<<

Lectures on Complex Analytic

by L. Schwartz

Tata Institute of Fundamental Research, Bombay 1955 (Reissued 1963) Lectures on Complex Analytic Manifolds

by L. Schwartz

Notes by M.S. Narasimhan

No part of this book may be reproduced in any form by print, microfilm or any other means without written permis- sion from the Tata Institute of Fundamental Research, Co- laba, Bombay 5.

Tata Institute of Fundamental Research, Bombay 1955 (Reissued 1963) Contents

1 Lecture 1 1

2 Lecture 2 7

3 Lecture 3 11

4 Lecture 4 19

5 Lecture 5 25

6 Lecture 6 31

7 Lecture 7 35

8 Lecture 8 41

9 Lecture 9 47

10 Lecture 10 51

11 Lecture 11 57

12 Lecture 12 63

13 Lecture 13 69

14 Lecture 14 75

iii iv Contents

15 Lecture 15 79

16 Lecture 16 83

17 Lecture 17 91

18 Lecture 18 97

19 Lecture 19 105

20 Lecture 20 113

21 Lecture 21 119

22 Lecture 22 125

23 Lecture 23 131

24 Lecture 24 139

25 Lecture 25 145

26 Appendix 153 Lecture 1

Introduction 1 We shall first give a brief account of the problems we shall be consider- ing. We wish to define a complex analytic , V(n), of complex dimension n and study how numerous (in a sense to be clarified later) are the holomorphic functions and differential forms on this manifold. If V(n) is compact, like the , we find by the maximum principle that there are no non-constant holomorphic functions on V(n). On the contrary if V(n) = Cn, the space of n complex variables, there are many non-constant holomorphic functions. On a compact complex an- alytic manifold the problem of the existence of holomorphic functions is trivial, as we have remarked; but not so the problem of holomorphic differential forms. On a compact complex manifold there exist a finite number, say hp, of linearly independent holomorphic differential forms of degree p. We have to study the relation between these forms and the algebraic groups of the manifold i.e., the relation be- tween hp and the pth Betti-number bp of the manifold. It is necessary not only to study holomorphic functions and forms but also meromor- phic functions and meromorphic forms. For instance, it is necessary to go into Cousin’s problem which is a generalization of the problem of Mittag-Leffler in the plane. [Mittag-Leffler’s problem is to find in the plane meromorphic functions with prescribed polar singularities and po- 2 lar developments]. We may look for the same problem on manifolds and the relation between this problem and the topological properties of the

1 2 Lecture 1

manifold. In the case of Stein manifolds the problem always admits of a solution as in the case of the . We thus observe the great variety of problems which can be exam- ined. The study of these problems can be divided into three fundamental parts: 1) Local study of functions, which is essentially the study of func- tions on Cn. Immediately afterwards one can pose some general problems for all complex analytic manifolds.

2) The study of compact complex manifolds; in particular, the study of compact Kahlerian¨ manifolds.

3) The study of Stein manifolds. We shall examine some properties of compact Kahlerian¨ manifolds in detail. Compact Riemann surfaces will appear as special case of these manifolds. We shall prove the Riemann-Roch theorem in the case of a compact Riemann . We shall not concern ourselves with the study of Stein manifolds.

Differentiable Manifolds 3 To start with, we shall work with real manifolds and examine later the situation on a complex manifold. We shall look upon a complex analytic manifold of complex dimension n as a 2n-dimensional real manifold having certain additional properties. We define an indefinitely differentiable (C∞) real manifold of di- mension N. (We reserve the symbol n for the complex dimension). It is first of all a locally compact topological space which is denumerable at infinity (i.e., a countable union of compact sets). On this space we are given a family of ‘maps’, exactly in the sense of a geographical map. Each map is a homeomorphism of an of the space onto an open set in RN, the N-dimensional Euclidean space. [For instance, if N = 2, we can imagine the manifold to be the surface of the earth and the image of a portion of the surface to be a region on the two dimensional page on which the map is drawn]. We require the domains of these maps to Lecture 1 3 cover the manifold so that every point of the manifold is represented by at least one point in the ‘atlas’. We also impose a condition of coherence in the overlap of the originals of two maps. Suppose that the originals of the maps have a non-empty intersection; we get a correspondence in RN between the images of this intersection if we make correspond to each other the points which are images of the same point of the mani- fold. We now have a correspondence between two open subsets of RN and we demand that this correspondence should be indefinitely differen- tiable i.e., defined by means of N indefinitely differentiable functions of 4 N variables. Thus an N-dimensional C∞ manifold is a locally de- numerable at infinity for which is given a covering by open sets U j, and N for each U j a homeomorphism ϕ j of U j onto an open set in R such that the map 1 ϕ j ϕ− : ϕi(Ui U j) ϕ j(Ui U j) ◦ i ∩ → ∩ is indefinitely differentiable. We call the pair (U j,ϕ j) a map, and we say that the family of maps given above defines a differentiable structure on the manifold. We call the family of maps an atlas. If (U,ϕ) is a map, by composing ϕ with N the coordinate functions on R we get N-functions x1,..., xN on U; the functions xi are called the coordinate functions and form the local coordinate system defined by the map (U,ϕ). From a theoretical point of view it is better to assume that the atlas we have is a maximal or complete atlas, in the sense that we cannot add more maps to the family still preserving the compatibility conditions on the overlaps. From any atlas we can obtain a unique complete atlas con- taining the given atlas; we say that two atlases are equivalent or define the same differentiable structure on the manifold if they give rise to the same complete atlas. N We have defined a C∞ manifold by requiring certain maps from R N to R to be C∞ maps; it is clear how we should define real analytic or quasi-analytic or k-times differentiable (Ck) manifolds. N We shall be able to put on a C∞ manifold all the notions in R that 5 are invariant under C∞ transformations. For instance we have the notion N of a C∞ on a C∞ manifold V . Let f be a real valued function 4 Lecture 1

defined on VN. Let(U,ϕ) be a map; by restriction of f to U we get a function on U; by transportation we get a function on an open subset of RN, namely the function f ϕ 1 on ϕ(U) and we know what it means ◦ − to say that such a function is a C∞ function. We define f to be a C∞ function if and only if for every choice of the map (U,ϕ) the function 1 f ϕ− defined on ϕ(U) is a C∞ function. We are sure that this notion is ◦ N correct, since the notion of a C∞ function on R is invariant under C∞ transformation. [We can define similarly C∞ functions on open subsets of VN (which are also manifolds!)].

If x1,..., xN are coordinate functions corresponding to (U,ϕ) and ∂ f ∂ f a U we define ,..., to be the partial of ∈ ∂x1 a ∂xN a f ϕ 1 at ϕ(a). ◦ − On a Ck manifold we can define the notion of a C p function for p k. ≤ The space of differentials and the tangent space at a point. We have now to define the notions of tangent vector and differential of a function at a point of the manifold. We define the differential of a C1 function at a point a of VN. For a C1 function f on RN the differential at a point a is the datum of the ∂ f ∂ f 6 system of values ,..., . We make an abstraction of this in ∂x1 a ∂xN a the case of a manifold. Let U be a fixed open neighbourhood containing a. All C1 functions defined on this neighbourhood form a , in fact an algebra, denoted by Ea,U . We say that a function f Ea,U ∈ is stationary at a (in the sense of maxima and minima!) if all the first partial derivatives vanish at the point a. The notion of a function being stationary at a has an intrinsic meaning; for if the partial derivatives at a vanish in one coordinate system they will do so in any other coordinate system. Let S a,U denote the subspace of functions stationary at a. We define the space of differentials at a to be the quotient space Ea,U /S a,U. If f Ea,U its differential at a, (d f )a, is defined to be its canonical ∈ image in the quotient space. We can prove two trivial properties: its independence of the neighbourhood U chosen and the fact that the space of differentials is of dimension N. If we choose a coordinate system Lecture 1 5

(x1,..., xN) at a we have the canonical basis (dx1)a,..., (dxN)a for the space of differentials; in terms of this basis the differential of a function f has the representation ∂ f (d f ) = (dx ) . a ∂ i a xi a We now proceed to define the tangent space at a. If x is a vector in RN we have the notion of derivation along x. The best way to define this f (a + tx) f (a) is to define the of a function f along x as lim − , t 0 t → if it exists. If x is the unit vector along the xi-axis we have the partial ∂ derivation . Thus in RN a vector defines a derivation. In a mani- 7 ∂xi fold, on the other hand, we may define a tangent vector as a derivation. We may define a tangent vector as a derivation of the algebra Ea,U with 1 1 values in R , i.e., as a L : Ea,U R which has the differen- → tiation property: L( f g) = L( f )g(a) + f (a)L(g). But there will be some difficulty and we prefer to give a definition which is related directly to the space of differentials defined above. When we consider only C∞ functions these two definitions coincide. We define a tangent vector or a derivation at a to be a linear function Ea,U R which is zero on stationary functions. Once again this defini- → ′ tion is independent of U; moreover the tangent space is N-dimensional. ∂ If (x1,..., xN) is a coordinate system at a, are tangent vectors ∂xi a at a. These are obviously linearly independent. Let L be any tangent vector at a. We may write any C1 function f as ∂ f f = f (a) + (x a ) + g i i ∂ − xi a where g is stationary at a so that we have ∂ f ∂ L( f ) = L(x ) or L = L(x ) i ∂ i ∂ xi a xi a Thus corresponding to a coordinate system (x1,..., xN) we have the ∂ ∂ canonical base ,..., of the tangent space at a. ∂x1 ∂xN 6 Lecture 1

Thus at a point a on the manifold VN we have two N-dimensional vector spaces: the tangent space at a, Ta(V) and the space of differentials 8 at a, Ta∗(V). They are duals of each other and the duality is given by the scalar product

L, (d f )a = L( f ), L Ta(V), (d f )a T ∗(V). ∈ ∈ a As it is, we first defined the space of differentials at a point and then the tangent space at that point; as we shall see, it is better to think of the tangent space as the original space and the space of differentials as its dual space. Lecture 2

C∞ maps, diffeomorphisms. Effect of a map 9 We define a C∞ map from a C∞ manifold U of dimension p to a C∞ manifold V of dimension q (p and q need not be equal). Let Φ be a continuous map of U into V. We say that Φ is a C∞ map if, for every choice of maps (u,ϕ) in U and (v,ψ) in V such that u Φ 1(v) is non ∩ − empty, the map

1 1 ψ Φ ϕ− : ϕ(u Φ− (v)) ψ(v) ◦ ◦ ∩ → p q which is a map of an open subset of R into one in R , is a C∞ map. We can define a Ck map of one Cn manifold into another if k < n. If U, V, W are C manifolds and Φ; U V and Ψ : V W are C maps ∞ → → ∞ then the map Ψ Φ : U W is also a C map. A map Φ : U V (U, ◦ → ∞ → V C∞ manifolds) is said to be a C∞ isomorphism or a diffeomorphism if Φ is (1, 1) and both Φ and Φ′ are C∞ maps. p Let Φ be a C∞ map of a C∞ manifold U into another C∞ manifold Vq. Let a be a point of U and b = Φ(a). We shall now describe how Φ gives rice to a linear map of Ta(U) into Tb(V) and linear map of Tb∗(V) into Ta∗(U). Let us choose an open neighbourhood A of a and an open neighbourhood B of b such that Φ(A) B. If f is a Ck function on an ⊂ open subset of V, f Φ is also a Ck function on an open subset of U. ◦ We call f Φ the reciprocal image of f with respect to Φ and denote it 1 ◦ 1 by Φ− ( f ) or Φ∗( f ). Now let f Eb,B; the restriction of Φ− ( f ) to A 10 ∈ 1 belongs to Ea;A. We thus have a natural linear map Φ : Eb,B Ea,A. − → By this map functions stationary at b go over into functions stationary

7 8 Lecture 2

1 at a : Φ : S b,B S a,A. So Φ induces a linear map of Eb,B/S b,B into − → Ea,A/S a,A i.e., a linear map of Tb∗(V) into Ta∗(U). We denote this map 1 also by Φ− . We now give this map in terms of coordinate functions x1,..., xp at a and y1,..., yq at b. Suppose the map is given by y j = Φ j(x1,..., xp). Φ j are C∞ functions of (x1,..., xp). The reciprocal image of the func- tion f (y1,..., yq) is the function

(x1,..., xp) f (Φ1(x1,..., xp),..., Φq(x1,..., xp)). → q ∂ f Let (d f )b = (dy j)b be a differential at b; then j=1 ∂y j b q p Φ 1 ∂ f ∂ i Φ− (d f )b = (dx j)a . ∂y  ∂x  i=1 i b j=1 j a      In particular, ∂Φ j Φ 1(dy ) = (dx ) − j b ∂x i a i i a

So in terms of the canonical basis (dy1)b,..., (dyq)b for Tb∗(V) and the canonical basis (dx1)a,..., (dxp)a for Ta∗(U) the linear transformation Φ 1 : T (V) T (U) is given by the Jacobian matrix − b∗ → a∗

∂Φ j

∂xi a We have a Jacobian matrix only if we choose coordinate systems at a and b. 1 11 The map Φ− goes in the direction opposite to that of the map Φ; we 1 now give a direct transformation. Associated with the linear map Φ− : T (V) T (U) we have the transpose of this map Φ : Ta(U) Tb(V). b∗ → a∗ → This mapping is called the differential of the mapping Φ at a. By the definition of the transpose we have

Φ(L), (d f )b = L, (d( f Φ)a ◦ Lecture 2 9 where L Ta(U); that is, we have Φ(L)( f ) = L( f Φ) and this gives ∈ ◦ a direct description of the map Φ : Ta(U) Tb(V), because if L is a → derivation at a, the above formula defines Φ(L) as a linear form on Eb;B which is obviously a derivation. We shall hereafter refer to a differential as a tangent co-vector. Invariance of dimension We now prove the theorem of invariance of dimension: tow dif- feomorphic manifolds have the same dimension. If Φ is a diffeomor- phism of U onto V, and Φ(a) = b, a U, we have the linear maps ∈ Φ : Ta(U) Tb(V) and Ψ : Tb(V) Ta(U) where Ψ denotes the in- → → verse of the map Φ : U V. Since Φ Ψ= identity and Ψ Φ= identity, → ◦ ◦ the same relations hold for the associated linear transformations Φ and Ψ. Consequently Ta(U) and Tb(V) are isomorphic; so U and V have the same dimension. Tensor fields and differential forms We proceed to examine the question of tensor fields and differential forms. Let a be a point on the C∞ manifold V. An element of Ta(V) 12 is called a contravariant vector at a; an element of Ta∗(V) is called a covariant vector at a. A tensor at a of contravariant order p and covariant order q is an element of the tensor product p q T (V) T (V) .  a   a∗  ⊗ ∧     Thus any tensor of any kind that can be defined on a vector space can be put at a point on the manifold. p We call an element of the p th exterior power Λ Ta(V), a p-vector at p Λ p p a and an element of Ta∗(V) a p-covector at a. If ω is a p-covector ω p p p and ω q a q-covector ωΛq (the exterior product of ω and q ) is a p + q ∧ ω ω covector and we have p q q p ωΛ = ( 1)pq Λω. ω − ω The exterior algebra,

N p 0 = Λ Λ = Ta∗(V) Ta∗(V), ( Ta∗(V) R) 0 10 Lecture 2

over T (V) will be of particular interest in the sequel. ∧ a∗ All this we have at a point of the manifold. Now we consider the whole manifold. A vector field (contravariant) on V is a map which assigns to every point a of the manifold V a tangent vector θ(a) at a; θ is a map of V into the set of all tangent vectors of V such that θ(a) Ta(V). ∈ Similarly a p times contravariant, q times covariant tensor field on V is a map which assigns to every point a of V a tensor at a of contravariant order p and covariant order q. A scalar field is an ordinary real valued function on V. If we assign to every point a V a p-covector ω(a) at a ∈ we obtain a differential form ω of degree p on V. ω(a) is called the value 13 of the differential form at the point a. Associated with a differentiable function f we have a differential form of degree 1 which assigns to every point a the differential of f at a,(d f )a. We state some trivial properties of tensor fields and differential forms. All tensor fields of a particular kind (of contravariant order p and covariant order q) form a vector space. We can add tensors at each point and multiply the tensor at each point by a scalar). The differential forms of degree p(p = 0, 1,..., N) form a vector space. We can multi- p ply a differential form of degree p, ω, and a differential form of degree q p q q, and obtain a differential form, ωΛ of degree p + q; we have only ω ω p q to take at every point a the exterior product ω(a)Λ (a). Thus the space ω of all differential forms is an algebra. Lecture 3

“The Tensor Bundles” 14 We have different kinds of tensor’s attached at each point of a C∞ mani- fold VN. We shall organize the system of tensors of a specified type into a new manifold. We first consider the set of all the tangent vectors at all points of V. We denote this set by T(V). We define on T(V) first a topological structure and then a differentiable structure so that T(V) becomes a 2N- dimensional C∞-manifold. We do this by means of the fundamental system of maps defining the manifold structure on V. We choose a map (U,ϕ). Let x1,..., xN be the coordinate functions corresponding to this ∂ map. Let a U. In terms of the canonical basis a tangent vector ∈ ∂xi L at a has the representation ∂ L = ξ j ∂x j a N Let ϕ(a) = (a1,..., aN) R . We can represent L by the point (a1, 2N∈ ..., aN,ξ1,...,ξN) in R . The set of all tangent vectors at points a ∈ U are in (1, 1) correspondence with the space ϕ(U) RN. We carry × over the topology in ϕ(U) RN to this set (by requiring the map L × → (a1,..., aN,ξ1,...,ξN) to be a homeomorphism). This we do for every map (U,ϕ). We have now to verify that the topology we introduced is consistent on the overlaps. Let (W,ψ) be another map and y1,..., yN the corresponding coordi- nate functions. Let a U W and ψ(a) = (b1,..., bN). Let the com- 15 ∈ ∩ 11 12 Lecture 3

ponents of L with respect to the canonical basis corresponding to the map (W,ψ)be(η1, . . . , ηN). b1,..., bN are C∞ functions of (a1,..., aN). Further ∂ L = ηi( )a ∂yi ∂ = ξ j ∂x j a ∂y ∂ = ξ i j ∂x ∂y j i j a i a so that ∂y η = ξ i i j ∂x j j a

So the η are C∞ functions of (a1,..., aN,ξ1,...,ξN). Hence the map

(a1,..., aN,ξ1,...,ξN) (b1,..., bN, η1, . . . , ηN) →

is a C∞ map; likewise the map

(b1,..., bN, η1, . . . , ηN) (a1,..., aN,ξ1,...,ξN) →

is a C∞ map. In particular the maps are continuous and this proves that the topology we defined is consistent on the overlaps. Now the above reasoning shows also that we have in fact defined a C∞ structure on T(V). Of course the atlas we have given is not complete. This is a very special atlas; in fact any map of V gives rise to a map in this atlas. If we have a Ck-manifold V we can put on T(V) only a (k 1) − times differentiable structure; for, the expressions for the ηi in terms of 16 the ξi involve the first partial derivatives which are only (k 1) times − differentiable. In particular if V is a C1-manifold T(V) will be only a topological manifold. The set of all p-times contra variant and q-times covariant tensors on a C∞ manifold V can be made similarly into a C∞ manifold p q T(V) T (V)    ∗  ⊗         Lecture 3 13

p We also have the manifold of p-covectors ΛT ∗(V) which is of dimension + N N P . The set of all tangent covectors, not-necessarily homogeneous, Λ (i.e., the union of Ta∗(V), a V) can also be endowed with a C∞ struc- ∈ N ture: this manifold, ΛT ∗(V), is of dimension N + 2 . All these manifolds built from V have quite a special structure; they have the structure of a vector fibre bundle. We shall not define here precisely the concept of a fibre bundle. Let us consider, for example, T(V). If a V, Ta(V) is called the fibre ∈ over the point a. T(V) is the collection of the fibres, each fibre being a vector space of dimension N. Actually T(V) is partitioned into the fibres. T(V) is called the tangent bundle of V. T(V) is the bundle space and V the base space. There are two important maps associated with T(V). One is the canonical projection π : T(V) V which associates → to every tangent vector its origin: if L Ta(V), π(L) = a. (This map ∈ is the projection of the bundle onto the base). The other map gives the canonical imbedding of V in T(V); this map V T(V) assigns to every → point a V the zero tangent vector at a. We can now consider V as a ∈ submanifold of T(V). (This canonical imbedding is possible in any fibre 17 bundle in which the fibre is a vector space). A section of a fibre bundle is a map which associates to every point a in the base space a point in the fibre over a. A section of T(V) is a p vector field. A section of ΛT ∗(V) is a differential form of degree p. A tensor field of a definite type is a section of the corresponding tensor bundle.

