Clifford Geometry: A Seminar Joseph C. Va´rilly and Jose´ M. Gracia-Bond´ıa Universidad de Costa Rica, 2060 San Jose´, Costa Rica Diciembre de 1995 Contents 1 Vector bundles and their classification 4 1.1 Generalities on manifolds 4 1.2 Principal bundles and vector bundles 5 1.3 Hermitian vector bundles 7 1.4 Operations on vector bundles 7 1.5 Equivalent bundles 8 1.6 Sections of vector bundles 9 1.7 Local sections and transition functions 10 1.8 Cˇech cocycles 12 1.9 The Cˇech cohomology of S2 13 1.10 Line bundle classification 14 1.11 The second cohomology group 15 1.12 Classification of Hermitian line bundles 16 1.13 Classification of vector bundles 17 2 Complex projective spaces 19 2.1 Complex manifolds 19 2.2 Local charts for complex projective spaces 20 2.3 The Ka¨hler form 21 2.4 The Fubini–Study metric 22 2.5 The Riemann sphere 24 3 The de Rham complex and Hodge duality 24 3.1 The de Rham complex 24 3.2 The Riemannian volume form 25 3.3 The Hodge star operator 26 3.4 The Hodge Laplacian 31 1 4 The Hodge Laplacian on the 2-sphere 33 4.1 The rotation group in three dimensions 33 4.2 The Hodge operators on the sphere 35 4.3 Eigenvectors for the Laplacian 36 4.4 Spectrum of the Hodge Laplacian 38 4.5 The Hodge–Dirac operator 40 5 Connections on vector bundles 42 5.1 Modules of vector-valued forms 42 5.2 Connections 45 5.3 Curvature of a connection 46 5.4 A curvature formula 47 5.5 From de Rham cohomology to Cech cohomology 48 5.6 Line bundles over CPm 51 5.7 Connections on the hyperplane bundle 53 5.8 Characteristic classes 54 5.9 Chern classes and the Chern character 57 5.10 Classification of line bundles over CPm 60 5.11 The Levi-Civita connection on the tangent bundle 61 6 Clifford algebras 62 6.1 Superspaces and superalgebras 62 6.2 Clifford algebras 65 6.3 Clifford actions 66 6.4 Complex Clifford algebras 68 6.5 The Fock space of spinors 70 6.6 The Pin and Spin groups 72 6.7 The spin representation 76 7 Global Clifford modules 77 7.1 Clifford algebra bundles 77 7.2 Existence of spin structures 80 7.3 Spinc structures 83 7.4 The spinor module 85 7.5 The spin connection 86 7.6 Local coordinate formulas 89 8 Dirac operators and Laplacians 91 8.1 Connections and differential forms 91 8.2 Divergence of a vector field 93 8.3 The Hodge–Dirac operator revisited 94 8.4 Dirac operators 95 8.5 Laplacians 98 2 8.6 The Lichnerowicz formula 101 9 The Dirac operator on the Riemann sphere 102 9.1 Coordinates on the Riemann sphere 102 9.2 Sections and gauge transformations 104 9.3 The spin connection over the sphere 106 9.4 The Dirac operator over the sphere 107 9.5 The spinor Laplacian 110 9.6 The SU(2) action on the spinor bundle 112 9.7 Equivariance of the Dirac operator 115 9.8 Angular momentum operators 117 9.9 Spinor harmonics 119 9.10 The spectrum of the Dirac operator 122 10 Construction of representations of SU(2) 124 10.1 Characters of the maximal torus 124 10.2 Twisting connections 126 10.3 Twisted Dirac operators 127 10.4 The group action on the twisted spinor bundles 128 10.5 The Borel–Weil theorem 131 A Calculus on manifolds 135 A.1 Differential manifolds 135 A.2 Tangent spaces 136 A.3 Vector fields 136 A.4 Lie groups 138 A.5 Fibre bundles 139 A.6 Tensors and differential forms 141 A.7 Calculus of differential forms 142 A.8 The de Rham complex 145 A.9 Volume forms and integrals 146 3 1 Vector bundles and their classification The procedure commonly known as geometric quantization associates a Hilbert space to certain symplectic manifolds, and thereby allows a bridge to be built from the algebra of functions on the manifold to an operator algebra. It has been used, with considerable success, to reconstruct unitary representations of Lie groups from their symplectic homogeneous spaces; this is known as the Kirillov orbit method. An alternative method of constructing these representations, at least for compact groups, arises from the study of Dirac operators on these same homogeneous manifolds. The purpose of these notes is to clarify the roˆle of the Dirac operator. Hopefully, this will enable us to build a bridge between geometric quantization and noncommutative geometry, in which the key concept is a generalization of the Dirac operator to “noncommutative manifolds”. The main geometric objects with which we shall be concerned are vector bundles with some extra algebraic structure, such as a metric or a symplectic structure (or both). Roughly speaking, a vector bundle consists of the disjoint union E of a collection of isomorphic vector spaces Ex, indexed by the points of a differential manifold M , together with some smoothness conditions relating E and M ; an example is the tangent bundle of the manifold M . The algebraic structure of the “fibres” Ex transfers to give an interesting algebraic structure to the manifold E. The set of all possible vector bundles of a given species on a manifold M , up to a suitable equivalence relation, holds important information about the topology of the manifold M (in fact, it holds all the topological information we need to know for the purposes of noncommutative geometry) and we begin our study with this classification. We have gathered in Appendix A a compendium of definitions and basic facts on differ- ential manifolds, vector fields and differential forms, with which we assume the reader to be acquainted. We refer to it for the notations used below. 1.1 Generalities on manifolds Recall that a differentiable manifold of (real) dimension n is a (paracompact) topological space M provided with an atlas, i.e., a collection of charts (Uj, φj), where the domains Uj form a (locally finite) open covering U of M , each φj : Uj → Rm is a homeomorphism, and the “transition maps” φi ◦ φ−1j are smooth functions on each φj(Ui ∩ Uj). If n = 2m, we can regard R2m as Cm; we say that M is a complex manifold if the transition maps are holomorphic (as multi-variable vector-valued complex functions); it suffices that they satisfy the Cauchy–Riemann equations. We shall be mainly interested in the case where M is compact; then its differentiable structure is defined by a finite atlas. A useful property of the chart domains Uj is that they be contractible, that is, that there exist x0 ∈ Uj and a smooth function f : [0, 1] × Uj → Uj with f(0, x) = x0, f(1, x) = x; informally, Uj may be “deformed” to a point {x0}. The intersection of two such domains need not be contractible: just think of the sphere S2 covered by two large polar caps which overlap at the equator. It is good to know, however, that by using another (equivalent) atlas, we can avoid this difficulty. What we need is that the chart domains form a “good covering” 4 of M [12]. Definition 1.1. An open covering U := {Uj} of a topological space is a good covering1 if every nonempty finite intersection Uj1 ∩ · · · ∩ Ujr is contractible. A differentiable manifold M always has a good covering: take a Euclidean metric on M and cover M by an atlas for which each Uj is “geodesically convex” (that is, any two of its points can be connected by a minimal geodesic lying within Uj). 2 All finite intersections are also geodesically convex, and hence are contractible (since one can deform to a point x0 by retreating along geodesics emanating from that point). So we can replace the original atlas of M by a (possibly larger) atlas whose domains form a good covering. From now on, we will assume that this has been done. Exercise 1.1. Show that the following recipe defines a good covering {U1, U2, U3, U4} of the sphere S2. Take four points on S2 that are vertices of a regular tetrahedron (e.g., if a = 1/ √ 3, take (±a,±a,±a) with either one or three positive signs). Connect these points by six great- circle arcs, which determine four spherical triangles Fj. Finally, let Uj := {x ∈ S2 : d(x, Fj) <  } for some small  (say,  = 1/4). 1.2 Principal bundles and vector bundles Recall that a fibre bundle E pi−→M or simply E−→M (see Appendix A for notation) is a triple (E,M, pi), where the manifolds E and M are its total space and base space respectively, and pi : E → M is a surjective submersion, subject to two conditions: (a) that each fibre Ex := pi −1({x}) is diffeomorphic to a fixed manifold F (the “typical fibre”), and (b) for some atlas {(Uj, φj)} of M there is a family of local trivializations of E, i.e., diffeomorphisms ψj : pi −1(Uj)→ Uj × F (1.1) such that pi(ψ−1j (x, v)) = x for all x ∈ Uj, v ∈ F . Definition 1.2. Let G be a Lie group. A principal G-bundle P η−→M is a fibre bundle whose fibres are diffeomorphic to G, together with a free right action of G on the total space P , whose orbits are the fibres Px = η −1(x). We use the notation χj : η−1(Uj)→ Uj ×G for the local trivializations. Exercise 1.2. Suppose that G is a Lie group and that H is a closed subgroup of G (so that H is also a Lie group); then H acts freely on G by right multiplication g · h := gh. If η : G→ G/H is the quotient map, check that G η−→G/H is a principal H-bundle. Before examining where principal bundles come from, let us describe a general recipe for manufacturing new fibre bundles using a given principal bundle. Indeed, this recipe is the main reason why principal bundles are used at all. 1This is also known as a “contractible covering” [38], or a “Leray covering” [58]. 2It is a basic proposition of Riemannian geometry that each point has a geodesically convex neighbour- hood; see [32] or [36] for the proof. 5 Definition 1.3. Let P η−→M be principal G-bundle, and let F be a manifold on which the Lie group G acts on the left. The product manifold P × F carries a right action of G, given by (p, v) · g := (p · g, g−1 · v). (1.2) Let E := P ×G F be the set of orbits of this action; we denote the orbit of (p, v) by [p, v]. Notice that [p · g, v] = [p, g · v] on account of (1.2). Define pi : E →M : [p, v] 7→ η(p). For each p ∈ P , the map v 7→ [p, v] is a diffeomorphism from F to Eη(p), and E pi−→M is a fibre bundle with typical fibre F , which is said to be associated to the given principal G-bundle. Definition 1.4. A vector bundle E pi−→M is a fibre bundle whose “typical fibre” is a (real or complex) vector space V , such that each fibre Ex is a vector space of dimension dimV , and such that for each x ∈ Uj, the map v 7→ ψ−1j (x, v) is a linear isomorphism from V onto Ex. A real vector bundle with typical fibre V = R is called a real line bundle; a complex vector bundle with typical fibre V = C is called a complex line bundle. When we say simply “line bundle”, we shall usually mean a complex line bundle. Examples of (real) vector bundles over M are the tangent bundle TM −→M , the cotan- gent bundle T ∗M −→M , and its exterior powers ΛrT ∗M −→M : see Appendix A. Given a representation ρ : G→ GL(V ) of a Lie group, and a principal G-bundle P η−→M , we can form the associated bundle by taking g · v ≡ ρ(g)v; thus [p · g, v] = [p, ρ(g)v] for p ∈ P, v ∈ V, g ∈ G. (1.3) The fibres Ex become vector spaces by taking [p, u]+ [p, v] := [p, u+v], and λ[p, v] := [p, λv]; and the linearity of each ρ(g) shows that E pi−→M is a vector bundle with typical fibre V . We can now reverse the recipe of Definition 1.3 in order to associate a principal bundle to a given vector bundle: Definition 1.5. Let E pi−→M be a vector bundle3 with typical fibre V . Let Px, for x ∈M , denote the set of linear isomorphisms p : V → Ex (each such p is called a frame for Ex). If g ∈ GL(V ), then p ◦ g is again a frame, so that GL(V ) acts on the right (freely and transitively) on Px by composition p 7→ p ◦ g. The disjoint union P = ⊎ x∈M Px, with η(p) := x for p ∈ Px, defines a principal GL(V )-bundle P η−→M , which is called the frame bundle of E pi−→M . Exercise 1.3. Using the identity representation of the group GL(V ) on V , check that the vector bundle associated to the frame bundle P η−→M by the recipe (1.2) is the original vector bundle E pi−→M . 3We use a notation which covers the real and complex cases simultaneously; thus GL(V ) denotes either GLR(V ) or GLC(V ), according as V is a real or complex vector space. 6 1.3 Hermitian vector bundles If V carries some extra structure (e.g., an inner product or an orientation), we can restrict to the subgroupG ≤ GL(V ) which preserves that structure. We need to impose a corresponding structure on the fibres Ex and consider only those frames p : V → Ex which are structure- preserving. Under p 7→ p ◦ g, these comprise a principal G-bundle associated to E pi−→M . More generally, if we are given a Lie group G with a representation ρ : G → GL(V ), we define a right action of G on frames by p · g := p ◦ ρ(g). If, and sometimes this is “a big if”, we can select a subset of frames for each Ex for which this action is transitive and free (due to some extra structure given on the vector bundle), these will form the fibres Qx of a principal G-bundle Q η′−→M . Since [p · g, v] = [p, ρ(g)v] for p ∈ Q, v ∈ V , the vector bundle associated, via ρ, to the new principal bundle is still E pi−→M . For instance, we could ask that a real vector bundle E−→M be Euclidean, i.e., that each fibre Ex carry a real inner product gx(·, ·), which depends smoothly on x. If V is a real vector space with a positive-definite inner product q(·, ·), we consider “orthogonal frames” p satisfying gx(p(u), p(v)) = q(u, v) for all u, v ∈ V . It is clear that such frames form a principal O(n)-bundle associated to E pi−→M . Such an isomorphism p is determined by choosing an orthonormal basis in (Ex, gx) (as the image under p of a fixed orthonormal basis in (V, q)), so that this bundle is often called the “orthonormal frame bundle”. Alternatively, if E pi−→M is a complex vector bundle, we could ask that it be Hermitian, i.e., that each fibre Ex carry a (sesquilinear) inner product hx(·, ·), depending smoothly on x. Then if V is a complex Hilbert space with inner product 〈· | ·〉, we consider “unitary frames” p satisfying hx(p(u), p(v)) = 〈u | v〉 for u, v ∈ V . Such frames form a principal U(n)-bundle associated to E pi−→M . 1.4 Operations on vector bundles Definition 1.6. Given any (real or complex) vector bundle E−→M , one can form a dual vector bundle E∗−→M whose fibre E∗x is the dual vector space4 of Ex. Fix a basis {v1, . . . , vn} for the typical fibre V . A frame p : V → Ex is defined by selecting a basis {e1, . . . , en} for Ex and setting p(vj) := ej for each j; matching the dual bases {v′1, . . . , v′n}, {e′1, . . . , e′n} gives a frame p′ : V ∗ → E∗x for the dual bundle. Notice that (p ◦ g)′ = p′ ◦ g−t by change-of-basis formulae, where g−t := (g−1)t is the contragredient matrix to g ∈ GL(V ). Exercise 1.4. Verify that E∗−→M is the vector bundle associated to the frame bundle of E−→M , via the representation ρ(g) := g−t of GL(V ) on V ∗. Exercise 1.5. Check that the cotangent bundle T ∗M −→M is the dual bundle to the tangent bundle TM −→M . Definition 1.7. Given any two vector bundles E−→M , E ′−→M over the same base space, we can form two new vector bundles over M : their Whitney sum E ⊕ E ′−→M and their 4The dual space of V is V ∗ := HomR(V,R) or V ∗ := HomC(V,C), the space of R-linear or C-linear forms on V , according as V is a real or complex vector space. 7 tensor product E⊗E ′−→M , whose fibres at x ∈M are respectively the direct sum Ex⊕E ′x and the algebraic tensor product Ex ⊗ E ′x. The k-th exterior power of E−→M is the vector bundle ΛkE−→M whose fibre at x is ΛkEx. (For k = 0, we take Λ 0E := M × R.) These operations can be combined; for instance, the exterior algebra bundle Λ•E−→M is the Whitney sum of all exterior powers, for k = 0, 1, . . . , r, where r = dimEx is the rank of E−→M . Definition 1.8. The complexification of a real vector bundle E−→M is the complex vector bundle EC−→M with EC := E ⊗R C, i.e., (Ex)C := Ex ⊗R C = Ex ⊕ i Ex for each x ∈M . We write TCM and T ∗ CM for the total spaces of the complexified tangent and cotangent bundles of M . 1.5 Equivalent bundles Definition 1.9. A morphism of two fibre bundles E pi−→M and E ′ pi′−→M ′ is a pair of smooth maps (τ, σ), with τ : E → E ′ and σ : M → M ′, such that pi′ ◦ τ = σ ◦ pi, i.e., such that the following diagram commutes: E τ−−−→ E ′ pi y ypi′ M σ−−−→ M ′ and, in particular, τ(Ex) ⊆ E ′σ(x) for each x ∈M . A morphism of vector bundles is a bundle morphism for which τ : Ex → E ′σ(x) is linear. When M ′ = M , we usually take σ = idM . Definition 1.10. Let E pi−→M be a vector bundle and let φ : N → M be a smooth map. Write φ∗E := { (u, y) ∈ E ×N : pi(u) = φ(y) } and define p¯i : φ∗E → N and φ˜ : φ∗E → E by p¯i(u, y) := y and φ˜(u, y) := u. Then p¯i−1(y) = Eφ(y), and so φ∗E p¯i−→N is a vector bundle, called the pullback bundle of E pi−→M via φ. Moreover, (φ˜, φ) is a bundle morphism; in other words, we have a commutative diagram of smooth maps: φ∗E φ˜−−−→ E p¯i y ypi N φ−−−→ M Exercise 1.6. Show that the pullback bundle has the following universal property: if E ′ pi ′−→N is a vector bundle over N and if ρ : E ′ → E is a map such that (ρ, φ) is a bundle morphism from this bundle to E pi−→M , then there is a unique bundle morphism (τ, idN), with τ : E ′ → φ∗E, so that ρ = φ˜ ◦ τ . (In other words, any bundle morphism with base map φ factors through the pullback bundle.) 8 Definition 1.11. Let E pi−→M , E ′ pi′−→M be two vector bundles over the same base space. A vector bundle equivalence between them is an invertible vector bundle morphism (τ, id), which is given by a diffeomorphism τ : E → E ′ satisfying pi′ ◦ τ = pi and such that τ : Ex → E ′x is a linear isomorphism for each x. We shall denote by [E] the equivalence class of the vector bundle E pi−→M ; for the set of equivalence classes with typical fibre V , we write Vect(M ;V ). A vector bundle E−→M is trivial if it is equivalent to the product bundle M×F pr1−→M . Note, from Exercise 1.6, that the pullback bundle is unique up to equivalence, and so defines a unique [φ∗E] ∈ Vect(N ;V ). Exercise 1.7. Show that the tangent bundle TS1 of the unit circle S1 is trivial, by producing a pair of local trivializations which together form a cylinder. Exercise 1.8. Let E−→M be an arbitrary vector bundle over M ; write Er = M × V with dimV = r, so that Er−→M denotes the trivial vector bundle of rank r. Show that E ⊗Er−→M is equivalent to the Whitney sum E ⊕ · · · ⊕E−→M of r copies of E−→M . Definition 1.12. Let P η−→M , P ′ η′−→M be two principal G-bundles over the same base space. An equivalence between them is an invertible bundle morphism (χ, id) which is G- equivariant, that is, χ(p·g) = χ(p)·g for any p ∈ P . Note that χ : P → P ′ is a diffeomorphism. We shall denote by [P ] the equivalence class of P η−→M ; for the set of such equivalence classes we write Prin(M ;G). Proposition 1.1. Two vector bundles over M with the same typical fibre V are equivalent if and only if their frame bundles are equivalent as principal GL(V )-bundles. The association recipe thus yields a bijection [E]↔ [P ] between Vect(M ;V ) and Prin(M ;GL(V )). Exercise 1.9. Prove this, using the defining relation (1.3) of an associated vector bundle. 1.6 Sections of vector bundles Definition 1.13. A smooth section of a vector bundle E pi−→M is a smooth map s : M → E such that pi ◦ s = idM , i.e., s(x) ∈ Ex for each x ∈M . The totality of smooth sections will be denoted by Γ(E), or by Γ(M,E) if it is necessary to specify the base space M . Notice that Γ(E) is a module for the commutative algebra of functions C∞(M); the action of C∞(M) is just scalar multiplication in each fibre:5 (fs)(x) := f(x)s(x), for s ∈ Γ(E), f ∈ C∞(M). 5It would perhaps be more convenient, in view of an eventual translation to the language of noncom- mutative geometry, to write the multiplication on the right : (sf)(x) = s(x)f(x); but as this conflicts with traditional habits, and could be confusing in the case that s is a vector field, we will for the moment retain the usual notation of multiplying by scalars f(x) on the left. 9 “Global” smooth sections s : M → E can sometimes be hard to find. For instance, a line bundle admits a nonvanishing global section only if it is trivial. To see this, notice that any v ∈ Ex is of the form λs(x), for a unique λ ∈ C, since s(x) 6= 0; so λs(x) 7→ (x, λ) is a vector bundle equivalence between E and the trivial line bundle M × C.6 Lemma 1.2. If L−→M is a line bundle, then its tensor product with its dual line bundle, namely L⊗ L∗−→M , is a trivial line bundle. Proof. Indeed, Lx⊗L∗x ' End(Lx), so the tensor product bundle has an obvious nonvanishing section s0, whose value at each x is the identity operator on the line Lx. Since each End(Lx) is a one-dimensional vector space, any L ⊗ L∗−→M is a line bundle, with a nonvanishing global section. Corollary 1.3. The set of equivalence classes of line bundles over M has the structure of an abelian group. Proof. Let L−→M , L′−→M be any two line bundles over M , and define: [L] [L′] := [L⊗ L′], [L]−1 := [L∗]. (1.4) Let L0 := M × C so that [L0] is the trivial bundle class,7 and note that [L ⊗ L0] = [L] by Exercise 1.8. Since [L ⊗ L∗] = [L0] by the previous Lemma, the inverse of [L] is [L∗]. Moreover, the flip map u⊗v 7→ v⊗u from Lx⊗L′x to L′x⊗Lx determines a bundle equivalence between L⊗ L′ and L′ ⊗ L, so that the product (1.4) is commutative. We shall soon identify this group with a cohomology group of M . 1.7 Local sections and transition functions The question that now arises is how to deal with bundles that are not trivial, and how to give an effective description of such bundles. Since there are no nonvanishing global sections, we must make use of nonvanishing local sections sj ∈ Γ(Uj, E), where U := {Uj} is an open covering of M by chart domains which admit local trivializations ψj as in (1.1). Thus ψj(sj(x)) ≡ (x, fj(x)) (1.5) where fj : Uj → V \{0} is a smooth nonvanishing function. (Indeed, to say that sj is smooth is the same as saying that its local representative fj is a smooth function.) A global section s ∈ Γ(M,E) is determined, via (1.5), by such a family of local represen- tatives fj : Uj → V (which may now take zero values). Let r = dimV be the rank of the vector bundle E pi−→M . Suppose we can find a set sj = (sj1, . . . , sjr) with each sjk ∈ Γ(Uj, E) so that {sj1(x), . . . , sjr(x)} is a basis for the 6Here we are falling into the sloppy habit of referring to a fibre bundle by naming its total space only. If the base space M is fixed, this does no harm. 7We use complex line bundles only to be specific; the argument for real line bundles is identical. 10 fibre Ex at each x ∈ Uj. Then if tj ∈ Γ(Uj, E) is any smooth local section, we can find smooth functions h1j , . . . , h r j ∈ C∞(Uj) such that tj = ∑r k=1 h k j sjk on Uj. Thus Γ(Uj, E) is determined by the set of local sections sj; and globally, the module Γ(M,E) is determined by a family {(Uj, sj)} of such sets, one for each local chart of M . Such a family is sometimes called a local system of sections [38] for the vector bundle E−→M . Note that sj is a local frame over Uj and may be regarded as a local section of the frame bundle P η−→M . To be precise, choose and fix a basis {v1, . . . , vr} for V , and let pj ∈ Px, for x ∈ Uj, be the linear isomorphism from V to Ex determined by pj(vk) := sjk(x) for k = 1, . . . , r. Then pj ∈ Γ(Uj, P ). Conversely, any such local section pj determines the local frame sj. Suppose that (Ui, si) is another local frame and that the chart domains Ui and Uj overlap: Ui ∩ Uj 6= ∅. Then for x ∈ Ui ∩ Uj, the isomorphisms pi, pj from V to Ex determined by pi(vk) := sik(x), pj(vk) := sjk(x) are related by pj = pi ◦ gij(x) for some gij(x) ∈ GL(V ). In the notation of associated bundles, we have [pj, v] = [pi ◦ gij(x), v] = [pi, gij(x)v] (1.6) for v ∈ V ; or equivalently, ψi ◦ ψ−1j (x, v) = (x, gij(x)v), (1.7) i.e., gij is the expression in local coordinates of the transition between the local trivializations ψi, ψj of the vector bundle. Thus each gij : Ui ∩ Uj → GL(V ) is a smooth function. Exercise 1.10. Show that a principal bundle P η−→M is trivial if and only if it admits a global smooth section q : M → P . Definition 1.14. Let E−→M be a vector bundle, with typical fibre V , for which { (Uj, sj) : j ∈ J } is a local system of pointwise linearly independent local sections. The family of smooth functions gij : Ui∩Uj → GL(V ), defined whenever Ui∩Uj 6= ∅, such that si = gij ·sj on Ui ∩ Uj, satisfies the consistency conditions gii = id on Ui, gijgjk = gik on Ui ∩ Uj ∩ Uk. (1.8) (The notation gij · sj denotes the natural action of GL(V ) on E, as expressed by (1.6) or (1.7).) The set {gij} is called a family of transition functions for the vector bundle E−→M . Suppose now that the vector bundle E−→M carries extra structure, for instance a Hermitian metric. Then we can assume that the basis sj(x) for Ex respects this structure (continuing the Hermitian-metric example, we may take it to be an orthonormal basis, with sjk(x) = pj(vk) for a fixed orthonormal basis of V ). Then the transition functions gij map Ui ∩ Uj into the subgroup G of GL(V ) which preserves the appropriate structure (in our example, the unitary group of V ). In summary, every gij(x) belongs to the structure group G of the principal bundle P η−→M to which the vector bundle is associated. This suggests that we should study vector bundles by first classifying the corresponding frame bundles, and then invoking Proposition 1.1 to pass the classification to vector bundles. 11 This procedure works because the frame bundle is entirely determined by the transition functions. Specifically, we have the following “patching-together” construction. Lemma 1.4. Let M be a manifold with an atlas of local charts { (Uj, φj) : j ∈ J }; let G be a Lie group and suppose we are given a family of smooth functions gij : Ui∩Uj → G, defined for Ui ∩ Uj 6= ∅, that satisfies (1.8). Then there exists a principal G-bundle P η−→M and a family of local sections pj ∈ Γ(Uj, P ) such that pi(x) · gij(x) = pj(x) whenever x ∈ Ui ∩ Uj. Proof. Let Q denote the disjoint union ⊎ j∈J Uj × G and let P be the quotient space of Q formed by identifying (x, h)i ∈ Ui × G with (x, gij(x)h)j ∈ Uj × G whenever x ∈ Ui ∩ Uj. The condition (1.8) simply says that (x, h)i ∼ (x, gij(x)h)j is an equivalence relation on Q, so the quotient space is well defined. Write η[(x, h)i] := x; then it may be checked that P inherits from Q the structure of a differential manifold, of dimension dimM + dimG, such that η : P → M is a submersion. It remains to check that G acts freely and transitively on the right on each fibre η−1(x); but it is obvious that the right action of G on Q given by (x, h)i · g := (x, hg)i preserves equivalence classes and drops to a right action on P whose orbits are the fibres of η. If we are also given a representation ρ : G→ GL(V ), we may then create a vector bundle with transition functions {gij} by association, using (1.3). Conclusion: one can always patch together a vector bundle with base M and structure group G from a set of transition functions satisfying (1.8) and a representation of G. The remaining question is how can one describe the equivalence of vector bundles in terms of the transition functions, in order to obtain a manageable classification. The answer is provided by the theory of Cˇech cohomology. 1.8 Cˇech cocycles A family of transition functions forms what is called a “Cˇech 1-cocycle” with values in the structure group G. Some difficulties arise in the cohomology theory of these objects for noncommutative structure groups, so we shall assume for the present that this group is abelian. This still covers many cases of interest, such as the groups C, C×, R, U(1), Z, and Z2. Definition 1.15. Let U = {Uj : j ∈ J } be an open covering of a manifold M and let A be an abelian group; we will write the group operation additively. For r ∈ N, a Cˇech r-cochain over U with coefficients in A is a family of elements cj0j1...jr ∈ A, indexed by the collections of (r+ 1) sets {Uj0 , Uj1 , . . . , Ujr} ⊂ U for which Uj0 ∩Uj1 ∩ · · · ∩Ujr 6= ∅. The set of all r-cochains is denoted Cr(U, A); it is an abelian group. One often needs a broader definition, where A is replaced by a family of abelian groups A := {Aj0j1...jr} satisfying certain compatibility relations.8 We will always take Aj0j1...jr to be 8To be precise, A should be a sheaf of abelian groups over M . The interested reader may consult [14] or [58] for the full story. 12 the collection of smooth functions from Uj0 ∩Uj1 ∩· · ·∩Ujr into a fixed abelian group A. For r ∈ N, a Cˇech r-cochain over U with coefficients in A is then a family of smooth functions fj0j1...jr : Uj0 ∩Uj1 ∩ · · · ∩Ujr → A, defined whenever Uj0 ∩Uj1 ∩ · · · ∩Ujr 6= ∅. The set of all such families is denoted Cr(U, A). For instance, the chart maps φj of a manifold M form a 0-cochain in C 0(U,Rn) (or in C0(U,Cm), if M is a complex manifold). The transition functions of a complex line bundle gij form a 1-cochain in C 1(U,C×); the transition functions of a Hermitian line bundle form a 1-cochain in C1(U, U(1)). Definition 1.16. The Cˇech complex is the cochain complex9 (C•(U, A), δ) is defined as follows. The coboundary operator δr : C r(U, A)→ Cr+1(U, A) is given by (δa)ij := ai − aj, (δb)ijk := bij − bik + bjk for a ∈ C0(U, A), b ∈ C1(U, A), and (δc)j0...jr := ∑r k=0(−1)kcj0...jr−k−1jr−k+1...jr in general. It is immediate that δ2 = 0 by cancellation of terms, so we have a complex C0(U, A) δ−→C1(U, A) δ−→C2(U, A) δ−→· · · which gives rise to cohomology groups in the standard way: Zr(U, A) := { c ∈ Cr(U, A) : δc = 0 } are the Cˇech cocycles, Br(U, A) := { δb : b ∈ Cr−1(U, A) } are the Cˇech cobound- aries, and the quotient group Hr(U, A) := Zr(U, A)/Br(U, A) is the rth Cˇech cohomology group of the covering U with coefficients in A. Open coverings of M form a directed set (under refinement); we can eliminate U by taking a direct limit: the rth Cˇech cohomology group of the manifold M (with coefficients in A) is defined as10 Hˇr(M,A) := lim−→ U Hr(U, A). An essential result from topology [14] is that this limit is already attained when U is a good covering: Hˇr(M,A) = Hr(U, A) in this case. 1.9 The Cˇech cohomology of S2 Let us compute the Cˇech cohomology (with real coefficients) of the sphere S2. We know (Exercise 1.1) that it has a good covering U = {U1, U2, U3, U4} by open neighbourhoods of the four spherical triangles Fj obtained by projecting an inscribed regular tetrahedron outward from the centre. Each Ui ∩ Uj is a neighbourhood of the edge Fi ∩ Fj, and each Ui ∩ Uj ∩ Uk is a neighbourhood of the vertex Fi ∩ Fj ∩ Fk; all are contractible. Thus 0- cochains are labelled by the faces of the tetrahedron, 1-cochains are labelled by its edges, and 2-cochains are labelled by its vertices. Therefore C0(U,R) ' R4, C1(U,R) ' R6, C2(U,R) ' R4, 9See Appendix A for generalities on cochain complexes and their cohomology groups. 10We use the notation Hˇ to distinguish Cˇech cohomology from singular or de Rham cohomology; although we shall see that this is often unnecessary. 13 and since U1 ∩ U2 ∩ U3 ∩ U4 = ∅, we have Cr(U,R) = 0 for r ≥ 3. Thus the Cˇech complex reduces to C0(U,R) δ0−→C1(U,R) δ1−→C2(U,R) δ2−→ 0. (1.9) Now a ∈ ker δ0 iff a1 = a2 = a3 = a4, so Hˇ0(S2,R) = Z0(U,R) ' R; and (by linearity of δ0) B1(U,R) = im δ0 ' R3. The elements b ∈ ker δ1 = Z1(U,R) satisfy bij − bik + bjk = 0 for {i, j, k} ⊂ {1, 2, 3, 4}; these 4 linear equations for the six bij form a system of rank 3; and hence Z 1(U,R) ' R3. Thus the sequence (1.9) is exact at C1(U,R), and so Hˇ1(S2,R) = 0. Finally, B2(U,R) = im δ1 ' R6/ ker δ1 ' R3. Since Z2(U,R) = C2(U,R) ' R4, we conclude that Hˇ2(S2,R) ' R. Exercise 1.11. Compute the de Rham cohomology (Appendix A) of S2. Show that closed 0-forms on S2 are constant functions, that the volume form Ω = sin θ dθ dφ is a closed 2-form with nonzero integral, and that if β is another 2-form with ∫ S2 β = ∫ S2 Ω then β−Ω is exact. If α = f(θ, φ) dθ + g(θ, φ) sin θ dφ is a closed 1-form, use the Poincare´ lemma to show that, away from either the north or the south pole, α is of the form d(sin θ h(θ, φ)), and hence show that α is exact. Conclude that Hˇr(S2,R) ' HrdR(S2) for every r. Is that just a coincidence? 1.10 Line bundle classification Let L−→M be a complex line bundle with a local system {(Uj, sj)} of nonvanishing local sections and with transition functions gij : Ui ∩ Uj → C× such that si = gijsj on Ui ∩ Uj. (Since the fibres are one-dimensional, we get11 G = GL(1,C) = C× and sj 7→ gijsj is just the module action of functions on sections; this simplifies the notation considerably.) We may and shall assume that U = {Uj} is a good covering. The fundamental result we need is the following [57]. Proposition 1.5. The family of transition functions g := {gij} is a Cˇech 1-cocycle for the good covering U; its cohomology class [g] ∈ Hˇ1(M,C×) is independent of the local system of sections sj, and depends only on the equivalence class [L] of the complex line bundle L−→M . Moreover, the correspondence [L] 7→ [g] ∈ Hˇ1(M,C×) is an isomorphism of abelian groups. Proof. From its definition, g is clearly a Cˇech 1-cochain in C1(U,C×). The consistency condition (1.8) says that (δg)ijk := gijgjk/gik ≡ 1 on Ui ∩ Uj ∩ Uk, which means that g is a Cˇech 1-cocycle. If a different local system {(Uj, tj)} is chosen, then ti = hisi, where hi : Ui → C× is smooth, i.e., h ∈ C0(U,C×). Clearly ti = (hi/hj)gijtj, so that the transition functions for the new local system form the 1-cocycle g + δh (in additive notation). Thus the line bundle determines the class [g] in Hˇ1(M,C×). The frame bundle P −→M of a line bundle is formed simply by deleting the zero section from L, i.e., by taking Px = Lx \ {0} for each x ∈ M . A nonvanishing local section sj of L−→M may thus also be regarded as a section of the frame bundle P −→M . From Exercise 1.4, we see that the transition functions for the dual line bundle L∗−→M are 11The notation C× denotes the multiplicative group of nonzero complex numbers; we revert to multiplica- tive notation when working with this group. 14 1/gij. On passing to additive notation, we conclude that the dual bundle determines the class −[g] ∈ Hˇ1(M,C×). Similarly, if g′ij are transition functions for another line bundle L ′−→M , then the tensor product bundle L ⊗ L′−→M has transition functions gijg′ij, and so determines the class [g] + [g′] ∈ Hˇ1(M,C×). Therefore [L] 7→ [g] is a homomorphism from the group of line bundle classes to the group Hˇ1(M,C×). If [g] = 0 in Hˇ1(M,C×), then g is a coboundary δf , i.e., gij = fi/fj where each fi : Ui → C× is a smooth function. But then f−1i si = f −1 j sj whenever Ui ∩ Uj 6= ∅, and so there is a global nonvanishing section s ∈ Γ(L) given by s := f−1j sj on Uj; and therefore L−→M is a trivial line bundle. Hence the kernel of homomorphism [L] 7→ [g] is zero. From Lemma 1.4, a line bundle can always be patched together from a given Cˇech 1-cocycle in C1(U,C×), having this cocycle as its set of transition functions; this shows that [L] 7→ [g] is surjective. 1.11 The second cohomology group Consider the short exact sequence of abelian groups: 0−→Z ι−→C −→C×−→ 0, (1.10) where ι is inclusion and (z) := e2piiz. One may form Cˇech cochains with values in any of these groups. If c ∈ Zr(U,C×), then c = (b) with b ∈ Cr(U,C); since (δb) = δc = 0, we can find a ∈ Cr+1(Z) with ι(a) = δb; since ι(δa) = δ(ιa) = 0, we have δa = 0 and so a ∈ Zr+1(U,Z). One checks that [c] 7→ [a] is a well-defined homomorphism12 ∂r : Hˇr(M,C×)→ Hˇr+1(M,Z), called the Bockstein homomorphism [23, 58]. We therefore get a long exact sequence in cohomology: · · · −→ Hˇ1(M,C)−→ Hˇ1(M,C×) ∂−→ Hˇ2(M,Z)−→ Hˇ2(M,C)−→· · · Notice that the discrete group Z need not be underlined: if U is a good covering, an element of Cr(U,Z) is a family of Z-valued smooth functions with connected domains, i.e., a family of constant functions; thus Cr(U,Z) = Cr(U,Z) for all r, and so Hˇr(M,Z) = Hˇr(M,Z). Exercise 1.12. Verify that ∂ does not depend on the choices of a, b and c within their cohomology classes, and that ker ∂r = imH r and im ∂r = kerH r+1ι. Exercise 1.13. If U is a good covering of M and {ψj} is a partition of unity subordinate to U, define, for c ∈ Zr+1(U,C), an element b ∈ Cr(U,C) by bj0...jr−1 := ∑ j cj0...jr−1jψj (this is a locally finite sum). Show that δb = c and conclude that Hˇk(M,C) = 0 for k > 0. Proposition 1.6. The Bockstein homomorphism ∂ is an isomorphism of abelian groups between Hˇ1(M,C×) and Hˇ2(M,Z). Proof. This follows from the preceding exercise, but it is instructive to produce the isomor- phism explicitly. Let g ∈ Z1(U,C×); since each Ui∩Uj is contractible, we can find a smooth 12This is just the standard construction of the “connecting homomorphism” in homological algebra. 15 function fij : Ui ∩ Uj → C such that  ◦ fij = gij; we could write fij = (2pii)−1 log gij, but this is not quite correct since the complex logarithm is “multi-valued”. Define aijk := fij − fik + fjk : Ui ∩ Uj ∩ Uk → C (1.11) whenever Ui ∩ Uj ∩ Uk 6= ∅. Then exp(2piiaijk) = gijgjk/gik ≡ 1, and so aijk is Z-valued. Thus a is an element of C2(U,Z). Now (1.11) says that a = δf in C2(U,C), and so (δa)ijkl := aijk−aikl+aijl−ajkl = 0 on Ui∩Uj ∩Uk∩Ul; these are algebraic relations among Z-valued functions, and so δa = 0 in C3(U,Z); hence a ∈ Z2(U,Z). If we take g′ij := (hi/hj)gij with h ∈ C0(U,C×), we can find a smooth function ki : Ui → C such that  ◦ ki = hi; then  ◦ (fij + ki − kj) = g′ij. In other words, we modify f to f + δk in C1(U,C), and a = δf is unchanged. Therefore we obtain a well-defined homomorphism ∂ : [g] 7→ [a] from Hˇ1(M,C×) to Hˇ2(M,Z). To see that this is an isomorphism, let {ψj} be a smooth partition of unity subordinate to U. Suppose a ∈ Z2(U,Z) is any Cˇech 2-cocycle; define f ∈ C1(U,C) by fij := ∑ r aijrψr, and define g ∈ C1(U,C×) by gij := exp(2piifij). Then fij − fik + fjk = ∑ r (aijr − aikr + ajkr)ψr = ∑ r aijkψr = aijk (1.12) on using δa = 0; hence δf = a. This shows that ∂ is onto. If ∂[g] = 0 in Hˇ2(U,Z), we can arrange, by suitably choosing a representative g in its class, that fij − fik + fjk ≡ 0 on each Ui ∩ Uj ∩ Uk. Define k ∈ C0(U,C) by ki := ∑ r firψr; then ki − kj = ∑ r(fir − fjr)ψr = ∑ r fijψr = fij, and so f = δk. If hi := exp(2piiki), then g = δh in C1(U,C×), and so [g] = 0 in Hˇ2(U,C×). Hence ∂ is one-to-one, which establishes that ∂ : Hˇ1(M,C×)→ Hˇ2(M,Z) is an isomorphism. The upshot is that complex line bundles over M are classified by the integral Cˇech cohomology group Hˇ2(M,Z), at the very small price of going up to the second level in cohomology. This is a first taste of “quantization”, that is, an unexpected discreteness which appears in a seemingly continuous family of objects. The culprit here is the exact sequence (1.10), due to the periodicity of the exponential function. 1.12 Classification of Hermitian line bundles We are particularly interested in Hermitian line bundles L−→M , which have an inner product in each fibre (and therefore the local sections sj can be chosen so that sj(x) ∈ Lx is a unit vector). These have structure group U(1), and the previous arguments show that the group of equivalence classes of Hermitian line bundles is isomorphic to the Cˇech cohomology group Hˇ1(M,U(1)). Exercise 1.14. Construct this isomorphism in detail. We have a short exact sequence of abelian groups: 0−→Z ι−→R −→U(1)−→ 0, 16 where (t) := e2piit for t ∈ R. The corresponding long exact sequence is · · · −→ Hˇ1(M,R)−→ Hˇ1(M,U(1)) ∂−→ Hˇ2(M,Z)−→ Hˇ2(M,R)−→· · · and again Hˇ1(M,R) = Hˇ2(M,R) = 0 by the partition-of-unity construction. Here also, the Bockstein homomorphism is an isomorphism between Hˇ1(M,U(1)) and Hˇ2(M,Z). In this case, the image of each gij is not the whole of the unit circle U(1), since Ui ∩ Uj is contractible; by passing a half-line from the origin through an omitted point of U(1), we can choose a branch of the logarithm for which fij := (2pii) −1 log gij is well-defined. We then have aijk := 1 2pii (log gij − log gik + log gjk) ∈ Z (1.13) since exp(2piiaijk) ≡ 1 (note that the three branches of the logarithm on the right hand side of (1.13) need not be the same). Once again, we get a ∈ C2(U,Z) with δa = 0, so a ∈ Z2(U,Z). This leads to the following result. Proposition 1.7. The Bockstein homomorphism ∂ is an isomorphism of abelian groups between Hˇ1(M,U(1)) and Hˇ2(M,Z). Exercise 1.15. Write out the proof, using the arguments of Proposition 1.6. The Cˇech cohomology groups obviously depend only on the topology of M , and are computed by elementary topological arguments. However, it is desirable to replace them by de Rham cohomology groups, so that one can work with differential forms. This we do in Section 3. 1.13 Classification of vector bundles In order to classify vector bundles of rank higher than 1 (up to equivalence), we face the obstacle that the group structure on the line bundles does not extend to the higher-rank case. In fact, it turns out that the most useful classification of vector bundles relies on a weaker equivalence relation, called stable equivalence, which does not distinguish between a vector bundle E−→M and the Whitney sum E⊕E0−→M whenever the second summand E0−→M is a trivial bundle. This notion arises from the following fundamental property of vector bundles. Proposition 1.8. Let E−→M be a vector bundle over a compact manifold M . Then we can find another vector bundle E ′−→M such that the Whitney sum E⊕E ′−→M is a trivial vector bundle. Proof. Let U = {U1, . . . , Um } be a finite open covering of M by chart domains, and let {s1j , . . . , skj} be linearly independent sections in Γ(Uj, E) (where k is the rank of E). Let {ψj}1≤j≤m be a smooth partition of unity subordinate to U. Let σrj ∈ Γ(E) be defined as ψjs r j on Uj and as 0 on the complement of Uj; notice that the vectors σ r j (x) span the fibre Ex, for any x ∈M . 