions<br />

along i are isomorphic to the induction RG ⊗ RH − <strong>and</strong> coinduction<br />

Hom RH (RG, −) (which happen to be isomorphic<br />

to each other).<br />

Let us now recall the history <strong>of</strong> (co)homology theory <strong>of</strong> small<br />

categories <strong>and</strong> related works. Homology theory <strong>of</strong> small categories,<br />

H∗(C; N), can be found in Gabriel-Zisman [28] (1967),<br />

as a special case <strong>of</strong> homology <strong>of</strong> simplicial sets with coefficients,<br />

in the appendices. There they also showed H∗(C; N) ∼ =<br />

lim<br />

−→ ∗ N, where lim<br />

C −→ ∗ are the derived functors <strong>of</strong> direct limit<br />

C<br />

functor −→C<br />

lim . By contrast, although the special case <strong>of</strong> partially<br />

ordered sets was studied by J. E. Roos [66] (1961), cohomology<br />

<strong>of</strong> small categories began with Baues-Wirsching’s<br />

theory [3] (1985). They actually studied H ∗ (C; N ), where N<br />

is not a functor from C. Instead, it is a functor from another<br />

category F (C), the category <strong>of</strong> factorizations in C. In their


Preface<br />

situation cohomology theory <strong>of</strong> small category can also be<br />

∗<br />

introduced by ←−<br />

lim , derived functors <strong>of</strong> inverse limit functor<br />

C<br />

lim . Simplicial methods are used to define various chain or<br />

←−C<br />

cochain complexes whose (co)homology is the (co)homology<br />

<strong>of</strong> C.<br />

In these theories the coefficients <strong>of</strong> (co)homology are provide<br />

by functors. Thus analyzing structures <strong>of</strong> functors should be<br />

<strong>of</strong> great importance. A significant achievement in this direction<br />

was obtained by Lück [51] <strong>and</strong> tom Dieck [] (1987),<br />

where they classified simple <strong>and</strong> projective functors under<br />

reasonable assumptions on the small category. An influential<br />

approach to functor categories interestingly began with P.<br />

Gabriel’s thesis [27] (1962). Given a small category, he made<br />

it into an additive category by linearizing all morphism sets.<br />

Then he showed that the category <strong>of</strong> addictive functors from<br />

such an additive category to a module category is equivalent<br />

to another module category, <strong>of</strong> modules over an algebra<br />

that now we may call the additive category algebra <strong>of</strong> a small<br />

category. Later on B. Mitchell [56] (1972) introduced a different<br />

construction <strong>and</strong> defined an associative ring over a small<br />

category. Consequently he established connections between<br />

(R-mod) C <strong>and</strong> RC-mod. This is the result we pictured in the<br />

diagram.<br />

At this point, one perhaps does not see a connection between<br />

Mitchell’s work <strong>and</strong> category (co)homology. These<br />

two seemingly unrelated subjects were pieced up together<br />

by P. J. Webb (around 2000) who named these rings appeared<br />

in Mitchell’s paper, as “category algebras”. Representation<br />

theory <strong>of</strong> categories generalizes both representation<br />

theory <strong>of</strong> groups <strong>and</strong> <strong>of</strong> quivers, <strong>and</strong> thus is <strong>of</strong> great interest.<br />

<strong>Representations</strong> <strong>of</strong> categories are important to category<br />

(co)homology in the same way as group representations to<br />

xi


xii<br />

Preface<br />

group (co)homology. From category algebra point <strong>of</strong> view, we<br />

can put all necessary ingredients <strong>of</strong> category (co)homology<br />

under one framework, as shown in the diagram. Furthermore<br />

from the intrinsic structure <strong>of</strong> a category algebra, one can<br />

see the similarities with <strong>and</strong> differences from a groups algebra.<br />

Hence it explains why some classical results in group<br />

(co)homology may be generalized to category (co)homology,<br />

while others may not.<br />

At the early stage <strong>of</strong> category (co)homology, it seemed<br />

like a purely theoretic construction with few calculations.<br />

The thrust <strong>of</strong> recent development was brought in by representation<br />

<strong>and</strong> homotopy theorist working on locally determined<br />

structures. Their work completely reshaped the whole<br />

(co)homology theory <strong>and</strong> provided many interesting concrete<br />

categories to work with. We shall comment on it in Chapter<br />

6. In the last decade many interesting results on both abstract<br />

<strong>and</strong> concrete small categories have been obtained. For<br />

instance, a pivotal discovery is that the category <strong>of</strong> factorizations,<br />

F (C), first used by Quillen to show homotopy equivalence<br />

between BC <strong>and</strong> BC op , <strong>and</strong> then by Baues-Wirsching<br />

to introduce their cohomology theory, possesses the property<br />

that all (co)homology theories we consider here are indeed<br />

(co)homology <strong>of</strong> F (C) with appropriate coefficients. However<br />

when one tries to teach oneself about category (co)homology,<br />

one finds materials scattered in the literature <strong>and</strong> different<br />

writers have different background <strong>and</strong> styles. As an example<br />

the existing introduction to this subject by Webb [80]<br />

emphasizes its module-theoretic aspects. Lack <strong>of</strong> a comprehensive<br />

treatment makes the theory daunting for whoever<br />

wants to learn, <strong>and</strong> even for a researcher who uses category<br />

(co)homology theory it may cause inconvenience. Thus<br />

a book, introducing basic ideas <strong>of</strong> category (co)homology the-


Preface<br />

ory, presenting st<strong>and</strong>ard techniques <strong>and</strong> addressing the interactions<br />

between representation <strong>and</strong> homotopy theories, seems<br />

necessary. Such a book should provide a clear view <strong>of</strong> basic<br />

ideas, key methods, known results <strong>and</strong> unsolved conjectures,<br />

being a h<strong>and</strong>y introduction that can be used to foster further<br />

investigations <strong>and</strong> to search for future applications.<br />

Finally we turn to the structure <strong>of</strong> this book. The first<br />

two chapters consists <strong>of</strong> preliminaries needed in category<br />

(co)homology. It means one can find them in various classical<br />

books. The reason why I collect them here are, firstly it<br />

is convenient for the reader, secondly I try to provide some<br />

concrete examples to illustrate many abstract constructions<br />

which are in the center <strong>of</strong> this book. The first chapter recalls<br />

some basic definitions from category theory. The main<br />

focuses are limits <strong>of</strong> functors <strong>and</strong> their generalizations, that<br />

is, the Kan extensions. The purpose <strong>of</strong> the second chapter<br />

is to equip the reader with necessary knowledge about simplicial<br />

methods. We begin with a review <strong>of</strong> chain complexes.<br />

It is followed by an introduction to simplicial sets <strong>and</strong> the<br />

nerve <strong>of</strong> a small category where many combinatorially constructed<br />

chain complexes appear. Simplicial (co)homology is<br />

defined <strong>and</strong> several examples are given. To tell how to compare<br />

(co)homology <strong>of</strong> small categories, we have to inform the<br />

reader how to compare small categories <strong>and</strong> their nerves as<br />

well as classifying spaces. Thus some important categorical<br />

constructions are provided, which are needed throughout this<br />

book, for example to state Quillen’s Theorem A. For future<br />

references, <strong>and</strong> for the interested reader, we end Chapter 2<br />

with bisimplicial sets <strong>and</strong> several key results. Although only<br />

the statements will be used in future, we nonetheless present<br />

their pro<strong>of</strong>s.<br />

xiii


xiv<br />

Preface<br />

The third chapter introduces category algebras <strong>and</strong> their<br />

representations. Examples are served at the beginning to motivate<br />

the reader. We shall classify projective <strong>and</strong> injective<br />

modules under mild assumptions. Moreover we give an intrinsic<br />

characterization <strong>of</strong> category algebras so that they are<br />

comparable with cocommutative bialgebras. It explains why<br />

category algebras possess interesting homological properties.<br />

The fourth chapter studies (ordinary) (co)homology <strong>of</strong> category<br />

algebras. We begin with the most economical way by<br />

using derived functors to define category (co)homology. Then<br />

we recall Baues-Wirsching’s construction on the way to introduce<br />

the bar resolution. The bar resolution is simplicially<br />

constructed <strong>and</strong> it leads to various important modules by applying<br />

Kan extensions on it. We will see they are the most<br />

powerful tools for us. In this chapter we also define the extension<br />

<strong>of</strong> a category by a group <strong>and</strong> the Grothendieck spectral<br />

sequences.<br />

The fifth chapter discusses Hochschild (co)homology <strong>of</strong> category<br />

algebras. The key result here is a theorem to interpret<br />

Hochschild (co)homology <strong>of</strong> category algebras by their ordinary<br />

(co)homology, <strong>and</strong> vice versa. A theorem we prove here<br />

shows that all the previously mentioned (co)homology theories<br />

are just (co)homology <strong>of</strong> F (C) with coefficients. Some<br />

examples are given to demonstrate explicit calculations.<br />

The sixth chapter talks about connections between category<br />

<strong>and</strong> group cohomology. It contains mostly unpublished<br />

results. We bring up the notion <strong>of</strong> a local category <strong>of</strong> a finite<br />

group. Local categories are the motivating cases for research<br />

in category representations <strong>and</strong> cohomology. We put<br />

it in the end because we do need techniques developed earlier.<br />

This chapter contains many concrete categories <strong>and</strong> we<br />

shall see clearly how one can compute using various abstract


Preface<br />

machineries introduced before. A key concept in this chapter<br />

is a transporter category. We show how closely related<br />

are the transporter categories to the groups on which they<br />

are defined. The most important result is perhaps the finite<br />

generation <strong>of</strong> cohomology <strong>of</strong> modules <strong>of</strong> finite transporter categories.<br />

Also we will construct transfer maps for ordinary <strong>and</strong><br />

Hochschild cohomology <strong>of</strong> transporter categories. This chapter<br />

should help the reader to carry on further readings in<br />

advanced research papers.<br />

There may be many errors or even mistakes in this unfinished<br />

manuscript, all <strong>of</strong> which are my responsibility.<br />

xv<br />

Universitat Autònoma de Barcelona,<br />

June 2011<br />

Fei Xu<br />

xu@mat.uab.cat


Contents<br />

1 Functors <strong>and</strong> their Kan extensions . . . . . . . . . . . 1<br />

1.1Functors <strong>and</strong> limits . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.1.1Basic category theory . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.1.2Adjoint functors . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.1.3Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

1.2Restriction <strong>and</strong> Kan extensions . . . . . . . . . . . . . . . . . 21<br />

1.2.1Restriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

1.2.2Overcategories <strong>and</strong> undercategories . . . . . . . . . . 23<br />

1.2.3Kan extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

2 Simplicial methods . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

2.1Complexes <strong>and</strong> homology . . . . . . . . . . . . . . . . . . . . . . 31<br />

2.1.1Chain complexes, homology <strong>and</strong> chain homotopy 32<br />

2.1.2Double complexes <strong>and</strong> operations on chain<br />

complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

2.2Nerves, classifying spaces <strong>and</strong> cohomology . . . . . . . . 39<br />

2.2.1Simplicial sets <strong>and</strong> nerves <strong>of</strong> small categories . . 39<br />

2.2.2Classifying spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

2.2.3Cup product <strong>and</strong> cohomology ring . . . . . . . . . . . 62<br />

2.3Quillen’s work on classifying spaces. . . . . . . . . . . . . . 67<br />

2.3.1Quillen’s Theorem A . . . . . . . . . . . . . . . . . . . . . . . 67<br />

2.3.2Constructions over categories <strong>and</strong> relevant<br />

functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

2.4Further categorical <strong>and</strong> simplicial constructions . . . 74


xviii<br />

Contents<br />

2.4.1Grothendieck constructions . . . . . . . . . . . . . . . . . 75<br />

2.4.2Bisimplicial sets <strong>and</strong> homotopy colimits . . . . . . . 80<br />

2.4.3Pro<strong>of</strong>s <strong>of</strong> Quillen’s Theorem A <strong>and</strong><br />

Thomason’s theorem . . . . . . . . . . . . . . . . . . . . . . . 83<br />

3 Category algebras <strong>and</strong> their representations . 89<br />

3.1Basic concepts <strong>and</strong> examples . . . . . . . . . . . . . . . . . . . 89<br />

3.1.1Category algebras . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

3.1.2<strong>Representations</strong> <strong>of</strong> categories <strong>and</strong> Mitchell’s<br />

Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

3.1.3Three examples . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

3.2A closed symmetric monoidal category . . . . . . . . . . . 96<br />

3.2.1Tensor structure <strong>and</strong> an intrinsic<br />

characterization <strong>of</strong> category algebras . . . . . . . . . 96<br />

3.2.2The internal hom . . . . . . . . . . . . . . . . . . . . . . . . . . 103<br />

3.3Functors between module categories . . . . . . . . . . . . . 105<br />

3.3.1Restriction on algebras <strong>and</strong> modules . . . . . . . . . 105<br />

3.3.2Kan extensions <strong>of</strong> modules . . . . . . . . . . . . . . . . . . 108<br />

3.3.3Dual modules <strong>and</strong> Kan extensions . . . . . . . . . . . 114<br />

3.4EI categories, projectives <strong>and</strong> simples . . . . . . . . . . . . 116<br />

3.4.1EI condition <strong>and</strong> its implications . . . . . . . . . . . . . 116<br />

3.4.2Some representation theory . . . . . . . . . . . . . . . . . 118<br />

3.4.3Classifications <strong>of</strong> projectives <strong>and</strong> simples . . . . . . 125<br />

3.4.4Projective covers, injective hulls <strong>and</strong> their<br />

restrictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules. . . . . . . 133<br />

4.1General theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

4.1.1<strong>Cohomology</strong> <strong>of</strong> modules . . . . . . . . . . . . . . . . . . . . 133<br />

4.1.2<strong>Cohomology</strong> <strong>of</strong> a small category with<br />

coefficients in a functor . . . . . . . . . . . . . . . . . . . . . 140


Contents<br />

4.1.3Extensions <strong>of</strong> categories <strong>and</strong> low dimension<br />

cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150<br />

4.2Classical methods for computation . . . . . . . . . . . . . . 157<br />

4.2.1Minimal resolutions <strong>and</strong> reduction . . . . . . . . . . . 157<br />

4.2.2Examples using classifying spaces . . . . . . . . . . . . 160<br />

4.3Computation via adjoint functors . . . . . . . . . . . . . . . 162<br />

4.3.1Adjoint restrictions . . . . . . . . . . . . . . . . . . . . . . . . 162<br />

4.3.2Kan extensions <strong>of</strong> resolutions . . . . . . . . . . . . . . . . 164<br />

4.4Grothendieck spectral sequences . . . . . . . . . . . . . . . . 170<br />

4.4.1Grothendieck spectral sequences for a functor . . 170<br />

4.4.2Spectral sequences <strong>of</strong> category extensions . . . . . 175<br />

5 Hochschild cohomology . . . . . . . . . . . . . . . . . . . . . . 185<br />

5.1Hochschild homology <strong>and</strong> cohomology . . . . . . . . . . . 185<br />

5.1.1Definition <strong>and</strong> general properties. . . . . . . . . . . . . 185<br />

5.1.2Ring homomorphisms from the Hochschild<br />

cohomology ring . . . . . . . . . . . . . . . . . . . . . . . . . . . 190<br />

5.2Hochschild (co)homology <strong>of</strong> category algebras . . . . . 193<br />

5.2.1Basic ideas <strong>and</strong> examples . . . . . . . . . . . . . . . . . . . 193<br />

5.2.2Hochschild (co)homology as ordinary<br />

(co)homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195<br />

5.2.3EI categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204<br />

5.3Examples <strong>of</strong> the Hochschild cohomology rings <strong>of</strong><br />

categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208<br />

5.3.1The category E 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . 208<br />

5.3.2The category E 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 211<br />

5.3.3The category E 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 212<br />

5.3.4The category E 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 214<br />

6 Connections with group representations <strong>and</strong><br />

cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217<br />

6.1Local categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217<br />

xix


xx<br />

Contents<br />

6.1.1G-categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219<br />

6.1.2Homology representations <strong>of</strong> kG . . . . . . . . . . . . . 220<br />

6.1.3Transporter categories as Grothendieck<br />

constructions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221<br />

6.1.4Local categories . . . . . . . . . . . . . . . . . . . . . . . . . . . 224<br />

6.2Properties <strong>of</strong> local categories . . . . . . . . . . . . . . . . . . . 226<br />

6.2.1Two diagrams <strong>of</strong> categories . . . . . . . . . . . . . . . . . 226<br />

6.2.2Frobenius Reciprocity . . . . . . . . . . . . . . . . . . . . . . 227<br />

6.3The functor π: group representations via<br />

transporter categories . . . . . . . . . . . . . . . . . . . . . . . . . 229<br />

6.3.1Homology representations via transporter<br />

categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229<br />

6.3.2On finite generation <strong>of</strong> cohomology . . . . . . . . . . . 238<br />

6.3.3Transfer for ordinary cohomology . . . . . . . . . . . . 242<br />

6.4The functor ρ: invariants <strong>and</strong> coinvariants . . . . . . . . 252<br />

6.4.1Orbit categories . . . . . . . . . . . . . . . . . . . . . . . . . . . 253<br />

6.4.2Brauer categories, fusion <strong>and</strong> linking systems . . 255<br />

6.4.3Puig categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258<br />

6.4.4Orbit categories <strong>of</strong> fusion systems . . . . . . . . . . . . 259<br />

6.5Hochschild cohomology . . . . . . . . . . . . . . . . . . . . . . . . 260<br />

6.5.1<strong>Finite</strong> generation . . . . . . . . . . . . . . . . . . . . . . . . . . 260<br />

6.5.2Transfer for Hochschild cohomology . . . . . . . . . . 265<br />

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271


List <strong>of</strong> symbols<br />

Conventions: <strong>Categories</strong> are denoted by C, D in general unless<br />

otherwise specified; x, y are objects; α, β are morphisms<br />

in a category; F, G are general functors; u, v are functors between<br />

small categories; natural transformations between two<br />

functors are Φ, Ψ etc. Groups are written as G, H even when<br />

they are regarded as categories with one object.<br />

In a functor category we normally use M, N to denote the<br />

objects.<br />

• the trivial category <strong>of</strong> only one object <strong>and</strong> one morphism<br />

Set the category <strong>of</strong> sets<br />

T op the category <strong>of</strong> topological spaces<br />

Cat the category <strong>of</strong> small categories<br />

T C the category <strong>of</strong> all (if not specified, covariant) functors<br />

from a small category C to a category T<br />

Res u the restriction along a functor u<br />

LK u , RK u the left <strong>and</strong> right Kan extensions along u<br />

u/x, x\u categories over <strong>and</strong> under x ∈ Ob C, for a given<br />

u : D → C<br />

△ n the st<strong>and</strong>ard n-simplex<br />

n the totally ordered set 0 < 1 < · · · < n, a combinatorial<br />

construction <strong>of</strong> △ n<br />

△ the category <strong>of</strong> st<strong>and</strong>ard simplicies


xxii<br />

List <strong>of</strong> symbols<br />

∆ the diagonal functor C → C × C <strong>and</strong> various maps<br />

induced by it<br />

C e = C × C op the enveloping category <strong>of</strong> C<br />

F (C) the category <strong>of</strong> factorizations in C<br />

∇ : F (C) → C e the skew diagonal functor<br />

NC ∗ the nerve <strong>of</strong> C, a simplicial set<br />

BC the classifying space <strong>of</strong> C, geometric realization <strong>of</strong><br />

NC ∗ , a CW complex<br />

C ∗ , D ∗ , ... chain complexes<br />

C ∗,∗ , D ∗,∗ , ... double complexes<br />

Gr C M Grothendieck construction over M : C → Cat<br />

G ∝ P transporter category over a G-poset P (a Grothendieck<br />

construction)<br />

R a commutative ring with identity<br />

R-Mod the category <strong>of</strong> all R-modules<br />

R-mod the category <strong>of</strong> finitely generated R-modules<br />

k a field, usually algebraically closed<br />

Vect k the category <strong>of</strong> all k-vector spaces<br />

V ect k the category <strong>of</strong> finite-dimensional k-vector spaces<br />

RC the R-category algebra <strong>of</strong> C<br />

R the trivial RC-module, a constant functor<br />

H n (C, R), Hn(C, R) the nth simplicial (co)homology <strong>of</strong> C with<br />

coefficients in R<br />

H n (C; M), Hn(C; M) the nth (co)homology <strong>of</strong> C with coefficients<br />

in a functor/module M<br />

H n (BC, R), Hn(BC, R) the nth singular (co)homology <strong>of</strong> BC<br />

with coefficients in R<br />

H ∗ (C, R) the simplicial cohomology ring<br />

H ∗ (BC, R) the singular cohomology ring<br />

Ext ∗ RC(R, R) the ordinary cohomology ring <strong>of</strong> RC<br />

HH ∗ (RC) = Ext ∗ RCe(RC, RC) the Hochschild cohomology ring<br />

<strong>of</strong> RC


Chapter 1<br />

Functors <strong>and</strong> their Kan extensions<br />

Abstract We recall some basics from category theory <strong>and</strong> homological<br />

algebra. In this way we set conventions. The main<br />

constructions we discuss here are limits <strong>and</strong> their generalizations,<br />

the Kan extensions. We shall include some examples to<br />

illustrate these abstract concepts.<br />

1.1 Functors <strong>and</strong> limits<br />

1.1.1 Basic category theory<br />

In this section we recall some st<strong>and</strong>ard definitions <strong>and</strong> constructions.<br />

One purpose is to settle the conventions <strong>and</strong> terminologies<br />

that we shall follow in these notes.<br />

Definition 1.1.1. A category C consists <strong>of</strong> a class <strong>of</strong> objects<br />

Ob C <strong>and</strong> a class <strong>of</strong> morphisms Mor C such that<br />

1. each morphism α ∈ Mor C has a source <strong>and</strong> a target, written<br />

as s(α) <strong>and</strong> t(α), which are two objects in Ob C.<br />

2. for any pair <strong>of</strong> object x, y ∈ Ob C,<br />

Hom C (x, y) := {α ∣ ∣ s(α) = x <strong>and</strong> t(α) = y}<br />

is a set.<br />

3. for any α ∈ Hom C (x, y), β ∈ Hom C (y, z), their composite<br />

exists <strong>and</strong> is written as β ◦ α ∈ Hom C (x, z);<br />

4. for each object x there is an identity morphism 1 x ∈<br />

Hom C (x, x) satisfying the condition that 1 x ◦ α = α <strong>and</strong>


2 1 Functors <strong>and</strong> their Kan extensions<br />

β ◦ 1 x = β, if α ∈ Hom C (y, x) <strong>and</strong> β ∈ Hom C (x, z) for some<br />

y, z ∈ Ob C;<br />

5. if α ∈ Hom C (w, x), β ∈ Hom C (x, y) <strong>and</strong> γ ∈ Hom C (y, z),<br />

then (γ ◦ β) ◦ α = γ ◦ (β ◦ α).<br />

If α ∈ Hom C (x, y) is a morphism in C, very <strong>of</strong>ten we will picture<br />

it as α : x → y. A morphism α ∈ Mor C is a monomorphism<br />

if αβ = αγ for two morphisms β, γ ∈ Mor C, then<br />

β = γ. A morphism α ∈ Mor C is an epimorphism if βα = γα<br />

for two morphisms β, γ ∈ Mor C, then β = γ. A morphism<br />

α ∈ Hom C (x, y) is an isomorphism if there exists a morphism<br />

β ∈ Hom C (y, x) such that αβ = 1 y <strong>and</strong> βα = 1 x . An<br />

isomorphism is always both monomorphic <strong>and</strong> epimorphic.<br />

An isomorphism is also called an invertible morphism.<br />

We say two objects x <strong>and</strong> y are isomorphic, written as x ∼ =<br />

y, if there exists an isomorphism α ∈ Hom C (x, y). Let x be<br />

an object in a category C. Then the isomorphism class <strong>of</strong> x<br />

consists <strong>of</strong> all y ∈ Ob C that are isomorphic to x. The class<br />

<strong>of</strong> objects isomorphic to x is denoted by [x] ⊂ Ob C.<br />

A category is called a groupoid if every morphism is an<br />

isomorphism.<br />

An object x in C is initial if to any other object y, there<br />

exists a unique morphism x → y. An object x in C is terminal<br />

if to any other object y, there exists a unique morphism y →<br />

x. An object is a zero object if it is both initial <strong>and</strong> terminal.<br />

All initial (or terminal or zero) objects are isomorphic.<br />

Suppose C has a zero object 0, then for any two objects<br />

x, y we call the composite x → 0 → y the zero morphism<br />

in Hom C (x, y), written as 0 x,y . Note that the composite <strong>of</strong><br />

a zero morphism with any other morphism is a zero morphism.<br />

Let α ∈ Hom C (x, y) be a morphism. A kernel <strong>of</strong> α is<br />

a morphism β ∈ Hom C (w, x) satisfying αβ = 0 w,y such that<br />

if β ′ ∈ Hom C (w ′ , x) satisfies αβ ′ = 0 w ′ ,x then there exists a


1.1 Functors <strong>and</strong> limits 3<br />

unique morphism µ : w ′ → w such that β ′ = βµ. A cokernel<br />

<strong>of</strong> α is a morphism γ ∈ Hom C (y, z) satisfying γα = 0 x,z such<br />

that if γ ′ ∈ Hom C (x, z ′ ) satisfies γ ′ α = 0 x,z<br />

′ then there exists<br />

a unique morphism ν : z → z ′ such that γ ′ = νγ.<br />

Let I ⊂ Ob C be a subset. Then the coproduct <strong>of</strong> objects<br />

in I, written as ⊕ I x i, is an object X ∈ Ob C such that each<br />

x i ∈ I is equipped with a morphism α i : x i → X <strong>and</strong> if<br />

Y is another object equipped with β i : x i → Y then there<br />

exists a unique morphism γ : X → Y such that β i = γα i .<br />

The product <strong>of</strong> objects in I, written as ∏ I x i, is an object<br />

X ′ ∈ Ob C such that each x i ∈ I is equipped with a morphism<br />

α ′ i : X → x i <strong>and</strong> if Y ′ is another object equipped with β ′ i :<br />

Y → x i then there exists a unique morphism γ ′ : Y ′ → X ′<br />

such that α ′ i = β ′ i γ′ .<br />

Definition 1.1.2. A category C is called preadditive if every<br />

Hom C (x, y) is an abelian group <strong>and</strong> compositions are bilinear.<br />

A preadditive category is additive if furthermore all finite<br />

coproducts <strong>and</strong> all products exists in C.<br />

An additive category C is said to be preabelian if<br />

1. it has a zero object;<br />

2. every morphism has a kernel <strong>and</strong> a cokernel.<br />

A preabelian category is abelian if every monomorphism is a<br />

kernel <strong>of</strong> some morphism <strong>and</strong> every epimorphism is a cokernel<br />

<strong>of</strong> some morphism.<br />

There are abundance <strong>of</strong> categories in mathematics, many <strong>of</strong><br />

them are very natural while one can also cook up all kinds <strong>of</strong><br />

abstract categories.<br />

Example 1.1.3.1. The trivial category • has exactly one object<br />

• <strong>and</strong> one morphism 1 • ;<br />

2. A group G gives rise to two categories: the first one has<br />

only one object • <strong>and</strong> its morphism set is G; the sec-


4 1 Functors <strong>and</strong> their Kan extensions<br />

ond, denoted by EG, has Ob EG = G <strong>and</strong> Mor EG =<br />

⊎ g1 ,g 2 ∈Ob EGHom EG (g 1 , g 2 ) with Hom EG (g 1 , g 2 ) = {g 2 g1 −1 }.<br />

The former will usually be written just as G while the latter<br />

is called the Cayley graph <strong>of</strong> G;<br />

3. A partially ordered set (poset in short) P is naturally a category,<br />

still named P, if we let Ob P be the set <strong>of</strong> elements in<br />

the poset <strong>and</strong> Mor P = ⊎ x,y∈Ob C Hom C (x, y) in which the set<br />

Hom P (x, y) = {x ≤ y} if the two objects are comparable, or<br />

empty otherwise. Posets can be characterized as categories<br />

such that there exists at most one morphism between any<br />

two objects.<br />

4. The category <strong>of</strong> sets <strong>and</strong> set maps is denoted by Set, ;<br />

5. Let R be a commutative ring with identity <strong>and</strong> A an associative<br />

R-algebra. Then A-Mod <strong>and</strong> A-mod are the categories<br />

<strong>of</strong> all left A-modules <strong>and</strong> finitely generated left A-modules;<br />

6. The category Z-Mod is <strong>of</strong>ten written as Ab, the category <strong>of</strong><br />

abelian groups;<br />

7. The category <strong>of</strong> topological spaces <strong>and</strong> continuous maps is<br />

written as T op.<br />

In these notes, for an R-algebra A, if we do not specify, any<br />

A-module will be a left A-module.<br />

Definition 1.1.4. A category C is a small category if Ob C<br />

is a set, <strong>and</strong> is a finite category if Mor C is a finite set.<br />

A finite category is necessarily small by Definition 1.1.1 (4).<br />

In Example 1.1.3, the first three are small while the rest are<br />

abelian <strong>and</strong> not small.<br />

Definition 1.1.5. Let C be a category. Then a subcategory<br />

D ⊂ C consists <strong>of</strong> a subclass <strong>of</strong> objects Ob D ⊂ Ob C <strong>and</strong> a<br />

subclass <strong>of</strong> morphisms Mor D ⊂ Mor C, satisfying the axioms<br />

<strong>of</strong> a category with composition laws inherited from C.


1.1 Functors <strong>and</strong> limits 5<br />

A subcategory D ⊂ C if full, if for any pair <strong>of</strong> objects x, y ∈<br />

Ob D, we always have Hom D (x, y) = Hom C (x, y).<br />

Definition 1.1.6. A covariant functor F from a category<br />

D to another category C assigns to each x ∈ Ob D an object<br />

F(x) ∈ Ob C <strong>and</strong> to each α ∈ Hom D (x, y) a morphism F(α) ∈<br />

Hom C (F(x), F(y)) satisfying the conditions that<br />

1. F(1 x ) = 1 F(x) for every x ∈ Ob D;<br />

2. if α, β ∈ Mor D <strong>and</strong> β◦α exists, then F(β◦α) = F(β)◦F(α).<br />

A contravariant functor F ′ from a category D to another<br />

category C assigns to each x ∈ Ob D an object F ′ (x) ∈<br />

Ob C <strong>and</strong> to each α ∈ Hom D (x, y) a morphism F ′ (α) ∈<br />

Hom C (F ′ (x), F ′ (y)) satisfying the conditions that<br />

1 ′ . F ′ (1 x ) = 1 F ′ (x) for every x ∈ Ob D;<br />

2 ′ . if α, β ∈ Mor C <strong>and</strong> β◦α exists, then F ′ (β◦α) = F ′ (α)◦F ′ (β).<br />

Example 1.1.7. Let C be a category <strong>and</strong> x ∈ Ob C an object.<br />

Then<br />

1. there exists a covariant functor Hom C (x, −) : C → Set such<br />

that for any y ∈ Ob C, Hom C (x, −)(y) := Hom C (x, y) <strong>and</strong><br />

for any α ∈ Hom C (y, y ′ ) the composition Hom C (x, −)(α) =<br />

α ◦ − : Hom C (x, y) → Hom C (x, y ′ ); <strong>and</strong><br />

2. there exists a contravariant functor Hom C (−, x) : C → Set<br />

such that for any y ∈ Ob C, Hom C (−, x)(y) := Hom C (y, x)<br />

<strong>and</strong> for any α ∈ Hom C (y, y ′ ) the composition Hom C (−, x)(α) =<br />

− ◦ α : Hom C (y ′ , x) → Hom C (y, x).<br />

An object x ∈ Ob C is projective if Hom C (x, −) : C → Set<br />

preserves epimorphisms. An object x ∈ Ob C is injective if<br />

Hom C (−, x) : C → Set preserves monomorphisms. If C is an<br />

abelian category, we say that C has enough projective objects<br />

if for any y ∈ Ob C there exists a projective object x along<br />

with an epimorphism x → y <strong>and</strong> that C has enough injective


6 1 Functors <strong>and</strong> their Kan extensions<br />

objects if for any y ∈ Ob C there exists an injective object x<br />

along with a monomorphism y → x.<br />

Definition 1.1.8. Suppose C is a category. Its opposite category,<br />

named C op , share the same objects with C. Each α ∈<br />

Hom C (x, y) defines a unique morphism α op ∈ Hom C op(y, x). If<br />

α op ∈ Hom C op(y, x) <strong>and</strong> β op ∈ Hom C op(z, y), their composite<br />

is α op ◦ β op := (β ◦ α) op ∈ Hom C op(z, x).<br />

Passing from a category to its opposite has a dualizing effect<br />

on many categorical concepts, constructions <strong>and</strong> properties.<br />

For instance, it interchanges initial objects with terminal objects,<br />

monomorphisms with epimorphisms, projective objects<br />

with injective objects, products with coproducts etc. Moreover<br />

one can readily verify that a covariant functor F : C → D<br />

naturally determines a contravariant functor, the dual functor,<br />

F ∧ : C op → D, <strong>and</strong> vice versa. Thus what we learn about<br />

covariant functors can be directly translated to contravariant<br />

functors. From now on, if not specified, all functors will be<br />

covariant in these notes.<br />

Definition 1.1.9. A functor F : D → C is full if F(D) ⊂ C<br />

is a full subcategory.<br />

A functor F : D → C is faithful if for any α, β ∈ Mor D<br />

such that F(α) = F(β) ∈ Mor C, then α = β.<br />

If D <strong>and</strong> C are two preadditive categories, then a functor F :<br />

D → C is called additive if Hom D (x, y) → Hom C (F(x), F(y))<br />

is a homomorphism between abelian groups.<br />

Definition 1.1.10. Suppose F, G : D → C are two functors.<br />

A natural transformation Φ : F → G assigns to each object<br />

x ∈ Ob D a morphism Φ x : F(x) → G(x) so that we have a<br />

commutative diagram


1.1 Functors <strong>and</strong> limits 7<br />

F(x) Φ x <br />

G(x)<br />

F(α)<br />

<br />

G(α)<br />

F(y) Φy<br />

<br />

G(y)<br />

for any given α ∈ Hom D (x, y). If every Φ x is an isomorphism<br />

in C, we call such Φ a natural equivalence <strong>and</strong> write F ∼ = G<br />

in this situation.<br />

Definition 1.1.11. Two categories C <strong>and</strong> D are equivalent<br />

, written as C ≃ D, if there exist functors F : C → D <strong>and</strong><br />

G : D → C such that FG ∼ = Id D <strong>and</strong> GF ∼ = Id C .<br />

Definition 1.1.12. Let C be a category. If we take exactly<br />

one object from each isomorphism class <strong>of</strong> objects in C, then<br />

we can form a full subcategory consisting <strong>of</strong> these chosen objects.<br />

This subcategory is called a skeleton <strong>of</strong> C.<br />

Since skeletons <strong>of</strong> a category C are naturally equivalent to<br />

each other, we can speak about the skeleton <strong>of</strong> C. We shall<br />

always denote by [C] the skeleton <strong>of</strong> a category C. Then one<br />

can verify that [C] ≃ C.<br />

It is a fact that when D is small all the covariant functors<br />

from D to C form a category whose objects are these functors<br />

<strong>and</strong> morphisms are the natural transformations. We call it a<br />

functor category, written as C D . Sometimes we call such a<br />

D an index category. Functor categories are <strong>of</strong> pivotal importance<br />

in these notes. We <strong>of</strong>ten consider functor categories<br />

whose index categories are finite (such as a finite group G)<br />

<strong>and</strong> whose target categories are large (such as R-mod), in order<br />

to underst<strong>and</strong> finite categories via their representations<br />

(see Chapter 3) in certain large categories.<br />

The functors in Example 1.1.7 are very important to us.<br />

Here we present a crucial property <strong>of</strong> those functors.


8 1 Functors <strong>and</strong> their Kan extensions<br />

Definition 1.1.13. A covariant (respectively, contravariant)<br />

functor F : C → Set is called representable if it is naturally<br />

equivalent to Hom C (x, −) (respectively, Hom C (−, x)) for some<br />

x ∈ Ob C.<br />

Lemma 1.1.14 (Yoneda Lemma). Let F : C → Set be a<br />

functor. Then we have<br />

for any x ∈ Ob C.<br />

Hom Set<br />

C(Hom C (x, −), F) ∼ = F(x),<br />

Pro<strong>of</strong>. Each natural transformation Φ in the left side is uniquely<br />

determined by the morphism<br />

Φ x : Hom C (x, x) → F(x)<br />

which is uniquely determined by the image <strong>of</strong> Φ x (1 x ) ∈ F(x).<br />

Hence we can define a bijection Φ → Φ x (1 x ) <strong>and</strong> then the<br />

statement follows.<br />

⊓⊔<br />

In the end, we record a construction that we will use in the<br />

next section.<br />

Definition 1.1.15. Suppose C <strong>and</strong> D are two categories.<br />

Then we can define their product category C × D, whose objects<br />

are {(x, y) ∣ x ∈ Ob C, y ∈ Ob D} <strong>and</strong> whose morphisms<br />

are {(α, β) ∣ α ∈ Mor C, β ∈ Mor D}.<br />

A functor from a product category C × D to some target<br />

category sometimes is called a bifunctor.<br />

1.1.2 Adjoint functors<br />

We take the opportunity to record the definition <strong>of</strong> an adjoint<br />

functor <strong>and</strong> its basic properties for future reference.<br />

Definition 1.1.16. Let F : C → D <strong>and</strong> G : D → C be two<br />

functors such that there is a natural equivalence


1.1 Functors <strong>and</strong> limits 9<br />

Ω : Hom D (F(−), −) ∼= →Hom C (−, G(−))<br />

<strong>of</strong> functors C op ×D → Set. We say that F is left adjoint to G,<br />

G is right adjoint to F <strong>and</strong> write Ω : F ⊣ G for the adjunct.<br />

Example 1.1.17. If F : C → D <strong>and</strong> G : D → C give rise to<br />

category equivalences as in Definition 1.1.11, then F is both<br />

a left <strong>and</strong> a right adjoint <strong>of</strong> G.<br />

Another example is for U, V, W ∈ V ect k , there is an isomorphism<br />

Hom k (U ⊗ k V, W ) ∼ = Hom k (U, Hom k (V, W )).<br />

Next we deduce some fundamental properties <strong>of</strong> the adjunction.<br />

1. Naturality <strong>of</strong> Ω: for any α : x ′ → x <strong>and</strong> β : y → y ′ the<br />

following diagram commutes<br />

Hom D (F(x), y) Ω x,y <br />

Hom C (x, G(y))<br />

β◦(−)◦F(α)<br />

<br />

G(β)◦(−)◦α<br />

Hom D (F(x ′ ), y ′ )<br />

<br />

Ωx<br />

Hom<br />

′ ,y ′ C (x ′ , G(y ′ )).<br />

Equivalently, we have for any ϕ : F(x) → y<br />

Ω x ′ ,y ′ (β ◦ ϕ ◦ F(α)) = G(β) ◦ Ω x,y (ϕ) ◦ α.<br />

We shall repeatedly use the above formula. Very <strong>of</strong>ten, the<br />

subscript <strong>of</strong> Ω will be omitted for the sake <strong>of</strong> convenience.<br />

For instance, when x ′ = x <strong>and</strong> α = 1 x , we have<br />

Ω(β ◦ ϕ) = G(β) ◦ Ω(ϕ),<br />

<strong>and</strong> when y = F(x) <strong>and</strong> ϕ = 1 F(x) , we get<br />

Ω(β ◦ F(α)) = G(β) ◦ α.


10 1 Functors <strong>and</strong> their Kan extensions<br />

One should try to deduce similar formulas for future applications.<br />

2. Unit <strong>and</strong> counit: we define, for each x ∈ Ob C, Σ x =<br />

Ω(1 F(x) ) : x → GF(x) <strong>and</strong>, for each y ∈ Ob D, Λ y =<br />

Ω −1 (1 G(y) ) : FG(y) → y. By naturality <strong>of</strong> Ω, we actually<br />

obtain two natural transformations<br />

Σ : Id C → GF <strong>and</strong> Λ : FG → Id D .<br />

For example, for any α : x ′ → x, using various formulas we<br />

obtain from the naturality <strong>of</strong> Ω, we get<br />

GF(α) ◦ Σ x<br />

′ = GF(α) ◦ Ω(1 F(x ′ ))<br />

= Ω(F(α) ◦ 1 F(x ′ ))<br />

= Ω(1 F(x) ◦ F(α))<br />

= Ω(1 F(x) ) ◦ α<br />

= Σ x ◦ α.<br />

Thus we have a commutative diagram<br />

x ′<br />

α<br />

x<br />

Σ x ′<br />

<br />

GF(x ′ )<br />

Σx<br />

GF(α)<br />

<br />

GF(x).<br />

We can deduce that the following<br />

F FΣ<br />

−→FGF ΛF<br />

−→F,<br />

G ΣG<br />

−→GFG GΛ<br />

−→G<br />

compose to identities. For example, given some x ∈ Ob C,<br />

one can verify Λ F(x) ◦ F(Σ x ) = 1 F(x) by showing<br />

Ω(Λ F(x) ◦ F(Σ x )) = Ω(Λ F(x) ) ◦ Σ x = Σ x = Ω(1 F(x) ).<br />

Moreover one can recover the adjunct by unit <strong>and</strong> counit:<br />

for any ϕ : F(x) → y <strong>and</strong> ψ : x → G(y)<br />

Ω(ϕ) = G(ϕ) ◦ Σ x<br />

<strong>and</strong> Ω −1 (ψ) = Λ y ◦ F(ψ).


1.1 Functors <strong>and</strong> limits 11<br />

Theorem 1.1.18. Suppose F : C → D <strong>and</strong> G : D → C<br />

are two functors. If there exist two natural transformations<br />

Σ : Id C → GF <strong>and</strong> Λ : FG → Id D such that the following<br />

two composites<br />

F FΣ<br />

−→FGF ΛF<br />

−→F,<br />

G ΣG<br />

−→GFG GΛ<br />

−→G<br />

are identities, then F is a left adjoint <strong>of</strong> G. Moreover<br />

Ω : Hom D (F(x), y) → Hom C (x, G(y)) defined by Ω(ϕ) =<br />

G(ϕ) ◦ Σ x is the adjunct, with Σ, Λ the unit <strong>and</strong> counit <strong>of</strong><br />

the adjunction.<br />

Pro<strong>of</strong>. For x ∈ Ob C, y ∈ Ob D, φ ∈ Hom D (F(x), y) <strong>and</strong><br />

ψ ∈ Hom C (x, G(y)), we define<br />

Ω x,y (φ) = G(φ) ◦ Σ x <strong>and</strong> ¯Ωx,y (ψ) = Λ y ◦ F(ψ).<br />

Then we can verify that they define two natural transformations<br />

<strong>of</strong> functors C op × D → Sets. In fact for any α ∈<br />

Hom C (x, x ′ ), β ∈ Hom D (y, y ′ ) <strong>and</strong> φ ∈ Hom D (F(x), y) we<br />

have<br />

G(β) ◦ Ω(φ) ◦ α = G(β) ◦ G(φ) ◦ Σ x ◦ α<br />

= G(β ◦ φ) ◦ GF(α) ◦ Σ x<br />

′<br />

= G(β ◦ φ ◦ F(α)) ◦ Σ x<br />

′<br />

= Ω(β ◦ φ ◦ F(α)),<br />

which implies Ω is a natural transformation. The assertion<br />

for ¯Ω is proved in a similar way.<br />

Moreover Ω <strong>and</strong> ¯Ω provide natural equivalences because we<br />

can show<br />

( ¯Ω ◦ Ω)(φ) = ¯Ω(G(φ) ◦ Σx )<br />

= Λ y ◦ F(G(φ) ◦ Σ x )<br />

= Λ y ◦ FG(φ) ◦ F(Σ x )<br />

= φ ◦ Λ F(x) F(Σ x )<br />

= φ,


12 1 Functors <strong>and</strong> their Kan extensions<br />

<strong>and</strong> similarly (Ω ◦ ¯Ω)(ψ) = ψ.<br />

This theorem also implies that the adjoint <strong>of</strong> a functor is<br />

unique up to natural equivalence.<br />

Corollary 1.1.19. If G <strong>and</strong> G ′ are right adjoints <strong>of</strong> F, then<br />

they are naturally equivalent. Similarly if F <strong>and</strong> F ′ are two<br />

left adjoints <strong>of</strong> G, then they are naturally equivalent.<br />

Pro<strong>of</strong>. We only prove the first assertion. For any x ∈ Ob C<br />

<strong>and</strong> y ∈ Ob D we have two isomorphisms<br />

Hom C (x, G(y)) Ω<br />

←−Hom D (F(x), y) Ω′<br />

−→Hom C (x, G ′ (y)).<br />

Then (Ω◦Ω ′−1 )(1 G (y)) : ′ G ′ (y) → G(y) <strong>and</strong> (Ω ′ ◦Ω −1 )(1 G(y) ) :<br />

G(y) → G ′ (y) induce a natural equivalence between G <strong>and</strong><br />

G ′ .<br />

⊓⊔<br />

⊓⊔<br />

1.1.3 Limits<br />

In these notes, direct <strong>and</strong> inverse limits <strong>of</strong> functors are key<br />

concepts so we introduce their definitions here. Suppose M ∈<br />

Ob(T C ) is an object in a functor category. Then it can be<br />

identified with a commutative diagram <strong>of</strong> objects {M(x) ∣ ∣<br />

x ∈ Ob C} <strong>and</strong> morphisms {M(α) ∣ ∣ α ∈ Mor C} in T . For<br />

convenience <strong>and</strong> future reference, we call this the diagram <strong>of</strong><br />

M.<br />

Definition 1.1.20. Let C be a small category <strong>and</strong> T an arbitrary<br />

category. Consider the functor category T C . Suppose<br />

M ∈ Ob(T C ). Then it has an inverse limit, denoted by<br />

lim M ∈ Ob T , if for each x ∈ Ob C there exists a morphism<br />

θ x : ←−C<br />

←−C<br />

lim M → M(x) such that, by adding the object<br />

lim M <strong>and</strong> the morphisms {θ ∣<br />

←−C<br />

x x ∈ Ob C} to the diagram<br />

<strong>of</strong> M, we obtain an enlarged commutative diagram, <strong>and</strong> this


1.1 Functors <strong>and</strong> limits 13<br />

object lim ←−C<br />

M is universal in the sense that if t is another object<br />

in Ob T that enjoys the same properties as lim ←−C<br />

M then<br />

we must have a unique morphism Θ t : t → lim ←−C<br />

M making<br />

the whole diagram commutative as pictured as follows<br />

t<br />

θ ′ y<br />

Θ t<br />

θ ′ x<br />

<br />

lim M θ x<br />

←−C<br />

θ y <br />

M(α)<br />

M(x)<br />

M(y) .<br />

Dually the functor M has a direct limit in T , denoted by<br />

lim M ∈ Ob T , if for each x ∈ Ob C there exists a morphism<br />

−→C<br />

τ x : M(x) → lim ←−C<br />

M such that, by adding the object −→C<br />

lim M<br />

∣<br />

<strong>and</strong> the morphisms {τ x x ∈ Ob C} to the diagram <strong>of</strong> M,<br />

we obtain an enlarged commutative diagram, <strong>and</strong> this object<br />

lim M is universal in the sense that if t is another object in<br />

−→C<br />

Ob T that enjoys the same properties as lim −→C<br />

M then we must<br />

have a unique morphism Ξ t : −→C<br />

lim M → t making the whole<br />

diagram commutative as pictured as follows<br />

M(α)<br />

M(y) τy<br />

<br />

M(x)<br />

τ ′ y<br />

τ x<br />

τ ′ x<br />

lim<br />

−→C M Ξ t<br />

t<br />

It follows directly from the definition that, when a limit<br />

exists, it is unique up to isomorphism. If either the index<br />

category or the functor itself is structurally simple, we may<br />

explicitly compute the limits.


14 1 Functors <strong>and</strong> their Kan extensions<br />

Example 1.1.21.1. Let C = Z <strong>and</strong> T = Set. Then a functor<br />

M : Z → Set is represented by a chain <strong>of</strong> set maps<br />

· · · → M(−1) → M(0) → M(1) → M(2) → · · · .<br />

In particular if M 0 is given by the following diagram in which<br />

each map is an inclusion<br />

· · · ⊂ M 0 (−1) ⊂ M 0 (0) ⊂ M 0 (1) ⊂ M 0 (2) ⊂ · · · ,<br />

then lim ←−Z<br />

M 0 = ⋂ i∈Z M 0(i) <strong>and</strong> −→Z<br />

lim M 0 = ⋃ i∈Z M 0(i).<br />

2. If C = Z <strong>and</strong> T = V ect k , the category <strong>of</strong> finite-dimensional<br />

k-vector spaces. Suppose N is represented by the following<br />

· · · → 0 N(1) N(2) N(3) · · · .<br />

Then −→Z<br />

lim N does not exist in V ect k , the category <strong>of</strong> finitedimensional<br />

k-vector spaces, but lim ←−Z<br />

N = {0}.<br />

3. Let T be a category <strong>and</strong> I a set considered as a discrete<br />

category. Then a functor M : I → T is simply an I-indexed<br />

set <strong>of</strong> objects in T . The the coproduct <strong>of</strong> these objects is<br />

defined as ⊕ I x i = −→I<br />

lim M <strong>and</strong> the product <strong>of</strong> these objects<br />

is ∏ I x i = ←−I<br />

lim M.<br />

4. Let C be the following poset<br />

x<br />

α<br />

<br />

y β<br />

<br />

z<br />

<strong>and</strong> T = Vect k , the category <strong>of</strong> all k-vector spaces. Each<br />

functor M ∈ Ob T C is represented by<br />

M(y) M(β)<br />

M(x)<br />

M(α)<br />

<br />

M(z).


1.1 Functors <strong>and</strong> limits 15<br />

Then ←−C<br />

lim M is called the pullback <strong>of</strong> the latter diagram <strong>and</strong><br />

lim M = M(z).<br />

−→C<br />

Dually if D is the following poset<br />

a f <br />

g<br />

c<br />

<strong>and</strong> N ∈ Ob T D then −→D<br />

lim N is called the pushout <strong>of</strong> the<br />

diagram<br />

N(a)<br />

N(g)<br />

<br />

N(c)<br />

b<br />

N(f) <br />

N(b)<br />

<strong>and</strong> lim ←−D<br />

N = N(a).<br />

The definition <strong>of</strong> limits can be rewritten by using a simple,<br />

yet very important, construction.<br />

Definition 1.1.22. There is a constant functor<br />

K : T → T C<br />

such that, for any t ∈ Ob T , K(t) is defined by K(t)(x) = t<br />

<strong>and</strong> K(α) = 1 t for any x ∈ Ob C <strong>and</strong> α ∈ Mor C.<br />

In the literature, the functor K : T → T C is <strong>of</strong>ten named the<br />

“diagonal functor”. Since the terminology is used for another<br />

purpose, see Definition 2.2.34, we shall stick with our notion<br />

which seems to be more appropriate.<br />

Now we present alternative characterizations <strong>of</strong> limits. The<br />

definition <strong>of</strong> an inverse limit can be rephrased as saying that<br />

there exists lim ←−C<br />

M ∈ Ob C with a natural transformation<br />

Γ : K(lim ←−C<br />

M) → M which is universal in the sense that if<br />

there is another object t, along with a natural transformation<br />

Υ : K(t) → M, then there exists a unique morphism θ t :


16 1 Functors <strong>and</strong> their Kan extensions<br />

t → lim ←−C<br />

M hence a natural transformation K(θ t ) : K(t) →<br />

K(lim ←−C<br />

M) such that Γ ◦ K(θ t ) = Υ .<br />

K(θ t )<br />

K(t)<br />

K(lim ←−C<br />

M) Γ<br />

Especially we obtain a morphism<br />

Ω : Hom T C(K(t), M) → Hom T (t, lim ←−C<br />

M),<br />

given by Ω(Υ ) = θ t . It is an isomorphism because each t →<br />

lim M extends to a functor K(t) → M. Thus if every M ∈<br />

←−C<br />

Ob(T C ) has a limit in T , then lim ←−<br />

: T C → T is the right<br />

adjoint <strong>of</strong> K <strong>and</strong> Ω becomes the corresponding adjunct. We<br />

characterize the direct limit lim −→C<br />

in a similar way but we leave<br />

it to the reader. When T is “large” enough with respect to C,<br />

we can introduce the limits in an economical way.<br />

Proposition 1.1.23. If K has a right adjoint R : T C → T ,<br />

then for each M ∈ Ob(T C ), R(M) is the inverse limit <strong>of</strong><br />

M. If K has a left adjoint L : T C → T , then for each<br />

M ∈ Ob(T C ), L(M) is the direct limit <strong>of</strong> M.<br />

For convenience, we introduce the following concepts.<br />

Definition 1.1.24. A category T is called complete if for<br />

any small category C <strong>and</strong> any functor M ∈ T C the inverse<br />

limit lim ←−C<br />

M exists. A category T is called cocomplete if for<br />

any small category C <strong>and</strong> any functor M ∈ T C the direct<br />

limit lim −→C<br />

M exists.<br />

A category T is called finitely complete (respectively finitely<br />

cocomplete) if for any finite category C <strong>and</strong> any functor<br />

Υ<br />

M


1.1 Functors <strong>and</strong> limits 17<br />

M ∈ T C the inverse limit ←−C<br />

lim M (respectively the direct limit<br />

lim M) exists.<br />

−→C<br />

The categories Set <strong>and</strong> R-Mod are both complete <strong>and</strong> cocomplete.<br />

By Example 1.1.21, a coproduct is a direct limit <strong>and</strong> a product<br />

in an inverse limit. In abelian categories, the existence <strong>of</strong><br />

coproducts, respectively product, is equivalent to the cocompleteness,<br />

respectively completeness, condition on a category<br />

T .<br />

Proposition 1.1.25. Let T be an abelian category. Then<br />

1. it is complete (or finitely complete) if <strong>and</strong> only if all products<br />

(or all finite products) exist; <strong>and</strong><br />

2. it is cocomplete (or finitely cocomplete) if <strong>and</strong> only if all<br />

coproducts (or all finite coproducts) exist.<br />

Pro<strong>of</strong>. We only prove (1). Since a product is an inverse limit,<br />

we just have to demonstrate the opposite direction. Suppose<br />

C is a small (or finite) category <strong>and</strong> M : C → T is a functor.<br />

Let<br />

C 0 (C; M) = {f : Ob C →<br />

∏<br />

M(x) ∣ f(x) ∈ M(x)}<br />

<strong>and</strong><br />

C 1 (C; M) = {f : Mor C →<br />

x∈Ob C<br />

∏<br />

α∈Mor C<br />

We define δ : C 0 (C; M) → C 1 (C; M) by<br />

M(y) ∣ ∣ f(x α →y) ∈ M(y)}.<br />

δ(f)(x α →y) = f(y) − M(α)[f(x)].<br />

Then we can verify that the kernel <strong>of</strong> δ is ←−C<br />

lim M. ⊓⊔<br />

For example the category <strong>of</strong> finitely generated R-modules,<br />

R-mod, is both finitely complete <strong>and</strong> finitely cocomplete.


18 1 Functors <strong>and</strong> their Kan extensions<br />

Completeness <strong>and</strong> cocompleteness pass to functor categories.<br />

We record this <strong>and</strong> another result that shall be useful in the<br />

next chapter.<br />

Proposition 1.1.26. Let C be a small category. If T is complete<br />

(resp. cocomplete), then T C is complete (resp. cocomplete).<br />

Pro<strong>of</strong>. Let D be another small category <strong>and</strong> M : D → T C<br />

a functor. Since (T C ) D ∼ = T C×D , M can be identified with a<br />

bifunctor from ˜M(−, −) : C × D → T .<br />

We may define a functor N : C → T by N(x) = ←−D<br />

lim ˜M(x, −)<br />

for every x ∈ Ob C. Since [lim ←−D ˜M(−, −)](x) ∼ = lim ←−D<br />

[ ˜M(x, −)]<br />

for all x ∈ Ob D, we can show N is the limit ←−D<br />

lim M. It means<br />

T C is complete. The cocompleteness <strong>of</strong> T C can be proved in<br />

the same way.<br />

⊓⊔<br />

The above result tells us that Set C <strong>and</strong> (R-Mod) C are both<br />

complete <strong>and</strong> cocomplete if C is small. Furthermore (R-mod) C<br />

is finitely complete <strong>and</strong> cocomplete if C is finite.<br />

We record two results on preserving limits.<br />

Proposition 1.1.27. Let T be both complete <strong>and</strong> cocomplete<br />

<strong>and</strong> C a small category. Given a functor M ∈ Ob T C<br />

<strong>and</strong> T ∈ Ob T , we have<br />

Hom T (T, lim ←−C<br />

M) ∼ = lim ←−C<br />

Hom T (T, M(−))<br />

<strong>and</strong><br />

Hom T (lim −→C<br />

M, T ) ∼ = lim ←−C<br />

Hom T (M(−), T )<br />

Pro<strong>of</strong>. To prove the first isomorphism, we take the diagram <strong>of</strong><br />

M, which determines lim ←−C<br />

M. Applying the covariant functor<br />

Hom T (T, −) to the diagram it results in the diagram <strong>of</strong> the<br />

functor Hom T (T, M(−)). From here one may complete the


1.1 Functors <strong>and</strong> limits 19<br />

pro<strong>of</strong> based by using the universal property <strong>of</strong> an inverse limit.<br />

The second isomorphism may be proved in the same way. ⊓⊔<br />

Proposition 1.1.28. Let F : T → T ′ be the left adjoint <strong>of</strong><br />

G : T ′ → T .<br />

1. Suppose M : C → T has a direct limit. Then F(lim −→C<br />

M) ∼ =<br />

lim F ◦ M. −→C<br />

2. Suppose N : D → T ′ has an inverse limit. Then G(lim ←−D<br />

N) ∼ =<br />

lim G ◦ N. ←−D<br />

Pro<strong>of</strong>. We only prove (1). It is obvious that F(lim −→C<br />

M) fits into<br />

the defining diagram <strong>of</strong> −→C<br />

lim F◦M. We need to show F(lim −→C<br />

M)<br />

is universal. Suppose t ′ ∈ Ob T ′ is another object satisfying<br />

the colimit defining diagram. It follows from the adjunction<br />

Hom T ′(F(lim −→C<br />

M), t ′ ) ∼ = Hom T (lim −→C<br />

M, G(t ′ )),<br />

along with the unit Id → F◦G that G(t ′ ) satisfies the defining<br />

diagram <strong>of</strong> −→C<br />

lim M. From the universal property <strong>of</strong> −→C<br />

lim M,<br />

there exists a unique morphism lim −→C<br />

M → G(t ′ ) which gives<br />

rise to a morphism F(lim −→C<br />

M) → t ′ . The universal property<br />

<strong>of</strong> F(lim −→C<br />

M) follows from it.<br />

⊓⊔<br />

In the end, we give a result that is used in the next section.<br />

Theorem 1.1.29. Let C be a small category. Then any<br />

functor M : C → Set is canonically a colimit <strong>of</strong> a diagram<br />

<strong>of</strong> representable functors.<br />

Pro<strong>of</strong>. We shall establish this result by firstly constructing a<br />

small category D from M : C → Set <strong>and</strong> secondly showing<br />

that M induces a functor ˜M : D → Set C satisfying −→D<br />

lim ˜M ∼ =<br />

M.<br />

The objects in the category D are pairs (x, a) such that x<br />

is an object <strong>of</strong> C <strong>and</strong> a is an element in the set M(x). A


20 1 Functors <strong>and</strong> their Kan extensions<br />

morphism f : (x, a) → (y, b) is a morphism f ∈ Hom C (x, y)<br />

satisfying M(f)(a) = b.<br />

We define a functor ˜M : D → Set C by ˜M(x, a) =<br />

Hom C (x, −). From the Yoneda Lemma, there exists for each<br />

x ∈ Ob C an isomorphism <strong>of</strong> sets Hom Set<br />

C(Hom C (x, −), M) ∼ =<br />

M(x). It implies that each m x ∈ M(x) determines uniquely<br />

a functor f (x,mx ) : ˜M(x, mx ) → M making the enlarged diagram<br />

<strong>of</strong> ˜M commutative<br />

Hom C (x, −) = ˜M(x, mx )<br />

f◦−<br />

˜M(y, m y ) = Hom C (y, −)<br />

f (x,m x)<br />

M<br />

f (y,m y)<br />

Note that the commutativity <strong>of</strong> the diagram forces M(f)(m x ) =<br />

m y .<br />

In order to prove that M is the direct limit we need to<br />

show it is universal. Suppose L is another functor fitting<br />

into the above diagram, <strong>and</strong> for each (x, m x ) ∈ Ob D it<br />

comes with a functor g (x,mx ) : Hom C (x, −) → L. Again by<br />

the Yoneda Lemma, g (x,mx ) determines uniquely an element<br />

l x ∈ L(x). From the commutativity <strong>of</strong> the diagram, we must<br />

have L(f)(l x ) = l y . Now we define a functor h : M → L by<br />

h z (m z ) = l z<br />

if z ∈ Ob C. We may readily verify that h z ′M(f) = L(f)h z<br />

for any f : z → z ′ . It implies that h is well defined. Since by<br />

definition hf (x,mx ) = g (x,mx ) for all objects (x, m x ) ∈ Ob D, M<br />

is universal.<br />

⊓⊔<br />

The construction <strong>of</strong> D from a functor to Set in the pro<strong>of</strong> is<br />

the predecessor <strong>of</strong> the Grothendieck constructions introduced<br />

in Section 2.4.


1.2 Restriction <strong>and</strong> Kan extensions 21<br />

1.2 Restriction <strong>and</strong> Kan extensions<br />

In last section we mentioned that, for a functor category T C ,<br />

if the target category are complete <strong>and</strong> cocomplete, then we<br />

may define the limits using adjoints <strong>of</strong> the constant functor<br />

K : T → T C . When we examine closely, we realize that K itself<br />

is induced by another functor pt : C → •. This observation<br />

generates new ideas for constructing some extremely powerful<br />

functors, called Kan extensions.<br />

1.2.1 Restriction<br />

Definition 1.2.1. Suppose u : D → C is a functor between<br />

two small categories. For any target category T , u induces a<br />

functor Res u : T C → T D via precomposition with u, called<br />

the restriction along u.<br />

Example 1.2.2. The canonical functor pt : C → • induces a<br />

restriction T ∼ = T • → T C , which is identical to K <strong>of</strong> Definition<br />

1.1.22.<br />

In last section, we know explicitly the left <strong>and</strong> right adjoints<br />

<strong>of</strong> K. As a generalization, we shall describe the adjoints <strong>of</strong><br />

an arbitrary restriction. But before doing that, we provide<br />

several simple properties <strong>of</strong> the restriction.<br />

Lemma 1.2.3. Suppose u : D → C is a functor. Then for<br />

any M ∈ Ob(T C ), there are canonical morphisms lim ←−C<br />

M →<br />

lim Res ←−D uM <strong>and</strong> lim −→D<br />

Res u M → lim −→C<br />

M.<br />

Pro<strong>of</strong>. The map between inverse limits follows from the universal<br />

property


22 1 Functors <strong>and</strong> their Kan extensions<br />

lim M ←−C α x<br />

Θ<br />

<br />

lim Res φ x<br />

α y ←−D uM<br />

<br />

Res u M(a) = M(u(a))<br />

φ y<br />

<br />

Res u M(φ)=M(u(φ))<br />

Res u M(b) = M(u(b)) ,<br />

in which φ : a → b is a morphism in D. The map between<br />

direct limits can be established similarly.<br />

⊓⊔<br />

Proposition 1.2.4. Suppose u : D → C has a left adjoint<br />

v : C → D. Then for any T , Res u has a right adjoint Res v .<br />

In particular, if u <strong>and</strong> v are category equivalences, Res u <strong>and</strong><br />

Res v are equivalences.<br />

If T is complete <strong>and</strong> cocomplete, then for any M ∈<br />

Ob(T C ), lim −→C<br />

M ∼ = lim −→D<br />

Res u M, <strong>and</strong> for any N ∈ Ob(T ) D ,<br />

lim N ∼ = lim<br />

←−D ←−C<br />

Res v N.<br />

Pro<strong>of</strong>. We shall prove that for M ∈ Ob(T C ) <strong>and</strong> N ∈ Ob(T D )<br />

there exists an isomorphism<br />

Hom T D(Res u M, N) ∼ = Hom T C(M, Res v N).<br />

Let Σ : Id C → uv <strong>and</strong> Λ : vu → Id D be the unit <strong>and</strong> counit<br />

for the adjunction between u <strong>and</strong> v. We define two natural<br />

transformations Σ ′ : Id T<br />

C → Res v Res u <strong>and</strong> Λ ′ : Res u Res v →<br />

Id T<br />

D as follows. Given M ∈ Ob(T C ), x ∈ Ob C, N ∈ Ob(T D )<br />

<strong>and</strong> a ∈ Ob D,<br />

(Σ ′ M) x := M(Σ x ) : M(x) → Res v Res u M(x) = M(uv(x)),<br />

<strong>and</strong><br />

(Λ ′ N) a := N(Λ a ) : Res u Res v N(a) = N(vu(a)) → N(a).<br />

One can verify that Σ ′ <strong>and</strong> Λ ′ provide the unit <strong>and</strong> counit <strong>of</strong><br />

an adjunction.


1.2 Restriction <strong>and</strong> Kan extensions 23<br />

As for the second statement, we have for all t ∈ Ob T<br />

<strong>and</strong><br />

Hom T (lim −→D<br />

Res u M, t) ∼ = Hom T D(Res u M, K(t))<br />

∼ = HomT C(M, Res v K(t))<br />

∼ = HomT C(M, K(t))<br />

∼ = HomT (lim −→C<br />

M, t).<br />

Hom T (t, lim ←−C<br />

Res v N) ∼ = Hom T C(K(t), Res v N)<br />

∼ = HomT D(Res u K(t), N)<br />

∼ = HomT D(K(t), N)<br />

∼ = HomT (t, lim ←−D<br />

N).<br />

Thus −→D<br />

lim Res u M ∼ = lim −→C<br />

M <strong>and</strong> ←−C<br />

lim Res v N ∼ = lim ←−D<br />

N. ⊓⊔<br />

In many places we will have to consider the adjoint functors<br />

<strong>of</strong> some restriction Res u . When u has an adjoint, we get an<br />

adjoint <strong>of</strong> Res u which is also a restriction, by Proposition 1.2.4.<br />

The truth is that even if u does not admit an adjoint, we can<br />

still construct the adjoints <strong>of</strong> Res u , <strong>and</strong> this is the main result<br />

<strong>of</strong> the upcoming two sections.<br />

1.2.2 Overcategories <strong>and</strong> undercategories<br />

In order to introduce the adjoints <strong>of</strong> a restriction, we have<br />

to provide some important categorical constructions. These<br />

categorical constructions are <strong>of</strong> great importance in both homological<br />

algebra <strong>and</strong> homotopy theory <strong>of</strong> classifying spaces.<br />

We shall be familiar with them as they appear almost everywhere<br />

throughout these notes.<br />

Definition 1.2.5. Let u : D → C be a functor between<br />

(small) categories <strong>and</strong> x ∈ Ob C. The category over x, u/x,<br />

consists <strong>of</strong> objects {(a, α) ∣ ∣ a ∈ Ob D, α ∈ Hom C (u(a), x)}.<br />

For any two objects (a, α), (b, β), a morphism from (a, α) to


24 1 Functors <strong>and</strong> their Kan extensions<br />

(b, β) is a morphism µ ∈ Hom D (a, b) making the following<br />

diagram commutative<br />

u(a)<br />

u(µ)<br />

u(b)<br />

The category under x, written as x\u, is defined in a dual<br />

fashion. It consists <strong>of</strong> objects {(α, a) ∣ ∣ a ∈ Ob D, α ∈<br />

Hom C (x, u(a))}. For any two objects (α, a), (β, b), a morphism<br />

from (α, a) to (β, b) is a morphism µ ∈ Hom D (a, b)<br />

making the following diagram commutative<br />

x<br />

α<br />

β<br />

α<br />

β<br />

u(a)<br />

x,<br />

u(µ)<br />

u(b)<br />

The categories defined above are customarily called overcategories<br />

<strong>and</strong> undercategories, associated with u : D → C. We<br />

will see later on that Id C /x <strong>and</strong> x\Id C , for any x ∈ Ob C, are<br />

already very interesting.<br />

Remark 1.2.6.1. From definition, an object in the overcategory<br />

u/x, (a, α), can be pictured as u(a) →x, α <strong>and</strong> consequently<br />

a morphism µ : (a, α) → (b, β) can be equivalently<br />

interpreted as a sequence u(a) u(µ)<br />

→ u(b) →x. β This kind<br />

<strong>of</strong> rewritings will be useful for us when dealing with chains<br />

<strong>of</strong> morphisms in u/x <strong>and</strong> we shall come back to this point<br />

in later chapters. Similar reinterpretation can be made for


1.2 Restriction <strong>and</strong> Kan extensions 25<br />

objects <strong>and</strong> morphisms in undercategories too but we leave<br />

it to the reader.<br />

2. There is a canonical functor P x : u/x → D (resp. P x :<br />

x\u → D given on objects as projection to the first (resp.<br />

the second) component <strong>and</strong> on morphisms as the identity.<br />

For simplicity, we shall denote such functors just as P.<br />

3. If γ : x → y is a morphism in C, then it naturally induces a<br />

functor γ ∗ : u/x → u/y <strong>and</strong> a functor γ ∗ : y\u → x\u.<br />

Example 1.2.7.1. Let pt : C → • be the canonical functor.<br />

Then pt/• ∼ = •\pt ∼ = C.<br />

2. Let G be a group <strong>and</strong> H a subgroup. Then the inclusion<br />

functor i H : H ↩→ G gives exactly one overcategory i H /•.<br />

By direct calculation, the objects are {(•, g)|g ∈ G}, <strong>and</strong><br />

biject with the elements <strong>of</strong> G. There is a morphism from<br />

one object (•, g 1 ) to another (•, g 2 ) if there exists a h ∈ H<br />

such that g 1 = g 2 h. Obviously h = g2 −1 g 1. Thus there is at<br />

most one morphism from an object to another. Since h is<br />

invertible, there exists a morphism between two objects if<br />

<strong>and</strong> only if their are isomorphic in i H /•. In other words, two<br />

objects (•, g 1 ) to another (•, g 2 ) are isomorphic if <strong>and</strong> only<br />

if g 1 H = g 2 H. Because the category i H /• consists <strong>of</strong> [G : H]<br />

many groupoids, each <strong>of</strong> which is equivalent to the trivial<br />

category, i H /• is equivalent to the discrete set G/H <strong>of</strong> left<br />

cosets (regarded as a category).<br />

Similarly the undercategory •\i H has objects {(g, •)|g ∈<br />

G}. There is a morphism from (g 1 , •) to (g 2 , •) if <strong>and</strong> only<br />

if there exists a (unique) h ∈ H such that hg 1 = g 2 or<br />

equivalently h = g 2 g1 −1 . The undercategory is equivalent<br />

to H\G, the set <strong>of</strong> right cosets. We have an isomorphism<br />

i H /• → •\i H given by<br />

(•, g) ↦→ (g −1 , •) <strong>and</strong> (•, g 1 ) g−1 2 g 1<br />

→ (•, g 2 ) ↦→ (g −1<br />

1 , •)g−1<br />

2 g 1<br />

→ (g2 −1 , •).


