MAS61015 Algebraic Topology

9. Homotopy

Video (Definition 9.1 to Example 9.6)

We would like to make the following definition:

Definition 9.1 (Bogus).

Let X and Y be topological spaces, and let f,g:XY be two continuous maps from X to Y, or in other words, two elements of Top(X,Y). We say that f and g are homotopic if there is a path u:[0,1]Top(X,Y) from f to g, so that f and g lie in the same path component of Top(X,Y). This is an equivalence relation, by Proposition 5.9; the equivalence classes are called homotopy classes of maps. We define [X,Y]=π0Top(X,Y), so [X,Y] is the set of all homotopy classes of maps from X to Y.

This is bogus, because we have not introduced a topology on the set Top(X,Y), so it is not meaningful to talk about continuous paths. As we mentioned in Remark 7.28, it is possible to introduce an appropriate topology on Top(X,Y), but this involves many subtle technicalities. We therefore reformulate the definition in a different way.

Definition 9.2.

Let X and Y be topological spaces, and let f,g:XY be two continuous maps from X to Y. A homotopy from f to g is a continuous map h:[0,1]×XY such that h(0,x)=f(x) and h(1,x)=g(x) for all xX. We say that f and g are homotopic if there is a homotopy between them, and we write fg in this case.

Proposition 9.3.

The relation of being homotopic is an equivalence relation.

Proof.

This is closely analogous to Proposition 5.9.

  • (a)

    We can define a homotopy h from f to f by h(t,x)=f(x) for all t. This proves that ff, so the relation is reflexive.

  • (b)

    Suppose that fg, so we can choose a homotopy h from f to g. The function h¯(t,x)=h(1-t,x) then gives a homotopy from g to f, proving that gf. Thus, the relation is symmetric.

  • (c)

    Now suppose that ef and fg, so we can choose a homotopy k from e to f, and another homotopy h from f to g. We then define k*h:[0,1]×XY by

    (k*h)(t,x)={k(2t,x) if 0t12h(2t-1,x) if 12t1.

    The two clauses are consistent, because when t=12 we have

    k(2t,x)=k(1,x)=f(x)=h(0,x)=h(2t-1,x).

    The combined map k*h is easily seen to be continuous on when restricted to the sets [0,12]×X and [12,1]×X. These sets are closed (in the product topology), and their union is all of [0,1]×X, so the full map k*h is continuous by Proposition 3.35. It gives a homotopy from e to g, proving that the homotopy relation is transitive.

Definition 9.4.

The equivalence class [f] of a continuous map f:XY is called the homotopy class of f. We define [X,Y]=Top(X,Y)/, so this is the set of all homotopy classes.

Example 9.5.

Consider a continuous map f:S1{0}. Then f represents a closed loop in the complex plane, which cannot pass through the origin, but which can wind around the origin. Let n(f) be the total number of times the curve winds around the origin, with anticlockwise turns counting positively, and clockwise turns counting negatively. We call this the winding number of f. (We will formulate the definition more carefully at a later stage.)

It can be shown that two maps from S1 to {0} are homotopic iff they have the same winding number. It follows that the map n:Top(S1,{0}) induces a bijection [S1,{0}].

Example 9.6.

Let X be an arbitrary topological space, and let Y be a subset of n, with the subspace topology. Let f,g:XY be continuous maps. We can then define a map h:[0,1]×Xn by

h(t,x)=(1-t)f(x)+tg(x).

In general, we have no right to expect that this will land in the subspace Yn, so it will probably not provide a homotopy between f and g. However, in some special cases we may be able to prove that h([0,1]×X)Y, and in that case we do get a homotopy from f to g, which we will call a straight line homotopy.

For the most extreme example of this, suppose that Y is convex (as in Example 5.11), so that for all a,bY, the line segment from a to b is contained wholly in Y. Let f and g be continuous maps from X to Y, and put h(t,x)=(1-t)f(x)+tg(x) as before. Then h(t,x) lies on the line segment from f(x)Y to g(x)Y, so h(t,x)Y for all t and x. Thus, h gives a homotopy from f to g, proving that [f]=[g] in [X,Y]. As f and g were arbitrary, it follows that |[X,Y]|=1.

We next check that the notion of homotopy is compatible with composition.

Video (Proposition 9.7 to Definition 9.9)

Proposition 9.7.

Suppose we have continuous maps

Suppose that f0 is homotopic to f1, and that g0 is homotopic to g1. Then g1f1 is homotopic to g0f0.

Proof.

We are assuming that f0 is homotopic to f1, which means that there is a homotopy h:[0,1]×XY with h(0,x)=f0(x) and h(1,x)=f1(x) for all xX. We are also assuming that g0 is homotopic to g1, which means that there is a homotopy k:[0,1]×YZ with k(0,y)=g0(y) and k(1,y)=g1(y) for all yY. We can therefore define m:[0,1]×XZ by

m(t,x)=k(t,h(t,x)).

