Witt Groups of Complex Varieties Marcus Zibrowius Trinity Hall Department of Pure Mathematics and Mathematical Statistics University of Cambridge A thesis submitted for the degree of Doctor of Philosophy April 2011 Declaration of Originality This dissertation is the result of my own work and includes nothing which is the outcome of work done in collaboration. No part of this dissertation has been submitted for any other qualification. i ii Acknowledgements First and foremost, I thank my supervisor Burt Totaro, for his initiative to adopt me as a PhD student, and for his continuous guidance and support. No line of this thesis would have been written without him. Discussions with other mathematicians have greatly enriched my studies. I would particularly like to thank Carl McTague and Julian Holstein for their efforts in establishing a weekly Geometry Tea. Special thanks are moreover due to Baptiste Calme`s and Marco Schlichting, both of whom I wish I had talked to more often and much earlier. Finally, I thank Julia Goedecke for persistent encouragement, as well as last-minute proofreading services. My PhD studies were generously funded by the Engineering and Physical Sciences Research Council. I gratefully acknowledge further financial support from the Cambridge Philosopical Society. iii iv Abstract The thesisWitt Groups of Complex Varieties studies and compares two related cohomology theories that arise in the areas of algebraic geometry and topology: the algebraic theory of Witt groups, and real topological K-theory. Specifically, we introduce comparison maps from the Grothendieck-Witt and Witt groups of a smooth complex variety to the KO-groups of the underlying topological space and analyse their behaviour. We focus on two particularly favourable situations. Firstly, we explicitly com- pute the Witt groups of smooth complex curves and surfaces. Using the theory of Stiefel-Whitney classes, we obtain a satisfactory description of the comparison maps in these low-dimensional cases. Secondly, we show that the comparison maps are isomorphisms for smooth cellular varieties. This result applies in particular to projective homogeneous spaces. By extending known computations in topology, we obtain an additive description of the Witt groups of all projective homogeneous varieties that fall within the class of hermitian symmetric spaces. v Contents Introduction 1 I Two Cohomology Theories 3 I.1 Witt groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1a Symmetric bundles . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1b Symmetric complexes . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 1c Witt groups of schemes . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1d Hermitian K-theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 I.2 KO-theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2a Real versus symmetric bundles . . . . . . . . . . . . . . . . . . . . . . 15 2b Representability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 2c Generalized cohomology theories . . . . . . . . . . . . . . . . . . . . . 21 2d K- and KO-theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 2e The Atiyah-Hirzebruch spectral sequence . . . . . . . . . . . . . . . . 27 II Comparison Maps 31 II.1 An elementary approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 II.2 A homotopy-theoretic approach . . . . . . . . . . . . . . . . . . . . . . . . 36 2a A1-homotopy theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 2b Representing algebraic and hermitian K-theory . . . . . . . . . . . . . 37 2c The comparison maps . . . . . . . . . . . . . . . . . . . . . . . . . . . 40 2d Comparison of the coefficient groups . . . . . . . . . . . . . . . . . . . 44 2e Comparison with Z/2-coefficients . . . . . . . . . . . . . . . . . . . . 45 III Curves and Surfaces 47 III.1 Stiefel-Whitney classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 III.2 Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 2a Grothendieck-Witt groups of curves . . . . . . . . . . . . . . . . . . . 54 2b KO-groups of curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 III.3 Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 3a Witt groups of surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . 59 3b KO/K-groups of surfaces . . . . . . . . . . . . . . . . . . . . . . . . . 65 III.4 Comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4a The classical Witt group . . . . . . . . . . . . . . . . . . . . . . . . . 72 4b Shifted Witt groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 4c Comparison with Z/2-coefficients . . . . . . . . . . . . . . . . . . . . 77 vi IV Cellular Varieties 81 IV.1 The comparison theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 IV.2 The Atiyah-Hirzebruch spectral sequence for cellular varieties . . . . . . . 86 IV.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 3a Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 3b Projective spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 3c Grassmannians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 3d Maximal symplectic Grassmannians . . . . . . . . . . . . . . . . . . . 95 3e Quadrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 3f Spinor varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 3g Exceptional hermitian symmetric spaces . . . . . . . . . . . . . . . . 103 vii viii Introduction In this thesis, we study and compare two related cohomology theories that arise in the areas of algebraic geometry and topology: the algebraic theory of Witt groups, and real topological K-theory. Specifically, we introduce comparison maps from the Grothendieck- Witt and Witt groups of a smooth complex variety to the KO-groups of the underlying topological space and analyse their behaviour. Satisfactory results are obtained in low dimensions and for cellular varieties. The set-up is analogous to the situation one finds in K-theory. Given a smooth complex variety X, we have on the one hand the algebraic K-group K0(X) and on the other the complex topological K-group K0(X). One is defined in terms of algebraic vector bundles over X, the other in terms of complex continuous vector bundles with respect to the analytic topology on X. Moreover, we have a natural map k : K0(X)→ K0(X) that sends an algebraic vector bundle to the underlying continuous bundle. Of course, a given continuous vector bundle may have several different algebraic structures. For exam- ple, the bundles O(1)⊕O(−1) and O⊕O over the projective line P1 are both topologically trivial but they are non-isomorphic as algebraic vector bundles. In this particular case, they nevertheless represent the same equivalence class in the algebraic K-group K0(X), and in fact the comparison map k is an isomorphism for X = P1. But more serious prob- lems already appear on curves of positive genus. Over any such curve, we have an infinite family of line bundles of degree zero, all of which are topologically trivial, but all of which define distinct elements in K0(X). It may also happen that a continuous complex vector bundle has no algebraic structure at all. This occurs, for example, over projective surfaces of positive geometric genus. In general, the comparison map k is neither surjective nor injective. The idea behind Witt groups is to study not simply vector bundles but vector bundles endowed with the additional structure of a symmetric bilinear form. In this context, the analogues of the K-groups are given by the Grothendieck-Witt group GW0(X) and the real topological K-group KO0(X). Again, we have a comparison map gw0 : GW0(X)→ KO0(X) The Witt group W0(X) may be defined as the cokernel of a natural map from K0(X) to GW0(X). Writing (KO0/K)(X) for the corresponding cokernel in topology, we obtain an 1 Witt groups of complex varities induced comparison map w0 : W0(X)→ (KO0/K)(X) These two comparison maps and their generalizations will be our main objects of study. In Chapter I we collect background material. Precise definitions of the objects dis- cussed so far can be found there, along with an introduction to Balmer and Walter’s shifted Grothendieck-Witt and Witt groups GWi(X) and Wi(X). In particular, we explain in what sense these groups constitute a cohomology theory. The aim of Chapter II is to generalize the comparison with topology to these shifted groups, so that we obtain maps gw i : GWi(X)→ KO2i(X) w i : Wi(X)→ (KO2i/K)(X) Two different approaches are discussed. The first is more elementary but has some short- comings. The second, more elegant approach requires the context of A1-homotopy theory. A drawback of our construction is that it relies on some recent results that have not hitherto been properly published. Nonetheless, we believe this is ultimately the more promising route to follow, not least because it places the comparison maps in their natural context. It is this approach that we will focus on. After these preliminaries, we show in Chapter III that the comparison maps behave well in low dimensions. In particular, the maps w i on the Witt groups are isomorphisms for all smooth complex curves. For a surface X, they are isomorphisms if and only if every continuous complex line bundle over X is algebraic. For example, this is the case for all projective surfaces of geometric genus zero. As a first step towards the proof, we explicitly compute the Witt groups of arbitrary smooth complex curves and surfaces. The behaviour of w0 is then analysed using the theory of Stiefel-Whitney classes and the usual comparison theorem for e´tale cohomology with finite coefficients. For the comparison maps on shifted groups, more work is required, and the advantages of the homotopy-theoretic approach to their construction become essential. To round off this chapter, we briefly discuss how the results are related to more general comparison theorems involving Grothendieck-Witt groups with torsion coefficients. In Chapter IV, we show that the comparison maps gw i and w i are isomorphisms for smooth cellular varieties. This result is illustrated with a series of examples to which it applies. Namely, by extending known computations in topology, we obtain a complete additive description of the Witt groups of all projective homogeneous varieties that fall into the class of hermitian symmetric spaces, such as complex Grassmannians, projective quadrics and spinor varieties. 2 Chapter I Two Cohomology Theories In this chapter, we introduce the two main cohomology theories that we will be dealing with: the theory of Witt groups and KO-theory. Witt groups originated in the study of quadratic forms over fields, but the concept was later extended to rings, varieties and schemes by Knebusch [Kne77]. They may be thought of as variants of K-groups: in the same sense that the K-group of a variety X classifies vector bundles over it, there is a Grothendieck-Witt group GW0(X) classifying vector bundles equipped with symmetric forms, of which the Witt group is a quotient. Thus, we begin our discussion by introducing such symmetric bundles and spelling out the necessary definitions. The construction of cohomology theories based on Witt groups is fairly recent. Roughly ten years ago, Balmer gave a purely algebraic construction of a Witt cohomology theory by introducing Witt groups of triangulated categories [Bal00,Bal01a]. Rather than giving a detailed account of the general theory, we concentrate here on those aspects relevant in the context of regular schemes. More comprehensive introductions may be found in [Bal01b] and [Bal05]. We also briefly touch on a more general cohomology theory known as hermitian K-theory, which parallels algebraic K-theory more closely. This theory is currently being developed mainly by Schlichting [HS04,Sch10a,Sch10b,Sch]. The corresponding cohomology theory in topology, known as real topological K-theory or simply as KO-theory, has been well-known and studied since the early work of Atiyah. Its construction, which parallels that of complex topological K-theory perfectly, is recalled in the second half of this chapter. We close with a description of the Atiyah-Hirzebruch spectral sequence, a standard tool in topology that allows us to relate the KO-groups of a space to its singular cohomology. 1 Witt groups The algebraic K-group K0(X) of a scheme X can be defined as the free abelian group on isomorphism classes of vector bundles over X modulo the following relation: for any short exact sequence of vector bundles 0→ E→ F → G→ 0 3 Chapter I. Two Cohomology Theories over X, we have [F] = [E] + [G] in K0(X). In particular, as far as K0(X) is concerned, we may pretend that all exact sequences of vector bundles over X split. Witt groups can be defined similarly, using the notion of symmetric bundles. 1a Symmetric bundles Let X be a scheme over Z[12 ]. A symmetric bundle (E, ε) over X is a vector bundle 1 E over X equipped with a non-degenerate symmetric bilinear form ε : E⊗ E→ O. Alternatively, we may view ε as an isomorphism from E to its dual bundle E∨. In this case, the symmetry of ε is encoded by the fact that it agrees with its dual ε∨ under the canonical identification ω of E with (E∨)∨. That is, the following triangle commutes: E ε '.VVVV VVVV ∼=ω  E∨ (E∨)∨ ε∨ 18iiiiii (1) Two symmetric bundles (E, ε) and (F, ϕ) are isometric if there is an isomorphism of vector bundles i : E→ F compatible with the symmetries, i. e. such that i∨ϕi = ε. The orthogonal sum of two symmetric bundles has the obvious definition (E, ε) ⊥ (F, ϕ) := (E⊕ F, ε⊕ ϕ). 1.1 Example (Symmetric line bundles). Let Pic(X)[2] be the subgroup of line bun- dles of order ≤ 2 in the Picard group Pic(X). Any line bundle L ∈ Pic(X)[2] defines a symmetric bundle over X. When X is a projective variety over an algebraically closed field, all symmetric line bundles arise in this way. In general, the set of isometry classes of symmetric line bundles over X is described by H1et(X;Z/2), the first e´tale cohomology group of X with coefficients in Z/2 = O(1). The Kummer sequence exhibits this group as an extension of Pic(X)[2]: 0→ O ∗(X) O∗(X)2 → H1et(X;Z/2)→ Pic(X)[2]→ 0 The additional contribution comes from symmetric line bundles of the form (O, ϕ), where ϕ is some invertible regular function which has no globally defined square root. For example, the trivial line bundle over the punctured disk A1−0 carries a non-trivial symmetric form given by multiplication with the standard coordinate function. 1.2 Example (Hyperbolic bundles). Any vector bundle E gives rise to a symmetric bundle H(E) := (E⊕ E∨, ( 0 1ω 0 )) over X, the hyperbolic bundle associated with E. Hyperbolic bundles are the simplest members of the wider class of metabolic bundles. A metabolic bundle is a symmetric bundle (M, µ) which contains a Lagrangian, a subbundle 1By convention, we identify a vector bundle with its sheaf of sections. Thus, the common notation O for the sheaf of regular functions on X will be used to denote the trivial line bundle over X. 4 1 Witt groups j : N ↪→M of half the rank ofM on which the symmetric form µ vanishes. In other words, (M, µ) is metabolic if it fits into a short exact sequence of the form 0→ N j−−→M j ∨µ−−→ N∨ → 0 (2) The sequence splits if and only if (M, µ) is isometric to H(N) [Bal05, Example 1.1.21]. This motivates the definition of the Grothendieck-Witt group. 1.3 Definition. [Kne77, § 4] The Grothendieck-Witt group GW0(X) of a scheme X over Z[12 ] is the free abelian group on isometry classes of symmetric bundles over X modulo the following two relations: • [(E, ε) ⊥ (G, γ)] = [(E, ε)] + [(G, γ)] • [(M,µ)] = [H(N)] for any metabolic bundle (M,µ) with Lagrangian N The Witt group W0(X) is defined similarly, except that the second relation reads [(M,µ)] = 0. Equivalently, we may define W0(X) by the exact sequence K0(X) H−→ GW0(X) −→W0(X)→ 0 In addition to the hyperbolic map H appearing here, we have a forgetful map F in the opposite direction, sending symmetric bundles to their underlying vector bundles: GW0(X) F−→ K0(X) The most basic invariant of a vector bundle is its rank. For a connected scheme X, it induces the following well-defined homomorphisms on the above groups: K0(X) rk−→ Z GW0(X) rk−→ Z W0(X) rk−→ Z/2 Of course, on the Witt group the rank is only well-defined modulo two, since arbitrary metabolic bundles are equivalent to zero. 1.4 Example. If k is a field (of characteristic not 2), we simply write K0(k), GW0(k) and W0(k) for the corresponding groups of Spec(k). Since short exact sequences of vector spaces always split, these groups may be defined more directly: if Vect(k) and Bil(k) denote the monoids of vector spaces and of symmetric forms over k, then K0(k) and GW0(k) are simply the Grothendieck groups of Vect(k) and Bil(k), respectively, and W0(k) may be identified with the quotient Bil(k)/(N ·H), where H is the hyperbolic form H = ( 0 11 0 ). Since vector spaces are uniquely determined by their rank, the rank homomorphism on the K-group yields an isomorphism K0(k) ∼=→ Z. In contrast, the rank homomorphisms 5 Chapter I. Two Cohomology Theories on the Grothendieck-Witt and Witt groups are not isomorphisms in general unless k is algebraically closed. For example, the Witt group of R can be identified with Z via the map that sends a real symmetric form to its signature. Variants. Returning to the world of schemes, we may more generally consider vector bundles equipped with non-degenerate symmetric bilinear forms with values in any fixed line bundle L over X. Such vector bundles may be viewed as symmetric bundles with respect to the twisted duality E∨L := Hom(E,L) Under this interpretation, all notions introduced above immediately generalize, leading to the definition of twisted groups GW0(X;L) and W0(X;L). If E is symmetric with respect to the twisted duality ∨L⊗L, then E⊗ L∨ is symmetric with respect to the usual duality. Thus, these groups depend only on the class of L in Pic(X)/2. We could also work with −ω in place of the usual double-dual identification, leading to the notion of anti -symmetric bundles. The corresponding Grothendieck-Witt and Witt groups are denoted by GW2(X;L) and W2(X;L). The choice of notation will become clear in the next section, where shifted groups GWi(X;L) and Wi(X;L) are introduced for arbitrary integers i. Witt groups of exact categories The situation may be formalized by introducing the notion of an exact category with duality (A,∨, ω). An exact category is essentially an additive category A equipped with an abstract notion of short exact sequences, and a duality is determined by an endofunctor ∨ : Aop → A together with a natural isomorphism ω : idA ∼=−−→ ∨ ◦ ∨ which one refers to as double-dual identification. These structures are required to satisfy certain compatibility conditions. Precise definitions may be found in [Sch10a, § 2] or [Bal05, Definitions 1.1.1 and 1.1.13]. The notions of symmetric and metabolic objects over an arbitrary exact category with duality (A,∨, ω) are completely analogous to the notions for bundles given above, and they can be used to define the Grothendieck-Witt and Witt groups GW0(A,∨, ω) and W0(A,∨, ω). From this point of view, GW0(X) and W0(X) are simply the Grothendieck-Witt and Witt groups of the exact category Vect(X) of vector bundles over X equipped with its usual duality and double-dual identification. For the twisted groups, we use the same category equipped with the twisted duality ∨L and the canonical identification ωL of E 6 1 Witt groups with (E∨L)∨L , while for the groups of anti-symmetric bundles we use ∨L and −ωL: GW0(X;L) := GW0(Vect(X),∨L, ωL) GW2(X;L) := GW0(Vect(X),∨L,−ωL) 1b Symmetric complexes The notions of duality and symmetry may be generalized to complexes. If we fix an exact category with duality (A,∨, ω), then the usual dual of a complex E• : · · · → E2 d2−→ E1 d1−→ E0 d0−→ E−1 d−1−→ E−2 → · · · over A is defined term-by-term, i. e. E∨• is given by E∨• : · · · → E∨−2 d∨−1−→ E∨−1 d∨0−→ E∨0 d∨1−→ E∨1 d∨2−→ E∨2 → · · · with E∨0 placed in degree zero. Likewise, the dual of a morphism ϕ• : E• → F• is given by (ϕ∨)l := ϕ∨−l. Alternative dualities may be defined by composing with the shift functor, as follows. By convention, the shifted complex E•[1] is obtained from E• by moving all terms by one position to the left and changing the signs of the differentials, while a morphism ϕ is shifted without any sign changes. Thus, (ϕ[1])l = ϕl−1, and E•[1] takes the form E•[1] : · · · → E1 −d1−→ E0 −d0−→E−1 −d−1−→ E−2 −d−2−→ E−3 → . . . with E−1 in degree zero [Wei94, 1.2.8]. Shifts [i] for arbitrary integers i are obtained by iterating this construction or its inverse. Thus, for any integer i, we can define a shifted duality ∨i on complexes by E∨i• := (E ∨ • )[i] The double-dual identification ω on A induces a canonical identification of E• with (E∨i• )∨i , which we denote by ωi. 1.5 Definition. An i-symmetric complex (E•, ε) is a bounded complex E• together with a quasi-isomorphism ε : E• '−−→ E∨i• which is symmetric with respect to ∨i and the double-dual identification (−1) i(i+1) 2 ·ωi. In other words, ε satisfies the condition ε = (−1) i(i+1)2 · ε∨i ◦ ωi Because the double-dual identification is modified by a sign, this definition generalizes 7 Chapter I. Two Cohomology Theories at the same time the notions of symmetric and anti-symmetric objects: while symmetric objects over A may be viewed as 0-symmetric complexes concentrated in degree zero, anti- symmetric objects may be viewed as 2-symmetric complexes concentrated in degree one. Given any i-symmetric complex (E•, ε), its two-fold shift (E•[2], ε[2]) is (i+4)-symmetric. Thus, the essential aspects of i-symmetry depend only on the value of i modulo 4. Simple examples of i-symmetric complexes for arbitrary i are given by hyperbolic complexes, i. e. complexes of the form H i(E) := (E ⊕ E∨i , ( 0 1 (−1) i(i+1) 2 ωi 0 ) ). A less trivial example is the following. 1.6 Example (A 1-symmetric complex over P1). Consider the complex of vector bundles O(−1) ·x−→ O over the projective line P1 with coordinates [x : y]. Place O in degree zero. Multiplication by y induces a symmetric quasi-isomorphism with the dual complex shifted one to the left, so that we obtain a 1-symmetric complex Ψ0 :=  O(−1) ·x ,2 ·y  O ·(−y)  O ·(−x) ,2 O(1)  (3) Analogues of this 1-symmetric complex over projective curves of higher genus are described in Remark III.2.2. Witt groups of triangulated categories The category Chb(A) of bounded complexes over A equipped with the exact structure inherited from A and any of the dualities defined above is again an exact category with duality. However, i-symmetric complexes are not symmetric objects over (Chb(A),∨i,±ωi) in the above sense, since the symmetries are only required to be quasi-isomorphisms. On the other hand, the derived category Db(A) obtained from Chb(A) by formally inverting all quasi-isomorphisms is no longer exact but rather triangulated. One is thus led to develop the above theory in the context of triangulated categories. The definition of the K-group of a triangulated category is straightfoward: one simply replaces the short exact sequence in the definition of K0 by an exact triangle. For any exact category A, we then have K0(Db(A)) ∼= K0(A). So this is a perfect generalization. Witt groups of triangulated categories are developed by Balmer in [Bal00] and [Bal01a]. Naturally enough, the idea is to replace the short exact sequence (2) defining metabolic objects by an appropriate exact triangle. Once the notions of a triangulated category with duality (D,∨, ω) and its Witt group W(D,∨, ω) are settled, the Witt groups of i-symmetric complexes over an exact category with duality (A,∨, ω) may be defined as Wi(A,∨, ω) := W(Db(A),∨i, (−1) i(i+1) 2 ωi) 8 1 Witt groups These groups are known as the shifted Witt groups of (A,∨, ω). They are 4-periodic in i, in the sense that we have canonical isomorphisms Wi(A,∨, ω) = Wi+4(A,∨, ω). For i = 0 and i = 2, we recover the Witt groups of symmetric and anti-symmetric objects over A defined above, at least when 2 is invertible in A (i. e. when all homomorphism groups of A are uniquely 2-divisible) [Bal01a, Theorem 4.3; BW02, Theorem 1.4]. As Walter notes in [Wal03a], Balmer’s approach already works on the level of Grothen- dieck-Witt groups, so that we also have shifted groups GWi(A,∨, ω). Just like the usual Grothendieck-Witt groups, these are equipped with hyperbolic and forgetful maps. They appear in the following exact sequences [Wal03a, Theorem 2.6]: GWi−1(A,∨, ω) F−→ K0(A) H i−→ GWi(A,∨, ω) −→W0(A,∨, ω)→ 0 Not only do these generalize the exact sequence of Definition 1.3 but they also extend it by one term to the left. 1c Witt groups of schemes The shifted Witt groups of a schemeX are, of course, defined in terms of the exact category (Vect(X),∨, ω), or more generally in terms of (Vect(X),∨L, ωL) for any line bundle L over X. The essence of the previous sections may be summarized as follows. • For any scheme X over Z[12 ], any line bundle L over X and any integer i, we have groups GWi(X;L) := GWi(Vect(X),∨L, ωL) Wi(X;L) := Wi(Vect(X),∨L, ωL) which generalize the Witt groups of symmetric and anti-symmetric bundles discussed earlier [Bal01a, Theorem 4.7]. They are referred to as shifted (Grothendieck-)Witt groups of X with coefficients in L, or as such groups twisted by L. When L is trivial, it is frequently dropped from the notation. • The groups are four-periodic in i and two-periodic in L. That is, for any X, any i and arbitrary line bundles L and M over X, we have canonical isomorphisms GWi(X;L) ∼= GWi+4(X;L) GWi(X;L) ∼= GWi(X;L⊗M⊗2) and similarly for the Witt groups. • We have hyperbolic maps H iL : K0(X) → GWi(X;L) and forgetful maps F in the 9 Chapter I. Two Cohomology Theories opposite direction. They fit into exact sequences of the form GWi−1(X;L) F−→ K0(X) HiL−→ GWi(X;L) −→Wi(X;L)→ 0 (4) These sequences are known as Karoubi sequences. • It follows easily from the definitions that all these constructions are natural. That is, for any morphism of schemes f : X ′ → X and any line bundle L over X, we have induced maps K0(X) f∗−→ K0(X ′) GWi(X;L) f∗−→ GWi(X ′; f∗L) Wi(X;L) f∗−→Wi(X ′; f∗L) which allow us to view K0(−), GWi(−;−) and Wi(−;−) as contravariant functors from schemes with line bundles to abelian groups. The maps f∗ commute with the maps in the Karoubi sequences. 1.7 Example (Witt groups of a geometric point). Let p = Spec(k), where k is an algebraically closed field of characteristic not 2. Its Grothendieck-Witt and Witt groups are as follows: GW0(p) = Z W0(p) = Z/2 GW1(p) = 0 W1(p) = 0 GW2(p) = Z W2(p) = 0 GW3(p) = Z/2 W3(p) = 0 The classical Grothendieck-Witt group GW0(p) is generated by the symmetric bundle (O, id), while GW2(p) is generated by the anti-symmetric bundle (O ⊕ O, ( 0 1−1 0 )). The group GW3(p) is generated by the hyperbolic bundleH3(O). The hyperbolic bundleH1(O) is trivial in GW1(p) because it shares its Lagrangian O with the following 1-symmetric metabolic complex, which is exact: O id ,2 id  O − id  O − id ,2 O It follows from the naturality property that the groups of an arbitrary schemeX contain the groups of a point as direct summands. Namely, the inclusion of a point j : p ↪→ X and the projection of X onto p induce decompositions K0(X) = Z ⊕ K˜0(X) GWi(X) = GWi(p)⊕ G˜Wi(X) Wi(X) = Wi(p) ⊕ W˜i(X) 10 1 Witt groups in which K˜0(X), G˜Wi(X) and W˜i(X) denote the kernels of j∗ on the respective groups. When X is connected, these kernels are independent of the choice of p, and we refer to them as the reduced groups of X. Over an algebraically closed field, the pullback maps j∗ on K0(X), GW0(X) and W0(X) can be identified with the rank homomorphisms. 1.8 Example (Witt groups of P1 [Ara80,Wal03b]). For the projective line P1 over k, where k is as in the previous example, one finds that GW0(P1) = [Z]⊕ Z/2 W0(P1) = [Z/2] GW1(P1) = Z W1(P1) = Z/2 GW2(P1) = [Z] W2(P1) = 0 GW3(P1) = [Z/2]⊕ Z W3(P1) = 0 Here, the summands in square brackets are those generated by constant bundles, i. e. those that disappear when passing to reduced groups. The groups GW1(P1) and W1(P1) are generated by the 1-symmetric complex Ψ0 given in Example 1.6. Grothendieck-Witt and Witt groups of projective spaces of arbitrary dimensions are discussed in IV.3b. In general, when working over an algebraically closed ground field, the Witt groups are always 2-torsion: 2Ψ = 0 for any Ψ ∈ Wi(X;L). This is a consequence of the following well-known lemma. 