• 沒有找到結果。

Show that the existence of maximal trees is equivalent to the Axiom of Choice

在文檔中 Allen Hatcher (頁 96-105)

Deck Transformations and Group Actions

14. Show that the existence of maximal trees is equivalent to the Axiom of Choice

In this section we introduce a class of spaces whose homotopy type depends only on their fundamental group. These spaces arise many places in topology, especially in its interactions with group theory.

A path-connected space whose fundamental group is isomorphic to a given group G and which has a contractible universal covering space is called a K ( G , 1) space. The

‘1’ here refers toπ1. More generalK(G, n) spaces are studied in§4.2. All these spaces are called Eilenberg–MacLane spaces, though in the casen= 1 they were studied by

Hurewicz before Eilenberg and MacLane took up the general case. Here are some examples:

E

xample 1B.1. S1 is a K(Z, 1). More generally, a connected graph is a K(G, 1) with G a free group, since by the results of§1.A its universal cover is a tree, hence con-tractible.

E

xample 1B.2. Closed surfaces with infinite π1, in other words, closed surfaces other than S2 and RP2, are K(G, 1) ’s. This will be shown in Example 1B.14 below. It also follows from the theorem in surface theory that the only simply-connected surfaces without boundary are S2 and R2, so the universal cover of a closed surface with infinite fundamental group must be R2 since it is noncompact. Nonclosed surfaces deformation retract onto graphs, so such surfaces areK(G, 1) ’s with G free.

E

xample 1B.3. The infinite-dimensional projective space RP is a K(Z2, 1) since its universal cover isS, which is contractible. To show the latter fact, a homotopy from the identity map ofS to a constant map can be constructed in two stages as follows.

First, define ft:R

R by ft(x1, x2,···) = (1 − t)(x1, x2,···) + t(0, x1, x2,···).

This takes nonzero vectors to nonzero vectors for all t∈ [0, 1], so ft/|ft| gives a ho-motopy from the identity map ofS to the map(x1, x2,···)

,

(0, x1, x2,···). Then a homotopy from this map to a constant map is given bygt/|gt| where gt(x1, x2,···) = (1− t)(0, x1, x2,···) + t(1, 0, 0, ···).

E

xample 1B.4. Generalizing the preceding example, we can construct a K(Zm, 1) as an infinite-dimensional lens space S/Zm, where Zm acts on S, regarded as the unit sphere inC, by scalar multiplication bymthroots of unity, a generator of this action being the map (z1, z2,···)

,

e2π i/m(z1, z2,···). It is not hard to check that this is a covering space action.

E

xample 1B.5. A product K(G, 1)×K(H, 1) is a K(G×H, 1) since its universal cover is the product of the universal covers ofK(G, 1) and K(H, 1) . By taking products of circles and infinite-dimensional lens spaces we therefore get K(G, 1) ’s for arbitrary finitely generated abelian groups G . For example the n dimensional torus Tn, the product of n circles, is a K(Zn, 1) .

E

xample 1B.6. For a closed connected subspace K of S3 that is nonempty, the com-plementS3−K is a K(G, 1). This is a theorem in 3 manifold theory, but in the special case that K is a torus knot the result follows from our study of torus knot comple-ments in Examples 1.24 and 1.35. Namely, we showed that forK the torus knot Km,n there is a deformation retraction of S3− K onto a certain 2 dimensional complex Xm,n having contractible universal cover. The homotopy lifting property then implies that the universal cover of S3− K is homotopy equivalent to the universal cover of Xm,n, hence is also contractible.

E

xample 1B.7. It is not hard to construct a K(G, 1) for an arbitrary group G , us-ing the notion of a ∆ complex defined in §2.1. Let EG be the ∆ complex whose n simplices are the ordered (n+ 1) tuples [g0,··· , gn] of elements of G . Such an n simplex attaches to the (n− 1) simplices [g0,··· , bgi,··· , gn] in the obvious way, just as a standard simplex attaches to its faces. (The notation gbi means that this vertex is deleted.) The complex EG is contractible by the homotopy ht that slides each point x∈ [g0,··· , gn] along the line segment in [e, g0,··· , gn] from x to the vertex [e] , where e is the identity element of G . This is well-defined in EG since when we restrict to a face[g0,··· , bgi,··· , gn] we have the linear deformation to [e]

in[e, g0,··· , bgi,··· , gn] . Note that ht carries[e] around the loop [e, e] , so ht is not actually a deformation retraction of EG onto [e] .

