the ordinary ro c -graded bredon cohomology of a point · eilenberg-steenrod axioms for cohomology...

60
The Ordinary RO(C 2 )-Graded Bredon Cohomology of a Point Tiago Duarte Guerreiro Thesis to obtain the Master of Science Degree in Mathematics and Applications Examination Committee Supervisor: Prof. Pedro Ferreira dos Santos Co-supervisor: Prof. Wojciech Chacholski Members of the Committee: Chairperson: Prof. Pedro Manuel Agostinho Resende Prof. Gustavo Rui Gonçalves Fernandes de Oliveira Granja May 2015

Upload: others

Post on 03-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

The Ordinary RO(C2)-Graded Bredon Cohomology of aPoint

Tiago Duarte Guerreiro

Thesis to obtain the Master of Science Degree in

Mathematics and Applications

Examination Committee

Supervisor: Prof. Pedro Ferreira dos SantosCo-supervisor: Prof. Wojciech Chacholski

Members of the Committee: Chairperson: Prof. Pedro Manuel Agostinho ResendeProf. Gustavo Rui Gonçalves Fernandes de Oliveira Granja

May 2015

Page 2: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

ii

Page 3: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Para a minha avo.

iii

Page 4: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

iv

Page 5: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Acknowledgments

This project would not have been possible without the invaluable help and support of my advisor,

Pedro Ferreira dos Santos, who always found the time and patience to walk me through the course of

this work. I would also like to thank my co-advisor, Wojciech Chacholski, who helped me in my first steps

in Algebraic Topology. I would also like to thank Gustavo Granja for reading my thesis and giving me

valuable suggestions which made everything much clearer. It was a pleasure to work with them.

To all my friends who supported me, specially Ricardo, from whom I have probably received more

constant support than from anyone else during the whole process of this work and, of course, to Feder-

ica.

To my family, who always trusted me and to whom I owe most of what I am.

I would like to mention my University, Instituto Superior Tecnico, which I was lucky enough to be part

of for the last years of my studies. It opened me an infinite number of doors and it felt like a second

home, if not the first.

This project was supported by an FCT grant, ref PTDC/MAT-GEO/0675/2012, for which I thank Mar-

garida Mendes Lopes and Pedro Ferreira dos Santos.

v

Page 6: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

vi

Page 7: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Resumo

Seja C2 o grupo cıclico com dois elementos. Calculamos a estrutura de grupo da cohomologia

ordinaria de Bredon graduada em RO(C2) de um ponto para qualquer functor de Mackey constante.

Os ingredientes principais para tal sao o uso de sucessoes de cofibracao adequadas juntamente com

a versao equivariante da Dualidade de Spanier-Whitehead para um ponto.

Quando M provem de um anel comutativo, existe uma estrutura de anel em H∗,∗(X;M) que tem

origem num produto cup generalizado. Calculamos explicitamente essa estrutura para os functores

de Mackey constantes R, Z and Z2, usando aplicacoes entre os aneis de cohomologia equivariante e

cohomologia singular e o conceito de espectro-G que representa teorias de cohomologia de Bredon

graduadas em RO(G).

No decurso deste trabalho introduzimos varias nocoes em teoria em homotopia equivariante, nomeada-

mente os conceitos centrais de complexo-CW-G, functores de Mackey e espectros-G.

Palavras-chave: Complexo-CW-G, Functor de Mackey, Espectro-G, Cohomologia Equivari-

ante.

vii

Page 8: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

viii

Page 9: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Abstract

Let C2 be the cyclic group with two elements. We compute the group structure of the ordinary

RO(C2)-graded Bredon cohomology of a point for any constant Mackey functor. The main ingredients

to do so are suitable cofiber sequences together with equivariant Spanier-Whitehead Duality for a point.

When M comes from a commutative ring, there is a ring structure arising from the existence of a

generalized cup product in H∗,∗(X;M). We compute this structure explicitly for the constant Mackey

functors R, Z and Z2 using existing ring maps between equivariant cohomology and singular cohomol-

ogy rings and the notion of G-spectra which represent RO(G)-graded Bredon cohomology theories.

Throughout this work we introduce various notions in equivariant homotopy theory, namely the central

concepts of G-CW-complex, Mackey functors and G-spectra.

Keywords: G-CW-complex, Mackey functor, G-spectra, Equivariant Cohomology.

ix

Page 10: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

x

Page 11: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Contents

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Resumo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1 Introduction 1

1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Organization of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

2 Preliminaries 4

2.1 The equivariant homotopy category . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.1.1 G-maps and G-spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.1.2 Group representations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.1.3 Recovering the non-equivariant setting . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.1.4 G-Homotopy theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.2 Bredon Homology and Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2.1 G-CW-complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.2.2 Definitions and examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2.3 Relations between Bredon and non-equivariant homologies . . . . . . . . . . . . . 18

2.3 Spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3.1 Non-equivariant spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

2.3.2 Equivariant spectra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21

2.4 Mackey Functors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.5 The RO(G)-grading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.5.1 Representability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26

3 Computation of the cohomology ring of a point 29

3.1 RO(G)-graded cohomology of a point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.1.1 The group structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.1.2 The forgetful map to singular cohomology . . . . . . . . . . . . . . . . . . . . . . . 33

xi

Page 12: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

3.1.3 The multiplicative structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

Bibliography 43

xii

Page 13: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

List of Tables

2.1 The bookkeping diagrams. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

3.1 Group structure of H∗,∗(pt;R). The first index runs horizontally and the second index runs

vertically. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.2 Group structure of H∗,∗(pt;Z). The first index runs horizontally and the second index runs

vertically. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.3 Group structure of H∗,∗(pt;Z2). The first index runs horizontally and the second index

runs vertically. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

xiii

Page 14: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

xiv

Page 15: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Chapter 1

Introduction

1.1 Motivation

In [Bredon, 1967], Bredon initiated the study of equivariant cohomology theory, extending the usual

Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a

finite group G. This is achieved using the language of coefficient systems which are functors from the

category of orbits of G to the category of abelian groups.

These axioms, however, assume that cohomology functors are graded in Z although equivariant

cohomology theories are more naturally graded by the free abelian group generated by isomorphism

classes of irreducible representations of G. This group is usually called RO(G) for historical reasons

even though RO(G) is actually the group of equivalence classes of formal sums of representations.

In fact, it is proven in [May, 1996] that if the coefficient system of Bredon cohomology extends to a

Mackey functor, or, in other words, is the underlying coefficient system of a Mackey functor, then Bredon

cohomology theory extends to an RO(G)-graded cohomology theory.

This extension comes, however, at a price. It is significantly harder to make explicit computations

even for very simple spaces like a point and the existing literature is often incomplete. As a result, many

calculations that ought to be known are not and many standard tools from the non-equivariant world are

yet to be generalized to this setting.

This thesis is an attempt to make a self-contained exposition of basic G-equivariant homotopy theory

and present some important explicit computations in the case when G = C2.

1.2 Organization of the thesis

Throughout this work, a topological space means a compactly generated weak Hausdorff space and

we use X and Y to denote topological spaces with an action by a finite group G, unless otherwise

stated. By a map between topological spaces we mean a continuous map. For any prime p, the cyclic

group with p elements is written Cp to denote the group of equivariance and Zp otherwise. We are

particularly interested in studying equivariant cohomology in the case G = C2. We denote the constant

1

Page 16: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Mackey functor, defined in example (2.17), by M , and we will be specially interested in the cases when

M = R, Z and Z2. Singular cohomology in degree p of a space X with coefficients in an abelian group

M is written as Hpsing(X;M) and similarly for singular homology.

This work is organized in two main parts, the ”Preliminaries”, where we present the concepts and

results needed for the second part, the ”Computation of the cohomology ring of a point”.

Chapter 2 starts by giving an overview and tries to lay the foundations of the world of equivariant

homotopy theory. There, we generalize the notion of homotopy groups using representation spheres,

SV , where V is a finite-dimensional G-representation, which are spheres with an action, and that of

Eilenberg-MacLane spaces. We state and prove a panoply of adjunctions which are going to be essential

for our purposes. In particular, we prove that there is an adjunction,

G/H+ ∧ − : H − Top G− Top : U,

where, for K a subgroup of G, K − Top denotes the category of K-spaces and K-maps. We use this

to obtain a natural isomorphism, in Proposition (2.16), between the equivariant cohomology of C2+ ∧X

and singular cohomology of X, where X is a pointed C2-space.

In section (2.2), we introduce the main objects of equivariant Bredon cohomology, theG-CW-complexes.

These are built in much the same way as usual CW-complexes, only now we attach equivariant cells,

G/H × Dn, where H runs over the subgroups of G. We are then in a position to speak about the or-

dinary Z-graded Bredon cohomology and homology and give some examples. We also outline a proof

of a theorem by Lima-Filho, see [Lawson et al., 2003], that relates ordinary Z-graded Bredon homology

with singular homology in subsection (2.2.3).

We define and give examples of Mackey functors in section (2.4). Of particular importance for us is

the constant Mackey functor M described in example (2.17).

The RO(G)-graded Bredon cohomology theories, introduced in section (2.5), are representable

through a generalization of spectra, the genuine G-spectra, which we explain in section (2.3). Later

on, dos Santos, [dos Santos, 2003], identified a simple model for the spaces that represent these the-

ories generalizing the Dold-Thom theorem that we outline in this section. He proved, in particular, that

there is a natural equivalence,

HGV (X;M) ∼=

[SV ,M ⊗X

]G,

where V is a finite-dimensional G-representation and M ⊗X is a specific G-space with the property of

being an equivariant infinite loop-space. This will also be helpful in establishing an equivariant version

of Spanier-Whitehead Duality for a point, which we heavily use in computations.

Finally, in chapter 3, we proceed to the computations of the cohomology of a point. Subsection

(3.1.1) deals with the computation of the group structure of Hp,q(pt;M). There we make extensive use

of the cofiber sequence S(Rq,q)+ → D(Rq,q)+ → Sq,q of pointed G-spaces which is sequence (2.1) in

the text. The induced long exact sequence in homology allows us to reduce the problem of computing

the cohomology group of a point to that of computing the homology or cohomology of some projective

space. The results are summarized in tables (3.1), (3.2) and (3.3) and in theorem (3.1).

An important tool will be the forgetful map to singular cohomology, described in section (3.1.2). For

2

Page 17: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

any pointed G-CW-complex X and constant Mackey functor M , the map,

[X,M ⊗ SV

]G

ϕ−→[C2+ ∧X,M ⊗ SV

]G,

induced by the folding map ∇ : C2+ ∧X → X, can be seen as a forgetful map to singular cohomology.

We include examples and a straightforward proof of the important fact that this map is multiplicative with

respect to the cup product in H∗,∗(X;M).

This work ends in section (3.1.3) with the computation of the cohomology ring of a point, H∗,∗(pt;M),

whenM is each of the above mentioned Mackey functors. As a crucial step we compute the cohomology

ring of C2 and use it to compute that of a point. It is there that we use the forgetful map to singular

cohomology and the machinery of G-spectra to get our final results.

3

Page 18: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Chapter 2

Preliminaries

2.1 The equivariant homotopy category

2.1.1 G-maps and G-spaces

Let G be a topological group with unit e and X be a topological space. A left G-action on X is a

continuous map

ρ : G×X −→ X

(g, x) 7−→ g · x

such that

1. e · x = x

2. g(h · x) = (gh) · x.

for all g, h ∈ G and x ∈ X.

A G-space is a topological space together with a left G-action. We usually drop the dot and write

ρ(g, x) = gx. A right G-action is a map X × G → X such that xe = x and (xh)g = x(hg). We usually

work with left actions since if (x, g) 7→ xg is a right action then, (g, x) 7→ xg−1 is a left action.

A (continuous) map f : X → Y between G-spaces is called a G-map or a G-equivariant map if

f(gx) = gf(x) for all g ∈ G, x ∈ X. G-spaces and G-maps form a category denoted by G − Top.

The morphisms in that category are denoted by homG(X,Y ) or homG−Top(X,Y ) and an isomorphism

between objects is called a G-homeomorphism.

A based G-space is a G-space with a basepoint fixed by G and a map between based G-spaces is

a G-map that preserves the basepoint. Notice that it is possible to turn any unbased G-space X into a

based one by letting X+ = X∐∗ where ∗ is fixed by G. Based G-spaces and based G-maps form a

category denoted by G− Top∗. If (X,x0) and (Y, y0) are G-spaces we define their smash product as

X ∧ Y = (X × Y )/(X × y0 ∪ Y × x0).

4

Page 19: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

The orbit space of a G-space X, denoted X/G, consists of equivalence classes of the form x ∼ gx

where g ∈ G and x ∈ X. The orbit space can be endowed with the quotient topology. See [tom Dieck,

1987].