26 1 Functors <strong>and</strong> their Kan extensions<br />

When G = H, i G /• = Id G /• ∼ = •\Id G = •\i G is the Cayley<br />

graph <strong>of</strong> G.<br />

Based on our observations, by Proposition 1.2.4, if M ∈<br />

Ob(V ect H k<br />

) (commonly called a k-representation <strong>of</strong> H),<br />

lim M ∼ −→i H<br />

= lim<br />

/• −→G/H<br />

M ∼ = ⊕ g∈G/H gH ⊗ k M. One can see that<br />

G permutes these direct summ<strong>and</strong>s <strong>and</strong> the limit lim M −→iH /•<br />

is isomorphic to kG ⊗ kH M.<br />

3. Suppose u : D → C is a functor between two posets. Then<br />

for any x ∈ Ob C, u/x is isomorphic to the subposet <strong>of</strong><br />

D consisting <strong>of</strong> objects {a ∈ Ob D ∣ Hom C (u(a), x) ≠ ∅},<br />

while x\u is isomorphic to the subposet <strong>of</strong> D consisting <strong>of</strong><br />

objects {b ∈ Ob D ∣ Hom C (x, u(b)) ≠ ∅}.<br />

The following observation follows directly from definitions<br />

<strong>and</strong> will be useful to us. For any functor u : D → C one can<br />

define a (covariant) opposite functor u op : D op → C op such<br />

that u op (x) = x <strong>and</strong> u op (α op ) = u(α) op . Be aware that it is<br />

different from the dual functor given before Definition 1.1.9.<br />

Lemma 1.2.8. Suppose u : D → C is a functor. Consider<br />

its opposite functor u op : D op → C op . Then for any<br />

x ∈ Ob C = Ob C op we have (u/x) op ∼ = x\u op <strong>and</strong> (x\u) op ∼ =<br />

u op /x.<br />

1.2.3 Kan extensions<br />

In this section, we assume T to be a complete <strong>and</strong> cocomplete<br />

abelian category. The reader should bear Example 1.2.7 (1)<br />

in mind in order to see that Kan extensions generalize direct<br />

<strong>and</strong> inverse limits.<br />

Theorem 1.2.9. Let u : D → C be a functor between small<br />

categories. Then the restriction Res u : T C → T D admits a<br />

left adjoint LK u , called the left Kan extension along u, as


1.2 Restriction <strong>and</strong> Kan extensions 27<br />

well as a right adjoint RK u , called the right Kan extension<br />

along u.<br />

Pro<strong>of</strong>. We only sketch the constructions <strong>of</strong> the Kan extensions<br />

<strong>and</strong> leave details to be filled by the reader.<br />

Given M ∈ Ob(T D ) we define its left <strong>and</strong> right Kan extensions<br />

along u as<br />

LK u M = lim −→u/−<br />

Res P M<br />

<strong>and</strong> RK u M = lim ←−−\u<br />

Res P M,<br />

where P is the functor in Remark 1.2.6 (2). Here we only<br />

prove the statement for the left Kan extension because the<br />

pro<strong>of</strong> for the right Kan extension follows the same pattern.<br />

Step 1, we show LK u M is a functor from C to T . If<br />

γ ∈ Hom C (x, y), then we have a functor γ ∗ : u/x → u/y. Since<br />

Res γ∗ Res P M = Res P M as functors over u/x, by Lemma<br />

1.2.3, it determines a canonical morphism<br />

lim<br />

−→ γ ∗ : LK u M(x) = −→u/x<br />

lim Res P M → LK u M(y) = −→u/y<br />

lim Res P M.<br />

Hence LK u M = lim −→u/−<br />

Res P M ∈ Ob(T C ).<br />

Step 2, we state that LK u is a functor from T D to T C . For<br />

any natural transformation Ψ : M → M ′ between two objects<br />

<strong>of</strong> T D . We can use the universal property to build a canonical<br />

natural transformation LK u Ψ : LK τ M → LK u M ′ .<br />

Step 3, we construct an adjunct<br />

Ω : Hom T C(LK u M, N) → Hom T D(M, Res u N).<br />

For each M ∈ Ob(T D ) we define the counit<br />

by the defining map <strong>of</strong> a limit<br />

Σ M : M → Res u LK u M<br />

(Σ M ) a : M(a) = Res P M[(a, 1 u(a) )] → lim −→u/u(a)<br />

Res P M,


28 1 Functors <strong>and</strong> their Kan extensions<br />

for any a ∈ Ob D. For each N ∈ Ob(T C ) we put the counit<br />

Λ N : LK u Res u N → N<br />

such that, for every x ∈ Ob C, lim −→u/x<br />

Res P Res u N → N(x)<br />

comes from the universal property <strong>of</strong> direct limit. We need to<br />

prove that the following are identities<br />

Res u<br />

Res u Σ<br />

−→ Res u LK u Res u<br />

ΛRes u<br />

−→Resu <strong>and</strong> LK u<br />

ΣLK u<br />

−→ LKu Res u LK u<br />

LK u Λ<br />

−→LK u<br />

For the first we compute for any N ∈ Ob(T C ) <strong>and</strong> a ∈ Ob D<br />

the following composite<br />

(Res u N)(a) [Res u(Σ N )] a<br />

−→ (Resu LK u Res u N)(a) (Λ ResuN) a<br />

−→ (Res u N)(a).<br />

This composite is<br />

(Res u N)(a) = (Res P Res u N)[(a, 1 a )] → lim −→u/u(a)<br />

Res P Res u N → (Res u N<br />

which is really an identity by the universal property <strong>of</strong> lim −→u/u(a)<br />

Res P Re<br />

For the second we compute for any M ∈ Ob(T D ) <strong>and</strong> x ∈<br />

Ob C that<br />

(LK u M)(x) [LK uΛ M ] x<br />

−→ (LKu Res u LK u M)(x) (Λ LKuM) x<br />

−→ (LK u M)(x)<br />

gives the identity. But it is rewritten as<br />

lim Res −→u/x<br />

PM → lim −→u/x<br />

(Res P Res u LK u M) → lim −→u/x<br />

Res P M,<br />

which in turn equals −→u/x<br />

lim Res P applying to<br />

M → Res u LK u M → M.<br />

However the preceding morphisms compose to the identity<br />

because the following composite<br />

M(a) → lim −→u/u(a)<br />

Res P M → M(a)


1.2 Restriction <strong>and</strong> Kan extensions 29<br />

is an identity for every a ∈ Ob D, due to the fact that (a, 1 u(x) )<br />

is a terminal object <strong>and</strong> that Res P M(a, 1 u(a) ) = M(a). ⊓⊔<br />

Remark 1.2.10.1. One may choose T to be Ab, or R-Mod etc,<br />

for practical applications. When the index categories are<br />

finite, we can even use R-mod, the category <strong>of</strong> finitely generated<br />

R-modules.<br />

2. When u : D → C is a full embedding <strong>and</strong> M ∈ Ob(Ab D ),<br />

then LK u M ∈ Ob(Ab C ) restricted on u(D) is identified with<br />

M, which means LK u M(u(d)) = M(d) for any d ∈ Ob D.<br />

This is why we call such functors discovered by D. M. Kan,<br />

“the Kan extensions”.<br />

Example 1.2.11. In Example 1.2.7 (2) where the over- <strong>and</strong><br />

under-categories associated with i : H ↩→ G are computed,<br />

we can continue to verify that Res i is the usual restriction<br />

↓ G H , LK i is equivalent to the induction ↑ G H = kG ⊗ kH − <strong>and</strong><br />

RK i is equivalent to the coinduction ⇑ G H = Hom kH(kG, −).<br />

If G is finite, the two Kan extensions are well known to be<br />

equivalent.<br />

Corollary 1.2.12. Let u : D → C <strong>and</strong> v : E → D be two<br />

functors between small categories. Suppose T is a complete<br />

<strong>and</strong> cocomplete category, M ∈ Ob(T E ) <strong>and</strong> N ∈ Ob(T C ).<br />

Then Res v Res u = Res uv <strong>and</strong> consequently LK u LK v<br />

∼ = LKuv ,<br />

RK u RK v<br />

∼ = RKuv .<br />

Pro<strong>of</strong>. The equality between restrictions follows directly from<br />

definition. Then we have<br />

Hom T E(M, Res v Res u N) ∼ = Hom T D(LK v M, Res u N)<br />

∼ = HomT C(LK u LK v M, N).<br />

Hence LK u LK v<br />

∼ = LKuv . The isomorphism between right<br />

Kan extensions can be proved similarly.<br />

⊓⊔


Chapter 2<br />

Simplicial methods<br />

Abstract We begin with chain complexes <strong>and</strong> their homology,<br />

the fundamental concepts <strong>of</strong> homological algebra. Then<br />

we review simplicial constructions in algebra <strong>and</strong> topology<br />

as they provide concrete examples <strong>and</strong> motivating ideas. Out<br />

main interest lies in the nerve <strong>of</strong> a small category. This particular<br />

simplicial set allows us to define simplicial homology<br />

<strong>and</strong> cohomology <strong>of</strong> a category with coefficients in a commutative<br />

ring. The nerve <strong>of</strong> a small category has a geometric<br />

realization, called the classifying space. Hence we can also<br />

consider the singular homology <strong>and</strong> cohomology <strong>of</strong> a classifying<br />

space. We shall show these two theories agree. We will<br />

introduce various important categorical constructions. Meanwhile<br />

we develop techniques for comparing categoires, their<br />

nerves <strong>and</strong> classifying spaces. A major theorem is Quillen’s<br />

Theorem A. For future references <strong>and</strong> better underst<strong>and</strong>ing<br />

<strong>of</strong> simplicial methods, we also include a description <strong>of</strong> bisimplicial<br />

sets <strong>and</strong> several relevant results.<br />

2.1 Complexes <strong>and</strong> homology<br />

Here we recall basics about chain <strong>and</strong> cochain complexes as<br />

well as operations on them.


32 2 Simplicial methods<br />

2.1.1 Chain complexes, homology <strong>and</strong> chain homotopy<br />

Suppose A is an associative ring with identity <strong>and</strong> Z is the<br />

totally ordered set <strong>of</strong> integers. From Example 1.1.3 (3) we<br />

may deem Z as a category.<br />

Definition 2.1.1. A chain complex <strong>of</strong> A-modules is an object<br />

C ∈ Ob(A-Mod) Zop such that C(n → n + 2) = 0 for<br />

all n ∈ Z. In other words it consists <strong>of</strong> a collection <strong>of</strong> objects<br />

in A-Mod, {C n = C(n)} n∈Z , together with morphisms<br />

∂ n = C((n − 1) → n) : C n → C n−1 , called the differentials,<br />

such that ∂ n ◦ ∂ n+1 = 0.<br />

A cochain complex <strong>of</strong> objects in A-Mod is an object C ∈<br />

Ob(A-Mod) Z such that C(n → n + 2) = 0, for all n ∈ Z.<br />

Alternatively it consists <strong>of</strong> a collection <strong>of</strong> objects in A, {C n =<br />

C(n)} n∈Z , together with morphisms ∂ n = C(n → (n + 1)) :<br />

C n → C n+1 , called the differentials, such that ∂ n ◦ ∂ n−1 = 0.<br />

If {C n , ∂ n } n∈Z (respectively, {C n , ∂ n } n∈Z ) is a chain complex<br />

(respectively, cochain complex) <strong>of</strong> A-modules, <strong>and</strong> x ∈ C n<br />

(respectively x ∈ C n ), then we say x has degree n, written as<br />

deg x = n.<br />

We call a complex bounded below if C n (or C n ) vanishes<br />

when n is sufficiently small. A complex bounded above if C n<br />

(or C n ) vanishes when n is sufficiently large. A complex if<br />

bounded if it is both bounded below <strong>and</strong> bounded above (that<br />

is, there are only finitely many non-zero terms). A complex is<br />

called a stalk complex if it has exactly one non-zero term.<br />

Suppose C = {C n , ∂ n } is a chain complex. Then we usually<br />

write Z n (C) = ker ∂ n ⊂ C n . Any element x ∈ Z n (C) is called<br />

an n-cycle. We also write B n (C) = Im∂ n+1 ⊂ C n . Any element<br />

x ∈ B n (C) is called an n-boundary. Similarly suppose<br />

C = {C n , ∂ n } is a cochain complex. Then we usually write<br />

Z n (C) = ker ∂ n ⊂ C n . Any element x ∈ Z n (C) is called an n-


2.1 Complexes <strong>and</strong> homology 33<br />

cocycle. We also write B n (C) = Im∂ n−1 ⊂ C n . Any element<br />

x ∈ B n (C) is called an n-coboundary.<br />

Definition 2.1.2. The i-th homology <strong>of</strong> a chain complex<br />

{C n , ∂ n } n∈Z is<br />

Hi(C) = Z i (C)/B i (C).<br />

The i-th cohomology <strong>of</strong> a cochain complex {C n , ∂ n } n∈Z is<br />

H i (C) = Z i (C)/B i (C).<br />

Definition 2.1.3. Let C <strong>and</strong> D be two chain complexes (respctively,<br />

cochain complexes). A chain map (respectively,<br />

cochain map) φ : D → C is a natural transformation. In<br />

other words, it consists <strong>of</strong> maps φ n : D n → C n (respectively,<br />

φ n : D n → C n ) such that the following diagram commutes<br />

D n<br />

φ n <br />

∂ D n <br />

D n−1<br />

φ n−1<br />

respectively, D n ∂n D <br />

φ n <br />

C n ∂<br />

C<br />

n<br />

<br />

C n−1 C n ∂ n C<br />

D n+1<br />

φ n+1<br />

<br />

C n+1<br />

The chain map φ is a chain isomorphism if it is a natural<br />

equivalence.<br />

Clearly a chain map between chain complexes φ : D → C<br />

induces maps between homology groups φ ∗ : Hn(D) → Hn(C)<br />

<strong>and</strong> a cochain map between cochain complexes φ : D → C<br />

induces maps between cohomology groups φ ∗ : H n (D) →<br />

H n (C).<br />

For convenience, we shall call both chain maps <strong>and</strong> cochain<br />

maps just chain maps. Given a chain complex {C n , ∂ n } n∈Z <strong>of</strong><br />

objects in an abelian category A, we can obtain a cochain<br />

complex {C n , ∂ n } n∈Z , simply by asking C n = C −n <strong>and</strong> ∂ n =<br />

∂ −n . Thus in some sense, chain <strong>and</strong> cochain complexes are<br />

the same. In practice, we will regard a complex as a chain


34 2 Simplicial methods<br />

complex or a cochain complex depending on where it comes<br />

from.<br />

Definition 2.1.4. If φ, ψ : D → C are two chain maps, we<br />

say φ <strong>and</strong> ψ are chain homotopic, written as φ ≃ ψ, if there<br />

are maps h n : D n → C n+1 such that Φ n − Ψ n = h n−1 ∂ D n +<br />

∂ C n+1h n· · ·<br />

<br />

D n+1<br />

∂ D n+1<br />

D n<br />

∂ D n<br />

D n−1<br />

· · ·<br />

φ n+1 −ψ n+1<br />

h n<br />

φ n −ψ n<br />

h n−1<br />

φ n−1 −ψ n−1<br />

· · ·<br />

<br />

C n+1 ∂<br />

C<br />

n+1<br />

<br />

C n<br />

∂ C n<br />

<br />

C n−1<br />

<br />

· · ·<br />

We say two complexes D <strong>and</strong> C are chain homotopy equivalent,<br />

written as D ≃ C if there are chain maps φ : D → C<br />

<strong>and</strong> ψ : C → D such that φ ◦ ψ ≃ Id C <strong>and</strong> ψ ◦ φ ≃ Id D .<br />

We say a complex C is contractible, if it is chain homotopy<br />

equivalent to the zero complex. We say a complex C is acyclic<br />

if Hn(C) = 0 for all n. Acyclicity is strictly weaker than contractibility.<br />

Proposition 2.1.5. If φ, ψ : D → C are chain homotopic,<br />

<strong>and</strong> if F : A-Mod → B-Mod is an additive functor, then<br />

1. φ ∗ = ψ ∗ : Hn(D) → Hn(C). Thus a chain homotopy equivalence<br />

D ≃ C induces isomorphisms Hn(D) ∼ = Hn(C).<br />

2. F(φ) ≃ F(ψ) : F(D) → F(C), <strong>and</strong> F(φ) ∗ = F(ψ) ∗ :<br />

Hn(F(D)) → Hn(F(C)).<br />

Let C be a bounded below chain complex <strong>of</strong> A-modules.<br />

A projective resolution <strong>of</strong> C is a (bounded below) chain<br />

complex P <strong>of</strong> projective A-modules, along with a chain map<br />

P → C which induces isomorphisms on homology groups. Any<br />

two projective resolutions <strong>of</strong> C are chain homotopy equiva-


2.1 Complexes <strong>and</strong> homology 35<br />

lent. Similarly we can consider an injective resolution I <strong>of</strong> a<br />

bounded above complex <strong>of</strong> A-modules D.<br />

2.1.2 Double complexes <strong>and</strong> operations on chain complexes<br />

Definition 2.1.6. A double complex <strong>of</strong> A-modules is an object<br />

C ∈ Ob(A-Mod) Zop ×Z op such that C(p → (p + 2), 1 q ) =<br />

C(1 p , q → (q + 2)) = 0 <strong>and</strong><br />

C[(1 p+1 , q → (q + 1))(p → (p + 1), 1 q )] + C[(p → (p + 1), 1 q+1 )(1 p , q → (q + 1))] = 0<br />

for all p, g ∈ Z.<br />

Alternatively it consists <strong>of</strong> a collection <strong>of</strong> objects {C p,q =<br />

C(p, q)} p,q∈Z , together with maps<br />

∂ h = C((p − 1) → p, 1 q ) : C p,q → C p−1,q ; ∂ v = C(1 p , (q − 1) → q) : C p,q → C p,q−1<br />

such that ∂ h ◦ ∂ h = ∂ v ◦ ∂ v = ∂ v ◦ ∂ h + ∂ h ◦ ∂ v = 0.<br />

A map between two double complexes D → C is just a<br />

natural transformation. It is an isomorphism if it is a natural<br />

equivalence.<br />

Double complexes are also called bicomplexes in the literature.<br />

A double complex C = {C p,q } p,q∈Z is called bounded if<br />

it has only finitely many non-zero terms along each diagonal<br />

line p + q = n. An example will be the first quadrant double<br />

complex where C p,q = 0 unless both p <strong>and</strong> q are non-negative.<br />

Definition 2.1.7. The total complexes <strong>of</strong> a double complex<br />

C are given by<br />

Tot ⊕ (C) n = ⊕<br />

<strong>and</strong> Tot ∏ (C) n = ∏<br />

p+q=n<br />

C p,q<br />

with differentials ∂ = ∂ h + ∂ v .<br />

p+q=n<br />

C p,q<br />

These two constructions are usually different, but when C is<br />

bounded they are equal. Now we present a classical construction.


36 2 Simplicial methods<br />

Let A be an associative ring. Suppose D is a chain complex <strong>of</strong><br />

right A-modules <strong>and</strong> C is a chain complex <strong>of</strong> left R-modules.<br />

We define their product complex D ⊗ A C by<br />

(D ⊗ A C) n = ⊕<br />

D p ⊗ A C q .<br />

p+q=n<br />

It is a double complex with ∂ h p,q = ∂ D p ⊗1 <strong>and</strong> ∂ v p,q = (−1) p (1⊗<br />

∂ C q ), pictured as<br />

D p−1 ⊗ A C q<br />

∂ h p,q<br />

D p ⊗ A C q<br />

∂ v p−1,q <br />

D p−1 ⊗ A C q−1<br />

∂ h p,q−1<br />

∂ v p,q<br />

D p ⊗ A C q−1<br />

Hence the differential <strong>of</strong> the total complex Tot ⊕ (D ⊗ A C) is<br />

given by<br />

∂(x ⊗ y) = ∂ D p x ⊗ y + (−1) p x ⊗ ∂ C q y,<br />

for x ∈ D p <strong>and</strong> y ∈ C q .<br />

Suppose both E <strong>and</strong> C are chain complexes <strong>of</strong> left A-<br />

modules. Then we introduce another double complex Hom A (E, C)<br />

by (there are different conventions in the literature)<br />

Hom A (E, C) n = ∏<br />

Hom A (E −p , C q ),<br />

pictured as<br />

q+p=n<br />

Hom A (E −p+1 , C q )<br />

∂ h p,q<br />

Hom A (E −p , C q )<br />

∂ v p−1,q <br />

∂ v p,q<br />

Hom A (E −p+1 , C q−1 ) Hom A (E −p , C q−1 )<br />

∂ h p,q−1<br />

with ∂ h p,q(f) = f∂ E −p+1 <strong>and</strong> ∂ v p,q(f) = (−1) p+q ∂ C q f.


2.1 Complexes <strong>and</strong> homology 37<br />

The differential <strong>of</strong> the total complex Tot ∏ (Hom A (E, C)) is<br />

∂ n = ∂ h p,q + ∂ v p,q : Hom A (E, C) n → Hom A (E, C) n−1 given by<br />

for any f ∈ Hom A (E −p , C q ).<br />

∂ n f = f∂ E −p+1 + (−1) n ∂ C q f<br />

Theorem 2.1.8. Suppose R is a commutative ring. Let C,<br />

D <strong>and</strong> E be complexes <strong>of</strong> R-modules. We have an isomorphism<br />

<strong>of</strong> double complexes<br />

Hom R (C ⊗ R D, E) ∼ = Hom R (C, Hom R (D, E))<br />

The following theorem is well known <strong>and</strong> its pro<strong>of</strong> can be<br />

found in many places. Here we record it for future references.<br />

Theorem 2.1.9 (Künneth formula). Suppose A is a ring.<br />

Let D be a chain complex <strong>of</strong> right A-modules <strong>and</strong> C, E chain<br />

complexes <strong>of</strong> left A-modules. Then<br />

1. if D n <strong>and</strong> ∂(D n ) are flat for all n, there is a short exact<br />

sequence<br />

0 → ⊕ p+q=n Hp(D) ⊗ A Hq(C) → Hn(D ⊗ A C)<br />

→ ⊕ p+q=n−1 TorA 1 (Hp(D), Hq(C))<br />

2. if E n <strong>and</strong> ∂(E n ) are projective for all n, there is a short<br />

exact sequence<br />

0 → ∏ q−p=n+1 Ext1 A(Hp(E), Hq(C)) → Hn(Hom A (E, C))<br />

→<br />

∏<br />

q−p=n Hom A(Hp(E), Hq(C)) →<br />

Corollary 2.1.10 (Universal Coefficient Theorem). Let<br />

D be a chain complex <strong>of</strong> right A-modules over a ring A, E<br />

a chain complex <strong>of</strong> left A-modules <strong>and</strong> M a left A-module<br />

considered as a complex concentrated in degree zero. Then


38 2 Simplicial methods<br />

1. if D n <strong>and</strong> ∂(D n ) are flat for all n, there is a short exact<br />

sequence<br />

0 → Hp(D) ⊗ A M → Hn(D ⊗ A M) → Tor A 1 (Hn−1(D), M) → 0,<br />

2. if E n <strong>and</strong> ∂(E n ) are projective for all n, there is a short<br />

exact sequence<br />

0 → Ext 1 A(Hn−1(E), M) → H−n(Hom A (E, M)) → Hom A (Hn(E), M)<br />

In Corollary 2.1.10 (2), the middle term is <strong>of</strong>ten written as<br />

cohomology H n (Hom A (E, M)).<br />

The dual complex <strong>of</strong> C, denoted by C ∧ , is defined to be<br />

Hom A (C, A), where the second A is regarded as a complex<br />

concentrated in degree zero. This is a complex <strong>of</strong> right<br />

A-modules. Note that if P is a projective A-module then<br />

Hom A (P, A) is a projective right module. We can define an<br />

evaluation map<br />

ev : C ∧ ⊗ A C → A.<br />

It is non-zero only at degree zero. More explicitly it is given<br />

by f ⊗ x ↦→ f(x) for any base elements f ⊗ x ∈ (C ∧ ⊗ C) 0 .<br />

Here x ∈ C n <strong>and</strong> f ∈ C ∧ −n.<br />

Proposition 2.1.11. Suppose k is a field. Let D <strong>and</strong> C be<br />

two complexes <strong>of</strong> finite-dimensional k-vector spaces. If for<br />

every n, Hom k (D, C) n is finite-dimensional, then we have<br />

an isomorphism <strong>of</strong> complexes<br />

D ∧ ⊗ k C ∼ = Hom k (D, C).<br />

Pro<strong>of</strong>. Note that if M, N are two stalk complexes <strong>of</strong> k-vector<br />

spaces concentrated in degrees m <strong>and</strong> n respectively, then we<br />

have M ∧ ⊗ k N ∼ = Hom k (M, N) as stalk complexes concentrated<br />

in degree n − m. Consequently we have an isomorphism<br />

for every integer n, (D ∧ ⊗ k C) n<br />

∼ = Homk (D, C) n since<br />

Hom k (D, C) n is finite-dimensional <strong>and</strong> then Hom k (D, C) n =


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 39<br />

⊕ q+p=n Hom k (D −p , C q ). Hence in order to finish the pro<strong>of</strong>, we<br />

only need to verify that the two differentials are identified under<br />

the vector space isomorphisms.<br />

⊓⊔<br />

In the next section, we will see many combinatorially constructed<br />

complexes <strong>and</strong> double complexes.<br />

2.2 Nerves, classifying spaces <strong>and</strong> cohomology<br />

By Dold-Kan Correspondence, the category <strong>of</strong> non-negatively<br />

graded complexes <strong>of</strong> A-modules is equivalent to the category<br />

<strong>of</strong> simplicial A-modules. It implies also that the category<br />

<strong>of</strong> first quadrant double complexes <strong>of</strong> A-modules is<br />

equivalent to the category <strong>of</strong> bisimplicial A-modules. Since in<br />

these notes, we are mainly interested in various non-negatively<br />

graded complexes such as projective resolutions <strong>of</strong> modules,<br />

<strong>and</strong> moreover these complexes come from corresponding simplicial<br />

sets, it is useful to introduce simplicial sets <strong>and</strong> modules.<br />

2.2.1 Simplicial sets <strong>and</strong> nerves <strong>of</strong> small categories<br />

We recall the fundamental idea in algebraic topology <strong>of</strong> singular<br />

(co)homology theory. From here we shall see how algebraic<br />

<strong>and</strong> combinatorial methods enter the study <strong>of</strong> spaces. Then<br />

we will apply the same methods, called simplicial methods, to<br />

investigate small categories (instead <strong>of</strong> spaces). Our first definition<br />

<strong>of</strong> category (co)homology is given soon after the basic<br />

definitions are recorded.<br />

Let us start with the topological simplicies. For every integer<br />

n ≥ 0, a st<strong>and</strong>ard n-simplex is defined as a subspace <strong>of</strong> the<br />

(n + 1)-dimensional real vector space R n+1


40 2 Simplicial methods<br />

△ n = {(x 0 , x 1 , · · · , x n ) ∈ R n+1 ∣ ∣ xi ≥ 0 <strong>and</strong><br />

n∑<br />

x i = 1}.<br />

An i-face <strong>of</strong> △ n for 0 ≤ i ≤ n is a subspace such that there are<br />

exactly i+1 chosen entries that are not constantly zero. Each<br />

i-face is isomorphic to the st<strong>and</strong>ard i-simplex. For example<br />

in the following picture, △ 2 has exactly one 2-face (e.g. △ 2 ),<br />

three 1-faces (isomorphic to △ 1 ) <strong>and</strong> three 0-faces (isomorphic<br />

to △ 0 which is a point).<br />

i=0<br />

X 3<br />

(0,0,1)<br />

△ 2 ⊂ R 3<br />

X 2<br />

(1,0,0) (0,1,0)<br />

X 1<br />

Fix an integer n, there are many natural maps among all<br />

the faces <strong>of</strong> △ n . Since each face is isomorphic to a st<strong>and</strong>ard<br />

simplex, these maps can be described as maps among all st<strong>and</strong>ard<br />

simplicies. The most distinguished are the face maps<br />

d i : △ n−1 → △ n (inserting a zero) <strong>and</strong> the degeneracy maps<br />

s i : △ n+1 → △ n (adding up two adjacent entries), given explicitly<br />

by<br />

d i (x 0 , x 1 , · · · , x n−1 ) = (x 0 , · · · , x i , 0, x i+1 , · · · , x n−1 ), 0 ≤ i ≤ n;<br />

s i (x 0 , x 1 , · · · , x n+1 ) = (x 0 , · · · , x i + x i+1 , · · · , x n+1 ), 0 ≤ i ≤ n.<br />

Given a topological space X, in order to define <strong>and</strong> compute<br />

its singular homology H∗(X, Z) we first form the sets <strong>of</strong><br />

continuous maps S(X) n = Hom T op (△ n , X) <strong>and</strong> then produce<br />

free abelian groups ZS(X) n on top <strong>of</strong> them. Each face map


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 41<br />

d i : △ n−1 → △ n induces a map d i : S(X) n → S(X) n−1 , <strong>and</strong><br />

moreover ∂ n = ∑ n<br />

i=0 (−1)i d i : ZS(X) n → ZS(X) n−1 satisfies<br />

∂ n+1 ∂ n = 0. In this way we obtain a non-negatively graded<br />

chain complex<br />

{ZS(X) ∗ ; ∂ ∗ } ∗≥0 ,<br />

in which every differential ∂ n is determined by an alternating<br />

sum <strong>of</strong> face maps. The homology <strong>of</strong> this complex is defined<br />

to be the singular homology <strong>of</strong> X, written as H∗(X, Z). The<br />

degeneracy maps are important to us as well, <strong>and</strong> we shall<br />

discuss their roles shortly.<br />

In order to allow further applications <strong>of</strong> such a fundamental<br />

construction, we propose some alternative descriptions <strong>of</strong> the<br />

st<strong>and</strong>ard simplices based on the following two observations.<br />

Firstly, to specify a st<strong>and</strong>ard complex, it suffices to provide its<br />

vertices because △ n is the convex hull <strong>of</strong> the set <strong>of</strong> its vertices.<br />

Moreover since, in R n+1 , one can give the lexicographic order<br />

to its elements by asking (x 0 , · · · , x n ) < (y 0 , · · · , y n ) if there<br />

exists an integer 0 ≤ k ≤ n such that x i = y i for i < k <strong>and</strong><br />

x k < y k , particularly there is an order on the set <strong>of</strong> vertices <strong>of</strong><br />

△ n . This totally ordered set <strong>of</strong> vertices uniquely determines<br />

△ n . For example in R 3 we have (0, 0, 1) < (0, 1, 0) < (1, 0, 0),<br />

<strong>and</strong> we can certainly identify this totally ordered set with △ 2 .<br />

Thus giving △ n is equivalent to giving the totally ordered set<br />

(0, · · · , 0, 1) < (0, · · · , 1, 0) < · · · < (1, 0, · · · , 0) <strong>of</strong> points in<br />

R n+1 . Moreover there exists a natural one-to-one correspondence<br />

between the set <strong>of</strong> all totally ordered subsets <strong>of</strong> vertices<br />

<strong>and</strong> the set <strong>of</strong> faces <strong>of</strong> △ n .<br />

Secondly, because the face <strong>and</strong> degeneracy maps on △ n<br />

are completely determined by their values on the vertices,<br />

we can translate the face <strong>and</strong> degeneracy maps accordingly.<br />

For example, we can write out values <strong>of</strong> d 0 : △ 2 → △ 3 <strong>and</strong><br />

s 0 : △ 2 → △ 1 on the vertices


42 2 Simplicial methods<br />

d 0 (0, 0, 1) = (0, 0, 0, 1), d 0 (0, 1, 0) = (0, 0, 1, 0), d 0 (1, 0, 0) = (0, 1, 0, 0<br />

s 0 (0, 0, 1) = (0, 1), s 0 (0, 1, 0) = (1, 0), s 0 (1, 0, 0) = (1, 0).<br />

In general it is easy to see that d i <strong>and</strong> s i always send vertices<br />

to vertices. Furthermore they (weakly) preserve the order, in<br />

the sense that if a ≤ b then d i (a) ≤ d i (b) <strong>and</strong> s i (a) ≤ s i (b).<br />

We shall illustrate it by an example. Let us examine △ 1 =<br />

{(0, 1) < (1, 0)} <strong>and</strong> △ 2 = {(0, 0, 1) < (0, 1, 0) < (1, 0, 0)}.<br />

The face maps △ 1 → △ 2 , adapted to our new combinatorial<br />

expression, are given by embeddings<br />

d 0 : (0, 1) < (1, 0) ↦→ (0, 0, 1) < (0, 1, 0)<br />

d 1 : (0, 1) < (1, 0) ↦→ (0, 0, 1) < (1, 0, 0)<br />

d 2 : (0, 1) < (1, 0) ↦→ (0, 1, 0) < (1, 0, 0).<br />

The degeneracy maps △ 2 → △ 1 are the same as projecting<br />

the vertex (0, 1, 0) to one <strong>of</strong> the other two, upon identifying<br />

△ 1 with the line segment (0, 0, 1) − (1, 0, 0)<br />

s 0 : (0, 0, 1) < (0, 1, 0) < (1, 0, 0) ↦→ (0, 1) < (1, 0) = (1, 0) (0, 1) <<br />

s 1 : (0, 0, 1) < (0, 1, 0) < (1, 0, 0) ↦→ (0, 1) = (0, 1) < (1, 0) (0, 1) <<br />

In summary, the face <strong>and</strong> degeneracy maps are indeed weakly<br />

monotonic maps among those totally ordered sets corresponding<br />

to the st<strong>and</strong>ard simplices. As a matter <strong>of</strong> fact, all weakly<br />

monotonic functions among those totally ordered sets, coming<br />

from st<strong>and</strong>ard simplices, are composites <strong>of</strong> these face <strong>and</strong><br />

degeneracy maps.<br />

For various good reasons we continue to work on the combinatorial<br />

characterizations <strong>of</strong> st<strong>and</strong>ard simplices. Previously<br />

we have identified △ n with the poset (0, · · · , 0, 1) < (0, · · · , 1, 0) <<br />

· · · < (1, 0, · · · , 0), <strong>and</strong> have rewritten the face <strong>and</strong> degeneracy<br />

maps. Now we abstract the totally ordered set as<br />

0 < 1 < · · · < n. Consequently the face <strong>and</strong> degeneracy


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 43<br />

maps can be reformulated <strong>and</strong> will be denoted by d i <strong>and</strong> s i ,<br />

respectively. This reformulation allows us to forget the geometric<br />

definition <strong>of</strong> △ n , <strong>and</strong> thus all relevant constructions<br />

can be made in an entirely combinatorial fashion. For future<br />

applications we write out the ith face map d i (corresponding<br />

to d i ) for 0 ≤ i ≤ n<br />

{0 < 1 < · · · < n − 1} → {0 < 1 < · · · < n}<br />

{<br />

d i j, if j < i<br />

(j) =<br />

j + 1 , if j ≥ i<br />

<strong>and</strong> the ith degeneracy map s i (corresponding to s i ) for 0 ≤<br />

i ≤ n<br />

{0 < 1 < · · · < n + 1} → {0 < 1 < · · · < n}<br />

{<br />

s i j, if j ≤ i<br />

(j) =<br />

j − 1 , if j > i<br />

One can verify that these maps satisfy the relations (the<br />

cosimplicial identities)<br />

d j d i = d i d j−1 i < j<br />

s j d i = d i s j−1 i < j<br />

s j d i = 1 i = j or j + 1<br />

s j d i = d i−1 s j i > j + 1<br />

s j s i = s i s j+1 i ≤ j .<br />

For each n ∈ {0} ∪ N, we define an ordered set (which happens<br />

to be a finite category) n = 0 < 1 < 2 < · · · < n<br />

(or rather 0 → 1 → 2 → · · · → n). We denote by △ the<br />

category consisting <strong>of</strong> all such n, in which morphisms are<br />

weakly monotonic functions (equivalently, functors) among<br />

these ordered sets. Recall the motivating example at the<br />

very beginning <strong>of</strong> this section. Given a space X, in order


44 2 Simplicial methods<br />

to make use <strong>of</strong> △, we associated to each n, the combinatorial<br />

model <strong>of</strong> △ n , a set Hom T op (△ n , X), as well as an abelian<br />

group ZHom T op (△ n , X). If we put all pieces together, we realize<br />

that we actually constructed a contravariant functor<br />

Hom T op (−, X) : △ → Set, as well as a ZHom T op (−, X) :<br />

△ → Ab. Since these functors give us the singular homology<br />

<strong>of</strong> X, contravariant functors from △ to other categories may<br />

lead to interesting constructions too.<br />

Definition 2.2.1. A simplicial object in a category T is a<br />

contravariant functor X : △ → T . Two simplicial objects<br />

X, Y in T are said to be isomorphic, written as X ∼ = Y , if<br />

as functors they are naturally equivalent.<br />

A simplicial set is a simplicial object in Set.<br />

Equivalently, since monotonic maps are compositions <strong>of</strong> d i<br />

<strong>and</strong> s i , a simplicial object X in T consists <strong>of</strong> a set <strong>of</strong> objects<br />

X n := X([n]) ∈ Ob T (an element <strong>of</strong> X n is called an n-<br />

simplex) <strong>and</strong> morphisms among these simplicies which are<br />

composites <strong>of</strong> two special kinds <strong>of</strong> morphisms: d i := X(d i ) :<br />

X n → X n−1 , 0 ≤ i ≤ n, <strong>and</strong> s i := X(s i ) : X n → X n+1 ,<br />

0 ≤ i ≤ n, satisfying (the simplicial identities)<br />

d i d j = d j−1 d i i < j<br />

d i s j = s j−1 d i i < j<br />

d i s j = 1 i = j or j + 1<br />

d i s j = s j d i−1 i > j + 1<br />

s i s j = s j+1 s i i ≤ j .<br />

In order to define a simplicial object X, it suffices to specify<br />

{X n } n≥0 ⊂ Ob T , together with some maps d i <strong>and</strong> s i satisfying<br />

the above relations.<br />

We shall be interested in the cases where T = Set or R-Mod<br />

for some commutative ring R with identity. By Proposition


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 45<br />

1.1.27, the two categories SimpSet <strong>and</strong> Simp(R-Mod) are<br />

both complete <strong>and</strong> cocomplete.<br />

Example 2.2.2.1. Let C be a small category. We define the<br />

nerve <strong>of</strong> C to be a simplicial set NC = Hom Cat (−, C) such<br />

that NC n = C n (indeed Ob(C n ) since we only need the underlying<br />

set). Alternatively NC n can be identified with the<br />

set <strong>of</strong> n-chains <strong>of</strong> morphisms in C if n > 0 <strong>and</strong> Ob C if n = 0.<br />

When n > 0, the ith face map d i : NC n+1 → NC n is<br />

α<br />

d i (x 0 → · · · → x<br />

i α i+1<br />

i−1→xi → xi+1 → · · · → x n+1 )<br />

α<br />

= x 0 → · · · → x<br />

i α i+1<br />

i−1→ ̂xi → xi+1 → · · · → x n+1 ,<br />

where ∧ means removing an object. For instance when n = 1<br />

α<br />

the three face maps NC 2 → NC 1 are d 0 (x<br />

1 α 2<br />

0→x1→x2 ) =<br />

α<br />

x<br />

2<br />

α<br />

1→x2 , d 1 (x<br />

1 α 2<br />

α<br />

0→x1→x2 ) = x<br />

2 α 1<br />

α<br />

0 → x2 <strong>and</strong> d 0 (x<br />

1 α 2<br />

0→x1→x2 ) =<br />

α<br />

x<br />

1<br />

0→x1 . The ith degeneracy map s i : NC n → NC n+1 is given<br />

by<br />

s i (x 0 → · · · → x i−1 → x i → x i+1 → · · · → x n )<br />

= x 0 → · · · → x i−1 → x i<br />

1 xi<br />

→xi → x i+1 → · · · → x n .<br />

One can verify that these maps satisfy the simplicial identities.<br />

2. We consider a special situation <strong>of</strong> the first example. It<br />

shall fill the gap which may have occurred during transition<br />

from the geometric definition <strong>of</strong> st<strong>and</strong>ard n-simplicies to<br />

the abstract category-theoretic reformulation. We fix a nonnegative<br />

integer m. The totally ordered set m is indeed a<br />

category 0 → 1 → · · · → m. Thus we can consider its nerve<br />

Nm. By the preceding example we have Nm n = Ob(m n ).<br />

In other words the combinatorial m-simplex m corresponds<br />

to the functor Hom △ (−, m) : △ → Set. This correspondence<br />

actually extends to a (covariant) functor, given by


46 2 Simplicial methods<br />

ι(m) = Hom △ (−, m),<br />

By Yondeda Lemma we have<br />

ι : △ → SimpSet = Set △ .<br />

Hom SimpSet (Hom △ (−, m), Hom △ (−, n)) ∼ = Hom △ (m, n).<br />

It means ι is fully faithful.<br />

In the next section we shall talk about the geometric realization<br />

<strong>of</strong> simplicial sets. Then we will see that the simplicial<br />

set Hom △ (−, m) precisely gives rise to △ m .<br />

3. Let A be an associative ring <strong>and</strong> A-Mod the category <strong>of</strong> all<br />

A-modules. A simplicial object in A-Mod is called a simplicial<br />

A-module. From Set to A-Mod, there exists a natural<br />

covariant functor, given by constructing free A-modules.<br />

Suppose X : △ → Set is a simplicial set. Then we naturally<br />

obtain a simplicial A-module<br />

A[X] : △ → Set → A-Mod.<br />

Given a simplicial A-module Y , one can construct a (nonnegatively<br />

graded) chain complex <strong>of</strong> A-modules by defining<br />

C n (Y ) = Y n <strong>and</strong> ∂ n = ∑ n<br />

i=0 (−1)i d i : C n (Y ) → C n−1 (Y ). If<br />

Y comes from a simplicial set X, that is, Y = A[X], then<br />

we write the chain complex as C ∗ (X, A) instead <strong>of</strong> C ∗ (Y )<br />

or C ∗ (A[X]) for various good reasons.<br />

The following characteristic statement is enlightening. Suppose<br />

X is a simplicial set. Then we define the corresponding<br />

simplex category to be ι/X, where ι : △ → SimpSet is the<br />

functor in Example 2.2.2 (2).<br />

Theorem 2.2.3. Let X be a simplicial set. Then we have<br />

an isomorphism<br />

X ∼ = lim −→ι/X<br />

P,


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 47<br />

where P : ι/X → SimpSet is given by P(n, Φ) = Hom △ (−, n).<br />

Pro<strong>of</strong>. From the contravariant version <strong>of</strong> Theorem 1.1.28 we<br />

know for the functor X : △ op → Set we can define a category<br />

D <strong>and</strong> a functor ˜X : D → Set<br />

△ op = SimpSet such that<br />

X ∼ = lim −→D ˜X.<br />

We show the category D is exactly ι/X <strong>and</strong> subsequently ˜X<br />

can be identified with P.<br />

The objects in the overcategory ι/X are <strong>of</strong> the form (n, Φ),<br />

where Φ : ι(n) = Hom △ (−, n) → X is a simplicial map. By<br />

the Yoneda Lemma Hom SimpSet (Hom △ (−, n), X) ∼ = X(n) =<br />

X n . Thus Φ can be regarded as an element <strong>of</strong> the set X(n).<br />

This provides a bijection between Ob(ι/X) <strong>and</strong> Ob D. From<br />

here one can continue to finish the pro<strong>of</strong>. We leave the details<br />

for the reader.<br />

⊓⊔<br />

Definition 2.2.4. Let C be a small category <strong>and</strong> NC its<br />

nerve. Suppose R is a commutative ring with identity. Then<br />

the simplicial R-module R[NC] gives rise to a complex <strong>of</strong> R-<br />

modules, written as C ∗ (C, R). The n-th homology <strong>of</strong> C ∗ (C, R),<br />

denoted by Hn(C, R), is called the n-th homology <strong>of</strong> C with<br />

coefficients in R. The n-th cohomology <strong>of</strong> the cochain complex<br />

C ∗ (C, R) := C ∗ (C, R) ∧ = Hom R (C ∗ (C, R), R),<br />

denoted by H n (C, R), is called the n-th cohomology <strong>of</strong> C with<br />

coefficients in R.<br />

By direct calculation, one can see that H0(C, R) ∼ = H 0 (C, R)<br />

is a free R-module with rank equal to the number <strong>of</strong> connected<br />

components <strong>of</strong> C. For any chain complex C ∗ (C, R), define as<br />

above, we can insert the base ring R at degree -1 <strong>and</strong> obtain<br />

the so-called reduced chain complex ˜C∗ (C, R) = C ∗ (C, R) →


48 2 Simplicial methods<br />

R → 0. Then the homology <strong>of</strong> the reduced chain complex is<br />

called the reduced homology <strong>of</strong> C, written as ˜H∗ (C, R). We<br />

see ˜H−1 (C, R) = 0 <strong>and</strong> ˜H0 (C, R) is a free R-module with rank<br />

equal to rk RH0(C, R) − 1. A small category C is called R-<br />

acyclic if ˜H∗ (C, R) vanishes.<br />

Now we address the issue <strong>of</strong> degeneracy maps. Suppose X is<br />

a simplicial object in an abelian category T . Then we have a<br />

complex <strong>of</strong> objects in T , denoted by X ∗ . It has a subcomplex<br />

X ′ ∗ such that X ′ n = ∑ i s i(X n−1 ) if n > 0 <strong>and</strong> X ′ 0 = 0. Then<br />

we continue to define a quotient complex X † ∗ = X ∗ /X ′ ∗. We<br />

can also define the normalized complex <strong>of</strong> X, N ∗ (X), by<br />

N n (X) =<br />

n−1<br />

⋂<br />

i=0<br />

Ker(d i )<br />

<strong>and</strong> ∂ n = (−1) n d n . This is a subcomplex <strong>of</strong> X ∗ .<br />

Theorem 2.2.5.1. We have X ∗ = N ∗ (X)⊕X ∗ ′ <strong>and</strong> N ∗ (X) ∼ =<br />

X ∗. † Furthermore the quotient map X ∗ ↠ X ∗<br />

† induces a<br />

chain homotopy equivalence <strong>and</strong> X ∗ ′ is contractible.<br />

2. (Dold-Kan Correspondence) Suppose T is an abelian<br />

category <strong>and</strong> Ch ≥0 (T ) is the category <strong>of</strong> non-negatively<br />

graded chain complexes in T . Then the normalized chain<br />

complex functor functor<br />

is a category equivalence.<br />

N : SimpT → Ch ≥0 (T ),<br />

Pro<strong>of</strong>. For the first part, see [53, Chapter VIII, Section] <strong>and</strong><br />

[84, Section 8.3], <strong>and</strong> for the second part, the reader is referred<br />

to [84, Section 8.4].<br />

⊓⊔<br />

Because <strong>of</strong> the above theorem, we also call X † ∗ the normalized<br />

complex <strong>of</strong> X ∗ .


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 49<br />

When T = A-Mod, <strong>and</strong> X is a simplicial A-module, then<br />

X ′ n, for any n > 0, consists <strong>of</strong> degenerate elements <strong>of</strong> the form<br />

s i (x ′ ) for some x ′ ∈ X n−1 . When X = R[NC], a simplicial R-<br />

module defined over a small category C <strong>and</strong> a commutative<br />

ring R, the degenerate elements in R[NC] n are all the linear<br />

combinations <strong>of</strong> the degenerate elements in NC n which are<br />

exactly those n-chains containing an identity morphism. In<br />

practice, when we compute (co-)homology <strong>of</strong> a category, we<br />

only need to use the normalized complex C † ∗(C, R).<br />

Example 2.2.6.1. Let C be the following category with two<br />

objects, two identity <strong>and</strong> non-identity morphisms<br />

x<br />

α<br />

β<br />

To compute H∗(C, Z), we only have to write down the normalized<br />

chain complex 0 → C † 1 (C, Z) → C† 0 (C, Z) → 0<br />

y<br />

0 → Z{α, β} → Z{x, y} → 0.<br />

(By comparison, C ∗ (C, Z) is infinite.) The non-trivial differential<br />

is given by α ↦→ y − x <strong>and</strong> β ↦→ y − x. Then<br />

the only non-trivial homology groups are H0(C, Z) ∼ = Z <strong>and</strong><br />

H1(C, Z) ∼ = Z. The calculation <strong>of</strong> cohomology groups is left<br />

to the reader.<br />

2. The second category D is slightly different from the first.<br />

One can easily verify that the normalized complex C † ∗(D, Z)<br />

is infinite, <strong>of</strong> dimension two at each degree. By direct computation<br />

H0(D, Z) ∼ = Z is the only non-trivial homology<br />

group.<br />

α<br />

<br />

x y<br />

α −1


50 2 Simplicial methods<br />

However there is an easy way to see it, if one notices that<br />

D ≃ • <strong>and</strong> knows that a category equivalence induces isomorphism<br />

on homology (see Proposition 2.2.19).<br />

Definition 2.2.7. Let D <strong>and</strong> C be two small categories.<br />

Then the join <strong>of</strong> D with C, denoted by D ∗ C, is a category<br />

whose objects are Ob D ∪ Ob C, <strong>and</strong> whose morphisms<br />

are Mor D ∪ Mor C, plus exactly one extra morphism γ a,x ∈<br />

Hom D∗C (a, x) introduced for every pair <strong>of</strong> objects a ∈ Ob D<br />

<strong>and</strong> x ∈ Ob C. The composition laws in D ∗ C are determined<br />

by the composition laws in C <strong>and</strong> D, plus the equalities<br />

αγ a,x = γ a,y , γ a,x β = γ b,x for any α ∈ Hom C (x, y) <strong>and</strong><br />

β ∈ Hom D (b, a).<br />

By definition D ∗ C <strong>and</strong> C ∗ D are two different categories.<br />

One can easily construct N(D ∗ C) from ND <strong>and</strong> NC.<br />

Proposition 2.2.8. Suppose both D <strong>and</strong> C are connected<br />

small categories. If R = k is a field, then<br />

˜Hn(D ∗ C, k) ∼ =<br />

⊕ ˜Hi(D, k) ⊗ ˜Hj (C, k)<br />

i+j=n−1<br />

if n ≥ 0. Particularly H∗(D ∗ C, k) ∼ = H∗(C ∗ D, k).<br />

Pro<strong>of</strong>. For n ≥ 0, C n (D ∗ C, k) = k[N(D ∗ C)] n has a basis<br />

consisting <strong>of</strong> the following elements<br />

γx i ,y i+1<br />

x 0 → · · · → x i −→ yi+1 → · · · → y n<br />

where x 1 , · · · , x i ∈ Ob D <strong>and</strong> y i+1 , · · · , y n ∈ Ob C, for 0 ≤<br />

i ≤ n. Here γ xi ,y i+1<br />

is the unique morphism in Hom D∗C (x i , y i+1 ).<br />

Its differential is


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 51<br />

∑ i−1<br />

j=0 (−1)j γx i ,y i+1<br />

x 0 → · · · → ˆx j → · · · → x i −→ yi+1 → · · · → y n<br />

+ (−1) i γx i−1 ,y i+1<br />

x 0 → · · · → x i−1 −→ yi+1 → · · · → y n<br />

+ (−1) i+1 γx i ,y i+2<br />

x 0 → · · · → x i −→ yi+2 → · · · → y n<br />

+ ∑ n<br />

l=i+2 (−1)l γx i ,y i+1<br />

x 0 → · · · → x i −→ yi+1 → · · · → ŷ l → · · · → y n .<br />

Suppose ˜C∗ (D ∗ C, k), ˜C∗ (D, k) <strong>and</strong> ˜C∗ (C, k) are the reduced<br />

chain complexes with k inserted in degree -1. Then we can<br />

define a degree -1 chain map<br />

˜C ∗ (D ∗ C, k) → ˜C∗ (D, k) ⊗ ˜C∗ (C, k)<br />

such that, at degree −1 it is identity <strong>and</strong> at degree n ≥ 0,<br />

˜C n (D ∗ C, k) → [˜C∗ (D, k) ⊗ ˜C∗ (C, k)] n−1 is given by<br />

γx i ,y i+1<br />

x 0 → · · · → x i −→ yi+1 → · · · → y n ↦→ [x 0 → · · · → x i ]⊗[y i+1 → · · ·<br />

along with<br />

<strong>and</strong><br />

x 0 → · · · → x n ↦→ [x 0 → · · · → x n ] ⊗ 1,<br />

y 0 → · · · → y n ↦→ 1 ⊗ [x 0 → · · · → x n ].<br />

Here 1 ∈ k = ˜C−1 (D, k) = ˜C−1 (C, k). This chain map is an<br />

isomorphism <strong>and</strong> thus by Künneth formula we get<br />

˜Hn(D ∗ C, k) ∼ = ⊕ i+j=n−1 Hi(˜C∗ (D, k)) ⊗ Hj(˜C∗ (C, k))<br />

∼ =<br />

⊕i+j=n−1 ˜Hi(D, k) ⊗ ˜Hj (C, k).<br />

Example 2.2.6 (2) tells us that it is important to know how to<br />

compare various categories or more generally simplicial sets.<br />

Definition 2.2.9. A natural transformation Φ : X → Y between<br />

simplicial objects in T is called a simplicial map. The<br />

simplicial objects in T form a category SimpT .<br />

⊓⊔


52 2 Simplicial methods<br />

By definition a simplicial map Φ consists <strong>of</strong> a sequence <strong>of</strong><br />

morphisms in T , Φ n : X n → Y n , commuting with the relevant<br />

face <strong>and</strong> degeneracy maps.<br />

Lemma 2.2.10. Let R be a commutative ring with identity.<br />

A simplicial map Φ : X → Y between simplicial sets induces<br />

a chain map CΦ : C ∗ (X, R) → C ∗ (Y, R).<br />

Pro<strong>of</strong>. Because every Φ n : X n → Y n commutes with d i , it commutes<br />

with the differentials <strong>of</strong> these chain complexes which<br />

are alternating sums <strong>of</strong> d i ’s.<br />

⊓⊔<br />

Lemma 2.2.11. Suppose D <strong>and</strong> C are two small categories<br />

<strong>and</strong> Φ : ND → NC is a simplicial map. Then Φ determines<br />

a functor u : D → C which in turn gives rise to Φ as Nu.<br />

Pro<strong>of</strong>. The map Φ 0 : ND 0 → NC 0 gives an assignment<br />

Ob D → Ob C while Φ 1 : ND 1 → NC 1 gives an assignment<br />

Mor D → Mor C. We show that these define a functor<br />

u : D → C such that u(d) = Φ 0 (d) if d ∈ Ob D <strong>and</strong><br />

u(f) = Φ 1 (f) if f ∈ Mor D.<br />

First <strong>of</strong> all, Φ 0 <strong>and</strong> Φ 1 are compatible in the sense that if<br />

c f →d ∈ ND 1 then u(f) = Φ 1 (f) is a morphism from u(c) =<br />

Φ 0 (c) to u(d) = Φ 0 (d). This follows from the commutative<br />

diagram by choosing i = 0 or 1<br />

Φ<br />

ND<br />

1 <br />

1 NC 1<br />

d i d i<br />

<br />

<br />

<br />

ND 0 Φ0<br />

NC 0<br />

Second <strong>of</strong> all, we have to demonstrate u(fg) = u(f)u(g) for<br />

any two composable morphisms in Mor D. This time we just<br />

use a similar commutative diagram involving d 1 , Φ 1 <strong>and</strong> Φ 2 .<br />

At last we can show u(1 d ) = 1 u(d) by invoking a degeneracy<br />

map


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 53<br />

ND 1<br />

Φ 1 <br />

NC 1<br />

s 0<br />

<br />

s 0<br />

ND<br />

0 Φ0<br />

NC 0<br />

Since Φ is completely determined by Φ 0 <strong>and</strong> Φ 1 , one can prove<br />

that Nu is exactly Φ.<br />

⊓⊔<br />

Example 2.2.12. Suppose u : D → C is a functor between<br />

small categories <strong>and</strong> R is a commutative ring with identity.<br />

Then it induces a simplicial map Nu : ND → NC. Furthermore<br />

it induces a chain map Cu : C ∗ (D, R) → C ∗ (C, R).<br />

Corollary 2.2.13. A functor u : D → C induces a map u ∗ :<br />

H∗(D, R) → H∗(C, R) as well as u ∗ : H ∗ (C, R) → H ∗ (D, R).<br />

Pro<strong>of</strong>. The map u ∗ is induced by Cu while u ∗ is induced by<br />

Hom R (Cu, R).<br />

⊓⊔<br />

We shall illustrate the previous comparison results by an<br />

example. Note that if there are two functors v : E → D<br />

<strong>and</strong> u : D → C, we can easily verify that (uv) ∗ = u ∗ v ∗ <strong>and</strong><br />

(uv) ∗ = v ∗ u ∗ .<br />

Example 2.2.14. Let us consider the following category C with<br />

four non-identity morphisms<br />

µ <br />

z<br />

Its opposite category is pictured as<br />

x<br />

α<br />

β<br />

γ<br />

y


54 2 Simplicial methods<br />

y op α op <br />

β op <br />

γ x op<br />

op <br />

z op µ op<br />

In order not to make confusion in the opposite category we<br />

write x op for x etc. By direct calculation we can find that<br />

H0(C, Z) ∼ = H0(C op , Z) ∼ = Z <strong>and</strong> H1(C, Z) ∼ = H1(C op , Z) ∼ =<br />

Z ⊕ Z. These are the only non-trivial homology groups.<br />

We can list all (covariant) functors from C → C op , since<br />

there are not many. Suppose u : C → C op is a functor. We<br />

write u(C) to be the image <strong>of</strong> C, a subcategory <strong>of</strong> C op .<br />

1. If u(x) = x op , then both u(y) <strong>and</strong> u(z) have to be x op . In this<br />

case u(C) is a trivial category <strong>and</strong> has the only non-trivial<br />

homology at degree zero, which is Z.<br />

2. If u(x) = y op , then there are several possibilities. In any case<br />

we must have u(y) = u(z). Firstly, we may have u(C) =<br />

{y op }. Then Hi(u(C), Z) ∼ = Z if i = 0 or zero otherwise.<br />

Secondly we may have u(C) equals<br />

or<br />

<br />

y op α op x op<br />

β op<br />

y op <br />

x op .<br />

In the former situation we will have H0(u(C), Z) ∼ = H1(u(C), Z) ∼ =<br />

Z <strong>and</strong> zero otherwise. In the latter situation we have only<br />

non-trivial homology H0(u(C), Z) ∼ = Z.<br />

3. If u(x) = z op , then it is similar to 2.<br />

Since u can be decomposed into a sequence C ↠ u(C) ↩→<br />

C op , u ∗ is also the composite <strong>of</strong> H∗(C, Z) → H∗(u(C), Z) →<br />

H∗(C op , Z). Under the circumstance we underst<strong>and</strong> u ∗ com-


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 55<br />

pletely. A crucial fact is that, our previous calculations assert<br />

that there exists no such functor u : C → C op that induces the<br />

isomorphism <strong>of</strong> graded abelian groups H∗(C, Z) ∼ = H∗(C op , Z).<br />

Simplicial maps are used to compare two simplicial sets (<strong>and</strong><br />

resulting complexes). Now we introduce a way to compare two<br />

simplicial maps.<br />

Definition 2.2.15. For any two simplicial sets X, Y , we can<br />

define the Cartesian product X × Y by (X × Y ) n = X n × Y n<br />

with face <strong>and</strong> degeneracy maps d i = (d X i , dY i ), s i = (s X i , sY i ).<br />

Example 2.2.16. If D <strong>and</strong> C are two small categories, then<br />

N(D × C) ∼ = ND × NC.<br />

For brevity we denote by {0, · · · , 0, 1, · · · , 1} the element<br />

0 = · · · = 0 < 1 = · · · = 1 <strong>of</strong> the set N1 n with exactly i<br />

copies <strong>of</strong> 1 for 0 ≤ i ≤ n + 1.<br />

Definition 2.2.17. Let Φ, Ψ : X → Y be two simplicial<br />

maps between simplicial sets. We say Φ is simplicially<br />

homotopic to Ψ if there exists a natural transformation<br />

: X × N1 → Y such that n | Xn ×{(0,··· ,0)} = Φ n <strong>and</strong><br />

n | Xn ×{(1,··· ,1)} = Ψ n for all n ≥ 0. We call a simplicial<br />

homotopy from Φ to Ψ, written as : Φ→Ψ ≃ or simply Φ ≃ Ψ.<br />

We say two simplicial sets X <strong>and</strong> Y are homotopic, written<br />

as X ≃ Y , if there exist natural transformations Φ : X → Y<br />

<strong>and</strong> Ψ : Y → X such that ΨΦ ≃ Id X <strong>and</strong> ΦΨ ≃ Id Y .<br />

There are two canonical simplicial maps ι 0 , ι 1 : N0 → N1<br />

by sending {0, · · · , 0} to {0, · · · , 0} or {1, · · · , 1}, respectively.<br />

Based on this observation, being a simplicial homotopy<br />

is equivalent to having the following commutative diagram


56 2 Simplicial methods<br />

X ∼ = X × N0<br />

Id X ×ι 0<br />

<br />

X × N1<br />

Id X ×ι 1<br />

X × N0 ∼ = X<br />

Φ<br />

<br />

Ψ<br />

Y .<br />

Since (X × N1) n consists <strong>of</strong> n + 2 copies <strong>of</strong> X n in the<br />

form <strong>of</strong> X n × {0, · · · , 1, · · · , 1}, combinatorially, the above<br />

definition is equivalent to saying that there exist maps ′ i =<br />

s i n | Xn ×{0,··· ,1,··· ,1} : X n → Y n+1 for 0 ≤ i ≤ n such that<br />

d 0 ′ 0 = Φ n<br />

d n+1 ′ n = Ψ n<br />

d i ′ j = ′ j−1 d i i < j<br />

d j+1 ′ j+1 = d j+1 ′ j<br />

d i ′ j = ′ j d i−1 i > j + 1<br />

s i ′ j = ′ j+1 s i i ≤ j<br />

s i ′ j = ′ j s i−1 i > j.<br />

Remember in the definition <strong>of</strong> ′ i, {0, · · · , 1, · · · , 1} denotes<br />

the element 0 = · · · = 0 < 1 = · · · = 1 in N1 with i copies<br />

<strong>of</strong> 1.<br />

Definition 2.2.18. If the induced simplicial map Npt : NC →<br />

N• <strong>of</strong> the canonical functor pt : C → • gives rise to a homotopy<br />

NC ≃ N•, then we say C is contractible.<br />

Note that there always exist various functors • → C. If<br />

C is contractible, then any functor ι : • → C will induce<br />

N• ≃ NC.<br />

Proposition 2.2.19. If Φ : u → u ′ is a natural transformation<br />

between two functors u, u ′ : D → C, then Nu is<br />

homotopic to Nu ′ . Consequently Cu <strong>and</strong> Cu ′ are chain homotopic.<br />

In particular, if ND ≃ NC, then C ∗ (D, R) <strong>and</strong> C ∗ (C, R)<br />

are chain homotopy equivalent. Hence H∗(D, R) ∼ = H∗(C, R)<br />

<strong>and</strong> H ∗ (D, R) ∼ = H ∗ (C, R).


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 57<br />

Pro<strong>of</strong>. We can define a functor ˜Φ : D × 1 → C by ˜Φ(a, 0) =<br />

u(a), ˜Φ(a, 1) = u ′ (a) <strong>and</strong> ˜Φ(α, 1{0} ) = u(α), ˜Φ(α, 1{1} ) =<br />

u ′ (α) <strong>and</strong> ˜Φ(1a ,


58 2 Simplicial methods<br />

when T possesses all limits. This is the historical context for<br />

the Kan extensions.<br />

2.2.2 Classifying spaces<br />

It is <strong>of</strong>ten very useful to have some knowledge about classifying<br />

spaces when computing the (co-)homology <strong>of</strong> a small<br />

category. In this section we recall several important results as<br />

facts, in order to obtain a balanced view towards key results<br />

in the preceding section.<br />

Definition 2.2.22. The geometric realization |X|, <strong>of</strong> a simplicial<br />

set X, is the quotient <strong>of</strong><br />

⋃<br />

X n × △ n<br />

n≥0<br />

by the equivalence relation given by (d i x, y) ∼ (x, d i y) <strong>and</strong><br />

(s i x, y) ∼ (x, s i y).<br />

Definition 2.2.23. The geometric realization <strong>of</strong> NC, the<br />

nerve <strong>of</strong> a small category C, is called the classifying space<br />

<strong>of</strong> C, customarily denoted by BC.<br />

In the literature the classifying space <strong>of</strong> a small category is<br />

<strong>of</strong>ten written as |C|. However in these notes we try to avoid<br />

this notation since when C is a finite group, it has been used<br />

to denote the order <strong>of</strong> the group.<br />

In the following examples, the first two are not hard to verify<br />

while the other two requires further knowledge from algebraic<br />

topology so we point out places where the reader may find<br />

pro<strong>of</strong>s.<br />

Example 2.2.24.1. The classifying space <strong>of</strong> n is the st<strong>and</strong>ard<br />

n-simplex △ n .<br />

2. The classifying space <strong>of</strong> the following category is S 1 . Recall<br />

its integral homology <strong>and</strong> compare with Example 2.2.6 (1).


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 59<br />

x<br />

3. The classifying space <strong>of</strong> Z 2 is RP ∞ . One can see this by using<br />

Milnor’s construction EZ 2 . Then up to homotopy BZ 2 ≃<br />

EZ 2 /Z 2 .<br />

4. Let D <strong>and</strong> C be two small categories. There is a natural way<br />

to define the join <strong>of</strong> ND <strong>and</strong> NC. Then N(D∗C) ∼ = ND∗NC<br />

<strong>and</strong> consequently B(D ∗ C) ≃ BD ∗ BC, see [26].<br />

The following theorem asserts that the classifying spaces are<br />

“good”.<br />

Theorem 2.2.25. The space |X| is a CW-complex having<br />

one n-cell for each non-degenerate n-simplex <strong>of</strong> X.<br />

It explains the structures <strong>of</strong> classifying spaces in Example<br />

2.2.24.<br />

Proposition 2.2.26. The singular homology <strong>and</strong> cohomology<br />

<strong>of</strong> BC with coefficients in R, namely H∗(BC, R) <strong>and</strong><br />

H ∗ (BC, R), are identified with H∗(C, R) <strong>and</strong> H ∗ (C, R).<br />

Pro<strong>of</strong>. The normalized complex C † ∗(C, R) is a cellular complex<br />

for computing the singular homology <strong>of</strong> H∗(BC, R).<br />

The essential connection between simplicial sets <strong>and</strong> topological<br />

spaces is given by the following result.<br />

Theorem 2.2.27 (Kan). Let X be a simplicial set <strong>and</strong> Y<br />

a topological space. Then we have an adjunction<br />

Hom T op (|X|, Y ) ∼ = Hom SimpSet (X, SY ).<br />

Moreover the adjunct preserves homotopies.<br />

Pro<strong>of</strong>. Let f : |X| → Y be a continuous map. Then we define<br />

a simplicial map Φ f by [(Φ f ) n (x n )](a n ) = f(x n , a n ) for n ≥ 0,<br />

x n ∈ X n <strong>and</strong> a n ∈ △ n . Conversely if Φ : X → SY is a<br />

α<br />

β<br />

y


60 2 Simplicial methods<br />

simplicial map, we define f Φ : |X| → Y by f Φ (x n , a n ) =<br />

[Φ n (x n )](a n ) for n ≥ 0, x n ∈ X n <strong>and</strong> a n ∈ △ n . These two<br />

assignments give the adjunction.<br />

⊓⊔<br />

As an example, every CW-complex is homotopy equivalent<br />

to the classifying space <strong>of</strong> the poset <strong>of</strong> its singular simplices.<br />

Directly from Theorem 2.2.3 we have an alternative characterization<br />

<strong>of</strong> the geometric realization <strong>of</strong> a simplicial set X<br />

as<br />

|X| ∼ = lim −→ι/X<br />

|P|,<br />

where |P| sends each object (n, Φ) to △ n , because a left adjoint<br />

functor preserves direct limits. Given this homeomorphism<br />

we can prove Theorem 2.2.27 alternatively as follows.<br />

Hom T op (|X|, Y ) ∼ = Hom T op (lim −→ι/X<br />

|P|, Y )<br />

∼ = lim ←−ι/X<br />

Hom T op (|P|, Y )<br />

∼ = lim ←−ι/X<br />

Hom SimpSet (P, SY )<br />

∼ = HomSimpSet (lim −→ι/X<br />

P, SY )<br />

∼ = HomSimpSet (X, SY ).<br />

The third isomorphism does not depend on Theorem 2.2.27<br />

<strong>and</strong> it is true because<br />

Hom T op (△ n , Y ) = SY n<br />

∼ = HomSimpSet (Hom △ (−, n), SY ).<br />

Theorem 2.2.28. Let X, Y be two simplicial sets. Then<br />

there exists a natural homeomorphism<br />

|X × Y | ∼ = |X| × |Y |,<br />

if |X| × |Y | is a CW-complex.<br />

The space |X| × |Y | is a CW-complex if both X <strong>and</strong> Y are<br />

countable (that is, both ∪ n≥0 X n <strong>and</strong> ∪ n≥0 Y n are countable)


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 61<br />

or if either |X| or |Y | is locally finite (that is, every point is<br />

an inner point <strong>of</strong> a finite subcomplex).<br />

We collect some useful statements from last section which<br />

are adapted to classifying spaces. The upshot is that a homotopy<br />

between two simplicial maps or sets does give rise to a<br />

homotopy between continuous maps or spaces, in the usual<br />

topological sense. A small category C is contractible if BC has<br />

the same homotopy type <strong>of</strong> a point. By Theorem 2.2.27 it is<br />

equivalent to saying that NC ≃ N•.<br />

Corollary 2.2.29. A functor u : D → C induces a continuous<br />

map Bu : BD → BC.<br />

1. If u ′ : D → C is another functor <strong>and</strong> : ND × N1 →<br />

NC is a homotopy between Nu <strong>and</strong> Nu ′ , then we obtain a<br />

homotopy B : BD × △ 1 → BC between Bu <strong>and</strong> Bu ′ .<br />

2. A natural transformation Φ between u, u ′ : D → C induces<br />

a homotopy B Φ : BD × △ 1 → BC between Bu <strong>and</strong> Bu ′ .<br />

3. If C has an initial or a terminal object, then BC is contractible.<br />

The topological map Bu : BD → BC induces the maps u ∗<br />

<strong>and</strong> u ∗ as in Corollary 2.2.13.<br />

Remark 2.2.30. We must emphasize that continuous maps between<br />

classifying spaces are not always realized by functors. In<br />

other words, not all topological maps BD → BC come from<br />

a simplicial map ND → NC by Lemma 2.2.11. For instance,<br />

there is a natural homeomorphism between BC <strong>and</strong> BC op , but<br />

one cannot construct a functor between C <strong>and</strong> C op realizing<br />

this homeomorphism, except in very special situations (e.g. C<br />

is a group). See Example 2.2.14.<br />

We shall come back to comparing classifying spaces via functors<br />

by introducing Quillen’s work.


62 2 Simplicial methods<br />

2.2.3 Cup product <strong>and</strong> cohomology ring<br />

In this section, we introduce cross product (or external product),<br />

cup product (or internal product) <strong>and</strong> the resulting multiplicative<br />

structure on simplicial cohomology.<br />

Theorem 2.2.31 (Eilenberg-Zilber). Let X, Y be two<br />

simplicial sets. There is a natural chain homotopy equivalence<br />

ξ : C ∗ (X × Y, R) → C ∗ (X, R) ⊗ C ∗ (Y, R).<br />

Pro<strong>of</strong>. The chain map, called the Alex<strong>and</strong>er-Whitney map,<br />

can be explicitly constructed as follows<br />

n∑<br />

ξ(a, b) = d i+1 · · · d n (a) ⊗ d i 0(b),<br />

i=0<br />

where (a, b) ∈ C n (X × Y, R) = R[X n ] ⊗ R[Y n ].<br />

The inverse <strong>of</strong> ξ (up to chain equivalence), called the Eilenber-<br />

Zilber map<br />

is defined by<br />

η(c ⊗ d) =<br />

η : C ∗ (X, R) ⊗ C ∗ (Y, R) → C ∗ (X × Y, R),<br />

∑<br />

(p,q)-shuffles<br />

σ<br />

(−1) |σ| (s σ(1) · · · s σ(p) c, s σ(p+1) · · · s σ(n) d)<br />

for c ∈ C p (X, R) <strong>and</strong> d ∈ C q (Y, R) with p + q = n. Here a<br />

(p, q)-shuffle σ is a permutation <strong>of</strong> n letters such that<br />

σ(1) < · · · < σ(p) <strong>and</strong> σ(p + 1) < · · · < σ(n).<br />

⊓⊔<br />

Remark 2.2.32. The Alex<strong>and</strong>er-Whitney <strong>and</strong> Eilenberg-Zilber<br />

maps lead to chain homotopy equivalences between normalized<br />

chain complexes, in light <strong>of</strong> Theorem 2.2.5.


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 63<br />

Let C be a small category <strong>and</strong> R a commutative ring with<br />

identity. For C ∗ (−, R) = C ∗ (−, R) ∧ , we have a cross product<br />

induced by the Alex<strong>and</strong>er-Whitney map<br />

× : C i (X, R)⊗C j (Y, R) → Hom R (C i (X, R)⊗C j (Y, R), R) ≃ →C i+j (X×<br />

with the left map given by<br />

(f × g)(a, b) = f(d i+1 · · · d n (a))g(d i 0(b)),<br />

if f ∈ C i (X, R), g ∈ C j (Y, R), a ∈ C n (X, R) <strong>and</strong> b ∈<br />

C n (Y, R). One can check that ∂ ∗ (f ×g) = ∂ ∗ (f)×g+(−1) i f ×<br />

∂ ∗ (g). Thus it induces a cross product on cohomology<br />

× : H i (X, R) ⊗ H j (Y, R) → H i+j (X × Y, R)<br />

When R is a field, the map × is an isomorphism by Künneth<br />

formula.<br />

Proposition 2.2.33. Let X, Y be two simplicial sets <strong>and</strong><br />

τ : X ×Y → Y ×X the twist map, defined by τ(a, b) = (b, a)<br />

for any a ∈ X n <strong>and</strong> b ∈ Y n . Then τ induces an isomorphism<br />

τ ∗ : H ∗ (X × Y, R) → H ∗ (Y × X, R).<br />

Suppose f ∈ H i (X, R) <strong>and</strong> g ∈ H j (Y, R). Then τ ∗ (f ×g) =<br />

(−1) ij g × f.<br />

Pro<strong>of</strong>. Since τ is an isomorphism <strong>of</strong> simplicial sets, it certainly<br />

induces an isomorphism on (co-)chain complexes <strong>and</strong><br />

(co-)homology. Now we calculate the value <strong>of</strong> τ ∗ .<br />

Define τ ∗<br />

′ = ξτ ∗ η : C ∗ (Y, R) ⊗ C ∗ (X, R) → C ∗ (X, R) ⊗<br />

C ∗ (Y, R), as shown in the following commutative diagram<br />

C ∗ (Y × X, R)<br />

η ≃<br />

τ ∗ <br />

C ∗ (X × Y, R)<br />

C ∗ (Y, R) ⊗ C ∗ (X, R)<br />

<br />

τ<br />

′ C ∗ (X, R) ⊗ C ∗ (Y, R).<br />

∗<br />

≃<br />

<br />

ξ


64 2 Simplicial methods<br />

Then by direct calculation, it follows τ ∗ ′ is chain homotopy<br />

equivalent to the canonical chain equivalence τ ∗ ′′ : C ∗ (Y, R) ⊗<br />

C ∗ (X, R) → C ∗ (X, R) ⊗ C ∗ (Y, R) determined by b ⊗ a ↦→<br />

(−1) ij a ⊗ b for any a ∈ X i <strong>and</strong> b ∈ Y j . Hence τ ∗ : H ∗ (X ×<br />

Y, R) → H ∗ (Y × X, R) is given by the chain equivalence<br />

induced by τ ∗<br />

′′<br />

Hom R (C ∗ (X, R)⊗C ∗ (Y, R), R) → Hom R (C ∗ (Y, R)⊗C ∗ (X, R), R)<br />

On the other h<strong>and</strong>, there is a canonical chain equivalence<br />

C ∗ (X, R) ⊗ C ∗ (Y, R) → C ∗ (Y, R) ⊗ C ∗ (X, R)<br />

defined in the same fashion as τ ′′<br />

∗ . Thus there is a commutative<br />

diagram <strong>of</strong> cochain complexes<br />

C ∗ (X, R) ⊗ C ∗ (Y, R)<br />

∼ =<br />

C ∗ (Y, R) ⊗ C ∗ (X, R)<br />

Hom R (C ∗ (X, R) ⊗ C ∗ (Y, R), R) ∼=<br />

<br />

Hom R (C ∗ (Y, R) ⊗ C ∗ (X, R), R.<br />

Since the vertical maps give rise to the cross products by<br />

sending f ⊗g to f ×g for any f ∈ C i (X, R) <strong>and</strong> g ∈ C j (Y, R),<br />

while the horizontal maps map f ⊗g <strong>and</strong> f ×g to (−1) ij g ⊗f<br />

<strong>and</strong> (−1) ij g × f, respectively, our formula for τ ∗ follows from<br />

it.<br />

⊓⊔<br />

Definition 2.2.34. There exists a diagonal functor ∆ : C →<br />

C × C such that ∆(x) = (x, x) <strong>and</strong> ∆(α) = (α, α) for any<br />

x ∈ Ob C <strong>and</strong> α ∈ Mor C, respectively.<br />

The diagonal functor induces a natural chain map, the simplicial<br />

diagonal map, also denoted by<br />

∆ : C ∗ (C, R) → C ∗ (C × C, R).<br />

Composing with the Alex<strong>and</strong>er-Whitney map<br />

C ∗ (C × C, R) → C ∗ (C, R) ⊗ C ∗ (C, R),


2.2 Nerves, classifying spaces <strong>and</strong> cohomology 65<br />

we get the diagonal map,<br />

∆ : C ∗ (C, R) → C ∗ (C, R) ⊗ C ∗ (C, R).<br />

More explicitly for any x 0 → x 1 → · · · → x n , a base element<br />

<strong>of</strong> C n (C, R),<br />

n∑<br />

∆(x 0 → · · · → x n ) = (x 0 → · · · → x i )⊗(x i → · · · → x n ).<br />

i=0<br />

By dualizing the diagonal map, we obtain the cup product on<br />

C ∗ (C, R) = C ∗ (C, R) ∧ as<br />

∪ : C i (C, R) ⊗ C j (C, R) ×<br />

−→C i+j (C × C, R) ∆<br />

−→C i+j (C, R).<br />

Suppose f ∈ C i (C, R) <strong>and</strong> g ∈ C j (C, R) with i + j = n. Then<br />

(f ∪ g)(x 0 → · · · → x n ) = ∑ n<br />

k=0 f(x 0 → · · · → x k )g(x k → · · · → x n )<br />

= f(x 0 → · · · → x i )g(x i → · · · → x n ).<br />

From direct calculation we get ∂ ∗ (f ∪ g) = ∂ ∗ (f) ∪ g +<br />

(−1) i f ∪ ∂ ∗ (g) for any f ∈ C i (C, R) <strong>and</strong> g ∈ C j (C, R). Consequently<br />

the cup product passes to cohomology<br />

∪ : H i (C, R) ⊗ H j (C, R) ×<br />

−→H ∗ (C × C, R) ∆<br />

−→H i+j (C, R),<br />

which makes the graded R-module H ∗ (C, R) = ⊕ i≥0 H i (C, R)<br />

a graded ring. This is the ordinary cohomology ring <strong>of</strong> C with<br />

coefficients in R. We may readily verify the following equality.<br />

Proposition 2.2.35. f ∪ g = (−1) ij g ∪ f for any f ∈<br />

H i (C, R) <strong>and</strong> g ∈ H j (C, R).<br />

Pro<strong>of</strong>. It follows from Proposition 2.2.33 because the cup<br />

product is induced by a cross product.<br />

⊓⊔<br />

It means that the ordinary cohomology ring <strong>of</strong> a small category<br />

is graded commutative. Since the cup product is exactly<br />

the same as the one for the topological cohomology ring


66 2 Simplicial methods<br />

H ∗ (BC, R) = ⊕ i≥0 H i (BC, R). We actually have a ring isomorphism<br />

H ∗ (C, R) ∼ = H ∗ (BC, R).<br />

Remark 2.2.36. When C is a (discrete) group, H ∗ (C, R) is the<br />

usual group cohomology ring.<br />

Corollary 2.2.37. Let u : D → C be a functor. Then u ∗ :<br />

H ∗ (C, R) → H ∗ (D, R) is a ring homomorphism.<br />

This ring homomorphism is usually called the restriction.<br />

Proposition 2.2.38. Let D <strong>and</strong> C be small categories. Suppose<br />

k is a field. Then the cup product <strong>of</strong> any two positive<br />

degree elements in H ∗ (D ∗ C, k) is constantly zero.<br />

Pro<strong>of</strong>. In Proposition 2.2.8 we computed homology groups <strong>of</strong><br />

the join <strong>of</strong> two small categories. Accordingly there are formulas<br />

for cochain complexes <strong>and</strong> cohomology as well, i.e. for<br />

each n > 0<br />

˜C n (D ∗ C, k) ∼ = [˜C∗ (D, k) ⊗ ˜C∗ (C, k)] n−1 ,<br />

<strong>and</strong><br />

H n (D ∗ C, k) =<br />

⊕<br />

i+j=n−1<br />

˜H i (D, k) ⊗ ˜Hj (C, k).<br />

It implies that H n (D ∗ C, k) is spanned over a set <strong>of</strong> cohomology<br />

classes represented by functions <strong>of</strong> the form η α such<br />

that η α (β) = 1 if β = α <strong>and</strong> zero otherwise, <strong>and</strong> such that<br />

α ∈ N n (D ∗ C) consists <strong>of</strong> exactly one morphism γ x,y where<br />

x ∈ Ob D <strong>and</strong> y ∈ Ob C, as introduced in Definition 2.2.7.<br />

By direct computation, the cup product <strong>of</strong> any two such functions<br />

is constantly zero. Hence we are done.<br />

⊓⊔<br />

There is a topological interpretation to the above result.<br />

In fact the cohomology ring <strong>of</strong> a join <strong>of</strong> two spaces always<br />

have trivial cup product. See Example 2.2.24 (4). There is a<br />

topological interpretation. Suppose X <strong>and</strong> Y are two spaces.