This satisfies m(0,x)=k(0,h(0,x))=k(0,f0(x))=g0(f0(x)) and m(1,x)=k(1,h(1,x))=k(1,f1(x))=g1(f1(x)). Thus, m gives the required homotopy, provided that we can check that it is continuous. The tidiest proof of continuity is as follows: we let p:[0,1]×X[0,1] be the projection, and note that m can be written as the composite

[0,1]×Xp,h[0,1]×Y𝑘Z.

Here p is continuous by Lemma 7.10, so p,h is continuous by Proposition 7.11, so m is continuous by Proposition 3.24. ∎

Corollary 9.8.

For u[X,Y] and v[Y,Z], there is a well-defined composite vu[X,Z], given by vu=[gf] for any choice of maps f,g with u=[f] and v=[g].

Proof.

Immediate from the proposition. ∎

Definition 9.9.

We can now define a new category hTop, called the homotopy category. The objects are topological spaces (just as for the category Top), but the morphisms are now homotopy classes of maps, so

hTop(X,Y)=Top(X,Y)/=[X,Y].

The composition rule is given by Corollary 9.8. The identity morphism in hTop for an object X is just the homotopy class [idX] corresponding to the identity function.

Video (Definition 9.10 to Proposition 9.20)

Definition 9.10.

A continuous map f:XY is a homotopy equivalence if the corresponding homotopy class [f] is an isomorphism in the homotopy category. Explicitly, this means that there must exist a continuous map g:YX such that gfidX and fgidY. Any such map g is called a homotopy inverse for f. If there exists a homotopy equivalence from X to Y, we will say that X and Y are homotopy equivalent. We will sometimes use the notation XY to indicate this.

Lemma 9.11.

Every homeomorphism is a homotopy equivalence.

Proof.

Let f:XY be a homeomorphism, with inverse g:YX. This means that gf=idX and fg=idY. As gf is equal to idX, it is certainly homotopic to idX. As fg is equal to idY, it is certainly homotopic to idY. Thus, g is also a homotopy inverse for f, so f is a homotopy equivalence. ∎

A basic example is as follows:

Proposition 9.12.

For n>0, the sphere X=Sn-1 is homotopy equivalent to the space Y=n{0}.

Proof.

Recall that X=Sn-1={xn|x=1}. We define f:XY to be the inclusion function, so we just have f(x)=x. We define g:YX by g(y)=y/y. (It is important here that Y is the space of nonzero vectors in n, so y>0 and it is valid to divide by y.) Note that for xX we have x=1 so g(f(x))=x/x=x, so gf=idX. As gf is equal to idX, it is certainly homotopic to idX. In the opposite direction, however, the composite gf is not equal to the identity. Nonetheless, we can define h:[0,1]×Yn by

h(t,y)=(1-t)yy+ty=((1-t)y-1+t)y.

Here (1-t)y-1+t is always strictly positive (for 0t1), so h(t,y) can never be zero, so we can regard h as a map [0,1]×YY. Now h(0,y)=f(g(y)) and h(1,y)=y, so h is a homotopy from gf to idY. This proves that g is a homotopy inverse for f, so that f is a homotopy equivalence.

Definition 9.13.

Let X be a topological space, and let a be a point in X. A contraction of X to a is a continuous map h:[0,1]×XX such that h(0,x)=a for all x, and h(1,x)=x for all x. We say that X is contractible if it has a contraction (to some point aX).

Example 9.14.

Let Xn be a convex space, and let a be a point in X. We can then define a linear contraction of X to a by h(t,x)=(1-t)a+tx, so X is contractible. In particular, the spaces Bn, Δn and [0,1]n are all contractible.

Example 9.15.

The sphere S2 is not contractible, as you will probably agree if you try to imagine a contraction. However, at the moment we do not have any way of proving this.

Definition 9.16.

We write 1 for the set {0} (or any other set with precisely one element). We regard this as a topological space using the discrete topology (which is in fact the only possible topology in this case).

Proposition 9.17.

A space X is contractible iff it is homotopy equivalent to 1.

Proof.

Suppose that X is homotopy equivalent to 1. This means that we have maps f:X1 and g:1X, and a homotopy h from gf to idX, and a homotopy k from fg to id1. As 1={0} there is no choice about f: we must have f(x)=0 for all x. Similarly, there is no choice about k: we must have k(t,0)=0 for all t. However, there is some choice about g and h. Put a=g(0)X. We then have g(f(x))=g(0)=a for all x. Moreover, h is a homotopy from gf to idX, so we have h(0,x)=g(f(x))=a and h(1,x)=x for all x. Thus, h is a contraction t a, showing that X is contractible.

Conversely, suppose that we start from the assumption that X is contractible. All the above steps can be reversed in a straightforward way to prove that X is homotopy equivalent to 1. ∎

Remark 9.18.

The map [0,1]1 is the most basic example of a homotopy equivalence that is not a homeomorphism.

Proposition 9.19.

The relation of being homotopy equivalent is an equivalence relation.

Proof.