17 Write n = km, and define f : M × Cn → E (in the case of complex fibres; the real-fibre case is analogous) by f(x, t) := ∑ j,r tjrσ r j (x); then f is a surjective bundle map, i.e., (f, idM) is a bundle morphism. Write E ′x := { (x, t) : t ∈ Cn, f(x, t) = 0 }. Choose some hermitian (or Riemannian) metric on M , and let Fx be the orthogonal complement of E ′ x in {x}×Cn; one checks that the Fx form the fibres of a vector bundle F −→M , and that (f, idM) is a bundle equivalence between this bundle and E−→M . Since F ⊕E ′ = M ×Cn, we thereby obtain an invertible bundle map from M × Cn to E ⊕ E ′. Definition 1.17. Let M be a compact manifold. We say that two vector bundles E−→M and F −→M are stably equivalent if there exists a trivial bundle E0−→M such that E ⊕ E0 and F ⊕ E0 are equivalent. We denote by [[E]] the stable equivalence class of E. Exercise 1.16. Show that E−→M and F −→M are stably equivalent iff [E ⊕G] = [F ⊕G] for some third vector bundle G−→M (which need not be trivial). The equivalence classes of vector bundles over M form an abelian semigroup with identity, under the operation [E] + [F ] := [E ⊕ F ]. One would like to embed this in an abelian group in some canonical way. In fact there is a standard construction of such a group, by abstract nonsense. For any abelian semigroup A with identity, let K(A) be the abelian group with the following universal property: there is a unital semigroup homomorphism θ : A→ K(A) such that, whenever G is a group and γ : A→ G is a unital semigroup homomorphism, there is a unique group homomorphism κ : K(A)→ G for which γ = κ◦θ. Clearly, K(A) is unique up to isomorphism, and it is called the Grothendieck group of A. Exercise 1.17. Check that a group with the desired universal property is given by the following construction. Define an equivalence relation on A × A by (a, b) ∼ (a′, b′) iff a + b′ + c = a′ + b + c for some c ∈ A, and let K(A) be the set of equivalence classes with the obvious sum operation, where the class of (b, a) is inverse to the class of (a, b); let θ(a) be the class of (a, 0). Check also that each element of K(A) is of the form θ(a)− θ(b) for some elements a, b ∈ A. Exercise 1.18. What is K(N)? What is K(A) if A is the multiplicative semigroup Z \ {0}? An abelian semigroup A is said to “allow cancellation” if a + c = b + c implies a = b, for any c ∈ A. Regrettably, the semigroup Vect(M) of (ordinary) equivalence classes of vector bundles over M does not usually allow cancellation. For example, the tangent bundle TS2 → S2 is not trivial, whereas the normal bundle (of lines from the origin of R3 through the sphere S2) is trivial, and their Whitney sum is the trivial bundle S2 × R3 → S2; thus [TS2] + [S2 ×R] = [S2 ×R2] + [S2 ×R], so that cancellation is not allowed. In this case, the homomorphism θ is not injective. Definition 1.18. The K-theory of a compact manifold M is the Grothendieck group K0(M) := K(Vect(M)). Two vector bundles E−→M and F −→M have the same image in K0(M) iff they are stably equivalent; thus any element of K0(M) can be written as a difference of two stable equivalence classes, [[E]]− [[F ]]. 18 Exercise 1.19. Prove the assertion that θ([E]) = θ([F ]) iff [[E]] = [[F ]]. Show also that any element of K0(M) can be written as [E] − [Ok] for some k, where Ok−→M denotes the trivial bundle of rank k. It turns out that K0(M) carries important topological information about the manifold M ; in that context, vector bundles may be viewed as auxiliary tools to study topological spaces. There is a companion group, called K1(M), which together with K0(M) forms a cohomology theory distinct from the Cˇech and de Rham cohomologies. This theory is fully developed in the monograph of Atiyah [3]. 2 Complex projective spaces We illustrate the general theory with a look at a particular class of manifolds, namely the complex projective spaces. These are complex manifolds, that is, differentiable mani- folds whose transition maps are holomorphic; thus we may use complex local coordinates to describe them. They possess three important features: (a) a Hermitian metric; (b) a distinguished nondegenerate closed 2-form, which is known as a “symplectic structure”; (c) a “complex structure”, which at each point specifies a linear automorphism of the tangent space whose square is −1. Moreover, any two of these three structures determine the third; a complex manifold with such a triple structure is called a Ka¨hler manifold. The complex projective spaces are perhaps the simplest examples of compact Ka¨hler manifolds. 2.1 Complex manifolds Definition 2.1. A complex manifold of complex dimension m is a differentiable manifold of real dimension n = 2m which has an atlas of local charts (Uj, φj), with φj : Uj → Cm, such that the transition maps φi ◦ φ−1j are holomorphic. If (x1, . . . , xm, y1, . . . , ym) are local (real) coordinates on Uj, write z k := xk + iyk, z¯k := xk − iyk; then (z1, . . . , zm, z¯1, . . . , z¯m) is an alternative system of local (real) coordinates on Uj. Complex-valued 1-forms1 are locally generated by dz1, . . . , dzm, dz¯1, . . . , dz¯m, where we write dzk := dxk + i dyk, dz¯k := dxk − i dyk. Complex-valued vector fields in X(M,C) are locally generated by ∂/∂z1, . . . , ∂/∂zm, ∂/∂z¯1, . . . , ∂/∂z¯m, where ∂/∂zk := 1 2 (∂/∂xk−i∂/∂yk) and ∂/∂z¯k := 1 2 (∂/∂xk+ i∂/∂yk); more precisely, X(Uj,C) is a C∞(Uj,C)-module with these generators. (The notation is chosen so that dzj(∂/∂zk) = δjk and dz j(∂/∂z¯k) = 0.) In particular, we may write the differential of f ∈ C∞(M,C) as df = ∂f ∂zk dzk + ∂f ∂z¯k dz¯k ≡ ∂f + ∂¯f, (2.1) where, here and in the future, we use the Einstein summation convention of summing over repeated upper and lower indices.2 1In this section, all vector fields and differential forms will be taken complex-valued unless stated other- wise. 2To avoid possible misunderstandings, we emphasize that we do not sum over repeated upper or repeated lower indices, unless a summation sign appears explicitly. 19 Exercise 2.1. Check that the decomposition df = ∂f + ∂¯f does not depend on the local coordinate system, on account of the holomorphicity of the transition maps. Exercise 2.2. A smooth map f : M → N is called holomorphic iff all its local expressions ψi◦ f ◦ φ−1j have holomorphic Cartesian components. Verify that f ∈ C∞(M,C) is holomorphic iff ∂¯f = 0. Definition 2.2. In view of (2.1), we have a splitting A1(M) = A1,0(M) ⊕ A0,1(M) of C∞(M,C)-modules, and more generally, Ar(M) = ⊕ p+q=r Ap,q(M), where each Ap,q(M) is locally spanned by r-forms of the type f(z1, . . . , zm, z¯1, . . . , z¯m) dzi1 ∧ · · · ∧ dzip ∧ dz¯j1 ∧ · · · ∧ dz¯jq . Thus the algebra of differential forms is bigraded : A•(M) = ⊕ p,q A p,q(M). Again by (2.1), the exterior derivative splits as d = ∂ + ∂¯, where ∂ : Ap,q(M) → Ap+1,q(M), ∂¯ : Ap,q(M) → Ap,q+1(M). The identity d2 = 0 yields the three identities ∂2 = 0, ∂∂¯ = −∂¯∂, ∂¯2 = 0, (2.2) on taking account of the grading degrees. (These identities say that the spaces Ap,q(M) form the vertices of a “double complex”.) Exercise 2.3. Verify that the conjugation ω 7→ ω¯ on A•(M,C) = A•(M,R)⊗RC is such that ∂ω = ∂¯ω¯ and that ω¯ ∈ Aq,p(M) whenever ω ∈ Ap,q(M). 2.2 Local charts for complex projective spaces Definition 2.3. The m-dimensional complex projective space CPm is the set of complex lines through the origin in Cm+1, i.e., the one-dimensional complex subspaces of Cm+1. For any nonzero v ∈ Cm+1, the line 〈v〉 ≡ Cv lies in CPm, and η : Cm+1 \ {0} → CPm : v 7→ 〈v〉 is a quotient map.3 If (z0, z1, . . . , zm) denotes coordinates in Cm+1 (with respect to the standard orthonormal basis), we may regard each zj as a linear form on Cm+1; these cannot vanish simultaneously on Cm+1 \ {0}, so we get the following chart domains for CPm: Uj := { 〈v〉 ∈ CPm : zj(v) 6= 0 }, j = 0, 1, . . . ,m. Since 〈v〉 /∈ Uj iff 〈v〉 ⊂ ker zj, we can identify the complement of Uj with the set of lines in the hyperplane ker zj, which is homeomorphic to CPm−1. In particular, each Uj is open and dense in CPm, since its complement is a lower-dimensional submanifold. 3This quotient map actually defines the topology of CPm. 20 For 〈v〉 ∈ Uj and k 6= j, we define wkj (〈v〉) := zk(v) zj(v) . (2.3) This is well-defined since the fraction is unchanged under v 7→ λv with λ ∈ C×). Now4 φ(〈v〉) := (w0j , . . j ∨. . , wmj ) (2.4) is a homeomorphism from Uj onto Cm, and (w0j , . . j ∨. . , wmj ) is a system of local coordinates for the chart (Uj, φj). The wkj are Cartesian coordinates corresponding to the “homogeneous” coordinates z k, that is, (w0j , w 1 j , . . . , w m j ) = [z 0 : z1 : · · · : zm] in the standard notations. On the overlap Ui ∩ Uj, we have wki (〈v〉) = wkj (〈v〉) zj(v) zi(v) = wkj (〈v〉) wij(〈v〉) for k /∈ {i, j}, so the transition map φi ◦ φ−1j , when written as u 7→ v, is given by rational functions: vj = 1/ui and vk = uk/ui for other k, and hence is holomorphic. Thus the atlas { (Uj, φj) : j = 0, 1, . . . ,m } makes CPm a complex manifold. Since CPm \ Um ≈ CPm−1 and Um ≈ Cm ≈ R2m, we have (by induction) that CPm is topologically a cell complex with one k-dimensional cell in every even dimension k = 0, 2, . . . , 2m and no odd-dimensional cells. This allows us to compute the singular homology groups of the complex projective spaces by standard topological techniques [23]: we get Hk(CPm,Z) = Z if k = 0, 2, . . . , 2m; Hk(CPm,Z) = 0 for all other k. Thus CPm is a “torsion-free” space, i.e., all its integral homology groups are free; it is known that then the singular cohomology groups with integer coefficients can be computed by duality: Hk(CPm,Z) = Hom(Hk(CPm,Z),Z) = { Z if k = 0, 2, . . . , 2m, 0 otherwise. (2.5) 2.3 The Ka¨hler form Definition 2.4. For v = (z0, z1, . . . , zm) ∈ Cm+1, we write ‖v‖2 = ∑k |zk|2 = ∑k zkz¯k, and define Φ0(v) := i ∂∂¯ log ‖v‖2 = i ∂ (‖v‖−2∑kzk dz¯k) = i ‖v‖4 ( ‖v‖2 ∑ k dzk ∧ dz¯k − ∑ r,s6=j z¯rzs dzr ∧ dz¯s ) , 4The notation j∨ indicates that the index j is omitted from the sequence 0, 1, . . . ,m. 21 which is a 2-form in A1,1(Cm+1 \ {0}). It is clear that Φ0 is homogeneous of degree 0, i.e., Φ0(λv) = Φ0(v) for λ ∈ C×. This means that there is a 2-form Φ ∈ A1,1(CPm) such that Φ0 = η ∗Φ. To find an expression for Φ in local coordinates on Uj, we may identify Uj with a subset of the hyperplane zj = 1 in Cm+1, by using (2.3) with denominator 1. With this convention, the norm squared of a vector v in this hyperplane defines a positive function Qj on Uj (which tends to infinity at the boundary); in fact, ‖v‖2 = Qj(〈v〉) := 1 + ∑ k 6=j wkj w¯ k j (2.6) for 〈v〉 ∈ Uj with local coordinates (2.4). Thus Φ = i ∂∂¯ logQj = i Q2j ( Qj ∑ k 6=j dwkj ∧ dw¯kj − ∑ r,s 6=j w¯rjw s j dw r j ∧ dw¯sj ) . (2.7) The 2-form Φ is in fact real-valued, as is evident from its local expression (2.7), or alternatively by noting that Φ = −i ∂¯∂ logQj = Φ on account of (2.2); this Φ ∈ A2(CPm,R) is called the Ka¨hler form on CPm. It is important to note that Φ = −∂∂¯Qj is a closed 2-form. This follows at once from (2.2), since dΦ = (∂ + ∂¯)Φ = −∂2(∂¯Qj) + ∂¯2(∂Qj) = 0. Exercise 2.4. A real-valued 2-form β ∈ A2(M) is called nondegenerate if each βx ∈ Λ2T ∗xM is nondegenerate as an alternating R-bilinear form, i.e., βx(u, v) = 0 for all v ∈ TxM implies u = 0 in TxM . Show that β is nondegenerate iff for any local expression β ∣∣ U = brs dx r ∧ dxs, the matrix of local coefficients [brs] is nonsingular at each point of U . Verify that the Ka¨hler form Φ is nondegenerate on CPm by showing that the matrix A with entries Qjδrs− w¯rjwsj is positive definite (apply the Schwarz inequality to show that z¯tAz > 0 for all z ∈ Cm+1\{0}). Exercise 2.5. Show that Φ∧m, the m-fold exterior power of Φ, is a volume form on CPm, i.e., a 2m-form which is nonzero at each point of CPm when regarded as a section of the line bundle Λ2mT ∗CPm → CPm. 2.4 The Fubini–Study metric Definition 2.5. Let M be a differential manifold of even (real) dimension n = 2m. An almost complex structure on M is an operator J : X(M) → X(M) which is C∞(M,R)- linear,5 and which satisfies J2 = − id. Thus (JX)x = Jx(Xx) where Jx ∈ EndR(TxM) with (Jx) 2 = − id for each x ∈ M .6 Almost complex structures need not exist; if one does exist, we say that (M, J) is an almost complex manifold. 5In other words, J can be identified with a tensor on M , of bidegree (1, 1), given by (X,α) 7→ α(J(X)). One also says that J is a “tensorial operator”. 6An almost complex structure may therefore also be defined as a bundle automorphism (J, id) of the tangent bundle TM −→M for which J2 = − id. 22 It turns out that M need not be a complex manifold in order to possess an almost complex structure; a known example is the sphere S6 which has an almost complex structure related to the Cayley numbers. See [15, 58] for some discussion of this. Here we will stick to complex manifolds. A complex structure can always be defined locally if dimM is even; it suffices to take local coordinates (x1, . . . , xm, y1, . . . , ym) on a chart domain U and let J ( ∂ ∂xk ) = ∂ ∂yk , J ( ∂ ∂yk ) = − ∂ ∂xk . (2.8) Now J may be amplified to a C∞(U,C)-linear operator on X(U,C) in the natural way; from (2.8) we derive J ( ∂ ∂zk ) = i ∂ ∂zk , J ( ∂ ∂z¯k ) = −i ∂ ∂z¯k . On a complex manifold, these local definitions patch together to give a global definition of J. For instance, on CPm we have J ( ∂ ∂wkj ) = i ∂ ∂wkj , J ( ∂ ∂w¯kj ) = −i ∂ ∂w¯kj , (2.9) which is independent of the chart (Uj, φj). Exercise 2.6. Verify this independence on an overlap Ui ∩ Uj of charts of CPm. Lemma 2.1. Let M be a manifold with an almost complex structure J, and let Φ ∈ A2(M) be a real-valued nondegenerate 2-form on M which is invariant under J, i.e., Φ(JX, JY ) = Φ(X, Y ) for X, Y ∈ X(M). Then the recipe g(X, Y ) := Φ(X, JY ) defines a symmetric tensor on M which is also J-invariant. Proof. Clearly g is a tensor of bidegree (2, 0); symmetry follows from invariance, since g(Y,X) = Φ(Y, JX) = −Φ(JX, Y ) = Φ(JX, J2Y ) = Φ(X, JY ) = g(X, Y ). When M = CPm, Φ is the Ka¨hler form (2.7), and J is given by (2.9), we obtain on Uj: g = 2 Q2j ( Qj ∑ k 6=j dwkj · dw¯kj − ∑ r,s 6=j w¯rjw s j dw r j · dw¯sj ) . (2.10) Exercise 2.7. Check that g is positive definite, and thus defines a Riemannian metric on CPm. (Recall that the matrix with entries Qjδ rs − w¯rjwsj is positive definite, by Exercise 2.4.) Definition 2.6. The metric (2.10) is the Fubini–Study metric on CPm. 23 2.5 The Riemann sphere When m = 1, the complex projective space CP1 is identified with the Riemann sphere C∞ = C unionmulti {∞}, by identifying [z0 : z1] with z = z1/z0 ∈ C if z0 6= 0, and [1 : 0] with ∞. Notice that z ≡ w10 is the local complex coordinate in U0 = C∞\{∞}. By writing z = x+iy, we can also identify C∞ with the two-sphere S2, regarding the latter as a submanifold of R3, via the stereographic projection f(z) := ( 2x 1 + x2 + y2 , 2y 1 + x2 + y2 , −1 + x2 + y2 1 + x2 + y2 ) (2.11) and f([1 : 0]) := (0, 0, 1). The Ka¨hler form on C∞ is easily found. We have Q0(〈v〉) = 1+ |z|2 for v = (z0, z1) ∈ C2, so (2.7) reduces to Φ = i dz ∧ dz¯ (1 + |z|2)2 = − 2 dx ∧ dy (1 + x2 + y2)2 . (2.12) Exercise 2.8. Write (u1, u2, u3) to denote the right hand side of (2.11). Check that Φ = f ∗Ω, where Ω ∈ A2(R3) is given by Ω = (du1 ∧ du2)/2u3. The usual spherical coordinates on S2 are defined by the map h : R2 → S2 where h(θ, φ) := (sin θ cosφ, sin θ sinφ, cos θ). Check that h∗Ω = −1 2 sin θ dθ ∧ dφ. The Riemannian metric (2.10) reduces to g = 2(1 + |z|2)−2 dz · dz¯ or equivalently g = 2(1 + x2 + y2)−2 (dx2 + dy2), where dx2 denotes the symmetric product dx · dx. Exercise 2.9. Show that g = f ∗(1 2 G) where G = (du1)2 + (du2)2 + (du3)2 is the standard Riemannian metric on R3, and that h∗G = dθ2 + sin2 θ dφ2. We may regard (θ, φ) as local coordinates on the Riemann sphere. Thus we write simply: Φ = −1 2 (sin θ dθ ∧ dφ), g = 1 2 (dθ2 + sin2 θ dφ2). (2.13) It may prove useful to have available some relations between the local coordinate systems (z, z¯) and (θ, φ). We list a few identities that follow from the previous formulae: z = sin θ 1− cos θe iφ, dz = eiφ 1− cos θ (−dθ + i sin θ dφ), Q0 = 2 1− cos θ . (2.14) Exercise 2.10. Verify the formulae (2.14) and derive (2.13) directly. Exercise 2.11. Show that the complex structure J on S2 satisfying g(X, Y ) = Φ(X, JY ) is given by J(∂/∂θ) = − csc θ ∂/∂φ and J(csc θ ∂/∂φ) = ∂/∂θ in spherical coordinates. 3 The de Rham complex and Hodge duality 3.1 The de Rham complex Definition 3.1. If M is an n-dimensional differential manifold, its de Rham complex is the cochain complex A0(M) d−→A1(M)→ · · · → Ak(M) d−→Ak+1(M)→ · · · → An(M) 24 where Ak(M) = Ak(M,R) denotes the real-valued differential forms on M , and d is the exterior derivation. We recall (see Appendix A) that its k-cocycles are the closed differential k-forms ZkdR(M) := {ω ∈ Ak(M) : dω = 0 }, and its k-coboundaries are the exact differential k-forms BkdR(M) := { dβ : β ∈ Ak−1(M) }. The k-th de Rham cohomology group HkdR(M) := Hk(A•(M), d) is a real vector space. The 0-cocycles in Z0dR(M) are locally constant smooth functions, so H 0 dR(M) = Rm if M has exactly m connected components. In particular, H0dR(M) 6= 0. If U is a contractible manifold, then by the Poincare´ lemma,1 HkdR(U) = 0 for each k = 1, 2, . . . , n. If M is an orientable compact manifold (without boundary), with volume form ν ∈ An(M), then dν = 0 and ν is not exact since ∫ M ν 6= 0. (If µ ∈ An(M) is exact, with µ = dβ, say, then ∫ M µ = ∫ M dβ = 0 by Stokes’ theorem, since ∂M = ∅; thus, a volume form cannot be exact.) Therefore HndR(M) 6= 0. Moreover, since the integral vanishes on exact (n−1)-forms, again by Stokes’ theorem, we see that [ω] 7→ ∫ M ω is a well-defined linear form on HndR(M). 3.2 The Riemannian volume form Definition 3.2. Let g be a Riemannian metric on the compact oriented manifold M . Then g : X(M)×X(M)→ C∞(M) is a symmetric C∞(M)-bilinear form such that each gx : TxM× TxM → R is positive definite. The pair (M, g) is called a Riemannian manifold. If U ⊆M is a chart domain with local coordinates x1, . . . , xn, let gij := g(∂/∂xi, ∂/∂xj) ∈ C∞(U); then [gij] is a symmetric matrix, whose determinant we write as det g (by a slight abuse of notation, since this determinant is coordinate-dependent), and we have g = gij dx i · dxj on U . For each x ∈M , there is a vector space isomorphism gˆx : TxM → T ∗xM given by gˆx(u) := [v 7→ gx(u, v)]; these determine a diffeomorphism gˆ : TM → T ∗M such that (gˆ, idM) is an equivalence between the tangent and cotangent bundles. They also determine tensorial operators X 7→ X[ : X(M)→ A1(M), given by X[(Y ) := g(X, Y ), and its inverse α 7→ α] : A1(M)→ X(M), given by α(Y ) =: g(α], Y ). These “musical isomorphisms” [8] allow us to identify vector fields and 1-forms and to use them interchangeably, as the occasion demands. The metric g defines thus defines bilinear pairings, not only of vector fields but also of 1- forms and indeed of differential forms of any degree; we will denote all these pairings by (· | ·) whenever a fixed g is given. Thus (X | Y ) := g(X, Y ) for X, Y ∈ X(M); (α | β) := g(α], β]) for α, β ∈ A1(M); and the pairing on Ak(M) (for k > 1) is determined by (α1 ∧ · · · ∧ αk | β1 ∧ · · · ∧ βk) := det[(αi | βj)] for α1, . . . , αk, β1, . . . , βk ∈ A1(M). Definition 3.3. Let (M, g) be an oriented Riemannian manifold. Choose a local orthonormal frame X1, . . . , Xn for X(U), i.e., vector fields such that g(Xr, Xs) = δrs, which is oriented, 1A complex (C•, d) is called “acyclic” if Hk(C•, d) = 0 for k > 0. Thus the Poincare´ lemma says that the de Rham complex of a contractible manifold is acyclic. 25 that is, ν(X1, . . . , Xn) > 0 where ν is a volume form on M which defines the orientation. Let θˆk := X[k ∈ A1(U), so that θˆ1, . . . , θˆn is a local (oriented) orthonormal frame for A1(U). Then Ω := θˆ1 ∧ θˆ2 ∧ · · · ∧ θˆn ∈ An(M) (3.1) is a volume form on M , since (Ω |Ω) = 1 and in particular Ωx 6= 0 in ΛnT ∗xM for all x ∈M . This Ω is called the Riemannian volume form on (M, g). To see that Ω is in fact independent of the choice of oriented orthonormal frame, we argue as follows. If Y1, . . . , Yn is another oriented orthonormal frame for X(V ), where U ∩ V 6= ∅, then Yi := a j iXj where a = [a j i ] is a smooth function on U ∩ V with values in SO(n), the group of orthogonal n × n matrices of determinant 1. (That is to say, a is a local section of the principal SO(n)-bundle of oriented orthonormal frames on M .) Now ϑˆk := Y [k gives the corresponding oriented orthonormal basis for A1(V ), and thus ϑˆk = akl θˆl, from which it follows that ϑˆ1 ∧ · · · ∧ ϑˆn = (det a) θˆ1 ∧ · · · ∧ θˆn = Ω on U ∩ V . Therefore, Ω is defined globally by (3.1). One can express Ω in terms of local coordinates x1, . . . , xn on U for which the ∂/∂xj need not be orthonormal, i.e., gij 6= δij in general. The local coordinates should, however, be compatible with the orientation, which means that det g > 0. If y1, . . . , yn are other local coordinates on V , and if g˜ij := g(∂/∂y i, ∂/∂yj) ∈ C∞(V ), then det g˜ = J2 det g, where J := det[∂yr/∂xj] is the Jacobian of the transition function (which is positive). Hence√ det g˜ dy1 ∧ · · · ∧ dyn = √ det g dx1 ∧ · · · ∧ dxn (3.2) on U ∩ V , and thus the coordinate-independent expression √det g dx1 ∧ · · · ∧ dxn defines a volume on M . In particular, one may choose y1, . . . , yn so that g˜ij = δij, so det g˜ = 1 and dy1, . . . , dyn is an orthonormal frame: so the volume form (3.2) coincides with Ω of (3.1). In other words, Ω = √ det g dx1 ∧ · · · ∧ dxn (3.3) in any oriented local coordinate system. 3.3 The Hodge star operator Definition 3.4. Let (M, g) be a compact oriented Riemannian manifold of dimension n, and write m := bn/2c (so n = 2m or else n = 2m − 1). Define the Hodge star operator ? : A•(M,C)→ A•(M,C) as follows. Choose a local orthonormal frame X1, . . . , Xn for X(U) on some chart domain U ⊂ M , and let θˆ1, . . . , θˆn be the corresponding local orthonormal frame for A1(U), determined by θˆi(Xj) := δ i j. Define ? on A •(U,C) by ? := im((θˆ1)− ι(X1)) . . . ((θˆn)− ι(Xn)), (3.4) where ι(X) denotes contraction with the vector field X and (α) : ω 7→ α ∧ ω denotes exterior product with the 1-form α. It is readily checked that the right hand side of (3.4) is independent of these local orthonormal frames, and therefore defines an operator on all of A•(M,C). 26 To simplify products of noncommuting quantities as in (3.4), we use the following nota- tion: if aj1 , aj2 , . . . , ajk are elements of some algebra, where the indices jr are in increasing order, write →∏ 1≤r≤k ajr := aj1aj2 . . . ajk . (3.5) Thus ? = im ∏→ 1≤r≤k (θˆ r)− ι(Xr). It is also convenient to write J := {j1, . . . , jk} and denote the right hand side of (3.5) by ∏→ j∈J aj, where it is understood that the indices j ∈ J are arranged in increasing order. Exercise 3.1. Let V be an n-dimensional oriented real vector space with a positive definite symmetric bilinear form q. For u ∈ V , α ∈ V ∗, define operators ι(u) : ΛkV ∗ → Λk−1V ∗ and (α) : ΛkV ∗ → Λk+1V ∗ by [ι(u)η](v1, . . . , vk−1) := η(u, v1, . . . , vk−1) and (α)η := α ∧ η. Let {e1, . . . , en}, {e′1, . . . , e′n} be oriented orthonormal bases for (V, q) (so the change-of-basis matrix [q(ei, e ′ j)] has determinant +1) and let {ζ1, . . . , ζn}, {ζ ′1, . . . , ζ ′n} be the respective dual bases for V ∗. Show that ∏→ 1≤r≤n (ζ ′r) − ι(e′r) = ∏→ 1≤r≤n (ζ r) − ι(er) as operators on Λ•V ∗. Exercise 3.2. Use the previous exercise to show that the right hand side of (3.4) is indepen- dent of the given local orthonormal frames {Xr} and {θˆk}. Notice that when m is even, i.e., when n is of the form 4k or 4k + 3 for some integer k, then it is not necessary to use complex-valued forms since ? takes A•(M,R) to A•(M,R). However, for n of the form 4k+ 1 or 4k+ 2, we have im = ±i and complex forms are needed. More traditional treatments of Hodge duality [1, 28, 33, 57] use different sign conventions, so that ? operates on A•(M,R) in all cases, but the important involutivity property (see Lemma 3.3 below) is then lost. Lemma 3.1. If J := {j1, . . . , jk} ⊆ {1, . . . , n} and if J ′ := {1, . . . , n} \ J = {i1, . . . , in−k} is its complement, with indices written in increasing order in both cases, let η(J, J ′) = ±1 de- note the sign of the shuffle permutation which reorders (1, 2, . . . , n) as (j1, . . . , jk, i1, . . . , in−k). Write, for brevity, θˆJ := θˆj1∧· · ·∧θˆjk , where θˆ1, . . . , θˆn is a local orthonormal basis of A1(M). Then the Hodge star operator is given explicitly on this basis by the recipe: ?θˆJ = im(−1)nk+k(k−1)/2η(J, J ′) θˆJ ′ . (3.6) 27 Proof. This identity results from the following calculation: ?θˆJ = im →∏ 1≤j≤n ((θˆj)− ι(Xj)) θˆj1 ∧ · · · ∧ θˆjk = im(−1)k(k−1)/2 →∏ 1≤j≤n ((θˆj)− ι(Xj)) θˆjk ∧ · · · ∧ θˆj1 = im(−1)k(k−1)/2η(J ′, J) →∏ i∈J ′ (θˆi) →∏ j∈J (−ι(Xj)) θˆjk ∧ · · · ∧ θˆj1 = im(−1)k(k−1)/2(−1)kη(J ′, J) →∏ i∈J ′ (θˆi) 1 = im(−1)k(k−1)/2(−1)k(−1)k(n−k)η(J, J ′) θˆJ ′ . Here, one first reverses the order of the exterior product of 1-forms θˆj1 ∧ · · · ∧ θˆjk , by a permutation of sign (−1)k(k−1)/2. Then the operators (θˆj)− ι(Xj) act successively on forms of type θˆj ∧ βj, yielding −βj at each stage, until only the constant (−1)k remains; next, the operators (θˆi)− ι(Xi), with i ∈ J ′, act successively on forms of type γj, yielding θˆj ∧ γj at each stage, until θˆJ ′ is created. Lastly, the identity η(J, J ′)η(J ′, J) = (−1)k(n−k) is used: this is just the observation that the permutation which interchanges the blocks of indices J and J ′ —while preserving the order within each block— is a product of k(n− k) transpositions. Now (3.6) follows on noticing that (−1)k(−1)k(n−k) = (−1)nk(−1)k(k+1) = (−1)nk. Corollary 3.2. The Hodge star operator maps Ak(M,C) into An−k(M,C). The next calculation shows that the presence of the factor im is what ensures that ? is an involution, i.e., an operator whose square is the identity. Lemma 3.3. ?? = id. Proof. From (3.6) it is clear that ??θˆJ = CJ θˆ J for some constant CJ . One computes: CJ = i 2m(−1)nk+k(k−1)/2(−1)n(n−k)+(n−k)(n−k−1)/2η(J, J ′)η(J ′, J) = (−1)m(−1)n2(−1)(k(k−1)+(n−k)(n−k−1))/2(−1)k(n−k) = (−1)m(−1)n2(−1)(n2−n(2k+1)+2k2)/2(−1)nk−k2 = (−1)m(−1)(3n2−n)/2 = (−1)m(−1)(n2+n)/2 = (−1)m(−1)m(2m±1) = +1. This establishes that ?? = id, which also implies that ?mapsAk(M,C) onto An−k(M,C). Exercise 3.3. If X ∈ X(M), show that ?(X[)? = (−1)nι(X) as operators from Ak(M) to Ak−1(M). (Reduce to the special case X = Xj.) 28 Consider the case M = R3; though not compact, it is an oriented Riemannian manifold with the usual euclidean metric; here θˆk = dxk, and ? is determined by ?1 = −Ω = −dx1 ∧ dx2 ∧ dx3, ?dx1 = dx2 ∧ dx3, ?dx2 = −dx1 ∧ dx3, ?dx3 = dx1 ∧ dx2, on account of (3.6). When M = C with local coordinates x, y, take z = x + iy. For the usual metric, we have θˆ1 = dx, θˆ2 = dy, and so ?dx = i dy, ?dy = −i dx. Using the coordinates z, z¯, we then have ?dz = dz, ?dz¯ = −dz¯. Thus, ?df = ?(∂f + ∂¯f) = ∂f − ∂¯f , so that f ∈ C∞(C,C) is holomorphic iff ∂¯f = 0 (by the Cauchy–Riemann equations) iff ?df = df . More generally, any α ∈ A•(M,C) may be written as α = α+ + α− where α+ is selfdual, i.e., ?α+ = α+, and α− is antiselfdual, i.e., ?α− = −α−. When n = 2m and α ∈ Am(M), we have α+, α− ∈ Am(M) also. When n = 2 and M is a compact complex manifold (i.e., a Riemann surface), f ∈ C∞(M,C) is holomorphic iff df is selfdual. An important example is M = S2 with the Riemannian metric g = dθ2 + sin2 θ dφ2. In spherical coordinates, a local orthonormal basis is given by θˆ1 = dθ, θˆ2 = sin θ dφ. The star operator on S2 is determined by ?1 = iΩ = i sin θ dθ ∧ dφ, ?dθ = i sin θ dφ, and thus ?Ω = −i and ?(sin θ dφ) = −i dθ. For a general system of coordinates, the star operator can be described as follows. On a chart domain U , a k-form ω ∈ Ak(U) can be written as ω = ωj1...jk dxj1 ∧ · · · ∧ dxjk . New coefficients with “raised indices” are defined by ωr1...rk = gr1j1 . . . grkjkωj1...jk where [g rs] denotes the matrix inverse to the matrix [gij]. Let t1...tn := 0 if t1, . . . , tn contains a repeated index, and otherwise let t1...tn := ±1 be the sign of the permutation (1, . . . , n) 7→ (t1, . . . , tn). Then ?ω is given by ?(ωj1...jk dx j1 ∧ · · · ∧ dxjk) = im(−1)nk+k(k−1)/2 √ det g (n− k)!i1...in−kr1...rnω r1...rk dxi1 ∧ · · · ∧ dxin−k . (3.7) Exercise 3.4. Check the validity of (3.7) by showing that its right hand side is invariant under a general change of (oriented) frame, and that it reduces to (3.6) for an orthonormal frame. Exercise 3.5. Work out the action of ? for the Fubini–Study metric on CPm. The Hodge star operator relates the exterior product of forms to the Riemannian volume form, as follows. 29 Lemma 3.4. For α, β ∈ Ak(M,C), α ∧ ?β = im(−1)nk+k(k−1)/2(α | β) Ω. (3.8) Proof. Since both sides of (3.8) are C∞(M)-bilinear in (α, β), it suffices to verify equality for a local basis; thus we may take α = θˆI , β = θˆJ where I and J are subsets of {1, . . . , n}. Now (θˆI | θˆJ) = 1 or 0 according as I = J or not. If I 6= J then θˆI ∧ ?θˆJ = 0 from (3.6), since I∩J ′ 6= ∅. Also, θˆJ ∧?θˆJ = im(−1)nk+k(k−1)/2η(J, J ′)θˆJ ∧ θˆJ ′ = im(−1)nk+k(k−1)/2Ω. Definition 3.5. The complex vector space Ak(M,C) becomes a prehilbert space under the positive definite Hermitian form 〈〈α | β〉〉 := ∫ M (α¯ | β) Ω, where the integral over M is normalized by the requirement that ∫ M Ω = 1. Its completion with respect to this “integrated inner product”2 is a Hilbert space which may be denoted L2,k(M). Notice that 〈〈α | β〉〉 = 〈〈β¯ | α¯〉〉, i.e., the conjugation of forms is antiunitary. The inner product may be extended to all of A•(M,C) by declaring that k-forms and l- forms be orthogonal for k 6= l; the completion of A•(M,C) is the Hilbert-space direct sum L2,•(M) := ⊕n k=0 L 2,k(M). From (3.8), it follows that 〈〈α | β〉〉 = i−m(−1)nk+k(k−1)/2 ∫ M α¯ ∧ ?β. Lemma 3.5. The Hodge star operator is isometric, i.e., if α, β ∈ Ak(M,C) then 〈〈?α | ?β〉〉 = 〈〈α | β〉〉. Proof. From (3.6) it follows that ?α¯ = (−1)m?α for any α ∈ A•(M,C). Applying (3.8) twice, 〈〈?α | ?β〉〉 = i−m(−1)n(n−k)+(n−k)(n−k−1)/2 ∫ M ?α ∧ β = im(−1)n2+n(n−1)/2+k(k+1)/2 ∫ M ?α¯ ∧ β = im(−1)n(n+1)/2+k(k+1)/2(−1)k(n−k) ∫ M β ∧ ?α¯ = (−1)m(−1)n(n+1)/2〈〈β¯ | α¯〉〉 = 〈〈α | β〉〉 since 1 2 n(n+ 1) = m(2m± 1) ≡ m mod 2. Since ?? = id on L2,k(M), this isometry also satisfies 〈〈?α|γ〉〉 = 〈〈α|?γ〉〉 for α ∈ Ak(M,C), γ ∈ An−k(M,C); thus the Hodge star operator extends to a selfadjoint unitary operator on L2,k(M)⊕ L2,n−k(M). 2The double brackets are intended to distinguished the integrated inner product from the C∞(M)-valued form (· | ·). 30 3.4 The Hodge Laplacian Definition 3.6. The codifferential δ : Ak(M,C)→ Ak−1(M,C) is the adjoint of the exte- rior derivative with respect to the integrated inner product. In other words, if α ∈ Ak(M,C), β ∈ Ak−1(M,C), δα is determined by the relation 〈〈δα | β〉〉 := 〈〈α | dβ〉〉. (3.9) The Riesz theorem “guarantees” that there is a unique element δα ∈ L2k−1(M) satisfy- ing (3.9), though it is not immediately clear that δα ∈ Ak−1(M,C). That this is indeed the case, and that δ takes Ak(M,R) to Ak−1(M,R), is a consequence of the following basic identity. Lemma 3.6. δ = (−1)n+1?d?. Proof. If α ∈ Ak(M,C) and β ∈ Ak−1(M,C), then dβ¯ ∧ ?α+ (−1)k−1β¯ ∧ d(?α) = d(β¯ ∧ ?α) is an exact n-form, whose integral is zero by Stokes’ theorem. Thus 〈〈β | δα〉〉 = 〈〈dβ | α〉〉 = i−m(−1)nk+k(k−1)/2 ∫ M dβ¯ ∧ ?α = i−m(−1)nk+k(k−1)/2 ∫ M (−1)kβ¯ ∧ d(?α) = (−1)nk+k(k−1)/2(−1)k(−1)n(k−1)+(k−1)(k−2)/2〈〈β | ?d?α〉〉 = (−1)n+1〈〈β | ?d?α〉〉, and so δα = (−1)n+1?d?α since β is arbitrary. Clearly δ2 = 0, and in particular (A•(M), δ) is a chain complex. We say that ω ∈ Ak(M) is coclosed if δω = 0, or coexact if ω = δζ for some ζ ∈ Ak+1(M). Definition 3.7. The Hodge Laplacian is the operator ∆ on A•(M) defined as ∆ := (d+ δ)2 = dδ + δd. Notice that ∆ takes Ak(M) to Ak(M), since d raises and δ lowers degrees by one. [The restriction of ∆ to 0-forms is called the Laplace–Beltrami operator on C∞(M).] A k-form γ is called harmonic if ∆γ = 0. We denote the vector space of harmonic k-forms by Harmk(M). Lemma 3.7. A k-form γ is harmonic iff dγ = δγ = 0. Proof. If γ is both closed and coclosed, then clearly ∆γ = d(δγ) + δ(dγ) = 0. On the other hand, for any ω ∈ A•(M), 〈〈ω |∆ω〉〉 = 〈〈ω | dδω〉〉+ 〈〈ω | δdω〉〉 = 〈〈δω | δω〉〉+ 〈〈dω | dω〉〉 ≥ 0, so that ∆γ = 0 implies dγ = δγ = 0. 31 The Laplacian ∆ commutes with the operators ?, d and δ. Indeed, ?dδ = (−1)n+1?d?d? = δd? and ?δd = (−1)n+1d?d = dδ?, so ?∆ = ?dδ + ?δd = δd?+ dδ? = ∆?. Also, d∆ = dδd = ∆d and δ∆ = δdδ = ∆δ since d2 = δ2 = 0. On the Hilbert space L2,•(M), ∆ = dδ+ δd = dd∗ + d∗d is a formally selfadjoint positive operator with domain A•(M,C); however, it is generally an unbounded operator. It can be made bounded by defining a larger Hilbert space norm on A•(M,C) and completing to obtain a smaller Hilbert space H2n,•(M), called a Sobolev space. We shall not go into this matter here; we refer to [28] for the details. It turns out that ∆ is then bounded as an operator from H2n,•(M) to L2,•(M), which is, in fact, an elliptic differential operator.3 Elliptic operators have two main properties. Firstly, they are Fredholm operators. This implies both that ker ∆ is finite-dimensional, and that ∆ has closed range in L2,•(M). Secondly, a generalized solution of an elliptic differential operator is in fact smooth.4 This means that ∆u = 0 for u ∈ H2n,•(M) only when u in fact lies in A•(M,C). In other words, ker ∆ = Harm•(M,C). The Fredholm property of ∆ now implies that Harm•(M,C) is a finite-dimensional vector space. The theory of the Hodge Laplacian culminates in the fundamental theorem of Hodge, which states that any k-form on a compact oriented Riemannian manifold can be uniquely decomposed as a sum of a closed k-form, a coclosed k-form, and a harmonic k-form. Theorem 3.8. (Hodge). Let (M, g) be a compact oriented Riemannian manifold. Then for each k = 0, 1, . . . , n there is an orthogonal direct sum Ak(M) = dAk−1(M)⊕ δAk+1(M)⊕ Harmk(M). (3.10) That the summands are orthogonal follows from the identities 〈〈dα |δβ〉〉 = 〈〈d2α |β〉〉 = 0, 〈〈dα | γ〉〉 = 〈〈α | δγ〉〉 = 0, 〈〈δβ | γ〉〉 = 〈〈β | dγ〉〉 = 0, for α ∈ Ak−1(M), β ∈ Ak+1(M), γ ∈ Harmk(M). If P denotes the orthogonal projector on the Hilbert space L2,k(M) whose range is Harmk(M), then if ω ∈ Ak(M) the form ω − Pω lies in the range of the selfadjoint operator ∆; it turns out that this range is closed, even if one restricts ∆ to its original domain Ak(M). Hence, ω−Pω = ∆η for some η ∈ Ak(M); and therefore ω = d(δη)+δ(dη)+Pω. The full proof of Hodge’s theorem depends on verifying that ∆ is elliptic and identifying carefully the range of ∆. For the original treatment of Hodge, one can examine his book [33]; for more modern proofs, in the somewhat more general setting of an “elliptic complex”, one may consult [23, 28]. The inverse operator ∆−1, defined on the orthogonal complement of Harmk(M), is an integral operator, namely the “Green operator” for the partial differential equation ∆α = 0. Corollary 3.9. Each de Rham class [ω] ∈ HdR(M) contains exactly one harmonic form; thus γ 7→ [γ] is an isomorphism of Harmk(M,R) onto HkdR(M), and therefore every de Rham cohomology space of M is finite dimensional. 3This means that the “principal symbol” of ∆, which is a certain matrix-valued function on T ∗M , becomes invertible after deleting a neighbourhood of the zero section of the cotangent bundle. In fact, the principal symbol σ∆ of the Laplacian is given by σ∆(ξx) = −(ξx|ξx) := −gx(ξ]x, ξ]x) for ξx ∈ TxM . 4A differential operator whose generalized solutions are automatically smooth is called “hypoelliptic”. Any elliptic operator is hypoelliptic. 32 Proof. If ω = dα + δβ + γ in Ak(M,R), with γ harmonic, then ω is closed iff dδβ = 0 iff (β | dδβ) = (δβ | δβ) = 0 iff δβ = 0. For ω = dα + γ ∈ ZkdR(M), we clearly have [ω] = [γ]. Moreover, if ω′ ∈ ZkdR(M) with [ω′] = [ω], then ω′ = ω + dζ for some ζ ∈ Ak−1(M), so ω′ = d(α + ζ) + γ; hence ω′ and ω have the same harmonic component γ. Finally, notice that ?d = (−1)n−1δ?, so that whenever ω = dα+ δβ + γ in Ak(M,C), we also have ?ω = d((−1)n−1?β) + δ((−1)n−1?α) + ?γ in An−k(M,C). The uniqueness of the Hodge decomposition says that the star operator is a linear isomorphism of Harmk(M,C) onto Harmn−k(M,C); if desired, one can match real harmonic forms by γ 7→ i−m?γ. Passing to cohomology, this yields a well-defined R-linear isomorphism [γ] 7→ [i−m?γ] : HkdR(M) ∼−→Hn−kdR (M). This isomorphism is called Poincare´ duality.5 4 The Hodge Laplacian on the 2-sphere In this section, we investigate in detail the Hodge Laplacian on the 2-sphere S2, both as an illustration of the Hodge theory in general, and in order to introduce a first example of a Dirac operator. The sphere is, of course, an oriented Riemannian manifold, but it is also round, i.e., it is a homogeneous space under the group of rotations of R3. The rotation invariance of the Laplacian ∆ facilitates a complete description of all its eigenvalues and eigenvectors, which in turn leads to a corresponding spectral description of the Dirac operator D/ = d+ δ. 4.1 The rotation group in three dimensions Definition 4.1. The rotation group SO(3) consists of all 3× 3 real matrices A satisfying AtA = 13 and detA = 1. Any rotation belongs to a one-parameter subgroup { exp tN : t ∈ R }, where N belongs to the Lie algebra so(3) of the rotation group, i.e., the 3× 3 real matrices satisfying N t +N = 0, TrN = 0. Write N =  0 −n3 n2n3 0 −n1 −n2 n1 0  , (4.1) and suppose (n1)2 + (n2)2 + (n3)2 = 1. Then if n = (n1, n2, n3) ∈ R3, the matrix identities N2 = nnt − 13, N3 = −N hold, and it follows that exp tN = I + N sin t + N2(1 − cos t) ∈ SO(3). Now the rotation action of SO(3) on R3 (or on S2) is given explicitly by (exp tN)x = (I +N sin t+N2(1− cos t))x = x+ (n× x) sin t+ (n(n · x)− x)(1− cos t) = x cos t+ (n× x) sin t+ n(n · x)(1− cos t). (4.2) 5For noncompact Riemannian manifolds, one can define “compactly supported de Rham cohomology” H•dR,c(M) by starting from the cochain complex of differential forms with compact support. Then Poincare´ duality is a family of isomorphisms between HkdR(M) and H n−k dR,c(M). In the compact case, both cohomologies coincide. 33 Let L1, L2, L3 be the generators of the rotation group, i.e., the elements of so(3) given by replacing n in (4.1) by the standard orthonormal basis vectors; so N = n1L1 + n 2L2 + n 3L3. It is immediate that [L1, L2] = L3, [L2, L3] = L1 and [L3, L1] = L2, or more compactly, [Li, Lj] = ij kLk. It is very convenient to introduce the matrices L± := L1 ± i L2 (which belong to the complexified Lie algebra so(3,C)). Then L21 + L 2 2 + L 2 3 = L−L+ + L 2 3 − iL3 (4.3) as 3× 3 complex matrices.1 Definition 4.2. The action of SO(3) on S2 induces an action of C∞(S2) by (R · f)(x) := f(R−1x); more generally, SO(3) acts on A•(S2) by R · ω := (R−1)∗ω. For the corresponding action of the Lie algebra so(3) on C∞(S2), the homomorphism property of the group action induces a Leibniz rule, so that N ∈ so(3) acts on C∞(S2) as a vector field N˜ , called the fundamental vector field of N , which is given explicitly by N˜f(x) := d dt ∣∣∣∣ t=0 f(exp(−tN)x). Since the rotation action on SO(3) is a restriction of an action on R3, the same formula yields fundamental vector fields in X(R3). In particular, since exp(−tL1)x = (x1, x2 − tx3, x3 + tx2) + O(t2) and similarly for exp(−tL2)x and exp(−tL3)x, on account of (4.2), the corresponding fundamental vector fields on R3 are L˜1 = x 3 ∂ ∂x2 − x2 ∂ ∂x3 , L˜2 = x 1 ∂ ∂x3 − x3 ∂ ∂x1 , L˜3 = x 2 ∂ ∂x1 − x1 ∂ ∂x2 . (4.4) On transforming (4.4) to spherical coordinates (r, θ, φ), one finds that the L˜j are inde- pendent of r and of ∂/∂r (as expected). Indeed, L˜± = e±iφ ( ∓i ∂ ∂θ + cot θ ∂ ∂φ ) , L˜3 = − ∂ ∂φ , (4.5) which may be regarded as vector fields either in X(R3 \ {0}) or in X(S2). Exercise 4.1. Verify the formulae (4.5) for the fundamental vector fields. The sphere S2 is a homogeneous space for the rotation group, and may be identified with the coset space SO(3)/SO(2), where SO(2) is the subgroup of east-west rotations which fix the north pole. One may therefore consider a system of local coordinates for S2 which 1Any Lie algebra g gives rise to an associative algebra, called its universal enveloping algebra U(g), whose elements are polynomial combinations of elements of g, reduced by the commutation relations among such elements. The Poincare´–Birkhoff–Witt theorem [35] proves that the natural map from g to U(g) is injective, so that g may be regarded as a subspace of U(g). Now (4.3) may be regarded as an identity in U(so(3,C)). 34 privileges the cartesian coordinate x3. We adopt the following local coordinates for the remainder of this chapter: ζ := x1 + ix2 = eiφ sin θ, x3 = cos θ. We also write ζ¯ := x1 − ix2 = e−iφ sin θ, which supplies a third local coordinate for R3. Let us abbreviate ∂3 := ∂/∂x 3, ∂ζ := ∂/∂ζ, ∂¯ζ := ∂/∂ζ¯. In these coordinates, the fundamental vector fields are given by L˜+ = −2ix3∂¯ζ + iζ∂3, L˜− = 2ix3∂ζ − iζ¯∂3, L˜3 = iζ¯∂¯ζ − iζ∂ζ . (4.6) A polynomial in the variables L˜j gives a differential operator on S2. For instance, the operator corresponding to (4.3) satisfies (L˜−L˜+ + L˜23 − iL˜3)ζ = (L˜3 − i)L˜3ζ = (L˜3 − i)(−iζ) = −2ζ, (4.7) since ∂¯ζ , and therefore also L˜+, vanishes on holomorphic functions of ζ. 4.2 The Hodge operators on the sphere Lemma 4.1. The Hodge star operator on A•(S2) is determined by the relations ?(dζ ∧ dx3) = ζ, ?(dζ) = x3 dζ − ζ dx3, in the (ζ, x3) coordinates. Proof. First observe that dζ ∧ dx3 = (ieiφ sin θ dφ+ eiφ cos θ dθ) ∧ (− sin θ dθ) = ieiφ sin2 θ dθ ∧ dφ = iζΩ, so that ?1 = iΩ = ζ−1 dζ ∧ dx3, and ?(ζ) = dζ ∧ dx3; reciprocally, ?(dζ ∧ dx3) = ζ. Hodge duals of 1-forms are given by ? dζ = eiφ ?(i sin θ dφ) + eiφ cos θ ?(dθ) = eiφ dθ + eiφ cos θ(i sin θ dφ), which simplifies to ?(dζ) = x3 dζ − ζ dx3, and reciprocally, ?(x3 dζ − ζ dx3) = dζ. The codifferential δ = −?d? is now easily found. For instance, δ(dζ) = ?d(ζ dx3−x3 dζ) = ?(2dζ ∧ dx3) = 2ζ, and since δ(ζ) = 0, it follows that ∆(ζ) = (dδ + δd)ζ = δ(dζ) = 2ζ. (4.8) 35 Definition 4.3. Since rotations act on differential forms through R · ω := (R−1)∗ω, their generators act as Lie derivatives : Ljω := d dt ∣∣∣∣ t=0 (exp(−tLj)∗ω) = LeLjω for j = 1, 2, 3. Lie derivatives commute with exterior derivation: LXd = dLX as operators on A•(S2), for any X ∈ X(S2); and, in particular, Ljd = dLj for j = 1, 2, 3. An operator T on A•(S2) is rotation-invariant if T (R · ω) = R · (Tω) for any R ∈ SO(3). Since the three one-parameter subgroups { exp(tLj) : t ∈ R } generate SO(3), this holds iff TLj = LjT for j = 1, 2, 3. For instance, the Hodge star operator ? = i((θˆ 1) − ι(θˆ1]))((θˆ2)−ι(θˆ2])) is unchanged if the local oriented orthonormal frame {θˆ1, θˆ2} is replaced by {R · θˆ1, R · θˆ2}; this says that ? is invariant under rotations, and therefore ?Lj = Lj? for j = 1, 2, 3. Since δ = −?d?, ∆ = dδ + δd, it follows that δLj = Ljδ and ∆Lj = Lj∆ for j = 1, 2, 3. In particular, the Hodge Laplacian is rotation-invariant. Exercise 4.2. Use the Cartan identity LX = ιXd+ dιX to show that LjΩ = 0 for j = 1, 2, 3. What can one conclude from this? The Laplace–Beltrami operator ∆0 (the restriction of ∆ to C ∞(S2)) is thus a second-order differential operator which commutes with rotations. It therefore represents a quadratic element in the centre of the algebra U(so(3)). Now it is known [34, 35] that this centre is the polynomial algebra R[C] generated by the “Casimir element” C = L21 + L22 + L23. Thus ∆0 = a(L˜ 2 1 + L˜ 2 2 + L˜ 2 3) for some constant a; ∆ and each Lj commutes with d, and so ∆ = a(L21 + L 2 2 + L 2 3) = a(L−L+ + L 2 3 − iL3). From (4.7) and (4.8) one sees that a = −1, and therefore ∆ = −L−L+ − L23 + iL3. (4.9) Exercise 4.3. Show directly, using only the commutation relations [Li, Lj] = ij kLk, that any quadratic polynomial in L1, L2, L3 commuting with each Lj must be a multiple of L 2 1+L 2 2+L 2 3. 