You are on page 1of 142

An Introduction to Algebraic Topology

Michael Alder

HeavenForBooks.com

An Introduction to Algebraic Topology


by

Michael D Alder

HeavenForBooks.com

This Edition Michael D. Alder, 2001

Warning: This edition is not to be copied, transmitted, excerpted or printed except on terms authorised by the publisher

HeavenForBooks.com

Contents

1 Preliminaries 1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1.1 1.1.2 1.1.3 1.1.4 What is Topology? . . . . . . . . . . . . . . . . . . . .

9 9 9

Why do we care? . . . . . . . . . . . . . . . . . . . . . 12 Invariants . . . . . . . . . . . . . . . . . . . . . . . . . 16 Summary . . . . . . . . . . . . . . . . . . . . . . . . . 19

1.2 Back to Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 1.2.1 1.2.2 1.2.3 1.2.4 1.2.5 Metric Spaces . . . . . . . . . . . . . . . . . . . . . . . 20 Topological Spaces . . . . . . . . . . . . . . . . . . . . 25 Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 Manifolds . . . . . . . . . . . . . . . . . . . . . . . . . 45 Odd . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

1.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 2 Categories and Functors and

53

CONTENTS 2.1 Categories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 2.2 The Fundamental Group Functor . . . . . . . . . . . . . . . . 58 2.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 3 Singular Homology 75

3.1 The Homology Groups . . . . . . . . . . . . . . . . . . . . . . 75 3.2 Hn is a functor . . . . . . . . . . . . . . . . . . . . . . . . . . 85 3.3 Homotopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 3.4 Homology of Pairs . . . . . . . . . . . . . . . . . . . . . . . . 96 3.5 Excision . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 3.5.1 3.5.2 3.5.3 Subdivision . . . . . . . . . . . . . . . . . . . . . . . . 112 Small Cubes . . . . . . . . . . . . . . . . . . . . . . . . 119 Excision at Last . . . . . . . . . . . . . . . . . . . . . . 122

3.6 The Mayer-Vietoris Theorem . . . . . . . . . . . . . . . . . . . 124 3.7 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 3.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138 Bibliography 141

HeavenForBooks.com

Introduction to Algebraic Topology

Preface
This book was used for the University of Western Australias Mathematics Departments Course in Algebraic Topology taken by Honours (Fourth) year students in 1996. Views expressed herein are those of the author, not those of the University nor the Department insofar as either can be said to have a view. The material covered represents a short (one semester, two lectures a week) and therefore minimalist sketch of Algebraic Topology. I had briey considered subtitling the book Algebraic Topology for Dummies but it would be unfair to dummies to let them nd out about Topology or indeed any other branch of Pure Mathematics. There are, after all, plenty of other subjects for them to pass their time with. The book was much inuenced by notes by Patrick Hew of lectures given in 1995 by Lyle Noakes, by Lyles own notes, and contain various borrowings, notably from Edwin Spaniers Algebraic Topology [1] and most extensively from William Masseys A basic Course in Algebraic Topology [2]. I also found that old favourite of mine, G.F. Simmons Introduction to Topology and Modern Analysis [3] useful for jogging my memory, after a rather long period of time. The treatment may also have been inuenced by CTC Walls course given over quarter of a century ago in Cambridge and Liverpool. This is a suitable place to thank Terry for giving me an inferiority complex which promises to last a lifetime1. The use of computers is having a profound eect on Mathematics and Science, turning parts of Pure Mathematics into an experimental science (it has always had an element of this, we all have to do sums, but it is vastly increased in recent times). There is an aspect however which has been made much more important, and that is qualitative theories. If you want to study the kinds of things that can happen, a computer simulation may be quite useless or positively misleading. If the orbits of a planet are elliptical and close
All human beings are ignorant and stupid, it is part of the design specication. The sooner they nd this out the better. I didnt nd out until relatively late, which is regrettable. I realise that this view is completely out of touch with modern educational thought, which believes in giving everybody a high self-esteem by never letting them do anything dicult. There are other good things about it, too.
HeavenForBooks.com
1

MICHAEL D. ALDER

to circular, a numerical approximation, as is produced by a computer, will usually produce orbits which spiral outwards. The ner the approximation, the smaller the eect and the tighter the spiral, but never, ever, do you get re-entrant curves. There is a rather considerable dierence between these and real orbits. So in order to use a computer intelligently (which is, admittedly, an uncommon and decidedly eccentric thing to do), the machine needs to be supplemented by some heavy thought about the nature of the possibilities. Do we really get outward spirals, or inward spirals (much more plausible physically) or do we get re-entrant curves? These qualitative considerations are quintessentially topological, and topology is essential to understanding them. As a geometer, by training and temperament, I want to be able to go from physical and geometric intuitions and make them respectable parts of mathematics. This means that a lot of topology consists of denitions, intended to articulate some intuitive idea. There are a lot of subtle complications which can arise in topology, and much experiment has gone into nding denitions which make your life easy, or at least as easy as possible. Unfortunately, the resulting denitions are not always obvious and intuitively natural to the beginner. So you have the choice of either starting o with natural and obvious denitions and then having a horrible mess later on, or taking some less than obvious denitions as a starting point, which take a while to get used to but save you eort in the longer run. The denition of a topological space is a case in point. Some trust in your teachers, although generally undesirable and often dicult to attain, is useful. Alternatively, a considerable quantity of patience is essential, because the construction of the heavy machinery needed to do the calculations takes a lot of preliminary work. I have tried to mitigate the awfulness of the abstraction by briey discussing motivational considerations, but it has to be faced that there is no Royal Road to Geometry. A mathematician has to be able to use both hemispheres of his brain in concert: he has to be able to think about his subject in three languages simultaneously, and to hop between them like a kangaroo on a pointed pogo stick. He must be able to simultaneously translate between Images, English and Algebra. The student who uses this book must beware of confusing the English or the pictorial parts with a new sort of Mathematics which is done by waving the arms. It is much harder than that. So although I may be chatty
HeavenForBooks.com

Introduction to Algebraic Topology

in part of the notes, I am stern and precise where it matters. Vagueness and sloppiness are useful as heuristics for developing the intuitions part of the way, but real mathematics culminates in a crystaline purity, seldom achieved but always aimed at. A look at any of the books on algebraic topology will make it clear that this course barely scratches the surface of the subject. Populist and commercial notions of what a University is for, have ensured that the preparation of students in these decadent times is such that more would be unattainable. Even so, it is, as warned, tough stu that will take all your eorts. On the other hand, the self-esteem of those who have mastered the subject is legendary. From the heights of algebraic topology, it is easy to look down on rather a lot. With which ambiguous comment, I close. The symbol denotes the end of a proof.

HeavenForBooks.com

MICHAEL D. ALDER

HeavenForBooks.com

Chapter 1 Preliminaries
1.1
1.1.1

Motivation
What is Topology?

Algebraic Topology takes problems of a more or less geometric sort, and translates them into problems of a clearly algebraic sort. The geometric sort needs some clarication. Everyone has heard that Topology is rubber sheet geometry, where we can deform the objects. In the kind of geometry one used to meet at school, one thought of a triangle as existing independent of its coordinates. Two triangles which diered by a shift or a rotation were, pretty much, the same triangle. We said they were congruent, and made no essential distinction between them. Any property of a rigid geometric object was shared by any congruent object. Such geometry can be thought of as a study of objects in Rn which are invariant under the Euclidean Group. Projective Geometry, not much now studied, extended the study to the Projective Group. Two objects are as near as dammit the same in Projective Geometry if one is a perspective view of the other from some angle. If ABC is a triangle in the plane, and abc is another triangle in a possibly dierent plane, both planes sitting in R3, so that the lines through Aa, Bb and Cc meet in a point P, then the two triangles look the same from the point P. 9

10

MICHAEL D. ALDER

Some of the theory of Projective Geometry has practical implications for recognising objects in images and computer graphics. In Topology, we go nearly all the way. The intuitive idea is, roughly, that we do not distinguish two objects which can be deformed into each other. We imagine a world in which objects are not rigid as in Euclidean geometry but are made of plasticene or chewing gum or something which can be deformed by stretching and compressing and twisting, but not tearing or gluing. We articulate this by saying that we deal with spaces which are regarded as equivalent if there is a homeomorphism between them: Denition 1.1.1 If U and V are topological spaces, f : U V is a homeomorphism i 1. it has an inverse map g : V U , such that f g is the the identity map on V and g f is the identity map on U . 2. Both f and g are continuous Thus the unit interval [0, 1] and that of double the length [0, 2] are obviously homeomorphic by the function y = 2x and its inverse x = y/2. For the present purposes, if you do not know what a topological space is, take it to be a metric space. If you do not know what a metric space is, you have problems, but I shall give denitions shortly. Exercise 1.1.1.1 Show that homeomorphism is an equivalence relation on any set of spaces. Show that a homeomorphism is 1-1 and onto. Hence the denition of a topologist as a man who cant tell a coee cup from a doughnut. This applies to the toroidal doughnut invented by Americans who
HeavenForBooks.com

Introduction to Algebraic Topology

11

cleverly charge us for the hole: if such a doughnut were made of plasticene, you could deform it progressively and continuously into a coee cup, the handle of the cup being the bit around the hole in the doughnut. It is plausible to the intuitions that you cannot however deform a solid ball into a doughnut. Just digging a hole through the ball would give a doughnut, but there would have to be some tearing to get the hole. This would make the requisite maps discontinuous. Similarly, if we take the real line, we can show that it is homeomorphic to the open interval (1, 1) by explicitly constructing maps, but it is not, one might feel intinctively, homeomorphic either to the closed interval [1, 1] or to R2 . These claims of what is intuitively plausible of course refer to ones intuitive feelings about what continuous maps can and cannot do. They need to be proved carefully, not just left to ones woolly feelings. Exercise 1.1.1.2 Construct an explicit map from (1, 1) to R which is a homeomorphism. Show that (1, 1) is not homeomorphic to [1, 1]. Note that if we relax the continuity conditions on the maps between objects, we have the condition that two objects are equivalent i they have the same cardinality, there are the same number of points in the two objects. All the above objects, the doughnut, the solid ball, R, (1, 1) and [1, 1] are equivalent in this sense. We are doing set theory if we take this approach, which is another and much less interesting subject. Exercise 1.1.1.3 Show that there is a bijection (1-1 and onto map) between R and R3. Show that there is a bijection between [1, 1] and (1, 1). Show that the exponential map exp : [0, 1) C given by exp(t) = cos(2t) + i sin(2t) is a continuous, 1-1 map from a half open interval onto the unit circle. Does this mean that the unit circle and the half open interval are homeomorphic?
HeavenForBooks.com

12

MICHAEL D. ALDER

1.1.2

Why do we care?

Why do we care about such things? Who gives a damn if spaces are homeomorphic or not? Why play with rubber sheet geometry? There are two distinct answers. The rst is that it is useful, it has applications. For example, it is important when one comes to study dierential equations on spheres and tori. This is perfectly intelligible at the intuitive level: You can imagine lots of little arrows dening a vector eld on a sphere or the surface of a doughnut, and you can probably credit that there is sense in having a smooth curve which is a solution to such a system. Moreover, these things can arise physically with no trouble at all: If we are studying vibrations in a crystal, we can take a unit cube of the crystal which is repeated regularly, and regard the whole crystal as, approximately, an innite replication of this cube. Thus we are looking at functions dened on the cube which are periodic. If we did this for a function on the real line which was periodic over [, ], we can treat this as a function dened on the circle. If it were a square in two dimensions, we would get a torus, and if a cube, we get a three dimensional version of a torus. Or consider functions or vector elds dened on R2 which have the property that they vanish at innitya condition common in Physical systems. Then we can extend the vector eld to one on a sphere in the way which you may have met in Complex Function Theory. Now suppose we perturb the vector eld a little bit. This may correspond to deforming the space it is on. Two vector elds may be deformed into each other continuously, or they may be of a dierent type. The qualitative type of a vector eld or ow on a sphere is determined, more or less, by the zeros of the vector eld and what type of zeros they are. It is the topological type of the space it is on which determines what kind of ows can exist. For example, it is easy to have a ow on a torus which has no xed points at all, but there is no such ow on a sphere. Deforming the sphere, taking any space homeomorphic to the standard sphere, does not change that statement. Denition 1.1.2 A ow on a space U is a continuous map v : R U U
HeavenForBooks.com

Introduction to Algebraic Topology with the following properties: u U, v (0, u) = u t R, v (t, ) : U U is a homeomorphism, and s, t R, u U, v (s, v (t, u)) = v (s + t, u)

13

(1.1) (1.2) (1.3)

Denition 1.1.3 A ow v on a space U has u U a xed point of v i t R, v (t, u) = u Exercise 1.1.2.1 Construct a ow on the unit circle which has no xed point. It might be simpler to construct a ow on R2 which takes the unit circle into itself so that the only xed point of the ow on R2 is the origin. Construct a ow on the torus which has no xed point. Construct a ow on the sphere which has precisely one xed point. So if we want to study qualitative properties of ows on such surfaces as spheres and tori, we need to know something of the properties of the underlying spaces which are topological properties. Since computers enable us to study numerically ows on all sorts of spaces, but since the qualitative features get thoroughly buggered up by such things as truncation and rounding errors, the qualitative theory of ows, also known as Dynamical Systems, is rather important. It can save you trusting your computer blindly, and help tackle issues where the computer behaves like a rather convoluted random number generator. Another example of an application of Topology which may appeal to some occurs in image analysis, when we want to nd straight lines in images. A favourite method used by Computer Scientists and Engineers is called the Hough Transform. What we do is to consider an image as a collection of points in R2, and then consider the space of all lines in the plane. This space is a new space as far as you are concerned, one of the Grassmannian Spaces. It is called Real Projective 2-Space, or RP2 . You can think of it as follows: Take the space of lines in the plane and write them as ax + by + c = 0. Now the triple (a, b, c) determines a line, which might lead you to the idea
HeavenForBooks.com

14

MICHAEL D. ALDER

Figure 1.1: Construction of Real Projective 2-Space that the space of lines in the plane is just R3 . This would be quite wrong, because if we take two such triples (a, b, c) and (A, B, C ) such that one is a non-zero multiple of the other, (A, B, C ) = (a, b, c) for some = 0, then they determine the same line. So we x up that a2 + b2 + c2 = 1. This might lead you to the belief that the space of lines in the plane is just the 2-sphere. This is still wrong. Opposite points on the 2-sphere correspond to the same line. So we throw away the southern hemisphere, c < 0. We still need to glue together opposite points on the equator. If we glue together one pair of opposite points, we get a peculiar looking bent disk with lips as in gure 1.1 The nal business of gluing together the two bounding circles is complicated by the fact that we cannot just stretch them around. The wrong points get glued. We have to twist one of the circles before gluing. This means that it is not possible to build the object in three dimensional space. It is, however, a perfectly respectable two dimensional surface, just like a sphere or a torus. Or a Klein Bottle. The somewhat woolly idea of gluing edges together sounds as though it needs to be made precise, but a woolly idea is often a good starting point1 .
This is still wrong as a representation of the lines in the plane. It contains a line (0, 0, 1), the North pole of the hemisphere. This, does not correspond to any actual line in the plane, the equation 0x + 0y + 1 = 0 is not the equation of any line unless you think the empty set is a line. If you look to see what happens as a, b get very close to zero and c compensates, you discover that we are getting lines which are further and further away
HeavenForBooks.com
1

Introduction to Algebraic Topology The following is an excerpt from a paper of mine on image analysis: The Hough Transform

15

It is possible to obtain a maximum likelihood estimate from a set of points drawn from a line or lines in R2 , by turning to the space, M, of all such lines in R2 . It is usual to adjoin to the space of all lines in the plane a line at innity, which produces the space RP2 , real projective 2-space. This may be regarded as unrealistic, since the line at innity is not going to appear in any image, but neither are lines suciently remote, since images are bounded. We may more plausibly remove a contractible neighbourhood of the line at innity and obtain the space M which is homeomorphic to RP2 with a disk removed from it. By parametrising this space, homeomorphic to a M obius strip, and observing that a point in the original image (by which is meant the compact rectangle in R2 in which our line is sought) generates a curve in the space of lines, the line in the image space is transformed into a set of curves intersecting in the point which parametrises the given line. If the points of the image are subject to noise, then the transform of them will yield a cluster of points the centroid of which may be construed as a maximum likelihood estimate of the parametrisation of the underlying line. This is generally known in the image processing literature as the Hough Transform. It transforms the problem of nding a line or lines in the image space into the problem of nding a cluster or clusters, or maxima of a density function, in the line space. Engineers try to parametrise the space of lines in various ways; if you were to write each line as y = mx + c, then the m c space is the transform space. This has a problem, namely that you miss out on the vertical lines. You can get around this by using two parametrisations, one with y = mx + c and one with x = m y + c . This has some complications because most lines are obtained twice. So engineers have sought a parametrisation which gets everything and gets it only once. Unfortunately there is no such parametrisation, because it would be a homeomorphism between a disk or rectangle in R2 and a M obius strip. And there isnt one. Note that parametrisations are local homeomorphisms. So Topology occurs quite naturally in understanding the limitations of algorithms in image analysis. It is rather useful, if you are an engineer searching hard for something, to be told it isnt there to be found.
from the origin. So the point is said to dene a line at innity.
HeavenForBooks.com

16

MICHAEL D. ALDER

You give up sooner, so you save time. These are some examples, but there are heaps of others. Topology comes up in Theoretical Physics, in many areas. It is central to modern mathematics. So Topology is useful. The second reason for studying the geometry of spaces up to homeomorphism, that is, only caring about properties which are preserved by every homeomorphism, is that it is fun. This is, of course, rather a subjective assessment, and maybe as fun goes it is pretty highbrow. Still, if it wasnt fun, people wouldnt do it unless they got paid much more money than academics get.

1.1.3

Invariants

It is clear then that there are some spaces which look dierent to the classical geometer, but which look to be equivalent to the topologist; but that not all spaces look equivalent to a topologist, or there would be little interest in studying them. One of the natural things to try to do when sorting out which objects are equivalent is to look for a simple thing like a number which cannot be dierent for two objects if they are homeomorphic. I claimed that R and R2 are not homeomorphic, and I assure you as a scholar and a gentleman2 that Rn and Rm are not homeomorphic unless n = m. The dimension, then is an example of an invariant, something which gives us a test to see if things are homeomorphic. If they are not, it may be the case that the dimensions are the same, as with [1, 1] and (1, 1). But if the dimensions are dierent, the spaces cannot be homeomorphic. Or to put it another way,
2 I hope you are not naive enough to trust me on this. A scholar, just maybe, but a gentleman only for short periods and in a bad light. As a mathematician, your job is not to trust anybody except on a purely temporary basis. You will therefore, as a good student, spend a lot of your spare time trying to prove that

Rn = Rm n = m where = denotes the relation of being homeomorphic. It looks fairly easy, does it not? Heh, heh, heh.
HeavenForBooks.com

Introduction to Algebraic Topology

17

a homeomorphism must preserve the dimension, or the dimension is invariant under the transformation which is a homeomorphism. So one way to tell, sometimes, if things are not homeomorphic, is to compute the dimension. If the same, could be, if dierent, no way. Of course, a sphere and a torus have the same dimension, so it is worth looking for other, more subtle invariants. Let us look at one which allows us to tell that a sphere and a torus are not homeomorphic. Denition 1.1.4 The unit 2-sphere is dened to be the set x R3 : x = 1 More generally, the unit n-sphere is dened to be x Rn+1 : x = 1 Denition 1.1.5 The unit 2-torus is dened to be the cartesian product space S1 S1 A triangulation of a surface is, intuitively, a process of chopping up the surface into triangles, so that the triangles intersect not at all or along their edges or vertices. This is not very precise, but if we take a tetrahedron regarded as four triangles sitting in R3, we get something which is homeomorphic to a 2-sphere and is triangulated. Exercise 1.1.3.1 Prove that a tetrahedron surface consisting of four triangles is in fact a sphere despite the pointy bits (known, technically, as the vertices). Now count the number of vertices (V), Edges (E) and faces (F) in the triangulation. For the sphere/tetrahedron we have V = 4, E= 6 F= 4. The Euler Characteristic is the number = V F + E = 2. Suppose we subdivide a
HeavenForBooks.com

18

MICHAEL D. ALDER

face of the tetrahedron. Then we add one new vertex, lose one face and gain three, so gain two faces, and also gain another three edges. Counting the edges as negative, this sums to zero, so the Euler characteristic is unchanged (invariant) under the operation of rening the triangulation. Actually, any triangulation of the sphere will produce an Euler characteristic of 2. This needs proof, but some experimenting will give you a conviction that the claim is defensible. A little thought shows that if we could show that any two triangulations of a space have a common renement, we would be done.

Exercise 1.1.3.2 Construct a formal denition of a triangulation of a surface. At least, make it as formal as you know how; this may require you to dene a surface formally, too. Construct a formal denition of a renement of a triangulation. Can you prove that any two triangulations have a common renement3 ?

Triangulating a torus is a little more complicated. Think of it as a square which has had two opposite edges glued together to give a tube, and then the two circular ends glued together to give the torus. Now you can triangulate the square, carefully, so that when you do the gluing, it stays a triangulation, but some of the edges have been glued together. There is a triangulation shown in Masseys book [2] with V=9, E=27 and F= 18, giving = 0. There is a somewhat more extremal triangulation with two triangular faces to give the original square, and gluings to give V=1, E= 3 and F = 2, again giving = 0. Again, the same argument as before shows that rening the triangulation doesnt change , and a rather careful analysis can establish that every triangulation of a torus (or indeed any surface) has the same Euler Characteristic. And since 2 = 0, the Torus and the 2-sphere are not homeomorphic.

Exercise 1.1.3.3 Triangulate the space RP2 and calculate its Euler characteristic.
The correct answer to this question is probably No. It rst requires you to dene a triangulation carefully, of course, and then to grapple with the possibility of some rather frightful ones.
HeavenForBooks.com
3

Introduction to Algebraic Topology

19

This argument may not altogether convince you, I hope it doesnt, because the details which I left out are quite dicult. Nevertheless. the idea of triangulating the space is an important one, because it turns it into a much more algebraic sort of object, consisting of vertices and edges and faces, all with certain incidence relationships between them, and the idea of being able to compute invariants for spaces is an important one.