The identity function from X to itself is clearly a homotopy equivalence, so the relation is reflexive. Suppose that X is homotopy equivalent to Y, so we can choose a homotopy equivalence f:XY and a homotopy inverse g:YX, so fg:YY and gf:XX are homotopic to the respective identity maps. This means that g is a homotopy equivalence with homotopy inverse f, so Y is homotopy equivalent to X. This proves that our relation is symmetric. Finally, suppose that X is homotopy equivalent to Y and Y is homotopy equivalent to Z. Let d and e be homotopy inverses for f and g, so df1X and fd1Yeg and ge1Z. Using Proposition 9.7, we deduce that

degf=d(eg)fd1Yf=df1X
gfde=g(fd)eg1Ye=ge1Z.

This proves that gf:XZ is a homotopy equivalence, with homotopy inverse de:ZX. This in turn shows that our relation is transitive, as required. ∎

Proposition 9.20.

Suppose that X0 is homotopy equivalent to Y0 and X1 is homotopy equivalent to Y1. Then X0×X1 is homotopy equivalent to Y0×Y1.

Proof.

For i=0,1 we let fi:XiYi be a homotopy equivalence, with homotopy inverse gi:YiXi. This means that we have homotopies hi from gifi to idXi, and homotopies ki from figi to idYi. Now make the following definitions:

f :X0×X1Y0×Y1 f(x0,x1) =(f0(x0),f1(x1))
g :Y0×Y1X0×X1 g(y0,y1) =(g0(y0),g1(y1))
h :[0,1]×X0×X1X0×X1 h(t,x0,x1) =(h0(t,x0),h1(t,x1))
k :[0,1]×Y0×Y1Y0×Y1 k(t,y0,y1) =(k0(t,y0),k1(t,y1)).

We find that h gives a homotopy from gf to the identity, and k gives a homotopy from fg to the identity, so we have a homotopy equivalence between X0×X1 and Y0×Y1, as required. ∎

Example 9.21.

The solid torus, the Möbius band, and {0} are all homotopy equivalent to S1. To explain this in more detail, let D be the vertical disc in the xz plane of radius 1 centred at (2,0,0). The “solid torus” is the space obtained by revolving D around the z-axis; this is easily seen to be homeomorphic to S1×D2. Now D2 is convex, so it is homotopy equivalent to 1. It follows that S1×D2 is homotopy equivalent to S1×1=S1.

Next, for θ[0,2π], let Pθ be the vertical plane through the z-axis in 3 that has angle θ with the xz-plane. Let Dθ be the intersection of Pθ with the solid torus, which is a vertical disc of radius 1 centred at (2cos(θ),2sin(θ),0). Let Iθ be the diameter of Dθ that makes an angle of θ/4 to the vertical, and let M be the union of all the sets Dθ. This is a version of the Möbius band. It is homeomorphic to the space

M={(z,w)S1×B2|w2/z is real and nonnegative }.

Define f:S1M and g:MS1 and h:I×MM by

f(z) =(z,0)
g(z,w) =z
h(t,(z,w)) =(z,tw).

Then gf=1 and h is a homotopy from fg to 1, so g is a homotopy inverse for f, so M is homotopy equivalent to S1 as claimed.

Finally, Proposition 9.12 tells us that 2{0} is homotopy equivalent to S1, and can be identified with 2, so {0} is also homotopy equivalent to S1.

Definition 9.22.

Let X and Y be topological spaces. We say that X is a homotopy retract of Y if there exist continuous maps X𝑓Y𝑔X such that gf is homotopic to idX. (We make no assumption about fg.) Any pair (f,g) with this property will be called a homotopy retraction pair for (X,Y).

Example 9.23.

Put

X =2{0}
Y = figure eight ={(x,y)|(x-1)2+y2=1 or (x+1)2+y2=1}.

Note that X is two-dimensional whereas Y is one-dimensional so there is no injective continuous map from X to Y, so X is not an actual retract of Y. However, it is a homotopy retract of Y. To see this, define maps X𝑓Y𝑔X by

f(x,y) =(1,0)+(x,y)/(x,y)
g(x,y) =(x-1,y).

The composite gf:XX is just (gf)(x,y)=(x,y)/(x,y). The straight line joining (x,y) to (x,y)/(x,y) does not pass through the origin, so gf is homotopic to the identity as required. (The map fg:YY is not homotopic to the identity, so we do not have a homotopy equivalence, but that is not important.)

Proposition 9.24.

Suppose that X is a homotopy retract of Y and that Y is contractible. Then X is also contractible.

Proof.

By hypothesis, we have continuous maps

such that gf, qp and pq are homotopic to the respective identity maps. Now put m=pf:X1 and n=gq:1X. We then have nm=gqpfgfid:XX. Also, mn is a map from 1 to 1, and the only map from 1 to 1 is the identity, so mn=id. Thus, m and n give a homotopy equivalence from X to 1, proving that X is contractible. ∎

Proposition 9.25.

Suppose we have two continuous maps f,g:XY, giving maps f*,g*:π0(X)π0(Y) as in Proposition 5.20. If f is homotopic to g, then f*=g*.

Proof.

Let h:[0,1]×XY be a homotopy between f and g, so h(0,x)=f(x) and h(1,x)=g(x) for all x. For any aX we have f*[a]=[f(a)] and g*[a]=[g(a)]. We need to prove that these are the same. Equivalently, we need to find a path u from f(a) to g(a) in Y. We can just take u(t)=h(t,a). ∎