4.3 Eigenvectors for the Laplacian Since ∆ is invariant under rotations, any subspace of differential forms which is stable under the SO(3) action is mapped by ∆ into another such subspace. The search for eigenvec- tors under ∆ should therefore start with the irreducible subspaces of the representation R 7→ (R−1)∗ of the compact group SO(3) on A•(S2). This is the point of view adopted by Folland [26], who obtained a complete spectral decomposition of the Hodge Laplacian on a sphere of any dimension. Here we show how this works for the 2-sphere. It is useful to recall some facts about representations of compact Lie groups, which may be found in many places, e.g. [13, 37, 53]. Any irreducible unitary representation of a com- pact group G acts on a finite dimensional Hilbert space, and the Hilbert space of any unitary representation may be written as a (possibly infinite) direct sum of irreducible subrepresen- tations. Moreover, by the Peter–Weyl theorem, all such irreducible representations already 36 occur in the decomposition of the regular representation of G on L2(G), and their carrier spaces are in fact subspaces of C∞(G). In the case G = SO(3), the identification SO(3)/SO(2) ≈ S2 goes as follows: elements of SO(3) are parametrized by three local coordinates (φ, θ, ψ), called “Euler angles”, and SO(2) is regarded as the one-parameter subgroup { expψL3 : ψ ∈ R }; the remaining angles (θ, φ) as the spherical coordinates on S2. Thus functions on S2 are identified with functions on SO(3) which are constant on SO(2)-cosets, i.e., functions which do not depend on the variable ψ; in this way the representation of SO(3) on C∞(S2) —or on its completion L2(S2)— becomes a subrepresentation of the regular representation on L2(SO(3)). Its irreducible subspaces are spanned by the spherical harmonics Ylm(θ, φ), where l = 0, 1, 2, . . . and m = −l, . . . , l− 1, l; indeed Hl := span{Ylm : m = −l, . . . , l } is an irreducible subspace of dimension 2l + 1, and the Ylm form an orthonormal basis for L 2(S2). It turns out, again by the Peter–Weyl theorem, that SO(3) has (up to equivalence) exactly one irreducible unitary representation of each odd dimension, so the orthonormality of the spherical harmonics shows that in the decomposition of the rotation action on C∞(S2), each such representation occurs once only, and that there are no other irreducible subrepresentations.2 There is a general principle for finding irreducible representations of a compact Lie group G, called the “theorem of the highest weight” [13, 35, 37]. One finds a maximal torus3 in G, that is, a subgroup T ≤ G which is a torus, i.e., isomorphic to Tk for some k, with k maximal; if G = SO(3), then k = 1 and T := SO(2) ' T will do. In a representation space for G, one looks for a joint eigenvector for the torus T ; when G = SO(3), this is just an eigenvector v for L3, whose eigenvalue is called a “weight” of T . Within gC one finds k “raising elements” which annihilate v (since the weight is “highest”) and k “lowering elements” which, applied successively to v, generate a basis for an irreducible representation space V ; the commutation relations in gC ensure that applying other generators does not enlarge the space V . When G = SO(3), there is one raising element, namely L+, and one lowering element, namely L−. The upshot of the general theory is this. Within each space of k-forms on the sphere (k = 0, 1, 2), we must find forms α satisfying L+α = 0 and L3α = cα for some eigenvalue c, and such that the vector space span{Lr−α : r ∈ N } is finite dimensional. The identity (4.9) guarantees that α is also an eigenvector for the Laplacian ∆. Definition 4.4. From (4.6), the identity L+f = 0 is satisfied whenever f ∈ C∞(S2) is holomorphic in ζ and independent of x3. Since (L3f)(ζ) = −iζ f ′(ζ), this f is an eigenvector for L3 iff it is of the form f(ζ) = ζ l for some l; since ζ l = eilφ sinl θ, the smoothness of f forces the condition l ∈ N. Define φ0l ∈ A0(S2) by φ0l(ζ, x3) := ζ l, for l = 0, 1, 2, . . . . Exercise 4.4. Show that Lk−φ0l is a linear combination of terms (x 3)k−2rζ¯rζ l−k+r and check that Lk−φ0l 6= 0 for k = 0, 1, . . . , 2l but L2l+1− φ0l = 0. Conclude that the functions {R · φ0l : 2For other compact groups, the completeness of the decomposition of the natural representation on 0- forms on a homogeneous space may be obtained from a counting argument based on Frobenius reciprocity. For the case of SO(n) and the sphere Sn−1, we refer to Folland [26]. 3Any compact connected Lie group is the union of all its maximal tori, and any two maximal tori are conjugate, by a theorem of Weyl: see [13] for a proof. 37 R ∈ SO(3) } span an irreducible representation space for SO(3), of dimension 2l + 1, for each l ∈ N. Exercise 4.5. Prove that the functions {Lr−φ0l : l ∈ N, r = 0, . . . , 2l } span a dense subset of L2(S2) by showing that any homogeneous polynomial in the variables ζ, ζ¯, x3 is a linear combination of these. Exercise 4.6. Check that 〈〈φ0l | φ0m〉〉 = 0 for l 6= m, and that 〈〈α | L−β〉〉 = 〈〈L+α | β〉〉 for α, β ∈ A0(S2), using (4.6). Show that [L+, L−] = −2iL3 and [L3, L−] = iL− and conclude that L+L r −φ0l = alrL r−1 − φ0l for some constant alr. Deduce that the functions L r −φ0l, when suitably normalized, yield an orthonormal basis for L2(S2). We have thus identified a complete set of irreducible subrepresentations of the rotation action on A0(S2). Note that L3(ζ l) = −ilζ l, so the L3-eigenvalue is −il. It is immediate from (4.9) that ∆(ζ l) = (l2 + l) ζ l. We take stock that L+φ0l = 0, L3φ0l = −il φ0l, ∆φ0l = l(l + 1)φ0l. (4.10) It is now clear that the Laplace–Beltrami operator ∆0 is a formally selfadjoint, positive operator on L2(S2), with spectrum sp(∆0) = { l(l + 1) : l ∈ N }. Since ∆0 commutes with L−, each Lr−φ0l (r = 0, . . . , 2l) is an eigenvector for the eigenvalue l(l + 1), which therefore has multiplicity 2l+ 1. The set sp(∆0), with these multiplicities, is sometimes referred to as the spectrum of the Riemannian manifold S2 [8]. Note, in particular, that ker ∆0 = Harm 0(S2) is the one-dimensional space spanned by φ00, i.e., the space of constant functions. Thus, the only harmonic functions on S2 are the constants. 4.4 Spectrum of the Hodge Laplacian Definition 4.5. Since ∆ commutes with the exterior derivative and the Hodge star operator, we can manufacture more eigenvectors by applying these operators to the φ0l. Since dφ00 = 0 because φ00 is constant, dφ0l is an eigenvector only for l ≥ 1. Define ψ1l := l −1 dφ0l = ζ l−1 dζ, φ1l := −?ψ1l = −ζ l−1?dζ = ζ l dx3 − ζ l−1x3 dζ, for l = 1, 2, 3, . . . . Since L+, L3 and ∆ each commutes with d and ?, we obtain from (4.10): L+ψ1l = 0, L3ψ1l = −il ψ1l, ∆ψ1l = l(l + 1)ψ1l, L+φ1l = 0, L3φ1l = −il φ1l, ∆φ1l = l(l + 1)φ1l. Thus the 1-forms ψ1l and φ1l are highest-weight vectors for irreducible representations of SO(3), of dimension 2l + 1; in fact, the representation spaces are spanned respectively by {Lr−ψ1l : r = 0, . . . , 2l } and {Lr−φ1l : r = 0, . . . , 2l }. Exercise 4.7. Show that 〈〈ψ1l | ψ1m〉〉 = 〈〈φ1l | φ1m〉〉 = 0 for l 6= m, and that 〈〈ψ1l | φ1m〉〉 = 0 for all l,m ≥ 1. 38 Exercise 4.8. Check that 〈〈α | L−β〉〉 = 〈〈L+α | β〉〉 for α, β ∈ A1(S2), and conclude that the 1-forms {Lr−ψ1l,Lr−φ1l : l ≥ 1, r = 0, . . . , 2l } are orthogonal. The 1-forms ψ1l are exact, so applying d to them yields only zero. However, the 1-forms φ1l are coexact, since φ1l = −?dφ0l = δ(?φ0l), so we may define ψ2l := (l + 1) −1 dφ1l = ζ l−1 dx3 ∧ dζ = iζ lΩ for l = 1, 2, 3, . . . . Since the expression iζ lΩ also makes sense for l = 0, we also define ψ20 := iΩ. Then L+ψ2l = 0, L3ψ2l = −il ψ2l, ∆ψ2l = l(l + 1)ψ2l for l = 0, 1, 2, . . . . Furthermore, since ?1 = iΩ, we have ?ψ2l = φ0l for l ∈ N. We may summarize this information with two commutative diagrams: φ0l ?←−−− ψ2l l−1d y x(l+1)−1d ψ1l −?−−−→ φ1l φ0l ?−−−→ ψ2l (l+1)−1δ x yl−1δ ψ1l −?←−−− φ1l (4.11) where l ≥ 1. The first diagram takes stock of the foregoing definitions, and the vertical arrows in the second diagram are formed by composing three arrows from the first, using the identity δ = −?d?. From this it is evident that the ψkl are exact forms and the φkl are coexact forms. A complete circuit of four arrows in either diagram corresponds to applying the operator (l(l + 1))−1(−?d?d− d?d?) = (l(l + 1))−1∆, which acts as the identity at each vertex. We now have a complete set of eigenforms for ∆. Indeed, we have shown already that the Lr−φ0l are a complete set of eigenforms of degree 0. Since ? is a bijection between 0-forms and 2-forms, the Lr−ψ2l, for l ∈ N, r = 0, . . . , 2l densely span A2(S2). For the 1-forms, we may use the Hodge decomposition (3.10); the diagrams (4.11) show that d : δA1(S2) → dA0(S2) and δ : dA1(S2) → δA2(S2) are bijections, that the ψ1l densely span dA0(S2), and that the φ1l densely span δA 2(S2). It remains only to observe that there are no nonzero harmonic 1-forms. This can be seen by noting that by writing any 1-form as α = f(ζ) dζ+g(x3) dx3 + h(ζ, x3)(x3dζ − ζ dx3), since f(ζ) dζ + g(x3) dx3 is exact, and h(x3dζ − ζ dx3) = ?(h dζ) is coexact. Thus, {Lr−ψ1l,Lr−φ1l : l ≥ 1, r = 0, . . . , 2l } forms an orthogonal basis for A1(S2). The spectrum of the Hodge Laplacian is therefore sp(∆) = { l(l + 1) : l ∈ N }, with multiplicities 4l(l + 1) for the eigenvalue l(l + 1) when l ≥ 1, and multiplicity 2 for the zero eigenvalue. Indeed, ker ∆ = span{φ00, ψ20} = Harm•(S2). If K denotes the operator on L2,•(S2) which is zero on Harm•(S2) and inverts ∆ on its orthogonal complement, then K is a bounded selfadjoint operator and, moreover, is compact, since its spectrum sp(K) = { (l(l+ 1))−1 : l ∈ N } accumulates only at 0 and consists of eigenvalues with finite multiplicities. Also, ∆K = K∆ = I − P , where P is the orthogonal projector of rank 2 with range Harm•(S2). 39 Corollary 4.2. The de Rham cohomology spaces of S2 are given by H•dR(S2) ' Harm•(S2) ' R⊕ 0⊕ R, (4.12) on decomposing ker ∆ by degrees of forms. Notice that this coincides with the Cˇech cohomology Hˇ•(S2,R), computed in subsec- tion 1.9. Definition 4.6. An important topological invariant of a manifold M is its Euler character- istic4 χ(M) := dimM∑ k=0 (−1)k dimHkdR(M). For the 2-sphere, (4.12) yields χ(S2) = 1− 0 + 1 = 2. 4.5 The Hodge–Dirac operator Definition 4.7. The Hodge–Dirac operator on the 2-sphere S2 is the operator D/ := d+δ, whose square is the Hodge Laplacian: D/ 2 = ∆. Let Aeven(S2) := A0(S2) ⊕ A2(S2) be the algebra of differential forms of even degree on S2, and write Aodd(S2) := A1(S2) to denote the odd-degree forms. Then D/ is an odd operator in the sense that it interchanges forms of even and odd parities: D/ (Aeven(S2)) ⊆ Aodd(S2) and D/ (Aodd(S2)) ⊆ Aeven(S2). From (4.11), the action of D/ is given explicitly by D/φ0l = l ψ1l, D/ψ1l = (l + 1)φ0l, D/ψ2l = l φ1l, D/ φ1l = (l + 1)ψ2l, (4.13) for l = 0, 1, 2, . . . on the left, and l = 1, 2, 3, . . . on the right. It follows that D/ (√ l + 1φ0l ± √ l ψ1l ) = ± √ l(l + 1) (√ l + 1φ0l ± √ l ψ1l ) , D/ (√ l + 1ψ2l ± √ l φ1l ) = ± √ l(l + 1) (√ l + 1ψ2l ± √ l φ1l ) . Since these eigenforms for D/ densely span L2,•(S2), one concludes that D/ is a formally selfadjoint5 unbounded linear operator on L2,•(S2), with spectrum sp(D/ ) = {±√l(l + 1) : l ∈ N }. Exercise 4.9. What are the multiplicities of the eigenvalues of D/ ? 4The Euler characteristic may be computed as the integral of a certain differential form over M , as we shall see later. 5The term “formally selfadjoint” means that 〈〈D/ω | η〉〉 = 〈〈ω |D/η〉〉 for ω, η ∈ A•(S2); operator theorists would say that D/ is “symmetric”. With some more work, one can check that D/ is essentially selfadjoint, which means that it has an extension to a larger domain in the Hilbert space L2,•(S2) which is a closed, unbounded selfadjoint operator. See [39] for a proof of this. 40 It also follows from (4.13) that kerD/ = span{φ00, ψ20}, and that D/ has dense range, so that cokerD/ = 0; moreover, the restriction of D/ to L2,•(S2) Harm•(S2) has a compact inverse. Thus D/ is a Fredholm operator.6 Its index is given by indD/ := dim(kerD/ )− dim(cokerD/ ) = 2− 0 = 2. Corollary 4.3. indD/ = χ(S2). This is a first example of an index theorem, wherein a certain integer obtained by an integral over the manifold (namely, the Euler characteristic) turns out to be equal to the Fredholm index of an operator (namely, the Hodge–Dirac operator) which is bound up with the geometric structure of the manifold. For a wider discussion of index theorems in modern geometry and topology, we refer to [9, 28, 39, 42]. There is an equivalent method of computing the index from the grading of A•(S2) into forms of even and odd degree. Write L2,•(S2) = L2,even(S2) ⊕ L2,odd(S2) where L2,even(S2) and L2,odd(S2) are the respective completions of Aeven(S2) and Aodd(S2); this is then a graded Hilbert space. The odd operator may be written as D/ = ( 0 D/ 1 D/ 0 0 ) , where D/ 0 : L 2,even(S2)→ L2,odd(S2) and D/ 1 : L2,odd(S2)→ L2,even(S2). The selfadjointness of D/ says that D/ 1 = D/ † 0, and this may be verified directly from (4.13) also. Now cokerD/ 1 ' (kerD/ 0) ⊥ and cokerD/ 0 ' (kerD/ 1)⊥, so the index of D/ equals indD/ = dim(kerD/ 0)− dim(kerD/ 1). (4.14) From (4.13), the right hand side of (4.14) equals 2− 0 = 2, as expected. Exercise 4.10. Show that l 〈〈ψ1l |ψ1l〉〉 = (l+1) 〈〈φ0l |φ0l〉〉 and that l 〈〈φ1l |φ1l〉〉 = (l+1) 〈〈ψ2l | ψ2l〉〉. Deduce from (4.13) that D/ 0 and D/ 1 are adjoints of each other. The Hodge–Dirac operator is one of many operators which are collectively known as Dirac operators. Some common properties are: (i) they are unbounded selfadjoint Fredholm oper- ators; (ii) the Hilbert spaces on which they act are graded into “even” and “odd” subspaces, which the Dirac operators interchange; (iii) their squares are “generalized Laplacians” (of which we shall have more to say later) acting on Riemannian manifolds. Definition 4.8. As a second example, let us redefine the grading of differential forms by splitting A•(S2) into the (±1)-eigenspaces for the Hodge star operator: thus A+(S2) := {ω : ?ω = ω } is the space of selfdual forms, and A−(S2) := { η : ?η = −η } is the set of antiselfdual forms. Now δ? = −?d implies that D/? = −?D/ , so the same operator D/ = d+ δ interchanges selfdual and antiselfdual forms: D/ (A±(S2)) ⊆ A∓(S2). 6Conventionally, Fredholm operators are taken to be bounded, whereas D/ and ∆ are not. One could remedy this by redefining the norm on the domain space, but then the domain and range would lie in different Hilbert spaces. Instead, we use the alternative definition [22] of a Fredholm operator as a closed, possibly unbounded, operator between Hilbert spaces which has dense domain, finite-dimensional kernel and finite-codimensional (therefore closed) range. In this sense, the operator closures of D/ and ∆ are unbounded Fredholm operators on L2,•(S2). 41 The diagrams (4.11) show that the forms ψ2l + φ0l (l ≥ 0) and φ1l − ψ1l (l ≥ 1) are selfdual, whereas ψ2l − φ0l (l ≥ 0) and φ1l + ψ1l (l ≥ 1) are antiselfdual. Since D/ commutes with L−, applying powers to L− to these forms generates a complete set [i.e., a set which spans a dense subspace of L2,•(S2)]. Thus we may rewrite (4.13) as D/ (ψ2l + φ0l) = l (φ1l + ψ1l), D/ (φ1l − ψ1l) = (l + 1) (ψ2l − φ0l), D/ (φ1l + ψ1l) = (l + 1) (ψ2l + φ0l), D/ (ψ2l − φ0l) = l (φ1l − ψ1l), where l ≥ 1 in all cases; and D/ (ψ20 ± φ00) = 0. Observe that, in the new grading of L2,•(S2), the odd operator D/ can be written as D/ = ( 0 D/ − D/ + 0 ) , (4.15) where D/ ± : L2,±(S2)→ L2,∓(S2). Now we define its index as indD/ := dim(kerD/ +)− dim(kerD/ −), (4.16) which equals 1− 1 = 0 since the kerD/ ± are one-dimensional spaces, spanned by (ψ20±φ00). The corresponding topological quantity [9] arises from the bilinear form s([α], [β]) :=∫ S2 α ∧ β on H1dR(S2). The integral depends only on the cohomology classes [α] and [β], and s is antisymmetric; we assign to s a “signature” of zero.7 The index theorem for this case is the equality of this zero signature with the zero index for D/ as defined by (4.15). 5 Connections on vector bundles Line bundles are classified by integral Cˇech 2-cocycles, according to the theory developed in Section 1. In this chapter, we show how to produce, for a given Hermitian line bundle E−→M , a Cˇech 2-cocycle which is associated to its class. This 2-cocycle comes from the curvature form of a connection on the line bundle. A connection, or covariant derivative, is a general structure which supplements that of a vector bundle with a notion of “parallel displacement” among neighbouring fibres. There are several possible ways to introduce connections; we shall adopt here an algebraic approach, in the spirit of Cartan calculus of vector fields and forms. 5.1 Modules of vector-valued forms Lemma 5.1. Suppose that E−→M and E ′−→M are two vector bundles over the (compact) manifold M . If τ : E → E ′ is a bundle map, i.e., a smooth map such that (τ, idM) is a vector bundle morphism, there is an C∞(M)-linear map τ∗ : Γ(E) → Γ(E ′) given by τ∗s := τ ◦ s; and the correspondence τ 7→ τ∗ satisfies (idE)∗ := idΓ(E) and (τ ◦ σ)∗ = τ∗ ◦ σ∗. 7A less trivial example arises on considering a Riemannian manifold M of dimension n = 4k; the corre- sponding formula yields a symmetric bilinear form q([α], [β]) := ∫ M α ∧ β on H2kdR(M), and the topological invariant is the signature of this bilinear form. 42 Proof. The C∞(M)-linearity of τ means that τ∗(fs) = fτ∗s for s ∈ Γ(E), f ∈ C∞(M), which follows from the linearity of τ : Ex → E ′x for each x ∈ M , since τ∗(fs)(x) := τ(f(x)s(x)) = f(x)τ(s(x)) = (f τ∗s)(x). The remaining assertions are obvious. Let A := C∞(M) be the algebra of smooth functions on M . One way to restate the preceding Lemma is to say that E 7→ Γ(E), τ 7→ τ∗ is a covariant functor Γ: Vect(M) → Mod(A) from the category of vector bundles over M to the category of A-modules. Exercise 5.1. Verify that Γ(E∗) = Γ(E)∗, where E∗ → M is the dual vector bundle to E−→M , and the notation E∗, for an A-module E, denotes the module of A-linear maps from E to A. Definition 5.1. If A is any (real or complex) algebra, a bimodule E over A is a vector space with bilinear operations A × E → E and E × A → E, usually simply as (a, s) 7→ as and (s, a) 7→ sa, satisfying 1 s = s 1 = s, a(a′s) = (aa′)s, and (sa′)a = s(a′a), for s ∈ E, a, a′ ∈ A. The tensor product E ⊗A E′ of two such bimodules is the bimodule1 whose elements are finite sums ∑ j sj ⊗ s′j with sj ∈ E and s′j ∈ E′, subject only to the relations (sa)⊗s′−s⊗ (as′) = 0, for s ∈ E, s′ ∈ E′, a ∈ A. When A is commutative, the identification sa = as makes each A-module into an A-bimodule; the foregoing recipe defines the tensor product of A-modules in that case. Proposition 5.2. Let A := C∞(M) be the algebra of smooth functions on a compact ma- nifold M , and let E−→M , E ′−→M be vector bundles over M . Then there is a canonical isomorphism of A-modules: Γ(E)⊗A Γ(E ′) ' Γ(E ⊗ E ′). Proof. If s ∈ Γ(E), s′ ∈ Γ(E ′), let s ⊗ s′ denote the section x 7→ s(x) ⊗ s′(x) of the tensor product bundle E⊗F −→M . We provisionally denote by s⊗As′ the element of Γ(E)⊗AΓ(E ′) given by the definition of the tensor product ofA-modules. Let θ : Γ(E)⊗AΓ(E ′)→ Γ(E⊗E ′) be the A-linear map determined by θ(s⊗As′) := s⊗s′; the claim is that θ is an isomorphism. If U is a chart domain in M over which the bundles E−→M and E ′−→M are trivial, then E ⊗ E ′−→M is also trivial over U . Indeed, we have seen in subsection 1.7 that any local section t ∈ Γ(U,E) is of the form t = ∑rk=1 hksk, where {s1, . . . , sr} is a local system of sections for E over U , and h1, . . . , hr ∈ C∞(U); in other words, the sections {s1, . . . , sr} generate Γ(U,E) freely over C∞(U). If {s′1, . . . , s′l} is a local system of sections for E ′ over U , they generate Γ(U,E ′) freely as a C∞(U)-module, and it is clear that { sj ⊗ s′k : j = 1, . . . , r; k = 1, . . . , l } generate Γ(U,E⊗E ′) as a free C∞(U)-module, since { sj(x)⊗ s′k(x) : j = 1, . . . , r; k = 1, . . . , l } is a basis for Ex ⊗ E ′x for each x ∈ U . In summary, θ is an isomorphism whenever the bundles E−→M and E ′−→M (and consequently E⊗E ′−→M) are trivial. In the general case, there are vector bundles F −→M and F ′−→M such that E ⊕ F −→M and E ′ ⊕ F ′−→M are trivial,2 by Proposition 1.8. Let ι : E → E ⊕ F and 1The bimodule operations on E⊗AE′ are, of course, defined by a(s⊗s′) := (as)⊗s′ and (s⊗s′)a := s⊗(s′a). 2The use of Proposition 1.8 (existence of the supplementary bundle) is the only point in this proof where the compactness of M is used. It should be said that, with some work to establish the existence of a finite trivialising open covering of M , the compactness assumption can be dropped: we refer to [17] for a proof. 43 σ : E ⊕ F → E be the extension and restriction maps: ι(u) := (u, 0), σ(u, v) := u, and let ι′ : E ′ → E ′⊕F ′, σ′ : E ′⊕F ′ → E ′ be similarly defined. Then σ ◦ ι = idE, so σ∗ ◦ ι∗ = idΓ(E) and also σ′∗ ◦ ι′∗ = idΓ(E′); thus, ι∗ and ι′∗ are injective, whereas σ∗ and σ′∗ are surjective. Now Ex⊗E ′x is a direct summand of the vector space (Ex⊕Fx)⊗(E ′x⊕F ′x) for each x ∈M ; this yields bundle maps ι′′ : E⊗E ′ → (E⊕F )⊗ (E ′⊕F ′), σ′′ : (E⊕F )⊗ (E ′⊕F ′)→ E⊗E ′ satisfying σ′′∗ ◦ ι′′∗ = idΓ(E⊗E′). Finally, let ι∗ ⊗ ι′∗ : Γ(E)⊗A Γ(E ′)→ Γ(E ⊕ F )⊗A Γ(E ′ ⊕ F ′) be defined by (ι∗ ⊗ ι′∗)(s ⊗ s′) := ι∗s ⊗ ι′∗s′, and let σ∗ ⊗ σ′∗ : Γ(E ⊕ F ) ⊗A Γ(E ′ ⊕ F ′) → Γ(E)⊗A Γ(E ′) be defined similarly. We thus have two commutative diagrams: Γ(E)⊗A Γ(E ′) θ−−−→ Γ(E ⊗ E ′) ι∗⊗ι′∗ y yι′′∗ Γ(E ⊕ F )⊗A Γ(E ′ ⊕ F ′) Θ−−−→ Γ((E ⊕ F )⊗ (E ′ ⊕ F ′)) (5.1) where Θ is the isomorphism of free A-modules already obtained, and Γ(E)⊗A Γ(E ′) θ−−−→ Γ(E ⊗ E ′) σ∗⊗σ′∗ x xσ′′∗ Γ(E ⊕ F )⊗A Γ(E ′ ⊕ F ′) Θ−−−→ Γ((E ⊕ F )⊗ (E ′ ⊕ F ′)) (5.2) From (5.1), θ is injective, since ι∗⊗ ι′∗ and ι′′∗ are injective and Θ is bijective; and (5.2) shows analogously that θ is surjective. Thus θ is an A-linear isomorphism in the general case. Corollary 5.3. Each A-linear map from Γ(E) to Γ(E ′) is of the form τ∗ for some bundle map τ : E → E ′. Proof. The maps τ : E → E ′ form the total space of the vector bundle Hom(E,E ′)−→M , whose fibres are Hom(Ex, E ′ x) ' E∗x ⊗ E ′x. Thus τ can be identified with a section of the vector bundle E∗ ⊗ E ′−→M . On the other hand, an A-linear map from Γ(E) to Γ(E ′) belongs to HomA(Γ(E),Γ(E ′)) ' Γ(E)∗ ⊗A Γ(E ′) ' Γ(E∗) ⊗A Γ(E ′); it is easily seen that τ∗ ∈ HomA(Γ(E),Γ(E ′)) corresponds to θ−1(τ) ∈ Γ(E∗)⊗AΓ(E ′) under these identifications. Since θ is bijective, these account for all A-linear maps from Γ(E) to Γ(E ′). Definition 5.2. Consider the vector bundle E ′ = ΛrT ∗M , whose smooth sections are the r-forms over M : Ar(M) = Γ(ΛrT ∗M). Define Ar(M,E) := Γ(E)⊗A Ar(M) ' Γ(E ⊗ ΛrT ∗M), (5.3) using Proposition 5.2. The elements of the A-module Ar(M,E) are finite sums of the form∑ k sk ⊗ ωk, where sk ∈ Γ(E), ωk ∈ Ar(M); we shall refer to them as “E-valued r-forms over M”. On a complex manifold, we may also define Ap,q(M,E) := Γ(E)⊗A Ap,q(M). 44 5.2 Connections Definition 5.3. A connection on a complex vector bundle E−→M is a C-linear3 map ∇ : Γ(E)→ A1(M,E) such that ∇(sf) = (∇s)f + s⊗ df for all s ∈ Γ(E), f ∈ C∞(M). (5.4) This Leibniz rule4 shows that the definition is local, i.e., that ∇s is determined by its restrictions to chart domains Uj. To see that, let {fj} be a smooth partition of unity subordinate to {Uj}, and notice that ∑ j dfj = d (∑ j fj ) = d(1) = 0, and thus ∇s =∑ j∇(sfj) by (5.4). This locality is immediately useful in showing that connections exist on any vector bundle. First, if E−→M is a trivial bundle, E ≈M×Cr, so that Γ(E) ' Ar, the exterior derivative d : Ar → A1(M)r : (h1, . . . , hr) 7→ (dh1, . . . , dhr) is a connection, since d(hkf) = dhk f+hk df for k = 1, . . . , r, i.e., d(h f) = (dh)f+h⊗df for h ∈ Ar. In the general case, over each chart domain Uj of M one can choose a local system of sections sj = (sj1, . . . , sjr) for E−→M ; the expansion t = ∑ k sjk h k gives an isomorphism ψ∗j : Γ(Uj, E) → C∞(Uj)r : t 7→ h. (This is just the pullback to sections of the local trivialization ψj of (1.1)). Since A 1(Uj, E) ' Γ(Uj, E)⊗C∞(Uj)A1(Uj), we get an isomorphism ψ∗j ⊗ id : A1(Uj, E)→ A1(Uj)r on tensoring with A1(Uj). Now define ∇(j) := (ψ∗j ⊗ id)−1 ◦d◦ψ∗j , so that ∇(j) : Γ(Uj, E)→ A1(Uj, E) is a connection on pi−1(Uj)−→Uj. Finally, take any smooth partition of unity {fj} subordinate to the cover {Uj}, and define ∇s := ∑ j∇(j)(sfj); one checks that ∇ is a connection on the vector bundle E−→M . Exercise 5.2. Carry out this check. The Leibniz rule (5.4) means that ∇ is not itself A-linear. However, if ∇0 and ∇1 are two connections on E−→M , it is immediate from (5.4) that (∇1−∇0)(sf) = (∇1s−∇0s)f for s ∈ Γ(E) and f ∈ A. By Corollary 5.3, ∇1 −∇0 = α∗ for some α ∈ Γ(EndE ⊗ T ∗M) = A1(M,EndE). Therefore, ∇1s = ∇0s+ α ◦ s for s ∈ Γ(E); (5.5) conversely, given ∇0, this equation defines a connection ∇1 when α ∈ A1(M,EndE). This says that the set of all connections on E−→M is an affine space, based on the vector space A1(M,EndE). Definition 5.4. If X ∈ X(M) is a vector field, the contraction ιX : A1(M)→ A is A-linear, since ιX(fβ) = f β(X) = f ιXβ for f ∈ A, β ∈ A1(M). Thus it extends to an A-linear map from A1(M,E) = Γ(E)⊗A A1(M) to Γ(E)⊗A A = Γ(E) by ιX(s⊗ β) := s(ιXβ) = s β(X). If ∇ : Γ(E)→ A1(M,E) is a connection, we write ∇X := ιX ◦ ∇ : Γ(E)→ Γ(E). This is a C-linear map satisfying the Leibniz rule: ∇X(sf) = (∇Xs)f + s(Xf). (5.6) 3For real vector bundles, we require only that ∇ be an R-linear map. 4It is convenient to regard Γ(E) and A1(M,E) as right A-modules; if one wishes to regard them as left A-modules by identifying fs = sf , the Leibniz rule can equivalently be written as ∇(fs) = f(∇s) + df ⊗ s. 45 Definition 5.5. If E−→M is a Hermitian vector bundle, the metric h on E defines a A- sesquilinear map Γ(E)×Γ(E)→ A by (s|t) : x 7→ hx(s(x), t(x)), for s, t ∈ Γ(E). This extends to a sesquilinear form Γ(E) × Ak(M,E) → Ak(M) by tensoring, i.e., (s | t ⊗ ω) := (s | t)ω when ω ∈ Ak(M). A connection ∇ on E−→M is compatible with the metric if (∇s | t) + (s | ∇t) = d(s | t) for all s, t ∈ Γ(E). (5.7) Notice that this is yet another variant of the Leibniz rule. If X ∈ X(M), a compatible connection satisfies (∇Xs | t) + (s | ∇Xt) = X(s | t). Exercise 5.3. In general, for any connection ∇ on a vector bundle E−→M , one may define a dual connection ∇∗ on the dual bundle E∗−→M by stipulating that the following Leibniz rule should hold: 〈∇∗ξ, s〉+〈ξ,∇s〉 = d〈ξ, s〉, where 〈·, ·〉 denotes the evaluation map Γ(E∗)× Γ(E) → A. Verify that, for given ∇ and ξ ∈ Γ(E∗), this Leibniz rule determines a well- defined element ∇∗ξ of A1(M,E∗), and that the operator ∇∗ thus obtained is a connection. Exercise 5.4. Suppose that E−→M and E ′−→M are equivalent vector bundles and that τ : E → E ′ is an invertible bundle map. Extend τ∗ to an operator from Ak(M,E) to Ak(M,E ′) by tensoring: τ∗(s⊗ω) := τ∗s⊗ω for s ∈ Γ(E), ω ∈ Ak(M). If ∇ is a connection on E−→M , show that ∇′ := τ∗ ◦ ∇ ◦ τ−1∗ is a connection on E ′−→M . 5.3 Curvature of a connection Definition 5.6. If ∇ : Γ(E)→ A1(M,E) is a connection on a vector bundle E−→M , there is a canonical way to extend∇ to a linear map∇ : Ak(M,E)→ Ak+1(M,E). SinceAk(M,E) is a vector space generated by elements of the form s ⊗ ω, for s ∈ Γ(E), ω ∈ Ak(M), it suffices to define ∇(s⊗ ω) := (∇s) ∧ ω + s⊗ dω for s ∈ Γ(E), ω ∈ Ak(M). (5.8) If η ∈ A•(M), it follows that ∇(s⊗ (ω ∧ η)) = (∇s) ∧ ω ∧ η + s⊗ dω ∧ η + (−1)k s⊗ ω ∧ dη, from which one obtains the “graded Leibniz rule”: ∇(ζ ∧ η) = (∇ζ) ∧ η + (−1)k ζ ∧ dη if ζ ∈ Ak(M,E), η ∈ A•(M). (5.9) Exercise 5.5. Verify that the extended map ∇ : Ak(M,E)→ Ak+1(M,E) is well-defined, i.e., if ∑ j sj ⊗ ωj = ∑ r tr ⊗ ηr, with sj, tr ∈ Γ(E) and ωj, ηr ∈ Ak(M), then ∑ j∇(sj) ⊗ ωj =∑ r∇(tr)⊗ ηr. The iterated map ∇2 : Γ(E)→ A2(M,E) satisfies ∇2(sf) = ∇((∇s)f)+∇(s⊗ df) = (∇2s)f −∇s ∧ df +∇s ∧ df + s⊗ d(df) = (∇2s)f, for s ∈ Γ(E), f ∈ A. Thus ∇2 is an A-linear map from Γ(E) to A2(M,E) = Γ(E⊗Λ2T ∗M); by Corollary 5.3, ∇2s = ωs with ω ∈ Γ(EndE ⊗ Λ2T ∗M) = A2(M,EndE). This “matrix- valued 2-form” ω is called the curvature of the connection ∇. 46 Exercise 5.6. If τ : E → E ′ is an invertible bundle map between equivalent vector bundles over M , and if ∇, ∇′ are connections on E and E ′ related by ∇′ = τ∗ ◦ ∇ ◦ τ−1∗ , show that their curvatures are likewise related by ω′ = τ∗ ◦ ω ◦ τ−1∗ . On the trivial bundle E = M ×Cr, let d be the connection (h1, . . . , hr) 7→ (dh1, . . . , dhr) given by the exterior derivative. From (5.5), we can write ∇s = ds + αs where α ∈ A1(M,EndE) = A1(M)r×r is a matrix of 1-forms on M . Explicitly, if s ∈ Ar has k-th component hk, then αs has k-th component αkl h l, so α = [αkl ] ∈ A1(M)r×r. In this case, the curvature ω is a matrix of 2-forms: ωs = ∇2s = d(ds+ αs) + α ∧ (ds+ αs) = ((dα)s− α ∧ ds) + (α ∧ ds+ (α ∧ α)s) = (dα + α ∧ α) s, (5.10) so that ω = dα + α ∧ α in A2(M)r×r. In components, ωkl = dαkl + αkm ∧ αml . 5.4 A curvature formula Lemma 5.4. The curvature ω of a connection ∇ on a line bundle E−→M satisfies ω(X, Y )s = ∇X∇Y s−∇Y∇Xs−∇[X,Y ]s (5.11) for all X, Y ∈ X(M) and s ∈ Γ(E). Proof. Since ω ∈ A2(M,EndE) = Γ(EndE)⊗AA2(M), the evaluation on the pair of vector fields X, Y yields ω(X, Y ) ∈ Γ(EndE), so that ω(X, Y )s ∈ Γ(E). Let the right hand side of (5.11) be denoted provisionally by F (X, Y )s. We claim that s 7→ F (X, Y )s is A- linear. To see that, denote by f˜ the operator on Γ(E) of right multiplication by f ∈ A; then the Leibniz rule (5.6) can be rewritten as [∇X , f˜ ] = X˜f . The desired A-linearity of F (X, Y ) = [∇X ,∇Y ]−∇[X,Y ] follows from [F (X, Y ), f˜ ] = [[∇X ,∇Y ], f˜ ]− [∇[X,Y ], f˜ ] = [∇X , [∇Y , f˜ ]] + [[∇X , f˜ ],∇Y ]− [∇[X,Y ], f˜ ] = [∇X , Y˜ f ] + [X˜f,∇Y ]− ([X, Y ]f )˜ = (X(Y f)− Y (Xf)− [X, Y ]f )˜ = 0, on using the Jacobi identity. Thus, by Corollary 5.3, F (X, Y ) lies in Γ(EndE), for each X, Y ∈ X(M). Furthermore, the formula ιhX = h˜ιX for h ∈ A, which entails ∇hX = h˜∇X , shows that F is A-bilinear in (X, Y ), and so F ∈ A2(M,EndE). We must now show that ω and F coincide. It is enough to check this locally, since if {fj} is a partition of unity on M , the identities ω(X, Y )s = ∑ j ω(X, Y )sfj, F (X, Y )s =∑ j F (X, Y )sfj show that it suffices to prove ω(X, Y )s = F (X, Y )s when X, Y and s vanish outside a chart domain Uj. Thus we may suppose that E−→M is a trivial bundle and indeed that E = M × Cr, whereupon ω = dα + α ∧ α if α ∈ A2(M)r×r is the matrix of 1-forms satisfying ∇ = d+ α. 47 Let us denote by h the element (hi, . . . , hr) of A r = Γ(E); then ∇Xh = ιX(dh+ αh) has k-th component Xhk + αkl (X)h l. Thus F (X, Y )h has k-th component X(Y hk + αkl (Y )h l) + αkl (X)Y h l + αkm(X)α m l (Y )h l − Y (Xhk + αkl (X)hl) − αkl (Y )Xhl − αkm(Y )αml (X)hl − [X, Y ]hk − αkl ([X, Y ])hl = X(αkl (Y ))h l − Y (αkl (X)hl)− αkl ([X, Y ])hl + (αkm(X)αml (Y )− αkm(Y )αml (X))hl = dαkl (X, Y )h l + (αkm ∧ αml )(X, Y )hl, that is, F (X, Y ) is a matrix of 2-forms with components (dαkl + α k m ∧ αml )(X, Y ). Hence F = dα + α ∧ α = ω in A2(M,EndE). Exercise 5.7. Verify that F (gX, hY )s = F (X, Y )s gh for g, h ∈ A. The connection ∇ is determined by local 1-forms αj ∈ A1(Uj,EndE), and by a local system sj in Γ(Uj, E) with respect to which the identification h↔ sjkhk is made. It should be carefully noted that the local 1-forms αj do not patch together to give a global 1-form α ∈ A1(M,EndE) unless E−→M is a trivial bundle, since there is no global d for which ∇ = d+α unless the vector bundle is trivial.5 However, the local 2-forms ωj = dαj+αj∧αj ∈ A2(Uj,EndE) do patch together, since they are restrictions to the Uj of the global 2-form ω. When L−→M is a line bundle, there is a simplification: by Lemma 1.2, the line bundle EndL−→M is trivial, so the curvature form ω belongs to A2(M), since A2(M,EndL) ' A2(M) ⊗A Γ(EndL) ' A2(M) ⊗A A = A2(M) via canonical isomorphisms. Furthermore, there are local 1-forms αj ∈ A1(M) such that ω = dαj on each chart domain Uj, so that the curvature ω is locally exact and hence is a closed 2-form on M . The curvature form ω depends on the connection ∇, but its de Rham class [ω] ∈ H2dR(M) does not. To see that, recall from (5.5) that if ∇0 and ∇1 are two connections on L, with respective curvatures ω0 and ω1, then ∇1s − ∇0s = α s for some α ∈ A1(M), and hence ω1−ω0 = dα, an exact 2-form. Thus the class [ω] depends only on the line bundle L−→M . Exercise 5.8. Suppose L−→M and L′−→M are equivalent line bundles and that τ : L → L′ is an invertible bundle map. Show that τ∗ ∈ Γ(Hom(L,L′)) intertwines the canonical isomorphisms Γ(EndL) ' A and Γ(EndL′) ' A, and deduce that the connections ∇ and τ∗ ◦ ∇ ◦ τ−1∗ have the same curvature. Conclude that the class [ω] depends only on the equivalence class [L] of the line bundle. 5.5 From de Rham cohomology to Cech cohomology We have now associated to any line bundle L−→M , by very different procedures, two second-degree cohomology classes, namely the Cˇech class obtained directly from its transition 5The local 1-forms αj may be globalized by pulling back to the frame bundle P η−→M ; recall that the local system sj may be regarded as a local section of the frame bundle. It is possible to construct a global 1-form α˜ ∈ A1(P,EndV ) which incorporates each η∗αj , and different connections given rise to different α˜: this is called the connection 1-form for ∇, for which one may consult [9, 39, 52] or any standard text on differential geometry. 48 functions and the de Rham class of the curvature of an arbitrary connection. We are led to suspect that these two cohomologies are related in some underlying fashion. This is indeed the case: one way of proving the de Rham theorem, which says that the cohomology of real-valued differential forms on the manifold M depends only on the topology of M (i.e., not on its differential structure) is to show that the de Rham cohomology is isomorphic to Cˇech cohomology with constant real coefficients.6 For the full proof, we refer to [12] or [17, Appendix E]. We need only the second-degree case of this isomorphism; however, its proof illustrates the general method. Proposition 5.5. Let M be a compact manifold. Then H2dR(M) ' Hˇ2(M,R) by a canonical isomorphism. Proof. Select a finite good covering7 U = {Uj} for M . We must show that H2dR(M) ' H2(U,R). We first define maps between Z2dR(M) and Z2(U,R), which have some ambiguities that can be removed by passing to cohomology, thereby yielding well-defined R-linear maps between H2dR(M) and H 2(U,R). Start with a closed 2-form ω in A2(M); denote by ω := {ωj} the set of restrictions of ω to each set Uj of the good covering; this is an element of C 0(U,A2), the set of Cˇech 0-cochains with 2-form coefficients. Since dω = 0 and each Uj is contractible, the Poincare´ lemma shows that there is αj ∈ A1(Uj) with dαj = ωj for each j; write α := {αj} ∈ C0(U,A1). Now on each nonvoid overlap Ui ∩ Uj we have d(αi − αj) = ωi − ωj = 0, so that αi − αj = dfij with f := {fij} ∈ C1(U,A0). On each nonvoid Ui ∩ Uj ∩ Uk we have d(fij − fik + fjk) = 0 by cancellation, so that aijk := fij−fik+fjk is a constant real-valued function (since Ui∩Uj∩Uk is connected); since aijk − aijl + aikl − ajkl = 0 on Ui ∩Uj ∩Uk ∩Ul by cancellation, we have δa = 0, i.e., a ∈ Z2(U,R). There are some ambiguities in the choices of αj and fij, so this process does not give a well-defined map ω 7→ a. Firstly, we could replace each αj by αj + dgj, where g := {gj} ∈ C0(U,A0); then fij becomes fij + gi − gj, which leaves aijk unchanged. Secondly, we could replace each fij by fij + cij, where the cij are constant functions, i.e., c := {cij} ∈ C1(U,R); this changes a to a + δc, and [a] ∈ H2(U,R) is left unchanged. Therefore ω 7→ [a] is well-defined. Thirdly, we could replace ω by ω+ dβ, adding an exact form; then αj becomes αj + βj, and fij = (αi + βi) − (αj + βj) is unchanged because βi and βj agree on Ui ∩ Uj. Therefore, [a] depends only on the de Rham class of ω, so [ω] 7→ [a] is a well-defined R-linear map from H2dR(M) to H 2(U,R). To go the other way, start from a ∈ Z2(U,R), and take a smooth partition of unity {ψj} subordinate to the covering U. As in the proof of Proposition 1.6, define f ∈ C1(U,A0) by fij := ∑ r aijrψr; then fij − fik + fjk = aijk just as in (1.12), since δa = 0; moreover, dfij − dfik + dfjk = 0 since the aijk are constant. Define α ∈ C0(U,A1) by αj := ∑ k ψk dfjk; now αi − αj = ∑ k ψk dfij = dfij on Ui ∩Uj, and so dαi − dαj = 0 there, which says that the 6Another way is to establish an isomorphism between de Rham cohomology and singular cohomology. This is thoroughly dealt with in [23]. 7This is the only point at which compactness is invoked; and the compactness assumption may be removed whenever the existence of a finite good covering can be established independently. 49 local 2-forms ωj := dαj patch together to give a global 2-form ω on M . Since dωj = 0, ω is closed, i.e., ω ∈ Z2dR(M). The correspondence a 7→ ω depends on the partition of unity {ψj}, but it is clear that a 7→ f 7→ α 7→ ω retraces the earlier path ω 7→ α 7→ f 7→ a; thus we have shown that the linear map [ω] 7→ [a] is surjective. Moreover, if [a] = 0, then a = δc with c ∈ C1(U,R), so that fij := ∑ r(cij−cir+cjr)ψr = cij−hi+hj where hj := ∑ r cjrψr and thus dfij = dhj−dhi; but then αj = ∑ k ψk(dhk − dhj) = βj − dhj, where βj is the restriction to Uj of the 1-form β := ∑ k ψk dhk ∈ A1(M); thus ω = dβ is exact, i.e., [ω] = 0. Hence [ω] 7→ [a] is injective, and is therefore an isomorphism (or real vector spaces) between H2dR(M) and H 2(U,R). The foregoing proof can be cast in a more algebraic framework, as follows. The abelian groups (actually, vector spaces) Cpq := Cp(U,Aq) of q-form-valued Cˇech p-cochains are related by two coboundary operators, δ : Cpq → Cp+1,q and d : Cpq → Cp,q+1, satisfying δd− dδ = 0; if we introduce ∂ := (−1)pd on Cpq, we get δ∂+ ∂δ = 0, so we obtain a “double complex”. By introducing D := δ + ∂ : Cpq → Cp+1,q ⊕ Cp,q+1, and Em := ⊕p+q=mCpq, we obtain a new cochain complex (E•, D), sometimes called the Cˇech–de Rham complex. (For instance, a 1-cochain for this complex is of the form α ⊕ f where α and f are as in the proof of Proposition 5.5, and D(α ⊕ f) = dα ⊕ (δα − df) ⊕ δf , which evaluates to ω ⊕ 0 ⊕ a. A general theorem [12, 17] asserts that the k-th cohomology groups for the complexes (A•(M), d), (E•, D), and (C•(U,M), δ) are isomorphic, for any k ∈ N. Definition 5.7. The standard inclusion ι : Z→ R induces an injection of Cˇech cohomology groups ι∗ : Hˇ2(M,Z)→ Hˇ2(M,R). We regard Hˇ2(M,Z) as a subset of Hˇ2(M,R) by identi- fying it with its image under ι∗. We say that a de Rham cohomology class [ω] ∈ H2dR(M) is integral if the corresponding Cˇech class [a] lies in Hˇ2(M,Z). Theorem 5.6. Let L−→M be a Hermitian line bundle over a compact manifold, and let ∇ be a connection on it, compatible with the metric, with curvature ω. Then [(2pii)−1ω] is an integral de Rham class, corresponding to the class [L] of the line bundle in Hˇ2(M,Z). Proof. Let U = {Uj} be a covering of M by chart domains, and let {sj} be a local system of nonvanishing sections for the line bundle L−→M . Then sj ∈ Γ(Uj, L) for each j. By the locality property of ∇, we have ∇sj ∈ A1(Uj, L) = Γ(Uj, L)⊗C∞(Uj)A1(Uj), so ∇sj = sj⊗αj for some αj ∈ A1(Uj) (because sj generates Γ(Uj, L) as a C∞(Uj)-module). We may take sj to be normalized with respect to the metric, i.e., (sj | sj) = 1 on Uj. Since ∇ is compatible, we find that αj is a purely imaginary 1-form, because α¯j + αj = (∇sj | sj) + (sj | ∇sj) = 0. By (5.8), ∇2sj = (sj ⊗ αj) ∧ αj + s ∧ dαj = s ∧ dαj, so that ω = dαj on Uj. Let us write βj := (2pii) −1αj; then the βj are real-valued 1-forms, satisfying dβj = (2pii)−1ω. Thus [(2pii)−1ω] ∈ H2dR(M) is a real de Rham class. The local sections sj are related by si = gijsj = sjgij on Ui ∩ Uj, where the gij are the U(1)-valued transition functions of the line bundle (see Definition 1.14). Then sj ⊗ gijαi = si ⊗ αi = ∇si = ∇(sjgij) = (sj ⊗ αj)gij + sj ⊗ dgij 50 on Ui ∩ Uj; since sj does not vanish there, we get αi = g −1 ij αjgij + g −1 ij dgij = αj + g −1 ij dgij on Ui ∩ Uj. This gives βi − βj = (2pii)−1g−1ij dgij = (2pii)−1 d(log gij) for a suitable branch of the logarithm, and fij := (2pii) −1 log gij is a real-valued function on Ui ∩ Uj satisfying βi − βj = dfij. Now aijk := fij − fik + fjk gives the Cˇech 2-cocycle a ∈ C2(U,R) such that [a] corresponds to [(2pii)−1ω] under the isomorphism of Proposition 5.5. On the other hand, the proof of Proposition 1.6 constructs this very same Cˇech class [a] as the element of Hˇ2(M,Z) which corresponds to [L] under the Bockstein isomorphism between Hˇ1(M,U(1)) and Hˇ2(M,Z). Hence, [(2pii)−1ω] is integral and corresponds to [L]. In particular, we see that a closed 2-form on M is the curvature of some compatible connection only if it equals 2pii times an integral 2-form.8 5.6 Line bundles over CPm Definition 5.8. Any element of CPm is a line through the origin in Cm+1, and two such lines intersect only at the origin. The tautological line bundle L−→CPm is given by taking the disjoint union of these lines; the fibre at any element of CPm is the very same line. More explicitly, we can take L to be L := { (`, v) ∈ CPm × Cm+1 : v ∈ ` }, (5.12) and define pi(`, v) := `. Here L−→CPm is manifestly a subbundle of a trivial vector bundle. Indeed, let `⊥ be the subspace of Cm+1 orthogonal to ` (with respect to the usual inner product on Cm+1), and if E := { (`, u) ∈ CPm × Cm+1 : u ∈ `⊥ }, then the Whitney sum L⊕ E = CPm × Cm+1 is trivial. Exercise 5.9. If η : Cm+1\{0} → CPm is the quotient map, Cm+1\{0} η−→CPm is a principal C×-bundle. Show that the tautological line bundle is associated to this principal bundle via the representation ρ of C× on C given by the multiplication ρ(λ)µ := λµ. Definition 5.9. The dual of the tautological line bundle on CPm is the hyperplane bundle H −→CPm, where H` = `∗ (the one-dimensional space of linear functionals on `). 