C∞ Tensor fields and C∞ Differential forms

A C∞ vector field is an indefinitely differentiable section of T(V) i.e. it is a C map θ : V T(V) such that for a V, θ(a) Ta(V). A C ∞ → ∈ ∈ ∞ differential form is a C map ω : V ΛT (V) such that ω(a) ΛT (V) ∞ → ∗ ∈ a∗ for a V. A C tensor field is defined similarly. ∈ ∞ It is good to come back to the coordinate system and see how a C∞ vector field looks. Let θ be a C∞ vector field. Let (U,ϕ) be a map of V. N Let a U and ϕ(a) = (a1,..., aN) R . Let the components of θ(a) in ∈ ∈ 14 Lecture 3

terms of the canonical basis (with respect to (U,ϕ)) at a be ξ1,...,ξN.

ξi = ξi(a1,..., aN) N are functions of (a1,..., aN). Thus exactly as in R the vector field is given by N-functions of the coordinates of the origin of the vector. If θ is a C∞ vector field the map

(a1,..., aN) (a1,..., aN,ξ1,...,ξN) → 18 is a C∞ map and hence the ξi’s are C∞ functions of (a1,..., aN). Con- versely if for every choice of the map (U,ϕ) the ξi’s are C∞ functions of (a1,..., aN), θ is a C∞ vector field, since the ai’s are always C∞ func- tions of (a1,..., aN). It will be useful in particular to know how to recognize by means of the coordinate systems whether a given differential form is C∞. Let 2 us consider, for simplicity, a differential form of degree two, ω. Let (U,ϕ) be a map and x1,..., xN the corresponding coordinate functions. If a U,(dx1)a,..., (dxN)a is the canonical basis for T (V). Once we ∈ a∗ know the canonical basis for Ta∗(V) we also know the canonical basis 2 2 Λ Λ for its second exterior power Ta∗(V). The canonical basis for Ta∗(V) is (dxi)aΛ(dx j)a , i < j. In terms of this basis we write 2 ω(a) = ωi j(a)(dxi)aΛ(dx j)a. i< j

2 ωi j(a) are functions of a ω is a C differential form if and only if for ∞ every choice of the map (U,ϕ) the functions ωi j are C∞ functions. In terms of the canonical basis (dxi)a, we can write ω as follows:

ω = ωi j dxiΛdx j i< j This is not ‘abuse of language’; the expression has a correct meaning. ωi j is a function i.e., a differential form of degree zero. dxi is the differ- 19 ential of the function xi i.e., the differential form of degree one which Lecture 3 15 assigns to the point a the differential (dxi)a and we have already defined the exterior multiplication of two differential forms. It is quite remarkable that we can define a manifold structure on the set of all tangent vectors; a priori there was no relation between tangent spaces defined abstractly. Also the notion of a vector varying continu- ously in a vector space which itself varies is a priori extraordinary.

C∞ Differential forms

By a form we shall always mean a C∞ differential form. When we consider Ck differential forms we will state it explicitly. We recall some p fundamental properties of the forms: Let ω be a p-form. When we have p a map we have a representation of ω in terms of the canonical basis:

p = Λ Λ ω ω j1... jp dx j1 ... dx jp . j1<...< jp This is only a local representation; we do not, in general, have a global representation. The following are the principal properties of the forms.

1. Differentiability.

2. The linear structure: We can add two p-forms and multiply a form by a scalar. All p-forms form a vector space.

3. Algebra structure: We can multiply two forms ω and ω and ob- tain the form ωΛω. The multiplication satisfies the anti-commuta- tivity rule.

4. The reciprocal image of a form: 20

This is a very important notion. Let U and V be two C∞ mani- p folds, and Φ : U V a C∞ map. Suppose ω is a differential form →p of degree p on V. ω gives rise to a differential form of degree p 1 p on U, Φ− (ω), which we call the reciprocal image of ω by Φ. Let a U and Φ(a) = b. The value of ω at b is a p-covector, ω(b), ∈ 1 p at b. Φ− (ω) is the differential form which assigns to every point 16 Lecture 3

1 p a U the p-covector Φ− (ω(b)). It is seen easily that if ω is a C∞ ∈ 1 form Φ− (ω) is also a C∞ form. Any kind of covariant tensor field has a reciprocal image. How- ever it is in general impossible to define the direct image of con- travariant tensor field. For it may be that a point b V is the ∈ image of no point of U or the image of an infinity of points of U. Of course we can define the direct image of a contravariant tensor field when Φ is a diffeomorphism. One of the reasons for the utility of differential forms is that they have a reciprocal image. Another is the possibility of exterior differentiation. 5. Exterior differentiation: p To a given Ck differential form ω of degree p we associate a differ- p k 1 p ential form dω of degree p+1 which is of class C − ; dω is called p 21 the exterior differential or the coboundary of ω. The important point to be noticed is that the exterior differential of a differential form of degree p is of degree p + 1.

We define the exterior differentiation by axiomatic properties. We p restrict our attention to C∞ forms only. Let E denote the space of all p p C∞ p-forms on the manifold (if p , 0, 1,..., N, E = 0). Then d : E p+1 → E (p = 0, 1,..., N) is a map which satisfies the following properties: 1) The operation d is purely local: if two forms ω and ω coincide on an open set U, then dω = dω on U. − 2) d is a linear operation: d(ω + ω) = dω + dω d(λω) = λdω, λ a constant.

3) With respect to the structure of algebra d has the following prop- erty: p q p q p q d(ω ω) = dω ω + ( 1)pω dω ∧ ∧ − ∧ Lecture 3 17

(The sign ( 1)p in the second term is to be expected as no symbol − can pass over a p-form without taking the sign ( 1)p). − 4) d2 = 0; i.e., d(dω) = 0.

5) If f is a form of degree zero i.e., a scalar function, then d f is the ordinary differential of the function which associates to every point a the differential of f at a.

Lecture 4

Existence and uniqueness of the exterior differenti- ation. The DGA E (V). 22 We give a sketch of the proof of the existence and uniqueness of d. First we assume the existence and prove uniqueness. Since d is a local operation it is enough to reason on an open subset U of RN. Let ω be a N p form on U and x1,..., xN the coordinate functions in R . Then

ω = ω j ... j dx j ... dx j . 1 p 1 ∧ ∧ p j1<...< jp

Let d′ be an operation satisfying conditions [(1) - (5)]. Since d′ is linear it is sufficient to consider the form

ω = ω j ... j dx j ... dx j 1 p 1 ∧ ∧ p By the

d′ω = d′(ω j ... j dx j ... dx j ) 1 p 1 ∧ ∧ p = d′ω j ... j dx j ... dx j 1 p 1 ∧ ∧ p + ω j ... j d′(dx j ... dx j ) 1 p 1 ∧ ∧ p By property 5, d f = d′ f for a function f . So

d′(dx j ... dx j ) = d′(d′ x j ... d′ x j ). 1 ∧ ∧ p 1 ∧ ∧ p Using the product rule it is seen by induction that

d′(d′ x j ... d′ x j ) 1 ∧ ∧ p 19 20 Lecture 4

p = i 1 ( 1) − d′ x j1 ... d′ x ji 1 d′(d′ x ji ) d′ x ji+1 ... d′ x jp − ∧ ∧ − ∧ ∧ ∧ ∧ i=1

2 23 Since d′ = 0, this sum is zero. So

d′ω = dω j ... j dx j ... dx j . 1 p ∧ 1 ∧ ∧ p

This proves that d′ is unique. This formula also shows how we should try to define the operation d to prove the existence. We define d locally by this formula. Let (U,ϕ) be a map and x1,..., xN the corresponding coordinate functions. Let ω be a p form. In U, ω can be written as

ω = ω j ... j dx j ... dx j 1 p 1 ∧ ∧ p j<...<jp In U we define

dω = dω j ... j dx j ... dx j . 1 p ∧ 1 ∧ ∧ p j1<...< jp It can be verified that d has the properties 1 - 5 in U. It follows from this and the uniqueness theorem we proved above that d is defined intrinsi- cally on the whole manifold (If Ui and U j are of two maps and di, d j the exterior differentiations defined in Ui, U j respectively by the above formula, then di = d j in Ui U j by the uniqueness theorem). ∩ Let us consider some examples. In R3 we have 0, 1, 2 and 3 forms. If f is a zero form, ∂ f ∂ f ∂ f d f = dx + dy + dz. ∂x ∂y ∂z

24 In R3 the canonical basis for two forms is usually taken to be dy dz, ∧ dz dx and dx dy. Let A dx + B dy + C dz be a 1-form. ∧ ∧ ∂c ∂B d(A dx + B dy + C dz) = dy dz ∂y − ∂z ∧ ∂A ∂C ∂B ∂A + dx dx + dx dy ∂z − ∂x ∧ ∂x − ∂y ∧ Lecture 4 21

If A dy dz + B dz dx + C dx dy be a two form, then ∧ ∧ ∧ ∂A dA dy dz = dx dy dz. ∧ ∧ ∂x ∧ ∧ and ∂A ∂B ∂C d(A dy dz + B dz dx + C dx dy) = + + dx dy dz. ∧ ∧ ∧ ∂x ∂y ∂Z ∧ ∧ The exterior derivatives of 0, 1 and 2 forms correspond respectively to the notions of the gradient of a function, and curl and divergence of a vector field. The formula d2 = 0 corresponds to curl grad = 0 and div curl = 0. p p We make some remarks on the space E (V) of p forms on V. E (V) 0 is an infinite dimensional vector space. For instance E (V) is infinite di- mensional since we can construct a C∞ function taking prescribed val- ues at arbitrary finite number of points. The space

N p E (V) = E (V) 0 is an algebra. (The product of two C∞ forms is C∞). We shall later put on E (V) a topological structure so that it becomes a topological vector 25 space. E (V) is a graded algebra: it is decomposed into the homoge- p neous pieces E (V) and the multiplication obeys the anti-commutativity rule. E (V) has an internal operation d : E (V) E (V) by which a ho- → mogeneous element of order p is taken into a homogeneous element of order p + 1 with the properties d2 = 0 and

p q p q p q d(ω ω) = dω ω + ( 1)pω dω. ∧ ∧ − ∧ E (V) has the structure of what is called a graded algebra with a differ- ential operator (DGA). We consider the behaviour of d with respect to mappings of mani- folds. Let Φ : U V be a C map of the two manifolds U and V. → ∞ Then we have a mapping Φ 1 : E (V) E (U) Φ 1 is a − → − 22 Lecture 4

of the graded algebra E (V) into the graded algebra E (U). We shall now 1 prove that Φ− is a homomorphism of the DGA’s. We have to prove that 1 Φ− commutes with the coboundary operator d:

1 1 dΦ− (ω) = Φ− (dω)

where ω is a form on V. (Strictly speaking the ‘d’s are different). We 1 shall prove this with the minimum possible calculation. Since d and Φ− have a local character we may assume U and V to be open subsets of Euclidean spaces. Since Φ 1 : E (V) E (U) is a homomorphism of − → the algebras it is sufficient to prove the result for a system of elements which generate E (V). If x1,..., xN are the coordinate functions in V, dx1,..., dxN and the C∞ functions on V generate E (V). So we have 26 only to prove in the case when ω is a 0-form and when ω is a 1-form which is the differential of a function. Let f be a 0-form; the result we wish to prove is just the definition of the reciprocal image:

1 Φ− (d f ) = d( f Φ) by definition ◦ 1 = d(Φ− ( f )).

1 Let ω = d f , f being a function. dω = 0, so Φ− (dω) = 0. We have to 1 prove that dΦ− (d f ) = 0; but we have just proved that

1 1 1 1 Φ− (d f ) = d(Φ− ( f )), so that dΦ− (d f ) = d(dΦ− ( f )) = 0.

Change of Variables.

The reciprocal image of a map gives a good method of obtaining the formula for . Let us consider for instance polar co- ordinates in R3 : x = r Sin θ Cos ϕ, y = r Sin θ Sin ϕ, z = r Cos ϕ. We wish to express A dx dy in terms of polar coordinates. We have a map ∧ Φ : R3 R3 : Φ(r,θ,ϕ) = (x, y, z). We have to find the reciprocal image → of A dx dy with respect to Φ. Since Φ 1 preserves products, ∧ − 1 1 1 1 Φ− (A dx dy) = Φ− (A)Φ− (dx) Φ− (dy) ∧ ∧ = A(r Sin θ Cos ϕ, r Sin θ Sin ϕ, r Cos ϕ) Lecture 4 23

= (Sin θ Cos ϕdr + r Cos θ Cos ϕdθ r Sin θ Sin ϕdϕ) − = (Sin θ Sin ϕdr + r Cos θ Sin ϕdθ + r Sin θ Cos ϕdϕ) = A(r2 Sin θ Cos θdθ dϕ r Sin2 θdϕ dr) ∧ − ∧ (We use dr dr = 0, dr dθ = dθ dr etc). ∧ − ∧ ∧ Poincare’s´ Theorem on differential forms in RN.

We mention here without proof a theorem of Poincare.´ We say that a 27 p p p 1 p 1 p-form ω is exact if ω = d ω− where ω− is a (p 1) form. For p = 0, p − this implies, by convention, that ω = 0. Since d2 = 0, a necessary p p condition for ω to be exact is that dω = 0. Poincare’s´ theorem is that, p in RN, this condition is also sufficient for ω to be exact, provided p 1. ≥ Thus a necessary and sufficient condition for a p-form (p 1) in RN to ≥ be exact is that its is zero.

Lecture 5

Manifolds with boundary 28 The upper hemisphere (with the rim) is an example of a manifold with boundary. Here the boundary is regular. But in the case of a closed triangle the boundary has singularities at the vertices. We consider only manifolds with good boundary. The manifolds that we have defined earlier are special cases of manifolds with boundary and will be referred to as manifolds without boundary. An N-dimensional C∞ manifold with boundary is a locally compact space VN, countable at infinity, for which is given two kinds of maps with the following properties. A map of the first kind maps an open subset of V homeomorphically onto an open subset of RN, exactly as in the case of the ordinary manifolds. A map of the second kind maps an open subset of V homeomorphically onto an open subset of the half- space N (x1,..., xN), x1 0 , in R . { ≥ } The domains of the maps cover V. In the overlaps the maps are related by C∞ functions. (We calculate the x1-derivatives at points on the hy- perplane x1 = 0 in the positive direction). As in the case of ordinary manifolds we require the family of maps to be complete. The notions of C∞ functions, tangent vectors, differential forms etc. can be defined as in the case of manifolds without boundary. Let a V. An tangent, L, at a is defined to be a positive ∈ derivation: if f 0 is of class C1 in a neighbourhood of a and f (a) = 0, 29 ≥ then L( f ) 0 and there exists at least one function f such that f 0, ≥ ≥ 25 26 Lecture 5

f (a) = 0 and L( f ) > 0. An exterior tangent at a is a negative derivation: if f 0, f (a) = 0 then L( f ) 0 and there exists at least one f such that ≥ ≤ f 0, f (a) = 0 and L( f ) < 0. ≥ Suppose a is a point which is in the interior of the domain of a map of the first kind, say (U,ϕ). Let f be a function of class C1 in a neighbourhood of a with f 0 and f (a) = 0. Since f attains the ≥ minimum at a, by considering the function f ϕ 1 at ϕ(a) we find that ◦ − f is stationary at a. So L( f ) = 0 for any tangent vector L at a. So at a point which is in the interior of the domain of at least one map of the first kind there are no interior and exterior tangents. On the other hand, let a be a point which is mapped by a map of the second kind onto a point in the hyperplane x1 = 0: let (U,ϕ) be a map of the second kind at a,(x1,..., xN) the corresponding coordinate functions and ϕ(a) a point on the hyperplane x1 = 0. Then the tangent vector

∂ ∂ L = ξ1 + + ξN ∂x1 a ∂xN a

is an interior tangent vector at a if ξ1 > 0. (and exterior tangent vector if ∂ f 1 ξ1 < 0). For, if f 0 and f (a) = 0, L( f ) = ξ1 since the f ϕ− ≥ ∂x1 a ◦ function is stationary at ϕ(a) on the hyperplane x1 = 0; L( f ) 0 so if ≥ ξ1 > 0 and for the function f = x1(x1 0, x1(a) = 0), L(x1) = ξ1 > 0. ≥ 30 It follows that a point which is in the interior of the domain of a map of the first kind is never mapped by a map of the second kind onto a point in the hyperplane x1 = 0 and that a point which is mapped by a map of the second kind onto a point in the hyperplane x1 = 0 is not in the interior of a map of the first kind. A point which is in the domain of at least one map of the first kind is called an interior point. A point which is mapped by a map of the second kind into a point on the hyperplane x1 = 0 is called a boundary point. • • N 1 N 1 Let V − denote the set of boundary points of V. V − is called • N 1 the boundary of V. V − is an N 1 dimensional manifold without • − N 1 boundary; the maps of V − are given by the restriction of the maps of Lecture 5 27

• • N 1 N 1 the second kind to V − . Moreover the tangent space to V − at a point • a VN 1 can be canonically identified with a subspace of the tangent ∈ − space at a to VN.

Oriented manifolds.