1.9 Lemma. Let X be a scheme over an algebraically closed field k of characteristic not 2. Then, for any Ψ ∈ GWi(X;L), we have H iL(F (Ψ)) ∼= 2Ψ. Proof. We may assume that Ψ is the class of some i-symmetric complex (E•, ε). Using the assumption that char(k) 6= 2, we define an isometry between the hyperbolic complex H iL(E•) and the direct sum (E•, ε) ⊕ (E•,−ε) by 12 ( 1 ε−1 1 −ε−1 ) . Since k contains a square root of −1, the symmetric complex (E•,−ε) is isometric to (E•, ε). Witt groups of regular schemes One of the key features of Balmer’s triangulated approach is that it elevates the theory of Witt groups into the realm of cohomology theories. This assertion is filled with meaning by the following paragraphs. By convention, a regular scheme will be a regular, noetherian, separated scheme over Z[12 ]. In order to avoid the regularity assumption, we would have to work with Witt groups defined in terms of coherent sheaves rather than vector bundles. Localization sequences. Let X be a regular scheme, in the above sense. Then, for any open subscheme U ⊂ X, we have a long exact sequence relating the Witt groups of X to those of U . These sequences are known as localization sequences. If we denote the closed complement of U by Z and the open inclusion by j : U ↪→ X, the assoicated sequence can 11 Chapter I. Two Cohomology Theories be written as follows: · · · →WiZ(X;L)→Wi(X;L) j ∗ →Wi(U ;L|U )→Wi+1Z (X;L)→ · · · (5) The groups WiZ(X;L) appearing here are the Witt groups of X with support on Z. The terminology comes from the fact that they are defined in terms of complexes whose cohomology is supported on Z. The groups with support on Z = X agree with the usual Witt groups of X. When Z is a smooth closed subvariety of a smooth quasi-projective variety X, we have a de´vissage or Thom isomorphism relating the Witt groups of X with support on Z to the Witt groups of Z. The precise form of these isomorphisms depends not only on the codimension c, but also on the normal bundle N of Z in X [Nen07, § 4]: Wi−c(Z;L|Z ⊗ detN) ∼=−→WiZ(X;L) (6) Thus, in the above situation, the localization sequence may be rewritten purely in terms of the Witt groups of X, U and Z. By periodicity, the localization sequences may be arranged as exact polygons with twelve vertices. We also have localization sequences involving Grothendieck-Witt groups, of the following form [Wal03a, Theorem 2.4]: GWiZ(X)→ GWi(X)→ GWi(U) →Wi+1Z (X)→Wi+1(X)→Wi+1(U)→Wi+2Z (X)→ · · · (7) These sequences agree with the localization sequences for Witt groups from the fourth term onwards, and may of course again be defined for arbitrary twists L. However, if one wishes to continue the sequences to the left, one has to revert to the methods of higher algebraic K-theory. This is discussed very briefly in Section 1d. Excision. Let f : X ′ → X be a morphism of regular schemes. Given a closed subscheme Z ⊂ X, write Z ′ for its preimage X ′×X Z under f . If f is flat and maps Z ′ isomorphically to Z, then f induces isomorphisms of Witt groups with support [Bal01b, Corollary 2.3]: f∗ : WiZ(X) ∼=−→WiZ′(X ′) (8) In particular, if Z is contained in an open subscheme U of X, then WiZ(X) ∼= WiZ(U). Mayer-Vietoris sequences. The existence of localization sequences and the excision property imply the exactness of Mayer-Vietoris sequences. That is, given a covering of a regular scheme X by two open subschemes U and V , we have a long exact sequence of the following form [Bal01b, Theorem 2.5]: · · · →Wi(X)→Wi(U)⊕Wi(V )→Wi(U ∩ V )→Wi+1(X)→ · · · 12 1 Witt groups Homotopy invariance. If E is the total space of a vector bundle over a regular scheme X, then the projection pi : E → X induces isomorphisms of Witt groups: pi∗ : Wi(X) ∼=−→Wi(E) More generally, this holds for any flat affine morphism of regular schemes pi : E → X whose fibres are affine spaces [Gil03, Corollary 4.2]. The properties of Witt groups discussed also hold on the level of Grothendieck-Witt groups. This may be deduced in each case from the corresponding properties of K- groups and the Karoubi sequences (4), using Karoubi induction (c. f. the proof of Propo- sition II.1.3 in the next chapter). Multiplication. Finally, we mention the multiplicative structure on shifted Witt groups. In the same way that the tensor product of vector bundles or complexes over X induces a ring structure on K0(X), we have a ring structure on GW0(X) and W0(X) induced by the tensor product of symmetric bundles. More generally, in [GN03] Gille and Nenashev develop pairings between the shifted Witt groups of X, of the following form: ? : WiZ(X;L)⊗WjZ′(X;M)→Wi+jZ∩Z′(X;L⊗M) Again, this product may be lifted to Grothendieck-Witt groups [Wal03a, end of § 2]. It is graded-commutative in the sense that Ψi ?Ψj = (OX ,− id)ij ?Ψj ?Ψi for Ψi ∈ WiZ(X;L) and Ψj ∈ WjZ′(X;M). Over an algebraically closed field, (OX ,− id) is isometric to the trivial symmetric bundle (OX , id), acting as the unit, so the product becomes honestly commutative. As usual, this “internal” product gives rise to an “external” product between the (Gro- thendieck-)Witt groups of two different schemes X and Y : if L and M are line bundles over X and Y , respectively, we have a cross product × : WiZ(X;L)⊗WjW (Y ;M)→Wi+jZ×W (X × Y ;pi∗XL⊗ pi∗YM) Here, piX and piY denote the respective projections from X × Y to X and Y . For Ψ ∈WiZ(X;L) and Φ ∈WjW (Y ;M), the product is defined as Ψ×Φ := pi∗X(Ψ) ? pi∗Y (Φ). 13 Chapter I. Two Cohomology Theories 1d Hermitian K-theory The algebraic K-group of a scheme fits into a family of higher algebraic K-groups Kn(X). In general, however, there seems to be no purely algebraic description of the higher groups. Rather, following Quillen [Qui73], one defines the higher K-groups of a scheme X as the homotopy groups of a topological space K(X) associated with X. By working with an appropriate spectrum in place of K(X), one may further define groups Kn(X) in all degrees n ∈ Z. For a regular scheme, however, the groups in negative degrees vanish [TT90, Proposition 6.8]. An analogous construction for Grothendieck-Witt groups, usually referred to as (higher) hermitian K-theory, is developed in [Sch10b, Section 10]. Given a scheme X and a line bundle L over it, Schlichting constructs a family of spectra GWi(X;L) from which her- mitian K-groups can be defined as GWin(X;L) := pin(GWi(X;L)) These groups satisfy many properties analogous to those discussed in the case of Witt groups above. The localization sequences now have the following form [Sch10b, Theo- rem 14]: · · · → GWin,Z(X)→ GWin(X)→ GWin(U)→ GWin−1,Z(X)→ · · · (9) In degree n = 0, one recovers Walter’s Grothendieck-Witt groups, while Balmer’s Witt groups appear as hermitian K-groups in negative degrees. More precisely, for any regular scheme X, we have natural identifications GWi0(X;L) ∼= GWi(X;L) GWin(X;L) ∼= Wi−n(X;L) for n < 0 (10) These identifications are due to appear in full generality in [Sch]. For affine varieties, the identifications of Witt groups may be found in [Hor05]: see Proposition A.4 and Corollary A.5. For a general regular scheme X, we can pass to a vector bundle torsor T over X such that T is affine [Jou73, Lemma 1.5; Hor05, Lemma 2.1].1 By homotopy invariance, the Witt groups of T may be identified with those of X, and the same is true for the hermitian K-groups: in this case, homotopy invariance follows from homotopy invariance in the affine case, as shown in [Hor05, Corollary 1.12], and the Mayer-Vietoris sequences established in [Sch10b, Theorem 1]. 1This step is known as Jouanolou’s trick. 14 2 KO-theory 2 KO-theory We now turn to the corresponding cohomology theories in topology. These are known as complex and real topological K-theory, or simply as K- and KO-theory. To ensure that the definitions given here are consistent with the literature, we restrict our attention to finite-dimensional CW complexes.1 Since we are ultimately only interested in topological spaces that arise as complex varieties, this is not a problem. So let X be such a CW complex. If we imitate the definitions of K0(X) and GW0(X) given in Section 1a, using continuous complex vector bundles over X in place of algebraic vector bundles, we obtain the complex and real topological K-groups K0(X) and KO0(X). However, since short exact sequences of vector bundles over CW complexes always split, the definitions may be simplified: 2.1 Definition. For a finite-dimensional CW complex X, the complex topological K- group K0(X) is the free abelian group on isomorphism classes of continuous complex vector bundles over X modulo the relation [E ⊕ G] = [E] + [G]. Similarly, the KO-group KO0(X) is the free abelian group on isometry classes of continuous complex vector bun- dles equipped with non-degenerate symmetric bilinear forms over X, modulo the relation [(E, ε) ⊥ (G, γ)] = [(E, ε)] + [(G, γ)]. In the following, continuous vector bundles will be referred to simply as vector bundles, and a vector bundle equipped with a non-degenerate symmetric bilinear form will be referred to as a symmetric bundle. As in the algebraic case, we may view K0 and KO0 as contravariant functors, defined in this case on the category of finite-dimensional CW complexes and continuous maps between them. Also, we again have natural transformations from K0 to KO0 and vice versa, corresponding to the hyperbolic and the forgetful map. These are traditionally referred to as “realification” and “complexification”, and written as r : K0(X)→ KO0(X) c : KO0(X)→ K0(X) The terminology is explained in the next section. 2a Real versus symmetric bundles There is a more common description of KO0(X) as the K-group of continuous real vector bundles over X. The equivalence with the definition given here can be traced back to the fact that the orthogonal group O(n) is a maximal compact subgroup of both GLn(R) and On(C). Alternatively, this equivalence may be seen concretely along the following lines. 1The key property we need is that any vector bundle over a finite-dimensional CW complex has a stable inverse. See the proof of Theorem 2.7. 15 Chapter I. Two Cohomology Theories We say that a complex bilinear form ε on a (real or complex) vector bundle F is real if ε : F ⊗ F → C factors through R. 2.2 Lemma. Any complex symmetric bundle (E, ε) has a unique real subbundle <(E, ε) ⊂ E such that <(E, ε)⊗RC = E and such that the restriction of ε to <(E, ε) is real and positive definite. Concretely, a fibre of <(E, ε) is given by the real span of any orthonormal basis of the corresponding fibre of E. To prove this lemma, we need the following fact: 2.3 Lemma. Any complex symmetric bundle (F, ϕ) over a topological space X is locally trivial. That is, any point of X has an open neighbourhood over which (F, ϕ) is isometric to the trivial symmetric bundle (CrkF, id). Proof. Let p be an arbitrary point of X. To prove the claim, we may assume without loss of generality that F is trivial as a complex vector bundle, and that the fibre of (F, ϕ) at p is the trivial symmetric space ( Cr, ( 1 0. . . 0 1 )) Let e1, . . . , er be the sections of the trivial bundle Cr that provide the standard basis at each fibre. Since ϕ is orthonormal at p, we may find an open neighbourhood of p on which the deviation of ϕ from orthonormality is arbitrarily small. That is, for any positive  ∈ R, we may find an open neighbourhood of p over which ‖ϕ(ei, ej)− δij‖ <  (∗) for all i and j, where ‖·‖ denotes the standard hermitian metric and δij is the Kronecker delta. The lemma follows from the existence of a Gram-Schmidt process for such “almost orthonormal” symmetric bilinear forms. More precisely, we have the following statement: Let ϕ be a non-degenerate symmetric bilinear form on an r-dimensional com- plex vector space with basis e1, . . . , er. Suppose that (∗) holds for some suffi- ciently small . Then there exists another basis e˜1, . . . , e˜r which is orthonormal with respect to ϕ, such that each e˜i may be expressed by a formula depending continuously on the coefficients ϕ(ei, ej) of ϕ. For lack of reference, we include the following proof of this statement. We take “suffi- ciently small” to mean that  ≤ 1 2r+1r! . Let √· denote the continuous endofunction of {z ∈ C | <(z) > 0} that assigns to a complex number with positive real part that square root whose real part is also positive. The orthonormal basis e˜1, . . . , e˜r is constructed inductively, starting with e˜1 := e1√ ϕ(e1, e1) By definition, ϕ(e˜1, e˜1) = 1. For l > 1, we have ‖ϕ(e˜1, el)‖ < 2. 16 2 KO-theory Now, suppose we have already constructed e˜1, . . . , e˜i satisfying the following two con- ditions: ϕ(e˜j , e˜k) = δjk for all j, k ≤ i (C1) ‖ϕ(e˜j , el)‖ < 2ii! ·  for all j ≤ i and l > i (C2) Set e′i+1 := ei+1 − ∑i k=1 ϕ(ei+1, e˜k)e˜k. Then ϕ(e ′ i+1, e˜k) = 0 for all k ≤ i, and ϕ(e′i+1, e ′ i+1) = ϕ(ei+1, ei+1)− i∑ k=1 ϕ(ei+1, e˜k)2 A crude estimation shows that ‖ϕ(e′i+1, e′i+1)− 1‖ < ‖ϕ(ei+1, ei+1)− 1‖+ i∑ k=1 ‖ϕ(ei+1, e˜k)‖2 < + i(2ii!)2 < 1/2 In particular, ϕ(e′i+1, e ′ i+1) has positive real part, so that we may normalize e ′ i+1 by setting e˜i+1 := e′i+1√ ϕ(e′i+1, e ′ i+1) Then ϕ(e˜j , e˜k) = δij for all k, j ≤ i+ 1, as desired. In other words, (C1) holds with i replaced by i+ 1. The same is true of condition (C2). Indeed, if we fix l > i+ 1, then for any j ≤ i we have ‖ϕ(e˜j , el)‖ < 2ii! < 2i+1(i+ 1)!, and for j = i+ 1 we find that ‖ϕ(e˜i+1, el)‖ ≤ 1 ‖ √ ϕ(e′i+1, e ′ i+1)‖ · ( ‖ϕ(ei+1, el)‖+ i∑ k=1 ‖ϕ(ei+1, e˜k)‖ · ‖ϕ(e˜k, el)‖ ) < 2 · (+ i(2ii!)2) < 2 · (2ii!+ i(2ii!)) = 2i+1(i+ 1)! Proof of Lemma 2.2. In the case of a vector bundle over a point we may assume as above that (E, ε) is isometric to (Cr, id). Clearly, the subspace Rr ⊂ Cr has the required proper- ties. To see uniqueness, suppose W is another r-dimensional real subspace of Cr such that ε|W is real and positive definite. Pick an orthonormal basis of W with respect to ε|W , e1 + if1, e2 + if2, . . . , er + ifr where ej and fj are vectors in Rr and i is the imaginary unit in C. Then we see that ε(ej , fk) = 0 for all j, k (D1) ε(ej , ek) = ε(fj , fk) for all j 6= k (D2) ε(ek, ek) = ε(fk, fk) + 1 for all k (D3) 17 Chapter I. Two Cohomology Theories It follows from (D2) and (D3) that the vectors e1, . . . , er are linearly independent. Indeed, if ∑ j λjej = 0 for certain λj ∈ R, then 0 = ε ( ∑ λjej , ∑ λjej) = ε( ∑ λjfj , ∑ λjfj) + ∑ λ2j Since ε is positive definite on Rr, all λj must be zero. On the other hand, (D1) implies that the real spans of e1, . . . , er and f1, . . . , fr in Rr are orthogonal. Thus, f1 = · · · = fr = 0, and W = Rr. If (E, ε) is an arbitrary complex symmetric bundle over a space X, then by Lemma 2.3 any point of X has some neighbourhood over which (E, ε) can be trivialized in the form above. We know how to define <(E, ε) over each such neighbourhood, and by uniqueness these local bundles can be glued together. 2.4 Corollary. For any CW complex X, the mapping (E, ε) 7→ <(E, ε) defines an iso- morphism between the monoid of isometry classes of complex symmetric bundles over X and the monoid of isomorphism classes of real vector bundles over X. Proof. Given a real vector bundle E overX, we choose an inner product σ on E and consider the complex symmetric bundle (E ⊗R C, σC), where σC denotes the C-linear extension of σ to E⊗R C. Since σ is defined uniquely up to isometry, so is (E⊗R C, σC), and we obtain a well-defined inverse to the map above. An alternative proof that avoids the uniqueness part of the preceding lemma is given in [MH73, Chapter V, § 2]. It follows that the definitions of KO0(X) in terms of complex symmetric bundles and in terms of real bundles agree. Moreover, we see from the concrete description of the correspondence between these types of bundles given in the corollary and its proof that the hyperbolic map K0(X)→ KO0(X) corresponds to the map sending a complex vector bundle to its underlying real bundle, whereas the forgetful map from KO0(X) to K0(X) corresponds to the map that sends a real vector bundle F to its complexification F ⊗R C. This explains the terminology and the notation used for these maps in topology. Real versus non-degenerate Grassmannians The correspondence between complex symmetric and real bundles is reflected by the fact that their projectivizations, or more generally their Grassmannian bundles, are homotopy equivalent. This will be used in the next section to describe a representing space for KO-theory, and it will also be relevant in the discussion of Stiefel-Whitney classes in Section III.1. For any real or complex vector bundle E over a topological spaceX, we have Grassman- nian bundles RGr(k,E) or Gr(k,E) over X whose fibres are given by the Grassmannians of real or complex k-planes in the corresponding fibres of E. There are universal k-bundles 18 2 KO-theory over these spaces, which we denote by UE. For complex symmetric bundles, Grassmanni- ans may be defined as follows. 2.5 Definition. The non-degenerate Grassmannian Grnd(k, (E, ε)) associated with a com- plex symmetric bundle (E, ε) is the open subbundle of the complex Grassmannian bundle Gr(k,E) given in each fibre by those k-planes T for which the restriction ε|T is non- degenerate. The universal symmetric bundle U(E,ε) is defined as the restriction of the universal bundle over Grnd(k,E) endowed with the symmetric form induced by ε. The non-degenerate Grassmannian Grnd(k, (E, ε)) contains the Grassmannian of k- planes of the real bundle <(E, ε). An inclusion j : RGr(k,<(E, ε)) ↪→ Gr(k, (E, ε)) (11) is given by sending k-dimensional subspaces T in the fibres of <(E, ε) to their complexifi- cations T ⊗R C. In fact, we will see in the next lemma that this inclusion is a homotopy equivalence. Moreover, the restriction of the universal bundle U(E,ε) to RGr(k,<(E, ε)) corresponds to the universal real bundle over this space in the sense that <(j∗U(E,ε)) = U<(E,ε) (12) 2.6 Lemma. For any complex symmetric bundle (E, ε), the inclusion j defined in (11) is a homotopy equivalence. A homotopy inverse is provided by a retract of Gr(k, (E, ε)) onto RGr(k,<(E, ε)). Proof. Consider the projection pi : E = <(E, ε)⊕ i<(E, ε) <(E, ε). We define a retract r of j by sending a complex k-plane T in a fibre of E to the subspace pi(<(T, ε|T )) in the corresponding fibre of <(E, ε). This is indeed a linear subspace of real dimension k: since ε is positive definite on <(T, ε|T ) but negative definite on i<(E, ε), the intersection <(T, ε|T ) ∩ i<(E, ε) is trivial. More generally, we define a family of endomorphisms of E parametrized by t ∈ [0, 1] by pit : <(E, ε)⊕ i<(E, ε) <(E, ε)⊕ i<(E, ε) (x, y) 7→ (x, ty) This family interpolates between the identity pi1 and the projection pi0, which we can identify with pi. We claim that for any k-plane T in a fibre of E the image pit(<(T, ε|T )) ⊂ E is a real linear subspace of dimension k on which ε is real and positive definite. The claim concerning the dimension has already been verified in the case t = 0 and follows 19 Chapter I. Two Cohomology Theories for non-zero t from the fact that pit is an isomorphism. Now take a non-zero vector v ∈ pit(<(T, ε|T )) and write it as v = x+ tiy, where x, y ∈ <(E, ε) and x+ iy ∈ <(T, ε|T ). Since ε(x, x), ε(y, y) and ε(x + iy, x + iy) are all real we deduce that ε(x, y) = 0; it follows that ε(v, v) is real as well. Moreover, since ε(x + iy, x + iy) is positive we have ε(x, x) > ε(y, y), so that ε(v, v) > (1− t2)ε(y, y). In particular, ε(v, v) > 0 for all t ∈ [0, 1], as claimed. It follows that T 7→ pit(<(T, ε|T )) ⊗R C defines a homotopy from j ◦ r to the identity on Grnd(k, (E, ε)). 2b Representability The homotopy classification of vector bundles implies the representability of the functors K0 and KO0 in the homotopy category H of topological spaces. More precisely, there are representable functors on H which agree with K0 and KO0 on all finite-dimensional CW complexes. We first recall how this works in the complex case. For KO0, we have two equivalent representing spaces, corresponding to the characterizations in terms of real vector bundles and complex symmetric bundles, respectively. Let us write Grr,n for the Grassmannian Gr(r,Cr+n) of complex r-bundles in Cr+n, and let Grr be the union of Grr,n ⊂ Grr,n+1 ⊂ · · · under the obvious inclusions. Denote the universal r-bundles over these spaces by Ur,n and Ur. For any connected paracompact Hausdorff space X, we have a one-to-one correspondence between the set Vectr(X) of isomorphism classes of rank r complex vector bundles over X and homotopy classes of maps from X to Grr: a homotopy class [f ] in H(X,Grr) corresponds to the pullback of Ur along f [Hus94, Chapter 3, Theorem 7.2]. To describe K0(X), we need to pass to Gr, the union of the Grr under the embeddings Grr ↪→ Grr+1 that send a complex r-plane W to C⊕W . 2.7 Theorem. For finite-dimensional CW complexes X, we have natural identifications K0(X) ∼= H(X,Z×Gr) (13) such that, for X = Grr,n, the class [Ur,n] + (d − r)[C] in K0(Grr,n) corresponds to the inclusion Grr,n ↪→ {d} ×Grr,n ↪→ Z×Gr. Proof. The theorem is of course well-known, see for example [Ada95, page 204]. To deduce it from the homotopy classification of vector bundles, we note first that any CW complex is paracompact and Hausdorff [Hat09, Proposition 1.20]. Moreover, we may assume that X is connected. The product Z×Gr can be viewed as the colimit of the inductive system∐ d≥0 {d} ×Grd ↪→ ∐ d≥−1 {d} ×Grd+1 ↪→ ∐ d≥−2 {d} ×Grd+2 ↪→ · · · ⊂ Z×Gr 20 2 KO-theory Any continuous map from X to Z × Gr must factor through one of the components colimn({d} × Grn). By cellular approximation, it is in fact homotopic to a map that factors through {d} ×Grn for some n. Thus, H(X,Z×Gr) ∼= ∐ d∈Z colimnVectn(X) where the colimit is taken over the maps Vectn(X)→ Vectn+1(X) sending a vector bundle E to C⊕E. We define a map from the coproduct to K0(X) by sending a vector bundle E in the dth component to the class [E]+(d−rkE)[C] in K0(X). To see that this is a bijection, we use the fact that every vector bundle E over a finite-dimensional CW complex has a stable inverse: a vector bundle E⊥ over X such that E⊕ E⊥ is a trivial bundle [Hus94, Chapter 3, Proposition 5.8]. If we replace the complex Grassmannians by real Grassmannians RGrr,n, we obtain the analogous statement that KO0 can be represented by Z×RGr. Equivalently, but more in the spirit of Definition 2.1, we could work with the non-degenerate Grassmannians defined in the previous section. So let Grndr,n abbreviate Gr nd(r,Hr+n), where H is the hyperbolic plane (C2, ( 0 11 0 )), and let Undr,n denote the restriction of the universal bundle over Gr(r,C2r+2n) to Grndr,n. Then colimits Grndr and Gr nd can be defined in the same way as for the usual Grassmannians. By (12) and Lemma 2.6 the homotopy classification of real vector bundles is equivalent to the homotopy classification of complex symmetric bundles in the sense that we have commutative diagrams( isomorphism classes of real vector bundles of rank r over X ) H(X,RGrr) ∼=lr ( isometry classes of complex symmetric bundles of rank r over X )< ∼= LR H(X,Grndr ) ∼=lr j∗ ∼= LR For finite-dimensional CW complexes X, we obtain natural identifications KO0(X) ∼= H(X,Z×Grnd) (14) Here, for even (d− r), the inclusion Grndr,n ↪→ {d} ×Grndr,n ↪→ Z×Grnd corresponds to the class of [Undr,n] + d−r 2 [H] in GW 0(Grndr,n). 2c Generalized cohomology theories The representable functors K0 and KO0 may be embedded into sequences of functors Ki and KOi defined for all integers i. These constituted the first examples of generalized coho- 21 Chapter I. Two Cohomology Theories mology theories in topology. By definition, a generalized cohomology theory is a sequence of functors Ei from topological spaces to abelian groups satisfying all the axioms of a cohomology theory formulated by Eilenberg and Steenrod except for one: the dimension axiom. That is, it is not required that Ei(point) vanishes for i 6= 0. Indeed, as we will see, the values of Ki(X) and KOi(X) are periodic in i, and in particular, there are non-zero groups Ki(point) and KOi(point) in arbitrarily high degrees. Here, we describe the general method of defining generalized cohomology theories via spectra. The construction of spectra Ktop and KOtop representing topological K- and KO-theory is sketched at the beginning of the next section. The classical book [Ada95] may be consulted both for general background on spectra and for details concerning the particular case of K-theory; see in particular Example 2.2. Generalized cohomology theories arise as representable functors on the stable homotopy category SH, the homotopy category of spectra. More precisely, a generalized cohomology theory E∗ may be obtained from a spectrum E as follows. Recall that SH is a triangulated category in which the shift functor is given by suspension. We will denote the suspension by S1 ∧ −. Given any spectrum X, one defines the cohomology groups of X with respect to E by E˜i(X) := SH(X, Si ∧ E) Since suspension is invertible, these groups are defined for all integers i. Using the functor Σ∞ : H• → SH from the pointed to the stable homotopy category that assigns to a pointed space (X,x) its suspension spectrum, the definition may be specialized to pointed spaces. That is, we define the (reduced) generalized cohomology theory on H• corresponding to E by E˜i(X,x) := Ei(Σ∞(X,x)) If X is connected, we often simply write E˜i(X) for E˜i(X,x), where x is an arbitrary point of X. Finally, the cohomology groups Ei(X) of an unpointed space are defined as the reduced groups of X+, the union of X with a disjoint base point: Ei(X) := E˜i(X+) For connected X, we have a canonical decomposition Ei(X) = Ei(point)⊕ E˜i(X). The fact that the functors E∗ define a generalized cohomology theory follows directly from the construction. We emphasize two properties: • For the reduced theory E˜∗, we have suspension isomorphisms σn : E˜i−n(X,x) ∼= E˜i(Sn ∧ (X,x)) (15) • Exact triangles in SH give rise to long exact sequences of cohomology groups. For 22 2 KO-theory example, any continuous map f : X ′ → X fits into an exact triangle involving its mapping cone C(f): Σ∞(X ′+) Σ∞(f+) ,2 Σ∞(X+) qxjjjj jj Σ∞(C(f)) fm This triangle induces the long exact sequence · · · → E˜i(C(f))→ Ei(X) f ∗ −→ Ei(X ′)→ E˜i+1(C(f))→ · · · 2d K- and KO-theory We now specialize the preceding discussion to the cases of K- and KO-theory. After sketch- ing the constructions of the corresponding spectra, we discuss some general consequences along with more specific aspects of these two theories. The key ingredient in the construction of a K-theory spectrum Ktop is Bott periodicity. One observes that the infinite Grassmannian Gr can be identified with the classifying space BU of the infinite unitary group, so that K0 is represented by Z×BU. By Bott periodicity, this space is equivalent to its own two-fold loop space Ω2(Z×BU). Thus, one may define a 2-periodic Ω-spectrum Ktop whose even terms are all given by Z× BU. Similarly, RGr is equivalent to the classifying space BO of the infinite orthogonal group. In this case, Bott periodicity says that Z×BO is equivalent to Ω8(Z×BO), so one obtains a spectrum KOtop which is 8-periodic. Following the definition in the previous section, we obtain 2- and 8-periodic cohomology theories on reduced spaces: K˜i(X,x) := SH(Σ∞(X,x), Si ∧Ktop) K˜Oi(X,x) := SH(Σ∞(X,x), Si ∧KOtop) The groups K˜i(X) and Ki(X) of unpointed spaces and the corresponding KO-groups are defined exactly as above. It follows from the construction that the functors K2i and KO8i agree with the functors K0 and KO0 that we started with. If X is a connected finite-dimensional CW complex, then K˜0(X) may be identified with the subgroup of K0(X) given by virtual bundles of rank zero, i. e. by those elements [E]− [F] of K0(X) for which rkE−rkF = 0. Moreover, the suspension isomorphisms allow us to express all the lower K-groups K−i(X) and KO−i(X) in terms of K˜0 and K˜O0. That is, for all i ≥ 0 we have the following isomorphisms: K−i(X) ∼= K˜0(Si ∧ (X+)) KO−i(X) ∼= K˜O0(Si ∧ (X+)) 23 Chapter I. Two Cohomology Theories Thus, whenX is a finite-dimensional CW complex, its K- and KO-groups may be described in terms of vector bundles over suspensions of X. 2.8 Example (K-groups of a point). The topological K-group of a point is the free abelian group generated by the trivial complex line bundle. Thus, K2i(point) ∼= Z for all integers i. On the other hand, the preceding isomorphisms imply that K−2(point) ∼= K˜0(S2) An explicit generator of K˜0(S2) is given by [τ ]− [C], where τ denotes the Hopf bundle over S2. If we identify S2 with the complex projective line CP1, the Hopf bundle corresponds to the tautological or universal bundle O(−1). Long exact cohomology sequences. As we have seen, any continuous map f : X ′ → X gives rise to long exact sequences of K- and KO-groups. By periodicity, these may be arranged as exact polygons with 6 and 24 vertices, respectively. We examine two particular cases. • If f : A→ X is a cofibration, for example an inclusion of a closed subcomplex A into a CW complex X, the mapping cone C(f) is homotopy equivalent to the quotient space X/A. The long exact sequences then take the form in which they appear most often in topology. The groups K˜Oi(X/A) are usually denoted KOi(X,A), and similarly for the K-groups. • Suppose X is a smooth manifold with a smooth submanifold Z. Let f : U ↪→ X denote the inclusion of the open complement of Z into X. Then the mapping cone C(f) is the homotopy quotient X/hU , which may be realized as the quotient of X by the closed complement of a tubular neighbourhood of Z in X. It follows that C(f) is homotopy equivalent to the Thom space of the normal bundle N of Z in X. Thus, the long exact sequences have the form · · · → K˜Oi(ThomZ N)→ KOi(X)→ KOi(U)→ K˜Oi+1(ThomZ N)→ · · · (16) Alternatively, we will sometimes denote the groups K˜i(X/hU) and K˜O i(X/hU) by KiZ(X) and KO i Z(X), in analogy with the notation used in algebraic geometry. Multiplication. The spectra Ktop and KOtop are ring spectra. For the associated coho- mology theories, this means that we have multiplication maps K˜Oi(X,x)⊗ K˜Oj(Y, y) ·→ K˜Oi+j((X,x) ∧ (Y, y)) KOi(X)⊗KOj(Y ) ×−→ KOi+j(X × Y ) on KO-groups and similarly for K-groups. These products generalize the products that may be defined in terms of vector bundles for finite-dimensional CW complexes [Gra75, 24 2 KO-theory Theorem 29.14]. They are natural and respect the suspension isomorphisms (15) in the sense that σn(x · y) = σn(x) · y. In particular, σn itself may be expressed as multiplication with σn(1), where 1 denotes the unit in K0(point) or KO0(point), respectively. The periodicity isomorphisms Ki(X) ∼= Ki−2(X) and KOi(X) ∼= KOi−8(X) are induced by multiplication with generators g ∈ K−2(point) and λ ∈ KO−8(point). The generator g is described in Example 2.8. Coefficient rings. The groups Ei(point) of a generalized cohomology theory E∗ are known as its coefficient groups. For K-theory, we simply have K0(point) ∼= Z K1(point) = 0 The coefficient groups of KO-theory are more complicated [Bot69, page 66]: KO0(point) = Z KO7(point) = Z/2 KO2(point) = 0 KO1(point) = 0 KO4(point) = Z KO3(point) = 0 KO6(point) = Z/2 KO5(point) = 0 (17) We have arranged them in a slightly unusual fashion so as to facilitate the comparison with Example 1.7. Since our theories are multiplicative, their coefficient groups may moreover be assembled to coefficient rings E∗(point) = ⊕Ei(point). These can be written as follows [Bot69, page 741]: K∗(point) = Z [ g, g−1 ] (18) KO∗(point) = Z [ η, α, λ, λ−1 ]/ (2η, η3, ηα, α2 − 4λ) (19) Here, g and λ are the generators of K−2(point) and KO−8(point) inducing the periodicity isomorphisms, as mentioned above. The generators η and α are of degrees −1 and −4, respectively. The Bott sequence. The spectra Ktop and KOtop fit into an exact triangle of the form KOtop ∧ S1 η ,2 KOtop szmmm mmm m Ktop fm This triangle induces long exact sequences known as Bott sequences [Bot69, pages 75 and 1122; BG10, 4.I.B]. Concretely, the Bott sequence of a topological space X looks as follows: 1Unfortunately, the relation ηα = 0 is missing here, and this omission seems to have pervaded much of the literature. Of course, the relation follows from the fact that KO−5(point) = 0. 