The group G acts on EG by left multiplication, an element g ∈ G taking the simplex [g0,··· , gn] linearly onto the simplex [gg0,··· , ggn] . Only the identity e takes any simplex to itself, so by an exercise at the end of this section, the action of G on EG is a covering space action. Hence the quotient map EG

EG/G is the universal cover of the orbit space BG= EG/G, and BG is a K(G, 1).

Since G acts on EG by freely permuting simplices, BG inherits a ∆ complex structure from EG . The action of G on EG identifies all the vertices of EG , so BG has just one vertex. To describe the ∆ complex structure on BG explicitly, note first that every n simplex of EG can be written uniquely in the form

[g0, g0g1, g0g1g2,··· , g0g1··· gn]= g0[e, g1, g1g2,··· , g1··· gn]

The image of this simplex in BG may be denoted unambiguously by the symbol [g1|g2| ··· |gn] . In this ‘bar’ notation the gi’s and their ordered products can be used to label edges, viewing an

edge label as the ratio between the two labels on the vertices at the endpoints of the edge, as indicated in the figure. With

g g

g

0 g0

g3 g3

g0

1 g1 g1 g0g1

g0g1 2 g0g1 2g g0g1 2g g3

g2 g3

g g1 2

g g1 2

g2 g2

g g1 2

this notation, the boundary of a simplex [g1| ··· |gn] of BG

consists of the simplices [g2| ··· |gn] , [g1| ··· |gn−1] , and [g1| ··· |gigi+1| ··· |gn] for i= 1, ··· , n − 1.

This construction of a K(G, 1) produces a rather large space, since BG is al-ways infinite-dimensional, and if G is infinite, BG has an infinite number of cells in each positive dimension. For example, BZ is much bigger than S1, the most efficient K(Z, 1). On the other hand, BG has the virtue of being functorial: A homomorphism f : G

H induces a map Bf : BG

BH sending a simplex [g1| ··· |gn] to the simplex [f (g1)| ··· |f (gn)] . A different construction of a K(G, 1) is given in§4.2. Here one starts with any 2 dimensional complex having fundamental group G , for example

the complex XG associated to a presentation of G , and then one attaches cells of di-mension 3 and higher to make the universal cover contractible without affecting π1. In general, it is hard to get any control on the number of higher-dimensional cells needed in this construction, so it too can be rather inefficient. Indeed, finding an efficientK(G, 1) for a given group G is often a difficult problem.

It is a curious and almost paradoxical fact that ifG contains any elements of finite order, then every K(G, 1) CW complex must be infinite-dimensional. This is shown in Proposition 2.45. In particular the infinite-dimensional lens space K(Zm, 1) ’s in Example 1B.4 cannot be replaced by any finite-dimensional complex.

In spite of the great latitude possible in the construction of K(G, 1) ’s, there is a very nice homotopical uniqueness property that accounts for much of the interest in K(G, 1) ’s:

T

heorem 1B.8. The homotopy type of a CW complex K(G, 1) is uniquely determined by G .

Having a unique homotopy type of K(G, 1) ’s associated to each group G means that algebraic invariants of spaces that depend only on homotopy type, such as ho-mology and cohoho-mology groups, become invariants of groups. This has proved to be a quite fruitful idea, and has been much studied both from the algebraic and topological viewpoints. The discussion following Proposition 2.45 gives a few references.

The preceding theorem will follow easily from:

P

roposition 1B.9. Let X be a connected CW complex and let Y be a K(G, 1) . Then every homomorphism π1(X, x0)

π1(Y , y0) is induced by a map (X, x0)

(Y , y0)

that is unique up to homotopy fixing x0.

To deduce the theorem from this, letX and Y be CW complex K(G, 1) ’s with iso-morphic fundamental groups. The proposition gives maps f : (X, x0)

(Y , y0) and

g : (Y , y0)

(X, x0) inducing inverse isomorphisms π1(X, x0)≈ π1(Y , y0) . Then f g and gf induce the identity on π1 and hence are homotopic to the identity maps.