1.1.4

Summary

This completes the motivational part of the chapter; I have indicated that Topology is about spaces and that there are important spaces which are topologically distinct, but where the problem of showing that they are not homeomorphic can be quite hard. I have indicated briey how there is an approach through invariants, one of which is the dimension, and one of which is the Euler Characteristic, , which can be obtained by triangulating the spaces, and then computing numbers which can be combined to give . I have been disgustingly woolly about details, and the next section is designed to establish the basic machinery that will be needed to make what I have sketched out intellectually respectable. It will also allow us to build far more powerful apparatus for investigating the highly non-trivial problem of deciding if two spaces are homeomorphic. We noted that some surfaces can be embedded in R3 and some cannot, and there is clearly a problem of deciding when a space can be embedded in another. This is another problem tackled by algebraic topology. There are many others, such as the question of which kinds of vector elds can exist on which kinds of manifolds, but even describing most of the problems which can be solved can be a lot of work, so I stop here.

1.2

Back to Basics

The following is material which is logically a prerequisite for this course, and I expect readers to know it thoroughly. It is possible to learn it here but not easy.
HeavenForBooks.com

20

MICHAEL D. ALDER

1.2.1

Metric Spaces

Denition 1.2.1 A Metric Space, (X,d), is an ordered pair the rst component of which is a set and the second of which is a map d : X X R satisfying the following conditions: x, y X, d(x, y ) 0 x X, d(x, x) = 0, and x, y X, d(x, y ) = 0 x = y x, y, z X, d(x, z ) d(x, y ) + d(y, z ) (1.4) (1.5) (1.6)

The map d is called the metric; it is sometimes known as a distance function. Denition 1.2.2 If (X, d) is a metric space and U X , then (U, d) is a subspace with the inherited metric obtained by restricting d to U U . Denition 1.2.3 For every Metric Space (X, d), for every positive real r and for every x X , the open ball B (x, r) X is the set {y X : d(x, y ) < r} Denition 1.2.4 For any metric space (X, d) and for any subset U X , U is open in X i U is a union of open balls. Exercise 1.2.1.1 Show that if ( X, d ) is a metric space, so is ( X, d), where x, y X, d(x, y ) = d(x, y ). Find a case where (X, d) is a metric but (X, d2 ) is not. Denition 1.2.5 If (x, d) and (Y, e) are two metric spaces, and if there is a map f : X Y , then we say that f is continuous at a X i R+ R+ x X : d(x, a) < d(f (x), f (a)) <
HeavenForBooks.com

Introduction to Algebraic Topology

21

Denition 1.2.6 If (x, d) and (Y, e) are two metric spaces, and if there is a map f : X Y , then we say that f is continuous i a X , f is continuous at a. This must look familiar! It generalises to arbitrary metric spaces the essential idea of continuity on R. We often take the metrics for granted and write f : X Y is a continuous map between metric spaces X and Y . Denition 1.2.7 For any map f : X Y , and for any V Y , f 1 V = {x X : f (x) V } Proposition 1.2.1 A map f : X Y between metric spaces (X, d) and (Y, e) is continuous i V Y is open in Y f 1 V is open in X . Proof: Let V be open in Y . Then a f 1 V, there exists an open ball B in V containing f (a), since f (a) V and V is a union of open balls. If the centre of B is at c and the radius of B is r1 , then a B d(a, c) < r1 , so there is an open ball B of radius (r1 d(a, c)) centred on f (a) contained in V . Since f is continuous it is continuous at a, so there is an open ball A on a, the image by f of which is inside B . Hence f 1 V is a union of open balls and is open in X. a X, R+ , let V be the open ball on f (a) in Y of radius . Then f 1 V is an open set in X and contains a. Hence it contains an open ball containing a, and hence it contains an open ball centred on a. Its radius is the required .

HeavenForBooks.com

22

MICHAEL D. ALDER

Corollary 1.2.1.1 The composite of continuous maps is continuous. One of the important things that has come out of Algebraic Topology is a realisation that it is the maps which are allowed between objects which are important, since there is some sort of structure preserved by the maps, and this is how you dene that structure. Hence: Denition 1.2.8 If d, d are two metrics on a space X , the metrics are said to be equivalent i the identity map IX : (X, d) (X, e) and the map IX : (X, e) (X, d) are both continuous. If two metrics on X are equivalent, then it is easy to see that all the maps out of X which are continuous with respect to one metric are continuous with respect to the other. This also holds for maps into X . Exercise 1.2.1.2 Prove the above assertion. Proposition 1.2.2 Two metrics d, e on a set X are equivalent i every set which is open in (X, d) is open in (X, e) and vice versa. Proof. The identity map on X, IX , takes every subset to itself; if U is open in (X, e) 1 then IX U = U is open in (X, d) since IX : (X, d) (X, e) is continuous. Similarly the other way round.

The map dogma which claims that it is the maps which determine the structure, tells us that if we are interested in the underlying structures on metric spaces, we should be looking at the open sets on the space. Proposition 1.2.3 In any metric space (X, d), the empty set is open in X , and X is open in X . The intersection of two open sets U, V is open, and for any set of open sets, the union of them all is open in X .
HeavenForBooks.com

Introduction to Algebraic Topology Proof.

23

That X is open in X is rather trivial: take every open ball on every point of X . That the empty set, , is open is surprising, since on the face of it it isnt a union of open balls. Given a collection C of open balls of X , the union over C is the set of points which are in at least one ball in the set C of balls. If C is empty, this is also empty. If U, V are open in X , they are both unions of open balls, and if x U V , then there is an open ball BU containing x and contained in U . Without loss of generality, we may take BU to be centred on x. Similarly there is an open ball centred on x, BV which is contained in V . The smaller of the two balls is in both, i.e. is in U V . Hence U V is open in X . If {Uj : j J } is a set of open sets of X , then x X, x
j J

{Uj } j, x Uj

Hence there is an open ball containing x which is contained in Uj and hence contained in the union. Consequently, the union is a union of open balls, and is therefore open. Note that this holds (vacuously) if J = .

Exercise 1.2.1.3 What is the intersection of an empty collection of sets?

The above theorem leads, like many theorems, fairly naturally to a denition: Denition 1.2.9 A Topological Space is a set X and a non-empty collection T = {Uj : j J }
HeavenForBooks.com

24

MICHAEL D. ALDER

of subsets of X satisfying the following conditions: T XT i, j J, Ui , Uj T Ui Uj T K, {Uj T : j K } ,


j K

(1.7) (1.8) (1.9) (1.10)

Uj T

We shall call T the collection of open sets of X in what follows. We say that the open sets are closed under pairwise intersections and arbitrary unions, and that and the whole space are always open. Then the last proposition asserts that every metric space is a topological space, with the open sets of the topology being the open sets of the metric. The converse is not true. Exercise 1.2.1.4 Take a two point set X = {x, y }. List all the topologies on X. Show that one, at least, cannot be derived from a metric. It is possible to go to even more primitive objects than topologies, but there are not many reasons for doing so. On the other hand, there are topological spaces which arise in the study of sets of functions where there is no good metric. So topological spaces are a Good Thing. On yet another hand, most of the spaces studied by algebraic topologists are embedded safely in Rn for some n, and consequently are metric spaces. On the fourth hand, life is often cleaner, simpler and altogether less hassle if we go to the basics and work with least machinery. Sometimes, all the extra machinery just gets in the way of seeing what is happening. For reasons of simplicity and austerity I shall therefore use topological spaces as the basic framework. The abstraction frightens some people to death, and charms others. I suggest that you see it as a kind of game and try to enjoy it. The next advice, superuous to the intelligent and wasted on the rest, is to make up an example of every new object dened. In fact several examples, some straightforward and some bizarre. This is (a) good fun, requiring ingenuity and perhaps creativity and (b) extremely useful in coming to grips with the abstract ideas.
HeavenForBooks.com

Introduction to Algebraic Topology

25

1.2.2

Topological Spaces

Repeating ourselves for completeness: Denition 1.2.10 If X is any set, a topology for X is a collection T of subsets of X satisfying: T and X T U, V X, U, V T U V T J, {Uj T : j J } ,
j J

(1.11) (1.12) (1.13)

Uj T

Denition 1.2.11 A topological space is a set X together with a topology T for X . Denition 1.2.12 If (X, T ) and (Y, S ) are topological spaces and if f : X Y is a map, then f is continuous i V Y, V S f 1 V T

Exercise 1.2.2.1 Show that the composite of continuous maps is continuous. Denition 1.2.13 For any set X , the map IX : X X is the identity, x X, IX (x) = x Exercise 1.2.2.2 Show that the identity map IX : (X, T ) (X, S ) is continuous i the topologies S , T are the same. Denition 1.2.14 A map f : X Y between topological spaces (X, T ) and (Y, S ) is a homeomorphism i it is continuous and has a continuous inverse, that is, i g : Y X : f g = IY , g f = IX and both f and g are continuous.
HeavenForBooks.com

26

MICHAEL D. ALDER

Denition 1.2.15 A map f : X Y between topological spaces (X, T ) and (Y, S ) is an embedding of X in Y i it is a homeomorphism onto its image. Thus S 1 normally comes embedded in C or R2 . Denition 1.2.16 The subsets of X which are elements of T are called the open sets of the topological space. A closed set is one which is the set complement of an open set. Denition 1.2.17 If (X, T ) is a topological space, and if V X is a subset, then we dene T |V = {Vj : j J } where {Uj : j J } = T and j J, Vj = V Uj Proposition 1.2.4 T |V is a topology for V . Proof. By denition, T |V , and V T |V . It is clear that for any two sets Vi , Vj T |V , Vi Vj = (Ui Uj ) V so T |V is closed under intersections. By distributivity, it is also closed under arbitrary unions.

Denition 1.2.18 The topology T |V on the subset V of the topological space (X, T ) is called the relative or subspace topology. Exercise 1.2.2.3 If (X, T ) is a topological space derived from a metric space (X, d) and V X , show that the relative topology on V is derived from the metric restricted to V .
HeavenForBooks.com

Introduction to Algebraic Topology

27

You will notice that there are such things as linear spaces, linear maps, linear subspaces and cartesian product linear spaces; there are groups, homomorphisms, subgroups and cartesian product groups. There are also metric spaces, continuous maps, metric subspaces and cartesian product metric spaces, and there are topological spaces, continuous maps and topological subspaces. You might also note that sometimes, there are quotient linear spaces, and quotient groups. The question comes up, are there quotient metrics? Are there quotient topologies? The answers are sometimes and yes, respectively. The question are there cartesian product metrics? is of course yes, with a choice of several, but there is a unique product topology on the cartesian product of two topological spaces. Denition 1.2.19 Let denote an equivalence relation on a set X , and let T be a topology on X . Let X/ denote the equivalence classes, and let : X X/ denote the projection which takes each point x of X to the class [x] containing x. Then let T / denote the collection of subsets of X/ : V X/ T / i 1V is open in X This collection of subsets of X/ is called the quotient topology on X/ . Calling something a topology doesnt make it one: Proposition 1.2.5 The quotient topology is a topology on X/ . Proof. T / since 1 = . Similarly, X/ T / . U, V X/ , U, V T / 1 U T and 1 V T 1 U 1V T 1 (U V ) T U V T/

HeavenForBooks.com

28 Similarly for unions.

MICHAEL D. ALDER

Note that this ensures that the projection map is continuous, just. If the topology on X had any fewer open sets, or the topology on the quotient set any more, it wouldnt be. Exercise 1.2.2.4 Construct an equivalence relation on a nite set, construct a topology on the set and the construct the quotient topology. When we talked rather vaguely of gluing points together to get a new space, we are implicitly dening a rather obvious equivalence relation on the points of the rst space, then taking the quotient space with the quotient topology. For example, take the topology induced from the standard metric on the unit square in R2 . Dene an equivalence relation on the points contained in the square as follows: a point in the interior is equivalent only to itself. A point on the bottom boundary is equivalent to itself and also to the point with the same x-coordinate on the top boundary, and the points on the left and right boundaries are paired if their y-coordinate is the same. Then the quotient space, with the quotient topology, is homeomorphic to a torus. Exercise 1.2.2.5 Show that there exists a metric space and an equivalence on it so that, regarding the metric space as having the induced topology, the quotient space is not a metric space. Exercise 1.2.2.6 If X is a space and A is a subspace, we write X/A to denote the quotient space obtained when every point of A is equivalent to every other point of A, and every point of X A is equivalent only to itself. If X = [0, 1] and A = {0, 1}, describe X/A. If X is a 2-sphere and A is the equator, describe X/A. If X is the two dimensional disk {x R2 : x 1}, and A is the boundary, S 1 , what is X/A?
HeavenForBooks.com

Introduction to Algebraic Topology

29

If X = [0, 1] and A is the sequence (1, 1/2, 1/4, , 1/2n , , 0), draw a picture of X/A. What dierence does it make if 0 is left out of A? Denition 1.2.20 If (X, T ) and (Y, S ) are topological spaces, and if X Y is the cartesian product of the underlying sets, with X : X Y X and Y : X Y Y the two projection maps, and if U T and V S , then I dene a topology (T S ) on X Y by stipulating:
1 1 U Y V (T S ) X

and every (open) set in T S is a union of such sets. This makes the projection maps just continuous. It is also an open map, that is, it takes open sets to open sets. Again, just. If the topology on the product had any more open sets, it would not be open. Since the set of topologies on a base set is partially ordered by inclusion, we could use the observations on just continuous and just open to characterise the quotient topology. Again, pious hopes are not enough; we have to show that there really is a topology here. Proposition 1.2.6 The product topology is a topology on the cartesian product. Proof: Do it as an exercise.

Exercise 1.2.2.7 Dene the product topology for the cartesian product of a nite collection of topological spaces. Find maps which are open and others which are not.
HeavenForBooks.com

30

MICHAEL D. ALDER

A map between topological spaces is closed i the image of every closed set is closed. Show that the projection from R2 to R which projects onto the rst component is a map which is (a) continuous (b) open and (c) not closed. Show that the space X Y and the space Y X , with the product topologies, are homeomorphic. Exercise 1.2.2.8 Construct a denition to make precise the uy notion of maps being just continuous. You may need to look at the family of all topologies and their obvious partial ordering. Give a denition of the product topology via this idea. There are several approaches. Exercise 1.2.2.9 Construct a denition of a product topology for any collection of topological spaces. Do so in such a way that the projection maps are continuous and open. Be careful! S 1 was dened earlier, and T 2 was dened via the cartesian product. Exercise 1.2.2.10 Let the space dened by gluing together opposite sides 2 of the unit square as above be called T1 . Observe that it is dened as a topological space after all these denitions. Show that it is homeomorphic to T 2. Note that the constructions we have got for making topological spaces from old ones are quite powerful, and easily extensible. For example, suppose I have two topological spaces which are surfaces, and I embed a disk in each. Now I chop out the interior of the disks, and glue the two circular boundaries together. This gives me a new object, called the connected sum of the two surfaces. There is a bit of a problem of ensuring that the homeomorphism class of resulting objects will not depend on the details of the embeddings and the gluing, so the idea needs a bit of work, but it is plausible that this can be done. Exercise 1.2.2.11 Take two copies of the real line R and R and let be the unique eld isomorphism between them, with x = (x). Identify x R
HeavenForBooks.com

Introduction to Algebraic Topology

31

with x R for every x except zero. The result is essentially a Real line with two distinct zeros. Show that there is a well dened topology, specify the open sets. Show the space is not metrisable, i.e. there is no metric on it which gives the topology. We can place additional requirements on topological spaces in order to make them at least a little more like the metric spaces we know and love. One is: Denition 1.2.21 A topological space (X, T ) is said to be hausdor or T2 i for every pair of distinct points x, y X there exist disjoint open sets U, V T with x U and y V . Exercise 1.2.2.12 Show that every metrisable space is hausdor. Exercise 1.2.2.13 Show that in a non-empty hausdor space, the points are closed. That is, x X, X {x} is open From now on, we shall simply take the topology for granted and simply refer to the open sets on X , and write that X is a topological space. Also, unless otherwise stated, we assume that all the spaces we shall be concerned with are hausdor. Many notions which you are used to from studying Rn make sense and are much simpler in the setting of topological spaces. Denition 1.2.22 Let S 0 denote the topological space with two points, and every subset open. You can probably see why the above space is so called. The topology where every subset is open, or alternatively every singleton set is open, is is called the discrete topology. Such spaces are usually disconnected in an obvious intuitive sense, and S 0 is the simplest of them, except for the empty set. The topology at the other end of the partially ordered set of topologies on a set X is the chaotic topology which has precisely two open sets if X = .
HeavenForBooks.com

32

MICHAEL D. ALDER

Denition 1.2.23 For any topological space X, we say that a map f : X S 0 which is onto is a disconnection of X. If there exists a disconnection of X, we say that X is disconnected. If X is not disconnected because there exists no disconnection, we say X is connected. Exercise 1.2.2.14 Show that a topological space X is disconnected i there exist two non-empty open sets U,V which are disjoint and contain X in the union. Prove that R {0} is disconnected, but that R is connected. Hence deduce that removing any point for R gives something not homeomorphic to R. Show that R {p} is homeomorphic to R {q } for any points p, q R. Denition 1.2.24 A path from a to b in X , for any two points a, b X , is a continuous map f : [0, 1] X such that f (0) = a, f (1) = b. The topological space [0, 1] has the obvious topology derived from the usual metric as a subspace of R. Exercise 1.2.2.15 Show that the open sets on R are the sets which are a union of open intervals. Show that [0, 1) is open in [0, 1] but not in R. Denition 1.2.25 A space X is path connected i a, b X , there is a path from a to b. Exercise 1.2.2.16 Show that R is path connected. Show that S 1 is path connected. Show that if X and Y are path connected, so is X Y . Exercise 1.2.2.17 Show that if a topological space X is path connected, then it is connected. Show that the converse is false. One of the useful properties that a subset of Rn could have was to be closed and bounded. This comes up in lots of approximation theorems. The generalisation of closed presents no problems:
HeavenForBooks.com

Introduction to Algebraic Topology

33

Denition 1.2.26 A subset U of a topological space X is closed i the complement, X U is open in X . Denition 1.2.27 A cover of a space X by open sets is a family of open sets the union of which contains X . A subcover is a selection of the sets from the family. The notion of boundedness seems to require a metric. We can however translate the Heine-Borel theorem intro a denition. Recall: Theorem 1.2.1 (Heine-Borel) A subset X of R which is closed and bounded has the property that every cover of X by open sets has a nite subcover. Proof. See any good textbook on point set topology or elementary analysis, e.g. [3]. This leads to the denition: Denition 1.2.28 A topological space X is compact i every cover of X by open sets has a nite subcover. Exercise 1.2.2.18 Show that every closed subspace of a compact space is compact. Show that the image of a compact space by a continuous map is compact. Exercise 1.2.2.19 Show that the product of hausdor spaces is hausdor. Exercise 1.2.2.20 Show that if a point x is in the complement X A of a compact subspace A of a hausdor space X, then there are disjoint open sets U and V such that A U and x V .
HeavenForBooks.com

34

MICHAEL D. ALDER

Exercise 1.2.2.21 Show that every compact subspace of a hausdor space is closed, but that compact subspaces need not be closed in general. Deduce that if we have a 1-1 continuous map of a compact space into a hausdor space, the inverse is also continuous. The following result is at rst sight surprising, but follows from a suitable choice of denition of the product topology: Theorem 1.2.2 (Tychonov) Any non-empty product of compact spaces is compact. Proof. See any good book on analytic topology. For example, G.F. Simmons Introduction to Topology and Modern Analysis [3]. Exercise 1.2.2.22 Prove the Tychonov theorem for the case of a nite product of spaces. We cannot leave the subject of topological spaces without saying a little about limit points. There are two big areas of mathematics, algebra and topology. Algebra is about algorithms for doing sums. Topology is about notions of approximation or closeness. So ultimately the point of studying analytic or point set topology is to understand what Newton was on about when he took limits. Denition 1.2.29 If A X for a topological space X , then a X is said to be a closure point of A i every open set containing a contains points of A. Denition 1.2.30 If A X for a topological space X , then the set of all and is called the closure of A. closure points of A is written A
HeavenForBooks.com

Introduction to Algebraic Topology

35

Denition 1.2.31 If A X for a topological space X , then a X is said to be a limit point of A i every open set containing a contains points of A other than a. Denition 1.2.32 A subset A X of a topological space, X , is said to be discrete i no point of X is a limit point of A. Exercise 1.2.2.23 Show that if X is a compact hausdor space, there can be no innite discrete subset. Alternatively, every innite subset of a compact hausdor space has a limit point. Exercise 1.2.2.24 Show that 0 is a limit point of R+. Denition 1.2.33 For any space X , and any subspace A X , the boundary of A in X is the set A of points which are in the closure of A and also in the closure of the complement of A. Denition 1.2.34 If A is a subset of a topological space X , then the interior of A is the union of all open subsets of X contained in A. The interior and closure operators give us (generally) new subspaces, and most of the foundations of analysis can be obtained by taking these operators as primitives instead of going through open sets. That is to say, I can dene a space by giving operators on its subspaces with certain axiomatic properties, and then derive a denition of an open set, instead of starting with an axiomatic denition of open sets and deriving a denition of the closure and interior operators. Exercise 1.2.2.25 Find a subset of R2 which is closed and has empty interior. Find another which is open and has empty closure. Find another which has the closure of the interior empty while the set is not empty. Find another which has the interior of the closure empty while the set is non-empty. Find another which has the closure of the interior the same as the interior of the closure. Let all your examples be dierent.
HeavenForBooks.com