Example 2.1. 1. Any topological group G is a G-space.

2. Let H < G. The group multiplication H ×G→ G is a left H-action.

3. If H / G, then G acts on H by conjugation.

4. G/H is the orbit space of the right action G × H → G. It is itself a G-space with action given by

(g′, gH) 7→ g′gH.

5. If X and Y are G-spaces, then X × Y is a G-space where the action is componentwise. Such an

action is called diagonal. If X and Y are based then X ∧ Y is a G-space with action induced by

the diagonal action.

6. The space of continuous maps between G-spaces X and Y endowed with the compact open

topology, homTop(X,Y ), is a G-space under conjugation: Let f : X → Y . Define (gf)(x) =

gf(g−1x). If X and Y are based we denote by F (X,Y ) the subspace of homTop(X,Y ) consisting

of based maps.

7. Suppose C2 acts on Sn via the antipodal involution. The orbit space of this action is RPn.

An action of G in X is called free if, for all x ∈ X, gx = x implies that g is the identity element of G. It

is trivial if all the points of X are fixed by the action.

2.1.2 Group representations

Let G be a group. A representation of G, or G-representation, on a vector space V over the field k

is a group homomorphism,

ρ : G→ GL(V ).

Thus, defining a representation of G on V is the same as defining a linear map,

ρ(g) : V → V g ∈ G,

such that ρ(e) = idV and ρ(g1g2) = ρ(g1)ρ(g2) for g1, g2 ∈ G and e the identity of G. Another equivalent

way to see a representation is as an action on V ,

ρ : G× V −→ V

(g, v) 7−→ g · v,

where g acts linearly.

5

Page 20: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

If V has the extra structure of an inner product, a G-representation is orthogonal if ρ(G) is a sub-

group of the orthogonal automorphisms of V , O(V ). An abuse of notation is frequent and the represen-

tation is sometimes denoted by (ρ, V ) or just by V when ρ is clear from the context. The dimension of

the representation is the dimension of V over k, and it is written dim(ρ). The representation is faithful if

ker(ρ) = e and it is called trivial if ρ(g) = idV for every g ∈ G.

Example 2.2. Let G = C2 and V be the vector space R2 over R. Choose the canonical basis for R2. We

have, for example, the representation given by

ρ(e) =

1 0

0 1

, ρ(g) =

−1 0

0 1

This is a two-dimensional faithful representation of C2 on R2.

Example 2.3. Let G be the additive group of integers, (Z,+), and ρ a representation of G on V . Then

ρ(0) = idV and, for every r ∈ Z, ρ(r) = ρ(1)r. The representation is then completely determined by the

invertible linear map ρ(1) : V → V .

It is a basic problem of representation theory to classify all the representations of a group. In order

to do that it is necessary to have an adequate notion of when two representations are essentially the

same. This leads us to define isomorphism of representations.

Two representations (ρ1, V1) and (ρ2, V2) are isomorphic if there is a linear isomorphism ϕ : V1 → V2

that interwines the actions of G, that is, such that

ϕ ρ1(g) = ρ2(g) ϕ, ∀ g ∈ G.

Let (ρ, V ) be a representation of the group G. A linear subspace W of V is G-invariant if ρ(g)(W ) ⊂

W , for every g ∈ G. In that case we can define a representation (ρW ,W ) by ρW (g)(w) = ρ(g)(w), for

every g ∈ G and w ∈ W and we call W a subrepresentation of V . A subrepresentation W of V is

called proper if W 6= 0 and W 6= V . Given two representations, (ρ1, V1) and (ρ2, V2), we can form a new

representation, (ρ1 ⊕ ρ2, V1 ⊕ V2) by

ρ1 ⊕ ρ2 : G→ GL(V1 ⊕ V2)

g 7→ ρ1(g)⊕ ρ2(g),

where V1 ⊕ V2 denotes the direct sum of the vector spaces V1 and V2.

A nonzero representation is irreducible if it has no proper subrepresentations and it is completely

reducible if it is a direct sum of irreducible subrepresentations. A central result in representation theory

for finite groups is Maschke’s theorem:

Theorem 2.1. Let G be a finite group and k a field whose characteristic does not divide the order of G.

Then every representation V is completely reducible.

Example 2.4. There are only two irreducible realC2-representations up to isomorphism. The one dimen-

sional trivial representation denoted by 1 or R1,0, and the one dimensional sign representation denoted

6

Page 21: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

by σ or R1,1. Any C2-representation, V , is thus isomorphic to a direct sum of these two. In other words,

V =(R1,0

)p ⊕ (R1,1)q. We simplify notation and write V = Rp+q,q.

To any representation V of the group G we can associate its representation sphere. The represen-

tation sphere of the representation V , written SV , is the one-point compactification of V . The action of

G on V extends to SV acting trivially on∞.

Given an orthogonal representation V , we define the unit disk in V as

D(V ) = v ∈ V | ‖v‖ ≤ 1

and the unit sphere in V as

S(V ) = v ∈ V | ‖v‖ = 1.

There is an important induced cofiber sequence of based G-spaces,

S(V )+ → D(V )+ → SV , (2.1)

arising from the homeomorphism

D(V )/S(V ) ∼= SV .

For more details see [Masulli, 2011].

Following the previous example, we write SV = Sp+q,q when V = Rp+q,q is a C2-representation. The

first index, p + q, is called the topological dimension of the sphere and the second index, q, is called

the weight of the representation.

2.1.3 Recovering the non-equivariant setting

In this section we explore an important tool when working within the equivariant world: It is often

possible to translate equivariant information into non-equivariant data. Roughly, the idea is to associate

to an equivariant map a set of maps between fixed point sets.

Let H < G be a subgroup of G. The H-fixed point set of X is the set

XH = x ∈ X |hx = x, ∀h ∈ H.

Example 2.5. The normalizer ofH inG, NG(H), acts naturally onXH . In fact, if x ∈ XH and g ∈ NG(H)

then g−1hg ∈ H for every h ∈ H. So, g−1hgx = x and hgx = gx, which implies that gx is also in XH .

Call NG(H)/H the Weyl group of H in G and denote it by WG(H). Since H acts trivially on XH the

action of the normalizer on XH induces an action by the Weyl group.

It is interesting to note that a G-map f : X → Y takes fixed point sets to fixed point sets. If H is

a subgroup of G and x ∈ XH , then hf(x) = f(hx) = f(x) and so f(x) ∈ Y H . Therefore we have a

fixed point functor , (−)H : G − Top → Top, that takes a G-map f : X → Y to the continuous map

fH : XH → Y H .

7

Page 22: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Recall from example (2.1) that we can see homTop(X,Y ) as a G-space under conjugation. There is

an equality of G-spaces,

homG−Top(X,Y ) = homTop(X,Y )G.

Proposition 2.2. Let X be a locally compact Hausdorff G-space and Y an Hausdorff G-space. Then,

for any G-space B, there is a natural homeomorphism,

homG−Top(X × Y,B)ψ→ homG−Top(Y,homTop(X,B)),

where, for f : X × Y → B we define ψ(f)(y)(x) = f(x, y) with x ∈ X and y ∈ Y .

Proposition 2.3. Let X be a G-space and K a space with the trivial G-action. Then, the following

natural homeomorphisms hold:

1. homG−Top(K,X) ∼= homTop(K,XG).

2. homG−Top(X,K) ∼= homTop(X/G,K).

The details of the previous propositions can be seen in [May, 1996].

Next, we construct the so called induced and coinduced spaces. Let H be a subgroup of G and Y

an H-space. We define the induced G-space as the space

G×H Y = (G× Y )/ ∼

where (gh, y) ∼ (g, hy) for all g ∈ G, h ∈ H and y ∈ Y . The G-action on G ×H Y is defined by

γ · [g, y] 7→ [γg, y]. Similarly, the coinduced G-space is a subspace of homTop(G, Y ) and is

mapH(G, Y ) = f : G→ Y | f(gh−1) = hf(g), h ∈ H, g ∈ G.

The G-action on mapH(G, Y ) is defined by (γf)(g) = f(γ−1g). Analogous constructions can be made

for based spaces. It is usual to write G+ ∧H Y for the induced based G-space and map∗H(G+, Y ) for the

coinduced based G-space.

Lemma 2.4. If X is a G-space and H is a subgroup of G, then there are G-homeomorphims:

1. G×H X ∼= G/H ×X

2. mapH(G,X) ∼= homTop(G/H,X).

Proof. For the first isomorphism we consider the two maps

G×H X → G/H ×X

([g, x]) 7→ ([g] , gx)

8

Page 23: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

and

G/H ×X → G×H X

([g] , x) 7→ ([g, g−1x

]).

These are well defined G-maps that are inverses of each other and the result follows. For the second

isomorphism we consider the two maps

mapH(G,X)→ homTop(G/H,X)

(f : G→ X) 7→ (f : G/H → X)

[g] 7→ gf(g)

and

homTop(G/H,X)→ mapH(G,X)

(f : G/H → X) 7→ (f : G→ X)

g 7→ g−1f([g]).

Again these two maps are well defined G-maps that are inverses of each other.

Recall that, given a pair of functors F : C → D and G : D → C between categories C and D, we say

that F and G are adjoint functors and, in particular, F is left adjoint to G and G is right adjoint to F , if

there are natural transformations η : 1C → G F and ε : F G→ 1D, such that,

1F = εF Fη 1G = Gε ηG.

We call η and ε the unit and counit of the adjunction, respectively. This definition is equivalent as requiring

the existence of isomorphisms,

homD(FC,D) ∼= homC(C,GD),

natural in C ∈ C and D ∈ D. See [Awodey, 2010].

Both constructions of induced and coinduced spaces are functorial and give adjoints to the forgetful

functor U : G-Top→ H-Top.

Proposition 2.5. Let H be a subgroup of G and U be the forgetful functor from G-Top to H-Top. The

induced space functor, G×H − is left adjoint to U and the coinduced space functor mapH(G,−) is right

adjoint to U .

Proof. The unit and counit for the first adjunction are the H-map given by η : Y → G ×H Y, y 7→ [e, y]

and the G-map defined by ε : G×H X → X, [g, x] 7→ gx, respectively. In fact, the compositions,

G×H Y G×H (U(G×H Y )) G×H Y

[g, y] [e, [g, y]] [g, y]

G×HηY εG×HY

9

Page 24: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

and,

U(X) U(G×H (U(X))) U(X)

x [e, x] x

ηU(X) Uε

are the identity maps idG×HY and idU(X), respectively. For the second adjunction, the unit and counit

are the G-maps η : X → mapH(G,X), x 7→ (fx(g) = x, ∀g ∈ G) and the H-map ε : mapH(G, Y ) →

Y, f 7→ f(e), respectively. The verification is analogous to the previous one and is omitted.

We can rewrite the previous proposition in the more useful way:

Proposition 2.6. Let H be a subgroup of G, X be a G-space and Y an H-space. Then we have the

following natural homeomorphisms

1. homG−Top(G×H Y,X) ∼= homH−Top(Y,X),

2. homG−Top(X,mapH(G, Y )) ∼= homH−Top(X,Y ).

Corollary 2.7. Let H be a subgroup of G. The functor G/H × − : Top → G − Top is left adjoint to the

fixed point functor (−)H : G− Top→ Top.

Proof. The following isomorphisms are all natural:

homG−Top(G/H ×K,X) ∼= homG−Top(G×H K,X) Lemma (2.4).

∼= homH−Top(K,X) Proposition (2.6).

∼= homTop(K,XH). Proposition (2.3).

The arguments used apply in the based case and, in particular, G/H+ ∧ − is left adjoint to (−)H .

This adjunction allows us to regard equivariant homotopy groups of a pointed space X, defined in the

next section, as non-equivariant homotopy groups of the H-fixed points of X.

2.1.4 G-Homotopy theory

In this subsection we introduce the basic notions of equivariant homotopy theory and define, in

particular, equivariant homotopy groups.

Let I = [0, 1] be the unit interval with the trivial action by G. Two G-maps f, g : X → Y are said to be

G-homotopic if there is a continuous G-map,

F : X × I → Y,

such that F (x, 0) = f(x) and F (x, 1) = g(x), for every x ∈ X. Here X × I has the diagonal action. As in

the non-equivariant case, G-homotopy is an equivalence relation. That being so, we define [X,Y ]G to be

the set of G-homotopy classes of G-maps X → Y and let [f ] denote the G-homotopy class represented

by f : X → Y . The homotopy category hG-Top has G-spaces as objects and G-homotopy classes of

10

Page 25: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

G-maps as morphisms. A G-space is also an H-space for every H < G, and we write [X,Y ]H for the

set of homotopy classes of H-maps X → Y .

