<<

3. Topological K-theory

Shortly after Grothendieck introduced K0(X) for an algebraic variety X, M. Atiyah and F. Hirzebrch introduced the analogous theory for topological spaces ∗ [2] based on topological vector bundles. We consider this theory Ktop(−) for two reasons: first, the theory developed by Atiyah and Hirzebruch has been a model for 40 years of effort in algebraic K-theory, effort that has recently produced significant advances; second, topological K-theory of the underlying analytic space of a com- plex variety X, Xan, provides a much more computable theory to which algebraic K-theory maps. A (complex) topological of rank r, p : E → T , on a space T is a continuous map with fibers each given the structure of a complex of dimension r such that there is an open covering {Ui ⊂ T } with the property that r ∼ there exist homeomorphisms φi : C ×Ui → E|Ui over Ui which are C-linear on each fiber. One readily verifies that such a topological vector bundle p : E → T , on T of rank r determines and is determined by patching data: a collection of continuous, fiber-wise linear homemorphisms for each i, j −1 r r φj ◦ φi = θi,j : C × (Ui ∩ Uj) ' (E|Ui )Uj = (E|Uj )Ui ' C × (Ui ∩ Uj). This in turn is equivalent to a 1-cocycle 1 {θi,j ∈ Mapscont(Ui,j, GL(r, C))} ∈ Z (T, GL(r, C)). Indeed, two such 1-cocycles determine isomorphic topological vector bundles if and only if they differ by a co-boundary. Thus, the topological vector bundles on T of rank r are “classified” by H1(T, GL(r, C)). It is instructive to recognize the distinction between an algebraic vector bun- dle on the complex algebraic variety X and a topological vector bundle on the associated analytic space Xan. If X is a projective smooth variety with the prop- erty that CH∗(X) ⊗ Q → H∗(Xan, Q) is not surjective, then we know that not every topological vector bundle on X admits an algebraic structure. We can view this in terms of patching functions as asserting that a given topological 1- 1 an cocycle {θi,j} ∈ Z (X , GL(r, C)) is not cohomologous to a cocycle for which each θi,j ∈ Mor(X, GLr,C). If C is a projective, smooth curve of genus g > 1, we have seen that K0(C) contains the uncountable abelian (complex torus) 0 g 0 an P ic (C) ' C /Λ every element of which maps to 0 ∈ Ktop(C ). This reflects the fact that algebraic vector bundles which are non-isomorphic as algebraic vec- tor bundles can become isomorphic when viewed as topological vector bundles. In terms of patching data, what occurs here is that some algebraic patching data are coboundaries of continuous but not algebraic 0 − cochains. Our examples of algebraic vector bundles immediately give us many examples of topological vector bundles as we observe in the following proposition. Proposition 3.1. Let X be a complex variety with underlying analytic space Xan. Then any algebraic vector bundle E on X of rank r determines naturally a rank r topological vector bundle on Xan. N Consider the Grassmannian Grassn(P ) of dimension n-planes in projective N- space, or equivalently of n dimensional linear quotients of CN+1. A map (i.e., N algebraic morphism) from an algebraic variety X to Grassn(P ) corresponds nat- urally to a rank n algebraic vector bundle E over X together a choice of N +1 global sections which determine a surjective bundle map OX → E. A map (i.e., continuous 1 2

N an function) from a topological space T to Grassn(P ) corresponds naturally to a rank n topological vector bundle on T together a choice of N + 1 global sections which determine a surjective bundle map (CN+1 × T ) → E. We shall use the notation Grass ( ∞)an ≡ lim Grass ( N )an. n P −→ n P N The following statements about topological vector bundles are not valid (in gen- eral) for algebraic vector bundles. These properties suggest that topological K- theory is better behaved than algebraic K-theory. Proposition 3.2. (cf. [1]) Let T be a compact Hausdorff space. If p : E → T is a topological vector bundle on T , then for some N > 0 there is a surjective map of bundles on T , (CN+1 × T ) → E. Any surjective map E → F of topological vector bundles on T admits a splitting over T . N an The set of homotopy classes of maps [T, Grass∞(P ) ] is in natural 1-1 corre- spondence with the set of isomorphism classes of rank n topological vector bundles on T . Proof. The first statement is proved using a partition of unity argument. The proof of the second statement is proved by establishing a Hermetian metric on E (so that E ' F ⊕ F ⊥), which is achieved by once again using a partition of unity argument. To prove the last statement, one verifies that if T ×I → G is a homotopy relating continuous maps f, g : T → G and if E is a topological vector bundle on G, then f ∗E ' g∗E as topological vector bundles on T . Once again, a partition of unity argument is the key ingredient in the proof.  Proposition 3.3. There is a natural homotopy equivalence ' ∞ an BU(n) → Grassn(P ) , where BUn is the classifying space of the Lie group Un of n × n unitary matrices. For any space T , the set of homotopy classes of maps [T,BU × ],BU = lim BU Z −→ n n admits a natural structure of an induced by block sum of matrices Un × Um → Un+m. We define 0 Ktop(T ) ≡ [T,BU × Z]. 0 For any compact, Hausdorff space T , Ktop(T ) is naturally isomorphic to the Grothendieck group of topological vector bundles on T :