36

MICHAEL D. ALDER

1.2.3

Groups

We are doing algebraic topology, which means we have to know the elements of algebra as well as of topology. Fortunately, algebra is easy (CTC Wall) and anyway the amount I shall assume you know is very small. Denition 1.2.35 A group is a set G together with a binary operation, , : G G G such that: a, b, c G : (a b) c = a (b c) e G a G : a e = e a = a a G a1 G : a a1 = e = a1 a Denition 1.2.36 A group G is abelian i a, b G : a b = b a (1.17) (1.14) (1.15) (1.16)

We usually suppress the and so write ab for a b, except when the group is abelian when we write a + b. A group may be given or presented via generators and relations. In the simplest case there are no (non-trivial) relations: Denition 1.2.37 The free group on a non-empty set X is the set of (nite length) strings n n n xi1i1 xi2i2 xikik under concatenation, where the xi are any elements of X but consecutive xi are dierent, and the ni are integers which may be zero, and where x0 is the group identity, e for any x X . When concatenating, the rule of indices xi xj = xi+j is followed, in particular x X : xx1 = e. This is a trivial relation, not to be compared with the non-trivial ones we consider subsequently. We shall write F (X ) for the free group on X .
HeavenForBooks.com

Introduction to Algebraic Topology

37

Example 1.2.3.1 The free group on a singleton set X = {x} has terms of the form xn for any integer n, and hence is isomorphic to Z, the group of integers. It is, of course, abelian. The free group on the set X = {x, y } having two elements has strings of the form: xn1 y n2 xn3 y n4 y n each of which is an element of F (X ), which is not abelian. A relation on a group is an equation of the form xi1i1 xi2i2 yik k = e where e is the identity of the group, and where the terms xi in the relation are elements of the group. Example 1.2.3.2 On the free group on one element x, take the relation x2 = e. Then the group satisfying this condition is just Z2 . Denition 1.2.38 S is a set of generators for the group G i S G and every element of G is in the free group on S It is rather taken for granted here that a string on the free group on S is interpreted as a product in the group G of the elements of G that are in S . This can be expressed rather easily using the idea of a homomorphism. Denition 1.2.39 If G, H are groups, a map f : G H is a (group) homomorphism i x, y G, f (xy ) = f (x)f (y ) where we have used concatenation of symbols to denote the (generally dierent) operations in each group.
HeavenForBooks.com n n ni

38

MICHAEL D. ALDER

Denition 1.2.40 An isomorphism between two groups G, H is a map f : G H which has an inverse g : H G both maps being homomorphisms. Exercise 1.2.3.1 Show that if f is a homomorphism with inverse h. Then h is also a homomorphism. Denition 1.2.41 The kernel of a homomorphism of groups, f : G H is the set ker(f ) = {g G : f (g ) = eH } where eH is the identity in H . Exercise 1.2.3.2 Show that the kernel of a homomorphism is a subgroup of the domain. Denition 1.2.42 The image of a homomorphism of groups, f : G H is the set im(f ) = {h H, g G : f (g ) = h} Exercise 1.2.3.3 Show that the image of a homomorphism is a subgroup of the codomain. Denition 1.2.43 If G, H are groups, we can make G H into a group by dening the binary operation on the cartesian product by the rule (g1 , h1 ) (g2 , h2) = (g1 g2 , h1 h2) where I have used juxtaposition for the two operations in G, H . Now observe that if S is any set and if f : S G is a map of the set into a group G, there is an obvious and unique extension of f to the free group on : F (S ) G, which is a group homomorphism. All we have to the set S , f on a string in the obvious way given that we have f dened do is to dene f (xn) = (f (x))n . on the basic terms. In particular, f
HeavenForBooks.com

Introduction to Algebraic Topology

39

rigorously. Show it is unique, i.e. Exercise 1.2.3.4 Dene the extension f show that there is only one homomorphism of the free group F (S ) into G which agrees with f on the elements of S . We may now give a more intelligible denition of a generating set: Denition 1.2.44 A non-empty subset S G of a group G is said to be a set of generators for G i i : F (S ) G is onto, where i is the inclusion map of S in G. In the case of the above denition, the kernel of the map i is a normal subgroup of the free group F (S ) and equating any element of this to the identity gives a relation on G. When this is done for a set of generators of the kernel, we say that G is presented in terms of generators and relations. If G is an abelian group, we often signify this by using additive notation; thus we write g1 + g2 instead of g1 g2 , and 0 instead of e for the identity, and 2g for g 2 . Example 1.2.3.3 For the case of a singleton set {1} which generates Z and the relation 2x = 0, we note that we have 2Z is the kernel, which is a free group on {2}, and that Z2 is the resulting group. Evidently all the nite cyclic groups can be presented in an analogous way. Exercise 1.2.3.5 Translate the above into multiplicative notation. It is a fact, although we do not give a proof, that: Proposition 1.2.7 Every subgroup of a free group is free. The trivial subgroup is, by a not unreasonable convention, generated by the empty set of generators. Any group can be presented in terms of generators and relations; more precisely:
HeavenForBooks.com

40

MICHAEL D. ALDER

Proposition 1.2.8 Any group G is the image by a homomorphism of a free group. That is, any group is a quotient group of a free group. Proof. Take as a set of generators of G, G itself. Then the unique extension of the identity map gives a homomorphism from F (G) to G which is onto.

Denition 1.2.45 The commutator subgroup of a group G is the smallest normal subgroup containing all elements of the form xyx1y 1 , for every pair of x, y G. The expression xyx1y 1 is denoted [x, y ] and the commutator subgroup written [G, G].

Proposition 1.2.9 The quotient group G/[G, G] is abelian. It is sometimes called the abelianisation of G. Proof. An exercise for the diligent student.

Just as we have dened the free group on a set S , we can dene the Free Abelian Group on S as the abelianisation of F (S ). An easier way to think of it is to take it that the strings of symbols from S which are elements of the Free Abelian Group on S have the symbols themselves commuting. Hence everything commutes.

Example 1.2.3.4 The free abelian group on a singleton set {x} is still, up to isomorphism, Z. The free abelian group on two generators is Z Z. The usual expression is in terms of the direct sum Z Z. The direct sum is the same as the direct product for a nite collection of abelian groups.
HeavenForBooks.com

Introduction to Algebraic Topology

41

Denition 1.2.46 If {Gj : j J } is a set of abelian groups, the direct sum Gj


j J

is the subset of the cartesian product j J Gj which contains only those product terms for which all but a nite number of the components are the identity 0j Gj . It is common to use direct sum notation for nite sets of abelian groups rather than the direct or cartesian product. We can see that there is a trivial abelian analogue of the result that any group can be presented in terms of generators and relations: Proposition 1.2.10 Every abelian group is the homomorphic image of a free abelian group. Proof. Take the free abelian group on the group G itself, let I : G G be the identity, and observe that, just as for the free group, there is a unique extension , a homomorphism from the free abelian group on G to G, which is of I to I rather trivially onto.

A particularly important case for us is when an abelian group is given as a homomorphic image of a free abelian group, and when the free abelian group is the free abelian group on some nite set. The examples of Z, Z Z and Z2 given above obviously satisfy this condition. Denition 1.2.47 An abelian group is said to be nitely generated when it has a nite generating set, alternatively when it is the homomorphic image of a free abelian group on a nite set. There are a number of intuitively obvious results which can be proved, and should be proved carefully for those unused to elementary group theory. In particular:
HeavenForBooks.com

42

MICHAEL D. ALDER

Proposition 1.2.11 The free abelian group on a set of k elements is isomorphic to Zk , and Zk is isomorphic to Z i k = . Denition 1.2.48 An element g of a group G has order n i g n = e and g k = e n|k We take the order to be positive. Proposition 1.2.12 The elements of an abelian group G having nite order form a subgroup of G. Denition 1.2.49 The subgroup of elements of an abelian group G of nite order is called the torsion subgroup of G. If the torsion subgroup of G is the identity element alone, then G is said to be torsion free. If every element of G has nite order, G is said to be a torsion group Proposition 1.2.13 If an abelian group G has torsion subgroup T , then the group G/T is torsion free. Proposition 1.2.14 If abelian groups G, H are isomorphic, then so are their torsion subgroups, T, U and the quotient free abelian groups G/T, H/U . None of the above results are particularly dicult. The following result is known as the classication theorem for nitely generated abelian groups and is relatively complicated. It need not be proved if you are not an algebraist, but it needs to be known. Fortunately it is easy to remember: Theorem 1.2.3 (Classication of nitely generated abelian groups) Every nitely generated abelian group G is isomorphic to the direct sum of the Torsion subgroup T with the quotient group G/T . The latter is isomorphic to the direct sum of k copies of Z, for non-negative integer k . The former may be written, up to isomorphism, as a sum C1 C2 Cn, where each Ci is a nite cyclic group of order mi, and each mi is a divisor of mi+1, and these orders are uniquely determined.
HeavenForBooks.com

Introduction to Algebraic Topology

43

Denition 1.2.50 The order of G/T for a nitely generated abelian group G with torsion group T is called the rank of G. The orders of the torsion subgroup cyclic components are called the torsion coecients of G. For a proof of the above theorem, see any reasonable book on group theory. (A reasonable book on group theory will contain a proof by denition of reasonable.) Exercise 1.2.3.6 Show that Z2 Z3 Z3 is a nitely generated abelian group, and nd the torsion coecients. Finally, a result we shall need later. Denition 1.2.51 If G is an abelian group and A, B are subgroups, A + B is the set {a + b}, a A, b B Proposition 1.2.15 If A, B are subgroups of an abelian group G, then A + B and A B are subgroups of G. Proof: It suces to prove closure, which is obvious.

Theorem 1.2.4 (First Isomorphism Theorem) If A, B are subgroups of an abelian group G, and if i : A A + B is the obvious inclusion, the map i : is an isomorphism.
HeavenForBooks.com

A A+B AB B

44 Proof.

MICHAEL D. ALDER

It is obvious that i(A B ) is a subgroup of B , so a A, a + (A B ) A AB i(a) = (a + 0) + B A+B B

So there really is an induced homomorphism i : A+B A AB B

a A, [a] = a + (A B ) ker(i ) i(a) B a AB [ a] = 0

A AB

Hence i is 1-1.

a A, b B, (a + b) + B

A+B B

A+B B A+B a+B B a + (b + B ) i(a + A B ) = a + (b + B ) A+B B

So i is onto. Hence i is an isomorphism.

HeavenForBooks.com

Introduction to Algebraic Topology

45

This concludes the algebra with which I shall assume familiarity. There is a bit more algebra to come, but if you are uent with the above material, it should not be dicult. If the above material looks deep and mysterious, do all the proofs of the propositions.

1.2.4

Manifolds

First, as for earlier subjects, a supercial, chatty, discursive outline of the general ideas with examples and no proofs. If this were a course on Zoology, this is where I should be showing you photographs of elephants prior to dening one. (This gives you some idea of how much I know bout Zoology.) It should be apparent that all the analytic or point set topology that has been discussed above, is technical machinery for studying something else. It is true that some mathematicians live their whole lives worrying about the machinery, but it was developed for a reason, in fact two separate reasons. The two major applications are, rst to spaces of functions so as to get a precise grip on dierential and integral operators, and second to real topology, the study of manifolds. It is the second that motivates us here. By way of warning, it should be said that other sorts of spaces are important; there are spaces which arise in order to get a grip on manifolds, and the natural spaces for algebraic topology are probably the complexes. There are several varieties of these, too, and we shall meet some of them later. Still, it was studying manifolds that led to the origins of real topology, and so it is desirable to have a good feeling for what they are. The simplest manifolds are the one dimensional curves, and the two dimensional surfaces. One of the important things about a manifold is that it has a dimension. Once we have dened them properly, the curves and surfaces are just particular examples. The one dimensional manifolds are pretty simple: up to homeomorphism there are just two: one is R and the other is S 1. This is too simple and a false statement as it stands, since a collection of manifolds of the same dimension can be a manifold, so I need to say that there are only two connected manifolds, R and S 1. Of these, only S 1 is compact. The two dimensional
HeavenForBooks.com

46

MICHAEL D. ALDER

compact, connected manifolds, otherwise known as compact, connected surfaces, are a bit more complicated; RP2 is one, S 2 is another, and T 2, the torus is a third. We can regard the torus as obtained by sticking a handle on a sphere. We can stick any nite number of handles on a sphere, the number of handles is called the genus, and when we do this we get dierent compact, connected surfaces. Or we can glue handles on to RP2 to get another lot of compact connected surfaces. And this is all the compact, connected two dimensional manifolds there are. The above assertion is is a very strong claim indeed, and a claim which requires proof; and a long and complicated proof it is, if done carefully. Of course we cant even begin to tackle it because proofs are relative to denitions and we havent got any. Yet. The three dimensional compact connected manifolds are too horrible to be discussed in this course, and for dimensions higher than three there is no useful way to describe them all. Still, much can be done with particular cases, and saying things about particular cases is both possible and highly desirable. Some reasons were discussed in the rst part of this chapter: manifolds and ows on them arise in nature. One of the ways manifolds tend to arise is by constraints on measurement variables, and these occur in various ways, some of them very physical. The simplest mathematical example is probably a linear relation between two variables as in ax + by = 0. These give linear manifolds, in general. More interesting is when we have non-linear constraints on two variables, as in x2 + y 2 1 = 0. This, of course, gives the unit circle. In general, you might conjecture that there is a generalisation of the rank-nullity theorem to the eect that if you have n variables and k independent dierentiable conditions on them, you get a smooth n k dimensional solution space, which is a smoooth manifold sitting in Rn . This, when stated correctly, is the implicit function theorem, and isnt as strong as the conjecture. We saw also that manifolds can arise dierently: we got RP2 by a quotient process of a rather dierent sort. The manifolds which arise by constraining n variables are, when they exist, embedded in Rn . But RP2 came dened in an intrinsic way which said nothing about an embedding. So it would be essential to have a denition which does not depend on an embedding. This
HeavenForBooks.com

Introduction to Algebraic Topology would allow us to dene a manifold without respect to where it is sitting.

47

To give an example of this, I can dene S 1 by saying that it is obtained by gluing two copies of [1, 1] together, joining the two pairs of end points. Alternatively, I can say it is [0, 1]/{0, 1}, the quotient of the unit interval by identifying its end points. These are an intrinsic denitions which do not say where, if anywhere, the space is embedded. Note, also, that the universe we live in is three dimensional: there are grounds for thinking that the entire universe is compact. We do not know whether we live in S 3 , or S 1 S 2 or S 1 S 1 S 1 , or conceivably something much nastier. And we have no reason to suppose that there is anything outside the Universe in which our Universe is embedded. So intrinsic denitions of manifolds are essential. To get a grasp of what can be done here, throw your mind back to an earlier stage of your mathematical development and the way in which you parametrised surfaces and curves in order to do integration of vector elds (or dierential forms if you did it properly). Basically, a parametrisation of a surface involved mapping some nice region in R or R2 onto the surface, or a part of it. You may have had to do this with several nice regions, as when you want to integrate over a sphere. Basically, you took squares or rectangles in the plane, and set one to the northern half and one to the southern half, or at least, this is one way which comes out of the spherical polar coordinate framework. There are other ways of tackling this, but a possible way is to start out by saying that we get a manifold by gluing cubes together by mapping the cubes into the manifold. We shall use this approach when dealing with Homology Theory, the primary tool of algebraic topology. In the case of one dimension, we glue intervals, in the case of surfaces we glue squares onto the manifold. In general the problem with this is that we can get rather a lot of other things besides manifolds. As well as having to talk about manifolds, we shall be interested in manifolds with boundary. The unit square in R2 I 2 = {(x, y ) R2 : 0 x 1&0 y 1} has a boundary to it, and we want to be able to describe these things as well. After that rather extensive bit of chat about what sort of things manifolds
HeavenForBooks.com

48

MICHAEL D. ALDER

are, why we love them and are curious about them, why we need an intrinsic denition rather than one which gives us the object as a subset of Rn , let us get to the nitty-gritty and give a denition which although imperfect will do for a rst pass:

Denition 1.2.52 Let X be a hausdor space, and let { Uj : j J } be a set of open subsets of Rn . Let {fj : Uj X } be a set of continuous maps from the open sets such that each fj is 1-1 and the images of the Uj cover X , that is x X, j J, u Uj : fj (u) = x Then if whenever Ui Uj = we have that the map fi1 fj is a homeomorphism on its domain, we say that the family of maps {fj : j J } is a family of charts, and that X together with this family of charts is a manifold of dimension n.

The family of charts is called an atlas. It is formally convenient to extend the atlas by sticking in all possible charts which are compatible with the atlas, but this is a technicality which must not distract you from the essentials. Again, we do not stipulate here that the domains of the charts, the Uj should be open disks or open cubes, but we dont say that they cant be either.

Exercise 1.2.4.1 Let U1 be the open interval (, ) which is an open set in R. Let f1 : U1 S 1 R2 be the map which sends t to (cos t sin t) R2. Let U2 be the same as U1, but let f2 : U2 R2 be dened by f2 (t) = ( cos t, sin t). Show explicitly that this denes an atlas for S 1 and hence that S 1 really is a manifold of dimension 1.

Note that our construction makes a manifold locally a metric space, indeed locally a linear space, Rn . This is cheering, but it does not necessarily make it a proper metric space, since distances in one chart may be dierent from distances in another. Still, we can hope for a good many properties of the nice spaces which are open subsets of Rn to be carried over.
HeavenForBooks.com

Introduction to Algebraic Topology

49

Exercise 1.2.4.2 Note that we started o with a hausdor space. Suppose we had dropped this condition, making it, say, any old set. Find a thoroughly nasty quasi-manifold which isnt metrisable. Exercise 1.2.4.3 Show that S 2 is a manifold, likewise T 2 Exercise 1.2.4.4 Show that the product of two manifolds is a manifold in a natural way. Exercise 1.2.4.5 Show that the quotient space of a manifold need not be a manifold. If we want to extend the idea to a manifold with boundary, which, be it noted is not a manifold, we need to distinguish between the interior points of the manifold which are treated as above, and the boundary points which look rather special. Take a half of Rn , H n = {(x1, x2, xn ) Rn : x1 0} and require that some of the points of the space X should be covered by the boundary of H n instead of an open set in Rn . Exercise 1.2.4.6 Dene the term manifold with boundary using the above idea. Show that [0, 1] is a one dimensional manifold with boundary, and that D2 = {x R2 : x 1} is a two dimensional manifold with boundary. Show that the product of a manifold with a manifold with boundary is always a manifold with boundary, and that the product of two manifolds with boundary is also a manifold with boundary. Show that the boundary of a manifold with boundary is a manifold, and show that the boundary of a compact connected manifold with boundary is a compact, but not generally connected, manifold. (This will require you to dene the boundary of a manifold with boundary)
HeavenForBooks.com

50

MICHAEL D. ALDER

The connected sum operation on surfaces was sketched earlier: we take an embedding of D2 , the unit disk in R2 into one surface (i.e. two dimensional manifold) and another of the same unit disk into another surface. We then remove the interior of the disks from both surfaces, and identify the matching points on the circle. It should be clear to those of a nervous disposition that there are some perfectly horrible maps of D2 into R2 , and that this isnt necessarily a straightforward job. For higher dimensions, it can only get worse- meditate upon the Alexander Wild Horn ed sphere, for example. If you feel foolhardy, try the next exercise:

Exercise 1.2.4.7 Show that if we do this operation on the two manifolds which are both copies of R2 with single charts in the atlas in both cases, we get a manifold. Similarly, show carefully that attaching a handle to R2 gives a manifold.

1.2.5

Odd

There is one result for compact metric spaces which is needed in the sequel and which doesnt logically belong anywhere we have been so far. First a few denitions. Denition 1.2.53 If U X is a subset of a metric space (X, d), the diameter of U is sup{d(u, v ) : u, v U } when this exists. Denition 1.2.54 Let {Uj : j J } be a cover of a metric space (X, d) by open sets. We say that the real number > 0 is a Lebesgue number for the cover, if for every subset A X of diameter less than , there is at least one j J such that A Uj .

It is clear that we can cover some spaces with open sets which get progressively smaller, and where the amount of intersection of the sets also gets smaller, so no Lebesgue number exists. On the other hand:
HeavenForBooks.com

Introduction to Algebraic Topology Theorem 1.2.5 (Lebegues Covering Lemma) In a compact metric space, every open cover has a Lebesgue number. There is a proof in [3].

51

1.3

Summary

The preceding back to basics section summarised all the material on metric spaces, topological spaces and groups which you are assumed to know. Whether you actually do know it is not altogether clear, but if not, the remedy is in your own hands. By now you can be assumed to be suciently adult to take whatever steps are necessary to make up for the deciencies of your education so far, when they become too painfully apparent. This summary isnt going to summarise the preceding section, because the preceding section is a summary. Sort that one out. We are now ready to start on the new stu which I shall assume you dont know about.