8This observation is the launching point for the theory of geometric quantization, which seeks to represent certain functions on a symplectic manifold (M,Ω), i.e., a manifold equipped with a closed nondegenerate 2-form, by operators on a Hilbert space. The elements of this Hilbert space come from sections of a certain Hermitian line bundle over M , equipped with a compatible connection whose curvature is (2pii~)−1Ω (where ~ is a positive constant, identified with Planck’s constant). Theorem 5.6 shows that such a line bundle exists iff [~−1Ω] is integral; this is a discreteness condition on the original symplectic form Ω, hence the use of the word “quantization”. 51 The tensor product bundles L⊗k−→CPm and H⊗l−→CPm, for k, l ∈ N, give more examples of line bundles over CPm. (We take L⊗0 = H⊗0 := CPm × C by convention.) We aim to show that these are distinct, and that any line bundle over CPm is equivalent to one on this list. Since [L]−1 = [H] in the group of line bundle classes, and since H2dR(CP m) = R, we have Hˇ2(CPm,Z) ' Z, so it suffices to verify that [H] corresponds to a generator of the infinite cyclic group Hˇ2(CPm,Z). Definition 5.10. A complex vector bundle E−→M on a complex manifold M is a holo- morphic vector bundle if its transition functions gij : Ui∩Uj → GL(m,C) are holomorphic. The space of holomorphic sections of this line bundle is denoted by O(M,E) or simply by O(E). Exercise 5.10. Compute the transition functions for the tautological and hyperplane bundles over CPm and hence show that these are holomorphic line bundles. Consider first the trivial bundle CPm×C−→CPm; a holomorphic section of this bundle is of the form s(x) = (x, f(x)), where f : CPm → C is a holomorphic function. If (U, φ) is a chart with φ(U) = Cm, then f ◦ φ−1 : Cm → C is an entire holomorphic function on Cm, which is bounded since CPm is compact and f is continuous; thus, by Liouville’s theorem, f is constant. More generally, any holomorphic function on a compact complex manifold is constant. It turns out that the tautological line bundle has no global holomorphic sections, other than the zero section. The hyperplane bundle H −→CPm, by contrast, has nontrivial global holomorphic sections. For any f ∈ (Cm+1)∗, define sf ∈ Γ(CPm, H) by sf (`) := f ∣∣ ` . Con- versely, any holomorphic section of H −→CPm is of the form s(`) := g∣∣ ` , where g : Cm+1 → C is a holomorphic function whose restriction to each one-dimensional subspace is linear, i.e., g is homogeneous of degree one. Since only first-degree terms can then occur in the Taylor series of g, the function g is itself linear, i.e., g ∈ (Cm+1)∗ and s = sg. Thus, f 7→ sf is a linear bijection between (Cm+1)∗ and O(H). In particular, O(H) is finite-dimensional, with dimO(H) = m + 1. A basis is given by σ0, . . . , σm, where σj = szj and z j ∈ (Cm+1)∗ be the j-th coordinate function, for j = 0, 1, . . . ,m. The standard inner product 〈〈· | ·〉〉 on Cm+1 gives metrics on the line bundles L and H. We may write (u | v)` := 〈〈u | v〉〉 for u, v ∈ L`, i.e., u, v ∈ ` ⊂ Cm+1. Now if φ, ψ ∈ `∗ are given by φ(v) = 〈〈u | v〉〉, ψ(v) = 〈〈w | v〉〉 for v ∈ `, then (φ | ψ)`∗ := 〈〈u | v〉〉〈〈w | v〉〉〈〈v | v〉〉 = 〈〈w | v〉〉〈〈v | u〉〉 〈〈v | v〉〉 , which is independent of any nonzero v ∈ `. When ` ∈ Uj, we have the equality 〈〈v | v〉〉 = z0z¯0 + · · ·+ zmz¯m = zj z¯j ( 1 + ∑ k 6=j wkj w¯ k j ) = |zj(v)|2Qj(`), 52 and in particular (σj | σj) = Q−1j = ( 1 + ∑ k 6=j wkj w¯ k j )−1 on Uj. (5.13) By continuity, (σj | σj) = 0 and thus σj vanishes on the complement of Uj. 5.7 Connections on the hyperplane bundle Definition 5.11. When M is a complex manifold and ∇ is a connection on a holomorphic vector bundle E−→M , we say that ∇ is compatible with the holomorphic structure if ∇s ∈ A1,0(E) whenever s ∈ O(E). If the vector bundle is also Hermitian, we say that ∇ is a canonical connection if it is compatible with both the metric and the holomorphic structure. Proposition 5.7. On the hyperplane bundle H −→CPm, there is a unique canonical con- nection. Proof. Let {Uj, σj} be the local system of holomorphic sections of the previous Section. Since σj is nonvanishing on Uj (e.g., on account of (5.13)), the argument of the proof of Theorem 5.6 shows that any connection ∇ on the hyperplane bundle is given on each Uj by ∇σj = σj ⊗ αj, for some αj ∈ A1(Uj). Thus ∇ is compatible with the holomorphic structure iff each αj lies in A1,0(Uj). We may write αj = ajk dw k j with each ajk ∈ C∞(Uj), so that ∇ is also compatible with the metric iff Q−1j ( ajk dw k j + a¯jk dw¯ k j ) = (σj | ∇σj) + (∇σj | σj) = d(σj | σj) = d(Q−1j ) = −Q−2j d (∑ k 6=j wkj w¯ k j ) iff ajk = −Q−1j w¯kj for all k 6= j. Thus ∇ satisfies ∇σj = −σj ⊗Q−1j ∑ k 6=j w¯kj dw k j , so this calculation establishes existence and uniqueness of the canonical connection. Proposition 5.8. The curvature of the canonical connection on the hyperplane bundle is −iΦ, where Φ is the Ka¨hler form on CPm. Proof. The curvature ω satisfies ω = dαj on Uj, so ω = dαj = Q −1 j ∑ k 6=j dwkj ∧ dw¯kj +Q−2j dQj ∧ ∑ k 6=j w¯kj dw k j = Q−2j ( Qj ∑ k 6=j dwkj ∧ dw¯kj − ∑ r,s6=j w¯rjw s j dw r j ∧ dw¯sj ) = −iΦ, by comparison with (2.7). 53 The element of H2dR(CP m) which corresponds to the equivalence class of the hyperplane bundle is therefore [−(2pi)−1Φ]. 5.8 Characteristic classes We have seen that the curvature of a Hermitian line bundles over M is a closed 2-form ω on M and [(2pii)−1ω] is an integral cohomology class. For vector bundles of higher rank, it is possible to obtain integral cohomology classes of even degree from the matrix-valued curvature ω by taking suitable traces of its exterior powers. These classes are topological invariants of the manifold M . Definition 5.12. Let q˜ : (Cr×r)k → C be a symmetric k-linear map on the vector space of complex r × r matrices. One says that q˜ is invariant (under the adjoint representation Ad(g)A := gAg−1 of GL(r,C)) if q˜(gA1g −1, . . . , gAkg−1) = q˜(A1, . . . , Ak) (5.14) for all g ∈ GL(r,C), A ∈ Cr×r. The function q : Cr×r → C defined by q(A) := q˜(A,A, . . . , A) is called a homogeneous polynomial of degree k; when q˜ is invariant, we call it an invariant polynomial on Cr×r. With the “polarization formula” q˜(A1, . . . , Ak) := (k!) −1∑ |J |=s(−1)k−sq(Aj1 + · · ·+Ajs), the map q˜ may be recovered from q. More generally, an invariant polynomial on Cr×r is a finite sum of homogeneous invariant polynomials of various degrees. (A constant function is an invariant polynomial of degree zero.) For instance, in the expansion det(t−A) = ∑rk=0(−1)kqk(A)tr−k, each qk is an invariant homogeneous polynomial; here qr(A) = detA, q1(A) = trA, and q2(A) = ∑ ij q˜(ωt, . . . , β, . . . , dωt, . . . , ωt) = − ∑ ij q˜(ωt, . . . , β, . . . , [[αt, ωt]], . . . , ωt) = ∑ j=1k q˜(ωt, . . . , [[αt, β]], . . . , ωt) + q ′(ωt; dβ) = q′(ωt; dβ + [[αt, β]]) = q′(ωt; (d/dt)ωt), (with dωt in the i-th position and β in the j-th position in the double summations). Here we have used the Bianchi identity dωt + [[αt, ωt]] = 0 and the invariance property (5.16) applied to the form q˜(ωt, . . . , dωt, . . . , β, . . . , ωt) and also to q˜(ωt, . . . , β, . . . , dωt, . . . , ωt). If E ′−→M is another vector bundle equivalent to E−→M , and it τ : E → E ′ is an invertible bundle map, then by Exercises 5.4 and 5.6, the recipes ∇′ := τ∗ ◦ ∇ ◦ τ−1∗ and ω′ := τ∗◦ω◦τ−1∗ match connections and curvatures on both vector bundles; by the invariance property (5.14) of the polynomial q, we have q(ω′) = q(ω) in A2k(M). Thus the cohomology class [q(ω)] depends only on the equivalence class of the vector bundle E−→M , and we may denote it by q([E]), or more simply by q(E) ∈ H2kdR(M)⊗R C. If E−→M is a Hermitian vector bundle, we consider only bundle maps τ : E → E ′ which preserve the fibre metrics. Thus we may use polynomials q which are only invariant under the unitary group U(r) rather than GL(r,C); in (5.16) we may only use forms β with either iβ or β itself real-valued. If a connection ∇ on E−→M is compatible with the metric, then −iω is real-valued, and the cohomology class [q(−iω)] belongs to H2kdR(M). 5.9 Chern classes and the Chern character Definition 5.13. Let E−→M be a Hermitian vector bundle, together with a compatible connection ∇ whose curvature is ω ∈ A2(M,EndE). Define a U(r)-invariant polynomial on Cr×r by c(A) := det ( 1r − 1 2pii A ) , which may be written as a sum of homogeneous polynomials c(A) = 1 + c1(A) + c2(A) + · · ·+ cr(A) (5.20) where c1(A) = (i/2pi) trA, cr(A) = (i/2pi) r detA; and if {iλ1, . . . , iλr} are the eigenvalues of A, then ck(A) = (−1/2pi)k ∑ λj1λj2 . . . λjk ; the invariant homogeneous polynomials ck are real-valued on the Lie algebra u(r) = iRr×r of the unitary group U(r). 57 The class ck(E) ∈ H2kdR(M) is called the k-th Chern class, and c(E) ∈ HevendR (M) is called the total Chern class of the vector bundle. Exercise 5.16. If E∗−→M is the dual vector bundle to E−→M , and if ∇∗ is the dual connection to ∇ (see Exercise 5.3), show that ∇∗ has curvature −ωt ∈ A2(M,End(E∗)), and conclude that ck(E ∗) = (−1)kck(E). Exercise 5.17. If φ : N →M is a smooth map, and if ∇ is a connection on a Hermitian vector bundle E−→M with curvature ω, find a connection ∇′ on the pullback bundle φ∗E−→N whose curvature is φ∗ω ∈ A2(N,End(φ∗E)). Conclude that ck(φ∗E) = φ∗ck(E) ∈ H2kdR(N). For Hermitian line bundles (r = 1), the total Chern class is just c(L) := 1 + c1(L). From Theorem 5.6, we know that c1(L) is an integral cohomology class, so c(L) is also integral. This integrality property holds for Chern classes of any Hermitian vector bundle. Indeed, if E−→M and E ′−→M are two Hermitian vector bundles with compatible connections ∇ and ∇′ and respective curvatures ω and ω′, then ∇⊕∇′ is a connection on E ⊕ E ′−→M , with curvature ω ⊕ ω′ ∈ A2(M,End(E ⊕ E ′)). Clearly c(ω ⊕ ω′) = det ( 1r − ω 2pii ) ∧ det ( 1r′ − ω ′ 2pii ) = c(ω) ∧ c(ω′), (5.21) on account of (5.15). Now the wedge product of closed forms induces a product of cohomology classes, since (η+ dζ)∧ (η′+ dζ ′) = η ∧ η′+ d(η ∧ ζ ′+ ζ ∧ (η′+ dζ)) if dη = 0 and dη′ = 0, so [η ∧ η′] is not affected by adding an exact form to either η or η′; in other words, the recipe [η] [η′] := [η ∧ η′] is a well-defined product10 making H•dR(M) into a ring. The even-degree classes form a commutative subring HevendR (M). On passing to cohomology, (5.20) becomes c(E ⊕ E ′) = c(E) c(E ′) in HevendR (M). (5.22) The integral 2-forms generate an integral subring Heven(M,Z). By (5.21), c(E) is integral, i.e., lies in Heven(M,Z), whenever E−→M is a Whitney sum of Hermitian line bundles. An important splitting principle (see [12] for a proof) asserts that some pullback bundle φ∗E−→N can be split (into a sum of line bundles) in such a way that c(E) 7→ φ∗(c(E)) = c(φ∗E) is injective, and therefore c(E) is integral since c(φ∗E) is. Definition 5.14. The Chern character ch(E) of the Hermitian vector bundle E−→M of rank r is given by the invariant power series ch(A) := tr ( exp((2pi)−1i A) ) = 1 + ch1(A) + ch2(A) + · · · , (5.23) or equivalently by the invariant polynomial obtained by discarding the terms chk with 2k > dimM , since ω∧k ∈ A2k(M,EndE) and hence chk(ω) = 0 for 2k > dimM . This polynomial is found explicitly by writing the eigenvalues of A as {2piiµ1, . . . , 2piiµr}, expanding the function e−µ1 + · · ·+ e−µr in a Taylor series, and discarding high-degree terms. 10This product is usually called the cup product in de Rham cohomology. 58 Exercise 5.18. Check that ch1(E) = c1(E), ch2(E) = 1 2 (c1(E) 2 − 2c2(E)), and ch3(E) = 1 6 (c1(E) 3 − 3c1(E)c2(E) + c3(E)). In general, the chk(E) are polynomial combinations of the Chern classes cj(E) with ratio- nal coefficients, on account of the 1/k! terms in the Taylor series expansion; thus they might not be integral classes, but rational classes, i.e., elements of Heven(M,Q) = Heven(M,Z)⊗ZQ. Proposition 5.12. The Chern character satisfies the homomorphism properties: ch(E ⊕ E ′) = ch(E) + ch(E ′), ch(E ⊗ E ′) = ch(E) ch(E ′). Proof. The invariant power series ch(A) of (5.22) satisfies ch(A⊕A′) = ch(A) + ch(A′) and ch(A⊗ 1 + 1⊗ A′) = ch(A) ch(A′); where ⊕ and ⊗ denote the usual direct sum and tensor product of matrices. These identities are simple to check for diagonal matrices, therefore hold for diagonalizable matrices by invariance, and hence hold generally, since the set of diagonalizable r × r matrices is dense in Cr×r. Given connections ∇, ∇′ with curvatures ω, ω′ on the respective vector bundles E−→M and E ′−→M , the curvatures of ∇⊕∇′ and ∇⊗∇′ are ω ⊕ ω′ ∈ A2(M,End(E ⊕E ′)) and ω ⊗ 1 + 1⊗ ω′ ∈ A2(M,End(E ⊗ E ′)). On replacing A, A′ by ω, ω′ in the foregoing matrix identities, bearing in mind (5.15), and on passing to cohomology, one obtains the desired formulae (5.23) for ch(E ⊕ E ′) and ch(E ⊗ E ′). Exercise 5.19. Explain how the tensor product connection ∇⊗∇′ is defined, and check the given formula for its curvature. A trivial bundle Er = M × Cr−→M has a “flat” connection (i.e., a connection with zero curvature) namely d, and thus ch(Er) = 1. Thus, if two vector bundles E−→M and F −→M are stably equivalent, then ch(E)−ch(F ) is an integral multiple of 1 in Heven(M,Q). If one defines the “reduced K-theory” K˜0(M) of M as the quotient of K0(M) by Z (on identifying r ∈ N with [[Er]] ∈ K0(M)), the tensor product of vector bundles makes K˜0(M) a commutative ring, and thus the Chern character defines a ring homomorphism ch: K˜0(M)→ Heven(M,Q). For Hermitian line bundles L−→M and L′−→M , Proposition 5.12 yields the identity c1(L⊗ L′) = c1(L) + c1(L′), (5.24) by extracting the component in H2dR(M) from the formula ch(L⊗ L′) = ch(L) ch(L), using ch1(L) = c1(L). Thus c1 determines a homomorphism from the group of line bundle classes to the additive group H2(M,Z) of integral de Rham classes. Since c1(L) = [(i/2pi)ω] when ω is the curvature of a compatible connection on L−→M , and c1(L∗) = −c1(L), we conclude that [L] 7→ c1(L∗) is the isomorphism described in Theorem 5.6. To show that [L] 6= [L′], it is enough to show that c1(L) 6= c1(L′). Moreover, if dimM = 2m, then c1(L) m = [(i/2pi)nω∧m] is an integral 2m-form, and so ∫ M (i/2pi)nω∧m ∈ Z, since the identification of H2mdR (M) with R is precisely the map [ν] 7→ ∫ M ν. 59 5.10 Classification of line bundles over CPm Proposition 5.13. The group of classes of line bundles over CPm is an infinite cyclic group, generated by the class [H] of the hyperplane bundle. Proof. Since we already know that Hˇ2(CPm,Z) ≡ Z, we need only check that [H] is a generator. Equivalently, we must check that c1(H) = [(2pi) −1Φ] is a generator for H2(M,Z). For this, it is enough to show that ∫ CPm(2pi) −mΦ∧m = 1. Since the complement of the chart domain U0 is a lower-dimensional submanifold (diffeo- morphic to CPm−1), we need only show that the integral over U0 equals 1. We may therefore use the formula (2.7) (with j = 0) for the Ka¨hler form Φ. Then∫ U0 Φ∧m = ∫ Cm ( iQ−10 m∑ k=1 dwk0 ∧ dw¯k0 − iQ−20 m∑ k,l=1 w¯k0w l 0 dw k 0 ∧ dw¯l0 )∧m . (5.25) At the point (r, 0, . . . , 0) ∈ Cm, Q0 simplifies to 1 + r2, and the integrand on the right hand side of (5.24) becomes( i(1 + r2)−2 dw10 ∧ dw¯10 + i(1 + r2)−1 m∑ k=2 dwk0 ∧ dw¯k0 )∧m = imm! (1 + r2)m+1 m∧ k=1 dwk0 ∧ dw¯k0 = 2mm! (1 + r2)m+1 m∧ k=1 dxk ∧ dyk = 2mm! (1 + r2)m+1 λ, (5.26) where λ is Lebesgue measure on Cm, and wk0 = xk + iyk give Cartesian coordinates on Cm. Interpreting r as a polar coordinate on Cm, one can write λ = r2m−1 dr d2m−1θ, with θ ∈ S2m−1 being the angular part. The right hand side of (5.25) is invariant under the unitary group U(m), as is the Ka¨hler form, so it represents the integrand at all points, not just at (r, 0, . . . , 0). The volume of the sphere S2m−1 is Ω2m = 2 pim/(m− 1)! (see [1, 8, 28], for instance), so the desired integral is∫ CPm (2pi)−mΦ∧m = m! pim ∫ Cm r2m−1 (1 + r2)m+1 dr d2m−1θ = m! pim Ω2m ∫ ∞ 0 r2m−1 (1 + r2)m+1 dr = ∫ ∞ 0 2mr2m−1 (1 + r2)m+1 dr = ∫ ∞ 0 mtm−1 (1 + t)m+1 dt = ∫ 1 0 mum−1 du = 1, on substituting t = r2, u = t/(1 + t). Corollary 5.14. The hyperplane bundle is not trivial, since c1(H) 6= 0. 60 This completes the classification of Hermitian line bundles over CPm, since any line bundle L′−→CPm satisfies c1(L′) = k c1(H) for some k ∈ Z; thus L′ ∼ H⊗k if k > 0, L′ ∼ L⊗(−k) if k < 0, and L′ is trivial iff k = 0. Furthermore, k is precisely the integral over CPm of (−2pii)−mω∧mL′ where ωL′ is the curvature of any compatible connection on L′−→CPm. 5.11 The Levi-Civita connection on the tangent bundle Definition 5.15. LetM be a Riemannian manifold, and let∇ be a connection on the tangent bundle TM −→M . The fundamental 1-form θ is the unique element of A1(M,TM) sat- isfying ιXθ = X for all X ∈ X(M) = Γ(TM). The torsion of ∇ is T := ∇θ ∈ A2(M,TM). Exercise 5.20. Show that the contraction map ιX : A 1(M,E) → Γ(E) of Definition 5.4 ex- tends to an A-linear map ιX : A k(M,E)→ Ak−1(E) such that ιX(ζ∧η) = (ιXζ)∧η+(−1)kζ∧ ιXη for ζ ∈ Ak(M,E), η ∈ A•(M), provided we define ιXs := 0 for s ∈ Γ(E) = A0(M,E). We can then define T (X, Y ) := ιY (ιXT ) ∈ X(M). Lemma 5.15. If X, Y ∈ X(M), then T (X, Y ) = ∇XY −∇YX − [X, Y ]. Proof. It is not hard to show that∇Xζ = ∇(ιXζ)+ιX(∇ζ) and∇X(ιY ζ) = ιY (∇Xζ)+ι[X,Y ]ζ for ζ ∈ A1(M,TM). For the particular case ζ = θ, these identities give T (X, Y ) = ιY (ιX∇θ) = ιY (∇Xθ −∇(ιXθ)) = ιY (∇Xθ −∇X) = ∇X(ιY θ)− ι[X,Y ]θ −∇YX = ∇XY − [X, Y ]−∇YX, (5.27) as claimed. Exercise 5.21. Verify the aforementioned formulae for ∇Xζ and ∇X(ιY ζ) by applying (5.8) and (5.9) (and their analogues for ιX) in the case ζ = Z ⊗ α with Z ∈ X(M), α ∈ A1(M). Proposition 5.16. If M is a Riemannian manifold, there is a unique connection ∇ on the tangent bundle TM −→M , which is compatible with the Euclidean metric on TM and is torsion-free. Proof. Compatibility with the metric demands that (∇X | Y ) + (X | ∇Y ) = d(X | Y ), as in (5.7), where (X | Y ) = g(X, Y ) denotes the bilinear pairing on X(M) = Γ(TM) determined by the Euclidean metric g. By Lemma 5.15, ∇ is torsion-free iff ∇XY −∇YX = [X, Y ] for X, Y ∈ X(M). Thus Z(X |Y ) = ιZ(∇X |Y )+ ιZ(X |∇Y ) = (∇ZX |Y )+(X |∇ZY ). A short calculation then shows that 2(∇ZX | Y ) = X(Y | Z)− Y (Z |X) + Z(X | Y ) + (X | [Y, Z]) + (Y | [Z,X])− (Z | [X, Y ]), (5.28) which establishes the uniqueness of ∇. On the other hand, it is easy to show that the right hand side is A-linear in Y and Z, hence is of the form ιZ(DX | Y ), where D : X(M) → A1(M,TM). By replacing X by hX (with h ∈ A) on the right hand side of (5.27) and then simplifying, one verifies the Leibniz rule for D, which proves the existence of the desired connection. 61 Definition 5.16. The unique metric-compatible torsion-free connection on the tangent bun- dle of a Riemannian manifold M is called its Levi-Civita connection. Its curvature, in A2(M,TM), is usually denoted by R, and is also called the Riemannian curvature ten- sor11 of M . Exercise 5.22. Verify the following property of the Riemannian curvature tensor: R(X, Y )Z +R(Y, Z)X +R(Z,X)Y = 0 for X, Y, Z ∈ X(M), (5.29) using only Lemma 5.4, Lemma 5.15, and the Jacobi identity. Exercise 5.23. Verify the following symmetry properties of the Riemannian curvature tensor: (W |R(X, Y )Z) = −(R(X, Y )W | Z) = (X |R(W,Z)Y ), for W,X, Y, Z ∈ X(M). 6 Clifford algebras A Clifford algebra is an associative algebra which is generated by starting with a real vector space and defining a product of vectors in such a way that the square of any vector is a scalar. This can be done consistently if the generating vector space is Euclidean, i.e., if it carries a symmetric bilinear form. For the zero bilinear form, the corresponding algebra is just the exterior algebra on the given vector space; otherwise, it has the same underlying vector space as the exterior algebra, but with a modified product operation. The Clifford algebra has interesting matrix representations, whose representation spaces are called “Clifford modules”. All of these spaces are graded into an “even” part and an “odd” part, so we begin with a general discussion of vector spaces and algebras which are Z2-graded. 6.1 Superspaces and superalgebras Definition 6.1. A superspace is just a vector space with a given Z2-grading: V = V +⊕V −; here V + is called the even subspace and V − is called the odd subspace.1 A superalgebra is an algebra whose underlying vector space is a superspace: A = A+ ⊕ A−, where the product respects the grading,2 i.e., A+ · A+ ⊆ A+, A− · A− ⊆ A+, A+ · A− ⊆ A−, and A− · A+ ⊆ A−. 11Since A2(M,TM) = X(M) ⊗A A2(M), the map (X,Y, α) 7→ α(R(X,Y )) is a tensor of bidegree (2, 1) on M . 1The unfortunate prefix “super”, which is simply a synonym for “Z2-graded”, was introduced by Feliks Berezin [7] about 30 years ago, and has since become fashionable. Berezin wished to extend the calculus of Gaussian integrals by regarding an exterior algebra as a “space of functions of anticommuting variables”. This crazy idea works astonishingly well. 2If we regard the exponents + and − as the elements of the additive group Z2, these four inclusions may be collected as the formula Ai · Aj ⊆ Ai+j . In this way we can define a G-graded algebra for any abelian group G, although only the cases G = Z2 and G = Z are commonly used. 62 Definition 6.2. The exterior algebra Λ•V of a vector space V is a Z-graded algebra, whose subspace of degree k is Λk(V ) for k = 0, 1, . . . , dimV (and for k < 0 or k > dimV , one sets Λk(V ) := {0}), since Λk(V ) ∧ Λl(V ) ⊆ Λk+l(V ) for k, l ∈ Z. But Λ•V is also a superalgebra, since we may define Λ+(V ) := ⊕ k even Λk(V ), Λ−(V ) := ⊕ k odd Λk(V ). Indeed, any Z-graded algebra becomes a superalgebra, by collecting the subspaces of even degree and of odd degree in this manner. Definition 6.3. If V = V + ⊕ V − is a superspace, then EndV is a superalgebra, with End+ V := End(V +)⊕ End(V −), End− V := Hom(V +, V −)⊕ Hom(V −, V +), Exercise 6.1. Guess the definition of a supermodule. Any superspace V is a supermodule for the superalgebra EndV . Definition 6.4. We say an element a of a superalgebra A is homogeneous if either a ∈ A+ or a ∈ A−; its parity #a is defined as #a := 0 if a ∈ A+, #a := 1 if a ∈ A−. Analogously, in a Z-graded algebra, we define the degree of a homogeneous element as #a := k if a ∈ Ak. There is an important, if somewhat informal, sign rule in superalgebra, which says that in any calculation in which the order of multiplication of homogeneous elements is reversed (i.e., ab is changed to ba), a sign factor of (−1)#a#b must be inserted. Thus, for example, we say that a superalgebra is “supercommutative” if ab = (−1)#a#bba for homogeneous elements a, b ∈ A; in other words, even elements commute with both even and odd elements, but two odd elements anticommute. Notice that the exterior algebra Λ•V is supercommutative. Definition 6.5. The superbracket in a superalgebra is the bilinear operation A×A→ A defined, for a, b homogeneous, by [[a, b]] := ab− (−1)#a#bba. A superalgebra is supercommutative iff all supercommutators [[a, b]] vanish, i.e., if the superbracket is trivial. The superbracket satisfies the properties: [[a, b]] + (−1)#a#b[[b, a]] = 0, [[a, [[b, c]]]] = [[[[a, b]], c]] + (−1)#a#b[[b, [[a, c]]]]. (6.1) A vector space with a bilinear operation (of any kind) which satisfies (6.1) is called a Lie superalgebra. Several bracket notations are in general use. Usually one writes [a, b] = ab − ba, and anticommutators ab+ ba are denoted {a, b} or sometimes [a, b]+; thus [[a, b]] = [a, b] if either a or b is even, and [[a, b]] = {a, b} if both a and b are odd. Warning : many superalgebraists use [a, b] to denote a supercommutator, even when a and b are odd. 63 Definition 6.6. A supertrace on a superalgebra A is a linear form τ which vanishes on supercommutators, i.e., τ([[a, b]]) = 0 for all a, b ∈ A. When A = EndE for E a superspace, we may write a ∈ A+ and b ∈ A− as a = ( a+ 0 0 a− ) , b = ( 0 b− b+ 0 ) , (6.2) and we define a supertrace on EndE by Str(a+ b) := Tr(a+)− Tr(a−). (6.3) Exercise 6.2. Write the supercommutator [[a, b]] for a, b homogeneous elements of EndE in the matrix notation (6.2) —there are four cases— and thus verify that (6.3) defines a supertrace. The tensor product U ⊗V of two superspaces U and V is a superspace, with the grading: (U ⊗ V )+ := (U+ ⊗ V +)⊕ (U− ⊗ V −), (U ⊗ V )− := (U+ ⊗ V −)⊕ (U− ⊗ V +). (6.4) The tensor product of two superalgebras A and B, with the standard product recipe (a ⊗ b)(a′ ⊗ b′) := aa′ ⊗ bb′ is not, however, a superalgebra, since this product does not respect the grading (6.4). However, one defines a graded tensor product, denoted A ⊗ B, whose underlying superspace is A⊗B, by using the multiplication rule: (a⊗ b)(a′ ⊗ b′) := (−1)#a′#baa′ ⊗ bb′ (6.5) (for a, a′, b, b′ homogeneous) in view of the passage of a′ to the left of b. Exercise 6.3. Verify that the product (6.5) is compatible with the grading (6.4) on the superspace A⊗B. Exercise 6.4. If U and V are (ungraded) vector spaces, show that Λ•(U ⊕ V ) ' Λ•U ⊗ Λ•V as superalgebras. Exercise 6.5. If U and V are superspaces, define an action of EndU ⊗ EndV on the super- space U ⊗ V in such a way that EndU ⊗ EndV ' End(U ⊗ V ) as superalgebras. Exercise 6.6. Let A be a supercommutative superalgebra, and let V be a superspace. Es- tablish the following identity for homogeneous elements of A ⊗ EndV : [[a⊗ T, b⊗ S]] = (−1)#T #bab⊗ [[T, S]]. Show that the linear map StrA : A ⊗ EndV → A given by StrA(a ⊗ T ) := a Str(T ) is an A-valued supertrace, in the sense that it vanishes on supercommutators. Definition 6.7. A superbundle over a smooth manifold M is a vector bundle E−→M which is a Whitney sum of two vector bundles: E = E+ ⊕ E−. The fibres Ex = E+x ⊕ E−x are superspaces. 64 The space of sections Γ(E) is a Z2-graded C∞(M)-module: Γ(E) = Γ(E+) ⊕ Γ(E−). The E-valued forms on M may also be graded by degree, and so, in accordance with (6.4), A•(M,E) carries the total Z2-grading: A+(M,E) := Aeven(M,E+)⊕Aodd(M,E−), A−(M,E) := Aeven(M,E−)⊕Aodd(M,E+). 6.2 Clifford algebras In this section we treat briefly the algebraic theory of Clifford algebras over Euclidean vector spaces. The presentation follows the treatment in [9], and the appendix to [54]. For a more detailed exposition of the algebraic theory, see [27] or [39]. All these rely on the fundamental paper of Atiyah, Bott and Shapiro [4]. Definition 6.8. Let V be a real vector space, with dimR V = n, and q a positive definite symmetric bilinear form on V ; we shall call the pair (V, q) a Euclidean vector space.3 The Clifford algebra C`(V ) = C`(V, q) is an associative algebra generated by the elements of V subject (only) to the relation v2 = −q(v, v) 1. More precisely, one defines C`(V, q) := T(V )/I(q), where T(V ) is the tensor algebra over V and I(q) is the ideal generated by { v⊗ v+ q(v, v) 1 : v ∈ E }. The canonical mapping of V into C`(V, q) is injective, so V may be regarded as a subspace of C`(V, q); we then have the fundamental relation uv + vu = −2 q(u, v) for u, v ∈ V. (6.6) Exercise 6.7. Check that (6.6) is a consequence of v2 = −q(v, v). Proposition 6.1. The algebra C`(V, q) satisfies the following universal property: if A is a real algebra with identity and f : V → A is a linear mapping such that f(v)2 = −q(v, v) 1, then there is a unique algebra homomorphism f˜ : C`(V, q)→ A such that f = f˜ ∣∣ V . Exercise 6.8. Give the proof of Proposition 6.1. A Clifford algebra is Z2-graded, as follows. By taking f(v) := −v in the previous propo- sition, with A = C`(V, q), we see that there is a unique automorphism α of C`(V, q) ex- tending f , such that α2 = id. (From the definition, each element of C`(V, q) is a linear combination of products v1v2 . . . vk with v1, . . . , vk ∈ V ; clearly we have α(v1v2 . . . vk) = (−1)k v1v2 . . . vk.) We write C`±(V, q) to denote the (±1)-eigenspaces of α. Thus C`(V, q) is a superalgebra. Notice that V ⊂ C`−(V, q). Now take A to be the opposite algebra of C`(V, q) (the same vector space with the product reversed), and f(v) := v; we get an involutive antiautomorphism of C`(V, q) given by (v1 . . . vk) ! := vk . . . v1. We define another antiautomorphism a 7→ a¯, called conjugation, by a¯ := α(a)! = α(a!). Many properties of the Clifford algebra come from the following simple proposition, due to Chevalley [16]. 3A Clifford algebra can be defined if q is a symmetric bilinear form of any signature. However, we shall use only forms which are either positive definite or (occasionally) negative definite. 65 Proposition 6.2. Let (V, q) and (W, r) be two Euclidean vector spaces, and let q ⊕ r be the symmetric bilinear form on V ⊕W given by (q⊕ r)(v1 +w1, v2 +w2) := q(v1, v2) + r(w1, w2). Then there is an isomorphism of superalgebras C`(V ⊕W, q ⊕ r) ' C`(V, q) ⊗ C`(W, r). Proof. Define f : V ⊕ W → C`(V, q) ⊗ C`(W, r) by f(v + w) := v ⊗ 1 + 1 ⊗ w. Using (6.5), we see that f(v + w)2 = −(q ⊕ r)(v + w, v + w) 1 ⊗ 1; therefore f extends to an algebra homomorphism f˜ : C`(V ⊕W, q ⊕ r) → C`(V, q) ⊗ C`(W, r). Since the elements v ⊗ 1 + 1 ⊗ w generate C`(V, q) ⊗ C`(W, r) as an algebra, f˜ is surjective; to see that it is injective, it suffices to compute f˜ on a basis for C`(V ⊕W, q ⊕ r) generated by bases for V and W . If e1, . . . , en is an orthonormal basis of (V, q), write eJ := ej1 . . . ejk for J = {j1, . . . , jk} ⊆ {1, 2, . . . , n} with 1 ≤ j1 < · · · < jk ≤ n; and let e∅ := 1. Then { eJ : J ⊆ {1, 2, . . . , n} } is a basis of C`(V, q). In particular, dim C`(V, q) = 2n and dim C`+(V, q) = dim C`−(V, q) = 2n−1. Exercise 6.9. Use the previous Proposition to verify this basis, by induction on n. Suppose V = Rn and that qn is the standard Euclidean inner product on Rn; in that case, we abbreviate C`(n) := C`(Rn, qn). Since C`(1) = span{1, e1} with e21 = −1, we have C`(1) ' C (as real algebras); with C`+(1) = R, C`−(1) = iR. Next, C`(2) = span{1, e1, e2, e1e2} ' H, the algebra of quaternions. In both these cases, a 7→ a¯ denotes the usual conjugation. Since C`(R,−q1) = span{1, eˆ1} with eˆ21 = +1, we have C`(R,−q1) ' R⊕ R, so C`(V, q) depends on the signature of the form q. Also, C`(R2,−q2) ' R2×2, by taking eˆ1 = ( 1 0 0 −1 ) , eˆ2 = ( 0 1 1 0 ) . A complete list of the algebras C`(Rn,±qn) is given in [4], and reproduced in [27, 39]. A basic fact is what is called Bott periodicity : C`(n + 8) ' C`(n) ⊗ C`(8) (ungraded tensor product). Since C`(8) ' R16×16 is a real matrix algebra, this means that C`(n+ 8) ' C`(n)16×16, so one only need determine C`(n) for n = 1, 2, . . . , 8. Exercise 6.10. Show that C`(3) ' H ⊕ H. The remaining cases are C`(4) ' H2×2, C`(5) ' C4×4, C`(6) ' R8×8, C`(7) ' R8×8 ⊕ R8×8. 6.3 Clifford actions Definition 6.9. A Clifford module for the Clifford algebra C`(V, q) is a superspace F = F+ ⊕ F− together with an even homomorphism (called a Clifford action) c : C`(V, q) → EndF . In other words, c(a)F± ⊆ F± if a ∈ C`+(V, q), and c(b)F± ⊆ F∓ if b ∈ C`−(V, q). Suppose that F has a (Euclidean or hermitian) inner product (·|·), and denote the adjoint of A ∈ EndF by A†, that is, (x |A†y) := (Ax |y)) for x, y ∈ F . We say that F is a selfadjoint Clifford module if c(a)† = c(a¯) for all a ∈ A. Notice that is enough to see that c(v)† = −c(v) for all v ∈ V , i.e., that each c(v) is skewadjoint. Write EndC`(V,q) F := {R ∈ EndF : [[c(a), R]] = 0, for a ∈ C`(V, q) } to denote the subalgebra of EndF consisting of operators which supercommute with the Clifford action. Exercise 6.11. Check that [[RS, c(v)]] = R[[S, c(v)]]+(−1)#S[[R, c(v)]]S for R, S homogeneous elements of EndF , and conclude that EndC`(V,q) F is indeed an algebra. 66 Definition 6.10. Let (V, q) be a finite-dimensional real vector space with a nondegenerate bilinear form.4 The dual space V ∗ of R-linear forms on V is a real vector space of the same dimension; we may define the so-called musical isomorphisms [8], namely [ : V → V ∗ and ] : V ∗ → V , by u[(v) := q(u, v), q(λ], v) := λ(v). They are mutually inverse: (u[)] = u, (λ])[ = λ. If W is a complex vector space with a hermitian inner product 〈〈· | ·〉〉 (or a Hilbert space,5 not necessarily finite-dimensional), we define [ : W → W ∗ and ] : W ∗ → W similarly by u[(v) := 〈〈u | v〉〉, 〈〈λ] | v〉〉 := λ(v). In the complex case, the musical isomorphisms are antilinear. Definition 6.11. Let (V, q) be an Euclidean vector space, and let Λ•V be its exterior algebra. For v ∈ V , define the exterior multiplication (v) and the contraction ι(v[) in End(Λ•V ) by (v)(v1 ∧ · · · ∧ vk) := v ∧ v1 ∧ · · · ∧ vk, ι(v[)(v1 ∧ · · · ∧ vk) := k∑ j−1 (−1)j−1q(v, vj)v1 ∧ · · j ∨· · ∧ vk. (6.7) (The notation j∨ indicates that the term with index j is missing from the exterior product.) For k = 0, we define (v)1 := v, ι(v[)1 := 0. Note that ι(v[) is a graded derivation:6 ι(v[)(α ∧ β) = ι(v[)α ∧ β + (−1)#αα ∧ ι(v[)β for α, β ∈ Λ•V. On Λ•V we use the real inner product (u1 ∧ · · · ∧ um, v1 ∧ · · · ∧ vn) := δmn det [ q(uk, vl) ] It is easily seen that (v)† = ι(v[) for v ∈ V . Define c(v)α := (v)α− ι(v[)α, for α ∈ Λ•V. (6.8) Exercise 6.12. Check that (v)† = ι(v[) and that (u)ι(v[) + ι(v[)(u) = q(v, u). Conclude that c(v)2 = −q(v, v) 1 for v ∈ V . By Proposition 6.1, c extends to a Clifford action of C`(V, q) on Λ•V . Since c(v)† = −c(v), this action is selfadjoint. 4This definition applies to both symmetric and antisymmetric bilinear forms. 5In the infinite-dimensional case, the bijectivity of the musical isomorphisms is a restatement of the Riesz theorem. 6It could have been defined as the unique graded derivation that takes u to q(v, u). 67 Definition 6.12. The R-linear map σ : C`(V, q)→ Λ•V given by σ(a) := c(a) 1 (6.9) for a ∈ C`(V, q), is a vector-space isomorphism (but not an algebra isomorphism) between the Clifford algebra and the exterior algebra, called [9] the symbol map. (It is easy to see that σ is surjective; since C`(V, q) and Λ•V have the same dimension 2n, it is also injective.) Note that σ(1) = 1 and σ(v) = v for v ∈ V . The inverse isomorphism Q : Λ•V → C`(V, q) may be called [9] a quantization map, since it maps a supercommutative algebra to an algebra which is no longer supercommuta- tive.7 Exercise 6.13. Show that σ(uv) = u∧v−q(u, v) 1, so that Q(u∧v) = uv+q(u, v) 1. Establish the general formula: Q(v1 ∧ · · · ∧ vk) = 1 k! ∑ τ∈Sk (−1)τ vτ(1) . . . vτ(k), (6.10) where (−1)τ denotes the sign of the permutation τ . In particular, Q(ej1∧· · ·∧ejk) = ej1 . . . ejk if {e1, . . . , en} is an orthonormal basis for V and j1 < · · · < jk. Exercise 6.14. The map Q does not preserve the Z-grading of Λ•V , but does preserve its filtration. A filtration of an algebra A is a system of subspaces Aj with Aj ⊆ Aj+1 and AiAj ⊆ Ai+j. For A = C`(V, q), take Aj to be the subspace generated by products at most j vectors in V . There is an associated graded algebra, namely grA := ⊕ j Aj/Aj−1 and a canonical “symbol map” σ : A → grA. Check that for A = C`(V, q), the associated graded algebra is none other than Λ•V with the symbol map given by (6.9). 6.4 Complex Clifford algebras Definition 6.13. The complexification C`(V, q) := C`(V, q) ⊗R C can be regarded as the Clifford algebra over C of the complexified vector space VC with the same quadratic form q amplified to VC: any vector in VC can be written uniquely as w = u + iv with u, v ∈ V , and one defines q(w,w′) := q(u, u′)− q(v, v′) + iq(u, v′)− iq(v, u′); notice that the amplified bilinear form q is not positive definite on VC. On VC all nondegenerate symmetric bilinear forms are equivalent, so we may abbreviate C`(V, q) to C`(V ); in particular, C`(V, q) ' C`(n) := C`(Cn, qn) if dimR V = n. Exercise 6.15. Prove that C`(1) ' C⊕ C and that C`(2) ' C2×2. 7Quantum field theory deals with Fermi fields, which are systems of operators {φ(v) : v ∈ V } satisfying an “anticommutation relation” akin to (6.6). It helps to imagine an analogous situation in which the anticommutators vanish, i.e., the algebra generated by the Fermi fields is replaced by a supercommutative algebra; restoration of the scalar terms q(u, v) then corresponds to quantizing the latter system. 68 Exercise 6.16. Prove that C`(n+2) ' C`(n)⊗C`(2) (ungraded tensor product)8 by showing that the C-linear map f : Cn+2 → C`(n) ⊗ C`(2) given by f(e1) := 1 ⊗ e1, f(e2) := 1 ⊗ e2, f(ej) := i ej−2⊗ e1e2 for j ≥ 3, satisfies {f(ej), f(ek)} = −2δjk(1⊗ 1), and hence extends to the desired isomorphism. From this we find that C`(2m) ' CN×N with N = 2m; and C`(2m+1) ' CN×N ⊕CN×N . In particular, C`(V, q) is a simple matrix algebra iff dimR V is even. From now on, we consider only the case that V has even dimension n = 2m. Definition 6.14. Any finite-dimensional module for the matrix algebra A = CN×N is of the form F = S1 ⊕ · · · ⊕ Sr where each Sk has dimension N and A acts irreducibly on each Sk. Another way to express this is to say that F = W ⊗ S, where dimC S = N and A acts trivially on W , that is, a · (w ⊗ s) = w ⊗ (a · s). Therefore, if dimV is even, any Clifford module for C`(V ) is of the form F = W ⊗ S, with c(a)(w ⊗ s) := w ⊗ c(a)s, where S is an irreducible Clifford module, unique up to equivalence, called the spinor module (or “spinor space”) for C`(V ). As algebras, C`(V ) and EndC(S) are isomorphic. Exercise 6.17. Define HomC`(V )(S, F ) to be the vector space of C-linear maps T : S → F which intertwine the Clifford actions of C`(V ) on S and on F , i.e., c(a)(Ts) := T (c(a)s) for all s ∈ S, a ∈ C`(V ). Check that the map T⊗s 7→ Ts gives an isomorphism HomC`(V )(S, F )⊗ S ' F , and deduce that HomC`(V )(S, F ) ' W and EndC`(V ) F ' EndCW . Definition 6.15. Let (V, q) be an oriented Euclidean vector space of even dimension n = 2m (over R), and let {e1, . . . , en} be an orthonormal basis for (V, q) which is compatible with the given orientation.9 We define the chirality element of C`(V ) as γ := im e1e2 . . . en. (6.11) If {e′1, . . . , e′n} is another oriented orthonormal basis, then e′j = ∑n k=1 gjkek for g ∈ SO(n) an orthogonal matrix of determinant +1, and so im e′1e ′ 2 . . . e ′ n = det(g)γ = γ: thus the recipe (6.11) is independent of the chosen basis. The chirality element γ satisfies γ2 = 1 (that is why the factor im belongs in the defini- tion), and it anticommutes with the copy of V within C`(V ). Indeed, ejγ = −γej since ej anticommutes with every ek except ej itself; by linearity, vγ = −γv for all v ∈ V . Therefore γvγ = −v for v ∈ V , so by Proposition 6.1 γaγ = α(a) for all a ∈ C`(V ). The Z2-grading on C`(V ) induced by γ (by declaring an element even or odd according as it commutes or anticommutes with γ) coincides, fortunately, with the original grading given by α. Notice that if n = 4k is a multiple of 4, then γ belongs to the real Clifford algebra C`(V, q), and it is not necessary to use the complexification C`(V ). 8This is the “Bott periodicity” identity for complex Clifford algebras. 9Recall that an orientation of V is a choice of a positive direction in the real line ΛnV ; a basis {v1, . . . , vn} for V is compatible with the orientation iff v1 ∧ · · · ∧ vn is positive. 69 6.5 The Fock space of spinors The availability of γ means that any (ordinary) complex module F under the Clifford algebra can be given a Z2-grading by taking F± to be the (±1)-eigenspaces of c(γ) ∈ EndC F . In particular, the spinor module is a superspace S = S+ ⊕ S−, with C`(V ) ' EndC S as superalgebras. We now give an explicit construction of the spinor module. Definition 6.16. A Euclidean vector space (V, q) of even real dimension n = 2m can be made into a complex Hilbert space by choosing an orthogonal complex structure J on (V, q); this is an R-linear operator on V satisfying: J2 = −1 and q(Ju, Jv) = q(u, v) for u, v ∈ V. (6.12) Now we regard V as a complex vector space via the rule (α+ iβ)v := αv+βJv for α, β ∈ R. The hermitian form 〈〈u | v〉〉 := q(u, v) + iq(Ju, v) is positive definite since q is. We shall denote the complex Hilbert space thus obtained by (V, q, J). Notice that i〈〈u | v〉〉 = 〈〈u | Jv〉〉 = −〈〈Ju | v〉〉 for u, v ∈ V . Exercise 6.18. For (V, q) = (R4, q4), find all J ∈ EndR V = R4×4 satisfying (6.12). Show that the set of such J has two connected components, each homeomorphic to a 2-sphere S2. Exercise 6.19. If g ∈ EndR(V ), check that g is a C-linear map on the Hilbert space (V, q, J) iff gJ = Jg. Conclude that { g ∈ GLR(V ) : gJ = Jg } = GLC(V, q, J). If O(V, q) denotes the orthogonal group of those g ∈ GLR(V ) for which q(gu, gv) = q(u, v) for all u, v ∈ V , show that UJ(V ) := O(V, q) ∩ GLC(V, q, J) is the unitary group of the Hilbert space (V, q, J), i.e., the group of complex automorphisms satisfying 〈〈gu | gv〉〉 = 〈〈u | v〉〉 for u, v ∈ V . Exercise 6.20. Let J(V, q) denote the set of all orthogonal complex structures on a Euclidean vector space (V, q). Show that J(V, q) is empty if dimR V is odd, and nonempty if dimR V is even (produce an example). Show that the orthogonal group O(V, q) acts transitively on J(V, q) by J 7→ gJg−1, and that the isotropy subgroup of any J is the unitary group UJ(V ). Exercise 6.21. Check that {u1, Ju1, . . . , um, Jum} is an orthonormal basis for (V, q) when {u1, . . . , um} is an orthonormal basis for (V, q, J), and deduce that O(V, q) ' O(R2m, q2m) ≡ O(2m) and that UJ(V ) ' Ui(Cm) ≡ U(m). Conclude that J(V, q) is diffeomorphic to the quotient manifold O(2m)/U(m) (which, as it happens, has two connected components).10 Definition 6.17. Choose and fix an orthogonal complex structure J on (V, q). Denote by FJ(V ) or simply by F(V ) the complex exterior algebra Λ •(V, q, J) over the Hilbert space (V, q, J). Then F(V ) is a complex vector space of dimension 2m, called the Fock space11 10The identity component SO(V, q) of the group O(V, q), which consists of those g with determinant +1, satisfies SO(V, q) ' SO(2m), so one of the two components of J(V, q) is the set of “orientation-preserving” J , which is diffeomorphic to SO(2m)/U(m). 11This is sometimes called a fermion Fock space to distinguish it from the Hilbert space formed from a symmetric algebra over V (suitably completed), which is known as a “boson Fock space”; these spaces describe many-particle states of fermions and bosons in quantum field theory. 