2.3 Quillen’s work on classifying spaces 67<br />

Then their join X∗Y is a suspension Σ(X×Y/X∨Y ). For any<br />

suspended space, its cohomology ring has trivial cup product.<br />

The structure <strong>of</strong> X ∗ Y also explains Proposition 2.2.8. In<br />

Example 2.2.6 (1), the cup product <strong>of</strong> any two elements <strong>of</strong><br />

positive degree in H ∗ (C, R) is zero.<br />

2.3 Quillen’s work on classifying spaces<br />

In order to assure a functor u : D → C inducing a simplicial<br />

homotopy equivalence ND ≃ NC, one must have a functor on<br />

the opposite direction. However one may very <strong>of</strong>ten obtain a<br />

(topological) homotopy equivalence Bu : BD → BC without<br />

the existence <strong>of</strong> such v : C → D that Bv is also a homotopy<br />

equivalence. Such homotopies certainly are not simplicial <strong>and</strong><br />

will be investigated now.<br />

2.3.1 Quillen’s Theorem A<br />

Quillen’s Theorem A (also called Quillen’s Fibre Theorem)<br />

provides sufficient conditions on a functor between two small<br />

categories which guarantee the functor inducing a homotopy<br />

equivalence between classifying spaces.<br />

Theorem 2.3.1 (Quillen’s Theorem A). Let u : D →<br />

C be a functor between two small categories. If either all<br />

the overcategories or all the undercategories are contractible.<br />

Then Bu is a homotopy equivalence.<br />

Pro<strong>of</strong>. The pro<strong>of</strong> is given in Section 2.4.3.<br />

Under certain circumstances, one may replace the over- or<br />

undercategories with some simpler constructions.<br />

Definition 2.3.2. Let u : D → C be a functor between small<br />

categories. For each x ∈ Ob C, we define the fibre <strong>of</strong> u over x<br />

⊓⊔


68 2 Simplicial methods<br />

to be the category u −1 (x) whose objects are preimages <strong>of</strong> x<br />

<strong>and</strong> whose morphisms are preimages <strong>of</strong> 1 x .<br />

Apart from intuition, the fibres are usually not good for<br />

comparing the homotopy types <strong>of</strong> two classifying spaces. For<br />

instance if G is a group <strong>and</strong> H is a subgroup, then the inclusion<br />

functor has fibre at • equal to the trivial category. But<br />

there is no reason to say that BH would be homotopy equivalent<br />

to BG, especially when we think about the extreme case<br />

where H is the trivial subgroup <strong>of</strong> G.<br />

Definition 2.3.3. Let u : D → C be a functor between small<br />

categories. We call D prefibred over C if the natural functor i x :<br />

u −1 (x) → x\u has a right adjoint h x for each x ∈ Ob C. We<br />

call D prec<strong>of</strong>ibred over C if the natural functor i x : u −1 (x) →<br />

u/x has a left adjoint h x for each x ∈ Ob C.<br />

Corollary 2.3.4. Suppose u : D → C such that D is either<br />

prefibred or prec<strong>of</strong>ibred over C. If the fibre u −1 (x) over every<br />

x ∈ Ob C is contractible, then Bu is a homotopy equivalence.<br />

Pro<strong>of</strong>. Since there are adjoint functors between u −1 (x) <strong>and</strong><br />

x\u (or u/x) for every x ∈ Ob C. The condition implies the<br />

contractibility <strong>of</strong> all the undercategories (or overcategories).<br />

Then we apply Quillen’s Theorem A.<br />

⊓⊔<br />

We will see some prefibred <strong>and</strong> prec<strong>of</strong>ibred examples in the<br />

next section. Note that, for G a group <strong>and</strong> H a subgroup,<br />

i : H ↩→ G is a simple example that H is neither prefibred<br />

nor prec<strong>of</strong>ibred over G, unless H = G.<br />

Suppose u : D → C makes D prec<strong>of</strong>ibred over C. Then we<br />

have<br />

Hom u −1 (x)(h x (a, α), b) ∼ = Hom u/x ((a, α), i x (b))<br />

for any b ∈ Ob u −1 (x) <strong>and</strong> (a, α) ∈ Ob u/x. Let us try<br />

to underst<strong>and</strong> h x . The set Hom u/x ((a, α), b) consists <strong>of</strong> ex-


2.3 Quillen’s work on classifying spaces 69<br />

actly those f ∈ Hom D (a, b) such that u(f) = α : u(a) →<br />

x = u(b). The identity 1 hx (a,α) corresponds to a morphism<br />

in Hom u/x ((a, α), i x h x (a, α)), which is some morphism ˜f ∈<br />

Hom D (a, h x (a, α)) such that u( ˜f) = α. Then each f ∈<br />

Hom D (a, b), satisfying u(f) = α, must uniquely factor through<br />

˜f. If f = g ˜f for g ∈ HomD (h x (a, α), b) such that u(g) = 1 x ,<br />

then g corresponds to f ∈ Hom u/x ((a, α), i x (b)) in the isomorphism.<br />

We have alternative descriptions <strong>of</strong> h x <strong>and</strong> h x .<br />

Lemma 2.3.5. Suppose u : D → C is as above <strong>and</strong> γ : x →<br />

y is a morphism in C.<br />

1. If D is prefibred over C, then it induces a functor<br />

γ ⋆ : u −1 (y) → y\u → x\u → u −1 (x).<br />

Furthermore the right adjoint <strong>of</strong> u −1 (x) → x\u is given by<br />

(α, a) ↦→ α ⋆ (a).<br />

2. If D is prec<strong>of</strong>ibred over C, then it induces a functor<br />

γ ⋆ : u −1 (x) → u/x → u/y → u −1 (y).<br />

Furthermore the left adjoint <strong>of</strong> u −1 (y) → u/y is given by<br />

(b, β) ↦→ β ⋆ (b).<br />

Pro<strong>of</strong>. We shall only prove (2) because the pro<strong>of</strong> for (1) is<br />

similar.<br />

Let (b, β) ∈ Ob u/y. Then<br />

Hence β ⋆ (b) = h y (b, β).<br />

β ⋆ : b ↦→ (b, 1 y ) ↦→ (b, β) ↦→ h y (b, β).<br />

Definition 2.3.6. Let u : D → C be a functor between small<br />

categories. We call D fibred over C if it is prefibred over C <strong>and</strong><br />

moreover (βα) ⋆ = α ⋆ β ⋆ for any composable α, β ∈ Mor C.<br />

⊓⊔


70 2 Simplicial methods<br />

We call D c<strong>of</strong>ibred over C if it is prec<strong>of</strong>ibred over C <strong>and</strong><br />

moreover (βα) ⋆ = β ⋆ α ⋆ for any composable α, β ∈ Mor C.<br />

2.3.2 Constructions over categories <strong>and</strong> relevant functors<br />

We describe several important constructions here, which also<br />

serve an examples to illustrate Quillen’s Theorem A.<br />

Definition 2.3.7. Suppose C is a small category. The (first)<br />

category <strong>of</strong> factorizations in C, denoted by F (C), has Ob F (C) =<br />

Mor C. An object admits a morphism to another object if <strong>and</strong><br />

only if as morphisms in C, the first object factors through the<br />

second.<br />

Iterating the construction, for non-negative integer n, we<br />

can define F n (C) = F (F n−1 (C)) as the n-th category <strong>of</strong> factorizations<br />

in C. Here F 0 (C) is defined to be C.<br />

We shall see that in order to underst<strong>and</strong> C very <strong>of</strong>ten we<br />

have to utilize F (C).<br />

In order not to cause confusions, when a morphism α ∈<br />

Mor C is considered as an object in F (C), we write it as [α].<br />

In the above definition, a morphism in F (C) is a pair (α ′ , α ′′ ) :<br />

[α] → [β] <strong>and</strong> is customarily pictured by<br />

y<br />

α ′ <br />

α<br />

x<br />

y ′ x ′ α ′′<br />

β<br />

Example 2.3.8.1. When C is a group, F (C) has all group elements<br />

as objects. The category F (C) is a groupoid whose<br />

skeleton is isomorphic to the group itself.<br />

2. When C is a poset, then F (C) is another poset. For 1, we<br />

have F (1) as follows


2.3 Quillen’s work on classifying spaces 71<br />

(α,1 0 )<br />

(1 1 ,1 0 )<br />

<br />

[1 ← 0]<br />

(1 1 ,α)<br />

[0 ← 0]<br />

[1 ← 1]<br />

<br />

(1 0 ,1 0 )<br />

(1 1 ,1 1 )<br />

For convenience we denote the non-identity morphism 1 ← 0<br />

by α.<br />

One can see that there are two natural covariant functors<br />

t : F (C) → C <strong>and</strong> s : F (C) → C op , where t st<strong>and</strong>s for target<br />

while s means source.<br />

Proposition 2.3.9. Suppose D ≃ C. Then F (D) ≃ F (C).<br />

Pro<strong>of</strong>. Let u : D → C <strong>and</strong> v : C → D be such that vu ∼ = Id D<br />

<strong>and</strong> uv ∼ = Id C . We construct two functors between F (C) <strong>and</strong><br />

F (D) which induce an equivalence.<br />

We define F (u) : F (D) → F (C) by F (u)([α]) = [u(α)],<br />

<strong>and</strong> F (u)(α ′ , α ′′ ) = (u(α ′ ), u(α ′′ )). One can verify this gives a<br />

functor. Similarly there exists a functor F (v) : F (C) → F (D).<br />

Since F (v)F (u)([α]) = [vu(α)] ∼ = [α], we have F (v)F (u) ∼ =<br />

Id F (D) . Similarly we get F (u)F (v) ∼ = Id F (C) . Thus F (D) ≃<br />

F (C).<br />

⊓⊔<br />

Proposition 2.3.10. Consider the canonical functors C←F t (C) →C s o<br />

Then<br />

1. F (C) is c<strong>of</strong>ibred over both C <strong>and</strong> C op . It implies, for any<br />

x ∈ Ob C, the natural functor i x : t −1 (x) → t/x has a<br />

left adjoint, <strong>and</strong>, for any y ∈ Ob C op , the natural functor<br />

i y : t −1 (y) → s/y has a left adjoint;<br />

2. for any x ∈ Ob C, the category t −1 (x) ∼ = (Id C /x) op has an<br />

initial object;


72 2 Simplicial methods<br />

3. for any y ∈ Ob C op , the category s −1 (y) ∼ = y\Id C has an<br />

initial object.<br />

Consequently both t <strong>and</strong> s induce homotopy equivalences.<br />

Pro<strong>of</strong>. The category F (C) is prec<strong>of</strong>ibred over C because we<br />

can define h x ([α], α ′ ) = [α ′ α] for any x ∈ Ob C <strong>and</strong> ([α], α ′ ) ∈<br />

Ob t/x such that<br />

Hom t −1 (x)(h x ([α], α ′ ), [β]) ∼ = Hom t/x (([α], α ′ ), i x (β)).<br />

We can compute t −1 (x) <strong>of</strong> t : F (C) → C for every x ∈ Ob C<br />

<strong>and</strong> find t −1 (x) ∼ = (Id C /x) op . It has an initial object [1 x ]<br />

<strong>and</strong> thus is contractible. This means that t/x is always contractible<br />

<strong>and</strong> thus t induces an equivalence by Quillen’s Theorem<br />

A. Moreover from the description <strong>of</strong> t −1 (x) we can see<br />

that F (C) is c<strong>of</strong>ibred over C.<br />

We can similarly prove that s : F (C) → C op induces an<br />

equivalence as well, with s −1 (y) ∼ = y\Id C .<br />

⊓⊔<br />

A hidden fact in the pro<strong>of</strong> is that whenever there is a category<br />

C, we can produce many over- <strong>and</strong> undercategories Id C /x<br />

<strong>and</strong> x\Id C , all <strong>of</strong> which are contractible. This will be used for<br />

an important construction later on.<br />

Remark 2.3.11. Here the homotopy equivalence Bt : BF (C) →<br />

BC usually is not a simplicial homotopy equivalence as we do<br />

not expect to have a functor from C to F (C) giving a homotopy<br />

equivalence BC → BF (C). If such a functor were to<br />

exist, then we would obtain a functor inducing the homotopy<br />

equivalence BC → BC op , <strong>and</strong> hence inducing isomorphism on<br />

homology H∗(C, R) → H∗(C op , R), which rarely happens. See<br />

Example 2.2.14.<br />

Definition 2.3.12. The functor ∇ = (t, s) : F (C) → C e =<br />

C × C op is called the skew diagonal functor. The category C e<br />

is called the enveloping category <strong>of</strong> C.


2.3 Quillen’s work on classifying spaces 73<br />

The above construction is a special case <strong>of</strong> a more general<br />

definition F (u) for any u : D → C, naturally equipped with a<br />

functor F (u) → D × C op . Indeed F (Id C ) = F (C), see Section<br />

2.4.3.<br />

Example 2.3.13. When C is a poset, both F (C) <strong>and</strong> C e are<br />

posets <strong>and</strong> ∇ : F (C) → C e is an embedding. For the poset<br />

C = 1, we have F (1) ↩→ 1 e as follows<br />

[1 ← 0] (1, 0)<br />

[0 ← 0]<br />

[1 ← 1]<br />

↩→ (0, 0)<br />

(1, 1)<br />

(0, 1)<br />

Proposition 2.3.14. The category F (C) is c<strong>of</strong>ibred over C e .<br />

It implies that, for any (y, x) ∈ Ob C e , the functor i (y,x) :<br />

∇ −1 (y, x) → ∇/(y, x) has a left adjoint. Thus N∇ −1 (y, x) ≃<br />

N∇/(y, x).<br />

Furthermore ∇ −1 (y, x) ∼ = Hom C (x, y).<br />

Pro<strong>of</strong>. We first prove that F (C) is prec<strong>of</strong>ibred over C e . To<br />

this end, we need to show that for every (y, x) ∈ Ob C e , the<br />

functor i (y,x) : ∇ −1 (y, x) → ∇/(y, x) has a left adjoint. It is<br />

easy to identify ∇ −1 (y, x) with the discrete set Hom C (x, y).<br />

Then we define h (y,x) ([α], (α ′ , α ′′ )) = [α ′ αα ′′ ] ∈ Hom C (x, y)<br />

which leads to an isomorphism<br />

Hom ∇ −1 (y,x)([α ′ αα ′′ ], [β]) ∼ = Hom ∇/(y,x) (([α], (α ′ , α ′′ )), ([β], (1 y , 1 x ))).<br />

Note that in any case either morphism set has at most one<br />

element.<br />

If (γ ′ , γ ′′ ) : (y, x) → (y ′ , x ′ ) is a morphism, then (γ ′ , γ ′′ ) ⋆ is<br />

given by<br />

[α] ↦→ ([α], (1 y , 1 x )) ↦→ ([α], (γ ′ , γ ′′ )) ↦→ [γ ′ αγ ′′ ],


74 2 Simplicial methods<br />

for any α ∈ Hom C (x, y). Hence one can verify that F (C) is<br />

c<strong>of</strong>ibred over C e .<br />

⊓⊔<br />

We note that some <strong>of</strong> the overcategories or undercategories<br />

or fibres can be empty. As a reminder, the homology <strong>of</strong> ∅ is<br />

constantly zero.<br />

Remark 2.3.15. In the previous two propositions F (C) are<br />

c<strong>of</strong>ibred over C, C op as well as C e . It is not fibred over these<br />

categories. For instance in Example 2.3.13, i x : ∇ −1 (1, 0) =<br />

[1 ← 0] → ∇/(1, 0) ∼ = F (1) has only a left adjoint <strong>and</strong> no<br />

right adjoint because [1 ← 0] is a terminal object in F (1). In<br />

these notes, it is very important to compute limits over various<br />

over <strong>and</strong> undercategories. We <strong>of</strong>ten do it by establishing<br />

adjoint functors between these categories <strong>and</strong> some simpler<br />

categories. Then by Proposition 1.2.4 we will know whether<br />

we can simplify the computations <strong>of</strong> either direct or inverse<br />

limits (usually not both!). We shall see many explicitly calculations<br />

from Chapter 3.<br />

2.4 Further categorical <strong>and</strong> simplicial constructions<br />

In this section we will introduce the definition <strong>of</strong> a Grothendieck<br />

construction which is important when we deal with finite<br />

categories constructed from groups. Grothendieck constructions<br />

are a certain kind <strong>of</strong> homotopy colimits. By Thomason’s<br />

Homotopy Colimit Theorem, every homotopy colimit<br />

is a Grothendieck construction. In order to provide the interested<br />

reader a necessary underst<strong>and</strong>ing <strong>of</strong> these constructions,<br />

we will define bisimplicial sets, <strong>and</strong> since bisimplicial sets are<br />

introduced, there is no reason not to give pro<strong>of</strong>s <strong>of</strong> Quillen’s<br />

<strong>and</strong> Thomason’s theorems. Although this chapter is the most<br />

suitable place to present these materials, it is possible to postpone<br />

reading them until they are needed. Indeed, only un-


2.4 Further categorical <strong>and</strong> simplicial constructions 75<br />

derst<strong>and</strong>ing the Grothendieck constructions <strong>and</strong> knowing the<br />

statements are sufficient for our purposes in these notes.<br />

2.4.1 Grothendieck constructions<br />

Let C be a small category <strong>and</strong> F : C → Cat a functor. Then<br />

we can construct a small category, called the Grothendieck<br />

construction Gr C F, as follows.<br />

The objects are pairs <strong>of</strong> the form (x, a) with x ∈ Ob C <strong>and</strong><br />

a ∈ Ob F(x). Let (x, a) <strong>and</strong> (y, b) be two objects. A morphism<br />

(x, a) → (y, b) is a pair (α, f) such that α ∈ Hom C (x, y) <strong>and</strong><br />

f ∈ Hom F(b) (F(α)(a), b). It can be pictured as<br />

x<br />

α<br />

α<br />

y A ′ <br />

B ′<br />

C ′<br />

D ′<br />

y A ′ <br />

B ′<br />

x<br />

Note that there always exists a functor π : Gr C F → C.<br />

Example 2.4.1. Let u : D → C be a functor between small<br />

categories. Then the over- <strong>and</strong> undercategories can be realized<br />

as Grothendieck constructions. Given an object x ∈ Ob C, we<br />

can define two functors u = Hom C (u(−), x) : D → Set <strong>and</strong><br />

u = Hom C (x, u(−)) : D → Set. One can verify that Gr D u<br />

<strong>and</strong> Gr D u are isomorphic to u/x <strong>and</strong> x\u, respectively.<br />

We record one more important Grothendieck construction,<br />

although it is not needed in this chapter.<br />

Definition 2.4.2. A small category C is an EI category if all<br />

endomorphisms are isomorphisms.<br />

C ′<br />

D ′


76 2 Simplicial methods<br />

Typical examples <strong>of</strong> EI categories are groups <strong>and</strong> posets.<br />

For each EI category C, one can introduce a partial order on<br />

the set <strong>of</strong> isomorphism classes <strong>of</strong> objects by [x] ≤ [y] if <strong>and</strong><br />

only if Hom C (x, y) ≠ ∅. Let us denote by [C] the poset <strong>of</strong><br />

isomorphism classes <strong>of</strong> objects in C. Obviously there exists a<br />

natural functor p : C → [C]. Since [C] is a poset, we know the<br />

barycentric subdivision <strong>of</strong> [C], written as S([C]), is another<br />

poset whose objects are chains <strong>of</strong> non-isomorphism objects<br />

[x] n = [x 0 ] → [x 1 ] → · · · → [x n ] <strong>and</strong> there is a morphism<br />

[x] n → [y] m if <strong>and</strong> only if [y] m is a subchain <strong>of</strong> [x] n . We define<br />

a functor<br />

˜p : S([C]) → Cat<br />

by asking ˜p([x] n ) to be a category whose objects are the y n =<br />

y 0 → y 1 → · · · → y n in NC n such that p(y i ) = [x i ] for<br />

all 0 ≤ i ≤ n, <strong>and</strong> whose morphisms are (n + 1)-tuples <strong>of</strong><br />

morphisms in C making the following diagram commutative<br />

α<br />

y 1 0 y 1<br />

α 2 <br />

· · · αn <br />

y n<br />

γ 0 <br />

γ 1 <br />

γ n <br />

y ′ 0 β 1<br />

<br />

y ′ 1 β 2<br />

<br />

· · · βn<br />

<br />

y ′ n<br />

Since C is EI, all γ i are isomorphism. This means that ˜p([x] n )<br />

is always a groupoid. One can readily verify that for any<br />

x n → y m in Mor(sd[C]), there is a well defined functor<br />

˜p(x n ) → ˜p(y m ) given by first removing objects in x n which<br />

are not isomorphic to any object in y m , <strong>and</strong> then composing<br />

appropriate morphisms in what is left <strong>of</strong> x n .<br />

Definition 2.4.3 (Slominska). Let C be an EI category.<br />

Then its subdivision, written as S(C), is defined to be Gr S([C])˜p.<br />

Alternatively S(C) is a category whose objects are chains <strong>of</strong><br />

non-isomorphisms x n = x 0 → x 1 → · · · → x n in C, <strong>and</strong> a<br />

morphism from x n to y m = y 0 → y 1 → · · · → y m is an (m +


2.4 Further categorical <strong>and</strong> simplicial constructions 77<br />

1)-tuple <strong>of</strong> isomorphisms (γ 0 , · · · , γ m ), making the following<br />

diagram commutative<br />

<br />

<br />

<br />

<br />

<br />

α 1<br />

x n1<br />

α 2<br />

· · · αn x nm<br />

y 0<br />

γ 1<br />

β1<br />

y 1<br />

γ n<br />

β2<br />

· · · βn<br />

y m<br />

x n0<br />

γ 0<br />

<br />

in which every x ni ∈ {x i } n i=0 <strong>and</strong> x ni<br />

∼ = yi in C. From this<br />

description there exists a functor s : S(C) → C by x n ↦→ x 0 .<br />

Thus there are maps s ∗ : H∗(S(C), R) → H∗(C, R) <strong>and</strong> s ∗ :<br />

H ∗ (C, R) → H ∗ (S(C), R).<br />

Proposition 2.4.4 (Slominska). The map s induces a homotopy<br />

equivalence BS(C) ≃ BC. Hence we have isomorphisms<br />

s ∗ : H∗(S(C), R) → H∗(C, R) <strong>and</strong> s ∗ : H ∗ (C, R) →<br />

H ∗ (S(C), R).<br />

Pro<strong>of</strong>. We shall prove all undercategories are contractible. In<br />

order to do so, we show S(C) is prefibred over C. Then one<br />

can easily verify that for each x ∈ Ob C, s −1 (x) has x as a<br />

terminal object <strong>and</strong> hence is contractible.<br />

To show S(C) is prefibred over C we only have to find a<br />

right adjoint to the inclusion i x : s −1 (x) → x\s. Suppose<br />

(α, x 0 → x 1 → · · · → x n ) is an object in x\s. Then we define<br />

a functor x\s → s −1 (x) by<br />

(x 0<br />

α 1<br />

→x1 → · · · → x n , α) ↦→ (x α 1α<br />

→x 1 → · · · → x n ).<br />

Moreover if (γ 0 , γ 1 , · · · , γ m ) is a morphism from (α, x 0 →<br />

x 1 → · · · → x n ) to (α ′ , x ′ 0 → x ′ 1 → · · · → x ′ m), then its image<br />

is defined as (1 x , γ 1 , · · · , γ n ). Now one can check this functor<br />

is right adjoint to i x . Hence we are done.<br />

⊓⊔<br />

These isomorphisms will be generalized in Chapter 4.


78 2 Simplicial methods<br />

There are two or three occasions in this book that we need<br />

to underst<strong>and</strong> the classifying spaces <strong>of</strong> some categories. Thus<br />

we provide two examples on subdivisions <strong>of</strong> categories.<br />

Example 2.4.5.1. Let C be the following EI category<br />

G<br />

x<br />

α <br />

y H .<br />

Then its underlying poset [C] is x → y, <strong>and</strong> S([C]) is the<br />

poset [x] ← [x → y] → [y]. One can easily find that ˜p[x]<br />

<strong>and</strong> ˜p[y] are<br />

G x , y H<br />

respectively. The groupoid ˜p[x → y] is<br />

G×H {x α <br />

y} ,<br />

in which we denote by α the only object in ˜p[x → y]. The two<br />

functors ˜p[x → y] → ˜p[x] ˜p[x → y] → ˜p[y] are the obvious<br />

projections. Thus the category S(C) has objects <strong>of</strong> the forms<br />

([x], x), ([y], y) <strong>and</strong> ([x → y], x α →y). All the morphisms in<br />

S(C) are <strong>of</strong> the following forms<br />

•([x → y] → x, g), ∀g ∈ Mor(˜p[x]) = Aut C (x) = G;<br />

•([x → y] → y, h), ∀h ∈ Mor(˜p[y]) = Aut C (y) = H;<br />

•(1 [x→y] , (g ′ , h ′ )), ∀(g ′ , h ′ ) ∈ Mor(˜p[x → y]) = Aut C (x) ×<br />

Aut C (y) = G × H.<br />

As one can see the subdivision S(C) is more complicated<br />

than C itself so it seems not helpful in terms <strong>of</strong> underst<strong>and</strong>ing<br />

BC through BS(C). However, in the next section, we shall<br />

see it does help if we combine with Thomason’s Homotopy<br />

Colimit Theorem.<br />

For future reference, we record two more finite EI categories,<br />

which share the same underlying poset with C. The reader<br />

may try to figure out their subdivisions. We shall give only<br />

partial descriptions <strong>of</strong> these constructions.


2.4 Further categorical <strong>and</strong> simplicial constructions 79<br />

2. Let D be the following category<br />

h<br />

x<br />

1 x<br />

α<br />

β<br />

y {1 y }<br />

,<br />

where h 2 = 1 x <strong>and</strong> βh = α. Then we also have poset [D] =<br />

x → y, <strong>and</strong> S([D]) = [x] ← [x → y] → [y]. We can find ˜p[x]<br />

<strong>and</strong> ˜p[y] are<br />

h<br />

1 x<br />

x , y {1 y }<br />

respectively. The third groupoid ˜p[x → y] has two isomorphic<br />

objects x α →y <strong>and</strong> x β →y. Its skeleton is the trivial category.<br />

3. Let E be the following EI category<br />

h<br />

g<br />

x<br />

1 x<br />

α<br />

β<br />

gh<br />

y {1 y }<br />

,<br />

where g 2 = h 2 = 1 x , gh = hg, αh = α, αg = β, βh =<br />

β, βg = β. Then we again have poset [E] = x → y, <strong>and</strong><br />

S([E]) = [x] ← [x → y] → [y]. By direct calculations, ˜p[x]<br />

<strong>and</strong> ˜p[y] are<br />

h<br />

g<br />

x<br />

1 x<br />

<br />

, y {1 y } <br />

gh<br />

respectively. The groupoid ˜p[x → y] has two isomorphic<br />

objects x→y α <strong>and</strong> x→y. β However in this case, its skeleton<br />

is isomorphic to C 2 = {1 x , h}, which is the subgroup <strong>of</strong><br />

Aut C (x) × Aut C (y) that stabilizes the set {α, β}.<br />

We will come back to these examples at the end <strong>of</strong> the next<br />

section.


80 2 Simplicial methods<br />

2.4.2 Bisimplicial sets <strong>and</strong> homotopy colimits<br />

Definition 2.4.6. A bisimplicial set is a contravariant functor<br />

△ × △ to Set.<br />

Because<br />

Hom Cat (△ op ×△ op , SimpSet) ∼ = Hom Cat (△ op , Hom Cat (△ op , SimpSet<br />

clearly we may regard a bisimplicial set as a simplicial object<br />

in the category <strong>of</strong> simplicial sets. Thus a bisimplicial set is<br />

indeed a simplicial simplicial set.<br />

Alternatively, a bisimplicial set is a bigraded sequence <strong>of</strong> sets<br />

X p,q for p, q ≥ 0, together with horizontal face <strong>and</strong> degeneracy<br />

maps d h : X p,q → X p−1,q <strong>and</strong> s h : X p,q → X p+1,q as well as<br />

vertical face <strong>and</strong> degeneracy maps d v : X p,q → X p,q−1 <strong>and</strong> s v :<br />

X p,q → X p,q+1 . These maps must satisfy simplicial identities<br />

in either direction, <strong>and</strong> in addition every horizontal map must<br />

commutes with every vertical map. The total complex <strong>of</strong> a<br />

bisimplicial set, Tot(X), is the simplicial set<br />

Tot(X) n := ⊕ n≥0<br />

X n,∗ × △ n / ∼,<br />

in which the equivalence is given by (d h i x, y) ∼ (x, di y) <strong>and</strong><br />

(s h i x, y) ∼ (x, si y).<br />

We can also consider bisimplicial objects in an abelian category<br />

A-Mod for some associative ring. Then by Dold-Kan<br />

Correspondence, the category <strong>of</strong> bisimplicial A-modules is<br />

equivalent to the category <strong>of</strong> first quadrant double complexes<br />

<strong>of</strong> A-modules.<br />

Fix a commutative ring R. Each bisimplicial set gives rise<br />

to a first quadrant double complex R[X] p,q with differentials<br />

∂ h p,q = ∑ i (−1)i d h i <strong>and</strong> ∂ v p,q = (−1) p ∑ i (−1)i d v i . Consequently


2.4 Further categorical <strong>and</strong> simplicial constructions 81<br />

the category <strong>of</strong> bisimplicial R-modules is equivalent to the category<br />

<strong>of</strong> first quadrant double chain complexes <strong>of</strong> R-modules.<br />

Definition 2.4.7. The diagonal, diagX, <strong>of</strong> a bisimplical set<br />

X is the simplicial set obtained via restriction along the diagonal<br />

functor △ → △ × △.<br />

Theorem 2.4.8 (Bousfield-Kan ). Let X, Y be two bisimplicial<br />

sets.<br />

1. There is a natural simplicial isomorphism Tot(X) ∼ = diagX.<br />

2. Given a bisimplicial map Φ p,q : X p,q → Y p,q , if for all q,<br />

X ∗,q → Y ∗,q is a weak homotopy equivalence, then so is<br />

diagΦ : diagX → diagY .<br />

Pro<strong>of</strong>. For the first part, we define a map Ψ, induced by the<br />

maps X p,∗ × △ p → diagX p given by<br />

(x, d p y) ↦→ d h px ∈ X p,p ,<br />

for any (x, d p y) ∈ X p,q × △ p n. One may readily check that Ψ<br />

is a simplicial isomorphism.<br />

As for Part (2), we will not provide a pro<strong>of</strong> here but it is<br />

not surprising from Part (1). See [9], [Tornehave], [Reedy] <strong>and</strong><br />

[8, 16].<br />

Definition 2.4.9. Suppose G : C → SimpSet is a functor.<br />

⊕Then the simplicial replacement <strong>of</strong> G is the bisimplicial set<br />

∗<br />

G which in dimension n consists <strong>of</strong> the coproduct<br />

( ⊕ G) ⊕<br />

n :=<br />

G(x 0 ).<br />

∗<br />

x 0 →···→x n ∈NC n<br />

The horizontal face map d i maps G(s(σ)) to G(s(d i σ)) by the<br />

identity map if i > 0 <strong>and</strong> by α 0 if i = 0. The horizontal degeneracy<br />

map s i maps the simplicial set G(s(σ)) to G(s(s i σ))<br />

by the identity map.


82 2 Simplicial methods<br />

Definition 2.4.10. Let G : C → SimpSet be a functor.<br />

The homotopy colimit <strong>of</strong> G, hocolim C G, is diag( ⊕ ∗ G).<br />

For example one can easily see that hocolim C Npt = NC.<br />

Theorem 2.4.11 (Thomason’s Homotopy Colimit Theorem).<br />

Suppose C is a small category <strong>and</strong> F : C → Cat is<br />

a functor. Let Gr C F be the Grothendieck construction on F.<br />

Then there is a natural weak equivalence<br />

NGr C F ≃ hocolim C NF.<br />

The pro<strong>of</strong> is in the next section. It is clear from the pro<strong>of</strong> <strong>of</strong><br />

Thomason’s theorem, <strong>and</strong> Definitions 2.2.2 <strong>and</strong> 2.4.9 that<br />

BGr C F ≃ hocolim C BF.<br />

Remark 2.4.12. The homotopy colimit is a replacement in<br />

T op <strong>of</strong> the colimit we introduced earlier. In the category T op,<br />

the colimit <strong>of</strong> a diagram <strong>of</strong> spaces does exit. However it does<br />

not behave very well <strong>and</strong> thus is not good to work with in algebraic<br />

topology. One can find a well-known example in [DH].<br />

There is also a dual notion <strong>of</strong> homotopy limit but we shall<br />

not touch it in this book.<br />

The following examples provide the only three homotopy<br />

colimits that the reader needs to know for our purposes. In<br />

fact, they are only used in Section 4.2.2, for computing certain<br />

ordinary (co)homology.<br />

Example 2.4.13. When C is the poset P : a ← b → c, for any<br />

F : P → Cat the homotopy colimit hocolim P BF is called a<br />

homotopy pushout because it fits into the right lower corner<br />

<strong>of</strong> the following diagram


2.4 Further categorical <strong>and</strong> simplicial constructions 83<br />

BF(b→a)<br />

<br />

BF(b) BF(b→c) <br />

BF(a)<br />

BF(c)<br />

<br />

hocolim P BF<br />

such that it is commutative up to homotopy <strong>of</strong> continuous<br />

maps, <strong>and</strong> hocolim C BF enjoys a certain universal property,<br />

analogous to that <strong>of</strong> a colimit.<br />

Example 2.4.14. In Example 2.4.5, the classifying spaces <strong>of</strong><br />

the subdivisions <strong>of</strong> all three categories can be realized as homotopy<br />

pushouts, according to Example 2.4.13. For instance,<br />

in Example 2.4.5 (3) we have BE ≃ B(C 2 ×C 2 )/BC 2 because<br />

we have the following homotopy pushout diagram<br />

inclusion<br />

B˜p([x → y]) = BC<br />

<br />

2 B(C 2 × C 2 ) = B˜p([x])<br />

Bpt<br />

<br />

B˜p([y]) = B•<br />

<br />

B(C 2 × C 2 )/BC 2 ≃ hocolim P B˜p ≃ BS(E<br />

Similarly, in Example 2.4.5 (1), we get BC ≃ BG ∗ BH (a<br />

join <strong>of</strong> spaces), <strong>and</strong> in Example 2.4.5 (2) BD ≃ BC 2 . At<br />

least in these small examples, subdivisions are useful to obtain<br />

the homotopy type <strong>of</strong> classifying spaces <strong>of</strong> categories. This is<br />

one <strong>of</strong> the the reasons why we introduced the Grothendieck<br />

construction earlier.<br />

2.4.3 Pro<strong>of</strong>s <strong>of</strong> Quillen’s Theorem A <strong>and</strong> Thomason’s theorem<br />

The pro<strong>of</strong>s are taken directly from the original papers <strong>of</strong><br />

Quillen [61] <strong>and</strong> Thomason [77].<br />

In order to prove Quillen’s Theorem A, we shall first generalize<br />

the construction <strong>of</strong> category <strong>of</strong> factorizations.<br />

Definition 2.4.15. Suppose u : D → C is a functor between<br />

two small categories. We define F (u) to be a category


84 2 Simplicial methods<br />

whose objects are (a, α, x) with a ∈ Ob D, x ∈ Ob C <strong>and</strong><br />

α ∈ Hom C (x, u(a)). A morphism from (a, α, x) to (b, β, y)<br />

is a pair <strong>of</strong> morphisms (f, γ) with f ∈ Hom D (a, b) <strong>and</strong><br />

γ ∈ HomC(y, x) such that β = u(f)αγ as in the diagram<br />

u(a)<br />

u(f)<br />

<br />

u(b)<br />

Note that there exists two natural functors P t : F (u) →<br />

D op <strong>and</strong> P s : F (u) → C, as well as P t × P s : F (u) →<br />

D × C op . For any category C, F (Id C ) is exactly the category<br />

<strong>of</strong> factorizations in C, F (C), such that P s = s <strong>and</strong> P t = t.<br />

Let us notice that any element (a 0 , α 0 , x 0 ) → (a 1 , α 1 , x 1 ) →<br />

· · · → (a n , α n , x n ) ∈ NF (u) n can be pictured as a commutative<br />

diagram<br />

x 0<br />

α 0 <br />

x 1<br />

α 1 <br />

α <br />

β<br />

x<br />

y<br />

γ<br />

· · ·<br />

x n<br />

α n <br />

u(a 0 )<br />

<br />

u(a 1 )<br />

<br />

· · ·<br />

<br />

u(a n )<br />

Pro<strong>of</strong> (<strong>of</strong> Quillen’s Theorem A). We will only prove the case<br />

where all undercategories associated wtih u are contractible.<br />

The other case can be shown similarly.<br />

Let F (u) be as above. We define a functor N[Id C /u(−)] ∗ :<br />

⊕D → SimpSet <strong>and</strong> consider its simplicial replacement X =<br />

∗ N[Id C/u(−)] ∗ the bisimplicial set with X p,q consists <strong>of</strong><br />

pairs <strong>of</strong> sequences <strong>of</strong> morphisms in C <strong>and</strong> D, respectively,<br />

(x p → · · · → x 0 → u(a 0 ), a 0 → a 1 → · · · → a q ).<br />

The horizontal <strong>and</strong> vertical face <strong>and</strong> degeneracy maps are<br />

given in the obvious way. We have a simplicial map X ∗q →<br />

ND q as well as a simplicial map X p∗ → NCp op . Since the


2.4 Further categorical <strong>and</strong> simplicial constructions 85<br />

diagonal <strong>of</strong> X is the nerve <strong>of</strong> F (u), these two functors provide<br />

two more simplicial maps<br />

⊕<br />

N(Id C /u(a 0 )) op<br />

q → ⊕<br />

N• q = ND q<br />

a 0 →···→a q a 0 →···→a q<br />

<strong>and</strong> ⊕<br />

N(x 0 \u) p →<br />

⊕<br />

N• p = NCp op .<br />

x p →···→x 0 x p →···→x 0<br />

The geometric realizations <strong>of</strong> these two maps are homotopy<br />

equivalences because both Id D /u(a 0 ) <strong>and</strong> x 0 \u are contractible<br />

(the latter by assumption). Hence by Theorem 2.4.7<br />

(2) we obtain two homotopy equivalences<br />

BD<br />

BP t<br />

BP<br />

BF (u)<br />

s<br />

BC op .<br />

By examining the following diagram<br />

C op<br />

P s<br />

C op<br />

P s =s<br />

F (u)<br />

u ′<br />

F (C)<br />

P t<br />

D<br />

u<br />

P t =t C<br />

where u ′ is defined by u ′ (a, α, x) = (u(a), α, x). It forces Bu ′<br />

<strong>and</strong> thus Bu to be homotopy equivalences.<br />

⊓⊔<br />

The pro<strong>of</strong> <strong>of</strong> Thomason’s homotopy colimit theorem is in<br />

spirit similar to that <strong>of</strong> Quillen’s Theorem A. We remind the<br />

reader that NF (u) ≃ diagX = hocolim D N[Id C /u(−)] in the<br />

above pro<strong>of</strong>, <strong>and</strong> one verifies that Gr D Id C /u(−) ∼ = F (u).<br />

Pro<strong>of</strong> (<strong>of</strong> Thomason’s theorem). We shall establish a natural<br />

map<br />

η : hocolim C NF → NGr C F<br />

<strong>and</strong> construct a new functor ˜F : C → Cat so that we have<br />

natural equivalences


86 2 Simplicial methods<br />

λ 1<br />

λ<br />

hocolim C NF<br />

<br />

hocolim C N ˜F 2 <br />

NGr C F<br />

satisfying ηλ 1 ≃ λ 2 .<br />

Step 1: Similar to the pro<strong>of</strong> <strong>of</strong> Quillen’s Theorem A, hocolim C NF<br />

is the diagonal <strong>of</strong> ⊕ ∗<br />

NF such that for any integer n ≥ 0,<br />

( ⊕ ∗ NF) n consists <strong>of</strong> n-simplicies<br />

α<br />

(x<br />

1 α<br />

0→ · · ·<br />

n f 1 f n<br />

→xn , a 0→ · · · →an )<br />

where x 0 → · · · → x n ∈ NC n <strong>and</strong> a 0 → · · · → a n ∈ NF(x 0 ) n .<br />

We define η by<br />

α<br />

η n (x<br />

1 α<br />

0→ · · ·<br />

n f 1 f n<br />

→xn , a 0→ · · · →an )<br />

= (x 0 , a 0 ) (α 1,F(α 1 )(f 1 ))<br />

−→ (x 1 , F(α 1 )(a 1 )) (α 2,F(α 2 α 1 )(f 2 ))<br />

−→ · · ·<br />

· · · (α n,F(α n···α 1 )(f n ))<br />

−→ (x n , F(α n · · · α 1 )(a n ))<br />

Step 2: Let us construct λ 1 . We define ˜F : C → Cat to be<br />

the functor<br />

˜F(x) = π/x<br />

for π : Gr C F → C. Recall that the objects in ˜F(x) are <strong>of</strong> the<br />

form ((y, a), α) such that y ∈ Ob C, a ∈ Ob F(y) <strong>and</strong> α : y →<br />

x, <strong>and</strong> moreover any α : y → x induces ˜F(α) : ˜F(y) → ˜F(x).<br />

There exists a canonical functor<br />

˜F(x) → F(x)<br />

for any x ∈ Ob C, given by ((y, a), α) ↦→ F(α)(a). This functor<br />

has a right adjoint F(x) → ˜F(x) by a ↦→ ((x, a), 1x ). Consequently<br />

N ˜F(x) → NF(x) is a simplicial homotopy equivalence.<br />

Since all functors constructed previously, ˜F(x) → F(x),<br />

assemble to a natural transformation


2.4 Further categorical <strong>and</strong> simplicial constructions 87<br />

˜F → F<br />

<strong>of</strong> functors C → Cat, we obtain a simplicial homotopy equivalence<br />

N ˜F → NF which gives rise to a homotopy equivalence,<br />

by Theorem 2.4.7 (2),<br />

λ 1 : hocolim C N ˜F<br />

≃<br />

→hocolimC NF.<br />

Step 3: Now we show there is a simplicial homotopy equivalence<br />

λ 2 : hocolim C N ˜F → GrC F.<br />

Similar to Step 1, hocolim C N ˜F is the diagonal <strong>of</strong><br />

⊕∗ N ˜F such<br />

that for any integer n ≥ 0, ( ⊕ ∗ N ˜F)n consists <strong>of</strong> n-simplicies<br />

α<br />

(x<br />

1 α<br />

0→ · · ·<br />

n (h 1 ,l 1 )<br />

→xn , c 0 → · · · (h n,l n )<br />

→ c n )<br />

where x 0 → · · · → x n ∈ NC n <strong>and</strong> c 0 → · · · → c n ∈<br />

N ˜F(x0 ) n = [N(π/x 0 )] n . Note that each c i is <strong>of</strong> the form<br />

((y i , b i ), β i ) such that y i ∈ Ob C, b i ∈ Ob F(y i ) <strong>and</strong> β i : y i →<br />

x 0 . Moreover, h i+1 : y i → y i+1 satisfies β i+1 h i+1 = β i , <strong>and</strong><br />

l i+1 : F(h i+1 )(b i ) → b i+1 , for every i ≥ 0. Note that there are<br />

morphisms α i · · · α 1 β i : y i → x i . We define λ 2 by<br />

α<br />

λ 2 (x<br />

1 α<br />

0→ · · ·<br />

n<br />

→xn , ((y 0 , b 0 ), β 0 ) (h 1,l 1 )<br />

→ · · · (h n,l n )<br />

→ ((y n , b n ), β n ))<br />

= (y 0 , b 0 ) (h 1,l 1 )<br />

→ · · · (h n,l n )<br />

→ (y n , b n ).<br />

Let us consider NGr C F as a bisimplicial set with (p, q)-<br />

simplicies NGr C F p,q = NGr C F q . This means it is constant<br />

in horizontal direction. Clearly NGr C F ∗ = diagNGr C F ∗,∗ .<br />

The simplicial map λ 2 is the diagonalization <strong>of</strong> an obvious<br />

bisimlicial map λ : ⊕ ∗ Gr C ˜F → NGr C F ∗,∗ .<br />

Thus, by Theorem 2.4.7 (1), we only have to show for each q,<br />

λ ∗,q : ⊕ ∗ N ˜F∗,q → NGr C F ∗,q is a homotopy equivalence. But<br />

λ ∗,q is the coproduct over all q-simplicies (x 0 , c 0 ) → · · · →


88 2 Simplicial methods<br />

(x q , c q ) <strong>of</strong> NGr C F <strong>of</strong> the map ⊕ p x q\Id C → ⊕ p<br />

•. Then by<br />

Theorem 2.4.7 (2) λ ∗,q , <strong>and</strong> so λ 2 , are homotopy equivalences.<br />

Step 4: Finally we establish a simplicial hopmotopy<br />

: hocolim C N ˜F × N1 → NGrC F<br />

from ηλ 1 to λ 2 . We define it by<br />

(i+1) zeros<br />

α<br />

(x<br />

1 α<br />

0→ · · ·<br />

n<br />

{ }} {<br />

→xn , ((y 0 , b 0 ), β 0 ) → · · · → ((y n , b n ), β n ))×( 0, · · · , 0, 1, · · ·<br />

= (y 0 , b 0 ) (h 1,l 1 )<br />

→ · · · (h i,l i )<br />

→ (y i , b i ) (α i+1···α 1 β i ,F(α i+1···α 1 β i+1 )(l i+1 ))<br />

−→<br />

(x i+1 , F(α i+1 · · · α 1 β i+1 )(b i+1 )) (α i+2,F(α i+2···α 1 β i+2 )(l i+2 ))<br />

−→<br />

(x i+2 , F(α i+2 · · · α 1 β i+2 )(b i+2 )) → · · · → (x n , F(α n · · · α 1 β n )(b n )).<br />

One may verify that all simplicial identities are met. Hence<br />

we are done.<br />

⊓⊔


Chapter 3<br />

Category algebras <strong>and</strong> their representations<br />

Abstract The concept <strong>of</strong> a category algebra is introduced<br />

here. The relation ship between the module category <strong>of</strong> a category<br />

algebra <strong>and</strong> an appropriate functor category is demonstrated<br />

at the very beginning. We shall characterize the intrinsic<br />

properties <strong>of</strong> category algebras. Especially we show<br />

how one may compare a category algebra with a cocommutative<br />

bialgebra so such algebras enjoy similar homological<br />

properties. We are particularly interested in finite EI category<br />

algebras because we may classify their indecomposable<br />

projective <strong>and</strong> simple modules. When dealing with category<br />

algebras, the usual homological tools, such as induction <strong>and</strong><br />

coinduction, are replaced by Kan extensions. Examples are<br />

supplied to show how one may compute Kan extensions <strong>of</strong><br />

various modules.<br />

Throughout this chapter, the base ring R is always a commutative<br />

ring with identity. A module will normally be a left<br />

module, unless otherwise specified. If S is a set, then RS<br />

st<strong>and</strong>s for the free R-module generated by S.<br />

3.1 Basic concepts <strong>and</strong> examples<br />

3.1.1 Category algebras<br />

The category algebra is a natural way to linearize a category.


90 3 Category algebras <strong>and</strong> their representations<br />

Definition 3.1.1. Let C be a small category <strong>and</strong> R a commutative<br />

ring. The category algebra RC is a free R-module<br />

whose basis is the set <strong>of</strong> morphisms <strong>of</strong> C. We define a product<br />

on the basis elements <strong>of</strong> RC by<br />

{<br />

α ◦ β, if α <strong>and</strong> β can be composed in C;<br />

α ∗ β =<br />

0 , otherwise<br />

<strong>and</strong> then extend this product linearly to all elements <strong>of</strong> RC.<br />

With this product, RC becomes an associative R-algebra.<br />

If Ob C is finite, it is easy to see that ∑ x∈Ob C 1 x is the identity<br />

<strong>of</strong> RC where 1 x is the identity <strong>of</strong> Aut C (x). If a category<br />

C is finite, then Ob C is finite <strong>and</strong> RC is <strong>of</strong> finite R-rank.<br />

Another way to linearize a category C over a ring R is to<br />

construct a new category C R whose objects are the same as C<br />

while the morphisms between any two objects x, y ∈ C R are<br />

Hom CR (x, y) := RHom C (x, y). This category C R is additive<br />

<strong>and</strong> one can obtain the category algebra RC by forgetting the<br />

category structure in an obvious way.<br />

Example 3.1.2. Suppose C is the following finite category<br />

1 x <br />

x<br />

with g 2 = 1 y <strong>and</strong> α = gα. Then the category algebra RC is<br />

<strong>of</strong> rank 4 with identity 1 RC = 1 x + 1 y . It contains two group<br />

algebras R{1 x } <strong>and</strong> R{1 y , g}.<br />

We say C is connected if C as a (directed) graph is connected.<br />

Every category C can be written as the disjoint union<br />

<strong>of</strong> connected components C = ⊎ i∈J C i , where each C i is a connected<br />

full subcategory <strong>and</strong> J is an index set. As a consequence<br />

the category algebra RC becomes a direct product <strong>of</strong><br />

α<br />

1 y<br />

<br />

<br />

y<br />

g


3.1 Basic concepts <strong>and</strong> examples 91<br />

ideals RC i , i ∈ J. Thus in order to study the properties <strong>of</strong> RC<br />

it suffices to study the properties <strong>of</strong> each RC i . For simplicity<br />

<strong>and</strong> some technical reasons we <strong>of</strong>ten make the connectedness<br />

assumption.<br />

3.1.2 <strong>Representations</strong> <strong>of</strong> categories <strong>and</strong> Mitchell’s Theorem<br />

We shall show that a fundamental property <strong>of</strong> the category<br />

algebra RC is that it provides a mechanism for investigating<br />

R-representations <strong>of</strong> C, which we define now.<br />

Definition 3.1.3. An (R-)representation <strong>of</strong> a category C is<br />

a (covariant) functor M : C → R-mod.<br />

When the base ring is understood, we <strong>of</strong>ten abbreviate an<br />

R-representation as a representation <strong>of</strong> C. All representations<br />

<strong>of</strong> C form a functor category (R-mod) C , which is an abelian<br />

category with enough projectives <strong>and</strong> injectives so we can<br />

talk about subfunctors <strong>and</strong> quotient functors. The following<br />

fundamental theorem <strong>of</strong> category representations tells us an<br />

alternative description <strong>of</strong> the functor category.<br />

Theorem 3.1.4 (Mitchell). For any small category C with<br />

finitely many objects, there exist functors ι : (R-mod) C →<br />

RC-mod <strong>and</strong> σ : RC-mod → (R-mod) C such that<br />

1. σ ◦ ι ∼ = Id (R-mod)<br />

C, <strong>and</strong><br />

2. ι is fully faithful, <strong>and</strong> moreover if Ob C is finite then ι◦σ ∼ =<br />

Id RC-mod .<br />

Thus if Ob C is finite, the R-representations <strong>of</strong> C can be<br />

identified with the unital RC-modules.<br />

Pro<strong>of</strong>. Assume F : C → R-mod is a representation <strong>of</strong> C.<br />

We construct a free R-module M F = ⊕ x∈Ob C<br />

F (x). For any<br />

m ∈ F (x) <strong>and</strong> morphism α ∈ Mor C, we ask α·m = F (α)(m),<br />

if t(α) = x, or α · m = 0 if t(α) ≠ x. By extending this


92 3 Category algebras <strong>and</strong> their representations<br />

operation linearly we obtain an RC-module structure on M F .<br />

This construction can be easily verified to define a functor<br />

ι : (R-mod) C → RC-mod.<br />

Conversely, if M is an RC-module, we may define a functor<br />

F M by F M (x) = 1 x · M. Since, if α ∈ Hom C (x, y) <strong>and</strong> m ∈<br />

1 x · M, α · m = (1 y ◦ α) · m = 1 y · (α · m) ∈ 1 y · M, we see<br />

that F M : C → R-mod is well defined. It induces a functor<br />

σ : RC-mod → (R-mod) C .<br />

We may readily check these two functors satisfy 1 <strong>and</strong> 2. ⊓⊔<br />

Similar statements can be made between right RC-modules<br />

<strong>and</strong> contravariant functors from C to R-mod.<br />

Remark 3.1.5. Throughout these notes we will be particularly<br />

interested in RC-modules lying in ι{(R-mod) C }. One reason<br />

is that many important modules are indeed <strong>of</strong> this form. Another<br />

reason is that, on top <strong>of</strong> module-theoretic methods, we<br />

may apply simplicial or homotopy-theoretic tools. Hence we<br />

want to focus on categories with finitely many objects, although<br />

many results make sense for arbitrary small categories.<br />

Moreover since there is a well developed representation theory<br />

<strong>of</strong> finite-dimensional algebras, from now on we will restrict<br />

our attention to the finite categories. Consequently for practical<br />

reasons we will not distinguish between RC-mod <strong>and</strong><br />

(R-mod) C , <strong>and</strong> normally refer to an object in these categories<br />

as an RC-module.<br />

Among all RC-modules, there is a distinguished one that we<br />

introduce below. It plays a key role throughout these notes.<br />

Definition 3.1.6. For any category C, the constant functor<br />

or trivial module R : C → R-mod, is defined by R(x) = R<br />

for all x ∈ Ob C <strong>and</strong> R(α) = Id R for all α ∈ Mor C.


3.1 Basic concepts <strong>and</strong> examples 93<br />

We provide several examples <strong>of</strong> simplicially constructed<br />

modules where both the module <strong>and</strong> functor aspects are described.<br />

Example 3.1.7.1. The constant functor R corresponds to the<br />

module R Ob C, on which RC acts via α · x = y, if α ∈<br />

Hom C (x, y), or zero otherwise.<br />

2. The regular module RC corresponds to a functor such that<br />

RC(x) = 1 x · RC = RHom C (−, x).<br />

3. The opposite algebra (RC) op is isomorphic to the algebra <strong>of</strong><br />

the opposite category C op , <strong>and</strong> thus a set <strong>of</strong> base elements<br />

are Mor C op . Consider the enveloping category <strong>of</strong> C, namely<br />

C e = C × C op . Then we have<br />

RC e ∼ = (RC) e = (RC) ⊗ R (RC) op ,<br />

the enveloping algebra <strong>of</strong> RC. Consequently RC is an RC e -<br />

module via (α, β op ) · γ = αγβ, for any α, β, γ ∈ Mor C. As<br />

a functor RC : C e → R-mod, we have<br />

RC(y, x) = RHom C (x, y)<br />

if (x, y) ∈ Ob C e .<br />

Occasionally we will need to discuss right RC-modules. A<br />

dual version <strong>of</strong> Mitchell’s theorem says that the right RCmodules<br />

corresponds to the contra-variant functors/representations<br />

<strong>of</strong> C. There exists a natural functor (−) ∧ = Hom R (−, R) :<br />

RC-mod → mod-RC. We give an explicit description <strong>of</strong> the<br />

dual module M ∧ = Hom R (M, R) (a right RC-module) <strong>of</strong><br />

M ∈ RC-mod. In case R is a field, ∧ induces an antiisomorphism<br />

between RC-mod <strong>and</strong> mod-RC.<br />

Lemma 3.1.8. Suppose M ∈mod-RC. Then its dual M ∧ ∈<br />

RC-mod has values M ∧ (x) = M(x) ∧ = Hom R (M(x), R),<br />

<strong>and</strong> each α ∈ Mor C acts via M(α) ∧ = Hom R (M(α)(−), R).


94 3 Category algebras <strong>and</strong> their representations<br />

Pro<strong>of</strong>. As a free R-module, M ∧ = ⊕ x∈Ob C M(x)∧ . Each α ∈<br />

Hom C (x, y) acts as M(α) : M(y) → M(x) <strong>and</strong> thus induces<br />

a map M(α) ∧ : M(x) ∧ → M(y) ∧ . One can readily check that<br />

M ∧ (x) is exactly M(x) ∧ so we know the structure <strong>of</strong> M ∧ as<br />

a right RC-module.<br />

⊓⊔<br />

The next statement follows from direct calculation.<br />

Corollary 3.1.9. We have R ∧ = R. Here the first R is a<br />

left module <strong>and</strong> the second is a right module.<br />

3.1.3 Three examples<br />

Let us consider three motivating examples.<br />

Example 3.1.10. Let G be a (discrete) group. Then the group<br />

can be regarded as a category with only one object •, whose<br />

morphisms are the elements <strong>of</strong> G. The group algebra RG is<br />

the same as the category algebra RG. A left RG-module M<br />

can be regarded as the representation <strong>of</strong> G given by a certain<br />

functor Φ : G → R-mod, sending • to M <strong>and</strong> Aut G (•) = G<br />

into Aut R (M). The trivial RG-module R is exactly the trivial<br />

module <strong>of</strong> RG.<br />

Note that in this case G ∼ = G op <strong>and</strong> thus RG ∼ = RG op <strong>and</strong><br />

RG e ∼ = RG ⊗ RG ∼ = R(G × G).<br />

A quiver q = (Γ 0 , Γ 1 ) is a directed graph having Γ 0 <strong>and</strong><br />

Γ 1 as the set <strong>of</strong> vertices <strong>and</strong> the set <strong>of</strong> arrows, respectively.<br />

The path algebra <strong>of</strong> q can be thought as a category algebra<br />

as follows. Any directed graph G = (Γ 0 , Γ 1 ) may be used to<br />

generate a category C G on the same set Γ 0 <strong>of</strong> objects, where<br />

the morphisms <strong>of</strong> this category are the strings <strong>of</strong> composable<br />

arrows <strong>of</strong> G. It is called the free category generated by G.<br />

Example 3.1.11. The path algebra <strong>of</strong> a quiver q is the category<br />

algebra RC q <strong>of</strong> the free category C q .


3.1 Basic concepts <strong>and</strong> examples 95<br />

It is necessary to point out that, given a category, its category<br />

algebra is usually different from its path algebra, if we<br />

consider the category as a quiver at the same time. Nevertheless,<br />

there is a relationship between these two algebras, as we<br />

now explain.<br />

Proposition 3.1.12. Let q be a category. We may regard q<br />

as a quiver <strong>and</strong> form the free category C q over q. There is a<br />

natural functor u : C q → q, which extends to a surjective homomorphism,<br />

still denoted by u : RC q → Rq, from the path<br />

algebra <strong>of</strong> q to the category algebra <strong>of</strong> q, such that its kernel<br />

I is generated by { α 1<br />

← α 2<br />

← − α ←<br />

1α 2<br />

}, where α1 <strong>and</strong> α 2 are arrows<br />

<strong>of</strong> q. This epimorphism induces a natural isomorphism <strong>of</strong><br />

R-algebras RC q /I <strong>and</strong> Rq.<br />

Pro<strong>of</strong>. The functor u is defined as follows. For each x ∈ Ob C q ,<br />

φ(x) = x. For each α ∈ Mor C q , u(α) = the composite <strong>of</strong><br />

the maps in the string α. It can be extended linearly to an<br />

epimorphism u : RC q → Rq, having the kernel I. ⊓⊔<br />

The last example <strong>of</strong> a category algebra is the incidence algebra<br />

<strong>of</strong> a locally finite poset (partially ordered set). A (closed)<br />

interval <strong>of</strong> P is a subposet [x, y] ⊂ P which consists <strong>of</strong> all<br />

objects z such that x ≤ z ≤ y for a given pair <strong>of</strong> objects<br />

x, y in P. The incidence algebra I(P, R) is an R-algebra <strong>of</strong> all<br />

functions f : Int(P) → R, where Int(P) is the set <strong>of</strong> intervals<br />

<strong>of</strong> P. The multiplication (also called convolution) if defined<br />

by<br />

(fg)([x, y]) = ∑<br />

f([x, z])g([z, y]).<br />

x≤z≤y<br />

It has an identity δ such that<br />

{<br />

1, if x = y;<br />

δ([x, y]) =<br />

0 , otherwise.


96 3 Category algebras <strong>and</strong> their representations<br />

On the other h<strong>and</strong> since whenever x ≤ y in a poset we can<br />

replace ≤ with an arrow x → y, a poset is always a category.<br />

Example 3.1.13. Let P be a finite poset. Then the incidence<br />

algebra I(P, R) <strong>of</strong> P is isomorphic to the category algebra<br />

RP.<br />

3.2 A closed symmetric monoidal category<br />

Suppose C is finite. We describe the intrinsic structure <strong>of</strong> RC<br />

which make it comparable with a cocommutative bialgebra.<br />

Later on we shall see RC-mod possesses similar homological<br />

properties as the module category <strong>of</strong> a cocommutative bialgebra.<br />

3.2.1 Tensor structure <strong>and</strong> an intrinsic characterization <strong>of</strong> category algebras<br />

Suppose G is a group. The category RG-mod is a symmetric<br />

monoidal category equipped with a tensor product − ⊗ R −<br />

<strong>and</strong> a tensor identity R. The reason why RG-mod enjoys such<br />

good properties is that RG is a cocommutative Hopf algebra.<br />

For general properties <strong>of</strong> Hopf algebras, we refer the reader<br />

to []. For an arbitrary unital associative R-algebra there are<br />

maps<br />

1. There is a map µ : A ⊗ R A → A, called the multiplication.<br />

2. The following diagram is commutative (associativity)<br />

A ⊗ A ⊗ A µ⊗Id<br />

Id⊗µ <br />

A ⊗ A µ<br />

A ⊗ A<br />

µ<br />

3. There is a map ι : R → A, called the unit, <strong>and</strong> the following<br />

diagram is commutative<br />

<br />

A


3.2 A closed symmetric monoidal category 97<br />

R ⊗ A ι⊗Id <br />

A ⊗ A<br />

∼ =<br />

µ<br />

Id⊗ι<br />

<br />

∼ =<br />

A ⊗ R<br />

A<br />

If A = RG, it is also a counital coalgebra, which means, on<br />

top <strong>of</strong> the previous maps, we have the following extra maps,<br />

reversing maps in 1-3.<br />

4. There is a comultiplication, RG → RG ⊗ R RG.<br />

6. The following diagram is commutative (coassociativity)<br />

∆<br />

RG<br />

<br />

RG ⊗ RG<br />

∆ <br />

RG ⊗ RG<br />

∆⊗Id<br />

Id⊗∆<br />

<br />

RG ⊗ RG ⊗ RG,<br />

5. The group algebra has an augmentation map ɛ : RG → R,<br />

also called the counit.<br />

R ⊗ RG<br />

∼ =<br />

ɛ⊗Id<br />

RG<br />

∆ <br />

∼ =<br />

RG ⊗ RGId⊗ɛ RG ⊗ R<br />

Any R-module satisfying conditions 4-6 is called a counital<br />

coassociative coalgebra. Any R-algebra satisfying 4-6 is<br />

named a bialgebra.<br />

A group algebra is usually not commutative, but<br />

7. there is a twist map τ : RG ⊗ R RG → RG ⊗ R RG, given<br />

by τ(a ⊗ b) = b ⊗ a, such that the following diagram is<br />

commutative<br />

∆<br />

RG ⊗ RG<br />

RG<br />

τ<br />

∆<br />

<br />

RG ⊗ RG.


98 3 Category algebras <strong>and</strong> their representations<br />

An R-coalgebra or bialgebra satisfying 7 is called cocommutative.<br />

8. Moreover there is an antipode η : RG → RG in the sense<br />

that if δ(a) = ∑ i b i ⊗ c i then ∑ i b iη(c i ) = ∑ i η(b i)c i =<br />

ɛ(a)e.<br />

If a bialgebra has an antipode, then it is called a Hopf algebra.<br />

Remark 3.2.1. For a group algebra, η is explicitly given by<br />

g → g −1 . The comultiplication is given by g ↦→ g ⊗ g.<br />

A cocommutative bialgebra A has the significant property<br />

that A-mod is a symmetric monoidal category with tensor<br />

identity R. More precisely it means that for any two A-<br />

modules M <strong>and</strong> N, M ⊗ R N is still an A-module, M ⊗ R N ∼ =<br />

N ⊗ R M, <strong>and</strong> M ⊗ R R ∼ = M. In other words, A-mod inherits<br />

various constructions on R-mod. Note that the tensor identity<br />

R plays the role <strong>of</strong> both the unit <strong>and</strong> co-unit <strong>of</strong> A. If A<br />

is Hopf, then even Hom R (M, N) becomes an A-module such<br />

that<br />

Hom A (L ⊗ M, N) ∼ = Hom A (L, Hom R (M, N)).<br />

In the literature, Hom R (M, N) is called a function object or<br />

the internal hom. Furthermore when R = k is a field, A<br />

is cocommutative Hopf <strong>and</strong> M, N are finite-dimensional, we<br />

have an isomorphism <strong>of</strong> A-modules<br />

M ∧ ⊗ N ∼ = Hom R (M, N).<br />

Since R ∈ A-mod, M ∧ is a left A-module under the circumstance,<br />

satisfying (M ∧ ) ∧ ∼ = M.<br />

In what follows, we begin with a description about how R-<br />

mod gives rise to a symmetric monoidal category structure<br />

on RC-mod. Then we characterize the structure <strong>of</strong> RC using


3.2 A closed symmetric monoidal category 99<br />

some structure maps comparable to those <strong>of</strong> a cocommutative<br />

bialgebra. In this way we demonstrate why a category algebra<br />

<strong>and</strong> a cocommutative bialgebra, as well as their module categories,<br />

are similar yet different. It is the intrinsic structure <strong>of</strong> a<br />

category algebra that makes it a natural <strong>and</strong> interesting subject<br />

<strong>of</strong> investigation. We note that R-mod itself is the module<br />

category <strong>of</strong> the R-category algebra <strong>of</strong> the trivial category •,<br />

based on Mitchell’s theorem.<br />

Fix a finite category C. The so-called internal product on<br />

(R-mod) C ≃ RC-mod, in which the tensor product is denoted<br />

by ˆ⊗ R , is defined by (M ˆ⊗ R N)(x) = M(x) ⊗ R N(x)<br />

for any M, N ∈ RC-mod ≃ (R-mod) C <strong>and</strong> x ∈ Ob C. The<br />

module structure <strong>of</strong> M ˆ⊗N can be viewed as given by the comultiplication<br />

∆ : RC → RC ⊗ R RC, induced by the canonical<br />

diagonal functor ∆ : C → C × C whose action on each<br />

α ∈ Mor C is ∆(α) = α ⊗ α. One can easily verify that the<br />

trivial module R is the tensor identity with respect to ˆ⊗ R . For<br />

the sake <strong>of</strong> simplicity, we shall write ⊗ for ⊗ R , <strong>and</strong> ˆ⊗ for ˆ⊗ R ,<br />

throughout this book. We note that in the literature (see for<br />

example [25]), the symbol ⊗ is <strong>of</strong>ten used instead <strong>of</strong> ˆ⊗ for the<br />

internal tensor product <strong>of</strong> functors. The new notation ˆ⊗ is introduced<br />

because we need to distinguish between M ⊗ N <strong>and</strong><br />

M ˆ⊗N. In fact, as R-modules, the inclusion M ˆ⊗N ⊂ M ⊗N,<br />

for any RC-modules M <strong>and</strong> N, is <strong>of</strong>ten strict.<br />

As we mentioned above, the diagonal functor ∆ : C → C ×C<br />

induces a co-multiplication on the category algebra RC. The<br />

co-multiplication ∆ : RC → RC ⊗ RC almost gives us a coalgebra<br />

structure on RC except that there is not a suitable<br />

choice <strong>of</strong> a map RC → R which would serve as the co-unitary<br />

map. Recall from Example 3.1.7 (1) that the trivial module<br />

R can be realized by R Ob C. The next result states a natural<br />

augmentation map from RC to R Ob C.


100 3 Category algebras <strong>and</strong> their representations<br />

Lemma 3.2.2. There exists a surjective linear map<br />

defined on base elements by<br />

where t(α) is the target <strong>of</strong> α.<br />

RC ɛ ↠R Ob C = R,<br />

ɛ(α) = t(α),<br />

Pro<strong>of</strong>. The surjective map induces a RC-module structure on<br />

R Ob C as follows. If x ∈ Ob C <strong>and</strong> β ∈ Hom C (x ′ , y), then<br />

βx = y when x = x ′ , <strong>and</strong> βx = 0 otherwise.<br />

⊓⊔<br />

If we return to ∆, we realize that the image <strong>of</strong> it really<br />

lies in RC ˆ⊗RC, a subspace <strong>of</strong> RC ⊗ RC, <strong>and</strong> moreover<br />

∆ : RC → RC ˆ⊗RC becomes a RC-map, since RC ˆ⊗RC, unlike<br />

RC ⊗ RC, is a well-defined RC-module. The above observations<br />

hint that, in order to get a sound “co-algebra” structure,<br />

one needs to use ˆ⊗ other than ⊗ to define the structure<br />

maps. This motivates us to write down the following maps<br />

which resemble almost all <strong>of</strong> the structure maps for a cocommutative<br />

bialgebra. Abusing terminology, we adopt the same<br />

names for the structure maps <strong>of</strong> a category algebra, such as<br />

co-multiplication <strong>and</strong> co-unit, as their counterparts for a bialgebra.<br />

In a coalgebra, the augmentation map <strong>and</strong> co-unit are<br />

the same, so by analogy the map ɛ in the preceding lemma<br />

will occasionally be called the co-unit. We emphasize that the<br />

unit, given by the natural inclusion map R ∼ ι<br />

= R · 1 RC →RC,<br />

is different from the co-unit.<br />

Proposition 3.2.3. Let RC be the category algebra <strong>of</strong> a finite<br />

category C. Then we have the following R-linear maps:<br />

a co-multiplication ∆ : RC → RC⊗RC, defined by ∆( ∑ ∑<br />

α λ αα) =<br />

α λ αα ⊗ α, a co-unit ɛ : RC → R defined as above, <strong>and</strong> a<br />

twist map τ : RC ⊗RC → RC ⊗RC defined on base elements


3.2 A closed symmetric monoidal category 101<br />

by τ(α ⊗ α ′ ) = α ′ ⊗ α, such that the following diagrams are<br />

commutative: 1. co-associativity<br />

∆<br />

RC<br />

<br />

RC ⊗ RC<br />

∆ <br />

RC ⊗ RC<br />

2. co-unitary property<br />

R ˆ⊗RC<br />

∼ =<br />

ɛ⊗Id<br />

<strong>and</strong> 3. co-commutativity<br />

∆<br />

RC ⊗ RC<br />

∆⊗Id<br />

Id⊗∆ RC ⊗ RC ⊗ RC,<br />

RC<br />

∆ <br />

∼ =<br />

<br />

RC ˆ⊗RC<br />

Id⊗ɛ RC ˆ⊗R<br />

RC<br />

τ<br />

∆<br />

<br />

RC ⊗ RC.<br />

If we denote by µ the multiplication, then we have furthermore<br />

three commutative diagrams: 4. Multiplication <strong>and</strong> comultiplication<br />

RC ⊗ RC<br />

∆⊗∆ <br />

RC ⊗ RC ⊗ RC ⊗ RC<br />

µ<br />

RC<br />

∆ <br />

RC ⊗ RC<br />

Id⊗τ⊗Id<br />

5. Unit <strong>and</strong> co-multiplication:<br />

RC<br />

<strong>and</strong> 6. Unit <strong>and</strong> co-unit<br />

µ<br />

R ⊗ R<br />

∆<br />

ι⊗ι<br />

<br />

µ⊗µ<br />

<br />

RC ⊗ RC ⊗ RC ⊗ RC,<br />

<br />

RC ˆ⊗RC,


102 3 Category algebras <strong>and</strong> their representations<br />

RC<br />

ι<br />

R<br />

ɛ<br />

where the R-linear map η : R → R is defined by 1 ↦→<br />

∑<br />

x∈Ob C x,<br />

We note that the only missing diagram is the compatibility<br />

<strong>of</strong> multiplication <strong>and</strong> co-unit. The reason is that one usually<br />

cannot give R a meaningful algebra structure so that<br />

ɛ : RC → R becomes an algebra homomorphism. The existence<br />

<strong>of</strong> the above tensor structure ˆ⊗ is well known. Notably<br />

it has been used to define the internal product in functor<br />

homology theory, see for example [25].<br />

Remark 3.2.4. If we remove the finiteness condition on C, then<br />

there is no identity in the algebra kC. Nevertheless many <strong>of</strong><br />

the constructions are still valid, although in this case, we are<br />

forced to use the full subcategory V ect C k<br />

because kC-mod does<br />

not have a tensor structure.<br />

Remark 3.2.5. If C = P happens to be a poset, then there<br />

is another co-multiplication that one can find in [72]. Let<br />

α ∈ Mor P, one has ¯∆ : kP → kP ⊗ kP such that<br />

¯∆(α) = ∑ {β,γ|βγ=α}<br />

β ⊗ γ. Nevertheless it is possible to give<br />

an incidence algebra a Hopf algebra structure. The augmentation<br />

map ¯ɛ : kP → k is given by ¯ɛ(α) = 1 if α is an identity,<br />

or ¯ɛ(α) = 0 otherwise. The antipode kP → kP can also be<br />

explicitly constructed. Suppose α ∈ Mor P is a morphism in<br />

the poset. Then a factorization <strong>of</strong> α is a way <strong>of</strong> writing α as<br />

a product α 1 α 2 · · · α n for some non-identity α i ∈ Mor P. The<br />

length l(α) <strong>of</strong> α is the maximal n for such a factorization to<br />

exist. Then the antipode is given by α ↦→ ∑ β (−1)l(β) β, in<br />

which β runs over all possible factors <strong>of</strong> α.<br />

η<br />

<br />

R,


3.2 A closed symmetric monoidal category 103<br />

Indeed one can do this for the category algebras <strong>of</strong> the socalled<br />

(finite) Möbius categories. A main feature <strong>of</strong> a Möbius<br />

category is that each object only has one automorphism.<br />

It seems that RC can only be a (cocommutative) Hopf algebra<br />

if C is <strong>of</strong> the two extreme structures: either a group<br />

(all morphisms are invertible), or a Möbius category (identity<br />

morphisms are the only invertible morphisms) .<br />

Remark 3.2.6. Let A be an R-algebra. Although A-mod is not<br />

monoidal in general, the module category A e -mod is always<br />

equipped with a tensor product ⊗ A such that A is the tensor<br />

identity.<br />

When A = RC is a category algebra, (RC) e ∼ = RC e becomes<br />

the category algebra <strong>of</strong> the category C e := C × C op , <strong>and</strong> hence<br />

there are two distinct monoidal structures on RC e -mod. The<br />

monoidal structure given by ⊗ RC is more interesting to us, <strong>and</strong><br />

we shall deal with it in Chapter 5 on Hochschild cohomology.<br />

3.2.2 The internal hom<br />

Usually one can not define an antipode for a category algebra.<br />

This causes a problem when one attempts to define<br />

Hom k (M, N) as a sensible kC-module (generalizing the group<br />

case). We shall give a remedy below. Another relevant fact is<br />

that the product <strong>of</strong> two projective kC-modules is in general<br />

not projective, see Example 3.. These make many homological<br />

properties, such as the cohomology theory, <strong>of</strong> a finitedimensional<br />

category algebra different from those <strong>of</strong> a finitedimensional<br />

cocommutative Hopf algebra.<br />

Proposition 3.2.7 (Swenson). Suppose C is a finite category<br />

(or at least Ob C is finite). Let M, N ∈ RC-mod. Then<br />

we can define an internal hom Hom(M, N) ∈ RC-mod.