HeavenForBooks.com

52

MICHAEL D. ALDER

HeavenForBooks.com

Chapter 2 Categories and Functors and


2.1 Categories

You will by now have noticed certain basic similarities which keep on coming back to haunt us. There are things called sets. Between most pairs of them, there may be several things called maps. Given any set, X , there are usually some subsets of the set X . Given two sets,X, Y there is a product set, X Y . If there is a map f : X Y and another g : Y Z , then there is a map called the composite and written g f : X Z . Particularly important kinds of maps occur between sets; for instance every set X is associated with an identity map, IX . If f : X Y is any map between sets, then f IX = IY f = f . If g : Y X is another map, and if f g = IY and g f = IX , then we say that f, g are inverses. In this case, there is an equivalence between some of the sets determined by the invertible maps. If I went through every occurrence of the word set and every occurrence of the word map and replaced them by group and group homomorphism, the paragraph would still be true. If I went through and replaced every occurrence of set by linear space and map became linear map, the same thing would hold. Other possibilities will doubtless occur to you. This leads 53

54

MICHAEL D. ALDER

us to a new stage of abstraction: Denition 2.1.1 A category is a collection of two sorts of things, Objects and maps. The collection of objects is generally a class which is not a set. Each object, X , is associated with a particular map IX . Each map f , is associated with a pair of objects, dom(f) and ran(f). In the case of the map IX , both of these are the object X . If X, Y, Z are three objects (not necessarily distinct), and if f, g are two maps (not necessarily distinct), and if dom(g) = ran(f) = Y and dom(f) = X, ran(g) = Z, then there is a map written gf , with dom(gf) = X, ran(gf) = Z. For composites of three or more, the composition is associative whenever it is dened. Exercise 2.1.0.1 List as many distinct categories as you can think of. About a dozen is average. Any sensible abstraction like this is intended to save eort by doing things in abstract categories rather than doing them one damn time after another in several dierent categories. Naturally, we write f : X Y whenever f is a map in the category under discussion and dom(f) = X, ran(f) = Y. Sometimes we write: f X Y

or X f

-Y

for this. The composite of maps f, g will be written gf : X f

-Y

-Z

HeavenForBooks.com

Introduction to Algebraic Topology Now we can dene a monomorphism in an abstract category as follows:

55

Denition 2.1.2 A map f : X Y is said to be a monomorphism i whenever g, h : A X are maps, fg = fh g = h It is clear that for the category of sets, if f is not 1-1, then it may be that fg = fh but g = h. g and h may dier, sending a point a to g (a) = h(a). But if f (g (a)) = f (h(a)) we may still have fg = fh. Conversely, if f is 1-1, then if g and h dier on some point a A, then so will fg and fh. So we have shown that in the category of sets, the 1-1 maps are monomorphisms. Dually: Denition 2.1.3 A map f : X Y is said to be an epimorphism i whenever g, h : Y Z are maps, gf = hf g=h

It is obvious that if, in the category of sets, the map f is not onto, then we could have g = h but gf = hf . Conversely, if f is onto, and if g (f (x)) = h(f (x)) for every x X , then g = h. Exercise 2.1.0.2 Show that monomorphisms and epimorphisms are what youd expect them to be in (a) the categories of Groups and group homomorphisms and (b) the category of topological spaces and continuous maps. Likewise, we can dene the product of two objects X, Y in any category; it is a new object X Y and a pair of maps, X : X Y X, Y : X Y Y with a certain property holding for all other objects in the category.
HeavenForBooks.com

56

MICHAEL D. ALDER

Exercise 2.1.0.3 Construct a denition of a product of two objects in an arbitrary category. It must, of course, come out to be the same as the cartesian product in the category of sets and the category of metric spaces, and the direct product in the category of groups.

The trick in doing the last exercise is to throw away points as elements of objects, and to replace them by maps of other things into the object. In category theory, the objects have no elements in general. We dont have a lot to work with; just maps, really1. With a small amount of ingenuity, you can however dene a subobject of an object X in an arbitrary category, it is essentially some object and a monomorphism to X . Since there might be several distinct subobjects which are distinguishable abstractly but might have the same image in X , it is more practical to dene a subobject as an equivalence class of monomorphisms to X .

Exercise 2.1.0.4 Construct a good denition of a subobject of an object X in an arbitrary category. Construct a good denition of a quotient object.

If you get hopelessly stuck with the above exercises, Peter Freyds elegant book [4],Abelian Categories, may help. The exercises are actually a lot of fun. Not all of my exercises are fun, but these are. This diagram drawing convention will be used a lot in more complicated diagrams later on, where we will have objects and maps looking like this:

-P 6 6 f j A g- C
B i
1

We can even get along without objects if necessary. Think about it.
HeavenForBooks.com

Introduction to Algebraic Topology

57

In the above diagram, we often say the diagram commutes; this means that if = jg , as maps from A to P . Diagrams of this sort, and indeed much more complicated ones, will be very useful to keep track of what is going on. The next thing to do with categories is to look at superdupermaps between them. These are called functors, and we are going to study a number of them. Functors take objects in one category to objects in another category, and maps between the objects get taken to maps between the corresponding objects. I shall spend some time constructing functors for the category of topological spaces and continuous maps, T op, to the category of groups and homomorphisms, G . This will be our machinery for translating geometric, topological problems into algebraic problems. This is a truly zappy idea of great importance; once seen in abstract form, it is easy to fall in love with this notion. Mathematicians hate work2 , and this can save so much! Denition 2.1.4 If C and D are categories, we say that a covariant functor F : C D assigns to every object X C an object F (X ) D, and to every map f : X Y in C , a map F (f ) : F (X ) F (Y ) in D, so that for all pairs of maps f, g in C which compose to fg in C , we have F (fg ) = F (f )F (g )

There is another kind of functor called a contravariant functor which reverses the direction of the map, but we dont have time to look at any of them in this course, so all our functors will be covariant, and I shall just use the term functor to mean covariant functor from now on. Well, all this abstraction is quiet fun which keeps us amused, but it might be that close investigation would show that we are merely playing an obscure game with no particular point to it, such as football. And for serious blokes like me, that would be intolerable. So it is time to build a really useful functor and to use it for getting some useful results about spaces. The second part of this chapter does that with a nice little functor called 1. It takes every
2

But we dont count thinking about things as work.


HeavenForBooks.com

58

MICHAEL D. ALDER

Figure 2.1: A few loops on x topological space X to a group, 1 (X ) known as the fundamental group of the space. All this section was algebra, the next will be nice and topological.

2.2

The Fundamental Group Functor

Recall that for any topological space X , for any two points, a, b X , a path from a to b is a continuous map f : [0, 1] X such that f (0) = a and f (1) = b. This leads to: Denition 2.2.1 For any topological space X , and for any point x X , a loop based at x is a continuous map f : [0, 1] X such that f (0) = f (1) = x.

I am getting fed up with writing continuous map all the time, so take it that we are inside the category of topological spaces and continuous maps for the rest of this chapter, unless contraindicated. So map means continuous map.
HeavenForBooks.com

Introduction to Algebraic Topology

59

There is always a rather unexciting constant loop which sends everything in [0, 1] to x, but there are, in general rather a lot of loops with a given base point x; I suggest you think of a disk with a hole in it, and x as a point somewhere in the annular region. Then some of the loops will go around the hole and others wont. More accurately, some of the loops could be shrunk back to the constant loop, and some couldnt because they go around the hole at least once. In gure 2.1, there are two loops based on x, one of which could be contracted to a point and one of which could not because the hole gets in the way. We care about holes in things, so we need to distinguish these kinds of loops. This leads to: Denition 2.2.2 If two maps between topological spaces are given, f, g : X Y , A homotopy between f and g is a map F : X [0, 1] Y such that x X, F (x, 0) = f (x), & F (x, 1) = g (x) If you think of the interval [0, 1] as being the time, then you start o at time t = 0 with f , and wind up at time t = 1 with g , and you make the transition continuously. That is to say, you are continuously deforming the map f into the map g . Exercise 2.2.0.5 Show that homotopy is an equivalence relation between the maps from X to Y . Denition 2.2.3 If there is a homotopy between f and g , we write f We need a slight variation: Denition 2.2.4 If X is a topological space and A is a non-empty subspace, and if Y is another space and B Y is a nonempty subspace, and if f : X Y is a continuous map such that f (A) B , then we say that f is a map between the pairs (X, A) and (Y, B ).
HeavenForBooks.com

g.

60

MICHAEL D. ALDER

Note that there is a category of pairs of spaces. If the second part of the pair is empty, we get the original space back again to all intents and purposes. This gives a functor from the category T op to the category P airs, of pairs of spaces. Denition 2.2.5 Suppose that both f, g : (X, A) (Y, B ) are maps between pairs of spaces, and suppose that there is some subspace X X such that f and g agree on X , i.e. x X , f (x) = g (x). We say that f is homotopic to g relative to X i there is a continuous map F : (X, A) [0, 1] (Y, B ) such that F (, 0) = f ; F (, 1) = g ; & x X t [0, 1] : F (x, t) = f (x) = g (x) Denition 2.2.6 If F is a homotopy relative to X between f and g , we write f f |X |. Sometimes we omit the |X | when it is clear that the homotopies considered are all relative to the same subspace X of X . This means we can now talk about homotopies between loops in a sensible way: we want X to be [0, 1] with A = {0, 1}, the end points. We want Y to be the space we are going to have loops in; and we want B to be the point of the space where the loops are based. And we want homotopies always relative to X = A which take one loop to another. Of course, the same language can be used to talk about lots of other useful things. Exercise 2.2.0.6 Let f : [0, 1] [0, 1] be any path from 0 to 1, and let g be the constant map which sends [0, 1] to 0 in [0, 1]. Show that these maps are homotopic. Example 2.2.0.1 For any points a, b R, let f : [0, 1] R be a path from a to b. Let g be another. Show that there is homotopy of f to g relative to {0, 1}. Solution: Dene F (x, t) = tg (x) + (1 t)f (x). Then this is a continuous map from [0, 1] [0, 1] into R since it is a sum of products of continuous maps. It has
HeavenForBooks.com

Introduction to Algebraic Topology

61

the property that F (0, t) = tg (0) + (1 t)f (0) = a, and F (1, t) = tg (1) + (1 t)f (1) = b, for any t [0, 1]. So it is a homotopy relative to the end points, as required. Note that a minor textual alteration makes this work for any maps f, g : n [0, 1] R for any n. Moreover, a, b do not have to be distinct. Denition 2.2.7 A space X is said to be contractible i the identity map is homotopic to a constant map. Exercise 2.2.0.7 Show that the unit ball in Rn is contractible. Exercise 2.2.0.8 Show that any convex subset of Rn is contractible. Exercise 2.2.0.9 Do you think that the unit circle or a sphere might be contractible? Denition 2.2.8 If f : X Y and g : Y X are maps between spaces, and if gf is homotopic to IX and fg is homotopic to IY , then we say that X, Y are of the same homotopy type. Exercise 2.2.0.10 Show that homotopy type is an equivalence relation on topological spaces. Show that homeomorphic spaces have the same homotopy type. Note that if X and Y are homeomorphic, they are certainly of the same homotopy type. The converse is false, since a contractible space has the homotopy type of a single point space. And these spaces have dierent cardinality so cannot be homeomorphic, except when they are both points. Exercise 2.2.0.11 Show that the unit circle has the same homotopy type as R2 {0}, the punctured plane.
HeavenForBooks.com

62

MICHAEL D. ALDER

We are now ready to dene the functor 1 . Denition 2.2.9 1 , the fundamental group Let X be a space, and x some xed point of X . We have that homotopy between the loops at x relative to the end point [0, 1] of the interval is an equivalence relation, trivially. Let be a loop, and let [ ] denote the class of all loops homotopic to . All homotopies are taken relative to the end points. Let 1(X, x) be the set of homotopy equivalence classes. If 1 is a loop at x and 2 is another, then we dene 1 2 to be the loop 3 given by t [0, 1/2], 3 (t) =
1 (2t); t

[1/2, 1], 3 (t) =

2 (2t)

We refer to this as the composition of loops. The composition of loops is easily visualised: think of t as the time. Loop 1 takes you on some circuit that brings you back to your starting point after one second. Loop 2 is a dierent path, but starting out at time zero, you get back to your starting point at time 1 second. All loops take exactly a second to traverse. So the composite loop does the rst one and then the second one. In order to get it done in a second, you have to travel twice as fast as the other guys. Note that there is nothing commutative about this. Reversing the order of the loops traversed may make a dierence. Proposition 2.2.1 The set 1 (X, x) is a group under composition of loops. Proof. First we have to note that if 1 1 and if 2 2 , then we have that 1 2 1 2 . The homotopies are all relative to the ends of [0, 1]. The details are left as an exercise for the diligent student. The composition is associative. ( 1 2 ) 3 goes around loop 1 in time 1/4, then it goes around loop 2 in time 1/4, and nally it goes around loop 3 in time 1/2. 1 ( 2 3 ) goes around loop 1 in time 1/2, and the remaining two loops in time 1/4. Speeding up the traverse of loop 1 while slowing down that of loop 3 is easily taken care of by an obvious homotopy.
HeavenForBooks.com

Introduction to Algebraic Topology

63

The constant map sending [0, 1] to x is in the class of the identity, and whenever f : [0, 1] X is a loop at x, the map f [0, 1] X dened by f (t) = f (1 t) gives a map which is an inverse to f with respect to the composition.

Exercise 2.2.0.12 Construct an explicit homotopy to show associativity. Construct an explicit homotopy between any loop folloed by its inverse and the constant map. Exercise 2.2.0.13 Suppose you calculate the group 1 (X, x) and I calculate the group 1 (X, y ) where x, y are dierent. Show that if X is path connected, your group will be isomorphic to mine. For obvious reasons, we are only concerned with the fundamental group of path connected spaces. If a space is not path connected, we have to look at the separate path components, since the existence of a path between points gives an equivalence relation. The above exercise gives us some justication for writing 1(X ) as a group whenever X is path connected. So we shall. Now the happy thing about 1 is that it is a functor! That is to say, it not only takes spaces to groups, it takes maps between them into homomorphisms. More precisely, let PCT denote the category of path connected topological spaces and continuous maps, and G the category of groups and group homomorphisms. Then: Proposition 2.2.2 1 : PCT G is a functor. Proof. Let f : X Y be a map in PCT , and let 1 (X ) and 1(Y ) be dened as above, with any choice of x X , and y = f (x) Y . Then any
HeavenForBooks.com

64

MICHAEL D. ALDER

loop in X is a map from [0, 1] to X with (0) = (1) = x. And f is a map , with homotopy from [0, 1] to Y with f (0) = f (1) = y . Moreover if F , then f f with homotopy f F . So each homotopy equivalence class of loops in X is taken to a well dened homotopy equivalence class of loops in Y . In other words, 1(f ) : 1(X ) 1 (Y ) is a well dened map. It remains to show it is a homomorphism of groups. It is immediate that it takes the identity loop class to the identity loop class; given loops 1 , 2 in X, we get f 1 and f 2 as loops in Y. The composite of these loops, as dened for any composite of loops, just goes around the rst loop at double speed, leaving time to go around the second loop at double speed before time runs out. This is precisely what we get if we take the composite loop of 1 with 2 , and then follow it by f .

It sounds more impressive if you say it in algebra: Exercise 2.2.0.14 Write out the above proof formally and rigorously so as to impress those who havent gured out the idea. Exercise 2.2.0.15 Show that the identity map I : X X for any path connected topological space X is taken to the identity map on 1 (X ).

Exercise 2.2.0.16 Show that if X f

-Y

-Z

is in the category PCT and if 1 takes it to:


HeavenForBooks.com

Introduction to Algebraic Topology ) - (Y ) (g (Z )


1 1 1

65

1(X )

1(f )

Then 1 (g f ) = 1(g )1 (f ) This exercise completes the last link in the chain. So at last we have a functor. Proposition 2.2.3 If the space X is contractible, the fundamental group is the trivial group. Proof. It is clear that if X amd Y have the same homotopy type, then 1(X ) and 1(Y ) are isomorphic, since the identity is taken to the identity. And if X is contractible, then there is a homotopy equivalence between X and a point of X . Finally, 1({x}) for any point {x} is the trivial group.

Corollary 2.2.3.1 1(B ) = {0} for any ball B in Rn . If a space is a cartesian product of two path connected spaces it is path connected, and a loop in the product space is a pair of loops in each component. The homotopies between loops also go componentwise. This and a small amount of thought leads to the following result: Exercise 2.2.0.17 Show that 1 (X Y ) is isomorphic to 1(X ) 1 (Y )
HeavenForBooks.com

66

MICHAEL D. ALDER

We next prove the useful result that says that 1 (S 1) = Z. This is intuitively obvious, but needs a certain amount of work. Because the technicalities tend to obscure the ideas, I give a very informal proof; your job is to carefully ll in the details and translate it into algebra. The algebraic proof can then be shown to analysts, algebraists and applied mathematicians to show how hard algebraic topology is. A fully comprehensive proof will be at least four times as long, and will contain lots less English and no pictures. Theorem 2.2.1 1 (S 1) = Z. (Sketch) Proof. Let f : [0, 1] S 1 be a continuous map. Without loss of generality, we suppose that f (0) = f (1) = (1, 0). The map exp : R S 1 which was dened earlier and has exp(t) = (cos(2t), sin(2t)), is called a covering map, and R is called a covering space for S 1 . We consider the diagram:

; exp ; ;; - ? [0, 1] S f
f
1

so that the diagram commutes, The idea is to lift the map f to the map f = f . We do not expect that f (0) = f (1), but since f (1) = i.e. so that exp f (1, 0), we must have that the value of f (1) is an integer. We may reasonably (0) = 0; this will make f (1) the winding number of f , which insist that f characterises the homotopy eqivalence class of f . We rst cover S 1 by two open sets, U and V , dened by: U = {(x, y ) S 1 : y > 1/10} ; V = {(x, y ) S 1 : y < 1/10} so that U covers the top half of the circle and a little bit over, and V the bottom half and a little bit over. U and V are obviously homeomorphic to open intervals, and can be lifted by exp1 to intervals in R. exp is not, of
HeavenForBooks.com

Introduction to Algebraic Topology

67

course, a homeomorphism, but it is one locally, meaning that if we restrict the domain suitably it becomes one onto its range. There are a lot of disjoint intervals in exp1 (U ). We now take f (0) = (0, 1) as our starting point and chop up the domain of f into closed subintervals {[0, t1], [t1, t2], [ti, ti+1 ], [tn1 , 1]} so that each interval has its image in either U or V but not both. [0, 1] We do this by noting that f (0) U , so there is some smallest t, t such that f (t) is not in U , or f ([0, 1]) U . In the latter case, f is clearly homotopic to the constant map sending everything to (1, 0) by a contraction of the interval U . ) has a negative x value or it has a In the more interesting case, either f (t positive x value, since we can easily see that it is in the closure of U and not in U , which gives only two possible locations. If it has a positive value, either , or it has remained in the component it has left V at some lower value of t < t of U V containing (1, 0). In the latter case, we restart the argument with ) has V instead of U . In the former case, we put t1 = t. In the case where f (t . We repeat starting with t1 and continue negative x value, we also put t1 = t until we nish with the value of f (1) which we know is back at (1, 0). The fact that this can be done in a nite number of steps follows from the compactness of [0, 1] and the Lebesgue covering theorem, since {f 1 U, f 1 V } is a covering of a compact metric space and has a Lebesgue number. So chopping up the unit interval into smaller regions than this is guaranteed to produce a partition of [0, 1] with every piece either in U or V , and our partition can be obtained by joining together consecutive bits of this Lebesgue partition. This means that we have an alternating sequence: it is not clear which one we start with, but the intervals of the partition alternate, so that the interior points are alternately sent to U and V . We can now lift each of these intervals to R. We do this with a careful choice is of which possible interval to lift U or V to. The aim is to ensure that f continuous: The starting point is going to lift to 0 in R, and if the image of [0, t1) is in U , and if f (t1 ) has negative x value, then the sheet of f 1 (U )
HeavenForBooks.com

68

MICHAEL D. ALDER

that we are interested in will be from a little less than zero to a little more (t1) to be a little more than 1/2 too, by continuity. than 1/2. This forces f (Any other possible choice would dier from this by an integer, so it isnt going to be possible to make a mistake here.) If, on the other hand, f (t1 ) has (t1) is a small negative number, about 1/10. Similarly if positive x, then f (t1 ). the image of [0, t1) is in V . Either way, we nail f (1) = n, We now move onto the next interval, and repeat until we nish with f for some integer n Z. g . We need to Suppose that g : [0, 1] S 1 is another map, and that f show that g lifts to a map g which has g (1) = n for the same n. We can do this by lifting the homotopy. The homotopy is a map F : [0, 1] [0, 1] R and the Lebesgue covering lemma applies to this space too. We have to chop the unit square up into bits so that each bit is a local homotopy between maps which go into U or into V , and an earlier problem showed that we can do to this. Joining up the homotopies of each subinterval to get a homotopy of f g which leaves the end points xed is messy but conceptually straightforward. The obvious complication in homotopy piecing is that the partition elements are dierent for the two functions. I leave you to work out how to deal with this issue. , g The last part is to show that if f, g lift to f , and if the two maps have the same value of n = f (1) = g (1), we want to show that f g . Now a homotopy and g between f which leaves the end points xed is easy to construct in R by an earlier exercise, and we merely follow such a homotopy by exp to get a homotopy between f and g . It is (fairly) clear then, that for each map there is an integer which corresponds to its homotopy class. So we assign each homotopy class to this integer, to get a map from 1(S 1 ) to Z. We have shown that this is a well dened map of homotopy classes, and also that it is 1-1. That it is onto is trivial. Finally, the composite of loops goes to the sum of their winding numbers. Hence 1(S 1 ) = Z.

HeavenForBooks.com

Introduction to Algebraic Topology

69

Exercise 2.2.0.18 Fill in the details. This is a long job, but will develop your condence something wonderful.