Similarly, we can define G-homotopies between based G-spaces. A based G-homotopy between

based G-maps, f, g : X → Y is a continuous based G-map,

F : X ∧ I+ → Y,

such that F (x, 0) = f(x) and F (x, 1) = g(x), for every x ∈ X. Based G-homotopy is also an equivalence

relation and we define [X,Y ]∗G to be the set of based G-homotopy classes of based G-maps X → Y .

The homotopy category hG-Top∗ is analogously defined. When the context is clear we use [X,Y ]G to

denote [X,Y ]∗G.

We now turn to one of the most important notions in equivariant homotopy theory, that of equivariant

homotopy group. Let X be a pointed G-space and H < G. How do we define πn(X) equivariantly? In

other words, how do we encode the homotopic information about X taking the G-action into consider-

ation? A natural place to start is from considering the homotopy classes of based G-maps Sn → X,

where Sn has the trivial G-action. This, however, is not enough to fully explore the equivariance since

it only captures the homotopy of XG. Thus, we look instead to homotopy classes of based G-maps

G/H+ ∧ Sn → X. Note that homG−Top∗(G/H+ ∧ Sn, X) ∼= homTop∗(Sn, XH).

Let n ∈ N and X a G-space. For each H < G, we define the component H of the nth equivariant

homotopy group of X, πHn (X), by

πHn (X) = [G/H+ ∧ Sn, X]∗G .

Proposition 2.8. Let X be a G-space, and H a subgroup of G. Then,

πHn (X) ∼= πn(XH).

The computation of the equivariant homotopy groups of a G-space consists of the computation of

the homotopy groups of its H-fixed points where H runs over the subgroups of G.

A G-map f : X → Y is said to be a weak G-equivalence if it induces isomorphisms

f∗ : πHn (X)∼=−→ πHn (Y )

for all H < G and n ≥ 0.

Notice that f : X → Y is a weak G-equivalence if and only if fH : XH → Y H is a weak equivalence

for all H < G. If such a map exists we say that X and Y are weakly G-equivalent and write X ' Y .

The notion of equivariant homotopy group can be further generalized by allowing spheres with

non-trivial G-actions, namely, by using non-trivial representation spheres. Let V be an orthogonal G-

11

Page 26: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

representation. In this case we define the V th equivariant homotopy group of X as

πHV (X) =[G/H+ ∧ SV , X

]∗G.

Notice that, from lemma (2.4) and proposition (2.6) it follows that πHV (X) ∼=[SV , X

]∗H

.

2.2 Bredon Homology and Cohomology

2.2.1 G-CW-complexes

We would like to extend the definition of CW-complex to one that accounts for the presence of equiv-

ariance. Such an extension, while meaningful, should also be restricting enough in order that analogs

of the basic theorems on CW-complexes hold.

Let Dn = x ∈ Rn| ‖x‖ ≤ 1 be the unit disk in Rn and Sn−1 its boundary, both with trivial G-action.

Set D0 as a point and S−1 = ∅ and call an equivariant n-cell of type G/H to the G-space G/H ×Dn.

The construction of a G-CW-complex is done inductively much like that of non-equivariant CW-

complexes. The main difference is that orbit spaces, G/H with H < G, play the role of 0-cells. These

are the equivariant ‘points’.

Start with a disjoint union of orbits, G/H, where H runs over the subgroups of G, as the 0-skeleton,

X0. Then, each skeleton Xn+1 is obtained from the previous one, Xn, by attaching equivariant cells,

G/Hα ×Dn+1 along the boundaries.

More precisely, a G-CW-complex X is a G-space together with a filtration Xn |n ≥ 0 such that:

1. X0 is a disjoint union of orbit spaces, G/Hα, where Hα < G.

2. Xn+1 is obtained from Xn by attaching equivariant (n + 1)-cells, G/Hα × Dn+1, via attaching

G-maps, ϕα : G/Hα × Sn → Xn. So,

Xn+1 =

(Xn

∐ϕα

G/Hα ×Dn+1

)/ ∼

where (gHα, x) ∼ ϕ(gHα, x) for (gHα, x) ∈ G/Hα × Sn.

3. X = ∪nXn and X carries the weak topology with respect to the family Xn. This means that

A ⊂ X is closed if and only if A ∩Xn is closed in Xn for every n.

Example 2.6. Let X be the the n-sphere with antipodal action by the group C2, S(Rn+1,n+1). We give

X the structure of a C2-CW-complex.

• X0 = C2 ×D0. Call its elements (e, x0) and (g, x0).

• To obtain X1 we attach C2 ×D1 to X0 via ϕ : C2 × S0 → X0. Let S0 = 0, 1. Then, we make the

choices, (e, 1) 7→ (e, x0) and (e, 0) 7→ (g, x0). By equivariance, we get that ϕ(g, 1) = g · ϕ(e, 1) =

(g, x0) and ϕ(g, 0) = g · ϕ(e, 0) = (e, x0). The space X1 is G-homeomorphic to S(R2,2).

12

Page 27: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

• We obtain X2 by attaching an equivariant 2-cell, C2 × D2 via a map ϕ : C2 × S1 → X1. Again,

ϕ(g, x) = g · ϕ(e, x), for every x ∈ S1. After identifying X1 with S(R2,2) we can regard ϕ(e,−) as

the identity of the underlying set of S(R2,2).

• We carry on like that until we attach an equivariant n-cell to Xn−1.

Then,

X =(C2 ×D0

)⋃(C2 ×D1

)⋃· · ·⋃

(C2 ×Dn) .

Example 2.7. Let X = Sn,n. We give X a C2-CW-complex structure. The action of C2 on X fixes the

north and south poles of Sn.

• X0 = C2/C2 ×D0∐C2/C2 ×D0. These are the two disjoint fixed 0-cells and we call them 0 and

∞.

• X1 is obtained from X0 by attaching to it a free 1-cell, C2 × D1 via ϕ : C2 × S0 → X0 such that

ϕ(e, 0) = ϕ(g, 0) = 0 and ϕ(e, 1) = ϕ(g, 1) =∞. We can identify X1 with S1,1.

• X2 is obtained from X1 by attaching to it a free 2-cell, C2 × D2 via ϕ : C2 × S1 → X1 such that

x 7→ ϕ(e, x) is a homeomorphism. This space is G-homeomorphic to a sphere of dimension 2 in

which C2 acts by rotating it by 180 denoted by S2,2.

• This process is carried until we attach a free n-cell, C2 ×Dn to Xn−1.

Then,

X =(C2/C2 ×D0

∐C2/C2 ×D0

)⋃(C2 ×D1

)⋃· · ·⋃

(C2 ×Dn) .

2.2.2 Definitions and examples

In [Bredon, 1967], Bredon describes the first equivariant cohomology theories. These theories are

defined for G-CW-complexes. The slogan here is ‘Orbits are equivariant points’.

The orbit category, denoted OG, is a category consisting of:

• Objects: The G-spaces G/H where H is a subgroup of G.

• Morphisms: G-maps between the objects.

Example 2.8. Let G = Cp where p is prime. We can represent OG through the diagram,

Cp

τ

where τ : Cp → Cp is multiplication by the generator and π : Cp → e is the projection. These satisfy

πτ = π and τp = id.

Recall that outside our equivariant world, the coefficients of the ordinary homology and cohomology

theories are defined to be H0(pt) or H0(pt), in the unreduced case. Such groups constitute the first

13

Page 28: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

distinctions between different homology and cohomology theories. Actually, these groups determine the

value of the (co)homology theory on any finite CW-complex. This is essentially because points constitute

the building blocks of a space with a CW structure.

In the equivariant setting, things get more complicated as the building blocks are now orbit spaces

G/H, for H < G. So, the coefficient system has to contain a family of groups, one for each orbit

space. Also, it is necessary to say how these blocks are related and that requires defining induced

homomorphisms between the family of groups corresponding to equivariant maps G/H → G/K.

More precisely, a contravariant coefficient system for G is a functor M : OopG −→ Ab and a

covariant coefficient system for G is a functor N : OG −→ Ab.

Example 2.9. Let X be a G-space.

• We have the contravariant constant coefficient system,

A : G/H 7→ A

f : G/K → G/H 7→ id : A→ A

where A is a fixed abelian group.

• Let H < K < G. The covariant constant coefficient system defined as,

A : G/H A

f : G/H → G/K f∗ : A A

r [K : H] r,

where A is an abelian group and [K : H] the index of H in K.

• Homotopy and homology groups are examples of contravariant coefficient systems:

πn(X)(G/H) := πn(XH).

and

Hn(X)(G/H) := Hn(XH).

The contravariant coefficient systems for a group G form an abelian category that we denote by

CG := Fun(OopG , Ab). See [Bredon, 1967] for more details. This means that it is possible to compute

kernels and cokernels and, in particular, to have chain complexes and homology.

Non-equivariantly, the cellular cohomology of a CW-complex X is defined to be the cohomology of

the cellular chain complex,

Cn(X) = Hn(Xn, Xn−1;Z),

where Xn denotes the n-skeleton of X.

14

Page 29: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Let X be a G-CW-complex. Based on the previous construction, we define contravariant coefficient

systems, Cn(X), called the cellular chains on X by,

Cn(X) := Hn(Xn, Xn−1;Z) : OG → Ab

G/H 7→ Hn((Xn)H , (Xn−1)H ;Z).

For any H < G, the H-fixed point set of the n-skeleton of X, is a CW-complex. Thus, for each H, we

have a long exact sequence of the triple ((Xn)H , (Xn−1)H , (Xn−2)H) with connecting homomorphism,

· · · → Hn((Xn)H , (Xn−1)H ;Z)→ Hn−1((Xn)H , (Xn−1)H ;Z)→ · · · .

This produces a natural transformation d : Cn(X) → Cn−1(X), such that the components, dH , H < G

satisfy d2H = 0 and, so, d2 = 0. We can, therefore, conclude that C∗(X) is a chain complex in CG. Given

M ∈ CG, homOG(Cn(X),M) is the group of natural transformations Cn(X) → M and C∗G(X;M) :=

homOG(C∗(X),M) is a cochain complex in Ab.

The ordinary Z-graded Bredon cohomology of a G-CW-complex X with coefficients in the con-

travariant coefficient systemM , denotedH∗G(X;M), is the cohomology of the cochain complexC∗G(X;M).

The coboundary map δ : CnG(X;M) → Cn+1G (X;M) is given by the composition f 7→ f d, for every

f ∈ CnG(X;M) with d : Cn+1(X)→ Cn(X) defined in the previous paragraph.

Example 2.10. We calculate the Bredon cohomology of X = C2+ ∧Sn where G = C2 acts by swapping

the summands. The first step is to give X a G-CW-structure. Let

• X0 = C2/C2 ×D0. This is a fixed 0-cell.

• Xi = X0, ∀i < n.

• Xn = C2 ×Dn. We attach a free n-cell via the constant map, ϕ : C2 × Sn−1 → X0.

Next, we calculate the values of the chain complex C∗(X) for each orbit:

Cn(X)(C2) = Hn(Xn, Xn−1;Z) = Z⊕ Z.

Cn(X)(C2/C2) = Hn((Xn)C2 , (Xn−1)C2 ;Z) = 0.

Ci(X)(C2) = Ci(X)(C2/C2) = Hi(Xi, Xi−1;Z) = 0 ∀1 ≤ i < n.

C0(X)(C2) = H0(X0, ∅) = Z.

C0(X)(C2/C2) = H0((X0)C2 , ∅) = Z.

For each 0 ≤ i ≤ n, we have to compute homOC2(Ci(X),M) where we choose M to be the constant

15

Page 30: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

coefficient system M = Z, given by the identity on orbits, Z id→ Z.

homOC2(Cn(X),Z) = Z⊕ Z;

homOC2(Ci(X),Z) = 0, ∀1 ≤ i < n;

homOC2(C0(X),Z) = Z.

We end up with a chain complex in Ab of the form:

0→ Z→ 0→ · · · → 0→ Z⊕ Z→ 0→ · · · .

Therefore, with such coefficient system, we have that HnC2

(X;M) = Z⊕ Z and H0C2

(X;M) = Z are the

only non-trivial Bredon cohomology groups for X.

The ordinary Z-graded Bredon homology of a G-CW-complex X with coefficients in the covariant

coefficient system N , denoted HG∗ (X;N), is the homology of the chain complex of abelian groups,

CG∗ (X;N), which in degree n is

Cn(X)⊗OG N =⊕

G/H∈OG

Cn(X)(G/H)⊗Z N(G/H)/ ∼ .