104 3 Category algebras <strong>and</strong> their representations<br />

Pro<strong>of</strong>. We want to define Hom(M, N) such that there is an<br />

isomorphism<br />

Hom RC (L ˆ⊗M, N) ∼ = Hom RC (L, Hom(M, N)).<br />

Let RC = RC · 1 x . Assume the above isomorphism. Then the<br />

right h<strong>and</strong> side is Hom RC (RC·1 x , Hom(M, N)) ∼ = Hom(M, N)(x),<br />

<strong>and</strong> the left h<strong>and</strong> side is Hom RC (RC · 1 x ˆ⊗M, N). Hence we<br />

may define Hom(M, N) by<br />

Hom(M, N)(x) = Hom RC (RC · 1 x ˆ⊗M, N).<br />

When C = G is a group, it is straightforward from Swenson’s<br />

definition that Hom(M, N) ∼ = Hom R (M, N) as RG-modules.<br />

Alternatively consider the diagonal functor ∆ : C → C × C.<br />

Because R(C ×C) ∼ = RC ⊗RC, for L, M, N ∈ RC-mod, L⊗M<br />

is a R(C × C)-module, <strong>and</strong><br />

Hom RC (L ˆ⊗M, N) = Hom RC (Res ∆ (L ⊗ M), N)<br />

∼ = HomR(C×C) (L ⊗ M, RK ∆ N)<br />

∼ = HomRC⊗RC (L ⊗ M, RK ∆ N)<br />

∼ = HomRC (L, Hom RC (M, RK ∆ N)).<br />

Here in Hom RC (M, RK ∆ N), RK ∆ N ∈ RC ⊗ RC-mod is<br />

considered as a (R · 1 RC ) ⊗ RC-module <strong>and</strong> hence a RCmodule.<br />

The RC ⊗ RC-, or rather the RC ⊗ (R · 1 RC )-, module<br />

structure on RK ∆ N provides a RC-module structure on<br />

Hom RC (M, RK ∆ N). Then we may verify that Hom(M, N) =<br />

Hom RC (M, RK ∆ N).<br />

With the internal hom, one might attempt to define a dual<br />

module <strong>of</strong> M by Hom(M, R) in order to generalize the group<br />

case. The following example explain why this is not a good<br />

idea.<br />

⊓⊔


3.3 Functors between module categories 105<br />

Example 3.2.8. Let k be a field <strong>of</strong> characteristic two <strong>and</strong> C<br />

the following category<br />

{1 x } <br />

x<br />

α<br />

β<br />

y<br />

{1 y ,g}<br />

with g 2 = 1 y , gα = α <strong>and</strong> gβ = β.<br />

We consider two modules S x,k <strong>and</strong> S y,k such that S x,k (x) =<br />

k, S x,k (y) = 0, S y,k (x) = 0 <strong>and</strong> S y,k (y) = k. Then Hom(S x,k , k)<br />

can be easily shown to be isomorphic to S x,k . However Hom(S y,k , k) ∼ =<br />

S 2 x,k ⊕ S y,k because kC1 x ˆ⊗S y,k<br />

∼ = S<br />

2<br />

y,k<br />

<strong>and</strong> kC1 y ˆ⊗S y,k<br />

∼ = kC1y .<br />

Thus Hom(Hom(S y,k , k), k) ≁ = Sy,k .<br />

3.3 Functors between module categories<br />

We investigate functors for comparing two module categories<br />

over category algebras.<br />

3.3.1 Restriction on algebras <strong>and</strong> modules<br />

We prove some basic properties, many <strong>of</strong> which follow directly<br />

from simple reasoning.<br />

Definition 3.3.1. Suppose u : D → C is a (covariant) functor.<br />

We define Res u : RC-mod → RD-mod to be the restriction<br />

along u. Given a module M ∈ RC-mod, we have<br />

Res u M = M ◦ u ∈ RD-mod.<br />

Lemma 3.3.2. Let u : D → C be a functor. Then Res u R =<br />

R.<br />

Since a functor u : D → C also extends linearly to a natural<br />

map <strong>of</strong> R-modules ū : RD → RC, it is reasonable to ask if<br />

ū is always an algebraic homomorphism because if it is, the<br />

so-called change-<strong>of</strong>-base-ring or the representation-theoretic<br />

restriction, ↓ RC<br />

RD : RC-mod → RD-mod, should coincide with


106 3 Category algebras <strong>and</strong> their representations<br />

Res u . The answer to the question is no, <strong>and</strong> here is a simple<br />

example. Let D = 1 <strong>and</strong> C = •. There is a unique functor pt :<br />

D → C. The induced map ¯pt : RD → RC is not an algebra<br />

homomorphism since the product <strong>of</strong> the two isomorphisms in<br />

D is zero while the product <strong>of</strong> their images is not.<br />

Proposition 3.3.3. A functor u : D → C extends linearly<br />

to an algebra homomorphism ū : RD → RC if <strong>and</strong> only<br />

if u is injective on Ob D. When this happens, the induced<br />

functor followed by 1 RD , 1 RD ◦ ↓ RC<br />

RD : RC-mod → RD-mod, is<br />

exactly Res u .<br />

Pro<strong>of</strong>. We know u(βα) = u(β)u(α) for any pair <strong>of</strong> composable<br />

morphisms α, β in D. The injectivity <strong>of</strong> u implies two<br />

morphisms α, β ∈ Mor(D) are composable if <strong>and</strong> only if<br />

u(α), u(β) ∈ Mor(C) are composable.<br />

If u is injective on Ob D, then we define a map ū : RD →<br />

RC as the linear extension <strong>of</strong> functor u, i.e., ū( ∑ ∑<br />

i r iα i ) =<br />

i r iū(α i ) for any r i ∈ R, α i ∈ Mor(D). This ū is indeed<br />

an algebra homomorphism because our previous observation<br />

<strong>of</strong> u implies ū(( ∑ j r jβ j )( ∑ i r iα i )) = ū( ∑ j r jβ j )ū( ∑ i r iα i ) is<br />

always true.<br />

On the other h<strong>and</strong> if the linear extension ū : RD → RC is<br />

an algebra homomorphism then we must have ū(0) = 0 <strong>and</strong><br />

then ū(1 x )ū(1 y ) = ū(1 x · 1 y ) = 0 unless x = y. This suggests<br />

that u is injective on Ob C ′ .<br />

When ū : RD → RC is an algebra homomorphism, we<br />

show 1 RD ◦ ↓ RC<br />

RD = Res u : RC-mod → RD-mod. Let M be an<br />

RC-module <strong>and</strong> γ ∈ Mor(D). Then it corresponds to some<br />

F M ∈ (R-mod) C . The RD-modules Res u M <strong>and</strong> 1 RD·(M ↓ RC<br />

RD )<br />

are isomorphic because Res u M = M Resu (F M ), γ · M = u(γ)M<br />

for any γ ∈ Mor D <strong>and</strong> 1 RD kills the elements <strong>of</strong> M ↓ RC<br />

RD that<br />

are not supported on any objects <strong>of</strong> D.<br />

⊓⊔


3.3 Functors between module categories 107<br />

Proposition 3.3.4. Let D <strong>and</strong> C be equivalent small categories.<br />

Then (R-mod) C ≃ (R-mod) D , an equivalence which<br />

sends the constant functor to the constant functor. If both<br />

Ob C <strong>and</strong> Ob D are finite then RC <strong>and</strong> RD are Morita equivalent.<br />

Pro<strong>of</strong>. We show that the two functor categories (R-mod) C <strong>and</strong><br />

(R-mod) D are equivalent. Then it implies the module categories<br />

RC-mod <strong>and</strong> RD-mod are equivalent, hence RC <strong>and</strong><br />

RD are Morita equivalent. In fact if u : D → C <strong>and</strong> v : C → D<br />

are equivalences, we have Res u Res v<br />

∼ = IdRD := Id RD-mod : (Rmod)<br />

D → (R-mod) D because <strong>of</strong> the following diagram<br />

M(vu(x)) = (Res u Res v M)(x) ∼ =<br />

(Id RD M)(x) = M(x)<br />

M(vu(α))=(Res u Res v M)(α)<br />

M(vu(y)) = (Res u Res v M)(y) ∼=<br />

(Id RD M)(α)=M(α)<br />

<br />

(Id RD M)(y) = M(y)<br />

where M ∈ (R-mod) D , α : x → y ∈ Mor D <strong>and</strong> Id RD is the<br />

identity functor. Similarly we can show Res v Res u<br />

∼ = IdRC : (Rmod)<br />

C → (R-mod) C . Clearly the constant functor restricts to<br />

the constant functor always.<br />

⊓⊔<br />

Remark 3.3.5. Recall from Chapter 2 that if D ≃ C then<br />

ND ≃ NC (<strong>and</strong> BD ≃ BC). The above result is an algebraic<br />

consequence <strong>of</strong> a category equivalence. On one h<strong>and</strong> the existence<br />

<strong>of</strong> a Morita equivalence implies that the two (category)<br />

algebras are the same, in terms <strong>of</strong> homological properties.<br />

On the other h<strong>and</strong> there are non-equivalent finite categories<br />

whose algebras are Morita equivalent, see Example 3.1.16 (5).<br />

By contrary a simplicial/topological equivalence does not<br />

provide a Morita equivalence. For instance the posets 0 <strong>and</strong><br />

1 are simplicial homotopy equivalent, because there are natural<br />

adjoint functors between them. However if k is a field<br />

then k0 <strong>and</strong> k1 are not equivalent as algebras. The reason


108 3 Category algebras <strong>and</strong> their representations<br />

is that k0 ∼ = k has only one simple module k while k1 has<br />

two simple modules because its identity can be written as a<br />

sum <strong>of</strong> orthogonal primitive idempotents 1 0 + 1 1 . Since their<br />

category algebras have different numbers <strong>of</strong> non-isomorphic<br />

simple modules, k0-mod <strong>and</strong> k1-mod cannot be equivalent<br />

as categories.<br />

3.3.2 Kan extensions <strong>of</strong> modules<br />

Since not every functor u : D → C has an adjoint, we have to<br />

look for other functors for comparing RD-mod <strong>and</strong> RC-mod.<br />

Clearly the restriction Res u possesses two adjoint functors,<br />

the left <strong>and</strong> right Kan extensions LK u <strong>and</strong> RK u . Here will<br />

be our first attempt to study Kan extensions <strong>of</strong> modules. We<br />

note that, because Res u is exact, LK u preserves projectives<br />

while RK u preserves injectives.<br />

In light <strong>of</strong> Proposition 3.2.3, if one is careful enough, the<br />

restriction can be written using the usual module-theoretic<br />

notation ↓ C D :=↓RC RD , <strong>and</strong> then in the literature one may see that<br />

its left adjoint <strong>and</strong> right adjoint are denoted by ↑ C D := RC ⊗ RD<br />

−, the induction, <strong>and</strong> ⇑ C D := Hom RD(RC, −), the coinduction.<br />

They are the Kan extensions in different forms. Since there<br />

is plenty <strong>of</strong> description <strong>of</strong> the induction <strong>and</strong> coinduction for<br />

algebra representations, in these notes we shall mainly discuss<br />

the functor-theoretic aspect <strong>of</strong> them through Kan extensions.<br />

A special situation <strong>of</strong> Kan extensions, the group case, can<br />

be found in Example 1.2.11. One can see that using moduletheoretic<br />

methods it is not easy to obtain most <strong>of</strong> the following<br />

results for general category algebras.<br />

Suppose M ∈ RD-mod <strong>and</strong> u : D → C is a functor. Then<br />

we obtain an RC-module LK u M. As a functor, for each x ∈<br />

Ob C, (LK u M)(x) is given by a direct limit, <strong>and</strong> if α : x → y


3.3 Functors between module categories 109<br />

is a morphism, it induces a map (LK u M)(x) → (LK u M)(y)<br />

by the universal property <strong>of</strong> −→<br />

lim. This specifies the RC-action<br />

on LK u M.<br />

Suppose P is a poset <strong>and</strong> Q is a subposet. We call Q an<br />

ideal <strong>of</strong> P if for a pair <strong>of</strong> object x ∈ Ob Q <strong>and</strong> y ∈ Ob P<br />

such that x ≤ y in P then y belongs to Q. Before we start<br />

the first calculation we mention that if D is a full subcategory<br />

<strong>of</strong> a small category C then any functor over D can be naively<br />

considered as a functor over C by asking its value to be zero<br />

on any objects that do not belong to D. In other words, any<br />

RD-module is naturally an RC-module under the assumption.<br />

Proposition 3.3.6. Suppose P is a poset <strong>and</strong> Q is an ideal<br />

together with the inclusion i : Q ↩→ P. Then for any M ∈<br />

RQ-mod, the RP-module LK i M is exactly M, regarded as<br />

an RP-module.<br />

Pro<strong>of</strong>. One can find that (LK i M)(x) ∼ = lim<br />

−→Q≤x<br />

M, where Q ≤x<br />

is a subposet <strong>of</strong> Q consisting <strong>of</strong> objects smaller or equal to x.<br />

By assumption this poset is either empty or has a terminal<br />

object x. Consequently (LK i M)(x) is either zero if x ∉ Ob Q<br />

or M(x) otherwise. Hence we are done.<br />

Alternatively we can compute RP ⊗ RQ M. For any x ∈<br />

Ob P, we have<br />

1 x·(RP⊗ RQ M) = ∑ y≤x<br />

RHom P (y, x)⊗M = ∑ y≤x<br />

RHom P (y, x)⊗1 y·M.<br />

This is not zero if <strong>and</strong> only if 1 y · M ≠ 0 which means y ∈<br />

Ob Q. By assumption, it forces x belongs to Q. Thus 1 x ·<br />

(RP⊗ RQ M) ≠ 0 if <strong>and</strong> only if x ∈ Ob Q. When this happens,<br />

1 x · (RP ⊗ RQ M) = 1 x · M. Thus we obtain the same result.<br />

⊓⊔


110 3 Category algebras <strong>and</strong> their representations<br />

The concept <strong>of</strong> an idea can be defined for a certain class <strong>of</strong><br />

categories, which are called EI-categories <strong>and</strong> will be introduced<br />

shortly. The above result stays true in this generality.<br />

We will also give a dual concept <strong>of</strong> coideal later on.<br />

Proposition 3.3.7. Suppose u : D → C satisfies the condition<br />

that for every x ∈ Ob C u/x is connected. Then<br />

LK u R = R.<br />

Especially for t : F (C) → C, we have LK t R ∼ = R as an<br />

RC-module.<br />

Pro<strong>of</strong>. When u/x is connected for an x ∈ Ob C, we have<br />

lim R = R. Then by the universal property <strong>of</strong> direct limit<br />

−→u/x<br />

we see any α : x → y must induce the identity morphism on<br />

R. Thus LK u R = R. ⊓⊔<br />

For any M ∈ RF (C)-mod we can compute LK t M explicitly.<br />

If x ∈ Ob C, then −→t/x<br />

lim M ∼ = lim −→t<br />

M by Propositions<br />

2.3.10 <strong>and</strong> 1.2.4. The category t −1 (x) has an initial ob-<br />

−1 (x)<br />

ject [1 x ] <strong>and</strong> thus in general we cannot simplify the direct<br />

limit. We want to construct an example based on this description<br />

showing that LK t (M ˆ⊗N) ≁ = LKt (M) ˆ⊗LK t (N) for<br />

M, N ∈ RF (C)-mod.<br />

Example 3.3.8. Consider the following poset P <strong>and</strong> its factorization<br />

category F (P) (a poset too)<br />

2<br />

α<br />

1 [α] [β]<br />

β<br />

3<br />

, [1 2 ]<br />

We define an RF (P)-module M by<br />

0<br />

R<br />

0<br />

R<br />

<br />

0<br />

[1 1 ]<br />

[1 3 ]


3.3 Functors between module categories 111<br />

Since we know explicitly t −1 (1), t −1 (2) <strong>and</strong> t −1 (3), we can<br />

compute directly that LK t M(1) = R 2 , LK t M(2) = LK t M(3) =<br />

0. Thus the RP-module LK t M ˆ⊗LK t M is pictured by<br />

0<br />

R 4<br />

On the other h<strong>and</strong>, since M ˆ⊗M = M, LK t (M ˆ⊗M) ≁ =<br />

LK t M ˆ⊗LK t M.<br />

Proposition 3.3.9. Let C be a small category <strong>and</strong> ∇ :<br />

F (C) → C e the skew diagonal functor. Then LK ∇ R ∼ = RC<br />

as an RC e -module.<br />

Pro<strong>of</strong>. We compute (LK ∇ R)(y, x) = lim −→∇/(y,x)<br />

R for any<br />

(y, x) ∈ Ob C e . By Proposition 2.3.13 the functor i (y,x) :<br />

∇ −1 (y, x) → ∇/(y, x) has a left adjoint. Hence Proposition<br />

1.2.4 implies that −→∇/(y,x)<br />

lim R ∼ = lim −→∇<br />

R. However<br />

−1 (y,x)<br />

∇ −1 (y, x) ∼ = Hom C (x, y) so we have −→∇<br />

lim R = RC(y, x),<br />

−1 (y,x)<br />

where RC is considered as a functor C e → R-mod. We are<br />

done.<br />

⊓⊔<br />

0<br />

For a finite group algebra RG <strong>and</strong> M, N ∈ RG-mod, we<br />

have a canonical isomorphism Hom RG (M, N) ∼ = Hom RG (R, Hom R (M,<br />

where Hom R (−, −) is the internal hom for RG-mod. The<br />

above proposition allows us to generalizes this isomorphism,<br />

but without using the previously defined internal hom for RCmod.<br />

Corollary 3.3.10. Let M, N ∈ RC-mod. Then Hom R (M, N) ∈<br />

RC e -mod <strong>and</strong><br />

Hom RC (M, N) ∼ = Hom RF (C) (R, Res ∇ Hom R (M, N))<br />

∼ = HomRC (R, RK t Res ∇ Hom R (M, N)).


112 3 Category algebras <strong>and</strong> their representations<br />

Pro<strong>of</strong>. We can define an RC e -module structure on Hom R (M, N)<br />

by<br />

[(α, β op ) · f](m) = αf(βm),<br />

for any f ∈ Hom R (M, N) <strong>and</strong> m ∈ M. Then<br />

Hom RC (M, N) ∼ = Hom RC e(RC, Hom R (M, N))<br />

∼ = HomRC e(LK ∇ R, Hom R (M, N))<br />

∼ = HomRF (C) (R, Res ∇ Hom R (M, N))<br />

∼ = HomRF (C) (Res t R, Res ∇ Hom R (M, N))<br />

∼ = HomRC (R, RK t Res ∇ Hom R (M, N)).<br />

Since F (G) is equivalent to G as categories, under Morita<br />

equivalence between RF (G) <strong>and</strong> RG, RK t Res ∇ Hom R (M, N) ∈<br />

RG-mod is exactly the usual RG-module Hom R (M, N). In<br />

Chapter 5, we shall give a higher version <strong>of</strong> this corollary<br />

in terms <strong>of</strong> Ext, based on Hom RC (M, N) = Ext 0 RC(M, N) =<br />

Ext 0 RC e(RC, Hom R(M, N)). The reader should compare the<br />

above isomorphism with Hom RC (M, N) ∼ = Hom RC (R, Hom(M, N)).<br />

We will explain later on that the latter usually does not admit<br />

an higher version using Ext.<br />

Proposition 3.3.11. Let p : C × C op → C be the projection<br />

to the first component. Then LK p RC = R as an RC-module.<br />

Pro<strong>of</strong>. For each x ∈ Ob C one can easily verify that p/x ∼ =<br />

Id C /x × C op . Since Id C /x has an terminal object (x, 1 x ),<br />

the inclusion {(x, 1 x )} → Id C /x has a left adjoint. Thus<br />

the natural functor C op ∼ = • × C op → Id C /x × C op ∼ = p/x<br />

also has a left adjoint. By Proposition 1.2.4 (LK p RC)(x) =<br />

lim RC ∼ = lim<br />

−→p/x −→C opRC. As a functor from C op to R-mod<br />

(RC)(x) = RHom C (x, −) <strong>and</strong> thus we can define a morphism<br />

(RC)(x) → R by ∑ i r iα i ↦→ ∑ i r i so that R fits into the<br />

defining diagram for direct limit. In fact if M fits into the<br />

⊓⊔


3.3 Functors between module categories 113<br />

limit defining diagram<br />

RHom C (y, −)<br />

RHom C (x, −)<br />

θ y<br />

θ x<br />

M ,<br />

then θ y (α) = θ x (β) for any β ∈ Hom C (y, −) <strong>and</strong> α ∈<br />

Hom C (x, −). Thus (LK p RC)(x) = R <strong>and</strong> we are done. ⊓⊔<br />

The last proposition can be generalized to a functor isomorphism<br />

LK p<br />

∼ = − ⊗RC R, <strong>and</strong> we shall come back to it in<br />

Chapter 4.<br />

Note that if we have two functors u : D → C <strong>and</strong> v :<br />

E → D then Res uv = Res v Res u <strong>and</strong> as a consequence<br />

LK uv = LK v LK u . It is enlightening if we summarize our<br />

three calculations in one diagram for t = p∇.<br />

R<br />

RF (C)-mod<br />

LK ∇<br />

RC e -mod<br />

RC<br />

LK t<br />

LK p<br />

∼ =−⊗RC R<br />

RC-mod .<br />

R<br />

Based on explicit formulas for computing LK ∇ <strong>and</strong> LK p ,<br />

one should be able to find examples, similar to Example 3.3.8,<br />

that they do not commute with corresponding tensor products.<br />

Definition 3.3.12. Let C be a small category <strong>and</strong> x an object.<br />

Then we define a full subcategory C x which consists <strong>of</strong><br />

one object x <strong>and</strong> all <strong>of</strong> its endomorphisms. We also define C [x]<br />

to be the full subcategory <strong>of</strong> C containing all objects that are<br />

isomorphic to x.


114 3 Category algebras <strong>and</strong> their representations<br />

Proposition 3.3.13. Let C be a small category <strong>and</strong> x ∈<br />

Ob C. Consider the inclusion ι : C x ↩→ C. Then LK ι [REnd C (x)] ∼ =<br />

RHom C (x, −).<br />

Pro<strong>of</strong>. For every y ∈ Ob C, the overcategory ι/y has objects<br />

{(x, α)} where α runs over the set Hom C (x, y). A morphism<br />

(x, α) → (x, β) is an endomorphism g ∈ End C (x) such that<br />

α = βg. Hence we can fit RHom C (x, y) into the commutative<br />

diagram<br />

[REnd C (x)](x, α)<br />

α◦−<br />

β◦−<br />

[REnd C (x)](x, β)<br />

RHom C (x, y) ,<br />

in which [REnd C (x)](x, α) = [REnd C (x)](x, β) = REnd C (x),<br />

<strong>and</strong> α ◦ −, β ◦ − are compositions. If M is another R-module<br />

that can replace RHom C (x, y) <strong>and</strong> make a new commutative<br />

diagram, then, by examining the images <strong>of</strong> End C (x), one can<br />

easily establish a canonical map RHom C (x, y) → M such<br />

that the whole diagram is commutative. This means that<br />

RHom C (x, y) ∼ = lim −→ι/y<br />

[REnd C (x)] <strong>and</strong> thus LK ι [REnd C (x)] ∼ =<br />

RHom C (x, −).<br />

⊓⊔<br />

3.3.3 Dual modules <strong>and</strong> Kan extensions<br />

In group representations, the two Kan extensions are isomorphic.<br />

Many important properties <strong>of</strong> group representations rely<br />

on this fact. Unfortunately it is not the case for category representations.<br />

To obtain a balanced underst<strong>and</strong>ing <strong>of</strong> the homological<br />

properties <strong>of</strong> category algebras, we describe some<br />

compatibility between the two Kan extensions. Here we assume<br />

the base ring is a field R = k.<br />

Let M ∈ mod-kC be a right module. Its dual M ∧ =<br />

Hom k (M, k) becomes a left kC-module. The functor (−) ∧ =


3.3 Functors between module categories 115<br />

Hom k (−, k) is an anti-isomorphism between the two module<br />

categories. Suppose G is a group <strong>and</strong> H is a subgroup. In<br />

group representation theory, there are two well-known isomorphisms<br />

(M ↓ G H )∧ ∼ = M ∧ ↓ G H , if M ∈ kG-mod, <strong>and</strong><br />

(N ↑ G H )∧ ∼ = N ∧ ↑ G H , if N ∈ kH-mod. We shall extend these<br />

to category representations.<br />

Lemma 3.3.14. Let τ : D → C be a functor between finite<br />

categories. Then we have Res τ M ∧ ∼ = (Res τ M) ∧ , for any<br />

M ∈mod-kC.<br />

Pro<strong>of</strong>. The isomorphism follows from Lemma 3.1.8.<br />

Lemma 3.3.15. With the same assumption as above, we<br />

have RK τ N ∧ ∼ = (LK τ N) ∧ , for any N ∈ mod-kD.<br />

Pro<strong>of</strong>.<br />

RK τ N ∧ ∼ = Hom kC -mod(kC, RK τ N ∧ )<br />

∼ = HomkD-mod(Res τ kC, N ∧ )<br />

∼ = Hommod-kD(N, (Res τ kC) ∧ )<br />

∼ = Hommod-kD(N, Res τ kC ∧ )<br />

∼ = Hommod-kC(LK τ N, kC ∧ )<br />

∼ = HomkC -mod(kC, (LK τ N) ∧ )<br />

∼ = (LKτ N) ∧ .<br />

The last isomorphism can be immediately combined with<br />

computations from the preceding section to get the right Kan<br />

extensions <strong>of</strong> certain modules.<br />

Since this duality exchanges projective <strong>and</strong> injective modules,<br />

whenever the base ring is a field or a complete local ring<br />

if we can characterize the projective modules then the structures<br />

<strong>of</strong> injective modules are understood through duality.<br />

⊓⊔<br />

⊓⊔


116 3 Category algebras <strong>and</strong> their representations<br />

3.4 EI categories, projectives <strong>and</strong> simples<br />

In this section, we investigate the representation theory <strong>of</strong> EIcategories.<br />

It will help us in computing various (co)homology<br />

groups. For technical reasons we always assume in this section<br />

the base ring R is a field or a complete discrete valuation ring,<br />

in order to have the unique decomposition property for every<br />

RC-module.<br />

Some <strong>of</strong> the general theory <strong>of</strong> EI-categories is given by tom<br />

Dieck in [15] (I.11), much <strong>of</strong> which was due to Lück, see also<br />

[51].<br />

3.4.1 EI condition <strong>and</strong> its implications<br />

Recall from Definition 2.4.2 that EI-category is a small category<br />

C in which all endomorphisms are isomorphisms.<br />

One <strong>of</strong> the important features <strong>of</strong> EI-categories is described<br />

as follows. Given an EI-category C, there is a preorder defined<br />

on Ob C, that is, y ≤ x if <strong>and</strong> only if Hom C (y, x) ≠ ∅. Let [y]<br />

be the isomorphism class <strong>of</strong> an object y ∈ Ob C. This preorder<br />

induces a partial order on the set Is C <strong>of</strong> isomorphism classes<br />

<strong>of</strong> Ob C (specified by [y] ≤ [x] if <strong>and</strong> only if Hom C (y, x) ≠<br />

∅), which plays an important role in studying representations<br />

<strong>and</strong> cohomology <strong>of</strong> EI-categories. Because <strong>of</strong> the existence <strong>of</strong><br />

an order for the isomorphism classes <strong>of</strong> objects in any EIcategory,<br />

EI-categories are sometimes referred to as ordered<br />

categories by some authors, see [57], [42].<br />

Definition 3.4.1. For any EI category C <strong>and</strong> any object<br />

x ∈ Ob C, we can define a full subcategory D ≤x ⊂ D consisting<br />

<strong>of</strong> all y ∈ Ob D such that [y] ≤ [x], or equivalently<br />

Hom C (y, x) ≠ ∅. Similarly we can define other full subcategories<br />

<strong>of</strong> D: D x .


3.4 EI categories, projectives <strong>and</strong> simples 117<br />

An object in an EI category C is called maximal if C >x = ∅,<br />

<strong>and</strong> is minimal if C


118 3 Category algebras <strong>and</strong> their representations<br />

<strong>and</strong> otherwise Mˆx (y) = 0. For each minimal x we have a short<br />

exact sequence<br />

0 → Mˆx → M → M/Mˆx → 0<br />

such that M/Mˆx is atomic. In either case, we can repeat the<br />

process for Mˆx . Since C is finite, we will eventually obtain a<br />

filtration <strong>of</strong> M with every factor atomic. We can further refine<br />

this filtration until every factor is simple.<br />

For each RC-module, even if C is not finite, the above analysis<br />

can visualize its structure, which will help us in future<br />

investigations. For example, every simple RC-module has to<br />

be atomic.<br />

3.4.2 Some representation theory<br />

Here we recall some st<strong>and</strong>ard facts in representation theory.<br />

Suppose A is a finite-dimensional algebra over a field k. An A-<br />

module M is indecomposable if M is not a direct sum <strong>of</strong> two<br />

non-trivial submodules. An element e ∈ A is an idempotent if<br />

e ≠ 0 <strong>and</strong> e 2 = e. For example 1 is always an idempotent <strong>and</strong><br />

if e ≠ 1 is an idempotent then so is 1 − e. Two idempotents<br />

e 1 <strong>and</strong> e 2 are orthogonal if e 1 e 2 = e 2 e 1 = 0. An idempotent<br />

e is primitive if one cannot write e = e 1 + e 2 such that both<br />

e 1 <strong>and</strong> e 2 are orthogonal idempotents.<br />

Suppose e ∈ A is a non-identity idempotent. Then e <strong>and</strong><br />

1−e are a pair <strong>of</strong> orthogonal idempotents with 1 = e+(1−e).<br />

It results in a decomposition<br />

A = Ae ⊕ A(1 − e).<br />

The module Ae is projective <strong>and</strong> one can check that it is<br />

indecomposable (not a direct sum <strong>of</strong> two non-trivial modules)<br />

if <strong>and</strong> only if e is primitive.


3.4 EI categories, projectives <strong>and</strong> simples 119<br />

Since A is finite-dimensional, 1 can be decomposed as a sum<br />

<strong>of</strong> finitely many pairwise orthogonal primitive idempotents<br />

(called a primitive decomposition)<br />

1 = e 1 + e 2 + · · · + e n .<br />

Consequently the regular module A becomes a direct sum<br />

A = Ae 1 ⊕ Ae 2 ⊕ · · · ⊕ Ae n .<br />

<strong>and</strong> we obtain a list <strong>of</strong> indecomposable projective A-modules,<br />

Ae i . Up to isomorphism this is a complete list <strong>of</strong> indecomposable<br />

projective A-modules.<br />

In fact if 1 = e ′ 1 + e ′ 2 + · · · + e ′ m is another sum <strong>of</strong> primitive<br />

idempotents in A, then m = n <strong>and</strong> there exists an invertible<br />

element a ∈ A such that 1 = a1a −1 = ae ′ 1a −1 +ae ′ 2a −1 +· · ·+<br />

ae ′ na −1 such that up to an reordering one has e i = ae ′ i a−1<br />

for all possible i. Following from this fact, the automorphism<br />

<strong>of</strong> A, induced by a, maps each direct summ<strong>and</strong> Ae ′ i <strong>of</strong> A =<br />

Ae ′ 1⊕Ae ′ 2⊕· · ·⊕Ae ′ n isomorphically to Ae i . Thus the number<br />

<strong>of</strong> idempotents in a primitive decomposition is equal to the<br />

number <strong>of</strong> indecomposable direct summ<strong>and</strong> <strong>of</strong> A.<br />

Example 3.4.4. Suppose C is a finite category <strong>and</strong> R is a commutative<br />

ring with identity. Then 1 RC is a sum <strong>of</strong> orthogonal<br />

idempotents ∑ x∈Ob C 1 x. Hence we have a decomposition <strong>of</strong><br />

the regular module<br />

RC = ⊕<br />

RC · 1 x = ⊕<br />

RHom C (x, −),<br />

x∈Ob C<br />

x∈Ob C<br />

<strong>and</strong> each summ<strong>and</strong> RHom C (x, −) is a projective RC-module.<br />

However since an idempotent 1 x may be decomposable in RC,<br />

the above decomposition can be refined, in a way depending<br />

on the choice <strong>of</strong> R. If this happens, the corresponding projec-


120 3 Category algebras <strong>and</strong> their representations<br />

tive module RHom C (x, −) is a direct sum <strong>of</strong> other projective<br />

modules. See Example 3.4.7.<br />

In order to describe the simple A-modules, we have to introduce<br />

a few more terminologies. Let J(A) be the Jacobson<br />

radical <strong>of</strong> A. It is the intersection <strong>of</strong> all left maximal ideals <strong>of</strong><br />

A. In particular it is nilpotent in the sense that there exists an<br />

integer n > 0 such that J(A) n = 0. By contrast an element<br />

a ∈ A is nilpotent if a n = 0 for some positive integer n. If an<br />

ideal satisfies the condition that every element is nilpotent,<br />

then this ideal is contained in the radical.<br />

Example 3.4.5. If C is an EI-category, the any non-isomorphism,<br />

as an element in the category algebra RC is nilpotent. Since<br />

they form an ideal <strong>of</strong> RC, it implies that all non-isomorphisms<br />

are contained in J(RC). A important fact is that every element<br />

in J(A) is nilpotent, but the converse if not true (see<br />

Example 3.4.7 (4)).<br />

For each A-module, the radical <strong>of</strong> M, RadM, is the intersection<br />

<strong>of</strong> all maximal submodules <strong>of</strong> M. For example the<br />

regular module has its radical RadA = J(A) because left<br />

ideal <strong>of</strong> A is exactly the same as a left submodule <strong>of</strong> A. In<br />

general RadM = J(A)M. It is easy to see that M/RadM has<br />

a trivial radical. Any A-module with trivial radical is called<br />

semi-simple. A semi-simple module is called simple if it is indecomposable.<br />

Equivalently an A-module is simple if it does<br />

not contain any non-trivial submodule. For example A/RadA<br />

is semi-simple. Every simple A-module occurs as a direct summ<strong>and</strong><br />

in this semi-simple module up to isomorphism.<br />

The quotient A/RadA is itself an algebra with identity ¯1,<br />

the image <strong>of</strong> 1 ∈ A. A pairwise-orthogonal primitive decomposition<br />

1 = e 1 + e 2 + · · · + e n gives rise to ¯1 ∈ A/RadA,<br />

¯1 = ē 1 + ē 2 + · · · + ē n , which is again a sum <strong>of</strong> pairwiseorthogonal<br />

primitive idempotents in A/RadA. This simple


3.4 EI categories, projectives <strong>and</strong> simples 121<br />

observation actually establishes a one-to-one correspondence<br />

between the sets <strong>of</strong> isomorphism classes <strong>of</strong> indecomposable<br />

projective A-modules <strong>and</strong> <strong>of</strong> simple A-modules.<br />

Proposition 3.4.6. Every indecomposable projective A-module,<br />

up to isomorphism, is <strong>of</strong> the form Ae, for some primitive<br />

idempotent e ∈ A. Moreover, Ae/Rad(Ae) is a simple A-<br />

module <strong>and</strong> every simple A-module arises in this way.<br />

Moreover the number <strong>of</strong> idempotents in a primitive decomposition<br />

<strong>of</strong> 1 ∈ A, the number <strong>of</strong> indecomposable summ<strong>and</strong>s<br />

<strong>of</strong> A <strong>and</strong> the number <strong>of</strong> indecomposable summ<strong>and</strong>s<br />

<strong>of</strong> A/Rad(A) equal to each other.<br />

There exists a large collection <strong>of</strong> good references on representation<br />

theory <strong>of</strong> associative algebras. However for those who<br />

do not plan to go over the whole theory, just bear in mind the<br />

basic constructions <strong>and</strong> important facts that we record here.<br />

Then through upcoming examples one can see how they work.<br />

We shall use them to classify projective <strong>and</strong> simple modules<br />

<strong>of</strong> certain finite category algebras. It will be sufficient for us<br />

to develop (co)homology theory <strong>of</strong> categories <strong>and</strong> modules.<br />

Example 3.4.7.1. Let G = {g ∣ g 2 = 1 • } be the cyclic group<br />

<strong>of</strong> order 2, regarded as a category with one object. If k = C<br />

is the field <strong>of</strong> complex numbers, the identity 1 • can be written<br />

as 1 •+g<br />

, a decomposition into a sum <strong>of</strong> orthogonal<br />

primitive idempotents. The regular module is a direct sum<br />

<strong>of</strong> two one dimensional modules CG ∼ = C(1 • +g)⊕C(1 • −g).<br />

Thus both C(1 • + g) <strong>and</strong> C(1 • − g) are projective. They are<br />

simple as well because they cannot have non-trivial submodules.<br />

It means CG is semi-simple with trivial radical. The<br />

module C(1 • +g) is the trivial module <strong>and</strong> C(1 • −g) is called<br />

the sign representation.<br />

However when k is a field <strong>of</strong> characteristic 2, 1 • is primitive.<br />

Hence kG is indecomposable. The regular module has ex-<br />

2<br />

+ 1 •−g<br />

2


122 3 Category algebras <strong>and</strong> their representations<br />

actly one non-trivial submodule k(1 • + g), which has to be<br />

the radical Rad(kG). Then kG/Rad(kG) is one-dimensional<br />

<strong>and</strong> is simple. It is the only simple kG-module, the trivial<br />

module.<br />

2. The poset 1 = 0 → 1 is a category with two objects 0 <strong>and</strong> 1.<br />

For any field k, the identity 1 k1 = 1 0 +1 1 in the category algebra<br />

k1. The two identity morphisms 1 0 <strong>and</strong> 1 1 are primitive<br />

orthogonal idempotents so k1 = k{1 0 , α} ⊕ k1 1 . The first<br />

indecomposable summ<strong>and</strong> has exactly one non-trivial submodule<br />

k{α}, the radical Rad(k{1 0 , α}) <strong>of</strong> k{1 0 , α}. Then<br />

it gives rise to a one-dimensional simple module S 0 . As a<br />

functor, S 0 (0) = k <strong>and</strong> S 0 (1) = 0. The second summ<strong>and</strong> is<br />

<strong>of</strong> dimension one so it is already simple. If we denote it by<br />

S 1 . As a functor S 1 (0) = 0 <strong>and</strong> S 1 (1) = k.<br />

3. Now we examine the category C <strong>of</strong> Example 3.1.2 that is<br />

neither a group nor a poset.<br />

1 x <br />

x<br />

α<br />

1 y<br />

<br />

<br />

y<br />

g<br />

with g 2 = 1 y <strong>and</strong> α = gα. We always have 1 kC = 1 x + 1 y so<br />

kC = k{1 x , α} ⊕ k{1 y , g}. Similar to 2, the first summ<strong>and</strong>,<br />

named P x , is indecomposable <strong>and</strong> has radical k{α}. The<br />

quotient <strong>of</strong> P x by its radical is a one-dimensional simple<br />

module S x (analogues to S 0 as above). According to 1, the<br />

second direct summ<strong>and</strong> is decomposable if k = C. Whence<br />

we have CG ∼ = C(1 • + g) ⊕ C(1 • − g). It means when CC<br />

has three indecomposable projective modules <strong>and</strong> the same<br />

number <strong>of</strong> simple modules.


3.4 EI categories, projectives <strong>and</strong> simples 123<br />

The category algebra CC<br />

Indecomposable projective modules Simples module<br />

P x,1 = C{1 x , α} S x,1 = C{1 x }<br />

P y,1 = C{1 y + g} S y,1 = C{1 y + g}<br />

P y,−1 = C{1 y − g} S y,−1 = C{1 y − g}<br />

When k is <strong>of</strong> characteristic 2, k{1 y , g} is indecomposable.<br />

Whence kC only has two indecomposable projective <strong>and</strong><br />

simple modules.<br />

The category algebra kC, chark = 2<br />

Indecomposable projective modules Simple modules<br />

P x,1 = k{1 x , α} S x,1 = k{1 x }<br />

P y,1 = k{1 y , g} S y,1 = P y,1 /k{1 y + g}<br />

4. A useful example to bear in mind is the following category<br />

(a groupoid) D that is equivalent to •<br />

x<br />

α<br />

α −1<br />

One can write RD = RD · 1 x ⊕ RD · 1 y = R{1 x , α} ⊕<br />

R{1 y , α −1 }. Observe that we have an isomorphism (−) ◦<br />

α −1 : R{1 x , α} → R{1 y , α −1 }. When R = k is a field,<br />

every non-zero element <strong>of</strong> a1 x + bα ∈ kD · 1 x generates the<br />

whole module. Hence kD · 1 x has no non-trivial submodule<br />

<strong>and</strong> thus is simple. Note that it corresponds to the functor<br />

S x,1 which takes values S x,1 (x) = k1 x <strong>and</strong> S x,1 (y) = kα.<br />

Similarly kD · 1 y is also a simple module, corresponding to<br />

S y,1 given by S y,1 (x) = kα −1 <strong>and</strong> S y,1 (y) = k1 y , which is<br />

isomorphic to S x,1 . The algebra RD is semi-simple but it<br />

contains two nilpotent elements α <strong>and</strong> α −1 .<br />

5. Finally let us consider the following category E<br />

y.


124 3 Category algebras <strong>and</strong> their representations<br />

x<br />

α<br />

β<br />

such that αβ = e, βe = β, eα = α <strong>and</strong> e 2 = e. Note that α<br />

<strong>and</strong> β are not invertible in E. Hence E is not a groupoid <strong>and</strong><br />

the two objects are not isomorphic in E. However the category<br />

algebra RE is isomorphic to R{1 x , α, β, e} × R{1 y } ∼ =<br />

RD × R•.<br />

This category is a simple example <strong>of</strong> inverse categories, see<br />

[51]. The category algebra <strong>of</strong> an inverse category is canonically<br />

isomorphic to a direct product <strong>of</strong> groupoid algebras.<br />

Sometimes we may want to use injective modules, so we finish<br />

this section with several remarks concerning the injectives.<br />

In general injective modules behave better than the projectives<br />

in the sense that for any ring A <strong>and</strong> any A-module M,<br />

there exists a minimal injective A-module I M such that M<br />

admits an injection into I M . This module is called the injective<br />

hull <strong>of</strong> M. There are module categories with enough<br />

injective but not with enough projectives. However if A is a<br />

finite-dimensional algebra, then for any M ∈ A-mod there<br />

exists a minimal projective module P M which admits a surjection<br />

onto M. It is called the projective cover <strong>of</strong> M. For<br />

instance, in the tables <strong>of</strong> Example 3.4.7 (3), each row consists<br />

<strong>of</strong> a simple module as well as its projective cover. When we do<br />

representation theory we <strong>of</strong>ten prefer working with projective<br />

modules because simple modules come from their quotients.<br />

As we pointed out earlier one has an anti-isomorphism<br />

(−) ∧ = Hom k (−, k) from A-mod to mod-A. Suppose P is a<br />

projective A-module, then P ∧ is an injective right A-module.<br />

The anti-isomorphism provides a bijection between projective<br />

left (resp. right) A-modules <strong>and</strong> injective right (resp. left) A-<br />

1 y<br />

<br />

<br />

y<br />

e


3.4 EI categories, projectives <strong>and</strong> simples 125<br />

modules. Thus knowing all projectives leads to getting all<br />

injectives. An important case is the group algebra <strong>of</strong> a finite<br />

group G, or more generally a finite-dimensional cocommutative<br />

Hopf algebra. They are self-injectives, which means the<br />

regular module is an injective module. In this case, the sets<br />

<strong>of</strong> projective <strong>and</strong> injective modules coincide.<br />

In the end we mention an important concept in algebra.<br />

Definition 3.4.8. Let A <strong>and</strong> B be two rings. Then A is<br />

Morita equivalent to B if A-mod is equivalent to B-mod.<br />

There are many interesting invariants under a Morita equivalence.<br />

For instance, a Morita equivalence preserves the number<br />

<strong>of</strong> isomorphism classes <strong>of</strong> simple modules. Hochschild<br />

(co)homology is invariant under Morita equivalence. Here we<br />

shall focus on category algebras only. In fact we will prove that<br />

a category equivalence induces a Morita equivalence between<br />

category algebras. Thus in Example 3.4.7 (4) RC is Morita<br />

equivalent to R = R•.<br />

3.4.3 Classifications <strong>of</strong> projectives <strong>and</strong> simples<br />

Now we start describing the projective <strong>and</strong> simple modules <strong>of</strong><br />

an EI category algebra. This part <strong>of</strong> the work is due to Lück<br />

[51], as is described by tom Dieck in [15]. The base ring R is<br />

assumed to be a field or a complete discrete valuation ring.<br />

Let C be a small category <strong>and</strong> x ∈ Ob C an object. Suppose<br />

P x is a projective RC x -module (or in other words a projective<br />

REnd C (x)-module). Then its left Kan extension LK ι P x ,<br />

along ι : C x ↩→ C, is a projective kC-module. Especially<br />

LK ι [REnd C (x)] ∼ = RHom C (x, −) by Proposition 3.2.11.<br />

Let us assume furthermore C is a finite EI category. Then<br />

End C (x) = Aut C (x) for every x ∈ Ob C. From Example 3.4.4<br />

we already learned that RC decomposes into a direct sum


126 3 Category algebras <strong>and</strong> their representations<br />

⊕ x∈Ob C RC · 1 x . Now we try to analyze the indecomposable<br />

direct summ<strong>and</strong>s.<br />

Suppose Is C is the full subcategory <strong>of</strong> C, consisting <strong>of</strong> all<br />

objects <strong>and</strong> all isomorphisms. Then R Is C is a subalgebra <strong>of</strong><br />

RC.<br />

Lemma 3.4.9. If 1 RC = ∑ n<br />

i=1 e i is a primitive decomposition<br />

<strong>of</strong> 1 RC in R Is C, then it is also a primitive decomposition<br />

<strong>of</strong> 1 RC in RC.<br />

Pro<strong>of</strong>. Given any primitive decomposition <strong>of</strong> 1 RC in RC, the<br />

number <strong>of</strong> idempotents in this decomposition is equal to the<br />

number <strong>of</strong> indecomposable direct summ<strong>and</strong>s <strong>of</strong> the regular<br />

module <strong>of</strong> RC, which is equal to the number <strong>of</strong> indecomposable<br />

direct summ<strong>and</strong>s <strong>of</strong> RC/Rad(RC) by Proposition 3.4.5.<br />

Let us take the decomposition 1 RC = ∑ n<br />

i=1 e i. We need to<br />

show it is primitive. To this end, we prove n equals the number<br />

<strong>of</strong> indecomposable direct summ<strong>and</strong>s in RC/Rad(RC).<br />

Since all non-isomorphisms generate an ideal I <strong>of</strong> RC, which<br />

is contained in Rad(RC) <strong>and</strong> which induces an algebra isomorphism<br />

RC/I ∼ = R Is C, from the isomorphism RC/Rad(RC) ∼ =<br />

(RC/I)/(Rad(RC)/I), we know the two sides have the same<br />

numbers <strong>of</strong> indecomposable direct summ<strong>and</strong>s. From definition<br />

one can check that Rad(RC)/I is the radical <strong>of</strong> RC/I. Then<br />

by Proposition 3.4.5, applied to both RC <strong>and</strong> RC/I ∼ = R Is C,<br />

we see the primitive decompositions <strong>of</strong> 1 RC in both RC <strong>and</strong><br />

R Is C must have the same number <strong>of</strong> idempotents. Hence we<br />

are done.<br />

⊓⊔<br />

The category Is C is a disjoint union <strong>of</strong> groupoids, each <strong>of</strong><br />

which comes from an isomorphism class <strong>of</strong> some object. Recall<br />

that, for each object x ∈ Ob C, we denote by [x] the set <strong>of</strong> objects<br />

isomorphic to x, <strong>and</strong> C [x] the full subcategory consisting<br />

<strong>of</strong> these objects.


3.4 EI categories, projectives <strong>and</strong> simples 127<br />

Lemma 3.4.10.1. If x ∼ = y are two isomorphic objects, <strong>and</strong><br />

f y ∈ Hom C (x, y) is an isomorphism, then the assignment<br />

α ↦→ α · f y for each α ∈ RC · e defines an isomorphism <strong>of</strong><br />

RC-modules RC · 1 y → RC · 1 x .<br />

2. If 1 x = ∑ n<br />

i=1 e i is a primitive decomposition in RAut C (x),<br />

then 1 y = ∑ n<br />

i=1 f ye i fy<br />

−1 is a primitive decomposition in<br />

RAut C (y). Furthermore if we fix for each y ∼ = x an isomorphism<br />

f y ∈ Hom C (x, y), then<br />

∑ n∑<br />

f y e i fy<br />

−1<br />

y∈Ob C [x]<br />

i=1<br />

is a primitive decomposition <strong>of</strong> the identity 1 RC[x] in the<br />

groupoid algebra RC [x] .<br />

Pro<strong>of</strong>. The isomorphism is straightforward to prove. Now if<br />

1 x = ∑ n<br />

i=1 e i is a primitive decomposition in RAut C (x), certainly<br />

1 y = ∑ n<br />

i=1 f ye i fy<br />

−1 is a decomposition in RAut C (y). It<br />

has to be primitive, because if it were not, then a primitive<br />

decomposition would be a sum <strong>of</strong> more than n idempotents<br />

in RAut C (y). However fy<br />

−1 (−)f y maps such a primitive decomposition<br />

<strong>of</strong> 1 y to a decomposition <strong>of</strong> 1 x , which contradicts<br />

with the assumption that 1 x = ∑ n<br />

i=1 e i is a primitive decomposition.<br />

⊓⊔<br />

The reader can compare the above statements with Example<br />

3.4.7 (4).<br />

Corollary 3.4.11. Let C be a finite EI category. One can<br />

write<br />

1 RC = ∑ ∑n x<br />

e xj ,<br />

x∈Ob C<br />

∑where n x is a positive integer for each x ∈ Ob C <strong>and</strong> 1 x =<br />

nx<br />

j=1 e xj. As a consequence, RC = ⊕ x∈Ob C ⊕ n x<br />

j=1 RC · e xj<br />

j=1


128 3 Category algebras <strong>and</strong> their representations<br />

for some primitive pairwise orthogonal idempotents e xj ∈<br />

RAut C (x), x ∈ Ob C.<br />

Any projective RC-module is isomorphic to a direct sum <strong>of</strong><br />

indecomposable projective modules <strong>of</strong> the form RC · e, where<br />

e ∈ RAut C (x) is a primitive idempotent, for some x ∈ Ob C.<br />

Given a primitive orthogonal decomposition 1 RC = ∑ i e i<br />

such that each e i belongs to some group algebra RAut C (x),<br />

each summ<strong>and</strong> <strong>of</strong> RC ∼ = ⊕ e i<br />

RC · e i is indeed a left Kan<br />

extension<br />

LK ι [RAut C (x)e i ] = {LK ι [RAut C (x)]}e i<br />

∼ = RHomC (x, −)e i = RCe i<br />

because LK ι commutes with direct sums. Here ι : C x ↩→ C is<br />

the inclusion. In each RCe i , its radical contains ∑ z ≁ =x<br />

RHom x,z e i<br />

which are linear combinations <strong>of</strong> non-isomorphisms in Hom C (x, −).<br />

We continue to characterize the simple RC-modules. Directly<br />

from the EI condition we have seen that a simple module<br />

S has to be atomic. It matches with our description <strong>of</strong><br />

indecomposable projective modules. The quotient <strong>of</strong> RCe i ,<br />

for a primitive idempotent e i ∈ RAut C (x), by its radical is an<br />

atomic module supported on C [x] . Moreover for each y ∼ = x,<br />

S(y) must be a simple RAut C (y)-module. In fact all simple<br />

RC-modules are exactly those simple modules <strong>of</strong> R Is C, which<br />

are obtained from simple modules <strong>of</strong> automorphism group algebras<br />

RAut C (x).<br />

Theorem 3.4.12 (Lück). Let C be a finite EI-category.<br />

The isomorphism classes <strong>of</strong> the simple RC-modules biject<br />

with the pairs ([x], V ), where x ∈ Ob C <strong>and</strong> V is a simple<br />

RAut C (x)-module, taken up to isomorphism.<br />

Pro<strong>of</strong>. First <strong>of</strong> all, we already know that all simple RCmodules<br />

are atomic. Thus simple RC-modules are exactly<br />

those simple RC [x] -modules, with x running over Ob C.


3.4 EI categories, projectives <strong>and</strong> simples 129<br />

Secondly for a fixed x, RC [x] -mod is equivalent to RAut C (y),<br />

for any y ∼ = x, through restrictions induced by the equivalences<br />

C y → C [x] <strong>and</strong> C [x] → C y . Hence simple RC [x] -modules<br />

biject with simple RC y -modules, for any y ∼ = x. Since C y is<br />

the group Aut C (y), we have proved the assertion. ⊓⊔<br />

Because <strong>of</strong> the above theorem, it is natural to denote a simple<br />

RC-module by S x,V , if it comes from a simple RAut C (x)-<br />

module V , for some x ∈ Ob C. For consistency, we use P x,V for<br />

the projective cover <strong>of</strong> S x,V , whose structure is determined by<br />

its value at the object x. If RAut C (x)·e is the projective cover<br />

<strong>of</strong> the simple RAut C (x)-module V , then RC · e is the projective<br />

cover <strong>of</strong> S x,V . The reader may revisit Example 3.4.7 to<br />

get better underst<strong>and</strong>ing <strong>of</strong> our results <strong>and</strong> notations in this<br />

section.<br />

Example 3.4.13. Let k be a field <strong>of</strong> characteristic two <strong>and</strong> C<br />

the following category<br />

{1 x } <br />

x<br />

α<br />

β<br />

y<br />

{1 y ,g}<br />

with g 2 = 1 y , gα = α <strong>and</strong> gβ = β. Indeed the algebra kC has<br />

two (one-dimensional) simples S x,k , S y,k <strong>and</strong> their projective<br />

covers are P x,k = k{1 x , α, β}, <strong>and</strong> P y,k = k{1 y , g}, respectively.<br />

The product P x,k ˆ⊗P x,k<br />

∼ = Px,k ⊕ Sy,k 2 is not projective<br />

because S y,k ≠ P y,k .<br />

Remark 3.4.14. Using the tensor product, one can introduce<br />

a “representation ring” <strong>of</strong> RC, namely a(RC), which consists<br />

<strong>of</strong> Z-linear combinations <strong>of</strong> symbols like [M], representing<br />

an isomorphism class <strong>of</strong> a simple RC-module M. For any<br />

two elements [M] <strong>and</strong> [N], the multiplication is defined by<br />

[M] · [N] = [M ˆ⊗N]. However this product does not exist in<br />

K 0 (RC), which is spanned over the set <strong>of</strong> isomorphism classes<br />

<strong>of</strong> indecomposable projectives.


130 3 Category algebras <strong>and</strong> their representations<br />

With the description <strong>of</strong> indecomposable projectives, we can<br />

show when the trivial module R is projective.<br />

Proposition 3.4.15. Let C be a finite EI-category. Then R<br />

is projective if <strong>and</strong> only if each connected component <strong>of</strong> C<br />

has a unique isomorphism class <strong>of</strong> minimal objects [x], with<br />

the properties that for all y in the same connected component<br />

as x, Aut C (x) acts transitively on Hom(x, y), <strong>and</strong><br />

|Aut C (x)| is invertible in R.<br />

Pro<strong>of</strong>. Without loss <strong>of</strong> generality, we may assume C is connected.<br />

If R is projective then R ∼ = ⊕ P y,W for certain indecomposable<br />

projective modules P y,W . Since R is indecomposable<br />

<strong>and</strong> takes constant value at all objects, we must have that<br />

R ∼ = P x,V for some x ∈ Ob C, x is minimal, <strong>and</strong> all minimal<br />

objects are isomorphic. Moreover because P x,V (x) = R,<br />

the projective cover <strong>of</strong> the simple kAut C (x)-module V , we get<br />

V = R <strong>and</strong> R must be projective as an RAut C (x)-module.<br />

Thus R is projective if <strong>and</strong> only if R ∼ = P x,R , all minimal<br />

objects are isomorphic to x <strong>and</strong> |Aut C (x)| −1 ∈ R.<br />

Now, assume all minimal objects are isomorphic to x <strong>and</strong><br />

|Aut C (x)| −1 ∈ R. By Proposition 3.2.11 RHom C (x, −) ∼ =<br />

LK ι [RAut C (x)]. Then R ∼ = P x,R can<br />

∑<br />

be explicitly constructed,<br />

1<br />

using the idempotent e =<br />

|Aut C (x)| g∈Aut C (x)<br />

g, as<br />

R ∼ = P x,R<br />

∼ = LKι [RAut C (x)e] = {LK ι [RAut C (x)]}e = RHom C (x, −)e<br />

because LK ι commutes with direct sum. It is equivalent to<br />

saying that at each y ∈ Ob C, R ∼ = RHom C (x, y)e. This happens<br />

if <strong>and</strong> only if Aut C (x) to act transitively on Hom C (x, y).<br />

⊓⊔


3.4 EI categories, projectives <strong>and</strong> simples 131<br />

3.4.4 Projective covers, injective hulls <strong>and</strong> their restrictions<br />

Let C be a finite category <strong>and</strong> R = k a field. Then any finitely<br />

generated kC-module M admits a minimal projective resolution<br />

P ∗ → M → 0<br />

in the sense that if P ′ ∗ → M → 0 is another projective resolution<br />

<strong>of</strong> M, then Id M induces a split injection <strong>of</strong> complexes<br />

from the minimal resolution to the latter. Similarly we can<br />

define a minimal injective resolution <strong>of</strong> M, 0 → M → I ∗ .<br />

Proposition 3.4.16. Suppose C is an EI category.<br />

1. If D ⊂ C is a coideal <strong>and</strong> P ∈ kC-mod is an indecomposable<br />

projective module, then Res ι P ∈ kD-mod is either an<br />

indecomposable projective or zero.<br />

2. If D ⊂ C is an ideal <strong>and</strong> I ∈ kC-mod is an indecomposable<br />

injective module, then Res ι I ∈ kD-mod is either an<br />

indecomposable injective or zero.<br />

Pro<strong>of</strong>. Let P = P x,V be an indecomposable projective kCmodule.<br />

If x ∈ Ob C, then Res i P x,V is brutally truncated from<br />

P x,V <strong>and</strong> is indecomposable projective as an kD-module. On<br />

the other h<strong>and</strong>, if x ∉ Ob D, then Res i P x,V = 0.<br />

For the case <strong>of</strong> injective modules we recall (Res i P ) ∧ ∼ =<br />

Res i P ∧ for any right projective module, by Lemma 3.2.12.<br />

Note that Statement 1 stays true for right projective modules<br />

if we replace the term “coideal” by “ideal”. Now we combine<br />

this with the duality between (indecomposable) right projectives<br />

<strong>and</strong> left injectives.<br />

⊓⊔<br />

For example if x is an object in C, then C ≤x is a coideal. Take<br />

any indecomposable projective module P x,V , then Res i P x,V<br />

∼ =<br />

P V , the projective cover <strong>of</strong> the simple kAut C (x)-module V .<br />

This module P V is an indecomposable projective kC ≤x -module.


132 3 Category algebras <strong>and</strong> their representations<br />

Definition 3.4.17. Let C be an EI category. Suppose M is a<br />

kC-module. Then we define C M ⊂ C to be the smallest ideal<br />

satisfying the condition that if x ∉ Ob C M then M(x) = 0.<br />

Suppose M is a kC-module. Then we define C M ⊂ C to be<br />

the smallest coideal satisfying the condition that if x ∉ Ob C M<br />

then M(x) = 0.<br />

Suppose (M, N) is an ordered pair <strong>of</strong> kC-modules. We define<br />

a full subcategory C N M to be C M ∩ C N .<br />

Obviously if N ⊂ M then C N ⊂ C M <strong>and</strong> C N ⊂ C M .<br />

Lemma 3.4.18. Let C be a finite EI category <strong>and</strong> M a kCmodule.<br />

Then the projective cover P M <strong>of</strong> M satisfies the<br />

condition that C PM ⊂ C M , <strong>and</strong> the injective hull I M <strong>of</strong> M<br />

satisfies the condition that C I M<br />

⊂ C M .<br />

This result allows us to give a characterization <strong>of</strong> the minimal<br />

projective <strong>and</strong> injective resolutions <strong>of</strong> a module.<br />

Corollary 3.4.19. Let C be a finite EI category <strong>and</strong> M ∈<br />

kC-mod. Suppose P ∗ <strong>and</strong> I ∗ are minimal projective <strong>and</strong> injective<br />

resolutions <strong>of</strong> M. Then for every n ≥ 0, C Pn ⊂ C M<br />

<strong>and</strong> C I n<br />

⊂ C M .<br />

Pro<strong>of</strong>. The kernel K 0 <strong>of</strong> P 0 ↠ M satisfies C K0 ⊂ C P0 ⊂ C M .<br />

We use the preceding lemma repeatedly. Similar argument<br />

can be made on I ∗ .<br />

⊓⊔<br />

The last corollary will be useful when we compute cohomology<br />

<strong>of</strong> modules.


Chapter 4<br />

<strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

Abstract We begin through investigation <strong>of</strong> (co)homology<br />

theories in this chapter. Extensions <strong>of</strong> modules over a category<br />

algebra is a concept <strong>of</strong> paramount importance here. We<br />

shall discuss various ways to examine Ext groups, multiplicative<br />

structure <strong>and</strong> their relationship with previously defined<br />

simplicial <strong>and</strong> singular (co)homology. A particular important<br />

situation is when the first module is trivial. In this case, on<br />

top <strong>of</strong> the module theoretic tools, simplicial methods are applicable.<br />

We shall provide a discussion <strong>of</strong> the bar resolution<br />

<strong>and</strong> its Kan extensions. Examples are used to illustrate various<br />

computational methods. In the end, the Grothendieck<br />

spectral sequences are introduced <strong>and</strong> we will study them in<br />

a couple <strong>of</strong> special cases.<br />

4.1 General theory<br />

4.1.1 <strong>Cohomology</strong> <strong>of</strong> modules<br />

Since RC-mod is an abelian category with enough projectives<br />

<strong>and</strong> injectives, for any two modules M, N ∈ RC-mod<br />

we can consider the Ext groups Ext i RC(M, N), with i ≥ 0.<br />

It is the i-th right derived functor <strong>of</strong> the left exact functor<br />

Hom RC (M, −) (or Hom RC (−, N)). In general for any<br />

M ∈ RC-mod Ext ∗ RC(M, M) has a ring structure with product<br />

given by the Yoneda splice. Usually it is not graded com-


134 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

mutative. However it is in the case <strong>of</strong> M = R. In this chapter<br />

we shall compare the Yoneda splice with the cup product<br />

introduced earlier.<br />

Similarly for a right RC-module M ′ <strong>and</strong> a left RC-module<br />

N ′ , we can study Tor RC<br />

i (M ′ , N ′ ) as the i-th right derived functors<br />

<strong>of</strong> M ′ ⊗ RC − (or − ⊗ RC N ′ ). In these notes we shall focus<br />

on cohomology. However we shall remark on homology whenever<br />

it is appropriate.<br />

We recall some basics about extensions <strong>of</strong> modules <strong>and</strong> their<br />

relationship with cohomology classes.<br />

Definition 4.1.1. Let M, N be A-modules. An n-fold extension<br />

, n ≥ 1, <strong>of</strong> M by N is an exact sequence <strong>of</strong> A-modules<br />

0 → N → L n−1 → · · · → L 0 → M → 0.<br />

Two n-fold extensions <strong>of</strong> M by N are equivalent if there is<br />

a commutative diagram<br />

0<br />

<br />

N L<br />

n−1 · · ·<br />

<br />

L<br />

0 M 0<br />

0<br />

<br />

N L ′ n−1<br />

<br />

· · ·<br />

<br />

L ′ 0<br />

<br />

M 0<br />

Then we can extend this by symmetry <strong>and</strong> transitivity to<br />

an equivalence relation among n-fold extensions <strong>of</strong> M by N.<br />

Proposition 4.1.2. There is an one-to-one correspondence<br />

between elements <strong>of</strong> Ext n A(M, N) <strong>and</strong> equivalent classes <strong>of</strong><br />

n-fold extensions <strong>of</strong> M by N.<br />

Pro<strong>of</strong>. Let P ∗ → M → 0 be a projective resolution. Then<br />

an extension determines an element in Ext n RC(M, N) by the<br />

following lifting


4.1 General theory 135<br />

· · ·<br />

<br />

P n+1<br />

<br />

P n<br />

∂ n <br />

P n−1<br />

· · ·<br />

<br />

P<br />

0 M 0<br />

f<br />

<br />

0<br />

<br />

N L n−1<br />

<br />

· · ·<br />

<br />

L 0<br />

<br />

M 0<br />

We see from here that two equivalent extensions give rise to<br />

the same element in Ext n A(M, N).<br />

Conversely if two n-fold extensions determine the same elements<br />

in the group Ext n A(M, N), then we can construct a<br />

commutative diagram (by enlarging P ∗ we may assume f is<br />

surjective)<br />

0 N<br />

<br />

L n−1<br />

<br />

L n−2<br />

<br />

· · ·<br />

<br />

L 0<br />

<br />

M 0<br />

0 N P n−1 /∂ n (Kerf)<br />

<br />

0 N<br />

<br />

L ′ n−1<br />

P n−2<br />

<br />

L ′ n−2<br />

· · ·<br />

<br />

P<br />

0 M 0<br />

<br />

· · ·<br />

<br />

L ′ 0<br />

<br />

M 0<br />

Definition 4.1.3. Let M ∈ A-mod. Then we can define the<br />

Yoneda splice on Ext ∗ A(M ′ , M) <strong>and</strong> Ext ∗ A(M, M ′′ ) so that<br />

for any η ∈ Ext i A(M ′ , M) <strong>and</strong> η ′ ∈ Ext j A (M, M ′′ ), η ′ ∗ η ∈<br />

Ext i+j<br />

A<br />

(M ′ , M ′′ ) is given by<br />

0 → M ′′ → N j−1 · · · → N 0 → N ′ i−1 → · · · → N ′ 0 → M ′ → 0,<br />

if η is represented by 0 → M → N ′ i−1 · · · → N ′ 0 → M ′ → 0<br />

<strong>and</strong> η ′ is represented by 0 → M ′′ → N j−1 · · · → N 0 → M →<br />

0.<br />

The Yoneda splice gives Ext ∗ A(M, M) a ring structure which<br />

is not graded commutative in general. Moreover Ext ∗ A(M, M)<br />

acts on Ext ∗ A(M, N) <strong>and</strong> Ext ∗ A(N, M) for another N ∈ A-<br />

mod.<br />

Definition 4.1.4. We call Ext ∗ RC(R, R) = ⊕ i≥0 Exti RC(R, R)<br />

the ordinary cohomology ring <strong>of</strong> the category algebra A. The<br />

⊓⊔


136 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

product in this ring is defined by the Yoneda splice <strong>of</strong> Ext<br />

classes.<br />

We shall show later on that this ring is isomorphic in a<br />

natural way to H ∗ (C, R) ∼ = H ∗ (BC, R) so it deserves the name.<br />

Now based on the tensor structure on RC-mod, we provide<br />

a module theoretic description to the ring Ext ∗ RC(R, R).<br />

In the meantime we pave the road to studying the ring action<br />

<strong>of</strong> Ext ∗ RC(R, R) on various Ext groups. We comment that<br />

since (RC-mod, ˆ⊗, R) is a monoidal category with an exact<br />

tensor product, it gives rise to a suspended monoidal category<br />

(D − (RC), ˆ⊗, R) <strong>and</strong> then following a general statement<br />

[71] on the endomorphisms <strong>of</strong> the identity in a suspended<br />

monoidal category, End D − (RC)(R) is a graded commutative<br />

ring. It will be clear in this section that this endomorphism<br />

ring is isomorphic to what we call the ordinary cohomology<br />

ring Ext ∗ RC(R, R).<br />

Let M, M ′ , N, N ′ ∈ RC-mod which are projective as R-<br />

modules. We will define the cup product to be<br />

∪ : Ext i RC(M, N)⊗Ext j RC (M ′ , N ′ ) → Ext i+j<br />

RC (M ˆ⊗M ′ , N ˆ⊗N ′ ).<br />

Since R is the identity with respect to ˆ⊗, this will give us<br />

a ring structure on Ext ∗ RC(R, R), as well as an action <strong>of</strong><br />

Ext ∗ RC(R, R) on Ext ∗ RC(M, N) for arbitrary M, N ∈ RC-mod.<br />

One shall compare our construction with [, Section 3.2] for<br />

cocommutative Hopf algebras.<br />

Now we are ready to give a precise definition to the cup<br />

product. Suppose ζ ∈ Ext m RC(M, N) is represented by an exact<br />

sequence<br />

0 → N → L m−1 → P m−2 → · · · → P 0 → M → 0,<br />

where P i ’s are projective RC-modules, <strong>and</strong> ζ ′ ∈ Ext n RC(M ′ , N ′ )<br />

is represented by an exact sequence


4.1 General theory 137<br />

0 → N ′ → L ′ n−1 → Q n−2 → · · · → Q 0 → M ′ → 0,<br />

where Q j ’s are projective RC-modules. Since projective RCmodules<br />

are projective R-modules, all modules in these two<br />

exact sequences are projective R-modules. Furthermore an<br />

RC-module L being a projective R-module is equivalent to<br />

L(x) being projective for all x ∈ Ob C. Then applying the<br />

Künnethe formula to the following two complexes <strong>of</strong> projective<br />

R-modules<br />

<strong>and</strong><br />

0 → N(x) → L m−1 (x) → P m−2 (x) → · · · → P 0 (x)<br />

0 → N ′ (x) → L ′ n−1(x) → Q n−2 (x) → · · · → Q 0 (x),<br />

for all x ∈ Ob C, we get exact sequences<br />

0 → (N ˆ⊗N ′ )(x) → (L m−1 ˆ⊗N)(x) ⊕ (N ˆ⊗L ′ n−1)(x) → · · ·<br />

→ (P 0 ˆ⊗Q 0 )(x) → (M ˆ⊗M ′ )(x) → 0,<br />

with x running over Ob C. Thus we get an exact sequence <strong>of</strong><br />

RC-modules<br />

0 → N ˆ⊗N ′ → (L m−1 ˆ⊗N ′ ) ⊕ (N ˆ⊗L ′ n−1) → · · · → P 0 ˆ⊗Q 0 → M ˆ⊗M<br />

which is defined to be the cup product <strong>of</strong> ζ <strong>and</strong> ζ ′ ,<br />

ζ ∪ ζ ′ ∈ Ext m+n<br />

RC (M ˆ⊗M ′ , N ˆ⊗N ′ ).<br />

Lemma 4.1.5. Let ζ, ζ ′ be as above. The cup product ζ ∪ ζ ′<br />

is the Yoneda splice <strong>of</strong><br />

with<br />

ζ ˆ⊗N ′ ∈ Ext i RC(M ˆ⊗N ′ , N ˆ⊗N ′ )<br />

M ˆ⊗ζ ′ ∈ Ext j RC (M ˆ⊗M ′ , M ˆ⊗N ′ ).