We now make use of the last result to prove the Brouwer xed point theorem in dimension two. This will need a few intermediate steps. Denition 2.2.10 If i : A X is an inclusion map of a subset A in X , a retraction of X onto A is a map r : X A which leaves A xed, that is, r i = IA . Exercise 2.2.0.19 Show that there is a retraction of the punctured plane, R2 {(0, 0)} onto the unit circle. Topological intuition assures us that there is no retraction of D2 onto its boundary, S 1. Such a map would surely have to tear the skin of the drum somewhere, and so would not be continuous. On the other hand, actually proving that a particular kind of map, among all the huge variety of maps, does not exist is rather hard. It isnt easy for non-topologists to see how to even start on the problem. We can do it however using functors, and in this case the functor 1 will do the job. Proposition 2.2.4 There is no retraction of D2 onto its boundary, S 1 . Proof. If S1 i r

-D

is a diagram showing the hypothetical retraction, then the functor 1 takes it to a diagram
HeavenForBooks.com

70

MICHAEL D. ALDER Z i r

- {0}

Where I have written i for 1(i) and r for 1 (r), and have used the fact that D2 is contractible and hence has trivial fundamental group. Now we have that r i = IS1 hence that ri = IZ . This asserts that we have Z i

- {0}

-Z

such that r i = IZ . But the identity map on Z cannot factor through the zero map. So no such diagram is possible. So the map r cannot exist in the topological category; no retraction of the disk onto its boundary exists. This is rather an impressive result if you think in terms of the diculty of proving it. The fact that it is intuitively obvious is neither here nor there. Whether intuitively obvious things have proofs is not particularly intuitively obvious. This one does. You may now cheer and throw your hats in the air. Finally, let f : D2 D2 be a continuous map. We say that f has a xed point a i f (a) = a. We show that every such map has a xed point. Theorem 2.2.2 (Brouwer, special case) Every continuous map f : D2 D2 has at least one xed point. Proof. Suppose not. Then we can construct a retraction of D2 onto its boundary by taking every x D2 and the (distinct) point f (x), and joining them by a straight line. This line intersects the boundary of D2 , S 1 in two points, choose that point p so that p, x, f (x) are in that order. Now dene r(x) = p. This is clearly continuous and leaves S 1 xed, so is a retraction of D2 onto its boundary. But there is no such retraction. Hence there is no map without a xed point.
HeavenForBooks.com

Introduction to Algebraic Topology

71

Exercise 2.2.0.20 The hypothetical retraction r(x) was said to be clearly continuous. How clear is it? Exercise 2.2.0.21 I claim 1 (S 2) = {0}. If f : [0, 1] S 2 s a map with f (0) = f (1) = (1, 0, 0) S 2, and f is not onto, choose a point p S 2 not in the image, remove it, and contract S 2 {p} to a point, giving a homotopy equivalence between the punctured sphere and a point. If f is onto S 2 , take the point (1, 0, 0)in S 2 and draw a little circle around it, of radius, say, 1/10. Find the rst value of t [0, 1] such that f (t) is on the boundary of the circle, and the next value of t for which it leaves the boundary circle. These may be the same. Now make a homotopy which pushes the part inside to the boundary of the circle. Repeat for all subsequent pairs of points. There are at most a nite number of these by lebesgue again. The result is a map, homotopic to the original map, which does not have the centre of the circle in the image. So we can contract it to the constant map. This sloppy argument shows that 1(S 2 ) = {0}. Make the above argument rigorous. Generalise it to S n , n 3. Use it to prove that a coee cup and a beachball are denitely dierent. Exercise 2.2.0.22 I claim that R is not homeomorphic to R2 . Suppose it were, by a map f : R R2 . Remove a point of R, p, and also remove f (p). Then the homeomorphism restricts to a homeomorphism between a punctured plane and R {0}. But the former is path connected and the latter is not. So the original homeomorphism cannot exist. Make the above argument rigorous. Denition 2.2.11 A space X is said to be simply connected if it is path connected and has trivial fundamental group.
HeavenForBooks.com

72

MICHAEL D. ALDER

Exercise 2.2.0.23 Show that R2 and R3 are not homeomorphic.

Exercise 2.2.0.24 Recalling the construction of RP2 , try making it as a subset of R3 as follows: join together one pair of opposite points on the boundary of the northern hemisphere, otherwise known as the equator. Now, joining all the other points on the two circles is impossible because the rest of the hemisphere gets in the way. So chop out a big disk centred on the north pole, to give you a band containing the equator and going up to latitude 20o or thereabouts. You can now successfully do all the gluing. Try it with a big strip of paper if you dont believe me. The resulting object is well known to be a M obius strip. The boundary of the M obius strip is just the bounding circle of the disk you cut away. It got a bit twisted when you did the gluing. It follows that if you take a M obius strip and measure the perimeter, and then you get a disk with the same perimeter, and then you glue the two bounding circles together, you will get RP2 back. Get your mother to knit you one for your birthday. Suppose you had a map of the circle into RP2 which went around the M obius strip part once. It is intuitively plausible that this is not contractible. (Your intuitions may be developed more if you do the knitting.) On the other hand, if you were to map a circle into the M obius strip part that went around twice, you could push it gently but rmly onto the boundary of the strip. From which you could collapse it to a point over the disk part of RP2 . This leads to the conjecture: 1(RP2 ) = Z2 This is in fact true, but not entirely easy to prove rigorously.

Exercise 2.2.0.25 Take a gure eight shape consisting, topologically, of two circles glued together at a single point. What would you expect its fundamental group to be?
HeavenForBooks.com

Introduction to Algebraic Topology

73

2.3

Summary

This chapter has introduced a very powerful idea with a lot of consequences and ramications; enough to make your hair curl. First we noted that we kept meeting dierent categories, so we dened the abstract idea of one, making it rather easy to see that we already knew at least a dozen diernt categories. Then we introduced the idea of a functor between categories. This was all abstract algebra, and you could either take it on faith that there was some sensible reason for doing so, or just enjoy it as good clean fun. In the second part, you saw that there was a good reason for the abstraction, it allowed us to discuss particular functors and, with a little work, to construct 1 . Computing it for any path connected space is a fairly terrifying prospect, but computing it for a few simple cases is not too bad, and once done, some nice results come out, results which are hideously messy if done almost any other way. You are now in a position to see more clearly what I meant when I said that algebraic topology is about translating geometric problems into algebraic ones. Functors are the translation machine. So far, you have only met one such functor, and the problems of calculating its values are formidable. If we take the pair of spaces (S n, ), where is a particular point, and look at the homotopy class of maps from this into a pair (X, ), where the second is a dierent point in general, we can get the higher homotopy groups, n(X, ). In other words, there is an innite collection of functors, all giving some information about the holes of various sorts in spaces. This is rather vague, and we do not pursue it further. There is another family of functors, the Homology functors, Hn , which go from a reasonably large class of spaces into the category of abelian groups, and also do something to nd holes in the space. This is a nicer category to work in, and we can go a long way to setting up a machine for actually computing all the homology groups. The general idea is exactly as for 1, but we can tackle higher dimensions without a qualm, and do the calculations much more easily than for the fundamental group. For the history and some background information which explains the terHeavenForBooks.com

74

MICHAEL D. ALDER

minology, the sixth chapter of Masseys book [2] is strongly recommended, particularly for those with some background in Calculus on Manifolds. We now proceed to establish the machinery of Homology Theory. It should be noted that there are several versions of Homology theory, all with their advantages and drawbacks. I follow Massey in going through the singular theory for cubes, partly because although the abstraction is greater, it relates better to the Calculus on Manifolds material you have done at an earlier stage of your education (I hope), and it is probably quicker to get to the main computational machinery. There is a tradition among Pure Mathematicians of intimidating Physicists, Engineers and others by not explaining the geometry of what you are doing, and just giving long complex algebraic proofs. In the case of algebraic topology, the scope for terrorising them is too great for it to be sporting. The trouble then is that if you explain the geometry to them, they assure you that it is all obvious. If this happens, ask them to tell you the homotopy groups k (S n ). Trying to intimidate Psychologists or Arts students is a waste of time because they all think that higher mathematics means longer and longer division. Better, perhaps, not to talk to them at all, except that they go on to highly paid jobs in the Public Service and play at running the country. Convincing the intelligent ones that they are too thick to handle anything complicated is just about possible, but this only means that the unintelligent ones go on to take charge. This is a hard problem, not covered in this book.

HeavenForBooks.com

Chapter 3 Singular Homology


3.1 The Homology Groups

You are used to the idea of parametrising surfaces and regions of R3 in order to integrate dierential forms (or vector elds, if you did it the old fashioned way) over the surfaces or regions. It was usually a matter of taking a rectangle or cube and mapping it onto the surface; whereupon you pulled back the form onto the rectangle or cube using properties of the parametrising map and its derivative, and then did the much nicer integration over the rectangle or cube. I am already crapped o with having to write rectangle or cube. They dier only in the dimension: Denition 3.1.1 I = [0, 1] R. I 0 is a set containing but a single point, I 0 = {0}, and I 1 = I . n Z, n 1 I n+1 is the cartesian product I I n. From now on, map means continuous map if we are in the category of topological spaces and continuous maps or one of its obvious subcategories, and it means homomorphism if we are in the category of groups and group 75

76

MICHAEL D. ALDER

Figure 3.1: A Singular 2-cube on X homomorphisms, or any convenient subcategory thereof (such as Abelian groups). I am a kindly bloke, and will use the longer terminology occasionally just to remind you of which category we are in, but you ought to keep track of that!

Denition 3.1.2 A singular n-cube in a topological space X , is a map f : I n X . Think of X = R3, and n = 1 for a curve in the space, n = 2 for a patch of a surface. Or, simpler, take X to be a disk in R2 with a hole in it, as in gure 3.1, and think of singular 1-cubes, i.e. curves in the space, or perhaps a singular 2-cube. The term singular means stretch it, twist it, but dont tear it, and then plonk it in the space you are interested in. Dont ask me why that particular name. From now on, I shall avoid f as a name for a singular n-cube, but will use q : I n X in the general case; q for qube. I want to preserve the generic sounding f for other general maps between any old topological spaces. We are now going to start on some denitions which are, perhaps, the least motivated denitions in the book, so I shall try to make some suggestions of what to think of as we go. This is where some faith in your teacher is called
HeavenForBooks.com

Introduction to Algebraic Topology

77

for, or a deal of patience. The lucky ones will have done some of this in some earlier stage of their education, but for the rest: Denition 3.1.3 Qn (X ) denotes the free abelian group on the singular cubes in X . This means that we take formal sums of curves in the space. An element of Q1(X ) looks something like 2c1 + 3c2 4c3 where the ci are curves in the space with end points. The curves dont have to have anything to do with each other. We only have a nite number of terms in each sum, even though there are lots of possible terms to put in them. Some of the curves have collapsed into points, or the squares into line segments. We want to get rid of them: Denition 3.1.4 A canonical projection is a map : Rn Rn1 which omits one of the variables. There are n such maps. Denition 3.1.5 A singular n-cube q : I n X is degenerate i there is a singular (n 1)-cube q : I n1 X and a canonical projection : I n I n1 such that q = q . Degenerate n-cubes have had one or more of the dimensions kicked out of them, rather like Obi-Wan Kenobi after Darth Vader had seen him o. I guess that was a literary allusion. Denition 3.1.6 If Dn (X ) denotes the subgroup of Qn (X ) generated by the degenerate n-cubes in X , let Cn (X ) denote the quotient abelian group Qn(X )/Dn (X ). This is called the group of cubical singular n-chains on X .
HeavenForBooks.com

78

MICHAEL D. ALDER

Exercise 3.1.0.26 Why do you think that we factor out the subgroup of degenerate n-cubes rather than simply take the free abelian group on the nondegenerate n-cubes? Imagine that you have a singular 2-cube, that is to say a square, on X , that is to say, you have taken a square and stretched it a bit and mapped it into X , as in gure 3.1. The boundary of the square is also mapped in, and consists of four singular 1-cubes or curves. There is a problem of specifying the orientation of these edges. If we change the sign, we reverse the orientation. It is clear that any non-degenerate singular n-cube has a boundary which is a singular (n 1)chain, and more generally, a singular n-chain has a boundary which can be obtained by taking the integer multiplicities of the n-cubes in the chain and applying the same multiplicities to the boundary elements to get an (n 1)chain. Some of the bizarre aspects of these denitions may be starting, although only starting, to make some sense. A singular n-cube has, as a bit of thought will show, 2n faces which come in matching pairs, the front and back: Denition 3.1.7 Let f : I n X be a singular n-cube on X . We dene some singular (n 1)-cubes, 2n of them, as follows:

Fi (f ) : I n1 X Fi(f )(x1 , , xn1 ) = f (x1 , , xi1 , 0, xi, , xn1 ) and Bi (f ) : I n1 X Bi (f )(x1 , , xn1 ) = f (x1, , xi1, 1, xi , , xn1 )

These give the Front and Back faces for 1 i n. Fi (f ) is called the front i-face and Bi (f ) the back i-face of the singular cube f . Note that although
HeavenForBooks.com

Introduction to Algebraic Topology

79

the terminology is three dimensional, the process of nding faces makes sense at all dimensions except the zero dimensional case. Exercise 3.1.0.27 Write down the inclusion map of the unit square in R2 as your singular 2-cube. Write out explicit formulae for the four singular faces. Because it makes sense in all non-zero dimensions, we can iterate to nd faces and then faces of them. And so on until we hit dimension 0. Exercise 3.1.0.28 With the above exercise, do it again to nd the vertices at the ends of the singular edges. Apply the maps Fi Bj (f ) to 0. Also Fi Fj and the other combinations. Write down the inclusion map for the unit cube in R3 and nd the faces and the faces of the faces and the faces of the faces of the faces. Exercise 3.1.0.29 Show that for 1 i < j n, Fi Fj Bi Bj Fi Bj Bi Fj = = = = Fj 1 Fi Bj 1 Bi Bj 1 Fi Fj 1 Bi

If you can work through one, you can do all of them, because they arent very dierent. If you cannot see how to do the rst one, verify that it works on the unit cube. We may now dene the boundary of a singular n-cube; it is an (n 1)-chain: Denition 3.1.8 For any nondegenerate n-cube, q , n > 0, the boundary of q is the singular (n 1)-chain: n (q ) = (1)i [Fi(q ) Bi (q )]
i=1,n

HeavenForBooks.com

80

MICHAEL D. ALDER

Given the fact that we can extend n to arbitrary chains by simply putting in the multiplicities and adding, we get: Denition 3.1.9 The boundary operator n denes a homomorphism n > 0 n : Cn(X ) Cn1 (X ) Exercise 3.1.0.30 Obtain the boundary of the unit cube in R3 regarded as a singular cube. Exercise 3.1.0.31 Show that for all n 2, and for every singular n-cube q , n1 n(q ) = 0 Awful Warning There is a temptation to think of the boundary of a singular cube as being the boundary of the image. It is no such thing. We do not have to have a singular n cube an embedding of I n . It may be, but it does not have to be. End of Awful Warning Note that what we get as the derived chain is an oriented boundary. It is worth drawing the following diagram of the sequence of abelian groups and homomorphisms: n+1

-C

-C

n1

n1

-C

n2

-C

There are going to be a lot of diagrams looking rather like this, so you may as well get used to the easy ones rst. Now there are chains in Cn1 which are boundaries of chains in Cn, and there are chains in Cn1 which are nothing of the kind. We can see that the chains
HeavenForBooks.com

Introduction to Algebraic Topology

81

in Cn1 which are boundaries are a subgroup of Cn1 consisting of the image of n. We dene: Denition 3.1.10 Bn (X ) = im(n+1 ) = n+1 (Cn+1 ) Cn This group is called the group of singular n-boundaries of X. Now think about a chain of 1-cubes (curves) which goes around a hole in X , for some two dimensional space X ; (draw such a chain in gure 3.1). There is no possible singular 2-cube (square) or chain of 2-cubes in X of which this is the boundary. On the other hand, if you were to ll in the hole by a singular 2-cube and take its boundary, it would be such a chain. So if we had a 1-chain c which went around a hole, it would be reasonable to expect that 1(c) = 0 would hold, just as if the hole were lled in by a singular 2cube. This follows from 12 = 0 when it is applied to the imaginary singular 2-cube which lls in the hole. After all, its the same 1-chain whether there is a singular 2-cube or a hole inside it. We therefore dene: Denition 3.1.11 Zn (X ) = ker(n ) Cn This group is called the group of singular n-cycles of X. Now both Bn and Zn are subgroups of Cn for n > 0. Since n1 n(f ) = 0 we see that Bn is a subgroup of Zn . The latter represents, loosely speaking, all the things which are boundaries of either singular (n + 1)-chains or of holes or collections of holes, while the former represents all the things which are actually boundaries of singular (n + 1)-chains. We therefore dene:
HeavenForBooks.com

82 Denition 3.1.12

MICHAEL D. ALDER

n 1, Hn (X ) = Zn (X )/Bn (X ) which is called the nth Homology Group of X. This has got us to the point of having a denition, and some sort of rationalisation, for thinking that the homology groups of a space tell us something about the holes in the space. An n-hole is what you get if you take a nondegenerate singular n-cube and cut its image out of the space. The groups Cn , Bn and Zn are likely to be fairly humungous groups with an awful lot of generators. On the other hand Hn (X ) might just about be reasonable in size, which is a sloppy way of saying nitely generated. At least for nice simple spaces like gure 3.1. Such is indeed the case. Exercise 3.1.0.32 Construct a cycle in Z1 (R2 ) by choosing ane isometries of I in R2 to be the 1-cubes. Suppose I give you a chain of 1-cubes in R2 or R3 . How could you tell by inspecting them whether or not they dened a cycle? Is it true that every chain which is a cycle has to have the same multiplicity for each constituent n-cube? Exercise 3.1.0.33 Show that a degenerate 1-cube d which is a constant map has d = 0. This would make degenerate 1-cubes non-trivial cycles if we had not factored them out. Is such a degenerate cycle a boundary of anything? What eect will this have on the homology? Take the 1-chain in R2 which contains two 1-cubes q1, q2, and is just q1 q2. q1 (x) = (x, 0), and q2(x) = (x, sin(x)). Draw the chain. Show that this chain is a cycle. Show that it is not a boundary unless we factor out at least some degenerate cubes. Exercise 3.1.0.34 Take the space X to be R2 with the interior of the unit square removed. Restrict the space of allowable functions to be not the continuous functions, of which there are rather a lot, but instead the ane isometries. Take the chain of 1-cubes to be the sequence of four straight line edges
HeavenForBooks.com

Introduction to Algebraic Topology

83

which go around the removed square in an anticlockwise sense. Show this chain is a cycle. Show that it is the only cycle, except for integral multiples of it, which does not bound a singular 2-cube. Note that we havent got a denition for H0 , since there is no following homomorphism to have a kernel. We have some freedom to dene it in several ways of which two are standard. Probably the simplest is to take the diagram n+1

Cn

-C

n1

n1

-C

n2

-C

{0}

and extend it to the right by putting in lots of zero groups, Ci (X ) = {0}, i < 0, as shown. All the n for n 0 are the zero map, Z0 = C0 and so: Denition 3.1.13 H0 (X ) = C0(X )/B0 (X ) Another approach is to dene a homomorphism called the augmentation : C0 (X ) Z by dening (q ) = 1 for any 0-cube (point) q . This makes C1 = Z, Ci (X ) = {0} if i < 1. There isnt any choice of what to make the homomorphisms past , which sends a 0-chain, which is a nite set of weighted points with integral weights, to the sum of the weights. It is clear that 1 = 0 when applied to any 1-cube, and hence applied to any chain. We now dene Denition 3.1.14 0 (X ) = ker() Z
HeavenForBooks.com

84 and hence:

MICHAEL D. ALDER

0 (X )/B0 (X ) 0 (X ) = Z H which latter is called the reduced 0-dimensional Homology Group of X. Exercise 3.1.0.35 Is the empty space path conncted? Example 3.1.0.2 Compute all the homology groups for the space consisting of a single point, p. Do it for the (hausdor) space consisting of two points. Solution Any singular n-cube mapping into p must be degenerate for n 1, so Cn = 0 if n 1, and hence Hn (p) = 0 for n 1. We need only look at the sequence: 3

- {0}

- {0}

C0

{0}

or the sequence 3

- {0}

- {0}

-C

{0}

for the reduced zero dimensional homology. For the rst case, C0 = Z0 is the set of all sums of n-cubes, and there is only the one 0-cube, q : {0} p, which is not degenerate. So every chain is of the form kq for some integer k . In other words, there is a single generator, and C0 = Z0 = Z. There are no boundaries, since every map into p is degenerate. Hence H0 (p) = Z up to isomorphism.
HeavenForBooks.com

Introduction to Algebraic Topology 0 (p) = {0}, the trivial group. By the same argument, H

85

For X = S 0 , a two point space, there are two possible 0-cubes, and the group C0(S 0 ) has two generators, and is Z Z, which is the same as H0 (S 0). Exercise 3.1.0.36 Compute H0 (I ) from the denition. The next few sections will rst show that Hn is a functor, and then that it has some rather convenient properties. It is the properties that matter, because we shall use them to establish some rules for computing homology of new spaces in terms of the known homology of simpler spaces; then armed with the results we already have, we shall be able to compute some useful results. All the results will come in a rush at the end, after a lot of eort spent in getting to the critical point. Rather like sex, really.

3.2

Hn is a functor

We have shown that for every non-negative integer n, there is an nth homology group. The next thing is to show that each Hn is actually a functor. That is to say, if f : X Y is a continuous map between spaces, Hn (f ) : Hn (X ) Hn (Y ) is, for any non-negative integer n, a homomorphism of abelian groups. This, I am pleased to tell you, is fairly easy. Denition 3.2.1 If f : X Y is a continuous map between spaces,and if q : I n X is a singular n-cube in X , then f q : I n Y is a singular n-cube in Y . And if q is degenerate, so is f q . It follows that f induces a homomorphism of the chain groups: f#,n : Cn(X ) Cn(Y )
HeavenForBooks.com

86

MICHAEL D. ALDER

for every non-negative integer n. Remember, getting from singular cubes to chains is the most trivial of matters: we just take nite integer linear combinations of some of the singular cubes, and thats a chain. That f#,n is a group homomorphism follows because we have simply dened it as a set map on the generators and gone to the unique extension. Proposition 3.2.1 The following diagram commutes for every n 1: Cn (X ) f#,n
n

- C (Y )
n

Cn1 (X ) f #,n1

? - C (Y )
n n1

Proof If you take a singular n-cube q in X , regard it as a rather insipid chain in Cn (X ) ( 1 q), map it into f#,n (q ), a rather insipid chain in Cn (Y ), then you observe that the boundary in our sense is taken to the boundary. (Not, of course, having anything to do with the boundary of the image set!) Each of the front and back faces are taken to front and back faces of the cube f#,n (q ). Go back to the denition of the face of a singular cube if you dont believe me. The proposition for general chains just consists of formally multiplying each cube q by an integer and adding it to a nite set of others, which have also been multiplied up by an integer. All the numbers pass through the system in a thoroughly uninteresting way. The result follows.

Exercise 3.2.0.37 Conrm the above rather imsy argument by doing all the algebra in a god-fearing and respectable manner.
HeavenForBooks.com

Introduction to Algebraic Topology Note the use of diagrams to express f#,n1 n = n f#,n neatly and perspicuously. It follows that:

87

Proposition 3.2.2 If c is a cycle in Cn(X ), i.e. c Zn (X ), then f#,n (c) Zn(Y ). Proof. We do what is called diagram chasing. Starting in the left hand corner with c and going down rst, n (c) = 0 Cn1 (X ), by denition of Z . And f#,n1 is a homomorphism and so takes 0 to 0. This means that f#,n (c) ker(n) = Zn(Y ), by going around the diagram the other way.