2There are misprints on both pages. In particular, the central group in the diagram on page 112 should be K0. 25 Chapter I. Two Cohomology Theories . . .→ KO2i−1X → KO2i−2X → K2i−2X → KO2iX → KO2i−1X → K2i−1X → KO2i+1X → KO2iX → K2iX → KO2i+2X → KO2i+1X → . . . (20) The maps appearing in this sequence may be described explicitly: • The maps KOjX → KjX are the complexification maps c. • The maps Kj−2X → KOjX are the composites of the periodicity isomorphisms Kj−2X ∼= Kj X and the realification maps r. • The maps between KO-groups are given by multiplication with η ∈ KO−1(point). Twisted KO-groups It is possible to define twisted KO-groups KO0(X;L) for complex line bundles L over X similarly to the way this is done for Grothendieck-Witt groups. It turns out, how- ever, that these groups may alternatively be expressed as the usual (reduced) KO-groups K˜O2(ThomL) of the Thom space of L [AR76, 3.8]. Here, we take this identification as our definition. More generally, we define KO-groups twisted by arbitrary complex bundles E over X as follows. 2.9 Definition. For any complex vector bundle E of constant rank r over a topological space X, we define KOp(X;E) := K˜Op+2r(ThomE) When E is a trivial bundle, its Thom space is just a suspension of X, so that KOp(X;E) agrees with the usual KO-group KOp(X). Moreover, the Thom isomorphisms for KO- theory of [ABS64] show that the groups KO(X;E) only depend on the determinant line bundle of E: 2.10 Lemma. For complex vector bundles E and F over a topological space X with iden- tical first Chern class modulo two, we have KOp(X;E) ∼= KOp(X;F) Proof. A complex vector bundle E whose first Chern class vanishes modulo two has a spin structure and is therefore oriented with respect to KO-theory [ABS64, § 12]. That is, we have a Thom isomorphism KOpX ∼=−→ K˜Op+2r(ThomE) Now suppose c1(E) ≡ c1(F) mod 2. We may view E⊕ E⊕ F both as a vector bundle over E and as a vector bundle over F, and by assumption it is oriented with respect 26 2 KO-theory to KO-theory in both cases. Thus, both groups in the lemma can be identified with KOp(X;E⊕ E⊕ F). Remark. In general, the identifications of Lemma 2.10 are non-canonical. Given a spin structure on a real vector bundle, the constructions in [ABS64] do yield a canonical Thom class, but there may be several different spin structures on the same bundle. Still, canonical identifications exist in many cases. For example, there is a canonical spin structure on the square of any complex line bundle, yielding canonical identifications KOp(X;L) ∼= KOp(X;L⊗M⊗2) for any two complex line bundles L and M over X. In general, different spin structures on a spin bundle over X are classified by H1(X;Z/2). 2e The Atiyah-Hirzebruch spectral sequence In topology, there is a standard computational tool for generalized cohomology theories known as the Atiyah-Hirzebruch spectral sequence (AHSS). Given a generalized cohomo- logy theory E∗ and a finite-dimensional CW complex X, this spectral sequence takes the form [Ada95, III.7; Koc96, Theorem 4.2.7] Ep,q2 = H p(X;Eq(point))⇒ Ep+q(X) It is concentrated in the half-plane p ≥ 0 and has differentials dr of bidegree (r,−r + 1). In good cases, it enables us to compute the cohomology E∗(X) from the coefficient groups of E∗ and the singular cohomology of X. The spectral sequence is natural with respect to continuous maps of finite-dimensional CW complexes. Moreover, if E∗ is a multiplicative cohomology theory represented by a ring spectrum, then the spectral sequence is also multiplicative [Koc96, Proposition 4.2.9]. That is, the multiplication on the E2-page induced by the cup product on singular coho- mology and the ring structure of E∗(point) descends to a multiplication on all subsequent pages, such that the multiplication on the E∞-page is compatible with the multiplica- tion on E∗(X). In particular, each page is a module over E∗(point). In this case, the differentials of the spectral sequence are derivations, i. e. they satisfy a Leibniz rule. In some situations, it is more natural to consider the spectral sequence for the reduced cohomology theory E˜∗ associated with E∗. Then the spectral sequence is written as E˜p,q2 = H˜ p(X;Eq(point))⇒ E˜p+q(X) 27 Chapter I. Two Cohomology Theories The AHSS for K- and KO-theory Specializing to the cases of K- and KO-theory, we obtain the following two multiplicative spectral sequences. Ep,q2 = H p(X; Kq(point))⇒ Kp+q(X) Ep,q2 = H p(X; KOq(point))⇒ KOp+q(X) For K-theory, the spectral sequence has the singular cohomology of X with integral coef- ficients in all even rows, while the odd rows vanish. For KO-theory, the spectral sequence is 8-periodic in q. We have the integral cohomology of X in rows q ≡ 0 and q ≡ −4 mod 8, its cohomology with Z/2-coefficients in rows q ≡ −1 and q ≡ −2, and all other rows are zero. RLq H0(X;Z) H1(X;Z) H2(X;Z) Sq2◦piYY YYYYY )/YYYYYY H3(X;Z) H4(X;Z) p ,2 H0(X;Z/2) H1(X;Z/2) H2(X;Z/2) Sq2 YYYYYY )/YYYYY Y H3(X;Z/2) H4(X;Z/2) H0(X;Z/2) H1(X;Z/2) H2(X;Z/2) H3(X;Z/2) H4(X;Z/2) 0 0 0 0 0 H0(X;Z) H1(X;Z) H2(X;Z) H3(X;Z) H4(X;Z) 0 0 0 0 0 ... ... ... ... ... Figure 1: The E2-page of the Atiyah-Hirzebruch spectral sequence computing KO∗(X) The low-order differentials of these spectral sequences may be described as follows. Let pi : H∗(X;Z)→ H∗(X;Z/2) Sq2 : H∗(X;Z/2)→ H∗+2(X;Z/2) β : H∗(X;Z/2)→ H∗+1(X;Z) denote reduction modulo two, the second Steenrod square, and the Bockstein homo- morphism, respectively. For K-theory, the differentials on the E3-page of the Atiyah- Hirzebruch spectral sequence are given by the composition β◦Sq2◦pi [HJJS08, Chapter 21, § 5]. For KO-theory, the differentials d∗,02 and d∗,−12 on the E2-page are given by Sq2 ◦pi and Sq2, respectively. The differential d∗,−23 on the E3-page can be identified with β ◦ Sq2 [Fuj67, 1.3]. 28 2 KO-theory The AHSS for Thom spaces In order to compute twisted KO-groups as discussed in Section 2d, we need to apply the Atiyah-Hirzebruch spectral sequence of KO-theory to Thom spaces. So let X be a finite- dimensional CW complex, and let pi : E → X be a vector bundle of constant rank over X. Though we will be mainly interested in the case when E is complex, we may more generally assume here that E is any real vector bundle which is oriented. Then the Thom isomorphism for singular cohomology tells us that the reduced cohomology of the Thom space ThomE is additively isomorphic to the cohomology of X itself, apart from a shift in degrees by r := rkRE. The isomorphism is given by multiplication with a Thom class θ in H˜r(ThomE;Z): H∗(X;Z) ∼=−→ H˜∗+r(ThomE;Z) x 7→ pi∗(x) · θ Similarly, the reduction of θ modulo two induces an isomorphism of the respective singular cohomology groups with Z/2-coefficients. Thus, apart from a shift of columns, the entries on the E2-page of the spectral sequence for K˜O∗(ThomE) are identical to those on the E2-page for KO∗(X). However, the differentials may differ. 2.11 Lemma. Let E pi→ X be a complex vector bundle of constant rank over a topological space X, with Thom class θ as above. The second Steenrod square on H˜∗(ThomE;Z/2) is given by “ Sq2 + c1(E)”, where c1(E) is the first Chern class of E modulo two. That is, Sq2(pi∗x · θ) = pi∗ (Sq2(x) + c1(E)x) · θ for any x ∈ H∗(X;Z/2). More generally, if E is a real oriented vector bundle, the second Steenrod square on the cohomology of its Thom space is given by “ Sq2 + w2(E)”, where w2 is the second Stiefel-Whitney class of E. Proof. This is a special case of the following well-known identity of Thom [MS74, page 91]: Sqi(pi∗x · θ) = pi∗ (Sqi(x) + wi(E)x) · θ The higher differentials in the spectral sequence for K˜O∗(ThomE) also depend only on the second Stiefel-Whitney class of E. This follows from the observation that the Atiyah- Hirzebruch spectral sequence is compatible with Thom isomorphisms, which is made more precise by the next lemma. 2.12 Lemma. Let E be a real oriented vector bundle over a finite-dimensional CW com- plex X, of constant rank r. Suppose E is oriented, and let θ be a Thom class as above. If E is moreover oriented with respect to KO-theory, then θ survives to the E∞-page of the Atiyah-Hirzebruch spectral sequence computing K˜O∗(ThomE), and the Thom iso- morphism for H∗ extends to an isomorphism of spectral sequences. That is, for each 29 Chapter I. Two Cohomology Theories page right multiplication with the class of θ in E˜r,0s (ThomE) gives an isomorphism of E∗,∗s (X)-modules E∗,∗s (X) ·θ−→∼= E˜ ∗+r,∗ s (ThomE) Moreover, any lift of θ ∈ E˜r,0∞ (ThomE) to K˜Or(ThomE) defines a Thom class of E with respect to KO-theory. The isomorphism of the E∞-pages of the spectral sequences is induced by the Thom isomorphism given by multiplication with any such class. Proof. We may assume without loss of generality that X is connected. Fix a point x on X. The inclusion of the fibre over x into E induces a map ix : Sr ↪→ ThomE. By assumption, the pullback i∗x on ordinary cohomology maps θ to a generator of H˜r(Sr), and the pullback on K˜O∗ gives a surjection K˜O∗(ThomE) i∗x K˜Or(Sr) Consider the pullback along ix on the E∞-pages of the spectral sequences for Sr and ThomE. Since we can identify E˜r,0∞ (ThomE) with a quotient of K˜Or(ThomE) and E˜r,0∞ (Sr) with K˜Or(Sr), we must have a surjection i∗x : E˜ r,0 ∞ (ThomE) E˜r,0∞ (Sr) On the other hand, the behaviour of i∗x on E˜ r,0∞ is determined by its behaviour on H˜r, whence we can only have such a surjection if θ survives to the E˜∞-page of ThomE. Thus, all differentials vanish on θ, and if multiplication by θ induces an isomorphism from E∗,∗s (X) to E˜∗+r,∗s on page s, it also induces an isomorphism on the next page. Lastly, consider any lift of θ to an element Θ of K˜Or(ThomE). It is clear by construction that right multiplication with Θ gives an isomorphism from E∞(X) to E˜∞(ThomE), and thus it also gives an isomorphism from KO∗(X) to K˜O∗(ThomE). Thus, Θ is a Thom class for E with respect to KO-theory. Lemma 2.12 allows the following strengthening of Lemma 2.10: 2.13 Corollary. For complex vector bundles E and F over X with identical first Chern class modulo two, the spectral sequences computing K˜O∗(ThomE) and K˜O∗(ThomF) can be identified up to a possible shift of columns when E and F have different ranks. 30 Chapter II Comparison Maps If X is a smooth complex variety, then we can consider both the algebraic and the topo- logical cohomology theories discussed in the previous chapter. That is, on the one hand we can study the algebraic K-group and the (Grothendieck-)Witt groups of X viewed as a variety, and on the other hand we can study the topological complex and real K-groups of X viewed as a topological space, i. e. the K-groups of the set X(C) of complex points of X equipped with the analytic topology. It follows directly from the definitions of these groups that we have natural maps k : K0(X)→K0(X(C)) (1) gw0 : GW0(X)→KO0(X(C)) (2) If we write (KO0/K)(X(C)) for the cokernel of the realification map from K0(X(C)) to KO0(X(C)), we moreover have an induced map w0 : W0(X)→(KO0/K)(X(C)) (3) Our aim in this chapter is to extend these maps to be defined on shifted groups GWi(X) and Wi(X) for arbitrary integers i in such a way that they are compatible with as much structure of the respective cohomology theories as possible. Two different approaches will be discussed. The first approach is the more elementary one. The idea is to use the multiplicative structure of the theories to define the maps on shifted Grothendieck-Witt groups in such a way that many of their properties can be checked by explicit calculations. This approach has previously been detailed in [Zib09]. We refrain from reproducing the full details here, but only sketch the key steps in the argument. The results obtained in [Zib09] only just suffice to prove the comparison theorem for cellular varieties included in Chapter IV of this thesis. However, as we will explain, the fundamental question of whether the comparison maps are in general compatible with the boundary maps appearing in long exact localization sequences remains open. The idea of the second approach is to use a construction of Grothendieck-Witt groups, or more generally of hermitian K-theory, that resembles the homotopy-theoretic definition of KO-theory so closely that all compatibility issues disappear. More precisely, our second 31 Chapter II. Comparison Maps definition of the comparison maps will rely on the following theorem from A1-homotopy theory: Hermitian K-theory is representable in the stable A1-homotopy category by a spectrum whose complex realization in the usual stable homotopy category represents KO-theory. Unfortunately, although this result seems to be well-known among experts, it has not yet been properly published. We therefore highlight both this and a related statement that we will be using in the form of Assumption 2.2. Proofs of both parts of the assumption are announced by Morel in [Mor06]. To add further credibility to the first statement, we explain in Section 2b how the analogous claim in the unstable homotopy category may be deduced from recent results of Schlichting and Tripathi. A brief summary of those superficial aspects of A1-homotopy theory that we will constantly be referring to is included in Section 2a. 1 An elementary approach Recall from I.2d that the suspension isomorphisms in KO-theory are given by multiplica- tion with generators of the KO-groups of spheres. That is, if e˜n denotes the generator of K˜On(Sn) corresponding to the unit in KO0(point) under suspension, then the suspension isomorphism for an arbitrary pointed space (X,x) is given by multiplication with e˜n: K˜Oi−n(X,x) ∼=−−→ K˜Oi(Sn ∧ (X,x)) Similarly, for an unpointed space X, we have isomorphisms KOi−n(X) ∼=−−→ K˜Oi(Sn ∧ (X+)) KOi(X)⊕KOi−n(X) ∼=−−→ KOi(Sn ×X) Explicitly, if we write en for the element of KOn(Sn) corresponding to e˜n and pi for the projection Sn×X  X, then the lower map is given by (x, y) 7→ pi∗x+ en× y. Analogous isomorphisms exist for Grothendieck-Witt groups: 1.1 Lemma. Let X be a smooth variety with some closed subset Z. Let Ψ0 be the gener- ator of GW1(P1) described in Example I.1.6. We have isomorphisms GWi−1Z (X) ∼=−−→ GWiZ(X × P1) GWi−1Z (X) ∼=−−→ GWiZ(X × A1) given by multiplication with the restriction of Ψ0 to GW1{0}(P 1) and its further restriction 32 1 An elementary approach to GW1{0}(A 1), respectively. Moreover, we have an isomorphism GWiZ(X)⊕GWi−1Z (X) ∼=−−→ GWiZ×P1(X × P1) given by (x, y) 7→ pi∗x+ y×Ψ0, where pi denotes the projection X ×P1  X. All of these isomorphisms also hold with Witt groups in place of Grothendieck-Witt groups. Proof. For Witt groups, the first isomorphism is a special case of Theorem 2.5 in [Nen07], the case when Z = X being Theorem 8.2 in [BG05]. The second isomorphism follows via excision. The decomposition of Wi(X × P1) into Wi(X) ⊕Wi−1(X) is a special case of Theorem 1.5 in [Wal03b]. It may be deduced from the second isomorphism by considering the localization sequences associated with the inclusion of X×A1 into X×P1, since these sequences split. A decomposition of WiZ(X × P1) can be obtained analogously, using the more general localization sequences of triples discussed in Theorem 1.6 of [Bal01b]. Lastly, since the corresponding isomorphisms also hold for K0, with F (Ψ0) ∈ K0(P1) in place of Ψ0, the corresponding isomorphisms of Grothendieck-Witt groups may be deduced via Karoubi induction (c. f. the proof of Lemma 1.3 below and Lemma 2.3 in [Zib09]). Now let X be a smooth complex variety. Then the two isomorphisms GWi(X)⊕GWi−1(X) ∼=−−−→ GWi(X × P1) KO2i(X(C))⊕KO2i−2(X(C)) ∼=−→KO2i(X(C)× S2) can be used to define comparison maps gw i : GWi(X)→ KO2i(X(C)) inductively for all non-positive i. Explicitly, if gw i is already defined for some i ≤ 0, then for x ∈ GWi−1(X) we define gw i−1(x) to be the unique element in KO2i−2(X(C)) satisfying1 gw i−1(x)× e2 = gw i(x× (−Ψ0)) The following properties of these maps can be checked via explicit calculations. Firstly, it is clear from the construction that they are natural with respect to morphisms of smooth varieties, and that they respect the multiplicative structures of Grothendieck-Witt groups and KO-theory. Moreover, they are compatible with the Karoubi and Bott sequences (see 1We use −Ψ0 instead of Ψ0 because the image of e2 under the complexification map KO2(S2) c−→ K2(S2) and the periodicity isomorphism K2(S2) ∼=−→ K0(S2) agrees with the image of −Ψ0 under the forgetful map GW1(P1) F−→ K0(X): we have k(F (−Ψ)) = k(O(−1) − O) = [τ ] − [C] = σ−2(g) = c(e2) × g. See Examples I.1.6 and I.2.8. 33 Chapter II. Comparison Maps (I.4) and (I.20)) in the sense that we have induced maps w i : Wi(X)→ (KO2i/K)(X(C)) and commutative diagrams of the following form: GWi−1(X) F ,2 gw i−1  K0(X) Hi ,2 k  GWi(X) ,2 gw i  Wi(X) ,2 w i  0 KO2i−2(X(C)) g i−1◦c ,2 K0(X(C)) r◦g −i ,2 KO2i(X(C)) ,2 (KO2i/K)(X(C)) ,2 0 To obtain comparison maps in positive degrees, one checks next that the maps already defined respect the periodicity isomorphisms. This can again be done by expressing these isomorphisms in a multiplicative way. We include a short proof to convey the flavour of these arguments. 1.2 Lemma. The maps gw i : GWi(X) → KO2i(X(C)) defined above respect the periodi- city isomorphisms GWi(X) ∼= GWi−4(X) and KO2i(X(C)) ∼= KO2i−8(X(C)). Proof. Recall from Section I.2d that the periodicity isomorphism in KO-theory is given by multiplication with a generator λ ∈ KO−8(point) whose sign is fixed by the relation α2 = 4λ for a generator α ∈ KO−4(point). The periodicity isomorphism for Grothendieck- Witt groups is induced by shifting complexes two positions to the right. This isomorphism may be interpreted as cross product with the element Λ := [O[−2], id] of GW−4(point), where O[−2] denotes the complex consisting of the trivial line bundle in degree−2. To show that gw−4 maps Λ to λ, we use the following square from the diagram above, evaluated on a point: K0(point) k  H−4 ,2 GW−4(point) gw−4  K0(point) r◦g4 ,2 KO−8(point) By Lemma I.1.9, we have H−4(O) = H−4(F (Λ)) = 2Λ, and similarly r(g4) = 2λ. Thus, gw−4 must map Λ to λ. We see similarly that the comparison maps gw i and w i, now defined for all shifts i ∈ Z, induce isomorphisms between the Grothendieck-Witt and KO-groups of a point. 1.3 Proposition. The comparison maps are isomorphism on a point: gw i : GWi(point) ∼=−−→ KO2i(point) w i : Wi(point) ∼=−−→ (KO2i/K)(point) 34 1 An elementary approach Proof. We can easily see that the corresponding groups of a point p are isomorphic by direct comparison: since K1(p) vanishes, the Bott sequence (I.20) implies that we can identify the quotients (KO2i/K)(p) with the odd KO-groups KO2i−1(p), so we can read off the values of all groups from Example I.1.7 and (I.17). Moreover, it is clear that the maps k, gw0 and w0 are isomorphisms. It follows by periodicity that gw i and w i are isomorphisms for all i divisible by four. For all other values of i, the groups Wi(p) are trivial, so w i is an isomorphism on a point in general. To see that the maps gw i are also isomorphisms for arbitrary i, we can use the comparison of the Karoubi and Bott sequences: . . . ,2 GWi−1(p) ,2 gw i−1  K0(p) ,2 ∼=  GWi(p) ,2 gw i  Wi(p) ,2 ∼=  0 ,2  . . . . . . ,2 KO2i−2(p) ,2 K0(p) ,2 KO2i(p) ,2 KO2i−1(p) ,2 0 ,2 . . . Given the periodicity of the Grothendieck-Witt groups, the claim concerning the maps gw i follows from repeated applications of the Five Lemma. This strategy of proof is known as “Karoubi induction”. As a next step, we would like to compare the localization sequence (I.7) of a smooth variety X with a smooth closed subvariety Z to the corresponding localization sequence (I.16) in KO-theory. Actually, this only makes sense for small parts of these sequences, and we have to compose the maps w i defined on Witt groups with multiplication by η ∈ KO−1(point): GWiZ(X) ,2 ?  GWi(X) ,2 gw i   GWi(U) ∂ ,2 gw i  Wi+1Z (X) ,2 ?  Wi+1(X) ,2 η◦w i+1   Wi+1(U) η◦w i+1  KO2iZ (X) ,2 KO 2i(X) ,2 KO2i(U) ∂ ,2 KO2i+1Z (X) ,2 KO 2i+1(X) ,2 KO2i+1(U) Here, we have written X for X(C) in the second row. The symbol “” marks the two squares that commute by naturality, and the question is how to fill in the dotted arrows such that all squares commute. Using Thom isomorphisms, we could replace the groups with support on Z by the actual cohomology groups of Z, and then we could use the comparison maps already defined for the dotted arrows. But we do not know a way of proving that the resulting squares are commutative without using comparison maps on groups with restricted support at some point. Unfortunately, we are similarly unaware of any straight-forward way of defining com- parison maps on groups GW0Z(X) with support Z 6= X. The solution in [Zib09] is to adapt a known construction of classes in relative K-groups detailed in [Seg68], in terms of complexes of vector bundles whose cohomology is supported on an open subspace. To construct a comparison map on GW0Z(X), we first choose a suitable open neighbourhood of 35 Chapter II. Comparison Maps Z in X and define a map to a KO-group of complexes with support on this neighbourhood. In a second step, we construct maps from such KO-groups defined in terms of complexes to the usual relative KO-groups. We do not know whether these maps are isomorphisms but, in any case, we obtain a comparison map gw0 : GW0Z(X)→ KO0Z(X(C)) by composition. Finally, we need to check that this map is independent of the choices made. Once the construction of comparison maps on GW0Z(X) is settled, maps on shifted groups GWiZ(X) can be defined in exactly the same way as before. Moreover, these constructions also work for groups twisted by line bundles. This allows us to check, again by explicit calculations, that the comparison maps are compatible with Nenashev’s Thom isomorphisms, and to conclude that two further squares in the above diagram commute. Lastly, we are left with the square involving the boundary maps. This square remains problematic. What is needed is either an algebraic description of the boundary morphism in KO-theory, or a homotopy-theoretic description of the boundary morphism for (Gro- thendieck-)Witt groups. The latter idea led us to consider a totally different approach to the comparison problem altogether, which we describe in the next section. 2 A homotopy-theoretic approach A homotopy theory of schemes emulating the situation for topological spaces has been developed over the past 20 years mainly by Morel and Voevodsky. Today, it is known as either A1-homotopy theory or motivic homotopy theory. The authoritative reference is [MV99]. Closely related texts by the same authors are [Voe98], [Mor99] and [Mor04]. A textbook introduction is provided by [DLØ+07]. Below, we summarize the main points relevant for us in just a few sentences. 2a A1-homotopy theory In short, the A1-homotopy category is constructed as follows. Consider the category Smk of smooth separated schemes of finite type over a field k. One of the problems with this category that arises if one tries to imitate topological arguments is that it does not have colimits: given two schemes X and Y with a common subscheme A, there is in general no smooth candidate for the union of X and Y “glued along A”. However, the category Smk can be embedded into some larger category Spck of “spaces over k” which is closed under small limits and colimits, and which can even be equipped with a model structure. The A1-homotopy category H(k) over k is the homotopy category associated with this model category. In fact, there are several possible choices for Spck and many possible model structures yielding the same homotopy category H(k). One possibility is to consider the category of simplicial presheaves over Smk, or the category of simplicial sheaves with respect to the Nisnevich topology. Both categories contain Smk as a full subcategory via the Yoneda 36 2 A homotopy-theoretic approach embedding, and they also contain simplicial sets viewed as constant (pre)sheaves. One may then apply a general recipe for equipping the category of simplicial (pre)sheaves over a site with a model structure (see [Jar87]). In a crucial last step, one forces the affine line A1 to become contractible by localizing with respect to the set of all projections A1×X  X. As in topology, we also have a pointed version H•(k) of H(k). Remarkably, these two categories contain several distinct “circles”: the simplicial circle S1, the “Tate circle” Gm = A1 − 0 (pointed at 1) and the projective line P1 (pointed at ∞). They are related by the intriguing formula P1 = S1 ∧Gm. A common notational convention which we will follow is to define Sp,q := (S1)∧(p−q) ∧G∧qm for any p ≥ q. In particular, we then have S1 = S1,0, Gm = S1,1 and P1 = S2,1. One can take the theory one step further by passing to the stable homotopy category SH(k), a triangulated category in which the suspension functors Sp,q ∧ − become inverti- ble. This category is usually constructed using P1-spectra. The triangulated shift functor is given by suspension with the simplicial sphere S1,0. Finally and crucially, the analogy with topology can be made precise: when we take our ground field k to be the complex numbers, or more general any subfield of C, we have a complex realization functor H(k)→ H (4) that sends a smooth schemeX to its set of complex pointsX(C) equipped with the analytic topology. There is also a pointed realization functor and, moreover, a triangulated functor of the stable homotopy categories SH(k)→ SH (5) which takes Σ∞(X+) to Σ∞(X(C)+) for any smooth scheme X [Rio06, The´ore`me I.123; Rio07a, The´ore`me 5.26]. 