138 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

There exists a natural map φ M = − ˆ⊗M : Ext ∗ RC(R, R) →<br />

Ext ∗ RC(M, M) lies in the graded center, for any M ∈ RCmod.<br />

Particularly, Ext ∗ RC(R, R) is graded commutative.<br />

Pro<strong>of</strong>. By using exact sequences representing ζ <strong>and</strong> ζ ′ , one<br />

can easily establish a map between (n + m)-fold extensions<br />

(ζ ˆ⊗Id N ′) ∗ (Id M ˆ⊗ζ ′ ) → ζ ∪ ζ ′ . Let<br />

<strong>and</strong><br />

D : 0 → N → P m−1 = L m−1 → P m−2 → · · · → P 0<br />

C : 0 → N ′ → Q n−1 = L ′ n−1 → Q n−2 → · · · → Q 0<br />

come from the given two extensions. Then we have a map <strong>of</strong><br />

(m + n)-fold extensions<br />

0 N ˆ⊗N ′ <br />

(P m−1 ˆ⊗N ′ ) ⊕ (N ˆ⊗Q n−1 )<br />

<br />

· · ·<br />

<br />

P 0 ˆ⊗Q<br />

0 M ˆ⊗M ′ <br />

0<br />

f m+n−1 <br />

0 N ˆ⊗N ′ <br />

P m−1 ˆ⊗N ′ <br />

· · ·<br />

<br />

M ˆ⊗Q 0<br />

<br />

M ˆ⊗M ′ <br />

0<br />

given by<br />

{<br />

(D ˆ⊗C)<br />

f i :<br />

i → P 0 ˆ⊗Q i → M ˆ⊗Q i , 0 ≤ i ≤ n − 1;<br />

(D ˆ⊗C) i → P i−n ˆ⊗N ′ , n ≤ i ≤ n + m − 1.<br />

Thus cup product is a Yondea splice.<br />

Next for an m-fold extension (all modules chosen to be R-<br />

projective)<br />

ζ : 0 → R → L m−1 → P m−2 → · · · → P 0 → R → 0,<br />

the tensor product with M<br />

0 → M → L m−1 ˆ⊗M → P m−2 ˆ⊗M → · · · → P 0 ˆ⊗M → M → 0,<br />

stays exact because<br />

0 → R = R(x) → L m−1 (x) → P m−2 (x) → · · · → P 0 (x) → R(x) = R<br />

f 0


4.1 General theory 139<br />

is split exact <strong>and</strong> thus<br />

0 → M(x) → L m−1 (x)⊗M(x) → P m−2 (x)⊗M(x) → · · · → P 0 (x)⊗M<br />

is exact for all x ∈ Ob C. Thus − ˆ⊗M gives us a map, commuting<br />

with Yoneda splice,<br />

φ M = − ˆ⊗M : Ext ∗ RC(R, R) → Ext ∗ RC(M, M),<br />

<strong>and</strong> hence is a ring homomorphism.<br />

If ζ ′ ∈ Ext n RC(M, M), then φ M (ζ) ∗ ζ ′ = ζ ∪ ζ ′ <strong>and</strong> ζ ′ ∗<br />

φ M (ζ) = ζ ′ ∪ ζ. We want to show ζ ∪ ζ ′ = (−1) mn ζ ′ ∪ ζ.<br />

This comes from the fact that, given the cocommutativity<br />

τ∆ = ∆, we can establish an isomorphism <strong>of</strong> complexes <strong>of</strong><br />

RC-modules<br />

C ˆ⊗D → D ˆ⊗C,<br />

by a ⊗ b ↦→ (−1) deg a deg b b ⊗ a for any two homogeneous elements.<br />

⊓⊔<br />

In terms <strong>of</strong> projective resolutions, we can describe the cup<br />

product as follows. Let M, M ′ , N, N ′ be RC-modules which<br />

are projective as R-modules. Take two projective resolutions<br />

P ∗ → M → 0 <strong>and</strong> Q ∗ → M ′ → 0. Then by our previous observation,<br />

P ∗ ˆ⊗Q ∗ → M ˆ⊗M ′ → 0 is an exact sequence. This<br />

is usually not a projective resolution as the tensor product <strong>of</strong><br />

two projective is not projective, in contrast to the case <strong>of</strong> a cocommutative<br />

Hopf algebra. However we can build a projective<br />

resolution, unique up to chain homotopy, R ∗ → M ˆ⊗M ′ → 0<br />

such that there exists a chain map δ : R ∗ → P ∗ ˆ⊗Q ∗ <strong>and</strong> a<br />

commutative diagram<br />

R ∗<br />

δ <br />

M ˆ⊗M ′ <br />

0<br />

P ∗ ˆ⊗Q ∗<br />

<br />

M ˆ⊗M ′ <br />

0.


140 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

If for two integers m, n, ζ ∈ Ext m RC(M, N) <strong>and</strong> ζ ′ ∈ Ext n RC(M ′ , N ′ )<br />

are represented by two cocycles f : P m → N <strong>and</strong> g : Q n →<br />

N ′ , then the product ζ ∪ ζ ′ is represented by (f ˆ⊗g) ◦ δ :<br />

R m+n → N ˆ⊗N ′ . In Theorem 4.1.13 we shall show how to establish<br />

the algebra isomorphism Ext ∗ RC(R, R) ∼ = H ∗ (BC, R).<br />

4.1.2 <strong>Cohomology</strong> <strong>of</strong> a small category with coefficients in a functor<br />

In this section we introduce a particular important case, the<br />

Baues-Wirsching cohomology theory <strong>of</strong> small categories, <strong>and</strong><br />

go over some basic properties. The cohomology theory <strong>of</strong> small<br />

categories has been discussed in various places in literature,<br />

see Baues-Wirsching [3], Generalov [29] <strong>and</strong> Oliver []. One<br />

can also find in Gabriel-Zisman [28] <strong>and</strong> Hilton-Stammbach<br />

[35] the homology theory <strong>of</strong> small categories.<br />

Definition 4.1.6. Let C be a small category <strong>and</strong> R a commutative<br />

ring. We define H n (C; M), the n-th cohomology <strong>of</strong><br />

C with coefficients in a covariant functor M : C → R-<br />

mod, as the homology <strong>of</strong> the following cochain complex<br />

{C i (C; M), δ i } i≥0 , where<br />

∏<br />

C i (C; M) = {f : NC i →<br />

M(x i ) ∣ f([x 0 → · · · → x i ]) ∈ M<br />

[x 0 →x 1 →···→x i ]<br />

for all i ≥ 0; <strong>and</strong> for f ∈ C i (C; M),<br />

δ n φ<br />

i∑<br />

(f)([x 0 → · · · → x i→xi+1 ]) = (−1) j f([x 0 → · · · → ˆx j → · · · →<br />

+(−1) i+1 M(φ)(f([x 0 → · · · → x i ]))<br />

We define Hn(C; M), the n-th homology <strong>of</strong> C with coefficients<br />

in a covariant functor M : C → R-mod, as the homology <strong>of</strong><br />

the following chain complex {C i (C; M), δ i } i≥0 , where<br />

j=0


4.1 General theory 141<br />

C i (C; M) =<br />

⊕<br />

[x 0 →x 1 →···→x i ]<br />

M(x 0 ),<br />

for all i ≥ 0; <strong>and</strong> for any m [x0 →x 1 →···→x i ] ∈ C i (C; M) which<br />

is an element in some m ∈ M(x 0 ) indexed by x 0 → · · · →<br />

ˆx j → · · · → x i ,<br />

δ i ( ⊕ m [x0<br />

φ<br />

→x1 →···→x i−1 →x i ] ) = [M(φ)(m)] [x 1 →···→x i−1 ]<br />

+<br />

i∑<br />

(−1) j m [x0 →···→ ˆx j →···→x i ],<br />

j=0<br />

in which m [x0 →···→ ˆx j →···→x i ] means m ∈ M(x 0 ) is considered<br />

as an element <strong>of</strong> C i−1 (C; M) indexed by the (i − 1)-sequence<br />

x 0 → · · · → ˆx j → · · · → x i .<br />

We now give different interpretations to the above homology<br />

<strong>and</strong> cohomology theory. They will give us better underst<strong>and</strong>ing<br />

<strong>of</strong> the preceding definitions <strong>and</strong> then lead us to a more general<br />

cohomology theory. Recall that for any functor u : D → C<br />

between two small categories, we can naturally define two<br />

functors u/− : C → Cat, the category <strong>of</strong> small categories,<br />

<strong>and</strong> B(u/−) : C → T op, the category <strong>of</strong> topological spaces.<br />

Thus for any u : D → C, we may write C ∗ (u/x, R), x ∈ Ob C<br />

as the chain complex coming from the simplicial set associated<br />

to the small category u/x, which can be regarded as the<br />

cellular chain complex on the space B(u/x).<br />

Definition 4.1.7. Let C be a small category. We define the<br />

bar resolution <strong>of</strong> R ∈ RC-mod as B∗<br />

C = C ∗ (Id C /−, R), a<br />

complex <strong>of</strong> RC-modules.<br />

The bar resolution is a complex reconstructed from C ∗ (C, R)<br />

<strong>and</strong> its name is justified as follows.


142 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

Proposition 4.1.8. For any small category C, B C ∗ is a complex<br />

<strong>of</strong> projective RC-modules such that<br />

1. B C 0 ∼ = RC;<br />

2. B C ∗ → R → 0 is an exact sequence <strong>of</strong> RC-modules;<br />

3. There is a degree 1 isomorphism <strong>of</strong> complexes <strong>of</strong> R-modules<br />

from<br />

C ∗ (Id C /−, R) → R → 0<br />

to R[NC] ∗ = C ∗ (C, R) → 0.<br />

4. Hom RC (B C ∗, M) ∼ = ∏ x 0 →···→x i<br />

M(x i ).<br />

Remark 4.1.9. We emphasize that there is no conflict between<br />

2 <strong>and</strong> 3. A complex <strong>of</strong> RC-modules is exact if <strong>and</strong> only if it is<br />

pointwise exact.<br />

Pro<strong>of</strong>. For each i ≥ 0, C i (Id C /−, R) : C → R-mod is a well<br />

defined functor <strong>and</strong> thus an RC-module. There is an RCmap<br />

∂ i : C i+1 (Id C /−, R) → C i (Id C /−, R) as follows. For any<br />

x ∈ Ob C,<br />

∂ i ((x 0 , α 0 ) → · · · → (x j , α j ) → · · · → (x i+1 , α i+1 ))<br />

= ∑ i+1<br />

j=0 (−1)j [(x 0 , α 0 ) → · · · → (x ̂ j , α j ) → · · · → (x i+1 , α i+1 )],<br />

where α j ∈ Hom C (x j , x). Thus for each x ∈ Ob C, C ∗ (Id C /x, R) →<br />

R → 0 is the augmented chain complex <strong>of</strong> Id C /x. When we assemble<br />

these augmented chain complexes together, C ∗ (Id C /−, R) →<br />

R → 0 becomes a complex <strong>of</strong> RC-modules. Since every Id C /x<br />

has a terminal object (x, 1 x ) <strong>and</strong> thus is contractible, the complex<br />

C ∗ (Id C /−, R) → R → 0 is acyclic because its evaluation<br />

at any x, i.e. C ∗ (Id C /x, R) → R → 0 computes<br />

the reduced homology ˜H∗ (Id C /x, R) <strong>and</strong> is acyclic. Moreover,<br />

we can show every C i (Id C /−, R) is projective. Indeed<br />

for each i > 0, C i (Id C /x, R) has base elements <strong>of</strong> the form<br />

(x 0 , α 0 ) → · · · → (x i , α i ). The following bijection provides a<br />

different presentation <strong>of</strong> these base elements


4.1 General theory 143<br />

{(x 0 , α 0 ) β 1<br />

→ · · · →(x βi<br />

α i β i···β 1 α i β i α<br />

i , α i )} {x 0 −→ · · ·<br />

i<br />

−→xi →x}.<br />

This bijection in fact induces an isomorphism <strong>of</strong> complexes <strong>of</strong><br />

R-modules (see Definitions 1.2.5 <strong>and</strong> 2.2.4)<br />

C i (Id C /−, R) ∼ =<br />

−→R[NC] i+1 = C i+1 (C, R), ∀i ≥ 0.<br />

We get immediately C 0 (Id C /−, R) ∼ = C 1 (C, R) ∼ = RC. In general<br />

let us take the new R-basis <strong>of</strong> C i (Id C /−, R), {x 0 → · · · →<br />

x i → x}, <strong>and</strong> regroup them by putting two such sequences<br />

into one subset <strong>of</strong> the basis if, after removing the rightmost<br />

object, they become identical. Then each subset determines<br />

an RC-module<br />

R{x 0 → · · · → x i<br />

α<br />

→−<br />

∣ ∣ s(α) = x i },<br />

isomorphic to the projective module RHom C (x i , −). It means<br />

as RC-modules<br />

C i (Id C /−, R) ∼ ⊕<br />

=<br />

x 0 →x 1 →···→x i<br />

RHom C (x i , −).<br />

Hence B∗<br />

C → R → 0 is a projective resolution. In the end<br />

for any M ∈ RC-mod, there is an isomorphism by Yoneda<br />

lemma<br />

Hom RC (C i (Id C /−, R), M) ∼ =<br />

∏<br />

x 0 →···→x i<br />

M(x i ).<br />

We remind the reader that R can be defined as R Ob C ∼ =<br />

R[NC] 0<br />

∼ = C0 (C, R). Thus we have a degree one isomorphism<br />

<strong>of</strong> chain complexes (<strong>of</strong> R-modules) from C ∗ (Id C /−, R) →<br />

R → 0 to R[NC] ∗ = C ∗ (C, R) → 0 if in the first complex<br />

we insert R in degree -1.<br />

⊓⊔<br />

Example 4.1.10. In Example 2.2.6 (1) we considered the following<br />

category with two objects, two identity <strong>and</strong> non-


144 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

identity morphisms<br />

x<br />

α<br />

β<br />

Its nerve gives rise to a non-normalized chain complex C ∗ (C, R)<br />

which is <strong>of</strong> rank 2 in degree zero, rank 4 in degree 1 <strong>and</strong> rank<br />

2n + 2 at degree n ≥ 2. It is certainly not exact.<br />

The bar resolution B C ∗ can be written down by computing<br />

the two overcategories Id C /x <strong>and</strong> Id C /y. In fact Id C /x is a<br />

trivial category with one object (x, 1 x ), <strong>and</strong> Id C /y is a poset<br />

with three objects (x, α), (x, β) <strong>and</strong> (y, 1 y ).<br />

(x, α)<br />

α<br />

(y, 1 y )<br />

y<br />

β<br />

(x, β)<br />

Hence the bar resolution is given by B C ∗(y) <strong>and</strong> B C ∗(x). The<br />

former has rank 3 in degree zero, rank 5 in degree 1, <strong>and</strong><br />

rank 2n + 3 in degree n ≥ 2; while the latter has rank 1 at<br />

every degree. Thus B C ∗ has rank 4 in degree zero, rank 6 in<br />

degree 1 <strong>and</strong> rank 2n + 4 in degree n ≥ 2. One can see that<br />

Rank(B C i ) =Rank(C i+1 (C, R)) for all i ≥ 0. This equality is<br />

actually induced by the R-isomorphism we explained in the<br />

pro<strong>of</strong> <strong>of</strong> Proposition 3.1.3. Note that both B C ∗(y) <strong>and</strong> B C ∗(x)<br />

are exact because they come from the nerves <strong>of</strong> the two overcategories<br />

Id C /y <strong>and</strong> Id C /x which are contractible. Furthermore<br />

we can explicitly verify that B C 0 ∼ = RC, B C 1 ∼ = P x,1 ⊕P 3 y,1,<br />

etc.<br />

If one examines the normalized complexes <strong>of</strong> C ∗ (C, R),<br />

C ∗ (Id C /y, R) <strong>and</strong> C ∗ (Id C /x, R) then the above equality between<br />

ranks are not true any more.<br />

Remark 4.1.11. Proposition 4.1.8 summarizes to a characterization<br />

<strong>of</strong> the bar resolution in terms <strong>of</strong> the nerve, as well as


4.1 General theory 145<br />

an isomorphism<br />

Hom RC (B C i , M) ∼ =<br />

∏<br />

x 0 →···→x i<br />

M(x i ).<br />

By contrast, when we consider R as a right RC-module. The<br />

bar resolution <strong>of</strong> it is C ∗ (−\Id C , R). We shall also denote it<br />

by B∗ C when it does not cause any confusion. For each i ≥ 0,<br />

C i (−\Id C , R) ∼ ⊕<br />

=<br />

x 0 →x 1 →···→x i<br />

RHom C (−, x 0 ).<br />

Since RHom C (−, x 0 ) = 1 x0 · RC, we can verify that<br />

Bi C ⊗ RC M ∼ ⊕<br />

=<br />

x 0 →x 1 →···→x i<br />

M(x 0 ).<br />

When we examine the special case for M = R, we get isomorphisms<br />

<strong>of</strong> complexes<br />

Hom RC (Bi C , R) ∼ =<br />

∏<br />

<strong>and</strong><br />

B C i ⊗ RC R ∼ =<br />

x 0 →···→x i<br />

R ∼ = Hom R (R[NC] i , R).<br />

⊕<br />

x 0 →···→x i<br />

R ∼ = R[NC] i .<br />

Proposition 4.1.12. Let C be a small category <strong>and</strong> R a<br />

commutative ring. If M : C → R-mod is a functor, then<br />

lim<br />

−→ n M ∼ = Tor RC<br />

C n (R, M) ∼ = Hn(C; M),<br />

<strong>and</strong><br />

lim<br />

←− n M ∼ = Ext n C RC(R, M) ∼ = H n (C; M).<br />

Pro<strong>of</strong>. Let U be an arbitrary R-module. Then we have<br />

Hom R (U, Hom RC (R, M)) ∼ = Hom RC (R⊗ R U, M) ∼ = Hom RC (U, M),


146 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

where U ∈ RC-mod is a constant functor. Hence by Proposition<br />

1.1.23, ←−C<br />

lim M ∼ = Hom RC (R, M).<br />

Similarly<br />

Hom R (R⊗ RC M, U) ∼ = Hom RC (M, Hom R (R, U)) ∼ = Hom RC (M, U),<br />

<strong>and</strong> consequently R ⊗ RC M ∼ = lim −→C<br />

M.<br />

We will only prove the isomorphisms for cohomology because<br />

the statements for homology can be obtained similarly.<br />

Since the kernel <strong>of</strong> the differential δ 0 : C 0 (C; M) → C 1 (C; M)<br />

can be identified with Hom RC (R, M), we get H 0 (C; M) ∼ =<br />

Hom RC (R, M).<br />

n<br />

Since lim ←−C<br />

is left exact <strong>and</strong> ←−<br />

lim is its n-th right derived<br />

C<br />

n<br />

functor, we have ←−<br />

lim M ∼ = Ext n C RC(R, M). Hence we only have<br />

to show Ext n RC(R, M) ∼ = H n (C; M). Because Ext n RC(R, M) can<br />

be computed by using any projective resolution <strong>of</strong> R in RCmod,<br />

we can use the bar resolution to do it. But we have an<br />

isomorphism for each i ≥ 0<br />

Hom RC (Bi C , M) ∼ =<br />

∏<br />

x 0 →···→x i<br />

M(x i ) ∼ = C i (C; M).<br />

Suppose C = G is a group <strong>and</strong> M, N are two left RGmodules,<br />

since a left RG-module can be naturally regarded as<br />

a right RG-module (<strong>and</strong> vice versa), it makes sense to consider<br />

both Tor RG<br />

∗ (M, N) <strong>and</strong> Tor RG<br />

∗ (N, M). Indeed we always have<br />

Tor RG<br />

∗ (M, N) ∼ = Tor RG<br />

∗ (N, M), see [12, Chapter 3]. It is not<br />

the case in general for category homology.<br />

We deduce now that the preceding homology <strong>and</strong> cohomology<br />

theories are the same as the simplicial <strong>and</strong> singular homology<br />

<strong>and</strong> cohomology theories we introduced in Chapter<br />

2.<br />

⊓⊔


4.1 General theory 147<br />

Theorem 4.1.13. We have isomorphisms<br />

<strong>and</strong><br />

Tor RC<br />

∗ (R, R) ∼ = H∗(C; R) ∼ = H∗(C, R) ∼ = H∗(BC, R)<br />

Ext n RC(R, R) ∼ = H ∗ (C; R) ∼ = H ∗ (C, R) ∼ = H ∗ (BC, R).<br />

The latter are algebra isomorphisms.<br />

Pro<strong>of</strong>. In order to prove the isomorphisms between graded R-<br />

modules, we only need to show that H∗(C; R) ∼ = H∗(C, R) <strong>and</strong><br />

H ∗ (C; R) ∼ = H ∗ (C, R). But these follow from Remark 4.1.11<br />

which compare Definitions 2.2.4 with 4.1.6.<br />

In order to prove the isomorphism between cohomology rings<br />

is an algebra isomorphism we just have to compare the cup<br />

products. Here we only need to explain the first algebra isomorphism<br />

because the second is clear.<br />

Using the bar resolution B∗<br />

C → R → 0, we can describe<br />

the cup product on Ext ∗ RC(R, R). In fact, we can explicitly<br />

write out a diagonal approximation map (unique up to chain<br />

homotopy)<br />

R 0<br />

B C ∗<br />

D <br />

B∗ C ˆ⊗B∗<br />

C <br />

R 0<br />

as a natural transformation, given by<br />

α<br />

D x (x<br />

1<br />

n∑<br />

0→x1 → · · · →x αn α α<br />

n→x) = (x<br />

1<br />

α···α i+1<br />

0→ · · · → xi<br />

i=0<br />

→ x)⊗(xi<br />

α i+1<br />

for any x ∈ Ob C <strong>and</strong> integer n. Since for each n ≥ 0,<br />

(B C ∗ ˆ⊗B C ∗) n = ⊕ i+j=n BC i ˆ⊗B C j <strong>and</strong> there is a natural map<br />

Hom RC (B C i , R) ⊗ Hom RC (B C j , R) → Hom RC (B C i ˆ⊗B C j , R),<br />

→ · · ·<br />

α<br />


148 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

one can easily see, under the isomorphism Hom RC (B C ∗, R) ∼ =<br />

Hom R (R[NC] ∗ , R), that the diagonal approximation map D<br />

induces a map<br />

Hom RC (B C ∗, R) ⊗ Hom RC (B C ∗, R) → Hom RC (B C ∗, R)<br />

identical to the following<br />

Hom R (R[NC] ∗ , R)⊗Hom R (R[NC] ∗ , R) → Hom R (R[NC] ∗ , R)<br />

induced by the Alex<strong>and</strong>er-Whitney map. Since these two<br />

maps are used to calculate cup products in the two cohomology<br />

rings, our observations imply that Ext ∗ RC(R, R) ∼ =<br />

H ∗ (BC, R) as algebras.<br />

⊓⊔<br />

Recall that the bar resolution can also be constructed via the<br />

nerve <strong>of</strong> overcategories associated with the identity functor<br />

Id C : C → C, as B∗ C ∼ = C ∗ (Id C /−, R) := RN ∗ (Id C /−). In this<br />

form, Bn(x) C ∼ = C n (Id C /x), for each x ∈ Ob C <strong>and</strong> integer<br />

n ≥ 0, consists <strong>of</strong> the following chains as base elements<br />

(x 0 , β 0 ) γ 1<br />

→(x 1 , β 1 ) → · · · → (x n−1 , β n−1 ) γ n<br />

→(x n , β n ),<br />

in which β i is a morphism in Hom C (x i , x), <strong>and</strong> γ i ∈ Hom C (x i−1 , x i )<br />

such that β i−1 = β i γ i−1 . The previously defined diagonal approximation<br />

map D is given by<br />

D x ((x 0 , β 0 ) γ 1<br />

→(x 1 , β 1 ) → · · · → (x n−1 , β n−1 ) γ n<br />

→(x n , β n ))<br />

= ∑ n<br />

i=0 [(x 0, β 0 ) γ 1<br />

→ · · · →(x γi<br />

i , β i )] ⊗ [(x i , β i ) γ i+1<br />

→ · · · →(x γn<br />

n , β n )]<br />

This is exactly the Alex<strong>and</strong>er-Whitney map for the chain<br />

complex coming from the nerve <strong>of</strong> Id C /x.<br />

All homology <strong>and</strong> cohomology theories <strong>of</strong> small categories<br />

that we have introduced so far coincide whenever they are<br />

comparable. Hence we shall only deal with the most general<br />

form Ext ∗ RC(M, N) <strong>and</strong> Tor RC<br />

∗ (M ′ , N) from now on, where<br />

M, N ∈ RC-mod <strong>and</strong> M ′ ∈ mod-RC.


4.1 General theory 149<br />

Proposition 4.1.14. Suppose u : D → C is a functor. Then<br />

we have a restriction Res u : RC-mod → RD-mod. Given two<br />

RC-modules M, N, Res u induces a natural map<br />

res u : Ext ∗ RC(M, N) → Ext ∗ RD(Res u M, Res u N),<br />

called the restriction.<br />

Pro<strong>of</strong>. Take a projective resolution P ∗ <strong>of</strong> M. Then Res u P ∗ →<br />

Res u M → 0 is an exact sequence <strong>of</strong> RD-modules. Hence<br />

for any projective resolution Q ∗ → Res u M → 0, the usual<br />

lifting, a chain map, Q ∗ → Res u P ∗ induces a cochain map<br />

Hom RC (P ∗ , N) → Hom RD (Q ∗ , Res u N) because we have a<br />

cochain map Hom RC (P ∗ , N) → Hom RD (Res u P ∗ , Res u N). This<br />

is a well defined map since it does not depend on the choice<br />

<strong>of</strong> P ∗ <strong>and</strong> Q ∗ .<br />

⊓⊔<br />

When M = N = R, by using bar resolutions, one can easily<br />

see that the above restriction is the same as the one we defined<br />

for simplicial cohomology in Chapter 2.<br />

Remark 4.1.15. Suppose u : D → C is a functor. Let us consider<br />

res u : Ext ∗ RC(R, N) → Ext ∗ RD(R, Res u N).<br />

One way to obtain some information about the restriction is to<br />

examine its action in degree zero. In fact, the restriction is the<br />

same as Hom RC (R, N) → Hom RC (LK u R, N) ∼= →Hom RD (R, Res u N)<br />

induced by the counit LK u R → R. When LK u R ∼ = R, res u<br />

becomes an isomorphism in degree zero. Giving LK u R ∼ = R<br />

is equivalent to saying that every overcategory <strong>of</strong> u is R-<br />

acyclic, that is, ˜H∗ (u/x, R) vanishes for every x ∈ Ob C. In<br />

other words, every overcategory u/x has to be connected, see<br />

Proposition 3.2.7.


150 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

4.1.3 Extensions <strong>of</strong> categories <strong>and</strong> low dimension cohomology<br />

We describe low dimension cohomology groups Ext n RC(M, N)<br />

for n = 1, 2. The results <strong>and</strong> their pro<strong>of</strong>s are taken from<br />

Webb’s notes [80]. Let us recall that there is an augmentation<br />

map ɛ : RC → R by sending α to t(α), its target. The kernel<br />

<strong>of</strong> ɛ is a left ideal <strong>of</strong> RC, which we call the (left) augmentation<br />

ideal, <strong>and</strong> write as IC := Kerɛ.<br />

Lemma 4.1.16. The augmentation ideal IC is a free R-<br />

module with basis the elements α − 1 t(α) , where α runs over<br />

all the non-identity morphisms in C.<br />

Definition 4.1.17. Let M ∈ RC-mod. We define a derivation<br />

d : C → M to be a mapping from Mor C to M so that<br />

d(α) ∈ M(t(α)) <strong>and</strong> so that d(αβ) = M(α)d(β) + d(α). The<br />

set <strong>of</strong> derivations forms an R-module Der(C, M).<br />

Given any set <strong>of</strong> elements {u x ∈ M(x) ∣ ∣x ∈ Ob C} we obtain<br />

a derivation specified by d(α) = M(α)(u s(α) ) − u t(α) . Any<br />

derivation obtained in this way is called an inner derivation.<br />

The inner derivations form an R-module IDer(C, M).<br />

Proposition 4.1.18.1. Der(C, M) ∼ = Hom RC (IC, M) as R-<br />

modules.<br />

2. H 1 (C; M) ∼ = Der(C, M)/IDer(C, M).<br />

Pro<strong>of</strong>. Given a homomorphism δ : IC → M, we can define<br />

a derivation d : C → M by d(α) = δ(α − 1 t(α) ). It makes<br />

sense because M(α)d(β) + d(α) = δ[(α(β − 1 t(β) )] + d(α) =<br />

δ(αβ − α) + d(α) = δ(αβ − 1 t(αβ) )) = d(αβ) since t(αβ) =<br />

t(α). Conversely given a derivation d we can define a RChomomorphism<br />

δ : IC → M by δ(α − 1 t(α) ) = d(α), <strong>and</strong> we<br />

can verify that δ is an RC-map. Hence we have proved (1).<br />

As for (2), consider the short exact sequence 0 → IC →<br />

RC → R → 0 <strong>and</strong> apply Ext ∗ RC(−, M). We obtain an exact


4.1 General theory 151<br />

sequence<br />

0 → Hom RC (R, M) → Hom RC (RC, M) → Hom RC (IC, M) → Ext 1 RC(R<br />

The image <strong>of</strong> Hom RC (RC, M) ∼ = M consists <strong>of</strong> f m | IC , where<br />

f m is the RC-homomorphism determined by some m ∈ M via<br />

the canonical map m ↦→ f m such that f m (1 RC ) = m. Because<br />

f m (α − 1 t(α) ) = [M(α)](m) − [M(1 t(α) )](m) = [M(α)](1 s(α) ·<br />

m) − [M(1 t(α) )](1 t(α) · m), f m | IC is an inner derivation determined<br />

by the set <strong>of</strong> elements {1 x · m ∈ M(x) ∣ ∣ x ∈ Ob C}.<br />

Moreover all inner derivations are <strong>of</strong> this form. We are done.<br />

⊓⊔<br />

In order to characterize H 2 (C; K), we have to introduce category<br />

extensions in the sense <strong>of</strong> G. H<strong>of</strong>f.<br />

Definition 4.1.19 (H<strong>of</strong>f). An extension E <strong>of</strong> a category C<br />

via a category K is a sequence <strong>of</strong> functors<br />

K ι<br />

−→E π<br />

−→C,<br />

which has the following properties:<br />

1. Ob K = Ob E = Ob C, ι is injective <strong>and</strong> π is surjective on<br />

morphisms; <strong>and</strong><br />

2. if π(α) = π(β), for two morphisms α, β ∈ Mor(E), if <strong>and</strong><br />

only if there is a unique g ∈ Mor(K) such that β = ι(g)α.<br />

We may readily deduce some useful information from the<br />

above definition.<br />

Proposition 4.1.20.3. If αι(h) exists for α ∈ Mor(E) <strong>and</strong><br />

h ∈ Mor(K), then there exists a unique h α ∈ Mor(K) such<br />

that ι(h α )α = αι(h). Moreover any α ∈ Hom E (x, y) induces<br />

a group homomorphism K(x) → K(y); <strong>and</strong><br />

4. for any α ∈ Hom C (x, y), K(y) acts regularly (i.e. freely<br />

<strong>and</strong> transitively) on π −1 (α).


152 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

Pro<strong>of</strong>. Suppose αι(h) exists for a given α ∈ Mor E. Then<br />

since π(αι(h)) = π(α), by Definition 4.1.19 (2), there exists<br />

a unique h α ∈ Mor K such that αι(h) = ι(h α )α. If h 1 , h 2 ∈<br />

K(x), then we have αι(h 1 h 2 ) = ι[(h 1 h 2 ) α ]α. On the other<br />

h<strong>and</strong>, we also get αι(h 1 h 2 ) = αι(h 1 )ι(h 2 ) = ι(h 1 ) α [αι(h 2 )] =<br />

ι(h α 1)ι(h α 2)α. Hence h α 1h α 2 = (h 1 h 2 ) α <strong>and</strong> α induces a group<br />

homomorphism.<br />

Now assume β ∈ Hom E (x, y) <strong>and</strong> π(β) = α. Let h ∈ K(y).<br />

If ι(h)β = β, then we have π(ι(h)β) = π(β) = α. Definition<br />

4.1.19 (2) forces h = 1 y because <strong>of</strong> the uniqueness property.<br />

⊓⊔<br />

One can see that K is indeed a disjoint union <strong>of</strong> the groups<br />

π −1 (1 x ) for all 1 x ∈ Mor(C) (regarded as categories), <strong>and</strong> can<br />

be identified with a functor K : E → Groups. Usually from<br />

the context, one knows when we take K to be a category <strong>and</strong><br />

when it is regarded as a functor.<br />

Since in practice one <strong>of</strong>ten crosses a concept dual to the<br />

category extension, for future reference, we first state its definition.<br />

Definition 4.1.21. An opposite extension E <strong>of</strong> C via K is<br />

a sequence <strong>of</strong> functors K ι →E π →C such that the following sequence<br />

is an extension <strong>of</strong> C op<br />

K op ι op<br />

−→E op π op<br />

−→C op ,<br />

Sometimes we just say K → E → C is an opposite extension.<br />

Dually there are characterizations as follows.<br />

1 op . Ob K = Ob E = Ob C, ι is injective <strong>and</strong> π is surjective on<br />

morphisms; <strong>and</strong><br />

2 op . if π(α) = π(β), for two morphisms α, β ∈ Mor(E), if <strong>and</strong><br />

only if there is a unique g ∈ Mor(K) such that β = αι(g).


4.1 General theory 153<br />

3 op . If ι(h)α exists for α ∈ Mor(E) <strong>and</strong> h ∈ Mor(K), then there<br />

exists a unique h ′ ∈ Mor(K) such that αι(h ′ ) = ι(h)α; <strong>and</strong><br />

4 op . for any α ∈ Hom C (x, y), K(x) acts regularly on π −1 (α).<br />

One <strong>of</strong> the main difference between extensions <strong>and</strong> opposite<br />

extensions is that, for an extension, K can be considered as<br />

a covariant functor from C to Groups while, for an opposite<br />

extension, K gives a contra-variant functor. When we study<br />

H 1 <strong>and</strong> H 2 , extensions are used because we deal with left modules.<br />

If dually we want to obtain corresponding statements for<br />

right modules, then we have to replace extensions by opposite<br />

extensions in wherever it applies.<br />

Example 4.1.22.1. Any category C is an extension <strong>and</strong> opposite<br />

extension <strong>of</strong> itself by ⊎ Ob C<br />

•, a disjoint union <strong>of</strong> some<br />

trivial groups.<br />

2. Any extension <strong>of</strong> a group is both an extension <strong>and</strong> an opposite<br />

extension if we consider the group <strong>and</strong> its extension as<br />

categories.<br />

3. The following sequence is an extension <strong>of</strong> the rightmost category<br />

by a group which is the disjoint union <strong>of</strong> a trivial group<br />

<strong>and</strong> a cyclic group <strong>of</strong> order 2:<br />

x<br />

C 2<br />

<br />

{1 y }<br />

y<br />

C 2<br />

<br />

y<br />

ι π<br />

α<br />

<br />

x<br />

C 2<br />

<br />

x<br />

{1 x }<br />

α<br />

C 2<br />

<br />

y<br />

.<br />

4. The following sequence is an opposite extension <strong>of</strong> the rightmost<br />

category by a group which is the disjoint union <strong>of</strong> a<br />

trivial group <strong>and</strong> a cyclic group <strong>of</strong> order 2:


154 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

C 2<br />

<br />

y<br />

C 2<br />

<br />

y<br />

{1 y }<br />

y<br />

x<br />

{1 x }<br />

ι π<br />

α<br />

<br />

x<br />

C 2<br />

<br />

α<br />

x .<br />

C 2<br />

One shall easily see the difference <strong>and</strong> relationship between<br />

the first two examples. We note that in both examples, the<br />

categories at the two ends have finitely generated cohomology<br />

rings, while the ones in the middle do not, see Example<br />

2.2.4 (4). <strong>and</strong> Proposition 2.2.38.<br />

5. Let G be a finite group <strong>and</strong> p a prime dividing the order <strong>of</strong> G.<br />

A collection P <strong>of</strong> p-subgroups <strong>of</strong> G is a set <strong>of</strong> p-subgroups<br />

which is closed under conjugations in G. The transporter<br />

category G ∝ P = Gr G P has various quotient categories,<br />

e.g. the orbit category O P (G), many <strong>of</strong> which form extension<br />

sequences, e.g.<br />

K s → G ∝ P → O P (G),<br />

where K s (H) = H for any H ∈ Ob P = Ob(G ∝ P). We<br />

shall come back to this in Chapter 6.<br />

Definition 4.1.23. Two extensions are equivalent if we have<br />

the following commutative diagram<br />

K E<br />

<br />

C<br />

K E ′ <br />

C<br />

Definition 4.1.24. An extension is split if it admits a functor<br />

s : C → E such that π ◦ s = 1 C .<br />

Proposition 4.1.25. An extension is split if <strong>and</strong> only if it<br />

is equivalent to a Grothendieck construction.


4.1 General theory 155<br />

Pro<strong>of</strong>. Assume an extension K → E → C is split. Then<br />

K can be restricted to a functor C → E → Groups. The<br />

Grothendieck construction Gr C K has the same objects <strong>of</strong> the<br />

form (x, • x ) where x ∈ Ob C <strong>and</strong> • x is the only object <strong>of</strong><br />

K(x). A morphism from (x, • x ) to (y, • y ) is a pair (α, h) with<br />

α ∈ Hom C (x, y) <strong>and</strong> h ∈ K(y). Hence we can define a funtor<br />

Gr C K → E by (x, • x ) ↦→ x <strong>and</strong> (α, h) ↦→ hα.<br />

On the other h<strong>and</strong> for an extension K → Gr C K → C, we can<br />

define a functor C → Gr C K by x ↦→ (x, • x ) <strong>and</strong> α ↦→ (α, 1 y )<br />

if α ∈ Hom C (x, y). Thus this extension is split.<br />

⊓⊔<br />

Proposition 4.1.26. The equivalence classes <strong>of</strong> extensions<br />

<strong>of</strong> C by K are in bijection with H 2 (C; K) in such a way that<br />

the zero element corresponds to the Grothendieck construction<br />

Gr C K.<br />

Pro<strong>of</strong>. We will provide correspondence between equivalence<br />

classes <strong>of</strong> extensions <strong>and</strong> elements <strong>of</strong> H 2 (C; K).<br />

Let K → E → C be an extension. We can choose a section for<br />

E → C, that is, for each morphism α : x → y in C a morphism<br />

q(α) : x → y in E with pq(α) = α. If α : x → y <strong>and</strong> β : y → z<br />

are two morphisms in C, then q(βα) = ι[τ(β, α)]q(β)q(α) for<br />

a unique τ(β, α) ∈ K(z). Thus we obtain a mapping τ :<br />

C × C → C. The associativity <strong>of</strong> morphisms implies the 2-<br />

cocycle condition by direct calculation<br />

τ(γβ, α) + τ(γ, β) = τ(γ, βα) + K(γ)τ(β, α).<br />

If f ∈ Hom RC (RC, M), then mappings <strong>of</strong> the form<br />

˜f(β, α) := f(βα) − f(β) − K(β)f(α)<br />

automatically satisfies the 2-cocycle condition. If we examine<br />

Hom RC (B C ∗, M), used to define H ∗ (C; M) in Definition 3.1.1,<br />

we can verify that all these mappings τ <strong>and</strong> ˜f biject with<br />

the 2-cocycles <strong>and</strong> 2-coboudaries in the the cochain complex


156 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

Hom RC (B∗, C M). In fact we already learned that<br />

Hom RC (B2, C M) ∼ =<br />

∏<br />

K(z)<br />

<strong>and</strong><br />

x α →y β →z<br />

Hom RC (B C 1, M) ∼ = ∏ v γ →w<br />

K(w).<br />

With explicitly given differential we can prove the above bijection.<br />

In the end, we may verify that 2-cocycles are homologous<br />

if <strong>and</strong> only if the extensions which produce them are equivalent,<br />

<strong>and</strong> also a change in the choice <strong>of</strong> a section gives rise to<br />

a cohomologous 2-cocycle. Hence we are done.<br />

⊓⊔<br />

To finish this section, we comment on the connection between<br />

H i (C; M) <strong>and</strong> Ext i RC(R, M), for i = 1, 2. In fact from<br />

the short exact sequence 0 → IC → RC → R → 0, we get<br />

Ext i RC(R, M) ∼ = Ext i−1<br />

RC (IC, M).<br />

Let 0 → M → E → IC → 0 be an RC-module extension<br />

representing an element in Ext 1 RC(IC, M). We can construct<br />

an extension <strong>of</strong> categories<br />

M → Gr C E → Gr C IC.<br />

There is a splitting functor C → Gr C IC, which is identity on<br />

objects <strong>and</strong> sends a morphism α to (α − 1 t(α) , α). Now we<br />

form a pullback diagram<br />

M<br />

<br />

E<br />

<br />

C<br />

M Gr C E Gr C IC


4.2 Classical methods for computation 157<br />

which gives rise to an extension M → E → C, an element in<br />

H 2 (C; M). Note that if 0 → M → E → IC → 0 splits, then<br />

Gr C E ∼ = Gr GrC ICM, in H 1 (C; M).<br />

Conversely if M → E → C is an extension <strong>and</strong> IM is the<br />

kernel <strong>of</strong> the algebra homomorphism RE → RC. Then we can<br />

define an extension <strong>of</strong> RC-modules<br />

0 → M → IE/(IM · IE) → IC → 0,<br />

representing an element in Ext 2 RC(R, M) ∼ = Ext 1 RC(IC, M).<br />

4.2 Classical methods for computation<br />

4.2.1 Minimal resolutions <strong>and</strong> reduction<br />

Recall in Section 3.3.3 we described the minimal projective<br />

<strong>and</strong> injective resolutions <strong>of</strong> a module. With Definition 3.3.13<br />

we establish the following isomorphism.<br />

Proposition 4.2.1. Let C be an EI category, M, N two kCmodules<br />

<strong>and</strong> M ′ a right kC-module. Then<br />

<strong>and</strong><br />

Ext ∗ kC(M, N) ∼ = Ext ∗ kC N M(Res i M, Res i N)<br />

Tor kC<br />

∗ (M ′ , N) ∼ = Tor ∗ (Res<br />

kC M′ i M ′ , Res i N).<br />

Pro<strong>of</strong>. Suppose P ∗ is a minimal projective resolution <strong>of</strong> M.<br />

Then each direct summ<strong>and</strong> P x,V <strong>of</strong> any P n must satisfy<br />

the condition that x ∈ Ob C M by Corollary 3.3.15. Hence<br />

Hom kC (M, N) ∼ = Hom kCM (Res i M, Res i N) <strong>and</strong> Ext ∗ kC(M, N) ∼ =<br />

Ext ∗ kC M<br />

(Res i M, Res i N). Now we apply Corollary 3.3.15 again<br />

to a minimal injective resolution <strong>of</strong> N when computing<br />

<strong>and</strong> the isomorphism follows.<br />

Ext ∗ kC M<br />

(Res i M, Res i N)<br />

N


158 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

To prove the isomorphism for Tor the method is the same.<br />

One just need to take the right projective resolution <strong>of</strong> M ′ .<br />

⊓⊔<br />

This result is useful because one may replace the original<br />

category by a full subcategory before carrying on any computations.<br />

It is particular helpful if both M <strong>and</strong> N are atomic.<br />

Since any module <strong>of</strong> an EI category algebra has a filtration by<br />

atomic modules. Repeatedly using the above corollary may<br />

significantly simplify things. In following examples we will<br />

shall how one combines resolutions, module filtrations <strong>and</strong><br />

reductions through Proposition 4.2.1 to compute various Ext<br />

groups.<br />

Example 4.2.2.1. Suppose C = n is a poset. Then k ∼ =<br />

kHom n (0, −) is a projective module. Thus Ext ∗ kn(k, M) ∼ =<br />

M(0) for any M ∈ kn-mod.<br />

2. Suppose C is the following category<br />

x<br />

α<br />

β<br />

y.<br />

We will only write down non-vanishing Ext groups in this<br />

example.<br />

The minimal projective resolution <strong>of</strong> k is 0 → P y,k →<br />

P x,k → k → 0. Thus for any M ∈ kC-mod Ext ∗ kC(k, M)<br />

is given by the homology <strong>of</strong> 0 → Hom kC (P x,k , M) →<br />

Hom kC (P y,k , M) → 0. If M = k, then Ext 0 kC(k, k) ∼ =<br />

Ext 1 kC(k, k) ∼ = k. If M = S x,k , then Ext 0 kC(k, S x,k ) ∼ = k.<br />

If M = S y,k , then Ext 1 kC(k, S y,k ) ∼ = k.<br />

The minimal projective resolution <strong>of</strong> S x,k is 0 → Py,k 2 →<br />

P x,k → S x,k → 0. We can use it for direct calculation, but<br />

we want to make use <strong>of</strong> Proposition 4.2.1. Ext ∗ kC(S x,k , S x,k ) ∼ =<br />

Ext ∗ k{x}(Res i S x,k , Res i S x,k ) ∼ = Ext ∗ k(k, k). Hence Ext 0 kC(S x,k , S x,k ) ∼ =


4.2 Classical methods for computation 159<br />

k. Proposition 4.2.1 is not useful for computing Ext ∗ kC(S x,k , S y,k )<br />

so we have to compute the homology <strong>of</strong> 0 → Hom kC (P x,k , S y,k ) →<br />

Hom kC (P 2 y,k , S y,k) → 0, which gives Ext 1 kC(S x,k , S y,k ) ∼ = k 2 .<br />

For Ext ∗ kC(S x,k , k) we can use a resolution but we can also<br />

use the long exact sequence<br />

0 → Ext 0 kC(S x,k , k) → Ext 0 kC(S x,k , S x,k ) → Ext 1 kC(S x,k , S y,k ) → Ext 1 kC<br />

from<br />

0 → S y,k → k → S x,k → 0<br />

after applying Ext ∗ kC(S x,k , −). Since Ext 0 kC(S x,k , k) ∼ = Hom kC (S x,k , k)<br />

k, we know immediately Ext 1 kC(S x,k , k) ∼ = k 2 .<br />

3. Suppose k is a field <strong>of</strong> characteristic 2. Consider the following<br />

category D<br />

α <br />

y {1 y ,g}<br />

<br />

{1 x } x<br />

with gα = α. Then here are two one-dimensional simple<br />

modules S x,k <strong>and</strong> S y,k , together with their projective<br />

covers P x,k = k{1 x , α} <strong>and</strong> P y,k = {1 y , g}. Particularly<br />

k ∼ = kHom D (x, −) = P x,k<br />

M ∈ kD-mod, Ext ∗ kD(k, M) ∼ = M(x).<br />

The minimal projective resolution <strong>of</strong> S x,k is<br />

is projective. Thus for any<br />

· · · → P y,k → P y,k → P x,k → S x,k → 0,<br />

which is infinite <strong>and</strong> consists <strong>of</strong> a copy <strong>of</strong> P y,k in each degree<br />

n > 0. Thus Ext 0 kD(S x,k , S x,k ) ∼ = k <strong>and</strong> Ext n kD(S x,k , S y,k ) ∼ = k<br />

for all n > 0.<br />

The minimal projective resolution <strong>of</strong> S y,k is<br />

· · · → P y,k → P y,k → P y,k → S y,k → 0,<br />

which is infinite <strong>and</strong> consists <strong>of</strong> a copy <strong>of</strong> P y,k in each degree.<br />

We get Ext n kD(S y,k , S y,k ) ∼ = k for every n ≥ 0. Moreover<br />

Ext ∗ kD(S y,k , S y,k ) ∼ = Ext ∗ kC 2<br />

(k, k) by Proposition 4.2.1. We


160 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

leave the computation <strong>of</strong> Ext ∗ kD(−, k) <strong>and</strong> <strong>of</strong> all above Ext<br />

groups when k = C to the reader.<br />

One can determine the ring structure <strong>of</strong> Ext ∗ kC 2<br />

(k, k). In<br />

fact C 2 has a very simple bar resolution so one may quickly<br />

find, using simplicial method, that Ext ∗ kC 2<br />

(k, k) ∼ = k[X] is a<br />

polynomial ring with deg X = 1.<br />

Since Ext ∗ kD(k, k) ∼ = k, both Ext ∗ kD(S x,k , S y,k ) <strong>and</strong> Ext ∗ kD(S y,k , S y,k )<br />

are not finitely generated modules.<br />

If we look at the opposite category D op , again we have<br />

Ext ∗ kD op(k, k) ∼ = Ext ∗ kD(k, k) ∼ = k, <strong>and</strong> Ext ∗ kD op(k, S y,k) ∼ =<br />

Ext ∗ kD op(S y,k, S y,k ) is <strong>of</strong> infinite dimensional. It means we do<br />

not have the finite generation <strong>of</strong> Ext ∗ kD(k, M) over Ext ∗ kD(k, k)<br />

even if they can be calculated by using the same projective<br />

resolution.<br />

4.2.2 Examples using classifying spaces<br />

Underst<strong>and</strong>ing the homotopy type <strong>of</strong> a classifying space certainly<br />

will be a great help for compute cohomology rings. Here<br />

we list several cases where direct calculation is possible.<br />

Example 4.2.3. In Example 4.2.2, part (1) is a contractible<br />

category n so we know Ext ∗ kn(k, k) ∼ = k. Part (2) is a category<br />

whose classifying space is the circle S 1 . Part (3) is a<br />

contractible category because it has an initial object.<br />

However, there are more things we can take from homotopy<br />

theory. If we know all Ext ∗ kC(k, −), then we may readily write<br />

down Ext ∗ kCop(k, −).<br />

Now we examine some more sophisticated examples. Although<br />

they do not tell much about cohomology with local<br />

coefficients, they do provide substantial information about the<br />

ordinary cohomology ring <strong>of</strong> a category.<br />

Example 4.2.4.1. Let C be the following category


4.2 Classical methods for computation 161<br />

G<br />

x<br />

α <br />

y H .<br />

Then this is the join <strong>of</strong> G <strong>and</strong> H. Thus its classifying space<br />

is a join <strong>of</strong> spaces BC ≃ BG ∗ BH. Thus its cohomology is<br />

completely determined by G <strong>and</strong> H. Also the cup product<br />

is trivial.<br />

2. Let D be the following category<br />

h<br />

x<br />

1 x<br />

α<br />

β<br />

y {1 y }<br />

,<br />

where h 2 = 1 x <strong>and</strong> βh = α. Then the inclusion i :<br />

Aut D (x) → D induces a homotopy equivalence between<br />

their classifying spaces. In particular, we have a ring isomorphism<br />

Ext ∗ RD(R, R) = Ext ∗ RAut D (x)(R, R) for any ring R.<br />

When R = C, the cohomology ring is just C, while when R<br />

is a field <strong>of</strong> characteristic 2 the ordinary cohomology ring is<br />

a polynomial algebra with indeterminant in degree one.<br />

3. Let E be the following category<br />

h<br />

g<br />

x<br />

1 x<br />

α<br />

β<br />

gh<br />

y {1 y }<br />

,<br />

where g 2 = h 2 = 1 x , gh = hg, αh = α, αg = β, βh =<br />

β, βg = β. Its classifying space is the homotopy pushout <strong>of</strong><br />

BC 2<br />

BAut E (y) ∼ = •<br />

BAut E (x) = B(C 2 × C 2 )<br />

where the cyclic group <strong>of</strong> order 2 is {1 x , h}, the stabilizer <strong>of</strong><br />

Hom E (x, y) = {α, β}. Thus BE ≃ BAut E (x)/BC 2 . In fact,<br />

by Definition 2.4.3 <strong>and</strong> Proposition 2.4.4, we can consider<br />

the subdivision <strong>of</strong> E, which is a Grothendieck construction


162 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

Gr sd[E] M. Then by Theorem 2.4.10, BGr sd[E] M, <strong>and</strong> thus<br />

BE, is homotopy equivalent to hocolim sd[E] BM. Since sd[E]<br />

is a poset • ← • → •, this homotopy colimit is just the<br />

homotopy pushout <strong>of</strong> the above diagram, if one follows Definition<br />

2.4.3 to compute BM.<br />

We can use the exact sequence for computing relative cohomology<br />

(see for instance [34])<br />

H ∗ (BAut E (x), BC 2 ; R)<br />

to find the positive degree cohomology <strong>of</strong> BE, while in degree<br />

zero we just have R because E is connected. In fact<br />

H ∗ (BE, k), for k a field <strong>of</strong> characteristic 2, is a subring <strong>of</strong><br />

the polynomial ring k[X, Y], with deg X = deg Y = 1 <strong>and</strong><br />

with all linear combinations <strong>of</strong> {X i } i>0 removed. This ring<br />

has no nilpotent elements <strong>and</strong> is not finitely generated. In<br />

fact, this is the smallest example in commutative algebra<br />

that a subring <strong>of</strong> a Noetherian ring is not finitely generated.<br />

However when R = C, C is projective (by direct calculation<br />

or by Proposition 3.3.11). That means H 0 (BE, C) = C is<br />

the only non-vanishing cohomology. Note that C as a RE op -<br />

module is not projective, although H ∗ (BE op , C) = C as well.<br />

4.3 Computation via adjoint functors<br />

4.3.1 Adjoint restrictions<br />

It is well-known that if a functor u : D → C has a left adjoint<br />

v : C → D, then Res v is also the left adjoint <strong>of</strong> Res u . These<br />

are very special Kan extensions <strong>and</strong> hence can be regarded<br />

as the first applications <strong>of</strong> Kan extension before we move to<br />

more subtle situations. Because both Res u <strong>and</strong> Res v are exact,<br />

Res u preserves injectives while Res v preserves projectives.


4.3 Computation via adjoint functors 163<br />

Lemma 4.3.1. If u : D → C has a left adjoint v : C → D,<br />

then<br />

Ext ∗ RC(Res v M, N) ∼ = Ext ∗ RD(M, Res u N)<br />

for any M ∈ RD-mod <strong>and</strong> N ∈ RC-mod. If M ′ is a right<br />

RD-module, then<br />

Tor RC<br />

∗ (Res v M ′ , N) ∼ = Tor RD<br />

∗ (M ′ , Res u N)<br />

If furthermore u is an equivalence then we have<br />

Ext ∗ RC(M, N) ∼ = Ext ∗ RD(Res u M, Res u N),<br />

for M, N ∈ RC-mod. In particular res u is a ring isomorphism<br />

Ext ∗ RC(R, R) ∼ = Ext ∗ RD(R, R). If M ′′ is a right RCmodule,<br />

then under the circumstance we get<br />

Tor RC<br />

∗ (M ′′ , N) ∼ = Tor RD<br />

∗ (Res u M ′′ , Res u N)<br />

Pro<strong>of</strong>. These follow directly from the definitions <strong>of</strong> Ext <strong>and</strong><br />

Tor, as well as the adjunction.<br />

⊓⊔<br />

Let us briefly go over an important approach <strong>of</strong> computing<br />

∗<br />

higher limits ←−<br />

lim M. The atomic modules are useful to us<br />

C<br />

because for many interesting category C every module in RCmod<br />

admits a filtration by submodules such that every quotient<br />

is an atomic module. Since the higher limits <strong>of</strong> a module<br />

can be computed via higher limits <strong>of</strong> its atomic subquotients,<br />

∗<br />

one can seek ways <strong>of</strong> computing ←−<br />

lim M for an atomic module<br />

C<br />

M. This method was introduced <strong>and</strong> studied by Oliver [].<br />

∗<br />

We note that usually ←−<br />

lim M, for an atomic module M concentrated<br />

on x ∈ Ob C, is quite different from the group<br />

C<br />

cohomology H ∗ (Aut C (x), M(x)). Given a projective resolution<br />

P <strong>of</strong> R, P(x) must be an exact sequence <strong>of</strong> RAut C (x)-<br />

modules, beginning with R, for any object x ∈ Ob C. But<br />

the modules in P(x) do not have to be projective, <strong>and</strong> even<br />

if P(x) → R → 0 is a projective resolution the cohomology


164 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

groups H ∗ (C, M) <strong>and</strong> H ∗ (Aut C (x), M(x)) are not necessarily<br />

isomorphic. We comment here that the modules in P(x) will<br />

be projective if RAut C (x) happens to be semi-simple. Another<br />

possible way to obtain a projective resolution for the<br />

RAut C (x)-module R is formulated as follows. If Aut C (x) acts<br />

freely on Hom C (y, x) for each y ∈ Ob C with Hom(y, x) ≠ ∅,<br />

then every RHom C (y, x) is a free RAut C (x)-module. As an example<br />

the st<strong>and</strong>ard resolution evaluated at x, P C (x), becomes<br />

a projective resolution for the RAut C (x)-module R.<br />

4.3.2 Kan extensions <strong>of</strong> resolutions<br />

For an arbitrary functor u : D → C we can still construct<br />

the left <strong>and</strong> right adjoints <strong>of</strong> Res u , i.e. the Kan extensions. In<br />

the situation that u admits a left (or right) adjoint v, then<br />

the left (or right) Kan extension enjoys a particularly simple<br />

form Res v , which is exact. In general, the Kan extensions cannot<br />

have simplified forms, <strong>and</strong> consequently computing Kan<br />

extensions <strong>of</strong> an arbitrary module is difficult <strong>and</strong> the new<br />

modules do not come with explicit descriptions. However if a<br />

module is simplicially constructed, in the sense that it comes<br />

from the nerve <strong>of</strong> a category, then we do have some satisfactory<br />

results on their Kan extensions. This is a fundamental<br />

step to study cohomology rings <strong>and</strong> higher limits.<br />

Proposition 4.3.2. Suppose u : D → C is a functor between<br />

small categories. Then<br />

LK u B D ∗<br />

= LK u C ∗ (Id D /−, R) ∼ = C ∗ (u/−, R),<br />

a complex <strong>of</strong> projective RC-modules.<br />

Particularly for pt : D → •, we have<br />

LK pt B D ∗<br />

= LK pt C ∗ (Id D /•, R) ∼ = C ∗ (pt/•, R) ∼ = C ∗ (D, R),


4.3 Computation via adjoint functors 165<br />

Pro<strong>of</strong>. For the sake <strong>of</strong> convenience, we suppress the base ring<br />

R in our notations. Suppose x ∈ Ob C. Fix an integer n ≥ 0.<br />

Then<br />

LK u C n (Id D /−) ∼ = lim −→u/x<br />

C n (Id D /−).<br />

If (y, α) ∈ Ob u/x, that is, there is an α : u(y) → x, then by<br />

definition [C n (Id D /−)](y, α) = C n (Id D /y), <strong>and</strong> we can define<br />

a morphism<br />

θ (y,α) : C n (Id D /y) → C n (u/x)<br />

by [(y 0 , α 0 ) β 1<br />

→ · · · →(y βn<br />

n , α n )] ↦→ [(y 0 , αα 0 ) β 1<br />

→ · · · →(y βn<br />

n , αα n )].<br />

One can easily verify that we have a commutative diagram<br />

for any γ : (y, α) → (z, β)<br />

C n (Id D /y) = [C n (Id D /−)](y, α)<br />

θ (y,α)<br />

γ ∗<br />

C n (u/x)<br />

[C n (Id D /−)](z, β) = C n (Id<br />

θ (z,β)<br />

so that C n (u/x) fits into the limit defining diagram. If L is<br />

another R-module that fits into the limit defining diagram,<br />

equipped with maps θ L (y,α) : [C n(Id D /−)](y, α) → L for all<br />

(y, α) ∈ Ob(u/x). Then we can introduce a morphism Θ :<br />

C n (u/x) → L such that<br />

Θ[(y 0 , αα 0 ) β 1<br />

→ · · · →(y βn<br />

n , αα n )]<br />

:= θ(y L n ,αα n ) [(y 0, u(β n · · · β 1 )) β 1<br />

→(y 0 , u(β n · · · β 2 )) β 2<br />

→ · · · →(y βn<br />

n , 1 u(yn ))].<br />

This is a well defined morphism. In the end, since α n :<br />

(y n , αα n ) → (y, α) in a morphism in u/x, it induces a functor<br />

C n (Id D /y n ) → C n (Id D /y). In particular this functor gives<br />

[(y 0 , u(β n · · · β 1 )) β 1<br />

→(y 0 , u(β n · · · β 2 )) β 2<br />

→ · · · →(y βn<br />

n , 1 u(yn ))]<br />

↦→ [(y 0 , α 0 ) β 1<br />

→ · · · →(y βn<br />

n , α n )],


166 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

by composing with α n . By assumptions on L we get<br />

Θ(y L n ,αα n ) [(y 0, u(β n · · · β 1 )) β 1<br />

→(y 0 , u(β n · · · β 2 )) β 2<br />

→ · · · →(y βn<br />

n , 1 u(yn ))]<br />

= Θ(y,α) L [(y 0, α 0 ) β 1<br />

→ · · · →(y βn<br />

n , α n )].<br />

This equality implies that we can insert Θ into the following<br />

diagram <strong>and</strong> make it commutative<br />

C n (Id D /y) = [C n (Id D /−)](y, α)<br />

θ (y,α)<br />

θ L (y,α)<br />

γ ∗<br />

C n (u/x)<br />

Θ <br />

L<br />

[C n (Id D /−)](z, β) = C n (Id<br />

θ (z,β)<br />

θ L (z,β)<br />

Consequently C n (u/x) ∼ = lim −→u/x<br />

C n (Id D /−) for all x ∈ Ob C.<br />

Since all these isomorphisms assemble to an isomorphism<br />

C ∗ (u/−) ∼ = LK u C ∗ (Id D /−), by naturality, we are done. ⊓⊔<br />

Note the second statement generalizes the obvious fact<br />

LK pt B∗<br />

D = LK pt RD ∼ = lim −→D<br />

RD ∼ = R⊗ RD RD ∼ = R = R Ob D = C 0 (D<br />

Remark 4.3.3. To underst<strong>and</strong> the complex <strong>of</strong> projective RCmodules<br />

C ∗ (u/−), we describe the structure <strong>of</strong> each C n (u/−).<br />

As we observed in the pro<strong>of</strong>, each base element (y 0 , α 0 ) →<br />

· · · → (y n , α n ) ∈ C n (u/−) can be written as α n·[(y 0 , u(β n · · · β 1 )) →<br />

· · · → (y n , 1 u(yn ))]. Thus similar to the special case <strong>of</strong> bar resolution<br />

where u = Id C , we know<br />

LK u Bn D ∼ = C n (u/−, R) ∼ ⊕<br />

=<br />

RHom C (u(y n ), −).<br />

(y 0 ,u(β n···β 1 )) β 1<br />

→···βn<br />

→(y n ,1 u(y n))<br />

For future reference, the dual version for the right bar resolution<br />

is


4.3 Computation via adjoint functors 167<br />

LK u B D n ∼ = C n (−\u, R) ∼ =<br />

⊕<br />

(1 u(y0 ),y 0 ) β 1<br />

→···βn<br />

→(u(β n···β 1 ),y n )<br />

RHom C (−, u(y 0 )).<br />

Recall that a category is R-acyclic if its reduced (simplicial)<br />

homology with coefficients in R vanishes.<br />

Corollary 4.3.4. Suppose u : D → C is a functor, M ∈<br />

RC-mod <strong>and</strong> N ∈ mod-RC. Then<br />

1. if u/x is R-acyclic for every x ∈ Ob C,<br />

Ext ∗ RC(R, M) ∼ = Ext ∗ RD(R, Res u M) <strong>and</strong> Tor RC<br />

∗<br />

2. if x\u is R-acyclic for every x ∈ Ob C,<br />

Ext ∗ RC(R, N) ∼ = Ext ∗ RD(R, Res u N) <strong>and</strong> Tor RC<br />

∗<br />

(N, R) ∼ = Tor RD<br />

∗ (Res<br />

(R, M) ∼ = Tor RD<br />

∗ (R, R<br />

Pro<strong>of</strong>. We prove part (1). Under the assumption, LK u R ∼ = R<br />

<strong>and</strong> LK u B∗<br />

D ∼ = C ∗ (u/−, R) → R → 0 becomes a projective<br />

resolution. Hence the isomorphism follows from<br />

Hom RC (C ∗ (u/−, R), M) ∼ = Hom RD (B D ∗ , Res u M).<br />

For the second isomorphism, we also have<br />

N ⊗ RC C ∗ (u/−, R) ∼ = ⊕ N ⊗ (y 0 ,u(β n···β 1 )) β 1<br />

RC RHom C (u(y n )<br />

→···βn<br />

→(y n ,1 u(y n))<br />

∼<br />

⊕ = (Res (y 0 ,u(β n···β 1 )) β 1<br />

uN)(y n )<br />

→···βn<br />

→(y n ,1 u(y n))<br />

∼<br />

⊕ = β<br />

y 1<br />

(Res u N)(y n )<br />

0 →···βn<br />

→y n<br />

∼ = Resu N ⊗ RD Bn D .<br />

⊓⊔<br />

The above isomorphisms are Eckmann-Shapiro type results.<br />

Example 4.3.5. Let D be the category from Example 4.2.4 (2)<br />

h<br />

x<br />

1 x<br />

α<br />

β<br />

y {1 y }<br />

,


168 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

where h 2 = 1 x <strong>and</strong> βh = α. The inclusion i : Aut D (x) → D<br />

has two overcategories i/x <strong>and</strong> i/y, both contractible. Thus<br />

we have<br />

Ext ∗ RAut D (x)(R, M) ∼ = Ext ∗ RD(R, Res i M).<br />

This calculation can be generalized to all categories that possess<br />

a unique minimal object x whose automorphism group<br />

acts regularly on every non-empty morphism set Hom D (x, y).<br />

Recall from Section 2.3.1 that we introduced the subdivision<br />

S(C) <strong>of</strong> an EI category C <strong>and</strong> showed that their classifying<br />

spaces are homotopy equivalent <strong>and</strong> hence they have the same<br />

simplicial (co)homology.<br />

Corollary 4.3.6. Let C be an EI category <strong>and</strong> S(C) its subdivision.<br />

Suppose M ∈ RC-mod <strong>and</strong> M ′ ∈ RC op -mod. Then<br />

Tor RC<br />

∗ (R, M) ∼ = Tor RS(C)<br />

∗ (R, Res t M)<br />

<strong>and</strong><br />

Ext ∗ RC op(R, M ′ ) ∼ = Tor RS(C)op<br />

∗ (R, Res s M ′ )<br />

where s : S(C) → C is the natural functor.<br />

Pro<strong>of</strong>. In the pro<strong>of</strong> <strong>of</strong> Proposition 2.3., we actually proved<br />

that x\s is contractible for all x ∈ Ob C. This is equivalent<br />

to having s op /x contractible, by Lemma 1.2.8. Hence we can<br />

apply Corollary 4.3.4.<br />

⊓⊔<br />

Remark 4.3.7. Let u : D → C be a functor. We can use LK u<br />

to see how one obtains the maps introduced in Chapter 2<br />

u ∗ : H∗(D, R) → H∗(C, R) <strong>and</strong> u ∗ : H ∗ (C, R) → H ∗ (D, R). In<br />

fact we have the following diagram<br />

D<br />

u <br />

pt<br />

C<br />

pt<br />


4.3 Computation via adjoint functors 169<br />

Thus by LK pt = LK pt LK u we get LK pt B∗ D = LK pt LK u B∗ D .<br />

Since LK u B∗<br />

D ∼ = C ∗ (u/−, R) is a complex <strong>of</strong> projective modules<br />

<strong>and</strong> meanwhile there is a canonical map LK u R → R<br />

between RC-modules, there exists a lifting, which is a chain<br />

map unique up to chain homotopy (Comparison Theorem),<br />

· · ·<br />

<br />

LK u B∗<br />

D<br />

LK u R<br />

0<br />

· · ·<br />

<br />

B C ∗<br />

<br />

R<br />

<br />

0<br />

Conceptually we see the left Kan extension induces chain<br />

maps<br />

R ⊗ RD B∗ D ∼ = R ⊗ RC LK u B∗ D → R ⊗ RC B∗<br />

C<br />

because R ⊗ RD B∗ D = LK pt B∗ D = LK pt LK u B∗<br />

D ∼ = R ⊗RC<br />

LK u B∗<br />

D <strong>and</strong><br />

Hom RC (B C ∗, R) → Hom RC (LK u B D ∗ , R) ∼ = Hom RD (B D ∗ , R).<br />

Hence these give exactly the same map we introduced in<br />

Chapter 2. Alternatively by applying LK pt to the previous<br />

commutative diagram we obtain a chain map<br />

<br />

· · · LK pt LK u B∗<br />

D<br />

LK pt LK u R<br />

0<br />

· · ·<br />

<br />

LK pt B C ∗<br />

<br />

LK pt R<br />

<br />

0<br />

The positive degree part LK pt B D ∗ = LK pt LK u B D ∗ → LK pt B C ∗<br />

is exactly the chain map induced by u in Definition 2.2.4, i.e.<br />

C ∗ (D, R) → C ∗ (C, R), in light <strong>of</strong> Proposition 4.3.2.<br />

Remark 4.3.8. The following descriptions balance the underst<strong>and</strong>ing<br />

<strong>of</strong> left <strong>and</strong> right Kan extensions. Especially we know<br />

the right Kan extension <strong>of</strong> a certain injective resolution <strong>of</strong> k.<br />

Since (−) ∧ establishes a one-to-one correspondence between


170 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

right projective kC-modules <strong>and</strong> left injective kC-modules, we<br />

can extend the above lemmas to the bar resolution <strong>of</strong> k ∈modkD<br />

<strong>and</strong> obtain a complex <strong>of</strong> left injective kC-modules. Suppose<br />

B∗<br />

D ∼ → k → 0 is the bar resolution <strong>of</strong> k ∈mod-kD, where<br />

= C ∗ (−\Id D ). Then<br />

B D ∗<br />

0 → k → (B D ∗ ) ∧ ∼ = C∗ (−\Id D ) ∧<br />

is an injective resolution <strong>of</strong> the left kD-module k. Applying<br />

RK u we get a complex <strong>of</strong> injective kC-modules, excluding<br />

RK u k, as follows<br />

0 → RK u k → RK u (B D ∗ ) ∧ .<br />

By Lemma 3.2.13 the complex RK u (B D ∗ ) ∧ ∼ = (LK u B D ∗ ) ∧ ∼ =<br />

C ∗ (−\u) ∧ . Thus we have a simplicially constructed injective<br />

resolution <strong>of</strong> k, <strong>and</strong> we have control on the complex after<br />

applying the right Kan extension<br />

0 → RK u k → (B D ∗ ) ∧ ∼ = C∗ (−\u) ∧ .<br />

Remark 4.3.9. At this stage, the reader shall be well prepared<br />

to go over Chapter 5 where the main results in Section 5.2<br />

are applications <strong>of</strong> Proposition 4.3.2.<br />

4.4 Grothendieck spectral sequences<br />

Let u : D → C be a functor. We can construct Grothendieck<br />

spectral sequences. They generalize the Lydon-Hochschil-<br />

Serre spectral sequences for group extensions as well as the<br />

Leray-Serre spectral sequences for fibrations.<br />

4.4.1 Grothendieck spectral sequences for a functor<br />

Let u : D → C be a functor between two small categories.<br />

Consider the composites <strong>of</strong> following functors


4.4 Grothendieck spectral sequences 171<br />

as well as<br />

RD-mod RK u <br />

RC-mod lim ←−C <br />

R-mod,<br />

RD-mod LK u <br />

RC-mod lim −→C<br />

<br />

R-mod.<br />

The second isomorphism <strong>of</strong> the following is used in Remark<br />

4.3.7.<br />

Lemma 4.4.1. lim ←−C<br />

RK u<br />

∼ = lim ←−D<br />

<strong>and</strong> lim −→C<br />

LK u<br />

∼ = lim −→D<br />

.<br />

Pro<strong>of</strong>. This is a special case <strong>of</strong> Corollary 1.2.12 for u : D → C<br />

<strong>and</strong> pt : C → •.<br />

⊓⊔<br />

Since RK u preserves injectives, there exists a Grothendieck<br />

cohomology spectral sequence for any functor u : D → C<br />

<strong>and</strong> M ∈ RD-mod, which comes from a double complex<br />

E ∗,∗<br />

0<br />

E ∗,∗<br />

2<br />

(M) that we will describe shortly. The spectral sequence<br />

(M) ⇒ E∗,∗ ∞ (M) is<br />

H ∗ (C; H ∗ (−\u; M)) ⇒ H ∗ (D; M).<br />

i<br />

Remember that for any small category E, one has ←−<br />

lim ∼<br />

E =<br />

H i (E; −) ∼ = Ext i RE(R, −). It means H ∗ (−\u; M) is some sort<br />

<strong>of</strong> higher right Kan extension <strong>of</strong> M. Since LK u preserves projectives,<br />

we get a Grothendieck homology spectral sequence<br />

as well. However we will only construct the cohomology spectral<br />

sequence as the construction for the homology spectral<br />

sequence is similar. For future reference, we record the<br />

Grothendieck homology spectral sequence for u : D → C <strong>and</strong><br />

M ∈ RD-mod<br />

H∗(C; H∗(u/−; M)) ⇒ H∗(D; M),<br />

in which H∗(u/−; M) should be considered as higher left Kan<br />

extensions <strong>of</strong> M.