Where the equivalence relation ∼ is induced by (f∗m,n) ∼ (m, f∗n) for all f : G/H → G/K and

m ∈ Cn(X)(G/K), n ∈ N(G/H). The boundary map, CGn (X;N)→ CGn−1(X;N), is defined as ∂ = d⊗1

where d : Cn(X)→ Cn−1(X) is the map defined above.

Example 2.11. Let X = S1,1 be the one point compactification of R with an involution. This space has a

C2-CW-structure given by example (2.7). We calculate the Bredon homology groups of X with covariant

constant coefficient system Z from example (2.9). Recall that Cn(X)(G/H) = Hn((Xn)H , (Xn−1)H ;Z).

Then,

C1(X)(C2) = H1(X1, X0;Z) = Z⊕ Z.

C1(X)(C2/C2) = H1((X1)C2 , (X0)C2 ;Z) = 0.

C0(X)(C2) = H0(X0, ∅) = Z⊕ Z.

C0(X)(C2/C2) = H0((X0)C2 , ∅) = Z⊕ Z.

The natural transformation d : C1(X)→ C0(X) corresponds to the middle horizontal maps in the follow-

ing commutative diagram,

0 Z⊕ Z Z⊕ Z 0

0 0 Z⊕ Z 0.

0 id

The first row corresponds to the cellular chain complex associated to the free orbit, C∗(X)(C2), and the

second row to the cellular chain complex of the fixed point, C∗(X)(pt). The vertical maps are induced by

16

Page 31: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

the equivariant map π : C2 → pt. There is additional structure coming from the action of the generator

τ ∈ C2 on C∗(X)(C2).

From example (2.9), the covariant constant coefficient system Z takes the projection π : C2 → pt to

π∗ which is multiplication by 2 in Z and takes τ : C2 → C2 to the identity map in Z.

To get the chain complex related to Bredon homology in degree zero we tensor the right vertical

groups with the covariant constant coefficient system Z and get

C0(X)⊗ Z = C0(X)(pt)⊗ Z(pt)⊕

C0(X)(C2)⊗ Z(C2)/ ∼

= (Z⊕ Z)⊗ Z⊕

(Z⊕ Z)⊗ Z/ ∼

= Z⊕ Z⊕

Z⊕ Z/ ∼

Notice that, in degree zero, τ∗ : Z ⊕ Z → Z ⊕ Z is the identity map in Z ⊕ Z because the action on the

0-cells is trivial. The same holds for π∗ : Z⊕Z→ Z⊕Z. This makes the relation induced by τ trivial and

that induced by π be

π∗(m1,m2)⊗ 1 ∼ (m1,m2)⊗ π∗(1), (m1,m2) ∈ C0(X)(pt) ∼= Z⊕ Z

in the first two components, or, more explicitly,

(m1,m2, 0, 0) ∼ (0, 0, 2m1, 2m2).

So an element of the quotient is represented by an element of the form (0, 0, c, d) for c, d ∈ Z and, finally,

Z⊕ Z⊕

Z⊕ Z/ ∼ ∼= Z⊕ Z.

To get the chain complex in degree one we tensor the left vertical groups with the constant coefficient

system Z and get

C1(X)⊗ Z = C1(X)(pt)⊗ Z(pt)⊕

C1(X)(C2)⊗ Z(C2)/ ∼

∼= 0⊗ Z⊕

(Z⊕ Z)⊗ Z/ ∼

∼= Z⊕ Z/ ∼ .

We have that π∗ : C1(X)(pt) → C1(X)(C2) is the zero map since C1(X)(pt) is trivial and τ∗ swaps

the summands. So we obtain the relation,

τ∗(m3,m4)⊗ 1 ∼ (m3,m4)⊗ τ∗(1) (m3,m4) ∈ C1(X)(C2) ∼= Z⊕ Z

or, in other words,

(m4,m3) ∼ (m3,m4).

17

Page 32: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

So we have an isomorphism between Z⊕Z/ ∼ and Z. The only thing missing to calculate the 0th Bredon

homology of X is to calculate the image of the map

· · · → 0→ Z ∼= C1(X)⊗ Z ∂−→ C0(X)⊗ Z ∼= Z⊕ Z→ 0→ · · · .

Notice that the equivariance is captured within this chain complex. We have indeed two equivariant

0-cells and one equivariant 1-cell. The boundary map, ∂, sends each non-equivariant 1-cell to the

difference of the vertices in its boundaries and so the image of ∂ is equal to the image of 2∆− where ∆−

is the anti-diagonal map. Since any element of Z⊕Z may be uniquely written as a sum (m, 0) + (n,−n),

we conclude that HG0 (S1,1;Z) ∼= Z⊕ Z2 and that HG

1 (S1,1;Z) ∼= 0.

Equivariant Bredon cohomology and homology theories have reduced versions and those can be

defined analogously to the non-equivariant case. Consider the constant map ε : X → pt for a G-CW-

complex X and let M be a contravariant coefficient system. The kernel of the induced map in homology

ε∗ : HG∗ (X;M) → HG

∗ (pt;M) is the equivarint reduced Bredon homology of the pointed G-space X

and is denoted by HG∗ (X;M).

If X is not pointed, then HG∗ (X+;M) = HG

∗ (X;M). Equivariant reduced Bredon cohomology is

defined similarly and we refer to [Bredon, 1967] for the details.

2.2.3 Relations between Bredon and non-equivariant homologies

There is an important long exact sequence that relates Bredon homology with singular homology.

The goal of this section is to explain the ingredients in the proof of the following theorem and give some

applications.

Theorem 2.9 ([Lawson et al., 2003]). Let G = C2 and let X be a G-CW-complex. Then, there is a long

exact sequence,

· · · → Hn(X/G;Z)→ HGn (X;Z)→ Hn(XG;Z2)→ Hn−1(X/G;Z)→ · · · .

There is an analogous long exact sequence in reduced homology for pointed G-spaces which can

be obtained in a similar way to what we explain next. See [Lawson et al., 2003] for more details. Let X

be a topological space and denote by Z[X] the free abelian group on X. This is a topological group with

topology induced by that of X. If, moreover, X is a G-space, then Z[X] has a natural G-action given by

g(∑

nixi

)=∑

nigxi, ni ∈ Z, xi ∈ X.

Let τ be the generator of G. Call

Z[X]av = c+ τc|c ∈ Z[X]

the subgroup of averaged cycles of Z[X]G. This is a proper subgroup of Z[X]G and, so, we have an

18

Page 33: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

exact sequence,

0→ (Z[X])av→ Z[X]G → Z[X]G/ (Z[X])

av → 0. (2.2)

H. Blaine Lawson, Jr., Paulo Lima-Filho and Marie-Louise Michelsohn proved the following lemma in

[Lawson et al., 2003].

Lemma 2.10. 1. Let X be a compact topological C2-space. Then, there is a natural homeomorphism

Z[X]av ∼= Z[X/C2], where the latter denotes the free abelian group on X/C2.

2. Sequence (2.2) is a fiber sequence.

Moreover, there is a natural homeomorphism,

Z[X]G/Z[X]av ∼= Z2[XG].

The main idea used in the proof of theorem (2.9) is that homotopy groups of the G-fixed points of Z[X]

compute Bredon homology. See [Lima-Filho, 1997] for further details.

Corollary 2.11. Let X be a G-space with a free G-action. Then, Bredon homology reduces to non-

equivariant homology of the orbit space. More concretely, we have an isomorphism,

HGi (X;Z) ∼= Hi(X/G;Z).

Proof. Immediate from theorem (2.9) since XG = ∅.

Example 2.12. Consider the n-sphere in Rn+1 with the action given by the antipodal map, S(Rn+1,n+1).

Then,

HC2i (S(Rn+1,n+1);Z) ∼= Hi(RPn;Z) ∼=

Z if i = 0 or i = n and n is odd,

Z2 if 0 < i < n and i odd,

0 otherwise.

Corollary 2.12. Let n > 0. Then,

HC2i (Sn,n;Z) ∼=

Z if i = n and n is even,

Z2 if 0 ≤ i < n and i even,

0 otherwise.

Proof. We use the reduced version of the long exact sequence from theorem (2.9) with X = Sn,n and

get,

0→ Hi(Sn,n/G;Z)→ HG

i (Sn,n;Z)→ Hi((Sn,n)

G;Z2)→ Hi−1(Sn,n/G;Z)→ · · · .

Let i ≥ 0. The first map Hi(Sn,n/G;Z) → HG

i (Sn,n;Z) is injective because G = C2 only fixes two

19

Page 34: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

points of Sn,n and, so, Hi+1((Sn,n)G

;Z2) ∼= 0. Moreover, we have that

Hi((Sn,n)

G;Z2) ∼= Hi(pt

∐pt;Z2) ∼=

Z2 i = 0,

0 otherwise.

Therefore there are isomorphisms,

HGi (Sn,n;Z) ∼= Hi(S

n,n/G;Z),

for every i ≥ 1. On the other hand, we have isomorphisms,

Hi(Sn,n/G;Z) ∼= Hi(ΣRPn−1;Z) ∼=

Z if i = n and n is even,

Z2 if 1 < i < n and i is even,

0 otherwise.

It follows that

HG0 (Sn,n;Z) ∼= H0((Sn,n)

G;Z2) ∼= Z2.

which altogether gives us the result.

2.3 Spectra

2.3.1 Non-equivariant spectra

We define a prespectrum E to be a collection of based topological spaces En, for n ∈ N, together

with structure maps,

σ : ΣEn → En+1, ∀n ∈ N.

Call En the nth level of E. A map of prespectra , f : E → E′, is a sequence of maps fn : En → E′n

that commute with the structure maps, i.e., such that the diagram

ΣEn En+1

ΣE′n E′n+1

Σfn fn

commutes. This forms the category of prespectra, denoted by P.

A spectrum is a prespectrum such that the adjoint maps

σ : En → ΩEn+1

are homeomorphisms. A map of spectra f : E → E′ is a map between E and E′ regarded as prespectra.

We obtain the category of spectra which we denote by S.

20

Page 35: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Example 2.13. Let X be a based space. The suspension prespectrum of X consists of spaces

Xn = ΣnX together with the identifications σ : Σ(ΣnX) → Σn+1X. If X = S0 this is called the sphere

prespectrum and is denoted by S.

Example 2.14. Let G be an abelian group. The Eilenberg-MacLane prespectrum is defined by putting

Xn = K(G,n), where K(G,n) is the usual Eilenberg-MacLane space with homotopy concentrated in

degree n. It can be proven that K(G,n) and ΩK(G,n+ 1) are weakly equivalent and it follows that there

exists an Ω-spectrum HG such that (HG)n is a K(G,n) for all n.

2.3.2 Equivariant spectra

We want to extend the notions of spectrum and prespectrum to our equivariant world. One way to

do so, while exploiting all the structure arising from equivariance, is to index spectra on representation

spheres, which come already equipped with G-actions. More formally, we want to index in a G-universe.

A G-universe, U , is a countably infinite dimensional representation of G with an inner product such

that

1. U contains the trivial representation.

2. U contains countably many copies of each finite dimensional subrepresentation of iteself.

U is usually constructed as a direct sum of (Vi)∞ :=

⊕k≥1 Vi, where each Vi is an irreducible represen-

tation. A G-universe is trivial if it contains only the trivial representation and complete if it contains all

finite dimensional irreducible representations of G.

Given aG-universe U , let V be a finite dimensional subrepresentation of U . We define the suspension

and loop functors by

ΣV (−) = SV ∧ − and ΩV (−) = F (SV ,−).

Recall from example (2.1) that SV ∧X and F (X,Y ) denote the smash product with diagonal action and

the space of pointed maps X → Y with the action given by conjugation, respectively.

A G-prespectrum indexed on U is a collection of based G-spaces EV for each finite dimensional

subrepresentation of U together with basepoint-preserving G-maps,

σV,W : ΣW−V EV → EW ,

whenever V ⊂ W ⊂ U , where W − V denotes the orthogonal complement of V in W . We require that

σV,V = idV and the commutativity of the transitivity diagram,

ΣW−UEU ∼= ΣW−V ΣV−UEU ΣW−V EV

EW ,

ΣW−V σU,V

σU,W

σV,W

for U ⊂ V ⊂W ⊂ U .

21

Page 36: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

A G-spectrum indexed on U is a G-prespectrum indexed on U where the adjoint structure maps

σU,V : EU → ΩV−UEV

are homeomorphisms.

A map of G-spectra (or G-prespectra) , f : D → E, is a collection of maps, fV : DV → EV , such that,

ΣW−VDV DW

ΣW−V EV EW

ΣW−V fV

σV,W

fW

σV,W

commutes for every V ⊂ W ⊂ U . The category of G-spectra and the category of G-prespectra

indexed on U are denoted by GSU and GPU respectively.

