• 沒有找到結果。

The Splitting Principle

在文檔中 1.2. Classifying Vector Bundles (頁 69-78)

Now we can use the operations ψ2 and ψ3 and the relation ψ2ψ3= ψ6= ψ3ψ2 to prove Adams’ theorem.

P

roof of Theorem 2.19: The definition of the Hopf invariant of a map f : S4n−1

S2n

involved elements α, β∈ eK(Cf) . By Proposition 2.21, ψk(α)= k2nα since α is the image of a generator of eK(S4n) . Similarly, ψk(β) = knβ+ µkα for some µk ∈ Z.

Therefore

ψkψ`(β)= ψk(`nβ+ µ`α)= kn`nβ+ (k2nµ`+ `nµk

Since ψkψ`= ψk`= ψ`ψk, the coefficient k2nµ`+ `nµk of α is unchanged when k and ` are switched. This gives the relation

k2nµ`+ `nµk= `2nµk+ knµ`, or (k2n− kn`= (`2n− `nk

By property (6) of ψ2, we have ψ2(β)≡ β2mod 2 . Since β2 = hα with h the Hopf invariant of f , the formula ψ2(β)= 2nβ+ µ2α implies that µ2≡ h mod 2, so µ2 is odd if we assume h= ±1. By the preceding displayed formula we have (22n−2n3= (32n− 3n2, or 2n(2n− 1)µ3= 3n(3n− 1)µ2, so 2n divides 3n(3n− 1)µ2. Since 3n and µ2 are odd, 2n must then divide 3n− 1. The proof is completed by the following

elementary number theory fact. tu

L

emma 2.22. If 2n divides 3n− 1 then n = 1, 2, or 4.

P

roof: Write n= 2`m with m odd. We will show that the highest power of 2 dividing 3n− 1 is 2 for ` = 0 and 2`+2 for ` > 0 . This implies the lemma since if 2n divides 3n− 1, then by this fact, n ≤ ` + 2, hence 2`≤ 2`m= n ≤ ` + 2, which implies ` ≤ 2 and n≤ 4. The cases n = 1, 2, 3, 4 can then be checked individually.

We find the highest power of 2 dividing 3n− 1 by induction on `. For ` = 0 we have 3n− 1 = 3m− 1 ≡ 2 mod 4 since 3 ≡ −1 mod 4 and m is odd. In the next case ` = 1 we have 3n− 1 = 32m− 1 = (3m− 1)(3m+ 1). The highest power of 2 dividing the first factor is 2 as we just showed, and the highest power of 2 dividing the second factor is 4 since 3m+ 1 ≡ 4 mod 8 because 32 ≡ 1 mod 8 and m is odd. So the highest power of 2 dividing the product (3m− 1)(3m+ 1) is 8. For the inductive step of passing from ` to `+ 1 with ` ≥ 1, or in other words from n to 2n with n even, write 32n− 1 = (3n− 1)(3n+ 1). Then 3n+ 1 ≡ 2 mod 4 since n is even, so the highest power of 2 dividing 3n+ 1 is 2. Thus the highest power of 2 dividing 32n− 1 is twice the highest power of 2 dividing 3n− 1. tu

P

roposition 2.23. If X is a finite cell complex with n cells, then K(X) is a finitely generated group with at most n generators. If all the cells of X have even dimension then K1(X)= 0 and K0(X) is free abelian with one basis element for each cell.

The phrase ‘finite cell complex’ would normally mean ‘finite CW complex’ but we can take it to be something slightly more general: a space built from a finite discrete set by attaching a finite number of cells in succession, with no conditions on the dimen-sions of these cells, so cells are not required to attach only to cells of lower dimension.

Finite cell complexes are always homotopy equivalent to finite CW complexes (by de-forming each successive attaching map to be cellular) so the only advantages of finite cell complexes are technical. In particular, it is easy to see that a space is a finite cell complex if it is a fiber bundle over a finite cell complex with fibers that are also finite cell complexes. This is shown in Proposition 2.28 in a brief appendix to this section.