0 Z[iso classes of top vector bundles on T ] Ktop(T ) ' . [E] = [E1] + [E2], whenever E ' E1 ⊕ E2 N Proof. To verify the first assertion, one observes that the Grassmannian Grass∞(P ) is a homogeneous space for GLN+1(C), with stabilizer the parabolic subgroup with N reductive factors GLN+1−n, GLn. Equivalently, Grass∞(P ) is the quotient by a free action of GLn on the quotient of the Stieffel variety

GLN+1(C)/GLN+1−n o MN+1−n,n. 3

The assertion now follows from the observation that the colimit with respect to N of the analytic spaces of these Stieffel varieties is contractible. (External) of matrices gives a structure on tnBUn which determines a (homotopy associative and commutative) H-space structure on BU×Z which we view as the mapping telescope of the self map

tnBUn → tnBUn,BUi × {? ∈ BU1} → BUi+1. The (abelian) group structure on [T,BU × Z] is then determined. To show that this mapping telescope is actually an H-space, one must verify that it has a 2-sided identity up to pointed homotopy: one must verify that product on the left with ? ∈ BU1 gives a self map of BU × Z which is related to the identity via a base-point preserving homotopy. (Such a verification is not difficult, but the analogous verification fails if we replace the topological groups Un by discrete groups GLn(A) for some unital A.) Finally, the previous proposition implies a natural isomorphism of ∼ iso classes top vector bundles → [T, tnBUn] whenever T is a compact Hausdorff space. One readily verifies that taking the colimit of the self-map given by translation [T, tnBUn] → [T, tnBU] has the effect of group completing this monoid. Thus, the isomorphism of monoids induces an isomorphism of group completions. 

Example 3.4. Since the Lie groups Un are connected, the spaces BUn are simply connected and thus 0 1 Ktop(S ) = π1(BU × Z) = 0. 0 It is useful to extend Ktop(−) to a relative theory which applies to pairs (T,A) of spaces (i.e., T is a topological space and A ⊂ T is a closed subset). In the special case that A = ∅, then T/∅ = T+/?, the pointed space obtained by taking the disjoint union of T with a point ? which we declare to be the basepoint.

Definition 3.5. If T is a pointed space with basepoint t0, we define the reduced K-theory of T by ˜ ∗ ∗ Ktop(T ) ≡ Ktop(T, t0). For any pair (T,A), we define 0 ˜ 0 Ktop(T,A) ≡ Ktop(T/A) 0 thereby extending our earlier definition of Ktop(T ). For any n > 0, we define n ˜ 0 n Ktop(T,A) ≡ Ktop(Σ (T/A)). In particular, for any n ≥ 0, we define −n −n ˜ 0 n Ktop (T ) ≡ Ktop (T, ∅) ≡ Ktop(Σ (T+)). Observe that ˜ 0 0 0 0 Ktop(S ∧ T ) = ker{Ktop(S × T ) → Ktop(S) ⊕ Ktop(T )}, so that (external) tensor product of bundles induces a natural pairing −i −j −i−j Ktop(S) ⊗ Ktop(T ) → Ktop (S × T ). 4