2b Representing algebraic and hermitian K-theory In the context of A1-homotopy theory, Theorem I.2.7 describing K0 in terms of homotopy classes of maps to Grassmannians has an algebraic analogue. Grassmannians of r-planes in kn+r can be constructed as smooth projective varieties over any field k. Viewing them as objects in Spck, we can form their colimits Grr and Gr in the same way as in topology. The following theorem is established in [MV99, § 4] and spelt out explicitly in [Rio06, The´ore`me III.3 and Assertion III.4]. 37 Chapter II. Comparison Maps 2.1 Theorem. For smooth schemes X over k we have natural identifications K0(X) ∼= H(k)(X,Z×Gr) (6) such that the inclusion Grr,n ↪→ {d} × Grr,n ↪→ Z × Gr corresponds to the class [Ur,n] + (d− r)[O] in K0(Grr,n). An analogous result for hermitian K-theory has recently been obtained by Schlichting and Tripathi1: Let Grndr,n denote the “non-degenerate Grassmannians” defined as open subvarieties of Grr,r+2n as above, and let Grndr and Gr nd be the respective colimits. Then for smooth schemes over k we have natural identifications GW0(X) ∼= H(k)(X,Z×Grnd) (7) It follows from the construction that, when (d − r) is even, the inclusion of Grndr,n ↪→ {d} ×Grndr,n ↪→ Z×Grnd corresponds to the class of [Undr,n] + d−r2 [H] in GW0(Grndr,n), where Undr,n is the universal symmetric bundle over Gr nd r,n. The fact that hermitian K-theory is representable in H(k) has been known for longer, see [Hor05]. One of the advantages of having a geometric description of a representing space, however, is that one can easily see what its complex realization is. In particular, this gives us an alternative way to define the comparison maps. For any smooth complex scheme X we have the following commutative squares, in which the left vertical arrows are the comparison maps (1) and (2), and the right vertical arrows are induced by the complex realization functor (4). K0(X) ∼=  H(C)(X,Z×Gr)  K0(X(C))∼=H(X(C),Z×Gr) GW0(X) ∼=  H(C)(X,Z×Grnd)  KO0(X(C))∼=H(X(C),Z×Grnd) Some of the results quoted here are in fact known in a much greater generality. Firstly, higher algebraic and hermitian K-groups of X are obtained by passing to suspensions of X in (6) and (7). Even better, algebraic and hermitian K-theory are representable in the stable A1-homotopy category SH(k). Let us make the statement a little more precise by fixing some notation. With any spectrum E in SH(k) one may associate a cohomology theory in exactly the same way as explained in Section I.2c, the only difference being that the theory obtained is now bigraded. Explicitly, a reduced cohomology theory E˜∗,∗ on 1Talk “Geometric representation of hermitian K-theory in A1-homotopy theory” at the Workshop “Geometric Aspects of Motivic Homotopy Theory”, 6.–10. September 2010 at the Hausdorff Center for Mathematics, Bonn 38 2 A homotopy-theoretic approach H•(k) and a corresponding unreduced theory E∗,∗ on H(k) are defined by setting E˜p,q(X) := SH(k)(Σ∞X, Sp,q ∧ E) for X ∈ H•(k) Ep,q(X) := E˜p,q(X+) for X ∈ H(k) A spectrum K representing algebraic K-theory was first constructed in [Voe98, § 6.2]; see [Rio06] or [Rio07b] for some further discussion. It is (2,1)-periodic, meaning that in SH(k) we have an isomorphism S2,1 ∧K ∼=→ K Thus, the bigrading of the corresponding cohomology theory Kp,q is slightly artificial. The identification with the usual notation for algebraic K-theory is given by Kp,q(X) = K2q−p(X) (8) For hermitian K-theory we have an (8,4)-periodic spectrum KO, and the corresponding cohomology groups KOp,q are honestly bigraded. The translation into the notation used for hermitian K-groups in Section I.1d is given by KOp,q(X) = GWq2q−p(X) (9) We will refer to the number 2q− p as the degree of the group KOp,q(X). The relation with Balmer’s Witt groups obtained by combining (9) and (I.10) is illustrated by the following table: KOp,q p = 0 1 2 3 4 5 6 7 q = 0 GW0 W1 W2 W3 W0 W1 W2 W3 q = 1 GW12 GW 1 1 GW 1 W2 W3 W0 W1 W2 q = 2 GW24 GW 2 3 GW 2 2 GW 2 1 GW 2 W3 W0 W1 q = 3 GW36 GW 3 5 GW 3 4 GW 3 3 GW 3 2 GW 3 1 GW 3 W0 As for the representing spaces in the unstable homotopy category, it is known that the complex realizations of KO and K represent real and complex topological K-theory. This is well-documented in the latter case, see for example [Rio06, Proposition VI.12]. However, for KO our references are slightly thin. Rather than shedding any further light on this problem, we will at this point succumb to an “axiomatic approach”. For clarity, we make a record of the precise properties of the spectra KO and K that we will be using. 2.2 Standing assumptions. There exist spectra K and KO in SH(C) representing alge- braic K-theory and hermitian K-theory in the sense described above, such that: (a) The complex realization functor (5) takes K to Ktop and KO to KOtop. 39 Chapter II. Comparison Maps (b) We have an exact triangle in SH(C) of the form KO ∧ S1,1 η ,2 KO t|ppp ppp K dl (10) which corresponds to the usual triangle in SH. These results are announced in [Mor06]. Independent constructions of spectra repre- senting hermitian K-theory can be found in [Hor05] and in a recent preprint of Panin and Walter [PW10]. 2c The comparison maps Given the assumptions above, the existence of comparison maps with good properties is immediate. In fact, by assumption (a), complex realization induces comparison maps in arbitrary bidegrees. Thus, we have comparison maps k˜p,q : K˜p,q(X)→ K˜p(X(C)) k˜p,qh : K˜O p,q(X)→ K˜Op(X(C)) for any pointed space X, and similarly comparison maps kp,q and kp,qh for any unpointed space, including in particular the case of a smooth complex scheme X. The maps we originally intended to study appear as special cases in degrees 0 and −1. They will be denoted by the same letters as before: k := k0,0 : K0(X)→ K0(X(C)) gw q := k2q,qh : GW q(X)→ KO2q(X(C)) w q := k2q−1,q−1h : W q(X)→ KO2q−1(X(C)) In the following, we will usually writeX forX(C) when this is unambiguous. The following properties of the comparison maps follow directly from the construction: • They commute with pullbacks along morphisms of smooth schemes. • They are compatible with suspension isomorphisms. • They are compatible with the periodicity isomorphisms, so that we may identify kp,qh with kp+8,q+4h (and hence w q with w q+4 and gw q with gw q+4). • They are compatible with long exact sequences arising from exact triangles in SH(C). The crucial advantage of the homotopy-theoretic approach is, of course, the last point. It applies in particular to the following two kinds of sequences. 40 2 A homotopy-theoretic approach Localization sequences. As in topology, any morphism of smooth schemes f : X ′ → X gives rise to long exact sequences relating the cohomology groups of X ′ to those of X. In particular, if X has a smooth closed subscheme Z with open complement U , we have an exact triangle of the form Σ∞(X − Z)+ ,2 Σ∞(X+) qxjjjj jj Σ∞(X/U) fm in SH(C), where X/U is obtained as the homotopy quotient of X by U in the category of spaces1. The induced long exact sequences of cohomology groups may be referred to as localization sequences. To clarify how the bigrading works, we write out one sequence explicitly in the example of hermitian K-theory: · · · → K˜Op,q(X/U)→ KOp,q(X)→ KOp,q(U)→ K˜Op+1,q(X/U)→ . . . (11) As a further analogy with topology, Morel and Voevodsky show that the space X/U only depends on the normal bundle N of Z in X. To make this more precise, we introduce the Thom space of a vector bundle E over an arbitrary smooth scheme Z, defined as the homotopy quotient of E by the complement of the zero section: ThomZ(E) := E / (E− Z) Using a geometric construction known as deformation to the normal bundle, Morel and Voevodsky show in Theorem 2.23 of [MV99, Chapter 3] thatX/U is canonically isomorphic to ThomZ(N) in the unstable pointed A1-homotopy category. The comparison maps induce commutative ladder diagrams of the following form: . . . ,2 K˜Op,q(ThomZ N) ,2  KOp,q(X) ,2  KOp,q(U) ,2  K˜Op+1,q(ThomZ N) ,2  . . . . . . ,2 K˜Op(ThomZ N) ,2 KOp(X) ,2 KOp(U) ,2 K˜Op+1(ThomZ N) ,2 . . . Of course, part of the statement here is that the complex realization of the Thom space of N over Z can be identified with the corresponding Thom space in topology, but since we are working in the stable category, there is no need to check this explicitly: the two spaces are isomorphic in SH because they fit into isomorphic exact triangles. To see that they are already isomorphic in the unstable homotopy category H•, we would need to be more careful and use a realization functor defined on the level of model categories. Such a functor is constructed in [Dug01]. 1In the model category of spaces used in [MV99], inclusions are cofibrations, so the homotopy quotient of X by U coincides with the actual quotient. 41 Chapter II. Comparison Maps Karoubi/Bott sequences. The Karoubi sequence arising from triangle (10) may be written explicitly as follows: . . .→ KOp−1,q(X)→ KOp−2,q−1(X)→ K2q−p(X)→ KOp,q(X)→ KOp−1,q−1(X)→ K2q−p−1(X)→ . . . By Assumption 2.2(b), the comparison maps induce a commutative ladder diagram with this sequence in the first row and the Bott sequence (I.20) in the second. Near degree zero, it takes the following familiar form: . . . ,2 KO2i−1,iX ,2 k2i−1,ih  GWi−1X ,2 gw i−1  K0X ,2 k  GWiX ,2 gw i  WiX ,2 w i  0 ,2  . . . . . . ,2 KO2i−1X ,2 KO2i−2X ,2 K0X ,2 KO2iX ,2 KO2i−1X ,2 K1X ,2 . . . (12) As a consequence, the comparison map w i factors through (KO2i/K)(X) η→ KO2i−1(X), so that we have an induced map w i : Wi(X)→ (KO2i/K)(X) We will often switch between these two versions of w i without further comment. Groups with restricted support. Comparing the localization sequences (I.7) and (I.9) discussed in Chapter I to the localization sequence (11) above, we see that the groups K˜Op,q(X/U) play the role of hermitian K-groups of X supported on Z. This should be viewed as part of any representability statement, see for example [PW10, Theorem 6.5]. Alternatively, a formal identification of the groups in degrees zero and below using only the minimal assumptions we have stated could be achieved as follows: 2.3 Lemma. Let Z be a smooth closed subvariety of a smooth variety X, with open complement U . We have the following isomorphisms: K˜O2q,q(X/U) ∼= GWqZ(X) K˜Op,q(X/U) ∼= Wp−qZ (X) for 2q − p < 0 Proof. Consider Z = Z×{0} as a subvariety of X×A1. Its open complement (X×A1)−Z contains X = X × {1} as a retract. Thus, the projection from X × A1 onto X induces a splitting of the localization sequences associated with (X × A1 − Z) ↪→ X × A1, and we have GWi+1Z (X × A1) ∼= coker ( GWi+11 (X × A1) ↪→ GWi+11 (X × A1 − Z) ) K˜O2i+2,i+1( X×A 1 X×A1−Z ) ∼= coker (KO2i+1,i+1(X × A1) ↪→ KO2i+1,i+1(X × A1 − Z)) By (9), we can identify the groups appearing on the right, so we obtain an induced 42 2 A homotopy-theoretic approach isomorphism of the cokernels. The quotient X ×A1/(X ×A1 − Z) can be identified with the suspension of X/U by S2,1, so we have an isomorphism K˜O2i+2,i+1( X×A 1 X×A1−Z ) ∼= K˜O2i,i(X/U) On the other hand, by Lemma 1.1, we have an isomorphism GWi+1Z (X × A1) ∼= GWiZ(X) The proof in lower degrees is analogous. Twisting by line bundles. In the homotopy-theoretic approach, we can define hermitian K-groups with twists in a line bundle, or more generally in any vector bundle, in the same way as we did for KO-theory in Section I.2d. 2.4 Definition. For a vector bundle E of constant rank r over a smooth scheme X, we define the hermitian K-groups of X with coefficients in E by KOp,q(X;E) := K˜Op+2r,q+r(ThomE) The twisted groups in degrees zero and below agree with the usual twisted Grothendieck- Witt and Witt groups: 2.5 Lemma. For a vector bundle E over a quasi-projective variety X, we have isomor- phisms KO2q,q(X;E) ∼= GWq(X; detE) KOp,q(X;E) ∼= Wp−q(X; detE) for 2q − p < 0 Proof. This follows from Lemma 2.3 and Nenashev’s Thom isomorphisms for Witt groups: for any vector bundle E of rank r there is a canonical Thom class in WrX(E) which induces an isomorphism Wi(X; detE) ∼= Wi+rX (E) by multiplication [Nen07, Theorem 2.5]. This Thom class actually comes from a class in GWrX(E). Using the Karoubi sequence (I.4), one sees that multiplication with this class also induces an isomorphism of the corresponding Grothendieck-Witt groups. Remark. The isomorphisms of Lemmas 2.3 and 2.5 are constructed here in a rather ad hoc fashion, and we have taken little care in recording their precise form. Whenever we give an argument concerning the comparison maps on “twisted groups” in the following, we do all constructions on the level of representable groups of Thom spaces. The identifications with the usual twisted groups are only needed to identify the final output of concrete calculations as in Section IV.3. 43 Chapter II. Comparison Maps 2d Comparison of the coefficient groups As before, the comparison maps gw i and w i in degrees 0 and −1 are isomorphisms on a point. Given that we now also have maps in arbitrary degrees at our disposal, it is natural to ask whether they, too, are isomorphisms. However, a comparison with the situation in K-theory shows that this is not what we should expect. Indeed, it is easy to find counterexamples: • In degrees −3 or less, the comparison map is necessarily zero. The problem is that while η : Wp−q(X)→Wp−q(X) is an isomorphism in all negative degrees, the topological η is nilpotent (η3 = 0). • In degree −2, we have KO0,−1(point) ∼= W1(point) mapping to KO0(point). The first group is zero while the second is isomorphic to Z. • In degree 1, it is known that KO−1,0(point) = Z/2 [Kar05, Example 18], from which we may deduce via the Karoubi sequence that KO1,1(point) ∼= C∗. In particular, KO1,1(point) cannot map isomorphically to KO1(point) = 0. We now show that the map in degree 1 is always surjective, while the map in degree −2 is always injective. 2.6 Proposition. For a complex point, the comparison maps in degrees 1, 0, −1 and −2 have the properties indicated by the following arrows. KO2q−1,q(point) KO2q−1(point) KO2q,q(point) ∼=→ KO2q(point) KO2q+1,q(point) ∼=→ KO2q+1(point) KO2q+2,q(point) KO2q+2(point) Proof. The isomorphisms in degree zero can be deduced exactly as in Proposition 1.3. To deal with the map in degree 1, we note that the odd KO-groups of a complex point p are all trivial except for KO−1(p), so k2q−1,qh is trivially a surjection unless q ≡ 0 mod 4. In that case, we claim that the surjectivity of k−1,0h is clear from the following diagram: . . . ,2 KO−1,0(p) ,2 k−1,0h  GW−1(p) ,2 ∼=  K0(p) ,2 ∼=  . . . . . . ,2 KO−1(p) ,2 KO−2(p) ,2 K0(p) ,2 . . . 44 2 A homotopy-theoretic approach Indeed, if we compute the groups appearing here, we obtain: . . . ,2 KO−1,0(p) // // k−1,0h  Z/2 0 ,2 ∼=  Z ,2 ∼=  . . . . . . ,2 Z/2 ∼= ,2 Z/2 0 ,2 Z ,2 . . . In degree −2, three out of four cases are again trivial as KO2q+2,q(p) = Wq+2(p) is zero unless q ≡ 2 mod 4. For the non-trivial case, consider the map η appearing in tri- angle (10). As the negative algebraic K-groups of p are zero, η yields automorphisms of Wp−q(p) in negative degrees. In topology, the corresponding maps are given by multipli- cation with a generator η of KO−1(p), and η2 generates KO−2(p). So the commutative square W0(p) ∼= ,2 ∼=  W0(p) k0,−2h  KO−1(p) η ∼= ,2 KO−2(p) shows that k0,−2h is an injection (in fact, an isomorphism), as claimed. 2e Comparison with Z/2-coefficients Another advantage of defining the comparison maps as above is that we can easily pass to cohomology groups with torsion coefficients. Although these will not lie within our main focus, it seems worthwhile to point out some simple observations that can be made in the case of Z/2-coefficients. In general, suppose E∗,∗ is some cohomology theory associated with a spectrum E in SH(k). A version of E∗,∗ with Z/2-coefficients can be defined in terms of the spectrum E/2 that fits into the following exact triangle: E 2 ,2 E t}qqq qqq E/2 aj Here, 2 denotes the twofold sum of the identity map on E. The triangle defines E/2 up to isomorphism. It follows from the definition that the associated cohomology groups E∗,∗(X;Z/2) fit into the following long exact sequences, which we refer to as Bockstein sequences: · · · → Ep,q(X) ·2−→ Ep,q(X)→ Ep,q(X;Z/2)→ Ep+1,q(X) ·2−→ · · · In particular, we can apply this definition to obtain algebraic and hermitian K-groups 45 Chapter II. Comparison Maps with Z/2-coefficients. Moreover, since the same definitions work in topology, we obtain comparison maps ki : Ki(X;Z/2)→ K−i(X;Z/2) kp,qh : KO p,q(X;Z/2)→ KOp(X;Z/2) These cohomology theories and the maps between them share all the formal properties of their integral counterparts. For example, the existence of Karoubi and Bott sequences with Z/2-coefficients may be deduced from the 3 × 3-Lemma in triangulated categories [May01, Lemma 2.6], and the comparison maps are again compatible with these. The behaviour of the comparison maps for K-theory with Z/2-coefficients was pre- dicted by the Quillen-Lichtenbaum conjectures: they are isomorphisms in all degrees i ≥ dim(X)− 1 and injective in degree dim(X)−2. Proofs may be found in [Lev99, Corol- lary 13.5 and Remark 13.2] and [Voe03b, Theorem 7.10]. In particular, on a complex point we have isomorphisms for all i ≥ 0. This implies the same statement for the hermitian comparison maps. Namely, it is not difficult to see that the comparison maps on Witt groups with Z/2-coefficients w i : Wi(X;Z/2)→ (KO2i/K)(X;Z/2) are also isomorphisms when X is a point. For example, this can be deduced from Propo- sition III.4.13 below, where we show more generally that these maps are isomorphisms whenever they are isomorphisms on the corresponding integral groups and the odd topo- logical K-groups of X contain no 2-torsion. Thus, the following result may be obtained via Karoubi-induction. 2.7 Corollary. The comparison maps kp,qh : KO p,q(point;Z/2)→ KOp(point;Z/2) are isomorphisms in all non-negative degrees, i. e. for all (p, q) with 2q − p ≥ 0. 46 Chapter III Curves and Surfaces In the previous chapter, we introduced comparison maps gw i : GWi(X)→ KO2i(X) w i : Wi(X)→ (KO2i/K)(X) for any smooth complex variety X and showed that they are isomorphisms when X is a point. Here, we analyse what happens when X has dimension one or two. Our final result, proved in Theorems 4.4 and 4.12, is the following: Theorem. If X is a smooth complex curve, the comparison maps gw i are surjective and the maps w i are isomorphisms. If X is a smooth complex surface, the same claim holds if and only if every continuous complex line bundle over X is algebraic, i. e. if and only if the natural map Pic(X)→ H2(X;Z) is surjective. Some more detailed statements concerning the situation for surfaces are given in Propo- sition 4.6. For projective surfaces, the condition on line bundles is equivalent to the con- dition that the geometric genus h2,0 of X is zero. Our proof of this result is largely computational. Namely, we explicitly compute both the Witt groups Wi(X) and the groups (KO2i/K)(X) of arbitrary smooth complex vari- eties of dimension at most two, from which we see that they are abstractly isomorphic in the cases claimed. In degree zero, we then deduce that the isomorphism is induced by the comparison map w0 from the fact that all elements in W0(X) are detected by the first two Stiefel-Whitney classes. The shifted groups W1(X) and W2(X) require more work. To obtain the assertions concerning the comparison maps, we decompose an arbitrary surface into a union of curves and an affine piece whose Picard group vanishes modulo two. The computation of the classical Witt group W0(X) of a complex curve or surface is not new. It was accomplished by Ferna´ndez-Carmena more than 20 years ago [FC87], and, at its core, his calculation is very similar to the one we present. However, we have found it convenient to rewrite the computation in terms of more general machinery developed over the past years by Balmer, Walter and Pardon. Not only does this make more transparent which steps become more difficult in higher dimensions, but also the values of the shifted Witt groups Wi(X) drop out of the calculation for free. It turns out that we may in fact work over any algebraically closed field of characteristic not two. 47 Chapter III. Curves and Surfaces The structure of this chapter is as follows. In the first section, we recall some facts about Stiefel-Whitney classes in algebraic geometry and topology, and explain why they coincide for complex varieties. In Section 2, we include a computation of the Grothendieck- Witt groups of smooth complex projective curves. The result is not needed to prove the above comparison theorem, but it completes the picture in this dimension. The main calculations of Witt groups and of the groups (KO2i/K)(X) are presented in Section 3, leading to a proof of the comparison theorem in Section 4. Finally, we briefly analyse how our result is related to the Quillen-Lichtenbaum conjecture and its analogue for hermitian K-theory, which has recently appeared in [Sch10]. 1 Stiefel-Whitney classes As we will see, all elements in the Witt groups W0(X) of a curve or surface can be detected by the rank homomorphism and the first two Stiefel-Whitney classes. We therefore include a brief account of the general theory of these characteristic classes in this section. In short, Stiefel-Whitney classes are invariants of symmetric bundles that take values in the e´tale cohomology groups ofX with Z/2-coefficients. First, we summarize a construction of these classes mimicking Grothendieck’s construction of Chern classes via projective bundles, and we state their main properties. We then specialize to the case of Stiefel-Whitney classes over fields, as studied by Milnor. The particular appeal of the first two classes stems from the fact that they can be used to identify certain graded pieces of the Witt group of a field with e´tale cohomology groups. The existence of similar identifications in higher degrees was one of the claims of the now famous Milnor conjectures. However, we sketch only the barest outlines of this story. Finally, we explain the relation of these classes over complex varieties to the classical Stiefel-Whitney classes used in topology. Stiefel-Whitney classes over varieties The following construction of Stiefel-Whitney classes due to Delzant [Del62] and Laborde [Lab76] is detailed in [EKV93, § 5]. It works for any scheme X over Z[12 ]. The first Stiefel-Whitney class w1 of a symmetric line bundle over X is defined by the correspondence of isometry classes of such bundles with elements in H1et(X;Z/2) (see Example I.1.1). For an arbitrary symmetric bundle (E, ε), one considers the scheme Pnd(E, ε) given by the complement of the quadric in P(E) defined by ε. The restriction of the universal line bundle O(−1) on P(E) to Pnd(E, ε) carries a canonical symmetric form. Let w be its first Stiefel-Whitney class in H1et(Pnd(E, ε);Z/2). Then the cohomology of Pnd(E, ε) can be decomposed as H∗et(Pnd(E, ε);Z/2) = r−1⊕ i=0 p∗H∗et(X;Z/2) · wi (1) 48 1 Stiefel-Whitney classes where p is the projection of Pnd(E, ε) onto X. In particular, wr is a linear combination of smaller powers of w, so that for certain coefficients wi(E, ε) ∈ H i(X;Z/2) we have wr + p∗w1(E, ε) · wr−1 + p∗w2(E, ε) · wr−2 + · · ·+ p∗wr(E, ε) = 0 These coefficients wi(E, ε) are defined to be the Stiefel-Whitney classes of (E, ε). They are characterized by the following axiomatic description [EKV93, § 1]: Normalization. The first Stiefel-Whitney class of a symmetric line bundle w1 is as defined above. Boundedness. wi(E, ε) = 0 for all i > rk(E) Naturality. For any morphism f : Y → X, we have f∗wi(E, ε) = wi(f∗(E, ε)). Whitney sum formula. The total Stiefel-Whitney class wt := 1 + ∑ i≥1wi(E, ε)t i satisfies wt((E, ε) ⊥ (F, ϕ)) = wt(E, ε) · wt(F, ϕ) The Stiefel-Whitney classes of a metabolic bundle only depend on the Chern classes of its Lagrangian. More precisely, Proposition 5.5 in [EKV93] gives the following formula for a metabolic bundle (M, µ) with Lagrangian L: wt(M, µ) = rk(L)∑ j=0 (1 + (−1)t)n−jcj(L)t2j (2) For example, for the first two Stiefel-Whitney classes we have w1(M, µ) = rk(L)(−1) (3) w2(M, µ) = ( rk(L) 2 ) (−1,−1) + c1(L) (4) It follows that wt descends to a well-defined homomorphism from the Grothendieck-Witt group of X to the multiplicative group of invertible elements in ⊕ iH i et(X;Z/2)ti: wt : GW0(X) −→ (⊕ i H iet(X;Z/2)ti )× In particular, the individual classes descend to well-defined maps wi : GW0(X)→ H iet(X;Z/2) 49 Chapter III. Curves and Surfaces In general, none of the individual Stiefel-Whitney classes apart from w1 define homo- morphisms on GW0(X). It does follow from the Whitney sum formula, however, that w2 restricts to a homomorphism on the kernel of w1, and in general wi restricts to a homomorphism on the kernel of (the restriction of) wi−1. It is not generally true either that the Stiefel-Whitney classes factor through the Witt group of X: the right-hand side of (2) may be non-zero. However, we may deduce from (3) that the restriction of the first Stiefel-Whitney class to the reduced group G˜W0(X) factors through W˜0(X), yielding a map w1 : W˜0(X) −→ H1et(X;Z/2) We use a different notation to emphasize that the values of w1 and w1 on a given symmetric bundle may differ. That is, if (E, ε) is a symmetric bundle of even rank defining an element Ψ in W˜0(X), then in general w1(Ψ) 6= w1(E, ε). Rather, w1 needs to be computed on a lift of (E, ε) to G˜W0(X). Explicitly, if H denotes the constant hyperbolic bundle (O⊕O, ( 0 11 0 )), then w1(Ψ) = w1 ( (E, ε)− rk(E)2 H ) = w1(E, ε)− rk(E)2 (−1) We see similarly from (4) that w2 induces a well-defined map w2 : W˜0(X) −→ H2et(X;Z/2) / Pic(X) As before, w1 is a surjective homomorphism, while w2 restricts to a homomorphism on ker(w1). Stiefel-Whitney classes over fields When X is a field F (of characteristic not 2), the Stiefel-Whitney classes factor through Milnor’s K-groups modulo two, which are commonly denoted kMi (F ). We will denote the classes with values in kMi (F ) by w M i : GW0(F ) wMi )K KKK KKK KKK wi ,2 H iet(F ;Z/2) kMi (F ) α 4=qqqqqqqqqq Both the groups kMi (F ) and the classes w M i were constructed in [Mil69]. In the same paper, Milnor asked whether the map α appearing in the factorization was an isomorphism, a question that later became known as one of the Milnor conjectures. For i ≤ 2, which 50 1 Stiefel-Whitney classes will be the range mainly relevant for us, an affirmative answer was given by Merkurjev [Mer81]. A general affirmation of the conjecture was found more recently by Voevodsky [Voe03b]. A second conjecture of Milnor, also to be found in [Mil69], concerned the relation of W0(F ) to kMi (F ). To state it, we introduce the fundamental filtration. If we view W 0(F ) as a ring, then the reduced Witt group W˜0(F ) becomes an ideal inside W0(F ), which is traditionally written as I(F ). The powers of this ideal yield a filtration W0(F ) ⊃ I(F ) ⊃ I2(F ) ⊃ I3(F ) ⊃ . . . on the Witt ring of F , known as the fundamental filtration. Milnor conjectured that the associated graded ring was isomorphic to kM∗ (F ) := ⊕ikMi (F ). As a first step towards a proof, he constructed maps kMi (F ) → Ii(F )/Ii+1(F ) in one direction. Moreover, in degrees one and two, Milnor could show that these are isomorphisms, with explicit inverses induced by the Stiefel-Whitney classes wM1 and w M 2 . In combination with the isomorphisms α above, one obtains the following identifications: rk : W 0(F ) / I(F ) ∼=−→ H0et(F ;Z/2) w1 : I(F ) / I2(F ) ∼=−→ H1et(F ;Z/2) w2 : I 2(F ) / I3(F ) ∼=−→ H2et(F ;Z/2) [Mil69, § 4]. Today, these isomorphisms are commonly denoted e0, e1, e2. It was clear from the outset, however, that the higher isomorphisms ei : I i(F ) / Ii+1(F ) ∼=−→ H iet(F ;Z/2) (5) conjectured by Milnor could not be induced by higher Stiefel-Whitney classes. Their existence was ultimately proved in [OVV07]. Unlike in the case of Stiefel-Whitney classes, it does not seem to be clear how these maps may be generalized to varieties. Stiefel-Whitney classes over complex varieties Here, we show that over a complex variety the e´tale Stiefel-Whitney classes of symmetric vector bundles are compatible with the Stiefel-Whitney classes of real vector bundles used in topology. Suppose first that Y is an arbitrary CW complex. If we follow the construction of Stiefel-Whitney classes described above, with singular cohomology in place of e´tale coho- mology, we obtain classes wi(E, ε) in H i(Y ;Z/2) for every complex symmetric bundle (E, ε) over Y . On the other hand, given a real vector bundle F over Y , we have the usual Stiefel-Whitney classes wi(F). We claim that these classes are compatible in the following 51 Chapter III. Curves and Surfaces sense. Recall from Corollary I.2.4 that we have a one-to-one correspondence < :  isometry classes ofcomplex symmetric bundles over Y →  isomorphism classesof real vector bundles over Y  1.1 Lemma. For any complex symmetric vector bundle (E, ε) over a CW complex Y , the classes wi(E, ε) and wi(<(E, ε)) in H i(Y ;Z/2) agree. Proof. For a complex symmetric line bundle (L, λ) over Y , the first Stiefel-Whitney classes of (L, λ) and <(L, λ) in H1(Y ;Z/2) agree because both bundles have the same associated principal Z/2-bundle. In general, the Stiefel-Whitney classes of a real vector bundle F may be defined in the same way as the classes wi(E, ε), using its real projectivization RP(F) in place of the space Pnd(E, ε) [Hus94, Theorem 2.5 and Definition 2.6]. Thus, the lemma follows from the considerations of Section I.2a. Indeed, the spaces Pnd(E, ε) are instances of the non-degenerate Grassmannians introduced there. In particular, for any complex symmetric bundle (E, ε), the space Pnd(E, ε) contains RP(<(E, ε)) as a homotopy equivalent subspace. Let us denote the inclusion and the bundle projections as in the following diagram: RP(<(E, ε)) p %% %%K KK KK KK KK K   j ,2 Pnd(E, ε) pnd zzzzvv vv vv vv v Y The universal symmetric bundle O(−1)nd over Pnd(E, ε) restricts to the universal real bundle O(−1) over RP(<(E, ε)) in the sense that <(j∗O(−1)nd) = O(−1). In particular, by the naturality of Stiefel-Whitney classes and the above observation concerning line bundles, we have j∗w1(O(−1)nd) = w1(O(−1)) Let r := rk(E). Then the Stiefel-Whitney classes of (E, ε) are defined by the following relations in the cohomology groups Hr(Pnd(E, ε);Z/2) and Hr(RP(<(E, ε);Z/2), respec- tively: w1(O(−1)nd)r = r∑ i=1 w1(O(−1)nd)r−ip∗ndwi(E, ε) w1(O(−1))r = r∑ i=1 w1(O(−1))r−ip∗wi(<(E, ε)) The claim follows by applying j∗ to the first line and comparing it with the second. Now let X be a complex variety. Then the above lemma implies that the e´tale and the topological Stiefel-Whitney classes are compatible. Moreover, if we specialize equation (2) for the Stiefel-Whitney classes of a metabolic bundle (M, µ) with Lagrangian L to the case 52 2 Curves of a complex variety, then since −1 is a square in C we find that w2i(M, µ) = ci(L) ∈ H2iet (X;Z/2) and all odd Stiefel-Whitney classes of (M, µ) vanish. This corresponds to the well-known fact in topology that the even Stiefel-Whitney classes of a complex vector bundle agree with its Chern classes modulo two, whereas its odd Stiefel-Whitney classes are zero [MS74, Problem 14.B]. It follows in particular that the odd Stiefel-Whitney classes factor through the (reduced) Witt group of X, while the even classes induce maps w2i : W˜0(X) −→ H 2i et (X;Z/2) im(ci) We summarize the situation as follows. 