It implies that the splitting principle can be applied staying within the realm of finite cell complexes.

P

roof: We show this by induction on the number of cells. The complex X is obtained from a subcomplex A by attaching a k cell, for some k . For the pair (X, A) we have an exact sequence eK(X/A)

-→

Ke(X)

-→

Ke(A) . Since X/A = Sk, we have Ke(X/A)≈ Z, and exactness implies that eK(X) requires at most one more generator than eK(A) .

The first term of the exact sequence K1(X/A)

K1(X)

K1(A) is zero if all cells of X are of even dimension, so induction on the number of cells implies that K1(X)= 0. Then there is a short exact sequence 0

Ke0(X/A)

Ke0(X)

Ke0(A)

0

with eK0(X/A)≈ Z. By induction eK(A) is free, so this sequence splits, hence K0(X)≈ Z⊕K0(A) and the final statement of the proposition follows. tu This proposition applies in particular toCPn, which has a cell structure with one cell in each dimension 0, 2, 4,··· , 2n, so K1(CPn)= 0 and K0(CPn)≈ Zn+1. The ring structure is as simple as one could hope for:

P

roposition 2.24. K(CPn) is the quotient ringZ[L]/(L−1)n+1 where L is the canon-ical line bundle overCPn.

Thus by the change of variable x= L−1 we see that K(CPn) is the truncated poly-nomial ringZ[x]/(xn+1) , with additive basis 1, x,··· , xn. It follows that 1, L,··· , Ln is also an additive basis.

P

roof: The exact sequence for the pair (CPn,CPn−1) gives a short exact sequence 0

-→

K(CPn,CPn−1)

-→

K(CPn)

---→

ρ K(CPn−1)

-→

0

We shall prove:

(an) (L− 1)n generates the kernel of the restriction map ρ .

Hence if we assume inductively that K(CPn−1)= Z[L]/(L − 1)n, then (an) and the preceding exact sequence imply that {1, L − 1, ··· , (L − 1)n} is an additive basis for K(CPn) . Since (L− 1)n+1 = 0 in K(CPn) by (an+1) , it follows that K(CPn) is the quotient ringZ[L]/(L − 1)n+1, completing the induction.

A reason one might expect (an) to be true is that the kernel of ρ can be identi-fied with K(CPn,CPn−1)= eK(S2n) by the short exact sequence, and Bott periodicity implies that the n fold reduced external product of the generator L− 1 of eK(S2) with itself generates eK(S2n) . To make this rough argument into a proof we will have to relate the external product eK(S2)⊗ ··· ⊗ eK(S2)

K(Se 2n) to the ‘internal’ product K(CPn)⊗ ··· ⊗K(CPn)

K(CPn) .

The space CPn is the quotient of the unit sphere S2n+1 in Cn+1 under multipli-cation by scalars in S1⊂ C. Instead of S2n+1 we could equally well take the boundary of the product D20× ··· ×D2nwhere D2i is the unit disk in the ithcoordinate of Cn+1, and we start the count with i= 0 for convenience. Then we have

∂(D20× ··· ×D2n)=S

i(D20× ··· ×∂Di2× ··· ×D2n)

The action of S1 by scalar multiplication respects this decomposition. The orbit space of D20× ··· ×∂D2i× ··· ×D2n under the action is a subspace Ci⊂ CPn homeomorphic to the product D02× ··· ×Dn2 with the factor Di2 deleted. Thus we have a decomposi-tion CPn=S

iCi with each Ci homeomorphic to D2n and with Ci∩ Cj = ∂Ci∩ ∂Cj for i≠ j .

