• 沒有找到結果。

The Fundamental Group of the Circle

在文檔中 Allen Hatcher (頁 38-50)

To show that Φ is injective, suppose Φ(m) = Φ(n), which means ωm ' ωn. Let ft be a homotopy from ωm = f0 to ωn = f1. By (b) this homotopy lifts to a homotopy eft of paths starting at 0 . The uniqueness part of (a) implies that ef0= fωm and ef1 = fωn. Since eft is a homotopy of paths, the endpoint eft(1) is independent oft . For t= 0 this endpoint is m and for t = 1 it is n, so m = n.

It remains to prove (a) and (b). Both statements can be deduced from a more general assertion:

(c) Given a map F : Y×I

S1 and a map eF : Y×{0}

R lifting F|Y ×{0}, then there is a unique map eF : Y×I

R lifting F and restricting to the given eF on Y×{0}.

Statement (a) is the special case that Y is a point, and (b) is obtained by applying (c) with Y = I in the following way. The homotopy ft in (b) gives a mapF : I×I

S1 by

setting F (s, t)= ft(s) as usual. A unique lift eF : I×{0}

R is obtained by an appli-cation of (a). Then (c) gives a unique lift eF : I×I

R. The restrictions eF|{0}×I and F|{1}×I are paths lifting the constant path at xe 0, hence they must also be constant by the uniqueness part of (a). So eft(s)= eF (s, t) is a homotopy of paths, and eft lifts ft since p eF= F .

We shall prove (c) using just one special property of the projection p :R

S1,

namely:

(∗)

There is an open cover {Uα} of S1 such that for each α , p−1(Uα) can be decomposed as a disjoint union of open sets each of which is mapped homeo-morphically onto Uα by p .

For example, we could take the cover {Uα} to consist of any two open arcs in S1 whose union isS1.

To prove (c) we will first construct a lift eF : N×I

R for N some neighborhood in Y of a given point y0∈ Y . Since F is continuous, every point (y0, t)∈ Y ×I has a product neighborhood Nt×(at, bt) such that F Nt×(at, bt)

⊂ Uα for some α . By compactness of {y0}×I , finitely many such products Nt×(at, bt) cover {y0}×I . This implies that we can choose a single neighborhood N of y0 and a partition 0 = t0 < t1 < ··· < tm = 1 of I so that for each i, F(N ×[ti, ti+1]) is contained in some Uα, which we denote Ui. Assume inductively that eF has been constructed on N×[0, ti] . We have F (N×[ti, ti+1])⊂ Ui, so by (∗) there is an open set eUi ⊂ R projecting homeomorphically ontoUi by p and containing the point eF (y0, ti) . After replacingN by a smaller neighborhood of y0 we may assume that eF (N×{ti}) is con-tained in eUi, namely, replaceN×{ti} by its intersection with ( eF ||N ×{ti})−1( eUi) . Now we can define eF on N×[ti, ti+1] to be the composition of F with the homeomorphism p−1:Ui

Uei. After finitely many repetitions of this induction step we eventually get a lift eF : N×I

R for some neighborhood N of y0.

Next we show the uniqueness part of (c) in the special case thatY is a point. In this case we can omitY from the notation. So suppose eF and eF0are two lifts ofF : I

S1

such that eF (0)= eF0(0) . As before, choose a partition 0= t0< t1<··· < tm = 1 of I so that for each i , F ([ti, ti+1]) is contained in some Ui. Assume inductively that Fe= eF0on[0, ti] . Since [ti, ti+1] is connected, so is eF ([ti, ti+1]) , which must therefore lie in a single one of the disjoint open sets eUi projecting homeomorphically to Ui as in(∗). By the same token, eF0([ti, ti+1]) lies in a single eUi, in fact in the same one that contains eF ([ti, ti+1]) since eF0(ti)= eF (ti) . Because p is injective on eUi andp eF = p eF0, it follows that eF = eF0on [ti, ti+1] , and the induction step is finished.