Here is another simple diagram chase: Exercise 3.2.0.38 Show that f#,n maps Bn (X ) into Bn (Y ). More complicated diagrams give the potential for more exciting diagram chases. We shall do at least one rather nice one later. From the last proposition and the last exercise, it follows that we can take the quotient of f#,n . Proposition 3.2.3 There is a homomorphism f,n : Hn(X ) Hn (Y ) induced from f#,n . Proof.
HeavenForBooks.com

88

MICHAEL D. ALDER

We dene f,n by taking an element of the quotient of Zn (X ) by Bn (X ), which is going to be represented by a chain c in Zn (X ). Then we take the image of this chain by f#,n , or equivalently, we compose with f for all the cubes in the chain and multiply by the multiplicities, to get a chain in Zn (Y ). If we had taken another representative of Hn (X ), c , we have that c c Bn (X ). Hence f#,n (c c ) Bn (Y ) So whatever representative of the equivalence class we choose, we get equivalent representatives in Zn (Y ). In other words we get a well dened element of Hn (Y ). This gives us the required group homomorphism. It is a homomorphism because f#,n is. The map f,n is the required Hn (F ). It takes up less space in diagrams if we write f,n rather than Hn (f ), and even less to simply write f for all the f,n . This means you have to keep track of which n we are talking about by looking at Hn (X ). Saving typing is very appealing to me just at present. Exercise 3.2.0.39 Show that the above argument goes throught for H0 without any changes. 0. Show that the result also holds for H Exercise 3.2.0.40 Show that if f : X Y is the identity map, then so is (are?) f . Exercise 3.2.0.41 Show that f g = (fg )

3.3

Homotopy

One of the properties that we want to establish is that: f g f = g

HeavenForBooks.com

Introduction to Algebraic Topology

89

This will be the rst of the steps needed to be able to compute the homology of spaces from knowing the homology of simpler spaces. This is a fairly easy one. We do it by carrying the homotopy over to the chains. We take, initially, Cn(X ) = {0} for every n < 0. This means that we can dene a chain complex which is a simple abstraction of all the things we might conceivably get as chain groups Cn(X ) together with the boundary operator: Denition 3.3.1 A chain complex is a set of abelian groups, {Cn }, one for each integer n, and for each integer n a map n : Cn Cn1 such that n Z, n1 n = 0 We shall leave out the from now on and write 2 = 0 for the above condition. Denition 3.3.2 A homomorphism of degree k between chain complexes {Cn, C } and {Dn , D } is a family of group homomorphisms fn : Cn Dn+k Denition 3.3.3 A chain map between chain complexes {Cn, C } and {Dn , D } is a family of group homomorphisms: fn : Cn Dn which commute with . This makes a chain map a particular sort of homomorphisms between chain complexes: it has degree zero and commutes with . A diagram of a chain map f is: n+1

-C

-C

n1

n1

-C

n2

fn n+1

fn1 n

fn2

? -D
n

? -D

n1

n1

? -D

n2

HeavenForBooks.com

90

MICHAEL D. ALDER

where all the squares commute, and I have used the same symbol in both chain complexes because there is no serious risk of confusion and it cuts down on the typing. Such chain maps arise, of course, from continuous maps between spaces. It is clear that there is a category of chain complexes and chain maps. It is also clear that there is a functor from topological spaces and continuous maps to this category, obtained by taking the free group on the n-cubes and factoring out the degenerate cubes. It is also clear that given a chain complex there is no particular diculty in dening a homology sequence. This allows us to consider taking the homology of other things besides spaces, but that is another story. If I have two maps f, g : X Y , they will generate two maps between the chain complexes C (X ) and C (Y ). In the following diagram, I have slithered to the category of chain complexes and taken two chain maps from one chain complex called C to another called D.

Cn+1

n+1

-C
fn

-C

n1

gn n

Dn+1

? ?

n+1

?? -D
n

? ? -D n1

I have also called these f, g , because they are just chain maps, and any connection with the maps between spaces is just your subconscious throwing up possibilities. Maybe the chain map f is actually h# for a continuous map h, and maybe it isnt. Of course, if itwere so obtained, we might ask the question, what relation should hold between the chain maps, if the original continuous maps were homotopic? There ought to be a chain homotopy between chain maps. It seems reasonable that there is something special about two chain maps which arise from homotopic maps. There must be some sort of relation between them. What should a chain homotopy of f to g , two chain maps of degree 0 be? It is plausible that it should be a homomorphism of degree 1, since a homotopy
HeavenForBooks.com

Introduction to Algebraic Topology

91

Figure 3.2: A Singular 2-cube on Y which is a homotopy between maps will lead to a homotopy between n-cubes in Y which is an (n + 1) cube in Y . (Think of two curves in a subspace of R2 , and a homotopy lling in the square, as in gure 3.2. The homotopy between the 1-cubes is a 2-cube.) The value of the homotopy at 1/2 is shown as a dotted curve. If we take q to be the standard 2-cube in R2 , by way of example, and if 3 f, g : R2 R are two homotopic maps, then we can regard the homotopy between f, g of q , as a mapping of a 3-cube into R3. If is the homotopy, then (q ) is the cube shown in gure 3.3. The boundary of this cube is (q ). It consists of a top part, f (q ), a bottom part with the opposite orientation, g (q ), and the empty box with the top and bottom removed, which is the Homotopy restricted to the boundary of q , with a minus sign to get the orientation right. Thus we get the formula: = f g We observe that this is intelligible as ocurring wholly on the chain complex, so it may be turned into a denition: Denition 3.3.4 A chain homotopy between a pair of chain maps f, g : C D is a chain homomorphism : C D of degree 1 such that: n Z, n+1 n + n1 n = fn gn
HeavenForBooks.com

92

MICHAEL D. ALDER

Figure 3.3: A Singular 3-cube in R3 which is a homotopy Drawing such a chain homotopy by adding to the above diagram:

-C -C ; ; ; ; ; ; ; f g ; ; ; ; ; ? ?; ??; ? ? - D -D -D Cn+1 n+1


n n n1 n n1 n n n+1 n n+1 n n1

Note that we do not have the triangular parts of the diagram commuting, in general. There are, after this sequence of denitions two things left to establish. The rst is that if there is a chain homotopy between chain maps f# , g# , then when we take the quotients to obtain the homology groups, the induced maps from f# and g# , namely f and g , on the homology groups, are the same. And the second thing to show is that a homotopy between f, g : X Y in the category of topological spaces and continuous maps, gives rise to a chain
HeavenForBooks.com

Introduction to Algebraic Topology

93

homotopy between f# , g# : C (X ) C (Y ), using the obvious notation. I write f# c g# to indicate that there is a chain homotopy between f# and g# As usual, a horrible sequence of denitions has been given with the aim of making life easy in the end. The problem is that we are actually working on something rather complicated. Stokes Theorem becomes rather trivial if the denitions are right, and just a muddle which it is hard to see how to generalise if they arent, so some time devoted to dening the issues is well spent. (If you are sadistic, try asking an old fashioned applied mathematician for a generalisation of Stokes theorem to four dimensions, and whether it actually holds there. Fuzziness and the warm glow that comes from an intuitive grasp of a theorem are a good start, but not much use on their own.) I have spared you the task of working out from rst principles what the denition of a chain homotopy ought to be: what it ought to be is something that (a) is always induced from a homotopy and (b) gives the result that going to the homology groups gives us the same homomorphism.

Proposition 3.3.1 If there is a chain homotopy between chain maps f, g : C D, then the induced maps of the homology groups are the same. Saying it in algebra: f
c

g f = g

Proof. Recall that n Z, n+1 n + n1 n = fn gn

Its a simple diagram chase, follow c carefully:

HeavenForBooks.com

94

MICHAEL D. ALDER c

-C -C ; ; ; ; ; ; f g ;; ; ; ; ; ? ?; ??; ? ? - D -D -D Cn+1 n+1


n n n1 n n1 n n n+1 n n+1 n n1

Write ZC , ZD for the kernels of the maps and BC , BD for the images. Take any chain c ZC n . Travelling right, we see that c = 0 by denition of ZC , and so n1 n (c) = 0, from which we deduce that n+1 n (c) = fn (c) gn (c) Now by denition of BD , we observe that n+1 n (q ) BDn , and so when we take equivalence classes, f,n (c) = g,n (c). Hence f = g .

Proposition 3.3.2 A homotopy between maps f, g : X Y induces a chain homotopy between f# , g# : C (X ) C (Y ) Proof. Let : I X Y be a homotopy between g, f : X Y . Note that it is convenient to put the I rst rather than second, and that it makes no dierence since the two products are homeomorphic. Note also that I have taken a homotopy from g to f , for reasons to do with signs of the terms. First we dene a map on the free group of singular cubes, hn : Qn(X ) Qn+1 (Y ) by hn(q )(x1, x2, , xn+1 ) = (x1 , q (x2, x3 , , xn+1 ))
HeavenForBooks.com

Introduction to Algebraic Topology

95

Now this makes hn (q ) a singular (n + 1)-cube on Y , and we can look at the front and back faces of this cube: F1(hn )(q ) = g# (q ) B1 (hn)(q ) = f# (q )

Fi(hn )(q ) = hn1 Fi1 (q ), 2 i n + 1 Bi (hn )(q ) = hn1 Bi1 (q ), 2 i n + 1 Recalling the expression for the operator, this gives: n+1 (hn )(q ) = f# (q ) g# (q ) +
j =1,n

(1)j +1 hn1 (Fj (q ) Bj (q ))

= f#(q ) g# (q ) hn1 n (q ) It therefore follows that for any singular n-cube q , (f#,n g#,n )(q ) = (n+1 hn + hn1 n )(q ) If q is degenerate, so is hn (q ), so we can extend hn on singular cubes to Hn on chains. This provides the required chain homotopy between f# and g# .

Exercise 3.3.0.42 Work through the above argument very carefully for the case of 1-cubes (curves) in the space of gure 3.2 for the space Y , and, say, the unit square in R2 as X . If you want, you can make the homotopy and hence the maps f, g particular concrete maps. Or you can leave them general. I suggest you get down to a level of concreteness with which you feel comfortable, and then do enough concrete examples to see the general pattern emerging. If still confused, work through a similar three dimensional example. Keep going until the proposition is obvious.
HeavenForBooks.com

96

MICHAEL D. ALDER

The two results above lead to the nal result: Theorem 3.3.1 If two maps f, g : X Y are homotopic, then n 0, Hn (f ) = Hn (g ), i.e. f = g

Exercise 3.3.0.43 Show that the above result also holds for the maps between the reduced 0-dimensional Homology Groups. Exercise 3.3.0.44 Show that the relation of being chain homotopic is an equivalence relation on the collection of chain maps between two chain complexes. Proposition 3.3.3 The homology of a contractible space X is trivial except for H0 (X ) which is, up to isomorphism, Z. Proof. This follows, as the similar result for the fundamental group, from the fact that any contractible space has the homotopy type of a point, and the preceding theorem, which implies that any two spaces of the same homotopy type have isomorphic homology.

3.4

Homology of Pairs and Exact Sequences

It is noticeable that there are striking similarities between the fundamental group functor and the homology functors; the homotopy type arguments that we used for showing that 1(B n) = {0} all go through as a result of the fact
HeavenForBooks.com

Introduction to Algebraic Topology

97

that both functors take homotopic maps to the same group homomorphism. In fact, H1 is the abelianisation of 1. It was convenient to extend the fundamental group to take account of pairs, (X, A), where A X were topological spaces. In particular, A = gave us back the original space as a particular pair, and the case where A is a single point is also important. We next look at the homology of pairs of spaces. It will appear that this extension gives us a good start on the business of computing homology of complicated spaces in terms of homology of simpler spaces. You might feel that this is long overdue, and that you have done quite enough to earn your reward already. I did warn you that algebraic topology is for the tough people. You must hang on longer. Dont give up. Never say die. And similar inspiring remarks. Proposition 3.4.1 If (X, A) is a pair of spaces the map i : A X being the inclusion, then n Z, i# : Cn (A) Cn(X ) is a monomorhism, i.e is 1-1. Proof. We may extend Cn(X ) in the usual two ways to the case for n negative; in either case the result holds since two distinct cubes in Cn (X ) which come from cubes in Cn (A) by composing with i must have been distinct in Cn (A). And if it holds for cubes, it must hold for chains.

It follows that we can consider Cn (A) to be a subgroup of Cn (X ). Exercise 3.4.0.45 Does it follow that if A is a proper subspace of X then Cn(A) is a proper subgroup of Cn (X )? Denition 3.4.1 We write Cn (X, A) to denote the quotient group Cn (X )/Cn (A) We call this group the group of n-chains of the pair (X, A).
HeavenForBooks.com

98

MICHAEL D. ALDER

It is immediate that n (Cn (A)) Cn1 (A), since A is a perfectly respectable space in its own right. It follows that : Cn (X ) Cn1 (X ) induces a homomorphism: n : Cn (X, A) Cn1 (X, A) Proposition 3.4.2 2=0 Proof.

Cn+1 (X )

n+1

Cn (X )

Cn1 (X )

n+1

n1

Cn+1 (X, A)

n+1

Cn (X, A)

Cn1 (X, A)

The projection maps down to the quotient groups commute with the boundary maps , , so the result follows by a simple diagram chase.

If f : (X, A) (Y, B ) is a map of pairs (that is a continuous map between spaces X and Y which takes A to a subset of B ), then f# : Cn (X )/Cn (A) Cn (Y )/Cn (B ) is a well dened homomorphism for each integer n. It follows that there is a well dened functor from pairs of spaces and pair maps to chains and chain maps.
HeavenForBooks.com

Introduction to Algebraic Topology

99

We can now proceed to do homology exactly as before. Using the same symbol for , we have that ker(n) is a subgroup of Cn (X, A) for all intgers n, called Zn (X, A) and im(n+1 ) is another, called Bn (X, A), with the latter a subgroup of the former since 2 = 0. Hence we dene Hn (X, A) = Zn (X, A)/Bn (X, A) for every intger n. By choosing C1 to be Z and using the augmentation map : C0 C1, we get the reduced homology. It is clear that a cycle in Cn (X, A) is just a chain in Cn (X ) which has boundary in A instead of having boundary zero. So a curve which has both ends in A(but is not wholly in A) is the relative version of a non-trivial loop.

Exercise 3.4.0.46 Show that Cn (X, A) is a free abelian group generated by the non-degenerate singular n-cubes which are not contained in A.

Exercise 3.4.0.47 You might conjecture that there is some connection between the homology of the pair (X, A) and the homology of X/A which is the space obtained by shrinking A to a single point. (Every point of A is equivalent to every other point of A, and no point of X A is equivalent to anything except itself.) Explore this with special cases.

Exercise 3.4.0.48 You might subsequently conjecture that shrinking A to a point by a quotient map is unnecesarily brutal. Take the space X A I which has the relation that every a A X is identied with (a, 0) A I , with the quotient topology. This puts a little tower on A. Now identify all the points of A 1. This makes the tower on A into a cone on A. What can you nd out about the homology of this space and the homology of (X, A)?

There is a line mising from the top of the last diagram; we can put in the chain complex obtained from A, and the chain map derived from i : A X , to get:
HeavenForBooks.com

100

MICHAEL D. ALDER

Cn+1 (A)

n+1

Cn (A)

Cn1 (A)

i#

i#

i#

Cn+1 (X )

n+1

Cn (X )

Cn1 (X )

n+1

n1

Cn+1 (X, A)

n+1

Cn (X, A)

Cn1 (X, A)

Note that im(i#) = ker(n) for every n, that all the squares commute, and that 2 = 0 and also 2 = 0, so we can do homology and get an induced diagram:

Hn+1 (A)

Hn (A)

Hn1 (A)

Hn+1 (X )

Hn (X )

Hn1 (X )

Hn+1 (X, A)

Hn (X, A)
HeavenForBooks.com

Hn1 (X, A)

Introduction to Algebraic Topology

101

There is a new homomorphism called, with studied lack of originality, , which turns this array of triples into a long, long sequence. It goes from Hn(X, A) to Hn1 (A), as shown in the following diagram:

Hn+1 (A)

Hn(A)

Hn1 (A)

Hn+1 (X )

Hn(X )

Hn1 (X )

Hn+1 (X, A)

Hn(X, A)

Hn1 (X, A)

Our next task is to construct this family of homomorphisms. Proposition 3.4.3 There is a homomorphism, called , from Hn (X, A) to Hn1 (A). Proof (construction). Take u Hn (X, A). There is a representative cycle, cu in Cn(X, A). n is onto, so there is some chain c in (Cn(X ) which is taken to cu by n. Now the homomorphism n must take cu to zero in Cn(X, A), since it corresponds to something in the homology group Hn (X, A) and is therefore in the kernel of n. So if we go by the lower path from Cn (X ) to Cn1 (X, A), c is sent to zero. Since the square commutes, n1 n(c ) = 0, that is, n (c ) ker(n1 ). Now the vertical sequence Cn1 (A) Cn1 (X ) Cn1 (X, A)
HeavenForBooks.com

102

MICHAEL D. ALDER

on the previous page has the property that the kernel of n1 is precisely the image of i# . So c is actually in the subgroup Cn1 (A). c determines some element of the homology of Hn1 (A) It is to this element that we send u. We have to show a number of things, all done by diagram chasing. I repeat the most crucial diagram so you can go through it with a pencil and keep track of elements of groups.

Cn+1 (A)

n+1

Cn (A)

Cn1 (A)

i#

i#

i#

Cn+1 (X )

n+1

Cn (X )

Cn1 (X )

n+1

n1

Cn+1 (X, A)

n+1

Cn (X, A)

Cn1 (X, A)

First, we must show that the result does not depend on which c Cn(X ) we choose. Suppose we had chosen c . Now c c is sent to zero by n, so must be in the image of i#,n . In other words, c c is actually some element of Cn (A). Hence n(c c ) is a boundary element in Cn1 (A), so c and c determine a unique homology class. We are using the commutativity of the top right hand square in the above diagram in a clandestine and underhand way, but it works out correctly. Second, we must show that if c, d are dierent elements of Cn(X, A) that are homologous, i.e. that represent the same u in Hn (X, A), we wind up with d homologous to c in Cn1 (A). So take du also in ker(n), such that cu du is in im(n+1 ). Then, if e Cn+1 (X, A) is sent to cu du Cn(X, A), since
HeavenForBooks.com

Introduction to Algebraic Topology

103

n+1 is onto, there is an element f of Cn+1 (X ) which gets sent to cu du by going around the lower part of the commutative square, so it gets sent by n+1 to g in Cn (X ). Since there may be several such f , there may be several such g . Well, we knew there was at least one because n is onto, but for all of them, n (g ) = 2(f ) = 0. In other words, c and d go to the same element of Cn1 (X ) and we have shown that there is a perfectly respectable map of some sort from Hn (X, A) to Hn1 (A). It remains only to show it is a homomorphism. Take u, v in Hn (X, A) and trace some representatives cu , cv throught, along with cu + cv . The resulting diagram chase uses the fact that all the other maps are homomorphisms, and is left as an easy exercise for the student.

Exercise 3.4.0.49 Work through the above argument very, very carefully. Doing diagram chases makes you a better and more wonderful person. The last part of this section requires the idea of an exact sequence of groups. This is perhaps the most recurring single idea in algebraic topology, to try to nd exact sequences. It tells you how innocent of the subject you are that this is your rst encounter with the little sweeties. Denition 3.4.2 A sequence of groups and homomorphisms fn+2

-G

fn+1

n+1

-G

fn

-G

n1

is said to be exact at Gn i im(fn+1 ) = ker(fn ). The sequence is exact i it is exact at Gn for every Gn in the sequence. Exact sequences dont have to be innite (although most of ours will be), and there is a special exact sequence:
HeavenForBooks.com

104

MICHAEL D. ALDER

Denition 3.4.3 A short exact sequence is a sequence:

{0}

-G

-G

-G

{0}

which is exact at all the non-zero groups.

Exercise 3.4.0.50 Show that for the above short exact sequence, f is 1-1 and g is onto.

Now we come to an important and useful result:

Theorem 3.4.1 The homology sequence

-H

n+1 (X, A)

- H (A)
n

- H (X )
n

- H (X, A)
n

is exact. Proof. I shall show that the sequence is exact at Hn (A). The other two places are left as an important exercise (which is much easier with a diagram!) First I show that im() ker(i ) Take [x] im(). Then there is some cycle, x, in Cn (A) such that [x] = x + Bn (A), and this x is obtainable by the process which dened : Hn+1 (X, A) Hn (A) applied to some cu Cn+1 (X, A). In particular, cu Zn+1 ((X, A).
HeavenForBooks.com

Introduction to Algebraic Topology

105

We obtained such an x, by arguing that : Cn+1 (X ) Cn+1 (x, A) is onto, so there is some d Cn+1 (X ) which projects by to cu . Now : Cn+1 (X ) Cn(X ) takes this into something, say f , which is in ker(n) and is hence in im(i# ) and hence gives us x in Cn(A). Now i(x) may be taken to be represented by f . But this is in the image of : Cn+1 (X ) Cn(X ), and is hence a boundary in Cn(X ) and is hence zero in Hn (X ). So im() ker(i) Next we show that ker(i) im() Take [y ] ker(i). Then take y Cn (A) to represent this class. i#(y ) Cn(X ) must therefore be a boundary, that is, it is in the image of : Cn1 (X ) Cn (X ). If it is the image of g Cn+1 (X ), (g ) Cn+1 (X, A), where g is, by construction, a cycle, then (g ) is also a cycle, and hence determines a homology class [ (g )]. It is clear from the construction of : Hn+1 (A) Hn (A) that ([ (g )]) = [y ] Hence ker(i) im(). The matter of checking that the constructions are independent of representatives was done for the construction of . This concludes the proof of excactness at Hn (A). Exactness at Hn (X, A) is similar, and at Hn (X ) is easier.