Consider now C0 = D21× ··· ×D2n. Its boundary is decomposed into the pieces

iC0 = D21× ··· ×∂D2i× ··· ×D2n. The inclusions (Di2, ∂D2i)⊂ (C0, ∂iC0)⊂ (CPn, Ci) give rise to a commutative diagram

D

K( 12, ∂D12) K(Dn2, ∂Dn2)

−−−−−−−−−−→

−−−−−→ −−−−−→

−−−−−−−−−−−−−−−−−−−−−−−−−−−−−−→

−−−−−−−−−−−−−−−−−→

− − − −−−−−

− − − −−−−−

−−→ −−→− −→ −−→

−−→ − →

...

C

K( 0, ∂1C0)

...

K(C0,∂nC0) K(C0,∂C0)

K( P ,C1) n

n

... ...

C K( P C ) K( P ,C1 n

C n C ) )

n, C

K(CPn)

...

K(CPn) K(CPn)

K(CPn,CPn 1

-⊗

where the maps from the first column to the second are the n fold products. The upper map in the middle column is an isomorphism because the inclusion C0

>

CPn

induces a homeomorphism C0/∂C0≈ CPn/(C1∪···∪Cn) . TheCPn−1at the right side of the diagram sits in CPn in the last n coordinates ofCn+1, so is disjoint from C0, hence the quotient mapCPn/CPn−1

CPn/(C1∪···∪Cn) is a homotopy equivalence.

The element xi∈ K(CPn, Ci) mapping downward to L−1 ∈ K(CPn) maps upward to a generator of K(C0, ∂iC0) ≈ K(Di2, ∂D2i) . By commutativity of the diagram, the product x1··· xn then generates K(CPn, C1∪ ··· ∪ Cn) . This means that (L− 1)n

generates the image of the map K(CPn,CPn−1)

K(CPn) , which equals the kernel of

ρ , proving (an) . tu

SinceCPnis the union of the n+1 balls Ci, Example 2.13 shows that all products of n+ 1 elements of eK(CPn) must be zero, in particular (L− 1)n+1= 0. But as we have just seen, (L− 1)n is nonzero, so the result in Example 2.13 is best possible in terms of the degree of nilpotency.

Now we come to the Leray-Hirsch theorem for K–theory, which will be the major theoretical ingredient in the proof of the splitting principle:

T

heorem 2.25. Let p : E

B be a fiber bundle with E and B compact Hausdorff and with fiber F such that K(F ) is free. Suppose that there exist classes c1,··· , ck K(E) that restrict to a basis for K(F ) in each fiber F . If either

(a) B is a finite cell complex, or

(b) F is a finite cell complex having all cells of even dimension, then K(E) , as a module over K(B) , is free with basis{c1,··· , ck}.

Here the K(B) module structure on K(E) is defined by β· γ = p(β)γ for β ∈ K(B) and γ ∈ K(E) . Another way to state the conclusion of the theorem is to say that the mapΦ : K(B)⊗K(F )

K(E) , Φ(Pibii(ci))=Pip(bi)ci for i

the inclusion F

>

E , is an isomorphism.

In the case of the product bundle E= F ×B the classes ci can be chosen to be the pullbacks under the projection E

F of a basis for K(F ) . The theorem then asserts that the external product K(F )⊗K(B)

K(F×B) is an isomorphism.

For most of our applications of the theorem either case (a) or case (b) will suffice.

The proof of (a) is somewhat simpler than (b), and we include (b) mainly to obtain the splitting principle for vector bundles over arbitrary compact Hausdorff base spaces.

P

roof: For a subspace B0⊂ B let E0= p−1(B0) . Then we have a diagram

−−→ −−→

−−→ −−−−−→

−−−−−→

0

Φ Φ Φ

B B K ( ) ( )

, K ( ) F K ( ) B K ( ) F

−−−−−→

K ( B0)K ( ) F

−−−−−−−−−−−−−−−→

−−−−−−−−−−−−→

K ( E,E0) K ( ) E

−−−−−−−−−−−−−−−−−→

K ( E0)

−−−−−−−

where the left-handΦ is defined by the same formula Φ(P

ibii(ci))=P

ip(bi)ci, but with p(bi)cireferring now to the relative product K(E, E0)×K(E)

K(E, E0) .