Let EN be an N-dimensional vector space over real numbers. The space N ΛEN is one dimensional and is isomorphic to R, but not canonically. N To orient EN is to decide which elements of ΛEN should be considered positive and which negative. If A , 0 and B , 0 are two elements of N N ΛEN, we have A = aB, a a . The non-zero elements of ΛEN fall into two classes defined as follows: two elements B and A = aB belong to the same class if a > 0 and to opposite classes if a < 0. Selecting one of these two classes as the class of positive N-vectors is called orienting the vector space EN. A vector space for which a choice 31 of one of the two classes has been made is said to be oriented. If EN is N N N N N oriented we may orient E∗ in a natural way: ΛE∗ is the dual of ΛE ; N N we say that a non-zero element in ΛE∗ is positive if its scalar product N with any positive vector in ΛEN is positive. We can orient only vector spaces over ordered fields. N Let V be a C∞ manifold (with or without boundary). By an orienta- N tion at a point a V we mean an orientation of the tangent space Ta(V). ∈ VN is said to be oriented if we have chosen at every point of VN an ori- entation satisfying the following coherence condition. Let (U,ϕ) be any connected map (i.e., a map whose domain is connected) and a U. The ∈ differential of the map ϕ at a gives rise to an isomorphism of Ta(V) onto N N Tϕ(a)(R ), which can be identified with R itself. Since we have ori- N ented Ta(V) this isomorphism induces an orientation on R . We require this induced orientation on RN to be the same for every point a U. ∈ If a manifold can be oriented it is said to be orientable; otherwise, non-orientable. Not every manifold is orientable. For example the Mobius-band (with or without the boundary) is non-orientable. The dif- 28 Lecture 5

ficulty in non-orientable manifolds is this: Let (U,ϕ) be a connected map. We can always choose a coherent orientation in U. If(U′,ϕ′) is another connected map such that U U is non-empty we can extend ∩ ′ the orientation in U to U′. In non-orientable manifolds it so happens 32 that when we continue the orientation like this along certain paths and come back to U we arrive at the opposite orientation. We can give a description of the orientation which uses only the N maps. Suppose V is oriented. Let (U,ϕ) be a map and (x1,..., xN) the coordinate functions of (U,ϕ). Then this map determines a unique orientation in RN. If this induced orientation is not the canonical ori- entation of RN (canonical orientation in RN is the orientation for which N e1 ... eN > 0 where (e1,..., eN) is the canonical basis for R ) the ∧ ∧ map given by (x2, x1, x3,..., xN) induces the canonical orientation in RN. Thus it is possible to cover VN by the domains of maps which in- N duce on R the canonical orientation; if (U,ϕ) and (U′,ϕ′) are two maps which induce on RN the canonical orientation in RN the coherence maps 1 1 ϕ ϕ − and ϕ ϕ have a positive Jacobian. Conversely if we have ◦ ′ ′ ◦ − a covering of VN by the domains of maps all of whose coherence maps have a positive Jacobian, these maps determine an orientation of VN. • N N 1 N Let V be a manifold with boundary and V − its boundary. If V • N 1 is oriented, V − is canonically oriented: the N 1 vector eN 1 tangent • • − − N 1 N 1 to V − at a point a V − will be said to be positive if, e1 being an ∈ N N exterior tangent vector to V at a, the N vector e1 eN 1 tangent to V ∧ − at a is positive. We consider only orientable manifolds and assume further that a definite orientation has been chosen for the manifold.

Integration on a Manifold.

33 Let VN be an oriented manifold (with or without boundary). Let ω be a continuous N-form on VN which vanishes outside a compact set. With ω we can associate a real number, ω called the integral of ω on V V possessing the following properties: Lecture 5 29

1) Let K be a compact set outside of which ω is zero. Let U be an open set containing K. Then

ω = ω. U V (The orientation on U is the one induced from the orientation on V)

2) If ω and ω′ are two continuous N-forms vanishing outside com- pact sets,

ω + ω′ = ω + ω′; λω = λ ω, λ V V V V V a constant.

3) If Φ : VN WN is an orientation preserving diffeomorphism of → VN onto WN and ω a continuous N-form on WN, then

1 − Φω = ω. V W

4) If V is an open subset of RN with the canonical orientation of RN and ω = f dx1 ... dxN, f a continuous function vanishing ∧ ∧ outside a compact set,

ω = ... f dx1 ... dxN where the term in the right side denotes the ordinary Riemann integral of f .

It can be proved that these four properties determine the integral uniquely.

Lecture 6

Integration on chains. 34 Let V be an oriented C∞ manifold without boundary. We shall now N define integrals of p-forms on some kind of p-dimensional submanifolds of V. N An elementary p-chain on V is a pair (W, Φ) where W is a p-dimen- p p sional oriented C∞ manifold with boundary and Φ : W V is a C∞ p → N map which is continuous at infinity (i.e., the inverse image of every compact set of V is a compact set of W). To avoid certain logical diffi- culties we shall assume that all the manifolds W are contained in RN0 . A p p-chain on V is a finite linear combination of elementary p-chains with real coefficients. The p-chains evidently form a vector space over R. The support of a p-form ω is the smallest closed set outside which the form in zero. Thus it is the closure of the set of points a such that ω(a) , 0. The support of an elementary chain (W, Φ) is the image of p the map Φ. The support of an arbitrary chain

Γp = a1Γ 1 + + akΓ k p p Γ (where are distinct elementary chains) is defined to be the union of the supports of Γi, i = 1, 2,..., k. The support of a chain is always a closed set (since the image of every closed set by a map continuous at infinity is a closed set). p W Let ω be a p-form on V and Γp = p, Φ . Let us suppose that the 35 31 32 Lecture 6

p intersection of the supports of ω and Γp is compact. Then we define the p p integral of ω on Γp, ω, by the formula Γ p

p 1 p ω = Φ− ω. Γ W p p

1 p The integral on the right is defined as Φ− (ω) has compact support. If Γ = Γ Γ p ai pi ( pi elementary chains) is an arbitrary p-chain such that p the intersection of the supports of Γp and ω is compact define the integral p of ω on Γp by:

p p ω = ai ω Γ Γ p pi

Stockes’ Formula

W • Let (p, Φ) be an elementary p-chain. Let W denote the boundary of p 1 • − • W oriented canonically and Φ W the restriction of Φ to W . Then p |p 1 p 1 • • − − ( W , Φ W ) is a p 1 chain. We define this chain to be the boundary p 1 |p 1 − of− the p-chain− (W, Φ). We define the boundary of an arbitrary p-chain p by linearity. We denote the boundary of Γp by bΓp. If b is the operator which maps a chain to its boundary then b2 = 0, since the boundary of 36 a manifold with boundary is a manifold without boundary. p 1 Let Γp be a p-chain and ω− a p 1 form such that the supports of p 1 − Γp and ω− have a compact intersection. Stokes’ formula asserts that

p 1 p 1 ω− = d ω− . bΓp Γp Lecture 6 33

Currents

Currents were introduced by de Rham to put the chains and the forms in the same stock. Currents are generalizations of distributions on RN. Currents are related to differential forms just the same way distributions are to functions. p Let E (V) denote the space of p-forms on V. We shall introduce p p a topology in E (V) so that E (V) becomes a : p We say that the of forms ω j tends to zero as j , if for { } → ∞ every map (U,ϕ) and for every compact set K U each derivative of p ⊂ each coefficient of ω j [expressed with the help of the coordinate func- tions (x1,..., xN) of the map (U,ϕ)] tends uniformly to zero on ϕ(K) as j . We may simply describe this topology as the topology of uni- → ∞ form convergence of the “coefficients” of the forms along with all their derivatives on every compact subset of V. Convergence in the sense de- scribed above is called convergence in the sense of E or in the sense of C . ∞ p Let D denote the space of p-forms with compact support. It is diffi- p p cult to introduce a topology on D adapted to the C∞ structure. Let DK 37 denote the space of p-forms whose supports are contained in the com- p pact set K. We consider DK as a topological space with the topology p induced from E . A current, T, of degree p (or of dimension N p) is a linear form on N p N p − − − D the restriction of which to every D K (K compact) is continuous. If ϕ1 and ϕ2 are N p forms with compact supports −

T(ϕ1 + ϕ2) = T(ϕ1) + T(ϕ2)

T(λϕ1) = λT(ϕ1), λ a constant

N p if ϕ j D have their supports in the same compact set K and if ∈ − ϕ j 0 in the sense of C as j then T(ϕ j) 0 as j . → ∞ →∞ → →∞ We shall write T,ϕ instead of T(ϕ).

Lecture 7

Some examples of currents 38 N p N p 1) A C p-form ω defines a current of degree p. For every ϕ− D− ∞ ∈ we define p N p p N p ω, ϕ− = ω ϕ− ∧ V

p N p (ω ϕ− has compact support). ∧ We have to verify that if ϕ j is a sequence of N p forms whose { } − supports are contained in the same compact set K and if ϕ j 0 → in the sense of C then ω ϕ j 0. By using a partition of ∞ ∧ → V unity we may assume that K is contained in the domain of a map. Then continuity follows from well-known properties of Riemann integrals on RN. If a p-form defines the zero current it can be proved easily that the form itself is zero. (This is a consequence of the existence of numerous N p forms). We can therefore identify a p-form with − the current it gives rise to. More generally a locally summable p-form defines a current of degree p. A differential form of degree p is said to be locally summable if, for every compact set K contained in the domain U of a map (U,ϕ), the coefficients of the differential form (expressed p in terms of the map (U,ϕ)) are summable on ϕ(K). If ω is a locally

35 36 Lecture 7

N p summable p-form and ϕ− an N p form with compact support p N p − 39 the integral ω ϕ− can be defined. Then the scalar product ∧ V

p N p p N p ω, ϕ− = ω ϕ− . ∧ V

defines a current of degree p. It can be proved that if a locally summable p-form defines the zero current the differential form is zero almost every where. (Though we have no notion of a Lebesgue measure on a manifold, the notion of a set of measure zero has an intrinsic meaning. A set on the manifold will be said to be of measure zero if its image by every map has Lebesgue measure zero on RN). So there is a (1, 1) correspondence between the space of currents of degree p defined by locally summable p- forms and the classes of locally summable p forms, a class being the set of all p-forms almost everywhere equal to the same form.

2) The second example of a current is of quite a different character. An N p chain ΓN p defines a current of degree p. We define, for N −p − ϕ D− ∈ ΓN p,ϕ = ϕ − ΓN p − (Since ϕ has compact support ϕ is defined). ΓN p − In this case it can happen that a chain ΓN p is not the zero chain, − nevertheless the integral of every N p form on ΓN p is zero. − − For this reason we shall consider two chains equivalent if for any N p form with compact support the integrals of the form on the − 40 two chains are equal. There is a (1, 1) correspondence between these equivalence classes and the space of currents of degree p defined by N p chains. − 3) Currents of Dirac. Lecture 7 37

This is the generalization of Dirac’s distribution on RN. Let a be a fixed point on V. The Dirac N-current at the point a, δ(a), is defined by δ(a),ϕ = ϕ(a) where ϕ is a C∞ function with compact support. More gener- N p ally, let X− be a fixed N p vector tangent to the manifold at a ϕ − being an N p form with compact support we define − N p N p p − = − δ N p , ϕ X ,ϕ(a) ( X− ) (ϕ(a) is the value of the form at a). The scalar product on the right N p N p Λ− Λ− is given by the duality between Ta(V) and Ta∗(V). This defines a current of degree p. This is called a Dirac current of degree p.

Partition of Unity

Suppose Ωi is an open covering of V. Then there exists a system of { } C scalar functions αi defined on V such that ∞ { } (i) αi 0 ≥ (ii) support of αi Ωi ⊂ (iii) αi are locally finite i.e., only a finite number of supports of αi 41 { } meet a given compact set. (All functions αi except a finite number vanish on a given compact set)

(iv) αi = 1 (This sum is finite at every point by condition (iii)). [If Ωi are relatively compact, then αi have compact supports.] The functions αi constitute a partition of unity subordinate to the covering Ωi a partition of the function 1 into non-negative C func- { }− ∞ tions with small supports. This theorem on partition of unity proves the existence of numerous non-trivial C∞ functions and forms on V. 38 Lecture 7

Support of a current

Let T be a current. We say that T is equal to zero in an open set Ω if T,ϕ = 0 for every form ϕ with compact support contained in Ω. Suppose Ωi is a system of open sets and Ωi = Ω. If a current T ∪i is zero in every Ωi then T = 0 in Ω. For, let ϕ be a form with compact support contained in Ω. Applying the theorem on partition of unity we can decompose the form ϕ into a finite sum of forms having supports in Ωi, as the support of ϕ is compact:

ϕ = αiϕ = ϕi, Supp ϕi Ωi. ⊂ consequently T,ϕ = T,ϕi = 0. 42 This result shows that there exists a largest open set in which a cur- rent T is zero, namely the union of all open sets in which T is zero. (This set may be empty). The complement of this set will be called the support of T. The support of the current defined by a form (C∞)ω coin- cides with the support of ω; the support of the current defined by a chain is not always identical with the support of the chain.

Main operations on currents

1) Addition of two currents and multiplication of a current by a scalar:

If T1 and T2 are two currents of degree p we define T1 + T2 and λT1,(λ a constant) by:

T1 + T2,ϕ = T1,ϕ + T2,ϕ T,ϕ = T,ϕ 2) Multiplication of a current by a form Just as in the case of distributions, we can not multiply two cur- p rents. However we can multiply a current by a form. If ω is a p Lecture 7 39

q N p q form and α a q form, we have for ϕ −D− ∈ ω α,ϕ = (ω α) ϕ ∧ ∧ ∧ V = ω (α ϕ) ∧ ∧ V = ω,α ϕ ∧ and ω α,ϕ = ( 1)pq α ω,ϕ . ∧ − ∧ p q Now for any current T of degree p and any q-form α we define 43 p q T α by: ∧ p q N p q p q N p q N p q N p q T α, −ϕ− = T, α −ϕ− , −ϕ− −D− ∧ ∧ ∈ q p p q We define α T as ( 1)pqT α. ∧ − ∧ 3) The coboundary of a current. Suppose ω is a p form and ϕ and N p 1 form with compact − − support. Since V is a manifold without boundary Stokes’ formula yields d(ω ϕ) = 0 ∧ V But d(ω ϕ) = dω ϕ + ( 1)pω dϕ so that ∧ ∧ − ∧ dω ϕ = ( 1)p+1 ω dϕ ∧ − ∧ V V or dω,ϕ = ( 1)p+1 ω, dϕ − Now for any current T of degree p we define the coboundary dT, which is a current of degree p + 1, by:

N p 1 + d, T,ϕ = ( 1)p 1 T, dϕ , ϕ −D− − ∈ 40 Lecture 7

Let D′(V) denote the direct sum of the spaces of currents of de- gree 0, 1,..., N D (V) is a graded vector space. d is a linear map ′ d : D (V) D (V) which raises the degree of every homoge- ′ → ′ neous element by 1. Moreover d2 = 0. For, T being a current of N p 2 degree p we have, for ϕ −D− , ∈ dd T,ϕ = ( 1)p+2 dT, dϕ − = ( 1)p+1( 1)p+2 T, ddϕ − − = 0.

44

Let us consider the coboundary of a current given by a chain ΓN p − p+1 dΓN p,ϕ = ( 1) Γ, dϕ − − = ( 1)p+1 dϕ − Γ

= ( 1)p+1 ϕ by Stokes’ formula, − Γ = ( 1)p+1 bΓ,ϕ . − So the coboundary of a current given by a chain is the current given by the boundary of the chain, but for sign. Thus the coboundary operator of differential forms and the boundary operator of chains appear as particular cases of the coboundary operator of currents. If T is a current of degree p and α a form, then

d(T α) = dT α + ( 1)pT dα. ∧ ∧ − ∧ Lecture 8

Currents with compact support 45 If T is a current of degree p and ϕ an N p form with arbitrary support − and if the supports of T and ϕ have compact intersection then T,ϕ can be defined. In particular if T has compact support T,ϕ can be defined for any C∞N p form. With this definition of T,ϕ T becomes a contin- − N p uous linear functional on E− . (i.e., a current of degree p with compact N p support can be extended to a continuous linear functional on E− ). Con- N p versely a continuous linear functional L on E− defines a current T of N p degree p by restriction to D− . It can be easily shown that this current has compact support and that

N p T,ϕ = L(ϕ) for every ϕ D − ∈ Consequently the space of p-current s with compact supports is identical p N p − with the dual space E ′ of E .

Cohomology spaces of a complex

A complex, E, is a graded vector space with a differential operator of degree 1: i) E is a vector space (over R) which is the direct sum of sub-spaces E p where p runs through non-negative (sometimes, all) integers.

41 42 Lecture 8

ii) E has a coboundary operator: there exists an endomorphism d : E E such that dE p E p+1 and d2 = 0. → ⊂ 46 The elements of E p are called elements of degree p. An element ω E is said to be a cocycle (or closed) if dω = 0. An ∈ element ω E is said to be a coboundary if there exists an element ω ∈ such that dω = ω. Let Z denote the vector space of cocycles and B the space of coboundaries. Since d2 = 0, B Z. The space Z/B = H (or ⊂ H(E)) is defined to be the cohomology vector space of E. Let Z p denote the space of cocycles of degree p and Bp the space of coboundaries of degree p. The space H p is called the pth cohomology vector space of E and the dimension of H p, bp, is called the pth Betti number of the complex. We have H = H p If a complex E is an algebra with respect to a multiplication ( ) ∧ satisfying the conditions.

i) E p Eq E p+q ∧ ⊂ pq p q ii) ω1 ω2 = ( 1) ω2 ω1, ω1 E , ω2 E . ∧ − ∧ ∈ ∈ p p q iii) d(ω1 ω2) = dω1 ω2 ( 1) ω1 dω2, ω1 E , ω2 E . ∧ ∧ − − ∧ ∈ ∈ is called a differential graded algebra. (D.G.A). If E is a D.G.A., H(E) can be endowed with the structure of an algebra. For Z is a subalgebra of E and B is a two sided ideal of Z. H(E) is known as the cohomology algebra of E.

Cohomology on a Manifold

Associated with a manifold VN we have a number of complexes and the corresponding cohomology groups.

47 i) E (V) = E p(V), where E p(V) is the space of all p forms on V. ii) E m(V) = E pm , where E pm denotes the space of m-times differ- p p ff entiable p-forms, ω, for which dω is also m-times di erentiable. Lecture 8 43

p p iii) E ′(V) = E ′(V), where E ′(V) is the space of currents of degree p with compact support.

p p N p m = m m − m iv) E ′ (V) E ′ (V) where E ′ (V) is the dual of E v) D(V) = D p(V), where D p(V) is the space of p forms with compact support.

p p vi) Dm(V) = Dm(V). where Dm′ (V) is the space of m times dif- ferentiable forms, ω, of degree p with compact support for which dω is also m times differentiable.

p p vii) D′(V) = D′(V), where D(V) is the space of currents of degree p.

p p N p m = m m − m viii) D′ (V) D′ (V), where D′ (V) is the dual space of D (V). Betti-numbers of E (RN), E (S N) and E (T N).

We shall examine the Betti numbers of RN, the N-sphere S N, and the N- T N.

p i) RN. By Poincare’s´ theorem a closed p-form ω is a coboundary if p 1. ≥ So 48 bp(E (RN)) = 0 if p 1. ≥ N If for a zero form f (i.e., a C∞ function f on R ) d f = 0, f is N N constant on R since R is connected. By convention B0(E (V)) (the space of boundaries of degree 0)= 0. So b0(E (RN)) = 1. 44 Lecture 8

ii) N-sphere S N (The set of points on RN+1 defined by the equation x2 + + x2 = 1). 1 N+1 Since S N is connected b0(E (S N)) = 1. It can be shown that

bp(E (S N)) = 0 for 1 p N 1 ≤ ≤ − and bN(E (S N)) = 1.

iii) Torus T N. (T N = RN/ZN, where ZN is the group of the integral lattice points in RN). It can be proved that

N bp(E (T N)) = . p In this case we can give the complete structure of the cohomol- N ogy ring. The differential forms dx1,..., dxN on R define N- N differential forms on T , which we still denote by dx1,..., dxN. It turns out that the classes of the forms

dxi ... dxi , i1 <...< ip 1 ∧ ∧ p 49 generate the pth cohomology group of T N.