1.2 Proposition. Let X be a complex variety. The Stiefel-Whitney classes factor through the reduced Grothendieck-Witt and KO-group of X to yield commutative diagrams G˜W0(X)  wi ,2 H iet(X;Z/2) ∼=  K˜O0(X) wi ,2 H i(X;Z/2) Moreover, for all odd i we have induced maps W˜0(X)  wi ,2 H iet(X;Z/2) ∼=  (K˜O0/K˜)(X) wi ,2 H i(X;Z/2) and for even i we have induced maps W˜0(X)  wi ,2 H iet(X;Z/2) im(ci)  (K˜O0/K˜)(X) wi ,2 H i(X;Z/2) im(ci) 2 Curves In the next section, we will compute the Witt groups and the groups (KO2i/K)(X) for smooth varieties of dimension at most two. Here, we briefly summarize the results we will obtain in the case of curves. Moreover, we give a complete description of their 53 Chapter III. Curves and Surfaces Grothendieck-Witt groups. This description will not be used in the proofs of the main comparison theorems of this chapter but is included merely for the purposes of illustration and completeness. 2a Grothendieck-Witt groups of curves Let C be a smooth curve over an algebraically closed field k of characteristic not 2, and let Pic(C) be its Picard group. If C is projective, say of genus g, we may write Pic(C) as Z ⊕ Jac(C). The free summand Z is generated by a line bundle O(p) associated with a point p on C, while Jac(C) denotes (the closed points of) the Jacobian of C, a g-dimensional abelian variety parametrizing line bundles of degree zero over C. As a group, Jac(C) is two-divisible, and Jac(C)[2] has rank 2g (e. g. [Mil08, Chapter 14]). In particular, when C is projective, we have H1et(C;Z/2) = Pic(C)[2] = (Z/2)2g H2et(C;Z/2) = Pic(C)/2 = Z/2 If C is not projective, it is affine and may be obtained from a smooth projective curve by removing a finite number of points. Note that when we remove a single point p from a projective curve C, the Picard group is reduced to Pic(C − p) = Jac(C). It follows that the Picard group of any affine curve is two-divisible. In particular, for any affine curve C we have H2et(C;Z/2) = Pic(C)/2 = 0 The reduced algebraic K-group of a smooth curve may be identified with its Picard group via the first Chern class, so that we have an isomorphism K0(C) ∼= Z⊕ Pic(C) (6) The following proposition shows that, similarly, the Grothendieck-Witt and Witt groups of C are completely determined by Pic(C) and the group H1et(C;Z/2) of symmetric line bundles. 2.1 Proposition. Let C be a smooth curve over an algebraically closed field of character- istic not 2. The Grothendieck-Witt and Witt groups of C are as follows: GW0(C) = [Z]⊕H1et(C;Z/2)⊕H2et(C;Z/2) W0(C) = [Z/2]⊕H1et(C;Z/2) GW1(C) = Pic(C) W1(C) = H2et(C;Z/2) GW2(C) = [Z] W2(C) = 0 GW3(C) = [Z/2]⊕ Pic(C) W3(C) = 0 54 2 Curves Here, the summands in square brackets are the trivial ones coming from a point, i. e. those that disappear when passing to reduced groups. In particular, for a projective curve of genus g we obtain: GW0(C) = [Z]⊕ (Z/2)2g+1 W0(C) = [Z/2]⊕ (Z/2)2g GW1(C) = Z⊕ Jac(C) W1(C) = Z/2 GW2(C) = [Z] W2(C) = 0 GW3(C) = [Z/2]⊕ Z⊕ Jac(C) W3(C) = 0 For affine curves, no non-trivial twists are possible. When C is projective, the groups twisted by a generator O(p) of Pic(C)/2 are as follows: GW0(C,O(p)) = Z⊕ (Z/2)2g W0(C,O(p)) = (Z/2)2g GW1(C,O(p)) = Z⊕ Jac(C) W1(C,O(p)) = 0 GW2(C,O(p)) = Z W2(C,O(p)) = 0 GW3(C,O(p)) = Z⊕ Jac(C) W3(C,O(p)) = 0 2.2 Remark (Explicit generators). The isomorphism between W˜0(C) andH1et(X;Z/2) is the obvious one, i. e. W˜0(C) is generated by symmetric line bundles (see Example I.1.1). When C is projective of genus g, the group W1(C) has a single generator Ψg, which for g = 0 we may take to be the 1-symmetric complex given in Example I.1.6. When g ≥ 1, a generator may be constructed as follows: Pick two distinct points p and q on C. Let s and t be sections O→ O(p) and O→ O(q) of the associated line bundles that vanish at p and q, respectively. Choose a line bundle L ∈ Pic(C) that squares to O(p− q). Then Ψg :=  L⊗ O(−p) s ,2 t  L −t  L⊗ O(q − p) −s ,2 L⊗ O(q)  (7) is a 1-symmetric complex over C generating W1(C). Proof of Proposition 2.1, assuming Proposition 3.1: The values of the untwisted Witt groups may be read off directly from Proposition 3.1 below: we simply note that c1 : Pic(C) → H2et(C;Z/2) is surjective and induces an iso- morphism Pic(C)/2 ∼= H2et(C;Z/2). The twisted groups Wi(C,O(p)) of a projective curve C may then be calculated from the localization sequence (I.5) associated with the inclusion of the complement of a point p of C into C. Indeed, since the line bundle O(p) is trivial over C − p, this sequences takes the following form: · · · →Wi−1(p)→Wi(C,O(p))→Wi(C − p)→Wi(p)→ · · · 55 Chapter III. Curves and Surfaces We now deduce the values of the Grothendieck-Witt groups, concentrating on the case when C is projective. The affine case is very similar, except that all extension problems disappear. Untwisted case. To compute the untwisted Grothendieck-Witt groups of a projective curve C, we use the Karoubi sequences (I.4) and their restrictions to reduced groups: G˜Wi−1(C) F→ K˜0(C) H i→ G˜Wi(C) W˜i(C)→ 0 Note first that the identification of the K-group of our projective curve C given in (6) may be written explicitly as K0(C) ∼=←→ Z⊕ Z⊕ Jac(C) [E] 7→ (rkE, d,detE(−d)) [L(d)] + (r − 1)[O]←[ (r, d,L) Here, in the map from left to right, we have written d for the degree of the determinant line bundle of E. Given this description, the endomorphisms FH i of K0(C) may easily be computed directly. For the restrictions to the reduced K-group we obtain: im(FH i) = 0 ⊂ K˜0(C) when i is even2Z⊕ Jac(C) ⊂ K˜0(C) when i is odd (8) To compute G˜W0(C), we first observe that W3(C) = 0 implies that K˜0(C) surjects onto G˜W3(C). Therefore, the image of G˜W3(C) in K˜0(C) coincides with the image of FH3, and by (8) we obtain a short exact sequence 0→ Z/2 −→ G˜W0(C) −→ W˜0(C) =Jac(C)[2] → 0 This sequence splits. To define a splitting of the surjection, pick line bundles L of order 2 defining additive generators eL := (L, id)−(O, id) of W˜0(C). We claim that the homomor- phism obtained by lifting these to the corresponding elements in GW0(C) is well-defined, i. e. that 2eL vanishes in GW0(C). Indeed, by Lemma I.1.9 we have 2eL = H0(L − O), and H0 vanishes on L− O since by (8) the latter is contained in F (G˜W3(X)). Next, the image of G˜W0(C) in K˜0(C) is Jac(C)[2]. Since the quotient Jac(C)/ Jac(C)[2] is isomorphic to Jac(C), we obtain a short exact sequence 0→ Z⊕ Jac(C) H1−→ G˜W1(C) −→ W˜1(C) = Z/2 → 0 Let Ψg be the 1-symmetric complex given by (I.3) or (7). As an element of K˜0(C) = 56 2 Curves Z ⊕ Jac(C), its underlying complex is equivalent to (1,O). Thus, by Lemma I.1.9 again, we have H1(1,O) = H1(F (Ψg)) = 2Ψg. It follows that the sequence does not split and that Ψg descends to a generator of W˜1(C). Thus, G˜W1(C) = Z⊕ Jac(C). Carrying on to compute G˜W2(C), we first note that by (8) the image of G˜W1(C) in K˜0(C) contains 2Z ⊕ Jac(C). Moreover, F (Ψ) = ψ = (1,O), so in fact G˜W1(C) surjects onto K˜0(C). It follows that G˜W2(C) = 0. Finally, since W˜3(C) also vanishes, we see that H3 gives an isomorphism between K˜0(C) and G˜W3(C). This completes the computations in the untwisted case. Twisted case. To compute the Grothendieck-Witt groups of C twisted by a generator O(p) of Pic(C)/2 it again seems easiest to compare them with those of the affine curve C−p, over which O(p) is trivial. More specifically, we will compare the respective Karoubi sequences via the following commutative diagram: GWi−1(C − p) F ,2 K0(C − p) H i ,2 GWi(C − p) ,2Wi(C − p) ,2 0 GWi−1(C,O(p)) F ,2 LR K0(C) Hi O(p) ,2 LR GWi(C,O(p)) ,2 LR Wi(C,O(p)) ,2 LR 0 (9) The restriction map K0(C)→ K0(C−p) is the projection Z⊕Z⊕Jac(C) Z⊕0⊕Jac(C) killing the free summand corresponding to line bundles of non-zero degrees. The twisted version of (8) reads as follows: im(FH iO(p)) = Z · (2, 1)⊕ 0 ⊂ K0(C) when i is even0⊕ Z⊕ Jac(C) ⊂ K0(C) when i is odd (10) Now consider (9) with i = 0. Arguing as in the untwisted case, we may compute the cokernels of the forgetful maps F to reduce the diagram to a comparison of two short exact sequences: 0 ,2 Z ( 20 ) ,2 Z⊕ Jac(C)[2] ,2 (Z/2)⊕ Jac(C)[2] ,2 0 0 ,2 Z ,2 GW0(C,O(p)) ,2 LR Jac(C)[2] ,2 ( 01 ) LR 0 Since the outer vertical maps are injective, so is the central vertical map. It follows from a diagram chase that the image of a generator of K0(C) in GW0(C,O(p)) cannot be divisible by 2. Thus, the lower sequence must split, yielding GW0(C,O(p)) = Z⊕ Jac(C)[2]. Next, we see by (10) that the free summand in GW0(C,O(p)) has image Z ·(2, 1)⊕0 in K0(C). On Jac(C)[2], on the other hand, F must be the inclusion Jac(C)[2] ↪→ Jac(C), by comparison with the case of C−p. This gives GW1(C,O(p)) = Z⊕ (Jac(C)/ Jac(C)[2]) ∼= Z⊕ Jac(C). 57 Chapter III. Curves and Surfaces The image of GW1(C,O(p)) in K0(C) may be read off directly from (10), and we may deduce that GW2(C,O(p)) is a free abelian group generated by the twisted hyperbolic bundle H2O(p)(O). Equally directly, we see that GW 3(C,O(p)) = Z ⊕ Jac(C). This com- pletes the computation of all Grothendieck-Witt groups of C. 2b KO-groups of curves Now suppose that C is a smooth curve defined over C. Its topological KO-groups may be computed from the Atiyah-Hirzebruch spectral sequence. 2.3 Proposition. Let C be a smooth projective complex curve of genus g as above. Then its KO-groups are given by KO0(C) = [Z]⊕ (Z/2)2g+1 KO−1(C) = [Z/2]⊕ (Z/2)2g KO2(C) = Z KO1(C) = Z2g ⊕ Z/2 KO4(C) = [Z] KO3(C) = 0 KO6(C) = [Z/2]⊕ Z KO5(C) = Z2g Here, the square brackets again indicate which summands vanish when passing to reduced groups. In Section 3b, we will explain in detail how the values of the groups (KO2i/K)(C) may also be obtained from the spectral sequence. They may be read off as special cases from Proposition 3.8. In fact, we can also read off the values of the twisted groups (KO2i/K)(C;O(p)) by applying the same proposition to the Thom space ThomC(O(p)) (see Definition I.2.9). We then find that these groups agree with the corresponding Witt groups in Proposition 2.1: (KO0/K)(C) = [Z/2]⊕ (Z/2)2g (KO0/K)(C,O(p)) = (Z/2)2g (KO2/K)(C) = Z/2 (KO2/K)(C,O(p)) = 0 (KO4/K)(C) = 0 (KO4/K)(C,O(p)) = 0 (KO6/K)(C) = 0 (KO6/K)(C,O(p)) = 0 3 Surfaces This section contains the main calculations of this chapter: the computation of the Witt groups of smooth curves and surfaces on the one hand, and the computation of the cor- 58 3 Surfaces responding topological groups (KO2i/K) on the other. The comparison of the results will be postponed until the next section. 3a Witt groups of surfaces Consider the algebraic K-group K0(X) of a smooth variety X, and let ci be the Chern classes ci : K0(X) → CHi(X) with values in the Chow groups of X. Filter K0(X) by K0(X) ⊃ K˜0(X) ⊃ ker(c1). If X has dimension at most two, then the map (rk, c1, c2) induces an isomorphism [Ful98, Example 15.3.6]: gr∗(K0(X)) ∼= Z⊕ CH1(X)⊕ CH2(X) (11) A similar statement holds for the Witt group. 3.1 Proposition. Let k be an algebraically closed field of characteristic char(k) 6= 2, and let X be a smooth variety over k of dimension at most two. Filter W0(X) by W0(X) ⊃ W˜0(X) ⊃ ker(w1). Then the map (rk, w1, w2) gives an isomorphism gr∗(W0(X)) ∼= Z/2⊕H1et(X;Z/2)⊕ ( H2et(X;Z/2) / Pic(X) ) Moreover, if we write S1 for the squaring map CH1(X)/2→ CH2(X)/2, then the shifted Witt groups are as follows: W1(X) ∼= ker(S1)⊕H3et(X;Z/2) W2(X) ∼= coker(S1) W3(X) = 0 3.2 Example. Suppose X is a smooth complex projective surface of Picard number ρ. Write bi for the Betti numbers of X and ν for the rank of H2(X;Z)[2]. Then the above result shows that the Witt groups of X are as follows: W0(X) = [Z/2]⊕ (Z/2)b1+b2−ρ+2ν W1(X) = (Z/2)b1+ρ+2ν if S1 = 0(Z/2)b1+ρ+2ν−1 if S1 6= 0 W2(X) = Z/2 if S1 = 00 if S1 6= 0 W3(X) = 0 Our computation will follow a route described in [Tot03]. Namely, there is a sub- tle relationship between Witt groups and e´tale cohomology groups with Z/2-coefficients, 59 Chapter III. Curves and Surfaces encoded by three spectral sequences: Es,t2,BO(X) = H s(X,Ht) ⇒ Hs+tet (X;Z/2) (BO) Es,t2,Par(X) = H s(X,Ht) ⇒ Hs(X,W) (P) Es,t2,GW(X) = H s(X,Wt) ⇒ Ws+t(X) (GW) Here, H is the Zariski sheaf attached to the presheaf that sends an open subset U ⊂ X to Hjet(U ;Z/2). Similarly, W denotes the Zariski sheaf attached to the presheaf sending U to W0(U), and we set Wt = W for t ≡ 0 mod 40 otherwise Before specializing to the case of a surface, we give a short discussion of each of these spectral sequences for an arbitrary dimensional variety X. For simplicity, we assume that X is a smooth variety over a field of characteristic not 2 throughout. More general statements may be found in the references. The Bloch-Ogus spectral sequence (BO). The first spectral sequence is the well- known Bloch-Ogus spectral sequence [BO74]. This is a first quadrant spectral sequence with differentials in the usual directions, i. e. of bidegree (r,−r + 1) on the rth page. The E2-page is concentrated above the main diagonal s = t, and the groups along the diagonal may be identified with the Chow groups of X modulo two [BO74, Corollary 6.2 and proof of Theorem 7.7]: Hs(X,Hs) ∼= CHs(X)/2 (12) When the ground field is algebraically closed, the sheaves Ht vanish for all t > dim(X), so that the spectral sequence is concentrated in rows 0 ≤ t ≤ dim(X). · · · · · · · · · · · · H0(X,H3) /6fffffffff mmm H1(X,H3) /6fffffffff mmm H2(X,H3) /6fffffffff mmm CH3(X)/2 gggg mmm H0(X,H2) 07gggggggg 3;nnnnnnn H1(X,H2) 07gggggggg 3;nnnnnnn CH2(X)/2 07ggggggggg 3;nnnnnnn · H0(X,H1) 07gggggggg 3;ooooooo 6@ CH1(X)/2 07ggggggggg 3;ooooooo 6@ · · CH0(X)/2 d2ggggg 07ggg d3ooo 3;ooood4 7A · · · s ,2 t LR Figure 1: Pardon’s spectral sequence. The entries below the diagonal are zero. 60 3 Surfaces Pardon’s spectral sequence (P). Pardon’s spectral sequence can be indexed to have the same E2-page as the Bloch-Ogus spectral sequence. The differentials on the rth page then have bidegree (1, r − 1), as illustrated in Figure 1. In [Tot03], Totaro shows that, under the identifications (12), the differential d2 on the main diagonal of the E2-page cor- responds to the Steenrod operation S1 : CHs(X)/2→ CHs+1(X)/2, as defined by Brosnan and Voevodsky [Bro03,Voe03a]. The sequence converges to the cohomology of X with respect to W in the usual sense that the ith column of the E∞-page is isomorphic to the associated graded module of H i(X,W) with respect to some filtration. In order to describe the filtration onH0(X,W) = W(X), we briefly summarize how the spectral sequence arises. Let X be as above, and let K be its function field. We denote the residue field of a scheme-theoretic point x of X by k(x). In [Par04], Pardon shows that the sheaf W has a flasque resolution by a Gersten-Witt complex W which on open subsets U ⊂ X takes the form W(U) : W0(K)→ ∐ x∈U(1) W0(k(x))→ · · · → ∐ x∈U(n) W0(k(x))→ 0 (13) Here, U (i) denotes the set of codimension i points of X contained in U , and W0(k(x)) is to be viewed as a constant sheaf supported on the closure of x in X. In particular, W is a subsheaf of the constant sheaf W0(K). Let I(K) ⊂W0(K) be the fundamental ideal of K (see Section 1). The fundamental filtration on W0(K), given by the powers of I(K), induces a filtration on W which we denote by It :=W ∩ It(K) (14) Pardon shows more generally that the fundamental filtrations of the Witt groups W0(k(x)) give rise to flasque resolutions of the sheaves It of the form It(U) : It(K)→ ∐ x∈U(1) It−1(k(x))→ ∐ x∈U(2) It−2(k(x))→ . . . Applying the standard construction of the spectral sequence of a filtered complex of abelian groups [GM03, Chapter III.7, Section 5] to the filtration of W(X) by It(X), we obtain a spectral sequence Hs(X, It/It+1)⇒ Hs(X,W) To conclude, Pardon uses the affirmation of the Milnor conjectures (see (5)) to identify the sheaves It/It+1 with the sheaves Ht. The following lemma is a direct consequence of this construction. 61 Chapter III. Curves and Surfaces 3.3 Lemma. The filtration on W(X) appearing on the zeroth column of the E∞-page of Pardon’s spectral sequence is given by the global sections of the sheaves It defined in (14): E0,t∞,Par(X) ∼= It(X)/It+1(X) The edge homomorphism including this column into the zeroth column of the E2-page is induced by the sheafification map from the presheaf quotient of It by It+1 to the quotient sheaf It/It+1. The group W(X), viewed as a subgroup of W0(K), is commonly referred to as the unramified Witt group of X. The Gersten-Witt spectral sequence (GW). The Gersten-Witt spectral sequence of Balmer and Walter, third on the list above, may be thought of as an analogue of the Bloch-Ogus spectral sequence for Witt groups. In [BW02], Balmer and Walter, too, construct a Gersten-Witt complex of the form (13). Although it seems likely that their construction agrees with Pardon’s, this does not yet seem to have been established. It is known in any case, however, that their complex also provides a flasque resolution of W [BGPW02, Lemma 4.2 and Theorem 6.1]. The spectral sequence constructed in [BW02] may therefore be rewritten in the form above: 3.4 Theorem. [BW02] Let X be a smooth variety over a field of characteristic not 2. There is a spectral sequence with E2-page Es,t2 = Hs(W) in rows t ≡ 0 mod 40 elsewhere The sequence has differentials dr of bidegree (r,−r + 1) and converges to Ws+t(X). Proof of Proposition 3.1: We now specialize to the case when dim(X) ≤ 2 and the ground field is algebraically closed. Then the Bloch-Ogus spectral sequence (BO) collapses immediately and we obtain descriptions of the groups Hs(X;Ht) in terms of the e´tale cohomology of X with Z/2- coefficients. The E2-page of Pardon’s spectral sequence therefore takes the following form: H2et(X;Z/2) / Pic(X) H 3 et(X;Z/2) H4et(X;Z/2)∼= CH2(X)/2 H1et(X;Z/2) d2gggg 07gggg Pic(X)/2 S1hhhh 07hhh · Z/2 · · s ,2 t LR 62 3 Surfaces On the other hand, we see that the Gersten-Witt spectral sequence also collapses, so that Wi(X) ∼= H i(X,W). Thus, the ith column of Pardon’s spectral sequence simply converges to Wi(X). In particular, if we write It(X) for the filtration on W0(X) corresponding to the filtration It(X) ofW(X) defined in (14), then by Lemma 3.3 we have E 0,t ∞,Par ∼= grIt W0(X). 3.5 Lemma. Let X be as above. Then, for t = 1 or 2, the edge homomorphisms grIt W 0(X) ∼= E0,t∞,Par ↪→ E0,t2,Par can be identified with the Stiefel-Whitney classes wt. More precisely, I1(X) = W˜0(X), I2(X) = ker(w1), and we have commutative diagrams E0,1∞,Par   ,2 E0,12,Par W˜0(X) / ker(w1) w1 ,2 H1et(X;Z/2) E0,2∞,Par   ,2 E0,22,Par ker(w1) w2 ,2 H2et(X;Z/2) / Pic(X) Proof. The statement of the lemma is not surprising: the only non-obvious maps that enter into the construction of Pardon’s spectral sequence are the isomorphisms et displayed in (5); for t = 1 or 2, these can be identified with Stiefel-Whitney classes, and the latter can be defined globally. In more detail, the various identifications arising from the Bloch-Ogus spectral sequence and Pardon’s spectral sequence fit into the following diagram: E0,t∞,Par   ,2 E0,t2,Par It(X) / It+1(X) ,2 ?  ( It / It+1 ) (X)   ,2 ∼=  It(K) / It+1(K) ∼=et  Htet(X;Z/2) sheafification ,2  Ht(X)   ,2 Htet(K;Z/2) E0,t∞,BO ∼= (BO collapses) ,2 E0,t2,BO The claim is that, for t = 1 or 2, the diagram commutes if we take the dotted map to be the tth Stiefel-Whitney class. Indeed, et can be identified with wt in these cases, and the horizontal compositions across the two middle lines in the diagram are given by pullback along the inclusion of the generic point into X. 63 Chapter III. Curves and Surfaces There are two potentially non-zero differentials in Pardon’s spectral sequence: the differential S1 : CH1(X)/2 → CH2(X)/2, which is simply the squaring operation, and another differential d from H1et(X;Z/2) to H3et(X;Z/2). Since the first Stiefel-Whitney class always surjects onto H1et(X;Z/2), we see from the previous lemma that H1et(X;Z/2) survives to the E∞-page. So the differential d must vanish. On the other hand, since E0,22 = E 0,2∞ , the lemma implies that the restriction of the second Stiefel-Whitney class to the kernel of the first gives an isomorphism with H2et(X;Z/2)/Pic(X). This proves the first statement of the proposition. Since S1 is the only non-trivial differential left, the values of the shifted Witt groups also follow easily from the spectral sequence. This completes the proof of Proposition 3.1. In the following corollary, the Chern and Stiefel-Whitney classes refer to the corresponding maps on reduced groups: ci : K˜0(X) −→ CHi(X) wi : G˜W0(X) −→ H iet(X;Z/2) 3.6 Corollary. Let X be as above. Then the second Stiefel-Whitney class is surjective and restricts to an epimorphism w2 : ker(w1) − H2et(X;Z/2) Its kernel is given by the image of ker(c1) under the hyperbolic map H0 : K˜0(X)→ G˜W0(X). Proof. Note first that H0 : K˜0(X)→ G˜W0(X) factors through the kernel of w1. In fact, we can restrict the Karoubi sequence to an exact sequence forming the first row of the following diagram. K˜0(X) H0 ,2 c1  ker(w1) // // w2  ker(w1) ,2 w2∼=  0 Pic(X) ,2 H2et(X;Z/2) // // H 2 et(X;Z/2) / Pic(X) ,2 0 The lower row is obtained from the Kummer sequence, hence also exact, and the diagram commutes. We know from the previous proposition that the map w2 on the right is an isomorphism. Since c1 is surjective, surjectivity of w2 follows. It remains to show that the restriction H0 : ker(c1)→ ker(w2) is surjective. By an instance of the Snake Lemma, this is equivalent to showing that the kernel of H0 on K˜0(X) maps surjectively to Pic(X)2 under c1. The kernel of H0 is given by the image of G˜W3(C) under the forgetful map F , which coincides with the image of K˜0(X) under the composition FH3. Thus, it suffices to show that c1 restricts to a surjection c1 : FH3(K0(X))→ Pic(X)2. This is done in the next lemma. 64 3 Surfaces 3.7 Lemma. Let X be as above. Then for odd i we have commutative diagrams K0(X) c1  Hi ,2 GWi(X) F ,2 K0(X) c1  Pic(X) ,2 Pic(X) L 7→ L2 For even i, the diagram commutes if we replace the map on Pic(X) by the constant map. Proof. This can be checked via a direct calculation. The first Chern class c1 takes a vector bundle E to its determinant line bundle detE. Under FH i, the class of E is mapped to the class E⊕ (−1)iE∨ in K0(X), which c1 then takes to (detE)⊗(1−(−1)i). Remark. More generally, the endomorphism of gr∗(K0(X)) ∼= Z ⊕ Pic(X) ⊕ CH2(X) induced by FH i is given by (2, 0, 2) if i is even, and by (0, 2, 0) if i is odd. 3b KO/K-groups of surfaces On a topological space X, let us write Sq2Z for the composition H2(X;Z)/2 ↪→ H2(X;Z/2) Sq 2 −→ H4(X;Z/2) where Sq2 is the squaring operation. 3.8 Proposition. Let X be a connected CW complex of dimension at most four. Fil- ter the group (KO0/K)(X) by (KO0/K)(X) ⊃ (K˜O0/K˜)(X) ⊃ ker(w1). Then the map (rk, w1, w2) induces an isomorphism gr ( (KO0/K)(X) ) ∼= Z/2⊕H1(X;Z/2)⊕ (H2(X;Z/2)/H2(X;Z)) The remaining groups (KO2i/K)(X) have the following values: (KO2/K)(X) ∼= ker(Sq2Z)⊕H3(X;Z/2) (KO4/K)(X) ∼= coker(Sq2Z) (KO6/K)(X) = 0 3.9 Example. Suppose X is a compact four-dimensional manifold. Write bi for its Betti numbers and ν for the rank of H2(X;Z)[2]. Then the above result shows that 65 Chapter III. Curves and Surfaces (KO0/K)(X) = [Z/2]⊕ (Z/2)b1+2ν (KO2/K)(X) = (Z/2)b1+b2+2ν if Sq2Z = 0(Z/2)b1+b2+2ν−1 if Sq2Z 6= 0 (KO4/K)(X) = Z/2 if Sq2Z = 00 if Sq2Z 6= 0 (KO6/K)(X) = 0 3.10 Remark. Note that both cases — Sq2Z = 0 and Sq 2 Z onto — can occur even for complex projective surfaces of geometric genus zero. For example, Sq2Z is onto for P2. On the other hand, the Wu formula [MS74, Theorem 11.14] shows that Sq2 is given by multiplication with c1(X) mod 2. So Sq2Z = 0 on any projective surface X of geometric genus zero whose canonical divisor is numerically trivial. Concretely, we could take X to be an Enriques surface (a quotient of a K3-surface by a fixed-point free involution): see [Bea96, page 90]. To calculate the groups (KO2i/K)(X), we use the Atiyah-Hirzebruch spectral sequences for the K- and KO-theory of X. We begin with some general facts. 3.11 Lemma. The differential d1,−23 : E 1,−2 3 −→ E4,−43 is trivial both in the Atiyah-Hirze- bruch spectral sequence for K-theory and in the spectral sequence for KO-theory. Proof. This is immediate from the descriptions of these differentials in terms of the second Steenrod square given in Section I.2e. 3.12 Lemma. Let X be a connected finite-dimensional CW complex. Denote the filtra- tions on the groups K0(X) and KOi(X) associated with the Atiyah-Hirzebruch spectral sequences by K0(X) ⊃ K1(X) = K2(X) ⊃ K3(X) = K4(X) ⊃ . . . KOi0(X) ⊃ KOi1(X) ⊃ KOi2(X) ⊃ KOi3(X) ⊃ . . . The initial layers of these filtrations have more intrinsic descriptions in terms of Chern and Stiefel-Whitney classes: • K2(X) = K1(X) = K˜0(X), and KOi1(X) = K˜Oi(X) • The map K2(X) → H2(X;Z) arising from the spectral sequence may be identified with the first Chern class, at least up to a sign. That is, we have the following commutative diagram: E2,−2∞   ,2 E2,−22 K˜0(X) OOOO ±c1 // // H2(X;Z) In particular, K4(X) = ker(c1). Moreover, since c1 is surjective, E 2,−2∞ = E2,−22 . 