Just to get the notation somewhat straight, let us take T to be a single point T = {t}. Then T+ = {t, ?}, the 2-point space with new point ? as base-point. Then 2 2 Σ (T+) is the 2-sphere S , and thus −2 0 2 0 Ktop({t}) = ker{Ktop(S )) → Ktop(?)}. We single out a special element, the Bott element −2 1 1 β = [OP (1)] − [OP ] ∈ Ktop(pt)), where we have abused notation by identifying (P1)an with S2 and the images of 1 0 1 an 1 algebraic vector bundles on P in Ktop((P ) ) have the same names as in K0(P ). The most famous, and most important, theorem in topological K-theory is the following formulation by Atiyah of R. Bott’s famous theorem that there is a natural homotopy equivalence Ω(U) → BU ×Z. Since U is naturally equivalent to Ω(BU × Z), we conclude that 2 Ω (BU × Z) ' BU × Z. Atiyah interprets this 2-fold periodicity in terms of K-theory as follows. Theorem 3.6. (Bott Periodicity) For any space T and any i ≥ 0, mutliplication by the Bott element induces a natural isomorphism −i −i−2 β : Ktop(T ) → Ktop (T ). ˜ 0 2 In particular, taking T to be a point, we conclude that Ktop(S ) = Z, generated by the Bott element.

0 Example 3.7. Let S denote {∗,?} = ∗+. According to our definitions, the K- 0 theory Ktop(∗), of a point equals the reduced K-theory of S . In particular, for n > 0, −n ˜ −n 0 ˜ 0 n Ktop (∗) = Ktop (S ) = Ktop(S ) = πn(BU). Thus, we conclude  Z if n is even Kn (∗) = top 0 if n is odd We can reformulate this by writing  Z if i + n is even Ki (Sn) = top 0 if i + n is odd As observed by Atiyah and Hirzebruch, topological K-theory enjoys all of the properties of singular cohomology (other than the fact that its value on a point is somewhat more complicated), and thus can be computed for many familiar spaces. This is formalized in the following theorem.

n Theorem 3.8. The sequence of contravariant Ktop(−, −) on pairs satisfies the following properties: • (Homotopy invariance) If f, g :(S, B) → (T,A) are continuous maps of pairs related by a homotopy F : S × I → T restricting to B × I → A, then f ∗ = g∗. • (Exactness) For every pair (T,A), there is a long n n n n+1 · · · → Ktop(X,A) → Ktop(X, ∅) → Ktop(A, ∅) → Ktop (X,A) → · · · 5

• (Excision) For every U ⊂ A whose closure is contained in the interior of A, the inclusion (T − U, A − U) → (T,A) induces an isomorphism ∗ ∗ Ktop(T,A) ' Ktop(T − U, A − U). ∗ ˜ ∗ Thus, Ktop(−) (respectively, Ktop(−)) is an unreduced (resp., reduced) generalized cohomology theory. Example 3.9. Assume that T is a finite cell complex with only even dimensional ∗ cells. Then our previous computation of Ktop(pt) and the preceding theorem imply 0 1 that Ktop(T ) is a on the cells of T and that Ktop(T ) is 0. As for algebraic K-theory, we can define Chern classes 0 2i ci : Ktop(T ) → H (T, Z) where the right hand side is the singular cohomology of the topological space T . Perhaps the simplest way to do this is to identify canonical classes 2i ci ∈ H (BU(n), Z), n ≥ i ∞ which arise from fundamental classes of algebraic cycles on Grassn(P ). Then, 0 since an element ζ ∈ Ktop(T ) corresponds to a homotopy class of maps fζ : T → BU × Z and since j j H (BU, Z) = H (BU(n), Z), j ≤ 2n ∗ we can define ci(ζ) = f (ci). Alternatively, we could approximate T by compact space and use the splitting principle together with the computation of H∗(P(E), Z) as a H∗(T, Z) module. Once we have Chern classes, we can define the Chern character either in terms of Newton polynomials in the Chern classes, in terms of power sums using the splitting principle, or simply by pulling back universal Chern character components 2i chi ∈ H (BU, Q defined universally (using one of these methods) on universal bundles over Grass- mannians. Another important computational tool is the following Atiyah-Hirzebruch spec- ∗ tral sequence which relates Ktop(−) to (singular) cohomology with coeffi- cients. Theorem 3.10. For any topological space T , there exists a spectral sequence p,q p ˜ p+q p p q p+q E2 = H (T, Ktop (S )) = H (T,Ktop(pt)) ⇒ Ktop (T ). Proof. There are two basic approaches to proving this spectral sequence. The first is to assume T is a cell complex, then consider T as a filtered space with Tn ⊂ T the ∗ union of cells of dimension ≤ n. The properties of Ktop(−) stated in the previous theorem give us an exact couple associated to the long exact sequences q n q q q q+1 n · · · → ⊕Ktop(S ) ' Ktop(Tn/Tn−1) → Ktop(Tn) → Ktop(Tn−1) → ⊕Ktop (S ) → · · · where the direct sum is indexed by the n-cells of T . The second approach applies to a general space T and uses the Postnikov tower of BU × Z. This is a tower of fibrations whose fibers are Eilenberg-MacLane spaces for the groups which occur as the homotopy groups of BU × Z.  6