We have seen that b0(E (RN)) = 1 and bp(E (RN)) = 0, p 1. The ≥ Betti numbers of D(RN) are not the same as those of E (RN). There is a theorem, which we may call Poincare’s´ theorem for compact supports, which asserts that in RN a closed p-form with compact support is the coboundary of a p 1 form with compact support if p N 1 and an N − ≤ − N-form ω with compact support is the coboundary of an N 1 form with − compact support if and only if

ω = 0 RN Lecture 8 45

N (The integral is defined since ω has compact support). It follows that

bp(D(RN)) = 0, 0 p N 1 ≤ ≤ − and it can be shown that

bN(D(RN)) = 1.

This example shows that there are at least two different kinds of co- homologies on a manifold - the cohomology with compact supports and cohomology with arbitrary supports. One part of de Rham’s theorem asserts that, of the cohomologies given by

m m m m E , E , D′ , D′; D, D , E ′ , E , only two cohomologies are distinct.

Lecture 9

Topology on E , E ′, D, D ′ 50 On E ′ we introduce the . The space D and D′ are in duality. In each of the spaces D and D′ we introduce the weak topology (which is a locally convex topology) defined by the other; the topologies on D and D′ are Hausdorff. We remark that if F is a topological vector space and F′ its dual with the weak topology, then the dual of F′ is F. de Rham’s Theorem

The first part of de Rham’s theorem. The first part of de Rham’s theorem gives canonical isomorphisms m m between the cohomology vector spaces of E , E , D′ and D′ and also canonical isomorphisms between the cohomology vector space of D, m m D , E ′ and E ′ . For instance, let us consider E (V) and D′(V). We define the canoni- p p p p cal isomorphism between H (E (V)) and H (D′(V)). If ω1 E (V) with p p ∈ dω1 = 0 it determines a closed current, ω1 of degree p (as the cobound- ary operator for currents is an extension of the coboundary operator for p p p 1 forms). If ω1 and ω2 are cohomologous (i.e., there is a p 1 form ω p p p 1 p p − with ω1 ω2 = d ω− ) the currents ω1 and ω2 are also cohomologous. − p p So we have in fact a linear map of H (E (V) into H (D′(V))). The first part of de Rham’s theorem asserts that this mapping is an isomorphism. 51

47 48 Lecture 9

That this mapping is (1, 1) means that if a form considered as a current is the coboundary of a current it is also the coboundary of a differen- tial form. That the map is onto means that in each cohomology class of currents there exists a current defined by a differential form (that is, we have no other cohomology classes of currents than the ones given by closed differential forms). We call the cohomology spaces given by any one of the complexes m m E , E , D′, D′ the cohomology spaces with arbitrary supports. The co- m homology spaces given by any one of the complexes , m, , are D D E ′ E ′ called the cohomology spaces with compact supports. We shall denote p p by H (V) and Hc (V) the pth cohomology spaces with arbitrary supports p p and compact supports respectively and by b and bc the dimensions of p p H and Hc . In this connection we shall give an example of a natural homomor- phism which is not an isomorphism in general. We have an obvious linear map from H p(D(V)) to H p(E (V)). In general this map is neither p (1, 1) nor onto. It may happen that a p-form ω with compact support is the coboundary of some p 1 form but not the coboundary of a p 1 − − form with compact support (as in the case of RN, p = N); and it may happen that there are cohomology classes of p-forms which contain no forms with compact support. The second part of de Rham’s theorem: the theorem of closure. In each of the spaces D, E etc., the space of co-cycles is closed: for if ω j ω then dω j dω. The theorem of closure asserts that in each → → of these spaces the space of coboundaries is also closed.

52 Some consequences of the theorem of closure

Let F be a locally convex topological vector space and F′ its dual endowed with the weak topology. Suppose G is a linear sub-space of F. 0 Let G◦ be the subspace of F′ orthogonal to G (An element of G is a linear form T on F such that T,ϕ = 0 for every ϕ G). Let G be ∈ ◦◦ the subspace of F orthogonal to G0. G00 is the biorthogonal of G. It is a simple consequence of Hahn-Banach theorem that G00 = G. Similarly the biorthogonal of a sub-space G of F′ is G¯. Let further H and G be subspaces of F such that H G; suppose H ⊂ Lecture 9 49 is closed in F. It can be shown that H0/G0 is canonically the topological 0 0 dual of G/H and conversely. (H and G are subspaces of F′ orthogonal to H and G respectively). These general considerations along with the closure theorem lead to two interesting consequences. i) Orthogonality relation Let Z be the space of cocycles and B the space of coboundaries in N p D − . Let further Z′ be the space of cocycles and B′ the space of p coboundaries in D′ . It is trivial to see that Z′ is the orthogonal space of B. For if dT = 0 then

T, dϕ = dT,ϕ = 0, ± N p 1 conversely if T, dϕ = 0 for every ϕ −D− , ∈ dT,ϕ = T, dϕ = 0 ± N p 1 for every ϕ −D− so that dT = 0. It is also true that the or- 53 ∈ thogonal space of Z is B′. To prove this we first notice that the orthogonal space of B is Z. For, if ϕ Z ′ ∈ dS,ϕ = S, dϕ ± p 1 − as dϕ = 0; conversely if dS,ϕ = 0 for every S D′ then p1 ∈ S, dϕ = 0 for every S D− so that dϕ = 0. So the biorthogonal ∈ ′ of B′ is the orthogonal space of Z. But the biorthogonal of B′ is the closure of B′ and by the closure theorem B′ = B′. So the orthogonal space of Z is B′. Thus the orthogonal space of the space of cocycles is the space of coboundaries and the orthogonal space of the space of cobound- aries is the space of cocycles (in D and D′ and so on). ii) Poincare’s´ duality theorem By the general result on the topological vector spaces stated above p N p it follows that H = Z′/B′ is canonically the dual of Hc − = Z/B. 50 Lecture 9

Thus the pth cohomology vector space with arbitrary supports is canonically the (topological) dual of the (N p) th cohomology − vector space with compact supports. This is the Poincare´ duality theorem. N p N p Suppose H − (D(V)) = Hc − is finite dimensional. Since the N p N p b − topology on Hc − is Hausdorff it is the usual topology on R c . N p So the topological dual and the algebraic dual of Hc − are the p N p same. So H is the algebraic dual of Hc − . Consequently

p N p b = bc − .

p N p It is to be remarked that H and Hc − are not canonically isomor- phic but only canonically dual. Lecture 10

Some applications 54 Since a closed 0-form is a function which is constant on each connected component, b0 is the number of connected components of V. A closed 0-form with compact support is a function which is constant on each compact connected component and zero on each non-compact compo- nent. It follows that 0 = bc number of compact connected components. By Poincare’s´ du- ality theorem we have N = 0 = b bc number of compact connected components, N = 0 = bc b number of connected components. N N Let ω be an N-form with compact support. If ω is to be the cobound- ary of an N 1 form with compact support it is necessary and sufficient N − that ω, f = 0 for every closed 0-form (orthogonality relations). If V is connected, a closed 0-form is a . So, in case V is N connected, for ω to be the coboundary of an N 1 form with compact − support it is necessary and sufficient that

N ω = 0. V Similarly we can prove that a necessary and sufficient condition for N an N-form ω with arbitrary support to be the coboundary of an N 1 55 N − form is that integral of ω on every compact connected component of the manifold should be zero.

51 52 Lecture 10

The third part of de Rham’s Theorem

The third part of de Rham’s theorem states that if V is compact all the Betti numbers of V are finite. In the compact case there is no difference between cohomology with compact supports and cohomology with arbitrary supports. The pth and (N p) th cohomology spaces are canonically the duals of each other − and we have the duality relation for the Betti-numbers:

p N p b = b −

Riemannian Manifolds

An N-dimensional Euclidean vector space EN over R is an N dimen- sional vector space over R with a positive definite bilinear form (or what is the same, a positive definite quadratic form). There is a scalar product (x, y) between any two elements x and y of EN which is bilinear and which has the properties

(x, y) = (y, x), and (x, x) > 0 for x , 0.

N Let e1,..., eN be a basis of E and gi j = (ei, e j). If x = xiei and = N y yiei are two vectors of E we have (x, y) = gi j xiy j. N A C∞ manifold V is called a C∞ Riemannian manifold if on each tan- 56 gent space of VN we have a positive definite bilinear form such that the twice covariant tensor field defined by these bilinear forms is a C∞ ten- sor field. Thus at each tangent space of a Riemannian manifold we have a Euclidean structure. The condition that the tensor field defined by the bilinear forms is a C∞ tensor field may be expressed in terms of local coordinate systems as follows: for every choice of the local coordinate system x1,..., xN the functions ∂ ∂ gi j(a) = , ∂xi a ∂x ja Lecture 10 53

are C∞ functions in the domain of the coordinate system, the scalar product ( , ) being given by the bilinear form on Ta(V).

Riemannian structure on an arbitrary C∞ manifold

We shall now show that we can introduce a C∞ Riemannian structure on any C∞ manifold. This result is of importance because the results on a Riemannian manifold which are of a purely topological nature can be proved for any manifold by introducing a Riemannian structure.

Let V be a C∞ manifold. If (Ui,ϕi) is a map we define a positive definite quadratic form at each Ta(V), a Ui by transporting the fun- 2 + + ∈2 N damental quadratic form “(dx1 dxN)” at Tϕ(a)(R ) by means of N the isomorphism between Ta(V) and Tϕ(a)(R ) given by the differential of the map ϕ at a. Let Qi(a) denote the positive definite quadratic form in Ta(V) given by the map (Ui,ϕi). Let αi be a partition of unity sub- 57 { } ordinate to the Ui . For a V, we define a quadratic form in Ta(V) { } ∈ by: Q(a) = αi(a)Qi(a) (the summation being over all Ui containing ff a; only a finite number of αi(a) are di erent from zero). Q(a) is a posi- tive definite quadratic form as the Qi(a)s are positive definite, αi(a) 0 ≥ and at least one αi(a) , 0. Since the αi are locally finite we can find a neighbourhood U of an arbitrary point of V such that

Q(a) = αi(a)Qi(a) (finite sum) for a U. ∈ where Qi(a) are C∞ quadratic forms on U. This proves that the quadratic forms Q(a) define a C∞ Riemannian structure on V. In the above argument we have made essential use of the positive definiteness of the quadratic form. The same construction would not succeed if we want to construct a C∞ indefinite metric with prescribed signature; for the sum of two quadratic forms with the same signature may not be a quadratic form with the same signature. In fact, we cannot put on an arbitrary C∞ manifold a C∞ indefinite metric with arbitrarily prescribed signature. 54 Lecture 10

Canonical Euclidean structures in p Λ Ta∗(V) and Ta∗(V)

Let VN be a Riemannian manifold. The positive definite quadratic form in Ta(V) defines a canonical isomorphism of Ta(V) onto Ta∗(V). For a fixed y Ta and any x Ta,(x, y) is a linear form on Ta. We denote this ∈ ∈ 58 linear form by γ(y). The linear map γ : y γ(y) is an isomorphism of → Ta onto Ta∗. We have the relation

(x, y) = x,γ(y)

connecting the Euclidean structure on Ta and the duality between Ta and Ta∗. A Euclidean structure in Ta defines a canonical Euclidean structure in Ta∗(V); we simply transport the positive definite quadratic form in Ta to Ta∗ by means of the canonical isomorphism γ. [Any linear map of Ta∗ to Ta defines a canonical bilinear form in Ta∗; the canonical bilinear form in Ta∗ may also be defined as the bilinear form given by the map 1 γ : T Ta. The quadratic forms in Ta and T are called inverses of − a∗ → a∗ each other. If (gi j) is the matrix of the quadratic form in Ta with respect to a basis in Ta the matrix of the canonical quadratic form in Ta∗ with 1 respect to the dual base is the matrix (gi j)− . A Euclidean Structure in Ta also defines canonical Euclidean struc- p p Λ Λ tures in Ta and Ta∗. Let e1,..., eN be an orthonormal basis in Ta (with respect to the quadratic form defining the Euclidean structure). Now, a positive definite quadratic form is uniquely determined if we specify a basis (x1,..., xN) as a system of orthonormal basis; the matrix of the quadratic form with respect to this basis is the identity matrix. We p take in ΛTa the positive definite quadratic form for which the elements ei ... ei , i1 <...< ip form an orthonormal basis. This quadratic 1 ∧ ∧ p 59 form ( , ) is intrinsic; for we have (x1 ... xp, y1 ... yp) ∧ ∧ ∧ ∧

(x1, y1) ... (x1, yp) ..... (xp, y1) ... (xp, yp) Lecture 10 55 for xi, yi Ta and on the right is intrinsically defined. ∈ The canonical isomorphism γ : Ta Ta∗ has a canonical extension p p → Λ Λ γ : Ta Ta∗, which is also an isomorphism. This isomorphism → p Λ defines the canonical Euclidean structure in Ta∗.

Lecture 11

The star operator 60 The star operator is defined for a Euclidean oriented vector space; this operator associates to every p-vector an (N p) vector. − Let us consider Ta∗(V) with the canonical Euclidean structure given by the Riemannian structure on V. Let Ta∗(V) be oriented. We first de- N fine the star ( ) operator on 0 - vectors i.e., scalars. ΛT (V) is a one ∗ a∗ dimensional space in which the class of positive vectors has been cho- N N Λ Λ sen. Ta∗(V) has a canonical Euclidean structure. Let τ Ta∗(V) be ∈ p p the unique positive vector of length 1. We define 1 = τ. For β ΛT ∗ ∈ a∗ we define β as the (N p) vector which satisfies the relation ∗ − p p p (α, β) = α ( βp). ∧ ∗ p p for every α ΛT . There exists one and only one element with this ∈ a∗ property. We choose an orthonormal basis e1,..., eN in Ta∗ such that e1 ... eN > 0 and define on the basis elements ei1 ... eip (i1 < ∧ ∧ p ∗ ∧ ∧ Λ ...< ip) of Ta∗ by: = (ei1 ... eip ) ǫek1 ... ekN p ∗ ∧ ∧ ∧ ∧ − where k1,..., kN p are the indices complementary to i1,..., ip and ǫ is the sign of the permutation− 1 2 ...... N i1 i2 ... ipk1 ... kn p − 57 58 Lecture 11

p 61 we define on the whole of ΛT by linearity. It is immediate that this is ∗ a∗ the only operation having the property

p p p p (α, β) = α ( β) ∧ ∗ We write 1β = ( 1)p(N p) β. It is easily verified that ∗− − − ∗ p p(N p) β = ( 1) − β. ∗∗ − p N p The operator gives an isomorphism of ΛT onto Λ− T . This iso- ∗ a∗ a∗ morphism carries an orthonormal basis into an orthonormal basis and hence preserves scalar products. If A and B are two vectors in R3 with the natural orientation, (A B) ∗ ∧ ( operation with respect to the natural Riemannian structure in R3) is ∗ what is usually called the vector product of A and B. In R2 the star operation for vector is essentially rotation through an angle π/2.

The star operator on differential forms

We suppose that VN is a oriented Riemannian manifold. The operation p ∗ Λ is then defined on each Ta∗(V). Suppose now that ω is a differential form of degree p. By taking at every point a the N p covector ω(α) we get a differential form of − ∗ degree N p, which we denote by ω. If ω is a C p-form ω is a − ∗ ∞ ∗ C (N p) form. In particular we have the N-form 1. This N form, ∞ − ∗ denoted by τ, defines the volume element on the Riemannian manifold. 62 If x1,..., xN is a local coordinate system for which dx1 ... dxN > 0 ∧ ∧ then τ = √g dx1 ... dxN, ∧ ∧ where √g is the positive root of the determinant g of the matrix (gi j). The star operator on differential forms, defined above, gives an iso- p N p morphism, called the star isomorphism, between E and E− and also p N p between D and D− . Lecture 11 59

The global scalar product of two C∞-forms

We define the scalar product of two p-forms α and β when the support of either α or β is compact by:

(α,β) = (α,β)aτ V

((α,β)a is the scalar product of the p-covectors α(a) and β(a) at a; τ is the volume element of the Riemannian manifold). We have

(α,β) = α β ∧∗ V = α, β ∗ 1 = − α,β ∗ The operator on currents ∗ Let T be a current. We define T by: ∗ 1 T,ϕ = T, − ϕ . ∗ ∗ We then have 63 1 − T,ϕ = T, ϕ . ∗ ∗ 1 The star operator on currents is the transpose of the operator − defined ∗ on forms with compact support.

The Riemannian scalar product of a p-current and a p-form

We define the scalar product of a current T of degree p and a p-form ϕ by: (T,ϕ) = T, ϕ ∗ if the support of either T or ϕ is compact. We shall call this scalar product the Riemannian scalar product between T and ϕ. 60 Lecture 11

The star coboundary operator ∂

We define the operator ∂ (“del”) on currents. If T is a current of degree p we define ∂T, a current of degree p 1, by: − p 1 p 1 (∂T, ϕ− ) = (T, d ϕ− )

∂ is the adjoint of the operator d with respect to the Hilbertian structure defined by the Riemannian scalar product. We have

(∂T p,ϕ) = (T, dϕ) 1 = − T, dϕ ∗ N p 1 1 = ( 1) − − d − T,ϕ − ∗ N p 1 1 1 = ( 1) − − − d − T,ϕ − ∗ ∗ ∗ N p+1 1 = ( 1) − ( d − T,ϕ) − ∗ ∗

64 so that p N p+1 1 ∂T = ( 1) − d − T − ∗ ∗ and p p 1 ∂T = ( 1) − d T. − ∗ ∗ From d2 = 0 it follows that ∂2 = 0. The operator ∂ defines a new differential graded structure in D′. The operator is also defined for the spaces E , E ′ and D. We now have a new cohomology, the cohomology. ∗ This cohomology is not different from the cohomology that we al- ready have. Consider, for instance, E . We define an isomorphism be- N p p tween H − (E ) and H (E ). Suppose ω is an N p form with ∂ω = 0. ∂ d − 1 1 Then d − ω = 0. Since is an isomorphism d − ω = 0 or d ω = 0 ±∗ ∗ ∗ ∗ ∗ i.e., ω is closed. The mapping ω ω∗ induces an isomorphism be- ∗ N p p → tween H − (E ) and H (E ). ∂ d Lecture 11 61

The Laplacian ∆

We define the operator ∆ by:

∆= d∂ + ∂d. − This is a differential operator of the second order. ∆ preserves de- grees. Since d∂ and ∂d are self-adjoint, ∆ is self-adjoint:

(∆T,ϕ) = (T, ∆ϕ). and ∆d commute: ∆d = d∂d = d∆. ∆ and ∂ also commute. ∆ also commutes with : ∗ ∆ = ∆ . ∗ ∗

Lecture 12

The operator ∆ on functions 65 Let x1,..., xN be a local coordinate system such that dx1 ... dxN > 0; ∧ ∧ ∂ ∂ i j let gi j = , and (g ) the matrix inverse to the matrix (gi j). Let ∂xi ∂x j ω jdx j be a 1-form. We calculate ∂( ω jdx j) j