66 3 Surfaces • The map K4(X) → H4(X;Z) can be identified with the restriction of the second Chern class to ker(c1), again up to a sign: E4,−4∞   ,2 E4,−42 ker(c1) OOOO ±c2 ,2 H4(X;Z) Thus, K6(X) = ker(c2). (Note that the upper horizontal map in this diagram is indeed an inclusion, by the previous lemma.) • Likewise, the map KO01(X) → H1(X;Z/2) can be identified with the first Stiefel- Whitney class: E1,−1∞   ,2 E1,−12 K˜O0(X) OOOO w1 // // H1(X;Z/2) Thus, KO02(X) = ker(w1) and E 1,−1∞ = E1,−12 . • Lastly, the map KO02(X)→ H2(X;Z/2) can be identified with the restriction of the second Stiefel-Whitney class, i. e. we have a commutative diagram E2,−2∞   ,2 E2,−22 ker(w1) OOOO w2 ,2 H2(X;Z/2) Thus, KO04(X) = ker(w2). Proof. The first statement is clear. The other assertions follow by viewing the maps in question as cohomology operations and computing them for a few spaces. For lack of reference, we include a more detailed proof of the statements concerning KO-theory. The case of complex K-theory can be dealt with analogously. First, we analyse the map K˜O0(X)→ H1(X;Z/2) arising from the spectral sequence. A priori, we have defined this map only for finite-dimensional CW complexes. But we can extend it to a natural transformation of functors on the homotopy category of all connected CW complexes, using the fact that the canonical map H1(X;Z/2)→ limiH1(Xi;Z/2) is an isomorphism for any CW complex X with i-skeletons Xi. On the homotopy category of connected CW complexes, the functor K˜O0(−) is represented by BO. Natural transfor- mations K˜O0(−)→ H1(−;Z/2) are therefore in one-to-one correspondence with elements of H1(BO;Z/2) = Z/2 · w1, where w1 is the first Stiefel-Whitney class of the universal bundle over BO. Thus, either the map in question is zero, or it is given by w1 as claimed. Since it is non-zero on S1, the first case may be discarded. To analyse the map KO02(X) → H2(X;Z/2), we note that the previous conclusion yields a functorial description of KO02(X) as the kernel of w1 on K˜O 0(X). Moreover, given 67 Chapter III. Curves and Surfaces this description, we may define a natural set-wise splitting of the inclusion of KO02(X) into K˜O0(X) for any finite-dimensional CW complex X as follows: K˜O0(X) KO02(X) [E]− rk(E)[R] 7→ [E]− rk(E)[R]− [detE] + [R] The composition K˜O0(−) → H2(−;Z/2) can be extended to a natural transformation of functors on the homotopy category of connected CW complexes in the same way as before, and it may thus be identified with an element of H2(BO;Z/2) = Z/2 · w21 ⊕ Z/2 · w2. Consequently, its restriction to KO02(X) is either 0 and w2. Again, the first possibility may easily be discarded, for example by considering S2. Lemma 3.12 has some immediate implications for low-dimensional spaces. In the fol- lowing corollary, all characteristic classes refer to the corresponding maps on reduced groups: ci : K˜0(X) −→ H2i(X;Z) wi : K˜O0(X) −→ H i(X;Z/2) 3.13 Corollary. Let X be a connected CW complex of dimension at most four. Then the second Chern class is surjective and restricts to an isomorphism c2 : ker(c1) ∼=−→ H4(X;Z) The second Stiefel-Whitney class is also surjective, and restricts to an epimorphism w2 : ker(w1) − H2(X;Z/2) Its kernel is given by the image of ker(c1) under realification. Moreover, the induced map w2 on (K˜O0/K˜)(X) restricts to an isomorphism w2 : ker(w1) ∼=−→ ( H2(X;Z/2) / H2(X;Z) ) Proof. Consider the Atiyah-Hirzebruch spectral sequences for K- and KO-theory. Lemma 3.11 shows that the sequence for K-theory collapses. Moreover, in the spectral sequence for KO-theory, no differentials affect the diagonal computing KO0(X). This implies the first two claims of the corollary. Next, we compare the two spectral sequences via realification. The description of the kernel of w2 may be obtained from the following row-exact commutative diagram: 68 3 Surfaces 0 ,2 H4(X;Z) ,2 ∼=  K˜(X) c1 ,2 r  H2(X;Z) ,2 reduction mod 2  0 0 ,2 H4(X;Z) ,2 ker(w1) w2 ,2 H2(X;Z/2) ,2 0 The final claim of the corollary also follows from this diagram, by identifying ker(w1) with ker(w1)/K˜(X). The situation for a connected CW complex of dimension at most four may now be summarized as follows. Firstly, by Lemma 3.12, the filtrations on the groups K0(X) and KO0(X) arising in the Atiyah-Hirzebruch spectral sequences can be written as K0(X) ⊃ K˜0(X) ⊃ ker(c1) KO0(X) ⊃ K˜O0(X) ⊃ ker(w1) Secondly, Corollary 3.13 implies that the maps (rk, c1, c2) and (rk, w1, w2) on the associated graded groups induce isomorphisms gr(K0(X)) ∼= Z ⊕H2(X;Z) ⊕ H4(X;Z) (15) gr ( (KO0/K)(X) ) ∼= Z/2⊕H1(X;Z/2)⊕ (H2(X;Z/2)/H2(X;Z)) (16) In particular, we have proved the first part of Proposition 3.8. Proof of the remaining claims of Proposition 3.8: To compute (KO2i/K)(X) for general i, we identify this quotient with the image of η : KO2i(X)→ KO2i−1(X) This image can be computed at each stage of the filtration KOj(X) = KOj0(X) ⊃ KOj1(X) ⊃ · · · ⊃ KOj4(X) To lighten the notation, we simply write KOjk for KO j k(X) in the following, and we write H∗(X) for the singular cohomology of X with integral coefficients. Since we are assuming that X is at most four-dimensional, there are only three possibly non-zero differentials in the spectral sequence computing KO∗(X). The first two are the differentials Sq2 ◦ pi : H2(X) −→ H4(X;Z/2) Sq2 : H2(X;Z/2)→ H4(X;Z/2) on the E2-page, depicted in Figure I.1. If E 4,−2 2 = H 4(X;Z/2) is not killed by Sq2, then 69 Chapter III. Curves and Surfaces we have a third possibly non-trivial differential on the E3-page: d1,03 : H 1(X)→ coker(Sq2) The differential d1,−23 vanishes by Lemma 3.11. Computation of KO2/K. For the last stage of the filtration, we have a commutative diagram 0 ,2 KO24 ,2  KO23 ,2 η  H3(X; KO−1(point)) ,2 ∼=  0 0 ,2 0 ,2 KO13 H 3(X; KO−2(point)) ,2 0 Thus, KO23 maps surjectively to KO 1 3 ∼= H3(X;Z/2). Next, we have 0 ,2 KO23 ,2 η  KO22 ,2 η  ker(Sq2 ◦ pi) ,2 pi  0 0 ,2 KO13 ,2 KO 1 2 ,2 ker(Sq2) ,2 0 where ker(Sq2 ◦pi) ⊂ H2(X; KO0(point)) and ker(Sq2) ⊂ H2(X; KO−1(point)). This gives a short exact sequence of images, which must split since all groups involved are killed by multiplication with 2. Thus, we have im ( KO22 η→ KO12 ) ∼= KO13⊕ im( ker(Sq2 ◦ pi)→ ker(Sq2)) ∼= H3(X;Z/2)⊕ ker(Sq2Z) where Sq2Z is the composition defined at the beginning of this section. Finally, the diagram 0 ,2 KO22 η  K˜O2(X) ,2 η  0 ,2  0 0 ,2 KO12 ,2 K˜O 1(X) ,2 E1,04 ,2 0 shows that im(K˜O2(X) η→ K˜O1(X)) ∼= im(KO22 η→ KO12). So (K˜O2/K˜)(X) is isomorphic to H3(X;Z/2)⊕ ker(Sq2Z), as claimed. Computation of KO4/K. Proceeding as in the previous case, we consider the diagram 0 ,2 H4(X)  K˜O4(X) ,2 η  0 ,2  0 0 ,2 KO34 ,2 K˜O 3(X) ,2 H3(X) ,2 0 70 4 Comparison where the vertical map on the left is the composition H4(X; KO0(point)) H4(X; KO−1(point)) coker(Sq2 ◦ pi) = KO34 Since Sq2 ◦ pi has the same cokernel as Sq2Z, we may deduce that (K˜O4/K˜)(X) ∼= im(K˜O4(X) η→ K˜O3(X)) ∼= coker(Sq2Z) Computation of KO6/K. In this case, the commutative diagram 0 ,2 H2(X)  K˜O6(X) ,2 η  0 ,2  0 0 ,2 0 ,2 K˜O5(X) H1(X) ,2 0 demonstrates that the map K˜O6(X) η→ K˜O5(X) is zero, so (K˜O6/K)(X) is trivial. This completes the computations of (K˜O2i/K)(X). 4 Comparison In this section, we finally compare our two sets of results for complex curves and surfaces and prove the comparison theorem mentioned in the introduction to this chapter. First, however, we summarize the situation one finds in K-theory. K-groups The algebraic and topological K-groups of a smooth complex curve C are given by K0(C) = Z⊕ Pic(C) K0(C) = Z⊕H2(C;Z) The comparison map K0(C) → K0(C) is always surjective. Indeed, the map is an iso- morphism on the first summand, and for a projective curve the map on reduced groups is simply the projection from Pic(C) ∼= Z⊕ Jac(C) onto the free part. It follows that the map is still surjective if we remove a finite number of points from C. For a smooth complex surface X, we have seen in (11) and (15) that we have filtrations on the K-groups such that gr∗(K0(X)) ∼= Z⊕ Pic(X)⊕ CH2(X) gr∗(K0(X)) ∼= Z⊕H2(X;Z)⊕H4(X;Z) 71 Chapter III. Curves and Surfaces Both isomorphisms can be written as (rk, c1, c2), and the comparison map K0(X) → K0(X) corresponds to the usual comparison maps on the filtration. Of course, on the first summand we again have the identity. Moreover, the map CH2(X)→ H4(X;Z) is always surjective: If X is projective, this follows from the fact that H4(X;Z) is generated by a point. In general, we can embed any smooth surface X into a projective surface X as an open subset with complement a divisor with simple normal crossings (c. f. Lemma 4.8). Then CH2(X) surjects onto CH2(X), and similarly H4(X;Z) surjects onto H4(X;Z), so that the claim follows. Since we will need this observation in a moment, we record it as a lemma. 4.1 Lemma. For any smooth complex surface X, the natural map CH2(X)→ H4(X;Z) is surjective. 4.2 Corollary. For any smooth complex surface X, the natural map K0(X)→ K0(X) is surjective if and only if the natural map Pic(X)→ H2(X;Z) is surjective. When X is projective, we see from the exponential sequence that we have surjections if and only if X has geometric genus zero. Equivalently, this happens if and only if X has full Picard rank, i. e. if and only if its Picard number ρ agrees with its second Betti number b2. More generally, the Picard group of any smooth complex surface can be written as Pic(X) = Zρ ⊕H2(X;Z)tors ⊕ Pic0(X) (17) where ρ ≤ b2 is an integer that generalizes the Picard number, H2(X;Z)tors is the torsion subgroup of H2(X;Z), and Pic0(X) is a divisible group [PW01, Corollary 6.2.1]. Again, the natural map Pic(X)→ H2(X;Z) is surjective if and only if ρ = b2. 4a The classical Witt group For the classical Witt group W0(X), the situation can be analysed in a similar way as in the case of K0(X), using our description in terms of Stiefel-Whitney classes. 4.3 Proposition. For a smooth complex curve C, the map gw0 : GW0(C) KO0(C) is surjective, and w0 : W0(C) ∼=→ (KO0/K)(C) is an isomorphism. Similarly, for a smooth complex surface X, both gw0 and w0 are surjective. If X is projective, then w0 is an isomorphism if and only if Pic(X) surjects onto H2(X;Z). Proof. It suffices to show the corresponding statements for reduced groups. Let X be a smooth complex variety of dimension at most two. Since the first Stiefel-Whitney classes 72 4 Comparison are always surjective, we have a row exact commutative diagram of the form 0 ,2 ker(w1) ,2  G˜W0(X) w1 ,2 gw0  H1et(X;Z/2) ,2 ∼=  0 0 ,2 ker(wtop1 ) ,2 K˜O 0(X) wtop1 ,2 H1(X;Z/2) ,2 0 Similarly, by Corollaries 3.6 and 3.13 we have a row exact commutative diagram 0 ,2 H0(ker(c1)) ,2  ker(w1)  w2 ,2 H2et(X;Z/2) ,2 ∼=  0 0 ,2 r(ker(ctop1 )) ,2 ker(w top 1 ) wtop2 ,2 H2(X;Z/2) ,2 0 In both diagrams, we have written wtopi and c top i for the topological characteristic classes to avoid ambiguous notation. The kernels of c1 and c top 1 can be identified with CH 2(X) and H4(X;Z), respectively, via the second Chern classes. Thus, Lemma 4.1 implies that the vertical map on the left of the lower diagram is a surjection. The surjectivity of the comparison map gw0 on GW0(X) follows. If we apply the same analysis to W0(X), then the second diagram reduces to a com- mutative square ker(w1)  w2 ∼= ,2 H2et(X;Z/2) / Pic(X)  ker(wtop1 ) wtop2 ∼= ,2 H2(X;Z/2) / H2(X;Z) When X is a curve, the vertical arrow on the right is an isomorphism, proving the claim. In general, we have a row-exact commutative diagram of the following form: 0 ,2 Pic(X) / 2 ,2   H2et(X;Z/2) ,2 ∼=  H2et(X;Z/2) / Pic(X) ,2  0 0 ,2 H 2(X;Z) / 2 ,2 H 2(X;Z/2) ,2 H2(X;Z/2) / H2(X;Z) ,2 0 If Pic(X) → H2(X;Z) is surjective, then the two outer maps become isomorphisms and it follows that the comparison map w0 is also an isomorphism. Using the description of the Picard group of X given by (17), we see that the converse is also true. 4b Shifted Witt groups We now generalize Proposition 4.3 to shifted groups. The final result is stated in Theo- rems 4.4 and 4.12. For curves, there is in fact very little left to be shown. We nevertheless 73 Chapter III. Curves and Surfaces give a detailed proof in preparation for a similar line of arguments in the case of surfaces. 4.4 Theorem. For any smooth complex curve C, the maps gw i : GWi(C)→ KOi(C) are surjective, and the maps w i : Wi(C)→ (KOi/K)(C) are isomorphisms. Proof. By Lemma II.2.6, the comparison maps appearing here are isomorphisms on a point, so the claims are equivalent to the corresponding claims involving reduced groups. It suffices to show that the maps gw i : G˜Wi(C) → K˜Oi(C) are surjective and that the maps w i : W˜i(C)→ K˜O2i−1(C) are injective. First, suppose C is affine. Then the cohomology of C is concentrated in degrees 0 and 1 and we see that W1(C), W2(C), W3(C) and K˜O2(X), K˜O4(C), K˜O6(C) all vanish. Thus, the claims are trivially true. The case that C is projective can be reduced to the affine case. Indeed, if p is any point on C, then C˜ := C − p is affine. By comparing the localization sequences arising from the inclusion of C˜ into C, we see that the comparison maps for C must also have the desired properties: . . . ,2 GWi−1(p) ,2 ∼=  G˜Wi(C) ,2  G˜Wi(C˜) ,2  Wi(p) ,2 ∼=  W˜i+1(C) ,2  W˜i+1(C˜) ,2   . . . . . . ,2 KO2i−2(p) ,2 K˜O2i(C) ,2 K˜O2i(C˜) ,2 KO2i−1(p) ,2 K˜O2i+1(C) ,2 K˜O2i+1(C˜) ,2 . . . (18) 4.5 Corollary. Theorem 4.4 also hold for groups with twists in any line bundle. Proof. Introducing a twist by the line bundle O(p) into the localization sequence (18) only affects the groups of C, so we can conclude as before. More generally, given a line bundle O(D) associated with a divisor D = ∑ nipi on C, we can similarly reduce to the case of a trivial line bundle over C −⋃i pi. We now want to imitate this proof for surfaces, replacing the role of points on the curve by curves on the surface. We first prove the following. 4.6 Proposition. For any smooth complex surface X, the comparison maps have the properties indicated by the following arrows:gw0 : GW0(X) KO0(X)w1 : W1(X) KO1(X) gw2 : GW2(X) KO4(X)w3 : W3(X) KO5(X) 4.7 Lemma. Proposition 4.6 is true when X is affine and Pic(X)/2 vanishes. 74 4 Comparison Proof. By the theorem of Andreotti and Frankel, a smooth complex affine variety of dimension n has the homotopy type of a CW complex of real dimension at most n [AF59,Laz04, 3.1]. In particular, its cohomology is concentrated in degrees ≤ n . For a smooth affine surface X, Pardon’s spectral sequence shows immediately that W1(X) = W2(X) = W3(X) = 0. Similarly, the Atiyah-Hirzebruch spectral sequence shows that K˜O4(X) vanishes. Thus, three of the four claims are trivially satisfied. The fact that gw0 : GW0(X) KO0(X) is surjective was already shown in 4.3. Before proceeding with the proof of Proposition 4.6, we make a note of two general facts that we will use. First, we will need the following standard consequence of Hironaka’s resolution of singularities: 4.8 Lemma. In characteristic zero, any smooth variety can be embedded into a smooth compact variety with complement a divisor with simple normal crossings. Proof. Let X be a smooth variety over a field of characteristic zero, and let X be some compactification. Any singularities of X may be resolved without changing the smooth locus [Kol07, Theorem 3.36], so we may assume that X is smooth. Applying the Princi- palization Theorem [Kol07, Theorem 3.26] to (the ideal sheaf of) the complement of X in X yields the variety we are looking for. Note that in dimension two there is no distinction between compactness and projectiv- ity: any smooth compact surface is projective [Har70, II.4.2]. It follows that an arbitrary smooth surface is at least quasi-projective. Secondly, we will need the following lemma concerning generators of the Picard group. 4.9 Lemma. Let X be a smooth quasi-projective variety over an algebraically closed field of characteristic zero. Then any element of Pic(X)/2 can be represented by a smooth prime divisor (i. e. by a smooth irreducible subvariety of codimension 1). If X is projective, we may moreover take the divisor to be very ample (i. e. to be given by a hyperplane section of X for some embedding of X into some PN ). Proof. We consider the case when X is projective first. If X is a projective curve, Pic(X)/2 ∼= Z/2 is generated by a point and there is nothing to show. So we may assume dim(X) ≥ 2. Fix a very ample line bundle L on X. Any element of Pic(X)/2 can be lifted to a line bundle M on X. Tensoring with any sufficiently high power of L will yield a very ample line bundleM⊗Lm [Laz04, 1.2.10]. In particular, if we take m large and even we obtain a very ample line bundle that maps to the class of M in Pic(X)/2. The claim now follows from Bertini’s theorem on hyperplane sections in characteristic zero [Har77, Corollary 10.9 and Remark 10.9.1]. 75 Chapter III. Curves and Surfaces In general, if X is quasi-projective, we may embed it as an open subset into a smooth resolution X of its projective closure. Then Pic(X) surjects onto Pic(X), and we obtain smooth prime divisors on X generating Pic(X)/2 by restriction. 4.10 Example. Consider P˜2, the blow-up of P2 at a point p. Its Picard group is given by Pic(P˜2) = Z[H]⊕ Z[E], where H is a hyperplane section of P2 that misses p and E is the exceptional divisor, both isomorphic to P1. A divisor a[H]− b[E] is ample if and only if a > b > 0. Thus, {[H], 2[H] − [E]} is a basis of Pic(P˜2) consisting of ample divisors. A smooth curve representing 2[H]− [E] is given by the birational transform of a smooth conic in P2 through p. Proof of Proposition 4.6. LetX be a smooth surface. By Lemma 4.8, we can find a projec- tive surface X and smooth curves D1, . . . , Dk on X whose union ⋃ Di is the complement of X in X. On the other hand, by Lemma 4.9, we can find smooth ample curves generating Pic(X)/2. Let C1, . . . , Cρ be a subset of these curves generating Pic(X)/2, and put Ui := X − C1 − C2 − · · · − Ci For sufficiently large k, the divisors ∑ j Dj + k(C1+ · · ·+Ci) are ample on X. Thus, each Ui is affine. Moreover, we see from the exact sequences Z[Ci+1|Ui ]→ Pic(Ui)→ Pic(Ui+1)→ 0 and the choice of the Ci that rkZ/2(Pic(Ui)/2) = ρ−i. Thus, by Lemma 4.7, Proposition 4.6 holds for Uρ. We can now proceed as in the proof of Theorem 4.4, by adding the curves back in to obtain X. Namely, consider the successive open inclusions Ui+1 ↪→ Ui. The closed complements of these are given by restrictions of the curves Ci, so we obtain a diagram similar to (18) with Ui playing the role of C, Ui+1 in the role of C˜ and Ci+1 in the role of a point. If the normal bundle of Ci+1 in Ui is not trivial, the sequences will in fact involve twisted groups of Ci, but in any case we can conclude using Lemma 4.5. 4.11 Corollary. Proposition 4.6 also holds for groups with twists in a line bundle over X. Proof. As we have seen, generators of Pic(X)/2 can be represented by smooth curves on X. Thus, we can argue as in the proof of Corollary 4.5. 4.12 Theorem. Suppose X is a smooth projective surface for which the natural map Pic(X)→ H2(X;Z) is surjective. Then the comparison maps gw i : GWi(X)→ KOi(X) are surjective, and the maps w i : Wi(X)→ (KO2i/K)(X) are isomorphisms. 76 4 Comparison By Proposition 4.3, the assumption on the map Pic(X)→ H2(X;Z) is clearly necessary. Proof. Consider the squaring operations S1 and Sq2Z appearing in the computations of the Witt and (KO/K)-groups. For any smooth complex variety X, we have a commutative diagram Pic(X)/2 S 1 ,2  _  CH2(X)/2  H2(X;Z)/2 ,2 _  H4(X;Z)/2 _  H2(X;Z/2) Sq 2 ,2 H4(X;Z/2) When X is a surface, the two vertical maps on the right are both isomorphisms, and the horizontal map in the middle is essentially Sq2Z. So whenever Pic(X) surjects onto H2(X;Z), we may identify S1 and Sq2Z. In particular, we may do so for any a smooth pro- jective surface of geometric genus zero. It then follows by comparison of Propositions 3.1 and 3.8 that the Witt groups of such a surface agree with the groups (KO2i/K)(X). On the other hand, we see from Proposition 4.6 that each of the maps w i : Wi(X)→ (KO2i/K)(X) is either surjective or injective. Given that we are dealing with finite groups, these maps must be isomorphisms. Moreover, since we know from Corollary 4.2 that we also have a surjection from the algebraic to the topological K-group of X, we may deduce via the Karoubi/Bott sequences that the maps gw i : GWi(X)→ KO2i(X) are surjective for all values of i. 4c Comparison with Z/2-coefficients As we have seen, the comparison maps on Witt groups are isomorphisms for all surfaces X for which Pic(X) surjects onto H2(X;Z). To obtain isomorphisms on the level of Grothendieck-Witt groups, we need to pass to Z/2-coefficients (see Section II.2e) and introduce one additional topological constraint on X. 4.13 Proposition. Let X be a smooth complex variety, of any dimension. If the odd topological K-groups of X contain no 2-torsion (i. e. if K1(X)[2] = 0), then the integral comparison maps Wi(X)→ (KO2i/K)(X) are isomorphisms for all i if and only if the comparison maps with Z/2-coefficients Wi(X;Z/2)→ (KO2i/K)(X;Z/2) are isomorphisms for all i. Before giving the proof, we make one preliminary observation. 77 Chapter III. Curves and Surfaces 4.14 Lemma. Let X be a topological space such that K1(X)[2] = 0. Then multiplication by η ∈ KO−1(point) induces an isomorphism (KO2i/K)(X) η−→∼= KO 2i−1(X)[2] Proof. In general, since 2η = 0, the following exact sequence may be extracted from the Bott sequence: 0→ (KO2i/K)(X) η−→ KO2i−1(X)[2]→ K1(X)[2] This proves the claim. Proof of Proposition 4.13. We claim that we have the following row-exact commutative diagram: 0 ,2 GW i(X) K0(X) ,2  GWi(X;Z/2) K0(X;Z/2) ,2  Wi+1(X) ,2  0 0 ,2 KO 2i(X) K0(X) ,2 KO 2i(X;Z/2) K0(X;Z/2) ,2 KO2i+1(X)[2] ,2 0 (∗) Indeed, the lower exact row may be obtained by applying the Snake Lemma to the following diagram of short exact sequences induced by the Bockstein sequences for K- and KO- theory: 0 ,2 K0(X)/2 ∼= ,2  K0(X;Z/2) ,2  0  0 ,2 KO2i(X)/2 ,2 KO2i(X;Z/2) ,2 KO2i+1(X)[2] ,2 0 The upper row of (∗) may be obtained similarly from the Bockstein sequences for algebraic and hermitian K-theory. The vertical maps are induced by the comparison maps in degrees 0, 0 and −1, respectively. Using the canonical identification of GWi(X)/K0(X) with Wi(X), we may however identify the first vertical map with the usual comparison map w i in degree −1. Likewise, the second vertical map may be identified with the comparison map w i for Witt groups with Z/2-coefficients. Lastly, by the previous lemma, the entry in the lower right corner may be identified with (KO2i+2/K)(X). Thus, diagram (∗) can be rewritten in a form from which both implications of the proposition may be deduced: 0 ,2Wi(X) ,2 w i  Wi(X;Z/2) ,2 w i  Wi+1(X) ,2 w i+1  0 0 ,2 KO2i(X) K(X) ,2 KO2i(X;Z/2) K(X;Z/2) ,2 KO2i+2(X) K(X) ,2 0 78 4 Comparison Recall from II.2e that, by the Quillen-Lichtenbaum conjectures, the comparison maps on K-groups of curves with Z/2-coefficients are isomorphisms in all non-negative degrees, while for a smooth complex surface X we have isomorphisms in all positive degrees and an inclusion in degree zero: Ki(X;Z/2) ∼=−−→ K−i(X;Z/2) for all i > 0 K0(X;Z/2) ↪−→ K0(X;Z/2) For these low-dimensional cases, proofs may also be found in [Sus95, 4.7; PW01, Theo- rem 2.2]. 4.15 Lemma. Let X be a smooth complex variety of dimension at most two. The map K0(X;Z/2) ↪→ K0(X;Z/2) is an isomorphism if and only if the map Pic(X)→ H2(X;Z) is surjective and K1(X)[2] = 0. Proof. We see from the description of the Picard group (17) that Pic(X) → H2(X;Z) is surjective if and only if it is surjective after tensoring with Z/2. Moreover, by a similar argument as in Corollary 4.2, the surjectivity of the map Pic(X)/2 → H2(X;Z)/2 is equivalent to the surjectivity of the map K0(X)/2 → K0(X)/2. The claim follows from these equivalences and a commutative diagram induced by the Bockstein sequences: 0 ,2 K0(X)/2 ∼= ,2  K0(X;Z/2) ,2  0  0 ,2 K0(X)/2 ,2 K0(X;Z/2) ,2 K1(X)[2] ,2 0 If we combine the results on K-theory with our results for Witt groups and Proposi- tion 4.13, we obtain the following corollary via Karoubi induction. 4.16 Corollary. Let X be a smooth complex variety of dimension at most two. Assume that the natural map Pic(X)→ H2(X;Z) is surjective and that K1(X) has no 2-torsion. Then the hermitian comparison maps KOp,q(X;Z/2) −→ KOp(X;Z/2) are isomorphisms in all non-negative degrees, i. e. for all (p, q) such that 2q − p ≥ 0. In particular, for all shifts i we have isomorphisms GWi(X;Z/2) ∼=−→ KO2i(X;Z/2) For example, the conditions of the corollary are satisfied for any smooth complex curve, 79 Chapter III. Curves and Surfaces and for any simply-connected projective surface of geometric genus zero. The condition that K1(X)[2] = 0 can be rephrased in terms of the integral cohomology group H3(X;Z): 4.17 Lemma. Let X be a smooth complex variety of dimension at most two. Then the torsion of K1(X) agrees with the torsion of H3(X;Z). Proof. By Lemma 3.11, the Atiyah-Hirzebruch spectral sequence for the K-theory of X collapses. Since the first integral cohomology group H1(X;Z) of a smooth complex curve or surface is free, we find that K1(X) ∼= H1(X;Z)⊕H3(X;Z), and the lemma follows. Remark. Conversely, if we assume the analogue of the Quillen-Lichtenbaum conjecture for hermitian K-theory, i. e. if we assume that the hermitian comparison maps with Z/2- coefficients are isomorphisms in high degrees, then we can recover our comparison theorem for Witt groups for all surfaces X with K1(X)[2] = 0. Said analogue appeared recently in [Sch10]. However, it does not seem possible to relate our result to the Quillen-Lichtenbaum conjecture for surfaces with 2-torsion in K1(X). Such surfaces do exist. In particular, if X is an Enriques surface, then Pic(X) surjects onto H2(X;Z) but K1(X)[2] ∼= pi1(X) ∼= Z/2 [Bea96, page 90]. 80 Chapter IV Cellular Varieties In this chapter, we turn our attention to cellular varieties. As we will see, the comparison maps behave particularly well on these, allowing us to identify all Grothendieck-Witt and Witt groups with the corresponding KO-groups. This will be used to compute the Witt groups of a series of projective homogeneous varieties. By definition, a smooth cellular variety is a smooth variety X with a filtration by closed subvarieties ∅ = Z0 ⊂ Z1 ⊂ Z2 ⊂ · · · ⊂ ZN = X such that the complement of Zk in Zk+1 is an open “cell” isomorphic to Ank for some nk. For example, projective n-space Pn contains an open cell isomorphic to An with closed complement Pn−1, so such a filtration can be obtained inductively. It is well-known that, when X is a smooth cellular variety over the complex numbers, the comparison map on K-groups is an isomorphism: K0(X) ∼=−→ K0(X) In fact, both sides are easy to compute. They decompose as direct sums of the K-groups of the cells, each of which is isomorphic to Z. However, the computation of Witt groups of cellular varieties does not seem to be as straight-forward. It is true, of course, that the Witt groups of complex varieties decompose into copies of Z/2, the Witt group of C, but even in the cellular case there is no general understanding of how many copies to expect. Nonetheless, it is possible to prove by an induction over the number of cells of X that the comparison maps on Grothendieck-Witt and Witt groups are also isomorphisms: GWi(X) ∼=−→ KO2i(X) Wi(X) ∼=−→ (KO2i/K)(X) The map (KO2i/K)(X) η→ KO2i−1(X) is an isomorphism in this situation (see Lemma 2.2), so the second line may equivalently be stated as Wi(X) ∼=−→ KO2i−1(X) Thus, the Grothendieck-Witt and Witt groups of smooth complex cellular varieties may be read off directly from their KO-groups. As in the previous chapter, we use the homotopy-theoretic approach to the comparison 81 Chapter IV. Cellular Varieties maps. As a by-product of the proof we give, we also obtain partial information on the maps in degrees 1 and −2. A slightly different proof using only the maps in degrees 0 and −1 and those properties that can be seen in a more elementary way may be found in [Zib09, Theorem 3.1]. The argument is sketched in Remark 1.5. After presenting our proof of the comparison result at the beginning of this chapter, we turn to concrete topological computations. As our main tool will be the Atiyah-Hirzebruch spectral sequence, we collect some general observations concerning its behaviour on cellular varieties in Section 2. We then run through the list of all projective homogeneous varieties that fall within the class of hermitian symmetric spaces. The untwisted KO-groups of these are known by several papers of Kono and Hara, and we extend their computations to twisted KO-groups. Combined with the comparison result, this furnishes us with a complete additive description of the Witt groups of these varieties. 