172 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

We shall construct the cohomology spectral sequence <strong>and</strong><br />

then assure the reader that there exists a natural pairing <strong>of</strong><br />

such double complexes E ∗,∗<br />

0 (M) ⊗ E∗,∗ 0 (N) → E∗,∗ 0 (M ˆ⊗N).<br />

With the pairing, we have a product on each page <strong>of</strong> the<br />

Grothendieck cohomology spectral sequences<br />

En i,j (M) ⊗ En s,t (N) → En<br />

i+s,j+t (M ˆ⊗N),<br />

<strong>and</strong><br />

E∞ i,j (M) ⊗ E∞ s,t (N) → E∞<br />

i+s,j+t (M ˆ⊗N),<br />

Thus we have a ring structure on<br />

H ∗ (C; H ∗ (−\u; R)) ⇒ H ∗ (D; R),<br />

over which the following is a module<br />

H ∗ (C; H ∗ (−\u; M)) ⇒ H ∗ (D; M).<br />

Fix an RD-module M we start with a double complex<br />

(M). First take an injective resolution <strong>of</strong> M<br />

E ∗,∗<br />

0<br />

0 → M → I 0 → I 1 → I 2 → · · · .<br />

Then we apply RK u to get a complex <strong>of</strong> injective kC-modules<br />

RK u I 0 → RK u I 1 → RK u I 2 → · · · ,<br />

<strong>and</strong> consequently a commutative diagram<br />

RK u I 0<br />

RK u I 1<br />

RK u I 2<br />

· · ·<br />

J 0,0<br />

J 1,0<br />

J 2,0<br />

· · ·<br />

J 0,1<br />

J 1,1<br />

J 2,1<br />

· · ·<br />

. . . ,


4.4 Grothendieck spectral sequences 173<br />

in which every column is an injective resolution <strong>of</strong> the top<br />

module. Now apply ←−C<br />

lim <strong>and</strong> we obtain a double cochain complex,<br />

denoted by E ∗,∗<br />

0 (M),<br />

lim<br />

←−C J 0,0<br />

lim<br />

←−C J 1,0<br />

lim<br />

←−C J 2,0<br />

· · ·<br />

lim<br />

←−C J 0,1<br />

lim<br />

←−C J 1,1<br />

lim<br />

←−C J 2,1<br />

· · ·<br />

lim<br />

←−C J 0,2<br />

lim<br />

←−C J 1,2<br />

lim<br />

←−C J 2,2<br />

· · ·<br />

. . . ,<br />

<strong>and</strong> it gives rise to the Grothendieck spectral sequence recorded<br />

above. We omit the details as the construction is st<strong>and</strong>ard <strong>and</strong><br />

we are more interested in finding a pairing. Suppose we also<br />

have a double complex E ∗,∗<br />

0 (N) for another RD-module N<br />

lim<br />

←−C J ′ 0,0<br />

lim<br />

←−C J ′ 1,0<br />

lim<br />

←−C J ′ 2,0<br />

· · ·<br />

lim<br />

←−C J ′ 0,1<br />

lim<br />

←−C J ′ 1,1<br />

lim<br />

←−C J ′ 2,1<br />

· · ·<br />

lim<br />

←−C J ′ 0,2<br />

lim<br />

←−C J ′ 1,2<br />

lim<br />

←−C J ′ 2,2<br />

· · ·<br />

. . . ,<br />

<strong>and</strong> furthermore a double complex E ∗,∗<br />

0 (M ˆ⊗N) for the RDmodule<br />

M ˆ⊗N


174 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

lim<br />

←−C J ′′<br />

0,0<br />

lim<br />

←−C J ′′<br />

1,0<br />

lim<br />

←−C J ′′<br />

2,0<br />

· · ·<br />

lim<br />

←−C J ′′<br />

0,1<br />

lim<br />

←−C J ′′<br />

1,1<br />

lim<br />

←−C J ′′<br />

2,1<br />

· · ·<br />

lim<br />

←−C J ′′<br />

0,2<br />

lim<br />

←−C J ′′<br />

1,2<br />

lim<br />

←−C J ′′<br />

2,2<br />

. . . .<br />

We want to establish a natural map<br />

lim J ←−C<br />

i,j ⊗ lim ←−C<br />

J s,t ′ → lim ←−C<br />

J i+s,j+t,<br />

′′<br />

which is compatible with the differentials. In fact, since there<br />

is a unique map, given by the universal property <strong>of</strong> ←−<br />

lim,<br />

lim J ←−C<br />

i,j ⊗ lim ←−C<br />

J s,t ′ → lim ←−C<br />

J i,j ˆ⊗J s,t,<br />

′<br />

we only need to construct a map ←−C<br />

lim J i,j ˆ⊗J s,t ′ → lim ←−C<br />

J i+s,j+t.<br />

′′<br />

Our definition is again based on the universal property <strong>of</strong><br />

lim<br />

←−<br />

, along with the tensor product <strong>of</strong> complexes <strong>of</strong> functors<br />

in Section 3.4.1. We emphasize that ←−C<br />

lim J i,j ⊗ lim ←−C<br />

J s,t<br />

′ →<br />

lim J ←−C<br />

i,j ˆ⊗J s,t ′ respects the differentials in E ∗,∗<br />

0 due to its construction<br />

via the universal property. This is the case when we<br />

define lim ←−C<br />

J i,j ˆ⊗J s,t ′ → lim ←−C<br />

J i+s,j+t ′′ <strong>and</strong> thus we will not verify<br />

the map we are about to construct does respect differentials.<br />

From the two injective resolutions0 → M → I ∗ <strong>and</strong> 0 →<br />

N → I ∗, ′ we can build a commutative diagram<br />

0<br />

<br />

M ˆ⊗N I ∗ ˆ⊗I ′ ∗<br />

<br />

0<br />

<br />

M ˆ⊗N<br />

<br />

I ′′<br />

∗ ,<br />

· · ·


4.4 Grothendieck spectral sequences 175<br />

in which the upper row is an exact sequence <strong>and</strong> the lower<br />

one is the injective resolution used to define E ∗,∗<br />

0 (M ˆ⊗N). Applying<br />

RK u we obtain a chain map RK u (I ∗ ˆ⊗I ∗) ′ → RK u I ∗ ′′ .<br />

Especially we have for any non-negative integers i <strong>and</strong> s a<br />

map RK u (I i ˆ⊗I s) ′ → RK u I i+s. ′′ The universal property <strong>of</strong> ←−<br />

lim<br />

provides a morphism RK u I i ˆ⊗RK u I s ′ → RK u (I i ˆ⊗I s). ′ Thus<br />

we have a natural map<br />

RK u I i ˆ⊗RK u I s ′ → RK u I i+s.<br />

′′<br />

Next we repeat the above tensor construction for the two injective<br />

resolutions 0 → RK u I i → J i,∗ <strong>and</strong> 0 → RK u I s ′ → J s,∗.<br />

′<br />

It follows from our discussions that there is a commutative diagram<br />

0<br />

<br />

RK u I i ˆ⊗RK u I s<br />

′ J i,∗ ˆ⊗J s,∗<br />

′<br />

0<br />

<br />

RK u I ′′<br />

i+s<br />

<br />

J i+s,∗+∗ ′′ .<br />

In particular there exists J i,j ˆ⊗J s,t ′ → J i+s,j+t, ′′ <strong>and</strong> consequently<br />

the desired map lim ←−C<br />

J i,j ˆ⊗J s,t ′ → lim ←−C<br />

J i+s,j+t. ′′ Hence<br />

we do obtain a pairing E ∗,∗<br />

0 (M) ⊗ E∗,∗ 0 (N) → E∗,∗ 0 (M ˆ⊗N).<br />

4.4.2 Spectral sequences <strong>of</strong> category extensions<br />

We introduced category extensions <strong>and</strong> opposite extensions<br />

in Section 4.1.3. Here we want to show that the Grothendieck<br />

spectral sequences for a functor have simpler forms when the<br />

functor fits into an extension sequence.<br />

Lemma 4.4.2. Let K → E → C be an extension. Then there<br />

exists a natural functor ι : K(y) → π/y such that every<br />

undercategory associated with it is contractible. Hence


176 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

1. H∗(π/−; M) ∼ = H∗(K(−); M(−)) in RC-mod for any M ∈<br />

RE-mod; <strong>and</strong><br />

2. H ∗ (−\π op ; N) ∼ = H ∗ (K op (−); N(−)) in RC op -mod for any<br />

N ∈ RE op -mod.<br />

Pro<strong>of</strong>. The category π/y has objects <strong>of</strong> the form (x, α), where<br />

x ∈ Ob E = Ob C <strong>and</strong> α ∈ Hom C (x, y). From the definition<br />

<strong>of</strong> π/y, one can see that the maximal objects are (y, g),<br />

g ∈ Aut C (y), which are isomorphic to each other <strong>and</strong> have<br />

automorphism groups isomorphic to K(y). Let us take the<br />

full subcategory [(y, 1 y )] <strong>of</strong> π/y consisting <strong>of</strong> all maximal objects.<br />

Its skeleton is isomorphic to the group K(y). Using<br />

Quillen’s Theorem A, we show the undercategories associated<br />

with ι : [(y, 1 y )] ↩→ π/y are contractible.<br />

Fix an object (x, α) ∈ Ob(π/y). The undercategory (x, α)\(π/y)<br />

has objects <strong>of</strong> the form (β, (y, g)), where β : (x, α) → (y, g)<br />

is an morphism in π/y satisfying gπ(β) = α. Since π(β) =<br />

g −1 α, by the definition <strong>of</strong> a category extension, β = g −1 αh<br />

for a unique h ∈ K(x). From here we can deduce that<br />

(β, (y, g)) ∼ = (β ′ , (y, g ′ )) for any (y, g ′ ) <strong>and</strong> β ′ : (x, α) →<br />

(y, g ′ ), <strong>and</strong> that (β, (y, g)) ∈ (x, α)\(π/y) has a trivial automorphism<br />

group. These imply (x, α)\(π/y) is equivalent to a<br />

point, <strong>and</strong> hence is contractible.<br />

The first isomorphism follows from Corollary 4.3.4 (2)<br />

Tor R(π/y)<br />

∗<br />

(R, M) ∼ = Tor RK(y) (R, Res π M),<br />

<strong>and</strong> the naturality in y.<br />

The pro<strong>of</strong> <strong>of</strong> the second is the same. By Lemma 1.2.8 the<br />

overcategories associated with ι op : K(y) op → (π/y) op are<br />

contractible. Then we apply Corollary 4.3.4 (1). ⊓⊔<br />

We can obtain similar statements for opposite extensions.<br />

Keep in mind that since K is a group, K ∼ = K op .<br />


4.4 Grothendieck spectral sequences 177<br />

Lemma 4.4.3. Let K → E → C be an opposite extension.<br />

Then there exists a natural functor ι : K(y) → y\π such<br />

that every undercategory associated with it is contractible.<br />

Hence<br />

1. H∗(π op /−; M) ∼ = H∗(K op (−); M(−)) in RC op -mod for any<br />

M ∈ RE op -mod; <strong>and</strong><br />

2. H ∗ (−\π; N) ∼ = H ∗ (K(−); N(−)) in RC-mod for any N ∈<br />

RE-mod.<br />

Pro<strong>of</strong>. Since K op → E op → C op is an extension, we know by<br />

Lemma 4.4.2 that there is a functor ι : K op (y) → π op /y such<br />

that all undercategories are contractible. From Lemma 1.2.8,<br />

this is equivalent to saying that all overcategories associated<br />

to ι op : K(y) → y\π are contractible.<br />

⊓⊔<br />

For brevity we write H ∗ (K; M) etc, instead <strong>of</strong> H ∗ (K(−); M(−))<br />

etc, for the functors in above lemmas.<br />

Proposition 4.4.4. Given a functor M ∈ RE-mod, there<br />

are two spectral sequences associated with an extension K →<br />

E → C as follows:<br />

1. a homology spectral sequence<br />

E 2 ij = Hi(C; Hj(K; M)) ⇒ Hi+j(E; M);<br />

<strong>and</strong><br />

2. a cohomology spectral sequence<br />

E ij<br />

2 = H i (C op ; H j (K op ; M op )) ⇒ H i+j (E op ; M op ).<br />

Dually we have results for opposite extensions. Recall that<br />

RE op -mod = mod-RE. If M ∈ RE-mod, it gives M op ∈ RE op -<br />

mod.<br />

Proposition 4.4.5. Given a functor M ∈ RE-mod, there<br />

are two spectral sequences associated with an opposite extension<br />

K → E → C as follows:


178 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

1. a homology spectral sequence<br />

E 2 ij = Hi(C op ; Hj(K op ; M op )) ⇒ Hi+j(E op ; M op );<br />

<strong>and</strong><br />

2. a cohomology spectral sequence<br />

E ij<br />

2 = H i (C; H j (K; M)) ⇒ H i+j (E; M).<br />

Spectral sequences naturally give rise to some long exact<br />

sequences, <strong>and</strong> we record them below.<br />

Remark 4.4.6. If K → E → C is an extension <strong>and</strong> M ∈ REmod,<br />

one can obtain two five term exact sequences<br />

H2(E; M) → H2(C; M) → H0(C; H1(K; M)) → H1(E; M) → H1(C; M<br />

<strong>and</strong><br />

0 → H 1 (C op ; M op ) → H 1 (E op ; M op ) → H 0 (C op ; H 1 (K op ; M op )) → H 2 (C o<br />

When K → E → C is a group extension then these two exact<br />

sequences are the usual five term sequence in group homology<br />

<strong>and</strong> cohomology.<br />

Similarly if K → E → C is an opposite extension, we have<br />

H2(E op ; M op ) → H2(C op ; M op ) → H0(C; H1(K op ; M op )) → H1(E op ; M op<br />

<strong>and</strong><br />

0 → H 1 (C; M) → H 1 (E; M) → H 0 (C; H 1 (K; M)) → H 2 (C; M) → H 2 (E<br />

In general the finite generation <strong>of</strong> cohomology rings <strong>of</strong> both<br />

K <strong>and</strong> C does not guarantee the cohomology ring <strong>of</strong> E has the<br />

same property. One <strong>of</strong> the examples, C 2 , we used to demonstrate<br />

that the cohomology rings <strong>of</strong> EI-categories are not<br />

finitely generated, is an extension <strong>of</strong> a contractible category:


4.4 Grothendieck spectral sequences 179<br />

1 y<br />

<br />

y<br />

h 2<br />

<br />

y<br />

h 2<br />

<br />

y<br />

ι π<br />

α<br />

<br />

α .<br />

x<br />

h 2<br />

<br />

x<br />

h 2<br />

<br />

However when K is cohomologically trivial, the cohomology<br />

rings <strong>of</strong> E <strong>and</strong> C are isomorphic.<br />

Corollary 4.4.7. Suppose K → E → C is an extension, <strong>and</strong><br />

|K(x)| is invertible in R for every object x. Then for any<br />

M ∈ RE-mod<br />

H∗(E; M) ∼ = H∗(C; H0(K; M)) ∼ = H∗(C; −→K<br />

lim M).<br />

H ∗ (E op ; M) ∼ = H ∗ (C op ; H 0 (K op ; M)) ∼ = H ∗ (C op ; ←−K<br />

lim<br />

opM).<br />

Since lim ←−K<br />

opR ∼ = R in RC op -mod, we have H ∗ (C; R) ∼ =<br />

H ∗ (C op ; R) ∼ = H ∗ (E op ; R) ∼ = H ∗ (E; R) as algebras.<br />

Pro<strong>of</strong>. Under the assumption, the E 2 (resp. E 2 ) page <strong>of</strong> the<br />

cohomology (resp. homology) spectral sequence collapses to<br />

the vertical (resp. horizontal) axis.<br />

⊓⊔<br />

Let K → E → C be an extension. Since there is a natural<br />

correspondence between the subcategories <strong>of</strong> C <strong>and</strong> those <strong>of</strong><br />

E, one would like to exploit further connections between the<br />

homological properties <strong>of</strong> C <strong>and</strong> E. Let D be a subcategory<br />

<strong>of</strong> C <strong>and</strong> E D its “preimage” in E. We show the undercategories<br />

(or overcategories) associated with the inclusions are<br />

equivalent, when K → E → C is an extension (or an opposite<br />

extension) <strong>of</strong> C.<br />

Definition 4.4.8. Let K → E → C be an extension <strong>and</strong><br />

D ⊂ C a subcategory. The subextension <strong>of</strong> D in E via K| D ,<br />

x<br />

1 x


180 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

named E D , is a subcategory <strong>of</strong> E whose object set is the same<br />

as D <strong>and</strong> whose morphism set consists <strong>of</strong> morphisms in E<br />

which are preimages <strong>of</strong> morphisms in D.<br />

If D is a full subcategory <strong>of</strong> C then E D is a full subcategory <strong>of</strong><br />

E. Given an extension K → E → C, Aut K (x) → Aut E (x) →<br />

Aut C (x) is a subextension for any x ∈ Ob C.<br />

Proposition 4.4.9. Let K → E → C a sequence <strong>of</strong> functors<br />

<strong>and</strong> D a full subcategory <strong>of</strong> C with the inclusion ι D : D → C.<br />

Then<br />

1. if E is an extension <strong>of</strong> C, K| D → E D → D is the subextension<br />

<strong>and</strong> ι ED : E D → E is the inclusion, then for any<br />

y ∈ Ob C = Ob E, the undercategory y\ι D is isomorphic to<br />

a subcategory <strong>of</strong> the undercategory y\ι ED , which is equivalent<br />

to y\ι ED ;<br />

2. if E is an opposite extension <strong>of</strong> C, K| D → E D → D is<br />

the opposite subextension <strong>and</strong> ι ED : E D → E is the inclusion,<br />

then for any y ∈ Ob C = Ob E, the overcategory ι D /y<br />

is isomorphic to a subcategory <strong>of</strong> the overcategory ι ED /y,<br />

which is equivalent to ι ED /y.<br />

Pro<strong>of</strong>. We will only prove (2). In ι ED /y, any two objects (x, α)<br />

<strong>and</strong> (x, β) are isomorphic if <strong>and</strong> only if π(α) = π(β). Let<br />

ι ED /y ⊂ ι ED /y be the full subcategory consisting <strong>of</strong> one object<br />

from each isomorphism class <strong>of</strong> objects described above.<br />

Then ι ED /y <strong>and</strong> ι ED /y are equivalent. We prove the former is<br />

isomorphic to ι D /y.<br />

There is a natural bijection between objects sets <strong>of</strong> these<br />

two categories (x, α) → (x, π(α)) (π is surjective on morphisms).<br />

We show there is a bijection between the morphism<br />

sets <strong>and</strong> the bijections extend to a functor which gives an<br />

isomorphism between two categories. Any (x, α) →(z, γ β) in<br />

Mor(ι ED /y) gives rise to a morphism (x, π(α)) π(γ)<br />

→ (z, π(β)) in


4.4 Grothendieck spectral sequences 181<br />

ι D /y. On the other h<strong>and</strong>, a morphism (x, π(α)) π(γ)<br />

→ (z, π(β))<br />

in ι D /y implies π(β)π(γ) = π(α), which means there exists a<br />

unique g ∈ K(x) such that βγ = αg. Thus we have a uniquely<br />

defined morphism (x, α) g−1<br />

→(x, αg) →(z, γ β) = (x, α) γg−1<br />

→ (z, β)<br />

in Mor(ι ED /y). Note that a different γ ′ such that π(γ ′ ) = π(γ)<br />

gives the same morphism (x, α) γg−1<br />

→ (z, β), so the map from<br />

Mor(ι D /y) to Mor(ι ED /y) is well-defined. It’s straightforward<br />

to check these two assignments on morphisms are mutually<br />

inverse to each other.<br />

In order to show the bijections on objects <strong>and</strong> morphisms<br />

defining an isomorphism between categories, we need to verify<br />

they preserve composition <strong>and</strong> identity. We will just prove<br />

the former <strong>and</strong> leave the pro<strong>of</strong> <strong>of</strong> preserving identity to the<br />

reader. Suppose (x, α) → (z, β) → (w, γ) is a composite<br />

<strong>of</strong> two morphisms in ι ED /y. Then our map naturally sends<br />

it to a composite <strong>of</strong> morphisms (x, π(α)) → (z, π(β)) →<br />

(w, π(γ)). Conversely, if (x, π(α)) π(u)<br />

→ (z, π(β)) π(v)<br />

→ (w, π(γ)) =<br />

(x, π(α)) π(vu)<br />

→ (w, π(γ)) is the composite <strong>of</strong> two morphisms in<br />

ι D /y, then we need to show the two morphisms (x, α) ug−1<br />

→ (z, β) vh−1<br />

→ (w, γ<br />

= (x, α) vh−1 ug<br />

→<br />

−1<br />

(w, γ) <strong>and</strong> (x, α) vut−1<br />

→ (w, γ) are equal, where<br />

g, h, t are isomorphisms, described in the preceding paragraph.<br />

Since π(vh −1 ug −1 ) = π(vut −1 ), there is a unique isomorphism<br />

s satisfying vh −1 ug −1 = vut −1 s. But then we have<br />

α = γvut −1 = γvh −1 ug −1 = γvut −1 s, <strong>and</strong> this forces s = 1<br />

because Definition 4.1.21 (4 op ). Hence we get vh −1 ug −1 =<br />

vut −1 .<br />

⊓⊔<br />

The following corollary is a useful outcome <strong>of</strong> the proposition.


182 4 <strong>Cohomology</strong> <strong>of</strong> categories <strong>and</strong> modules<br />

Corollary 4.4.10. Let K → E → C be a sequence <strong>of</strong><br />

functors <strong>and</strong> D ⊂ C a full subcategory with the inclusion<br />

ι D : D → C. Then<br />

1. if E is an extension <strong>of</strong> C, then y\ι D is contractible (or<br />

R-acyclic or connected) if <strong>and</strong> only if y\ι ED is;<br />

2. if E is an opposite extension <strong>of</strong> C, then ι D /y is contractible<br />

(or R-acyclic or connected) if <strong>and</strong> only if ι ED /y is.<br />

Example 4.4.11. Let K → E → C be an extension with a<br />

unique maximal object x such that Aut C (x) acts freely <strong>and</strong><br />

transitively on Hom C (y, x) for any y ∈ Ob C. Then it’s easy<br />

to check that ι : Aut C (x) ↩→ C induces a homotopy equivalence<br />

since all undercategories associated to it are contractible.<br />

Hence we know Aut E (x) ↩→ E is a homotopy equivalence as<br />

well.<br />

Since any category can be regarded as a trivial extension<br />

<strong>of</strong> itself, the following result is some sort <strong>of</strong> generalization <strong>of</strong><br />

Corollary 4.3.4.<br />

Corollary 4.4.12. Suppose there is an extension K → E →<br />

C. If ι D : D ↩→ C is an inclusion such that y\ι D is contractible<br />

for every y ∈ Ob C, then H ∗ (E; M) ∼ = H ∗ (E D ; M)<br />

for any M ∈ RE op -mod, <strong>and</strong> H∗(E; N) ∼ = H∗(E D ; N) for any<br />

N ∈ RE-mod. Here E D is the subextension corresponding to<br />

D.<br />

Suppose there is an opposite extension K → E → C. If<br />

ι D : D ↩→ C is an inclusion such that ι D /y is contractible for<br />

every y ∈ Ob C, then H ∗ (E; M) ∼ = H ∗ (E D ; M) for any M ∈<br />

RE-mod, <strong>and</strong> H∗(E; N) ∼ = H∗(E D ; N) for any N ∈ RE op -<br />

mod. Here E D is the opposite subextension corresponding to<br />

D.<br />

Pro<strong>of</strong>. We prove the statements for cohomology. Since y\ι D<br />

is contractible for every y ∈ Ob C, y\ι ED is contractible for


4.4 Grothendieck spectral sequences 183<br />

every y ∈ Ob E as well by Corollary 4.4.10. Then we apply<br />

Corollary 4.3.4.<br />

When we have an opposite extension, using Corollary 4.4.10<br />

again we get that ι ED /y is contractible for every y ∈ Ob E.<br />

Hence Corollary 4.3.4 can be used to obtain the isomorphism.<br />

⊓⊔<br />

As an example when K → E → C is an extension (resp.<br />

an opposite extension) <strong>and</strong> C has a unique maximal (resp.<br />

minimal) object x <strong>and</strong> Aut C (x) acts regularly on Hom C (y, x)<br />

(or Hom C (x, y)) for any y ∈ Ob C, we have H ∗ (C; M) ∼ =<br />

H ∗ (Aut C (x), M(x)) hence H ∗ (E; M) ∼ = H ∗ (Aut E (x), M(x))<br />

for any contra-variant (resp. covariant) functor M.


Chapter 5<br />

Hochschild cohomology<br />

Abstract We consider Hochschild cohomology <strong>of</strong> category<br />

algebras here. The upshot is that we can translate Hochschild<br />

cohomology into functor cohomology, which provide a context<br />

for comparing Hochschild <strong>and</strong> ordinary cohoomology <strong>of</strong><br />

a small category. A very general theorem on the relationship<br />

<strong>of</strong> these two cohomology theories will be stated <strong>and</strong> proved,<br />

combining module theoretic <strong>and</strong> simplicial methods. Some<br />

examples are computed to help the reader to underst<strong>and</strong>.<br />

5.1 Hochschild homology <strong>and</strong> cohomology<br />

Let A be an associative R-algebra with identity. Then we<br />

consider A as an A e -module. We shall assume A is a finitely<br />

presented flat R-module. This implies that A is a projective<br />

R-module.<br />

5.1.1 Definition <strong>and</strong> general properties<br />

Definition 5.1.1. The acyclic Hochschild complex <strong>of</strong> A is<br />

{B A n } n≥−1 such that<br />

B A n = A ⊗ R · · · ⊗ R A<br />

} {{ }<br />

(n+2)−copies<br />

<strong>and</strong> such that the differential ∂ n : A ⊗n → A ⊗(n−1) is given by


186 5 Hochschild cohomology<br />

∂ n (a 0 ⊗ a 1 ⊗ · · · ⊗ a n ) ↦→<br />

n∑<br />

(−1) i a 0 ⊗ · · · ⊗ a i a i+1 ⊗ · · · ⊗ a n .<br />

i=0<br />

By direct calculations, one can see ∂ 2 = 0 <strong>and</strong> so we do have<br />

a complex <strong>of</strong> A e -modules. Moreover since we can establish a<br />

chain contraction s n : B A n−1 → B A n by s n (a 0 ⊗ · · · ⊗ a n ) =<br />

a 0 ⊗ · · · ⊗ a n ⊗ 1, this complex is exact.<br />

Suppose M ∈ A e -mod. Then we can regard it as a right A e -<br />

module by m·(a, a ′ ) := a ′·m·a. Thus we shall not distinguish<br />

left <strong>and</strong> right A e -modules in this chapter.<br />

Definition 5.1.2. For any M ∈ A e -mod, we define<br />

1. the Hochschild homology HH ∗ (A, M) to be the homology <strong>of</strong><br />

· · · → M ⊗ A<br />

e B A n → · · · → M ⊗ A<br />

e B A 1 → M ⊗ A<br />

e B A 0 → 0;<br />

<strong>and</strong><br />

2. the Hochschild cohomology HH ∗ (A, M) to be the homology<br />

<strong>of</strong><br />

0 → Hom A e(B A 0 , M) → Hom A e(B A 1 , M) → · · · → Hom A e(B A n , M) →<br />

Remark 5.1.3. Since for any n ≥ 0 there is an isomorphism<br />

<strong>of</strong> A e -modules<br />

B A n → A e ⊗ R ˜BA n ,<br />

in which ˜B A 0 = R <strong>and</strong> ˜B A n for n ≥ 1 is<br />

}<br />

A ⊗ R ·<br />

{{<br />

· · ⊗ R A<br />

}<br />

,<br />

n−copies<br />

<strong>and</strong> furthermore because ˜B A n is a projective R-module B A n is<br />

a projective A e -module, we actually obtain a complex <strong>of</strong> A e -<br />

modules with A in degree −1<br />

· · · → B A n → · · · → B A 1 → B A 0 → A → 0.


5.1 Hochschild homology <strong>and</strong> cohomology 187<br />

We call B A ∗ the bar resolution <strong>of</strong> the A e -module A. Note that<br />

M ⊗ A eBn A ∼ = M ⊗ R ˜BA n <strong>and</strong> Hom A e(Bn A , M) ∼ = Hom R (˜B n A , M).<br />

Because <strong>of</strong> our assumption on A we can also define Hochschild<br />

homology <strong>and</strong> cohomology as Tor Ae<br />

n (M, A) <strong>and</strong> Ext n Ae(A, M),<br />

respectively. When M = A, we usually write HH n (A) <strong>and</strong><br />

HH n (A) for the Hochschild homology <strong>and</strong> cohomology.<br />

When we regard the bar resolution B∗<br />

A → A → 0 as a<br />

complex <strong>of</strong> right A-modules, it splits. Hence for any left A-<br />

module M, Bn A ⊗ A M → A⊗ A M → 0 is a projective resolution<br />

<strong>of</strong> the left A-module M. Based on this observation, we prove<br />

the following well-known results [14].<br />

Proposition 5.1.4. Let M, N ∈ A-mod <strong>and</strong> M ′ ∈ mod-A.<br />

Suppose R is a field. Then Hom R (M, N) <strong>and</strong> N ⊗ R M ′ are<br />

A e -modules such that<br />

Ext ∗ A(M, N) ∼ = H ∗ (A, Hom R (M, N)) ∼ = Ext ∗ A e(A, Hom R(M, N)).<br />

<strong>and</strong><br />

Tor A ∗ (M ′ , N) ∼ = H∗(A, N ⊗ R M ′ ) ∼ = Tor Ae<br />

∗ (N ⊗ R M ′ , A).<br />

Pro<strong>of</strong>. We only prove the isomorphism for Ext. The A e -<br />

module structure on Hom R (M, N) is given by (a 1 , a 2 )f(m) =<br />

a 1 f(a 2 m) for any (a 1 , a 2 ) ∈ A e <strong>and</strong> m ∈ M. Let Bn<br />

A → A<br />

be the previously introduced projective resolution <strong>of</strong> the A e -<br />

module A. Then<br />

Hom A e(B A n , Hom R (M, N)) = Hom R (˜B A n , Hom R (M, N))<br />

∼ = HomR (˜B A n ⊗ R M, N)<br />

∼ = HomA (B A n ⊗ A M, N).


188 5 Hochschild cohomology<br />

Since R is a field, Bn A ⊗ A M is projective <strong>and</strong> thus B∗ A ⊗ A M<br />

becomes a projective resolution <strong>of</strong> M. Consequently the last<br />

term above computes Ext ∗ A(M, N).<br />

⊓⊔<br />

Let B∗<br />

A → A → 0 be the bar resolution <strong>of</strong> the A e -module<br />

A. There is a chain map D : B∗ A → B∗ A ⊗ A B∗ A given by<br />

n∑<br />

D(a 0 ⊗a 1 ⊗· · ·⊗a n+1 ) = (a 0 ⊗· · · a i ⊗1)⊗(1⊗a i+1 ⊗· · ·⊗a n+1 ).<br />

i=0<br />

Let ζ ∈ H n (A) <strong>and</strong> ζ ′ ∈ H m (A). Then the cup product ζ ∪ ζ ′<br />

is given by<br />

B∗<br />

A D<br />

→B∗ A ⊗ A B∗<br />

A ζ⊗ζ<br />

−→A ′<br />

⊗ A A = A.<br />

More explicitly ζ ∪ ζ ′ : Bn+m A → A is given by<br />

(ζ∪ζ ′ )(a 0 ⊗a 1 ⊗· · ·⊗a n+m+1 ) = ζ(a 0 ⊗· · · a n ⊗1)ζ ′ (1⊗a n+1 ⊗· · ·⊗a n+m+<br />

Based on previous construction, HH ∗ (A) forms a ring, called<br />

the Hochschild cohomology ring <strong>of</strong> A. The multiplicative<br />

structure was known to Yoneda when R is a field.<br />

By [67], one can use an arbitrary projective resolution P ∗ →<br />

A → 0 to construct the cup product, <strong>and</strong> there exists a unique<br />

chain map D : P ∗ → P ∗ ⊗ A P ∗ such that (∂ 0 ⊗ ∂ 0 )D = ∂ 0 . In<br />

fact one just have to show that P ∗ ⊗ A P ∗ → A → 0 is also a<br />

projective resolution.<br />

Theorem 5.1.5 (Gerstenhaber). With the above cup product<br />

the Hochschild cohomology ring<br />

HH ∗ (A) = ⊕ i≥0<br />

HH i (A)<br />

is graded commutative. Moreover the cup product coincides<br />

with the Yondea splice.


5.1 Hochschild homology <strong>and</strong> cohomology 189<br />

Pro<strong>of</strong>. Let 0 → A → L n−1 → · · · → L 0 → A → 0 <strong>and</strong><br />

0 → A → M m−1 → · · · → M 0 → A → 0 be two n-fold <strong>and</strong><br />

m-fold extensions. They uniquely determine two cohomology<br />

classes by liftings<br />

· · ·<br />

<br />

B A n<br />

B A n−1<br />

· · ·<br />

<br />

B A 0<br />

A 0<br />

<strong>and</strong><br />

ζ=ζ n <br />

ζ n−1<br />

<br />

0<br />

<br />

A = L n<br />

<br />

L n−1<br />

<br />

· · ·<br />

<br />

L 0<br />

<br />

A 0<br />

· · ·<br />

<br />

B A m<br />

B A m−1<br />

ζ 0 <br />

· · ·<br />

<br />

B A 0<br />

A<br />

<br />

0<br />

ξ=ξ m <br />

ξ m−1<br />

<br />

0<br />

<br />

A = M m<br />

<br />

M m−1<br />

<br />

· · ·<br />

<br />

M 0<br />

<br />

A 0.<br />

The Yoneda splice ζ ∗ ξ is given by any lifting <strong>of</strong> ξ = ξ m<br />

· · ·<br />

<br />

Bn+m A Bn+m−1<br />

A · · ·<br />

<br />

Bm<br />

A<br />

θ n <br />

θ n−1 <br />

θ 0 ξ<br />

0<br />

<br />

A<br />

<br />

L<br />

n−1 · · ·<br />

<br />

L 0<br />

<br />

A<br />

ξ 0 <br />

B A m−1<br />

· · ·<br />

<br />

A 0<br />

<br />

M m−1<br />

<br />

· · ·<br />

<br />

A 0<br />

0 0 .<br />

To finish our pro<strong>of</strong>, we will define two explicit liftings. On one<br />

h<strong>and</strong>, the composition <strong>of</strong> the following chain maps<br />

θ ∗ ′ : B∗<br />

A D<br />

−→B∗ A ⊗ A B∗<br />

A Id⊗ξ<br />

−→B∗ A ⊗ A A[m] → B∗ A [m].<br />

provides a lifting <strong>of</strong> ξ by<br />

θ i ′ : Bi+m<br />

A D<br />

−→(B ∗ A ⊗ A B∗ A Id⊗ξ<br />

) i+m −→Bi<br />

A ⊗ A A[m] → Bi A [m] ξ i<br />

→L i ,<br />

for each 0 ≤ i ≤ n. Here [m] denotes a shift <strong>of</strong> the chain<br />

complex. From its definition we can check that ζ ∗ ξ = ζ ◦<br />

(Id ⊗ θ n) ′ ◦ D = ζ ∪ ξ.


190 5 Hochschild cohomology<br />

On the other h<strong>and</strong> we can define ξ ′′<br />

i : B A n+i → BA i<br />

ξ ′′<br />

i (a 0 ⊗· · ·⊗a n+i+1 ) = (−1) ni ξ(a 0 ⊗· · ·⊗a n ⊗1)⊗a n+1 ⊗· · ·⊗a n+i+1 .<br />

They give rise to another lifting <strong>of</strong> ξ. Then we have<br />

ζ ∗ ξ(a 0 ⊗ · · · ⊗ a n+m+1 ) = ζθ ′′ m(a 0 ⊗ · · · ⊗ a n+m+1 )<br />

= (−1) nm ζ(ξ(a 0 ⊗ · · · ⊗ a n ⊗ 1) ⊗ a n+1 ⊗ · ·<br />

= (−1) nm ξ(a 0 ⊗ · · · ⊗ a n ⊗ 1)ζ(a n+1 ⊗ · · · ⊗<br />

= (−1) nm (ξ ∪ ζ)(a 0 ⊗ · · · ⊗ a n+m+1 ).<br />

The pro<strong>of</strong> is taken from [70].<br />

Theorem 5.1.6 (Dennis,). If A <strong>and</strong> B are Morita equivalent,<br />

then HH ∗ (A) ∼ = HH ∗ (B) <strong>and</strong> there is a ring isomorphism<br />

HH ∗ (A) ∼ = HH ∗ (B).<br />

Pro<strong>of</strong>. The pro<strong>of</strong> to the general situation is not difficult, but<br />

one needs to know the functors realizing the given equivalence<br />

between two module categories. See []<br />

⊓⊔<br />

In the situation <strong>of</strong> category algebras, if D ≃ C are two categories,<br />

then the Hochschild (co)homology <strong>of</strong> RD <strong>and</strong> RC are<br />

isomorphic. These isomorphisms can be seen after we express<br />

Hochschild (co)homology as certain ordinary (co)homology.<br />

Hochschild homology <strong>and</strong> cohomology are important invariants<br />

<strong>of</strong> rings. However they are very difficult to compute. The<br />

main result in this chapter is that when A = RC is a category<br />

algebra, we can express HH ∗ (RC) by ordinary cohomology,<br />

<strong>and</strong> this allows effective calculation.<br />

by<br />

⊓⊔<br />

5.1.2 Ring homomorphisms from the Hochschild cohomology ring<br />

In general, if A <strong>and</strong> B are two associative k-algebras <strong>and</strong> M is<br />

a A ⊗ k B op -module, or equivalently a A-B-bimodule, we can


5.1 Hochschild homology <strong>and</strong> cohomology 191<br />

define a ring homomorphism induced by the tensor product<br />

− ⊗ A M<br />

φ M : Ext ∗ A e(Λ, Λ) → Ext∗ A⊗ R Bop(M, M).<br />

Let 0 → A → L n−1 → P n−2 → · · · → P 0 → A → 0 represent<br />

an element in Ext n A e(A, A). We may assume P i are projective<br />

A e -modules. Then considered as a complex <strong>of</strong> right<br />

A-modules, it is split exact. Thus tensoring with M gives an<br />

exact sequence <strong>of</strong> A-B-modules<br />

0 → M → L n−1 ⊗ A M → P n−2 ⊗−AM → · · · → P 0 ⊗ A M → M → 0.<br />

This induces a ring homomorphism φ M : Ext ∗ Ae(A, A) →<br />

Ext ∗ A⊗ R B op(M, M). If N is another A ⊗ R B op -module, we<br />

see Ext ∗ A⊗ R B op(M, N) has an Ext∗ Ae(A, A)-module structure<br />

via the ring homomorphisms φ M <strong>and</strong> φ N together with the<br />

Yoneda splice. We quote the following theorem <strong>of</strong> Snashall<br />

<strong>and</strong> Solberg [], which generalizes Gerstenhaber’s theorem.<br />

Theorem 5.1.7. Let A <strong>and</strong> B be two associative R-algebras.<br />

Let η be an element in Ext n Ae(A, A) <strong>and</strong> θ an element in<br />

Ext m A⊗ R Bop(M, N) for two A-B-bimodules M <strong>and</strong> N. Then<br />

φ N (η)θ = (−1) mn θφ M (η).<br />

Pro<strong>of</strong>. Let us fix an element η for n ≥ 1<br />

0 → A → L n−1 → P n−2 → · · · → P 0 → A → 0<br />

with P i projective.<br />

First we consider the case for θ ∈ Hom A⊗B op(M, N) (m = 0).<br />

When n = 0, Ext 0 Ae(A, A) = Z(A). It means each element in<br />

Ext 0 Ae(A, A) is a multiplication by some a ∈ Z(A). Such a<br />

map induces an element in Hom A⊗B op(M, M) by m ↦→ am.<br />

One can easily verify that (η ⊗ A N)θ = θ(η ⊗ A M).<br />

When n = 1, we can construct the following commutative<br />

diagram


192 5 Hochschild cohomology<br />

0<br />

<br />

A ⊗ A M L 0 ⊗ A M<br />

A⊗θ <br />

L 0 ⊗θ<br />

A ⊗ A M 0<br />

<br />

0 A ⊗ A N X A ⊗ A M<br />

A⊗θ<br />

0<br />

<br />

A ⊗ A N<br />

<br />

L 0 ⊗ A N<br />

<br />

A ⊗ A N 0<br />

where the middle row is (η ⊗ A N)θ = θ(η ⊗ A M).<br />

Next let θ ∈ Ext 1 A⊗ R Bop(M, N) represented by 0 → N →<br />

Y → M → 0. Since all syzygies <strong>of</strong> A are projective as right<br />

A-modules, we have the following commutative diagram<br />

0<br />

<br />

Ω i (A) ⊗ A N<br />

<br />

P i−1 ⊗ A N<br />

0<br />

Ω i−1 (A) ⊗ A N<br />

0<br />

0<br />

<br />

Ω i (A) ⊗ A Y<br />

<br />

P i−1 ⊗ A Y<br />

Ω i−1 (A) ⊗ A Y<br />

0<br />

0<br />

<br />

Ω i (A) ⊗ A M P i−1 ⊗ A M Ω i−1 (A) ⊗ A M 0<br />

for all i with Ω 0 (A) = A. Denote the upper row by σ i , the<br />

rightmost column by θ i , the leftmost column by θ i+1 <strong>and</strong> the<br />

lower row by σ ′ i. Then by the 3 × 3-splice <strong>of</strong> [52, Lemma 3.3],<br />

we get σ i θ i = −θ i+1 σ ′ i for all i ≥ 1. Since<br />

η⊗ A N = (0 → A⊗ A N → L 0 ⊗ A N → Ω n−1 (A)⊗ A N → 0)σ n−1 · · · σ 1<br />

<strong>and</strong><br />

η⊗ A M = (0 → A⊗ A M → L 0 ⊗ A M → Ω n−1 (A)⊗ A M → 0)σ ′ n−1 · · · σ<br />

we obtain an equality (η ⊗ A N)θ = (−1) n θ(η ⊗ A M) by combining<br />

all above.<br />

When m > 1, since every θ ∈ Ext m Ae(M, N) is the Yoneda<br />

splice <strong>of</strong> m extensions, it follows directly that (η ⊗ A N)θ =<br />

(−1) mn θ(η ⊗ A M). We are done.<br />

⊓⊔


5.2 Hochschild (co)homology <strong>of</strong> category algebras 193<br />

5.2 Hochschild (co)homology <strong>of</strong> category algebras<br />

5.2.1 Basic ideas <strong>and</strong> examples<br />

Let C be a small category. Recall from Section 3.1.1 that we<br />

call C e = C × C op the enveloping category <strong>of</strong> a small category<br />

C. We also showed in Example 3.1.5 that there is a natural isomorphism<br />

kC e ∼ = (kC) e . As a functor, kC(x, y) = kHom C (y, x)<br />

if Hom C (y, x) ≠ ∅ <strong>and</strong> kC(x, y) = 0 otherwise. Here (x, y) ∈<br />

Ob C e . This result is just a simple observation. It implies the<br />

enveloping algebra <strong>of</strong> a category algebra <strong>of</strong> C is the category<br />

algebra <strong>of</strong> its enveloping category, so later on we will just use<br />

the terminology kC e when dealing with Hochschild cohomology.<br />

This identification enables us to apply functor cohomology<br />

theory to the investigation <strong>of</strong> the Hochschild cohomology<br />

theory <strong>of</strong> category algebras.<br />

A key ingredient is F (C), the category <strong>of</strong> factorizations in<br />

C. Recall that the category F (C) has the morphisms in C as<br />

its objects. In order to avoid confusion, we write an object in<br />

F (C) as [α], whenever α ∈ Mor C. A morphism from [α] ∈<br />

Ob F (C) to [α ′ ] ∈ Ob F (C) is given by a pair <strong>of</strong> u, v ∈ Mor C,<br />

making the following diagram commutative<br />

x<br />

u <br />

α<br />

y<br />

x ′ y ′ .<br />

α ′<br />

In other words, there is an morphism from [α] to [α ′ ] if <strong>and</strong><br />

only if α ′ = uαv for some u, v ∈ Mor C, or equivalently α is a<br />

factor <strong>of</strong> α ′ in Mor C. The category F (C) admits two natural<br />

covariant functors to C <strong>and</strong> C op<br />

v op<br />

C<br />

t s<br />

F (C) <br />

C op ,


194 5 Hochschild cohomology<br />

where t <strong>and</strong> s send an object [α] to its target <strong>and</strong> source,<br />

respectively. Using his Theorem A <strong>and</strong> its corollary, Quillen<br />

showed these two functors induce homotopy equivalences <strong>of</strong><br />

the classifying spaces. We will be interested in the functor<br />

∇ = (t, s) : F (C) → C e = C × C op ,<br />

sending an [α] ∈ Ob F (C) to (x, y) ∈ Ob C e if α ∈ Hom C (y, x)<br />

<strong>and</strong> a morphism (u, v op ) ∈ Mor F (C) to (u, v op ) ∈ Mor(C e ).<br />

The importance <strong>of</strong> the functor ∇ : F (C) → C e lies in the<br />

fact that its target category gives rise to the Hochschild cohomology<br />

ring <strong>of</strong> C, while its source category determines the<br />

ordinary cohomology ring <strong>of</strong> C ≃ F (C). In the situation <strong>of</strong><br />

(finite) posets <strong>and</strong> groups, the functor is well-understood <strong>and</strong><br />

in the group case it has been implicitly used to establish the<br />

homomorphism from the Hochschild cohomology ring to the<br />

ordinary cohomology ring.<br />

Example 5.2.1.1. When C is a poset, ∇ : F (C) → C e sends<br />

F (C) isomorphically onto a full category C e ∆ ⊂ Ce , where<br />

Ob C e ∆ = {(x, y) ∈ Ob C e ∣ ∣ Hom C (y, x) ≠ ∅}<br />

(the full subcategory C∆ e is well-defined whenever C is EI).<br />

One can easily see that RC as a functor only takes nonzero<br />

values at objects in Ob C∆ e . Furthermore as a RCe ∆ -<br />

module, RC ∼ = R is the trivial module. Since C∆ e ∼ = F (C)<br />

is a co-ideal in the poset C e , we obtain Ext ∗ RC e(RC, RC) ∼ =<br />

Ext ∗ RC (RC, RC) ∼ = Ext ∗<br />

∆ e RF (C)(k, k) ∼ = Ext ∗ kC(k, k), where the<br />

last isomorphism comes from the fact that BF (C) ≃ BC.<br />

This isomorphism between the two cohomology rings was<br />

first established in [30];<br />

2. When C is a group, the category F (C) is a groupoid <strong>and</strong> is<br />

equivalent to a subcategory <strong>of</strong> the one object category C e<br />

with morphism set


5.2 Hochschild (co)homology <strong>of</strong> category algebras 195<br />

{(g, g −1op ) ∣ ∣ g ∈ Mor C} ⊂ Mor C e .<br />

Based on this description, one can prove the existence <strong>of</strong> the<br />

surjective homomorphism from the Hochschild cohomology<br />

ring to the ordinary cohomology ring <strong>of</strong> a group, which is<br />

basically the same as the classical approach. See for example<br />

[].<br />

5.2.2 Hochschild (co)homology as ordinary (co)homology<br />

The two examples <strong>of</strong> last section hint that we may express<br />

Hochschild (co)homology in terms <strong>of</strong> ordinary (co)homology.<br />

Moreover they show us the way to establish such an expression.<br />

Let us examine the following commutative diagram <strong>of</strong><br />

small categories<br />

F (C)<br />

t<br />

∇<br />

<br />

C e = C × C op<br />

p<br />

C ,<br />

where p is the projection onto the first component. Recall from<br />

Section 3.2.2 that the preceding diagram leads to another<br />

R<br />

RF (C)-mod<br />

LK ∇<br />

RC e -mod<br />

RC<br />

LK t<br />

LK p<br />

∼ =−⊗RC R<br />

RC-mod .<br />

R<br />

In the rest <strong>of</strong> this section, we will establish <strong>and</strong> describe<br />

the following maps, induced by the three left Kan extensions<br />

LK t , LK p <strong>and</strong> LK ∇ respectively,


196 5 Hochschild cohomology<br />

t ∗ : Ext ∗ RF (C)(R, R) → Ext ∗ RC(R, R),<br />

p ∗ : Ext ∗ RC e(RC, RC) → Ext∗ RC(R, R)<br />

∇ ∗ : Ext ∗ RF (C)(R, R) → Ext ∗ RCe(RC, RC).<br />

Theorem 5.2.2. Let C be a small category. For any functor<br />

M ∈ RC e -mod, we have<br />

Ext ∗ RC e(RC, M) ∼ = Ext ∗ RF (C)(R, Res ∇ M).<br />

The maps t ∗ is an ring isomorphism, p ∗ ∼ = φ C is a split<br />

surjective ring homomorphism <strong>and</strong> ∇ ∗ is a split injective<br />

ring homomorphism.<br />

The theorem is proved in a series <strong>of</strong> lemmas.<br />

Lemma 5.2.3. The following complex <strong>of</strong> RC e -module<br />

LK ∇ B F (C)<br />

∗ → LK ∇ R → 0<br />

is a projective resolution <strong>of</strong> the RC e -module RC.<br />

Pro<strong>of</strong>. Let B∗<br />

F (C) → R → 0 be the bar resolution. By Proposition<br />

4.3.2, LK ∇ B∗<br />

F (C) ∼ = C∗ (∇/−, R). In Proposition 2.3.13,<br />

due to Quillen, we asserted that the category F (C) is a c<strong>of</strong>ibred<br />

category over C e . More explicitly for any (x, y) ∈ Ob C e<br />

the overcategory ∇/(x, y) is homotopy equivalent to the fibre<br />

∇ −1 (x, y), which is the discrete category Hom C (y, x). In other<br />

words, for any (x, y) ∈ Ob C e ,<br />

is exact. Thus<br />

C ∗ (∇/(x, y), R) → RHom C (y, x) → 0<br />

C ∗ (∇/−, R) → RC → 0<br />

is a projective resolution <strong>of</strong> the RC e -module RC. Furthermore<br />

we have LK ∇ R ∼ = RC by direct calculation<br />

(LK ∇ R)(x, y) = lim −→∇/(x,y)<br />

R ∼ = lim −→HomC (y,x) R ∼ = RHom C (y, x).


5.2 Hochschild (co)homology <strong>of</strong> category algebras 197<br />

⊓⊔<br />

For any functor M ∈ RC e -mod, the isomorphism<br />

Ext ∗ RC e(RC, M) ∼ = Ext ∗ RF (C)(R, Res ∇ M)<br />

is a direct consequence <strong>of</strong> Lemma 5.2.3.<br />

Before we study the ring homomorphisms, we give a interesting<br />

result on Res ∇ RC. For an arbitrary u : D → C, since<br />

LK u is the left adjoint <strong>of</strong> Res u , there are natural transformations<br />

Id → Res u LK u <strong>and</strong> LK u Res u → Id. We pay attention<br />

to the case <strong>of</strong> ∇ : F (C) → C e . There exists an RF (C)-<br />

homomorphism R → Res ∇ LK ∇ (R) = Res ∇ (RC) as well as<br />

a RC e -homomorphism RC = LK ∇ Res ∇ (R) → R. The latter<br />

gives rise to a RF (C)-homomorphism<br />

Res ∇ (RC) = Res ∇ LK ∇ Res ∇ (R) → R = Res ∇ R.<br />

In case that C is a poset, one has R = Res ∇ (RC). When C is<br />

a group, F (C) is a groupoid, equivalent to the automorphism<br />

group <strong>of</strong> [1 C ] ∈ Ob F (C), that is, {(g, g −1op ) ∣ g ∈ Mor C}.<br />

If we name the full subcategory <strong>of</strong> F (C), consisting <strong>of</strong> one<br />

object [1 C ], by ˜∆C <strong>and</strong> the inclusion (an equivalence) by i :<br />

˜∆C ↩→ F (C). Then Res ∇i (RC) = Res ∇ (RC)([1 C ]) is a R ˜∆Cmodule<br />

with the action (g, g −1op )·a = gag −1 , a ∈ Res ∇i (RC).<br />

Thus Res ∇i (RC) = ⊕ Rc g , where c g is the conjugacy class <strong>of</strong><br />

g ∈ Mor C. In particular R = Rc 1C is a direct summ<strong>and</strong> <strong>of</strong><br />

Res ∇i (RC) <strong>and</strong> it implies R ∣ Res ∇ (RC) as RF (C)-modules<br />

because i is an equivalence <strong>of</strong> categories.<br />

Lemma 5.2.4. Let C be a small category. Then R ∣ ∣ Res ∇ (RC)<br />

as RF (C)-modules.<br />

Pro<strong>of</strong>. One needs to keep in mind that the restriction <strong>of</strong> a<br />

module usually has a larger R-rank than the module itself<br />

since ∇ is not injective on objects. We define a RF (C)-


198 5 Hochschild cohomology<br />

homomorphism (a natural transformation) ι : R → Res ∇ (RC)<br />

by the assignments ι [α] (1 R ) = α ∈ Res ∇ (RC)([α]) for each<br />

[α] ∈ Ob F (C). If [β] is another object in Ob F (C) <strong>and</strong><br />

(u, v op ) ∈ Hom F (C) ([α], [β]) is an arbitrary morphism, then<br />

by the definition <strong>of</strong> an F (C)-morphism, (u, v op ) · α = uαv =<br />

β. Hence ι maps R isomorphically onto a submodule <strong>of</strong><br />

Res ∇ (RC). On the other h<strong>and</strong>, we may define a RF (C)-<br />

homomorphism ɛ : Res ∇ (RC) → R such that, for any<br />

[α] ∈ Ob F (C), ɛ [α] : Res ∇ (RC)([α]) → R([α]) = R sends<br />

each base element in Res ∇ (RC)([α]) = RHom C (y, x) to 1 R .<br />

One can readily check the composite <strong>of</strong> these two maps is the<br />

identity<br />

R ι →Res ∇ (RC) ɛ →R,<br />

<strong>and</strong> this means R ∣ Res ∇ (RC) or Res ∇ (RC) = R ⊕ N C for<br />

some RF (C)-module N C .<br />

⊓⊔<br />

The module N C as a functor can be described by<br />

N C ([α]) = R{β − γ ∣ ∣ β, γ ∈ Hom C (y, x)},<br />

if [α] ∈ Ob F (C) <strong>and</strong> α ∈ Hom C (y, x). It will be useful to our<br />

computation since it determines the “difference” between the<br />

Hochschild <strong>and</strong> ordinary cohomology rings <strong>of</strong> a category. The<br />

next lemma finishes <strong>of</strong>f our pro<strong>of</strong> <strong>of</strong> the main theorem.<br />

Using adjunction between LK ∇ <strong>and</strong> Res ∇ , along with Lemmas<br />

5.2.3 <strong>and</strong> 5.2.4, we get a commutative diagram<br />

Hom RF (C) (B∗ F (C) , R)<br />

splitting<br />

Hom RF (C) (B F (C)<br />

∗<br />

, Res ∇ RC)<br />

LK ∇<br />

∼ =<br />

<br />

Hom RC e(LK ∇ B F (C)<br />

∗ , LK ∇ R)<br />

<br />

Hom RC e(LK ∇ B∗<br />

F (C) , RC)<br />

The top row gives rise to ∇ ∗ . From this diagram one can see<br />

it can also obtained as


5.2 Hochschild (co)homology <strong>of</strong> category algebras 199<br />

Ext ∗ RF (C)(R, R) → Ext ∗ RF (C)(R, Res ∇ RC) ∼ = Ext ∗ RCe(RC, RC).<br />

Now we turn to establish the ring homomorphisms.<br />

Lemma 5.2.5. The map t ∗ : Ext ∗ RF (C)(R, R) → Ext ∗ RC(R, R)<br />

is a ring isomorphism.<br />

Pro<strong>of</strong>. As we explained in Remark 4.3.7, t ∗ is induced by<br />

LK t , <strong>and</strong> is the same as the map between simplicial/singular<br />

cohomology, induced by the functor t in Chapter 2. Since<br />

F (C) ≃ C, it is an isomorphism.<br />

⊓⊔<br />

Lemma 5.2.6. The map p ∗ : Ext ∗ RC e(RC, RC) → Ext∗ RC(R, R)<br />

is equal to φ C , induced by − ⊗ RC R.<br />

Pro<strong>of</strong>. Suppose M ∈ RC e -mod. We show LK p M ∼ = M ⊗ RC R<br />

as RC-modules. Let x ∈ Ob C. Then<br />

1 x · LK p M = (LK p M)(x) = lim −→p/x<br />

M.<br />

Because p/x ∼ = (Id C /x) × C op ≃ {(x, 1 x )} × C op , we have<br />

lim M ∼ = lim<br />

−→p/x −→IdC<br />

M<br />

/x×C<br />

∼ op = lim −→{(x,1x<br />

M<br />

)}×C<br />

∼ op = lim −→C opM(x, −)<br />

∼ = lim −→C<br />

op1 x · M<br />

∼ = R ⊗RC op (1 x · M)<br />

∼ = 1x · M ⊗ RC R.<br />

In particular it implies LK p (RC e ) ∼ = RC e ⊗ RC R ∼ = RC ⊗ R<br />

R. Thus if ˜P∗ → RC → 0 is the projective resolution in<br />

Section 5.1.1, it splits when regarded as a complex <strong>of</strong> right RCmodules.<br />

Consequently LK p<br />

∼ = − ⊗RC R maps any projective<br />

resolution <strong>of</strong> RC to an exact sequence <strong>of</strong> left RC-modules<br />

LK p ˜P∗ → LK p (RC) ∼ = R → 0,


200 5 Hochschild cohomology<br />

which is a projective resolution. Hence LK p induces a chain<br />

map<br />

Hom RC e( ˜P∗ , RC) → Hom RC (LK p ˜P∗ , R),<br />

<strong>and</strong> the ring homomorphism<br />

p ∗ = φ C : Ext ∗ RC e(RC, RC) → Ext∗ RC(R, R).<br />

Lemma 5.2.7. Let C be a small category. Then ∇ ∗ is a ring<br />

homomorphism <strong>and</strong> there is another ring homomorphism ɛ ∗<br />

Ext ∗ RC e(RC, RC) ↠ Ext∗ RF (C)(R, R),<br />

such that ɛ ∗ ∇ ∗ ∼ = 1 <strong>and</strong> p ∗ ∼ = t ∗ ɛ ∗ . It means ∇ ∗ <strong>and</strong> ∇ ∗ (t ∗ ) −1<br />

are injective ring homomorphisms while p ∗ <strong>and</strong> ɛ ∗ = (t ∗ ) −1 p ∗<br />

are surjective ring homomorphisms.<br />

Pro<strong>of</strong>. We prove the map ∇ ∗ is a ring homomorphism. Then<br />

from p ∗ ∇ ∗ = t ∗ we get [(t ∗ ) −1 p ∗ ]∇ ∗ = 1 Ext<br />

∗<br />

RF (C)<br />

(R,R) <strong>and</strong> thus<br />

we can define ɛ ∗ = (t ∗ ) −1 p ∗ which is a surjective ring homomorphism.<br />

Take the bar resolution B∗<br />

F (C) → R → 0. We have seen that<br />

LK ∇ B∗<br />

F (C) → LK ∇ R ∼ = RC → 0 is a projective resolution <strong>of</strong><br />

the RC e -module RC. Let f, g be two cocycles representing two<br />

cohomology classes. Then we construct the following diagram<br />

B F (C)<br />

∗<br />

LK ∇ B F (C)<br />

∗<br />

D F (C)<br />

<br />

B F (C)<br />

∗<br />

LK ∇ D F (C)<br />

<br />

LK ∇ (B F (C)<br />

LK ∇ B∗<br />

F (C) LK<br />

D Ce ∇ B∗<br />

F (C)<br />

∗<br />

Θ ∇<br />

<br />

ˆ⊗B F (C)<br />

∗<br />

ˆ⊗B F (C)<br />

∗<br />

⊗ RC LK ∇ B F (C)<br />

∗<br />

f⊗g<br />

⇓LK ∇<br />

) LK ∇(f⊗g)<br />

LK ∇ (f)⊗LK ∇ (g)<br />

<br />

R ˆ⊗R<br />

⊓⊔<br />

∼ =<br />

<br />

LK ∇ (R ˆ⊗R) ∼ =<br />

<br />

Θ 0<br />

∼ =<br />

LK ∇ (R) ⊗ RC LK ∇ (R)<br />

The first row represents f ∪ g, the cup product <strong>of</strong> f <strong>and</strong> g.<br />

The left Kan extension LK ∇ maps it to the second row <strong>of</strong><br />

<br />

R<br />

<br />

RC<br />

∼ =<br />

<br />

RC


5.2 Hochschild (co)homology <strong>of</strong> category algebras 201<br />

RC e -modules which represents the image <strong>of</strong> the cup product,<br />

<strong>and</strong> we want to show it gives rise to the cup product <strong>of</strong><br />

LK ∇ (f) <strong>and</strong> LK ∇ (g) as Hochschild cohomology classes. Since<br />

we have LK ∇ (D F (C) ) : LK ∇ B∗<br />

F (C) → LK ∇ (B∗<br />

F (C) ˆ⊗B∗ F (C) ),<br />

<strong>and</strong> LK ∇ (B∗<br />

F (C) ) <strong>and</strong> LK ∇ (B∗<br />

F (C) ) ⊗ RC LK ∇ (B∗<br />

F (C) ) are chain<br />

homotopy equivalent as both <strong>of</strong> them are projective resolutions<br />

<strong>of</strong> RC, we can construct chain maps D Ce <strong>and</strong> Θ ∇ , unique<br />

up to chain homotopy, such that the above diagram is commutative.<br />

Because the Hochschild diagonal approximation map<br />

always exists <strong>and</strong> is unique up to chain homotopy, independent<br />

<strong>of</strong> the choice <strong>of</strong> a projective resolution <strong>of</strong> RC [67], D Ce will<br />

serve as the Hochschild diagonal approximation map. Then<br />

since the lower two rows form a commutative diagram, we<br />

know they represent the same cohomology class, i.e. the cup<br />

product LK ∇ (f) ∪ LK ∇ (g), in Ext ∗ RCe(RC, RC). ⊓⊔<br />

The surjective ring homomorphism<br />

ɛ ∗ = (t ∗ ) −1 p ∗ : Ext ∗ RC e(RC, RC) → Ext∗ RF (C)(R, R)<br />

is the composite <strong>of</strong><br />

Ext ∗ RC e(RC, RC) ∼= →Ext ∗ RF (C)(R, Res ∇ RC)↠Ext ∗ RF (C)(R, R).<br />

Remark 5.2.8. Slightly modifying the previous argument, we<br />

can also demonstrate the action <strong>of</strong> Ext ∗ RC e(RC, RC) on Ext∗ RCe(RC, M)<br />

alternatively via ˆ⊗ on RF (C)-mod. For any M ∈ RC e -mod,<br />

one gets<br />

Ext ∗ RC e(RC, M) ∼ = Ext ∗ RF (C)(R, Res ∇ M).<br />

It was also shown that the RF (C)-module Res ∇ RC natural<br />

splits as R ⊕ N C for some N C ∈ RF (C)-mod. This provides<br />

a surjective homomorphism ρ : Res τ RC ˆ⊗Res ∇ M → Res ∇ M,<br />

<strong>and</strong> hence a map


202 5 Hochschild cohomology<br />

ρ ∗ : Ext ∗ RF (C)(R, Res ∇ RC ˆ⊗Res ∇ M) → Ext ∗ RF (C)(R, Res ∇ M).<br />

The latter fits into the following commutative diagram<br />

Ext ∗ RF (C) (R, Res ∇RC) ⊗ Ext ∗ RF (C) (R, Res ∇M)<br />

∪<br />

<br />

Ext ∗ RC e(RC, RC) ⊗ Ext∗ RCe(RC, M)<br />

Ext ∗ RF (C) (R, Res ∇RC ˆ⊗Res ∇ M)<br />

ρ ∗<br />

<br />

Ext ∗ RF (C) (R, Res ∇M)<br />

Ext ∗ RC<br />

∪<br />

e(RC, M),<br />

which reduces to [67, Proposition 3.1] when C = G is a<br />

group. The top row is the so-called cup product with respect<br />

to the pairing ρ. Since Ext ∗ RF (C)(R, R) is a direct summ<strong>and</strong><br />

<strong>of</strong> Ext ∗ RF (C)(R, Res ∇ RC), it also exhibits the action <strong>of</strong><br />

Ext ∗ RC(R, R) on Ext ∗ RCe(RC, M), via its identification with<br />

Ext ∗ RF (C)(R, R).<br />

Note that when C is an abelian group, we obtain an isomorphism<br />

[36, ]<br />

Ext ∗ RC e(RC, RC) ∼ = Ext ∗ RF (C)(R, Res ∇ (RC)) ∼ = RC⊗ R Ext ∗ RC(R, R).<br />

Finally we comment on Hochschild homology. It is a direct<br />

consequence <strong>of</strong> Lemma 5.2.3. Its pro<strong>of</strong> is entirely analogues<br />

to that <strong>of</strong> Corollary 4.3.4 (1). We do not have a counterpart<br />

for Corollary 4.3.4 (2) because we do not know the structure<br />

<strong>of</strong> (y, x)\∇ ( for posets ).<br />

Theorem 5.2.9. Suppose N ∈ RC e -mod. Then we have<br />

Particularly<br />

Tor RCe<br />

∗<br />

Thus Tor RC<br />

∗<br />

Tor RCe<br />

∗<br />

Tor RCe<br />

∗<br />

(RC, RC) ∼ = Tor<br />

(N, RC) ∼ = Tor<br />

RF (C)<br />

∗<br />

RF (C)<br />

∗<br />

(Res ∇ N, R).<br />

(Res ∇ RC, R) ∼ = Tor<br />

RF (C)<br />

∗<br />

(R, R)⊕Tor<br />

(R, R) is isomorphic to a direct summ<strong>and</strong> <strong>of</strong><br />

(RC, RC).<br />

RF (C)<br />


5.2 Hochschild (co)homology <strong>of</strong> category algebras 203<br />

Remark 5.2.10. Suppose u : D → C is an equivalence. By<br />

Proposition 2.3.9 we naturally obtain another equivalence<br />

F (u) : F (D) → F (C). Because we have a commutative diagram<br />

F (D) F (u) <br />

F (C)<br />

∇ <br />

D e u e <br />

C e<br />

we can deduce that the Hochschild (co)homology <strong>of</strong> RD <strong>and</strong><br />

RC are isomorphic. For instance, given any M ∈ RC e -mod,<br />

we have<br />

Tor RCe<br />

∗<br />

(RC, RC) ∼ RF (C)<br />

= Tor∗ (Res ∇ RC, R)<br />

∼ = Tor<br />

RF (D)<br />

∇<br />

∗ (Res F (u) Res ∇ RC, R)<br />

∼ = Tor<br />

RF (D)<br />

∗ (Res ∇ Res u eRC, R)<br />

∼ = Tor<br />

RD e<br />

∗ (Res u eRC, RD).<br />

Finally using Hochschild (co)homology, we can establish the<br />

following useful isomorphisms between ordinary (co)homology.<br />

Part (1) extends Proposition 3.3.10.<br />

Theorem 5.2.11. Suppose k is a field <strong>and</strong> C is a small category.<br />

Let M, N ∈ kC-mod <strong>and</strong> M ′ ∈ mod-kC. Then we<br />

have<br />

1. Ext ∗ kC(M, N) ∼ = Ext ∗ kF (C)(k, Res ∇ Hom k (M, N)); <strong>and</strong><br />

2. Tor kC<br />

∗ (M ′ , N) ∼ kF (C)<br />

= Tor∗<br />

(Res ∇ (N ⊗ k M ′ ), k).<br />

Pro<strong>of</strong>. By Proposition 5.1.4, Ext ∗ kC(M, N) ∼ = Ext ∗ kC e(kC, Hom k(M, N))<br />

From Theorem 5.2.2, Ext ∗ kC e(kC, Hom k(M, N)) ∼ = Ext ∗ kF (C)(k, Res ∇ Ho<br />

Similarly by Proposition 5.1.4 <strong>and</strong> Theorem 5.2.9, we get<br />

the isomorphism for homology. Note that it is natural to give<br />

Res ∇ (N ⊗ M ′ ) a right RF (C)-module structure. ⊓⊔


204 5 Hochschild cohomology<br />

The above isomorphisms generalize the well known isomorphisms<br />

in group cohomology Ext ∗ kG(M, N) ∼ = Ext ∗ kG(k, Hom k (M, N))<br />

<strong>and</strong> Tor kG<br />

∗ (M, N) ∼ = Tor kG<br />

∗ (N ⊗ k M, k).<br />

Proposition 3.3.10 does not generalize to cohomology. Let<br />

B∗ C → k → 0 be the bar resolution. Then the lefting B∗<br />

F (C) →<br />

Res t B∗ C induces<br />

Hom kC (B C ∗, RK t Res ∇ Hom k (M, N)) ∼ = Hom kF (C) (Res t B C ∗, Res ∇ Hom k (M<br />

→ Hom kF (C) (B F (C)<br />

∗ , Res ∇ Hom R (M<br />

<strong>and</strong> thus<br />

Ext ∗ kC(k, RK t Res ∇ Hom k (M, N)) → Ext ∗ kF (C)(k, Res ∇ Hom R (M, N))<br />

∼ = Ext<br />

∗<br />

kC (M, N).<br />

These maps have no reason to be isomorphisms in general.<br />

Similarly we have a map<br />

Tor kC<br />

∗ (M ′ , N) ∼ kF (C)<br />

= Tor∗ (Res ∇ (N ⊗ k M ′ ), k)<br />

→ Tor kC<br />

∗ (LK t Res ∇ (N ⊗ k M ′ ), k),<br />

by Corollary 4.3.4 (1) <strong>and</strong> Lemma 4.4.1.<br />

5.2.3 EI categories<br />

In this section, we assume our categories are EI. The purpose<br />

is to compare Hochschild (co)homology <strong>of</strong> C with that <strong>of</strong> an<br />

automorphism group <strong>of</strong> an object.<br />

Suppose A is the full subcategory <strong>of</strong> C which consists <strong>of</strong> all<br />

objects <strong>and</strong> all isomorphisms in C. The category A is a disjoint<br />

union <strong>of</strong> finitely many finite groups. Its category algebra<br />

RA = ⊕ x∈Ob C RAut C(x) is an RC e -module, a direct sum <strong>of</strong><br />

atomic modules supported on minimal objects in C RC . There<br />

is a surjective RC e -morphism π : RC → RA, with kernel written<br />

as ker π. Considered as a functor ker π ⊂ RC takes non-


5.2 Hochschild (co)homology <strong>of</strong> category algebras 205<br />

zero values only at (x, y) for which there exists a C-morphism<br />

from y to x <strong>and</strong> x ≁ = y. Since there is an inclusion functor<br />

i : A → C, we have maps between ordinary (co)homology i ∗ :<br />

H ∗ (C, R) → H ∗ (A, R) <strong>and</strong> i ∗ : H∗(A, R) → H∗(C, R). Here we<br />

show there are maps between their Hochschild (co)homology<br />

as well.<br />

The short exact sequence <strong>of</strong> RC e -modules<br />

induces a long exact sequence<br />

0 → ker π → RC π →RA → 0<br />

· · · → Ext n RC e(RC, ker π) → Extn RC e(RC, RC) ˜π →Ext n RC e(RC, RA) η → · · ·<br />

By Proposition 4.2.1, one can see Ext ∗ RCe(RC, RA) is naturally<br />

isomorphic to<br />

Ext ∗ RAe(RA, RA),<br />

which is isomorphic to the direct product <strong>of</strong> the Hochschild<br />

cohomology rings <strong>of</strong> the automorphism groups <strong>of</strong> objects in<br />

C: ∏<br />

Ext ∗ RAut C (x) e(RAut C(x), RAut C (x)),<br />

[x]⊂Ob C<br />

where [x] st<strong>and</strong>s for the isomorphism class <strong>of</strong> x ∈ Ob C. The<br />

following map will still be written as ˜π<br />

˜π : Ext ∗ RC e(RC, RC) → Ext∗ RAe(RA, RA).<br />

We show ˜π can be identified with the algebra homomorphism<br />

induced by − ⊗ RC RA, where as a left RC-module RA is a<br />

direct sum <strong>of</strong> atomic modules.<br />

φ A : Ext ∗ RC e(RC, RC) → Ext∗ RC e(RA, RA) ∼ = Ext ∗ RAe(RA, RA).<br />

Hence we do not need to distinguish the maps φ A <strong>and</strong> ˜π.<br />

Lemma 5.2.12. The following diagram is commutative


206 5 Hochschild cohomology<br />

Ext ∗ RC e(RC, RC) ˜π <br />

Ext ∗ RCe(RC, RA)<br />

φ A<br />

∼<br />

<br />

=<br />

Ext ∗ RC e(RA, RA) <br />

∼ =<br />

Ext ∗ RAe(RA, RA).<br />

Pro<strong>of</strong>. This can be seen on the cochain level. Suppose R ∗ →<br />

RC → 0 is the minimal projective resolution <strong>of</strong> the RC e -<br />

module RC. Then Ext ∗ RCe(RC, RC) is the cohomology <strong>of</strong> the<br />

cochain complex Hom RC e(R ∗ , RC). The tensor product −⊗ RC<br />

RA induces a map<br />

Hom RC e(R ∗ , RC) → Hom RC−RA (R ∗ ⊗ RC RA, RA).<br />

Since B∗<br />

RC ⊗ RC RA is a projective resolution <strong>of</strong> the RC-RAmodule<br />

RA, R ∗ ⊗ RC RA is also a projective resolution <strong>of</strong> RA.<br />

Moreover because R ∗ is minimal, it is supported on C RC . It<br />

implies R ∗ ⊗ RC RA is also supported on C RC . But the RC e -<br />

module RA is supported on minimal objects <strong>of</strong> C RC , we have<br />

Hom RC−RA (R ∗ ⊗ RC RA, RA) ∼ = Hom RA e(Res i (R ∗ ⊗ RC RA), RA)<br />

∼ = HomRA e(Res i R ∗ , RA),<br />

which gives rise to φ A . Here i : A e → C e is the inclusion,<br />

induced by another inclusion, also written as i : A → C. On<br />

the other h<strong>and</strong> ˜π is exactly given by the same chain map<br />

Hom RC e(R ∗ , RC) → Hom RC e(R ∗ , RA) ∼ = Hom RA e(Res i R ∗ , RA).<br />

Thus we are done.<br />

We have the following commutative diagram, involving four<br />

cohomology rings.<br />

Theorem 5.2.13. Let C be an EI-category. Then we have<br />

the following commutative diagram<br />

⊓⊔


5.2 Hochschild (co)homology <strong>of</strong> category algebras 207<br />

Ext ∗ RC e(RC, RC) φ A =˜π <br />

φ C <br />

Ext ∗ RAe(RA, RA)<br />

φ A<br />

Ext ∗ RC(R, R)<br />

Res i<br />

<br />

Ext ∗ RA(R, R).<br />

Pro<strong>of</strong>. We prove it on the cochain level. Let R ∗ → RC → 0<br />

be the minimal projective resolution <strong>of</strong> the RC e -module RC.<br />

Then we have the following commutative diagram<br />

Hom RC e(R ∗ , RC) −⊗ RCRA <br />

Hom RC−RA (R ∗ ⊗ RC RA, RA)<br />

−⊗ RC R<br />

<br />

Hom RC (R ∗ ⊗ RC R, R)<br />

−⊗ RA R<br />

Hom RC (R ∗ ⊗ RC R, R)<br />

Hom RC (R ′ ∗, R)<br />

Res i<br />

<br />

Hom RA (Res i R ′ ∗, R)<br />

<br />

Hom RA (R ′′<br />

∗, R),<br />

in which R ′ ∗ → R → 0 <strong>and</strong> R ′′<br />

∗ → R → 0 are the projective<br />

resolutions <strong>of</strong> the trivial RC- <strong>and</strong> RA-modules satisfying<br />

the following commutative diagrams <strong>of</strong> RC-modules <strong>and</strong> RAmodules,<br />

respectively,<br />

R ′ ∗<br />

R 0 R ′′<br />

∗<br />

R ∗ ⊗ RC R R 0 <strong>and</strong> Res i R ′ ∗<br />

R<br />

<br />

0<br />

<br />

R 0.<br />

In the main diagram, upper left cochain complex computes<br />

Ext ∗ RC e(RC, RC), upper right corner computes Ext∗ RAe(RA, RA)<br />

by Lemma 5.2.12, lower left corner computes Ext ∗ RC(R, R) <strong>and</strong><br />

lower right corner computes Ext ∗ RA(R, R). Hence our statement<br />

follows.<br />

⊓⊔


208 5 Hochschild cohomology<br />

We note that in the theorem the category A may be replaced<br />

by any full subcategory <strong>of</strong> it. Especially, we have a<br />

commutative diagram for each Aut C (x) ⊂ A<br />

Ext ∗ RC e(RC, RC)φ RAut C (x)<br />

Ext ∗ RAut C (x) e(RAut C(x), RAut C (x))<br />

φ C <br />

φ AutC (x)<br />

<br />

Ext ∗ RC(R, R)<br />

Res i<br />

<br />

Ext ∗ RAut C (x)(R, R).<br />

5.3 Examples <strong>of</strong> the Hochschild cohomology rings <strong>of</strong> categories<br />

In this section we calculate the Hochschild cohomology rings<br />

for four finite EI-categories, with base field k <strong>of</strong> characteristic<br />

2.<br />

5.3.1 The category E 0<br />

In [87] we presented an example, by Aurélien Djament, Laurent<br />

Piriou <strong>and</strong> the author, <strong>of</strong> the mod-2 ordinary cohomology<br />

ring <strong>of</strong> the following category E 0<br />

h<br />

g<br />

x<br />

1 x<br />

α<br />

β<br />

gh<br />

y {1 y }<br />

,<br />

where g 2 = h 2 = 1 x , gh = hg, αh = βg = α, <strong>and</strong> αg = βh =<br />

β. The ordinary cohomology ring Ext ∗ kE 0<br />

(k, k) is a subring <strong>of</strong><br />

the polynomial ring H ∗ (Z 2 × Z 2 , k) ∼ = k[u, v], removing all<br />

u n , n ≥ 1, <strong>and</strong> their scalar multiples. It has no nilpotents <strong>and</strong><br />

is not finitely generated. By Theorem 2.3.4, it implies that the<br />

Hochschild cohomology ring Ext ∗ kE e 0 (kE 0, kE 0 ) is not finitely<br />

generated either. We compute its Hochschild cohomology ring<br />

using Proposition 2.3.5.