70 over (V, q, J). The Fock space is itself a complex Hilbert space, with the inner product determined by 〈〈u1 ∧ · · · ∧ um | v1 ∧ · · · ∧ vn〉〉 := δmn det [〈〈uk | vl〉〉]. We can choose (by induction on dimV ) an orthonormal basis {e1, . . . , en} for V so that Je2r−1 = e2r for r = 1, . . . ,m. It is useful to identify (V, q, J) with an m-dimensional complex subspace W of the com- plexification VC = V ⊗R C; take W := { v − iJv ∈ VC : v ∈ V }. Notice that W is isotropic for the amplified bilinear form q on VC, i.e., q(u − iJu, v − iJv) = 0 for u, v ∈ V , on ac- count of (6.12); and moreover, dimCW = m = 1 2 dimC VC, so W is a maximally isotropic subspace.12 The conjugate subspace W := { v + iJv ∈ VC : v ∈ V } satisfies W ∩W = 0, W ⊕W = VC, and indeed W is the orthogonal complement of W under the inner product on VC given by 〈〈w | z〉〉 := 2 q(w¯, z). The operator P+ := 1 2 (I − iJ) is then a unitary isomorphism between (V, q, J) and W , so we may identify these Hilbert spaces. The Riesz theorem allows us to identify the dual space V ∗ with W , and if w = P+v, we identify v[ with w¯ := P−v := 12(I + iJ)v. An orthonormal basis for W is {z1, . . . , zm}, where zk := P+e2k−1; notice that P+e2k = P+(Je2k−1) = izk. We may also identify F(V ) with the complex exterior algebra Λ•W . Now define c(v) on Λ•W , for v ∈ V , by c(v)α := (P+v)α− ι(P−v)α, (6.13) where ι(z¯)(w1 ∧ · · · ∧ wk) = ∑k j=1(−1)j−1〈〈z | wj〉〉w1 ∧ · · j ∨· · ∧ wk, in accordance with (6.7), and extend to VC by c(v1 + iv2) := c(v1) + ic(v2). Exercise 6.22. Show that {(w), ι(z¯)} = 〈〈z | w〉〉 for w, z ∈ W . Exercise 6.23. Show that c(w) = (w) and c(z¯) = −ι(z¯) for w, z ∈ W . Conclude that c extends to a selfadjoint Clifford action on Λ•(W ). Proposition 6.3. The operator c(γ) on the Fock space is the grading operator on the super- space Λ•W . Proof. The assertion is that the (±1)-eigenspaces of c(γ) are the even and odd subspaces Λ±W ; in other words, c(γ)α = (−1)kα for all α ∈ ΛkW . First of all, notice that zkz¯k − z¯kzk = i e2k−1e2k, so that γ = ∏→ 1≤k≤m(zkz¯k − z¯kzk). Moreover, c(zkz¯k− z¯kzk) = [c(zk), c(z¯k)] = [ι(z¯k), (zk)]. Now a direct calculation shows that, for α = zj1 ∧ · · · ∧ zjk , we have [ι(z¯r), (zr)]α = ±α, with negative sign iff r ∈ {j1, . . . , jk}. Since Λ•W is generated (as a complex vector space) by such α, we obtain c(γ)α = (−1)kα. 12If Z ≤ VC has dimension k, the subspace of vectors w with q(w, z) = 0 for all z ∈ Z has dimension 2m − k, since q is nondegenerate; thus an isotropic subspace can have dimension at most m. A maximally isotropic subspace is also called a polarization for q. 71 6.6 The Pin and Spin groups Definition 6.18. Let (V, q) be a Euclidean vector space with q positive definite. The invertible elements of C`(V, q) form a group which includes all nonzero scalars and all nonzero v ∈ V (since v−1 = −v/q(v, v)); and the “twisted conjugation” φ(a)b := α(a)ba−1 provides a linear action of this group on C`(V, q). The invertible elements which preserve the subspace V under this action form the Clifford group Γ(V, q) := { a : φ(a)(V ) ⊂ V }. The nonzero scalars t ∈ R× lie in Γ(V, q), and so do the nonzero vectors u ∈ V \ {0}, since if v ∈ V , φ(u)v = −uvu−1 = (vu− 2q(u, v))u−1 = v − 2q(u, v) q(u, u) u, (6.14) so u ∈ Γ(V, q). Geometrically, (6.14) is the reflection in the hyperplane orthogonal to u, since φ(u)u = −u and φ(u)v = v iff q(u, v) = 0. Lemma 6.4. The kernel of φ : Γ(V, q)→ GLR(V ) is the subgroup R× of nonzero scalars. Proof. If a ∈ kerφ, let a = a+ + a− with a± ∈ C`±(V, q); taking even and odd parts of the equation α(a)v = va gives a±v = ±va± for v ∈ V . If {e1, . . . , en} is an orthonormal basis for V , and if V1 denotes the hyperplane in V orthogonal to e1, then a + = b+ + e1b − with b± ∈ C`±(V1, q). Now the equation a+e1 = e1a+ yields b+e1 + b− = e1b+ − b−, so b− = 0; hence a+ ∈ C`(V1, q). A similar argument shows that a− ∈ C`(V1, q). An obvious induction argument now shows that a has only a scalar component when expanded in the basis { eJ : J ⊆ {1, . . . , n} } of C`(V, q), and so a ∈ R×. If v ∈ V , then vv¯ = vα(v) = −v2 = q(v, v) 1, so the quadratic form q(·, ·) extends to the map a 7→ aa¯ of C`(V, q) into itself. If a = ∑J aJeJ is the expansion of a with respect to the basis { eJ : J ⊆ {1, 2, . . . , n} } of C`(V, q), then aa¯ and a¯a have the same scalar component ∑ J a 2 J . If a ∈ Γ(V, q) and v ∈ V , then φ(a)v = u implies u = α(a)−1va, and so u = u! = a!va¯−1 = φ(a¯)v; in consequence, φ(aa¯)v = φ(a)φ(a¯)v = v, so that aa¯ ∈ kerφ = R×. Thus a 7→ aa¯ = a¯a yields a group homomorphism of Γ(V, q) into R×. Its kernel is Pin(V, q) := { a ∈ Γ(V, q) : aa¯ = 1 }. If a ∈ Γ(V, q) and v ∈ V , then q(φ(a)v, φ(a)v) = α(a)va−1a¯−1v¯α(a¯) = α(aa¯)(a¯a)−1vv¯ = −v2 = q(v, v), since aa¯ = a¯a is a scalar, and so φ(a) lies in the orthogonal group O(V, q). Restriction to Pin(V, q) yields a homomorphism φ : Pin(V, q)→ O(V, q) which is surjective since the orthog- onal group is generated by the reflections13 (6.14), and v ∈ Pin(V, q) whenever q(v, v) = 1. Definition 6.19. The Spin group of the Euclidean vector space (V, q) is defined as the even part of Pin(V, q), i.e., Spin(V, q) := C`+(V, q) ∩ Pin(V, q) := { a ∈ Γ(V, q) ∩ C`+(V, q) : aa¯ = 1 }. 13This is the Cartan–Dieudonne´ theorem, which can be proved by writing an orthogonal matrix in normal form as a direct of k plane rotation matrices and possibly a −1 diagonal entry, and by recalling that a plane rotation is a product of two reflections in lines. 72 The image φ(Spin(V, q)) is the subgroup SO(V, q) of O(V, q) generated by products of an even number of reflections. This is the rotation group, also called the special orthogonal group, since g ∈ O(V, q) lies in SO(V, q) iff det g = 1. The homomorphisms φ : Pin(V, q) → O(V, q) and φ : Spin(V, q) → SO(V, q) have the same kernel {±1} ' Z2, by Lemma 6.4. In particular, we have a short exact sequence of Lie groups: 1−→Z2−→ Spin(V, q) φ−→SO(V, q)−→ 1, so that Spin(V, q) is a double covering of the rotation group. Exercise 6.24. Show that Pin(V, q) is the set of products {v1 . . . vk} of unit vectors (i.e., vectors vj ∈ V such that q(vj, vj) = 1 for each j), by showing that these products form a normal subgroup which φ maps onto O(V, q). Deduce that Spin(V, q) is the set of products {v1 . . . v2r} of an even number of unit vectors. Exercise 6.25. If u, v ∈ V are unit vectors in V , with u 6= ±v, and if a(t) := 1 + q(u, v) sin 2t for 0 ≤ t ≤ pi 2 , define w(t) := a(t)−1/2(cos t u + sin t v), z(t) := a(t)−1/2(cos t v + sin t u), and b := (1 − q(u, v)2)−1/2(uv + q(u, v) 1). Show that t 7→ −w(t)w(−t) is a continuous path in Spin(V, q) from 1 to −1 through b, and that t 7→ w(t)z(−t) is a continuous loop in Spin(V, q) from uv to b and back. Deduce that Spin(V, q) is a connected group.14 On passing to the complexification, we may regard Spin(V, q) as a subgroup of the in- vertible elements of the complex Clifford algebra C`(V ); the circle group U(1), regarded as complex scalars is another such subgroup. Now if λ, µ ∈ U(1) and a, b ∈ Spin(V, q), then λa = µb in C`(V ) iff (µ, b) = (λ, a) or else (µ, b) = (−λ,−a); thus Spin(V, q) and U(1) generate the subgroup Spinc(V ) ' (Spin(V, q)× U(1))/Z2, where the quotient map is defined by the relation (λ, a) ∼ (−λ,−a). Define φc(λa)v := λ2φ(a)v for λa ∈ Spinc(V ); then φc maps Spinc(V ) onto SO(V, q)× U(1) with kernel {±1}, so there is another short exact sequence of Lie groups: 1−→Z2−→ Spinc(V ) φ c−→SO(V, q)× U(1)−→ 1. The Lie algebra of the spin group Spin(V, q) is readily identified as a subspace of the Clifford algebra C`(V, q). Indeed, let A+2 be the subspace of C` +(V, q) with basis {1}∪{ eiej : i < j } with {e1, . . . , en} an (arbitrary) orthonormal basis of V (compare with Exercise 6.14), and let C2(V, q) be the image of Λ2V under the quantization map Q. Lemma 6.5. C2(V, q) = { b ∈ A+2 : b¯ = −b }. Proof. Since Q(u ∧ v) = uv + q(u, v) 1, it follows that C2(V, q) ⊆ A+2 . Counting dimensions, dimC2(V, q) = dim Λ2V = 1 2 n(n− 1) = dimA+2 − 1, so C2(V, q) is a hyperplane in A+2 . Now 14Since the fundamental group of SO(V, q) is Z2, this shows that the covering map φ is nontrivial and hence that Spin(V, q) is the universal covering group of SO(V, q). 73 eiej = ejei = −eiej for i < j, so the map b 7→ b¯ + b is a scalar-valued R-linear form on A+2 , whose kernel is a hyperplane; and since Q(u ∧ v) = vu+ q(u, v) 1 = −uv − q(u, v) 1 = −Q(u ∧ v) for u, v ∈ V , it coincides with C2(V, q). If b ∈ C2(V, q) and w ∈ V , then [[b, w]] = [b, w] lies in V also; to see that, it suffices to take b = uv + q(u, v) 1, whereupon [b, w] = uvw − wuv = uvw + uwv − uwv − wuv = 2q(u,w)v − 2q(v, w)u ∈ V. (6.15) The only elements of A+2 which commute with V are scalars, so [b, w] vanishes for all w ∈ V iff b = 0 in C2(V, q). Hence τ(b)v := [b, v] defines an injective R-linear map τ : C2(V, q) → EndV . Proposition 6.6. The map τ is a Lie algebra isomorphism from C2(V, q) onto the Lie algebra of antisymmetric operators so(V, q) := {A ∈ EndV : q(Au, v) ≡ −q(u,Av) }. Proof. First notice that C2(V, q) is a Lie algebra under the bracket [b, d] := bd − db of C`+(V, q). Indeed, on taking b = uv + q(u, v) 1, d = wz + q(w, z) 1, one finds that [b, d] = [uv, wz] = [uv, w]z + w[uv, z] ∈ A+2 using (6.15), and [b, d] = [d¯, b¯] = [−d,−b] = −[b, d], so [b, d] ∈ C2(V, q). Also from (6.15), q([b, w], z) = 2q(u,w)q(v, z)− 2q(v, w)q(u, z) = −q(w, [b, z]), which shows that τ(b) ∈ so(V, q). Since dim so(V, q) = 1 2 n(n − 1) = dimC2(V, q) and τ is injective, its image is all of so(V, q). Finally, the Jacobi identity yields τ([b, d])v = [[b, d], v] = [[b, v], d] + [b, [d, v]] = −τ(d)τ(b)v + τ(b)τ(d)v, so τ([b, d]) = [τ(b), τ(d)] where the latter bracket is the commutator bracket in EndV (which preserves so(V, q)); thus, τ is a Lie algebra isomorphism. Lemma 6.7. If A ∈ so(V, q) and {e1, . . . , en} is an orthonormal basis for V , then τ−1(A) = 1 2 ∑ js q(Aes, ek) ek − ∑ j 0 for all i, j. A Euclidean vector bundle has transition functions with values in the orthogonal group O(r); thus it is orientable iff one can choose the gij to satisfy det gij = +1. (This may be expressed by saying that the structure group of E can be reduced from O(r) to SO(r).) To see whether a Euclidean vector bundle is orientable or not, take a good covering U = {Uj} for M and choose transition functions gij for U. Then det gij : Ui ∩ Uj → {±1} = Z2, and since gijgjk = gik on each nonempty Ui ∩ Uj ∩ Uk, we conclude that det g is a Cˇech 1-cocycle2 in C1(U,Z2). Any set of transition functions is of the form g′ij = figijf−1j with fj : Uj → O(r), as follows from the proof of Proposition 1.5. (If a local system of sections sj for E transforms as si = gijsj, and an equivalent system of local sections is given by s′j := fjsj, the new transition functions satisfy g ′ ijfj = figij.) Hence detf ∈ C0(U,Z2) and det g′ = det g + δ(detf). Thus the class w1(E) := [det g] ∈ Hˇ1(M,Z2) depends only on the equivalence class [E] of the vector bundle. It is usually called the first Stiefel–Whitney class of E [23, 39]. It is also customary to write w1(M) := w1(TM) when (M, g) is a Riemannian manifold. We may summarize the situation as follows. Lemma 7.2. A Euclidean vector bundle E−→M is orientable iff w1(E) = 0 in Hˇ1(M,Z2). A Riemannian manifold (M, g) is orientable iff w1(M) = 0. Proof. The condition w1(E) = 0 means that transition functions g ′ ij can be chosen to satisfy det g′ij = +1 identically. Suppose now that E−→M is an oriented Euclidean vector bundle (i.e., that w1(E) = 0), with transition functions gij : Ui∩Uj → SO(r) for a good covering U of M . Since each Ui∩Uj is contractible, one can lift these maps to the double covering Spin(r) of SO(r), i.e., we can find smooth functions hij : Ui ∩ Uj → Spin(r) such that φ(hij) = gij for all i, j. The relation φ(hijhjkh −1 ik ) = gijgjkg −1 ik = 1 in SO(r) shows that aijk := hijhjkh −1 ik = ±1 2We identify {±1} with Z2, in order to use additive notation when combining Cˇech cocycles. 80 in Spin(r). Thus a is a Cˇech 2-cochain in C2(U,Z2); indeed, a is a Cˇech 2-cocycle, as a short calculation shows. Exercise 7.4. Compute δa, to check that it is trivial in C3(U,Z2). We can make random changes of signs of some of the hij, i.e., we can let h ′ ij := bijhij where bij : Ui ∩Uj → {±1} are arbitrary (but constant) sign functions. Then b ∈ C1(U,Z2), and a′ijk := h ′ ijh ′ jk(h ′ ik) −1 yields a′ ∈ C2(U,Z2) with a′ = a + δb. Therefore, the class w2(E) := [a] ∈ Hˇ2(U,Z2) depends only on the equivalence class [E] of the vector bundle; it is called the second Stiefel–Whitney class of E. One writes w2(M) := w2(TM) when (M, g) is an oriented Riemannian manifold. Definition 7.6. A spin structure on an oriented Riemannian manifold (M, g) of dimen- sion n is a principal Spin(n)-bundle P η−→M together with a bundle map τ : P → Q where Q θ−→M is the principal SO(n)-bundle of oriented orthonormal frames for the tangent bun- dle TM −→M , such that τ(p · g) = τ(p) ◦ φ(g) for p ∈ P , g ∈ Spin(n). We say that M is a spin manifold if it admits at least one spin structure. By Lemma 1.4, such a principal Spin(n)-bundle may be assembled from transition func- tions hij : Ui∩Uj → Spin(n) satisfying φ(hij) = gij, where the gij are the transition functions of the frame bundle, provided only that hijhjk = hik on each nonempty Ui ∩ Uj ∩ Uk. Thus M admits a spin structure iff w2(M) = 0 in Hˇ 2(M,Z2). Since τ : P → Q reproduces the double covering φ : Spin(n)→ SO(n) on each fibre, the map τ is two-to-one. Inequivalent spin structures may arise from different choices of hij covering the same gij. The only freedom here comes from the sign changes bij = ±1 mentioned above, where h′ij = bijhij determine another spin structure on M . Since hijhjk = hik and similarly for the h′ij, it follows that bijbjk = bik also; thus b is a Cˇech 1-cocycle. Now if b = δc with c ∈ C0(U,Z2), i.e., if bij = ci/cj with all cj = ±1, then h′ij = (ci/cj)hij as functions from Ui ∩Uj to the group Spin(n), so the corresponding principal Spin(n)-bundles are equivalent. Conversely, a principal bundle equivalence yields relations h′ij = (ai/aj)hij for some smooth functions aj : Uj → Spin(n) satisfying ai = ±aj on overlaps; thus, signs cj can be chosen so that ciai = cjaj on overlaps, and consequently h ′ ij = (ci/cj)hij. In summary, the two spin structures are equivalent iff b = δc for some c, which gives the following result. Lemma 7.3. If M is a Riemannian manifold with w1(M) = 0 and w2(M) = 0, the inequiv- alent spin structures on M are classified by the Cˇech cohomology group Hˇ1(M,Z2). The computation of the Stiefel–Whitney classes for particular manifolds involves either a good deal of combinatorial calculation (see [28], for instance) or general theorems from topology [23, 39, 41]. One such general theorem is particularly useful: if E−→M is a complex vector bundle and ER−→M is the underlying real vector bundle, then w1(ER) = 0 and w2(ER) is the image of the first Chern class c1(E) under the canonical homomorphism from Hˇ2(M,Z) to Hˇ2(M,Z2) obtained from the standard homomorphism j : Z → Z2 (namely, 81 reduction modulo 2).3 That w1(ER) = 0 should be no surprise, because a complex vector bundle is always orientable. To see that, notice firstly that the unitary group U(r) can be regarded as a subgroup of SO(2r), since U(r) permutes orthonormal bases in Cr, and any such basis {e1, . . . , er} yields an orthonormal basis {e1, f1, . . . , er, fr} for R2r, by setting fr := ier; and since U(r) is connected, it is therefore contained 4 in the identity component SO(2r) of the orthogonal group of R2r. Now, with respect to any Hermitian metric on E, a local orthonormal basis of sections for Γ(U,E) gives a local orthonormal basis of sections for Γ(U,ER), and the U(r)-valued transition functions of E may thus be regarded as SO(2r)- valued transition functions of ER. Since the first Chern class c1(H) of the hyperplane bundle H → CPm is a generator of Hˇ2(CPm,Z) ' Z, so that c1(H)↔ 1 under this group isomorphism, its modulo-2 reduction is not zero: therefore, w2(HR) 6= 0 in Hˇ2(CPm,Z2). Now it can be shown [28] that the tangent bundle TCPm has the following property: if Er−→CPm denotes the trivial real bundle of rank r, then TCPm ⊕ E2 and HR ⊕ · · · ⊕HR (with (m + 1) summands) are equivalent real bundles over CPm. These facts suffice to decide which of the CPm are spin manifolds. Proposition 7.4. The complex projective space CPm is a spin manifold iff m is odd; and for odd m, the spin structure on CPm is unique. Proof. The isomorphism C`(V, q) ⊗ C`(W, r) ' C`(V ⊕W, q⊕r) of Proposition 6.2, restricted to the spin subgroups, shows that Spin(V, q) and Spin(W, r) may be regarded as commuting subgroups of Spin(V ⊕W, q⊕r). Thus we may embed the direct product Spin(k)×Spin(l) as a subgroup of Spin(k+l); and φmaps this to the usual embedding of SO(k)×SO(l) in SO(k+l). Thus, whenever E−→M and F −→M are oriented Euclidean vector bundles, the transition functions gij⊕g′ij of E⊕F take values in SO(k)×SO(l), and lift to hij⊕h′ij in Spin(k)×Spin(l). The second Stiefel–Whitney class therefore satisfies5 the relation w2(E⊕F ) = w2(E)+w2(F ). When M = CPm, we thereby obtain w2(CPm) + 0 = w2(TCPm ⊕ E2) = w2(HR ⊕ · · · ⊕HR) = (m+ 1)w2(HR). (7.5) Any nontrivial element of the group Hˇ2(CPm,Z2) is of order two,6 and w2(HR) 6= 0, so the right hand side of (7.5) vanishes iff m is odd. To get uniqueness, we must show that Hˇ1(CPm,Z2) = 0. This follows from the Bockstein homomorphism construction of subsection 1.11, applied to the exact sequence of abelian groups 0−→Z 2−→Z j−→Z2−→ 0, 3See Appendix B of [39] for a proof. 4If one chooses an orthonormal basis in Cr for which g ∈ U(r) is diagonal, i.e., g(ej) = eiαjej , then g(ej) = cosαj ej + sinαj fj , g(fj) = − sinαj ej + cosαj fj , so the image of g in SO(2r) is a direct sum of 2× 2 rotation blocks. 5When E and F are not orientable, this additivity breaks down; there is a “product formula” of Whitney [41] which yields the relation w2(E ⊕ F ) = w2(E) + w1(E)w1(F ) + w2(F ). 6Indeed, Hˇ2(CPm,Z2) is a vector space over the field Z2. 82 where ‘2’ denotes multiplication by two, and j : Z→ Z2 is reduction modulo 2. The Bockstein construction yields a long exact sequence in Cˇech cohomology · · · → Hˇ1(M,Z2) ∂−→ Hˇ2(M,Z) 2∗−→ Hˇ2(M,Z) j∗−→ Hˇ2(M,Z2) ∂−→ Hˇ3(M,Z)→ · · · for any compact manifold M . Let us chase this diagram backwards from Hˇ3(M,Z) in the case M = CPm. We know that Hˇ3(CPm,Z) = 0 from (2.5), so j∗ is surjective.7 We know also that Hˇ2(CPm,Z) = Z (these are the Chern classes found in subsection 5.10), so the image of 2∗, which equals the kernel of j∗, is the subgroup 2Z of “even” Chern classes, and 2∗ is just multiplication by 2, which is injective. This forces ∂ : Hˇ1(CPm,Z2)→ Hˇ2(CPm,Z) to be the zero homomorphism, and thus j∗ : Hˇ1(CPm,Z)→ Hˇ1(CPm,Z2) is onto; but Hˇ1(CPm,Z) = 0, again by (2.5), so we conclude that Hˇ1(CPm,Z2) = 0. 7.3 Spinc structures What can be done about manifolds like CP2, which are complex manifolds but do not admit a spin structure? Since our immediate aim is to describe an irreducible module for the complex Clifford bundle over such manifolds, we could relax our requirements slightly by replacing the required structure group of the tangent bundle by Spinc(n) rather than Spin(n). (Recall that Spinc(n) is the product of the group Spin(n) and the unitary scalars U(1) within the complex Clifford algebra C`(V ).) We say that a (compact, oriented, Riemannian) manifold M admits a spinc structure if there is a principal Spinc(n)-bundle P c η−→M and a bundle map τ : P c → Qc, satisfying τ(p · g) = τ(p) ◦ φc(g) for p ∈ P and g ∈ Spinc(n); here Qc is a principal SO(n) × U(1)-bundle of the form Q × R−→M , where8 Q is the SO(n)-bundle of oriented orthonormal frames of the tangent bundle, as before, and R−→M is a principal U(1)-bundle which we may choose as we please. Now a principal U(1)-bundle is just the frame bundle of a complex line bundle L→ M , so (up to equivalence) these are classified by Hˇ2(M,Z). The modulo-2 reduction j∗[L] is an element of Hˇ2(M,Z2), where the second Stiefel-Whitney class also lives. Suppose that w2(M) = j∗[L∗] in H2(M,Z2), i.e., that w2(M) + j∗[L] = 0; then one can find local sections h˜ij : Ui ∩ Uj → Spinc(n) such that φc(h˜ij) = λijgij are the transition functions of TM ⊕ L−→M , which patch together properly to give the required principal bundle P c−→M . Exercise 7.5. Assume that w2(M) + j∗[L] = 0 and construct P c−→M as indicated. If a spinc structure exists, M is called a spinc manifold. The foregoing argument says that this is the case iff w2(M) ∈ j∗(Hˇ2(M,Z)). Moreover, inequivalent spinc structures are parametrized by the classes of complex line bundles, i.e., by Hˇ2(M,Z). 7Actually, the formula (2.5) refers to singular or de Rham cohomology; but we can establish an iso- morphism between de Rham and Cˇech cohomology in degree 3 by a simple modification of the proof of Proposition 5.5. For the isomorphism between the de Rham–Cˇech cohomologies in any degree, see [12, 17]. 8The fibrewise product of principal bundles corresponds, by association, to the Whitney sum of vector bundles; thus, if Q and R are the frame bundles of vector bundles E and F over M , with respective structure groups G and H, then Q×R−→M may be defined as the frame bundle of E⊕F −→M , which is a principal G×H-bundle. 83 Lemma 7.5. The complex projective space CPm is a spinc manifold for any positive inte- ger m. Proof. In the proof of Proposition 7.4, we verified the surjectivity of the homomorphism j∗ : Hˇ2(CPm,Z)→ Hˇ2(CPm,Z2), without reference to the parity of m. In fact, any compact complex manifold carries a spinc structure. This follows from the relation w2(M) = j∗(c1(TholM)), where TholM −→M is the holomorphic tangent bundle, i.e., the complex vector bundle whose fibres Tx,holM are spanned by the tangent vectors (∂/∂z1)x, . . . , (∂/∂z m)x. To get a more concrete construction, we first regard U(m) as a subgroup of SO(2m) and consider the homomorphisms τ : U(m)→ SO(2m)×U(1) given by τ(g) := (g, det g), and σ : U(m) → Spinc(2m) defined as follows. For each g ∈ U(m) there is an orthonormal basis {e1, . . . , em} of Cm which diagonalizes g, i.e., g(ej) = eiαjej; write fj := iej and aj := e iαj/2(cos 1 2 αj + (sin 1 2 αj)ejfj) ∈ C`+(2m); then σ(g) := a1a2 . . . am ∈ Spinc(2m). Exercise 7.6. Check that all the aj commute, 9 that σ is a well-defined homomorphism, and that φc ◦ σ = τ . Exercise 7.7. The linear map on R2m determined by Jej := fj, Jfj := −ej is a complex structure; and U(m) = { g ∈ SO(2m) : gJ = Jg }. Write k := e1f1 + · · ·+ emfm ∈ C`+(2m) and check that exp(pik) = (−1)m and φ(exp(pi 4 k)) = J . Define the metaunitary group MU(m) as { a ∈ Spin(2m) : ak = ka } and verify that φ(MU(m)) = U(m). Show that the centre of the group MU(m) equals { exp(θk) : −pi < θ ≤ pi } ' U(1) if m is odd, whereas the centre is {± exp(θk) : 0 ≤ θ < pi } ' U(1)× Z2 if m is even. Now we can exhibit a spinc structure on a complex manifold M . Take any Hermitian metric on M and let Q′−→M be the unitary frame bundle of the holomorphic tangent bundle TholM −→M . Associate to it, firstly, a principal Spinc(2m)-bundle P −→M via the homomorphism σ, as in (1.3); and secondly, a principal U(1)-bundle R−→M via the homomorphism det : U(m)→ U(1). The latter is just the frame bundle of the complex line bundle K∗−→M where K −→M is the so-called canonical line bundle on M , defined by Γ(K) := Am,0(M). Since φc ◦ σ = τ , the principal bundle P −→M yields the required spinc structure. Of course, a spin manifold is automatically a spinc manifold. It suffices to use the trivial U(1) bundle R = M × U(1) to build a spinc structure from a given spin structure. Exercise 7.8. Complete the following construction of the spin structure for CPm, with m odd, due to Dabrowski and Trautman [21]. Show that CPm is diffeomorphic to the homogeneous space U(m+ 1)/(U(1)×U(m)) and that the unitary frame bundle is given by Q′ := U(m+ 9For a fixed orthonormal basis {e1, . . . , em} of Cm, the elements a1, . . . , am and the scalars eiθ1 generate a subgroup T of Spinc(2m) which is isomorphic to an (m + 1)-dimensional torus Tm+1; moreover, since φc(a1a2 . . . am) is a block diagonal matrix over R2m, the centralizer of T in Spinc(2m) is T itself, so it is a maximal torus. If b ∈ Spinc(2m), then by pulling back φc(b) to U(m) via τ and diagonalizing the resulting unitary matrix, one sees that b lies in some conjugate subgroup aTa−1. This exemplifies the well-known theorem of Weyl [13] that the maximal tori in a compact connected Lie group are conjugate and cover the whole group. 84 1)/U(1). Let MU(m) be the metaunitary group of Exercise 7.7 and let P ′ := MU(m + 1)/U(1), where the subgroup U(1) of MU(m + 1) is given as { exp(θk) : 0 ≤ θ < pi }, the identity component of the centre. Check that the inclusion MU(m) ⊂MU(m+1) drops to a free right action of MU(m) on P ′, and that the double covering φ : MU(m+ 1)→ U(m+ 1) drops to a double covering τ ′ : P ′ → Q′, which intertwines the respective actions of MU(m) and U(m). Finally, show how the inclusions MU(m) ↪→ Spin(2m) and U(m) ↪→ SO(2m) associate to these new principal bundles P and Q together with a double covering τ : P → Q which yields the desired spin structure. Catalogues of spin manifolds and spinc manifolds are given in several places, e.g., [29, 39]. As well as the classes of manifolds just discussed, it is worth mentioning that any sphere Sn is spin; any simply connected Lie group is spin; any orientable 2-dimensional manifold is spin (e.g., any Riemann surface); any 3-dimensional manifold is spin; and any orientable 4-dimensional manifold is spinc. As counterexamples, we note that CP2 is an orientable 4-dimensional manifold which is not spin; and the homogeneous space SU(3)/SO(3) is an orientable 5-dimensional manifold which is not spinc. 7.4 The spinor module Definition 7.7. Let M be a spin manifold of even dimension n = 2m, and let P −→M be the principal Spin(n)-bundle defining its spin structure.10 Let S = Λ•CW be an irreducible Clifford module for C`(n), i.e., a Fock space of complex dimension 2m, and let c : Spin(n)→ EndC S denote the spin representation. Let S(M)−→M be the complex vector bundle associated to the spin structure P via the spin representation c. It is called the spinor bundle over M . Its module of sections Γ(S(M)) is an irreducible Clifford module, and will be called the spinor module for the algebra Γ(C`(M)). Proposition 7.6. If M is an even-dimensional spin manifold, any Clifford module Γ(F ) is of the form Γ(W ⊗ S(M)), where Γ(W ) is a trivial Clifford module, i.e., the Clifford action is (κ,w) 7→ w for κ ∈ Γ(C`(M)), w ∈ Γ(W ). Proof. Let Γ(F ) be given. If A = C∞(M), any A-linear map from Γ(S(M)) to Γ(F ) is of the form τ∗ where τ : S(M)→ F is a bundle map, i.e., τ∗ is an element of Γ(Hom(S(M), F )). Thus any such map which intertwines the two Clifford actions belongs to Γ(W ), where W := HomC`(M)(S(M), F ) is the vector bundle over M with fibres Wx = HomC`(n)(S(M)x, Fx). From Exercise 6.17 it follows that wx ⊗ σx 7→ wx(σx) gives an vector space isomorphism Wx ⊗ S(M)x ' Fx. (If wx(σx) = 0 for any nonzero σx, then 0 = c(a)[wx(σx)] = wx(c(a)σx) for all a, so wx = 0 by irreducibility of S(M)x.) The intertwining property wx(c(a)σx) = c(a)[wx(σx)] shows that Wx⊗S(M)x becomes a C`(n)-module via the recipe c(a)[wx⊗σx] := wx ⊗ c(a)σx, and the aforementioned isomorphism intertwines this action with the given action on Fx. Globally, we obtain an invertible bundle map from W ⊗S(M) to F and hence an A-linear isomorphism from Γ(W ⊗ S(M)) to Γ(F ) which intertwines the action c(κ)[w ⊗ σ] := w ⊗ c(κ)σ (7.6) 10If M carries more than one spin structure, we choose and fix a particular one. 85 with the given Clifford action on Γ(F ). Exercise 7.9. Prove that Γ(EndC`(M) F ) ' Γ(EndCW ); in other words, match A-linear op- erators commuting with the Clifford action on Γ(F ) to A-linear operators on Γ(W ). Definition 7.8. The passage from the irreducible spinor module Γ(S(M)) to a Clifford module of the form Γ(W ⊗S(M)) whose Clifford action is given by (7.6) is called a twisting by the vector bundle W −→M (which in principle may be any complex vector bundle). We refer to Γ(W ⊗ S(M)) as a twisted Clifford module. Notice, in particular, that any Clifford module Γ(F ) where F −→M has minimal rank 2m is obtained by twisting the spinor module with a complex line bundle. Clearly, then, the twisting may affect the global topology of F , but for algebraic properties of the Clifford action it is enough to study the spinor module. 7.5 The spin connection The next task is to show that the spinor module admits a distinguished connection, to be called the “spin connection”, which satisfies a “Leibniz rule” with respect to the Clifford action. This is not a trivial matter, as the spinor module is not “tensorial”, that is, its existence has topological obstructions not faced by modules of tensors, vector fields, or differential forms. Therefore the algebraic techniques used in Section 5 to construct new connections from old ones are not enough to produce a suitable connection on the spinor bundle. What is needed is to use the spin representation. This we have defined, in Section 6, as a group representation of Spin(V, q) as unitary operators on S which does not drop to a representation of the special orthogonal group SO(V, q). It may, however, be regarded as a projective (or “double-valued”) representation of SO(V, q). The corresponding infinitesimal representations of Lie algebras are not troubled by this topological problem,11 and there is a linear isomorphism τ : C2(V, q)→ so(V, q) between the Lie algebras. Definition 7.9. Let µ˙ : so(V, q)→ End+C(S) be the linear map given by µ˙(A) := c(τ−1(A)), (7.7) Note that µ˙(A) is skewhermitian since the Clifford action is selfadjoint, so c(b)† = c(b¯) = −c(b) for b ∈ C2(V, q) by Lemma 6.5. We call µ˙ the derived spin representation. Lemma 7.7. The derived spin representation satisfies the identity [µ˙(A), c(v)] = c(Av) (7.8) for all A ∈ so(V, q), v ∈ V . 11This is because V is finite-dimensional. In the infinite-dimensional case, the covering map φ has a 1-dimensional kernel (a circle), which gives rise to a “spin anomaly” in the infinitesimal representation of SO(V, q). For details, consult [31]. 86 Proof. It is equivalent to show that [τ−1(A), v] = Av as elements of the Clifford algebra C`(V, q). However, this follows immediately from the formula of Lemma 6.7 for τ−1(A), and from (6.15): [τ−1(A), v] = 1 4 n∑ j,k=1 q(Aej, ek) [ejek, v] = 1 2 n∑ j,k=1 q(Aej, ek) ( q(ej, v)ek − q(ek, v)ej ) = 1 2 n∑ j=1 q(ej, v)Aej + 1 2 n∑ k=1 q(ek, v)Aek = Av, on using the antisymmetry of A. It would be helpful to have a version of (7.8) which is valid when vectors v ∈ V are replaced by arbitrary elements b ∈ C`(V, q). To accomplish this, one needs to extend the operator A on V to a linear operator on C`(V, q). In fact, A extends as a derivation of the Clifford algebra: define Aˆ(1) := 0, Aˆ(v) := Av for v ∈ V , and Aˆ(v1 . . . vk) := k∑ j=1 v1 . . . vj−1(Avj)vj+1 . . . vk (7.9) for v1, . . . , vk ∈ V . Exercise 7.10. Check that the definition (7.9) is consistent by first verifying that Aˆ(uv) + Aˆ(vu) = 0 for u, v ∈ V . Then show that (7.8) extends to the identity [µ˙(A), c(b)] = c(Aˆb) for all A ∈ so(V, q), b ∈ C`(V, q). Now we can translate these algebraic identities to relations among sections of bundles over a spin manifold M , by applying them at each fibre. Thus, if A ∈ Γ(EndE) where E−→M is a Euclidean vector bundle, and if (As | t) = −(s | At) for s, t ∈ Γ(E), then A(x) ∈ so(Ex) for x ∈ M , so we may write A ∈ Γ(so(E)). In particular, if E = TM is the tangent bundle, the notation A ∈ Γ(so(TM)) means that A is an operator on Γ(TM) = X(M) satisfying g(AX, Y ) = −g(X,AY ) for X, Y ∈ X(M). We write µ˙(A) to denote the section x 7→ µ˙(A(x)) of the spinor bundle S(M)−→M . This defines a map µ˙ : Γ(so(TM)) → Γ(S(M)). Furthermore, by tensoring these A-modules of sections with Ak(M), we obtain maps µ˙ : Ak(M, so(TM))→ Ak(M,S(M)). If ∇ is a connection on the tangent bundle compatible with the metric, such as the Levi- Civita connection for instance, then, locally at least, ∇ = d + α where α ∈ A1(U, so(TM)) in view of the metric compatibility condition (5.7). Thus µ˙(α) makes sense as an element of A1(U, S(M)), where U is the domain of α. Theorem 7.8. Let M be an even-dimensional spin manifold and let ∇ be the Levi-Civita connection on the complex Clifford bundle C`(M)−→M . Then there is a connection ∇S on 87 the spinor bundle S(M)−→M satisfying the following Leibniz rule: ∇S(c(κ)σ) = c(∇κ)σ + c(κ)(∇Sσ), (7.10) for all κ ∈ Γ(C`(M)), σ ∈ Γ(S(M)). Proof. Let U = {Uj} be a covering of M by chart domains and let {fj} be a smooth partition of unity subordinate to U, with fj(x) > 0 iff x ∈ Uj. Then fj∇ is a connection on Uj compatible with the metric fjg. Suppose for the moment that there exist connections ∇Sj on each restriction of the spinor bundle S(M)−→M to Uj, satisfying ∇Sj (c(κ)σ) = c(fj∇κ)σ+c(κ)(∇Sj σ), whenever σ vanishes outside Uj; then the recipe ∇Sτ := ∑ j∇Sj (fjτ), for any τ ∈ Γ(S(M)), defines a connection satisfying (7.10). Therefore, we may replace M by any chart domain Uj; or we may equivalently suppose that the vector bundles C`(M)−→M and S(M)−→M are trivial, and that the Levi-Civita connection may be written as ∇ = d + α with α ∈ A1(M, so(TM)). The formula ∇ = d+α, defined initially on vector fields, is also applicable to Clifford products of vector fields, provided ακ is interpreted as the derivation action of α on κ ∈ Γ(C`(M)), in view of the Leibniz rule (7.4). Now introduce a connection ∇S on S(M) by defining ∇S := d+ µ˙(α). (7.11) From (7.8), extended to Clifford algebra, we get [µ˙(α), c(κ)] = c(ακ) on Γ(S(M)), and then ∇S(c(κ)σ) = d(c(κ)σ) + µ˙(α)(c(κ)σ) = c(dκ)σ + c(κ) dσ + [µ˙(α), c(κ)]σ + c(κ)µ˙(α)σ = c(dκ)σ + c(ακ)σ + c(κ) ( dσ + µ˙(α)σ ) = c(∇κ)σ + c(κ)(∇Sσ), verifying (7.10). We have constructed one solution ∇S to (7.10). If ∇˜S is another connection on S(M) sat- isfying the same module-derivation property, then ∇˜S−∇S is given by β ∈ A1(M,EndS(M)) such that β(c(κ)σ) ≡ c(κ)(βσ). That is to say, β lies in A1(M,EndC`(M) S(M)) ' A1(M) in view of Exercise 7.9 and the irreducibility of S(M). Therefore, ∇S is unique up to addition of a scalar action by a 1-form on M . We may apply the same reasoning to any Clifford module, with the following result. Proposition 7.9. Let Γ(F ) = Γ(W ⊗ S(M)) be a Clifford module over a spin manifold M , and let ∇˜ be a connection on F which is a module derivation, i.e., ∇˜(c(κ)τ) ≡ c(∇κ)τ + c(κ)(∇˜τ). Then there is a unique connection ∇W on W such that ∇W ⊗ 1 + 1⊗∇S = ∇˜. Proof. Since τ ∈ Γ(F ) is a finite sum of the form ∑j wj ⊗ σj, we may rewrite the module derivation property of ∇˜ as ∇˜(w ⊗ c(κ)σ) = w ⊗ c(∇κ)σ + c(κ)(∇˜(w ⊗ σ)). Let ∇W0 be an arbitrary connection on W . Then we also get [∇W0 ⊗ 1 + 1⊗∇S, 1⊗ c(κ)](w ⊗ σ) = w ⊗ c(∇κ)σ, 88 so β := ∇˜− (∇W0 ⊗1 + 1⊗∇S) lies in A1(M,EndF ) and commutes with the Clifford action, that is, β ∈ A1(M,EndC`(M) F ). By Exercise 7.9 it is of the form β = α ⊗ 1 for a unique α ∈ A1(M,EndW ). Writing ∇W := ∇W0 +α now gives ∇˜ = ∇W ⊗1+1⊗∇S, as desired. Lemma 7.10. The curvature of the spin connection is µ˙(R), where R denotes the Rieman- nian curvature, R ∈ A2(M, so(TM)). Proof. It suffices to check this on a chart domain U , where we may write ∇S = d+µ˙(α). The curvature form of the spin connection is given by (5.10) as ωS := µ˙(dα+α∧α) = µ˙(R). 7.6 Local coordinate formulas Definition 7.10. On a chart domain U of M with local coordinates {x1, . . . , xn}, we shall write the basic vector fields as ∂j ≡ ∂/∂xj. If M is Riemannian, the Christoffel symbols Γkij of the Levi-Civita connection on the chart domain U are the functions in C ∞(U) defined by ∇∂i∂j = Γkij∂k, (7.12) or equivalently, ∇∂j = Γkij dxi ⊗ ∂k. In particular, Γkij = dxk (∇∂i∂j). Notice also that Γk•j give the matrix components αkj of α ∈ A1(U,EndTM). Exercise 7.11. Check that the Levi-Civita connection on the cotangent bundle is determined locally by ∇ dxk = −Γkij dxi ⊗ dxj, or equivalently, ∇∂i dxk = −Γkij dxj. Exercise 7.12. From the definition (5.27) of the Levi-Civita connection, show that Γkij = 1 2 gkl(∂igjl + ∂jgil − ∂lgij), (7.13) using the relations [∂i, ∂j] = 0. Exercise 7.13. Show that the Christoffel symbols on the sphere S2 are given in spherical coordinates by Γφθφ = Γ φ φθ = cot θ, Γ θ φφ = − sin θ cos θ, Γkij = 0 otherwise, (7.14) by applying (7.13) to the metric g := dθ2 + sin2 θ dφ2. Exercise 7.14. The components of the Riemann curvature tensor may be written [52] as Rijkl := dx i(R(∂k, ∂l)∂j). Verify the “Ricci identities”: Rijkl = ∂kΓ i lj − ∂lΓikj + ΓmljΓikm − ΓmkjΓilm, by using the curvature formula (5.11). It is somewhat more common to use the components Rijkl = gimR m jkl = (∂i |R(∂k, ∂l∂j)). One sees from (5.11) that, for k, l fixed, the matrix with (i, j)-entry Rijkl is antisymmetric. Exercise 7.15. Verify that the curvature of the spin connection is given locally by ωS(∂k, ∂l) = −14Rijkl c(dxi) c(dxj), (7.15) on account of (7.7). 89 We now express the spin connection itself in local coordinates. Fix a chart domain U ⊂ M , and write g = gij dxi · dxj there. Since the matrix G = [gij] is positive definite, we can find a matrix H = [hαj ] of functions in C ∞(U) such that H tH = G. Indeed, if G1/2 is the positive definite square root of G, we may take H = AG1/2 where A is an orthogonal matrix at each point of U with A : U → SO(n) smooth.12 Choose and fix such an H, and let H−1 = [h˜rβ] be its inverse matrix. Recall that g −1 = grs ∂r · ∂s is the metric on the cotangent bundle T ∗M −→M ; thus we have hαi δαβh β j = gij, h˜ i αδ αβh˜jβ = g ij. Orthonormal bases for A1(U) and X(U) are then given by θα := hαj dx j, Eβ := h˜ r β∂r. (We use the convention of reserving Latin indices for coordinate bases and Greek indices for orthonormal bases.) By construction, we get g(Eα, Eβ) = δαβ, g −1(θα, θβ) = δαβ, and also (θα)] = Eα, (Eβ) [ = θβ on U .13 Locally, a smooth section of the spinor bundle looks like a smooth map from an open subset U of M (which can actually be taken as the complement of the closure of an arbitrarily small open set in M [19]) to the Hilbert space S of the spinor representation. Let { γα ≡ γα : α = 1, . . . , n } be a fixed set of unitary skewadjoint operators on S with the property γαγβ + γβγα = −2 δαβ. [For instance, take γα := c(eα) where {e1, . . . , en} is an orthonormal basis for (V, q).] We set c(dxr) := h˜rβγ β. (7.16) From (7.7) it is immediate that c(dxr)c(dxs) + c(dxs)c(dxr) = −2grs = −2(dxr | dxs), (7.17) so that (7.16) in fact defines a local Clifford action of A1(U) on Γ(U, S). Conversely, a given Clifford action on the spinor bundle, which necessarily satisfies (7.17), together with a given fixed set of γα, defines via (7.16) matrices [h˜rβ] and [h α j ] satisfying (7.7). 14 It is convenient to introduce “mixed” or “orthogonal” Christoffel symbols Γ˜βiα, Γ̂ β α in C∞(U) by ∇∂iEα =: Γ˜βiαEβ; ∇EEα =: Γ̂βαEβ. 12We require detA = 1 so as to preserve the orientation when passing to orthonormal bases in (7.7). 13These last relations give rise to the common phrasing: “we need not worry about raising and lowering indices, so long as we deal with orthonormal bases”. 14The possibility of varying the Clifford action by premultiplying H by arbitrary SO(n)-valued functions, provided only that these be compatible with changes of local charts, reflects the noncanonical nature of the spinor bundle. 90 The Γ˜βiα and Γ̂ β α are antisymmetric in the indices α, β; for instance, Γ˜βiα + Γ˜ α iβ = g(∇∂iEα, Eβ) + g(Eα,∇∂iEβ) = ∂i(δαβ) = 0, on account of the compatibility with the metric: α ∈ A1(U, so(TM)). The components of the spin connection are given by the EndS-valued functions ωi := 1 4 Γ˜βiα γ αγβ, ν := 1 4 Γ̂βα γ αγβ = 1 4 h˜iΓ˜ β iα γ αγβ. (7.18) Exercise 7.16. Show that hαj Γ˜ β iα = −∂i(hβj ) + Γkijhβk . Deduce the formula ωi = 1 4 ( Γkijgkl − ∂i(hαj )δαβhβl ) c(dxj) c(dxl), that expresses the spin connection in a coordinate basis. Exercise 7.17. Let X = aj∂j be the local expression of a vector field on M . Show that the contraction of X with the spin connection is given locally as an operator on C∞(U, S) by: ∇SXσ = aj(∂jσ + ωj)σ, and, in particular, that ∇SEσ = Eσ + ν(x)σ. Exercise 7.18. Check that ∇S is a hermitian connection on the spinor module. 8 Dirac operators and Laplacians A Dirac operator is an operator of odd parity on a Clifford module over a spin (or spinc) ma- nifold, which is a first-order differential operator whose corresponding Leibniz rule involves Clifford multiplication by differentials of functions. Its square is therefore an even-parity second-order differential operator, which gives a far-reaching generalization of the Laplace- Beltrami operator on a Riemannian manifold: we may call this square a “generalized Lapla- cian”, using the terminology of [9]. Moreover, any Dirac operator over a compact manifold is elliptic, which implies that its inverse (off its finite-dimensional kernel) is a compact op- erator, so it has a discrete spectrum of eigenvalues which give precise information about the geometry of the manifold. Even more importantly, as Connes [18] has pointed out, the Dirac operator determines the Riemannian metric, and therefore serves as a gateway for reformu- lating the entire corpus of Riemannian geometry in analytic terms, allowing its extension to discrete spaces, fractals, spaces of tilings and group orbits, and many other contexts whose geometric structure seemed only a few short years ago to be hopelessly beyond reach. 8.1 Connections and differential forms Definition 8.1. Let ∇ be a connection on the cotangent bundle T ∗M −→M of a compact manifold M ; the Leibniz rule ∇(ξ ∧ ω) = (∇ξ)∧ ω+ ξ ∧ (∇ω) extends it to a connection on 91 the exterior product bundle Λ•T ∗M −→M , so that ∇ maps A•(M) to A1(M)⊗A A•(M).1 The exterior product defines an A-linear map ˆ : A1(M)⊗AA•(M)→ A•(M) by ˆ(ω⊗η) := ω ∧ η. The composition ˆ ◦ ∇ is an R-linear endomorphism of A•(M). In local coordinates ∇ω = dxj ⊗∇∂jω, and so ˆ(∇ω) = (dxj)∇∂jω where (dxj) denotes exterior multiplication by dxj. We introduce also a contraction map ιˆ : X(M)⊗A A•(M)→ A•(M) by ιˆ(X ⊗ η) := ιXη. If (M, g) is a Riemannian manifold, we identify X(M) with A1(M) via the metric g, and write ιˆ(α⊗η) := ια]η. The composed map ιˆ◦∇ is another R-linear endomorphism of A•(M). Lemma 8.1. Let ∇ be a connection on the cotangent bundle T ∗M −→M ; and let T ∈ A2(M,TM) be the torsion of the dual connection ∇∗ on the tangent bundle. Then for any ω ∈ A•(M), ˆ(∇ω) = dω − ιTω, (8.1) where ιT : A k(M)→ Ak+1(M) denotes contraction with the torsion. Proof. Write ∇ω = βk ⊗ ηk ∈ A1(M) ⊗ A•(M); then ∇Xω = βk(X) ηk by contraction. Therefore, if X, Y ∈ X(M) and ω ∈ A1(M), ˆ(∇ω)(X, Y ) = (βk ∧ ηk)(X, Y ) = βk(X)ηk(Y )− βk(Y )ηk(X) = (∇Xω)(Y )− (∇Y ω)(X) = X(ω(Y ))− Y (ω(X))− ω(∇∗XY −∇∗YX) = dω(X, Y )− ω(∇∗XY −∇∗YX − [X, Y ]) = dω(X, Y )− ω(T (X, Y )). (8.2) This establishes (8.1) for 1-forms. Moreover, ιY ιX(ˆ(∇ω)) = ιY ιX(dω)− ιT (X,Y )ω, for a form ω ∈ Ak(M) of any degree, by an easy extension of the above argument. Exercise 8.1. Generalize (8.2) to higher-degree forms. Corollary 8.2. If ∇ is a connection on the cotangent bundle T ∗M −→M whose dual con- nection is torsion-free, then ˆ ◦ ∇ equals the exterior derivation, d. In particular, ˆ◦∇LC = d for the Levi-Civita connection ∇LC (on the cotangent bundle). There is a kind of dual formula relating the Levi-Civita connection with the codifferen- tial δ. Before revealing that, it is useful to consider the divergence of a vector field on a Riemannian manifold. 