The last step in the proof of (c) is to observe that since the eF ’s constructed above on sets of the form N×I are unique when restricted to each segment {y}×I , they must agree whenever two such sets N×I overlap. So we obtain a well-defined lift eF on all of Y×I . This eF is continuous since it is continuous on each N×I , and it is

unique since it is unique on each segment {y}×I . tu

Now we turn to some applications of this theorem. Although algebraic topology is usually ‘algebra serving topology,’ the roles are reversed in the following proof of the Fundamental Theorem of Algebra.

T

heorem 1.8. Every nonconstant polynomial with coefficients in C has a root in C.

P

roof: We may assume the polynomial is of the form p(z)= zn+ a1zn−1+ ··· + an. If p(z) has no roots in C, then for each real number r ≥ 0 the formula

fr(s)= p(r e2π is)/p(r )

|p(r e2π is)/p(r )|

defines a loop in the unit circleS1⊂ C based at 1. As r varies, fr is a homotopy of loops based at 1 . Since f0 is the trivial loop, we deduce that the class [fr]∈ π1(S1) is zero for all r . Now fix a large value of r , bigger than|a1| + ··· + |an| and bigger than 1 . Then for |z| = r we have

|zn| = rn= r · rn−1> (|a1| + ··· + |an|)|zn−1| ≥ |a1zn−1+ ··· + an|

From the inequality|zn| > |a1zn−1+ ··· + an| it follows that the polynomial pt(z)= zn+t(a1zn−1+···+an) has no roots on the circle|z| = r when 0 ≤ t ≤ 1. Replacing p by pt in the formula forfr above and letting t go from 1 to 0 , we obtain a homo-topy from the loop fr to the loop ωn(s)= e2π ins. By Theorem 1.7, ωn represents n times a generator of the infinite cyclic group π1(S1) . Since we have shown that n]= [fr]= 0, we conclude that n = 0. Thus the only polynomials without roots

in C are constants. tu

Our next application is the Brouwer fixed point theorem in dimension 2 .

T

heorem 1.9. Every continuous map h : D2

D2 has a fixed point, that is, a point x with h(x)= x .

Here we are using the standard notation Dn for the closed unit disk in Rn, all vectorsx of length|x| ≤ 1. Thus the boundary of Dn is the unit sphereSn−1.

P

roof: Suppose on the contrary that h(x)≠ x for all x ∈ D2. Then we can define a map r : D2

S1 by lettingr (x) be the point of S1 where the ray inR2 starting at h(x) and passing

throughx leaves D2. Continuity ofr is clear since small per- x r(x)

h(x)

turbations of x produce small perturbations of h(x) , hence also small perturbations of the ray through these two points.

The crucial property of r , besides continuity, is that r (x)= x if x ∈ S1. Thus r is a retraction ofD2 ontoS1. We will show that no such retraction can exist.

Let f0 be any loop inS1. InD2 there is a homotopy of f0 to a constant loop, for example the linear homotopy ft(s)= (1 − t)f0(s)+ tx0 where x0 is the basepoint of f0. Since the retraction r is the identity on S1, the composition r ft is then a homotopy in S1 from r f0= f0 to the constant loop at x0. But this contradicts the

fact thatπ1(S1) is nonzero. tu

This theorem was first proved by Brouwer around 1910, one of the early triumphs of algebraic topology. Brouwer in fact proved the corresponding result for Dn, and we shall obtain this generalization in Corollary 2.11 using homology groups in place of π1. One could also use the higher homotopy group πn. Brouwer’s original proof used neither homology nor homotopy groups, which had not been invented at the time. Instead it used the notion of degree for mapsSn

Sn, which we shall define in

§2.2 using homology but which Brouwer defined directly in more geometric terms.

These proofs are all arguments by contradiction, and so they show just the exis-tence of fixed points without giving any clue as to how to find one in explicit cases.