P

roof of 1B.9: Let us first consider the case that X has a single 0 cell, the base-pointx0. Given a homomorphism ϕ : π1(X, x0)

π1(Y , y0) , we begin the construc-tion of a map f : (X, x0)

(Y , y0) with f = ϕ by setting f (x0)= y0. Each 1 cell eα1 of X has closure a circle determining an element

[e1α] ∈ π1(X, x0) , and we let f on the closure of e1α π1 X

1

( x0)

ϕ

,

π1(X,x0)

π1(Y,y0)

f

i

−−−−→

−−−−−−−−−→ −−−−−−−−−→

be a map representingϕ([e1α]) . If i : X1

>

X denotes

the inclusion, then ϕi= f since π1(X1, x0) is gen-erated by the elements [e1α] .

To extendf over a cell e2β with attaching mapψβ:S1

X1, all we need is for the compositionf ψβto be nullhomotopic. Choosing a basepoints0∈ S1and a path inX1 fromψβ(s0) to x0,ψβ determines an elementβ]∈ π1(X1, x0) , and the existence

of a nullhomotopy of f ψβ is equivalent to f([ψβ]) being zero in π1(Y , y0) . We have i([ψβ]) = 0 since the cell e2β provides a nullhomotopy of ψβ in X . Hence f([ψβ])= ϕi([ψβ])= 0, and so f can be extended over e2β.

Extendingf inductively over cells enγ with n > 2 is possible since the attaching maps ψγ:Sn−1

Xn−1 have nullhomotopic compositions f ψγ:Sn−1

Y . This is because f ψγ lifts to the universal cover of Y if n > 2 , and this cover is contractible by hypothesis, so the lift of f ψγ is nullhomotopic, hence alsof ψγ itself.

Turning to the uniqueness statement, if two maps f0, f1:(X, x0)

(Y , y0)

in-duce the same homomorphism onπ1, then we see immediately that their restrictions to X1 are homotopic, fixing x0. To extend the resulting map X1×I ∪ X×∂I

Y

over the remaining cells en×(0, 1) of X×I we can proceed just as in the preceding paragraph since these cells have dimension n+ 1 > 2. Thus we obtain a homotopy ft:(X, x0)

(Y , y0) , finishing the proof in the case that X has a single 0 cell.

The case that X has more than one 0 cell can be treated by a small elaboration on this argument. Choose a maximal tree T ⊂ X . To construct a map f realizing a given ϕ , begin by setting f (T ) = y0. Then each edge e1α in X − T determines an element [e1α]∈ π1(X, x0) , and we let f on the closure of e1α be a map representing ϕ([e1α]) . Extending f over higher-dimensional cells then proceeds just as before.

Constructing a homotopy ft joining two given maps f0 and f1 with f0∗= f1∗ also has an extra step. Letht:X1

X1 be a homotopy starting withh0=11 and restricting to a deformation retraction of T onto x0. (It is easy to extend such a deformation retraction to a homotopy defined on all of X1.) We can construct a homotopy from f0|X1tof1|X1by first deformingf0|X1andf1|X1to takeT to y0by composing with ht, then applying the earlier argument to obtain a homotopy between the modified f0|X1 and f1|X1. Having a homotopyf0|X1' f1|X1 we extend this over all ofX in

the same way as before. tu

The first part of the preceding proof also works for the 2 dimensional complexes XG associated to presentations of groups. Thus every homomorphism G

H is

re-alized as the induced homomorphism of some map XG

XH. However, there is no uniqueness statement for this map, and it can easily happen that different presenta-tions of a group G give XG’s that are not homotopy equivalent.

Graphs of Groups

As an illustration of how K(G, 1) spaces can be useful in group theory, we shall describe a procedure for assembling a collection of K(G, 1) ’s together into a K(G, 1) for a larger groupG . Group-theoretically, this gives a method for assembling smaller groups together to form a larger group, generalizing the notion of free products.