5.3 Examples <strong>of</strong> the Hochschild cohomology rings <strong>of</strong> categories 209<br />

The category <strong>of</strong> factorizations in E 0 , F (E 0 ), has the following<br />

shape<br />

[α]<br />

[β]<br />

[1 x ]<br />

[1 y ]<br />

[h]<br />

[gh]<br />

[g]<br />

in which [1 x ] ∼ = [h] ∼ = [g] ∼ = [gh] <strong>and</strong> [α] ∼ = [β]. For the purpose<br />

<strong>of</strong> computation, we use the skeleton F ′ (E 0 ) <strong>of</strong> F (E 0 ) (which<br />

is equivalent to F (E 0 ) hence the two category algebras <strong>and</strong><br />

their module categories are Morita equivalent)<br />

{(α,1 op<br />

x ),(α,h op ),(β,g op ),(β,(gh) op )}<br />

{(1 y ,1 op<br />

x )}<br />

<br />

[α]<br />

[1 x ]<br />

[1<br />

y ].<br />

{(1 x ,1 op<br />

x ),(h,h op ),(g,g op ),(gh,(gh) op )}<br />

{(1 y ,α op )}<br />

{(1 y ,1 op<br />

y )}<br />

In the above category, next to each arrow is the set <strong>of</strong> homomorphisms<br />

in F ′ (E 0 ) from one object to another. The module<br />

N E0 ∈ kF ′ (E 0 )-mod (see Proposition 2.3.5) takes the following<br />

values<br />

N C ([1 x ]) = k{1 x + h, g + gh, 1 x + g} , N C ([h]) = k{1 x + h, g + gh, 1 x<br />

N C ([g]) = k{1 x + h, g + gh, 1 x + g} , N C ([gh]) = k{1 x + h, g + gh, 1 x<br />

N C ([α]) = k{α + β}<br />

, N C ([β]) = k{α + β},<br />

N C ([1 y ]) = 0.<br />

,


210 5 Hochschild cohomology<br />

Thus N E0 = S [1x ],k(1 x +h)⊕S [1x ],k(g+gh)⊕k ′ 1 x +g, where S [1x ],k(1 x +h)<br />

<strong>and</strong> S [1x ],k(g+gh) are simple kF ′ (E 0 )-modules such that S [1x ],k(1 x +h)([1 x ])<br />

k(1 x + h) <strong>and</strong> S [1x ],k(g+gh)([1 x ]) = k(g + gh), <strong>and</strong> k ′ 1 x +g is a<br />

kF ′ (E 0 )-module such that k ′ 1 x +g([1 x ]) = k(1 x +g), k ′ 1 x +g([α]) =<br />

k(α + β) <strong>and</strong> k ′ 1 x +g([1 y ]) = 0. Note that S [1x ],k(1 x +h)([1 x ]) =<br />

k(1 x + h), S [1x ],k(g+gh)([1 x ]) = k(g + gh) <strong>and</strong> k ′ 1 x +g([1 x ]) =<br />

k(1 x + g) are all isomorphic to the trivial kAut F ′ (E 0 )([1 x ])-<br />

module, <strong>and</strong> have the same trivial ring structure in the sense<br />

that the product <strong>of</strong> any two elements is zero. Hence we have<br />

(along with the result quoted in Section 2.4, paragraph two)<br />

Ext ∗ kF ′ (E 0 )(k, S [1x ],k(1 x +h)) ∼ = k(1 x + h) ⊗ k Ext ∗ kAut F ′ (E0 ) ([1 x])(k, k)<br />

<strong>and</strong><br />

Ext ∗ kF ′ (E 0 )(k, S [1x ],k(g+gh)) ∼ = k(g + gh) ⊗ k Ext ∗ kAut F ′ (E0 ) ([1 x])(k, k)<br />

as rings, in which k(1 x + h) <strong>and</strong> k(g + gh) are concentrated<br />

in degree zero in each ring. From the structure <strong>of</strong> F (E 0 ), one<br />

has Aut F ′ (E 0 )([1 x ]) ∼ = Z 2 × Z 2 .<br />

For computing Ext ∗ kF ′ (E 0 )(k, k ′ 1 x +g), we use the following<br />

short exact sequence <strong>of</strong> kF (E 0 )-modules<br />

0 → k ′ 1 x +g → k → S [1y ],k → 0.<br />

It induces a long exact sequence in which one can find<br />

Ext 0 kF ′ (E 0 )(k, S [1y ],k) = k <strong>and</strong> Ext n kF ′ (E 0 )(k, S [1y ],k) = 0 if n ≥ 1.<br />

Thus Ext 0 kF ′ (E 0 )(k, k ′ 1+g) = 0 while Ext n kF ′ (E 0 )(k, k) ∼ = Ext n kF ′ (E 0 )(k, k ′ 1+g)<br />

for each n ≥ 1. Hence as a ring<br />

Ext ∗ kF ′ (E 0 )(k, k ′ 1 x +g) ∼ = k(1 x +g)⊗ k Ext ∗>0<br />

kF ′ (E 0 ) (k, k) ∼ = k(1 x +g)⊗ k Ext ∗>0<br />

kE 0<br />

(<br />

All in all, we have<br />

Ext 0 kE e 0 (kE 0, kE 0 ) ∼ = Ext 0 kE 0<br />

(k, k) ⊕ k(1 x + h) ⊕ k(g + gh),


5.3 Examples <strong>of</strong> the Hochschild cohomology rings <strong>of</strong> categories 211<br />

<strong>and</strong> if n ≥ 1<br />

Ext n kE e 0 (kE 0, kE 0 )<br />

∼ = Ext<br />

n<br />

kE0 (k, k) ⊕ {k(1 x + g) ⊗ k Ext n kE 0<br />

(k, k)}<br />

⊕{k(1 x + h) ⊗ k Ext ∗ k(Z 2 ×Z 2 )(k, k)} ⊕ {k(g + gh) ⊗ k Ext n k(Z 2 ×Z 2 )(k,<br />

Combining all the information we obtained, the surjective ring<br />

homomorphism<br />

φ E0 : Ext ∗ kE e 0 (kE 0, kE 0 ) ↠ Ext ∗ kE 0<br />

(k, k)<br />

has its kernel consisting <strong>of</strong> all nilpotents. Consequently this<br />

Hochschild cohomology ring modulo nilpotents is not finitely<br />

generated, against the finite generation conjecture in []. We<br />

comment that the category algebra kE 0 is not a self-injective<br />

algebra (hence is not Hopf, by []). Nicole Snashall points out<br />

to the author that this algebra is Koszul since both kE 0 <strong>and</strong><br />

Ext ∗ kE 0<br />

(kE 0 , kE 0 ) as graded algebras are generated in degrees<br />

zero <strong>and</strong> one, where kE 0 = kE 0 /Rad(kE 0 ) ∼ = S x,k ⊕ S y,k .<br />

5.3.2 The category E 1<br />

The following category E 1 has a terminal object <strong>and</strong> hence is<br />

contractible:<br />

h<br />

x<br />

1 x<br />

α <br />

y {1 y } ,<br />

where h 2 = 1 x <strong>and</strong> αh = α. The contractibility implies the<br />

ordinary cohomology ring is simply the base field k. In this<br />

case F (E 1 ) is the following category


212 5 Hochschild cohomology<br />

[1 x ]<br />

(h,1 op<br />

y )<br />

(α,Aut E1 (x) op )<br />

(1 x ,1 op<br />

x ) <br />

(h,h op )<br />

(1 x ,h op )<br />

(h,1 op<br />

x )<br />

[α]<br />

[h]<br />

(1 x ,h op )<br />

(1 x ,1 op<br />

y )<br />

(h,1 op<br />

x )<br />

(1 y ,α op )<br />

(α,Aut E1 (x) op )<br />

[1 y ]<br />

(1 y ,1 op<br />

y )<br />

We calculate its Hochschild cohomology ring. By proposition<br />

2.3.5, we only need to compute Ext ∗ kF (E 1 )(k, N E1 ), where N E1<br />

has the following value at objects <strong>of</strong> F (E 1 )<br />

N E1 ([1 x ]) = k{1 x + h} , N E1 ([h]) = k{1 x + h},<br />

N E1 ([1 y ]) = 0 , N E1 ([α]) = 0.<br />

One can easily see that N E1 = S [1x ],k(1 x +h) is a simple module <strong>of</strong><br />

dimension one with a specified value k(1 x + h) at [1 x ]. Since<br />

[1 x ] ∼ = [h] ∈ Ob F (E1 ) are minimal objects, using quoted<br />

result in Section 2.4 paragraph two, we get<br />

Ext ∗ kF (E 1 )(k, N E1 ) ∼ = Ext ∗ kAut F (E1 )([1 x ])(k, k(1 x +h)) ∼ = k(1 x +h)⊗ k Ext ∗ kZ 2<br />

(<br />

which is isomorphic to k(1 x + h) ⊗ k k[u]. Here k[u] is a polynomial<br />

algebra with an indeterminant u at degree one <strong>and</strong><br />

k(1 x + h) is at degree zero. Thus<br />

Ext ∗ kE e 1 (kE 1, kE 1 ) ∼ = Ext ∗ kE 1<br />

(k, k)⊕Ext ∗ kF (E 1 )(k, N E1 ) ∼ = k⊕{k(1 x +h)⊗ k k<br />

The kernel <strong>of</strong> φ E1<br />

cohomology ring.<br />

consists <strong>of</strong> all nilpotents in the Hochschild<br />

5.3.3 The category E 2<br />

The following category has its classifying space homotopy<br />

equivalent to the join, BZ 2 ∗ BZ 2 = Σ(BZ 2 ∧ BZ 2 ) =


5.3 Examples <strong>of</strong> the Hochschild cohomology rings <strong>of</strong> categories 213<br />

Σ[B(Z 2 × Z 2 )/(BZ 2 ∨ BZ 2 )], <strong>of</strong> the classifying spaces <strong>of</strong> the<br />

two automorphism groups:<br />

h<br />

x<br />

1 x<br />

α<br />

g<br />

1 y<br />

<br />

y ,<br />

<br />

where h 2 = 1 x , αh = α = gα <strong>and</strong> g 2 = 1 y . As direct consequences,<br />

its ordinary cohomology groups are equal to k, 0, 0<br />

at degrees zero, one <strong>and</strong> two, <strong>and</strong> k n−2 at each degree n ≥ 3,<br />

<strong>and</strong> furthermore the cup product in this ring is trivial [87].<br />

We compute its Hochschild cohomology ring. The category<br />

F (E 2 ) is as follows<br />

(1 x ,1 op<br />

x ) <br />

(h,h op )<br />

[1 x ]<br />

(α,Aut E2 (x) op )<br />

(α,Aut E2 (x) op )<br />

(1 x ,h op )<br />

[h]<br />

(Aut E2 (x),Aut E2 (y) op )<br />

<br />

(h,1 op<br />

x )<br />

[α]<br />

(1 y ,1 op<br />

y )<br />

(Aut E2 (y),α op )<br />

(Aut E2 (y),α op )<br />

[1 y ]<br />

(g,g op )<br />

[g]<br />

(1 y ,g op )<br />

(g,1 op<br />

y )<br />

By Proposition 2.3.5, we need to compute Ext ∗ kE 2<br />

(k, N E2 ). In<br />

this case we have<br />

N E2 ([1 x ]) = k{1 x + h} , N E2 ([h]) = k{1 x + h},<br />

N E2 ([1 y ]) = k{1 y + g} , N E2 ([g]) = k{1 y + g},<br />

N E2 ([α]) = 0.<br />

It means N E2 = S [1x ],k(1 x +h) ⊕ S [1y ],k(1 y +g) <strong>and</strong> thus by Proposition<br />

2.2.5<br />

Ext ∗ kE 2<br />

(k, N E2 ) ∼ = Ext ∗ kAut F (E2 )([1 x ])(k, k(1 x + h)) ⊕ Ext ∗ kAut F (E2 )([1 y ])(k, k(1<br />

∼ = {k(1x + h) ⊗ k Ext ∗ kZ 2<br />

(k, k)} ⊕ {k(1 y + g) ⊗ k Ext ∗ kZ 2<br />

Hence


214 5 Hochschild cohomology<br />

Ext ∗ kE e 2 (kE 2, kE 2 ) ∼ = Ext ∗ kE 2<br />

(k, k)⊕{k(1 x +h)⊗ k k[u]}⊕{k(1 y +g)⊗ k k[v]}<br />

where k[u] <strong>and</strong> k[v] are two polynomial algebras with indeterminants<br />

in degree one. Both the Hochschild <strong>and</strong> ordinary<br />

cohomology rings modulo nilpotents are isomorphic to the<br />

base field k.<br />

5.3.4 The category E 3<br />

The following category has a classifying space homotopy<br />

equivalent to that <strong>of</strong> Aut E3 (x) ∼ = Z 2 (by Quillen’s Theorem A<br />

[], or see [86])<br />

h<br />

x<br />

1 x<br />

α<br />

β<br />

y {1 y }<br />

,<br />

where h 2 = 1 x <strong>and</strong> αh = β. We compute its Hochschild<br />

cohomology ring. The category F (E 3 ) is as follows (not all<br />

morphisms are presented since only its skeleton is needed)<br />

(1 y ,1 op<br />

x )<br />

(α,1 op<br />

x ),(β,h op )<br />

(h,1 op<br />

[α]<br />

(1 y ,h op )<br />

[β]<br />

(1 y ,1 op<br />

x )<br />

<br />

x ) [1 x ]<br />

[1 y ]<br />

(1 x ,h op )<br />

(1 x ,h op )<br />

[h]<br />

(h,1 op<br />

x )<br />

(1 y ,β op )<br />

.<br />

The module N E3 takes the following values<br />

(1 y ,1 op<br />

y )<br />

N E2 ([1 x ]) = k{1 x + h} , N E2 ([h]) = k{1 x + h},<br />

N E2 ([1 y ]) = 0 , N E2 ([α]) = k{α + β},<br />

N E2 ([α]) = k{α + β}.


5.3 Examples <strong>of</strong> the Hochschild cohomology rings <strong>of</strong> categories 215<br />

Thus N E3 fits into the following short exact sequence <strong>of</strong><br />

kF (E 3 )-modules<br />

0 → N E3 → k → S [1y ],k → 0.<br />

Just like in our first example, using the long exact sequence<br />

coming from it, we know Ext 0 kE 2<br />

(k, N E3 ) = 0 <strong>and</strong> Ext ∗>0<br />

kE 2<br />

(k, N E3 ) ∼ =<br />

k(1 x +h)⊗ k Ext ∗>0<br />

kF (E 3 ) (k, k) ∼ = k(1 x +h)⊗ k Ext ∗>0<br />

kE 3<br />

(k, k). Hence<br />

Ext ∗ kE e 3 (kE 3, kE 3 ) ∼ = Ext ∗ kE 3<br />

(k, k)⊕{k(1 x +h)⊗ k Ext ∗>0<br />

kE 3<br />

(k, k)}.<br />

The kernel <strong>of</strong> φ E3<br />

cohomology ring.<br />

contains all nilpotents in the Hochschild


Chapter 6<br />

Connections with group representations <strong>and</strong> cohomology<br />

Abstract We study various local categories <strong>of</strong> finite groups.<br />

The purpose is to compare representations <strong>and</strong> cohomology<br />

<strong>of</strong> groups <strong>and</strong> <strong>of</strong> these categories. The transporter categories<br />

play a key role to bridge up these two concepts. In fact a<br />

group <strong>and</strong> all its subgroups are some sort <strong>of</strong> transporter categories.<br />

We pay attention to the transporter categories <strong>and</strong><br />

try to demonstrate their similarities with <strong>and</strong> differences from<br />

groups, in terms <strong>of</strong> homological properties. We shall apply<br />

tools developed in the preceding two chapters to investigate<br />

transporter categories, a special kind <strong>of</strong> finite EI categories<br />

which partially motivate modern research on category representations<br />

<strong>and</strong> cohomology. In this chapter all categories are<br />

finite <strong>and</strong> all modules are finitely generated.<br />

6.1 Local categories<br />

Let G be a finite group <strong>and</strong> p a positive prime that divides<br />

|G|, the order <strong>of</strong> G. We study various collections <strong>of</strong> subgroups<br />

<strong>of</strong> G <strong>and</strong> resulting categories. <strong>Finite</strong> categories naturally appear<br />

in both group representation <strong>and</strong> homotopy theory. For<br />

a finite group G <strong>and</strong> a positive prime p that divides the order<br />

<strong>of</strong> G, various posets <strong>of</strong> p-subgroups have been investigated intensively<br />

to establish connections between representations <strong>of</strong><br />

G <strong>and</strong> those <strong>of</strong> its local subgroups [61, 7, 76, 80, 81, 69]. Over<br />

the past two decades, mathematicians realized that categories


218 6 Connections with group representations <strong>and</strong> cohomology<br />

build upon those previously mentioned posets should play an<br />

important role in modular representation theory [60]. Indeed,<br />

to any such poset P, it naturally comes with a G-action.<br />

There is a Grothendieck construction on P which results in<br />

a category G ∝ P, called a transporter category, containing<br />

P as a subcategory. A transporter category admits some interesting<br />

quotient categories, namely Brauer categories, Puig<br />

categories <strong>and</strong> orbit categories [75]. For many good reasons,<br />

we shall call any quotient category <strong>of</strong> a transporter category a<br />

local category, as it reveals some p-local information about G.<br />

<strong>Representations</strong> <strong>and</strong> cohomology <strong>of</strong> these local categories are<br />

currently in the center <strong>of</strong> group representation theory. In homotopy<br />

theory <strong>of</strong> classifying spaces, the geometric realization<br />

<strong>of</strong> the nerve <strong>of</strong> a transporter category is a Borel construction,<br />

EG × G BP. It was shown by several authors, especially<br />

Dwyer [17], that transporter categories play a key role in homology<br />

decompositions <strong>of</strong> classifying spaces. Recently, Broto,<br />

Levi <strong>and</strong> Oliver [10, 11] developed a theory <strong>of</strong> p-local finite<br />

groups, which is motivated by the use <strong>of</strong> local categories associated<br />

with BG (or its p-completion). In their terminology,<br />

a p-local finite group is a triple (S, F, L) such that S is a<br />

finite p-group, F is a finite category called a fusion system<br />

on S, <strong>and</strong> L is an extension <strong>of</strong> a full subcategory <strong>of</strong> F. The<br />

(p-completion <strong>of</strong>) classifying space <strong>of</strong> L behaves like the classifying<br />

space <strong>of</strong> a finite group. In this theory, cohomology <strong>of</strong><br />

small categories is an essential ingredients. See the new book<br />

[2] for fusion systems in group theory, homotopy theory <strong>and</strong><br />

representation theory.


6.1 Local categories 219<br />

6.1.1 G-categories<br />

Definition 6.1.1. Consider G as a category with one object<br />

•. A G-category is a functor F : G → Cat.<br />

In other words, a G-category is a category C, equipped with<br />

a group homomorphism G = Aut G (•) → Aut Cat (C). The<br />

simplest example is a point with trivial action by G. Recall<br />

that a set is regarded as a poset with trivial relations, <strong>and</strong> a<br />

poset is regarded as a category.<br />

Example 6.1.2. Suppose H ⊂ G is a subgroup. Consider the<br />

discrete set <strong>of</strong> left cosets G/H = {gH ∣ ∣ g ∈ G}. Then G acts<br />

on it by left multiplication, permuting these cosets.<br />

Example 6.1.3. Suppose H ⊂ G is a subgroup. Then the discrete<br />

set <strong>of</strong> conjugacy class G H = { g H ∣ ∣ g ∈ G} is also a<br />

G-category with G acting by conjugations.<br />

The above two examples are the examples <strong>of</strong> G-sets.<br />

Definition 6.1.4. A collection <strong>of</strong> subgroups <strong>of</strong> G is a set <strong>of</strong><br />

subgroups <strong>of</strong> G, closed under conjugations by elements in G.<br />

Example 6.1.5. A collection <strong>of</strong> subgroups <strong>of</strong> G is naturally<br />

a poset with inclusions as relations. Hence any collection <strong>of</strong><br />

subgroups is a G-poset.<br />

There are various interesting collections <strong>of</strong> subgroups <strong>of</strong> G.<br />

Example 6.1.6.1. The collection <strong>of</strong> all p-subgroups <strong>of</strong> G, denoted<br />

by S e p.<br />

2. The collection <strong>of</strong> all non-identity p-subgroups <strong>of</strong> G, denoted<br />

by S p . Below is a concrete example <strong>of</strong> the poset <strong>of</strong> all nontrivial<br />

2-subgroups <strong>of</strong> Σ 4 , S 2 (Σ 4 ):


220 6 Connections with group representations <strong>and</strong> cohomology<br />

· · ·<br />

· · ·<br />

D 8 D 8 D 8<br />

<br />

C 2 × C 2<br />

<br />

C 4<br />

<br />

V<br />

C 2<br />

<br />

C 2<br />

Here D 8 is the dihedral group <strong>of</strong> order 8, V is a Kleine four<br />

group <strong>and</strong> those C are cyclic groups with order specified<br />

in the subscripts. Due to the size, we do not record the full<br />

poset. In deed we only write down the subgroups <strong>of</strong> one <strong>of</strong><br />

the three D 8 . However, in order to obtain the full poset, one<br />

just needs to copy the same thing under the leftmost D 8 <strong>and</strong><br />

put it for each omitted part, where the dots appear. Note<br />

that V is a subgroup <strong>of</strong> all three D 8 .<br />

3. The collection <strong>of</strong> all non-identity elementary p-subgroups <strong>of</strong><br />

G, denoted by A p .<br />

4. The collection <strong>of</strong> all p-radical subgroups <strong>of</strong> G, denoted by<br />

B p .<br />

Moreover, there are various refinements <strong>of</strong> some <strong>of</strong> the above<br />

posets.<br />

Example 6.1.7.1. Let b be a p-block <strong>of</strong> kG. The the b-Brauer<br />

pairs S b form a G-poset. When b is the principal block, it is<br />

isomorphic to S e p.<br />

2. Let A be an interior G-algebra. Then the pointed subgroups<br />

<strong>of</strong> G form a G-poset.<br />

C 2<br />

C 2<br />

C 2<br />

6.1.2 Homology representations <strong>of</strong> kG<br />

Suppose C is a small G-category. Its nerve N ∗ C is a simplicial<br />

set, from which one can construct a chain complex C ∗ (C, k)


6.1 Local categories 221<br />

such that, for each i > 0, C i (C, k) is a k-vector space with a<br />

basis the set <strong>of</strong> i-chains <strong>of</strong> morphisms in C, while C 0 (C, k) =<br />

k Ob C. Then one can see that G acts on each C i (C, k) <strong>and</strong><br />

C i (C, k) by permutating its base elements. Consequently we<br />

obtains G-modules Hi(C, k) <strong>and</strong> H i (C, k).<br />

Definition 6.1.8. Let C be a small category. Assume the normalized<br />

complex C † ∗(C, k) is finite. Then the Euler characteristic<br />

χ(C) = χ(C, k) is defined to be ∑ i≥0 (−1)i dim k C † i (C, k).<br />

Note that C † ∗(C, k) being finite is equivalent to BC being<br />

a finite CW complex. Since ∑ i≥0 (−1)i dim k C † i (C, k) =<br />

∑<br />

i≥0 (−1)i dim k Hi(C, k), χ(C, k) is invariant under homotopy<br />

equivalence.<br />

Example 6.1.9.1. In Example 6.1.2, C ∗ (G/H) is a stalk complex<br />

C ∗ (G/H) = C 0 (G/H) ∼ = k ↑ G H with χ(G/H) = [G :<br />

H].<br />

2. In Example 6.1.3, C ∗ ( G H) is a stalk complex C ∗ ( G H) =<br />

C 0 ( G H) ∼ = k ↑ G N G (H) with χ(G H) = [G : N G (H)].<br />

3. In Example 6.1.4 (1), Sp e is not a stalk complex. However<br />

since Sp e has an initial object <strong>and</strong> thus is contractible,<br />

χ(Sp) e = 1.<br />

4. The inclusions A p ⊂ S p <strong>and</strong> B p ⊂ S p induce homotopy<br />

equivalences, by using Quillen’s Theorem A. Hence<br />

χ(S p , k) = χ(A p , k) = χ(B p , k).<br />

6.1.3 Transporter categories as Grothendieck constructions<br />

Various G-posets <strong>of</strong> subgroups <strong>of</strong> G demonstrate certain local<br />

structural information <strong>of</strong> G. However it is not complete. For<br />

example given H ∈ G <strong>and</strong> the discrete set G H, any two nonidentical<br />

objects are conjugate in G while this relationship is


222 6 Connections with group representations <strong>and</strong> cohomology<br />

not seen in G H. Thus we want a category that presents all intrinsic<br />

connections among any chosen collection <strong>of</strong> subgroups<br />

<strong>of</strong> G.<br />

Definition 6.1.10. Let G be a group <strong>and</strong> P a G-poset. We<br />

define the transporter category on P to be a Grothendieck<br />

construction G ∝ P := Gr G P. More precisely, G ∝ P has<br />

the same objects as P, that is, Ob(G ∝ P) = Ob P. For<br />

x, y ∈ Ob(G ∝ P), a morphism from x to y is a pair (g, gx ≤<br />

y) for some g ∈ G.<br />

In the literature the transporter categories are mostly considered<br />

as auxiliary constructions before passing to various<br />

quotient categories <strong>of</strong> them. Here we want to stress on the perhaps<br />

unique property, among various categories constructed<br />

from a group, that transporter categories admit natural functors<br />

to the group itself. It singles out this particular type <strong>of</strong><br />

categories <strong>and</strong> is the starting point <strong>of</strong> this chapter. Here in<br />

order to emphasize the similarities <strong>and</strong> connections between<br />

transporter categories <strong>and</strong> subgroups, we follow a definition<br />

which is well known to some algebraic topologists. The symbol<br />

G ∝ P is not st<strong>and</strong>ard <strong>and</strong> is used because this particular<br />

Grothendieck construction resembles a semidirect product,<br />

yet is different.<br />

This neat but seemingly abstract definition can be easily<br />

seen to give the usual transporter categories. For example,<br />

when P is the poset <strong>of</strong> non-trivial p-subgroups, we get<br />

G ∝ P = Tr p (G), the p-transporter category <strong>of</strong> G. The advantage<br />

<strong>of</strong> taking our approach is shown by the upcoming<br />

examples, where each subgroup <strong>of</strong> G is identified as a transporter<br />

category, up to a category equivalence.<br />

From Definition 6.1.10 one can easily see that there is a<br />

natural embedding ι P : P ↩→ G ∝ P via (x ≤ y) ↦→ (e, x ≤<br />

y). On the other h<strong>and</strong>, the transporter category admits a


6.1 Local categories 223<br />

natural functor π P : G ∝ P → G, given by x ↦→ • <strong>and</strong><br />

(g, gx ≤ y) ↦→ g. Thus we always have a sequence <strong>of</strong> functors<br />

P ι P<br />

↩→G ∝ P π P<br />

→G<br />

such that π P ◦ ι P (P) is the trivial subgroup or subcategory<br />

<strong>of</strong> G. For convenience, in the rest <strong>of</strong> this chapter we <strong>of</strong>ten<br />

neglect the subscript P <strong>and</strong> write ι = ι P , π = π P .<br />

Example 6.1.11.1. If G acts trivially on P, then G ∝ P =<br />

G × P.<br />

2. Let G be a finite group <strong>and</strong> H a subgroup. We consider<br />

the set <strong>of</strong> left cosets G/H which can be regarded as a G-<br />

poset: G acts via left multiplication. The transporter category<br />

G ∝ (G/H) is a connected groupoid whose skeleton<br />

is isomorphic to H. In this way one can recover all subgroups<br />

<strong>of</strong> G, up to category equivalences. Consequently we<br />

have k(G ∝ G/H) ≃ kH as well as a homotopy equivalence<br />

B(G ∝ G/H) ≃ BH.<br />

3. From Example 6.1.6 (2) we build a concrete transporter category<br />

Σ 4 ∝ S 2 (Σ 4 ) :<br />

C 2<br />

4<br />

8<br />

C 2 × C 2<br />

8<br />

4 <br />

4<br />

8<br />

C 2<br />

4<br />

8<br />

8<br />

· · ·<br />

· · ·<br />

<br />

8<br />

D 8<br />

8<br />

8 D 8<br />

8<br />

8 D 8<br />

8 24 24 24<br />

C 4 8 V 24<br />

8<br />

C 2<br />

8<br />

24<br />

8 <br />

8<br />

24<br />

C 2<br />

Here the numbers are the numbers <strong>of</strong> morphisms from one<br />

object to another. Note that again this is part <strong>of</strong> the whole<br />

category. However, it contains a skeleton so we know what<br />

is missing.<br />

8<br />

24<br />

8 <br />

8<br />

C 2<br />

8


224 6 Connections with group representations <strong>and</strong> cohomology<br />

Remark 6.1.12. One can check directly that if Hom G∝P (x, y) ≠<br />

∅ then both Aut G∝P (x) <strong>and</strong> Aut G∝P (y) act freely on Hom G∝P (x, y).<br />

Another notable structural fact is that G ∝ P is a category<br />

with subobjects, which means each morphism can be<br />

uniquely factorized as an isomorphism followed by a morphism<br />

in P (regarded as a subcategory <strong>of</strong> G ∝ P). In fact we<br />

have (g, gx ≤ y) = (e, gx ≤ y) ◦ (g, gx = gx) as shown in the<br />

diagram<br />

y<br />

(g,gx≤y)<br />

x<br />

(e,gx≤y)<br />

(g,gx=gx) gx.<br />

Definition 6.1.13. We call B(G ∝ P) the Borel construction<br />

on P.<br />

For any G-space X, one can define a Borel construction<br />

EG × G X. In our definition, we actually have B(G ∝ P) ≃<br />

EG × G BP. It explains the concept. In particular B(G ∝<br />

(G/H)) ≃ BH. Forming the transporter category over a G-<br />

poset eliminates the G-action, as an algebraic analogy <strong>of</strong> introducing<br />

a Borel construction over a G-space.<br />

6.1.4 Local categories<br />

Our concept <strong>of</strong> a transporter category is quite general because<br />

each subgroup H <strong>of</strong> G can be recovered as a transporter category<br />

G ∝ (G/H), for the G-poset G/H, up to a category<br />

equivalence. Thus it makes sense if we deem transporter categories<br />

as generalized subgroups for a fixed finite group.<br />

Definition 6.1.14. Any quotient category C <strong>of</strong> G ∝ P is<br />

called a local category <strong>of</strong> G. When P consists <strong>of</strong> p-subgroups<br />

<strong>of</strong> G, for a prime p ∣ ∣ |G|, we also call such a quotient category<br />

a p-local category.


6.1 Local categories 225<br />

A local category is connected with the group by the following<br />

diagram<br />

G ∝ P<br />

π<br />

G<br />

C<br />

Transporter categories were implicitly considered by Mark<br />

Ronan <strong>and</strong> Steve Smith [65] in the 1980s for constructing<br />

group modules, <strong>and</strong> later on played a key role in Bill Dwyer’s<br />

work [17] on homology decomposition <strong>of</strong> classifying spaces.<br />

Dwyer used this diagram to establish connections among various<br />

homotopy colimits (e.g. classifying spaces), while Ronan<br />

<strong>and</strong> Smith constructed kG-modules via representations<br />

<strong>of</strong> G ∝ P (using the language <strong>of</strong> G-presheaves on P).<br />

Example 6.1.15.1. The p-transporter category Tr p (G) = G ∝<br />

S p has all non-identity subgroups as its objects. For any<br />

p-subgroups P, Q, the morphism set is <strong>of</strong>ten written as<br />

Hom G (P, Q) = { g P ⊂ Q ∣ ∣ g ∈ G}. In particular AutG (P ) =<br />

N G (P ), the normalizer.<br />

2. The p-fusion system F p (G) is a quotient category <strong>of</strong> G ∝ S p ,<br />

given by Hom Fp (G)(P, Q) = Hom G (P, Q)/C G (P ).<br />

3. The p-orbit category O p (G) is a quotient category <strong>of</strong> G ∝ P,<br />

given by Hom Op (G)(P, Q) = Q\Hom G (P, Q).<br />

Example 6.1.16. Let b be a p-block <strong>of</strong> kG.<br />

1. The b-transporter category Tr b (G) = G ∝ S b has all nonidentity<br />

subgroups as its objects. For any p-subgroups P, Q,<br />

the morphism set is <strong>of</strong>ten written as Hom G∝Sb ((P, e P ), (Q, e Q )) =<br />

{ g (P, e P ) ⊂ (Q, e Q ) ∣ ∣ g ∈ G}. In particular AutG∝Sb (P ) =<br />

N G (P, e P ).<br />

2. The Brauer category, or p-fusion system, F b (G) is a quotient<br />

category <strong>of</strong> G ∝ S b , given by Hom Fb (G)(P, Q) =<br />

Hom G∝Sb (P, Q)/C G (P ).<br />

ρ


226 6 Connections with group representations <strong>and</strong> cohomology<br />

3. The b-orbit category O b (G) <strong>of</strong> F b (G) is a quotient category<br />

<strong>of</strong> F b (G), given by Hom Ob (G)(P, Q) = Q\Hom Fb (G)(P, Q).<br />

When b is the principal block, then the first two categories<br />

are isomorphic to the first two in Example 6.1.15. The b-orbit<br />

category is quite different from the p-orbit category because it<br />

is commonly believed that in non-principal block situation the<br />

latter construction is better than the former in capturing p-<br />

local information <strong>of</strong> the block b. This explains why transporter<br />

categories are important but still not enough <strong>and</strong> so we have<br />

to examine various quotient categories.<br />

6.2 Properties <strong>of</strong> local categories<br />

In this section, we illustrate the role <strong>of</strong> transporter categories<br />

in group representations. Then we continue to give several results<br />

on representations <strong>of</strong> transporter categories, which further<br />

demonstrate close connections between representations <strong>of</strong><br />

groups <strong>and</strong> transporter categories. This section will end with<br />

two applications, <strong>of</strong> representation theory, to cohomology <strong>of</strong><br />

transporter category algebras.<br />

6.2.1 Two diagrams <strong>of</strong> categories<br />

We have seen that given a G-poset P there exists a diagram<br />

<strong>of</strong> categories <strong>and</strong> functors<br />

π<br />

G ∝ P<br />

G<br />

C<br />

for any given local category C. Diagrams similar to this<br />

have been (implicitly) considered by Ronan-Smith [65, 6] <strong>and</strong><br />

Dwyer [17]. Here we pursue a direction that is closely related<br />

to the work <strong>of</strong> Ronan <strong>and</strong> Smith. The reader is referred to [17]<br />

ρ


6.2 Properties <strong>of</strong> local categories 227<br />

for Dwyer’s beautiful results on homology decompositions <strong>of</strong><br />

classifying spaces <strong>of</strong> groups.<br />

In practice one <strong>of</strong>ten finds that the functor G ∝ P → C<br />

is part <strong>of</strong> an extension (or an opposite extension) sequence<br />

<strong>of</strong> categories. (Among examples are various orbit categories,<br />

Brauer categories <strong>and</strong> Puig categories.) It means that for such<br />

a quotient category C there exists a category K which is a<br />

disjoint union <strong>of</strong> subgroups <strong>of</strong> Aut G∝P (x), x running over<br />

Ob P = Ob(G ∝ P), such that we can add K into the picture<br />

K<br />

π<br />

<br />

G ∝ P<br />

G<br />

C<br />

<strong>and</strong> moreover K ↩→ G ∝ P ↠ C satisfies some natural conditions<br />

provided in Section 4.1.. It will helps us to underst<strong>and</strong><br />

relationship between k(G ∝ P)-mod <strong>and</strong> kC-mod.<br />

For any M ∈ kG-mod, sometimes we denote by κ M the<br />

restriction Res π M <strong>and</strong> call it a constant value module.<br />

ρ<br />

6.2.2 Frobenius Reciprocity<br />

Applying classical tools in homological algebra, particularly<br />

the Kan extensions, we obtain a Frobenius Reciprocity between<br />

kG-mod <strong>and</strong> kC-mod, where kC is the (k-)category<br />

algebra <strong>of</strong> C. It implies, to some extent, comparing kG-mod<br />

with kC-mod provides an extended context for local representation<br />

theory. This observation illustrates a possible way to<br />

underst<strong>and</strong> group representations <strong>and</strong> cohomology via those<br />

<strong>of</strong> suitable categories. From the preceding diagram we obtain<br />

a diagram <strong>of</strong> module categories


228 6 Connections with group representations <strong>and</strong> cohomology<br />

LK π ,RK π<br />

k(G ∝ P)-mod<br />

LK ρ ,RK ρ<br />

<br />

Res π<br />

Res ρ<br />

kG-mod<br />

kC-mod.<br />

The adjunctions between the restrictions <strong>and</strong> Kan extensions<br />

have the following consequences.<br />

Proposition 6.2.1 (Frobenius Reciprocity). Suppose P<br />

is a G-poset <strong>and</strong> C is a quotient category <strong>of</strong> G ∝ P as in<br />

the preceding diagrams. Let M, N ∈ kG-mod <strong>and</strong> m, n ∈ kCmod.<br />

Then<br />

1. Hom kG (M, RK π Res ρ n) ∼ = Hom kC (LK ρ Res π M, n);<br />

2. Hom kG (LK π Res ρ m, N) ∼ = Hom kC (m, RK ρ Res π N).<br />

By direct calculations, these particular Kan extensions in<br />

Proposition 6.2.1 are simplified:<br />

•LK π<br />

∼ = lim −→P<br />

, RK π<br />

∼ = lim ←−P<br />

(used by Ronan-Smith. See<br />

Corollary 6.3. for a pro<strong>of</strong>);<br />

•LK ρ<br />

∼ =↑<br />

kC<br />

k(G∝P)<br />

(the induction), RK ρ<br />

∼ =⇑<br />

kC<br />

k(G∝P)<br />

(the coinduction),<br />

since ρ induces an algebra homomorphism k(G ∝<br />

P) → kC when C is a quotient category.<br />

Remark 6.2.2. For any n ∈ kC-mod, we shall write the k(G ∝<br />

P)-module Res ρ n as n because they share the same underlying<br />

vector space. Recall that κ M = Res π M. Then the Frobenius<br />

Reciprocity can be rewritten as<br />

(i’) Hom kG (M, lim ←−P<br />

n) ∼ = Hom kC (κ M ↑ kC<br />

k(G∝P) , n);<br />

(ii’) Hom kG (lim −→P<br />

m, N) ∼ = Hom kC (m, κ N ⇑ kC<br />

k(G∝P) ).<br />

When P = G/H for some subgroup H, we have natural<br />

isomorphisms ←−G/H<br />

lim ∼ = lim ∼<br />

−→G/H<br />

=↑<br />

G<br />

H . Then the above<br />

isomorphisms certainly become the usual adjunctions between<br />

↑ G H <strong>and</strong> ↓G H (the usual Frobenius Reciprocity) with


6.3 The functor π: group representations via transporter categories 229<br />

C = G ∝ (G/H) <strong>and</strong> ρ = Id, in light <strong>of</strong> the Morita equivalence<br />

between kC <strong>and</strong> kH.<br />

Remark 6.2.3. Our Frobenius reciprocity is different from a<br />

similar result <strong>of</strong> Ronan-Smith, see [6, 7.2.4], where they (implicitly)<br />

had a diagram <strong>of</strong> the same shape. However their<br />

C = G ∝ Q, not necessarily a quotient <strong>of</strong> G ∝ P, is another<br />

transporter category <strong>and</strong> ρ is induced by a G-map P → Q.<br />

This prohibits us from considering various quotients <strong>of</strong> transporter<br />

categories. Moreover since a G-map P → Q usually<br />

does not induce an algebra homomorphism from k(G ∝ P)<br />

to k(G ∝ Q), their Kan extensions cannot be interpreted as<br />

induction <strong>and</strong> coinduction.<br />

The functors ↑ kC<br />

k(G∝P)<br />

<strong>and</strong> ⇑kC<br />

k(G∝P)<br />

admit interesting interpretations<br />

when G ∝ P → C is part <strong>of</strong> an extension (or an<br />

opposite extension) sequence <strong>of</strong> categories. Under the circumstance<br />

↑ kC<br />

k(G∝P)<br />

<strong>and</strong> ⇑kC<br />

k(G∝P)<br />

on certain k(G ∝ P)-modules can<br />

be very well understood. We shall discuss it in Section 5.<br />

6.3 The functor π: group representations via transporter categories<br />

6.3.1 Homology representations via transporter categories<br />

Suppose P is a G-poset <strong>and</strong> π is the natural functor from the<br />

transporter category G ∝ P to G, regarded as a category with<br />

one object •. The naive fibre <strong>of</strong> the functor π, i.e. π −1 (•), is<br />

exactly P. On the other h<strong>and</strong>, the overcategory π/• provides<br />

the (topological) fibre <strong>of</strong> the map Bπ in the sense that the<br />

sequence <strong>of</strong> categories<br />

π/• → G ∝ P → G<br />

corresponds to a fibration after passing to classifying spaces<br />

B(π/•) → B(G ∝ P) → BG.


230 6 Connections with group representations <strong>and</strong> cohomology<br />

It leads to an action <strong>of</strong> G = π 1 (BG) on B(π/•), which is<br />

realized by a G-action on the category π/•, described shortly<br />

after Proposition 6.3.1. We shall see that there exists a functor<br />

π/• → P as well as an inclusion P → π/•, inducing a<br />

category equivalence. Interestingly the former is a G-functor,<br />

while the latter is not.<br />

The objects <strong>of</strong> π/• are <strong>of</strong> the form (x, h), in which x ∈<br />

Ob(G ∝ P) = Ob P <strong>and</strong> h ∈ G. A morphism from (x, h)<br />

to (x ′ , h ′ ) is a morphism (g, gx ≤ x ′ ) ∈ Mor(G ∝ P) such<br />

that h ′ g = h or equivalently g = h ′−1 h. It implies that each<br />

object (x, h) ∼ = (x ′ , h ′ ) if <strong>and</strong> only if x ′ = gx <strong>and</strong> h ′ = hg −1<br />

for some g ∈ G. Indeed the objects isomorphic to (x, h) are<br />

{(gx, hg −1 ) ∣ ∣ g ∈ G}. Particularly (x, h) ∼ = (hx, e) for the<br />

identity e ∈ G.<br />

There is also an undercategory •\π. One can put it in place<br />

<strong>of</strong> the overcategory <strong>and</strong> all our observations stay true. The<br />

objects <strong>of</strong> •\π are <strong>of</strong> the form (h, x), in which x ∈ Ob(G ∝<br />

P) = Ob P <strong>and</strong> h ∈ G. A morphism from (h, x) to (h ′ , x ′ ) is<br />

a morphism (g, gx ≤ x ′ ) ∈ Mor(G ∝ P) such that gh ′ = h.<br />

It implies that each object (h, x) ∼ = (h ′ , x ′ ) if <strong>and</strong> only if<br />

x ′ = gx <strong>and</strong> h ′ = g −1 h for some g ∈ G. Indeed the objects<br />

isomorphic to (h, x) are {(g −1 h, gx) ∣ ∣ g ∈ G}. Particularly<br />

(h, x) ∼ = (e, hx) for the identity e ∈ G.<br />

When P = •, G ∝ • ∼ = G <strong>and</strong> the functor π can be identified<br />

with Id G . Consequently π/• ∼ = Id G /•. The following<br />

result generalizes this special situation <strong>and</strong> gives a characterization<br />

<strong>of</strong> the overcategory <strong>and</strong> undercategory coming from<br />

G ∝ P → G.<br />

Proposition 6.3.1. The category π/• is isomorphic to P ×<br />

(Id G /•), <strong>and</strong> •\π is isomorphic to P×(•\Id G ). Consequently<br />

π/• ∼ = •\π.


6.3 The functor π: group representations via transporter categories 231<br />

Pro<strong>of</strong>. We establish an isomorphism φ : P × (Id G /•) → π/•<br />

as follows. The objects <strong>of</strong> P × (Id G /•) are {(x, (•, g)) ∣ g ∈ G, x ∈ Ob P}. We define φ((x, (•, g))) = (g −1 x, g) ∈<br />

Ob(π/•). For a morphism (x 1 ≤ x 2 , g2 −1 g 1) : (x 1 , (•, g 1 )) →<br />

(x 2 , (•, g 2 )) we put<br />

φ((x 1 ≤ x 2 , g2 −1 g 1)) := (g2 −1 g 1, (g2 −1 g 1)(g1 −1 x 1) ≤ g2 −1 x 2),<br />

a morphism from (g1 −1 x 1, g 1 ) to (g2 −1 x 2, g 2 ). We can write<br />

down its inverse ψ : π/• → P ×(Id G /•), given by ψ((y, h)) :=<br />

(hy, (•, h)). For any morphism in π/•, (h −1<br />

2 h 1, h −1<br />

2 h 1y 1 ≤ y 2 ) :<br />

(y 1 , h 1 ) → (y 2 , h 2 ), we define ψ((h −1<br />

2 h 1, h −1<br />

2 h 1y 1 ≤ y 2 )) :=<br />

(h 1 y 1 ≤ h 2 y 2 , h2 −1 h 1) : (h 1 y 1 , (•, h 1 )) → (h 2 y 2 , (•, h 2 )).<br />

The isomorphism for •\π can be similarly obtained.<br />

In the end, from Example 1.2.7 (1), we have an isomorphism<br />

Id G /• ∼ = •\Id G <strong>and</strong> hence the isomorphism between the over<br />

category <strong>and</strong> undercategory.<br />

⊓⊔<br />

As we have shown in Example 1.2.7 (2) that both •\Id G <strong>and</strong><br />

Id G /• are equivalent to •. For instance, there is the canonical<br />

functor pt : Id G /• → • as well as a functor • → Id G /•, given<br />

by • ↦→ (•, e), where e is the identity <strong>of</strong> G. The reader can<br />

quickly verify that they provide a category equivalence. These<br />

two categories are actually the Cayley graph.<br />

Corollary 6.3.2. There exists a natural embedding P ↩→<br />

π/• (or P ↩→ •\π) making P a skeleton <strong>of</strong> π/• (or •\π).<br />

Consequently LK π m ∼ = lim −→P<br />

m <strong>and</strong> RK π m ∼ = lim ←−P<br />

m for any<br />

m ∈ k(G ∝ P)-mod.<br />

Pro<strong>of</strong>. By the preceding proposition, π/• ∼ = P × Id G /•. Since<br />

we know explicitly the equivalence-inducing functors between<br />

Id G /• <strong>and</strong> •, we can easily translate them for π/•.<br />

The natural functor P → π/•, given by x ↦→ (x, e) <strong>and</strong><br />

x ≤ y ↦→ (e, x ≤ y), is an embedding, sending P to a skeleton


232 6 Connections with group representations <strong>and</strong> cohomology<br />

<strong>of</strong> π/•. It is straightforward to check that there is a natural<br />

surjective functor π/• → P, induced by (x, h) ↦→ hx. For any<br />

morphism in π/•, (h −1<br />

2 h 1, h −1<br />

2 h 1y 1 ≤ y 2 ) : (y 1 , h 1 ) → (y 2 , h 2 ),<br />

we define ψ((h −1<br />

2 h 1, h2 −1 h 1y 1 ≤ y 2 )) := h 1 y 1 ≤ h 2 y 2 .<br />

These two functors provide an equivalence between π/• <strong>and</strong><br />

P. Same equivalence can be established between P <strong>and</strong> •\π.<br />

The existence <strong>of</strong> equivalences between these categories forces<br />

lim m ∼ = lim<br />

−→π/• −→P<br />

m <strong>and</strong> ←−•\π<br />

lim m ∼ = lim ←−P<br />

m.<br />

⊓⊔<br />

Next we shall show that both π/• <strong>and</strong> •\π are G-categories.<br />

Moreover we study whether or not the above functors, inducing<br />

equivalences with P, are compatible with G-action. Before<br />

we do so, we need to specify what we mean by this compatibility.<br />

Definition 6.3.3. Let D, C be two G-categories <strong>and</strong> u : D →<br />

C a functor. We say u is a G-functor if for any g ∈ G <strong>and</strong><br />

x ∈ Ob D we have u(gx) = gu(x), <strong>and</strong> for any α ∈ Mor D,<br />

gu(α) = u(gα).<br />

We first describe the G-actions on Id G /• <strong>and</strong> on •\Id G . They<br />

are given on objects by u · (•, g ′ ) = (•, ug ′ ) <strong>and</strong> u · (g ′′ , •) =<br />

(g ′′ u −1 , •), respectively. On morphisms,<br />

while<br />

u · [(•, g 1 ) g−1 2 g 1<br />

−→(•, g 2 )] = (•, ug 1 ) g−1 2 g 1<br />

−→(•, ug 2 )<br />

u · [(g 1, ′ •) g′ 2 g′ −1<br />

1<br />

−→ (g 2, ′ •)] = (g 1u ′ −1 , •) g′ 2 g′ −1<br />

1<br />

−→ (g 2u ′ −1 , •).<br />

Immediately we see that the surjection Id G /• → • is a G-<br />

functor, <strong>and</strong> by contrast the embedding • ↩→ Id G /• is not a<br />

G-functor, because g• = • ↦→ (•, e) by the embedding while<br />

g(•, e) = (•, g), for any g ∈ G. Hence we do not expect the<br />

embedding P ↩→ π/• to be a G-functor.


6.3 The functor π: group representations via transporter categories 233<br />

Based on the G-actions on Id G /• <strong>and</strong> •\Id G <strong>and</strong> Proposition<br />

6.3.1, we can define G-actions on π/• <strong>and</strong> •\π. For any object<br />

(x, h) ∈ Ob(π/•) <strong>and</strong> u ∈ G, we have u · (x, h) = (x, uh),<br />

<strong>and</strong> for any morphism (g, gx ≤ x ′ ) : (x, h) → (x ′ , h ′ ) we<br />

have u · (g, gx ≤ x ′ ) = (g, gx ≤ x ′ ) : (x, uh) → (x ′ , uh ′ ).<br />

Note that g = h ′−1 h = (uh ′ ) −1 (uh). Similarly For any object<br />

(x, h) ∈ Ob(•\π) <strong>and</strong> u ∈ G, we have u · (x, h) = (x, hu −1 ),<br />

<strong>and</strong> for any morphism (g, gx ≤ x ′ ) : (x, h) → (x ′ , h ′ ) we<br />

have u · (g, gx ≤ x ′ ) = (g, gx ≤ x ′ ) : (x, hu −1 ) → (x ′ , h ′ u −1 ),<br />

with g = h ′ h −1 = (h ′ u −1 )(hu −1 ) −1 . From our constructions <strong>of</strong><br />

G-actions on π/• <strong>and</strong> •\π, we reach the following statements.<br />

Corollary 6.3.4.1. The isomorphisms π/• ∼ = P ×Id G /• <strong>and</strong><br />

•\π ∼ = P × •\Id G are G-isomorphisms.<br />

2. The natural functors π/• ↠ P <strong>and</strong> •\π ↠ P are G-<br />

functors.<br />

As we mentioned earlier, a G-category C gives rise to a complex<br />

<strong>of</strong> kG-modules C ∗ (C, k). Furthermore if u : D → C is<br />

a G-functor between two G-categories, then it induces a G-<br />

simplicial map u ∗ : ND ∗ → NC ∗ , <strong>and</strong> hence a chain map<br />

between complexes <strong>of</strong> kG-modules u : C ∗ (D, k) → C ∗ (C, k).<br />

Example 6.3.5. If H ⊂ K are subgroups <strong>of</strong> G, then there<br />

is a G-functor G/H → G/K. Hence we have a chain map<br />

C ∗ (G/H, k) → C ∗ (G/K, k). Since both complexes concentrate<br />

in degree zero, this chain map consists <strong>of</strong> only one kGmap:<br />

k(G/H) = C 0 (G/H, k) → k(G/K) = C 0 (G/K, k).<br />

Corollary 6.3.6. The G-functors π/• → P <strong>and</strong> •\π → P<br />

induce chain maps between complexes <strong>of</strong> kG-modules. The<br />

complex LK π B∗<br />

G∝P ∼ = C∗ (π/•, k) is a projective resolution<br />

<strong>of</strong> the finite complex <strong>of</strong> kG-modules C ∗ (P).<br />

Pro<strong>of</strong>. Since LK π preserves projectives, LK π B∗<br />

G∝P ∼ = C∗ (π/•, k)<br />

is a complex <strong>of</strong> projective kG-modules. The existing G-


234 6 Connections with group representations <strong>and</strong> cohomology<br />

functor π/• → P gives rise to a chain map <strong>of</strong> complexes<br />

<strong>of</strong> kG-modules LK π B∗<br />

G∝P ∼ = C∗ (π/•, k) → C ∗ (P, k). However<br />

since that G-functor is a category equivalence, it induces<br />

an isomorphism between the homology <strong>of</strong> complexes. ⊓⊔<br />

Remark 6.3.7. Any left kG-module is naturally a right kGmodule.<br />

If we consider C ∗ (P, k) as a complex <strong>of</strong> right kGmodules<br />

through (−) · g := g −1 · (−), then a projective resolution<br />

can be obtained as C ∗ (•\π, k) ∼ = RK π B∗<br />

G in which<br />

we take the right bar resolution <strong>of</strong> k ∈mod-k(G ∝ P). If we<br />

regard the complex <strong>of</strong> right modules C ∗ (•\π, k) as a complex<br />

<strong>of</strong> left modules via g · (−) := (−) · g −1 , then it is isomorphic<br />

to C ∗ (π/•, k). In fact, the isomorphism comes from the<br />

isomorphism <strong>of</strong> categories π/• ∼ = •\π.<br />

It should be useful to underst<strong>and</strong> the complexes C ∗ (π/•, k)<br />

<strong>and</strong> C ∗ (•\π, k).<br />

Example 6.3.8. By Proposition 6.3.1, if P = G/G = •, then<br />

π G/G = Id G <strong>and</strong> π G/G /• = Id G /• is the Cayley graph, giving<br />

rise to the total space EG whose complex is the bar resolution<br />

B G ∗ . More generally for P = G/H, we have an isomorphism<br />

<strong>of</strong> complexes <strong>of</strong> kG-modules C ∗ (π/•) ∼ = B G ∗ ⊗k(G/H), which<br />

provide a projective resolution <strong>of</strong> k(G/H) = C 0 (G/H) =<br />

C ∗ (G/H).<br />

For any small category C, there is an augmentation map<br />

ɛ : C ∗ (C, k) → k, which is zero on positive degrees <strong>and</strong> which<br />

maps every base element in C 0 (C, k) = k Ob C to 1 ∈ k.<br />

Corollary 6.3.9.1. The G-isomorphisms π/• ∼ = P × Id G /•<br />

<strong>and</strong> •\π ∼ = P × •\Id G induce isomorphisms between complexes<br />

<strong>of</strong> kG-modules.<br />

2. We have equivalences <strong>of</strong> complexes <strong>of</strong> kG-modules C ∗ (π/•, k) ≃<br />

C ∗ (P, k) ⊗ C ∗ (Id G /•, k) <strong>and</strong> C ∗ (•\π, k) ≃ C ∗ (P, k) ⊗


6.3 The functor π: group representations via transporter categories 235<br />

C ∗ (•\Id G , k). Furthermore C ∗ (π/•) ≃ C ∗ (P) ⊗ B∗<br />

G <strong>and</strong><br />

C ∗ (•\π) ≃ C ∗ (P) ⊗ B∗ G .<br />

3. The chain map C ∗ (P × Id G /•, k) → C ∗ (Id G /•, k), induced<br />

by P → •, corresponds to the chain map C ∗ (P, k) ⊗<br />

C ∗ (Id G /•, k) → C ∗ (Id G /•, k), induced by the augmentation<br />

map C ∗ (P, k) → k.<br />

Pro<strong>of</strong>. The first statement is a consequence <strong>of</strong> the preceding<br />

corollary.<br />

Since π/• ∼ = P × Id G /•, we have isomorphism C ∗ (π/•, k) ∼ =<br />

C ∗ (P × Id G /•, k). By Theorem 2.2.31, there is a natural<br />

chain homotopy equivalence C ∗ (P × Id G /•, k) → C ∗ (P, k) ⊗<br />

C ∗ (Id G /•, k), called the Alex<strong>and</strong>er-Whitney map. From its<br />

definition, one can see it is a chain map <strong>of</strong> kG-modules.<br />

To see the third statement, we draw a commutative diagram<br />

C ∗ (P × Id G /•, k)<br />

≃ <br />

C ∗ (P, k) ⊗ C ∗ (Id G /•, k)<br />

ɛ⊗1<br />

C ∗ (Id G /•, k) k ⊗ C ∗ (Id G /•, k)<br />

Suppose (x 0 → · · · → x n ) ⊗ (g 0 → · · · → g n ) in a base<br />

element <strong>of</strong> C n (P × Id G /•, k). Then the Alex<strong>and</strong>er-Whitney<br />

map sends it to<br />

n∑<br />

(x 0 → · · · → x i ) ⊗ (g i → · · · → g n ),<br />

i=0<br />

whose image under the right vertical map equals ɛ ⊗ 1(x 0 ⊗<br />

(g 0 → · · · → g n )) = g 0 → · · · → g n . This is exactly the image<br />

<strong>of</strong> (x 0 → · · · → x n ) ⊗ (g 0 → · · · → g n ) under the left vertical<br />

map, which is a projection map.<br />

⊓⊔<br />

The last corollary <strong>of</strong> Proposition 6.3.1 is actually a consequence<br />

<strong>of</strong> Corollary 6.3.2. Since it is closely related to the<br />

second main result in this section, we put it here.


236 6 Connections with group representations <strong>and</strong> cohomology<br />

Corollary 6.3.10. Suppose M ∈ kG-mod <strong>and</strong> n ∈ k(G ∝<br />

P)-mod. Then LK π (κ M ˆ⊗n) ∼ = M⊗LK π n <strong>and</strong> RK π (κ M ˆ⊗n) ∼ =<br />

M ⊗ RK π n as kG-modules. In particular LK π (κ M ) ∼ = M ⊗<br />

LK π k <strong>and</strong> RK π (κ M ) ∼ = M⊗RK π k, where LK π k ∼ = H0(BP, k) ∼ =<br />

H 0 (BP, k) ∼ = RK π k <strong>of</strong> dimension equal to the number <strong>of</strong><br />

connected components <strong>of</strong> P.<br />

Pro<strong>of</strong>. For the left Kan extension we have<br />

LK π (κ M ˆ⊗n) ∼ = lim −→P<br />

(κ M ˆ⊗n) ∼ = M ⊗ lim −→P<br />

n ∼ = M ⊗ LK π n.<br />

The second isomorphism is true because κ M as a kP-module<br />

admits trivial action. The statement for the right Kan extension<br />

is similar.<br />

⊓⊔<br />

The above corollary implies that LK π κ M<br />

∼ = RKπ κ M , suggesting<br />

the existence <strong>of</strong> a transfer map, which we will construct<br />

later on.<br />

The next result is a direct generalization <strong>of</strong> the fact that the<br />

two obvious kG-module structures on P ⊗ M are isomorphic,<br />

for P, M ∈ kG-mod with P projective. It reveals another connection<br />

between representations <strong>of</strong> groups <strong>and</strong> <strong>of</strong> transporter<br />

categories.<br />

Theorem 6.3.11. Let P ∈ k(G ∝ P)-mod be a projective<br />

module <strong>and</strong> κ M = Res π M for some M ∈ kG-mod. Then<br />

P ˆ⊗κ M is a projective k(G ∝ P)-module. Consequently<br />

ˆ⊗κ M → k ˆ⊗κ M = κ M → 0 is a projective resolution.<br />

B G∝P<br />

∗<br />

Pro<strong>of</strong>. We will prove P ˆ⊗κ M<br />

∼ = P⊗M, with k(G ∝ P) acting<br />

on the latter via left multiplication. To this end, we assume<br />

P = kHom G∝P (x, −). The pro<strong>of</strong> is entirely analogues to the<br />

case when P = •, i.e. when G ∝ • = G.<br />

We define a k-linear map ϕ : kHom G∝P (x, −) ⊗ M →<br />

kHom G∝P (x, −) ˆ⊗κ M as follows. On base elements ϕ((g, gx ≤<br />

y) ⊗ m) = (g, gx ≤ y) ⊗ (g, gx ≤ y)m, where the latter m is


6.3 The functor π: group representations via transporter categories 237<br />

considered as an element in κ M (x). For any (h, hy ≤ z), we<br />

readily verify<br />

(h, hy ≤ z)ϕ[(g, gx ≤ y)⊗m] = ϕ[(h, hy ≤ z)((g, gx ≤ y)⊗m)].<br />

Thus ϕ is a homomorphism <strong>of</strong> k(G ∝ P)-modules.<br />

We remind the reader that, following definition, given any<br />

pair <strong>of</strong> x, y ∈ Ob(G ∝ P) with Hom G∝P (x, y) non-empty,<br />

both Aut G∝P (x) <strong>and</strong> Aut G∝P (y) act freely on Hom G∝P (x, y).<br />

This implies that kHom G∝P (x, y) is a free kAut G∝P (x)- or<br />

kAut G∝P (y)-module. Consequently ϕ restricts on each y to<br />

the classical kAut G∝P (y)-isomorphism (see [5, 3.1.5])<br />

kHom G∝P (x, y)⊗M → (kHom G∝P (x, −) ˆ⊗κ M )(y) = kHom G∝P (x, y)⊗<br />

Furthermore because as vector spaces<br />

kHom G∝P (x, −) ˆ⊗κ M =<br />

⊕<br />

kHom G∝P (x, y)⊗M = kHom G∝P (x<br />

y∈Ob(G∝P)<br />

the linear map ϕ is actually one-to-one <strong>and</strong> hence an isomorphism<br />

<strong>of</strong> k(G ∝ P)-modules.<br />

⊓⊔<br />

For an arbitrary category algebra, the tensor product <strong>of</strong> a<br />

projective module with a non-trivial module usually does not<br />

stay projective, as one can find a counter-example in [, 2.5].<br />

The key point here is that the structure <strong>of</strong> G ∝ P allows us<br />

to apply some results on group algebras.<br />

Assume C ∗ is a complex <strong>of</strong> kG-modules. Then we naturally<br />

obtain a complex <strong>of</strong> k(G ∝ P)-modules via restriction, written<br />

as κ C∗ .<br />

Lemma 6.3.12. The functor π : G ∝ P → G induces a<br />

natural chain map Π from B∗ G∝P → k → 0 to κ B<br />

G ∗<br />

→ k → 0.<br />

More generally for any M ∈ kG-mod, it naturally induces<br />

a chain map between exact sequences <strong>of</strong> k(G ∝ P)-modules


238 6 Connections with group representations <strong>and</strong> cohomology<br />

Π M : {B∗ G∝P ˆ⊗κ M → κ M → 0} → {κ B<br />

G ∗<br />

ˆ⊗κ M = κ B<br />

G ∗ ⊗M → κ M → 0}.<br />

Pro<strong>of</strong>. The reason is that π induces a natural map between<br />

the nerves <strong>of</strong> these categories, while the bar resolutions are<br />

constructed from the chain complexes from the nerves. More<br />

explicitly the complexes B∗ G∝P → k → 0 <strong>and</strong> κ B<br />

G ∗<br />

→ k → 0<br />

evaluated at any x ∈ Ob(G ∝ P) are the augmented chain<br />

complexes C ∗ (Id G∝P /x) → k → 0 <strong>and</strong> B∗<br />

G → k → 0, respectively.<br />

The chain map Π is induced by π[(g, ga ≤ b)] = g.<br />

The construction <strong>of</strong> Π M for a fixed M ∈ kG-mod is similar.<br />

⊓⊔<br />

6.3.2 On finite generation <strong>of</strong> cohomology<br />

Subsections 6.4.2 <strong>and</strong> 6.4.3 contains two applications <strong>of</strong> results<br />

from Subsection 6.4.1 to transporter category cohomology<br />

<strong>and</strong> its connection with group cohomology. To study the<br />

functor ρ : G ∝ P → C, we will start a new Section 6.5.<br />

Let G be a finite group <strong>and</strong> P a finite G-poset. Then there<br />

exists a sequence <strong>of</strong> functors<br />

P ι →G ∝ P π →G,<br />

where ι is the natural embedding, <strong>and</strong> whose topological realization<br />

is a fibration<br />

BP<br />

=<br />

Bι <br />

B(G ∝ P) Bπ<br />

≃<br />

<br />

BG<br />

≃<br />

<br />

BP<br />

<br />

EG × G BP<br />

<br />

EG × G •<br />

where • is a point fixed by G. In order to study the finite<br />

generation <strong>of</strong> cohomology rings, we need to recall the<br />

Grothendieck cohomology spectral sequence for a functor<br />

u : D → C


6.3 The functor π: group representations via transporter categories 239<br />

H i (C; H j (\u; N)) ⇒ H i+j (D; N),<br />

where x\u is the undercategory for each x ∈ Ob C <strong>and</strong> N is an<br />

kD-module which can be regarded as a x\u-module through<br />

the forgetful functor x\u → D. Since G only has one object,<br />

the Grothendieck spectral sequence for π : G ∝ P → G reads<br />

as follows<br />

H i (G; H j (•\π; N)) ⇒ H i+j (G ∝ P; N),<br />

or, for being consistent with our Ext notation,<br />

Ext i kG(k, Ext j k(•\π)<br />

(k, N)) ⇒ Exti+j<br />

k(G∝P)<br />

(k, N),<br />

for any N ∈ k(G ∝ P)-mod, in which •\π has its skeleton is<br />

isomorphic to the poset P. By Section 4., the above spectral<br />

sequence is a module over<br />

Ext i kG(k, Ext j k(•\π)<br />

(k, k)) ⇒ Exti+j<br />

k(G∝P)<br />

(k, k).<br />

Meanwhile we have a morphism between the following Grothendieck<br />

spectral sequences, induced by<br />

P<br />

ι <br />

G ∝ P π<br />

π<br />

<br />

G<br />

•<br />

<br />

G ∝ • ∼=<br />

<br />

G,<br />

from which we obtain a diagram <strong>of</strong> spectral sequences<br />

Ext i kG(k, Ext j collapses<br />

<br />

k•<br />

(k, k))<br />

Ext i kG(k, Ext j k(•\π) (k, k)) <br />

Ext i+j<br />

Ext i+j<br />

k(G∝•)<br />

(k, k) .<br />

k(G∝P) (k, k)<br />

More precisely it makes the lower spectral sequence, <strong>and</strong> hence<br />

Ext i kG(k, Ext j k(•\π)<br />

(k, N)) ⇒ Exti+j<br />

k(G∝P)<br />

(k, N),


240 6 Connections with group representations <strong>and</strong> cohomology<br />

modules over Ext ∗ k(G∝•)(k, k) = Ext ∗ kG(k, k). We point out<br />

that the group cohomology ring H ∗ (G, k) ∼ = Ext ∗ kG(k, k) acts<br />

on<br />

H ∗ (G ∝ P; N) ∼ = Ext ∗ k(G∝P)(k, N)<br />

via the algebra homomorphism induced by π (or Bπ),<br />

Ext ∗ kG(k, k) ∼ = H ∗ (G, k) → H ∗ (|G ∝ P|, k) ∼ = H ∗ (G ∝ P; k) ∼ = Ext ∗ k(G<br />

Since P has the property that k is <strong>of</strong> finite projective dimension,<br />

we know H j (•\π; −) ∼ = Ext j k(•\π)<br />

(k, −) vanishes for large<br />

j. Furthermore, the well-known theorem <strong>of</strong> Evens <strong>and</strong> Venkov<br />

says that, for each j, the module Ext ∗ kG(k, Ext j k(•\π)<br />

(k, N)) is<br />

a finitely generated over Ext ∗ kG(k, k). Since E ∞ is a subquotient<br />

<strong>of</strong> E 2 <strong>of</strong> a cohomology spectral sequence, we have the<br />

following statement.<br />

Lemma 6.3.13. For any N ∈ kTr P (G)-mod, Ext ∗ k(G∝P)(k, N)<br />

is a finitely generated Ext ∗ kG(k, k)- <strong>and</strong> Ext ∗ k(G∝P)(k, k)-module.<br />

Theorem 6.3.11 helps us to extend this finite generation<br />

property.<br />

Proposition 6.3.14. For any M ∈ kG-mod <strong>and</strong> n ∈ k(G ∝<br />

P)-mod, Ext ∗ k(G∝P)(κ M , n) is finitely generated over Ext ∗ k(G∝P)(k, k).<br />

Pro<strong>of</strong>. One can easily deduce from Theorem 6.4.11, together<br />

with the internal hom in Section 3.4., an Eckmann-Sharpiro<br />

type isomorphism<br />

Ext ∗ k(G∝P)(κ M , n) ∼ = Ext ∗ k(G∝P)(k, Hom(κ M , n)).<br />

Then we apply the finite generation result that we just quoted.<br />

⊓⊔<br />

The ultimate goal is to establish the finite generation <strong>of</strong><br />

Ext ∗ k(G∝P)(m, n), for any m, n ∈ k(G ∝ P)-mod. However


6.3 The functor π: group representations via transporter categories 241<br />

the existence <strong>of</strong> an internal hom is not enough because, unless<br />

m is induced from a kG-module, Ext ∗ k(G∝P)(m, n) ≁ =<br />

Ext ∗ k(G∝P)(k, Hom(m, n)). We shall establish the finite generation<br />

theorem, Theorem 6.5.3, in the very last section because<br />

we have to rely on the Hochschild cohomology <strong>of</strong> transporter<br />

categories.<br />

Example 6.3.15. Let G = C 2 = {g ∣ ∣ g 2 = e} <strong>and</strong> chark =<br />

2. Then the G-poset S e 2 is {e} → C 2 , <strong>and</strong> the transporter<br />

category C := C 2 ∝ S e 2 is<br />

e <br />

{e}<br />

e<br />

g<br />

g<br />

C 2 .<br />

There are two simple modules S {e} = S {e},k <strong>and</strong> S C2 =<br />

S C2 ,k, both <strong>of</strong> dimension 1. Furthermore there is a short<br />

exact sequence 0 → S C2 → P {e} → S {e} → 0. If we<br />

apply Ext ∗ kC(−, S C2 ) to it, then we get Ext i kC(S {e} , S C2 ) ∼ =<br />

Ext i−1<br />

kC (S {e}, S {e} ) for i > 0 <strong>and</strong> Ext 0 kC(S {e} , S C2 ) = 0. The<br />

cohomology ring is finitely generated because<br />

Ext ∗ kC(k, k) ∼ = Ext ∗ kC(S {e} , S {e} ) ∼ = Ext ∗ kC 2<br />

(k, k) ∼ = k[x]<br />

it is a polynomial ring with one indeterminant x <strong>of</strong> degree 1.<br />

We can see that Ext i kC(S {e} , S C2 ) is <strong>of</strong> dimension 1 for i > 0<br />

<strong>and</strong> Ext ∗ kC(S {e} , S C2 ) is finitely generated over Ext ∗ kC(k, k). By<br />

contrast,<br />

Ext ∗ kC(k, Hom(S {e} , S C2 )) = 0<br />

since Hom(S {e} , S C2 ) = 0 by direct calculation.<br />

Remark 6.3.16. Using Dan Swenson’s definition [72] <strong>of</strong> the<br />

internal hom one verifies that<br />

Consequently,<br />

κ Homk (M,N) ∼ = Hom(κ M , κ N ).<br />

e


242 6 Connections with group representations <strong>and</strong> cohomology<br />

Ext ∗ k(G∝P)(k, κ Homk (M,N)) ∼ = Ext ∗ k(G∝P)(k, Hom(κ M , κ N )) ∼ = Ext ∗ k(G∝P)<br />

Before moving to the next section, we record a connection<br />

between cohomology <strong>of</strong> transporter categories <strong>and</strong> equivariant<br />

cohomology, which is perhaps known to the experts.<br />

Proposition 6.3.17. The left Kan extension induces an<br />

isomorphism<br />

λ M : Ext ∗ k(G∝P)(k, κ M ) ∼ = H ∗ G(BP, M),<br />

where H ∗ G (BP, M) is the equivariant cohomology group for<br />

some M ∈ kG-mod.<br />

Pro<strong>of</strong>. Take the bar resolution B∗<br />

G∝P → k → 0 <strong>and</strong> consider<br />

the complex Hom k(G∝P) (B∗<br />

G∝P , κ M ). The left Kan extension<br />

induces a chain map<br />

Hom k(G∝P) (B∗<br />

G∝P , κ M ) ∼ = Hom kG (LK π B∗ G∝P , M) ∼ = Hom kG (B∗ G ⊗C ∗ (P<br />

But the rightmost term is<br />

Hom kG (B G ∗ ⊗ C ∗ (P), M) ∼ = Hom kG (B G ∗ , Hom k (C ∗ (P), M)),<br />

which gives rise to H ∗ G (BP, M).<br />

⊓⊔<br />

In light <strong>of</strong> the above proposition, we may introduce the Tate<br />

cohomology <strong>of</strong> transporter categories as Tate equivariant cohomology.<br />

With Remark 3 in mind, one can further define<br />

negative degree Ext groups Ext ∗ k(G∝P)(κ M , κ N ).<br />

6.3.3 Transfer for ordinary cohomology<br />

Let u : D → C be a functor between small categories. There<br />

is always a restriction res u : H ∗ (C; k) → H ∗ (D; k). However<br />

usually one cannot construct a map in the opposite direction,<br />

unless the two Kan extensions are connected by a natural<br />

transformation. In this section, based on our knowledge about


6.3 The functor π: group representations via transporter categories 243<br />

representations <strong>of</strong> k(G ∝ P), we establish various transfer<br />

maps, including the Becker-Gottlieb transfer map with respect<br />

to π P : G ∝ P → G. Here we provide an algebraic<br />

alternative to the construction <strong>of</strong> Becker-Gottlieb [4], <strong>and</strong> the<br />

core idea is taken from Dwyer-Wilkerson [21, 9.13]). Essentially<br />

our construction incorporates [21, 9.13] in an entirely<br />

representation-theoretic setting. The upshot is that our construction<br />

is analogues to the classical situation, see for instance<br />

[5, 3.6.17]. Keep in mind that Res π <strong>and</strong> LK π generalize<br />

↓ G H <strong>and</strong> ↑G H , respectively, used in group cohomology.<br />

In Corollary 6.3.9 (3), we described a chain map<br />

LK π B∗<br />

G∝P ∼ = C∗ (π/•, k) ∼ = C ∗ (P×Id G /•, k) −→ C ∗ (Id G /•, k) ∼ = B∗ G ,<br />

induced by P → •. This chain map gives rise to a cochain<br />

map<br />

Hom kG (B∗ G , k) → Hom kG (LK π B∗<br />

G∝P , k) ∼ = Hom k(G∝P) (B∗ G∝P , k)<br />

<strong>and</strong> hence the restriction<br />

res π : Ext ∗ kG(k, k) → Ext ∗ k(G∝P)(k, k).<br />

We want to establish a map, called transfer, on the opposite<br />

direction<br />

Ext ∗ k(G∝P)(k, k) → Ext ∗ kG(k, k).<br />

Example 6.3.18. Before we construct the transfer, we examine<br />

the special case for group cohomology. Suppose H ⊂ G is<br />

a subgroup. We want to define a map tr G H : H ∗ (H, k) →<br />

H ∗ (G, k).<br />

For the functor π G/H : G ∝ (G/H) → G. We get<br />

LK πG/H B G ∗ = C ∗ (π G/H /•, k) ∼ = C ∗ ((G/H) × Id G /•, k)<br />

as complexes <strong>of</strong> kG-modules. The rightmost admits a map<br />

to C ∗ (Id G /•, k), induced by the G-functor pt : G/H → •. If


244 6 Connections with group representations <strong>and</strong> cohomology<br />

we invoke C ∗ ((G/H) × Id G /•, k) ≃ k(G/H) ⊗ B G ∗ , then this<br />

chain map is identified with the chain map k(G/H) ⊗ B G ∗ →<br />

C ∗ (Id G /•, k), induced by the augmentation<br />

ɛ : k(G/H) = k ↑ G H→ k.<br />

By applying Hom kG (−, k), this chain map induces the restriction<br />

res πG/H : H ∗ (G, k) → H ∗ (G ∝ (G/H), k) ∼ = H ∗ (H, k).<br />

To produce another map in the opposite direction, we notice<br />

that although there is no G-functor • → G/H (unless G =<br />

H), we can always build a kG-map after linearization<br />

It induces a chain map<br />

k = k• → k(G/H) = k ↑ G H .<br />

B G ∗ → k ↑ G H ⊗B G ∗ .<br />

When we apply Hom kG (−, k) to the following <strong>and</strong> take cohomology<br />

B∗ G → k ↑ G H ⊗B∗<br />

G ɛ⊗1<br />

−→B∗ G ,<br />

since<br />

k ↑ G H ⊗B∗ G ∼ = (k ⊗ B∗ G ↓ G H) ↑ G H= B∗ G ↓ G H↑ G H,<br />

we get Hom kG (B∗ G ↓ G H ↑G H , k) ∼ = Hom kH (B∗ G ↓ G H , k) <strong>and</strong> consequently<br />

two induced maps<br />

H ∗ (G, k) trG H<br />

←−H ∗ (H, k) resG H<br />

←−H ∗ (G, k).<br />

Since the composite k → k(G/H) = k ↑ G H→ k equals a scalar<br />

multiplication by [G : H], we know tr G H ◦ resG H = [G : H] =<br />

|G/H|. Here the restriction res G H is identified with res π G/H<br />

upon<br />

the isomorphism H ∗ (G ∝ (G/H), k) ∼ = H ∗ (H, k).