Our proof of the Fundamental Theorem of Algebra was similar in this regard. There exist other proofs of the Brouwer fixed point theorem that are somewhat more con-structive, for example the elegant and quite elementary proof by Sperner in 1928, which is explained very nicely in [Aigner-Ziegler 1999].

The techniques used to calculateπ1(S1) can be applied to prove the Borsuk–Ulam theorem in dimension two:

T

heorem 1.10. For every continuous map f : S2

R2there exists a pair of antipodal points x and−x in S2 with f (x)= f (−x).

It may be that there is only one such pair of antipodal points x ,−x , for example iff is simply orthogonal projection of the standard sphere S2⊂ R3 onto a plane.

The Borsuk–Ulam theorem holds also for maps Sn

Rn, as we show in Proposi-tion 2B.6. The proof forn= 1 is easy since the difference f (x)−f (−x) changes sign as x goes halfway around the circle, hence this difference must be zero for some x . For n ≥ 2 the theorem is certainly less obvious. Is it apparent, for example, that at every instant there must be a pair of antipodal points on the surface of the earth having the same temperature and the same barometric pressure?

The theorem says in particular that there is no one-to-one continuous map from S2 toR2, soS2 is not homeomorphic to a subspace of R2, an intuitively obvious fact that is not easy to prove directly.

P

roof: If the conclusion is false for f : S2

R2, we can define a map g : S2

S1 by

g(x) = f (x) − f (−x)

/|f (x) − f (−x)|. Define a loop η circling the equator of S2⊂ R3 by η(s)= (cos 2πs, sin 2πs, 0), and let h : I

S1 be the composed loop gη . Sinceg(−x) = −g(x), we have the relation h(s +1/2)= −h(s) for all s in the interval [0,1/2]. As we showed in the calculation of π1(S1) , the loop h can be lifted to a path eh : I

R. The equation h(s +1/2) = −h(s) implies that eh(s +1/2) = eh(s) +q/2 for some odd integer q that might conceivably depend on s ∈ [0,1/2]. But in fact q is independent ofs since by solving the equation eh(s+1/2)= eh(s)+q/2forq we see that q depends continuously on s∈ [0,1/2], so q must be a constant since it is constrained to integer values. In particular, we have eh(1)= eh(1/2)+q/2 = eh(0) + q. This means thath represents q times a generator of π1(S1) . Since q is odd, we conclude that h is not nullhomotopic. Buth was the composition gη : I

S2

S1, andη is obviously nullhomotopic in S2, sogη is nullhomotopic in S1 by composing a nullhomotopy of η with g . Thus we have arrived at a contradiction. tu

C

orollary 1.11. Whenever S2 is expressed as the union of three closed sets A1, A2, and A3, then at least one of these sets must contain a pair of antipodal points{x, −x}.

P

roof: Let di:S2

R measure distance to Ai, that is, di(x)= infy∈Ai|x − y|. This is a continuous function, so we may apply the Borsuk–Ulam theorem to the map S2

R2, x

,

d1(x), d2(x), obtaining a pair of antipodal points x and −x with d1(x)= d1(−x) and d2(x)= d2(−x). If either of these two distances is zero, then x and−x both lie in the same set A1 or A2 since these are closed sets. On the other hand, if the distances from x and −x to A1 and A2 are both strictly positive, then x and −x lie in neither A1 nor A2 so they must lie in A3. tu

To see that the number ‘three’ in this result is best possible, consider a sphere inscribed in a tetrahedron. Projecting the four faces of the tetrahedron radially onto the sphere, we obtain a cover ofS2 by four closed sets, none of which contains a pair of antipodal points.

Assuming the higher-dimensional version of the Borsuk–Ulam theorem, the same arguments show that Sn cannot be covered by n+ 1 closed sets without antipodal pairs of points, though it can be covered byn+2 such sets, as the higher-dimensional analog of a tetrahedron shows. Even the case n= 1 is somewhat interesting: If the circle is covered by two closed sets, one of them must contain a pair of antipodal points. This is of course false for nonclosed sets since the circle is the union of two disjoint half-open semicircles.