A G-spectrum indexed on a trivial universe is called naive and one indexed in a complete universe is

called genuine. Notice that naive G-spectra are equivalent to ordinary spectra such that the spaces are

based G-spaces and the structure maps are equivariant. If, moreover, G is trivial, G-spectra, reduces to

non-equivariant spectra.

Example 2.15. Let X be a G-space and define EV = SV ∧X. Together with the obvious structure maps

these form the suspension prespectrum of X.

Example 2.16. For each constant ”coefficient system” M there is an Eilenberg-MacLane G-spectrum

denoted by HM . To each finite-dimensional subrepresentation V it associates an equivariant Eilenberg-

MacLane space that we define in section (2.4).

The forgetful functor U : GSU → GPU has a left adjoint,

L : GPU → GSU ,

called the spectrification functor. See [May, 1996] for further details.

A map of spectra f : D → E is said to be a G-spectra weak equivalence , or weak G-equivalence

if the context is clear, if for each finite dimensional subrepresentation V of U , fV : DV → EV is a weak

G-equivalence.

2.4 Mackey Functors

Let G be a finite group and let G− Set denote the category of finite G-sets and G-maps. A Mackey

functor for G is a pair of functors (M∗,M∗) from G− Set to Ab satisfying the following axioms:

1. M∗ is a contravariant functor and M∗ and is a covariant functor.

2. M∗(S) = M∗(S) for all G-sets S. For a morphism ϕ ∈ G − Set we use the notation M∗(ϕ) = ϕ∗

and M∗(ϕ) = ϕ∗ and these are called the restriction and transfer maps respectively.

22

Page 37: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

3. For every pullback square in G− Set

U S

T V

γ

δ

α

β

it is required that α∗ β∗ = δ∗ γ∗.

4. Both M∗ and M∗ take disjoint unions to direct sums.

In the case G = Cp, the description of Mackey functors over G can be significantly simplified and we

include it for convenience. This is due to the fact that the orbit category of G is very simple, as shown in

example (2.8).

Lemma 2.13 ([Shulman, 2010]). A Mackey functor M over G = Cp for any prime p consists of abelian

groups M(Cp) and M(e) together with the corresponding restriction and transfer maps such that

M(e) M(Cp)π∗

π∗M(Cp) M(Cp)

τ∗

τ∗

satisfy

1. Contravariant functoriality: (τ∗)p = id and τ∗π∗ = π∗.

2. Covariant functoriality: (τ∗)p = id and π∗τ∗ = π∗.

3. τ∗τ∗ = id.

4. π∗π∗x =∑p−1i=0 (τ i)∗x, for all x ∈M(Cp).

A Mackey functor M for C2 consists of abelian groups M(C2) and M(e) together with the corre-

sponding restriction and transfer maps satisfying the conditions from lemma (2.13). We can specify it

as

M(C2)

τ∗

M(e)π∗

π∗

Example 2.17. For G = C2, the constant coefficient functor, Z, extends to a Mackey functor,

Z

id

Z×2

id

This is a Mackey functor since π∗π∗x = 2x = ex+ τ∗x, for every x ∈M(C2).

Example 2.18. More generally, such a Mackey functor exists over any finite group G and any abelian

groupM in place of Z with the restriction maps being the identity maps and the transfer mapsM(G/H)→

M(G/K) given by multiplication by the index of H in K, [K : H]. This is called the constant Mackey

functor for G with value M and denote it by M .

23

Page 38: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Example 2.19. The Burnside ring Mackey functor B associates to each orbit space G/H the burnside

ring of H. For G = C2 it is described by the diagram,

Z

id

Z⊕ Zπ∗

π∗

where π∗(a, b) = a + 2b and π∗(a) = (0, a). It is easy to see that this functor satisfies the conditions on

lemma (2.13) and so it is a Mackey functor. See [Shulman, 2010] for more details on this functor.

Denote by MG the category of Mackey functors over the finite group G. The morphisms in this

category are the natural transformations.

2.5 The RO(G)-grading

So far we have only seen equivariant cohomology theories graded on the integers. Even non-

equivariantly, the usual Eilenberg-Steenrod axioms assume the cohomology functors to be graded on

Z. It happens, however, that equivariant cohomology theories are more naturally given by a collection of

functors indexed on the free abelian group on isomorphism classes of irreducible representations of a

fixed group G. We refer the reader to [Shulman, 2010] and to [May, 1996] for many examples of this. In

fact, if we understand G-spheres to be representation spheres, SV , then we must adapt the suspension

axiom to allow suspension by such spheres and this force us to grade on representations.

Such a theory is usually called RO(G)-graded even though the underlying abelian group of RO(G)

actually consists of equivalence classes of formal sums of representations.

Fix a group of equivariance, G, and let M be a contravariant coefficient system. It is a result

due to Gaunce Lewis, Peter May and James MacClure that the ordinary Z-graded cohomology the-

ory, H∗G(−;M), extends to an RO(G)-graded cohomology theory if and only if M extends to a Mackey

functor, or in other words, M is the underlying contravariant coefficient system of a Mackey functor.

Mackey functors are, thus, the right coefficient systems to be used in RO(G)-graded cohomology

theories. In fact, every Mackey functor has an associated RO(G)-graded cohomology theory, denoted

V 7→ HVG (−;M), where V runs over a complete G-universe. This cohomology theory is uniquely char-

acterized by

• Hn(G/H;M) =

M(G/H) if n = 0,

0 otherwise.

• The restriction maps H0(G/K;M)→ H0(G/H;M) induced by i : G/H → G/K are the restriction

maps i∗ in the Mackey functor M .

There are transfer maps H0(G/H;M)→ H0(G/K;M) and these coincide with the transfer maps of

the Mackey functor.

Let us be more precise and define axiomatically a cohomology theory graded on representations.

Recall that we are thinking of RO(G) not as equivalence classes of representations but as the free

24

Page 39: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

HαG(X;M) Hα+V

G (ΣVX;M)

Hα+WG (ΣWX;M) Hα+W

G (ΣVX;M)

ΣW

ΣV

Hid+f (id)

(f ∧ id)∗

HαG(X;M) Hα+V

G (ΣVX;M)

Hα+V+WG (ΣV+WX;M).

ΣV

ΣV+W

ΣW

Table 2.1: The bookkeping diagrams.

abelian group on isomorphism classes of irreducible representations of G. In other words, every α ∈

RO(G) is of the form α =∑aiVi, where ai are integers and Vi distinct irreducible representations. The

following definition is adapted from [Shulman, 2010].

An ordinary RO(G)-graded reduced cohomology theory consists of a collection of functors,

HαG(−;M) : hG− Top∗ → Ab satisfying the following axioms:

1. Weak Equivalence: if f : X → Y is weak G-equivalence then HαG(f) := f∗ : Hα

G(Y ;M) →

HαG(X;M) is an isomorphism.

2. Exactness: HαG(−;M) is exact on cofiber sequences. See sequence (2.3) below to see what this

means when G = C2.

3. Additivity: the inclusions ιi : Xi →∨iXi of based spaces induce an isomorphism, ι∗ : Hα

G(∨Xi;M)→∏

i HαG(Xi;M).

4. Suspension: for each α ∈ RO(G) and G-representation V , there is a natural isomorphism

ΣV : HαG(X;M)→ Hα+V

G (ΣVX;M),

covariant in V and contravariant in X, with Σ0 the identity natural transformation.

5. Dimension: HnG(G/H+;M) = 0 for every nonzero integer n and H0

G(G/H+;M) = M(G/H). The

restriction maps H0(G/K+;M) → H0(G/H+;M) induced by i : G/H → G/K are the maps i∗ in

the Mackey functor M .

6. Bookkeeping: Let f : SV → SW be the equivariant map induced by an orthogonal isomorphism

between representations V and W . Then the diagrams in table (2.1) commute.

Similar axioms characterize ordinary RO(G)-graded reduced Bredon homology , HGα (−;M). For

an unbased space X we define HαG(X;M) to be Hα

G(X+;M) and call it the ordinary RO(G)-graded

unreduced Bredon cohomology of X and similarly for homology. When the group of equivariance, G,

is clear we often write Hα(−;M) to mean HαG(−;M) and similarly for homology.

We set up some notation for our case of interest, when G = C2. Recall that we think of RO(G) as

formal sums of distinct irreducible representations with coefficients in Z. When G = C2, there are only

two such representations up to isomorphism, the trivial representation, 1, and the sign representation,

σ, both one dimensional. We can, therefore, make the identification RO(C2) ∼= 1 · Z ⊕ σ · Z ∼= Z ⊕ Z.

25

Page 40: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Let V = Rp+q,q be the C2-representation corresponding to (p, q). In this case, we use the following

convenient notation,

HV (X;M) = HRp+q,q

(X;M) = Hp+q,q(X;M),

and say that RO(C2)-graded cohomology is bi-graded. Moreover, let V = Rp+q,q and V ′ = Rp′+q′,q′

be two C2-representations. Then, V ⊕ V ′ ∼= Rp+q+p′+q′,q+q′ . Following this observation, if α = (r, s) ∈

RO(C2), we write Hα(X;M) = Hr,s(X;M) and Hα+V (X;M) = Hr+p+q,s+q(X;M).

When we say that RO(C2)-graded Bredon cohomology extends Z-graded Bredon cohomology we

mean that

Hp,0G (X;M) = Hp

G(X;M)

and analogously for homology.

Let

Σp+q,q(−) = Sp+q,q ∧ −.

This allows us to write the suspension axiom in the following form:

Suspension: For each α = (r, s) ∈ RO(C2) and C2-representation Rp+q,q, there is a natural isomor-

phism,

Σp+q,q : Hr,s(X;M)→ Hp+q+r,q+s(Σp+q,qX;M),

covariant in the representation and contravariant in X, with Σ0,0 the identity natural transformation.

The exactness in cofiber sequences can also be more explicitly written as:

Exactness: Consider the cofiber sequence A → X → X/A. For every fixed integer s, there is a long

exact sequence in bi-graded cohomology,

· · · → Hr,s(X/A;M)→ Hr,s(X;M)→ Hr,s(A;M)→ Hr+1,s(X/A;M)→ · · · . (2.3)

Define,

H∗,∗(X;M) =⊕

(r,s)∈Z×Z

Hr,s(X;M).

WhenM is a commutative ring, there is a cup product^: Hp,q(−;M)×Hp′,q′(−;M)→ Hp+p′,q+q′(−;M)

that endows H∗,∗(X;M) with a graded commutative ring structure. See [Dugger, 2005].

2.5.1 Representability

As in the non-equivariant case, the RO(G)-graded cohomology theories can be represented. Actu-

ally, G-spectra represent such theories. In this section we identify a model for the Eilenberg-MacLane

G-spectra and prove an equivariant version of Spanier-Whitehead duality. Most of the content of this

section is taken from [dos Santos, 2003].

Let M be a discrete abelian group, and (X,x0) a pointed G-space. We define M ⊗X as the set of

26

Page 41: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

finite formal sums

M ⊗X :=

∑i

mixi | mi ∈M, xi ∈ X

/ m · x0 | m ∈M .

This is a G-space with action given by (g,∑mixi) 7→

∑imigxi. Notice that Z ⊗X+ is the pointed free

abelian group generated by X, Z[X]+.

In [dos Santos, 2003], dos Santos generalizes the equivariant Dold-Thom theorem, proved by Lima-

Filho, to the setting of ordinary RO(G)-graded Bredon cohomology. This result will be often used in our

computations in the next chapter. Let α ∈ RO(G). An equivariant Eilenberg-MacLane space of type

(M,α), denoted by K(M,α), is a classifying space for the functor HαG(−;M).

Theorem 2.14 (dos Santos). Let X be a based G-CW-complex and let V be a finite-dimensional G-

representation. Then, there is a natural equivalence,

HGV (X;M) ∼=

[SV ,M ⊗X

]G.

We will call this result the Equivariant Dold-Thom Theorem. As a consequence we see thatM⊗SV

is a K(M,V ) space. As mentioned in [dos Santos, 2003], this allows us to identify a simple model for

the Eilenberg-MacLane spectrum, HM . In fact, since any K(M,V ) space is a classifying space for the

cohomology functor HVG (−;M), it follows that, under the same conditions as in the previous theorem,

Corollary 2.15. There is a natural isomorphism,

HVG (X;M) ∼=

[X,M ⊗ SV

]G.

Corollary 2.16. There is a natural isomorphism,

HVG (C2+ ∧X;M) ∼= H

dim(V )sing (X;M),

where the right hand-side group is non-equivariant cohomology with coefficients in the abelian group M .