Remark 3.11. The Postnikov tower argument together with a knowledge of the k-invariants of BU × Z shows that after tenoring with Q this Atiyah-Hirzebruch ∗,∗ ∗,∗ spectral sequence collapses; in other words, that E2 ⊗ Q = E∞ ⊗ Q. We give one more important formal property of K-theory, thereby asserting that K-theory satisfies the properties necessary to be an “oriented multiplicative generalized cohomology theory.” Let p : E → T be a rank r, topological vector bundle on T and let π : P(E ⊕1) → T denote the associated projectivized bundle; the fiber over a point t ∈ T is a copy of Pr with p−1(t) ' Ar ⊂ Pr; equivalently, we view this fiber as consisting of lines in the fiber of E ⊕ 1 above t. Observe that there is a tautological map of bundles ∗ on P(E ⊕ 1), OP(E⊕1)(−1) → π (E ⊕ 1), whose restriction to the line above a point t˜ ∈ π−1(t) is simply the embedding of that line into E ⊕ 1. We compose this map with the projection π∗(E ⊕1) → π∗E and consider the corresponding global section ∗ s ∈ OP(E⊕1)(1) ⊗ π E whose zero locus is precisely T ' P(1) ⊂ P(E ⊕ 1). External wedge product with s gives maps i ∗ i+1 ∗ s ∧ (−):Λ (OP(E⊕1)(1) ⊗ π E) → Λ (OP(E⊕1)(1) ⊗ π E), ∗ ∗ thereby endowing Λ (OP(E⊕1)(1) ⊗ π E) with the structure of a cochain complex of vector bundles exact over P(E) (but not exact over T ' P(1) ⊂ P(E ⊕ 1)). Above any point t ∈ T , the restriction of this complex can be viewed as a class in 0 r r−1 0 2r Ktop(P , P ) = Ktop(S ) = Z which is readily seen to be a generator. Thus, λE is a topological K-theory orientation class for P(E ⊕ 1)/P(E), the Thom space for the vector bundle p : E → T . Theorem 3.12. (Thom isomorphism) Let p : E → T be a topological vector bundle on T and let π : P(E ⊕ 1) → T denote the associated projectivized bundle. Then multiplication by λE, the class of X i i ∗ 0 (−1) Λ (OP(E⊕1)(1) ⊗ π E) ∈ Ktop(P(E ⊕ 1), P(E)), i determines an isomorphism ∗ ∼ ∗ Ktop(T ) → Ktop(P(E ⊕ 1), P(E)). Remark 3.13. Observe that this construction applies in the context of an algebraic vector bundle E over an algebraic variety X, giving us a Thom class in K0(P(E ⊕ 1), P(E)). Once again, multiplication by this class is an isomorphism. There are several reasons why topological K-theory has sometimes proved to be a more useful computational tool than singular cohomology. 0 ev • Ktop(−) can be torsion free, even though H (−, Z) might have torsion. This is the case, for example, for compact Lie groups. ∗ • Ktop(−) is essentially Z/2-graded rather than graded by the natural num- bers. ∗ • Ktop(−) has interesting cohomology operations not seen in cohomology. Definition 3.14. Let T be a compact Hausforff space and E → T be a topological vector bundle of rank r. Define r X i i 0 λt(E) = [Λ E]t ∈ Ktop(T )[t], i=0 7