1 ∂ ω jdx j = − d ω jdx j   − ∗ ∗   j  j          Now ω jdx j = ω j dx j. Suppose ∗ ∗

dx j = ωi jdx1 ... d/xi ... dxN; ∗ ∧ ∧ ∧ ∧ i from the relation (dxk, dx j)τ = dxk dx j ∧∗ we obtain

k j k 1 g √gdx1 ... dxN = ( 1) − ωk jdx1 ... dxN ∧ ∧ − ∧ ∧ k 1 k j so that ωk j = ( 1) g √g. − − i 1 i j d ω jdx j = d ( 1) − ω jg √gdx1 ... d/xi ... dxN ∗    − ∧ ∧ ∧ ∧   j  i, j            63 64 Lecture 12

i 1 ∂ i j = ( 1) − (ω jg √g)dxi − ∂x i, j i

dx1 ... d/xi ... dxN ∧ ∧ ∧ ∧ ∧ ∂ i j = (ω jg √g)dx1 ... dxN ∂x ∧ ∧ i, j i 1 ∂ = (ω gi j √g)τ g ∂x j √ i, j i

1 66 For N forms − = . Hence ∗ ∗ 1 = 1 ∂ i j − d  ω jdx j (ω jg √g)τ ∗ ∗ √g ∂xi j  i, j     1 ∂ i j = (ω g √g) g ∂x j √ i, j i Finally = 1 ∂ i j ∂  ω jdx j (ω jg √g) − √g ∂xi j  i, j   We shall now calculate ∆u where u is a 0-form. We have ∂u ∂u = 0, ∆u = ∂du and du = dx j. − ∂x j By the calculation made above we find that 1 ∂ ∂u ∆u = gi j √g . g ∂x ∂x i, j √ i j This is the well-known Laplace operator on functions. In particular if VN = RN with the natural metric and the natural orientation, the matrix (gi j) is the unit matrix and

∂2u ∆u = . 2 i ∂xi Lecture 12 65

The elliptic character of ∆. Harmonic forms

In the case of a partial differential equation of the second order an elliptic operator is usually defined by considering the nature of the quadratic form given by the coefficients of the derivatives of the second order. However it is found more convenient to define an elliptic operator by intrinsic properties of the operator. This definition is valid for systems of 67 differential equations and also for differential equations of higher order. A local or differential operator D is defined to be a linear continuous operator on currents (D : D D ) taking forms into forms1 and hav- ′ → ′ ing the local character: DT (T a current) is known in an open set Ω if T is known in Ω. A differential operator D is called an elliptic operator2 if the following condition is satisfied: If T is a current such that DT = α is a C∞ form, in an open set Ω then T itself is a C∞ form in Ω. If D is elliptic every solution of the homogeneous equation DT = 0 is a C∞ form. The operator

m m 1 d d − D = + a1 + + am, ai E dx dx ∈

∂2 ∂2 on the distributions in R is elliptic. In R2 the operator + is ∂x2 ∂y2 ∂2 ∂2 elliptic while the wave operator is not elliptic. The function ∂x2 − ∂y2 ψ(x, y) = f (x + y) + g(x y) where f and g are continuous is a solution − of the wave equation (as a distribution), even though the functions may not be differentiable. We shall admit without proof the important theorem which states that the operator ∆ is elliptic. From the elliptic character of ∆ we can deduce that d is elliptic on 68 0 0 0 D ∂ = 0 on D . So ∆ = ∂d on D . Now if D1 and D2 are differential ′ ′ − ′ operators and D1D2 is elliptic then D2 is an elliptic operator (but not

1Actually this condition is superfluous; it can be proved that it is a consequence of the other conditions. 2In current literature such differential operators are referred to as hypoelliptic oper- ators. 66 Lecture 12

necessarily D1). For let D2T = α be a C∞ form; since D1D2T = D1α, D1α is a C∞ form and D1D2 is elliptic it follows that T is a C∞ form. 0 Since ∆= d∂ is elliptic, d is elliptic on D . − ′ A form ω which satisfies the equation ∆ω = 0 is called a harmonic form.

Compact Riemannian manifolds

We shall assume henceforth that V is a compact, oriented Riemannian manifold. If V is compact every harmonic form is closed and closed. (This ∗ result is false when V is not compact; for example, a closed 0-form in RN is a constant while there exist non-constant harmonic functions). For let ω be a harmonic form

(∆ω,ω) = (d∂ω,ω) + (∂dω,ω) = (∂ω,∂ω) + (dω, dω).

Since (∂ω,∂ω) 0, (dω, dω) 0 and (∆ω,ω) = 0 it follows that ≥ ≥ (∂ω,∂ω) = 0 and (dω, dω) = 0. But if f is a continuous non-negative function such that f τ = 0 then f 0. Since the Riemannian scalar ≡ V product is positive definite it follows that

dω = 0 and ∂ω = 0.

The of square summable forms 69 We now define the Hilbert space H p of square summable differential forms of degree p. A form ω is said to be measurable if its coefficients are measurable on every map. An element of H is a class, a class consisting of all measurable forms which are equal almost every-where to a form ω for which (ω,ω) = (ω,ω)aτ is finite. If ω and ω are elements of H then (ω, ω) is V Lecture 12 67 defined and this defines a positive definite scalar product in H . We can prove that H is complete with respect to the given by the scalar product. So H is a Hilbert space.

Lecture 13

Intrinsic characterization of H 70 H , considered as a topological vector space, is intrinsically attached to the manifold i.e., H is independent of the Riemannian metric. We shall now give an intrinsic characterization of H . Suppose U is the domain of a local coordinate system x1,..., xN and K a compact set contained in U. Suppose ω is a measurable p form such that (ω,ω) < . Suppose ∞

ω = ωIdxI on U. (I is a system of p indices written in the increasing order). Let further IJ (ω,ω)a = g (a)ωI(a)ωJ(a). If m(a) is the smallest eigen value of the matrix (gIJ(a)), m(a) is a con- tinuous function in K and hence has a lower bound m1 in K; m1 > 0, since (gIJ(a)) are positive definite. Since,

IJ 2 g (a)ωI(a)ωJ(a) m(a) [ωI(a)] ≥ I,J I,J it follows that

IJ (ω,ω) g ωIωJ √gdx1 ... dxN ≥   I,J  K     69 70 Lecture 13

2 Cm1 ωI dx1 ... dxN ≥ | | K

where C 0 is the lower bound of √g in K. Thus, if (ω,ω) < , ≥ ∞

2 ωI dx1 ... dxN < . | | ∞ K 71 Conversely, suppose ω is a measurable p-form such that for every compact K contained in the domain U of a map

2 ωI dx1 ... dxN < ; | | ∞ K then (ω,ω) < . We choose a finite covering of the manifold by do- ∞ mains Uλ of maps and a partition of unity αλ subordinate to the cover- { } ing Uλ. Let Kλ be the support of αλ. Then, with the obvious notation,

(ω,ω) = (ω,ω)aτ V

= (ω,ω)aαλτ λ V

= (ω,ω)aαλτ λ Uλ = IJ (λ) (λ) (λ) (λ) (λ) (λ) g (a) ωI ωJ αλ √g dx1 ... dxN . λ Uλ

IJ (λ) Let Mλ(a) be the greatest eigenvalue of (g (a) ). The functions Mλ, (λ) αλ and √g are bounded in Kλ so that

(λ) 2 (λ) (λ) (ω,ω) cλ ωI dx1 ... dxN , Cλ a constant, ≤ | | Kλ < . ∞ Lecture 13 71

72 Thus the elements of H can be characterized as classes, a class being the set of all measurable forms almost everywhere equal to a form whose coefficients are square summable on every compact set contained in the domain of a map. A simple application of the Fisher-Riesz theorem would now show that H is complete. It may be remarked convergence in H implies convergence in D′.

Decomposition of H

We shall now decompose the space H into the direct sum of three fun- damental spaces which are mutually orthogonal. Let H1 be the subspace of the elements ω H such that ∆ω = 0 (in the sense of currents). It ∈ follows from the elliptic character of ∆ that H1 is exactly the space of harmonic forms. Since d and ∂ are continuous operators on currents, H1 is closed. In fact we shall see later that H1 is finite dimensional.

We define H2 to be the space of elements ω H such that (ω, ϕ) = 73 ∈ ∗ 0 for every ϕ D with ∂ϕ = 0. H2 is a closed subspace because of the ∈ continuity of the scalar product. We shall now give another interpreta- tion of the space H2. If ω H2 then ω, ϕ = 0 for every ϕ such that ∈ ∗ ∗ d( ϕ) = 0 or ω is orthogonal to all N p forms which are closed. By ∗ − the orthogonality theorem ω is the coboundary of a current. Now let H be the subspace of elements ω of H for which dω H (the H ’s have ∈ 72 Lecture 13

different degrees). Then d operates on H and this gives a cohomol- ogy. There is a natural homomorphism of H(H ) into H(D ). Another ′ de Rham’s theorem states that this homomorphism is an isomorphism onto. This implies, in particular, that if ω H is the coboundary of a ∈ current it is also the coboundary of an element ω H . So H2 is the ∈ space of forms of H which are coboundaries of forms of H : H2 = dH H . ∩ We define H3 to be the space of elements ω H such that (ω,ϕ) = ∈ 0 for every ϕ D with dϕ = 0. H3 is closed. If ω H3, ω H2 and ∈ ∈ ∗ ∈ therefore = ∂ω, ω H . So H3 is the space of forms of H which are ∈ star coboundaries of forms of H : H3 = ∂H H . ∩ 74 We shall prove that H1, H2 and H3 are mutually orthogonal . Sup- pose α H1 and β H2. Then ∈ ∈ (α,β) = (α, dω), ω H ∈ = (∂α, ω) since α is a C form ∞ = 0 as ∂α = 0. Similarly H1 and H3 are orthogonal. To prove that H2 and H3 are orthogonal we shall first prove that

H2 = dD and H3 = ∂D

Evidently dD H2 and as H2 is closed dD H2. If H2 , dD ⊂ ⊂ as H2 is a closed subspace we can find a vector λ H2, λ , 0 such ∈ that λ is orthogonal to dD or (λ, dϕ) = 0 for every ϕ D or (∂λ,ϕ) = 0 ∈ for every ϕ D. This implies that ∂λ = 0. Already dλ = 0 since λ is ∈ derived. So λ H1 and λ H2. As H1 and H2 are orthogonal λ = 0; ∈ ∈ this proves H2 = dD. Similarly H3 = ∂D. If α, β D ∈ (dα,∂β) = (ddα,β) = 0

i.e., dD and ∂D are orthogonal. Since dD = H2 and ∂D = H3 it follows that H2 and H3 are orthogonal. Lecture 13 73

We next show that H1, H2 and H3 span H . Suppose ω is orthog- onal to H2 and H3; then (ω, dα) = 0 for every α D or (∂ω,α) = 0 for ∈ every α D or ∂ω = 0; similarly dω = 0. Hence ω H1. This proves ∈ ∈ that H = H1 + H2 + H3.

We denote by π1, π2 and π3 the projections on H1, H2 and H3 respec- 75 tively. H1 + H2 is the orthogonal of H3 or ∂D, so H1 + H2 is the space of forms ω with dω = 0 (in the sense of the currents). H1 +H3 is the space of forms ω with ∂ω = 0 and H2 + H3 is the space of forms orthogonal to the space of harmonic forms.

Cohomology and harmonic forms. Hodge’s theo- rem

In every cohomology class of C∞ forms there exists one and only one harmonic form. Let ω be a representative of the cohomology class; ω H1 +H2. π1ω is a harmonic form and π2ω which is the coboundary ∈ of a current is the coboundary of a form by de Rham’s theorem. As

ω π1ω = π2ω − ω and π1ω are cohomologous. So π1ω is a harmonic form belonging to the class of ω. If ω1 and ω2 are two harmonic forms in the same cohomology class, ω1 ω2 H1 H2 = 0 or ω1 = ω2. Thus we have, − ∈ ∩ in fact, a canonical isomorphism between the cohomology space of C∞ forms and the space of harmonic forms. The dimension of the space of harmonic forms of degree p is the pth Betti number of the manifold which is finite by the third part of de Rham’s theorem. Suppose ω H2; then ω = dω, ω H . There is one and only 76 ∈ ∈ one choice of the primitive ω which belongs to H3. For, if ω = dθ, θ H put ω = π3θ; then ω = dω. If ω = dω1 = dω2, ω1, ω2 H3, ∈ ∈ d(ω1 ω2) = 0, which implies that ω1 ω2 H3 H2 so that ω1 = ω2. − − ∈ ∩ Similarly any element of H3 is the star coboundary of one and only one element of H2. In both the cases the choice of the unique primitive is linear but not yet continuous. 74 Lecture 13

Now let ω H2; ω = dθ, θ H3. But θ = ∂ω, ω H2. Conse- ∈ ∈ ∈ quently for every ω H2,ω = d∂ω, ω H2. ∈ ∈

This ω is unique. Every element ω ofH2 is thus expressed in one and only one way in the form ω = d∂ω, ω H2. ∈

However, for ω H2, ∂dω = 0 so that d∂ = ∆. So any form ω H2 ∈ − ∈ can be written uniquely in the form

ω = ∆ω, ω H2. − ∈

Similarly every form in H3 is uniquely expressible as the Laplacian of an element of H3. As a consequence every form orthogonal to all the harmonic forms is the Laplacian of one and only one element of H2 +H3. Conversely if ω is the Laplacian of an element, then it belongs to H2 + H3. 77 We have thus completely solved the equations dX = A, ∂X = A and ∆ X = A in H . The equation dX = A is solvable if A H2. If ∈ A H2 there exists a unique solution X H3. Any general solution ∈ ∈ is given by X = X0+ closed form. The equation ∂X = A is solvable if A H3 and there exists a unique solution X0 H2. A general solution ∈ ∈ is obtained by X = X0+ closed form. The equation ∆X = A is solvable if A H2 + H3; there exists a unique solution orthogonal to the space ∈ of harmonic forms and a general solution is obtained by adding to this particular solution any harmonic form. These results constitute Hodge’s theorem. Lecture 14

Green’s Operator G 78 Suppose A H2 +H3. We know that there exists a unique X H2 +H3 ∈ ∈ such that ∆X = A. We write X = GA; we then have an operator − G : A GA on H2 + H3 with the property ∆G = I on H2 + H3. (I → − is the identity operator). If α H2 + H3 and ∆α H2 + H3 we have ∈ ∈ G∆α = α. We extend G to the whole space H by putting Gα = 0 for − α H1. G : H H is an operator which is zero on H1 and which ∈ → leaves the spaces H2 and H3 invariant. We shall prove a little later that G is continuous. G is called the Green’s operator. Let ω H . We can write ω uniquely as ∈

ω = π1ω + ω, ω H2 + H3. ∈ But ω = ∆Gω = ∆Gω as ω H2 + H3 and G = 0 on H1. So − − ∈ ω = π1ω ∆Gω. − Thus we have the formula

I = π1 ∆G − = π1 + d∂ + ∂dG.

Since I = π1 + π2 + π3, it follows that π2 = d∂G and π3 = ∂dG.

75 76 Lecture 14

Decomposition of D 79 H has certain fundamental defects; d, ∂ and ∆ do not operate on H . But these operators operate on D. We now consider D. G operates on D. For let ω be a C form. ω = π1ω ∆Gω or ∞ − ∆Gω = π1ω ω. Since π1ω ω is a C form it follows from the elliptic − − ∞ character of ∆ that Gω is a C∞ form. π1 evidently operates on D. Since π2 = d∂G and π3 = ∂dG, π2 and π3 also operate on D. Let D1 = D H1, D2 = D H2, D3 = D H3 ∩ ∩ ∩ D1 is the space of all harmonic forms. D2 = dD and D3 = ∂D by de Rham’s theorem. D1, D2 and D3 re closed subspaces of D. (D = E has a genuine topology). The linear maps π1, π2 and π3 from D onto the spaces D1, D2, D3 respectively have the following properties:

πiπ j = 0 for i , j 2 = πi πi I = π1 + π2 + π3,

Consequently D is the direct sum of the closed subspaces D1, 80 D2 and D3. For any element ω D we have the decomposition for- ∈ mula: ω = π1ω + d∂Gω + ∂dGω.

Continuity of G

Let F be a Frechet´ space i.e., a complete topological vector space with a denumerable basis of neighbourhoods of 0. Banach’s closed graph theorem states that if a linear map G : F F is discontinuous then there → exists a sequence of elements ϕ j 0 such that Gϕ j tends to a non-zero → element θ. So in order to prove that a linear map G of a Frechet´ space into itself is continuous it is sufficient to show that ϕ j 0 and Gϕ j θ → → together imply that θ = 0. [We can give an example of a normed vector space in which the closed graph theorem is not true. Let E be the space of in the closed (0, 1) with the topology of uniform Lecture 14 77

d convergence in (0, 1) (Norm f = Maxx (0,1) f (x) ). The operator is a ∈ | | dx discontinuous operator on this space, as a sequence of polynomials may tend uniformly to zero in (0, 1) while their derivatives may not. However the closed graph theorem is not true for this operator; for if a sequence dP j of polynomials P j 0 uniformly in (0, 1) and θ uniformly in → dx → (0, 1), we must have θ = 0. Here the space is not complete. In fact the completion of E in the norm defined above is the space of continuous d functions in (0, 1) (Weierstrass approximation theorem) and can not dx be extended to this space]. We shall now use the closed graph theorem to prove the continuity of G in D. D is a Frechet´ space. (In general E is a Frechet´ space; 81 here since the manifold is compact D = E ). Let ϕ j be a sequence of { } elements of D such that ϕ j 0 and Gϕ j θ; we have to show that → → θ = 0. Write ϕ j = π1ϕ j ∆Gϕ j − π1 is continuous in D. For if ω D ∈

π1ω = (ω, θk)θk k where θk is an orthonormal base for H1. If ϕ j ϕ in D(ϕ j, θk) → → (ϕ, θk). Therefore if ϕ j 0 in D, π1ϕ j 0 in D. Now since Gϕ j θ → → → and ∆ is continuous ∆Gϕ j ∆θ. So ∆θ = 0 or θ H1; already → ∈ θ H2 + H3; consequently θ = 0. ∈ Similarly it can be proved that G is continuous in H .

Self-adjointness of G

We shall now show that G is self-adjoint in H ;

(Gϕ,ψ) = (ϕ, Gψ),ϕ,ψ H ∈ If we put Gϕ = α and Gψ = β we have

ψ = π1ψ ∆β and ϕ = π1ϕ ∆α − − 78 Lecture 14

so that

(Gϕ,ψ) = (α,ψ)

= (α, π1ψ) (α, ∆β) − 82 and

(ϕ, Gψ) = (ϕ,β)

= (π1ϕ,β) (∆α,β) −

Since (α, π1ψ) = (π1ϕ,β) = 0(α,β H2 + H3 while π1ψ, π1ϕ H1) ∈ ∈ we have only to prove that (α, ∆β) = (∆α,β). But this is evident when ϕ or ψ (and hence α or β) is a C∞ form. So we have the relation (Gϕ,ψ) = (ϕ, Gψ) when ϕ or ψ is a C∞ form. Since D is dense in H and G is continuous we obtain

(Gϕ,ψ) = (ϕ, Gψ),ϕ,ψ H . ∈ G is hermitian positive (Gϕ,ϕ) 0 and (Gϕ,ϕ) = 0 if and only if ϕ is ≥ harmonic. For if Gϕ = α

(Gϕ,ϕ) = (α, π1ϕ ∆α) − = (α, ∆α) − 0 ≥ and (Gϕ,ϕ) = 0 if and only if ∆α = 0 or ∆ϕ = 0. Lecture 15

Decomposition of D ′ 83 We shall now extend the operator G to D′. For a current T we define GT by: (GT,ϕ) = (T, Gϕ). (This definition is consistent) Since G is continuous on D, it is easy to verify that GT is continuous on D. For if ϕ j 0 in D, Gϕ j 0 and → → (GT,ϕ j) = (T, Gϕ j) 0. The operator G is continuous on D endowed → ′ with the weak topology. For if T j 0 in the weak topology, for a fixed → ϕ D (GT j,ϕ) = (T j, Gϕ) 0. We define the operators π1, π2 and π3 ∈ → on D′ by:

π1 = I + ∆G

π2 = d∂G, π3 = ∂dG.