The relation between the fundamental group of a product space and the funda-mental groups of its factors is as simple as one could wish:

P

roposition 1.12. π1(X×Y ) is isomorphic to π1(X)×π1(Y ) if X and Y are path-connected.

P

roof: A basic property of the product topology is that a map f : Z

X×Y is

con-tinuous iff the mapsg : Z

X and h : Z

Y defined by f (z)= (g(z), h(z)) are both continuous. Hence a loopf in X×Y based at (x0, y0) is equivalent to a pair of loops g in X and h in Y based at x0andy0respectively. Similarly, a homotopyft of a loop in X×Y is equivalent to a pair of homotopies gt and ht of the corresponding loops in X and Y . Thus we obtain a bijection π1 X×Y , (x0, y0)

≈ π1(X, x0)×π1(Y , y0) , [f ]

,

([g], [h]) . This is obviously a group homomorphism, and hence an

isomor-phism. tu

E

xample 1.13: The Torus. By the proposition we have an isomorphism π1(S1×S1)≈ Z×Z. Under this isomorphism a pair (p, q) ∈ Z×Z corresponds to a loop that winds p times around one S1 factor of the torus andq times around the

other S1 factor, for example the loop ωpq(s)= (ωp(s), ωq(s)) . Interestingly, this loop can be knotted, as the figure shows for the case p= 3, q = 2. The knots that arise in this fashion, the so-calledtorus knots, are studied in Example 1.24.

More generally, the n dimensional torus, which is the product of n circles, has fundamental group isomorphic to the product of n copies of Z. This follows by induction onn .

Induced Homomorphisms

Supposeϕ : X

Y is a map taking the basepoint x0∈ X to the basepoint y0∈ Y . For brevity we write ϕ : (X, x0)

(Y , y0) in this situation. Then ϕ induces a homo-morphismϕ:π1(X, x0)

π1(Y , y0) , defined by composing loops f : I

X based at

x0 with ϕ , that is, ϕ[f ] = [ϕf ]. This induced map ϕ is well-defined since a homotopyft of loops based at x0 yields a composed homotopy ϕft of loops based aty0, so ϕ[f0]= [ϕf0]= [ϕf1]= ϕ[f1] . Furthermore, ϕ is a homomorphism since ϕ(f g)= (ϕf ) (ϕg), both functions having the value ϕf (2s) for 0 ≤ s ≤1/2 and the value ϕg(2s− 1) for1/2≤ s ≤ 1.

Two basic properties of induced homomorphisms are:

(ϕψ)= ϕψ for a composition (X, x0)

---→

ψ (Y , y0)

---→

ϕ (Z, z0) .

11=11 , which is a concise way of saying that the identity map11 :X

X induces

the identity map 11 :π1(X, x0)

π1(X, x0) .

The first of these follows from the fact that composition of maps is associative, so (ϕψ)f = ϕ(ψf ), and the second is obvious. These two properties of induced homo-morphisms are what makes the fundamental group a functor. The formal definition

of a functor requires the introduction of certain other preliminary concepts, however, so we postpone this until it is needed in§2.3.

Ifϕ is a homeomorphism with inverse ψ then ϕis an isomorphism with inverse ψ sinceϕψ= (ϕψ)=11=11 and similarly ψϕ=11 . We will use this fact in the following calculation of the fundamental groups of higher-dimensional spheres:

P

roposition 1.14. π1(Sn)= 0 if n ≥ 2.