6.3 The functor π: group representations via transporter categories 245<br />

Since for arbitrary G ∝ P, we have constructed the restriction<br />

res π : H ∗ (G, k) → H ∗ (G ∝ P, k).<br />

Now we would like to have a generalized transfer map in the<br />

opposite direction. As in the group case, it should be induced<br />

by some B∗ G → LK π B∗<br />

G∝P , or rather<br />

C ∗ (Id G /•, k) → C ∗ (π/•, k) ∼ = C ∗ (P×Id G /•, k) ≃ C ∗ (P, k)⊗C ∗ (Id G /<br />

Indeed, it should be reduced to establishing a map<br />

k = k• = C ∗ (•, k) → C ∗ (P, k).<br />

Note that, although there are many functors from • → P, all<br />

<strong>of</strong> them have the problem that they are not G-functors. Thus<br />

they will not directly produce chain maps between the above<br />

two complexes <strong>of</strong> kG-modules. Now we do have to work on the<br />

chain level. If such a chain map exists, then we immediately<br />

obtain a chain map<br />

C ∗ (•, k)⊗C ∗ (Id G /•, k) → C ∗ (P, k)⊗C ∗ (Id G /•, k) ≃ C ∗ (P×Id G /•, k)<br />

As the first attempt, we record the following observation.<br />

Lemma 6.3.19. For any n ≥ 0 <strong>and</strong> any x 0 → · · · →<br />

x n ∈ NP n , one can construct a chain map c i : C ∗ (•, k) →<br />

C ∗ (P, k), for each 0 ≤ i ≤ n.<br />

Pro<strong>of</strong>. Since C ∗ (•, k) = k, we define the chain maps as<br />

C 0 (•, k) → C 0 (P, k), 1 ↦→ ∑ I<br />

gx i ,<br />

Here I is the G-orbit <strong>of</strong> x i .<br />

Obviously various linear combinations <strong>of</strong> the above maps<br />

are still kG-maps.<br />

⊓⊔


246 6 Connections with group representations <strong>and</strong> cohomology<br />

Example 6.3.20. Suppose x ∈ NP 0 . Then C 0 (•, k) maps isomorphically<br />

to a 1-dimensional submodule <strong>of</strong> C 0 (P, k), that<br />

is, kO + (x) for O + (x) = ∑ y∈O(x)<br />

y where O(x) is the G-orbit<br />

<strong>of</strong> x in G ∝ P. Hence we get<br />

C ∗ (•, k)<br />

<br />

kO + (x) ɛ <br />

<br />

C ∗ (•, k)<br />

C ∗ (•, k)<br />

<br />

kO(x)<br />

<br />

<br />

C ∗ (•, k)<br />

C ∗ (•, k)<br />

<br />

C ∗ (P, k) ɛ<br />

<br />

C ∗ (•, k)<br />

Here kO(x) is isomorphic to k(G/Stab G (x)) ∼ = k ↑ G Stab G (x) .<br />

Tensoring with C ∗ (Id G /•, k) we get<br />

C ∗ (Id G /•, k)<br />

∼ =<br />

<br />

kO + (x) ⊗ C ∗ (Id G /•, k) ɛ<br />

C ∗ (Id G /•, k)<br />

<br />

kO(x) ⊗ C ∗ (Id G /•, k) ɛ <br />

<br />

<br />

C ∗ (Id G /•, k)<br />

C ∗ (Id G /•, k)<br />

C ∗ (Id G /•, k)<br />

<br />

C ∗ (P, k) ⊗ C ∗ (Id G /•, k)<br />

<br />

ɛ C ∗ (Id G /•, k)<br />

In the central column, the top one is isomorphic to C ∗ (Id G /•, k),<br />

<strong>and</strong> the middle is isomorphic to C ∗ (π O(x) /•, k). Thus by applying<br />

Hom kG (−, k) <strong>and</strong> taking cohomology we a commutative<br />

diagram<br />

H ∗ (G, k)<br />

<br />

H ∗ (G, k) H ∗ (G, k)<br />

∼ =<br />

H ∗ (G, k) H ∗ (G ∝ O(x), k)<br />

tr O(x)<br />

tr<br />

res<br />

H ∗ (G, k) H ∗ (G ∝ P, k)<br />

∼ =<br />

res O(x)<br />

res π<br />

<br />

H ∗ (G, k)<br />

H ∗ (G, k)


6.3 The functor π: group representations via transporter categories 247<br />

The transfer map is the usual one for group cohomology, see<br />

Example 6.3.17. The restrictions are induced by the functors<br />

G ∝ O(x) ↩→ G ∝ P → G. The above diagram reduces to a<br />

sequence <strong>of</strong> maps<br />

H ∗ (G, k) → H ∗ (G ∝ P, k) → H ∗ (G ∝ O(x), k) → H ∗ (G, k),<br />

or equally<br />

H ∗ (G, k) → H ∗ (G ∝ P, k) → H ∗ (Stab G (x), k) → H ∗ (G, k).<br />

We can actually take any G-set sitting inside P. The natural<br />

c<strong>and</strong>idates are all the sets NP n , n ≥ 0. For instance if α =<br />

x 0 < · · · < x n ∈ NP n <strong>and</strong> O(α) denotes the G-orbit <strong>of</strong> α,<br />

then we will get a sequence <strong>of</strong> maps<br />

H ∗ (G, k) → H ∗ (G ∝ P, k) → H ∗ (G ∝ O(α), k) → H ∗ (G, k),<br />

or equally<br />

H ∗ (G, k) → H ∗ (G ∝ P, k) → H ∗ (Stab G (α), k) → H ∗ (G, k).<br />

Note that Stab G (α) = ⋂ i Stab G(x i ).<br />

The preceding example implies that using naively constructed<br />

chain maps C ∗ (•, k) → C ∗ (P, k) it is unlikely to<br />

obtain a novel H ∗ (G ∝ P, k) → H ∗ (G, k) which does not depend<br />

on the transfer in group cohomology H ∗ (Stab G (α), k) →<br />

H ∗ (G, k). Thus we must try something more complicated.<br />

In the pro<strong>of</strong> <strong>of</strong> the next theorem, we construct the Becker-<br />

Gottlieb transfer, also seen in Dwyer-Wilkerson.<br />

It seems like many important theorems in group cohomology<br />

share the same nature <strong>and</strong> live in the same context. Using<br />

the double complex Hom kG (C † ∗(P) ⊗ B G ∗ , k) to obtain Webb’s<br />

Theorem, Dwyer’s sharp decompositions results


248 6 Connections with group representations <strong>and</strong> cohomology<br />

When P = S p , by Brown’s theorem, χ(P) = 1 in k. By<br />

Quillen’s() Theorem res π : H ∗ (G, k) → H ∗ (G ∝ P, k) is an<br />

isomorphism. What about A p etc<br />

Theorem 6.3.21. Suppose Res π : kG-mod → k(G ∝ P)-<br />

mod is the restriction along π <strong>and</strong> write κ M = Res π M for<br />

any M ∈ kG-mod. Then we have the following two maps,<br />

restriction <strong>and</strong> transfer,<br />

Ext ∗ kG(M, N) res P<br />

−→Ext ∗ k(G∝P)(κ M , κ N ) tr P<br />

−→Ext ∗ kG(M, N),<br />

which compose to χ(P; k) · 1, multiplication by the Euler<br />

characteristic <strong>of</strong> (the order complex <strong>of</strong>) P.<br />

Pro<strong>of</strong>. We shall construct these two maps. Then in the sequel<br />

we can deduce the statement on their composite. The restriction<br />

is a generalized version <strong>of</strong> the one shown at the beginning<br />

<strong>of</strong> this section.<br />

Suppose B∗<br />

G <strong>and</strong> B∗<br />

G∝P are the bar resolutions <strong>of</strong> k ∈ kGmod<br />

<strong>and</strong> k ∈ k(G ∝ P)-mod, respectively. Then B∗<br />

G ⊗ M<br />

is a projective resolution <strong>of</strong> a fixed M ∈ kG-mod. For<br />

another kG-module N, the cochain complex Hom kG (B∗<br />

G ⊗<br />

M, N) computes Ext ∗ kG(M, N). The exact functor Res π sends<br />

this cochain complex to Hom k(G∝P) (κ B<br />

G ∗ ⊗M, κ N ). Applying<br />

Lemma 6.3.12 we obtain the composite <strong>of</strong> two chain maps,<br />

unique up to chain homotopy,<br />

Hom kG (B∗ G ⊗M, N) Res π<br />

→ Hom k(G∝P) (κ B<br />

G ∗ ⊗M, κ N ) Π∗ M<br />

→Hom k(G∝P) (B∗<br />

G∝P<br />

These seemingly abstract maps can be written down explicitly,<br />

but we will leave it to the interested reader. The composite<br />

ΠM ∗ ◦ Res π induces a map on cohomology, which we call<br />

the restriction,<br />

res P : Ext ∗ kG(M, N) → Ext ∗ k(G∝P)(κ M , κ N ).<br />

ˆ⊗κ


6.3 The functor π: group representations via transporter categories 249<br />

It is helpful to have a different characterization <strong>of</strong> the<br />

restriction. In order to do so we use a series <strong>of</strong> obvious<br />

isomorphisms to rewrite the previously mentioned complex<br />

Hom k(G∝P) (B∗<br />

G∝P ˆ⊗κ M , κ N ). Firstly by adjunction, it is isomorphic<br />

to<br />

Hom kG (LK π (B∗<br />

G∝P ˆ⊗κ M ), N).<br />

Here (ΠM ∗ ◦ Res π)α is mapped to (Λ N ◦ LK π )(ΠM ∗ ◦ Res πα),<br />

where Λ N : LK π κ N → N, determined by the counit <strong>of</strong><br />

adjunction, is isomorphic to ¯ɛ ⊗ Id N . The natural map ¯ɛ :<br />

H0(P) → k is induced by the augmentation map ɛ : C 0 (P) →<br />

k (or rather P → •), by Corollary 6.3.10. Secondly from the<br />

same corollary our cochain complex is canonically isomorphic<br />

to<br />

Hom kG (LK π (B∗<br />

G∝P ) ⊗ M, N).<br />

Thirdly by Corollary 6.3.9 LK π B∗ G∝P ≃ B∗ G ⊗ C ∗ (P), the<br />

above complex is<br />

Hom kG (LK π (B G∝P<br />

∗ ) ⊗ M, N) ≃ Hom kG (B G ∗ ⊗ C ∗ (P) ⊗ M, N)<br />

∼ = HomkG (B G ∗ ⊗ M ⊗ C ∗ (P), N).<br />

The observations imply that res P : Ext ∗ kG(M, N) → Ext ∗ k(G∝P)(κ M , κ N<br />

is the same as the map induced by the following chain map<br />

Hom kG (B G ∗ ⊗ M, N) → Hom kG (B G ∗ ⊗ M ⊗ C ∗ (P), N),<br />

which is given by the augmentation ɛ : C ∗ (P) → k. Now we<br />

are ready to build the transfer map. From the cochain complex<br />

Hom kG (B G ∗ ⊗ M ⊗ C ∗ (P), N) we continue to establish a map<br />

which leads back to group cohomology<br />

Θ ∗ : Hom kG (B G ∗ ⊗ M ⊗ C ∗ (P), N) → Hom kG (B G ∗ ⊗ M, N).<br />

The map Θ ∗ is induce by the following chain map Θ : k →<br />

C ∗ (P) <strong>of</strong> Dwyer-Wilkerson [21, 9.13], as the composite <strong>of</strong>


250 6 Connections with group representations <strong>and</strong> cohomology<br />

k a↦→a·Id<br />

→ Hom k (C ∗ (P), C ∗ (P)) ∼= →C ∗ (P) ∧ ⊗C ∗ (P) Id⊗∆<br />

→ C ∗ (P) ∧ ⊗C ∗ (P)⊗C<br />

Here C ∗ (P) ∧ = Hom k (C ∗ (P), k), non-positively graded, is the<br />

k-dual <strong>of</strong> C ∗ (P). The composite <strong>of</strong><br />

Hom k(G∝P) (B∗<br />

G∝P<br />

defines a map<br />

ˆ⊗κ M , κ N ) ∼= →Hom kG (LK π (B G∝P<br />

∗<br />

tr P : Ext ∗ k(G∝P)(κ M , κ N ) → Ext ∗ kG(M, N).<br />

ˆ⊗κ M ), N) →Hom Θ∗<br />

kG<br />

which is called the transfer.<br />

In fact we have the following commutative diagram <strong>of</strong><br />

cochain complexes<br />

Hom kG (B∗ G ⊗ M, N) Hom k(G∝P) (B∗ G∝P ˆ⊗κ M , κ N )<br />

<br />

Hom kG (B∗ G ⊗ M<br />

= <br />

≃ <br />

=<br />

<br />

Hom kG (B∗ G ⊗ M, N) Hom kG (B∗ G ⊗ M ⊗ C ∗ (P), N) Hom kG (B∗ G ⊗ M<br />

Upon passing to cohomology, both rows give rise to<br />

Ext ∗ kG(M, N) res P<br />

→Ext ∗ k(G∝P)(κ M , κ N ) tr P<br />

→Ext ∗ kG(M, N).<br />

In the end we prove that the composite res P ◦tr P = χ(P)·1.<br />

Since the normalization C ∗ (P, k) → C † ∗(P, k) is a G-chain<br />

homotopy equivalence, we can replace C ∗ (P, k) by C † ∗(P, k) in<br />

our calculation. The normalized chain complex is finite so we<br />

can find an integer d such that C † d (P, k) ≠ 0 but C† n(P, k) = 0<br />

for all n > d. The following pro<strong>of</strong> is due to Dwyer-Wilkerson<br />

[21, 9.13] too, which shows by direct calculation that<br />

k → C † ∗(P) → k<br />

is a scalar multiplication by χ(P). Write the natural basis <strong>of</strong><br />

C † n(P) as {c i n} d n<br />

i=1 for d n = dim k C † n(P) (see Section 6.2.2).<br />

The step-by-step images <strong>of</strong> 1 ∈ k under Θ are


6.3 The functor π: group representations via transporter categories 251<br />

1 ↦→ Id C<br />

†<br />

∗ (P)<br />

↦→ ∑ d<br />

n=0 (−1)n { ∑ d n<br />

i=1 (ci n) ∧ ⊗ c i n}<br />

↦→ ∑ d<br />

n=0 (−1)n { ∑ d n<br />

= ∑ d<br />

n=0 (−1)n { ∑ d n<br />

↦→ ∑ d<br />

n=0 (−1)n { ∑ d n<br />

↦→ ∑ d<br />

n=0 (−1)n d n<br />

= χ(P).<br />

i=0 (ci n) ∧ ⊗ [c i n ⊗ t(c i n)]}<br />

i=1 [(ci n) ∧ ⊗ c i n] ⊗ t(c i n)}<br />

i=1 t(ci n)}<br />

Here t(c i n) ∈ C † 0 (P) denotes the last object, i.e. the target, <strong>of</strong><br />

the n-chain <strong>of</strong> morphisms c i n ∈ C † n(P).<br />

⊓⊔<br />

Remark 6.3.22. In fact, by Remark 6.3.15, the above restriction<br />

<strong>and</strong> transfer coincide with<br />

Ext ∗ kG(k, Hom k (M, N)) res P<br />

→Ext ∗ k(G∝P)(k, κ Homk (M,N)) tr P<br />

−→Ext ∗ kG(k, Hom<br />

When M = N = k, our construction is exactly the Becker-<br />

Gottlieb transfer ([4],[21]), because Ext ∗ k(G∝P)(k, κ M ) ∼ = H ∗ G (BP, M).<br />

We emphasize that if either M ∈ kG-mod is not acted trivially<br />

by kG or if G sends a connected component <strong>of</strong> P to a different<br />

one, then the constantly valued κ M ∈ k(G ∝ P)-mod<br />

is not truly constant since H 0 (G ∝ P; κ M ) ∼ = lim ←−G∝P<br />

κ M<br />

∼ =<br />

lim lim κ ∼ ←−G ←−P<br />

M = (M ⊗ H 0 (P)) G .<br />

Corollary 6.3.23. If χ(P; k) is invertible in k, then res P is<br />

an injective homomorphism.<br />

According to Dwyer, a collection <strong>of</strong> subgroups <strong>of</strong> G is a<br />

set <strong>of</strong> subgroups that is closed under conjugation. If P is a<br />

collection then it is naturally a G-poset in which the relations<br />

are inclusions <strong>and</strong> G-acts by conjugation. A collection is called<br />

ample if for a fixed prime p <strong>and</strong> a field k <strong>of</strong> characteristic p<br />

the restriction res P : Ext ∗ k(G∝P)(k, k) → Ext ∗ kG(k, k) is an


252 6 Connections with group representations <strong>and</strong> cohomology<br />

isomorphism. There is an extensive discussion on such posets<br />

in [17] or [18]. When P is ample, we get tr P = χ(P)res −1<br />

P .<br />

6.4 The functor ρ: invariants <strong>and</strong> coinvariants<br />

We have exploited the functor π : G ∝ P → G. Now we<br />

turn to study the other functor ρ : G ∝ P → C, where<br />

C is a quotient category <strong>of</strong> G ∝ P. In general situation it<br />

seems hard to make group theoretic interpretation <strong>of</strong> ↑ kC<br />

k(G∝P)<br />

<strong>and</strong> ⇑ kC<br />

k(G∝P)<br />

. However we can do so when we have certain<br />

quotient categories, which are part <strong>of</strong> some category extension<br />

sequences in the sense <strong>of</strong> H<strong>of</strong>f.<br />

An extension E <strong>of</strong> a category C via a category K is a sequence<br />

<strong>of</strong> functors<br />

K ι<br />

−→E ρ<br />

−→C,<br />

which satisfies certain properties. A sequence K ↩→ E ↠ C<br />

is called an opposite extension if K op ↩→ E op ↠ C op is an<br />

extension.<br />

The advantage <strong>of</strong> considering π : G ∝ P → C which is part<br />

<strong>of</strong> an extension (or opposite extension) is that it enables us<br />

to provide a good characterization <strong>of</strong> the left (or right) Kan<br />

extension. Indeed it is the case for many familiar category<br />

constructions in representation theory <strong>and</strong> homotopy theory.<br />

Remark 6.4.1. We emphasize that for any quotient category<br />

C <strong>of</strong> G ∝ P, P is naturally a subcategory <strong>of</strong> C. Thus for<br />

any kC-module, it makes sense to consider its limits. Indeed<br />

if m ∈ kC-mod then both ←−P<br />

lim m ∼ = lim ←−P<br />

Res ρ m <strong>and</strong> −→P<br />

lim m ∼ =<br />

lim Res −→P ρm are kG-modules.<br />

Lemma 6.4.2. Let K → E → C a sequence <strong>of</strong> three EIcategories<br />

<strong>and</strong> m ∈ kE-mod.


6.4 The functor ρ: invariants <strong>and</strong> coinvariants 253<br />

1. Suppose K → E → C is an extension. Then LK ρ m ∼ = m K ,<br />

where m K as a functor over C is given by m K (x) = m(x) K(x)<br />

(K(x)-coinvariants <strong>of</strong> the kAut E (x)-module m(x)), for any<br />

x ∈ Ob C = Ob E = Ob K.<br />

2. Suppose K → E → C is an opposite extension. Then<br />

RK ρ m ∼ = m K , where m K as a functor over C is given<br />

by m K (x) = m(x) K(x) (K(x)-invariants <strong>of</strong> the kAut E (x)-<br />

module m(x)), for any x ∈ Ob C = Ob E = Ob K.<br />

In the above lemma, there is another way to express the<br />

Kan extensions. Under the same assumptions, they are m K =<br />

H0(K; m) <strong>and</strong> m K = H 0 (K; m) respectively.<br />

In what follows, we shall apply the above statements to various<br />

local categories <strong>of</strong> G, in combination with the Frobenius<br />

reciprocity<br />

(i’) Hom kG (M, lim ←−P<br />

n) ∼ = Hom kC (LK ρ κ M , n);<br />

(ii’) Hom kG (lim −→P<br />

m, N) ∼ = Hom kC (m, LK ρ κ N ).<br />

6.4.1 Orbit categories<br />

Suppose P is a collection <strong>of</strong> subgroups <strong>of</strong> G on which G acts<br />

by conjugation. Then it forms a G-poset <strong>and</strong> we can define<br />

an orbit category O P as the quotient category <strong>of</strong> G ∝ P by<br />

asking<br />

Hom OP (P, Q) = Q\N G (P, Q).<br />

Then we have an extension sequence<br />

S ↩→ G ∝ P ↠ O P ,<br />

where S is the disjoint union <strong>of</strong> all objects in P, regarded as<br />

a subcategory <strong>of</strong> G ∝ P.<br />

When m = κ M for some M ∈ kG-mod, (LK ρ κ M )(P ) =<br />

M P for any P ∈ Ob P. We denote such a kO P -module by


254 6 Connections with group representations <strong>and</strong> cohomology<br />

M S := (κ M ) S = LK ρ κ M . Since giving a morphism (g, g P ≤<br />

Q) is the same as giving a group homomorphism P → Q, the<br />

conjugation induced by g, there is a natural way to construct<br />

a map M P → M Q , identical to the natural map H0(P ; M) =<br />

k ⊗ kP M → k ⊗ kQ M = H0(Q; M). Hence we know how kO P<br />

acts on M S .<br />

Proposition 6.4.3. Let M ∈ kG-mod <strong>and</strong> n ∈ kO P -mod.<br />

Then<br />

Hom kG (M, lim ←−P<br />

n) ∼ = Hom kOP (M S , n),<br />

where M S is as above.<br />

Corollary 6.4.4. lim ←−P<br />

n ∼ = Hom kOP (kG S , n) <strong>and</strong> Hom kG (k, lim ←−P<br />

n) ∼ =<br />

Hom kOP (k, n).<br />

As an example we let H be a subgroup <strong>of</strong> G <strong>and</strong> P the<br />

subgroups that are conjugate to H. The size <strong>of</strong> the discrete<br />

poset P is G/N G (H). Note that both G ∝ P <strong>and</strong> O P are<br />

connected groupoids, the former equivalent to N G (H) <strong>and</strong><br />

the latter N G (H)/H. Thus the above isomorphism can be<br />

interpreted as<br />

Hom kG (M, Hom kNG (H)(kG, N)) ∼ = Hom kNG (H)(M, N)<br />

∼ = Homk(NG (H)/H)(k(N G (H)/H) ⊗ kNG<br />

∼=<br />

Homk(NG (H)/H)(M H , N),<br />

where M ∈ kG-mod <strong>and</strong> N ∈ k(N G (H)/H)-mod.<br />

By looking at the special case we have just mentioned, the<br />

following statement implies that Proposition 5.2 may be useful<br />

in a greater generality.<br />

Lemma 6.4.5. Suppose M is an indecomposable kG-module.<br />

Let P be a p-subgroup <strong>and</strong> N a projective simple k(N G (P )/P )-<br />

module. Then Hom kNG (P )/P(M P , N) ≠ 0 if <strong>and</strong> only if there


6.4 The functor ρ: invariants <strong>and</strong> coinvariants 255<br />

exists a surjective map f : M → N ↑ G N G (P ). In this case<br />

N ↑ G N G (P ) is indecomposable <strong>and</strong> f(M) ∼ = k ↑ G P .<br />

Pro<strong>of</strong>. If Hom kNG (P )/P(M P , N) ≠ 0, then N has to be a direct<br />

summ<strong>and</strong> <strong>of</strong> M P because N is projective simple. By adjunction<br />

Hom kG (M, N ↑ G N G (P ) ) ∼ = Hom kG (M, Hom kNG (P )(kG, N)) ∼ = Hom k(NG (P<br />

we know there is a non-trivial map f : M → N ↑ G N G (P ). If we<br />

restrict this map back to N G (P ), the right side is a semisimple<br />

module <strong>and</strong> thus the image <strong>of</strong> M ↓ NG (P ) contains at least a<br />

copy <strong>of</strong> g ⊗ N for some g ∈ G. It forces the map f : M ↠<br />

N ↑ G N G (P ) to be surjective which implies that N ↑G N G (P ) = f(M)<br />

is indecomposable. Furthermore since as a kN G (P )-module<br />

N has P as a vertex, we get f(M) ∣ (N ↓ P ↑ N G(P ) ) ↑ G . But<br />

N ↓ N G(P )<br />

P<br />

is a direct sum <strong>of</strong> trivial modules. Simultaneously<br />

we obtain f(M) ∼ = k ↑ G P .<br />

The converse is straightforward by the adjunction. ⊓⊔<br />

6.4.2 Brauer categories, fusion <strong>and</strong> linking systems<br />

Suppose b is a p-block <strong>of</strong> the group algebra kG <strong>and</strong> P b is the<br />

set <strong>of</strong> b-Brauer pairs. Then for any G-subposet P ⊂ P b we can<br />

introduce the Brauer category B P as the quotient category <strong>of</strong><br />

G ∝ P such that<br />

Hom BP (P, Q) = Hom G∝P (P, Q)/C G (P ).<br />

This gives us an opposite extension, which means that the<br />

following sequence<br />

C G ↩→ (G ∝ P) op → B op<br />

P


256 6 Connections with group representations <strong>and</strong> cohomology<br />

is an extension sequence, given that C G is the disjoint union<br />

<strong>of</strong> all C G (P ), P ∈ Ob P. Dual to the extension situation we<br />

examined before, now we are able to describe the right Kan<br />

extension <strong>of</strong> modules.<br />

If m = κ M for some M ∈ kG-mod, we denote by M C G the<br />

kF P -module RK ρ κ M . Since a morphism (g, g P ≤ Q) provides<br />

a group homomorphism P → Q <strong>and</strong> thus induces an injection<br />

c g<br />

−1 : C G (Q) → C G (P ), we obtain an injection M C G(P ) →<br />

M CG(Q) . This leads to the kB P -action on M C G .<br />

Proposition 6.4.6. Let m ∈ kB P -mod <strong>and</strong> N ∈ kG-mod.<br />

Then<br />

Hom kG (lim −→P<br />

m, N) ∼ = Hom kBP (m, N C G<br />

),<br />

where N C G<br />

is as above.<br />

As an example we assume b is the principal block b 0 <strong>and</strong><br />

P is the conjugacy class <strong>of</strong> a fixed p-subgroup H. Then the<br />

discrete poset P has |G/N G (H)| objects. Both G ∝ P <strong>and</strong><br />

B P are connected groupoids, the former equivalent to N G (H)<br />

<strong>and</strong> the latter N G (H)/C G (H). Thus the above isomorphism<br />

can be interpreted as<br />

Hom kG (kG ⊗ kNG (H) M, N) ∼ = Hom kNG (H)(M, N)<br />

∼ = Homk(NG (H)/C G (H))(M, Hom kNG (H)(k(N G<br />

∼ = Homk(NG (H)/C G (H))(M, N C G(H) ),<br />

where N ∈ kG-mod <strong>and</strong> M ∈ k(N G (H)/C G (H))-mod.<br />

Corollary 6.4.7. We have Hom kG (lim −→P<br />

m, k) ∼ = Hom kBP (m, k)<br />

for any m ∈ kB P -mod.<br />

Suppose b is a nilpotent block. Then, for every b-Brauer pair<br />

(H, e), N G (H, e)/C G (H) is a p-group.<br />

Let B b = B Pb . If we fix a maximal object (S, e S ) <strong>and</strong> take<br />

all objects (Q, e Q ) with Q ⊂ S, then the full subcategory <strong>of</strong>


6.4 The functor ρ: invariants <strong>and</strong> coinvariants 257<br />

B b , consisting <strong>of</strong> all these objects, is a fusion system, usually<br />

written as F b or F S . Note that the inclusion F b ⊂ B b is an<br />

equivalence. There is a general theory <strong>of</strong> p-local finite groups<br />

introduced by Broto, Levi <strong>and</strong> Oliver. A p-local finite group<br />

consists <strong>of</strong> three categories (S, F, L), where S is a p-group, F<br />

is an (abstract) fusion system–a finite category whose objects<br />

are subgroups <strong>of</strong> S. Let us take the full subcategory F c ⊂ F<br />

consisting <strong>of</strong> F-centric subgroups <strong>of</strong> S. (If F = F b , the F-<br />

centric subgroups correspond to the so-called self-centralizing<br />

b-Brauer pairs in modular representation theory.) A centric<br />

linking system L c , if it exists, situates in the middle <strong>of</strong> a<br />

sequence<br />

Z ↩→ L c ↠ F c<br />

which is an opposite extension. Here Z is the disjoint union<br />

<strong>of</strong> the centers <strong>of</strong> all F-centric subgroups.<br />

Proposition 6.4.8. Let n ∈ kL c -mod <strong>and</strong> m ∈ kF c -mod.<br />

Then<br />

Hom kL c(Res ρ m, n) ∼ = Hom kF c(m, n Z ),<br />

where n Z is defined by n Z (P ) = n(P ) Z(P ) .<br />

Let Bb c be the full subcategory <strong>of</strong> B b for a block b, consisting<br />

<strong>of</strong> self-centralizing b-Brauer pairs. Since Fb c naturally identifies<br />

with a full subcategory <strong>of</strong> Bb c which induces an equivalence,<br />

we similarly can consider an opposite extension<br />

Z ↩→ ˜Lc b ↠ B c b.<br />

If ˜Lc b<br />

exists, then so is L c b<br />

, <strong>and</strong> vice versa. Moreover between<br />

the corresponding extensions there exists a natural embedding<br />

L c b → ˜Lc b<br />

inducing an category equivalence. By taking the<br />

larger category ˜Lc b<br />

(but essentially the same as L c b<br />

), we can<br />

write down<br />

C G /Z ↩→ G ∝ Pb c ↠ ˜Lc b ,


258 6 Connections with group representations <strong>and</strong> cohomology<br />

another opposite extension. Here C G /Z is the disjoint union<br />

<strong>of</strong> C G (P )/Z(P ) in which P runs over all F-centric subgroups.<br />

For the sake <strong>of</strong> convenience, we introduce a notation C G ′ =<br />

C G /Z so that C G ′ (P ) = C G(P )/Z(P ) ∼ = P C G (P )/P for each<br />

P .<br />

When m = κ M for some M ∈ kG-mod, we write the k ˜Lc b<br />

-<br />

module RK ρ κ M as M C′ G . Given a morphism (g, g P ≤ Q)<br />

it induces an injection C G ′ (Q) → C′ G (P ) thus a morphism<br />

M C′ G (P ) → M C′ G (Q) . Hence we get the k ˜Lc b<br />

-action on M C′ G<br />

Proposition 6.4.9. Let m ∈ k(G ∝ Pb c )-mod <strong>and</strong> N ∈ kGmod.<br />

Then<br />

Hom kG (lim −→P<br />

m, N) ∼ = Hom k ˜Lc b<br />

(m, N C′ G ),<br />

where N C′ G is defined as above.<br />

In particular if m = Res ρ m ′ for some m ′ ∈ kBb c -mod then<br />

Hom kG (lim −→P<br />

m ′ , N) ∼ = Hom kB<br />

c<br />

b<br />

(m ′ , N C G<br />

),<br />

a special situation <strong>of</strong> Proposition 6.5.3.<br />

Pro<strong>of</strong>. The first part is a direct consequence <strong>of</strong> Proposition<br />

5.4. As for the special case We need to notice that<br />

lim Res −→P ρm ′ ∼ = lim −→P<br />

m ′ as kG-modules. Then<br />

Hom kG (lim −→P<br />

Res ρ m ′ , N) ∼ = Hom k ˜Lc b<br />

(Res ρ m ′ , N C′ G )<br />

∼ = HomkB c<br />

b<br />

(m ′ , (N C′ G ) Z )<br />

∼ = HomkB c<br />

b<br />

(m ′ , N C G ).<br />

⊓⊔<br />

6.4.3 Puig categories<br />

If we take P A to be the poset <strong>of</strong> pointed subgroups on an interior<br />

G-algebra A, then analogues to the Brauer category for


6.4 The functor ρ: invariants <strong>and</strong> coinvariants 259<br />

any G-subposet P ⊂ P A we can introduce the Puig category<br />

L P as a quotient category <strong>of</strong> G ∝ P such that<br />

Hom LP (P γ , Q δ ) = Hom G∝P (P γ , Q δ )/C G (P ).<br />

Then some results in last section can be obtained accordingly.<br />

If m = κ M for some M ∈ kG-mod, we denote by M C G the<br />

kL P -module RK ρ κ M . Since a morphism (g, g P ≤ Q) provides<br />

a group homomorphism P → Q <strong>and</strong> thus induces an injection<br />

c g<br />

−1 : C G (Q) → C G (P ), we obtain an injection M C G(P ) →<br />

M CG(Q) . This leads to the kL P -action on M C G .<br />

Proposition 6.4.10. Let m ∈ kL P -mod <strong>and</strong> N ∈ kG-mod.<br />

Then<br />

Hom kG (lim −→P<br />

m, N) ∼ = Hom kLP (m, N C G<br />

),<br />

where N C G<br />

is as above.<br />

6.4.4 Orbit categories <strong>of</strong> fusion systems<br />

This method also works for the orbit category <strong>of</strong> a fusion<br />

system. Suppose F is an abstract fusion system. The one may<br />

define the orbit category O F in a similar fashion as above by<br />

Hom OF (P, Q) = Q\Hom F (P, Q).<br />

Again we obtain an extension sequence<br />

S ↩→ F ↠ O F ,<br />

where S is the disjoint union <strong>of</strong> objects in F.<br />

Proposition 6.4.11. If m ∈ kF-mod <strong>and</strong> n ∈ kO F -mod,<br />

then<br />

Hom kF (m, Res ρ n) ∼ = Hom kOF (m S , n).<br />

Moreover for F = Fb c, Hom kG(lim −→P<br />

m, k) ∼ = Hom kOF c(m S , k).<br />

b<br />

Pro<strong>of</strong>. When F = Fb c , we have isomorphisms


260 6 Connections with group representations <strong>and</strong> cohomology<br />

Hom kG (lim −→P<br />

m, k) ∼ = Hom kF (m, k)<br />

∼ = HomkOF c<br />

b<br />

(m S , k).<br />

⊓⊔<br />

6.5 Hochschild cohomology<br />

In this section we continue to demonstrate the similarities between<br />

transporter categories <strong>and</strong> groups. The first main assertion<br />

is the the Hochschild cohomology ring <strong>of</strong> a finite transporter<br />

category is finitely generated. From here we will establish<br />

the finite generation <strong>of</strong> cohomology. The second is that we<br />

can construct a transfer map between Hochschild cohomology.<br />

Both results are established by passing between Hochschild<br />

cohomology <strong>and</strong> ordinary cohomology <strong>of</strong> F (G ∝ P) <strong>and</strong><br />

G ∝ P. Thus computing various over <strong>and</strong> undercategories<br />

for functors from F (G ∝ P) to appropriate categories are the<br />

major auxiliary statements.<br />

6.5.1 <strong>Finite</strong> generation<br />

In Section 4.2.2 we have seen that the ordinary cohomology<br />

ring <strong>of</strong> a finite category can be infinitely generated even after<br />

quotient out nilpotent elements. Based on the Theorem<br />

5.2.2, the Hochschild cohomology ring <strong>of</strong> such a finite category<br />

algebra is not finitely generated either. So the question<br />

reduces to finding out whether or not Ext ∗ kCe(kC, kC) modulo<br />

nilpotents is finitely generated over Ext ∗ kC(k, k) if the latter is<br />

Noetherian. On the first attempt to solve this question, one<br />

may want to check if the Evens-Venkov Theorem on the finite<br />

generation <strong>of</strong> group cohomology could be generalized to category<br />

cohomology. This is not true since Example 4.2.2 (3)<br />

implies that we can not expect a finite generation property


6.5 Hochschild cohomology 261<br />

<strong>of</strong> Ext ∗ kC(M, N) over a finitely generated Ext ∗ kC(k, k). Thus<br />

we have to look at particular families <strong>of</strong> finite categories for<br />

the finite generation property. In what follows, we show finite<br />

transporter categories constructed over a finite group are very<br />

close to what we expect.<br />

To examine the finite generation <strong>of</strong> Hochschild cohomology<br />

ring, we use the isomorphism<br />

Ext ∗ k(G∝P) e(k(G ∝ P), k(G ∝ P)) ∼ = Ext ∗ kF (G∝P)(k, Res ∇ (k(G ∝ P)))<br />

because it allows us to use the Grothendieck spectral sequence.<br />

Let us introduce a functor ˜π = π ◦ t<br />

t<br />

F (G ∝ P)<br />

<br />

G ∝ P<br />

˜π<br />

π<br />

G .<br />

When we investigated the finite generation <strong>of</strong> H ∗ (G ∝ P, k)<br />

we used the Grothendieck spectral sequence. Based on the fact<br />

that •\π ≃ P, we obtained the finite generation. Naturally<br />

we want to look at the Grothendieck spectral sequence for<br />

˜π. In order to underst<strong>and</strong> the spectral sequence, we have to<br />

know the undercategory •\˜π. Since t induces a homotopy<br />

equivalence, the undercategory •\˜π should be similar to •\π.<br />

If we take P = • in the above diagram, then π is a canonical<br />

isomorphism <strong>and</strong> •\π ≃ •. By contrast ˜π can be identified<br />

with t G : F (G) → G <strong>and</strong> we also have •\˜π = •\t G ≃ •.<br />

Lemma 6.5.1. Let t G : F (G) → G be the target functor.<br />

Then we have two isomorphic categories<br />

1. •\˜π ∼ = F (P) × •\t G .<br />

2. ˜π/• ∼ = F (P) × t G /•.<br />

3. •\t G<br />

∼ = tG /• are equivalent to •.<br />

Pro<strong>of</strong>. We only prove Parts 1 <strong>and</strong> 3. First we prove (1).<br />

The objects in •\˜π are (•, [(g, gx ≤ y)]). A morphism from


262 6 Connections with group representations <strong>and</strong> cohomology<br />

(h, [(g, gx ≤ y)]) to (h ′ , [(g ′ , g ′ x ′ ≤ y ′ )]) is a morphism<br />

((l 1 , l 1 y ≤ y ′ ), (l 2 , l 2 x ′ ≤ x)) : [(g, gx ≤ y)] → [(g ′ , g ′ x ′ ≤ y ′ )]<br />

such that l 1 h = h ′ <strong>and</strong> g ′ = l 1 gl 2 . Thus from one object in<br />

•\˜π to another there is at most one morphism. It implies<br />

this finite undercategory is equivalent to a finite poset. Furthermore<br />

we notice that (h, [(g, gx ≤ y)]) is isomorphic to<br />

(e, [(h −1 g, (h −1 g)x ≤ h −1 y)]). But [(h −1 g, (h −1 g)x ≤ h −1 y)]<br />

is isomorphic to [(e, e(h −1 gx) ≤ h −1 y)] in F (G ∝ P). Now<br />

we define a functor •\˜π → F (P) × •\t G by<br />

(h, [(g, gx ≤ y)]) ↦→ [h −1 gx ≤ h −1 y] × (h, [g]).<br />

Its inverse is given by<br />

[x ≤ y] × (h, [g]) ↦→ (h, [(g, g(g −1 hx) ≤ hy)]).<br />

The isomorphism in (3) is easy to write down as (h, [g]) ↦→<br />

([g], h −1 ). Furthermore, any object ([g], h) ∈ Ob(t G /•) is isomorphic<br />

to ([e], e) which has only one endomorphism. ⊓⊔<br />

Now we can state the consequence <strong>of</strong> Lemma 6.5.1 (1). It is<br />

similar to the main result in Section 6.3.2.<br />

Proposition 6.5.2. Let G be a finite group <strong>and</strong> P a finite<br />

G-poset. Then for any M ∈ kF (G ∝ P)-mod, Ext ∗ kF (G∝P)(k, M)<br />

becomes a finitely generated Ext ∗ kG(k, k)- <strong>and</strong> Ext ∗ kF (G∝P)(k, k)-<br />

module. Especially the Hochschild cohomology ring Ext ∗ k(G∝P) e(k(G ∝<br />

P), k(G ∝ P)) is finitely generated.<br />

Pro<strong>of</strong>. We apply the Grothendieck spectral sequence to ˜π.<br />

Since •\˜π is equivalent to a finite poset F (P), H j (•\˜π, M)<br />

vanishes for all j larger than a chosen positive integer. Consequently<br />

the E 2 page <strong>of</strong> the spectral sequence only has finitely<br />

many non-zero rows in the first quadrant, <strong>and</strong> thus we have<br />

the finite generation <strong>of</strong> Ext ∗ kF (G∝P)(k, M).<br />

⊓⊔


6.5 Hochschild cohomology 263<br />

The preceding result enables us to prove a finite generation<br />

theorem. It will be the foundation for developing a support<br />

variety theory over the ring Ext ∗ k(G∝P)(k, k).<br />

Theorem 6.5.3. Suppose M, N are two k(G ∝ P)-modules.<br />

Then the module Ext ∗ k(G∝P)(M, N) is finitely generated over<br />

Ext ∗ k(G∝P)(k, k).<br />

Pro<strong>of</strong>. By Theorem 5.2.11,<br />

Ext ∗ k(G∝P)(M, N) ∼ = Ext ∗ kF (G∝P)(k, Res ∇ Hom k (M, N)).<br />

Hence by Proposition 6.5.2, Ext ∗ kF (G∝P)(k, Res ∇ Hom k (M, N))<br />

is finitely generated over Ext ∗ kF (G∝P)(k, k) ∼ = Ext ∗ k(G∝P)(k, k).<br />

We are done.<br />

⊓⊔<br />

If P = •, we get the usual assertion that Ext ∗ kGe(kG, kG)<br />

is finitely generated over Ext ∗ kG(k, k). If P = S b as in Example<br />

6.1., we have Ext ∗ kG(k, k) acting on Ext ∗ k(G∝S b ) e(k(G ∝<br />

S b ), k(G ∝ S b )) via Ext ∗ k(G∝S b )(k, k). Especially when b = b 0 ,<br />

S b<br />

∼ = Sp <strong>and</strong> Ext ∗ k(G∝S p )(k, k) ∼ = Ext ∗ kG(k, k).<br />

Corollary 6.5.4. If k is a field with positive characteristic<br />

p ∣ |G|, <strong>and</strong> b is a p-block, then, for any full subcategory<br />

Tr ⊂ Tr b (G) whose objects are closed under G-conjugation,<br />

Ext ∗ kTr<br />

e(kTr, kTr) is a finitely generated algebra.<br />

Given the principal block b 0 <strong>of</strong> a group algebra kG, we have<br />

a fusion system F b0 = F p over a fixed Sylow p-subgroup S.<br />

As we mentioned earlier, there exists a centric linking system<br />

L b0 = L c p which is determined by the full subcategory<br />

Tr c p(G) ≤S <strong>of</strong> the transporter category Tr p (G), consisting <strong>of</strong> all<br />

p-centric subgroups contained in S [10]. In fact L c p is a quotient<br />

category <strong>of</strong> Tr c p(G) ≤S by some p ′ -groups. In other words,<br />

if one looks at the canonical functor π : Tr c p(G) ≤S → L c p, each


264 6 Connections with group representations <strong>and</strong> cohomology<br />

undercategory has the property such that it has a minimal object<br />

whose automorphism group is p ′ <strong>and</strong>, if one regards this<br />

p ′ -automorphism group as a subcategory, the left Kan extension<br />

along the inclusion is exact. Furthermore the left Kan<br />

extension <strong>of</strong> the trivial group module is the trivial module <strong>of</strong><br />

the undercategory. This implies, by an Eckmann-Shapiro type<br />

result, the cohomology <strong>of</strong> each undercategory can be reduced<br />

to the cohomology <strong>of</strong> the automorphism group <strong>of</strong> the abovespecified<br />

minimal object in it. Consequently, the mod-p cohomology<br />

<strong>of</strong> each undercategory <strong>of</strong> π with arbitary coefficients<br />

vanishes in positive degrees. To summarize, since Tr c p(G) ≤S is<br />

equivalent to Tr c p(G) <strong>and</strong> the Grothendieck spectral sequence<br />

for π collapses, we have an isomorphism<br />

Ext ∗ kTr c p(G)(k, V ) ∼ = Ext ∗ kTr c p(G) ≤S<br />

(k, V ) ∼ = Ext ∗ kL c p (k, RK πV ),<br />

where RK π V = H 0 (\π; V ) ∼ = lim ←−\π<br />

V is the right Kan extension<br />

along π <strong>of</strong> V . It is similar to [10, Lemma 1.3 (iii)]<br />

in which right modules are considered <strong>and</strong> thus the left Kan<br />

extension is applied. Especially, the functors G ← Tr c p(G) ←↪<br />

Tr c p(G) ≤S → L c p induce isomorphisms <strong>of</strong> mod-p ordinary cohomology<br />

rings.<br />

Proposition 6.5.5. Let L c p be the centric linking system<br />

associated to the principal block <strong>of</strong> a finite group algebra<br />

kG. Then Ext ∗ kL (k, RK πV ) is finitely generated as an<br />

c p<br />

Ext ∗ kL (k, k)-module.<br />

c p<br />

However this is still far from underst<strong>and</strong>ing the finite generation<br />

<strong>of</strong> the Hochschild cohomology ring Ext ∗ k(L c p) e(kLc p, kL c p) ∼ =<br />

Ext ∗ kF (Lp)(k, Res c τ (kL c p)).


6.5 Hochschild cohomology 265<br />

6.5.2 Transfer for Hochschild cohomology<br />

We end this chapter with a transfer map between Hochschild<br />

cohomology. Consider the following commutative diagram <strong>of</strong><br />

functors<br />

(G ∝ P) e πe <br />

G e<br />

F (P)<br />

t <br />

∇<br />

<br />

F (G ∝ P)<br />

t <br />

F (π) <br />

∇<br />

F (G)<br />

P<br />

<br />

G ∝ P π<br />

<br />

G<br />

Recall that in Section 6.3.3, based on the bottom row, we<br />

were able to establish a transfer map between ordinary cohomology.<br />

Since the three target functors all induce homotopy<br />

equivalences, the middle row may as well give a transfer map<br />

Ext ∗ kF (G∝P)(−, −) → Ext ∗ kF (G)(−, −), for suitable modules.<br />

Since furthermore the upper right square demonstrates connections<br />

between the ordinary cohomology over kF (G ∝ P)<br />

<strong>and</strong> kF (G) with the Hochschild cohomology <strong>of</strong> kF (G ∝ P)<br />

<strong>and</strong> kF (G), respectively, we automatically wonder if there<br />

would be a transfer between the Hochschild cohomology <strong>of</strong><br />

kF (G ∝ P) <strong>and</strong> kF (G) The answer is yes, <strong>and</strong> we shall<br />

provide the construction here, which in spirit is similar to the<br />

transfer we built between the ordinary cohomology. A predecessor<br />

<strong>of</strong> this construction is due to Lincklemann [45]. For G<br />

a group <strong>and</strong> H a subgroup, he developed a method over symmetric<br />

algebras <strong>and</strong> defined a transfer map from HH ∗ (kH) to<br />

HH ∗ (kG). Recall that H can be recovered as the transporter<br />

category G ∝ (G/H), our construction generalizes his.<br />

Lemma 6.5.6. Let kG ∈ kG e -mod <strong>and</strong> k(G ∝ P) ∈ k(G ∝<br />

P) e -mod. Then the kF (G ∝ P)-module Res F (π) Res ∇ kG contains<br />

Res ∇ k(G ∝ P) as a submodule.<br />

t


266 6 Connections with group representations <strong>and</strong> cohomology<br />

Pro<strong>of</strong>. Let (x, y) ∈ Ob(G ∝ P) e . Then<br />

k(G ∝ P)(x, y) = kHom G∝P (y, x)<br />

= k{(h, hy ≤ x) ∣ ∣ h ∈ G}<br />

⊂ kG<br />

= (Res π ekG)(x, y).<br />

From here we can verify that k(G ∝ P) is a k(G ∝ P) e -<br />

submodule <strong>of</strong> Res π ekG. Hence the result follows. ⊓⊔<br />

If P = G/H, then G ∝ (G/H) ≃ H <strong>and</strong> F (G ∝ (G/H)) ≃<br />

H. The above inclusion says that the kH-module kG ↓ G H contains<br />

kH as a submodule. Here kG <strong>and</strong> kH are acted by G<br />

<strong>and</strong> H, respectively, by conjugation.<br />

Lemma 6.5.7. Consider the functor F (π) : F (G ∝ P) →<br />

F (G). Then F (π)/− ∼ = F (P) − × Id F (G) /−, where F (P) [g]<br />

denotes a poset, indexed by [g] ∈ Ob F (G), which is canonically<br />

isomorphic to F (P). Consequently<br />

C ∗ (F (π)/−) ∼ = C ∗ (F (P) − × Id F (G) /−)<br />

as complexes <strong>of</strong> kF (G)-modules. Since F (G) ≃ G,<br />

C ∗ (F (π)/[e]) ∼ = C ∗ (F (P) [e] × Id F (G) /[e]) ≃ C ∗ (F (P)) ⊗ B G ∗<br />

is a projective resolution <strong>of</strong> the complex <strong>of</strong> kG-modules<br />

C ∗ (F (P)).<br />

Pro<strong>of</strong>. Suppose [g] is an object in F (G). The the overcategory<br />

F (G)/[g] has objects {([(h, hx ≤ y)], (h 1 , h 2 ))}, in which<br />

[(h, hx ≤ y)] ∈ Ob F (G ∝ P). It implies g = h 1 hh 2 . As a<br />

consequence, hx ≤ y is equivalent to h −1<br />

2 x ≤ g−1 h 1 y. Morphisms<br />

((v 1 , v 1 y ≤ y ′ ), (v 2 , v 2 x ′ ≤ x)) are given by


6.5 Hochschild cohomology 267<br />

(v 1 ,v 2 )=F (π)((v 1 ,v 1 y≤y ′ ),(v 2 ,v 2 x ′ ≤x))<br />

[h] = F (π)([(h, hx ≤ y)])<br />

(h 1 ,h 2 )<br />

(h ′ 1 ,h′ 2 ) <br />

[h ′ ] = F (π)([(h ′ , h ′ x ′ ≤ y ′ )])<br />

It implies various identities: g = h 1 hh 2 = h ′ 1h ′ h ′ 2, h 1 = h ′ 1v 1<br />

<strong>and</strong> h 2 = v 2 h ′ 2. We define a functor F (π)/[g] → F (P) ×<br />

Id F (G) /[g] such that on objects<br />

([(h, hx ≤ y)], (h 1 , h 2 )) ↦→ ([h −1<br />

2 x ≤ g−1 h 1 y], ([h], (h 1 , h 2 ))),<br />

<strong>and</strong> on morphisms ((v 1 , v 1 y ≤ y ′ ), (v 2 , v 2 x ′ ≤ x)) ↦→ ((e, e), (v 1 , v 2 )),<br />

because v 1 y ≤ y ′ implies g −1 h 1 y ≤ g −1 h ′ 1y ′ while v 2 x ′ ≤ x implies<br />

h ′ −1 2 x ′ ≤ h2 −1 x. The inverse <strong>of</strong> this functor is defined by<br />

([x ≤ y], ([h], (h 1 , h 2 ))) ↦→ ([(h, h(h 2 x) ≤ h −1<br />

1 gy)], (h 1, h 2 ))<br />

on objects, <strong>and</strong> on morphisms is defined by ((e, e), (v 1 , v 2 )) ↦→<br />

((v 1 , v 1 h −1<br />

1 gy ≤ h′ 1−1 gy ′ ), (v 2 , v 2 h ′ 2x ′ ≤ h 2 x)). Thus we obtain<br />

an isomorphism<br />

F (π)/[g] ∼ = F (P) [g] × Id F (G)/[g],<br />

where F (P) [g] denotes a copy <strong>of</strong> F (P) indexed by the object<br />

[g]. If (l 1 , l 2 ) : [g] → [g ′ ] is a morphism in F (G). Then it<br />

induces a functor F (π)/[g] → F (π)/[g ′ ] given by<br />

<strong>and</strong><br />

([(h, hx ≤ y)], (h 1 , h 2 )) ↦→ ([(h, hx ≤ y)], (l 1 h 1 , h 2 l 2 ))<br />

((v 1 , v 1 y ≤ y ′ ), (v 2 , v 2 x ′ ≤ x)) ↦→ ((v 1 , v 1 y ≤ y ′ ), (v 2 , v 2 x ′ ≤ x))<br />

Using the isomorphisms F (π)/[g] ∼ = F (P) [g] × Id F (G) /[g] <strong>and</strong><br />

F (π)/[g ′ ] ∼ = F (P) [g ′ ] × Id F (G) /[g ′ ], one can see it induces an<br />

isomorphism F (P) [g] → F (P) [g ′ ].<br />

[g]


268 6 Connections with group representations <strong>and</strong> cohomology<br />

Finally since G is isomorphic to the automorphism group <strong>of</strong><br />

[e] in the groupoid F (G), we have an equivalence F (G) ≃ G.<br />

Thus C ∗ (F (P) [e] × Id F (G) /[e]) ≃ C ∗ (F (P)) ⊗ C ∗ (Id F (G) /[e])<br />

is a complex <strong>of</strong> projective kG-modules. Furthermore because<br />

([e]), it has to be a projective resolution<br />

<strong>of</strong> the trivial kG-module k. Hence we get the chain<br />

homotopy equivalence as stated.<br />

⊓⊔<br />

C ∗ (Id F (G) /[e]) = B F (G)<br />

∗<br />

This lemma allows us to describe the transfer map in Theorem<br />

5.2.2 in terms <strong>of</strong> factorization categories:<br />

tr P : Ext ∗ kF (G∝P)(k, k) → Ext ∗ kF (G)(k, k).<br />

Indeed we have a commutative diagram<br />

F (G ∝ P) F (π) <br />

F (G)<br />

t <br />

G ∝ P π<br />

<br />

G<br />

Thus we get a commutative diagram <strong>of</strong> cochain complexes<br />

Hom kF (G∝P) (B F (G∝P)<br />

∗ , k)<br />

∼ = <br />

∼ = <br />

Hom k(G∝P) (LK t B F (G∝P)<br />

∗ , k) ∼=<br />

<br />

≃<br />

<br />

Hom k(G∝P) (B G∝P<br />

∗ , k) ≃<br />

t<br />

Hom kF (G) (LK F (π) B F (G∝P)<br />

∗ , k)<br />

∼ =<br />

Hom kG (LK˜π B∗ F (G∝P) , k)<br />

≃<br />

<br />

Hom kG (C ∗ (P) ⊗ B G ∗ , k)<br />

Hom k(G∝P) (B∗ G∝P , k)<br />

<br />

Hom kG (B∗ G , k)<br />

≃<br />

<br />

Hom kG (LK t B F (G)<br />

∗ , k)<br />

∼ =<br />

<br />

Hom kF (G) (B F (G) , k)


6.5 Hochschild cohomology 269<br />

The isomorphisms are given by adjunctions <strong>and</strong> the chain homotopy<br />

equivalences are induced by changing projective resolutions.<br />

From the upper left corner to the lower right corner<br />

is the transfer that we want to describe. It factors through<br />

the lowest square, including the chain map which gives rise to<br />

the transfer constructed in Theorem 6.3.21. Note that F (G)<br />

is a groupoid <strong>and</strong> G is a skeleton <strong>of</strong> F (G). Comparing with<br />

[45, Section 4], F (G) plays the role <strong>of</strong> ∆G if one identifies G e<br />

with G × G.<br />

Theorem 6.5.8. There exists a map<br />

htr P : Ext ∗ k(G∝P) e(k(G ∝ P), k(G ∝ P)) → Ext∗ kGe(kG, kG).<br />

Pro<strong>of</strong>. Using Lemma 6.5.6 we have cochain maps<br />

Hom kF (G∝P) (B∗<br />

G∝P , Res F (π) Res ∇ kG) ∼ = Hom kF (G) (LK F (π) B∗<br />

G∝P , Res ∇<br />

∼ = HomkF (G) (C ∗ (F (π)/−), Res ∇<br />

∼ = HomkG (C ∗ (F (π)/[e]), kG)<br />

≃ Hom kG (C ∗ (F (P)) ⊗ B∗ G , kG)<br />

→ Hom kG (B∗ G , kG)<br />

Here the module kG in Hom kG (−, kG) is acted by kG via<br />

conjugations. The last map is induced by k → C ∗ (F (P)),<br />

a chain map constructed in the same way as the one in the<br />

pro<strong>of</strong> <strong>of</strong> Theorem 6.3.21 k → C ∗ (P). Indeed since t : F (P) →<br />

P is a G-functor, it induces an chain homotopy equivalence<br />

C ∗ (F (P)) ≃ C ∗ (P).<br />

By Lemma 6.5.5 we also have a chain map<br />

Hom kF (G∝P) (B G∝P<br />

∗<br />

, Res ∇ kF (G ∝ P)) → Hom kF (G∝P) (B∗<br />

G∝P , Res F (π)<br />

induced by the inclusion Res ∇ kF (G ∝ P) → Res F (π) Res ∇ kG.<br />

Hence altogether we obtain a chain map<br />

Hom kF (G∝P) (B∗<br />

G∝P , Res ∇ kF (G ∝ P)) → Hom kG (B∗ G , kG).


270 6 Connections with group representations <strong>and</strong> cohomology<br />

Passing to cohomology we get a map between Hochschild cohomology<br />

by Theorem 5.2.2<br />

Ext ∗ kF (G∝P)(k, Res ∇ kF (G ∝ P)) htr P<br />

∼ =<br />

<br />

Ext ∗ kG(k, kG)<br />

∼ =<br />

Ext ∗ k(G∝P) e(kF (G ∝ P), kF (G ∝ P)) htr P<br />

<br />

Ext ∗ kG e(kG, kG) ⊓⊔<br />

The above map deserves to be called a transfer since<br />

k ∣ ∣ Res ∇ kF (G ∝ P), k ∣ ∣ kG <strong>and</strong> k ∣ ∣ Res ∇ kF (G). In the<br />

construction <strong>of</strong> htr P if we replace the second module in all<br />

Hom − (−, −) <strong>and</strong> Ext ∗ −(−, −) by k then it is exactly the transfer<br />

created in Theorem 6.3.21 <strong>and</strong> reinterpreted before Theorem<br />

6.5.7. In other words, we have a commutative diagram<br />

Ext ∗ k(G∝P)(k, k)<br />

∼ =<br />

<br />

Ext ∗ kF (G∝P)(k, k)<br />

injection<br />

Ext ∗ kF (G∝P)(k, Res ∇ kF (G ∝ P))<br />

∼ =<br />

<br />

tr P<br />

tr P<br />

Ext ∗ kG(k, k)<br />

∼ =<br />

Ext ∗ kF (G)(k, k)<br />

injection<br />

Ext ∗ kF (G)(k, Res ∇ kF (G))<br />

Ext ∗ k(G∝P) e(k(G ∝ P), k(G ∝ P)) <br />

htr P<br />

Ext ∗ kGe(kG, kG)<br />

∼ =


References 271<br />

References<br />

1. J. Alperin, Local Representaion Theory, Cambridge Studies in Adv. Math. 11, Cambridge University Press (1986).<br />

2. M. Aschbacher, R. Kessar, B. Oliver, Fusion Systems in Algebra <strong>and</strong> Topology, in print 2011.<br />

3. H.-J. Baues, G. Wirsching, <strong>Cohomology</strong> <strong>of</strong> small categories, J. Pure Appl. Algebra 38 (1985) 187-211.<br />

4. J. C. Becker, D. H. Gottlieb, Transfer maps for fibrations <strong>and</strong> duality, Compositio Math. 33 (1976), 107-133.<br />

5. D. Benson, <strong>Representations</strong> <strong>and</strong> <strong>Cohomology</strong> I, Cambridge Studies in Adv. Math. 30, Cambridge University Press (1998).<br />

6. D. Benson, <strong>Representations</strong> <strong>and</strong> <strong>Cohomology</strong> II, Cambridge Studies in Adv. Math. 31, Cambridge University Press (1998).<br />

7. S. Bouc, Homologie de certains ensembles ordonnés, Comptes Rendues Acad. Sci. Paris 299, Série I (1984), 49-52.<br />

8. A. Bousfield, E. Friedl<strong>and</strong>er,<br />

9. A. Bousfield, D. Kan, Homotopy Limites, Completions <strong>and</strong> Localizations, Lecture Notes in Math 304 (1972) Springer.<br />

10. C. Broto, R. Levi, B. Oliver, Homotopy equivalences <strong>of</strong> p-completed classifying spaces <strong>of</strong> finite groups, Invent. Math.<br />

11. C. Broto, R. Levi, B. Oliver, The homotopy theory <strong>of</strong> fusion systems, J. Amer. Math. Soc. 16 (2003) 779-856.<br />

12. K. Brown, <strong>Cohomology</strong> <strong>of</strong> Groups, GTM87, Springer 1982.<br />

13. L. Budach, B. Graw, C. Meinel, S. Waack, Algebraic <strong>and</strong> Topological Properties <strong>of</strong> <strong>Finite</strong> Partially Ordered Sets, Teubner-<br />

Texte zur Mathematik, B<strong>and</strong> 109, Teubner Publishing House, 1987.<br />

14. H. Cartan, S. Eilenberg, Homological Algebra, Princeton University Press 1956.<br />

15. T. tom Dieck, Transformation groups, de Gruyter Studies in Math. 8, Walter de Gruyter, 1987.<br />

16. A. Dold, D. Puppe, Homologie nicht-additiver funktoren, Ann. Inst. Fourier (Grenoble) 11 (1961) 201-312.<br />

17. W. G. Dwyer, Homology decompositions for classifying spaces <strong>of</strong> finite groups, Topology 36 (1997), 783-804.<br />

18. W. G. Dwyer, H.-W. Henn, Homotopy Theoretic Methods in Group <strong>Cohomology</strong> (Birkhäuser 2001).<br />

19. W. G. Dwyer, D. Kan, A classification theorem for diagrams <strong>of</strong> simplicial sets, Topology, 23 (1984) 139-155.<br />

20. W. G. Dwyer, D. Kan,<br />

21. W. G. Dwyer, C. W. Wilkerson, Homotopy fixed-point methods for Lie groups <strong>and</strong> finite loop spaces, Ann. <strong>of</strong> Math. (2) 139<br />

(1994), 395-442.<br />

22. S. Eilenberg, Homological dimension <strong>and</strong> syzygies, Ann. <strong>of</strong> Math., 64 (1956) 328-336.<br />

23. P. Eting<strong>of</strong>, V. Ostrik, <strong>Finite</strong> tensor categories, Mosc. Math. J. 4 (2004) 627-654, 782-783.<br />

24. L. Evens, The cohomology ring <strong>of</strong> a finite group, Trans. Amer. Math. Soc., 101, (1961) 224-239.<br />

25. E. Friedl<strong>and</strong>er, V. Franjou, T. Pirashivili, L. Schwartz, Functor Homology Theory, Panoramas et Synthèses 16, Soc. Math.<br />

France 2003.<br />

26. R. Fritsch, M. Golasiński, Topological, simplicial <strong>and</strong> categorical joins, Arch. Math. (Basel) 82 (2004) 468-480.<br />

27. P. Gabriel, Des Catégories Abélinnes, Bull. Soc. Math. France 90 (1962) 323-448.<br />

28. P. Gabriel, M. Zisman, Calculus <strong>of</strong> Fractions <strong>and</strong> Homotopy Theory, Ergebnisse der Mathematik und ihrer Grenzgebiete,<br />

New Series, 35 Springer-Verlag New York (1967).<br />

29. A. I. Generalov, Relative homological algebra, cohomology <strong>of</strong> categories, posets <strong>and</strong> coalgebras, H<strong>and</strong>book <strong>of</strong> Algebra, Vol 1,<br />

M. Hazewinkel ed., North-Holl<strong>and</strong>, 1996.<br />

30. M. Gerstenhaber, S. D. Schack, Simplicial cohomology is Hochschild cohomology, J. Pure <strong>and</strong> App. Algebra, 30 (1983)<br />

143-156.<br />

31. J. Grodal, Higher limits via subgroup complexes, Ann. <strong>of</strong> Math. 155 (2002) 405-457.<br />

32. K. Gruenberg, Cohomological topics in group theory, Lecture Notes in Math. 143, Springer-Verlag, 1970.<br />

33. D. Happel, Hochschild cohomology <strong>of</strong> finite-dimensional algebras, Lecture Notes in Math. 1404 (1989) 108-126.<br />

34. A. Hatcher, Algebraic Topology, Cambridge University Press 2002.<br />

35. P. Hilton, U. Stammbach, A Course in Homological Algebra (Second edition), GTM 4 Springer (1997).<br />

36. G. H<strong>of</strong>f, Cohomologies et extensions de categories, Math. Sc<strong>and</strong>. 74 (1994) 191-207.<br />

37. J. Hollender, R. M. Vogt, Modules <strong>of</strong> topological spaces, applications to homotopy limits <strong>and</strong> E ∞ structures, Arch. Math.<br />

59, 115-129 (1992).<br />

38. M. Hovey, J. Palmieri, N. Strickl<strong>and</strong>, Axiomatic Stable Homotopy Theory, Amer. Math. Soc. Memoirs 610 (1997).<br />

39. K. Igusa, D. Zacharia, On the cohomology <strong>of</strong> incidence algebras <strong>of</strong> partially ordered sets, Comm. Algebra, 18 (1990) 873-887.<br />

40. S. Jackowski, J. McClure, Homotopy decomposition <strong>of</strong> classifying spaces via elementary abelian subgroups, Topology 31<br />

(1992), 113-132.<br />

41. S. Jackowski, J. McClure, B. Oliver, Homotopy decomposition <strong>of</strong> self-maps <strong>of</strong> BG via G-actions, I <strong>and</strong> II, Ann. <strong>of</strong> Math.<br />

135 (1992) 183-226; 227-270.<br />

42. S. Jackowski, J. Slomińska, G-functors, G-posets <strong>and</strong> homotopy decompositions <strong>of</strong> G-spaces, Fundamenta Mathematicae 169<br />

(2001) 249-287.<br />

43. R. Knörr, G. Robinson, Some remarks on a conjecture <strong>of</strong> Alperin, J. London Math. Soc. (2) 39 (1989), 48-60.<br />

44. N. Kuhn, The generic representation theory <strong>of</strong> finite fields: a survey <strong>of</strong> basic structure, Infinite Length Modules, Bielefeld<br />

1998.<br />

45. M. Linckelmann, Transfer in Hochschild cohomology <strong>of</strong> blocks <strong>of</strong> finite groups, Alg. Rep. Theo. 2 (1999) 107-135.<br />

46. M. Linckelmann, Varieties in block theory, J. Algebra 215 (1999) 460-480.<br />

47. M. Linckelmann, Fusion category algebras, J. Algebra 277 (2004) 222-235.<br />

48. M. Linckelmann, Alperin’s conjecture in terms <strong>of</strong> equivariant Bredon cohomology, Math. Z. 250 (2005), 495-513.<br />

49. M. Linckelmann, Hochschild <strong>and</strong> block cohomology varieties are isomorphic, Preprint 2009.<br />

50. M. Linckelmann, On inverse categories <strong>and</strong> transfer in cohomology, Preprint 2010.<br />

51. W. Lück, Transformation groups <strong>and</strong> algebraic K-Theory, Lecture Notes in Math. 1408, Springer-Verlag, 1989.<br />

52. S. Mac Lane, <strong>Categories</strong> for the Working Mathematician (Second edition), GTM 5 (Springer 1998).<br />

53. S. Mac Lane, Homology, Classics in Mathematics, Springer-Verlag, 1995.<br />

54. J. P. May, Simplicial Objects in Algebraic Topology, Chicago University Press 1967.<br />

55. J. McCleary, A User’s Guide to Spectral Sequences Cambridge Studies in Adv. Math. 58, Cambridge University Press 2000.<br />

56. B. Mitchell, Rings with several objects, Adv. Math. 8 (1972) 1-161.<br />

57. B. Oliver, Higher limits via Steinberg representations, Comm. Algebra 22 (1994) 1381-1393.<br />

58. B. Oliver, J. Ventura, Extensions <strong>of</strong> linking systems with p-group kernel, Math. Ann. 338 (2007) 983-1043.


272 6 Connections with group representations <strong>and</strong> cohomology<br />

59. R. S. Pierce, Associative Algebras, GTM 88, Springer-Verlag New York 1982.<br />

60. L. Puig, Structure locale dans les groupes finis, Bull. Soc. Math. France Suppl. Mém. 47 (1976).<br />

61. D. Quillen, Higher algebraic K-theory I, in: Lecture Notes in Math. 341 (Springer-Verlag, 1973), pp. 85-147.<br />

62. D. Quillen, Homotopy properties <strong>of</strong> the poset <strong>of</strong> nontrivial p-subgroups <strong>of</strong> a group, Adv. Math. 28 (1978), 101-128.<br />

63. Reedy,<br />

64. V. Reiner, P. J. Webb, Combinatorics <strong>of</strong> the bar resolution in group cohomology, J. Pure Appl. Algebra 190 (2004), no. 1-3,<br />

291-327.<br />

65. M. A. Ronan, S. D. Smith, Sheaves on buildings <strong>and</strong> modular representations <strong>of</strong> Chevalley groups, J. Algebra 96 (1985),<br />

319-346.<br />

66. J.-E. Roos, Sur les foncteurs dérivés de lim, Applications, Compte Rendue Acad. Sci. Paris 252 (1961) 3702-3704.<br />

←−<br />

67. S. Siegel, S. Witherspoon, The Hochschild cohomology ring <strong>of</strong> a groups algebra, Proc. London Math. Soc. 79 (1999) 131-157.<br />

68. J. Slomińska, Decompositions <strong>of</strong> categories over posets <strong>and</strong> cohomology <strong>of</strong> categories, Manuscripta Math. 104 (2001) 21-38.<br />

69. S. D. Smith, Subgroup Complexes, Book draft.<br />

70. O. Solberg, Support varieties for modules <strong>and</strong> complexes, Comtep. Math. 406, Amer. Math. Soc. (2006) 239-270.<br />

71. M. Suarez-Alvarez, The Hilton-Eckmann argument for the anti-commutativity <strong>of</strong> cup product, Proc. Amer. Math. Soc. 132<br />

(2004) 2241-2246.<br />

72. M. Sweedler, Hopf Algebras, Math. Lecture Notes, No. 44, Addison-Wesley Pub Co 1969.<br />

73. D. Swenson, The Steinberg Complex <strong>of</strong> an Arbitrary <strong>Finite</strong> Group in Arbitrary Positive Characteristic, PhD thesis, University<br />

<strong>of</strong> Minnesota 2009.<br />

74. P. Symonds, The Bredon cohomology <strong>of</strong> subgroup complexes, J. Pure Appl. Algebra 199 (2005), no. 1-3, 261-298.<br />

75. J. Thévenaz, G-algebras <strong>and</strong> Modular Representation Theory, Oxford University Press 1995.<br />

76. J. Thévenaz, P. J. Webb, Homotopy equivalence <strong>of</strong> posets with a group action, J. Combin. Theory Ser. A 56 (1991), 173-181.<br />

77. R. W. Thomason, Homotopy colimits in the category <strong>of</strong> small categories, Math. Proc. Cambridge Philoso. Soc. 85 (1979),<br />

91-109.<br />

78. Tornehave,<br />

79. B. B. Venkov, <strong>Cohomology</strong> algebras for some classifying spaces, Dokl. Akad. Nauk SSSR vol. 127 (1959) 943-944.<br />

80. P. J. Webb, Supgroup complexes, Arcata Conf. on <strong>Representations</strong> <strong>of</strong> <strong>Finite</strong> Groups, Proc. Symp. Pure Math. 47 (Part I),<br />

Amer. Math. Soc. 1987, 349-365.<br />

81. P. J. Webb, A local method in group cohomology, Comment. Math. Helvetivi 62 (1987), 135-167.<br />

82. P. J. Webb, An introduction to the representations <strong>and</strong> cohomology <strong>of</strong> categories, in: Group Representation Theory, (EPFL<br />

Press 2007), pp. 149-173.<br />

83. P. J. Webb, St<strong>and</strong>ard stratifications <strong>of</strong> EI categories <strong>and</strong> Alperin’s weight conjecture, J. Algebra 320 (2008) 4073-4091.<br />

84. C. Weible, An Introduction to Homological Algebra, Cambridge Studies in Adv. Math. 38, Cambridge University Press 1994.<br />

85. V. Welker, G. Ziegler, R. Zivaljević, Homotopy colimits-comparison lemmas for combinatorial applications, J. Reine Angew.<br />

Math. 509 (1999) 117-149.<br />

86. F. Xu, <strong>Representations</strong> <strong>of</strong> small categories <strong>and</strong> their applications, J. Algebra 317 (2007) 153-183.<br />

87. F. Xu, Hochschild <strong>and</strong> ordinary cohomology rings <strong>of</strong> small categories, Adv. Math. 219 (2008) 1872-1893.<br />

88. F. Xu, Tensor structure on kC-mod <strong>and</strong> cohomology, to appear in Skye 2009 Conference Proceedings.<br />

89. F. Xu, On local categories <strong>of</strong> finite groups, Preprint 2010.<br />

90. F. Xu, Becker-Gottlieb transfer for Hochschild cohomology, Preprint 2011.<br />

91. F. Xu, Support varieties for transporter category algebras, Preprint 2011.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!