The operators π1, π2 and π3 verify the relations:

πiπ j = 0 for i , j 2 = πi πi I = π1 + π2 + π3. = D′ is the direct sum of D1′ (the space of harmonic forms), D2′ dD′ = and D3′ ∂D′. For a current T we have the decomposition formula

T = π1T + d∂GT + ∂dGT.

79 80 Lecture 15

Commutativity of an operator with ∆, π1 and G 84 If an operator A : D D commutes with ∆ it also commutes with → π1 and G. If A∆ω = ∆Aω we have to show that π1Aω = Aπ1ω and GAω = AGω. If π1Aω = ω, ω is characterised by the properties: ∆ = i) ω 0 ii) Aω ω = ∆η for some η D. − ∈ = We shall verify that ω Aπ1ω also possesses these properties. ∆ = ∆ = ∆ = ω Aπ1ω A π1ω 0

as π1ω is a harmonic form;

Aω ω = Aω Aπ1ω − − = A(ω π1ω) − = A(∆η′)

= ∆(Aη′).

So π1Aω = Aπ1ω ω1 = GAω is characterised by: i) ∆ω1 = Aω π1Aω − − ii) π1ω1 = 0

We shall show that ω1 = AGω has these properties. ∆AGω = A∆Gω − − = A(ω π1ω) − = (Aω π1Aω) (as A and π1 commute) − 85 and

π1AGω = Aπ1Gω = 0.

The operator ∆ commutes with each of the operators d, ∂, ∆, , π1, ∗ π2, π3, G. Consequently π1 and G also commute with each of these operators. Lecture 15 81

Operators on currents as operators on cohomology spaces

Suppose A is an operator on currents with the following properties:

0 i) for every cohomology class α there exists at least one element of the class such that Aα is closed.

ii) If α is a coboundary and Aα is closed, then Aα is a coboundary. Then we can define A on the cohomology vector spaces intrinsi- 0 cally: for each cohomology class α we choose a representative 0 α such that Aα is closed and map α onto the cohomology class determined by Aα. Of course, this definition makes no use of any particular metric. If there exists at least one Riemannian metric on the manifold for which A and ∆ commute then the conditions i) and ii) are verified. 0 In a cohomology class α we choose the harmonic form α. Then Aα is closed, in fact, harmonic. ∆Aα = A∆α = 0. To verify the second condition we notice that a necessary and sufficient condi- tion for a closed element ω to be a coboundary is that π1ω = 0. If α is a coboundary and A α is closed,

π1Aα = Aπ1ω ( as A and π1 commute) = 0.

So A operates intrinsically on the cohomology spaces. 86

[We may simply identify the cohomology space with the space of harmonic forms and let A operate on the space of harmonic forms. (A operates on the space of harmonic forms as A and ∆ commute).]

Complex differential forms on a manifold

N Let V be a C∞ manifold. Just as we considered real valued C∞ func- N N tions on V we may also consider complex-valued C∞ functions on V ; 82 Lecture 15

a complex valued C∞ function is of the form ϕ = f + ig where f and g are real valued C∞ functions. N ff N At a point a of V we can define the space of di erentials Ta∗(V ) with respect to the complex valued functions, just the same way we N did in the case of real valued functions. The space Ta∗(V ) is of com- N plex dimension N. (Henceforth, Ta∗(V ) will always denote the space N of complex differentials at a. However Ta(V ) will only denote the real tangent space at a). Now if F is a vector space over the reals, the space F + iF is called the complexification of F; the complex dimension of F + iF is equal to the real dimension of F. We shall always consider N ff Ta∗(V ) as the complexification of the space of real di erentials at a. p Λ Similarly the space Ta∗(V) will be considered as the complexification 87 of the real tangent p-covectors at a. Thus an element ω T (V) has the ∈ a∗ canonical decomposition ω = ω1 +iω2 where ω1 and ω2 are real tangent p-covectors at a. If a complex vector space is obtained as the complexification of a real vector space, we have the notion of in this space. For if G = F + iF is the complexification of the real vector space F any element ω G has the canonical decomposition ω = ω1 + iω2 where ∈ ω1, ω2 F; the complex conjugate of ω is the element ω1 iω2. ∈ p p − Λ Λ The manifold TN∗ (V) of all Ta∗(V) is a C∞ manifold with real + N dimension N 2 p . We extend the operators d, ∂, ∆ to the complex differential forms by linearity d(ω1 + iω2) = dω1 + idω2 and a scalar product in the space of real differentials at a canonically to a Hermitian scalar product in its complexification i.e., in T (V). We extend by anti-linearity: a∗ ∗

(ω1 + iω2) = ω1 i ω2 ∗ ∗ − ∗ so that the relation (α,β)τ = αΛ β ∗ is preserved. The space of (complex) square summable forms becomes a Hilbert space over complex numbers. Lecture 16

Real vector spaces with a J-Structure 88 Suppose G(n) is a vector space over the complex numbers. G can also be considered as a vector space over the real numbers. If e1,..., en is n a basis of G over C, then e1,..., en, ie1,..., ien is a basis of G over R. The multiplication by i is a linear transformation of G, considered as a vector space over R, whose square is I, where I is the identity map. − The vectors e and ie are dependent when G is considered as a vector space over C but are independent over R. To avoid confusion we shall introduce the notion of a real vector space with a J-structure. A 2n-dimensional real vector space G along with a linear transfor- mation J : G G with J2 = I will be a called a real vector space with → − a J-structure. (A real vector space G with a J-structure can be considered as a vector space over complex numbers by defining

(α + iβ)e = αe + βJe where α and β are real numbers and e is an element of G). Let G be a real vector space with a J-structure and G + iG its com- plexification; the operator J is extended canonically to G + iG by defin- ing: J(X + iY) = JX + iJY, X, Y G ∈ J is a linear transformation on the complex vector space G + iG. (We 89 notice that the operations J and multiplication by i are different on G +

83 84 Lecture 16

iG; if x is a vector in G, Jx is a vector in G while ix is a vector in iG). The operator J on G + iG has the eigen values i and i. − J is an isomorphism of G + iG onto itself. Canonically we have an isomorphism of the dual space of G + iG, (G + iG)∗, onto itself con- tragradient to J (this isomorphism is the inverse of the transpose of J). We denote this operator on the dual space also by J. If α G + iG and ∈ β (G + iG) then ∈ ∗ α,β = Jα, Jβ 1 Jα,β = α, J− β = α, Jβ . −

J is also defined canonically on the exterior products of (G + iG)∗. On the p-th exterior product we have

J2 = ( 1)pI. − ∂ ∂ The operators and ∂z j ∂z j

∂ 1 ∂ ∂ = + i ∂z j 2 ∂x j ∂y j

90 We have

∂ = ∂ + ∂ ∂x j ∂z j ∂z j ∂ ∂ ∂ = i ∂y j ∂z j − ∂z j

The reason for such a definition is as follows. If we take an of 2n real variables x1, y1,..., xn, yn it can be extended to an ∂ ∂ 0We now define the operators and on differentiable functions on ∂z j ∂z j n n C /(z1,..., zn), z j = x j + iy j are the coordinate functions on C ). A priori they do not make sense. We define ∂ 1 ∂ ∂ = i , ∂z j 2 ∂x j − ∂y j Lecture 16 85 analytic function of 2n complex variables, which we still denote by x1, y1,..., xn, yn. Consider

z j = x j + iy j, z j = x j iy j − as 2n-independent complex variables (Here z j does not mean the com- plex conjugate of z j; this is so only if x j and y j are real). By changing the variables x j, y j to z j, z j we get the above expressions for the par- ∂ ∂ tial derivatives and . Thus these relations are true for analytic ∂z j ∂z j functions of real variables prolonged into the complex field. So we take these as general definitions. We have ∂ ∂ (zk) = δ jk, (z j) = 0. ∂z j ∂z j

∂ ∂ ∂ f ∂ f are the complex derivations. and are defined for any 91 ∂z j ∂z j ∂z j ∂z j differentiable function f . n Suppose f is a C∞ function on C , We have

∂ f ∂ f d f dx j + dy j. − ∂x ∂y j j j j

But this may be written as

∂ f ∂ f df = dz + dz ∂z j ∂z j j j j j

∂ f ∂ f (Here and are the complex derivatives of f defined above. dz j ∂z j ∂z j and dz j are the differentials of the functions z j and z j). Thus we have an expression for df as though z j and z j were independent variables. Similarly the formula for the coboundary of a differential form continues to hold as though z j and z j were independent variables. 86 Lecture 16

Holomorphic functions on Cn

If f is a complex valued function defined on an open subset Ω of C′ we say that f is holomorphic in Ω if

f (z + ζ) f (z) lim − , ζ 0 ζ → exists at every point z of Ω. We know that if f is holomorphic it satisfies ∂ f the Cauchy rule; the Cauchy rule can be written as = 0. Conversely, ∂z ∂ f if f is C and = 0, f is holomorphic. So a ′ ∂z 92 of one complex variable may be defined as a C1 function of x and y for ∂ f which = 0. (We may say that a holomorphic function is independent ∂z of z). For functions of several variables we adopt a similar definition. We say that a complex valued function f defined on an open subset n of C is holomorphic if f is a C∞ function with respect to the 2n real coordinates and ∂ f = 0( j = 1, 2,..., n). ∂z j

Transformation formulae

Suppose we have a diffeomorphism

(z1,..., zn) (ζ1,...,ζn) → n between two open subsets of C given by the functions ζk = ζk(x1, y1, ..., xn, yn), k = 1,..., n. We then have the following formulae:

∂ζ ∂ζ dζ = k dz + k dz k ∂z j ∂z j j j j

= ∂ζk + ∂ζk dζk dz j dz j .  ∂z j ∂z j  j     Lecture 16 87

In particular if the ζi are holomorphic functions of z1,..., zn we have ∂ζ dζ = k dz k ∂z j j j ∂ζk (Here is the ordinary partial derivative of ζk with respect to z j de- 93 ∂z j fined in the theory of holomorphic functions; and thus our notation is coherent). In this case we also have

= ∂ζk dζk dz j  ∂z j  j     ∂ = ∂ζk ∂ ∂z ∂z ∂ζ j k j k ∂ ∂ζ ∂ = k . ∂z ∂z j k j ∂ζk

Canonical complex structure on R2n

2n Let (x1,..., x2n) be the coordinate functions on R . We shall identify R2n with Cn by the map

(x1,..., x2n) (z1,..., zn) → where z j = x2 j 1 + ix2 j( j = 1, 2,..., n) − 2n We thus have a canonical complex structure on R . If(e1,..., e2n) is the canonical basis for R2n then the canonical complex structure on R2n is given by the J operator defined by:

Je2 j 1 = e2 j, Je2 j = e2 j 1( j = 1,..., n) − − − Let U be an open subset of R2n and Φ : U R2n be a C map. 94 → ∞ In terms of the canonical complex coordinates on R2n we may give this map by: (z1,..., zn) (ζ1,...,ζn) → 88 Lecture 16

We say that the map Φ is complex analytic (with respect to the canoni- 2n cal complex structure on R ) if ζ1,...,ζn are holomorphic functions of (z1,..., zn).

Complex analytic manifolds

(n) A complex analytic manifold, V , of complex dimension n is a C∞ manifold of real dimension 2n with an atlas (Ui,ϕi) (which is incom- { } plete with respect to the C∞ structure) having the following property: for any two maps (Ui,ϕi) and (U j,ϕ j) of the atlas, the map

1 ϕ j ϕ− : ϕi(Ui U j) ϕ j(Ui U j) ◦ i ∩ → ∩ is complex analytic (with respect to the canonical complex structure on R2n). We assume that the atlas is complete with respect to the complex analytic structure.

Some examples of complex analytic manifolds

i) Cn. The simplest example of a complex analytic manifold is Cn itself.

ii) The Riemann sphere S 2. Consider S 2 as the one point compacti- fication of C1 : S 2 = C1 . Take for one map the identity map ∪∞ of C1. For the second map take the map ζ defined by

ζ(z) = 1/z, z , ∞ ζ( ) = 0 ∞ 95 in the complementary set of 0. The intersection of these two maps is the complement of the points 0 and and here ζ = 1/z ∞ is a holomorphic function of z. It is known that the only spheres on which we may have a com- plex analytic structure are S 2 and S 6. On S 2 we have a complex analytic structure. For S 6 we do not know. Lecture 16 89

iii) The complex PCn The right generalization of the Riemann sphere is the n-dimen- sional , PCn. The n-dimensional com- plex projective space is defined as follows. We take Cn+1 and omit 0. In Cn+1 (0) we introduce an equivalence relation two = − = points z (z1,..., zn+1) and z′ (z1′ ,..., zn′ +1) are equivalent if = = + = , zi′ λzi(i 1,..., n 1)(z′ λz) for some λ 0. The quotient space of Cn+1 (0) by this relation (with the quotient topology) − is the n-dimensional complex projective space, PCn PCn is a compact complex analytic manifold of complex dimension n. We introduce complex analytic coordinate systems in PCn as follows. For a fixed i consider the set of points a in PCn whose represen- n+1 tatives in C are of the form (z1,..., zn+1), zi , 0. Then the mapping 96

z1 zi 1 z j+1 zn+1 a ,..., − , ,..., . → zi zi zi zi gives a map. The maps obtained for i = 1, 2,..., n + 1 cover PCn and are related by holomorphic functions on the overlaps.

iv) The complex torus.

Lecture 17

The operator J 97 We now pass on to some intrinsic properties of a complex analytic man- ifold V(n) of complex dimension n. The (real) tangent space to V at a, Ta(V), is a vector space of dimension 2n over R. We shall now introduce 2n on Ta(V) an intrinsic J-structure. Take a map at a into R . This map 2n 2n gives an isomorphism between Ta(V) and R . In R we have a J cor- responding to the canonical complex structure in R2n; by the canonical 2n isomorphism between Ta(V) and R (given by the map) we also have a J in Ta(V). We shall now prove that this J in Ta(V) is intrinsic. If

Jz1,...,zn and Jζ1,...,ζn are the J operators on Ta(V) corresponding to the local coordinates (z1,..., zn) and (ζ1,...,ζn) we have to prove that J and Y J are the same. To prove this we consider the complexification of Ta(V), Ta(V) + iTa(V). We extend J and J to Ta(V) + iTa(V); J and J are operators in Ta(V) + iTa(V) with eigenvalues i. To prove that J ± and J are the same it is sufficient to show that the corresponding eigen spaces are the same. From the relations

∂ ∂ ∂ ∂ J = i , J = i ∂zk ∂zk ∂zk − ∂zk it follows that for J the eigenspace corresponding to the eigenvalue i is the space spanned by ∂ ∂ ,..., ∂z1 ∂zn and the eigen-space corresponding to i is the space spanned by 98 − 91 92 Lecture 17

∂ ∂ ,..., ; ∂z1 ∂zn for J the corresponding spaces are spanned by ∂ ∂ ∂ ∂ ,..., and ,..., , ∂ζ ∂ζ 1 n ∂ζ1 ∂ζn respectively. Since the maps are related on the overlaps by holomorphic functions we have the relations ∂ ∂ζ ∂ = k , ∂z ∂z ∂ζ j k j k ∂ ∂z ∂ = k , ∂ζ ∂ζ ∂z j k j k ∂ ∂ζ ∂ = k , ∂z ∂z j k j ∂ζk ∂ ∂z ∂ = k ; ∂z ∂ζ j k ∂ζ j k these relations show that the eigen space of J and J corresponding to the eigen values i and i are the same. − J operates on the space of tangent covectors real or complex; we 99 have the relations Jdzk = idzk Jdzk = idzk. J also operates on the − space of tangent p-covectors (real or complex) at a; consequently J op- erates on the space differential forms. In fact J is a real operator i.e., J takes real differential forms into real differential forms. For forms of degree P we have J2 = ( 1)pI − Bigradation for differential forms

A differential form ω is said to be of bidegree or of type (p, q) if in every map, ω has the form

ω = ω j ... j k ...k dz j ... dz j dzk ... dzk . 1 p 1 q 1 ∧ ∧ p ∧ 1 ∧ ∧ q j1<...< jp k1<...

This definition is correct since if two maps overlap and a form defined in the overlap be of type (p, q) in one of the maps it is also of type (p, q) in the other map. (This follows easily from the fact that the maps are related by holomorphic functions on the overlaps). p is called the z degree and q the z degree of ω. Any differential form ω of total degree r can be uniquely written in the form (p,q) ω = ω p+q=r (p,q) where ω is a form of bidegree (p, q).

Holomorphic functions and forms 100 A (complex valued) function f on V(n) is said to be holomorphic if f is holomorphic on every map; this definition is correct as a holomorphic function of n holomorphic functions on Cn is again holomorphic. A holomorphic differential form of degree p is a form of type (p, 0) whose coefficients in every map are holomorphic.

The operators dz and dz

Suppose ω is a form of bidegree (p, q). A priori dω is the sum of forms of all bidgree (r, s) with r + s = p + q + 1. However we shall show that dω is the sum of a form of bidegree (p + 1, q) and a form of bidegree (p, q + 1). Let ω = ωJ,KdzJ dzK ∧ J,K in a map. J = ( j1,..., jp), K = (k1,..., kq) denote a system of indices in

dω = dωJ,KdzJ dzK ∧ ∂ωJ,K = dz1 dzJ dzK ∂z1 ∧ ∧ 94 Lecture 17

∂ωJ,K + dz1 dzJ dzJ ∂z1 ∧ ∧ Here the first form is of bidegree (p + 1, q) and the second of bidegree 101 (p, q + 1); therefore these two forms have an intrinsic meaning. Thus

+ (p,q) (p+1q) (p,q 1) d ω = α + β

intrinsically, where α and β are forms of bidegree (p+1, q) and (p, q+1) respectively. We now define the operators dz and dz by:

dzω = α, dzω = β.

We have dω = d2ω + dzω.

We observe that, in a map, dzω involves only the partial derivatives with respect to z while dzω involves only the partial derivatives with respect to z dz increases the degree corresponding to z by one while d increases z the degree corresponding to z by one. We extend the operators dz and dz to all forms by linearity. We shall now consider some properties of dz and d dz and d are z z complex operators (If ω is real dzω is complex). dz is of type (1, 0) and dz is of type (0, 1). [An operator is said to be of type (r, s) if it takes a form of bigradation (p, q) into a form of bigradation (p+r, q+ s)]. These operators are local operators; they are linear. If ω is a form of degree r we have the formulae:

r dz(ω ω) = dzω ω + ( 1) ω dzω, ∧ ∧ − ∧ d (ω ω) = d ω ω + ( 1)rω d ω. z ∧ z ∧ − ∧ z 102 It is enough to prove this for homogeneous forms. We take d(ω ω) ∧ and decompose it: p,q s,t d( ω ω) = dω ω + ( 1)p+qω dω ∧ ∧ − p+q ∧ = (dzω ω + ( 1) ω dzω) ∧ − ∧ Lecture 17 95

+ (d ω ω + ( 1)p+qω d ω); z ∧ − ∧ z the first term is of bidegree (p + s + 1, q + t) while the second is of bidegree (p + s, q + t + 1) and this proves the result. Further we have the relations

dzdz = 0, dzdz = 0 and dzdz + dzdz = 0. For, from the relations

2 (dz + dz)(dz + dz) = d = 0 we obtain dzdz + (dzdz + dzdz) + dzdz = 0. To conclude the above relations from this we have only to observe that p,q for any form ω the forms dzdzω, (dzdzω + dzdzω) and dzdzω are of different bidegrees namely of the bidegrees (p + 2, q), (p + 1, q + 1) and (p, q + 2) respectively. z and z cohomologies 103 We have now two new coboundary operators dz and dz and hence two new cohomologies, z and z cohomologies. We shall confine our attention to the z cohomology; this will give all the information about holomor- phic forms. We shall see that the construction of holomorphic forms p,q depends on the z cohomology. We shall denote by Hz (E (V)) the space Z p,q/Bp,q where Z p,q and Bp,q are the subspaces of E p,q(V) (the space of forms of bidegree (p, q)) consisting of z cocycles and z coboundaries respectively. p,q Similarly we have the space Hz (D(V)).