Proof. Let Xe denote the the space X with trivial action. There is a natural C2-homeomorphism, C2+ ∧

X → C2+ ∧ Xe given by [e, x] 7→ [e, x] and [τ, x] 7→ [τ, τx], where τ is the generator of C2. We have a

chain of natural isomorphisms,

HVG (C2+ ∧X;M) ∼=

[C2+ ∧X,M ⊗ SV

]C2

Corollary (2.15)

∼=[C2+ ∧Xe,M ⊗ SV

]C2

∼=[Xe,M ⊗ Sdim(V )

]Proposition (2.6)

∼= Hdim(V )sing (Xe;M) Corollary (2.15).

In the next chapter we will need to use an equivariant version of the Spanier-Whitehead Duality that

we introduce here.

27

Page 42: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Corollary 2.17 (Spanier-Whitehead Duality for a point). There is a natural isomorphism,

Hr,s(pt;M) ∼= H−r,−s(pt;M).

Proof. Let Rp+q,q be a C2-representation such that r + p+ q ≥ s+ q ≥ 0. Then,

Hr,s(pt;M) ∼= Hr+p+q,s+q(Sp+q,q;M)

∼= [Sp+q,q,M ⊗ Sp+q+r,q+s]G Corollary (2.15)

∼= Hp+q,q(Sp+q+r,q+s;M) Theorem (2.14)

∼= H−r,−s(pt;M).

28

Page 43: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Chapter 3

Computation of the cohomology ring

of a point

3.1 RO(G)-graded cohomology of a point

Throughout this chapter we consider G = C2. We compute the ordinary RO(G)-graded cohomology

of a point for the constant Mackey functors M = R, Z and Z2. The idea is to reduce these computa-

tions to those of the (Z-graded) singular cohomology or homology of some space, for each underlying

coefficient group.

In order to do so, we heavily use the suspension axiom present in both equivariant cohomology and

homology theories. For certain degrees, the use of equivariant Spanier-Whitehead duality is critical in

order to arrive at singular homology. Applying these tools, together with suitable cofiber sequences

allows us to compute almost all of the group structure for each Mackey functor.

Furthermore, for any G-space X and constant Mackey functor M , where M is a commutative ring,

there is a cup product turning H∗,∗(X;M) into a graded commutative ring. This ring structure is also

explicitly computed for each of the Mackey functors mentioned above. The main ingredient here is the

use of an existing forgetful map from RO(G)-graded cohomology to singular cohomology giving ring

homomorphisms H∗,∗(X;M)→ H∗,∗(C2+ ∧X;M).

3.1.1 The group structure

Let M be an abelian group. Recall from example (2.18) that the constant Mackey functor is given by,

M : OC2−→ Ab

G/H 7−→M,

where M takes τ : C2 → C2 to identity maps τ∗ : M → M and τ∗ : M → M and takes the projection

π : C2 → pt to the identity map π∗ : M(pt)→M(C2) and π∗ : M(C2)→M(pt) to multiplication by 2.

29

Page 44: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

We start by computing the group structure of the ordinary RO(C2)-graded Bredon cohomology of a

point with the constant Mackey functor M , Hp,q(pt;M), for every p, q ∈ Z. We refer to H∗,q(pt;M) as

the positive cone if q runs over the non-negative integers and as the negative cone when q runs over the

negative integers.

If q = 0 the dimension axiom tells us that Hp,0(pt;M) = M if and only if p = 0.

Let q > 0. Then,

Hp−q,−q(pt;M) ∼= Hp−q,−q(S0,0;M)

∼= Hp,0(Sq,q;M)

∼= Hpsing(S

q,q/C2;M).

The last isomorphism holds since, for any pointed G-space X, H∗sing(X/G;M) is a Bredon cohomol-

ogy theory graded on the integers and, therefore, it is completely determined by the dimension axiom.

In fact we have that H∗sing((G/H)/G;M) ∼= H∗(G/H;M) ∼= M .

The space Sq,q is the unreduced suspension of the (q − 1)-sphere inside Rq,q, S(Rq,q). So it follows

that

Hpsing(S

q,q/C2;M) ∼= Hpsing(ΣRP q−1;M)

∼= Hp−1sing(RP

q−1;M),

and we conclude that the negative cone is the singular cohomology of the (q − 1)-dimensional real

projective space,

Hp−q,−q(pt;M) ∼= Hp−1sing(RP

q−1;M), q > 0.

We proceed with the calculation of the positive cone. This case is more difficult and, in particular, we

have to use equivariant Spanier-Whitehead duality for a point which is corollary (2.17) in the text.

Hp+q,+q(pt;M) ∼= Hp+q,+q(S0,0;M)

∼= H−p−q,−q(S0,0;M) Corollary (2.17)

∼= H−p,0(Sq,q;M).

By the exactness axiom for homology, the homology functor H∗,0(−;M) applied to the cofiber se-

quence S(Rq,q)+ → D(Rq,q)+ → Sq,q, which is an instance of sequence (2.1), yields a long exact

sequence,

· · · → H−p,0(S(Rq,q);M)→ H−p,0(D(Rq,q);M)→ H−p,0(Sq,q;M)→

→ H−p−1,0(S(Rq,q);M)→ H−p−1,0(D(Rq,q);M)→ · · · . (3.1)

The disc D(Rq,q) is weakly C2-equivalent to a point and by the weak equivalence axiom for homology,

30

Page 45: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

for p 6= 0 and p 6= −1, we have H−p,0(D(Rq,q);M) ∼= H−p−1,0(D(Rq,q);M) ∼= 0. Hence,

H−p,0(Sq,q;M) ∼= H−p−1,0(S(Rq,q);M), whenever p 6= 0 and p 6= −1.

Since S(Rq,q) has a free C2-action, then, by corollary (2.11),

H−p−1,0(S(Rq,q);M) ∼= Hsing−p−1(S(Rq,q)/C2;M) ∼= Hsing

−p−1(RP q−1;M),

and, finally,

Hp+q,q(pt;M) ∼= Hsing−p−1(RP q−1;M), whenever p 6= 0 and p 6= −1.

By now we are able to compute all the cohomology groups Hp,q(pt;M) in the positive cone except

when p = q or p = q − 1.

Notice that, in order to get the remaining results, one has to compute H0,0(Sq,q;M) ∼= Hq,q(pt;M)

and H1,0(Sq,q;M) ∼= Hq−1,q(pt;M). This follows from equivariant Spanier-Whitehead duality and the

suspension axiom.

Sequence (3.1) yields the exact sequence,

0→ H1,0(Sq,q;M)→ H0,0(S(Rq,q);M)→ H0,0(D(Rq,q);M)→ H0,0(Sq,q;M)→ 0 (3.2)

and, as we shall see, its analysis for each abelian group M suffices to get the remaining cases.

It follows from the equivariant Dold-Thom theorem that the mapH0,0(S(Rq,q);M)→ H0,0(D(Rq,q);M)

in sequence (3.2) is the map M ×2−−→M . In fact, π0(X⊗M) is the set of connected components of X⊗M

and it is easy to see that, given x ∈ S(Rq,q), the element x ⊗m can be connected by a path to 0 ⊗ 2m

with 0 ∈ D(Rq,q) the origin.

Thus, we get the exact sequence,

0→ Hq−1,q(pt;M)→M×2−−→M → Hq,q(pt;M)→ 0,

and we see that Hq−1,q(pt;M) and Hq,q(pt;M) are isomorphic to the kernel and cokernel of M ×2−−→M ,

respectively. More concretely,

Hq−1,q(pt;M) ∼= M2, if q > 0,

where M2 denotes de 2-torsion subgroup of M and,

Hq,q(pt;M) ∼= M/2M, if q > 0.

We have thus proved the following theorem, which summarizes the group structure of the bigraded

cohomology of a point for any constant Mackey functor M ,

Theorem 3.1. Let M be a constant Mackey functor. Then, there are group isomorphisms,

1. Hp,0(pt;M) ∼= M if p = 0 and is trivial otherwise.

31

Page 46: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

6 R54 R32 R10 R-1-2 R-3-4 R-5-6 R

-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6

Table 3.1: Group structure of H∗,∗(pt;R). The first index runs horizontally and the second index runsvertically.

2. Hp,q(pt;M) ∼= Hp−q−1sing (RP−q−1;M), for q < 0.

3. Hp,q(pt;M) ∼= Hsingq−p−1(RP q−1;M), for q > 0, q 6= p, p+ 1. The remaining cases are

Hp,p+1(pt;M) ∼= M2 and Hp,p(pt;M) ∼= M/2M,

where p ≥ 0 and p > 0, respectively.

We use theorem (3.1) to provide explicit computations of the cohomology groups Hp,q(pt;M) where

M = R, Z and Z2.

The group structure of Hp,q(pt;R).

Hp,0(pt;R) ∼= R if p = 0 and trivial otherwise.

For q < 0 we have, Hp,q(pt;R) ∼= Hp−q−1sing (RP−q−1;R) ∼=

R if p = 0 and q is even,

0 otherwise.

For q > 0, Hp,q(pt;R) ∼= Hsingq−p−1(RP q−1;R) ∼=

R if p = 0 and q is even,

0 otherwise,where p 6= q and p 6= q − 1.

Moreover, Hp,p+1(pt;R) = Hp,p(pt;R) = 0, for every p > 0. The results are gathered in table (3.1).

The group structure of Hp,q(pt;Z).

Hp,0(pt;Z) ∼= Z if p = 0 and trivial otherwise.

For q < 0,

Hp,q(pt;Z) ∼= Hp−q−1sing (RP−q−1;Z) ∼=

Z if p = 0 and q is even,

Z2 if q + 1 < p < 0 and p 6≡ q mod 2 or p = 0 and q < −1 odd,

0 otherwise,

32

Page 47: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

6 Z Z2 Z2 Z2

5 Z2 Z2 Z2

4 Z Z2 Z2

3 Z2 Z2

2 Z Z2

1 Z2

0 Z-1-2 Z-3 Z2

-4 Z2 Z-5 Z2 Z2

-6 Z2 Z2 Z-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6

Table 3.2: Group structure of H∗,∗(pt;Z). The first index runs horizontally and the second index runsvertically.

and, for q > 0,

Hp,q(pt;Z) ∼= Hsingq−p−1(RP q−1;Z) ∼=

Z if p = 0 and q is even,

Z2 if q − 1 > p > 0 and p ≡ q mod 2,

0 otherwise,

where p 6= q and p 6= q − 1.

The remaining cases are Hp,p+1(pt;Z) = 0 and Hp,p(pt;Z) ∼= Z2 for p > 0. The results are gathered

in table (3.2).

The group structure of Hp,q(pt;Z2).

Hp,0(pt;Z2) ∼= Z2 if p = 0 and trivial otherwise.

In this case, for q < 0, Hp,q(pt;Z2) ∼= Hp−q−1sing (RP−q−1;Z2) ∼=

Z2 if q + 2 ≤ p ≤ 0,

0 otherwise,

and, for q > 0, Hp,q(pt;Z2) ∼= Hsingq−p−1(RP q−1;Z2) ∼=

Z2 if q − 2 ≥ p ≥ 0,

0 otherwise,where p 6= q and

p 6= q − 1.

Moreover, Hp,p+1(pt;Z2) ∼= Hp,p(pt;Z2) ∼= Z2 for p > 0. The results are gathered in table (3.3).

3.1.2 The forgetful map to singular cohomology

In this section we introduce an important map that allows us, often, to relate RO(C2)-graded coho-

mology with singular cohomology. We will make extensive use of this map when computing the ring

structure of the cohomology of a point for different constant Mackey functors.

Let Rp+q,q be a C2-representation and let X be a pointed C2-space. Recall from corollary (2.16) that

there are natural isomorphisms,

Hp+q,q(C2+ ∧X;M)∼=−→ Hp+q(X;M). (3.3)

33

Page 48: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

6 Z2 Z2 Z2 Z2 Z2 Z2 Z2

5 Z2 Z2 Z2 Z2 Z2 Z2

4 Z2 Z2 Z2 Z2 Z2

3 Z2 Z2 Z2 Z2

2 Z2 Z2 Z2

1 Z2 Z2

0 Z2

-1-2 Z2

-3 Z2 Z2

-4 Z2 Z2 Z2

-5 Z2 Z2 Z2 Z2

-6 Z2 Z2 Z2 Z2 Z2

-6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6

Table 3.3: Group structure of H∗,∗(pt;Z2). The first index runs horizontally and the second index runsvertically.