P

roof: Let f be a loop in Sn at a chosen basepoint x0. If the image off is disjoint from some other point x ∈ Sn then f is nullhomotopic since Sn− {x} is homeo-morphic to Rn, which is simply-connected. So it will suffice to homotope f to be nonsurjective. To do this we will look at a small open ball B in Sn about any point x ≠ x0 and see that the number of times that f enters B , passes

throughx , and leaves B is finite, and each of these portions of f can be pushed offx without changing the rest of f . At first glance this might appear to be a difficult task to achieve since the parts off in B could be quite complicated geometrically, for example space-filling curves. But in fact it turns out to be rather easy.

The setf−1(B) is open in (0, 1) , hence is the union of a possibly infinite collection of disjoint open intervals(ai, bi) . The compact set f−1(x) is contained in the union of these intervals, so it must be contained in the union of finitely many of them. Consider one of the intervals(ai, bi) meeting f−1(x) . The path fi obtained by restrictingf to the closed interval[ai, bi] lies in the closure of B , and its endpoints f (ai) and f (bi) lie in the boundary of B . If n≥ 2, we can choose a path gi from f (ai) to f (bi) in the closure of B but disjoint from x . For example, we could choose gi to lie in the boundary ofB , which is a sphere of dimension n− 1, hence path-connected if n ≥ 2.

Since the closure of B is homeomorphic to a convex set in Rn and hence simply-connected, the pathfi is homotopic to gi by Proposition 1.6, so we may homotopef by deformingfi to gi. After repeating this process for each of the intervals (ai, bi) that meet f−1(x) , we obtain a loop g homotopic to the original f and with g(I)

disjoint fromx . tu

E

xample 1.15. For a point x in Rn, the complementRn− {x} is homeomorphic to Sn−1×R, so by Proposition 1.12 π1(Rn− {x}) is isomorphic to π1(Sn−1)×π1(R).

Hence π1(Rn− {x}) is Z for n = 2 and trivial for n > 2.

Here is an application of this calculation:

C

orollary 1.16. R2 is not homeomorphic to Rn for n≠ 2.

P

roof: Suppose f :R2

Rn is a homeomorphism. The casen= 1 is easily disposed of since R2− {0} is path-connected but the homeomorphic space Rn− {f (0)} is not path-connected when n = 1. When n > 2 we cannot distinguish R2− {0} from

Rn− {f (0)} by the number of path-components, but by the preceding calculation of π1(Rn− {x}) we can distinguish them by their fundamental groups. tu

The more general statement that Rm is not homeomorphic to Rn if m≠ n can be proved in the same way using either the higher homotopy groups or homology groups. In fact, nonempty open sets in Rm and Rn can be homeomorphic only if m= n, as we will show in Theorem 2.19 using homology.

Induced homomorphisms allow relations between spaces to be transformed into relations between their fundamental groups. Here is an illustration of this principle:

P

roposition 1.17. If a space X retracts onto a subspace A , then the homomorphism i:π1(A, x0)

π1(X, x0) induced by the inclusion i : A

>

X is injective. If A is a deformation retract of X , then i is an isomorphism.

P

roof: If r : X

A is a retraction, then r i=11 , henceri=11 , which implies thati is injective. Ifrt:X

X is a deformation retraction of X onto A , so r0=11 ,rt|A =11 , andr1(X)⊂ A, then for any loop f : I

X based at x0∈ A the composition rtf gives a homotopy of f to a loop in A , so i is also surjective. tu

This gives another way of seeing thatS1 is not a retract ofD2, a fact we showed earlier in the proof of the Brouwer fixed point theorem, since the inclusion-induced map π1(S1)

π1(D2) is a homomorphismZ

0 that cannot be injective.

The exact group-theoretic analog of a retraction is a homomorphismρ of a group G onto a subgroup H such that ρ restricts to the identity on H . In the notation above, if we identifyπ1(A) with its image under i, thenris such a homomorphism fromπ1(X) onto the subgroup π1(A) . The existence of a retracting homomorphism ρ : G

H is quite a strong condition on H . If H is a normal subgroup, it implies that G is the direct product of H and the kernel of ρ . If H is not normal, then G is what is called in group theory the semi-direct product of H and the kernel of ρ .