Intrinsic characterization of holomorphic forms

Let ω be a C∞ form of type (p, 0); a necessary and sufficient condition for ω to be holomorphic is that

dzω = 0. 96 Lecture 17

Let ω = ωKdzK K in a map. ∂ωK d ω = dz1 dzK z ∂z ∧ K,1 1

∂ωK In ω is holomorphic, ωK are holomorphic; hence = 0 and dzω = 0. ∂z1 If dzω = 0 ∂ωK dz1 dzK = 0; ∂z ∧ K,1 1

104 in the sum on the left side all the terms dz1 dzK are different (i.e., if ∧ ∂ω , , K = (K, 1) (K′, l′), dz1 dzK dzl′ dzK′ ). Consequently 0. This ∧ ∧ ∂z1 proves that the ωK are all holomorphic.

Holomorphic forms and z cohomology

p,0 p,0 Let us now consider the space H (E (V)) H (E (V)) is just the space z z of (p, 0) forms which are z cocycles (since a (p, 0) from which is z p,0 coboundary is trivial). i.e., Hz (E (V)) is the space of holomorphic p- p,0 forms. Similarly Hz (D(V)) is the space of holomorphic forms with compact support. A holomorphic form with compact support is zero on each non-compact connected component. ,q p,q ,q Consider the space E = E of C∞ forms of z degree q. E p q is a complex with the differential operator dz. This gives rise to the ,q ,0 cohomology groups Hz . Hz is the space holomorphic forms. Lecture 18

The canonical orientation of a complex manifold 105 We shall now show that a complex analytic manifold V(n) (considered as a 2n-dimensional real manifold) is orientable and has a canonical orientation. The orientation on V2n is determined by the maps giving the complex analytic structure on V2n; we have to verify that two such maps, considered as real coordinate systems, have a positive Jacobian on the overlaps. To prove this, let (z1,..., zn) (ζ1,...,ζn) be a holomorphic n n → map of C to C . Let D be the Jacobian of ζ1,...,ζn with respect to z1,..., zn; let further zi = xi + iyi, ζi = ξi + iηi and J the Jacobian of the functions (ξ1, η1,...,ξn, ηn) with respect to (x1, y1,..., xn, yn). We shall 2 prove that J = D . Now dζi = dξi + id ηi and dζ = dξi id ηi so that | | i − 1 dξi dηi = dζ dζ ; ∧ −2i ∧ i Similarly 1 dxi dyi = − dzi dzii1 ∧ 2i ∧ We have

dξ1 dη1 ... dξn dηn dζ1 dζ ... dζn dζ J = ∧ ∧ ∧ ∧ = ∧ 1 ∧ ∧ ∧ n dx1 dy1 ... dxn dyn dz1 dz1 ... dzn dzn ∧ ∧ ∧ ∧ ∧ ∧ ∧ ∧ dζ1 ... dζn dζ ... dζ = ∧ ∧ ∧ 1 ∧ ∧ n dz1 ... dzn dz1 ... dzn ∧ ∧ ∧ ∧ ∧ = DD = D 2 | |

97 98 Lecture 18

as 106 dζ1 ... dζn dζ ... dζ ∧ ∧ = D, 1 ∧ ∧ n = D. dz1 ... dzn dz1 ... dzn ∧ ∧ ∧ ∧ In the case of our maps D , 0 and therefore J > 0.

Currents

We shall now define bigradation for currents and the operators J, dz and dz on currents. A current T is said to be of bidegree (p, q) if

r,s T, ϕ = 0 r,s whenever (p, q) , (n r, n s). (ϕ is a form of bidegree (r, s)). A current p,q − − T of bidegree (p, q) can be considered as a continuous linear functional n p,n q on −D − , the space of forms of bidegree (n p, n q) with compact − − support. We define the operator J on currents by:

1 JT,ϕ = T, J− ϕ . We do this because, when ω is a form we have

1 Jω,ϕ = ω, J− ϕ . 1 Writing J− ϕ = ψ, we have to prove that

Jω Jψ = ω ψ or J(ω ψ) = ω ψ ∧ ∧ ∧ ∧

107 but J(ω ψ) = ω ψ ∧ ∧ since ω ψ is a form of degree 2n and J = I on a form of degree 2n. (In ∧ general p,q p q q p J ω = ( i) i ω = i − ω). − Lecture 18 99

We define dz and dz on currents by:

p,q n p 1,n q p+q+1 dz T , − −ϕ − = ( 1) T, d ϕ − z p,q n p,n q 1 + + d T , − ϕ− − = ( 1)p q 1 T, d ϕ . z − z These relations are true for forms. In fact when ω is a form of total degree p we have

p+1 dzω ϕ = ( 1) ω dzϕ. ∧ − ∧ V V For, ϕ is of type (n 1, n) so that −

d (ω ϕ) = 0 and dz(ω ϕ) = d(ω ϕ) z ∧ ∧ ∧

By Stokes’ formula d(ω ϕ) = 0; hence dz(ω ϕ) = 0 or ∧ ∧ V V

p (dzω ϕ) + ( 1) ω dzϕ) = 0). ∧ − ∧ V

We now have some more cohomologies and there is need to prove 108 some kind of de Rham’s theorems. It can be proved that the z cohomolo- gies of D′ and E are the same and those of D and E ′ are the same.

Ellipticity of the system ∂/∂zk

∂ We shall show that the system in Cn is elliptic i.e., if ∂zk ∂T = αk (k = 1, 2,..., n) ∂zk where T is a current and αk are C∞ forms then T is a C∞ form. We have ∂ 1 ∂ ∂ = i ∂zk 2 ∂xk − ∂yk 100 Lecture 18

∂ 1 ∂ ∂ = + i ∂zk 2 ∂xk ∂yk 2 2 ∂ ∂ = 1 ∂ + ∂ 2 2 ∂zk ∂zk 4 ∂x ∂y   k k    so that   ∆ ∂ ∂ = ∂z ∂z 4 k k k where ∆ is the usual Laplacian in R2n. ∂T Since = αk E , ∆T E . By the elliptic character of ∆, T is ∂zk ∈ ∈ C∞. 109 It follows in particular that a distribution T on Cn which satisfies the Cauchy relations ∂T = 0(k = 1, 2,..., n) ∂zk is a holomorphic function. Similarly if a current T of bidegree (p, 0) on Cn satisfies the system of partial differential equations ∂T = 0(k = 1, 2,..., n) ∂zk then T is a holomorphic form of degree p.

0′ Ellipticity of dz on D

Let V(n) be a complex analytic manifold. If T is a current of degree zero and dzT = α is a C∞ form, then T is a C∞ function. For, in a map, the relation dzT = α implies ∂T = αk ∂zk

where αk are C∞ functions. By the result proved earlier T is a C∞ function. p,0 As a consequence we obtain that if T is a current of bidegree (p, 0) p,q p,0 such that dz T = 0 then T is a holomorphic form. Lecture 18 101

J-Hermitian forms 110 Let G2n be a vector space over R with a J-structure. A positive definite J-Hermitian form on G is a map H of G G into the complex numbers × with the following properties: 1) H is R-bilinear. 2) H(JX, Y) = H(X, JY) = iH(X, Y) for X, Y G (Hence H(JX, − ∈ JY) = H(X, Y)). 3) H(X, X) > 0 for X , 0. The real part of the positive definite Hermitian form, (X, Y) = Rl H(X, Y), defines a Euclidean structure in G. Since H(X, Y) is invariant under J,(X, Y) is also invariant under J:

(JX, JY) = (X, Y)

Since (JX, X) = 0 the vectors X and JX are orthogonal with respect to the Euclidean structure. From the relation H(JX, Y) = iH(X, Y), we have (JX, Y) = Im H(X, Y). − Hence

H(X, Y) = (X, Y) + i Im H(X, Y) = (X, Y) i(JX, Y) − = (X, Y) + i(X, JY).

This shows that H is completely determined by its real part. If (X, Y) is a positive definite quadratic form on G which is invariant under J this 111 determines a positive definite J-Hermitian form on G if we set

H(X, Y) = (X, Y) + i(X, JY).

Since (JX, X) = 0,( )(JX, Y) is an anti-symmetric bilinear form on G. 2 This defines an element Ω of G (G is the dual of G) by the formula ∧ ∗ ∗ Ω, X Y = (JX, Y), X, Y G. ∧ ∈ 102 Lecture 18

G can be considered (canonically) as a vector space over complex numbers, as we have already remarked. If (u1,..., un), (v1,..., vn) are the coordinates of the vectors U and V of G with respect to a complex basis (e1,..., en)., = = H(U, V) g jku jvk, g jk gk j, (gi j) > 0. j,k

Let us compute Ω in terms of this basis. Ω is given by

H(U, V) H(V, U) Ω, U V = − ∧ − 2i H(U, V) H(U, V) = − − 2i 1 = g jk(u jvk ukv j). −2i − j,k

If (e1∗,..., en∗) is the dual basis of (e1,..., en)

= u j, uk e∗j ek∗, U V ∧ ∧ v j, vk 112 so that Ω= 1 g jke∗j ek∗. −2i ∧ If we choose an orthonormal basis (e1,..., en) for the hermitian form,

Ω= 1 e∗j ek∗. −2i ∧ Hermitial Manifolds

(n) Let V be a complex analytic manifold. At each tangent space Ta(V) we have a canonical J-structure. V(n) will be called a Hermitian mani- fold if on each Ta(V) we have a positive definite J-Hermitial form such that the twice covariant tensor field defined by these forms is a C∞ tensor field. Lecture 18 103

On every complex analytic manifold we can put a Hermitian struc- ture, just the same way we introduced a Riemannian structure on a C∞ manifold. In a Hermitian manifold V(n) the real part of the J-Hermitian form on each Ta(V) gives rise to a Riemannian structure on the manifold; the imaginary part gives rise to a real C∞ differential form Ω of bidegree (1, 1). If in a coordinate system (z1,..., zn) the Hermitian form is given by

g jkdz jdzk, then in this map, 113 1 Ω= g jkdz j dzk. −2i ∧ If τ is the volume element associated with the Riemannian structure we n have the relation Ω = n!τ. For, if (z1,..., zn), z j = x j + iy j, is a local coordinate system at a V in which the Hermitian form at a is given by ∈ (dzk)a(dzk)a, we have, 1 Ωa = (dzk)a (dzk)a −2i ∧ k

= (dxk)a (dyk)a; ∧ n Ω = (dx1)a (dy1)a ... (dxn)a (dyn)a n! a ∧ ∧ ∧ ∧ ∧ = n!τa.

Kahlerian Manifolds

A Hermitian manifold is called a Kahlerian manifold if dΩ= 0. There exist manifolds with an infinity of Hermitian structures but with no Kahlerian structure. For a compact Kahlerian manifold V,

b2p 1. ≥ 104 Lecture 18

To prove this we notice that

Ωn = n!τ> 0. V V

114 Ωn is a cocycle but not a coboundary, since Ωn , 0. It follows that the V forms Ωp(1 p n), which are cocycles, are not coboundaries. Hence ≤ ≤ the 2p-th cohomology groups are different form zero or b2p 1. ≥ Since b4(S 6) = 0, if S 6 is a complex analytic manifold it is not Kahlerian. Every complex analytic manifold of complex dimension 1 is Kahle- rian because and 2-form is closed. Lecture 19

Some more operators 115 Let d = idz + id − z Then 1 d = JdJ− p,q q p 1 p q For, if ω is a form of bidegree (p, q), Jω = i − ω and J− ω = i − ω so that

1 q (p+1) p q Jdz J− ω = i − i − dzω

= idzω; − thus 1 Jdz J− = idz − and similarly 1 Jdz J− = idz so that 1 d = JdJ− . d is the transform of d by the automorphism J. Since J and d are real operators (i.e., take real forms into real forms) d is a real operator. Now we can decompose the operators d and d into real and imaginary parts z z as follows: 1 d = (d + id) z 2 105 106 Lecture 19

1 d = (d id). z 2 − We have evidently dd = 0 and dd + dd = 0. The operators ∂ and ∂z are the adjoints (with respect to the Rieman- z nian structure) of the operators dz and dz respectively:

(dzα,β) = (α,∂zβ)

(dzα,β) = (α,∂zβ)

116 the scalar product being the global Riemannian scalar product. ∂z is an operator of type ( 1, 0) while ∂ is an operator of type (0, 1). If ∂ is − z − the adjoint of d ∂ = i∂z i∂ ; − z we have 1 ∂z = (∂ i∂) 2 − 1 ∂ = (∂ + i∂). z 2 The following relations are easily verified:

∂z∂z = 0,∂z∂z = 0,∂z∂z + ∂z∂z = 0 ∂∂ = 0, ∂∂ = 0,∂∂ + ∂∂ = 0.

We now introduce two more operators, L and Λ. L is simply multi- plication by Ω : Lω =Ω ω. ∧ L is an operator of type (1, 1). Λ is the adjoint of L. We have

1 Λ= − L . ∗ ∗ For

(Λα,β) = (α, Lβ) 1 = − α Lβ ∗ ∧ Lecture 19 107

1 = − α Ω β ∗ ∧ ∧ 1 = Ω − α β ∧ ∗ ∧ 1 = (L − α) β ∗ ∧ 1 1 = − [ L − α] β ∗ ∗ ∗ ∧ 1 = ( L − α,β) ∗ ∗ 1 = (− L α,β) ∗ ∗ We have used above the fact that Ω is a real form of degree 2. 117

Commutativity relations in a Kahlerian manifold

We shall now consider the commutativity properties of the operators on a Kahlerian manifold. L commutes with the operators d, d, dz and dz; we denote this by

L d, d, dz, dz For, d(Ω ω) =Ω dω as dΩ= 0; ∧ ∧ since dzΩ+ dzΩ= 0 and dzΩ and dzΩ are forms of bidegree (2, 1) and (1, 2) we have dzΩ= 0 and dzΩ= 0; it follows that

dz(Ω ω) =Ω dzω ∧ ∧ d (Ω ω) =Ω d ω z ∧ ∧ z By taking the adjoints we find that 118

Λ ∂, ∂, ∂z,∂z 108 Lecture 19

L does not commute with ∂, ∂, ∂z and ∂z. Λ does not commute with d, d, d and dz. z L∂ω =Ω ∂ω, ∂Lω = ∂(Ω ω) ∧ ∧ and we have no rule for the ∂ of a product so that it is not possible to compare L∂ and ∂L directly. We have the following formula which gives the defect of commutativity of d and Λ: writing

[Λ, d] = Λd dΛ we have − [Λ, d] = ∂ − (Consequently Λ and d do not commute). This formula follows from the following formulae:

[Λ, dz] = i∂z

[Λ, d ] = i∂z z − which we shall prove in the next lecture. From these relations we have at once [Λ, d] = ∂.

119 By taking the adjoints we find that

[L,∂] = d = [L,∂z] idz [L,∂ ] = idz z − [L, ∂] = d. − We shall derive some important formulae from the above formulae.

d∂ = d(dΛ Λd) − = dΛd − ∂d = (dΛ Λd)d − = Λ d d. Lecture 19 109

Adding we find that d∂ + ∂d = 0 i.e., d and ∂ anti-commute (In a Riemannian structure d and ∂ have no commutativity relation; in the case of a Kahlerian¨ manifold d and ∂ anti- commute). We have further the following formulae: dd + dd = 0 ∂∂+ ∂∂ = 0 d∂ + ∂d = 0 ∂d + d∂ = 0 + = dzdz dzdz 0 ∂z∂z + ∂z∂z = 0

dz∂z + ∂zdz = 0

∂zdz + dz∂z = 0.

We now consider ∆. 120

d∂ = (dz + dz)(∂z + ∂z)

= dz∂z + dz∂z + dz∂z + dz∂z;

∂d = ∂zdz + ∂zdz + ∂zdz + ∂zdz. By addition, ∆= ∆z ∆ , − − − z where

∆z = dz∂z + ∂zdz − ∆ = d ∂ + ∂ d . − z z z z z Since ∆z and ∆z are pure operators (i.e., operators of type (0, 0)), ∆ is also a pure operator; in other words, ∆ does not change the bigradation. If ∆= d∂ + ∂d then ∆= ∆z ∆ − − − z 110 Lecture 19

so that ∆= ∆.

121 However, 1 ∆= J∆J− . So 1 ∆= J∆J− i.e., J and ∆ commute. Moreover,

1 1 dz∂z = (d + id) (∂ i∂) 2 2 − 1 = [d∂ +d∂ + id∂ id∂] 4 − and 1 ∂zdz = [∂d + ∂d + i∂d i∂d] 4 − so that ∆ ∆ ∆z = . − − 4 − 4 Similarly ∆ ∆ ∆ = . − z − 4 − 4 Consequently

∆z = ∆z and

∆= 2∆z = 2∆z = ∆

This formula shows that ∆ can be obtained in terms of dz and ∂z alone or in terms dz and ∂z alone. We shall now see that ∆ commutes with all the operators we have introduced. Let Pr,s be the projection of the space of forms on the space 122 of forms of bidegree (r, s). (Pr,s maps a form on its homogeneous com- Lecture 19 111 ponent of bidegree (r, s)). Since ∆ is a pure operator, ∆ commutes with Pr,s. ∆ commutes with L.

d∂L Ld∂ = d∂L dL∂ + dL∂ Ld∂ − − − = d[∂, L] + [d, L]∂ = dd − (d and L commute as Ω is a closed form of even degree). Similarly

[∂d, L] = dd − and hence [ ∆, L] = 0. − Since ∆ and L commute, ∆ and Λ also commute. From the relation

∆(Ω ω) =Ω ∆ω ∧ ∧ we see that if a form ω is harmonic the form Ω ω is also harmonic. ∧ In particular the form Ω itself is harmonic. In fact Ω is closed (it is ∗ already closed). To prove this we observe that

[L,∂] = d or Ω ∂ω ∂(Ω ω) = dω ∧ − ∧ (In general we do not have a formula for the ∂ of a product of two forms; however this formula gives an expression for the ∂ of the product of a differential form by Ω). Taking ω = 1, we find that ∂Ω = 0. Thus Ω is 123 closed with respect to d, ∂dz, dz, ∂z and ∂z. 1 Since d = JdJ− , ∆ commutes with d; hence commutes with ∂. Since ∆ commutes with d, d, ∂ and ∂, ∆ commutes with d , d , ∂ and z z z ∂ . That ∆ commutes with d and d is very important, as this result z z z connects harmonic forms with z and z cohomologies. Thus in a Kahlerian¨ manifold ∆ commutes with all the operators, in particular with dz and dz. 112 Lecture 19

In the case of a compact Kahlerian¨ manifold we have a decomposi- tion of D (and D′) as the direct sum of the space of harmonic forms, the space of z (or z) coboundaries and the space of z (resp z) star cobound- aries. For we have the decomposition

I = π1 + ∆G;

replacing ∆ by the new expressions we have

I = π1 + d∂G + ∂dG, = + + I π1 2dz∂zG2∂zdzG, I = π1 + 2dz∂zG + 2∂zdzG.