0 a polynomial with constant term 1 and thus an invertible element in Ktop[[t]]. Extend this to a homomorphism 0 0 ∗ λt : Ktop(T ) → (1 + Ktop(X)[[t]]) , i 0 0 (using the fact that λt(E ⊕ F ) = λt(E) · λt(F )) and define λ : Ktop(T ) → Ktop(T ) i to be the coefficient of t of λt. 0 More generally, we define these λ operations on Ktop(T ) for any T by defin- ing them first on the universal vector bundles over Grassmannians and using the 0 functoriality of Ktop(−). Definition 3.15. For any topological space T , define X d ψ (x) = ψi(X)ti ≡ rank(x) − t · (logλ (x)) t dt −t i≥0 0 for any x ∈ Ktop(T ). The Adams operations ψk satisfy many good properties, some of which we list below. 0 Proposition 3.16. For any topological space T , any x, y ∈ Ktop(T ), any k > 0 • ψk(x + y) = ψk(x) + ψk(y). • ψk(xy) = ψk(x)ψk(y). • ψk(ψ`(x) = ψk`(x). k q 2q • chq(ψ (x)) = k chq(x) ∈ H (T, Q). • ψp(x) is congruent modulo p to xp if p is a prime number. • ψk(x) = xk whenever x is a line bundle k k In particular, if E is a sum of line bundles ⊕iLi, then ψ (E) = ⊕((Li) ), the k-th power sum. By the splitting principle, this property alone uniquely determines ψk. top We introduce further operations, the γ-operations on K0 (T ). Definition 3.17. For any topological space T , define X i i γt(x) = γ (X)t ≡ λt/1−t(x) i≥0 0 for any x ∈ Ktop(T ). Basic properties of these γ-oerations include the following

(1) γt(x + y) = γt(x)γt(y) (2) γ([L] − 1) = 1 + t([L] − 1). (3) λs(x) = γs/1+s(x) 0 Using these γ operations, we define the γ filtration on Ktop(T ) as follows. γ,1 Definition 3.18. For any topological space T , define Ktop (T ) as the kernel of the rank map γ,1 0 0 Ktop (T ) ≡ ker{rank : Ktop(T ) → Ktop(π0(T ))}. For n > 1, define 0 γ,n γ,0 0 Ktop(T ) ⊂ Ktop (T ) ≡ Ktop(T ) i1 ik P to be the subgroup generated by monomials γ (x1) ··· γ (xk) with j ij ≥ n, xi ∈ γ,1 Ktop (T ). 8

Finally, we use the Adams operations and the γ-filtration to describe in the 0 following theorem the relationship between Ktop(T ), a group which has no natural grading, and the graded group Hev(T, Q). Theorem 3.19. Let T be a finite cell complex. Then for any k > 0, ψk restricts to γ,n a self-map of each Ktop (T ) and satisfies the property that it induces multiplication by kn on the quotient k n γ,n γ,n+1 ψ (x) = k · x, x ∈ Ktop (T )/Ktop (T )). Furthermore, the Chern character ch induces an isomorphism γ,n γ,n+1 2n chn : Ktop (T )/Ktop (T )) ⊗ Q ' H (T, Q). In particular, the preceding theorem gives us a K-theoretic way to define the 0 grading on Ktop(T ) ⊗ Q induced by the Chern character isomorphism. The graded 0 2n piece of (the associated graded of) Ktop(T )⊗Q corresponding to H (T, Q) consists of those classes x for which ψk(x) = knx for some (or all) k > 0.

References A [1] M. Atiyah, K-thoery, Benjamin 1967. AH [2] M. Atiyah and F. Hirzebruch, Vector bundles and homogeneous spaces, Proc. Symp in Pure Math, vol 3, A.M.S. (1961). B [3] R. Bott. Math Review MR0116022 (22 # 6817) of Le th´eor`emede Riemann-Roch by Armand Borel and Jean-Pierre Serre. FL [4] E. Friedlander and H.B. Lawson, Moving algebraic cycles of bounded degree. Inventiones Math. 132 (1998), 91-119. Fu [5] W. Fulton, Intersection Theory, Springer 1984. H [6] R. Hartshorne, Algebraic Geometry, Springer 1977. W [7] C. Weibel, Algebraic K-theory. http://math.rutgers.edu/ weibel/Kbook.html

Department of Mathematics, Northwestern University, Evanston, IL 60208 E-mail address: [email protected]