Recall from Chapter 0 the general definition of a homotopy as a familyϕt:X

Y ,

t∈ I , such that the associated map Φ : X×I

Y ,Φ(x, t) = ϕt(x) , is continuous. If ϕt takes a subspaceA⊂ X to a subspace B ⊂ Y for all t , then we speak of a homotopy of maps of pairs, ϕt:(X, A)

(Y , B) . In particular, a basepoint-preserving homotopy ϕt:(X, x0)

(Y , y0) is the case that ϕt(x0)= y0 for all t . Another basic property of induced homomorphisms is their invariance under such homotopies:

If ϕt:(X, x0)

(Y , y0) is a basepoint-preserving homotopy, then ϕ0∗= ϕ1∗. This holds since ϕ0∗[f ] = [ϕ0f ]= [ϕ1f ]= ϕ1∗[f ] , the middle equality coming from the homotopy ϕtf .

There is a notion of homotopy equivalence for spaces with basepoints. One says (X, x0) ' (Y , y0) if there are maps ϕ : (X, x0)

(Y , y0) and ψ : (Y , y0)

(X, x0)

with homotopies ϕψ ' 11 and ψϕ ' 11 through maps fixing the basepoints. In this case the induced maps on π1 satisfy ϕψ = (ϕψ) =11 =11 and likewise ψϕ =11 , so ϕ and ψ are inverse isomorphisms π1(X, x0) ≈ π1(Y , y0) . This somewhat formal argument gives another proof that a deformation retraction induces an isomorphism on fundamental groups, since ifX deformation retracts onto A then (X, x0)' (A, x0) for any choice of basepoint x0∈ A.

Having to pay so much attention to basepoints when dealing with the fundamental group is something of a nuisance. For homotopy equivalences one does not have to be quite so careful, as the conditions on basepoints can actually be dropped:

P

roposition 1.18. If ϕ : X

Y is a homotopy equivalence, then the induced homo-morphism ϕ:π1(X, x0)

π1 Y , ϕ(x0)is an isomorphism for all x0∈ X .

The proof will use a simple fact about homotopies that do not fix the basepoint:

L

emma 1.19. If ϕt:X

Y is a homotopy and h is the path ϕt(x0) formed by the images of a basepoint x0∈ X , then the three maps in the diagram at the right satisfy ϕ0∗= βhϕ1∗.

π1(X,x0)

π1(Y,ϕ1(x0))

ϕ1

π1(Y,ϕ0(x0))

ϕ0

βh

−−−−−→

−−−−−−−−→ −−−−−−−−→

P

roof: Let ht be the restriction ofh to the interval [0, t] , with a reparametrization so that the domain ofht is still [0, 1] . Explicitly, we can take ht(s)= h(ts). Then if f is a loop inX at the basepoint x0, the productht tf ) ht

gives a homotopy of loops at ϕ0(x0) . Restricting this x

0

ϕ0

0f ϕ

tf ϕ

1f ϕ

( ) x0

1

ht

t

ϕ ( )

x0 ϕ ( )

homotopy to t = 0 and t = 1, we see that ϕ0∗([f ])= βh ϕ1∗([f ])

. tu

P

roof of 1.18: Let ψ : Y

X be a homotopy-inverse for ϕ , so that ϕψ ' 11 and ψϕ'11 . Consider the maps

π1(X, x0)

---→

ϕ π1 Y , ϕ(x0)

---→

ψ π1 X, ψϕ(x0)

---→

ϕ π1 Y , ϕψϕ(x0)

The composition of the first two maps is an isomorphism sinceψϕ'11 implies that ψϕ= βh for someh , by the lemma. In particular, since ψϕ is an isomorphism, ϕ is injective. The same reasoning with the second and third maps shows that ψ is injective. Thus the first two of the three maps are injections and their composition is an isomorphism, so the first mapϕ must be surjective as well as injective. tu

Exercises

1. Show that composition of paths satisfies the following cancellation property: If f0 g0' f1 g1 and g0' g1 then f0' f1.

2. Show that the change-of-basepoint homomorphism βh depends only on the homo-topy class of h .

3. For a path-connected space X , show that π1(X) is abelian iff all basepoint-change homomorphismsβh depend only on the endpoints of the path h .