Using the suspension axiom in ordinary RO(C2)-graded Bredon cohomology, it is possible to extend the

natural isomorphism (3.3) for any integers r and s. More concretely, there are natural isomorphisms,

Hr,s(C2+ ∧X;M)∼=−→ Hr(X;M)

Therefore, we define the map

ϕ : Hr,s(X;M)→ Hr,s(C2+ ∧X;M),

induced by the folding map ∇ : C2+∧X → X, which restricts to the identity in X, and call it the forgetful

map to singular cohomology.

Consider the usual cofiber sequence, S(Rq,q)+ → D(Rq,q)+ → Sq,q. When q = 1 we have the cofiber

sequence C2+ ∧ S0,0 ∇−→ S0,0 → S1,1.

Applying H∗,q(−;M) we get the exact sequence in cohomology,

0→ H0,q(S1,1;M)→ H0,q(S0,0;M)ϕ−→ H0,q(C2+ ∧ S0,0;M)→ H1,q(S1,1;M)→ H1,q(S0,0;M)→ 0,

which after appropriate shiftings becomes,

0→ H−1,q−1(pt;M)→ H0,q(S0,0;M)ϕ−→ H0,q(C2+ ∧ S0,0;M)→ H0,q−1(pt;M)→ H1,q(pt;M)→ 0.

(3.4)

We use sequence (3.4) to compute a few examples of forgetful maps.

Example 3.1. We compute the forgetful map ϕ : H0,−2n(S0,0;M)→ H0,−2n(C2+ ∧S0,0;M), with n > 0.

When q = −2n, sequence (3.4) yields,

0→ H−1,−2n−1(pt;M)→ H0,−2n(S0,0;M)ϕ−→ H0,−2n(C2+ ∧ S0,0;M)→ H0,−2n−1(pt;M)→ 0.

Referring to tables (3.1), (3.2) and (3.3), and proposition (3.3) we can conclude the nature of the

34

Page 49: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

forgetful maps in each case. For M = R, we get that ϕ is an isomorphism. For the Mackey functor Z,

we have the short exact sequence,

0→ Z ϕ−→ Z→ Z2 → 0,

and ϕ is the inclusion 2Z ⊂ Z. Finally, for Z2, we have

0→ Z2 → Z2ϕ−→ Z2 → Z2 → 0.

In this case, ϕ is the zero map. Actually, we will see later that, for q < 0, ϕ : H0,q(S0,0;Z2)→ H0,q(C2+ ∧

S0,0;Z2) is always the zero map as a consequence of the multiplicative structure of the forgetful map.

Example 3.2. We compute the forgetful map ϕ : H0,2n(S0,0;M)→ H0,2n(C2+ ∧ S0,0;M) with n ≥ 0.

When q = 2n, sequence (3.4) yields,

0→ H−1,2n−1(pt;M)→ H0,2n(S0,0;M)ϕ−→ H0,2n(C2+∧S0,0;M)→ H0,2n−1(pt;M)→ H1,2n(pt;M)→ 0.

Again, we just have to refer to tables (3.1), (3.2) and (3.3) and proposition (3.3) to understand what the

forgetful maps ϕ : H0,2n(S0,0;M)→ H0,2n(C2+ ∧ S0,0;M) are for each Mackey functor. For M = R, we

get that ϕ is an isomorphism and similarly for the Mackey functor Z, and, so, ϕ : Z→ Z is multiplication

by either 1 or −1. For the case of Z2, we get

0→ Z2ϕ−→ Z2 → · · · ,

and we conclude that the forgetful map is the identify in Z2.

Example 3.3. The argument in example (3.2) allows us to conclude, more generally, that ϕ : H0,n(S0,0;Z2)→

H0,n(C2+ ∧ S0,0;Z2) is an isomorphism for n ≥ 0.

Proposition 3.2. The forgetful maps give ring homomorphisms,

ϕ : H∗,∗(X;M)→ H∗,∗(C2+ ∧X;M).

Proof. The forgetful map in cohomology is induced by a continuous map of G-spaces ∇ : C2+∧X → X.

The result follows from functoriality of the ring structure.

3.1.3 The multiplicative structure

Recall that, if M is a commutative ring, there is a cup product, ^: Hr,s(−;M) × Hr′,s′(−;M) →

Hr+r′,s+s′(−;M), where M is the constant Mackey functor, that endows H∗,∗(X;M) with a graded

commutative ring structure. In fact, consider the map,

µ : (M ⊗X) ∧ (M ⊗ Y )→M ⊗ (X ∧ Y )

(∑i

mixi) ∧ (∑j

njyj) 7−→∑i,j

minj(xi ∧ yj).

35

Page 50: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Let V and W be C2-representations. Given representatives α : Z → M ⊗ SV and β : Z → M ⊗ SW

of HV (Z;M) and HW (Z;M), respectively, the composition,

Z∆−→ Z ∧ Z α∧β−−−→M ⊗ SV ∧M ⊗ SW µ−→M ⊗ SV⊕W ,

where ∆: Z → Z ∧ Z is the diagonal map, classifies the cup product between α ∈ HV (Z;M) and

β ∈ HW (Z;M). We henceforth refer to the cup product, ^, has multiplication.

In this section we compute the ring structure of H∗,∗(pt;M) when M is R, Z or Z2. We begin by

computing the ring structure of H∗,∗(C2;M) and use this computation to analyze the forgetful maps to

singular cohomology studied in examples (3.1), (3.2) and (3.3). These will be critical in understanding

the cohomology ring of a point.

Proposition 3.3. Let M be a ring. There is a ring isomorphism,

H∗,∗(C2;M) ∼= M[u, u−1

],

where u has bi-degree (0, 1).

Proof. For all non-negative integers n, we have isomorphisms,

Hp,q(C2) ∼= Hp,q(C2+) ∼= Hp+n,q+n(Σn,nC2+).

Recall from the proof of corollary (2.16) that there are C2-homeomorphisms, Σn,nC2+∼= Σn,0C2+ which

are multiplication by τ ∈ C2,

Hp+n,q+n(Σn,nC2+) ∼= Hp+n,q+n(Σn,0C2+) ∼= Hp,q+n(C2+) ∼= Hp,q+n(C2).

We conclude that Hp,q(C2) does not depend on the weight of the representation but only on its the

topological dimension and so Hp,q+n(C2) ∼= M for p = 0 and is trivial otherwise.

Let α : Sk,0 ∧ C2+ → M ⊗ Sk,k and β : Sk′,0 ∧ C2+ → M ⊗ Sk′,k′ be generators of H0,k(C2;M) and

H0,k′(C2;M), respectively, where k and k′ are non-negative integers. By corollary (2.16), the equivariant

map α is determined by the inclusion Sk,0 →M ⊗ Sk,k such that (x, e) 7→ 1x and similarly for β.

The product is given by composition µ (α ∧ β) (id ∧∆): Sk+k′,0 ∧ C2+ → M ⊗ Sk+k′,k+k′ which

sends (x ∧ y) ∧ e 7→ 1(x ∧ y). This map represents a generator of H0,k+k′(C2;M).

Using the suspension axioms we can easily generalize to the case where k and k′ are any integers

and the result follows.

The computation of the ring structure for a point is more complicated in general and uses that of G.

Theorem 3.4. There is a ring isomorphism,

H∗,∗(pt;R) ∼= R[v, v−1

],

36

Page 51: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

where v ∈ H∗,∗(C2;R) has bidegree (0, 2).

Proof. Recall the exact sequence (3.4),

0→ H−1,q−1(pt;M)→ H0,q(S0,0;M)ϕ−→ H0,q(C2+ ∧ S0,0;M)→ H0,q−1(pt;M)→ H1,q(pt;M)→ 0.

From the analysis of table (3.1) we know that, for all integers n, H−1,2n−1(pt;R) = H0,2n−1(pt;R) = 0.

Therefore, the forgetful map is an isomorphism of abelian groups, ϕ : H0,2n(S0,0;R) → H0,2n(C2+ ∧

S0,0;R).

But ϕ : H∗,∗(S0,0;R) → H∗,∗(C2+ ∧ S0,0;R) is a ring map and so it is a ring isomorphism onto its

image. Therefore,

H0,2∗(pt;R) ∼= R[v, v−1

].

Since all the remaining structure is trivial the result follows.

This result can be obtained in a different way using the forgetful map more explicitly. We include it

here as an example.

Example 3.4. For the case M = R and X = pt, there are forgetful maps ϕ : H0,2q(S0,0)→ H0,2q(C2+ ∧

S0,0) for every integer q. Let v be a generator of H0,2(pt), that is, a non-zero real number. Then v2 is a

generator of H0,4(pt) since ϕ(v2) = ϕ(v) ^ ϕ(v) which is nonzero because ϕ(v) is nonzero by example

(3.2). In the same way, v−2 is a generator of H0,−4(pt) if v−1 generates H0,−2(pt) and so H∗,∗(pt) is a

finitely generated ring, which is commutative because its groups are located in even dimensions. Finally,

v ^ v−1 is also nonzero because ϕ(v ^ v−1) = ϕ(v) ^ ϕ(v−1) 6= 0.

Let X and Y be pointed G-CW-complexes and M a discrete abelian group. Define the maps,

ηX : X → Z⊗X σX,Y : (Z⊗X) ∧ (M ⊗ Y )→M ⊗ (X ∧ Y )

x 7→ 1x, (∑i zixi) ∧ (

∑jmjyj) 7→

∑i,j zimj(xi ∧ yj).

The correspondence V 7→ M ⊗ ΣVX, where V runs over an indexing set of a complete G-universe U ,

defines a G-prespectrum, whose structural maps are given by the compositions,

SW ∧ (M ⊗ ΣVX) (Z⊗ SW ) ∧ (M ⊗ ΣVX) M ⊗ ΣV+WX

x ∧ (∑imixi) 1x ∧ (

∑imixi)

∑imi(xi ∧ x)

ηSW ∧idM⊗ΣV XσSW ,ΣV X

The associated G-spectrum is denoted by M ⊗ Σ∞X except when X = S0, in which case we write

M ⊗ S.

In what follows we will also make use of the natural map, N : X ∧ F (W,Y ) → F (W,X ∧ Y ) defined

such that N(x ∧ ϕ)(w) = x ∧ ϕ(w).

Recall that HM is the Eilenberg-MacLane G-spectrum that represents reduced RO(G)-graded Bre-

don cohomology with coefficients in M . Let k > 0. We denote by Σ0,−kHM the shifted Eilenberg-

MacLane spectrum so that (Σ0,−kHM)p,q := HMp,q−k for any integers p and q such that q − k ≥ 0.

37

Page 52: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Lemma 3.5. Let M be a discrete ring. Let f : Sk,0 → M ⊗ Sk,k be a representative of an element α in

H0,k(pt;M) where k > 0 and and f its adjoint. Then, the map

h : Σ0,−kHM → HM

given for each representation Rp,q by the composition,

M⊗Sp,q−k id∧f−−−→M⊗Sp,q−k∧Ωk,0(M⊗Sk,k)N−→ Ωk,0(M⊗Sp,q−k∧M⊗Sk,k)

Ωk,0µ−−−−→ Ωk,0(M⊗Sp+k,q),

is a map of spectra and it represents multiplication by f in H∗,∗(pt;M).

Proof. Fix a C2-universe U . We have to prove that h commutes with the structure maps of each spectra,

that is, that the diagram,

ΣW−V (Σ0,−kHM)V (Σ0,−kHM)W

ΣW−VHMV HMW

ΣW−V hV

σuV,W

hW

σlV,W

commutes for every V ⊂W ⊂ U .

Let V = Rp′,q′ and W = Rp,q be such that q′ − k ≥ 0, p ≥ p′ and q ≥ q′ and let Σh = idSp−p′,q−q′ ∧ h.

The upper structure map is given by the composition

σuRp′,q′ ,Rp,q = σSp−p′,q−q′ ,Sp′,q′−k (ηSp−p′,q−q′ ∧ idM⊗Sp′,q′−k)

and the lower structure map is

σlRp′,q′ ,Rp,q = σSp−p′,q−q′ ,Sp′,q′ (ηSp−p′,q−q′ ∧ idM⊗Sp′+k,q′ ),

Let x∧∑imixi ∈ Sp−p

′,q−q′∧M⊗Sp′,q′−k. Then, σuRp′,q′ ,Rp,q (x∧∑imixi) =

∑imi(xi∧x) ∈M⊗Sp,q−k.

Let z ∈ Sk,0. Then,

hW

(∑i

mi(xi ∧ x)

)=(Ωk,0µ Ωk,0(id ∧ f)

)(z 7→

∑i

mi(xi ∧ x) ∧ z)

= Ωk,0µ(z 7→∑i

mi(xi ∧ x) ∧ f(z))

= Ωk,0µ(z 7→∑i

mi(xi ∧ x) ∧∑j

njzj) Setting f(z) =∑j

njzj .