Exercise 3.4.0.51 Produce the other two exactness proofs. You understand what is involved by doing these things, not by reading about them. Because it is rather more useless than usual to read a proof without following it through yourself in this area, I give the following important result as an
HeavenForBooks.com

106

MICHAEL D. ALDER

exercise. It is a very nice diagram chase which every educated person has done once. I said there would be a nice diagram chase coming up, and this is it.

Exercise 3.4.0.52 The Five Lemma In the following commutative diagram of abelian groups, the rows are exact and fi is an isomorphism for i = 1, 2, 4, 5, and a homomorphism for i = 3. Show that f3 in the middle is also an isomorphism.

C1 f1

i1

-C

i2

-C

i3

-C

i4

-C

f2 j1

f3 j2

f4 j3

f5 j4

D1

? -D
2

? -D
3

? -D
4

-? D

Your argument should have needed somewhat weaker conditions on the fi than actually provided. What are they?

Voyaging strange seas of thought in this present company, we have discovered the category of doubly innite exact sequences of abelian groups. It is a place where you wind up if you do homology on pairs of spaces from the category P airs. It seems reasonable to name this category, like Captain Cook naming Australia. There is no universally acknowledged name, so I shall call this category Diesag. This doesnt have the romance of Easter Island, or even Australia, but it is easy to remember. There isnt actually a category unless we have said what the maps are, but it is easy to see that there is a fairly reasonable candidate map; it is a doubly innite family of group homomorphismsfn :
HeavenForBooks.com

Introduction to Algebraic Topology in+1

107

-C

in

-C

n1

in1

-C

n2

fn jn+1

fn1 jn

fn2

? -D
n

? -D

n1

jn1

? -D

n2

such that all the squares commute. From now on, if we have two diesags C and D, a map f : C D in the category Diesag, is just such a doubly innite sequence of group homomorphisms which commute with the homomorphisms in each diesag in this way. When you travel strange seas of thought you discover new and bizarre things sticking out of said seas, and there is a temptation to stop and explore them. What strange and exotic beasties live in this category? What properties does it have in common with more familiar categories, and what properties are peculiar to it alone? Of course, one mostly does this exploration in a prosaic sort of way by looking at individual objects and maps of the category, and nding out what they are like by doing sums. But there are general questions one can ask: are there products in this category? Subobjects? Quotient objects? When can one say anything about the sub-diesags of a diesag? All these possibilities occur to the reective man or woman, and to be dragged past, not just a country but a whole category, without time to explore and get familiar with the place, is rather frustrating. Still, we are in a bit of a rush, and it seems unlikely that you can yet prove that R4 and R3 are not homeomorphic. So exploration will have to wait. The last thing that ought to be done before we close the section introducing pairs, is to show that there is a functor from the category of pairs of spaces to the category Diesag. We have shown how to take an object in the category of pairs of spaces to an object consisting of an innite exact sequence of abelian groups. It is reasonable to hope that a map in the category P airs is taken to a map in the (just discovered) category of doubly innite exact sequences of abelian groups. This is indeed the case (which is likely to make you more inclined to think that Diesag and P airs are sensible categories to take a look at.
HeavenForBooks.com

108

MICHAEL D. ALDER

Proposition 3.4.4 For any continuous map f : (X, A) (Y, B ) in the category of pairs of spaces, there is a family of homomorphisms induced between the exact sequence of relative homology groups which commute with the maps of the exact sequence. That is to say, there is in Diesag a map induced from f . Proof. In the following diagram

-H

n+1 (X, A)

- H (A)
n

- H (X )
n

- H (X, A)
n

- H ? (Y, B )
n+1

-H? (B )
n

-H? (Y )
n

-H? (Y, B )
n

we know that the homomorphisms f from Hn (X ) to Hn (Y ) and from Hn (A) to Hn (B ) exist, from the ordinary homology for spaces. That there is an induced homomorphism f# : Cn (X, A) Cn(Y, B ) from the map f follows from the denition of Cn(X, A) and the fact that f (A) B . It also follows that the map n : Cn (X, A) Cn1 (X, A) is well dened and that it commutes with f#. We want to show there is a well dened map f : Hn (X, A) Hn (Y, B ) The commutativity of f# with tells us that cycles are taken to cycles and boundaries to boundaries by f# , so the map f is well dened. This means that all the rungs of the ladder are in place, as prematurely indicated in the diagram. It is easy to conrm that fg = f g , as for single spaces. The only thing left is to show that every square of the diagram above commutes. The squares come in triples for every n, and two of the three are very straightforward. It is only the rst square which looks as though it might give trouble. It requires bashing through the denition of : Hn (X, A) Hn (A)
HeavenForBooks.com

Introduction to Algebraic Topology but is straightforward and is left as an exercise for the reader.

109

The only thing left to establish is the homotopy property for pairs. We dene a homotopy between two maps f, g : (X, A) (Y, B ) so as to make this work: Denition 3.4.4 Two maps of pairs f, g : (X, A) (Y, B ) are homotopic in the category of pairs i : (X I, A I ) (Y, B ) in the category of pairs such that (x, 0) = f (x) and (x, 1) = g (x), for every x X . Again we write f g for homotopic maps in the category of pairs. With this denition, it is not hard to show: Proposition 3.4.5 If f g, f, g : (X, A) (Y, B ) f = g : Hn (X, A) Hn (Y, B )

Exercise 3.4.0.53 Prove the above result carefully. These results extend the idea of a homology functor to pairs in a satisfactory way, and the fact that we get out a diesag is grounds for thinking that pairs are Good Things. (It is true that our diesag is rather uninteresting for integers less than 1.) The reader is invited to contemplate the reduced homology for pairs and the corresponding diesag of homology groups, and to conrm that its behaviour under mappings is properly functorial. He or she will nd it on page 172 of Masseys A basic course in Algebraic Topology
HeavenForBooks.com

110

MICHAEL D. ALDER

3.5

Excision

We are now on the home straight; there is one more important property of the homology functor from P airs to Diesag which we need to discuss. It is important in that with it, there is no great diculty in proving a theorem called the Mayer-Vietoris theorem. This tells us that a certain sequence is exact and enables us to compute the homology of spaces given homology of subspaces. By building up from easy spaces, we can therefore compute the homology of complicated spaces. This should make even the purest soul greedy, since it allows us to nail down all the surfaces and classify them completely, as well as giving us the result that Rn = Rm n = m where = denotes the relation of being homeomorphic. I hope you have been trying to prove this so simple looking theorem since the beginning of the book. There are a great many other useful results which can be obtained, some of which we will investigate. Before we can achieve this state of bliss, it is necessary to investigate what happens to the homology of a pair (X, A) if we cut a hole out of the A. If you brood on this for a few moments, you would probably guess that nothing at all happens to the homology, since A gets more or less taken out of account anyway. We factor out chains in A to get the pair chain Cn (X, A) = Cn (X )/Cn (A). Once rst thought then is that it seems quite likely that W A, Hn (X W, A W ) = Hn (X, A) This is actually false. Take X = I , A = I = {0, 1} and W = {0}. Your instinctive guess should be that Hn (X, A) is the homology of a circle, whatever that might be1 , but it is hard to believe this is always the same as Hn ((0, 1], {1}).
Actually you have some grounds for thinking that H1 (S 1 ) = Z is rather likely. Can you remember why?
1

HeavenForBooks.com

Introduction to Algebraic Topology

111

is in the interior of A. This is the It is, however true in the case where W Excision Theorem, and is the point of this section. I state it here so you will know where we are headed for the duration of this section, although proving it will take a little time: Theorem 3.5.1 Excision For any pair (X, A), and any W such that the closure of W is contained in the interior of A, the inclusion map i : (X W, A W ) (X, A) induces the map in Diesag i : H (X W, A W ) H (X, A) which is an isomorphism. It is easy to see that for the map i in Diesag to be an isomorphism in the category, all the i,n have to be isomorphisms, and vice versa. There is a tendency, of which you should beware, to think if examples of spaces and pairs which are rather fat and jolly and simple. If your space X is a disk with a hole taken out of the middle, and A is a smaller disk somewhere well away from the boundary of X , and W is a small star shaped hole taken out of the middle of A and well away from the boundary, then it doesnt look too hard to show the result. The trouble arises when the subspaces are rather nasty and tortuous. Think of the Alexander wild horned sphere (not related to the Rag Time Band) as a possible A when R3 = X . Think of all the nasty sets you get out of doing things to the cantor set. Think of the Mandelbrot set. Behold these thoroughly (from the point of someone who nds a torus complicated) horrid things and Despair! The problem of showing that any two triangulations of a manifold have a common renement runs into similar diculties. One way out is to restrict the class of maps we allow, so that the triangulations have to be smooth, or piecewise ane, Similarly, we can restrict our singular cubes to have particularly nice maps. This usually puts restrictions on those spaces for which the homology is dened, and even so can be quite tricky. Our basic approach is to show rst that we can insist that all our singular cubes must have their
HeavenForBooks.com

112

MICHAEL D. ALDER

images contained in small subsets. If you want to do the homology of (I 2, A), you have to be prepared to take the boundary as expressible in terms of very little line segments all following it around, rather than four big ones for the outside, or for that matter, one big one for the outside. The argument that we follow falls into three separate parts. The rst is to do with the idea of a subdivision of a singular chain. The second part applies this idea to show that the homology of a space or a pair is unchanged if instead of allowing arbitrary continuous maps in our n cubes, we restrict ourselves to small maps, where the images can be contained in small open sets. What, precisely this means, will be seen a little later. And nally, we go from this result to the excision theorem. I follow the line of ideas of Massey, [2] in A Basic Course in Algebraic Topology here. I amplify a few points, but if you prefer his proofs to mine, thats ne with me.

3.5.1

Subdivision

It is easy to chop the unit n-cube I n into 2n equal size bits by taking each axis and the hyperplane orthogonal to it and of distance 1/2 along that axis. If we do this for the domain of a singular n-cube q , then the result is 2n singular n-cubes, which we can regard as a chain by assigning each singular sub-cube multiplicity 1. This denes a map called subdivision, and written sd, from n cubes to n-chains, and by extension it denes a map sdn : Cn (X ) Cn(X ) for any X . There are two important results to be established. The rst is the intuitively obvious result that the subdivision of the boundary is the boundary of the subdivision. This should remind you of Stokes Theorem proofs: The internal boundaries of the subcubes of a cube cancel out, leaving only the outside. The second is that the identity map In : Cn (X ) Cn (X ) is chain homotopic to the subdivision map sdn , for every n Z. This is intelligible, but not immediately obvious. If we can show these results, then there is nothing to stop us subdividing
HeavenForBooks.com

Introduction to Algebraic Topology

113

again; the composite of chain homotopies is easily seen to be a chain homotopy, and by a trivial induction, it doesnt matter how many subdivisions we do, sdk n commutes with and is chain homotopic to the identity. The main task in this is to take the intuitively simple idea of a subdivision and turn the idea into algebra so we can prove the claims instead of appealing to an intuition which leaves out all the details. Each cube in the subdivision is associated with one corner of the original cube; a corner has coordinates a vector, each entry being either 0 or 1. Denition 3.5.1 En denotes the set of vertices of I n , and e = (e1 , , e2, en ) En iff I [1 n], ei {0, 1} Denition 3.5.2 q Qn (X ), e En , Se (q ) Qn (X ) is dene by Se (q )(x) = q ( e+x ) 2

This gives the singular subcube associated with the corner e. Now we dene Denition 3.5.3 For n > 1, sdn : Qn(X ) Cn (X ) by sdn (q ) =
eEn

Se (q )

and if q is a 0-cube, sd0 (q ) = q . We note that:


HeavenForBooks.com

114

MICHAEL D. ALDER

Proposition 3.5.1 If q is degenerate, so is sd(q ). Proof. Trivial.

This means we can extend sdn to sdn : Cn(X ) Cn (X ) by linearity. Proposition 3.5.2 sd commutes with the boundary operator: n sdn = sdn1 n Proof. If e, e are such that they dier in only one place, j and ej = 1, ej = 0, then Fj Se = Bj Se If e En and ej = 0, and e En1 = (e1 , e2, , ej 1 , ej +1 , en) then Fj Se = Se Fj and If e En and ej = 1, and e En1 = (e1 , e2, , ej 1 , ej +1 , en) then Bj Se = Se Bj
HeavenForBooks.com

Introduction to Algebraic Topology These identities are easily veried, and lead to the result n sdn (q ) = sdn1 n (q )

115

from which it follows that the extension to chains also commutes with the boundary operator.

The algebra follows the geometry precisely. We note also, and rather trivially, that Proposition 3.5.3 (sd0 (q )) = (q ) for any 0-cube q , where is the augmentation map.

This gives us the rst result of this subsection, that the boundary of a subdivision is the subdivision of the boundary. Next we establish that: Proposition 3.5.4 There is a chain homotopy between the identity chain map and the subdivision operator. Proof. We need to construct, for every integer n, a map n : Cn (X ) Cn+1 (X ) with the property that for any chain c Cn (X ), n+1 n (c) + n1 n(c) = sdn (c) c First we dene n : Cn (X ) Cn+1 (X )
HeavenForBooks.com

116

MICHAEL D. ALDER

Figure 3.4: A chain homotopy between a 1-cube and its subdivision to be the zero map, for n 0. For n = 1 we need, for every 1-cube, q , some 2-chain, the boundary of which is sd(q ) q , as sketched in the diagram in a somewhat impressionistic way. 2 For the general 1-cube, q , we put:
2 In gure 3.4, I am taking the obvious inclusion map for my cube q , with q (t) = (t, o) in X = R2 . In order to avoid everything getting confused with everything else, I have shown the boundary of the two 2-cubes I have chosen for (q ) as embedded in X = R2 , but the vertical progression is intended to convey that we have to project down onto the X-axis to get the required two 2-cubes. And, of course, I am trying to represent each chain by the images of the constituent cubes. The left hand cube is g0 (q ) dened for the top part of the diagram by:

(g0 (q ))(x, y) = ( and the right hand cube is g1 (q ) dened by: x+y 1 x+y 1

x , y) 2y

x+1 , y) 2y (g1 (q ))(x, y) = (1, y) (g1 (q ))(x, y) = (

Since we want the boundary to be the sum q0 + q1 q , where in our example q0 (t) = (t/2, 0) and q1 (t) = (1/2 + t/2, 0), we squash the y value to zero in the second component to get the bottom part of the gure, which is supposed to lie along the X-axis.

HeavenForBooks.com

Introduction to Algebraic Topology

117

1(q ) = G0 (q ) + G1 (q ) where (G0 (q ))(x, y ) = q ( and x+y 1 x+y 1 Then ( (q ))(x) = = = (G0(q ))(x) + (G1 (q ))(x)) ((F1G0 q )(x) (B1G0 q ))(x) + ((F2G0 q )(x) (B2 G0 q ))(x) ((F1G1 q )(x) (B1 G1 q ))(x) + ((F2G1 q )(x) (B2 G1 q ))(x) ((G0q )(0, x) (G0 q )(1, x)) + ((G0 q )(x, 0) (G0 q )(x, 1)) ((G1 q )(0, x) (G1 q )(1, x)) + ((G1 q )(x, 0) (G1 q )(x, 1)) 1 x )) + (q ( ) q (x)) = (q (0) q ( 2x 2 1 x+1 (q ( ) q (1)) + q ( ) q (1) 2x 2 x x+1 ) q (x) q (0) = q( ) + q( 2 2 x+1 ) 2y (G1 (q ))(x, y ) = q (1) (G1 (q ))(x, y ) = q ( x ) 2y

Which is the right answer, since q (0) is degenerate and gets factored out. The contribution from 0 is zero, of course. Note that the diagonal bit of boundary was traversed both ways and so cancelled out. This shows that there is a sensible denition of 1 , it is now necessary to dene n in full generality, and to verify, as in the above calculation, that the equation: n+1 n (q ) + n1 n(q ) = sdn (c) q holds. Then extends to arbitrary chains by linearity again.
HeavenForBooks.com

118

MICHAEL D. ALDER

Figure 3.5: A second attempt. We leave this to the diligent student: the details may be found in [2].

Exercise 3.5.1.1 Some experimenting is called for. It should be obvious that the above proof uses a rather arbitrary looking couple of maps which are used to produce the two n-cubes that sum to the chain homotopy. It does not look to be out of the question to do it with only one singular 2-cube in the case where n = 1. One imagines squashing the square to a triangle by contracting the base line to the midway point, and then projecting down to get the required chain homotopy, as in gure 3.5. Try it.

Exercise 3.5.1.2 Do the explicit calculation for the given proof from Massey, checking that the chain homotopy works for 2-cubes, and draw as many pictures as you can.

Exercise 3.5.1.3 Dene the (potential) chain homotopy in dimension one on a 1-cube q by (q )(t, x) = tx t/2 + 1/2
HeavenForBooks.com

Introduction to Algebraic Topology

119

and in dimension zero to be the zero map, and a bisection of the 1-cube q as the chain q ((1 x)/2) + q ((1 + x)/2) Show that the chain homotopy is a well dened chain homotopy between the identity and a bisection in dimension one. Dene a bisection on an n-cube q by Bq (x1, x2, xn ) = q ((1 x)/2, x2 , x3, , xn) + q ((1 + x)/2, x2, , xn) Can you extend the map to any dimension to get a chain homotopy between the bisection operator and the identity? If you could, or if you could nd any such chain homotopy, the proof of the last theorem could be reduced to induction on bisections, or at least, we could obtain a similar result, close enough for all practical purposes. Is there any such chain homotopy? Could you modify the bisection operator so as to preserve the essential element of bisecting I n , so that n iterations of it applied to dierent axes will give something close to the subdivision operator?

3.5.2

Small Cubes

Denition 3.5.4 If U = {Uj : j J } is a family of subsets of a topological space X such that the interiors of the Uj cover X , then a singular n-cube q : I n X is said to be small of order U if j J, q (I n) Uj .

In a metric space, think of a cover by the balls of radius some small number . Denition 3.5.5 Qn (X, U )) denotes the free abelian group generated by the n-cubes which are small of order U . The degenerate n-cubes which are small of order U generate a subgroup Dn (X, U ).
HeavenForBooks.com

120

MICHAEL D. ALDER

Since the degenerate n-cubes which are small of order U are just Dn (X ) Qn (X, U ), we have: Denition 3.5.6 The chain group Cn(X, U ) is the quotient group Qn (X, U )) Dn (X, U ) referred to as the chain group of elements small of order U . The relative chain group for pairs of spaces (X, A) is similarly dened. We note that the boundary of a small n-cube is a chain of small (n 1)-cubes so we can nish the job and dene: n : Cn (X, A, U ) Cn1 (X, A, U ) and hence Zn (X, A, U ) and Bn (X, A, U ) and nally Hn (X, A, U ). The details are left to the reader; they can be done as a cure for insomnia in bed at night in the dark. Note that Q0(X, U ) = Q0 (X ). Exercise 3.5.2.1 To make this functorial, we need maps from (X, U ) to Y, V ) which behave sensibly with respect to the two covers. Write down a sensible condition and show that Hn makes sense for the category of spaces with covers, and the category of pairs of spaces with covers. There is an inclusion map Qn(X, U ) in Qn (X ). This induces a homomorphism n : Cn(X, A, U ) Cn(X, A) which clearly commutes with the boundary operator, and hence induces : Hn (X, A, U ) Hn (X, A) for every integer n.
HeavenForBooks.com

Introduction to Algebraic Topology

121

The result of this section is that this map is an isomorphism: we do not have to worry about n-cubes that are big, we can restrict our attention to small ones. This, if you reect on it, is not surprising. If you have a cycle, you could subdivide it into very small cubes, but it would be, in a certain sense the same cycle. Well, up to a chain homotopy. Proposition 3.5.5 For any cover U of X , for any n-chain u Cn(X ), there is a positive integer k such that sdk n (u) Cn (X, U ) Proof. Since u is a nite sum of n-cubes, it suces to show the result for n-cubes. The result then follows from the Lebesgue Covering theorem. Exercise 3.5.2.2 Write out the details of the argument above. Exercise 3.5.2.3 Show that the argument that shows that the subdivision operator is homotopic to the identity behaves itself under open covers; in particular that the homotopies also behave properly when all the cubes are small of order U . Theorem 3.5.2 The map induced from the inclusion of small chains in all chains : Hn(X, U ) Hn (X ) is an isomorphism. Proof. To show is onto, let [u] be an element of Hn (X ) with representative cycle u Cn (X ). Subdivide every cube in u by a suitable subdivision until sdk n (u) Cn(X, U ), and since subdivision commures with the boundar operator, sdk n (u) is a cycle. Since it is chain homotopic to u, it is in the same homology class. Hence is onto.
HeavenForBooks.com

122

MICHAEL D. ALDER

To show is 1-1, take an element v in Cn (X, U ), which has ([v ]) = 0. Then n (v ) must be a boundary, so there is a u Cn+1 (X ) such that (u) = n (v ). By the conclusion that is onto, we can take it that there is a u Cn+1 (X, U ) that is chain homotopic to u. We have that u = 0, and so v is homotopic to a boundary, so has [v ] = 0. So is 1-1.

This has not established the result for pairs, only for spaces. However: Exercise 3.5.2.4 Show that the exact sequence for pairs together with the ve-lemma extends the above result to pairs.