The right-hand Φ is defined using the restrictions of the ci’s to the subspace E0. To see that the diagram (∗) commutes, we can interpolate between its two rows the row

-→

K(E, E0)⊗K(F )

-→

K(E)⊗K(F )

-→

K(E0)⊗K(F )

-→

by factoring Φ as the compositionP

ibii(ci)

,

Pip(bi)i(ci)

,

Pip(bi)ci.

The upper squares of the enlarged diagram then commute trivially, and the lower squares commute by Proposition 2.15. The lower row of the diagram is of course

exact. The upper row is also exact since we assume K(F ) is free, and tensoring an exact sequence with a free abelian group preserves exactness, the result of the tensoring operation being simply to replace the given exact sequence by the direct sum of a number of copies of itself.

The proof in case (a) will be by a double induction, first on the dimension of B , then within a given dimension, on the number of cells. The induction starts with the trivial case that B is zero-dimensional, hence a finite discrete set. For the induction step, suppose B is obtained from a subcomplex B0 by attaching a cell en, and let E0= p−1(B0) as above. By induction on the number of cells of B we may assume the right-hand Φ in (∗) is an isomorphism. If the left-hand Φ is also an isomorphism, then the five-lemma will imply that the middle Φ is an isomorphism, finishing the induction step.

Let ϕ : (Dn, Sn−1)

(B, B0) be a characteristic map for the attached n cell. Since Dn is contractible, the pullback bundle ϕ(E) is a product, and so we have a com-mutative diagram

−−→

−−→ −−−−−→

0

Φ Φ Φ

B B

K ( , ) K ( ) F K ( D S, )⊗K ( ) F

−−−−−−−−−−−−→

0 0

E E

K ( , ) K

(

(E),

−−−−−−−−−−−−−→

K (

ϕ ϕ(E )

)

× ×

n n 1

-, F

F

Dn Sn 1- )

The two horizontal maps are isomorphisms since ϕ restricts to a homeomorphism on the interior of Dn, hence induces homeomorphisms B/B0≈ Dn/Sn−1 and E/E0 ϕ(E)/ϕ(E0) . Thus the diagram reduces the proof to showing that the right-hand Φ, involving the product bundle Dn×F

Dn, is an isomorphism.

Consider the diagram (∗) with (B, B0) replaced by (Dn, Sn−1) . We may assume the right-hand Φ in (∗) is an isomorphism since Sn−1 has smaller dimension than the original cell complex B . The middle Φ is an isomorphism by the case of zero-dimensional B since Dndeformation retracts to a point. Therefore by the five-lemma the left-hand Φ in (∗) is an isomorphism for (B, B0)= (Dn, Sn−1) . This finishes the proof in case (a).

In case (b) let us first prove the result for a product bundle E= F ×B . In this case Ψ is just the external product, so we are free to interchange the roles of F and B . Thus we may use the diagram (∗) with F an arbitrary compact Hausdorff space and B a finite cell complex having all cells of even dimension, obtained by attaching a cell en to a subcomplex B0. The upper row of (∗) is then an exact sequence since it is obtained from the split short exact sequence 0

K(B, B0)

K(B)

K(B0)

0 by

tensoring with the fixed group K(F ) . If we can show that the left-hand Φ in (∗) is an isomorphism, then by induction on the number of cells of B we may assume the right-hand Φ is an isomorphism, so the five-lemma will imply that the middle Φ is also an isomorphism.

To show the left-handΦ is an isomorphism, note first that B/B0= Sn so we may

as well take the pair (B, B0) to be (Dn, Sn−1) . Then the middle Φ in (∗) is obviously an isomorphism, so the left-hand Φ will be an isomorphism iff the right-hand Φ is an isomorphism. When the sphere Sn−1 is even-dimensional we have already shown that Φ is an isomorphism in the remarks following the proof of Lemma 2.17, and the same argument applies also when the sphere is odd-dimensional, since K1 of an odd-dimensional sphere is K0 of an even-dimensional sphere.