4. A subspace X⊂ Rn is said to bestar-shaped if there is a point x0∈ X such that, for each x ∈ X , the line segment from x0 to x lies in X . Show that if a subspace X ⊂ Rn is locally star-shaped, in the sense that every point of X has a star-shaped neighborhood in X , then every path in X is homotopic in X to a piecewise linear path, that is, a path consisting of a finite number of straight line segments traversed at constant speed. Show this applies in particular when X is open or when X is a union of finitely many closed convex sets.

5. Show that for a space X , the following three conditions are equivalent:

(a) Every map S1

X is homotopic to a constant map, with image a point.

(b) Every map S1

X extends to a map D2

X .

(c) π1(X, x0)= 0 for all x0∈ X .

Deduce that a space X is simply-connected iff all maps S1

X are homotopic. [In this problem, ‘homotopic’ means ‘homotopic without regard to basepoints.’]

6. We can regard π1(X, x0) as the set of basepoint-preserving homotopy classes of maps (S1, s0)

(X, x0) . Let [S1, X] be the set of homotopy classes of maps S1

X ,

with no conditions on basepoints. Thus there is a natural mapΦ : π1(X, x0)

[S1, X]

obtained by ignoring basepoints. Show thatΦ is onto if X is path-connected, and that Φ([f ]) = Φ([g]) iff [f ] and [g] are conjugate in π1(X, x0) . Hence Φ induces a one-to-one correspondence between [S1, X] and the set of conjugacy classes in π1(X) , whenX is path-connected.

7. Define f : S1×I

S1×I by f (θ, s) = (θ + 2πs, s), so f restricts to the identity on the two boundary circles of S1×I . Show that f is homotopic to the identity by a homotopy ft that is stationary on one of the boundary circles, but not by any ho-motopyft that is stationary on both boundary circles. [Consider what f does to the path s

,

0, s) for fixed θ0∈ S1.]

8. Does the Borsuk–Ulam theorem hold for the torus? In other words, for every map f : S1×S1

R2 must there exist(x, y)∈ S1×S1 such thatf (x, y)= f (−x, −y)?

9. Let A1, A2, A3 be compact sets in R3. Use the Borsuk–Ulam theorem to show that there is one planeP ⊂ R3 that simultaneously divides eachAi into two pieces of equal measure.

10. From the isomorphism π1 X×Y , (x0, y0)

≈ π1(X, x0)×π1(Y , y0) it follows that loops inX×{y0} and {x0}×Y represent commuting elements of π1 X×Y , (x0, y0)

. Construct an explicit homotopy demonstrating this.

11. If X0 is the path-component of a spaceX containing the basepoint x0, show that the inclusion X0

>

X induces an isomorphism π1(X0, x0)

π1(X, x0) .

12. Show that every homomorphism π1(S1)

π1(S1) can be realized as the induced homomorphismϕ of a map ϕ : S1

S1.

13. Given a space X and a path-connected subspace A containing the basepoint x0, show that the mapπ1(A, x0)

π1(X, x0) induced by the inclusion A

>

X is surjective iff every path in X with endpoints in A is homotopic to a path in A .

14. Show that the isomorphism π1(X×Y ) ≈ π1(X)×π1(Y ) in Proposition 1.12 is given by [f ]

,

(p1∗([f ]), p2∗([f ])) where p1 and p2 are the projections of X×Y onto its two factors.

15. Given a map f : X

Y and a path h : I

X

from x0 to x1, show that fβh = βf hf in the diagram at the right.