1 The comparison theorem 1.1 Theorem. For a smooth cellular complex variety X, the following comparison maps are isomorphisms: k : K0(X) ∼=−→ K0(X) gw q : GWq(X) ∼=−→ KO2q(X) w q : Wq(X) ∼=−→ KO2q−1(X) More generally, the maps gw q and w q are isomorphisms for the corresponding groups twisted by an arbitrary vector bundle over X (see Section II.2c). In the proof, we concentrate on the hermitian case. The case of algebraic/complex K-theory could be dealt with similarly, or deduced from the hermitian case using trian- gle (II.10). As indicated in the introduction, it will be helpful to consider not only the maps gw q = k2q,qh and w q+1 = k2q+1,qh in degrees 0 and −1, respectively, but also the maps k2q−1,qh in degree 1 and the maps k 2q+2,q h in degree −2. In fact, we will prove the following extended version of the theorem, generalizing Proposition II.2.6: 1.2 Theorem. For X as above, the hermitian comparison maps in degrees 1, 0, −1 and −2 have the properties indicated: KO2q−1,q(X) KO2q−1(X) KO2q,q(X) ∼=→ KO2q(X) KO2q+1,q(X) ∼=→ KO2q+1(X) KO2q+2,q(X) KO2q+2(X) The analogous statements for twisted groups are also true. 82 1 The comparison theorem Proof. The proof will proceed by induction over the number of cells of X. To begin the induction, we need to consider the case of only one cell, which immediately reduces to the case of a point by homotopy invariance. This case was dealt with in Proposition II.2.6. Spheres. Since the theorem holds for a point, the compatibility of the comparison maps with suspensions immediately shows that it is also true for the reduced cohomology of the spheres (P1)∧d = S2d,d. In other words, the following maps in degrees 1, 0, −1 and −2 have the properties indicated: K˜O2q−1,q(S2d,d) K˜O2q−1(S2d) K˜O2q,q(S2d,d) ∼=→ K˜O2q(S2d) K˜O2q+1,q(S2d,d) ∼=→ K˜O2q+1(S2d) K˜O2q+2,q(S2d,d) K˜O2q+2(S2d) Cellular varieties. Now let X be a smooth cellular variety. By definition, X has a filtration by closed subvarieties ∅ = Z0 ⊂ Z1 ⊂ Z2 · · · ⊂ ZN = X such that the open complement of Zk in Zk+1 is isomorphic to Ank for some nk. In general, the subvarieties Zk will not be smooth. Their complements Uk := X − Zk inX, however, are always smooth as they are open in X. So we obtain an alternative filtration X = U0 ⊃ U1 ⊃ U2 · · · ⊃ UN = ∅ of X by smooth open subvarieties Uk. Each Uk contains a closed cell Ck ∼= Ank with open complement Uk+1. Our inductive hypothesis is that we have already proved the theorem for Uk+1, and we now want to prove it for Uk. We can use the following exact triangle in SH(C): Σ∞((Uk+1)+) ,2 Σ∞((Uk)+) pwggggg gggg Σ∞Thom(NCk\Uk) gn As Ck is a cell, the Quillen-Suslin theorem tells us that the normal bundle NCk\Uk of Ck in Uk has to be trivial. Thus, Thom(NCk\Uk) is A 1-weakly equivalent to S2d,d, where d is the codimension of Ck in Uk. Figure 1 on page 85 displays the comparison between the long exact cohomology sequences induced by this triangle. The inductive step is completed by applying the Five Lemma to each dotted map in the diagram. The twisted case. To obtain the theorem in the case of coefficients in a vector bundle E over X, we replace the exact triangle above by the triangle Σ∞Thom(E|Uk+1) ,2 Σ ∞Thom(E|Uk) ovffffff fffff Σ∞Thom(E|Ck ⊕NCk\Uk) ho The existence of this exact triangle is shown in the next lemma. The Thom space on the right is again just a sphere, so we can proceed as in the untwisted case. 83 Chapter IV. Cellular Varieties 1.3 Lemma. Given a smooth subvariety Z of a smooth variety X with complement U , and given any vector bundle E over X, we have an exact triangle Σ∞Thom(E|U ) ,2 Σ∞ThomE pwggggg ggggg gg Σ∞Thom(E|Z ⊕NZ\X) ho Proof. From the Thom isomorphism theorem we know that the Thom space of a vector bundle over a smooth base is A1-weakly equivalent to the quotient of the vector bundle by the complement of the zero section. Consider the closed embeddings U ↪→ (E − Z), X ↪→ E and Z ↪→ E. Computing the normal bundles, we obtain (E− Z)/(E−X) ∼= ThomU (E|U ) E/(E−X) ∼= ThomX E E/(E− Z) ∼= ThomZ(E|Z ⊕NZ\X) The claim follows by passing to the stable homotopy category and applying the octahedral axiom to the composition of the embeddings (E−X) ⊆(E− Z) ⊆ E. 1.4 Remark (Comparison in degree −2). As we saw in Section II.2d, the comparison map in degrees 1 and −2 are not isomorphisms even when X is a point, and in degrees below −2 the comparison maps are necessarily zero. For cellular varieties, the map in degree −2 may be identified with the inclusion of the 2-torsion subgroup of KO2q+2(X) into KO2q+2(X). This follows from the theorem and the description of the KO-groups of cellular varieties given in Lemma 2.2. 1.5 Remark. We indicate briefly how Theorem 1.1 can alternatively be obtained by working only with the maps in degrees 0 and −1 defined by more elementary means. The basic strategy — comparing the localization sequences arising from the inclusion of a closed cell Ck into the union of “higher” cells Uk — still works. But we cannot deduce that the comparison maps are isomorphisms on Uk from the fact that they are isomorphisms on Uk+1 because the parts of the sequences that we can actually compare are now too short. We can, however, still deduce that the maps in degree 0 with domains the Grothendieck- Witt groups of Uk are surjective, and that the maps in degree −1 with domains the Witt groups of Uk are injective. The inductive step can then be completed with the help of the Bott/Karoubi sequences. This argument works even without assuming that the comparison maps are compatible with the boundary maps in localization sequences in general: in the relevant cases the cohomology groups involved are so simple that this property can be checked by hand. 84 1 The comparison theorem  ···  K˜O2q−1,q(S2d,d)  // // K˜O2q−1(S2d)  KO2q−1,q(Uk)  k2q−1,qh ,2____ KO2q−1(Uk)  KO2q−1,q(Uk+1)  // // KO2q−1(Uk+1)  K˜O2q,q(S2d,d)  ∼= ,2 K˜O2q(S2d)  KO2q,q(Uk)  k2q,qh ,2_____ KO2q(Uk)  KO2q,q(Uk+1)  ∼= ,2 KO2q(Uk+1)  K˜O2q+1,q(S2d,d)  ∼= ,2 K˜O2q+1(S2d)  KO2q+1,q(Uk)  k2q+1,qh ,2____ KO2q+1(Uk)  KO2q+1,q(Uk+1)  ∼= ,2 KO2q+1(Uk+1)  K˜O2q+2,q(S2d,d)  ,2 ,2 K˜O2q+2(S2d)  KO2q+2,q(Uk)  k2q+2,qh ,2____ KO2q+2(Uk)  KO2q+2,q(Uk+1)  ,2 ,2 KO2q+2(Uk+1) ··· Figure 1: The inductive step. 85 Chapter IV. Cellular Varieties 2 The Atiyah-Hirzebruch spectral sequence for cellular vari- eties We now prepare the ground for the discussion of the KO-theory of some examples of cellular spaces in the next section. Our main tool will be the Atiyah-Hirzebruch spectral sequence described in Section I.2e. For cellular varieties, or more generally for CW complexes with only even-dimensional cells, this spectral sequence becomes simple enough to make some general deductions. So let X be a CW complex with cells only in even dimensions. Then the cohomology of X is free on generators given by the cells and concentrated in even degrees. The Atiyah- Hirzebruch spectral sequence for the K-theory of such a space collapses immediately: the entries of all columns and rows of odd degrees are zero, so there cannot be any non-zero differentials. We arrive at the following well-known lemma: 2.1 Lemma. Let X be a CW complex with cells only in even dimensions. Then K0(X) is a free abelian group of rank equal to the number of cells of X. The K-groups of odd degrees, on the other hand, vanish. After this preliminary observation, we turn to KO-theory. The Atiyah-Hirzebruch spectral sequence for the KO-theory of a CW complex with cells only in even dimen- sions does not necessarily collapse, but one can still make some general deductions. We summarize some lemmas of Hoggar and Kono and Hara. 2.2 Lemma. [Hog69, 2.1 and 2.2] Let X be a CW complex with only even-dimensional cells. Then: • The ranks of the free parts of KO0X and KO4X are equal to the number t0 of cells of X of dimension a multiple of 4. • The ranks of the free parts of KO2X and KO6X are equal to the number t1 of cells of X of dimension 2 modulo 4. • The groups of odd degrees are two-torsion, i. e. KO2i−1X = (Z/2)si for some si. • KO2iX is isomorphic to the direct sum of its free part and KO2i+1X. Moreover, multiplication by η ∈ KO−1(point) induces an isomorphism η : (KO2i/K)(X) ∼=−→ KO2i−1(X) The bullet points are summarized by Table 2 in Section 3a. Proof. The first two statements can be seen directly from the Atiyah-Hirzebruch spectral sequence for KO-theory (e. g. after tensoring with Q). The remaining statements may then be deduced from the above description of the K-groups and the Bott sequence (I.20). 86 2 The Atiyah-Hirzebruch spectral sequence for cellular varieties The free part of KO∗ is thus very simple. In good cases, the spectral sequence also provides a nice description of the 2-torsion. To see this, note that Sq2 Sq2 = Sq3 Sq1 must vanish when the cohomology of X with Z/2-coefficients is concentrated in even degrees. So we can view (H∗(X;Z/2),Sq2) as a differential graded algebra over Z/2. To lighten notation, we will write H∗(X, Sq2) := H∗(H∗(X;Z/2),Sq2) for the cohomology of this algebra. We keep the same grading as before, so that it is concentrated in even degrees. The row q ≡ −1 on the E3-page is given by H∗(X, Sq2) · η, where η is the generator of KO−1(point). Since it is the only row that contributes to KO∗ in odd degrees, we arrive at the following lemma, which will be central to our computations. 2.3 Lemma. Let X be as above. If the Atiyah-Hirzebruch spectral sequence of KO∗(X) degenerates on the E3-page then KO2i−1(X) ∼= ⊕ k H2i+8k(X, Sq2) Now suppose that E is a complex vector bundle over X. Then the twisted KO-groups KO∗(X;E) are computed by the Atiyah-Hirzebruch spectral sequence of the Thom space ThomE. Recall that the reduced cohomology of this space is isomorphic to the cohomology of X up to a shift in degrees. By Lemma I.2.11, the Steenrod square on H∗(ThomE;Z/2) is given by Sq2 + c1(E). If, as before, X has cells only in even dimensions, then Sq2+c1 may again be viewed as a differential on H∗(X;Z/2) for any c1 in H2(X;Z/2). Extending our previous notation, we denote cohomology with respect to this differential by H∗(X, Sq2 + c1) := H∗(H∗(X;Z/2),Sq2 + c1) (1) Lemma 2.3 has the following corollary. 2.4 Corollary. Let X and E be as above. If the Atiyah-Hirzebruch spectral sequence of K˜O∗(ThomE) degenerates on the E3-page, then KO2i−1(X;E) ∼= ⊕ k H2i+8k(X, Sq2 + c1E) In all the examples we consider below, the spectral sequence does indeed degenerate at this stage. However, showing that it does can be tricky. One step in the right direction is the following observation of Kono and Hara [KH91, Proposition 1]. 2.5 Lemma. Let X be as above. If the differentials d3, d4, . . ., dr−1 are trivial and dr is 87 Chapter IV. Cellular Varieties non-trivial, then r ≡ 2 mod 8. In other words, the first non-trivial differential after d2 can only appear on a page Er with page number r ≡ 2 mod 8. Such a differential is non-zero only on rows q ≡ 0 and q ≡ −1 mod 8. If it is non-zero on some x in row q ≡ 0, then it is also non-zero on ηx in row q ≡ −1. Conversely, if it is non-zero on some y in row q ≡ −1, there exists some x in row q ≡ 0 such that y = xη and dr is non-zero on x. Proof. We see from the spectral sequence of a point that drη = 0 for all differentials. Thus, multiplication by η gives a map of bidegree (0,−1) on the spectral sequence that commutes with the differentials. On the E2-page this map is mod-2 reduction from row q ≡ 0 to row q ≡ −1 and the identity between rows q ≡ −1 and q ≡ −2. It follows that on the E3-page multiplication by η induces a surjection from row q ≡ 0 to row q ≡ −1 and an injection of row q ≡ −1 into row q ≡ −2. This implies all statements above. We derive a corollary that we will use to deduce that the spectral sequence collapses for certain Thom spaces: 2.6 Corollary. Suppose we have a continuous map p : X → T of CW complexes with only even-dimensional cells. Suppose further that the Atiyah-Hirzebruch spectral sequence for KO∗(X) collapses on the E3-page, and that p∗ induces an injection in row q ≡ −1: p∗ : H∗(T,Sq2) ↪→ H∗(X, Sq2) Then the spectral sequence for KO∗(T ) also collapses at this stage. Proof. Write dr for the first non-trivial higher differential, so r ≡ 2 mod 8. Then, for any element x in row q ≡ 0, we have p∗(drx) = drp∗(x) = 0 since the spectral sequence for X collapses. From our assumption on p∗ we can deduce that drx = 0. By the preceding lemma, this is all we need to show. 88 3 Examples 3 Examples We now turn to the study of projective homogeneous varieties, that is, varieties of the form G/P for some complex simple linear algebraic group G with a parabolic subgroup P . Any such variety has a cell decomposition [BGG73, Proposition 5.1], so that our comparison theorem applies. As far as we are only interested in the topology of G/P , we may alternatively view it as a homogeneous space for the compact real Lie group Gc corresponding to G: 3.1 Proposition. Let P be a parabolic subgroup of a simple complex algebraic group G. Then we have a diffeomorphism G / P ∼= Gc / K where K is a compact subgroup of maximal rank in a maximal compact subgroup Gc of G. More precisely, K is a maximal compact subgroup of a Levi subgroup of P . Proof. The Iwasawa decomposition for G viewed as a real Lie group implies that we have a diffeomorphism G ∼= Gc · P [GOV94, Ch. 6, Prop. 1.7], inducing a diffeomorphism of quotients as claimed for K = Gc ∩ P . Since Gc ↪→ G is a homotopy equivalence, so is the inclusion Gc ∩ P ↪→ P . On the other hand, if L is a Levi subgroup of P then P = U o L, where U is unipotent and hence contractible. So the inclusion L ↪→ P is also a homotopy equivalence. It follows that any maximal compact subgroup Lc of L is also maximal compact in P , and conversely that any maximal compact subgroup of P will be contained as a maximal compact subgroup in some Levi subgroup of P . We may therefore assume that K ⊂ Lc ⊂ L ⊂ P and conclude that K ↪→ Lc is a homotopy equivalence. Since both groups are compact, it follows that in fact K ∼= Lc. The KO-theory of homogeneous varieties has been studied intensively. In particular, the papers [KH91] and [KH92] of Kono and Hara provide complete computations of the (untwisted) KO-theory of all compact irreducible hermitian symmetric spaces, which we list in Table 1. For the convenience of the reader, we indicate how each of these arises as a quotient of a simple complex algebraic group G by a parabolic subgroup P , describing the latter in terms of marked nodes on the Dynkin diagram of G as in [FH91, § 23.3]. The last column gives an alternative description of each space as a quotient of a compact real Lie group. On the following pages, we will run through this list of examples and, in each case, extend Kono and Hara’s computations to include KO-groups twisted by a line bundle. Since each of these spaces is a “Grassmannian” in the sense that the parabolic subgroup P in G is maximal, its Picard group is free on a single generator. Thus, there is exactly one non-trivial twist that we need to consider. In most cases, we — reassuringly — recover results for Witt groups that are already known. In a few other cases, we consider our results as new. 89 Chapter IV. Cellular Varieties G / P G Diagram of P G c / K Grassmannians (AIII) Grm,n SLm+n ◦ · · · ◦ • ◦ · · · ◦ 1 n n+m–1 U(m+ n) U(m)×U(n) Maximal symplectic Grassmannians (CI) Xn Sp2n ◦ ◦ · · · ◦ ◦ < • Sp(n) / U(n) Projective quadrics of dimension n ≥ 3 (BDI) Qn SOn+2 • ◦ · · · ◦ > ◦ (n odd) ◦ • ◦ · · · ◦}A (n even)◦ SO(n+ 2) SO(n)× SO(2) Spinor varieties (DIII) Sn SO2n ◦ ◦ ◦ · · · ◦}A• SO(2n) / U(n) Exceptional hermitian symmetric spaces: EIII E6 ◦ ◦ ◦ ◦ ◦ • Ec6 Spin(10) · S1 (Spin(10)∩S1=Z/4) EVII E7 ◦ ◦ ◦ ◦ ◦ ◦ • Ec7 Ec6 · S1 (Ec6∩S1=Z/3) Table 1: List of irreducible compact hermitian symmetric spaces. The symbols AIII, CI, . . . refer to E. Cartan’s classification. In the description of Grm,n we use U(m+n) instead of Gc = SU(m+ n). The untwisted KO-theory of complete flag varieties is also known in all three classical cases thanks to Kishimoto, Kono and Ohsita. We do not reproduce their result here but instead refer the reader directly to [KKO04]. By a recent result of Calme`s and Fasel, all Witt groups with non-trivial twists vanish for these varieties [CF11]. 3a Notation Topologically, a cellular variety is a CW complex with cells only in even (real) dimensions. For such a CW complex X the KO-groups can be written in the form displayed in Table 2 below. This was shown in Section 2 in the case when the twist L is trivial, and the general case follows: if X is a CW complex with only even-dimensional cells, so is the Thom space of any complex vector bundle over X [MS74, Lemma 18.1]. In the following examples, results on KO∗ will be displayed by listing the values of the ti and si. Since the ti are just given by counting cells, and since the numbers of odd- and even-dimensional cells of a Thom space ThomXE only depend on X and the rank of E, the ti are in fact independent of L. The si, on the other hand, certainly will depend on the twist, and we will sometimes acknowledge this by writing si(L). 90 3 Examples KO6(X;L) = Zt1 ⊕ (Z/2)s0 = GW3(X;L) KO7(X;L) = (Z/2)s0 = W0(X;L) KO0(X;L) = Zt0 ⊕ (Z/2)s1 = GW0(X;L) KO1(X;L) = (Z/2)s1 = W1(X;L) KO2(X;L) = Zt1 ⊕ (Z/2)s2 = GW1(X;L) KO3(X;L) = (Z/2)s2 = W2(X;L) KO4(X;L) = Zt0 ⊕ (Z/2)s3 = GW2(X;L) KO5(X;L) = (Z/2)s3 = W3(X;L) Table 2: Notational conventions in the examples. Only the si depend on L. 3b Projective spaces Complex projective spaces are perhaps the simplest examples for which Theorem 1.1 asserts something non-trivial, so we describe the results here separately before turning to complex Grassmannians in general. The computations of the Witt groups of projective spaces were landmark events in the history of the theory. In 1980, Arason was able to show that the Witt group W0(Pn) of Pn over a field k agrees with the Witt group of k [Ara80]. The shifted Witt groups of projective spaces, and more generally of arbitrary projective bundles, were first computed by Walter in [Wal03b]. Quite recently, Nenashev deduced the same results via different methods [Nen09]. In the topological world, complete computations of KOi(CPn) were first published in a 1967 paper by Fujii [Fuj67]. It is not difficult to deduce the values of the twisted groups KOi(CPn;O(1)) from these: the space Thom(OCPn(1)) can be identified with CPn+1, so KOi(CPn;O(1)) = K˜Oi+2(Thom(O(1))) = K˜Oi+2(CPn+1) Alternatively, we could do all required computations directly following the methods out- lined in Section 2. The result, in any case, is displayed in Table 3, coinciding with the known results for the (Grothendieck-)Witt groups. KO∗(CPn;L) L ≡ O L ≡ O(1) t0 t1 s0 s1 s2 s3 s0 s1 s2 s3 n ≡ 0 mod 4 (n/2) + 1 n/2 1 0 0 0 1 0 0 0 n ≡ 1 (n+ 1)/2 (n+ 1)/2 1 1 0 0 0 0 0 0 n ≡ 2 (n/2) + 1 n/2 1 0 0 0 0 0 1 0 n ≡ 3 (n+ 1)/2 (n+ 1)/2 1 0 0 1 0 0 0 0 Table 3: KO-groups of projective spaces 91 Chapter IV. Cellular Varieties 3c Grassmannians We now consider the Grassmannians Grm,n of complex m-planes in Cm+n. Again both the Witt groups and the untwisted KO-groups are already known: the latter by Kono and Hara [KH91], the former by the work of Balmer and Calme`s [BC08]. A detailed comparison of the two sets of results in the untwisted case has been carried out by Yagita [Yag09]. We provide here an alternative topological computation of the twisted groups. Balmer and Calme`s state their result by describing an additive basis of the total Witt group of Grm,n in terms of certain “even Young diagrams”. This is probably the most elegant approach, but needs some space to explain. We will stick instead to the tabular exposition used in the other examples. Let O(1) be a generator of Pic(Grm,n), say the dual of the determinant line bundle of the universal m-bundle over Grm,n. The result is displayed in Table 4. KO∗(Grm,n;L) L ≡ O L ≡ O(1) t0 t1 s0 s1 s2 s3 s0 s1 s2 s3 m and n odd s.t. m ≡ n a 2 a 2 b b 0 0 0 0 0 0 m and n odd s.t. m 6≡ n a 2 a 2 b 0 0 b 0 0 0 0 m ≡ n ≡ 0 m ≡ 0 and n odd n ≡ 0 and m odd a+ b 2 a− b 2 b 0 0 0 b 0 0 0  m ≡ n ≡ 2 m ≡ 2 and n odd n ≡ 2 and m odd a+ b 2 a− b 2 b 0 0 0 0 0 b 0 m ≡ 0 and n ≡ 2 a+ b 2 a− b 2 b 0 0 0 b1 0 b2 0 m ≡ 2 and n ≡ 0 a+ b 2 a− b 2 b 0 0 0 b2 0 b1 0 All equivalences (≡) are modulo 4. For the values of a and b = b1 + b2, put k := bm/2c and l := bn/2c. Then a := ( m+n m ) b := ( k+l k ) b1 := ( k+l−1 k ) b2 := ( k+l−1 k−1 ) Table 4: KO-groups of Grassmannians Our computation is based on the following geometric observation. Let Um,n and U⊥m,n be the universal m-bundle and the orthogonal n-bundle over Grm,n, so that U ⊕ U⊥ = 92 3 Examples O⊕(m+n). We have various natural inclusions between the Grassmannians of different dimensions, of which we fix two: Grm,n−1 ↪→ Grm,n via the inclusion of the first m+ n− 1 coordinates into Cm+n Grm−1,n ↪→ Grm,n by sending an (m − 1)-plane Λ to the m-plane Λ ⊕ 〈em+n〉, where e1, e2, . . . , em+n are the canonical basis vectors of Cm+n 3.2 Lemma. The normal bundle of Grm,n−1 in Grm,n is the dual U∨m,n−1 of the universal m-bundle. Similarly, the normal bundle of Grm−1,n in Grm,n is given by U⊥m−1,n. In both cases, the embeddings of the subspaces extend to embeddings of their normal bundles, such that one subspace is the closed complement of the normal bundle of the other. This gives us two cofibration sequences of pointed spaces: Grm−1,n+ i ↪→ Grm,n+ p  Thom(U∨m,n−1) (2) Grm,n−1+ i ↪→ Grm,n+ p  Thom(U⊥m−1,n) (3) These sequences are the key to relating the untwisted KO-groups to the twisted ones. Following the notation in [KH91], we write Am,n for the cohomology of Grm,n with Z/2- coefficients, denoting the Chern classes of U and U⊥ by ai and bi, respectively, and the total Chern classes 1 + a1 + · · ·+ am and 1 + b1 + · · ·+ bn by a and b: Am,n = Z/2 [a1, a2, . . . , am, b1, b2, . . . bn] a · b = 1 We write d for the differential given by the second Steenrod square Sq2, and d′ for Sq2+a1. To describe the cohomology of Am,n with respect to these differentials, it is convenient to introduce the algebra Bk,l = Z/2 [ a22, a 2 4, . . . , a 2 2k, b 2 2, b 2 4, . . . , b 2 2l ] (1 + a22 + · · ·+ a22k)(1 + b22 + · · ·+ b22l) = 1 Note that this subquotient of A2k,2l is isomorphic to Ak,l up to a “dilatation” in grading. Proposition 2 in [KH91] tells us that H∗(Am,n, d) = Bk,l if (m,n) = (2k, 2l), (2k + 1, 2l) or (2k, 2l + 1)Bk,l ⊕Bk,l · ambn−1 if (m,n) = (2k + 1, 2l + 1) Here, the algebra structure in the case where both m and n are odd is determined by (ambn−1)2 = 0. 93 Chapter IV. Cellular Varieties 3.3 Lemma. The cohomology of Am,n with respect to the twisted differential d′ is as follows: H∗(Am,n, d′) =  Bk,l−1 · am ⊕Bk−1,l · bn if (m,n) = (2k, 2l) Bk,l · am if (m,n) = (2k, 2l + 1) Bk,l · bn if (m,n) = (2k + 1, 2l) 0 if (m,n) = (2k + 1, 2l + 1) Proof. Let us shift the dimensions in the cofibration sequences (2) and (3) in such a way that we have the Thom spaces of U∨m,n and U⊥m,n on the right. Since the cohomologies of the spaces involved are concentrated in even degrees, the associated long exact sequence of cohomology groups falls apart into short exact sequences. Reassembling these, we obtain two short exact sequences of differential (Am,n+1, d)- and (Am+1,n, d)-modules, respectively: 0→ (Am,n, d′) · θ∨ p ∗ −→(Am,n+1, d) i ∗−→ (Am−1,n+1, d)→ 0 (4) 0→ (Am,n, d′) · θ⊥ p ∗ −→(Am+1,n, d) i ∗−→ (Am+1,n−1, d)→ 0 (5) Here, θ∨ and θ⊥ are the respective Thom classes of U∨m,n and U⊥m,n. The map i∗ in the first row is the obvious quotient map annihilating am. Its kernel, the image of Am,n under multiplication by am, is generated as an Am,n+1-module by its unique element in degree 2m, and thus we must have p∗(θ∨) = am. Likewise, in the second row we have p∗(θ⊥) = bn. The lemma can be deduced from here case by case. For example, when both m and n are even, i∗ maps H∗(Am,n+1, d) = Bk,l to the first summand of H∗(Am−1,n+1, d) = Bk−1,l ⊕ Bk−1,l · am−1bn by annihilating a2m. We know by comparison with the short exact sequences for the Am,n that the kernel of this map is Bk,l−1 mapping to Bk,l under multiplication by a2m. Thus, we obtain a short exact sequence 0→ Bk−1,l · am−1bn ∂−→ H∗(Am,n, d′) · θ∨ p ∗ −→ Bk,l−1 · a2m → 0 (6) For the Steenrod square Sq2 of the top Chern class am of U, we have Sq2(am) = a1am. This can be checked, for example, by expressing am as the product of the Chern roots of U. Consequently, d′(am) = 0. Together with the fact that H∗(Am,n, d′) is a module over H∗(Am,n+1, d), this shows that we can define a splitting of p∗ by sending a2m to amθ∨. Thus, H∗(Am,n, d′) contains Bk,l−1 · am as a direct summand. If instead of working with sequence (4) we work with sequence (5), we see that H∗(Am,n, d′) also contains a direct summand Bk−1,l · bn. These two summands intersect trivially, and a dimension count shows that together they encompass all of H∗(Am,n, d′). Alternatively, one may check explicitly that the boundary map ∂ above sends am−1bn to bnθ. The other cases are simpler. 94 3 Examples 3.4 Lemma. The Atiyah-Hirzebruch spectral sequence for K˜O∗(ThomU∨m,n) collapses at the E3-page. Proof. By Proposition 4 of [KH91] we know that the spectral sequence for KO∗(Grm,n) collapses as this stage, for any m and n. Now, if both m and n are even, we have (Bk,l−1 · am ⊕Bk−1,l · bn) · θ in the (−1)st row of the E3-pages of the spectral sequences for ThomU∨ and ThomU⊥, where θ = θ∨ or θ⊥, respectively. In the case of U∨ we see from (6) that p∗ maps the second summand injectively to the E3-page of the spectral sequence for KO∗(Grm,n+1). Similarly, in the case of U⊥, the first summand is mapped injectively to the E3-page of KO∗(Grm+1,n). Since the spectral sequences for ThomU∨ and ThomU⊥ can be identified via Corollary I.2.13, we can argue as in Corollary 2.6 to see that they must collapse at this stage. Again, the cases when at least one of m, n is odd are similar but simpler. We may now apply Corollary 2.4. The entries of Table 4 that do not appear in [KH91], i. e. those of the last four columns, follow from Lemma 3.3 by noting that Bk,l is concentrated in degrees 8i and of dimension dimBk,l = dimAk,l = ( k+l k ) . 3d Maximal symplectic Grassmannians The Grassmannian of isotropic n-planes in C2n with respect to a non-degenerate skew- symmetric bilinear form is given by Xn = Sp(n)/U(n). The universal bundle U over the usual Grassmannian Gr(n, 2n) restricts to the universal bundle over Xn, and so does the orthogonal complement bundle U⊥. We will continue to denote these restrictions by the same letters. Thus, U⊕U⊥ ∼= C2n over Xn, and the fibres of U are orthogonal to those of U⊥ with respect to the standard hermitian metric on C2n. The determinant line bundles of U and U⊥ give dual generators O(1) and O(−1) of the Picard group of Xn. 3.5 Theorem. The additive structure of KO∗(Xn;L) is as follows: t0 t1 si(O) si(O(1)) n even 2n−1 2n−1 ρ(n2 , i) ρ( n 2 , i− n) n odd 2n−1 2n−1 ρ(n+12 , i) 0 Here, for any i ∈ Z/4 we write ρ(n, i) for the dimension of the i-graded piece of a Z/4- graded exterior algebra ΛZ/2(g1, g2, . . . , gn) on n homogeneous generators g1, g2, . . . , gn of degree 1, i. e. ρ(n, i) = ∑ d≡i mod 4 ( n d ) A table of the values of ρ(n, i) can be found in [KH92, Proposition 4.1]. 95 Chapter IV. Cellular Varieties It turns out to be convenient to work with the vector bundle U⊥ ⊕ O for the compu- tation of the twisted groups KO∗(Xn;O(1)). Namely, we have the following analogue of Lemma 3.2. 