Now we turn to case (b) for nonproducts. The proof will once again be inductive, but this time we need a more subtle inductive statement since B is just a compact Hausdorff space, not a cell complex. Consider the following condition on a compact subspace U⊂ B :

For all compact V ⊂ U the map Φ : K(V )⊗K(F )

K(p−1(V )) is an isomor-phism.

If this is satisfied, let us call U good. By the special case already proved, each point of B has a compact neighborhood U that is good. Since B is compact, a finite number of these neighborhoods cover B , so by induction it will be enough to show that if U1 and U2 are good, then so is U1∪ U2.

A compact V ⊂ U1∪ U2 is the union of V1= V ∩ U1 and V2= V ∩ U2. Consider the diagram like (∗) for the pair (V , V2) . Since K(F ) is free, the upper row of this diagram is exact. Assuming U2 is good, the map Φ is an isomorphism for V2, so Φ will be an isomorphism for V if it is an isomorphism for (V , V2) . The quotient V /V2 is homeomorphic to V1/(V1∩ V2) so Φ will be an isomorphism for (V, V2) if it is an isomorphism for (V1, V1∩ V2) . Now look at the diagram like (∗) for (V1, V1∩ V2) . Assuming U1 is good, the mapsΦ are isomorphisms for V1 and V1∩ V2. Hence Φ is an isomorphism for (V1, V1∩ V2) , and the induction step is finished. tu

E

xample 2.26. Let E

X be a vector bundle with fibers Cn and compact base X . Then we have an associated projective bundle p : P (E)

X with fibersCPn−1, where P (E) is the space of lines in E , that is, one-dimensional linear subspaces of fibers of E . Over P (E) there is the canonical line bundle L

P (E) consisting of the vectors in the lines of P (E) . In each fiber CPn−1 of P (E) the classes 1, L,··· , Ln−1 in K(P (E)) restrict to a basis for K(CPn−1) by Proposition 2.24. From the Leray-Hirsch theorem we deduce that K(P (E)) is a free K(X) module with basis 1, L,··· , Ln−1.

P

roof of the Splitting Principle: In the preceding example, the fact that 1 is among the basis elements implies that p: K(X)

K(P (E)) is injective. The pullback bundle p(E)

P (E) contains the line bundle L as a subbundle, hence splits as L⊕E0 for E0

P (E) the subbundle of p(E) orthogonal to L with respect to some choice of inner product. Now repeat the process by forming P (E0) , splitting off another line bundle from the pullback of E0 over P (E0) . Note that P (E0) is the space of pairs of orthogonal lines in fibers of E . After a finite number of repetitions we obtain the flag bundle F (E)

X described at the end of§1.1, whose points are n tuples of

orthogonal lines in fibers of E , where n is the dimension of E . (If the fibers of E have different dimensions over different components of X , we do the construction for each component separately.) The pullback of E over F (E) splits as a sum of line bundles, and the map F (E)

X induces an injection on K since it is a composition

of maps with this property. tu

In the preceding Example 2.26 we saw that K(P (E)) is free as a K(X) module, with basis 1, L,··· , Ln−1. In order to describe the multiplication in K(P (E)) one therefore needs only a relation expressing Ln in terms of lower powers of L . Such a relation can be found as follows. The pullback of E over P (E) splits as L⊕E0 for some bundle E0 of dimension n− 1, and the desired relation will be λn(E0)= 0. To compute λn(E0)= 0 we use the formula λt(E)= λt(L)λt(E0) in K(P (E))[t] , where to simplify notation we let ‘ E ’ also denote the pullback of E over P (E) . The equation λt(E) = λt(L)λt(E0) can be rewritten as λt(E0) = λt(E)λt(L)−1 where λt(L)−1 = P

i(−1)iLiti since λt(L) = 1 + Lt . Equating coefficients of tn in the two sides of λt(E0)= λt(E)λt(L)−1, we get λn(E0)=P

i(−1)n−iλi(E)Ln−i. The relation λn(E0)= 0 can be written as P

i(−1)iλi(E)Ln−i = 0, with the coefficient of Ln equal to 1 , as desired. The result can be stated in the following form:

P

roposition 2.27. For an n dimensional vector bundle E

X the ring K(P (E)) is isomorphic to the quotient ring K(X)[L]/ P

i(−1)iλi(E)Ln−i

. tu

For example when X is a point we have P (E)= CPn−1and λi(E)= Ck for k=

n i

 , so the polynomialP

i(−1)iλi(E)Ln−ibecomes (L−1)nand we see that the proposition generalizes the isomorphism K(CPn−1)≈ Z[L]/(L − 1)n) .

Appendix: Finite Cell Complexes

As we mentioned in the remarks following Proposition 2.23 it is convenient for purposes of the splitting principle to work with spaces slightly more general than finite CW complexes. By a finite cell complex we mean a space which has a finite filtration X0 ⊂ X1⊂ ··· ⊂ Xk = X where X0 is a finite discrete set and Xi+1 is obtained from Xi by attaching a cell eni via a map ϕi: Sni−1

Xi. Thus Xi+1 is the quotient space of the disjoint union of Xi and a disk Dni under the identifications x ∼ ϕi(x) for x∈ ∂Dni = Sni−1.

P

roposition 2.28. If p : E

B is a fiber bundle whose fiber F and base B are both finite cell complexes, then E is also a finite cell complex, whose cells are products of cells in B with cells in F .

P

roof: Suppose B is obtained from a subcomplex B0 by attaching a cell en. By induc-tion on the number of cells of B we may assume that p−1(B0) is a finite cell complex.

If Φ : Dn

B is a characteristic map for en then the pullback bundle Φ(E)

Dn is

a product since Dn is contractible. Since F is a finite cell complex, this means that

we may obtainΦ(E) from its restriction over Sn−1 by attaching cells. Hence we may

obtain E from p−1(B0) by attaching cells. tu

bundles E

B by some general rule which applies to all base spaces B . The four classical types of characteristic classes are:

1. Stiefel-Whitney classes wi(E)∈ Hi(B;Z2) for a real vector bundle E . 2. Chern classes ci(E)∈ H2i(B;Z) for a complex vector bundle E . 3. Pontryagin classes pi(E)∈ H4i(B;Z) for a real vector bundle E .

4. The Euler class e(E)∈ Hn(B;Z) when E is an oriented n dimensional real vector bundle.

The Stiefel-Whitney and Chern classes are formally quite similar. Pontryagin classes can be regarded as a refinement of Stiefel-Whitney classes when one takes Z rather thanZ2 coefficients, and the Euler class is a further refinement in the orientable case.

Stiefel-Whitney and Chern classes lend themselves well to axiomatization since in most applications it is the formal properties encoded in the axioms which one uses rather than any particular construction of these classes. The construction we give, using the Leray-Hirsch theorem (proved in§4.D of [AT]), has the virtues of simplicity and elegance, though perhaps at the expense of geometric intuition into what prop-erties of vector bundles these characteristic classes are measuring. There is another definition via obstruction theory which does provide some geometric insights, and this will be described in the Appendix to this chapter.

3.1. Stiefel-Whitney and Chern Classes

Stiefel-Whitney classes are defined for real vector bundles, Chern classes for com-plex vector bundles. The two cases are quite similar, but for concreteness we shall em-phasize the real case, with occasional comments on the minor modifications needed to treat the complex case.

A technical point before we begin: We shall assume without further mention that all base spaces of vector bundles are paracompact, so that the fundamental results of Chapter 1 apply. For the study of characteristic classes this is not an essential restriction since one can always pass to pullbacks over a CW approximation to a given base space, and CW complexes are paracompact.

在文檔中 1.2. Classifying Vector Bundles (頁 69-78)

相關文件