1We find it more convenient in this Section to write Ar(M,E) = Ar(M)⊗A Γ(E), reversing the order of the tensor product given in (5.3). The Leibniz rule (5.4) is adapted accordingly, and the contraction operator ιX : A1(M,E)→ Γ(E) is now given by ιX(β ⊗ s) := β(X) s. 92 8.2 Divergence of a vector field Definition 8.2. Let X be a vector field on an oriented manifold M with volume form ν. The divergence of X relative to ν is the function divν X ∈ C∞(M) determined by (divν X)ν := LXν. If f ∈ C∞(M) is never zero, fν is another volume form with (divfν X) fν = LX(fν) = (Xf)ν + fLXν = (Xf + f divν X)ν, so that divfν X = divν X + Xf f . (8.3) If (M, g) is a Riemannian manifold with Riemannian volume form Ω, we write divX ≡ divΩX. Since LXΩ = d(ιXΩ) by Cartan’s identity, the identity ∫ M (divX)Ω = 0 (the divergence theorem) is an immediate consequence of Stokes’ theorem. Over a chart domain U , we may consider the local volume form µ = dx1 ∧ · · · ∧ dxn. If X = Xj ∂j ∈ X(U), then by computing LX(dx1 ∧ · · · ∧ dxn) we obtain the standard formula divµX = ∂jX j. Since Ω = √ det g µ on U by (3.3), we derive from (8.3) a local formula for the Riemannian divergence: divX = ∂jX j +Xj ∂j(ln √ det g) = 1√ det g ∂j(X j √ det g). Lemma 8.3. The Riemannian divergence of a vector field is obtained from the Levi-Civita connection on the tangent bundle via the local formula divX = dxj(∇LC∂j X). (8.4) Proof. The connection ∇LC is determined over a chart domain U by the Christoffel symbols Γkij of (7.12), satisfying Γ k ij = dx k (∇LC∂i ∂j). The Leibniz rule now yields, for X = Xj ∂j ∈ X(U), dxj(∇LC∂j X) = ∂jXj + ΓjjkXk. The verification of (8.4) thus reduces to checking the formula Γjjk = ∂k(ln √ det g). This relation is well known [36], but it is instructive to see how it goes. Recall that the Christoffel symbols are given by (7.13): Γkij = 1 2 gkl(∂igjl + ∂jgil − ∂lgij). 93 If [Gij] denotes the adjugate matrix to [gij] (that is, the matrix of cofactors), then by Cramer’s rule gij = Gij/ det g. Now gjl(∂jgkl − ∂lgkj) = 0 since the matrix [gij] is symmetric, and so Γjjk = 1 2 gjl ∂kgjl = 1 2 det g Gjl ∂kgjl = 1 2 det g ∂ det g ∂gjl ∂kgjl = 1 2 det g ∂k(det g) = 1 2 ∂k(ln det g), on using the Cramer expansion det g = gjlG jl. If we choose local orthonormal bases {θα} for A1(U) and {Eβ} for X(U) with Eα = (θα)], we may write, as in (7.7), θα = hαj dx j and Eβ = h˜ k β∂k. Then θ α(∇LCEαX) = hαj dxj(h˜kα∇LC∂k X), so we obtain the alternative formula to (8.4), using local orthonormal bases: divX = θα(∇LCEαX). (8.5) Yet another divergence formula, for a particular set of vector fields, is the following. Lemma 8.4. Let M be an oriented Riemannian manifold. Given ζ ∈ Ak(M) and η ∈ Ak−1(M), define Z ∈ X(M) by ω(Z) := (ζ | ω ∧ η) for ω ∈ A1(M). The divergence of this vector field is given locally by divZ = Eα(ζ | θα ∧ η), where {θ1, . . . , θn} and {E1, . . . , En} are local orthonormal bases of 1-forms and vector fields respectively, with Eα = (θ α)]. Proof. Since ω 7→ (ζ |ω∧η) is A-linear on A1(M), it indeed defines a vector field Z ∈ X(M). Since Ω = θ1 ∧ · · · ∧ θn locally, we find that ιZΩ = ∑n α=1(−1)α−1(ζ | θα ∧ η) θ1 ∧ · · α∨· · ∧ θn, and it follows that LZΩ = d(ιZΩ) = Eα(ζ | θα ∧ η) Ω. 8.3 The Hodge–Dirac operator revisited Lemma 8.5. If ∇LC denotes the Levi-Civita connection on the cotangent bundle of an oriented Riemannian manifold M , and δ : A•(M) → A•(M) is the codifferential, then ιˆ ◦ ∇LC = −δ. Proof. Choose ζ ∈ Ak(M) and η ∈ Ak−1(M) with k ≥ 1, and define Z ∈ X(M) by ω(Z) := (ζ | ω ∧ η). Choose local orthonormal bases {θ1, . . . , θn} and {E1, . . . , En} of 1-forms and vector fields, with Eα ≡ Eα = (θα)]. Since (ξ | β ∧ ω) = (ιβ]ξ | ω) for any ξ, ω ∈ A•(M) and β ∈ A1(M), we obtain (ζ | dη) = (ζ | ˆ(∇LCη)) = (ζ | (θα)∇LCEα η) = (ιEαζ | ∇LCEα η) = Eα(ιEαζ | η)− (∇LCEα ιEαζ | η) = Eα(ζ | θα ∧ η)− (ιEα∇LCEα ζ | η) = divZ − (ιˆ(∇LCζ) | η), 94 where we have used Corollary 8.2, the metric compatibility of ∇LC , and the commutation relation [∇X , ιY ] = ι[X,Y ] (see Exercise 5.21). On multiplying both sides by Ω and integrating, the term divZ is killed by Stokes’ theorem, and we obtain 〈〈δζ |η〉〉 := 〈〈ζ |dη〉〉 = −〈〈ιˆ(∇LCζ)| η〉〉; since η is arbitrary, this gives ιˆ(∇LCζ) = −δζ. Corollary 8.6. If cˆ : Γ(C`(M)) ⊗ A•(M) → A•(M) denotes the Clifford action on the de Rham algebra of an oriented Riemannian manifold, and if ∇LC denotes the Levi-Civita connection, then cˆ ◦ ∇LC = d+ δ. (8.6) Proof. If ∇ω = βk ⊗ ηk, then cˆ(∇ω) = c(βk)ηk = (βk)ηk − ι(βk)]ηk, so that cˆ = ˆ− ιˆ. Thus (8.6) follows immediately from Corollary 8.2 and Lemma 8.5. In subsection 4.5, we referred to d+δ as the “Hodge–Dirac operator”, which we obtained, in a somewhat ad-hoc fashion, as a square root of the Hodge Laplacian. The formula (8.6) reveals its true nature as the composition of a Clifford action and a connection which is compatible with this action (recall Proposition 7.1). This opens the way for the general definition of Dirac operators. 8.4 Dirac operators Definition 8.3. Let Γ(F ) be a selfadjoint Clifford module over a compact oriented Rieman- nian manifold M , that is, let F −→M be a Hermitian vector bundle and let there be given a selfadjoint Clifford action cˆ : Γ(C`(M) ⊗ F ) → Γ(F ), where we write cˆ(κ ⊗ ψ) := c(κ)ψ. Let ∇ be a Hermitian connection on F −→M that satisfies the Leibniz rule ∇(c(κ)ψ) = c(∇LCκ)ψ+ c(κ)(∇ψ), for κ ∈ Γ(C`(E)), ψ ∈ Γ(F ). The Dirac operator associated to the connection ∇ and the Clifford action cˆ is D/ := cˆ ◦ ∇. Via Γ(F ) ∇−→A1C(M)⊗AΓ(F ) cˆ−→Γ(F ), this D/ is a C-linear endomorphism of Γ(F ). In local coordinates, D/ψ = c(dxj)∇∂jψ. There are many Dirac operators to be found. First of all, if M is a spin manifold, we may consider the irreducible Clifford module Γ(S), i.e., the spinor module, with its spin connection ∇S. Then D/ S := cˆ ◦ ∇S is the original Dirac operator.2 Any other Clifford module Γ(S) on M is obtained by twisting, that is, S = W ⊗ F where W −→M is a Hermitian vector bundle carrying the trivial Clifford action, and the compatible connection is determined, via Proposition 7.9, by an arbitrary Hermitian connection on W . The Dirac operator d+ δ on the de Rham algebra arises in this way, since Λ•T ∗M ' S ⊗ S ′ where S ′−→M is the superbundle obtained from the spinor bundle by taking the opposite grading: (S ′)± := S∓. 2D/ S is usually called the Dirac operator on the spinor module. 95 Exercise 8.2. If c : C`(V )→ EndC Λ•CW is the Clifford action on Fock space, and if v ∈ V is a unit vector, show that c′(a) := −c(vav) is another selfadjoint Clifford action, whose restric- tion to Spin(V, q) switches the two irreducible subrepresentations of the spin representation. Conclude that there is an irreducible Clifford module S ′−→M on any spin manifold, whose “even” subspace Γ(S ′+) is the “odd” subspace Γ(S−) of the spinor module, and viceversa.3 Exercise 8.3. Define a Clifford action on the conjugate Fock space c¯ : C`(V ) → EndC Λ•CW by setting c¯(v)α¯ := (P−v)α¯ − ι(P+v)α¯ for v ∈ VC, α¯ ∈ W [compare with (6.13)]. Check that c¯(w) = −ι(w) for w ∈ W and c¯(z¯) := (z¯) for z ∈ W . If γ denotes, as usual, the chirality element of C`(V ), show that c¯(γ)α¯ = (−1)k+1α¯ whenever α¯ ∈ ΛkCW [compare with Proposition 6.3]. If S = Λ•CW is the spinor module for C`(V ), the dual Hilbert space S∗ is identified with Λ•CW via 〈z¯1 ∧ · · · ∧ z¯r, w1 ∧ · · · ∧ ws〉 := δrs det [ 2q(z¯k, wl) ] , that is, the conjugation C : w1∧· · ·∧ws 7→ w¯1∧· · ·∧w¯s is the (antilinear) Riesz isomorphism of S with S∗. Check that C intertwines the action c¯ on S∗ with the action c on S, and conclude that S∗ is an irreducible Clifford module for C`(V ) whose Z2-grading is given by (S∗)+ = C(S−) and (S∗)− = C(S+).4 Even if M is not spin, there may be many Clifford modules with compatible connections; the de Rham algebra with the Levi-Civita connection is again the prime example. In any case, since the obstruction to the existence of a spin structure is global, we can always write Γ(U, F ) = Γ(U,W ⊗ S) over a chart domain (where W may depend on U) and from any one compatible connection we can manufacture others by altering ∇W over U only. Alternatively, we may vary the Clifford action by redefining (7.16) (which amounts to an action of the SO(n) frame bundle), as explained in subsection 7.6. Moreover, we have the further freedom of making a smooth change in the metric g and thereby changing the Clifford action c on Γ(F ) (and also the connection, if necessary, to preserve compatibility). Thus any one Dirac operator gives rise to a large family of “smoothly perturbed” Dirac operators on the same Clifford module. Any Dirac operator is a first-order differential operator. To avoid any possible confusion of terminology, we make a formal definition. Definition 8.4. Let E−→M be a vector bundle. Any A ∈ Γ(EndE) defines a C-linear operator on Γ(E) by left multiplication, that is, (As)x := Ax(sx) for s ∈ Γ(E), x ∈M . Any connection5 ∇ on E−→M , when contracted by vector fields X ∈ X(M), provides other C-linear operators ∇X on Γ(E). The differential operators on E−→M are defined as the elements of the subalgebra D(M,E) generated by Γ(EndE) and by {∇X : X ∈ X(M) }. 3In more sophisticated terms, the spinor bundles S and S′ over a spin manifold are associated to the spin structure P −→M via the representations c and c′ of Spin(2m); these vector bundles are inequivalent since the representations c and c′ are inequivalent. 4We are indebted to William Ugalde for clarification on this point. 5In view of (5.5), two such connections differ by the action of an element of Γ(EndE), so only a single connection is needed here. 96 A finite sum of operators of the form A∇X1 . . .∇Xr with r ≤ k (and r = k for at least one summand) is called a differential operator of order k. By a partition-of-unity argument, a differential operator is of order k if it can be written as such a sum in each chart domain separately. Thus a Dirac operator D/ = c(dxj)∇∂j is a differential operator of first order. As an operator on the superspace Γ(F ), the Dirac operator D/ is odd, that is, D/ (Γ(F±)) ⊆ Γ(F∓), since cˆ : A1C(M)⊗A Γ(F±)→ Γ(F∓). Thus we may write D/ =: ( 0 D/ − D/ + 0 ) , (8.7) where D/ ± : Γ(F±)→ Γ(F∓). Its square is an even operator: D/ 2± : Γ(F±)→ Γ(F±); and D/ 2 = ( D/ −D/ + 0 0 D/ +D/ − ) . Last but not least, a Dirac operator is essentially selfadjoint. This means that D/ : Γ(F )→ Γ(F ) extends uniquely to a selfadjoint operator on the Hilbert space L2(F ) obtained by completing the space of smooth sections Γ(F ) with respect to the inner product 〈〈φ | ψ〉〉 :=∫ M (φ | ψ) Ω. Regrettably, D/ is always an unbounded operator, so this selfadjoint extension is still only densely defined. Rather than worry about identifying the precise domain of the extended D/ , we will stick to the original domain Γ(F ); a full proof of essential selfadjoint- ness (see [39], for instance) shows that nothing is thereby lost, as the closure of D/ on this domain is the full selfadjoint extension.6 We shall therefore show merely that D/ is formally selfadjoint, that is, 〈〈D/φ | ψ〉〉 = 〈〈φ |D/ψ〉〉 for φ, ψ ∈ Γ(F ). (8.8) Proposition 8.7. The Dirac operator D/ is formally selfadjoint. Proof. The argument is similar to that of Lemma 8.5. By invoking partitions of unity, we can reduce the problem to verifying (8.8) when φ, ψ are smooth sections of F −→M which vanish outside some chart domain; thus we write D/ = c(θα)∇Eα , where {θ1, . . . , θn} and {E1, . . . , En} are local orthonormal bases of 1-forms and vector fields, with Eα ≡ Eα = (θα)]. Now (φ |D/ψ) = (φ | c(θα)∇Eαψ) = −(c(θα)φ | ∇Eαψ) = −Eα(c(θα)φ | ψ) + (∇Eαc(θα)φ | ψ) = −Eα(c(θα)φ | ψ) + (c(∇LCEα θα)φ | ψ) + (c(θα)∇Eαφ | ψ) = −Eα(c(θα)φ | ψ) + (c(∇LCEα θα)φ | ψ) + (D/φ | ψ), (8.9) 6In other words, Γ(F ) is a “core” for D/ [45]. The compactness of M is not indispensable here: essential selfadjointness can be proven under the weaker assumption that the Riemannian manifold M is complete [39, 53, 61]. 97 where we have used the skewadjointness of c(θα), the hermiticity of the connection ∇, and the compatibility of ∇ with the Levi-Civita connection on the cotangent bundle. We claim that we can find a vector field Z ∈ X(M), depending on φ and ψ, such that divZ = −Eα(c(θα)φ|ψ)+(c(∇LCEα θα)φ|ψ). Then integration over M of both sides of (8.9) (multiplied by the Riemannian volume form Ω) yields the desired relation (8.8). The vector field Z is defined simply by ω(Z) := (φ | c(ω)ψ) = −(c(ω)φ | ψ), since the right hand side is A-linear in ω ∈ A1(M). Now −Eα(c(θα)φ | ψ) + (c(∇LCEα θα)φ | ψ) = Eα(θα(Z))− (∇LCEα θα)(Z) = θα(∇LCEαZ) from the definition of a dual connection. The divergence formula (8.5) says that the right hand side equals divZ, as claimed. 8.5 Laplacians Definition 8.5. Let ∇E : Γ(E) → A1(M,E) = Γ(T ∗M ⊗ E) be a connection on a vector bundle E−→M over a Riemannian manifold (M, g), and let ∇˜E := ∇LC ⊗∇E be its tensor product with the Levi-Civita connection on the cotangent bundle of M . Then ∇˜E maps Γ(T ∗M ⊗ E) to A1(M,T ∗M ⊗ E) = Γ(T ∗M ⊗ T ∗M ⊗ E), so it can be composed with ∇E. Contraction with the metric g−1 on T ∗M gives an A-linear map Trg : Γ(T ∗M⊗T ∗M⊗E)→ Γ(E). The composition of these three maps yields the following operator on Γ(E): ∆E := −Trg ◦∇˜E ◦ ∇E, called the Laplacian associated to the connection ∇E. The minus sign is a convention7 which assures that ∆E is a positive operator whenever ∇E is a Hermitian connection. To express ∆E in a more tractable form, we shall compute the result of contracting the operator ∇˜E ◦ ∇E with two vector fields X, Y ∈ X(M). If s ∈ Γ(E), we can write ∇Es = βk ⊗ sk, and so ιY ιX(∇˜E∇Es) = ιY ∇˜EX(βk ⊗ sk) = ιY (βk ⊗∇EXsk +∇Xβk ⊗ sk) = βk(Y )∇EXsk + (∇Xβk)(Y ) sk = ∇EX(βk(Y )sk)−X(βk(Y )) + (∇Xβk)(Y ) sk = ∇EX(βk(Y )sk)− βk(∇XY ) sk = ∇EX(∇EY s)− ι(∇XY )∇Es where we have written the Levi-Civita connection simply as ∇. In other words, ιY ιX(∇˜E ◦ ∇E) = ∇EX∇EY −∇E∇XY . 7Actually, the usual convention is to use the opposite sign; however, this results in operators, such as∑ j ∂ 2 j on Rn, which are negative definite. 98 Since g−1 = gij∂i ·∂j on a chart domain, we get immediately from (7.12) the local expression for the Laplacian: ∆E = −gij(∇E∂i∇E∂j − Γkij∇E∂k). (8.10) With a local orthonormal basis of vector fields {e1, . . . , en}, this takes a slightly simpler form:8 ∆E = − n∑ α=1 (∇Eeα∇Eeα −∇E∇eαeα). (8.11) It is immediate from (8.10) or (8.11) that the Laplacian ∆E is a second-order differential operator. Exercise 8.4. Use the formulas of Exercise 7.11 to give an alternative derivation of (8.10). Exercise 8.5. Show that the Laplacian ∆0 associated to the standard connection d on the trivial line bundle S2 × C−→ S2 is ∆0 = − ( ∂2 ∂θ2 + cot θ ∂ ∂θ + ∂2 ∂φ2 ) , i.e., the Laplace–Beltrami operator on A0(S2). Proposition 8.8. Let ∇E be a Hermitian connection on a vector bundle E over a Rie- mannian manifold (M, g). Then its Laplacian is a formally selfadjoint and positive operator on Γ(E). Proof. The strategy of the proof should by now be familiar: we choose sections s, t ∈ Γ(E) and with them construct a vector field Z ∈ X(M) such that (s|∆Et) = divZ+Tr(∇Es|∇Et). (The last term is locally expressed as ∑n α=1(∇Eeαs |∇Eeαt), where {e1, . . . , en} are orthonormal vector fields.) Multiplying these functions by Ω and integrating over M , we get the relation 〈〈s |∆Et〉〉 = 〈〈∇Es | ∇Et〉〉, (8.12) where these brackets denote integrated inner products on Γ(E) and A1(M,E) respectively.9 Positivity follows from setting s = t, and formal selfadjointness follows by repeating the argument with the roˆles of s and t interchanged. The vector field Z is defined by g(Z,X) := −(s | ∇EXt) for X ∈ X(M). Its divergence may be computed with the formula (8.5): divZ = θα(∇eαZ) = g(∇eαZ, eα) = −g(Z,∇eαeα) + eα g(Z, eα) = (s | ∇E∇eαeαt)− eα(s | ∇Eeαt) = (s |∆Et)− (∇Eeαs | ∇Eeαt), (with summation over α), so divZ = (s |∆Et)− Tr(∇Es | ∇Et) as claimed. 8We rename the vector fields Eα to eα temporarily to avoid a notational clash with the exponent E denoting a vector bundle. 9On account of (8.12), the Laplacian is often denoted ∇∗∇, as in [39], for instance. 99 We already know a second-order differential operator on the bundle Λ•T ∗M −→M , namely the “Hodge Laplacian”, which we now write ∆Hodge. It is natural to ask whether this operator equals the Laplacian ∆LC associated to the Levi-Civita connection on the exterior bundle. It turns out that they are not the same: indeed, they differ by a differential operator of order zero, that is essentially the Ricci tensor of Riemannian geometry [8, 36, 39]. Definition 8.6. The Ricci tensor on a Riemannian manifold is the symmetric tensor Ric of bidegree (2, 0) obtained from the Riemann curvature tensor R by defining Ric(X, Y ) as the trace of the A-bilinear form (W,Z) 7→ (W | R(Z, Y )X) on X(M). Locally, Ric(X, Y ) = dxk(R(∂k, Y )X). In terms of the components of the Riemann curvature tensor, we get Ric(∂i, ∂j) = R k ikj. By contracting with the metric, we obtain the curvature scalar10 K := Trg ◦Ric of the Riemannian manifold (M, g); notice that K ∈ C∞(M). Locally, K = gijRkikj. Exercise 8.6. Show that the Riemannian curvature tensor on the sphere S2 satisfies Rφθφθ = 1 and Rθφθφ = sin 2 θ, and deduce that Ric = g: the Ricci tensor coincides with the metric on S2. Conclude that K = 2 on S2: the sphere is a surface of constant (Gaussian) curvature. We can identify Ric with a tensor of bidegree (0, 2) on M , also called Ric, via the metric g; or better, with the element of Γ(EndT ∗M) defined by Ric(ω)(X) := Ric(ω], X). The relation between the Hodge Laplacian and the connection Laplacian, as operators on A•(M), is then given by the so-called Weitzenbo¨ck formula: ∆Hodge = ∆ LC + Ric . (8.13) We shall not prove this here, though we have the means to do so; consult [9] or [39] for the details. Two things are notable about the formula (8.13). First of all, ∆Hodge = (d + δ) 2 is the square of the Dirac operator d+ δ on the Clifford module A•(M), so the formula says that, for the de Rham complex at least, the square of the Dirac operator is almost, but not quite, a Laplacian; more precisely, it differs from the Laplacian by a differential operator of lower order that, by the way, depends only on the curvature of the connection which defines both the Dirac operator and the Laplacian. This is one of a family of formulae due variously to Bochner, Weitzenbo¨ck and Lichnerowicz, which say that D/ 2 − ∆ is a curvature-dependent multiplication operator. The second noteworthy feature is that ∆Hodge and ∆ LC are positive (formally) selfadjoint operators. What happens if the Ricci operator is positive too? For one thing, both ∆LC and Ric must then vanish on the kernel of the Hodge Laplacian, namely, on the harmonic forms. Thus, for example, on S2 the Ricci operator kills 0-forms and 2-forms but acts as the identity on 1-forms since the Ricci tensor coincides with the metric; the upshot of (8.13) is then that any harmonic 1-form on S2 must be zero (as we already noted in Section 4), and therefore H1dR(S2) = 0. This result is an example of a vanishing theorem, whereby certain cohomology groups reduce to zero —a topological result— on account of positivity properties of certain analytic operators. 10The curvature scalar is also called the Gaussian curvature when dimM = 2. 100 8.6 The Lichnerowicz formula Proposition 8.9. Let M be a compact spin manifold, let D/ S = cˆ ◦ ∇S denote the Dirac operator on the irreducible spinor module Γ(S), and let ∆S be the Laplacian associated to the spin connection ∇S. Then (D/ S)2 = ∆S + K 4 , (8.14) where K is the curvature scalar of M . Proof. It suffices to prove this on any chart domain; that is, we must show that, in a local coordinate basis, ( c(dxj)∇S∂j )2 = −gij(∇S∂i∇S∂j − Γkij∇S∂k)+ 14K. The left hand side is c(dxi)∇S∂ic(dxj)∇S∂j = c(dxi) c(dxj)∇S∂i∇S∂j + c(dxi) c(∇∂idxk)∇S∂k = c(dxi) c(dxj) (∇S∂i∇S∂j − Γkij∇S∂k) (8.15) = 1 2 [[c(dxi), c(dxj)]] (∇S∂i∇S∂j − Γkij∇S∂k)+ 12c(dxi) c(dxj)[∇S∂i ,∇S∂j ], where we have used the symmetry Γkij = Γ k ji due to the zero torsion ∇∂i∂j = ∇∂j∂i of the Levi-Civita connection. Now [[c(dxi), c(dxj)]] = −2gij —recall (7.17)— and [∇S∂k ,∇S∂l ] = ωS(∂k, ∂l) = −14Rijkl c(dxi) c(dxj) by (7.15), so by taking (8.10) into account we arrive at (D/ S)2 = ∆S − 1 8 Rijkl c(dx k) c(dxl) c(dxi) c(dxj). It remains only to check that the second term on the right reduces to 1 4 K. We may rewrite it as 1 8 Rjikl c(dx k) c(dxl) c(dxi) c(dxj) since Rijkl is antisymmetric in the indices i, j by (5.29). Since the antisymmetrization of Rjikl in the indices i, k, l vanishes by (5.28), it follows from Exercise 8.7 below that this term reduces to 1 8 Rjikl (−gkl c(dxi) c(dxj) + gil c(dxk) c(dxj)− gik c(dxl) c(dxj)), and since Rjiklg kl = 0 by antisymmetry of Rjikl in k, l, again by (5.29), this in turn reduces to 1 4 Rijklg ik c(dxl) c(dxj) = 1 4 Rkjki c(dx i) c(dxj) = 1 4 gijRkikj = 1 4 K, due to the symmetry Ricij := R k ikj of the Ricci tensor. Exercise 8.7. If u, v, w are three vectors in a Euclidean vector space (V, q), denote their antisymmetrized product by a := 1 6 (uvw+vwu+wuv−uwv−wvu−vuw) ∈ C`(V, q). Show that uvw = a− q(v, w)u+ q(u,w)v − q(u, v)w. The identity (8.14) is due to Lichnerowicz [40]. The proof generalizes in a straight- forward manner [9, 27, 39] to the case of a twisted Clifford module. The only differences consist in replacing the Laplacian ∆S by ∆F , where F = W ⊗ S, and the curvature term 1 2 c(dxi) c(dxj)ωS(∂i, ∂j) in (8.15) by 1 2 c(dxi) c(dxj)ωF (∂i, ∂j). Now by Proposition 7.9, the 101 curvature ωF can be decomposed as ωF = ωW +ωS, where ωW is the curvature of some com- patible connection on W . This yields an extra term of the form 1 2 c(dxi) c(dxj)ωW (∂i, ∂j), which, in view of (6.10), we may write as Q(ωW ), where Q : A•(M,W )→ C`(M) is the ex- tension of the quantization map of Definition 6.12 to the (trivial) Clifford module A•(M,W ). The result is a generalization of (8.14) known as the Bochner–Weitzenbo¨ck formula: (D/ F )2 = ∆F + 1 4 K +Q(ωW ). Exercise 8.8. When F = Λ•T ∗M ' S ⊗ S, verify (8.13) by showing that Ric−1 4 K = Q(ω0) for a suitable ω0 ∈ A2(M,S). 9 The Dirac operator on the Riemann sphere This section is devoted to a detailed exploration of a single but fundamental example: the Dirac operator on the irreducible spinor module over the sphere S2. While the sphere is undoubtedly the simplest possible even-dimensional compact spin manifold, its Dirac oper- ator exemplifies the full complexity of the general case while remaining directly accessible by elementary computations. We give here an account of its action on spinors, show its equivariance under the Lie group SU(2) of symmetries of the spinor module, compute its spectrum and exhibit a full set of eigenspinors. Surprisingly, such an account is not to be found in the current literature on spinors, so this exposition breaks some new ground. The ingredients have been available for a long time, and there is no reason why this story could not have been told thirty years ago. Indeed, before the geometrical theory of Dirac operators was developed at all, the eigenspinors for the Dirac operator on the sphere were considered (in 1938) by Schro¨dinger [47], who put his finger on the basic module property (7.10) of the spin connection (albeit not in so many words). A generation later, Newman and Penrose [43] introduced a family of functions on the sphere that they called “spinor harmonics”, which generalize the ordinary spherical harmonics and constitute the eigenspinors, as we shall see. In order to keep the development of the Dirac operator on S2 fairly self-contained, we begin by reviewing some elementary notions. Everything can be done with elementary calculus, provided one takes great care to get the signs and the constants right from the beginning. 9.1 Coordinates on the Riemann sphere The most direct way to reveal the Dirac operator on the sphere is to regard it as the Riemann sphere CP1, and use complex coordinates. Recall that CP1 may be described by homogeneous coordinates [z0 : z1] with z0, z1 not both zero, and is covered by two chart domains U0 := { [z0 : z1] : z0 6= 0 } and U1 := { [z0 : z1] : z1 6= 0 }. We shall use the local complex coordinates z := z1/z0 on U0, ζ := z 0/z1 on U1. The coordinate change is just z = ζ−1, ζ = z−1 on U0 ∩ U1. 102 We identify CP1 with the sphere S2 of unit vectors (u1, u2, u3) ∈ R3 by the stereographic projections : z = u1 + iu2 1− u3 on S 2 \ {N}, ζ = u 1 − iu2 1 + u3 on S2 \ {S} where N = (0, 0, 1) and S = (0, 0,−1) are the north and south poles. This identifies U0 with S2 \ {N} and U1 with S2 \ {S}. It is important to notice that both stereographic projections are orientation reversing. We may also use the standard spherical coordinates (θ, φ), satisfying u1 = sin θ cosφ, u2 = sin θ sinφ, u3 = cos θ. The stereographic projections are then expressed as z = eiφ cot θ 2 = eiφ sin θ 1− cos θ = e iφ1 + cos θ sin θ , ζ = e−iφ tan θ 2 = e−iφ sin θ 1 + cos θ = e−iφ 1− cos θ sin θ . (9.1) The positive functions (2.6) on U0 and U1 are given by Q0(z) := 1 + zz¯ = 2 1− cos θ , Q1(ζ) := 1 + ζζ¯ = 2 1 + cos θ . (9.2) Observe that Q0(z)/Q1(z −1) = zz¯. The local 1-forms dz, dz¯ on U0 and dζ, dζ¯ on U1 are given —recall (2.14)— by dz = eiφ 1− cos θ (−dθ + i sin θ dφ), dζ = e−iφ 1 + cos θ (dθ − i sin θ dφ), and their complex conjugates; notice that dζ = −e−2iφQ1/Q0 dz on U0 ∩U1. The basic local vector fields are given by ∂ ∂z = e−iφ 1− cos θ 2 ( − ∂ ∂θ − i sin θ ∂ ∂φ ) , ∂ ∂ζ = eiφ 1 + cos θ 2 ( ∂ ∂θ + i sin θ ∂ ∂φ ) , (9.3) and their complex conjugates; now ∂/∂ζ = −e+2iφQ0/Q1 ∂/∂z on U0 ∩ U1. The Riemannian metric g on S2 is the usual one: g = dθ2 + sin2 θ dφ2 = 4dz · dz¯(1 + zz¯)2 = 4dζ · dζ¯(1 + ζζ¯)2 (9.4) (although (2.13) differs from this by a factor of two). The Riemannian volume form is Ω = sin θ dθ ∧ dφ = −2i Q−20 dz ∧ dz¯ = −2i Q−21 dζ ∧ dζ¯, where the minus sign [compare with (2.12)] indicates the reversal of orientation in passing from (θ, φ) coordinates to (z, z¯) or (ζ, ζ¯) coordinates. 103 9.2 Sections and gauge transformations A section of a complex line bundle L−→ S2 is given by a pair of local sections over U0, U1 respectively. Once we have chosen basic sections s0 ∈ Γ(U0, L) and s1 ∈ Γ(U1, L) which are nonvanishing, any global section s ∈ Γ(L) is determined by a pair of smooth functions f0, f1 on C such that s(x) = f0(z) s0(x) for x ∈ U0, s(x) = f1(ζ) s1(x) for x ∈ U1, where z, ζ are the respective coordinates of the point x ∈ S2. The basic local sections are related by the transition function of the line bundle s0 = g01s1 (since the sphere is covered by only two charts, one transition function suffices), which implies a corresponding relation between f0(z) and f1(ζ), called a gauge transformation. In fine: a global section is determined by a related pair of functions, and the line bundle is identified by the particular gauge transformation relating them. We start with a brief mention of the holomorphic line bundles. The tautological line bundle L−→CP1 is defined, as in (5.12), by the fibres Lx := { (λz0, λz1) ∈ C2 : λ ∈ C } for x = [z0 : z1]; its basic sections are defined as s˜0(x) := (1, z) for x ∈ U0, s˜1(x) := (ζ, 1) for x ∈ U1. This yields s˜1(x) = z−1s˜0(x) for x ∈ U0 ∩ U1; the transition functions are therefore g˜10(z) = z −1, g˜01(ζ) = ζ−1, which of course are holomorphic on U0 ∩ U1. The hermitian metric on L is obtained from the natural inclusion of the fibres in C2; this means that (s˜0 | s˜0) = ‖(1, z)‖2 = 1 + zz¯ = Q0(z) and (s˜1 | s˜1) = ‖(ζ, 1)‖2 = 1 + ζζ¯ = Q1(ζ). Its dual, the hyperplane bundle H −→CP1 has local sections σ˜0, σ˜1 with (σ˜0 | σ˜0) = Q−10 on U0, (σ˜1 | σ˜1) = Q−11 on U1 —recall (5.13)— and so they extend smoothly to global sections on S2 just by setting σ˜0(N) := 0, σ˜1(S) := 0. The (holomorphic) transition functions are g˜01(ζ) = ζ and g˜10(z) = z. A global holomorphic section σ˜ ∈ O(CP1, H) is given by a pair of holomorphic functions f0, f1 with σ˜(x) = f0(z) σ˜0(x) and σ˜(x) = f1(ζ) σ˜1(x) for all x. That means that f0, f1 are entire functions whose possible singularities at infinity can only be poles, and f0(z) = zf1(z −1) for z ∈ C×; this relation identifies a Taylor series and a Laurent series, and can only hold if f0, f1 are of the form f0(z) = a + bz, f1(ζ) = b + aζ for some a, b ∈ C: we have once again established that O(CP1, H) ' C2. From now on, we shall normalize all basic sections and thus use only U(1)-valued tran- sition functions. Thus we replace the basic sections of L by s0 := Q0(z) −1/2s˜0, s1 := Q1(ζ)−1/2s˜1, and likewise σ0 := Q0(z) 1/2σ˜0, σ1 := Q0(ζ) 1/2σ˜1. This gives s1(x) = z −1 √ Q0(z) Q1(z−1) s0(x) = √ zz¯ z s0(x) = (z¯/z) 1/2s0(x), so the transition function is g10(z) = (z¯/z) 1/2 = e−iφ. A section s : CP1 → L is then given by a pair of functions (f0, f1) such that f0(z)s0(x) ≡ f1(ζ)s1(x) on U0 ∩ U1, i.e., such that f0(z) ≡ (z¯/z)1/2f1(z−1), f1(ζ) ≡ (ζ¯/ζ)1/2f0(ζ−1), (9.5) 104 where the second equation is of course redundant. The formula (9.5) exhibits the U(1) gauge transformation of the tautological line bundle. It is clear from (9.5) that f0 and f1 cannot both be holomorphic, and in general neither is holomorphic. Therefore, it would be more correct to write f0(z, z¯) and f1(ζ, ζ¯) to signal the dependence of these smooth functions on both real coordinates. We shall do so whenever the need arises. For the hyperplane bundle H −→CP1, the very same argument shows that a global section is given by a pair of functions (h0, h1) such that h0(z)σ0(x) ≡ h1(ζ)σ1(x) on U0 ∩U1, and which are therefore related by the gauge transformation h0(z) ≡ (z/z¯)1/2h1(z−1), h1(ζ) ≡ (ζ/ζ¯)1/2h0(ζ−1). (9.6) Exercise 9.1. Show that for any Hermitian line bundle E−→CP1, the sections in Γ(E) are described by pairs of functions (f0, f1) satisfying the relation f0(z) ≡ (z/z¯)k/2f1(z−1), where k ∈ Z and k[H] is the Chern class of E. We may decompose the complexified tangent bundle as TCS2 = T 1,0S2 ⊕ T 0,1S2, where the sections of the holomorphic tangent bundle T 1,0S2 are locally of the form f0(z, z¯) ∂/∂z or f1(ζ, ζ¯) ∂/∂ζ. [The sections of the “antiholomorphic tangent bundle” T 0,1S2 are of the form h0(z, z¯) ∂/∂z¯ or h1(ζ, ζ¯) ∂/∂ζ¯.] Then T 1,0S2−→ S2 is a Hermitian line bundle, under the metric determined by 〈∂/∂z | ∂/∂z〉 := g(∂/∂z¯, ∂/∂z) = 4 (1 + zz¯)2 . As normalized local sections we take Ez := 1 2 Q0(z) ∂ ∂z over U0, −Eζ := −12Q1(ζ) ∂ ∂ζ over U1. Since z = ζ−1 gives ∂/∂z = −ζ2 ∂/∂ζ, we find that Ez = −ζ2Q0(z)/Q1(ζ)Eζ = (z¯/z)(−Eζ). Thus a section of the holomorphic tangent bundle is given by a pair of functions f0, f1 satisfying f0(z)Ez ≡ f1(ζ)(−Eζ) on U0 ∩ U1, or equivalently f0(z) ≡ (z/z¯) f1(z−1). This establishes that [T 1,0S2] = 2[H] in Hˇ(S2,Z), i.e., the line bundles T 1,0S2 and H ⊗ H are equivalent. Exercise 9.2. Conclude from Ez = (z/z¯)(−Eζ) that the complex line bundle T 0,1S2 is equiv- alent to L⊗ L. Exercise 9.3. Write T ∗CS2 = Λ1,0T ∗S2⊕Λ0,1T ∗S2, where A1,0(S2) = Γ(Λ1,0T ∗S2) has elements f0(z, z¯) dz = f1(ζ, ζ¯) dζ and A 0,1(S2) = Γ(Λ0,1T ∗S2) consists of all h0(z, z¯) dz¯ = h1(ζ, ζ¯) dζ¯. Use g−1 to define Hermitian metrics on both these line bundles, write down suitable normal- ized local sections, and verify that Λ1,0T ∗S2 ∼ L⊗ L and Λ0,1T ∗S2 ∼ H ⊗H. 105 Definition 9.1. Let S−→ S2 denote the irreducible spinor bundle. Since T ∗CS2 ∼ S⊗S∗, we expect that S ∼ L⊕H, and that S+ ∼ L, S− ∼ H. Due to the noncanonical nature of the spinor bundle, we bypass the construction of basic local sections and define a spinor ψ over S2 directly as a pair of functions on each chart, denoted ψ±N(z, z¯) and ψ ± S (ζ, ζ¯) respectively, satisfying the gauge transformation rules: ψ+N(z, z¯) ≡ (z¯/z)1/2ψ+S (z−1, z¯−1), ψ−N(z, z¯) ≡ (z/z¯)1/2ψ−S (z−1, z¯−1), ψ+S (ζ, ζ¯) ≡ (ζ¯/ζ)1/2ψ+N(ζ−1, ζ¯−1), ψ−S (ζ, ζ¯) ≡ (ζ/ζ¯)1/2ψ−N(ζ−1, ζ¯−1). (9.7) It is immediate from (9.5) and (9.6) that (ψ+N , ψ + S ) determines a section of L and (ψ − N , ψ − S ) determines a section of H.1 9.3 The spin connection over the sphere Lemma 9.1. The Levi-Civita connection on the sphere is determined by the local formulae ∇∂i∂j = − 2 1 + x21 + x 2 2 (xi∂j + xj∂i − δijxk ∂k), (9.8) with x1 ≡ x1, x2 ≡ x2, where x1 + ix2 = z on U0 and x1 + ix2 = ζ on U1. Proof. The metric (9.4) is given by g = 4(1 + x21 + x 2 2) −2(dx21 + dx 2 2) on both charts, so gij = 4(1 + x 2 1 + x 2 2) −2δij and gij = 14(1 + x 2 1 + x 2 2) 2δij. Also, ∂kgij = −16xk(1 + x21 + x22)−3δij. From (7.13) we get at once Γkij = − 2 1 + x21 + x 2 2 (xiδ k j + xjδ k i − xkδij), from which (9.8) follows. Local orthonormal bases of vector fields (on both charts) are given by E1 := 1 2 (1 + x21 + x22)∂/∂x 1, E2 := 1 2 (1 + x21 + x 2 2)∂/∂x 2. With these bases, we get ∇∂iEα = xi∂α + 12(1 + x21 + x22)∇∂i∂α = 2 1 + x21 + x 2 2 (δiαx β Eβ − xαEi), or equivalently, Γ˜βiα = 2 1 + x21 + x 2 2 (δiαx β − δβi xα). The spin connection components are given by (7.18) as ωi = 1 4 Γ˜βiα γ αγβ, which yields ω1 = 1 2(1 + x21 + x 2 2) (xβγ1γβ − xαγαγ1) = x2 1 + x21 + x 2 2 γ1γ2, ω2 = 1 2(1 + x21 + x 2 2) (xβγ2γβ − xαγαγ2) = − x1 1 + x21 + x 2 2 γ1γ2, (9.9) 1We label the local spinor coefficients N and S (north and south) in order to reduce the clutter of numerical indices. 106 To return to complex notation, we notice that (ω1 + iω2)(z) = − iz1+zz¯γ1γ2 over U0 while (ω1 + iω2)(ζ) = − iζ1+ζζ¯γ1γ2 over U1. These are related by a gauge transformation: (ω1 + iω2)(z) = (z/z¯) (ω1 + iω2)(ζ). 9.4 The Dirac operator over the sphere Definition 9.2. We fix a local Clifford action on the spinor module by choosing a particular matrix function H˜ := [h˜rβ(z)] such that H˜ tH˜ = G−1 ≡ [gij(z)] = 1 4 (1 + zz¯)2 I over U0. Let us (arbitrarily) choose the positive square root H˜ := G−1/2 = 1 2 (1+zz¯) I. The Dirac operator on the chart U0 is then defined as D/N := c(dx j)∇S∂j = h˜jβ(z)γβ(∂j + ωj(z)) = 1 2 (1 + zz¯)(γβ∂β + γ βωβ(z)) = 1 2 (1 + zz¯)(γ1 ∂/∂x1 + γ2 ∂/∂x2)− 1 2 (x1γ 1 + x2γ 2). Here we have used the identities γ1γ1γ2 = −γ2 and γ2γ1γ2 = +γ1. It turns out that the Dirac operator over the chart U1, which we shall write as D/ S, is now completely determined: the Dirac operator on the spinor module is determined by its restriction to a single chart, no matter how small!2 Indeed, this is a characteristic feature of the spinor module [19]. However, in order to preserve the illusion of a symmetrical treatment of both charts, we shall anticipate the corresponding formulas for D/ S. We therefore choose the matrix [h˜rβ(ζ)] to be the negative square root of G −1 ≡ [gij(ζ)] = 1 4 (1 + ζζ¯)2 I, that is, h˜rβ(ζ) := −12(1 + ζζ¯)δrβ. This leads to D/ S = −12(1 + ζζ¯)(γ1 ∂/∂x1 + γ2 ∂/∂x2) + 12(x1γ1 + x2γ2). (9.10) Explicitly, ∂/∂x1 = ∂/∂z+∂/∂z¯ and ∂/∂x2 = i(∂/∂z−∂/∂z¯) on U0, and similar formulas hold on U1 with z replaced by ζ. This allows us to express the Dirac operators properly in complex coordinates: D/N = 1 2 (1 + zz¯) ( (γ1 + iγ2) ∂ ∂z + (γ1 − iγ2) ∂ ∂z¯ ) − 1 4 ( z(γ1 − iγ2) + z¯(γ1 + iγ2)), D/ S = −12(1 + ζζ¯) ( (γ1 + iγ2) ∂ ∂ζ + (γ1 − iγ2) ∂ ∂ζ¯ ) + 1 4 ( ζ(γ1 − iγ2) + ζ¯(γ1 + iγ2)). (9.11) These formulae are still a bit cumbersome; to obtain a simpler picture, we need to select a particular representation for the operators γ1 and γ2. 2On more general compact spin manifolds, the restriction of the Dirac operator to a single chart determines its restriction to neighbouring charts through its interaction with the gauge transformations; this in turn determine their neighbours, and so on. 107 Definition 9.3. The Fock space of R2 is the two-dimensional complex superspace Λ0C ⊕ Λ1C ' C⊕ C. We may represent γ1 and γ2 as anticommuting odd operators of square −1; a suitable choice is γ1 := ( 0 1 −1 0 ) , γ2 := ( 0 −i −i 0 ) . (9.12) The grading operator is iγ1γ2 = ( 1 0 0 −1 ) . Since γ1+iγ2 = ( 0 2 0 0 ) and γ1−iγ2 = ( 0 0 −2 0 ) , the expressions (9.11) simplify to D/N = ( 0 ðz −ð¯z 0 ) , D/ S = ( 0 −ðζ ð¯ζ 0 ) , (9.13) where ðz is the first-order differential operator3 ðz := (1 + zz¯) ∂ ∂z − 1 2 z¯. (9.14) The operator ðz was introduced by Newman and Penrose [43] and further studied by Goldberg et al. [30], with particular attention to its eigenfunctions. Since ðzψ = (1 + zz¯) ∂ψ ∂z − 1 2 ∂ ∂z (1 + zz¯)ψ = (1 + zz¯)3/2 ∂ ∂z ( (1 + zz¯)−1/2ψ ) , (9.15) we get the identity ðz = Q3/20 (∂/∂z)Q −1/2 0 in the algebra of differential operators on U0. The formal selfadjointness of D/ is not perhaps apparent from (9.13), since the term −1 2 z¯ in (9.14) seems to have the adjoint −1 2 z rather than +1 2 z. However, the inner product of spinors 〈〈φ |ψ〉〉 involves integration over the sphere, i.e., integration over C with respect to to the area form 2i(1 + zz¯)−2 dz ∧ dz¯. Thus, if φ, ψ are two spinors, then (9.15) gives (1 + zz¯)−2φ+Nðzψ − N = (1 + zz¯) −1/2φ+N ∂ ∂z ( (1 + zz¯)−1/2ψ−N ) , and it follows that 〈〈φ+ | ðzψ−〉〉 = −〈〈ð¯zφ+ | ψ−〉〉 on integrating by parts. Proposition 9.2. The restriction of the Dirac operator to U1 is determined by its restriction to U0. Proof. We must show that the form (9.13) of the operator D/ S is completely determined by that of D/N . This is possible because these operators act respectively on functions ψ ± S and ψ±N which are linked by the gauge transformations (9.7), and because the Dirac operator is odd, so the gauge transformations for both spinor parities must be invoked. 3The letter ð, from the Icelandic alphabet, is pronounced “edth”. 108 Given a spinor ψ with components ψ+ and ψ−, let φ = D/ψ. Then (ðzψ−N)(z, z¯) = φ+N(z, z¯) = (z¯/z) 1/2φ+S (z −1, z¯−1). On the other hand, with ζ = z−1 we get ∂ψ−N ∂z (z, z¯) = ∂ ∂z ( (z/z¯)1/2ψ−S (z −1, z¯−1) ) = 1 2 (zz¯)−1/2ψ−S (z −1, z¯−1)− (z/z¯)1/2z−2∂ψ − S ∂ζ (z−1, z¯−1) = (ζζ¯)1/2 ( −ζ ∂ψ − S ∂ζ (ζ, ζ¯) + 1 2 ψ−S (ζ, ζ¯) ) , and so the operator ðz transforms as follows: (ðzψ−N)(z, z¯) = (1 + zz¯) ∂ψ−N ∂z (z, z¯)− 1 2 z¯ψ−N(z, z¯) = (ζζ¯)−1/2(1 + ζζ¯) ( −ζ ∂ψ − S ∂ζ (ζ, ζ¯) + 1 2 ψ−S (ζ, ζ¯) ) − (ζ¯/ζ) 1/2 2ζ¯ ψ−S (ζ, ζ¯) = (ζ/ζ¯)1/2 ( −(1 + ζζ¯)∂ψ − S ∂ζ (ζ, ζ¯) + 1 2 (ζ−1 + ζ¯)ψ−S (ζ, ζ¯)− 12ζ−1ψ−S (ζ, ζ¯) ) = −(ζ/ζ¯)1/2ðζψ−S (ζ, ζ¯) = −(z¯/z)1/2ðζψ−S (z−1, z¯−1). (9.16) We conclude that φ+S = −ðζψ−S . If we apply complex conjugation to (9.16) and replace ψ−N by ψ+N and ψ − S by ψ + S , we find that φ−N(z, z¯) = −(ð¯zψ+N)(z, z¯) = (z/z¯)1/2ð¯ζψ+S (z−1, z¯−1), and it follows that φ−S = ð¯ζψ + S . We have recovered the expression of (9.13) for D/ S from the corresponding for D/N and from (9.7), without using the recipe (9.10) for D/ S. This means that the “choice” h˜ r β(ζ) := −1 2 (1 + ζζ¯)δrβ which led to (9.10) is actually forced by the action of D/ on spinors. Exercise 9.4. Use (9.1) and (9.3) to obtain expressions for ðz and ðζ in spherical coordinates (θ, φ). Check that eiφðz = −e−iφðζ − csc θ on U0 ∩ U1. Proposition 9.2 shows that the only freedom in the choice of the Dirac operator, once the metric and the spin connection are given, lies in selecting the matrix function H˜ = [h˜rβ(z)] which gives the local Clifford action on U0. The condition H˜ tH˜ = G−1 fixes the matrix H˜ up to premultiplication by an arbitrary SO(2) matrix function on U0. We may alternatively think of this as the freedom to select any local orthonormal bases of tangent vectors compatible with the orientation of S2, since this amounts to picking a section of the SO(2)-frame bundle. To express this in our complex-coordinate notation, we must bear in mind that the stereographic projections (θ, φ) 7→ (z, z¯) and (θ, φ) 7→ (ζ, ζ¯) are orientation reversing ; such a local orthonormal basis is therefore given by( E1 E2 ) := (− cosα sinα sinα cosα )( ∂/∂θ csc θ ∂/∂φ ) , (9.17) where α(θ, φ) is a smooth real-valued function which we call the spin gauge [25]. 109 Exercise 9.5. Check that Ez = 1 2 (E1 − iE2) if and only if α ≡ −φ, and −Eζ = 12(E1 − iE2) if and only if α ≡ φ. Exercise 9.6. Compute the mixed Christoffel symbols Γ˜βiα with the spin gauge (9.17) (use Exercise 7.16) and verify that Γ˜2θ1 = ∂α/∂θ, Γ˜ 2 φ1 = ∂α/∂φ− cos θ. Exercise 9.7. Use the results of the Exercises 9.5 and 9.6 to obtain the following local expres- sions in spherical coordinates for the spin connection ∇S on the chart domains U0 and U1: ∇Sθ = ∂ ∂θ , ∇Sφ = ∂ ∂φ ∓ 1 2 (1± cos θ)γ1γ2 (9.18) where the upper signs are for U0 and the lower signs are for U1. Exercise 9.8. Compute the Dirac operator in spherical coordinates with the spin gauge (9.17), using the results of Exercise 7.16: check that D/ is given by D/ − = eiα ( − ∂ ∂θ − i sin θ ∂ ∂φ + 1 2 sin θ (∂α ∂φ − cos θ ) − i 2 ∂α ∂θ ) , D/ + = e −iα ( ∂ ∂θ − i sin θ ∂ ∂φ − 1 2 sin θ (∂α ∂φ − cos θ ) − i 2 ∂α ∂θ ) , with the conventions of (8.7) and (9.12).4 9.5 The spinor Laplacian Lemma 9.3. The spinor Laplacian ∆S on the sphere S2 is given locally by ∆S = − ( ∂2 ∂θ2 + cot θ ∂ ∂θ + ∂2 ∂φ2 ) + ( 1± cos θ 2 sin θ )2 ± 1± cos θ sin2 θ γ1γ2 ∂ ∂φ , (9.19) where the upper signs are for U0 and the lower signs are for U1. Proof. It suffices to check that the general local formula (8.10) for the Laplacian specializes, in view of (7.14) and (9.18), to ∆S = − ∂ 2 ∂θ2 − 1 sin2 θ ( ∂ ∂φ ∓ 1 2 (1± cos θ)γ1γ2 )2 − 1 sin2 θ ( sin θ cos θ ∂ ∂θ ) , from which (9.19) is immediate. It is convenient to rewrite the spinor Laplacian in complex coordinates. From (9.3) we derive ∂ ∂φ = i ( z ∂ ∂z − z¯ ∂ ∂z¯ ) = −i ( ζ ∂ ∂ζ − ζ¯ ∂ ∂ζ¯ ) on U0 ∩ U1. 4The formulae (9.18) have been obtained by Dray [25] as the expressions of the Newman–Penrose operators −ð and ð¯ on quantities of “spin-weights” − 12 and + 12 respectively. 