Let Γ be a graph that is connected and oriented, that is, its edges are viewed as arrows, each edge having a specified direction. Suppose that at each vertexv ofΓ we

place a group Gv and along each edgee of Γ we put a homomorphism ϕe from the group at the tail of the edge to the group at the head of the edge. We call this data a graph of groups. Now build a space BΓ by putting the space BGv from Example 1B.7 at each vertex v of Γ and then filling in a mapping cylinder of the map Bϕe along each edgee ofΓ , identifying the two ends of the mapping cylinder with the two BGv’s at the ends of e . The resulting space BΓ is then a CW complex since the maps Bϕe

take n cells homeomorphically onto n cells. In fact, the cell structure on BΓ can be canonically subdivided into a∆ complex structure using the prism construction from the proof of Theorem 2.10, but we will not need to do this here.

More generally, instead of BGv one could take any CW complex K(Gv, 1) at the vertex v , and then along edges put mapping cylinders of maps realizing the homo-morphisms ϕe. We leave it for the reader to check that the resulting space KΓ is homotopy equivalent to the BΓ constructed above.

E

xample 1B.10. Suppose Γ consists of one central vertex with a number of edges radiating out from it, and the group Gv at this central vertex is trivial, hence also all the edge homomorphisms. Then van Kampen’s theorem implies that π1(KΓ ) is the free product of the groups at all the outer vertices.

In view of this example, we shall callπ1(KΓ ) for a general graph of groups Γ the graph product of the vertex groups Gv with respect to the edge homomorphismsϕe. The name for π1(KΓ ) that is generally used in the literature is the rather awkward phrase, ‘the fundamental group of the graph of groups.’

Here is the main result we shall prove about graphs of groups:

T

heorem 1B.11. If all the edge homomorphisms ϕe are injective, then KΓ is a K(G, 1) and the inclusions K(Gv, 1)

>

KΓ induce injective maps on π1.

Before giving the proof, let us look at some interesting special cases:

E

xample 1B.12: Free Products with Amalgamation. Suppose the graph of groups is A

C

B , with the two maps monomorphisms. One can regard this data as speci-fying embeddings of C as subgroups of A and B . Applying van Kampen’s theorem to the decomposition of KΓ into its two mapping cylinders, we see that π1(KΓ ) is the quotient ofA∗ B obtained by identifying the subgroup C ⊂ A with the subgroup C ⊂ B . The standard notation for this group is A ∗CB , the free product of A and B amalgamated along the subgroup C . According to the theorem, ACB contains bothA and B as subgroups.

For example, a free product with amalgamation Z ∗ZZ can be realized by map-ping cylinders of the mapsS1

S1

S1 that arem sheeted and n sheeted covering spaces, respectively. We studied this case in Examples 1.24 and 1.35 where we showed that the complex KΓ is a deformation retract of the complement of a torus knot in S3 ifm and n are relatively prime. It is a basic result in 3 manifold theory that the

complement of every smooth knot in S3 can be built up by iterated graph of groups constructions with injective edge homomorphisms, starting with free groups, so the theorem implies that these knot complements are K(G, 1) ’s. Their universal covers are all R3, in fact.

E

xample 1B.13: HNN Extensions. Consider a graph of groups C A

ψ

ϕ with ϕ and ψ both monomorphisms. This is analogous to the previous case A

C

B ,

but with the two groups A and B coalesced to a single group. The group π1(KΓ ), which was denoted A∗CB in the previous case, is now denoted A∗C. To see what this group looks like, let us regard KΓ as being obtained from K(A, 1) by attaching K(C, 1)×I along the two ends K(C, 1)×∂I via maps realizing the monomorphisms ϕ and ψ . Using a K(C, 1) with a single 0 cell, we see that KΓ can be obtained from K(A, 1)∨ S1 by attaching cells of dimension two and greater, soπ1(KΓ ) is a quotient ofA∗Z, and it is not hard to figure out that the relations defining this quotient are of the form tϕ(c)t−1= ψ(c) where t is a generator of the Z factor and c ranges over C , or a set of generators for C . We leave the verification of this for the Exercises.

As a very special case, taking ϕ= ψ =11 gives A∗A= A×Z since we can take KΓ = K(A, 1)×S1 in this case. More generally, taking ϕ = 11 with ψ an arbitrary automorphism of A , we realize any semidirect product of A and Z as A∗A. For example, the Klein bottle occurs this way, with ϕ realized by the identity map of S1 and ψ by a reflection. In these cases when ϕ=11 we could realize the same group π1(KΓ ) using a slightly simpler graph of groups, with a single vertex, labeled A, and a single edge, labeledψ .