X,

π1( x1) π1(X,x0)

f

Y, π1( f(x0)) Y,

π1( f(x1))

βh

βfh

−−−−−→ −−−−−−−−−−−→ −−−−−→

f

16. Show that there are no retractions r : X

A in the following cases:

(a) X = R3 with A any subspace homeomorphic to S1. (b) X = S1×D2 with A its boundary torus S1×S1. (c) X = S1×D2 and A the circle shown in the figure.

(d) X = D2∨ D2 with A its boundary S1∨ S1.

(e) X a disk with two points on its boundary identified and A its boundary S1∨ S1. (f) X the M¨obius band andA its boundary circle.

17. Construct infinitely many nonhomotopic retractions S1∨ S1

S1.

18. Using the technique in the proof of Proposition 1.14, show that if a space X is obtained from a path-connected subspace A by attaching a cell en with n≥ 2, then the inclusion A

>

X induces a surjection on π1. Apply this to show:

(a) The wedge sumS1∨ S2 has fundamental group Z.

(b) For a path-connected CW complexX the inclusion map X1

>

X of its 1 skeleton induces a surjection π1(X1)

π1(X) . [For the case that X has infinitely many cells, see Proposition A.1 in the Appendix.]

19. Modify the proof of Proposition 1.14 to show that if X is a path-connected 1 dimensional CW complex with basepoint x0 a 0 cell, then every loop in X is ho-motopic to a loop consisting of a finite sequence of edges traversed monotonically.

[This gives an elementary proof that π1(S1) is cyclic, generated by the standard loop winding once around the circle. The more difficult part of the calculation of π1(S1) is therefore the fact that no iterate of this loop is nullhomotopic.]

20. Suppose ft:X

X is a homotopy such that f0 andf1 are each the identity map.

Use Lemma 1.19 to show that for anyx0∈ X , the loop ft(x0) represents an element of the center ofπ1(X, x0) . [One can interpret the result as saying that a loop represents an element of the center ofπ1(X) if it extends to a loop of maps X

X .]

The van Kampen theorem gives a method for computing the fundamental groups of spaces that can be decomposed into simpler spaces whose fundamental groups are already known. By systematic use of this theorem one can compute the fundamental groups of a very large number of spaces. We shall see for example that for every group G there is a space XG whose fundamental group is isomorphic toG .

To give some idea of how one might hope to compute fundamental groups by decomposing spaces into simpler pieces, let us look at an example. Consider the space X formed by two circles A and B intersecting in a single point, which we choose as the basepoint x0. By our preceding calculations we know that π1(A) is infinite cyclic, generated by a loopa that goes once around A .

Similarly, π1(B) is a copy of Z generated by a loop b going b a once aroundB . Each product of powers of a and b then gives

an element of π1(X) . For example, the product a5b2a−3ba2 is the loop that goes five times aroundA , then twice around B , then three times around A in the opposite direction, then once around B , then twice around A . The set of all words like this consisting of powers ofa alternating with powers of b forms a group usually denoted Z ∗ Z. Multiplication in this group is defined just as one would expect, for example (b4a5b2a−3)(a4b−1ab3)= b4a5b2ab−1ab3. The identity element is the empty word, and inverses are what they have to be, for example (ab2a−3b−4)−1 = b4a3b−2a−1. It would be very nice if such words in a and b corresponded exactly to elements of π1(X) , so that π1(X) was isomorphic to the group Z ∗ Z. The van Kampen theorem will imply that this is indeed the case.

Similarly, if X is the union of three circles touching at a single point, the van Kampen theorem will imply that π1(X) is Z ∗ Z ∗ Z, the group consisting of words in powers of three letters a , b , c . The generalization to a union of any number of circles touching at one point will also follow.

The group Z ∗ Z is an example of a general construction called the free product of groups. The statement of van Kampen’s theorem will be in terms of free products, so before stating the theorem we will make an algebraic digression to describe the construction of free products in some detail.

在文檔中 Allen Hatcher (頁 38-50)