110 Using (9.2) and (9.3) it is readily checked that (1 + zz¯)2 ∂2 ∂z ∂z¯ = (1 + ζζ¯)2 ∂2 ∂ζ ∂ζ¯ = ∂2 ∂θ2 + cot θ ∂ ∂θ + ∂2 ∂φ2 . Since (1 + cos θ/ sin θ)2 = zz¯ and (1− cos θ/ sin θ)2 = ζζ¯ by (9.1), we arrive at ∆S = −(1 + zz¯)2 ∂ 2 ∂z ∂z¯ + 1 4 zz¯ + 1 2 (1 + zz¯) iγ1γ2 ( z ∂ ∂z − z¯ ∂ ∂z¯ ) over U0, = −(1 + ζζ¯)2 ∂ 2 ∂ζ ∂ζ¯ + 1 4 ζζ¯ + 1 2 (1 + ζζ¯) iγ1γ2 ( ζ ∂ ∂ζ − ζ¯ ∂ ∂ζ¯ ) over U1. (9.20) Lemma 9.4. The Dirac operator and the spinor Laplacian on S2 are related by D/ 2 = ∆S + 1 2 . Proof. This follows, of course, from the Lichnerowicz formula (8.14), since the sphere has constant scalar curvature K ≡ 2; but it is instructive to make a direct verification. From (9.14) we get −ðzð¯z = ( (1 + zz¯) ∂ ∂z − 1 2 z¯ )(−(1 + zz¯) ∂ ∂z¯ + 1 2 z ) = −(1 + zz¯)2 ∂ 2 ∂z ∂z¯ + 1 2 (1 + zz¯) ( z ∂ ∂z + z¯ ∂ ∂z¯ ) −1 4 zz¯ + (1 + zz¯) ( 1 2 − z¯ ∂ ∂z¯ ) = −(1 + zz¯)2 ∂ 2 ∂z ∂z¯ + 1 2 + 1 4 zz¯ + 1 2 (1 + zz¯) ( z ∂ ∂z − z¯ ∂ ∂z¯ ) , (9.21) whereas −ð¯zðz = (−(1 + zz¯) ∂ ∂z¯ + 1 2 z )( (1 + zz¯) ∂ ∂z − 1 2 z¯ ) = −(1 + zz¯)2 ∂ 2 ∂z ∂z¯ + 1 2 + 1 4 zz¯ − 1 2 (1 + zz¯) ( z ∂ ∂z − z¯ ∂ ∂z¯ ) (9.22) by a similar calculation. The different signs of the last terms on the right in (9.21) and (9.22) signify the presence of the grading operator iγ1γ2 in the representation (9.12). Thus D/ 2N = (−ðzð¯z 0 0 −ð¯zðz ) = −(1 + zz¯)2 ∂ 2 ∂z ∂z¯ + 1 2 + 1 4 zz¯ + 1 2 (1 + zz¯) iγ1γ2 ( z ∂ ∂z − z¯ ∂ ∂z¯ ) . (9.23) There is an identical formula for D/ S, on replacing z by ζ. Now a glance at (9.20) gives the Lichnerowicz formula D/ 2 = ∆S + 1 2 . 111 9.6 The SU(2) action on the spinor bundle Definition 9.4. The Lie group SU(2) of unitary matrices of determinant 1, g = ( α β −β¯ α¯ ) with αα¯ + ββ¯ = 1, acts transitively on the sphere S2 by rotations. The identification g ↔ (α, β) ∈ C2 shows that SU(2) is topologically the sphere S3, so it is compact. If [z0 : z1] are homogeneous coordinates of a point in CP1, the action is given by g · [z0 : z1] = [αz0 + βz1 : −β¯z0 + α¯z1]. On the charts U0 and U1, the action of SU(2) is given by Mo¨bius transformations : g · z = αz + β−β¯z + α¯ , g ′ · ζ = α¯ζ − β¯ βζ + α , (9.24) which clearly satisfies (g · z)−1 = g′ · ζ. Notice that g 7→ g′ is an (inner) automorphism of SU(2), since g′ := ( α¯ −β¯ β α ) = ( 0 i i 0 )( α β −β¯ α¯ )( 0 −i −i 0 ) . Exercise 9.9. Any isometry (rotation or reflection) of the sphere S2 takes circles to circles and takes antipodal pairs of points to antipodal pairs of points. Conversely, any smooth bijective transformation of S2 with these two geometrical properties is either a rotation or a reflection. Prove this converse, using the fact that a circle-preserving transformation of S2 corresponds, under stereographic projection, to a circle-preserving transformation of C∞, which is either a Mo¨bius transformation z 7→ (αz + β)/(γz + δ) or a conjugate Mo¨bius transformation z 7→ (αz¯ + β)/(γz¯ + δ); and that a Mo¨bius transformation is determined by the images of three points in C∞. Exercise 9.10. The antipode of z = eiφ cot 1 2 θ is ei(pi+φ) cot 1 2 (pi−θ) = −1/z¯. Use this fact and the preceding exercise to prove that any rotation of S2 is given by a Mo¨bius transformation of the form z 7→ (αz + β)/(−β¯z + α¯). Definition 9.5. The elements of SU(2) are conveniently described in terms of the Pauli matrices : σ1 := ( 0 1 1 0 ) , σ2 := ( 0 −i i 0 ) , σ3 := ( 1 0 0 −1 ) . If ~n = (n1, n2, n3) denote Cartesian coordinates in R3 of a point ~n ∈ S2, we write ~n · ~σ := n1σ1 + n2σ2 + n3σ3; then any g ∈ SU(2) may be written in the form g = exp ( 1 2 iψ ~n · ~σ) = cos 1 2 ψ + i sin 1 2 ψ ~n · ~σ (9.25) with ~n ∈ S2 and −pi < ψ ≤ pi. This is called the angle-axis parametrization of SU(2). We may identify the sphere S2 with the submanifold { g : Tr g = 0 } of SU(2), where ~n ∈ S2 corresponds to i~n · ~σ ∈ SU(2); the rotation action ρ of SU(2) on the sphere is given by conjugation: i~n · ~σ 7→ g(i~n · ~σ)g−1 =: iρ(g)~n · ~σ. The isotropy subgroup of the point ~n consists of elements of the form (9.25) with ψ arbitrary, which form a subgroup isomorphic to U(1). Thus S2 ≈ SU(2)/U(1) as a homogeneous space. 112 Exercise 9.11. Show that ρ(g)~n = ~n if and only if g is of the form (9.25) for some ψ ∈ R. Exercise 9.12. Show that there is a group isomorphism between Spin(3) and SU(2), such that if {e1, e2, e3} is an orthonormal basis of R3, then the elements e2e3, e3e1, e1e2 ∈ Spin(3) correspond respectively to iσ1, iσ2, iσ3 ∈ SU(2). The quotient mapping η : SU(2)→ S2 may be described more economically by regarding S2 as CP1 and proceeding as follows. Definition 9.6. The Hopf fibration η : SU(2)→ CP1 is the map given by η ( α β −β¯ α¯ ) := β α¯ . (9.26) It is immediate that ( α β −β¯ α¯ )( eiψ/2 0 0 e−iψ/2 ) η7−→ β α¯ , so SU(2) η−→CP1 is a principal U(1)-bundle, where the free right action of U(1) on the fibres η−1(β/α¯) is given simply by multiplication on the right by the diagonal elements of SU(2). Exercise 9.13. The Hopf fibration decomposes the sphere S3 ≈ SU(2) into a disjoint union of circles (the fibres), any two of which are linked. Regard S3 as R3unionmulti{∞} via the stereographic projection (α, β) 7→ (w, t) where w := α 1−=β ∈ C, t := <β 1−=β ∈ R. Check that χ(w, t) := (2t+ i(ww¯ + t2 − 1))/2w¯ ∈ C∞ is the expression in (w, t)-coordinates of the Hopf fibration. Deduce that the equation χ(w, t) = z represents the circle obtained by cutting the sphere ww¯+ t2− i(z¯w− zw¯) = 1 with the equatorial plane z¯w+ zw¯− 2t = 0; in particular, χ(w, t) = ∞ is the t-axis (including ∞) and χ(w, t) = 0 is the unit circle in the ww¯-plane. Show that any other circle χ(w, t) = z cuts the ww¯-plane obliquely at two points, one inside and one outside the unit circle; conclude that the circles χ(w, t) = 0 and χ(w, t) = z are linked. The group SU(2) acts on itself by left translations λ(g) : h 7→ gh; since these commute with right translations by U(1), they define a left action on the quotient manifold CP1, which is just the aforementioned rotation action. Indeed, if h = (1 + zz¯)−1 ( 1 z −z¯ 1 ) , then η(h) = z, and from 1 1 + zz¯ ( α β −β¯ α¯ )( 1 z −z¯ 1 ) = 1 1 + zz¯ ( α− βz¯ αz + β −β¯ − α¯z¯ −β¯z + α¯ ) it follows that η(gh) = g · η(h) ≡ ρ(g)η(h) for g, h ∈ SU(2). Schematically, we get a commutative diagram SU(2) λ(g)−−−→ SU(2) η y xη S2 ρ(g)−−−→ S2 113 which says that each pair of maps (λ(g), ρ(g)) is a morphism of the principal U(1)-bundle SU(2) η−→CP1. We may say that the group SU(2) acts “equivariantly” on the principal bundle. The corresponding type of group action on vector bundles is defined as follows. Definition 9.7. A homogeneous vector bundle with symmetry group G (a Lie group)5 is a vector bundle E−→M together with a pair of (left) actions τ : G×E → E and ρ : G×M → M such that each (τ(g), ρ(g)) is a vector bundle morphism on E−→M . We shall call such a pair a bundle action of G. A Hermitian homogeneous vector bundle is a Hermitian vector bundle with a bundle action for which each τ(g)x ∈ End(Ex) is unitary. If E = E+⊕E− is a superbundle, we say the bundle action of G is even if τ(g)x ∈ End+(Ex) for each x ∈M ; in other words, if both subbundles E±−→M are G-homogeneous under the bundle action (τ, ρ). We seek to define an even action of SU(2) on the spinor bundle S−→ S2. This can be pictured as a commutative diagram S± τ(g)−−−→ S± pi± y xpi± S2 ρ(g)−−−→ S2 which, in terms of the spinor components ψ± ∈ Γ(S±), means that τ(g)ψ±N(z, z¯) = A ± N(g, z)ψ ± N(g −1 · z, (g−1 · z)−), τ(g)ψ±S (ζ, ζ¯) = A ± S (g ′, ζ)ψ±S (g ′−1 · ζ, (g′−1 · ζ)−), (9.27) for g ∈ SU(2). By unitarity, the multipliers A± must be U(1)-valued functions; and they must satisfy the following consistency conditions in order that (9.27) define a group action: A±N(gh, z) = A ± N(g, z)A ± N(h, g −1 · z), (9.28) A±S (g ′h′, ζ) = A±S (g ′, ζ)A±S (h ′, g′−1 · ζ). (9.29) Exercise 9.14. Show that A(g, z) := (βz¯+ α¯)k/(β¯z+α)k is a formal solution to the equation A(gh, z) = A(g, z)A(h, g−1 · z), for g, h ∈ SU(2), z ∈ C∞; and that this solution is well- defined provided 2k is an integer. Bearing in mind that g−1 · z = α¯z − β β¯z + α , g′−1 · ζ = αζ + β¯−βζ + α¯ , the gauge transformation rule (9.7) for the spinors τ(g)ψ yields A+S (g ′, ζ)ψ+S ( αζ + β¯ −βζ + α¯ , α¯ζ¯ + β −β¯ζ¯ + α ) = (ζ¯/ζ)1/2A+N(g, ζ −1)ψ+N ( α¯− βζ β¯ + αζ , α− β¯ζ¯ β + α¯ζ¯ ) . (9.30) 5This is often called a G-vector bundle, for short. 114 The gauge transformation ψ+S ( αζ + β¯ −βζ + α¯ , α¯ζ¯ + β −β¯ζ¯ + α ) = ( α¯ζ¯ + β −β¯ζ¯ + α / αζ + β¯ −βζ + α¯ )1/2 ψ+N ( α¯− βζ β¯ + αζ , α− β¯ζ¯ β + α¯ζ¯ ) shows that the coefficients in (9.30) must satisfy the relation A+S (g ′, ζ) A+N(g, ζ −1) = ( ζ¯(αζ + β¯)(−β¯ζ¯ + α) ζ(−βζ + α¯)(α¯ζ¯ + β) )1/2 = ( (α + β¯ζ−1)(−β¯ζ¯ + α) (−βζ + α¯)(α¯ + βζ¯−1) )1/2 , or equivalently A+S (g ′, ζ) A+N(g, z) = (−β¯ζ¯ + α −βζ + α¯ )1/2 / ( βz¯ + α¯ β¯z + α )1/2 . (9.31) The following solution of (9.29) is therefore consistent with the spinor gauge transforma- tions: A+N(g, z) := ( βz¯ + α¯ β¯z + α )1/2 , A+S (g ′, ζ) := (−β¯ζ¯ + α −βζ + α¯ )1/2 . (9.32) Substituting these in (9.27) yields a bundle action of SU(2) on S+−→ S2. The same procedure leads to a bundle action of SU(2) on S−−→ S2. One need only replace the term (ζ¯/ζ)1/2 in (9.30) by (ζ/ζ¯)1/2 when invoking the gauge transformation rule (9.7); this leads to the choice of A−N(g, z) and A − S (g ′, ζ) as the complex conjugates of (9.32): A−N(g, z) := ( βz¯ + α¯ β¯z + α )−1/2 , A−S (g ′, ζ) := (−β¯ζ¯ + α −βζ + α¯ )−1/2 . (9.33) We summarize the foregoing in a definition. Definition 9.8. The Lie group SU(2) acts on the spinor bundle via (τ, ρ), where ρ is the rotation action (9.24) on the Riemann sphere, and τ is given by (9.27), where the multipliers A±N and A ± S are defined by (9.32) and (9.33). 9.7 Equivariance of the Dirac operator Lemma 9.5. Let T ∈ End+(Γ(U0, S)) be a transformation of the form (Tψ+N)(z, z¯) := a(z, z¯) 1/2ψ+N ( b(z), b(z) ) , (Tψ−N)(z, z¯) := a(z, z¯) −1/2ψ−N ( b(z), b(z) ) . (9.34) where a is a smooth function on C× and b is a rational function on C∞. Then Tðz = ðzT as operators from Γ(U0, S −) to Γ(U0, S+) if and only if a, b satisfy the pair of differential equations (1 + zz¯) db dz = a(z, z¯)(1 + b(z)b(z)), (1 + zz¯) ∂a ∂z = a(z, z¯)2b(z)− z¯a(z, z¯). (9.35) 115 Proof. It is enough to notice that T (ðzψ−N)(z, z¯) = a 1/2(1 + bb¯) ∂ψ−N ∂z (b, b¯)− 1 2 a1/2b¯ψ−N(b, b¯), whereas ðz(Tψ−N)(z, z¯) = (1 + zz¯) { −1 2 a−3/2 ∂a ∂z ψ−N(b, b¯) + a −1/2 db dz ∂ψ−N ∂z (b, b¯) } − 1 2 z¯a−1/2ψ−N(b, b¯), using both halves of (9.34); and then to equate coefficients of ψ−N and ∂ψ − N/∂z. Proposition 9.6. The Dirac operator on the sphere is equivariant under the action of SU(2) on the spinor bundle, i.e., τ(g)D/ = D/ τ(g) on Γ(S), for all g ∈ SU(2). Proof. To show that τ(g)D/ = D/ τ(g) for all g, it is enough to check this for g belonging to a collection of one-parameter subgroups which generate SU(2). Since αα¯ + ββ¯ = 1, we can write α = exp( i 2 φ+ i 2 ψ) cos 1 2 θ, β = exp( i 2 φ− i 2 ψ) sin 1 2 θ; we thereby see that any g ∈ SU(2) is of the form k(φ)h(θ)k(ψ), where6 k(φ) = ( eiφ/2 0 0 e−iφ/2 ) , h(θ) = ( cos 1 2 θ sin 1 2 θ − sin 1 2 θ cos 1 2 θ ) . (9.36) Now τ(k(t)) is of the form (9.34) with a(z, z¯) = e−it, b(z) = e−itz, so the equations (9.35) reduce to the identities (1 + zz¯)e−it = e−it(1 + zz¯), 0 = e−2iteitz¯ − z¯e−it. On the other hand, τ(h(t)) is of the form (9.34) with a(z, z¯) = z¯ sin 1 2 t+ cos 1 2 t z sin 1 2 t+ cos 1 2 t , b(z) = z cos 1 2 t− sin 1 2 t z sin 1 2 t+ cos 1 2 t . (9.37) for which ∂a ∂z = − sin 1 2 t z¯ sin 1 2 t+ cos 1 2 t (z sin 1 2 t+ cos 1 2 t)2 , db dz = 1 (z sin 1 2 t+ cos 1 2 t)2 . (9.38) From this it is easy to check that (9.37) satisfies the equations (9.35) for all t ∈ R. Thus τ(g)ðzψ−N = ðzτ(g)ψ − N for all g ∈ SU(2). By applying complex conjugation, we obtain ð¯zτ(g−1) = τ(g−1)ð¯z on functions ψ+N , and so τ(g)D/N = D/Nτ(g) for all g. Replacing z by ζ and g by g′ and adjusting a few signs, the same calculations show that τ(g)D/ S = D/ Sτ(g) for all g, as expected. Exercise 9.15. Verify directly that (9.37) satisfies the equations (9.35). 6The parameters (φ, θ, ψ) in this product are the so-called Euler angles for the group SU(2). 116 9.8 Angular momentum operators Definition 9.9. The homomorphisms t 7→ g(~n; t) := exp(1 2 it ~n · ~σ) = cos 1 2 t + i sin 1 2 t ~n · ~σ, for each ~n ∈ S2, yield all one-parameter subgroups of SU(2). The infinitesimal generators of these subgroups are − i 2 ~n · ~σ ∈ su(2), where su(2) is the Lie algebra of antihermitian 2 × 2 matrices. The corresponding generator J~n of the spinor action of this subgroup is −i(J~nψ±N)(z, z¯) := d dt ∣∣∣∣ t=0 τ(g(~n; t))ψ±N(z, z¯) = d dt ∣∣∣∣ t=0 at(z, z¯) ±1/2ψ±N ( bt(z), bt(z) ) , (9.39) where the coefficient (−i) is inserted for convenience, so that J~n is formally selfadjoint (rather than skewadjoint). For the three cardinal directions, where ~n · ~σ = n1σ1 + n2σ2 + n3σ3, we write the gen- erators simply as J1, J2, J3 respectively; these are commonly called the angular momentum generators.7 As before, we write J± := J1 ± iJ2. With the notations a˙0(z, z¯) := d dt ∣∣ t=0 at(z, z¯) and b˙0(z) := d dt ∣∣ t=0 bt(z), the definition (9.39) simplifies to J~n = −i ( b˙0(z) ∂ ∂z + b˙0(z) ∂ ∂z¯ ) + 1 2 a˙0(z, z¯) γ 1γ2 (9.40) on recalling that iγ1γ2 = ±1 on Γ(S±). Thus, for the one-parameter subgroup { k(−t) : t ∈ R }, where at(z, z¯) = eit and bt(z) = eitz, we get a˙0(z, z¯) = i, b˙0(z) = iz, and so J3 = z ∂ ∂z − z¯ ∂ ∂z¯ + 1 2 iγ1γ2 = −i ∂ ∂φ + 1 2 iγ1γ2. (9.41) For the one-parameter subgroup {h(−t) : t ∈ R }, at(z, z¯) and bt(z) are given by (9.37) with t replaced by −t. In this case a˙0(z, z¯) = 12(z − z¯) and b˙0(z) = 12(z2 + 1), leading to J2 = i 2 (z2 + 1) ∂ ∂z + i 2 (z¯2 + 1) ∂ ∂z¯ − 1 4 (z − z¯) γ1γ2. The one-parameter subgroup { cos 1 2 t+ i sin 1 2 t σ1 : t ∈ R } is generated by i2σ1, and at(z, z¯), bt(z) are now given by at(z, z¯) = cos 1 2 t+ iz¯ sin 1 2 t cos 1 2 t− iz sin 1 2 t , bt(z) = z cos 1 2 t− i sin 1 2 t cos 1 2 t− iz sin 1 2 t , for which a˙0(z, z¯) = i 2 (z¯ + z) and b˙0(z) = i 2 (z2 − 1). Therefore J1 = −12(z2 − 1) ∂ ∂z + 1 2 (z¯2 − 1) ∂ ∂z¯ − i 4 (z + z¯) γ1γ2. 7The suitability of interpreting these generators as angular momentum operators for a magnetic monopole is discussed at length in [10]. 117 The operators J± are therefore given by: J+ = J1 + iJ2 = −z2 ∂ ∂z − ∂ ∂z¯ − 1 2 z iγ1γ2, J− = J1 − iJ2 = ∂ ∂z + z¯2 ∂ ∂z¯ − 1 2 z¯ iγ1γ2. (9.42) It follows from this and from (9.41) that [J+, J−] = 2J3. One also obtains that J21 + J 2 2 = J−J+ + 12 [J+, J−] = J−J+ + J3. Exercise 9.16. Show that the operators J± are given in spherical coordinates by J± = e±iφ ( ± ∂ ∂θ + i cot θ ∂ ∂φ − 1 + cos θ 2 sin θ iγ1γ2 ) . These operators arise in the theory of the magnetic monopole [10, 25] when the monopole parameter8 µ = eg/~c takes the value µ = 1 2 . Exercise 9.17. Show that over U1 the angular momentum generators satisfy the analogue of (9.40), with z replaced by ζ. Verify that J3 = −ζ ∂ ∂ζ + ζ¯ ∂ ∂ζ¯ − 1 2 iγ1γ2, J+ = ∂ ∂ζ + ζ¯2 ∂ ∂ζ¯ − 1 2 ζ¯ iγ1γ2, J− = −ζ2 ∂ ∂ζ − ∂ ∂ζ¯ − 1 2 ζ iγ1γ2, (9.43) in the coordinates (ζ, ζ¯) over U1. There is one more operator worthy of mention, namely that corresponding to the “Casimir element” X21 +X 2 2 +X 2 3 , where {X1, X2, X3} is an orthonormal basis for the Lie algebra su(2). [The Casimir element belongs to the enveloping algebra U(su(2)).] The image of this element under a representation of the Lie algebra is called a “Casimir operator”. Definition 9.10. The Casimir operator for the spinor bundle action of SU(2) is the operator defined by C := J21 + J 2 2 + J 2 3 = J−J+ + J3(J3 + 1) on Γ(S). 9 Proposition 9.7. The Casimir operator satisfies the relations C = ∆S + 1 4 = D/ 2 − 1 4 . (9.44) 8The constants are c, the speed of light; ~, Planck’s constant; e, the electric charge of a particle whose total angular momentum is J ; and g, the magnetic charge of the monopole. The condition that 2µ be an integer is the quantization condition of Dirac [24], and arises from the necessary description of monopoles by complex line bundles over the sphere. 9Strictly speaking the Casimir operator should be −C, on account of the factor (−i) in (9.39); but we change the sign to obtain a positive operator. On Γ(S), it is formally selfadjoint. 118 Proof. Using (9.41) and (9.42), we compute that C = J−J+ + J3(J3 + 1) = −(1 + zz¯)2 ∂ 2 ∂z ∂z¯ + 1 4 (1 + zz¯) + 1 2 (1 + zz¯) iγ1γ2 ( z ∂ ∂z − z¯ ∂ ∂z¯ ) over U0, with an analogous formula over U1. A glance at (9.20) and (9.23) is enough to verify (9.44). This result shows that even in the simplest example of a spinor bundle over a compact manifold, the three operators commonly referred to as “the Laplacian” are distinct, and must be carefully distinguished. Though they only differ by constants, this has the important consequence that their spectra are not the same. A similar shifting of the Laplacian occurs in harmonic analysis on compact Lie groups [59], where the Casimir satisfies C = ∆+ 1 12 dimG. For SU(2), a 3-dimensional group, this leads us to expect C = ∆ + 1 4 , which is nicely confirmed by Proposition 9.7. 9.9 Spinor harmonics We come, finally, to the matter of diagonalizing the Dirac operator by finding an explicit basis of eigenspinors for D/ . In view of the SU(2) symmetry, we may suspect, by analogy with the diagonalization of the Hodge–Dirac operator in Section 4, that the members of this basis should be closely related to the spherical harmonics Ylm(θ, φ) on S2. We may also anticipate that at some point we shall need to use some heavy artillery from the representation theory of SU(2). However, we begin in a fairly pedestrian manner, with some polynomial calculations over the chart U0. Lemma 9.8. The identity ðz ( (1 + zz¯)−lzr(−z¯)s)= (1 + zz¯)−l((l + 1 2 − r)zr(−z¯)s+1 + rzr−1(−z¯)s) holds for all r, s ∈ N and l ∈ R. Exercise 9.18. Prove Lemma 9.8, and show also that the conjugate identity holds: −ð¯z ( (1 + zz¯)−lzr(−z¯)s)= (1 + zz¯)−l((l + 1 2 − s)zr+1(−z¯)s + szr(−z¯)s−1). These calculations show one method of finding eigenspinors: take for ψ−N a linear combi- nation of several terms of the form (1+zz¯)−lzr(−z¯)s, with a common value for the difference of exponents (r−s), and choose the coefficients cleverly enough that the result of applying ðz closely resembles a multiple of the original function. However, since we wish these functions to be components of spinors, we must first consider the effect of the gauge transformation rules. Lemma 9.9. Let φ : C→ C be a smooth function of the form φ(z, z¯) := (1 + zz¯)−l ∑ r,s∈N a(r, s)zr(−z¯)s. 119 Then φ represents a section in Γ(U0, S ±) if and only if l + 1 2 is a positive integer, and a(r, s) = 0 for r > l ∓ 1 2 or s > l ± 1 2 . Moreover, the coefficients must satisfy the symmetry relations a(r, s) = (−1)l± 12a(l ∓ 1 2 − r, l ± 1 2 − s). Proof. Suppose that φ represents a section in Γ(U0, S +). Then by (9.7) we obtain φ(z, z¯) = (z¯/z)1/2φ(z−1, z¯−1) = z−1/2z¯1/2(zz¯)l(1 + zz¯)−l ∑ r,s∈N a(r, s)z−r(−z¯)−s = (1 + zz¯)−l(−1)l+ 12 ∑ r,s∈N a(r, s)zl− 1 2 −r(−z¯)l+ 12−s, where the exponents in the sum on the right hand side must also be nonnegative integers. Thus l − 1 2 ∈ N, and the nonnegativity of the exponents on the right guarantees that r ∈ {0, 1, . . . , l − 1 2 } while s ∈ {0, 1, . . . , l + 1 2 }. The argument for sections in Γ(U0, S−) is similar. The structure of the symmetry relations among the coefficients, and the allowed ranges of the exponents, suggests the introduction of the following spinors. Definition 9.11. For each l ∈ N+ 1 2 = {1 2 , 3 2 , 5 2 , . . . }, and for each m ∈ {−l,−l + 1, . . . , l − 1, l},10 let Y ′lm ∈ Γ(S) be the spinors whose components over U0 are 2−1/2Y ±lm(z, z¯), where Y +lm(z, z¯) := Clm(1 + zz¯) −l ∑ r−s=m− 1 2 ( l − 1 2 r )( l + 1 2 s ) zr(−z¯)s, Y −lm(z, z¯) := Clm(1 + zz¯) −l ∑ r−s=m+ 1 2 ( l + 1 2 r )( l − 1 2 s ) zr(−z¯)s, (9.45) where the constants Clm are defined as 11 Clm := (−1)l−m √ 2l + 1 4pi √ (l +m)! (l −m)! (l + 1 2 )! (l − 1 2 )! . (9.46) Also, let Y ′′lm ∈ Γ(S) be the spinor whose components are 2−1/2Y +lm(z, z¯) and −2−1/2Y −lm(z, z¯). The simplest examples of (9.45) are Y ′1 2 , 1 2 := 1√ 4pi ( 1/ √ 1 + zz¯ z/ √ 1 + zz¯ ) , Y ′1 2 ,− 1 2 := 1√ 4pi ( z¯/ √ 1 + zz¯ −1/√1 + zz¯ ) . 10The range of allowed values of m is obtained by listing the possibilities for (r−s) in (9.45) and adjusting by 12 . Notice that each m is a half-integer. 11The precise form of the constants Clm is obtained by looking ahead to the normalization 〈〈Y ′lm |Y ′lm〉〉 = 1; for the present, we need only that the constants for Y +lm and Y − lm be the same. 120 The functions Y +lm and Y − lm were introduced by Newman and Penrose [43], using the notations − 1 2 Ylm and 1 2 Ylm respectively. In fact, these appear as a subfamily of functions sYlm with s ∈ {−l,−l+1, . . . , l−1, l} which, for l,m, s integers, they called “spin-s spherical harmonics”; for s = 0 they reduce to the everyday spherical harmonics Ylm on S2. Newman and Penrose also noted that their formulas make sense when l,m, s are all half-integers, and christened such functions “spinor harmonics”; they were investigated further by Goldberg et al [30]. Later, Dray [25] showed that these same functions occur as the spinor components in the theory of the magnetic monopole [10]. Lemma 9.10. ðzY −lm = (l + 1 2 )Y +lm and −ð¯zY +lm = (l + 12)Y −lm. Proof. Using Lemma 9.8, we find that ðzY −lm(z, z¯) equals Clm(1 + zz¯) −l ∑ r−s=m+ 1 2 ( l + 1 2 r )( l − 1 2 s ) {(l + 1 2 − r)zr(−z¯)s+1 + rzr−1(−z¯)s} = Clm(1 + zz¯) −l ∑ j−k=m− 1 2 { (l + 1 2 − j) ( l + 1 2 j )( l − 1 2 k − 1 ) + (j + 1) ( l + 1 2 j + 1 )( l − 1 2 k )} zj(−z¯)k. (9.47) The term in braces can be simplified, using the binomial identities k ( r k ) = r ( r − 1 k − 1 ) , (r − k) ( r k ) = r ( r − 1 k ) , to the form (l + 1 2 ) ( l − 1 2 j )( l − 1 2 k − 1 ) + (l + 1 2 ) ( l − 1 2 j )( l − 1 2 k ) = (l + 1 2 ) ( l − 1 2 j )( l + 1 2 k ) , so the right hand side of (9.47) equals (l + 1 2 )Y +lm. Corollary 9.11. The spinors Y ′lm and Y ′′ lm are eigenspinors for the Dirac operator, with nonzero integer eigenvalues ±(l + 1 2 ): D/Y ′lm = (l + 1 2 )Y ′lm, D/ Y ′′ lm = −(l + 12)Y ′′lm, and each eigenvalue ±(l + 1 2 ) has multiplicity (2l + 1). Proof. Just observe that( 0 ðz −ð¯z 0 )( Y +lm Y −lm ) = ( (l + 1 2 )Y +lm (l + 1 2 )Y −lm ) , ( 0 ðz −ð¯z 0 )( Y +lm −Y −lm ) = (−(l + 1 2 )Y +lm (l + 1 2 )Y −lm ) . The multiplicity is just the number of possibilities for the index m, i.e., the (2l+ 1) elements of {−l,−l + 1, . . . , l − 1, l}. Exercise 9.19. For a fixed l ∈ N+ 1 2 , express J3Y ′ lm, J+Y ′ lm and J−Y ′ lm as linear combinations of the spinors Y ′ln with n ∈ {−l,−l + 1, . . . , l − 1, l}. 121 9.10 The spectrum of the Dirac operator Definition 9.12. A representative function for the group SU(2) is a function in L2(SU(2)) which may appear as a matrix element in some finite-dimensional unitary representation of the group. It is a linear combination of the functions Djmn, indexed by j ∈ 12N ={0, 1 2 , 1, 3 2 , 2, . . . } and m,n ∈ {−j,−j + 1, . . . , j − 1, j}, which are defined in terms of the Euler-angle presentation g = k(α)h(β)k(γ) ∈ SU(2) by Djmn(α, β, γ) := √ (j + n)! (j − n)! (j +m)! (j −m)! e i(nα+mγ)(sin 1 2 β)2j × ∑ r (−1)j+m−r ( j +m r )( j −m r −m− n ) (cot 1 2 β)2r−m−n. (9.48) The Hilbert space L2(SU(2)) is described by defining the Haar measure on SU(2) in terms of the Euler angles as dg = (16pi2)−1 sin β dα dβ dγ, and the Parseval–Plancherel formula [37] shows that the functions Djmn form an orthogonal basis for this Hilbert space:∫ SU(2) |h(g)|2 dg = ∞∑ 2j=0 (2j + 1) j∑ m,n=−j ∣∣(Djmn | h)∣∣2. On comparing the definitions (9.45), (9.46) of the functions Y ±lm(z, z¯) with (9.48), we see from z = eiφ cot 1 2 θ that Y +lm(z, z¯) = √ 2l + 1 4pi Dl− 1 2 ,m (φ, θ, φ), Y −lm(z, z¯) = √ 2l + 1 4pi Dl1 2 ,m (φ, θ, φ). Exercise 9.20. Show that, with ζ = e−iφ tan 1 2 θ, the formulae Y +lm(ζ, ζ¯) = √ 2l + 1 4pi Dl− 1 2 ,m (φ, θ,−φ), Y −lm(ζ, ζ¯) = √ 2l + 1 4pi Dl1 2 ,m (φ, θ,−φ) express Y ±lm in terms of the SU(2) representative functions over U1. We now (at last!) fix the normalization of the inner product of spinors by ‖ψ‖2 = 〈〈ψ | ψ〉〉 := 1 4pi ∫ S2 (ψ+ψ+ + ψ−ψ−) sin θ dθ ∧ dφ. (9.49) Proposition 9.12. The spinors {Y ′lm, Y ′′lm : l ∈ N+ 12 ,m ∈ {−l, . . . , l} } form an orthonormal basis for the Hilbert space L2(S). Proof. We associate to each spinor ψ ∈ Γ(S) a pair of functions h± on SU(2) by h±(φ, θ, ψ) := √ 4pie±i(φ−ψ)/2 ψ±N(z, z¯) = √ 4pie∓i(φ+ψ)/2 ψ±S (ζ, ζ¯). 122 By integration over the ψ variable, we see that h+ is orthogonal to Djmn unless m = −12 , and h− is orthogonal to Djmn unless m = + 1 2 . The Parseval–Plancherel formula then shows that ‖ψ‖2 = 1 4pi ∫ SU(2) (|h+(g)|2 + |h−(g)|2) dg = ∑ l,m (2l + 1) 4pi (∣∣(Dl− 1 2 ,m | h+)∣∣2 + ∣∣(Dl1 2 ,m | h−)∣∣2) = ∑ l,m ∣∣∣∣ 14pi ∫ S2 Y +lmψ + Ω ∣∣∣∣2 + ∣∣∣∣ 14pi ∫ S2 Y −lmψ −Ω ∣∣∣∣2 = ∑ l,m ∣∣〈〈Y ′lm | ψ〉〉∣∣2 + ∣∣〈〈Y ′′lm | ψ〉〉∣∣2. (9.50) This is a Parseval identity for the orthonormal family {Y ′lm, Y ′′lm}, and so establishes com- pleteness of this family in L2(S). Corollary 9.13. The spectra of the Dirac operator, its square, the Casimir operator and the spinor Laplacian are given by sp(D/ ) = {±(l + 1 2 ) : l ∈ N+ 1 2 } = Z \ {0}, sp(D/ 2) = { (l + 1 2 )2 : l ∈ N+ 1 2 }, sp(C) = { l(l + 1) : l ∈ N+ 1 2 }, sp(∆S) = { l2 + l − 1 4 : l ∈ N+ 1 2 }. The respective multiplicities are: 2l + 1 for the eigenvalue ±(l + 1 2 ) of D/ , and 2(2l + 1) for each listed eigenvalue of D/ 2, C and ∆S. Proof. The eigenvalues of D/ are those given by Corollary 9.11; the completeness relation (9.50) shows that there are no others. The eigenvalues of C and ∆S follow from (9.43). Notice that the Casimir eigenvalues have the form l(l + 1) (compare the spectrum of the Hodge Laplacian), but with l half-integral in the present case. We recall from (4.16) the definition of the index of (any) Dirac operator: indD/ := dim(kerD/ +)− dim(kerD/ −), We end with an important result. Corollary 9.14. For the spinor module over S2, the index of the Dirac operator is zero. The Atiyah–Singer index theorem [9, 28, 39, 42] asserts the existence of a characteristic class whose integral coincides with the index of the Dirac operator. In fact, the characteristic form is given by Aˆ(R) := det−1/2(j(R)), where R is the Riemannian curvature and j(x) := (sinh 1 2 x)/1 2 x. Since the power series j(x) is even, however, only forms of degree 4k appear in 123 the expansion of Aˆ(R) by degrees; in particular, the second-degree component is zero. Thus for S2 or indeed any compact two-dimensional manifold M , the integral (2pii)−1 ∫ M Aˆ(R) (called the “Aˆ-genus of M”) vanishes, and on that basis the vanishing of the index of the Dirac operator is to be expected. To get a Dirac operator with a nontrivial kernel, one only has to twist the irreducible spinor module with a complex line bundle, such as H or L. In that context, the eigenspinors (9.45) yield explicit solutions of the Seiberg–Witten equations [60]. 10 Construction of representations of SU(2) The theory of the Dirac operator on the Riemann sphere, developed in the previous chapter, has a direct application to the construction of the irreducible unitary representations of the group SU(2). There are three main routes to that goal. In view of the Peter–Weyl theorem, an irreducible representation of a compact Lie group is finite-dimensional, unitarizable (that is, the representation space may be provided with an inner product that is invariant under the group action) and lies within the left regular representation of the group. The first route to the irreducible unitary representations is an algebraic construction called the “theorem of the highest weight” [13, 35, 37] which yields an abstract description of all such representa- tions up to unitary equivalence, but does not exhibit them concretely. Secondly, the famous Borel–Weil theorem [11, 50] constructs the representation spaces as spaces of holomorphic sections of certain line bundles over coadjoint orbits of the compact Lie group in question; this construction depends on the fact that these orbits are complex manifolds, and therefore carry a spinc structure. It turns out that the maximal-dimensional coadjoint orbits in fact carry a spin structure, compatible with the group action, and this opens a third path to the construction of representations, whereby the representation spaces are the kernel spaces of certain Dirac operators. (Indeed, this last recipe is related to the Borel–Weil–Bott construc- tion by a “twisting” with a certain line bundle, so the second and third routes are thereby equivalent.) Here we shall not attempt to give the construction for all compact groups, but rather shall build up the particular example of SU(2), that contains all the features of the general theory. 10.1 Characters of the maximal torus Any compact connected Lie group G contains a maximal torus T , and any two such tori are conjugate, by a basic theorem of Weyl [13]. For G = SU(2), we may take T to be the subgroup of diagonal unitary matrices k(φ) of (9.36), so that T ' U(1). Its characters are χm(k(φ)) = χm ( eiφ/2 0 0 e−iφ/2 ) = eimφ/2 (10.1) where m is necessarily an integer in order that χm be a well-defined homomorphism from T to U(1). The Hopf map η of (9.26) allows us to regard S2 as the homogeneous space SU(2)/T , whereby SU(2)−→ S2 becomes a principal U(1)-bundle. 124 Let x ∈ S2; if it is not the north or south pole, let ζ = z−1 be its local coordi- nates. Two particular local sections of this principal bundle are γ0 ∈ Γ(U0, SU(2)) and γ1 ∈ Γ(U1, SU(2)), given by γ0(x) := 1√ 1 + zz¯ ( z¯ 1 −1 z ) , γ1(x) := 1√ 1 + ζζ¯ ( 1 ζ −ζ¯ 1 ) . It is immediate that η(γ0(x)) = z −1 = ζ = η(γ1(x)) for x ∈ U0∩U1, which verifies that these are sections and implies that γ0(x) = γ1(x)h(ζ) with h(ζ) ∈ T ; a quick calculation yields h(ζ) = (√ ζ/ζ¯ 0 0 √ ζ¯/ζ ) , so that χm(h(ζ)) = (ζ/ζ¯) m/2. Every complex line bundle over S2 may be constructed as an associated bundle to this principal bundle via a suitable character of the maximal torus T . This can be seen directly by exhibiting these associated bundles, since we have already classified all line bundles over S2 in Proposition 5.13. However, it should be said that one can take a more principled point of view. On each such line bundle, we shall eventually construct an equivariant bundle action of SU(2); the equivalence classes of such actions generate a commutative semiring (under the Whitney sum and tensor product of equivariant vector bundles), and by formal subtraction (the Grothendieck construction) they generate an abelian group, denoted KSU(2)(S2) [48]. The characters of the maximal torus T form a ring R(T ), isomorphic to Z, and the matching of equivariant line bundles and characters yields a canonical isomorphism KSU(2)(S2) ' R(T ). The line bundle Lm−→ S2 associated to SU(2)−→ S2 via the character χm of T is given by (1.2): the space Lm has elements [g, v] with g ∈ SU(2), v ∈ C, and [gh, v] = [g, χm(h)v] whenever h ∈ T . The local sections s0 ∈ Γ(U0, Lm) and s1 ∈ Γ(U1, Lm) defined by sj(x) := [γj(x), 1] suffice to determine the class of L m. Indeed, s0(x) = [γ1(x)h(ζ), 1] = [γ1(x), χm(h(ζ))] = χm(h(ζ))s1(x) = (ζ/ζ¯) m/2s1(x), so the transition function of Lm is g01(ζ) = (ζ/ζ¯) m/2 = e−imφ/2. As before, we represent a general section s : S2 → Lm by a pair of functions (f0, f1) satisfying f0(z)s0(x) = f1(ζ)s1(x) on U0 ∩ U1. We therefore get the gauge transformation rule f1(ζ) ≡ (ζ¯/ζ)m/2f0(ζ−1). From Exercise 9.1 (compare also (9.5) and (9.6), which are particular cases), we obtain that the Chern class of Lm is m[L] = −m[H], so that Lm is indeed equivalent to the m-th tensor power of the tautological line bundle, as the notation had anticipated.1 1The isomorphism KSU(2)(S2) ' R(T ) is thus given by m[L] 7→ χm. 125 10.2 Twisting connections In subsection 9.3 we have seen that the spin connection ∇S is defined over U1 (say) by ∇S∂i = ∂i + ωi, where ζ = x1 + ix2 and ω1, ω2 are given by (9.9). Since ∂/∂ζ = 12(∂1 − i∂2), ∂/∂ζ¯ = 1 2 (∂1 + i∂2), we may rewrite (9.9) as 2 ∇S∂/∂ζ = ∂ ∂ζ + iζ¯ 2(1 + ζζ¯) γ1γ2, ∇S∂/∂ζ¯ = ∂ ∂ζ¯ − iζ 2(1 + ζζ¯) γ1γ2. If we use the presentation (9.12) of γ1 and γ2, in which the grading operator iγ1γ2 is a diagonal matrix, these become ∇S±∂/∂ζ = ∂ ∂ζ ± ζ¯ 2(1 + ζζ¯) , ∇S±∂/∂ζ¯ = ∂ ∂ζ¯ ∓ ζ 2(1 + ζζ¯) . (10.2) Now S+ = L, the tautological bundle over S2. If f1, . . . , fm, with m ∈ N, are local sections in Γ(U1, L), then their product is a local section in Γ(U1, L m). Denote by ∇(m) the tensor product of m copies of ∇S+ ; this is a connection on the vector bundle Lm−→ S2, for which ∇(m)X (f1 . . . fm) = m∑ j=1 f1 . . . fj−1∇S±X (fj) fj+1 . . . fm, from which it follows that ∇(m)∂/∂ζ = ∂ ∂ζ + mζ¯ 2(1 + ζζ¯) , ∇(m) ∂/∂ζ¯ = ∂ ∂ζ¯ − mζ 2(1 + ζζ¯) (10.3) on Γ(U1, L m). Exercise 10.1. If m is a negative integer, let ∇(m) be the tensor product of (−m) copies of ∇S− , i.e., the connection on Lm = H−m dual to the connection ∇(−m) on L−m; show that (10.3) holds also in this case. Exercise 10.2. The curvature ωm of the connection ∇(m) is determined by the relation ωm(∂/∂ζ, ∂/∂ζ¯) = [∇(m)∂/∂ζ ,∇(m)∂/∂ζ¯ ]. Check that ωm = −m(1 + ζζ¯)−2 dζ ∧ dζ¯. Definition 10.1. Let S−→ S2 denote the irreducible spinor bundle; let Lm−→ S2 be the complex line bundle with first Chern class m[L]; we call S⊗Lm−→ S2 a twisted spinor bundle. Clearly S ⊗ Lm ∼ Lm+1 ⊕ Lm−1. The twisted spin connection ∇˜ := ∇S ⊗ 1 + 1 ⊗∇(m) is determined, in view of (10.2) and (10.3), by ∇˜±∂/∂ζ = ∂ ∂ζ + (m± 1)ζ¯ 2(1 + ζζ¯) , ∇˜± ∂/∂ζ¯ = ∂ ∂ζ¯ − (m± 1)ζ 2(1 + ζζ¯) . (10.4) 2In this Section, we shall write explicit formulas over the chart domain U1 with local coordinates (ζ, ζ¯); we leave the corresponding formulas for U0 to the reader. 126 10.3 Twisted Dirac operators Definition 10.2. The Dirac operator D/m on the twisted spinor bundle S ⊗ Lm−→ S2 is given, in accordance with Definition 8.3, by D/m := c(dx 1)∇˜∂1 + c(dx2)∇˜∂2 = c(dζ)∇˜∂/∂ζ + c(dζ¯)∇˜∂/∂ζ¯ . From (9.10), we get c(dζ) = −1 2 (1 + ζζ¯)(γ1 + iγ2) = ( 0 −(1 + ζζ¯) 0 0 ) , c(dζ¯) = −1 2 (1 + ζζ¯)(γ1 − iγ2) = ( 0 0 1 + ζζ¯ 0 ) , so that D/m = c(dζ)∇˜−∂/∂ζ + c(dζ¯)∇˜+∂/∂ζ¯ ; explicitly, D/m = ( 0 −(1 + ζζ¯)∂/∂ζ − 1 2 (m− 1)ζ¯ (1 + ζζ¯)∂/∂ζ¯ − 1 2 (m+ 1)ζ 0 ) = ( 0 −ðζ − 12mζ¯ ð¯ζ − 12mζ 0 ) =: ( 0 D/ −m D/ +m 0 ) . (10.5) From (9.15) we get ðζψ + 12mζ¯ψ = Q1(ζ)∂ψ/∂ζ + 1 2 (m − 1)(∂Q1/∂ζ)ψ, where we have written Q1(ζ) = 1 + ζζ¯. Also, ð¯ζψ − 12mζψ = Q1(ζ)∂ψ/∂ζ¯ − 12(m+ 1)(∂Q1/∂ζ¯)ψ. Thus, D/ −m = −Q1(ζ)−(m−3)/2 ∂ ∂ζ Q1(ζ) (m−1)/2, D/ +m = Q1(ζ) (m+3)/2 ∂ ∂ζ¯ Q1(ζ) −(m+1)/2. (10.6) The kernel of D/ −m is a subspace of Γ(L m−1) and the kernel of D/ +m is a subspace of Γ(Lm+1). Suppose that ψ ∈ Γ(Lm+1), and let ψS(ζ, ζ¯) and ψN(z, z¯) be its component functions over U1 and U0 respectively. These are related by the gauge transformation rule ψN(ζ −1, ζ¯−1) ≡ (ζ/ζ¯)(m+1)/2ψS(ζ, ζ¯). From (10.6), ψ ∈ kerD/ +m iff a(ζ) := (1 + ζζ¯)−(m+1)/2ψS(ζ, ζ¯) (10.7) is an entire holomorphic function on C. Moreover, ψN(ζ −1, ζ¯−1) = ((1 + ζζ¯)ζ/ζ¯)(m+1)/2a(ζ) must be regular at ζ = ∞. Since Q1(ζ)(m+1)/2 = O(|ζ|)m+1, for nonnegative m this is only possible if a(ζ) ≡ 0. If m is negative, the entire function a(ζ) is O(|ζ|)|m|−1 as |ζ| → ∞, so that a(ζ) is a polynomial of degree at most |m| − 1. To sum up: dim kerD/ +m = { 0, if m ≥ 0, |m|, if m < 0. 127 The kernel of D/ −m is found similarly. Suppose that φ ∈ Γ(Lm−1), with component func- tions φS(ζ, ζ¯) and φN(z, z¯) related by φN(ζ −1, ζ¯−1) ≡ (ζ/ζ¯)(m−1)/2φS(ζ, ζ¯). Then φ ∈ kerD/ −m iff b(ζ¯) := (1 + ζζ¯)(m−1)/2φS(ζ, ζ¯) (10.8) is antiholomorphic. The regularity at∞ of φN(ζ−1, ζ¯−1) = ((1+ζζ¯)ζ¯/ζ)−(m−1)/2b(ζ¯), together with Q1(ζ) −(m−1)/2 = O(|ζ|)1−m, shows that b(ζ¯) ≡ 0 for m negative or zero, while for m positive b(ζ¯) is a polynomial in ζ¯ of degree at most m− 1. In fine: dim kerD/ −m = { m, if m > 0, 0, if m ≤ 0. We have in particular shown the following result. Proposition 10.1. The index of the Dirac operator D/m on the twisted spinor bundle S ⊗ Lm−→ S2 is given by indD/m = dim kerD/ + m − dim kerD/ −m = −m, that is, by the integer that labels the first Chern class of the twisting line bundle Lm. 10.4 The group action on the twisted spinor bundles Proposition 10.2. A bundle action of SU(2) on S ⊗ Lm−→ S2 is given by the formula (9.27), where ψ± ∈ Γ(Lm±1), and where the multipliers A±N , A±S are determined by A±N(g, z) := ( βz¯ + α¯ β¯z + α )(m±1)/2 , A±S (g ′, ζ) := (−β¯ζ¯ + α −βζ + α¯ )(m±1)/2 . Proof. The analysis of the SU(2)-action on the untwisted spinor bundle, given in Section 9.6, may be repeated verbatim, except that the gauge transformation factor (ζ¯/ζ)1/2 in (9.30) must be replaced by (ζ¯/ζ)(m±1)/2. This leads to the necessary and sufficient condition (9.31), with the exponent 1/2 replaced by (m± 1)/2, and the cocycles (10.2) satisfy that condition. Proposition 10.3. The Dirac operator D/m is equivariant under the action of SU(2) on the twisted spinor bundle S ⊗ Lm−→ S2 determined by (10.2). Proof. Let T ∈ End+(Γ(S⊗Lm)) be an operator that commutes with D/m and is of the form (Tψ±S )(ζ, ζ¯) = c±(ζ, ζ¯)ψ ± S ( b(ζ), b(ζ) ) . Then, as in Lemma 9.5, the identity TD/ −m = D/ − mT —in Hom(Γ(Lm−1),Γ(Lm+1))— yields the relations Q1(ζ)c−(ζ, ζ¯)b′(ζ) = Q1(b(ζ))c+(ζ, ζ¯), Q1(ζ) ∂c− ∂ζ = 1 2 (m− 1)(b(ζ)c+(ζ, ζ¯)− ζ¯c−(ζ, ζ¯)). (10.9) 128 Similarly, from TD/ +m = D/ + mT in Hom(Γ(L m+1),Γ(Lm−1)) we get Q1(ζ)c+(ζ, ζ¯)b′(ζ) = Q1(b(ζ))c−(ζ, ζ¯), Q1(ζ) ∂c+ ∂ζ¯ = −1 2 (m+ 1) ( b(ζ)c−(ζ, ζ¯)− ζc+(ζ, ζ¯) ) . (10.10) We check that these conditions are satisfied when T = T (g) is the operator corresponding to the diagonal element g = k(φ) of SU(2), determined by (10.2). Here b(ζ) = g′−1 · ζ = eiφζ. Clearly Q1(b(ζ)) ≡ Q1(ζ), so that (10.9) and (10.10) reduce to the conditions c+(ζ, ζ¯) = e iφc−(ζ, ζ¯), ∂c−/∂ζ = 0, and ∂c+/∂ζ¯ = 0, which together imply that c− and c+ are independent of ζ and ζ¯. Since the prescription c±(ζ, ζ¯) := ei(m±1)φ/2, dictated by (10.2), obeys c+ = e iφc−, we see that T (g) commutes with D/m. In the same way, it is easy to show that the operator T (h(θ)), associated to h(θ) of (9.36) by b(ζ) = h(θ)′−1 · ζ and by (10.2), commutes with D/m. Since elements of the form k(φ) and h(θ) generate SU(2), and since g 7→ T (g) is a homomorphism on account of the cocycle relations (9.29), we conclude that T (g)D/m = D/mT (g) for all g ∈ SU(2). Exercise 10.3. Complete the proof of the previous Proposition by verifying that the coefficient functions of T (h(θ)) satisfy (10.9) and (10.10). Corollary 10.4. The bundle action of SU(2) on S ⊗ Lm−→ S2 determined by (9.27) and (10.2) restricts to a finite-dimensional unitary representation ρm of SU(2) on the kernel of the Dirac operator D/m. To exhibit the representation ρm, it is useful to notice that with g ′−1 ·ζ = (αζ+β¯)/(−βζ+ α¯), one has Q1(g ′−1 · ζ) = Q1(ζ)| − βζ + α¯|−2 on account of αα¯+ ββ¯ = 1. If m > 0, then for ψ ∈ kerD/ −m ⊂ Γ(Lm−1), the bundle action (9.27) specializes to ρm ( α β −β¯ α¯ ) ψS(ζ, ζ¯) = (−β¯ζ¯ + α −βζ + α¯ )(m−1)/2 ψS ( αζ + β¯ −βζ + α¯ , α¯ζ¯ + β −β¯ζ¯ + α ) . (10.11) Using (10.8), we can write ψS(ζ, ζ¯) = Q1(ζ) −(m−1)/2b(ζ¯) with b a polynomial in ζ¯ of degree less than m, and the previous formula simplifies to ρm ( α β −β¯ α¯ )[ Q1(ζ) −(m−1)/2b(ζ¯) ] = Q1(ζ) −(m−1)/2(−β¯ζ¯ + α)m−1 b ( α¯ζ¯ + β −β¯ζ¯ + α ) . The right hand side is clearly also Q −(m−1)/2 1 times a polynomial in ζ¯ of degree less than m. If we define ψk(ζ, ζ¯) := Q1(ζ) −(m−1)/2ζ¯k, for k = 0, 1, . . . ,m − 1, these form an orthogonal basis for kerD/ −m. For these basis vectors, ρm ( α β −β¯ α¯ ) ψk(ζ, ζ¯) = Q1(ζ) −(m−1)/2(α¯ζ¯ + β)k(−β¯ζ¯ + α)m−k−1, (10.12) from which the matrix elements of the representation may be computed at once. 129 When m < 0, we obtain similar formulas for ρm. Indeed, for ψ ∈ kerD/ +m ⊂ Γ(Lm+1), we obtain (10.11) with the exponent (m − 1)/2 replaced by (m + 1)/2. Since, by (10.7), ψS(ζ, ζ¯) = Q1(ζ) (m+1)/2a(ζ) with a a polynomial in ζ of degree less than m, we get ρm ( α β −β¯ α¯ )[ Q1(ζ) (m+1)/2a(ζ) ] = Q1(ζ) (m+1)/2(−βζ + α¯)|m|−1 a ( αζ + β¯ −βζ + α¯ ) , (10.13) and on the orthogonal basis ψk(ζ, ζ¯) := Q1(ζ) −(|m|−1)/2ζk (k = 0, 1, . . . ,m − 1) for kerD/ +m, we find that ρm ( α β −β¯ α¯ ) ψk(ζ, ζ¯) = Q1(ζ) −(|m|−1)/2(αζ + β¯)k(−βζ + α¯)|m|−k−1. (10.14) From the explicit formulae (10.12) and (10.14), it is evident that ρ−m is the conjugate representation to ρm. Proposition 10.5. The representation ρm is irreducible. Proof. If m = 0, there is nothing to prove. The case m < 0 mirrors the case m > 0, on interchanging D/ +m with D/ − m; thus we may take m > 0. Then kerD/m = kerD/ − m is an m-dimensional Hilbert space with the orthogonal basis {ψk : k = 0, 1, . . . ,m − 1 }. It is immediate from (10.12) that this basis consists of joint eigenvectors for { ρm(g) : g ∈ T }; indeed, with g = k(φ) we get ρm(k(φ))ψk = e i(m−1)φ/2e−ikφψk, (10.15) so the eigenvalues are generally distinct, and therefore any operator S on kerD/ −m that com- mutes with each ρm(g) must have a diagonal matrix in this basis. On the other hand, the one-parameter subgroup of operators t 7→ ρm(h(t)) mingles the basis elements ψk: for h(t), one has α = cos 1 2 t, β = sin 1 2 t, and from (10.12) we get (d/dt) ∣∣ t=0 ρm(h(t))ψk = 1 2 kψk−1− 12(m−k−1)ψk+1 if 1 ≤ k ≤ m−2. A similar calculation with the subgroup h′(t) for which α = cos 1 2 t and β = i sin 1 2 t gives (d/dt) ∣∣ t=0 ρm(h ′(t))ψk = 12ikψk−1 + 1 2 i(m−k−1)ψk+1. Thus if S commutes with every ρm(g), it commutes with the “ladder operator” ψk 7→ kψk−1, and hence S is scalar. Irreducibility of ρm now follows from Schur’s lemma. In order to identify the unitary irreducible representation ρm of SU(2), it suffices to com- pute its character. Since the character is a class function, it is determined by its restriction to a maximal torus, so for m > 0 we obtain it immediately from (10.15): χm(k(φ)) = m−1∑ k=0 ei(m−1)φ/2e−ikφ = eimφ/2 eiφ/2 1− e−imφ 1− e−iφ = eimφ/2 − e−imφ/2 eiφ/2 − e−iφ/2 . When m < 0, we similarly get from (10.14) ρm(k(φ))ψk = e i(m+1)φ/2eikφψk, 130 and so its character is: χm(k(φ)) = |m|−1∑ k=0 ei(m+1)φ/2eikφ = eimφ/2 e−iφ/2 1− eimφ 1− eiφ = eimφ/2 − e−imφ/2 eiφ/2 − e−iφ/2 . Therefore the representations ρ−m and ρm are equivalent. This equivalence is to be expected on general grounds. The Weyl group of SU(2), namely the normalizer subgroup of the maximal torus T modulo T itself, is just the two-element group W ' Z2, since the only nontrivial isomorphism σ of the diagonal subgroup T is the interchange of diagonal elements, that can be implemented by conjugating with ( 0 1 −1 0 ) . By (10.1), χm ◦ σ = χ−m; so two characters of T lead to equivalent representations of SU(2) iff they lie in the same orbit under the action of the Weyl group. This exemplifies the general relation [13, 49]: R(G) ' R(T )W . This construction of the unitary irreducible representations of SU(2) directly from the equivariant Dirac operator exemplifies the “universal quantization map” of Vergne [55]: Q : KSU(2)(S2)→ R(SU(2)), which is already an instance of the index theorem of Atiyah, Segal and Singer [5, 6]. 