3.6 Lemma. There is an open embedding of the bundle U⊥ ⊕ O over the symplectic Grassmannian Xn into the symplectic Grassmannian Xn+1 whose closed complement is again isomorphic to Xn. Proof. To fix notation, let e1, e2 be the first two canonical basis vectors of C2n+2, and embed C2n into C2n+2 via the remaining coordinates. Assuming Xn is defined in terms of a skew-symmetric form Q2n, define Xn+1 with respect to the form Q2n+2 :=  0 1 0−1 0 0 0 0 Q2n  Then we have embeddings i1 and i2 of Xn into Xn+1 sending an n-plane Λ ⊂ C2n to e1⊕Λ or e2 ⊕ Λ in C2n+2, respectively. We extend i1 to an embedding of U⊥⊕O by sending an n-plane Λ ∈ Xn together with a vector v in Λ⊥ ⊂ C2n and a complex scalar z to the graph ΓΛ,v,z ⊂ C2n+2 of the linear map ( z Q2n(−, v) v 0 ) : 〈e1〉 ⊕ Λ→ 〈e2〉 ⊕ Λ⊥ To avoid confusion, we emphasize that v is orthogonal to Λ with respect to a hermitian metric on C2n. The value of Q2n(−, v), on the other hand, may well be non-zero on Λ. Consider the above embedding of U⊥ ⊕ O together with the embedding i2: U⊥ ⊕ O ↪−−→ Xn+1 Xn←−−↩i2 (Λ, v, z) 7→ ΓΛ,v,z 〈e2〉 ⊕ Λ Λ←[ To see that the two embeddings are complementary, take an arbitrary (n+1)-plane W in Xn+1. If e2 ∈W then we can consider a basis e2, ( a1 0 v1 ) , . . . , ( an 0 vn ) of W , and the fact that Q2n+2 vanishes on W implies that all ai are zero. Thus W can be identified with i2(〈v1, . . . , vn〉). If, on the other hand, e2 is not contained in W then we must have a vector of the form t(1, z′, v′) in W , for some z′ ∈ C and v′ ∈ C2n. Extend this vector to a basis of W of the form ( 1 z′ v′ ) , ( 0 b1 v1 ) , . . . , ( 0 bn vn ) 96 3 Examples and let Λ := 〈v1, . . . , vn〉. The condition that Q2n+2 vanishes onW implies that Q vanishes on Λ and that bi = Q2n(vi, v′) for each i. In particular, Λ is n-dimensional. Moreover, we can replace the first vector of our basis by a vector t(1, z, v) with v ∈ Λ⊥, by subtracting appropriate multiples of the remaining basis vectors. Since Q vanishes on Λ we have Q2n(vi, v′) = Q2n(vi, v) and our new basis has the form( 1 z v ) , ( 0 Q(v1,v) v1 ) , . . . , ( 0 Q(vn,v) vn ) This shows that W = ΓΛ,v,z. 3.7 Corollary. We have a cofibration sequence Xn+ i ↪→ Xn+1+ p  ThomXn(U⊥ ⊕ O) The associated long exact cohomology sequence splits into a short exact sequence of H∗(Xn+1)-modules since all cohomology here is concentrated in even degrees: 0→ H˜∗(ThomXn(U⊥ ⊕ O)) p ∗ → H∗(Xn+1) i ∗→ H∗(Xn)→ 0 (7) 3.8 Lemma. Let ci denote the ith Chern classes of U on Xn. We have H∗(Xn,Sq2) = Λ(a1, a5, a9, . . . , a4m−3) if n = 2mΛ(a1, a5, a9, . . . , a4m−3, a4m+1) if n = 2m+ 1 H∗(Xn,Sq2 + c1) = Λ(a1, a5, . . . , a4m−3) · c2m if n = 2m0 if n is odd for certain generators ai of degree 2i. Proof. Consider the short exact sequence (7). The mod-2 cohomology of Xn is an exterior algebra on the Chern classes ci of U, H∗(Xn;Z/2) = Λ(c1, c2, . . . , cn) and i∗ is given by sending cn+1 to zero. Thus, p∗ is the unique morphism ofH∗(Xn+1;Z/2)- modules that sends the Thom class θ of U⊥ ⊕ O to cn+1. This short exact sequence induces a long exact sequence of cohomology groups with respect to the Steenrod square Sq2. The algebra H∗(Xn,Sq2) was computed in [KH92, 2–2], with the result displayed above, so we already know two thirds of this sequence. Explicitly, we have a4i+1 = c2ic2i+1,1 so i∗ is the obvious surjection sending ai to ai (or to zero). Thus, the long exact sequence once again splits. 1In [KH92] the generators are written as c2ic ′ 2i+1 with c ′ 2i+1 = c2i+1 + c1c2i. 97 Chapter IV. Cellular Varieties If n = 2m we obtain a short exact sequence 0 → H∗(X2m,Sq2 + c1) · θ p ∗ → Λ(a1, . . . , a4m−3, a4m+1) i ∗→ Λ(a1, . . . , a4m−3) → 0 We see that H∗(X2m,Sq2+c1)·θ is isomorphic to Λ(a1, . . . , a4m−3)·a4m+1 as a module over Λ(a1, . . . , a4m+1). It is thus generated by a single element, which is the unique element of degree 8m + 2. Since p∗(c2mθ) = a4m+1, the class of c2mθ is the element we are looking for, and the result displayed above follows. If, on the other hand, n is odd, then i∗ is an isomorphism and H∗(Xn,Sq2 + c1) must be trivial. We see from the proof that p∗ induces an injection ofH∗(Xn,Sq2 + c1) · θ intoH∗(Xn,Sq2). Since we already know from [KH92, Theorem 2.1] that the Atiyah-Hirzebruch spectral sequence for KO∗(Xn) collapses, we can apply Corollary 2.6 to deduce that the spectral sequence for K˜O∗(ThomXn(U⊥ ⊕ O)) collapses at the E3-page as well. This completes the proof of Theorem 3.5. 3e Quadrics We next consider smooth complex quadrics Qn in Pn+1. As far as we are aware, the first complete results on (shifted) Witt groups of split quadrics were due to Walter: they are mentioned together with the results for projective bundles in [Wal03a] as the main appli- cations of that paper. Unfortunately, they seem to have remained unpublished. Partial results are also included in Yagita’s preprint [Yag04], see Corollary 8.3. More recently, Nenashev obtained almost complete results by considering the localization sequences aris- ing from the inclusion of a linear subspace of maximal dimension [Nen09]. Calme`s informs me that the geometric description of the boundary map given in [BC09] can be used to show that these localization sequences split in general, yielding a complete computation. The calculation described here is completely independent of these results. For n ≥ 3 the Picard group of Qn is free on a single generator given by the restriction of the universal line bundle O(1) on Pn+1. We will use the same notation O(1) for this restriction. 3.9 Theorem. The KO-theory of a smooth complex quadric Qn of dimension n ≥ 3 is as described in Table 5. Untwisted KO-groups. Before turning to KO∗(Qn;O(1)) we review the initial steps in the computation of the untwisted KO-groups. The integral cohomology of Qn is well- known: If n is even, write n = 2m. We have a class x in H2(Qn) given by a hyperplane section, and two classes a and b in Hn(Qn) represented by linear subspaces of Q of maximal 98 3 Examples KO∗(Qn;L) L ≡ O L ≡ O(1) t0 t1 s0 s1 s2 s3 s0 s1 s2 s3 n ≡ 0 mod 8 (n/2) + 2 n/2 2 0 0 0 2 0 0 0 n ≡ 1 (n+ 1)/2 (n+ 1)/2 1 1 0 0 1 1 0 0 n ≡ 2 (n/2) + 1 (n/2) + 1 1 2 1 0 0 0 0 0 n ≡ 3 (n+ 1)/2 (n+ 1)/2 1 1 0 0 0 0 1 1 n ≡ 4 (n/2) + 2 n/2 2 0 0 0 0 0 2 0 n ≡ 5 (n+ 1)/2 (n+ 1)/2 1 0 0 1 0 1 1 0 n ≡ 6 (n/2) + 1 (n/2) + 1 1 0 1 2 0 0 0 0 n ≡ 7 (n+ 1)/2 (n+ 1)/2 1 0 0 1 1 0 0 1 Table 5: KO-groups of projective quadrics (n ≥ 3) dimension. These three classes generate the cohomology multiplicatively, modulo the relations xm = a+ b xm+1 = 2ax ab = 0 if n ≡ 0axm if n ≡ 2 a2 = b2 = axm if n ≡ 0 mod 40 if n ≡ 2 mod 4 Additive generators can thus be given as follows: d 0 2 4 . . . n− 2 n n+ 2 n+ 4 . . . 2n Hd(Qn) 1 x x2 . . . xm−1 a, b ax ax2 . . . axm If n is odd, write n = 2m + 1. Then similarly multiplicative generators are given by the class of a hyperplane section x in H2(Qn) and the class of a linear subspace a in Hn+1(Qn) modulo the relations xm+1 = 2a and a2 = 0. d 0 2 4 . . . n− 1 n+ 1 n+ 3 n+ 5 . . . 2n Hd(Qn) 1 x x2 . . . xm a ax ax2 . . . axm The action of the Steenrod square on H∗(Qn;Z/2) is also well-known; see for example [Ish92, Theorem 1.4 and Corollary 1.5] or [EKM08, § 78]: Sq2(x) = x2 Sq2(a) = ax if n ≡ 0 or 3 mod 40 if n ≡ 1 or 2 Sq2(b) = Sq2(a) (for even n) As before, we write H∗(Qn,Sq2) for the cohomology of H∗(Qn;Z/2) with respect to the differential Sq2. 99 Chapter IV. Cellular Varieties 3.10 Lemma. Write n = 2m or n = 2m + 1 as above. The following table gives a complete list of the additive generators of H∗(Qn,Sq2). d 0 . . . n− 1 n n+ 1 . . . 2n Hd(Qn,Sq2) 1 axm if n ≡ 0 mod 4 1 a if n ≡ 1 1 a, b ab if n ≡ 2 1 xm if n ≡ 3 The results of Kono and Hara on KO∗(Q) follow from here provided there are no non- trivial higher differentials in the Atiyah-Hirzebruch spectral sequence. This is fairly clear in all cases except for the case n ≡ 2 mod 4. In that case, the class a + b = xm can be pulled back from Qn+1, and therefore all higher differentials must vanish on a+b. But one has to work harder to see that all higher differentials vanish on a (or b). Kono and Hara proceed by relating the KO-theory of Qn to that of the spinor variety Sn 2 +1 discussed in Section 3f. Twisted KO-groups. We now compute KO∗(Qn;O(1)). Let θ ∈ H2(ThomQnO(1)) be the Thom class of O(1), so that multiplication by θ maps the cohomology of Qn isomorphically to the reduced cohomology of ThomQnO(1). The Steenrod square on H˜∗(ThomQnO(1);Z/2) is determined by Lemma I.2.11: for any y ∈ H∗(Qn;Z/2) we have Sq2(y · θ) = (Sq2 y + xy) · θ. We thus arrive at 3.11 Lemma. A complete list of the additive generators of H˜∗(ThomQnO(1),Sq2) is pro- vided by the following table: d . . . n+ 1 n+ 2 n+ 3 . . . 2n+ 2 H˜d(. . . ) aθ, bθ if n ≡ 0 mod 4 xmθ axmθ if n ≡ 1 if n ≡ 2 aθ axmθ if n ≡ 3 We claim that all higher differentials in the Atiyah-Hirzebruch spectral sequence for K˜O∗(ThomQnO(1)) vanish. For even n this is clear. But for n = 8k + 1 the differential d8k+2 might a priori take xmθ to axmθ, and for n = 8k + 3 the differential d8k+2 might take aθ to axmθ. We therefore need some geometric considerations. Namely, the space ThomQnO(1) can be identified with the projective cone over Qn embedded in Pn+2. This projective cone can be realized as the intersection of a smooth quadric Qn+2 ⊂ Pn+3 with its projective tangent space at the vertex of the cone [Har92, p. 283]. Thus, we can consider the following inclusions: Qn j ↪→ ThomQnO(1) i↪→ Qn+2 100 3 Examples The composition is the inclusion of the intersection of Qn+2 with two transversal hyper- planes. 3.12 Lemma. All higher differentials ( dk with k > 2) in the Atiyah-Hirzebruch spectral sequence for KO∗(ThomQnO(1)) vanish. Proof. We need only consider the cases when n is odd. Write n = 2m+ 1. When n ≡ 1 mod 4 we claim that i∗ maps xm+1 in Hn+1(Qn+2,Sq2) to xmθ in Hn+1(ThomQnO(1),Sq2). Indeed, j∗i∗ maps the class of the hyperplane section x in H2(Qn+2) to the class of the hyperplane section x in H2(Qn). So i∗x in H2(ThomQnO(1)) must be non-zero, hence equal to θ modulo 2. It follows that i∗(xm+1) = θm+1. Since θ2 = Sq2(θ) = xθ, we have θm+1 = xmθ, proving the claim. As we already know that all higher differentials vanish on H∗(Qn+2,Sq2), we may now deduce that they also vanish on H∗(ThomQnO(1),Sq2). When n ≡ 3 mod 4 we claim that i∗ maps the element a in Hn+3(Qn+2,Sq2) to aθ in Hn+3(ThomQnO(1),Sq2). Indeed, a represents a linear subspace of codimension m+ 2 in Qn+2 and is thus mapped to the class of a linear subspace of the same codimension in Qn: j∗i∗(a) = ax in Hn+3(Qn). Thus, i∗(a) is non-zero in Hn+3(ThomQnO(1)), equal to aθ modulo 2. Again, this implies that all higher differentials vanish on H∗(ThomQnO(1),Sq2) since they vanish on H∗(Qn+2,Sq2). The additive structure of KO∗(Qn;O(1)) thus follows directly from the result for Hd(Qn,Sq2 + x) = H˜d+2(ThomQnO(1)) displayed in Lemma 3.11 via Corollary 2.4. 3f Spinor varieties Let GrSO(n,N) be the Grassmannian of n-planes in CN isotropic with respect to a fixed non-degenerate symmetric bilinear form, or, equivalently, the Fano variety of projective (n − 1)-planes contained in the quadric QN−2. For each N > 2n, this is an irreducible homogeneous variety. In particular, for N = 2n + 1 we obtain the spinor variety Sn+1 = GrSO(n, 2n+1). The variety GrSO(n, 2n) falls apart into two connected components, both of which are isomorphic to Sn. This is reflected by the fact that we can equivalently identify Sn with SO(2n− 1)/U(n− 1) or SO(2n)/U(n). As for all Grassmannians, the Picard group of Sn is isomorphic to Z; we fix a line bundle S which generates it. The KO-theory twisted by S vanishes: 3.13 Theorem. For all n ≥ 2 the additive structure of KO∗(Sn;L) is as follows: t0 t1 si(O) si(S) n ≡ 2 mod 4 2n−2 2n−2 ρ(n2 , 1− i) 0 otherwise 2n−2 2n−2 ρ(bn2 c,−i) 0 101 Chapter IV. Cellular Varieties The values ρ(n, i) are defined as in Theorem 3.5. Proof. The cohomology of Sn with Z/2-coefficients has simple generators e2, e4, . . . , e2n−2, i. e. it is additively generated by products of distinct elements of this list. Its multiplicative structure is determined by the rule e22i = e4i, and the second Steenrod square is given by Sq2(e2i) = ie2i+2 [Ish92, Proposition 1.1]. In both formulae it is of course understood that e2j = 0 for j ≥ n. What we need to show is that for all n ≥ 2 we have H∗(Sn,Sq2 + e2) = 0 Let us abbreviate H∗(Sn,Sq2+ e2) to (Hn, d′). We claim that we have the following short exact sequence of differential Z/2-modules: 0→ (Hn, d′) ·e2n→ (Hn+1, d′)→ (Hn, d′)→ 0 (8) This can be checked by a direct calculation. Alternatively, it can be deduced from the geometric considerations below. Namely, it follows from the cofibration sequence of Corol- lary 3.15 that we have such an exact sequence of Z/2-modules with maps respecting the differentials given by Sq2 on all three modules. Since they also commute with multiplica- tion by e2, they likewise respect the differential d′ = Sq2 + e2. The long exact cohomology sequence associated with (8) allows us to argue by induc- tion: if H∗(Hn, d′) = 0 then also H∗(Hn+1, d′) = 0. Since we can see by hand that H∗(H2, d′) = 0, this completes the proof. We close with a geometric interpretation of the exact sequence (8), via an analogue of Lemmas 3.2 and 3.6. Let us write U for the universal bundle over Sn, i. e. for the restriction of the universal bundle over Gr(n− 1, 2n− 1) to Sn, and U⊥ for the restriction of the orthogonal complement bundle, so that U⊕ U⊥ is the trivial (2n− 1)-bundle over Sn. As in Section 3d, we emphasize that under these conventions the fibres of U and U⊥ are perpendicular with respect to a hermitian metric on C2n−1 — they are not orthogonal with respect to the chosen symmetric form. 3.14 Lemma. The spinor variety Sn embeds into the spinor variety Sn+1 with normal bundle U⊥ such that the embedding extends to an embedding of this bundle. The closed complement of U⊥ in Sn+1 is again isomorphic to Sn. 3.15 Corollary. We have a cofibration sequence Sn+ i ↪→ Sn+1+ p  ThomSn U⊥ Note however that, unlike in the symplectic case, the first Chern classes of U and U⊥ pull back to twice a generator of the Picard group of Sn. For example, the embedding of S2 into Gr(1, 3) can be identified with the embedding of the one-dimensional smooth 102 3 Examples quadric into the projective plane, of degree 2, and the higher dimensional cases can be reduced to this example. Thus, c1(U) and c1(U⊥) are trivial in Pic(Sn)/2. proof of Lemma 3.14. The proof is similar to the proof of Lemma 3.6. Let e1, e2 be the first two canonical basis vectors of C2n+1, and let C2n−1 be embedded into C2n+1 via the remaining coordinates. Let Sn be defined in terms of a symmetric form Q on C2n−1, and define Sn+1 in terms of Q2n+1 := 0 1 01 0 0 0 0 Q  Let i1 and i2 be the embeddings of Sn into Sn+1 sending an (n − 1)-plane Λ ⊂ C2n−1 to e1 ⊕ Λ or e2 ⊕ Λ in C2n+1, respectively. Given an (n − 1)-plane Λ ∈ Sn together with a vector v in Λ⊥ ⊂ C2n−1, consider the linear map( −12Q(v, v) −Q(−, v) v 0 ) : 〈e1〉 ⊕ Λ→ 〈e2〉 ⊕ Λ⊥ Sending (Λ, v) to the graph of this function defines an open embedding of U⊥ whose closed complement is the image of i2. 3g Exceptional hermitian symmetric spaces Lastly, we turn to the exceptional hermitian symmetric spaces EIII and EVII. We write O(1) for a generator of the Picard group in both cases. 3.16 Theorem. The KO-groups of the exceptional hermitian symmetric spaces EIII and EVII are as follows: L ≡ O L ≡ O(1) t0 t1 s0 s1 s2 s3 s0 s1 s2 s3 KO∗(EIII;L) 15 12 3 0 0 0 3 0 0 0 KO∗(EVII;L) 28 28 1 3 3 1 0 0 0 0 Proof. The untwisted KO-groups have been computed in [KH92], the main difficulty as always being to prove that the Atiyah-Hirzebruch spectral sequence collapses. For the twisted groups, however, there are no problems. We quote from § 3 of said paper that the cohomologies of the spaces in question can be written as H∗(EIII;Z/2) = Z/2 [ t, u ]/ (u2t, u3 + t12) H∗(EVII;Z/2) = Z/2 [ t, v, w ]/ (t14, v2, w2) 103 Chapter IV. Cellular Varieties with t of degree 2 in both cases, and u, v and w of degrees 8, 10 and 18, respectively. The Steenrod squares are determined by Sq2 u = ut and Sq2 v = Sq2w = 0. Thus, we find H∗(EIII,Sq2 + t) = Z/2 · u⊕ Z/2 · u2 ⊕ Z/2 · u3 H∗(EVII,Sq2 + t) = 0 By Lemma 2.5, the Atiyah-Hirzebruch spectral sequence for EIII must collapse. This gives the result displayed above. 104 References [Ada95] Frank Adams, Stable homotopy and generalised homology, Chicago Lectures in Mathematics, University of Chicago Press, Chicago, IL, 1995. Reprint of the 1974 original. [AF59] Aldo Andreotti and Theodore Frankel, The Lefschetz theorem on hyperplane sections, Ann. of Math. (2) 69 (1959), 713–717. [Ara80] Jo´n Kristinn Arason, Der Wittring projektiver Ra¨ume, Math. Ann. 253 (1980), no. 3, 205–212. [ABS64] Michael Atiyah, Raoul Bott, and Arnold Shapiro, Clifford modules, Topology 3 (1964), no. suppl. 1, 3–38. [AR76] Michael Atiyah and Elmer Rees, Vector bundles on projective 3-space, Invent. Math. 35 (1976), 131–153. [Bal00] Paul Balmer, Triangular Witt groups. I. The 12-term localization exact sequence, K-Theory 19 (2000), no. 4, 311–363. [Bal01a] , Triangular Witt groups. II. From usual to derived, Math. Z. 236 (2001), no. 2, 351–382. [Bal01b] , Witt cohomology, Mayer-Vietoris, homotopy invariance and the Gersten conjecture, K-Theory 23 (2001), no. 1, 15–30. [Bal05] , Witt groups, Handbook of K-theory. Vol. 1, 2, Springer, Berlin, 2005, pp. 539–576. [BC09] Paul Balmer and Baptiste Calme`s, Geometric description of the connecting homomorphism for Witt groups, Doc. Math. 14 (2009), 525–550. [BC08] , Witt groups of Grassmann varieties (2008), available at www.math.uiuc.edu/ K-theory/0903/. Preprint. [BG05] Paul Balmer and Stefan Gille, Koszul complexes and symmetric forms over the punctured affine space, Proc. London Math. Soc. (3) 91 (2005), no. 2, 273–299. [BGPW02] Paul Balmer, Stefan Gille, Ivan Panin, and Charles Walter, The Gersten conjecture for Witt groups in the equicharacteristic case, Doc. Math. 7 (2002), 203–217. [BW02] Paul Balmer and Charles Walter, A Gersten-Witt spectral sequence for regular schemes, Ann. Sci. E´cole Norm. Sup. (4) 35 (2002), no. 1, 127–152. [Bea96] Arnaud Beauville, Complex algebraic surfaces, 2nd ed., London Mathematical Society Student Texts, vol. 34, Cambridge University Press, Cambridge, 1996. [BGG73] Joseph Bernstein, Israel Gelfand, and Sergei Gelfand, Schubert cells, and the cohomology of the spaces G/P , Uspehi Mat. Nauk 28 (1973), no. 3(171), 3–26 (Russian); English transl., Russian Math. Surveys 28 (1973), no. 3, 1–26. [BO74] Spencer Bloch and Arthur Ogus, Gersten’s conjecture and the homology of schemes, Ann. Sci. E´cole Norm. Sup. (4) 7 (1974), 181–201 (1975). [Bot69] Raoul Bott, Lectures on K(X), Mathematics Lecture Note Series, W. A. Benjamin, Inc., New York-Amsterdam, 1969. [Bro03] Patrick Brosnan, Steenrod operations in Chow theory, Trans. Amer. Math. Soc. 355 (2003), no. 5, 1869–1903. [BG10] Robert Bruner and John Greenlees, Connective real K-theory of finite groups, Mathematical Surveys and Monographs, vol. 169, American Mathematical Society, Providence, RI, 2010. 105 Chapter IV. Cellular Varieties [CF11] Baptiste Calme`s and Jean Fasel, Trivial Witt groups of flag varieties (2011), available at arxiv.org/abs/1103.4412. Preprint. [Del62] Antoine Delzant, De´finition des classes de Stiefel-Whitney d’un module quadratique sur un corps de caracte´ristique diffe´rente de 2, C. R. Acad. Sci. Paris 255 (1962). [Dug01] Daniel Dugger, Universal homotopy theories, Adv. Math. 164 (2001), no. 1, 144–176. [DLØ+07] Bjørn Ian Dundas, Marc Levine, Paul Arne Østvær, Oliver Ro¨ndigs, and Vladimir Voevod- sky, Motivic homotopy theory, Universitext, Springer-Verlag, Berlin, 2007. Lectures from the Summer School held in Nordfjordeid, August 2002. [EKM08] Richard Elman, Nikita Karpenko, and Alexander Merkurjev, The algebraic and geometric theory of quadratic forms, American Mathematical Society Colloquium Publications, vol. 56, American Mathematical Society, Providence, RI, 2008. [EKV93] He´le`ne Esnault, Bruno Kahn, and Eckart Viehweg, Coverings with odd ramification and Stiefel- Whitney classes, J. Reine Angew. Math. 441 (1993), 145–188. [FC87] Fernando Ferna´ndez-Carmena, The Witt group of a smooth complex surface, Math. Ann. 277 (1987), no. 3, 469–481. [Fuj67] Michikazu Fujii, K0-groups of projective spaces, Osaka J. Math. 4 (1967), 141–149. [Ful98] William Fulton, Intersection theory, 2nd ed., Ergebnisse der Mathematik und ihrer Grenzge- biete. 3. Folge. A Series of Modern Surveys in Mathematics, vol. 2, Springer-Verlag, Berlin, 1998. [FH91] William Fulton and Joe Harris, Representation theory, Graduate Texts in Mathematics, vol. 129, Springer-Verlag, New York, 1991. [GM03] Sergei Gelfand and Yuri Manin, Methods of homological algebra, 2nd ed., Springer Monographs in Mathematics, Springer-Verlag, Berlin, 2003. [Gil03] Stefan Gille, Homotopy invariance of coherent Witt groups, Math. Z. 244 (2003), no. 2, 211– 233. [GN03] Stefan Gille and Alexander Nenashev, Pairings in triangular Witt theory, J. Algebra 261 (2003), no. 2, 292–309. [GOV94] Vladimir Gorbatsevich, Arkadi˘ı Onishchik, and E`rnest Vinberg, Lie Groups and Lie Algebras III: Structure of Lie Groups and Lie Algebras, Encyclopaedia of Mathematical Sciences, vol. 41, Springer-Verlag, Berlin, 1994. [Gra75] Brayton Gray, Homotopy theory, Academic Press [Harcourt Brace Jovanovich Publishers], New York, 1975. An introduction to algebraic topology; Pure and Applied Mathematics, Vol. 64. [Har92] Joe Harris, Algebraic geometry, Graduate Texts in Mathematics, vol. 133, Springer-Verlag, New York, 1992. A first course. [Har70] Robin Hartshorne, Ample subvarieties of algebraic varieties, Notes written in collaboration with C. Musili. Lecture Notes in Mathematics, Vol. 156, Springer-Verlag, Berlin, 1970. [Har77] , Algebraic geometry, Springer-Verlag, New York, 1977. Graduate Texts in Mathematics, No. 52. [Hat09] Allen Hatcher, Vector bundles and K-theory, 2009. Version 2.1, available at www.math.cornell. edu/∼hatcher/VBKT/VBpage.html. [Hog69] Stuart Hoggar, On KO theory of Grassmannians, Quart. J. Math. Oxford Ser. (2) 20 (1969), 447–463. [Hor05] Jens Hornbostel, A1-representability of Hermitian K-theory and Witt groups, Topology 44 (2005), no. 3, 661–687. 106 3 Examples [HS04] Jens Hornbostel and Marco Schlichting, Localization in Hermitian K-theory of rings, J. London Math. Soc. (2) 70 (2004), no. 1, 77–124. [Hus94] Dale Husemoller, Fibre bundles, 3rd ed., Graduate Texts in Mathematics, vol. 20, Springer- Verlag, New York, 1994. [HJJS08] Dale Husemo¨ller, Michael Joachim, Branislav Jurcˇo, and Martin Schottenloher, Basic bundle theory and K-cohomology invariants, Lecture Notes in Physics, vol. 726, Springer, Berlin, 2008. [Ish92] Kiminao Ishitoya, Squaring operations in the Hermitian symmetric spaces, J. Math. Kyoto Univ. 32 (1992), no. 1, 235–244. [Jar87] John Frederick Jardine, Simplicial presheaves, J. Pure Appl. Algebra 47 (1987), no. 1, 35–87. [Jou73] Jean-Pierre Jouanolou, Une suite exacte de Mayer-Vietoris en K-the´orie alge´brique, Algebraic K-theory, I: Higher K-theories (Proc. Conf., Battelle Memorial Inst., Seattle, Wash., 1972), Springer, Berlin, 1973, pp. 293–316. Lecture Notes in Math., Vol. 341. [Kar05] Max Karoubi, Bott periodicity in topological, algebraic and Hermitian K-theory, Handbook of K-theory. Vol. 1, 2, Springer, Berlin, 2005, pp. 111–137. [KKO04] Daisuke Kishimoto, Akira Kono, and Akihiro Ohsita, KO-theory of flag manifolds, J. Math. Kyoto Univ. 44 (2004), no. 1, 217–227. [Kne77] Manfred Knebusch, Symmetric bilinear forms over algebraic varieties, Queen’s Univ., Kingston, Ont., 1977, pp. 103–283. Queen’s Papers in Pure and Appl. Math., No. 46. Also available from epub.uni-regensburg.de/12783/. [Koc96] Stanley Kochman, Bordism, stable homotopy and Adams spectral sequences, Fields Institute Monographs, vol. 7, American Mathematical Society, Providence, RI, 1996. [Kol07] Ja´nos Kolla´r, Lectures on resolution of singularities, Annals of Mathematics Studies, vol. 166, Princeton University Press, Princeton, NJ, 2007. [KH91] Akira Kono and Shin-ichiro Hara,KO-theory of complex Grassmannians, J. Math. Kyoto Univ. 31 (1991), no. 3, 827–833. [KH92] , KO-theory of Hermitian symmetric spaces, Hokkaido Math. J. 21 (1992), no. 1, 103– 116. [Lab76] Olivier Laborde, Classes de Stiefel-Whitney en cohomologie e´tale, Bull. Soc. Math. France Suppl. Mem. (1976), no. 48, 47–51. Colloque sur les Formes Quadratiques (Montpellier, 1975). [Laz04] Robert Lazarsfeld, Positivity in algebraic geometry. I, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics, vol. 48, Springer-Verlag, Berlin, 2004. Classical setting: line bundles and linear series. [Lev99] Marc Levine, K-theory and motivic cohomology of schemes (1999), available at www.math. uiuc.edu/K-theory/0336/. Preprint. [May01] J. Peter May, The additivity of traces in triangulated categories, Adv. Math. 163 (2001), no. 1, 34–73. [Mer81] Alexander Merkurjev, On the norm residue symbol of degree 2, Dokl. Akad. Nauk SSSR 261 (1981), no. 3, 542–547. [Mil08] James Milne, Lectures on Etale Cohomology (v2.10) (2008), available at www.jmilne.org/ math/. [Mil69] John Milnor, Algebraic K-theory and quadratic forms, Invent. Math. 9 (1969/1970), 318–344. [MH73] John Milnor and Dale Husemoller, Symmetric bilinear forms, Springer-Verlag, New York, 1973. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 73. 107 Chapter IV. Cellular Varieties [MS74] John Milnor and James Stasheff, Characteristic classes, Princeton University Press, Princeton, N. J., 1974. Annals of Mathematics Studies, No. 76. [Mor99] Fabien Morel, The´orie homotopique des sche´mas, Aste´risque (1999), no. 256. [Mor04] , An introduction to A1-homotopy theory, Contemporary developments in algebraic K- theory, ICTP Lect. Notes, XV, Abdus Salam Int. Cent. Theoret. Phys., Trieste, 2004, pp. 357– 441. [Mor06] , Rational stable splitting of Grassmanians and rational motivic sphere spectrum (2006), available at www.mathematik.uni-muenchen.de/∼morel/preprint.html. Draft. [MV99] Fabien Morel and Vladimir Voevodsky, A1-homotopy theory of schemes, Inst. Hautes E´tudes Sci. Publ. Math. (1999), no. 90, 45–143 (2001). [Nen07] Alexander Nenashev, Gysin maps in Balmer-Witt theory, J. Pure Appl. Algebra 211 (2007), no. 1, 203–221. [Nen09] , On the Witt groups of projective bundles and split quadrics: geometric reasoning, J. K-Theory 3 (2009), no. 3, 533–546. [OVV07] Dmitri Orlov, Alexander Vishik, and Vladimir Voevodsky, An exact sequence for KM∗ /2 with applications to quadratic forms, Ann. of Math. (2) 165 (2007), no. 1, 1–13. [PW10] Ivan Panin and Charles Walter, On the motivic commutative ring spectrum BO (2010), avail- able at www.math.uiuc.edu/K-theory/0978/. Preprint. [Par04] William Pardon, The filtered Gersten-Witt resolution for regular schemes (2004), available at www.math.uiuc.edu/K-theory/0419/. Preprint. [PW01] Claudio Pedrini and Charles Weibel, The higher K-theory of a complex surface, Compositio Math. 129 (2001), no. 3, 239–271. [Qui73] Daniel Quillen, Higher algebraic K-theory. I, Algebraic K-theory, I: Higher K-theories (Proc. Conf., Battelle Memorial Inst., Seattle, Wash., 1972), Springer, Berlin, 1973, pp. 85–147. Lec- ture Notes in Math., Vol. 341. [Rio06] Joe¨l Riou, Ope´rations sur la K-the´orie alge´brique et re´gulateurs via la the´orie homotopique des sche´mas, Ph.D. Thesis, l’Universite´ Paris 7 – Denis Diderot, 2006, www.math.u-psud.fr/ ∼riou/these/these.pdf. [Rio07a] , Cate´gorie homotopique stable d’un site suspendu avec intervalle, Bull. Soc. Math. France 135 (2007), no. 4, 495–547. [Rio07b] , Ope´rations sur la K-the´orie alge´brique et re´gulateurs via la the´orie homotopique des sche´mas, C. R. Math. Acad. Sci. Paris 344 (2007), no. 1, 27–32. [Sch10a] Marco Schlichting, Hermitian K-theory of exact categories, J. K-Theory 5 (2010), no. 1, 105– 165. [Sch10b] , The Mayer-Vietoris principle for Grothendieck-Witt groups of schemes, Invent. Math. 179 (2010), no. 2, 349–433. [Sch] , Hermitian K-theory, derived equivalences and Karoubi’s fundamental theorem. In preparation. [Sch10] , The homotopy limit problem and (e´tale) hermitian K-theory (2010), available at arxiv.org/abs/1011.4977. Preprint. [Seg68] Graeme Segal, Equivariant K-theory, Inst. Hautes E´tudes Sci. Publ. Math. (1968), no. 34, 129–151. [Sus95] Andrei Suslin, Algebraic K-theory and motivic cohomology, 2 (Zu¨rich, 1994), Birkha¨user, Basel, 1995. 108 3 Examples [TT90] Robert Wayne Thomason and Thomas Trobaugh, Higher algebraic K-theory of schemes and of derived categories, The Grothendieck Festschrift, Vol. III, Progr. Math., vol. 88, Birkha¨user Boston, Boston, MA, 1990, pp. 247–435. [Tot03] Burt Totaro, Non-injectivity of the map from the Witt group of a variety to the Witt group of its function field, J. Inst. Math. Jussieu 2 (2003), no. 3, 483–493. [Voe98] Vladimir Voevodsky, A1-homotopy theory, Proceedings of the International Congress of Math- ematicians, Vol. I (Berlin, 1998), 1998, pp. 579–604. [Voe03a] , Reduced power operations in motivic cohomology, Publ. Math. Inst. Hautes E´tudes Sci. (2003), no. 98, 1–57. [Voe03b] ,Motivic cohomology with Z/2-coefficients, Publ. Math. Inst. Hautes E´tudes Sci. (2003), no. 98, 59–104. [Wal03a] Charles Walter, Grothendieck-Witt groups of triangulated categories (2003), available at www. math.uiuc.edu/K-theory/0643/. Preprint. [Wal03b] , Grothendieck-Witt groups of projective bundles (2003), available at www.math.uiuc. edu/K-theory/0644/. Preprint. [Wei94] Charles A. Weibel, An introduction to homological algebra, Cambridge Studies in Advanced Mathematics, vol. 38, Cambridge University Press, Cambridge, 1994. [Yag04] Nobuaki Yagita, Coniveau spectral sequence and motivic cohomology of quadrics and classifying spaces (2004), available at www.math.uiuc.edu/K-theory/0709/. Preprint. [Yag09] , Note on Witt group and KO-theory of complex Grassmannians (2009), available at www.math.uni-bielefeld.de/LAG/man/325.html. Preprint. [Zib09] Marcus Zibrowius, Comparing Grothendieck-Witt groups of a complex variety to its real topo- logical K-groups (2009), available at www.dspace.cam.ac.uk/handle/1810/236624. Smith- Knight/Rayleigh-Knight Essay. 109