= (z 7→∑i,j

minj(xi ∧ x ∧ zj))

38

Page 53: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

On the other hand,

ΣW−V hV (x ∧∑i

mixi) = (idSp−p′,q−q′ ∧ hV )(x ∧∑i

mixi)

= x ∧ hV (∑i

mixi)

= x ∧ (z 7→∑i,j

minj(xi ∧ zj)).

Finally, applying σlRp′,q′ ,Rp,q to x ∧ (z 7→∑i,jminj(xi ∧ zj)), we get z 7→

∑i,jminj(xi ∧ x ∧ zj) and

the diagram commutes.

Theorem 3.6. The ring structure of H∗,∗(pt;Z) is completely determined by the following properties:

1. It is commutative.

Let y, ρ and α generate H0,2(pt), H1,1(pt) and H0,−2(pt) respectively. Then,

2. For any generator κ ∈ Hp,q(pt), κ ^ ρ generates Hp+1,q+1(pt) unless Hp,q(pt) is trivial.

3. Multiplication by y, y ^ − : H0,q−2(pt)→ H0,q(pt) is an isomorphism except when q = 0 or q = −1.

4. y ^ α = 2.

Proof. 1. The commutativity of H∗,∗(pt;Z) comes from the fact that the groups consist either of Z2’s

or are in even degrees. In this case the graded commutativity becomes commutativity.

2. The inclusion S1,1 → Z ⊗ S1,1 is a weak C2-equivalence. In fact, by the equivariant Dold-Thom

theorem we know the homotopy groups of Z ⊗ S1,1 and of its C2-fixed points and they agree with

those of S1,1. Let ρ be the inclusion S0,0 → S1,1 '−→ Z ⊗ S1,1 and consider the cofiber sequence

S0,0 → S1,1 → C2+ ∧ S1,0. We get the exact diagram,

H∗,∗(C2+ ∧ S1,0)

H∗,∗(S1,1) H∗,∗(S0,0)ρ∗

δ(3.5)

where δ : Hr,∗(S0,0) → Hr+1,∗(C2+ ∧ S1,0) is the connecting homomorphism in cohomology. It

can, thus, be seen that ρ∗ : Hp,q(S1,1) → Hp,q(pt) is an isomorphism whenever Hp−1,q−1(pt) and

Hp,q(pt) are both nontrivial and p 6= 1 and p 6= 0. We will see that, for the remaining cases,

multiplication by ρ may not be an isomorphism but is, at least, surjective and that is enough to

conclude that it takes generators to generators. In fact, a portion of the exact triangle (3.5) can be

translated into the exact sequence for every integer q,

0→ H0,q(S1,1)ρ∗−→ H0,q(S0,0)→ H1,q(C2+ ∧ S1,0)→ H1,q(S1,1)

ρ∗−→ H1,q(S0,0)→ 0.

Referring to table (3.2), we conclude that the maps ρ∗ : H−1,q−1(pt) → H0,q(pt) are isomor-

phisms for all q of the form q = −2r − 3, where r is a non-negative integer and that the maps

39

Page 54: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

ρ∗ : H0,q−1(pt) → H1,q are surjective for all q of the form q = 2r + 1, where r is a non-negative

integer and the result follows. The map ρ∗ is multiplication by ρ in cohomology. In fact, it sends the

identity representative of H∗,∗(S1,1) in ρ and it is a map between free modules over the cohomology

of a point.

3. By example (3.2), ϕ : H0,2q(S0,0)→ H0,2q(C2+ ∧ S0,0) is an isomorphism. By proposition (3.3), yn

generates H0,2n(pt) for every n > 0.

For the negative cone we verify that H0,−n−2(pt)^y−−→ H0,−n(pt) is an isomorphism for n ≥ 2. The

following technique is used in [Dugger, 2005].

Let E be the G-spectrum defined by the cofiber sequence,

Σ0,−2HZ y^−−→ HZ→ E.

The element y can be represented by a class f : S2,0 → Z⊗ S2,2.

By lemma (3.5) we know that multiplication by y is a map of spectra. Thus, for every pointed

G-CW-complex X we have a long exact sequence,

· · · → Hp,q−2(X)→ Hp,q(X)→ Ep,q(X)→ Hp−1,q−2(X)→ · · · . (3.6)

Then, analyzing sequence (3.6) and table (3.2) we get En,0(pt) ∼= Z2 for n = 0 and 0 otherwise.

Also, En,0(C2) = 0 for every n. So we conclude that the Z-graded equivariant cohomology E∗,0

is the Z-graded cohomology theory corresponding to the Mackey functor E0,0(−). It follows that

En,0(X) = Hn(XC2 ;Z2). Therefore, for n > 0, E0,−n(pt) ∼= E0,−n(S0,0) ∼= En,0(Sn,n) ∼= Hn(S0) ∼=

0. In the same way, E1,−n(pt) is trivial and the isomorphism H0,−n−2(pt)^y−−→ H0,−n(pt) follows

for n ≥ 2.

4. From example (3.1), we know that the forgetful map ϕ : H0,−2(S0,0) → H0,−2(C2+ ∧ S0,0) is

multiplication by 2. Consider the composition,

ϕ (^ y) : H0,−2(S0,0)→ H0,0(C2+ ∧ S0,0)

α 7→ ϕ(α ^ y)

We have that ϕ is an isormophism in degree (0, 0) by example (3.2). On the other hand, given that

ϕ is multiplicative, we have that ϕ(α ^ y) = ϕ(α) ^ ϕ(y) is two times a generator by proposition

(3.3).

Theorem 3.7. The ring structure of H∗,∗(pt;Z2) is completely determined by the following properties:

1. It is commutative.

Let τ , ρ and θ be generators of H0,1(pt), H1,1(pt) and H0,−2 respectively. Then,

40

Page 55: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

2. For any generator κ ∈ Hp,q(pt), κ ^ ρ generates Hp+1,q+1(pt), except when this group is trivial.

3. Multiplication by τ , τ ^ : H0,q−1(pt) → H0,q(pt) is an isomorphism except when q = 0 or q = −1

for dimensional reasons.

4. θ2 = 0.

Proof. 1. H∗,∗(pt;Z2) is commutative because it consists of Z2’s only.

2. Following the proof of point 2 of the last theorem, let ρ be the inclusion S0,0 → S1,1 → Z2 ⊗ S1,1.

Recall that we have a diagram in cohomology,

H∗,∗(C2+ ∧ S1,0)

H∗,∗(S1,1) H∗,∗(S0,0)ρ∗

δ(3.7)

where δ : Hr,∗(S0,0) → Hr+1,∗(C2+ ∧ S1,0) is the connecting homomorphism in cohomology. The

lower map ρ∗ assigns to S1,1 → Z2 ⊗ S1,1 the restriction S0,0 → S1,1 → Z2 ⊗ S1,1. This is a

map between free modules over the cohomology of a point, H∗,∗(pt), and, hence, it is multiplica-

tion by ρ. From the exactness of diagram (3.7), we conclude that ρ∗ : Hp−1,q−1(pt) → Hp,q(pt)

is an isomorphism whenever p 6= 1 and p 6= 0. The maps ρ∗ : H−1,q−1(pt) → H0,q(pt) and

ρ∗ : H0,q−1(pt) → H1,q(pt) are also isomorphisms in some cases. In fact, for every integer q

there is an exact sequence,

0→ H0,q(S1,1)ρ∗−→ H0,q(S0,0)→ H1,q(C2+ ∧ S1,0)→ H1,q(S1,1)

ρ∗−→ H1,q(S0,0)→ 0,

coming from exactness of diagram (3.7) as in theorem (3.6). Then, referring to table (3.3), we con-

clude that ρ∗ : H−1,q−1(pt)→ H0,q(pt) is an isomorphism whenever q ≤ −2 and ρ∗ : H0,q−1(pt)→

H1,q(pt) is an isomorphism whenever q ≥ 1.

3. The forgetful map gives a ring homomorphism ϕ : H∗,∗(S0,0) → H∗,∗(C2+ ∧ S0,0) ∼= Z2 in each

degree. Therefore, since ϕ(τ) is nonzero by example (3.3), ϕ(τ2) = ϕ(τ) ^ ϕ(τ) is nonzero.

So, by proposition (3.3), τ2 is nonzero and similarly for τn, for n > 2, that is, τn is a generator of

H0,n(pt). The multiplicative structure on the positive cone follows from this and the previous item.

To get the multiplicative structure on the negative cone we use the same approach as used in the

previous theorem. Let E be the C2-spectrum defined by the cofiber sequence,

Σ0,−1HZ2τ^−−→ HZ2 → E.

By lemma (3.5), the map τ ^ is a map of C2-spectra. Thus, for every pointed C2-CW-complex X

we get a long exact sequence,

· · · → Hp,q−1(X)→ Hp,q(X)→ Ep,q(X)→ Hp−1,q−1(X)→ · · · .

41

Page 56: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

It is easy to see that En,0(pt) ∼= Z2 if n = 0 and is 0 otherwise, as well as En,0(C2) = 0 for every n.

So, just as in the last theorem we get that En,0(X) = Hn(XC2 ;Z2). Since E1,−n(pt) = E0,−n(pt) =

0 we have that multiplication by τ is an isomorphism H0,−n−1(pt)→ H0,−n(pt) whenever n ≥ 2.

4. Suppose θ2 is the generator of H0,−4(pt). We know that multiplication by τ2 takes generators to

generators and, hence, we would have θ2 ^ τ2 = θ. However, we arrive at a contradiction since

θ ^ τ = 0 by degree reasons.

42

Page 57: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Bibliography

S. Awodey. Category Theory. Oxford Logic Guides, 2nd edition, 2010.

G. E. Bredon. Equivariant Cohomology Theories. Springer, 1967.

P. F. dos Santos. A note on the equivariant Dold-Thom theorem. J. Pure Appl. Algebra, 183(1-3):

299–312, 2003. ISSN 0022-4049.

D. Dugger. An Atiyah-Hirzebruch spectral sequence for KR-theory. K-theory, 2005.

H. B. Lawson, P. L.-F. Jr., and M.-L. Michelsohn. Algebraic cycles and the classical groups part I, real

cycles. Topology, 2003.

P. Lima-Filho. On The Equivariant Homotopy Of Free Abelian Groups On G-Spaces And G-Spectra.

Mathematische Zeitschrift, 1997.

P. Masulli. Equivariant homotopy: KR-theory. Master’s thesis, University of Copenhagen, 2011.

J. P. May. Equivariant Homotopy and Cohomology Theory, volume 91. CBMS, 2nd edition, 1996.

M. E. Shulman. Equivariant local coefficients and the RO(G)-graded cohomology of classifying spaces.

PhD thesis, University of Chicago, June 2010.

T. tom Dieck. Transformation Groups. De Gruyter Studies in mathematics ; 8, 1987.

43

Page 58: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

44

Page 59: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

Index

G-

CW-complex, 12

homeomorphism, 4

homotopy, 10

based, 11

map, 4

space, 4

based, 4

coinduced, 8

induced, 8

universe, 21

complete, 21

trivial, 21

action, 4

free, 5

trivial, 5

averaged cycle, 18

cellular chain, 15

coefficient system, 14

cohomology

RO(G)-graded, 25

Z-graded, 15

reduced

RO(G)-graded, 25

Z-graded, 18

cup product, 26

equivariant

n-cell, 12

Dold-Thom Theorem, 27

Eilenberg-MacLane space, 27

homotopy group, 11

Spanier-Whitehead Duality, 28

fixed point set, 7

functor

adjoint, 9

fixed point, 7

forgetful, 9

Mackey, 22

Burnside ring, 24

constant, 23

homology

RO(G)-graded, 25

Z-graded, 16

reduced

RO(G)-graded, 25

Z-graded, 18

invariant subspace, 6

map

folding, 34

forgetful, 34

restriction, 22

transfer, 22

orbit

category, 13

space, 5

prespectrum, 20

G-, 21

category, 20

Eilenberg-MacLane, 21

map, 20, 22

sphere, 21

suspension, 21, 22

45

Page 60: The Ordinary RO C -Graded Bredon Cohomology of a Point · Eilenberg-Steenrod axioms for cohomology to account for equivariance coming from the action of a finite group G. This is

representation, 5

completely reducible, 6

dimension, 6

faithful, 6

irreducible, 6

isomorphic, 6

orthogonal, 6

sphere, 7

smash product, 4

spectrum, 20

G-, 22

category, 20

genuine, 22

map, 22

naive, 22

subrepresentation, 6

proper, 6

topological dimension, 7

weak G-equivalence, 11

G-spectra, 22

weight, 7

Weyl group, 7

46