Here is another special case. Suppose we take a torus, delete a small open disk, then identify the resulting boundary circle with a longitudinal circle of the torus. This produces a space X that happens to be homeomorphic to a subspace of the stan-dard picture of a Klein bottle in R3; see Exercise 12 of§1.2. The fundamental group π1(X) has the form (Z ∗ Z) ∗ZZ with the defining relation tb±1t−1 = aba−1b−1 where a is a meridional loop and b is a longitudinal loop on the torus. The sign of the exponent in the term b±1 is immaterial since the two ways of glueing the boundary circle to the longitude produce homeomorphic spaces. The groupπ1(X)= a, b, t |||| tbt−1aba−1b−1

abelianizes to Z×Z, but to show that π1(X) is not iso-morphic to Z ∗ Z takes some work. There is a surjection π1(X)

Z ∗ Z obtained by setting b = 1. This has nontrivial kernel since b is nontrivial in π1(X) by the pre-ceding theorem. If π1(X) were isomorphic toZ ∗ Z we would then have a surjective homomorphismZ ∗ Z

Z ∗ Z that was not an isomorphism. However, it is a theorem in group theory that a free group F is hopfian — every surjective homomorphism F

F must be injective. Hence π1(X) is not free.

E

xample 1B.14: Closed Surfaces. A closed orientable surface M of genus two or greater can be cut along a circle into two compact surfacesM1 andM2 such that the

closed surfaces obtained from M1 and M2 by filling in their boundary circle with a disk have smaller genus than M . Each of M1 and M2 is the mapping cylinder of a map from S1 to a finite graph. Namely, view Mi as obtained from a closed surface by deleting an open disk in the interior of the 2 cell in the standard CW structure described in Chapter 0, so that Mi becomes the mapping cylinder of the attaching map of the 2 cell. This attaching map is not nullhomotopic, so it induces an injection on π1 since free groups are torsionfree. Thus we have realized the original surface M as KΓ for Γ a graph of groups of the form F1

←---

Z

-→

F2 with F1 and F2 free and

the two maps injective. The theorem then says that M is a K(G, 1) .

A similar argument works for closed nonorientable surfaces other thanRP2. For example, the Klein bottle is obtained from two M¨obius bands by identifying their boundary circles, and a M¨obius band is the mapping cylinder of the 2 sheeted covering spaceS1

S1.

P

roof of 1B.11: We shall construct a covering space eK

KΓ by gluing together copies of the universal covering spaces of the various mapping cylinders inKΓ in such a way that eK will be contractible. Hence eK will be the universal cover of KΓ , which will therefore be aK(G, 1) .

First a preliminary observation: Given a universal covering spacep : eX

X and a

connected, locally path-connected subspaceA⊂ X such that the inclusion A

>

X

in-duces an injection onπ1, then each component eA of p−1(A) is a universal cover of A . To see this, note thatp : eA

A is a covering space, so the induced map π1( eA)

π1(A)

is injective, and this map factors through π1( eX)= 0, hence π1( eA)= 0. For exam-ple, if X is the torus S1×S1 and A is the circle S1×{x0}, then p−1(A) consists of infinitely many parallel lines in R2, each of which is a universal cover ofA .

For a map f : A

B between connected CW complexes, let p : fMf

Mf be the

universal cover of the mapping cylinderMf. Then fMf is itself the mapping cylinder of a map ef : p−1(A)

p−1(B) since the line segments in the mapping cylinder struc-ture on Mf lift to line segments in fMf defining a mapping cylinder structure. Since Mff is a mapping cylinder, it deformation retracts onto p−1(B) , so p−1(B) is also simply-connected, hence is the universal cover ofB . If f induces an injection on π1, then the remarks in the preceding paragraph apply, and the components of p−1(A) are universal covers ofA . If we assume further that A and B are K(G, 1) ’s, then fMf and the components of p−1(A) are contractible, and we claim that fMf deformation retracts onto each component eA of A . Namely, the inclusion eA

>

Mff is a

homo-topy equivalence since both spaces are contractible, and then Corollary 0.20 implies that fMf deformation retracts onto eA since the pair (fMf, eA) satisfies the homotopy extension property, as shown in Example 0.15.