10.5 The Borel–Weil theorem The construction of the irreducible unitary representations of SU(2) on the kernel spaces of Dirac operators, developed in the preceding sections, produces two families of equivalent representations, according as one twists the standard bundle with a tensor power of the tautological bundle (Lm, with m > 0), or with a tensor power of the hyperplane bundle (Hn ∼ L−n, with n > 0). Now only the latter line bundles admit nonzero holomorphic sections, while only the former admit nonzero antiholomorphic sections. The celebrated construction of Borel and Weil does not deal with spinor bundles, but rather uses the fact that the coadjoint orbits of compact Lie groups are complex manifolds and constructs the desired representations on the spaces of holomorphic sections of holomorphic line bundles over those orbits. By a simple application of Liouville’s theorem —see Section 5.6 for the details in the case M = CP(m)— the compactness of the orbit implies that the space of holomorphic sections is finite-dimensional. Indeed, in many cases, that of the trivial line bundle for instance, the space of holomorphic sections reduces to zero. A variant of the Borel–Weil construction produces representations on space of antiholomorphic sections of the respective dual line bundles. There is a simple relationship between the spinor-based and the Borel–Weil constructions: one passes from one to the other by twisting or untwisting with a fixed line bundle that can be obtained directly from the spin structure of the maximal coadjoint orbit G/T . Indeed, this relationship is parallel to the replacement of the spin structure by a spinc structure 131 on a spin manifold: for that, one needs to identify a particular principal U(1)-bundle that combines with the spin structure to yield a spinc structure. (See the discussion in Section 7.3, and also Appendix D of [39].) If M is a complex manifold of complex dimension m, let K := Λm,0T ∗M ; then K −→M is the so-called canonical line bundle, and Γ(K) = Am,0(M). The first Chern class of the dual bundle K∗, turns out [41] to be the element of Hˇ2(M,Z) whose modulo-2 reduction is the Stiefel–Whitney class w2(M); thus w2(M) + j∗(c1(K)) = 0 in Hˇ2(M,Z2), in the notation of Section 7.3. If M is spin, so that w2(M) = 0, then j∗(c1(K)) = 0 also, so that c1(K) is even; in other words, there is a complex line bundle K1/2−→M such that K1/2 ⊗K1/2 ∼ K. In principle, the “square root of the canonical bundle” K1/2 may depend on the spin structure chosen for M . For the case M = S2, we know that K ∼ L⊗ L = L2 by Exercise 9.3. Thus we can take K1/2 := L. Its dual line bundle is K−1/2 := H. Therefore we find that S ⊗K−1/2 = (L⊕H)⊗H ∼ E ⊕H2 ∼ Λ0,•T ∗S2 where E = S2×C denotes the trivial line bundle, and Λ0,•T ∗S2−→ S2 is the complex vector bundle whose smooth sections form the module A0,•(S2) = A0,0(S2) ⊕ A0,1(S2). Over U1, such a section may be written as f(ζ, ζ¯) + h(ζ, ζ¯) dζ¯. Thus A0,•(S2) = Γ(S⊗L−1) is the domain of the twisted Dirac operator D/ −1. Of course, (10.5) for m = −1 simplifies to D/ −1 = ( 0 −Q1(ζ)∂/∂ζ + ζ¯ Q1(ζ)∂/∂ζ¯ 0 ) . The operatorD/ +−1 : A 0,0(S2)→ A0,1(S2) has as kernel the holomorphic functions inA0,0(S2) = C∞(S2), which are just the constant functions, by Liouville’s theorem. The factor Q1(ζ) = 1 + ζζ¯ is a normalization factor, that takes account of the metric on A0,1(S2). To see that, consider the twisted spinor bundle S ⊗ Lm for any negative m, say m = −(2j+1) with j a nonnegative half-integer.3 Clearly S⊗Lm = Λ0,•T ∗S2⊗H2j = H2j⊕H2j+2, so that S+⊗Lm = H2j admits holomorphic sections. If s1 denotes a section of H normalized by (s1 |s1) = 1, then s2j1 is a normalized section of H2j. We may select s1 as s1 := Q1(ζ)1/2σ1, where σ1 is the holomorphic section of H −→CP(1) of Section 5.6: the normalization follows from (5.13), whereby we have (σ1 | σ1) = Q1(ζ)−1. It is convenient4 to choose the metric 4g−1 = Q1(ζ)2(∂/∂ζ) · (∂/∂ζ¯) on T ∗CM , so that the normalized section of A0,1(S2) is just s21 = Q1(ζ)−1 dζ¯. It follows that σ21 = Q1(ζ)−2 dζ¯. Any section of A0,1(S2) is of the form h(ζ, ζ¯) dζ¯ = Q1(ζ)h(ζ, ζ¯) s21, and a section of S⊗L−1 = L0⊕L−2 can be written as φ = f(ζ, ζ¯) +h(ζ, ζ¯) dζ¯. Then the Dirac operator D/ −1 can be identified as follows. Firstly, D/ +−1f = ( Q1(ζ) ∂f ∂ζ¯ ) s21 = ∂f ∂ζ¯ dζ¯ = ∂¯f 3By now it should be clear that it is preferable to label representations of SU(2) by nonnegative half- integers j rather than the integers m. 4Without this normalization, the operator ∂¯ in the following formulae should be replaced by √ 2 ∂¯; this convention is adopted in [28], for instance. 132 where ∂¯ is the Dolbeault operator of Section 2.1. The twisted Dirac operator D/m, with m = −(2j + 1), is D/m = ( 0 −Q1(ζ)∂/∂ζ + (j + 1)ζ¯ Q1(ζ)∂/∂ζ¯ + jζ 0 ) , by rewriting (10.5). An element of Γ(H2j) can be written (over U1) as f(ζ, ζ¯)σ 2j 1 . Then D/ +m(f(ζ, ζ¯)σ 2j 1 ) = D/ + m(Q1(ζ) −jf(ζ, ζ¯) s2j1 ) = ( Q1(ζ) ∂ ∂ζ¯ + jζ )[ Q1(ζ) −jf(ζ, ζ¯) ] s2j+21 = Q1(ζ) −j+1∂f ∂ζ s2j+21 = Q1(ζ) 2∂f ∂ζ σ2j+21 = ∂f ∂ζ dζ¯ ⊗ σ2j1 = ∂¯f ⊗ σ2j1 . (10.16) The Dolbeault operator extends to A0,•(S2, H2j) := A0,•(S2)⊗Γ(H2j) by setting ∂¯(ω⊗s) := (∂¯ω) ⊗ s for s ∈ Γ(H2j). With this understanding,5 we have the relation D/ +−2j−1 = ∂¯ as operators from A0,0(S2, H2j) to A0,1(S2, H2j). In order to identify D/ −m, we must compute the adjoint of the Dolbeault operator. We consider first the case m = −1. Definition 10.3. The adjoint of ∂¯ : A0,0(S2) → A0,1(S2) is the operator ∂¯∗ : A0,1(S2) → A0,0(S2) given by 〈〈∂¯∗(h dζ¯) | f〉〉 := 〈〈h dζ¯ | ∂¯f〉〉 where 〈〈· | ·〉〉 denotes the integrated inner product in either space of sections, defined as in (9.49) by integrating the appropriate pairing of sections over S2 with respect to the volume form (4pi)−1 sin θ dθ ∧ dφ = (2pii)−1Q1(ζ)−2 dζ ∧ dζ¯. We find that 〈〈∂¯∗(h dζ¯) | f〉〉 = 〈〈h dζ¯ | ∂¯f〉〉 = 1 2pii ∫ C 4g−1 ( h¯ dζ, ∂f ∂ζ¯ dζ¯ ) Q1(ζ) −2 dζ ∧ dζ¯ = 1 2pii ∫ C h¯ ∂f ∂ζ¯ dζ ∧ dζ¯ = − 1 2pii ∫ C ∂h ∂ζ f dζ ∧ dζ¯ = −〈〈Q1(ζ)2 ∂h/∂ζ¯ | f〉〉, on integrating by parts, so that ∂¯∗(h dζ¯) = −(1 + ζζ¯)2 ∂h/∂ζ¯. From this it follows easily that D/ −−1(h dζ¯) = D/ − −1(Q1h s 2 1) = (−Q1∂/∂ζ + ζ¯)(Q1h) = −Q21 ∂h/∂ζ¯ = ∂¯∗(h dζ¯). (10.17) Thus D/ −−1 = ∂¯ ∗ on A0,1(S2). Exercise 10.4. Show D/ −−2j−1 = ∂¯ ∗ as operators from A0,1(S2, H2j) to A0,0(S2, H2j). 5We could have written ∂¯(2j) to denote the extended operator, but we suppress the index to reduce notational clutter. 133 We summarize these calculations as follows. Proposition 10.6. The Dirac operator on the twisted spinor bundle S ⊗ L−2j−1 equals the sum of the Dolbeault operator and its adjoint on the twisted bundle Λ0,•T ∗S2 ⊗H2j. Proof. The formulae (10.16) and (10.17), plus the previous exercise, establish that D/ +−2j−1 = ∂¯ and D/ −−2j−1 = ∂¯ ∗, thus D/ −2j−1 = ∂¯ + ∂¯∗ on the module A0,•(S2, H2j). The kernel of the Dirac operator D/ −2j−1 thus coincides with ker ∂¯ ⊕ ker ∂¯∗. The sec- ond summand is zero since kerD/ −−2j−1 = 0. Thus kerD/ + −2j−1 = ker ∂¯ consists of sections ψS(ζ, ζ¯) s 2j 1 = f(ζ, ζ¯)σ 2j 1 ∈ A0,0(S2, H2j) for which ∂f/∂ζ¯ = 0; these are precisely the holo- morphic sections f(ζ)σ2j1 ∈ O(H2j). Since ψS(ζ, ζ¯) s 2j 1 = f(ζ)σ 2j 1 = f(ζ)Q1(ζ) −j s2j1 it follows from (10.7) that f(ζ) is a polynomial of degree at most 2j. That is to say, we have shown that dimO(H2j) = 2j+ 1. By suppressing the factors Q1(ζ) −j in (10.13) and (10.14), we arrive at the following representation pij of SU(2) on O(H 2j), that is by construction equivalent to ρ−2j−1: pij ( α β −β¯ α¯ )[ f(ζ) ] = (−βζ + α¯)2j f ( αζ + β¯ −βζ + α¯ ) , and on the orthogonal basis { ξk := ζk σ2j1 : k = 0, 1, . . . , 2j } for O(H2j), we find that pij ( α β −β¯ α¯ ) ξk = (αζ + β¯) k(−βζ + α¯)2j−k σ2j1 . This completes the passage from the Dirac-operator-kernel construction of the irreducible representations of SU(2) to their realization on finite-dimensional spaces of holomorphic sections of line bundles over S2. The content of the foregoing construction is the Borel– Weil theorem for the compact group SU(2), that we may now state as follows. Theorem 10.7. Every irreducible unitary representation of SU(2) can be realized on a space of holomorphic sections of a holomorphic Hermitian line bundle over S2. This line bundle is determined, up to equivalence, by a character χ of a maximal torus of SU(2) modulo the action of the Weyl group Z2. Moreover, if E−→ S2 is a holomorphic Hermitian line bundle whose Chern class is m[H] with m > 0, then O(E) carries the representation corresponding to the character χm of the maximal torus. This result is of course well known; what we have done is to show how it arises from the equivariant index of the Dirac operator on the flag manifold S2. We remark that our construction is equivalent to the standard induced representation recipe for each pij, though this is usually achieved by studying the complexification of the compact Lie group in question [56]. 134 A Calculus on manifolds In this Appendix, we briefly review the concepts and notations of calculus on manifolds, with emphasis on the algebraic formulae which arise in differential geometry. Proofs are left to the reader. General references are the books of Abraham, Marsden and Ratiu [1], Crampin and Pirani [20], Singer and Thorpe [51], and Spivak [52]. A.1 Differential manifolds Definition A.1. A differential manifold of finite dimension n is a paracompact Hausdorff topological space M together with a family (or “atlas”) of local charts { (Uj, φj) : j ∈ J } such that U := {Uj : j ∈ J } is a locally finite open covering of M , φj : Uj → Rn is a homeomorphism onto an open subset of Rn, and the transition functions φi ◦ φ−1j : φj(Ui ∩ Uj)→ φi(Ui ∩ Uj) are smooth. A second atlas { (Vk, ψk) : k ∈ K } is declared equivalent to the first if every φi ◦ ψ−1k is smooth (the “differentiable structure” of M is actually an equivalence class of atlases). If M is compact, a finite atlas may be chosen. If n = 2m is even, we can regard the chart maps φj as having images in Cm. We say that M is a complex manifold if the transition functions are holomorphic maps between open subsets of Cm. Definition A.2. If M , N are two differential manifolds, a continuous map f : M → N is smooth if for any pair of local charts (U, φ) for M and (V, ψ) for N , the composite map ψ ◦f ◦φ−1 : φ(U ∩f−1(V ))→ ψ(V ) is smooth. When N = R, the set of all smooth functions on M is a commutative algebra over R, which we denote by C∞(M,R); when N = C, the smooth complex-valued functions on M forms a commutative C-algebra, C∞(M,C). We often write simply C∞(M), if it is clear from the context whether real-valued or complex- valued functions are to be used. A diffeomorphism between M and N is a bijective smooth function f : M → N whose inverse f−1 : N → M is also smooth. If such an f exists, we say that M and N are diffeomorphic. If (Uj, φj) is a local chart for M , we define x 1, . . . , xn ∈ C∞(Uj,R) by xk := prk ◦φj, where prk : Rn → R is the k-th coordinate projector. We say (x1, . . . , xn) is a system of local coordinates for M on the chart domain Uj. The following two lemmas show that smooth functions are abundant. Lemma A.1. If M is a differential manifold, and if V , W are two open subsets of M with V ⊂ W , there exists f ∈ C∞(M,R) such that supp f ⊂ W , f ≡ 1 on V , and 0 ≤ f ≤ 1 on W \ V . Lemma A.2. If M is a differential manifold, with atlas { (Uj, φj) : j ∈ J }, there exists a smooth partition of unity subordinate to the locally finite covering U, that is, a family { fj : j ∈ J } ⊂ C∞(M,R) with 0 ≤ fj ≤ 1 and supp fj ⊂ Uj for each j, such that∑ j∈J fj(x) = 1 for all x ∈M (the sum is finite for each x). 135 If M , N are two manifolds with respective atlases {(Uj, φj)}, {(Vk, ψk)}, the product manifold is the cartesian product M × N with atlas {(Uj × Vk, φj × ψk)}; its dimension is dim(M ×N) = dimM + dimN . A.2 Tangent spaces Definition A.3. Let M be a differential manifold, x ∈ M . Let C∞(M,x) be the set of all smooth functions f : Vf → R whose domain is an open neighbourhood of x in M ; this is a commutative algebra that includes C∞(M). Indeed, C∞(M) = ⋂ x∈M C ∞(M,x). A tangent vector at x is an R-linear map v : C∞(M,x) → R which satisfies the “local Leibniz rule”: v(fg) = v(f) g(x) + f(x) v(g), for all f, g ∈ C∞(M,x). These form a real vector space TxM . If (Uj, φj) is a local chart for M , with local coordinates (x 1, . . . , xn), then the directional derivatives ∂ ∂xj ∣∣ x : f 7→ Dj(f ◦φ−1)(φ(x)) form a basis for TxM ; in particular, dimTxM = n. If γ : I → M is a smooth curve, whose domain is an interval I ⊆ R, with γ(t0) = x, its velocity vector at x is γ˙(t0) ∈ TxM defined by γ˙(t0)(f) := (f ◦ γ)′(t0). Definition A.4. If f : M → N is smooth, and x ∈ M , the tangent mapping Txf : TxM → Tf(x)N is the R-linear map Txf(v) : h 7→ v(h ◦ f), for v ∈ TxM , h ∈ C∞(N, f(x)). If (y1, . . . , yr) is a system of local coordinates near f(x) ∈ N , the matrix of Txf is has entries ∂f k/∂xj|x := ∂∂xj ∣∣ x (yk ◦ f). Definition A.5. A smooth mapping f : M → N is an immersion if for all x ∈ M , the tangent map Txf : TxM → Tf(x)N is injective. If each Txf is surjective, f is a submersion. If f is both an immersion and a submersion, then dimM = dimN , and the Jacobian matrix of Txf (in local coordinates) is invertible, for each x; by the inverse function theorem, f is a diffeomorphism between a neighbourhood of x and a neighbourhood of f(x), for each x; we say f is a local diffeomorphism. However, f need not be injective or surjective on all of M , so it need not be a global diffeomorphism. A.3 Vector fields Definition A.6. Let M be a differential manifold. A vector field on M is a R-linear operator X : C∞(M)→ C∞(M), which is a derivation of this algebra, that is, it satisfies the Leibniz rule: X(fg) = (Xf) g + f (Xg), for all f, g ∈ C∞(M). These derivations form a vector space denoted by X(M). This is in fact a module for the algebra C∞(M), if we define1 fX ∈ X(M) by fX(h) := f(Xh). 1There is a notational difficulty here, since expressions like fXh, while not ambiguous, are confusing. For this reason, the derivation action of the vector field X on the function h is sometimes written X · h rather than Xh; then the module structure can be defined by (fX) · h := f(X · h), and so forth. 136 If Y ∈ X(M) and x ∈M , the recipe Yxf := (Y f)(x) defines a tangent vector Yx ∈ TxM . If (U, φ) is a local chart of M , any vector field X ∈ X(M) determines a vector field X|U ∈ X(U) by restriction. If (x1, . . . , xn) is the local coordinate system for this chart, we obtain n linearly independent local vector fields ∂ ∂xj ∈ X(U) by writing ∂ ∂xj (f) := Dj(f ◦ φ−1) ◦ φ. These form a basis for the C∞(U)-module X(U), that is, every X ∈ X(U) is of the form X = ∑n j=1 a j ∂ ∂xj , with a1, . . . , an ∈ C∞(U). Definition A.7. The Lie bracket of two vector fields X, Y ∈ X(M) is defined as [X, Y ] : f 7→ X(Y f)− Y (Xf). It is easy to check that this is a derivation of C∞(M); it is clearly skewsymmetric, and it satisfies the Jacobi identity : [X, [Y, Z]] + [Y, [Z,X]] + [Z, [X, Y ]] = 0. (A.1) Thus X(M) is an (infinite-dimensional) Lie algebra. Definition A.8. If τ : M → N is a diffeomorphism, and X ∈ X(M), we define the pushout τ∗X ∈ X(N) by: τ∗X(h) := X(h ◦ τ) ◦ τ−1 for all h ∈ C∞(N). For each x ∈M , we find that (τ∗X)τ(x) = Txτ(Xx). Note that this last formula makes sense if τ : M → N is smooth and surjective, but not necessary invertible. Lemma A.3. The pushout τ∗ : X(M) → X(N) is a Lie algebra homomorphism, i.e., τ∗ is linear and τ∗[X, Y ] = [τ∗X, τ∗Y ] for X, Y ∈ X(M). If σ : N → R is a diffeomorphism, then (σ ◦ τ)∗ = σ∗ ◦ τ∗. Definition A.9. An integral curve of a vector fieldX ∈ X(M) is a smooth curve γ : I →M such that γ˙(t) = Xγ(t) for all t ∈ I. One can always find a unique integral curve γx for X satisfying γ(0) = x in some maximal interval Ix 3 0, by the existence and uniqueness theorem for first-order ordinary differential equations. We say the vector field X is complete if Ix = R for all x ∈ M ; if M is compact, every vector field is complete. Write φt(x) := γx(t); then φt : M → M is a diffeomorphism for all t ∈ R, and φt ◦ φs = φt+s for all t, s. The one-parameter group of diffeomorphisms {φt} is called the flow generated by the vector field X. The vector field may be recovered from the flow by noticing that Xf = lim t→0 f ◦ φt − f t = d dt ∣∣∣∣ t=0 (f ◦ φt). 137 A.4 Lie groups Definition A.10. A Lie group is a differential manifold G which is also a group, for which the multiplication (g, h) 7→ gh : G×G→ G and the inversion g 7→ g−1 : G→ G are smooth maps. The left translations λg : h 7→ gh and the right translations ρg : h 7→ hg are diffeomor- phisms from G onto G. We usual write e to denote the identity element of G. A finite-dimensional (real or complex) vector space V is an additive Lie group. The group GLR(V ) of invertible linear operators on V is a Lie group (since it is a dense open subset of the vector space EndR(V )); if V is a complex vector space, GLC(V ) is a Lie group. We write GL(n,R) := GLR(Rn) and GL(m,C) := GLC(Cm). Definition A.11. A vector field X ∈ X(G) on a Lie group G is left-invariant if (λg)∗X = X for all g ∈ G. In that case, X is determined by its value at the identity, Xe ∈ TeG. A Lie algebra is a real vector space with a skewsymmetric bilinear operation [·, ·] sat- isfying the Jacobi identity (A.1). Since the Lie bracket [X, Y ] of two left-invariant vector fields X, Y is also left-invariant, TeG becomes a (finite-dimensional) Lie algebra by defining [Xe, Ye] := [X, Y ]e. We usually write g := TeG to denote this Lie algebra. Definition A.12. If G is a Lie group with Lie algebra g, and if X ∈ g, let γX be the integral curve of the corresponding left-invariant vector field such that γX(0) = e. Then γX(s+ t) = γX(s) γX(t) for all s, t ∈ R, so that t 7→ γX(t) is a one-parameter subgroup of G; also γtX(1) = γX(t) for t ∈ R. We write expX := γX(1). This defines the exponential map exp: g → G, which satisfies exp tX = γX(t), and thus t 7→ exp tX is a homomorphism. Unless G is abelian, exp is not a homomorphism of the additive group g into G. However, there is the important Campbell–Baker–Hausdorff formula: exp tX exp tY = exp ( t(X + Y ) + 1 2 t2[X, Y ] +O(t3) ) , and its corollary: exp tX exp tY exp(−tX) = exp(tY + t2[X, Y ] +O(t3)). (A.2) If G is a closed subgroup of GLR(V ), the Lie algebra can be identified with the subspace of operators {X ∈ EndR(V ) : exp tX ∈ G for all t ∈ R }, and the Lie bracket becomes [X, Y ] = XY − Y X ∈ EndR(V ). If W is a complex vector space, the Lie algebra of a subgroup of GLC(W ) is likewise identified to a subspace of EndC(W ). Definition A.13. A smooth left action of a Lie group G on a differential manifold M is a smooth map Φ: G × M → M : (g, x) 7→ Φ(g, x) ≡ g · x, such that e · x = x and g · (h · x) = (gh) · x, for g, h ∈ G, x ∈M . Thus Φg : x 7→ g · x is a diffeomorphism of M for each g ∈ G. If x ∈ M , the isotropy subgroup for x is Gx := {h ∈ G : h · x = x }, and the orbit of x is G · x := { g · x ∈ M : g ∈ G } ⊆ M . The natural bijection g · x 7→ gGx between G · x and the left-coset space G/Gx is a diffeomorphism when G/Gx is given a natural differential structure for which the quotient map η : G→ G/Gx is a submersion. 138 We say that the action of G on M is free if no group element except e leaves any point fixed; in that case, all isotropy groups are trivial and all orbits are diffeomorphic to G. We say that that the action of G on M is transitive if there is only one orbit. If so, and if H is the isotropy subgroup of some point of M , then M is diffeomorphic to G/H (and the action of G on M corresponds to permutation of the left-cosets g′H 7→ gg′H). A manifold with a transitive G-action is called a homogeneous space for the group G. Definition A.14. A smooth right action of a Lie group G on a differential manifold M is similarly defined, as a smooth map M × G → M : (x, g) 7→ x · g such that x · e = x and (x · g) · h = x · (gh), for g, h ∈ G, x ∈M . If H is a closed subgroup of G, then H acts (on the right) on G by right translations g 7→ gh; the orbits of this action are the cosets gH. Definition A.15. The adjoint action of a Lie group G on its Lie algebra g is the map (g,X) 7→ Ad(g)X given by Ad(g)X := d dt ∣∣∣∣ t=0 g(exp tX)g−1. It can be deduced from (A.2) that Ad(g)X ∈ g. The map g 7→ Ad(g) is a homomorphism Ad: G→ GL(g). Definition A.16. A representation of a Lie group G on a vector space V is a homomor- phism ρ : G→ GL(V ) for which (g, v) 7→ ρ(g)v is a smooth map from G× V to V . The derived representation of its Lie algebra g is the linear map ρ˙ : g→ End(V ) given by ρ˙(X)v := d dt ∣∣∣∣ t=0 ρ(exp tX)v. In particular, the derived representation ad of the adjoint representation Ad is given, on account of (A.2), as ad(X)Y = [X, Y ]. A.5 Fibre bundles Definition A.17. A fibre bundle is a triple (E,M, pi), more usually written as E pi−→M , where E and M are differential manifolds (called, respectively, the total space and the base space), and pi : E → B is a surjective submersion,2 such that each fibre Ex := pi−1({x}) is diffeomorphic to a fixed manifold F , called the “typical fibre”, and which is locally trivial in the following sense. If U = {Uj : j ∈ J } is a covering of the base space by chart domains, there are diffeomorphisms ψj : pi −1(Uj)→ Uj × F (A.3) (called “local trivializations”) such that pi(ψ−1j (x, v)) = x for all x ∈ Uj, v ∈ F . 2We shall often omit explicit mention of the submersion pi and simply write E−→M to denote the fibre bundle. 139 Definition A.18. If E pi−→M and E ′ pi′−→M ′ are two fibre bundles, a bundle morphism is a pair of smooth maps (τ, σ), with τ : E → E ′ and σ : M →M ′, such that pi′ ◦ τ = σ ◦ pi. (We also say that τ is a lifting of the map σ between the base spaces.) A bundle equivalence is a bundle morphism (τ, σ) such that both τ and σ are diffeo- morphisms. A fibre bundle E−→M is trivial if it is equivalent to the product bundle M × F pr1−→M via a bundle morphism (τ, idM). Thus (A.3) says that any fibre bundle is locally a trivial bundle. Definition A.19. A vector bundle is a fibre bundle E pi−→M whose typical fibre is a (real or complex) vector space V , and whose fibres Ex are vector spaces of the same dimension, such that the maps V → Ex : v 7→ ψ−1j (x, v) are linear isomorphisms. The dimension dimV is called the rank of the vector bundle. The tangent bundle TM −→M of a manifold M has total space TM := { (x, v) : v ∈ TxM }, with pi(x, v) := x. The local trivialization ψj : pi−1(Uj) → Uj × Rn is given by ψj(x, v) := (x; v 1, . . . , vn), with v = ∑ k v k ∂ ∂xk ∣∣ x . The atlas { (pi−1(Uj), (φj×id)◦ψj) : j ∈ J } makes TM a 2n-dimensional manifold. The fibre at x ∈M is the tangent space TxM . The cotangent bundle T ∗M −→M is formed similarly; its fibres are the dual spaces T ∗xM := (TxM) ∗. We define {dx1|x, . . . , dxn|x} as the dual basis in T ∗xM to the basis { ∂ ∂x1 ∣∣ x , . . . , ∂ ∂xn ∣∣ x } of TxM , and if ξ = ∑ k ξk dx k|x, then ψj(x, ξ) := (x; ξ1, . . . , ξn). Definition A.20. A smooth section of a fibre bundle E pi−→M is a smooth map s : M → E such that pi ◦ s = idM , i.e., s(x) ∈ Ex for each x ∈ M . We denote the totality of smooth sections by Γ(M,E), or simply by Γ(E) if the base space M is understood; it is a C∞(M)- module, where the action of C∞(M) is just scalar multiplication on each fibre: (fs)(x) := f(x) s(x), for s ∈ Γ(E), f ∈ C∞(M). If U ⊂ M is open, a smooth map s : U → pi−1(U) satisfying pi(s(x)) = x for x ∈ U is called a local section of E pi−→M ; all such maps form a vector space Γ(U,E), which is a module over C∞(U). From the definition, it is easy to see that a smooth section of the tangent bundle TM −→M can be written in local coordinates on a chart domain Uj as X = ∑n k=1 a k ∂ ∂xk , with each ak ∈ C∞(U). Thus X is nothing other than a vector field on M ; we have Γ(TM) = X(M) as C∞(M)-modules. A section of a trivial fibre bundle M × F pr1−→M is of the form s(x) = (x, f(x)) where f : M → F is a smooth map. Thus sections of more general bundles can be thought of as “functions” which take values in different sets at each point of their domains. 140 A.6 Tensors and differential forms Definition A.21. A differential 1-form on a differential manifold M is a map α : X(M)→ C∞(M) that is C∞(M)-linear, that is: α(X + Y ) = α(X) + α(Y ), α(fX) = f α(X), for X, Y ∈ X(M) and f ∈ C∞(M). These form a real vector space A1(M), which becomes a C∞(M)-module on defining fα : X 7→ f α(X). If x ∈ M , then α(Y )(x) = fα(Y )(x) if f is any smooth function with f(x) = 1, whose support is an (arbitrary small) neighbourhood of x. Thus α(Y )(x) depends only on Yx, and so Yx 7→ α(Y )(x) is an element αx of the dual space T ∗xM of TxM ; by definition, α(Y )(x) = αx(Yx). Hence α can be identified with the section x 7→ αx of the cotangent bundle T ∗M −→M ; and Γ(T ∗M) = A1(M) as C∞(M)-modules. In local coordinates over U , we can write α = ∑n k=1 fk dx k with each fk ∈ C∞(U). Definition A.22. A tensor of bidegree (p, q) on a manifold M is a multilinear map T : X(M)p ×A1(M)q → C∞(M) such that T (X1, . . . , Xp, α 1, . . . , αq) is C∞(M)-linear in each Xj and each αk. The tensor is called covariant if q = 0, or contravariant if p = 0. Any such tensor T defines a smooth section of a vector bundle over M whose fibre at x ∈M is (T ∗xM)⊗p ⊗ (TxM)⊗q. Definition A.23. A Riemannian metric on M is a tensor g of bidegree (2, 0) that is symmetric, i.e., g(X, Y ) = g(Y,X) for all X, Y ∈ X(M), and positive definite: g(X,X) > 0 for nonzero X ∈ X(M). Locally, we may take g = gij dxi · dxj, where [gij] is a positive- definite symmetric matrix of elements of C∞(U) and dxi · dxj := 1 2 (dxi ⊗ dxj + dxj ⊗ dxi); a Riemannian metric may be defined globally on M by taking g = ∑ j fjg (j), where g(j) is a metric on the chart domain Uj, and the functions fj form a smooth partition of unity on M . The pair (M, g), consisting of a differential manifold with a Riemannian metric g, is called a Riemannian manifold. A Hermitian metric on M is a tensor h of bidegree (2, 0) with values in C∞(M,C), such that h(X, Y ) = h(Y,X) for X, Y ∈ X(M) and h is positive definite. One often writes (X |Y ) instead of h(X, Y ). The same partition-of-unity argument shows that any manifold can be given a Hermitian metric. Definition A.24. A differential k-form on M is a covariant tensor ω : X(M)k → C∞(M) that is alternating, which means that ω(Xσ(1), . . . , Xσ(k)) = (−1)σω(X1, . . . , Xk) for σ ∈ Sk. The totality of k-forms on M is denoted Ak(M), and is a C∞(M)-module. Let ΛkT ∗M −→M be the vector bundle whose fibre at x is ΛkT ∗xM , the k-th exterior power of T ∗xM ; then A k(M) = Γ(ΛkT ∗M). The direct sum A•(M) := ⊕n k=0A k(M) = Γ(Λ•T ∗M) is a Z-graded C∞(M)-module. The zero-degree term isA0(M) := C∞(M). Under the exterior product, A•(M) is an algebra; 141 in fact, it is Z2-graded by the parity of the degree k, and is a supercommutative superalgebra, since this property holds in each fibre of Λ•T ∗M −→M . Thus ω ∧ η = (−1)]ω ]ηη ∧ ω where ]ω = k for ω ∈ Ak(M). Thus if ω ∈ Ak(M), η ∈ Al(M), and X1, . . . , Xk+l ∈ X(M), then (ω ∧ η)(X1, . . . , Xk+l) = 1 k! l! ∑ σ∈Sk+l (−1)σω(Xσ(1), . . . , Xσ(k)) η(Xσ(k+1), . . . , Xσ(k+l)), and in particular, (α ∧ β)(X, Y ) = α(X)β(Y ) − α(Y )β(X) for α, β ∈ A1(M). Moreover, if α1, . . . , αk ∈ A1(M) then α1 ∧ · · · ∧ αk ∈ Ak(M), with (α1 ∧ · · · ∧ αk)(X1, . . . , Xk) = det[αi(Xj)]. In local coordinates, an element of Ak(U) is of the form ω = ∑ |J |=k fJ dx j1 ∧ · · · ∧ dxjk , where J = {j1, . . . , jk} ⊆ {1, . . . , n}. A.7 Calculus of differential forms Definition A.25. If τ : M → N is a diffeomorphism, and ω ∈ Ak(N), we define the pull- back τ ∗ω ∈ Ak(M) by: τ ∗ω(X1, . . . , Xk) := ω(τ∗X1, . . . , τ∗Xk) ◦ τ. In particular, τ ∗β(X) := β(τ∗X) ◦ τ for β ∈ A1(N). Thus (τ ∗β)x(Xx) := βτ(x)(Txτ(Xx)), so that the linear map βτ(x) 7→ (τ ∗β)x : T ∗τ(x)N → T ∗xM is the transpose of the tangent map Txτ : TxM → Tτ(x)N . If τ : M → N is any smooth map, not necessarily a diffeomorphism, the pullback τ ∗ω of ω ∈ Ak(N) is likewise defined by transposition: (τ ∗ω)x((X1)x, . . . , (Xk)x) := ωτ(x)(Txτ((X1)x), . . . , Txτ((Xk)x)). For k = 0, we get simply: τ ∗f := f ◦ τ . Lemma A.4. The pullback τ ∗ : A•(N) → A•(M) is a degree-preserving homomorphism of exterior algebras, that is, τ ∗ is linear and τ ∗(ω ∧ η) = τ ∗ω ∧ τ ∗η for ω, η ∈ A•(N). If σ : N → R is a smooth map, then (σ ◦ τ)∗ = τ ∗ ◦ σ∗. Definition A.26. The contraction of a k-form ω ∈ Ak(M) with a vector field X ∈ X(M) is the (k − 1)-form ι(X)ω ≡ ιXω defined as ιXω(X1, . . . , Xk−1) := ω(X,X1, . . . , Xk−1). For f ∈ A0(M), we set ιXf := 0. If α ∈ A1(M), then ιXα = α(X) ∈ C∞(M). 142 Lemma A.5. Contraction with X ∈ X(M) is an odd derivation of the graded algebra A•(M), that is, ιX(ω ∧ η) = ιXω ∧ η + (−1)]ωω ∧ ιXη. In particular, ιX(α 1 ∧ · · · ∧ αk) = k∑ j=1 (−1)j−1αj(X) (α1 ∧ · · j ∨· · ∧ αk) for α1, . . . , αk ∈ A1(M). Proposition A.6. There is a unique operator d : A•(M)→ A•(M), called exterior deriva- tion, such that: 1. d is an odd derivation of degree +1, that is, d(Ak(M)) ⊂ Ak+1(M), and d(ω ∧ η) = dω ∧ η + (−1)]ωω ∧ dη; 2. df(X) = Xf for f ∈ A0(M) = C∞(M); 3. d2 ≡ d ◦ d = 0; 4. d is natural with respect to restrictions, that is, if U ⊂M is open and ω ∈ A•(U), then d(ω|U) = (dω) ∣∣ U . In local coordinates, d (∑ J fJ dx j1 ∧ · · · ∧ dxjk) = ∑J dfJ ∧ dxj1 ∧ · · · ∧ dxjk , where, for f ∈ C∞(U), df = ∑nj=1(∂f/∂xj) dxj. Lemma A.7. If ω ∈ Ak(M), its exterior derivative dω ∈ Ak+1(M) is given by dω(X1, . . . , Xk+1) = k+1∑ j=1 (−1)j−1Xj(ω(X1, . . j ∨. . , Xk+1)) + ∑ i N .) We usually suppress the index of d and write simply that d2 = 0 at all stages. The elements of Cn are called “n-cochains”: c ∈ Cn is an n-cocycle if dc = 0; it is an n-coboundary if c = db for some b ∈ Cn−1. The condition d2 = 0 says that the totality of n-coboundaries Bn(C•) is a subgroup of the group Zn(C•) of n-cocycles. The quotient group Hn(C•) := Zn(C•)/Bn(C•) is called the n-th cohomology group of the complex. A morphism between two complexes (C•, d) and (K•, d′) is a set of homomorphisms fn : C n → Kn which intertwines the d-maps, i.e., fn+1 ◦ dn = d′n ◦ fn for all n. Thus fn(Z n(C•)) ⊆ Zn(K•) and fn(Bn(C•)) ⊆ Bn(K•), so that f induces a homomorphism Hnf : Hn(C•)→ Hn(K•). The components Cn of a complex may have more structure than that of an abelian group: they could be vector spaces, modules over a commutative ring, etc.3 The cohomology groups Hn(C•, d) inherit a similar structure. Definition A.29. The de Rham complex of an n-dimensional differential manifold M is the terminating complex A0(M) d−→A1(M)→ · · · → Ak(M) d−→Ak+1(M)→ · · · d−→An(M) of C∞(M)-modules; here d is the exterior derivation. We say a k-form ω is closed if dω = 0, and that ω is exact if ω = dη for some (k − 1)-form η; thus ZkdR(M) := Zk(A•(M), d) comprises the closed k-forms and BkdR(M) := B k(A•(M), d) comprises the exact k-forms. The k-th de Rham cohomology group HkdR(M) := H k(A•(M), d) is a real vector space. We shall denote by [ω] ∈ HkdR(M) the class of ω ∈ ZkdR(M). By the de Rham theorems —see, for instance, [23]— if M is compact then HkdR(M) ' Hk(M,R), where the latter is the k-th singular cohomology group, which is a finite-dimen- sional real vector space depending only on the topology of M . The most important single fact about de Rham cohomology is the following proposition, often called the Poincare´ lemma. Proposition A.13. Suppose that U is a contractible manifold, i.e., for some x0 ∈ U , there is a smooth map f : [0, 1]×U → U such that f(0, x) = x0 and f(1, x) = x, for x ∈ U . Then H0dR(U) = R and HkdR(U) = 0 for k > 0. 3In fact, a complex can be formed from objects Cn and morphisms dn in any abelian category. 145 Proof. A contractible manifold has an atlas with a single chart (U, φ), so we can suppose that U ⊆ Rn, that U is star-shaped about x0, and that f(t, x) = (1 − t)x0 + tx. If ω ∈ Ak(U), let η := ι∂/∂t(f ∗ω) ∈ Ak−1([0, 1] × U). Now define hk : Ak(M) → Ak−1(M) by hk(ω) := ∫ 1 0 η dt. Then one can check that the maps hk form a “cochain homotopy”, i.e., that hk+1 ◦ d + d ◦ hk = id for each k > 0, and h1 ◦ d = id−x0. The triviality of the cohomology groups follows at once, since dω = 0 implies ω = d(hkω) for k > 0, and df = 0 implies f ≡ f(x0) for f ∈ A0(M). A.9 Volume forms and integrals Definition A.30. A volume form on an n-dimensional manifold M is a real n-form ν ∈ An(M) which is nonvanishing, i.e., νx 6= 0 in ΛnT ∗xM for all x ∈ M . A volume form need not exist; we say that M is orientable if one exists. We say that two volume forms µ, ν on M are equivalent if µ = fν for some f ∈ C∞(M) with f(x) > 0 for all x ∈ M . (This is clearly an equivalence relation.) An equivalence class for this relation is called an orientation on M ; a pair (M, ν) consisting of a manifold and a volume form ν in a given equivalence class is called an oriented manifold. In a local coordinate system on a chart domain U , we have ν = h dx1 ∧ · · · ∧ dxn with h ∈ C∞(U,R) nonvanishing. Under a change of local coordinates in the overlap of two chart domains, h is multiplied by a Jacobian factor, det[∂yi/∂xj]. Lemma A.14. A differential manifold M is orientable if and only if M has an atlas {(Uj, φj)} all of whose transition functions φi ◦ φ−1j have positive Jacobians. On an oriented manifold (M, ν), we therefore may and shall always choose an atlas such that in every local coordinate system we have ν = h dx1 ∧ · · · ∧ dxn with h > 0. (We say that the corresponding charts are “positively oriented”.) Proposition A.15. Let (M, ν) be an oriented manifold. Then there is a unique linear form∫ M : An(M)→ R, called the integral over M , such that if η ∈ An(M) vanishes outside the domain of a positively oriented chart (U, φ) with local coordinate system (x1, . . . , xn), and if η = f dx1 ∧ · · · ∧ dxn, then∫ M η = ∫ φ(U) f(x1, . . . , xn) dx1 . . . dxn, where the right hand side is a Lebesgue integral on Rn. Proof. Uniqueness of ∫ M follows from the change-of-variables formula for multiple Lebesgue integrals; existence follows by writing η = ∑ j fjηj where each ηj is supported in a chart domain and {fj} is a partition of unity. Lemma A.16. If (M, ν) and (N, ρ) are two oriented n-dimensional manifolds and if τ : M → N is an orientation-preserving diffeomorphism (i.e., τ ∗ρ is equivalent to ν), then ∫ M τ ∗η =∫ N η for all η ∈ An(N). 146 An n-form on M is closed, since An+1(M) = 0. If η = dζ is an exact n-form, with compact support in the domain of an oriented chart (U, φ), then∫ M η = ∫ U dζ = ∫ φ(U) ψ∗(dζ) = ∫ φ(U) d(ψ∗ζ), where ψ = φ−1 : φ(U)→ U . Since ψ∗ζ = ∑j gj dx1∧· ·j∨· ·∧dxn for some gj ∈ C∞(U) having compact supports in U , we conclude that d(ψ∗ζ) = (∑ j ∂gj/∂x j ) dx1 ∧ · · · ∧ dxn; therefore,∫ U dζ = 0 by the fundamental theorem of calculus. By a partition-of-unity argument, we obtain ∫ M dζ = 0 for any ζ ∈ An−1(M), so the integral vanishes on exact n-forms. (This is the “boundaryless” case of Stokes’ theorem.) In consequence, [η] 7→ ∫ M η is a well-defined linear form on HndR(M). 147 References [1] R. Abraham, J. E. Marsden and T. Ratiu, Manifolds, Tensor Analysis, and Applications, Springer, Berlin, 1988. [2] H. Araki, “Bogoliubov automorphisms and Fock representations of canonical anticom- mutation relations”, Contemp. Math. 62 (1987), 23–141. [3] M. F. Atiyah, K-Theory, Benjamin, New York, 1967. [4] M. F. Atiyah, R. Bott and A. Shapiro, “Clifford Modules”, Topology 3 (1964), 3–38. [5] M. F. Atiyah and I. M. Singer, “The index of elliptic operators. I”, Ann. Math. 87 (1968), 484–530. [6] M. F. Atiyah and G. B. Segal, “The index of elliptic operators. II”, Ann. Math. 87 (1968), 531–545. [7] F. A. Berezin, The Method of Second Quantization, Academic Press, New York, 1966. [8] M. Berger, P. Gauduchon and E. Mazet, Le Spectre d’une Varie´te´ Riemannienne, Lec- ture Notes in Math. 194, Springer, Berlin, 1971. [9] N. Berline, E. Getzler and M. Vergne, Heat Kernels and Dirac Operators, Springer, Berlin, 1992. [10] L. C. Biedenharn and J. D. Louck, The Racah–Wigner Algebra in Quantum Theory, Addison-Wesley, Reading, MA, 1981. [11] R. Bott, “Homogeneous vector bundles”, Ann. Math. 66 (1957), 203–248. [12] R. Bott and L. W. Tu, Differential Forms in Algebraic Topology, Springer, Berlin, 1982. [13] T. Bro¨cker and T. tom Dieck, Representations of Compact Lie Groups, Springer, Berlin, 1985. [14] J.-L. Brylinski, Loop Spaces, Characteristic Classes, and Geometric Quantization, Birk- ha¨user, Boston, 1993. [15] P. Candelas, “Lectures on Complex Manifolds”, in Superstrings and Grand Unification, T. Pradhan, ed., World Scientific, Singapore, 1988. [16] C. Chevalley, The Algebraic Theory of Spinors, Columbia Univ. Press, New York, 1954. [17] L. Conlon, Differentiable Manifolds: A First Course, Birkha¨user, Boston, 1993. [18] A. Connes, Noncommutative Geometry, Academic Press, London, 1994. [19] A. Connes, private communication. 148 [20] M. Crampin and F. A. E. Pirani, Applicable Differential Geometry, Cambridge Univ. Press, Cambridge, 1986. [21] L. Dabrowski and A. Trautman, “Spinor structures on spheres and projective spaces”, J. Math. Phys. 27 (1986), 2022–2028. [22] J. A. Dieudonne´, Treatise on Analysis, Vol. 2, Academic Press, New York, 1970. [23] J. A. Dieudonne´, Ele´ments d’Analyse, Vol. 9, Gauthier–Villars, Paris, 1982. [24] P. A. M. Dirac, “Quantized singularities in the electromagnetic field”, Proc. Roy. Soc. A 133 (1931), 60–72. [25] T. Dray, “The relationship between monopole harmonics and spin-weighted spherical harmonics”, J. Math. Phys. 26 (1985), 1030–1033. [26] G. B. Folland, “Harmonic analysis of the de Rham complex on the sphere”, J. reine angew. Math. 398 (1989), 130–143. [27] J. E. Gilbert and M. A. M. Murray, Clifford Algebras and Dirac Operators in Harmonic Analysis, Cambridge Univ. Press, Cambridge, 1991. [28] P. B. Gilkey, Invariance Theory, the Heat Equation, and the Atiyah–Singer Index The- orem, Publish or Perish, Wilmington, DL, 1984. [29] P. B. Gilkey, “The index theorem and Clifford modules”, in Forty More Years of Rami- fications: Spectral Asymptotics and its Applications, S. A. Fulling and F. J. Narcowich, eds., Dept. of Math., Texas A&M University, College Station, TX, 1991; pp. 265–317. [30] J. N. Goldberg, A. J. Macfarlane, E. T. Newman, F. Rohrlich and E. C. G. Sudarshan, “Spin-s spherical harmonics and ð”, J. Math. Phys. 8 (1967), 2155–2161. [31] J. M. Gracia-Bond´ıa and J. C. Va´rilly, “QED in external fields from the spin represen- tation”, J. Math. Phys. 35 (1994), 3340–3367. [32] S. Helgason, Differential Geometry, Lie Groups, and Symmetric Spaces, Academic Press, New York, 1978. [33] W. V. D. Hodge, The Theory and Applications of Harmonic Integrals, Cambridge Univ. Press, Cambridge, 1952. [34] R. Howe and E. C. Tan, Nonabelian Harmonic Analysis: Applications of SL(2,R), Springer, Berlin, 1992. [35] J. E. Humphreys, Introduction to Lie Algebras and Representation Theory, Springer, Berlin, 1970. [36] W. Klingenberg, Riemannian Geometry, de Gruyter, Berlin, 1982. 149 [37] A. W. Knapp, Representation Theory of Semisimple Groups, Princeton Univ. Press, Princeton, NJ, 1986. [38] B. Kostant, “Quantization and unitary representations”, in Lectures in Modern Analysis and Applications III, Lecture Notes in Mathematics 170, Springer, Berlin, 1970; pp. 87– 208. [39] H. B. Lawson and M. L. Michelsohn, Spin Geometry, Princeton Univ. Press, Princeton, NJ, 1989. [40] A. Lichnerowicz, “Spineurs harmoniques”, C. R. Acad. Sci. Paris 257A (1963), 7–9. [41] J. Milnor and J. D. Stasheff, Characteristic Classes, Princeton Univ. Press, Princeton, NJ, 1974. [42] C. Nash, Differential Topology and Quantum Field Theory, Academic Press, London, 1991. [43] E. T. Newman and R. Penrose, “Note on the Bondi–Metzner–Sachs group”, J. Math. Phys. 7 (1966), 863–870. [44] A. Pressley and G. Segal, Loop Groups, Clarendon Press, Oxford, 1986. [45] M. Reed and B. Simon, Methods of Modern Mathematical Physics, I: Functional Anal- ysis, Academic Press, New York, 1972. [46] M. A. Rieffel, “Morita equivalence for C∗-algebras and W ∗-algebras”, J. Pure Appl. Algebra 5 (1974), 51–96. [47] E. Schro¨dinger, “Eigenschwingungen des spha¨rischen Raumes”, Commentationes Pon- tificia Academia Scientiarum 2 (1938), 321–364. [48] G. B. Segal, “Equivariant K-theory”, Publ. Math. IHES 34 (1968), 129–151. [49] G. B. Segal, “The representation ring of a compact Lie group”, Publ. Math. IHES 34 (1968), 113–128. [50] J.-P. Serre, “Repre´sentations line´aires et espaces homoge´nes ka¨hle´riens des groupes de Lie compacts”, in Se´minaire Bourbaki, Expose´ 100, 1954. [51] I. M. Singer and J. Thorpe, Lecture Notes on Elementary Topology and Geometry, Scott–Foresman, Glenview, IL, 1967. [52] M. D. Spivak, A Comprehensive Introduction to Differential Geometry, 5 vols., Publish or Perish, Berkeley, 1979. [53] M. E. Taylor, Noncommutative Harmonic Analysis, Amer. Math. Soc., Providence, RI, 1986. 150 [54] J. C. Va´rilly and J. M. Gracia-Bond´ıa, “Connes’ noncommutative differential geometry and the Standard Model”, J. Geom. Phys. 12 (1993), 223–301. [55] M. Vergne, “Geometric quantization and equivariant cohomology”, in ECM: Proceedings of the First European Congress of Mathematics, A. Joseph, F. Mignot, F. Murat, B. Prum and R. Rentschler, eds., Progress in Mathematics 119, Birkha¨user, Boston, 1994; pp. 249–295. [56] N. R. Wallach, Harmonic Analysis on Homogeneous Spaces, Marcel Dekker, New York, 1973. [57] A. Weil, Introduction a` l’e´tude des varie´te´s ka¨hle´riennes, 2nd edition, Hermann, Paris, 1971. [58] R. O. Wells, Differential Analysis on Complex Manifolds, Springer, Berlin, 1980. [59] N. J. Wildberger, “Hypergroups, symmetric spaces and wrapping maps”, preprint, Syd- ney, 1995. [60] E. Witten, “Monopoles and four-manifolds”, Math. Res. Lett. 1 (1994), 769–796. [61] J. Wolf, “Essential selfadjointness for the Dirac operator and its square”, Indiana Univ. Math. J. 22 (1972), 611–640. 151