3.5.3

Excision at Last

Theorem 3.5.3 Excision Property is contained in the For any pair of spaces (X, A), and any W such that W interior of A, the inclusion map i : (X W, A W ) (X, A) induces the map i : H (X W, A W ) H (X, A) which is an isomorphism. Proof. Given (X W, A W ) as specied, there is a cover of X by sets U = {X W, A} of the sort we have been considering; that is, the sets need not be open but their interiors form an open cover. For the rest of this proof, U refers specically to this cover, as in the expression: Cn(X, A, U )
HeavenForBooks.com

Introduction to Algebraic Topology

123

Note that Cn(A) is a subgroup of Cn(X, U ), and so is Cn(X W ). Moreover, adding elements of one to elements of the other subgroup gives us the whole group: Cn (X, U ) = Cn (A) + Cn(X W ) This says that the chains which are either contained in A or in X W is the sum of the chains which are contained in A and those which are contained in X W . The sum is not isomorphic to the direct sum because there may be chains which are in both. Now Cn (X W, A W ) = Cn (X W )/Cn (A W ) = Cn (X W )/[Cn (X W ) Cn (A)] by denition, and there is an inclusion map : Cn (X W ) Cn(X, U ) which takes Cn (A W ) into Cn (A, U ), and hence induces a map [] : Cn (X W, A W ) Cn (X, A, U ) This can more usefully be expanded to: [] : Cn (X W ) Cn (X W ) + Cn(A) Cn (X W ) Cn(A) Cn(A)

which is well know to be an isomorphism (The rst isomorphism theorem of group theory). This therefore induces an isomorphism [] : Hn (X W, A W ) Hn (X, A, U ) for every integer n. And we already know that there is an isomorphism : Hn (X, A, U ) Hn (X, A) so the composition of these isomorphisms establishes the result.

HeavenForBooks.com

124

MICHAEL D. ALDER

This gives us the last of the fundamental properties of the homology functor. The next stage is to prove the Mayer Vietoris Theorem, and then we need only derive some results.

3.6

The Mayer-Vietoris Theorem

In this section we explore (since we now have all the abstract machinery in place) the question of the homology of a space X = A B in terms of the homology of the two subspaces A, B and the homology of A B . If we can get a sensible result out, we can calculate just about any homology of the manifolds, because they are built up out of patches which are contractible and hence have rather simple homology. We can, indeed, go much further than that, but shant. The aim is to make this introduction pretty minimal. The rst step is to observe that given X = AB there is an inclusion i : A B A, and another j : A B B , and also inclusions k : A X and : B X , and that these induce homomorphisms i# : Cn (A B ) j# : Cn (A B ) k# : Cn (A) # : Cn (B ) Cn (A) Cn (B ) Cn (X ) Cn (X )

and from them we also obtain induced homomorphisms: i : Hn (A B ) j : Hn (A B ) k : Hn (A) : Hn (B ) Hn (A) Hn (B ) Hn (X ) Hn (X )

HeavenForBooks.com

Introduction to Algebraic Topology These allow us to dene homomorphisms: : Cn(A B ) Cn(A) Cn (B ) : Cn (A) Cn (B ) Cn(X ) by the formulae: (x) = (i# (x), j# (x)) x Cn (A B ) (u, v ) = k# (u) # (v ) u Cn(A), v Cn (B )

125

and similarly, at the homology quotient level we can dene homomorphisms: [] : Hn (A B ) Hn (A) Hn (B ) [ ] : Hn (A) Hn (B ) Hn (X ) by the formulae: [] (x) = (i(x), j(x)) x Hn (A B ) [ ] (u, v ) = k (u) (v ) u Hn (A), v Hn (B ) Observe my habit of putting brackets [ ] around an object or a map to indicate that we have taken equivalence classes. It would violate too many established conventions to write [i#] for i. Anyway, there are rather a lot of factorings out of equivalence classes or cosets going on around here. You will recall that we constructed an exact sequence for pairs of spaces; we use a similar construction to get a homomorphism like the homomorphism: there is a homomorphism: n : Hn (X ) Hn1 (A B ) With this (yet to be dened homomorphism) we have: Theorem 3.6.1 Mayer Vietoris If X is a topological space, and if A X , B X and if X is contained in the union of the interiors of A and B , then there is, for each n Z, a homomorphism n : Hn (X ) Hn1 (A B )
HeavenForBooks.com

126 such that the sequence

MICHAEL D. ALDER

Hn (A B ) Hn (A) Hn(B ) Hn(X ) Hn1 (A B ) is exact. If A B = , the sequence remains exact if we substitute reduced homology in dimension zero. Proof. We take a cover by the sets U = {A, B }. For the rest of this proof, U denotes this cover and no other. We have that : Hn (X, U ) Hn (X ) is an isomorphism for all n Z, and that Cn(X, U ) = Cn (A) + Cn (B ) Indeed, it is easy to see that we have commutative diagrams: Cn (X, U ) Cn (X, U )

[]

[]

[]

; ; C (A) @@k @@ @@ R
n #

; ; ;;
k#

Cn (B )

; ; ;; ;;
#

Cn (X )

@@ @@ @@ R
#

?
Cn (X )

An n-cube in A is a cube in X that is contained in one of the elements of U and is hence an element of Cn(X, U ). It is slightly embarrassing to have to
HeavenForBooks.com

Introduction to Algebraic Topology

127

say such simple things in algebra. We slither over the distinction between k and k and between and in what follows. We now have a diagram which bears a passing resemblance to one which we saw in the section on homology of pairs of spaces. 0 0 0

?
Cn+1 (A B )

?
Cn(A B )

?
Cn1 (A B )

-C

n+1 (A)

Cn+1 (B )

- C (A) C (B ) - C

n n

n1 (A)

Cn1 (B )

Cn+1 (X, U )

? ?

Cn(X, U )

? ?

Cn1 (X, U )

? ?

The rst thing to note is that the vertical sequences are all exact. This merely means that is 1-1 and is onto and that im() = ker( ). That is 1-1 is obvious. That is onto is also obvious. That im() = ker( ) is also obvious, so its all obvious. The next thing to observe is that every square commutes. This can be seen by following some cube through the diagram.
HeavenForBooks.com

128 We now construct a map

MICHAEL D. ALDER

n : Hn(X, U ) Hn1 (A B ) by exactly the same method that we used with pairs. Take [u] in Hn (X, U ) with u Cn (X, U ) as representative. Then u is a cycle, so it gets sent to 0 Cn1 (X, U ) by . It is also sitting under an onto map , and so is (a) for some a in Cn (A) Cn(B ). Further, a is in ker( ) Cn1 (A)Cn1 (B ). Hence it is the image of some b Cn1 (AB ). We send [u] to [b] in dening . It remains to show that this gives a well dened map independent of the various choices made. This is an exercise for the student. Likewise the proof of exactness at the various points. That leaves us with an exact sequence: Hn (AB ) Hn (A)Hn (B ) Hn (X, U ) Hn1 (AB ) All that needs to be done is to replace Hn (X, U ) with the isomorphic Hn (X ). It is left to the diligent student to obtain the exact sequence for the reduced homology in the case where A B = .
[] [ ] []

3.7

Applications

After what may seem a humungous quantity of algebra, most of it nit-picking trivialities as far as the time spent doing it is concerned, we are ready to obtain some geometric results. The machinery constructed is fairly powerful, and if we consider only simple applications, that is in keeping with the plan to keep this book minimalist. Even so, it is possible to establish some results which are far from trivial. n to In order to keep the results simple, I dene the reduced homology H be Hn when n > 0 and the usual denition via the augmentation map when
HeavenForBooks.com

Introduction to Algebraic Topology

129

n = 0. It is consistently zero for negative integers, and so not very interesting. I write Hn (X ) = Z as shorthand for saying that Hn (X ) is innite cyclic. The next result is consistent with my remark that H1 is the abelianisation of 1 and extends it a long way: Proposition 3.7.1 If S n denotes the n-sphere, {x Rn+1 : x = 1}, k (S n ) = 0 k=n&k=0 H and n (S n) = Z n 0 H Proof. We have the result rather trivially for S 0 the two point space. The reduced zero homology group is Z (up to isomorphism, the way I dene all my groups), and the higher homology groups are all trivial because all chains are degenerate. For S 1 , the unit circle, Take A to be the circle with the point (0, 1) removed. I shall call this point the South Pole of the space S 1. Similarly B is the circle with the North Pole, (0, 1) removed. Now it is clear that A B = S 1. It is also clear that A and B are both contractible and have therefore k (A) = H k (B ) = {0} k 0, H Now the intersection A B is a circle with two points missing, and it is easy to see that this is of the same homotopy type as S 0 . There is a retraction of A B onto the points {(1, 0), (1, 0)} which we can think of as the equator of S 1. We now plug the known facts into the exact sequence : k (A)H k (B ) H k (S 1) H k1 (AB ) H k1 (A)H k1 (B ) H
HeavenForBooks.com [] [ ] []

130 We get

MICHAEL D. ALDER

k (S 1 ) 0 0 0 H for k > 1. At k = 1, we have:


[] [ ] [] 1 (S 1 ) 0 H Z 0

[]

[]

[]

k (S 1) = 0 Now the exactness of these sequences tells us immediately that H 1 (S 1 ) = Z . for k > 1 and that H 0 (S 1) = 0 also follows by the same arguments. The conclusion that H This gets the result for S 1 . We note that in general, there is a canonical embedding of S n1 as the equator of S n . The space S n with the north pole (0, 1, 0, , 0) removed is contractible. Likewise the space with the south pole removed. They therefore have consistently got trivial reduced homology in all dimensions. The intersection A B of the covers for the n-sphere has the homotopy type of S n1 . There is a simple induction argument to complete the proof.

Corollary 3.7.1.1 The n-sphere is not contractible to a point. Proof. If it were, the contraction would induce an isomorphism of homology groups between S n and the point. Which is not the case.

Corollary 3.7.1.2 n, m Z+ , n = m, S n = Sm
HeavenForBooks.com

Introduction to Algebraic Topology Proof.

131

A homeomorphism would induce isomorphic homology groups, which they aint.

Corollary 3.7.1.3 n, m Z+ , n = m, Rn = Rm Proof. Suppose there were a homeomorphism from Rn to Rm . Take a point out of Rn , and remove also the image of the point under the alleged homeomorphism. Then the homeomorphism restricts to a homeomorphism of the pointless spaces. But these are of the homotopy type of S n1 and S m1 respectively, and hence would have to have isomorphic homology groups. Which they dont. So no such homeomorphism exists.

Corollary 3.7.1.4 Brouwers Fixed Point Theorem If B = B n = {x Rn : x 1} is any ball, then any continuous map f : B B has a xed point, i.e. x B : f (x) = x Proof. As for the case n = 2 which we have already done; a xed-point free map f would lead to a retraction of the ball onto its boundary, that is the diagram S n1 B n S n1
i r

with ri = ISn1 the identity. This would lead to homology:


Hn1 (S n1 ) Hn1 (B n ) Hn1 (S n1 )

HeavenForBooks.com

132

MICHAEL D. ALDER

with r i = IZ which is impossible. Hence there is no such retraction, hence no xed-point free continuous map.

It is worth remarking that the fact that we have this result for all dimensions allows the possibility of transferring the ideas to function spaces. LeraySchauder theory does this to prove the existence of solutions to a large set of dierential equations. I must admit, that if the design of an aeroplane depended on Leray-Schauder theory, which in turn involves Brouwers theorem which in turn reqires the axiom of choice, I wouldnt want to y in it. Fortunately, there are more constructive existence proofs around. But they are long. When I was a graduate student, I worked next door to a sta member who had been a student of Brouwer in Holland. He told me that Brouwer was notorious for having aairs with his female students, and kept this up until well into his eighties. An example to us all. We can also get some mileage out of the homology results for spheres by looking to see what happens to the maps of an n-sphere into itself. Since any such map f : S n S n induces a map f : Z Z and since the only endomorphisms of Z consist of multiplication by an integer, we can use the integer to say something about f . It is called the degree of f . It is obvious that homotopic maps have the same degree. We met this idea in the case of S 1 earlier:

Exercise 3.7.0.1 Show that the previous denition of the degree of a map f : S 1 S 1 agrees with the present denition.
HeavenForBooks.com

Introduction to Algebraic Topology There is a special map which makes sense for any n-sphere: Denition 3.7.1 The antipodal map a : S n S n sends x to x. The following proposition is very cute: Proposition 3.7.2 If f : S n S n has no xed point, then f is homotopic to the antipodal map. Proof. We construct a homotopy H : S n I S n by H (x, t) = (1 t)f (x) tx (1 t)f (x) tx

133

This is well dened so long as the denominator is non zero; it could only be zero if (1 t)f (x) = tx for some value of x and of t. In this case we should certainly have (1 t)f (x) = (1 t) f (x) = tx = t x But f (x) = 1 = x , so t = 1/2 and f (x) = x, which is not the case. So the homotopy is well dened.

It is also clear that the antipodal map has no xed point. It follows that: Proposition 3.7.3 a map f : S n S n
HeavenForBooks.com

134

MICHAEL D. ALDER

has no xed point i it has degree (1)n+1 Proof. Given that we know it has to be homotopic to the antipodal map, all that remains is to show that the antipodal map has degree (1)n+1 . Well, the case n = 0 has degree 1 as required. An antipodal map on S n is an antipodal map on its equator, S n1 , extended in the obvious way to Rn+1 , followed by a reection in the plane of the equator x1 = 0. It therefore suces to show that the reection of the n-sphere in the x1 = 0 hyperplane has degree 1. Now for n = 1 we have that the antipodal map is just rotation by , which is obviously homotopic to the identity map (Rotate it back, slowly!), and hence has degree one. Since degrees of composites of maps multiply (!) we have that the degree of the reection in the plane x1 = 0 for S 1 must be 1. Indeed, the reection of S 1 in any hyperplane x1 = 0 where we embed the space R2 in Rn for n 2 in the obvious way, is still 1. Now a reection of S n in the plane x1 = 0 restricts to a reection of the equator S n1 in the same plane, for n 2. Considering the exact sequence

k (A)H k (B ) H k (S n ) H k1 (AB ) H k1 (A)H k1 (B ) H

[]

[]

[]

and putting A, B to be the sphere with the poles removed, and taking k = n, we obtain:

[] [ ] [] n (S n ) n1 (S n1 ) 0 H H 0

giving the result that is an isomorphism. If R denotes the reection map, we have an induced diagram:
HeavenForBooks.com

Introduction to Algebraic Topology []

135 []

[ ]

-H (S
n

-H

n1 (S

n1

[]

[ ]

? -H (S
n

R
n

? -H (S
n1

n1

[]

The map R has a particularly simple eect on chains: it merely reverses the sign of the rst component. It follows that the central square commutes. From this it follows that the degree of R : S n S n is the same as the degree of the restriction of R. Since it has degree 1 for n = 0 and n = 1, it has degree 1 for any S n. This gives the desired result.

Exercise 3.7.0.2 There are a number of details that need checking. Can you show, for instance, that the stereographic projection of the space consisting of an n-sphere with a pole removed (by projecting down onto a suitable hyperplane in Rn+1 ) is a homeomorphism between the space and Rn? Can you show that the degree of a composite of maps S n S n is the product of the degrees of the separate maps? Can you provide an explicit retraction of A B S n to the equator, S n1 ? All of these things are nice easy examination questions.

Denition 3.7.2 A vector eld on a space is a map which assigns a tangent vector to each point of the space. A non-zero vector eld is one where the vector assigned is never the zero vector.
HeavenForBooks.com

136

MICHAEL D. ALDER

Theorem 3.7.1 There exists a non-zero vector eld on S n i n is odd. Proof. There is clearly a non-zero vector eld on S 1 : assign to the point (x, y ) the vector (y, x). Then the vector is orthogonal to the line joining the origin to its tail, so is a genuine tangent vector to the circle, and it is never zero because x2 + y 2 = 1. If n is odd, to the point (x1, x2, xn+1 ) assign the vector (x2 , x1, x4, x3, xn+1 , xn) Then the same argument shows that this is a tangent vector and has non-zero norm. We now need to show that if n is even, there is no such vector eld. Suppose there were, call the vector eld V . Then there is a map fV : S n S n dened by fV (x) = V (x) V (x)

Since fV (x) is always orthogonal to x and neither are zero, the map fV has no xed point. It therefore has degree 1n+1 , which is 1 since n is even. But there is a homotopy between fV and the identity map, given by H (x, t) = tx + (1 t)fv (x) tx + (1 t)fv (x)

This is well dened so long as the denominator is never zero; for that to be the case we would need to have some t [0, 1] such that tx = (1 t)fV (x) Taking the inner product of both sides with x we obtain t x
2

= (1 t) < x, fV (x) >


HeavenForBooks.com

Introduction to Algebraic Topology

137

and the right hand side is zero for every value of t while the left hand side is zero only at t = 0. But in this case we still have tx + (1 t)fV (x) = fV (x) = 0 So the homotopy is well dened. It follows that fV has degree 1. But 1 = 1. Contradiction. Hence no such V exists. In dimension two this is known as the hairy ball theorem. The ows on a manifold arise as solutions to vector elds, as in Rn . It is therefore no surprise that we have the same result, so to speak, for ows: Theorem 3.7.2 If n is even, every ow on S n has a xed point. Proof. A ow v : R S n S n determines for each t R a map vt : S n S n First we show that if v has no xed points, then there is some t R such that vt has no xed points. Now if t = 1, we may have xed points of v1. We note that from the denition of a ow, all points which are xed under v1 are also xed under vk for any integer k . We therefore go the other direction: If F1 is the set of points which are xed under v1, let F2 be the set of points which are xed under the map v1/2. We note that F2 F1, and that both sets are closed, by continuity of v . We proceed in this way to construct a nested sequence of sets Fk which are the xed point sets for v 1k
2

HeavenForBooks.com

138

MICHAEL D. ALDER

There are two cases to consider; either at some positive integer , F = which gives us the required t, or for every positive integer k , Fk = . In the latter case, we note that we have a sequence of non-empty nested closed subsets of a compact space, S n , which must have non-empty intersection. (Or the complements are a nested collection of open sets which give a cover for the compact space having no nite subcover.) Dene the set F =
kZ+

Fk

Now each point of F is xed under vp k


2

for any positive integers p, k , which constitute a set which is dense in the positive reals. Indeed, p can be negative, so the set F is left xed under all vt in a dense subset of R, known as the dyadic rationals. And by continuity, F is xed under vt, t R, in other words F is a non-empty set of xed points of the ow v . Contradicting our hypothesis. Hence if v is a ow on S n without a xed point, t R : vt : S n S n has no xed point. Now there is a rather simple homotopy between vt and the identity on S n : H (s, x) = vst (x) and so vt has degree 1. But we know that any xed point free map S n S n has to have degree (1)n+1 , which is 1 for n even. The theorem follows.

3.8

Summary

If you reect on what we have covered, you will see that quite a lot of the mathematics of an undergraduate degree course has been used, implicitly or
HeavenForBooks.com

Introduction to Algebraic Topology

139

explicitly, in getting to this point. Some of it has been used in the obvious sense that you couldnt prove a result without the background denitions, but other material has some relevance too: you wouldnt want to prove the hairy ball theorem unless you were concerned with the qualitative aspects of systems of ordinary dierential equations on manifolds, and for that it helps to have considered systems on Rn . We have, in the rst chapter, summarised rather briey all the formal mathematics which you can be expected to know which has a bearing on this book. In the second chapter I gave some additional algebraic machinery, the language of Category Theory, and then exemplied the kind of application with which we are concerned by establishing some elementary properties of the Fundamental Group of a space and of a pair of spaces. This is a functor from the category of topological spaces to the category of groups. This allowed us to deduce the Brouwer Fixed Point Theorem in dimension two. It was a start. The Brouwer theorem in dimension three is easy to state in terms which a small child can grasp. Fill a matchbox up with plasticene. Now take the plasticene out of the box, and twist it, stretch it, tie knots in it, compress it (but do not tear it) and nally stu it back, any old how, into the match box. Then there is at least one point of the plasticene which is in the same position relative to the matchbox after the deformation, as it was before. I dont know if this is obvious to your intuitions. It isnt even believable to mine. What is even less believable is that it can be proved. And not only can it be proved, it can be proved to hold for matchboxes of arbitrary dimension. The third and last chapter was a brutal attack on the general class of problems indicated in low dimensions in chapter two. Instead of extending to the higher homotopy groups (because it is too hard) we constructed new functors, the homology functors which take value in abelian groups. We went through some rather humungous abelian groups, but came out in the end with nitely generated groups for the case of the spaces we were concerned with, actually just spheres and contractible spaces. Even though we wound up only studying spheres, we were able to deduce some results which are horribly dicult to prove any other way. And they
HeavenForBooks.com

140

MICHAEL D. ALDER

emerged with relatively little eort after the general machinery for tackling lots of problems had been built. This ends the book, but not the homology theory of topological spaces, still less algebraic topology. The above result could save you a lot of time trying to construct something that doesnt exist; it can be quite time consuming trying to construct something that does, so you should be properly grateful. The result extends to a whole family of spaces, giving a computable number for a map to decide if it can have a xed point. Some of the results of algebraic topology may seem to prove things which are intuitively obvious, such as n = m Rn = Rm A result that is as easy to state as this and which agrees with (some of) our intuitions, seems a trie inconsequential. This is quite wrong. First, intuitions can be quite misleading, as, say, the Banach-Tarski paradox shows3 . It was quite conceivable at one time that the result would turn out to be false. Whether you would have interpreted this as telling you that the foundations of the subject need to be cleaned up, or that the universe of mathematics is more bizarre than even the real world, is a matter of taste and judgement, but your intuitions based upon classical calculus are a poor guide to what is provable about manifolds or more exotic spaces. The fact that the above result takes a long time to prove tells you that it is deeper and more subtle than may appear. The fact that the machinery needed to prove it is capable of proving other less intutitive results adds weight to this conclusion. If you feel vaguely cheated because all we have proved is a small number of results which you believed in advance of reading this book, you are quite mistaken; you should feel pleased at having mastered such a powerful machine. There is nothing quite so destructive of careful analysis of hard problems, as the conviction that they are rather easy.

Human intuition is very valuable, but not to be trusted as a source of truth. Adolph Hitler and many other loonies throughout history have thought otherwise, and most of them have caused enormous trouble. Doubt everything, particularly the things you would very much like to believe
HeavenForBooks.com

Bibliography
[1] Edwin Spanier. Algebraic Topology. McGraw-Hill, New York, 1966. [2] William Massey. A Basic Course in Algebraic Topology. Springer-Verlag, New York 1991. [3] George F Simmons. Introduction to Topology and Modern Analysis. McGraw-Hill, New York, 1963. [4] Peter Freyd Abelian Categories. Harper, New York, 1965.

141

You might also like