Now we can describe the construction of the covering space eK of KΓ . It will be the union of an increasing sequence of spaces eK1 ⊂ eK2 ⊂ ···. For the first stage, let eK1 be the universal cover of one of the mapping cylinders Mf of KΓ . By the

preceding remarks, this contains various disjoint copies of universal covers of the twoK(Gv, 1) ’s at the ends of Mf. We build eK2 from eK1 by attaching to each of these universal covers ofK(Gv, 1) ’s a copy of the universal cover of each mapping cylinder Mg of KΓ meeting Mf at the end of Mf in question. Now repeat the process to construct eK3 by attaching universal covers of mapping cylinders at all the universal covers ofK(Gv, 1) ’s created in the previous step. In the same way, we construct eKn+1 from eKn for all n , and then we set eK=S

nKen.

Note that eKn+1 deformation retracts onto eKn since it is formed by attaching pieces to eKn that deformation retract onto the subspaces along which they attach, by our earlier remarks. It follows that eK is contractible since we can deformation retract eKn+1 onto eKn during the time interval[1/2n+1, 1/2n] , and then finish with a contraction of eK1 to a point during the time interval[1/2, 1].

The natural projection eK

KΓ is clearly a covering space, so this finishes the proof that KΓ is a K(G, 1).

The remaining statement that each inclusionK(Gv, 1)

>

KΓ induces an injection on π1 can easily be deduced from the preceding constructions. For suppose a loop γ : S1

K(Gv, 1) is nullhomotopic in KΓ . By the lifting criterion for covering spaces, there is a lift eγ : S1

K . This has image contained in one of the copies of the universale cover of K(Gv, 1) , so eγ is nullhomotopic in this universal cover, and hence γ is

nullhomotopic inK(Gv, 1) . tu

The various mapping cylinders that make up the universal cover of KΓ are ar-ranged in a treelike pattern. The tree in question, call it TΓ , has one vertex for each copy of a universal cover of a K(Gv, 1) in eK , and two vertices are joined by an edge whenever the two universal covers of K(Gv, 1) ’s corresponding to these vertices are connected by a line segment lifting a line segment in the mapping cylinder structure of a mapping cylinder ofKΓ . The inductive construction of eK is reflected in an inductive construction of TΓ as a union of an increasing sequence of subtrees T1 ⊂ T2 ⊂ ···.

Corresponding to eK1 is a subtreeT1⊂ T Γ consisting of a central vertex with a number of edges radiating out from it, an ‘asterisk’ with possibly an infinite number of edges.

When we enlarge eK1 to eK2, T1 is correspondingly enlarged to a tree T2 by attaching a similar asterisk at the end of each outer vertex ofT1, and each subsequent enlarge-ment is handled in the same way. The action ofπ1(KΓ ) on eK as deck transformations induces an action onTΓ , permuting its vertices and edges, and the orbit space of T Γ under this action is just the original graph Γ . The action on T Γ will not generally be a free action since the elements of a subgroup Gv ⊂ π1(KΓ ) fix the vertex of T Γ corresponding to one of the universal covers ofK(Gv, 1) .

There is in fact an exact correspondence between graphs of groups and groups acting on trees. See [Scott & Wall 1979] for an exposition of this rather nice theory.

From the viewpoint of groups acting on trees, the definition of a graph of groups is

usually taken to be slightly more restrictive than the one we have given here, namely, one considers only oriented graphs obtained from an unoriented graph by subdividing each edge by adding a vertex at its midpoint, then orienting the two resulting edges outward, away from the new vertex.

Exercises

1. Suppose a group G acts simplicially on a ∆ complex X , where ‘simplicially’ means that each element of G takes each simplex of X onto another simplex by a linear homeomorphism. If the action is free, show it is a covering space action.

2. Let X be a connected CW complex and G a group such that every homomorphism π1(X)

G is trivial. Show that every map X

K(G, 1) is nullhomotopic.

在文檔中 Allen Hatcher (頁 96-105)