These formulae will have important consequences in connection with the z and z cohomologies.

∆, G, π1 r,s , d,∂,π1, G,..., J, P , L, Λ, d, ∂, dz, d ,∂z,∂ . ∗ z z 124

Holomorphic forms on a Kahlerian¨ manifold

We have seen that every holomorphic form on Cn is harmonic. In an arbitrary Hermitian manifold it is not true that every holomorphic form is harmonic. However in a Kahlerian manifold every holomorphic form is harmonic. To prove this we use the fact that

∆= 2∆z, ∆z = dz∂z + ∂zdz.

p,0 If a form ω is holomorphic, then dzω = 0 and therefore ∂zdzω = 0; since ∂z decreases the z degree by 1, ∂zω = 0 and hence dz∂zω = 0. Therefore ∆ω = 0. Lecture 20

Proof of the formula [∆, dz] = i∂z 125 To prove the bracket relation

[Λ, dz] = i∂z in the case of a Kahlerian¨ manifold, we first verify this relation in the case of Cn with the canonical Kahlerian¨ metric

n dzkdzk. k=1 In this case 1 Ω= dzk dzk. −2i ∧ k The real part of the Hermitian form is given by

2 + 2 (dxk dyk). k The Euclidean structure given by the metric on the real tangent space at a point a induces a Euclidean structure on the real co-tangent space at a. We extend this Euclidean structure to a Hermitian structure in the complex co-tangent vector space. The 2n vectors dx1, dy1,..., dxn, dyn form an orthonormal basis for the real cotangent vector space; the vectors dz1, dz1,..., dzn, dzn, form an orthogonal basis for the complex co-tangent vector space; each of these vectors is of length √2 as dzk = 126

113 114 Lecture 20

j+k dxk + idyk and dzk = dxk idyk. The vector dzJ dzK is of length √2 − ∧ where j and k are the number of elements in J and K. We shall now introduce some elementary operators in Cn and ex- press the operators Λ, dz, ∂z etc. in terms of these operators. The opera- tor ek = dzkΛ

operates on forms by multiplying every form on the left by dzk ek is an operator of the type (1, 0). ek is the operator

dzkΛ.

ek is an operator of the type (0, 1). ik and ik are defined to be the adjoints of ek and ek respectively. ik is an operator of the type ( 1, 0) while ik is − an operator of the type (0, 1). We shall prove that the linear operator ik − is given by the formula

ik[ωdzJ dzK] = 0 ∧ ik[ω dzk dzJ dzK] = 2ω dzJ dzK ∧ ∧ ′ ∧ ∧ ′ ∧

(J′ is a set of indices without the index k). (this amounts essentially to the suppression of dzk). We shall verify that the operator defined by these formulae is the adjoint of ek. For any two forms α and β we shall verify that (ekα,β)a = (α, ikβ)a.

127 It is sufficient to verify these for the elementary forms i.e., to verify that

(ekα′dzJ dzK,β′dzL dzM)a = (α′dzJ dzK, ikβ′dzL dzM)a ∧ ∧ ∧ ∧ or (ekdzJ dzK, dzL dzM)a = (dzJ dzK, ikdzL dzM)a. ∧ ∧ ∧ ∧ The elements dZJ dzK are orthogonal. The right and left sides both { ∧ } vanish except when L = k + J and M = K. In this case both the sides are equal to 21+r+s where r and s are the number of indices in J and K respectively. Lecture 20 115

We shall now introduce two more elementary operators: ∂ ∂ ∂k = , ∂k = . ∂zk ∂zk

We can prove that the adjoint of ∂k is ∂k and the adjoint of ∂k is ∂k in a − − similar way. We can express all the operators in the Kahlerian¨ structure by means of these operators. We have 1 L = ekek −2i 1 Λ= ikik (taking the adjoint). 2i dz = ∂kek = ek∂k(ek and ∂k commute) dz = ∂kek ∂z = ∂kik. − 128 Now we can prove the bracket relation. We have 1 Λd = i i ∂ e , z 2i k k l l k,l 1 d Λ= ∂ e i i . z 2i 1 1 k k k,l

Since ∂1 commutes with ik and ik, 1 Λd = ∂ i i e . z 2i 1 k k 1 k,l e1 and ik do not commute. For k , 1 ike1 = e1ik so that − 1 1 ∂1ikike1 = ∂1ike1ik 2i −2i k,1 k,1 k,1 k=,1 1 = ∂ e i i . 2i 1 1 k k k,1 k,1 116 Lecture 20

Now 129 ikek + ekik = 2

(as ikek is zero for a term which contains dzk as factor while ekik is zero in the contrary case). So 1 1 2 ∂kikikek = ∂kikekik + ∂kik 2i −2i 2i 1 = ∂kekikik i∂kik. 2i − Consequently, 1 Λd = ∂ i i e z 2i 1 k k 1 k,1 1 = ∂1e1ikik i ∂kik 2i − k,1

= dzΛ+ i∂z which proves the bracket relation. We shall now derive the bracket relation in the case of an arbitrary Kahlerian¨ manifold by using a theorem of differential geometry. Suppose we have two C tensor fields and of the same kind on ∞ ⊕ ⊕′ a C manifold V. We shall say that , coincide upto the order m at ∞ ⊕ ⊕′ 130 a point a V if the coefficients as well as the partial derivatives upto ∈ order m of the coefficients of and coincide at a. This has an intrinsic ⊕ ⊕′ meaning: for if the property is true for one map at a it is true for any other map at a. Let now V be a Riemannian manifold with the field of positive-definite quadratic forms Q. Let a be a point of V. According to a theorem of Riemannian geometry we can find another C∞ field Q′ of positive definite quadratic forms defined in a neighbourhood of a such that Q and Q′ coincide upto the first order at a and such that Q′ gives the Euclidean structure in a neighbourhood of a (i.e., there exists a map in which Q′ = δi jdxidx j where δi j is the Kronecker Symbol). Q′ is said to be an osculating Eu- clidean structure for Q at a. Lecture 20 117

We now consider the analogous question for Hermitian manifolds. 2n Suppose a C∞ manifold V has two different complex structures, (1) and (2). We shall say that the complex structures (1) and (2) coincide at a V2n upto order m if for every function f holomorphic in (1) the ∈ field (dz)a f is zero upto order m at a. Equivalently, we may say that the structures (1) and (2) coincide upto order m at a if the intrinsic J- operators J1 and J2 coincide upto order m on the tangent space at a. Let V(n) be a Hermitian manifold. Let H denote the field of Hermitian forms giving the Hermitian structure and Ω the associated exterior 2-form. Let 131 (n) a be a point of V . Is it possible to find a Hermitian structure, H′ on V (n) (with the same underlying C∞ structure as V ) such that i) the two analytic structures coincide upto order 1 at a

ii) the fields of Hermitian forms H and H′ coincide upto order 1 at a.

n iii) H′ is the canonical hermitian structure on C for some map at a?. There is one trivial necessary condition. If Ω′ is the exterior 2- form corresponding to H′, Ω and Ω′ coincide upto order 1 at a so that dΩ and dΩ′ coincide at a; but dΩ′ = 0 at a. Therefore, dΩ(a) = 0 is a necessary condition. This condition can also be proved to be sufficient; this is the difficult part. In particular in a Kahlerian¨ manifold there exists an osculating Hermitian structure at every point. Let V(n) be a Kahlerian manifold. At a V(n) we choose an ∈ osculating Hermitian structure. The operators J corresponding to the two structures coincide at a up to order 1. The operators 1 d = JdJ− coincide at a. Since dz is a linear combination of d and d the operators d coincide at a; similarly the operators d z z coincide at a. The operators Λ coincide upto order 1 at a and the operators [Λ, dz] coincide at a. The operators i∂z also coincide at a. Since we have proved the relation

[Λ, dz] = i∂z

for the canonical Hermitian structure in Cn the same relation holds 132 118 Lecture 20

also for any Kahlerian¨ manifold.

The relation [Λ, d ] = i∂z z − is proved similarly. Lecture 21

Compact Manifolds with a Kahlerian¨ structure 133 Let V be a compact complex analytic manifold. We shall assume that there exists a Kahlerian¨ metric on V. From this assumption we shall derive some intrinsic properties of V i.e., properties of V which depend only on the complex analytic structure of V but not on any particular Kahlerian¨ metric. Let H(r,s) denote the quotient space of the space of the d-closed forms of bidegree (r, s), by the space of forms of bidegree (r, s) which are coboundaries (not necessarily of homogeneous forms). Then the pth cohomology space H p is the direct sum of the spaces H(r,s), r + s = p. To prove this we observe that, as P(r,s) commutes with ∆, P(r,s) operates canonically (i.e., independent of Kahlerian¨ metric) on H p (see lecture 15). To define this map, we choose in each cohomology class the har- monic form, ω, belonging to this class; since ∆ and P(r,s) commute, each homogeneous component of the harmonic form is harmonic and hence closed. Thus the homogeneous components of the harmonic form define cohomology classes and P(r,s) is the map which maps the cohomology class of ω into the cohomology class determined by the (r, s) component of ω. If pr,sH is the image of H by P(r,s), P(r,s)H may be identified with the space of harmonic forms of bidegree (r, s) and

H p = P(r,s)H. But P(r,s)H is just the space H(r,s) defined above and hence 134

119 120 Lecture 21

H p = H(r,s) r+s=p This decomposition is intrinsic; but we have used harmonic forms to prove this. If we define the double Betti number b(r,s) to be the dimension of the space H(r,s) we have bp = b(r,s) r+s=p Since the mapping ω ω, which assigns to every form its complex → conjugate, induces an isomorphism of H(r,s) onto H(s,r), we find that b(r,s) = b(s,r); so, when p is odd bp is the sum of numbers which are pairwise equal and hence bp is even. So the Betti numbers for odd di- (n r,n s) (r,s) mensions are even. Moreover H − − is dual to the space H so that (n r,n s) (r,s) b − − = b . Thus we have

(r,s) (s,r) (n r,n s) (n s,n r) b = b = b − − = b − −

We can also prove that bp is even for odd p by introducing a canon- ical complex structure on the pth cohomology space H p(R) formed from the real forms alone. [H p is the complexification of H p(R) and the com- 135 plex dimension of H p is equal to the real dimension of H p(R)]. Since J commutes with ∆, J operates canonically on H p(R). Since p is odd J2 = I and J gives a complex structure on H p(R). So the (real) dimen- − sion of H p(R) is even; and hence bp is even. The next result on the Betti-numbers is the following:

(r,s) (r 1,s 1) b b − − if r + s n + 1 ≥ ≤ (From this it follows at once that bp bp 2 if p n + 1). To prove this ≥ − ≤ we need the following result: the map

r 1,s 1 r,s − − Ω : Λ T ∗(V) ΛT ∗(V) a → a (multiplication by Ω) is one to one (i.e., Ω ω = 0 if and only if ω = 0) ∧ provided r + s n + 1. Since L commutes with ∆, L gives a map of ≤ the space of harmonic forms of bidegree (r 1, s 1) into the space of − − Lecture 21 121 harmonic forms of bidegree (r, s); by the algebraic result stated above this map is one to one if r + s n + 1; consequently b(r 1,s 1) br,s, if ≤ − − ≤ r + s n + 1. ≤ r 1,s 1 r,s The map Ω : − Λ− Λ → r 1,s 1 r,s We shall now prove that the map Ω : − Λ− Λ is one to one for → r + s n + 1. Since Ω is an operator of type (1, 1) it is sufficient to prove ≤ that q 2 q Ω : Λ− Λ → is one to one for q n + 1. This would follow if we prove that the map 136 ≤ n p n+p Ωp : Λ− Λ → is one to one for p 1. For, if q n + 1, q 2 = n p, for some p 1 ≥ ≤ − − ≥ and Ω ω = 0 Ωp ω = 0 ω = 0 ∧ ⇒ ∧ ⇒ since n p n+p Ωp : Λ− Λ n p n+p → is one to one. Since Λ− and Λ are of the same dimension it is enough p to show that the map Ω is onto. Let (x1, y1,..., xn, yn) be a basis for the space of real differentials at a such that

Ω= xi yi ∧ n+p Put xi yi = αi. The elements of Λ are generated by elements of the ∧ form xA yB, where the set of indices A and B have at least p indices ∧ in common. Consequently it is sufficient to prove that these elements p ω = xA yB are divisible by Ω . We may assume that ∧ ω = α1 ... αp+s xC yD ∧ ∧ ∧ ∧ where the indices C and D have no elements in common. Since the transformation xk yk, yk xk for indices k in D does not affect Ω → → − we may assume that ω is of the form

α1 ... αp+s xp+s+1 ... xn s. ∧ ∧ ∧ ∧ ∧ − 122 Lecture 21

Since 137 p+s (p + s)!α1 ... αp+s = (α1 + + αp+s) ∧ ∧ (the exponent represents power with respect to the exterior product), we have

ω p+s = (α1 + + αp+s) xp+s+1 ... xn s (p + s)! ∧ ∧ ∧ − p+s = (α1 + + αp + αp+1 + + αp+s + + αn s) − xp+s+1 ... xn s ∧ ∧ ∧ −

If we put γ = αn s+1 + + αn − ω p+s = (Ω γ) xp+s+1 ... xn s (p + s)! − ∧ ∧ ∧ − + p+s p+s 1 s p s p s = [Ω (p + s)Ω − γ + + ( 1) Ω γ ]. − 1 − s

. xp+s+1 ... xn s ∧ ∧ ∧ − as γs+1 = 0. The left side containing Ωp as a factor.

The space H(p,0)

We shall now show that the space H(p,0) is just the space of holomor- phic differential forms of degree p. A closed differential form of degree (p, 0) is z closed (by homogeneity) and hence holomorphic; and a holo- morphic form (of degree p) is harmonic and hence closed. On the other hand, since a holomorphic differential form is harmonic, a closed dif- 138 ferential form of bidegree (p, 0) cannot be a coboundary unless it is the zero form. From this we derive at once a majorant (in terms of the pth Betti- number) for the number of linearly independent holomorphic p-forms. Since bp = b(p,0) + + b(0,p) and b(p,0) = b(0,p) we have 2b(p,0) bp(for p , 0) ≤ Lecture 21 123

For differential forms of degree 1 we have

b1 = b(1,0) + b(0,1) = 2b(1,0) i.e., the dimension of the space of holomorphic differential forms of de- gree 1 is equal to half the first Betti number.

Compact Riemann Surfaces

A complex analytic manifold of complex dimension 1 is usually called a . We can always introduce a Kahlerian¨ metric on a Riemann surface. Let V(1) be a compact, connected Riemann surface; b2 = b0 = 1. Let b1 = 2g. The number g (= half the first Betti number) is called the genus of the Riemann surface. The number of linearly independent holomorphic forms of degree 1 is equal to the genus of the surface, by what we have seen. Since there are 2g independent 1-cycles and g independent holomorphic 1-forms the periods of a holomorphic 1-form cannot be prescribed arbitrarily on a basis of 1-cycles. However it can be proved that there exists a unique holomorphic differential form 139 with prescribed real parts of the periods. The Riemann sphere, S 2, is of genus zero. So there are no holomor- phic differential forms of degree 1 apart from the 0-form. Of course, this can be proved directly. Let ω be a holomorphic 1-form on S 2. ω can be written as f (z)dz in the plane, where f (z) is an in the plane. Using the map given by 1/z at we find that f (1/z). 1/z2 ∞ should be holomorphic at the origin. If a f (z) = a zn then f (1/z)1/z2 = n n n+2 z so that f (1/z). 1/z2 has a pole at the origin unless f 0. ≡ Next we consider a torus with the complex structure induced from C1. Here the genus is 1. So the differential dz (which is well defined on the torus) is, but for a constant multiple, the only holomorphic 1-form on the torus. 124 Lecture 21

All other compact Riemann surfaces can be considered as the quo- tient spaces of the unit circle by certain Fuchsian groups. Lecture 22

The identity between d, z and z cohomologies: de Rham theorem (Compact Kahlerian¨ manifolds) 140 We shall now prove that the cohomology spaces defined by the operators d, dz, dz and d are canonically isomorphic. That the ordinary cohomol- ogy and the z cohomology are the same will be of importance in the study of holomorphic and meromorphic forms on a compact Kahlerian¨ manifold. We shall give the proof in the case of the ordinary cohomology and the z cohomology. Let H be a cohomology space with respect to d and Hz the corresponding z cohomology space. We have a canonical 0 mapping from H to Hz. In each cohomology class ω (with respect to 0 d) we choose a form, ω, which is z closed and map the class ω to the z cohomology class determined by ω. In each d-cohomology class such a form exists, the harmonic form belonging to the class. We have to verify that if a form, α, which is a d-coboundary is z closed then α is a z coboundary. Since α is a d-coboundary πα1 = 0. By the decomposition formula

α = π1α + 2dz∂zGα + 2∂zdzGα

= 2dz∂zα so that ω is a z coboundary. The mapping so defined is actually an isomorphism. For the space Hz can be identified with the space of har- monic forms and we know that the space H also can be identified with 141

125 126 Lecture 22

the space of harmonic forms. We have thus a canonical isomorphism between d and z cohomologies, which is independent of the Kahlerian¨ metric. Thus the z cohomology for a compact complex analytic manifold with a Kahlerian¨ metric is just the ordinary cohomology. The first and third parts of de Rham’s theorem for z cohomology for such manifolds follow immediately. The projections π1, π = 2d ∂ G, π are contin- 2,z z − z 3,z uous and the image spaces corresponding to these projections are closed because they are kernels. This proves the second part of de Rham’s theorem for z cohomology.

de Rham theorems for z cohomology of an arbitrary complex analytic manifold

The first part of de Rham’s theorem for z cohomology in the case of an arbitrary manifold was proved only recently. The third part of de Rham’s theorem is also true: the z Betti-numbers are finite for a compact mani- fold. However it is not true in general that the spaces of z coboundaries are closed. On the other hand if we assume that all the Betti-numbers are finite this theorem can be restored.

The complex projective space

Let PCn denote the complex projective space of n dimensions PCn is a compact Kahlerian¨ manifold (see the appendix). Algebraic manifolds 142 imbedded without singularities in PCn are also compact Kahlerian¨ man- ifolds. (In general it is true that a complex analytic manifold regularly imbedded in a Kahlerian¨ manifold is a Kahlerian¨ manifold). For PCn we have bp = 1 for even p and bp = 0 for odd p. The form Ωk (Ω is the exterior 2-form associated with the Kahlerian¨ metric) gives an element , 0 of the pth cohomology space. In this connection we shall state another part of de Rham’s theo- rem (arbitrary C∞ manifold, d-cohomology). This part of de Rham theorem states that each cohomology class of currents contains a cy- Lecture 22 127 cle (closed chain) and that a cycle which is the coboundary of a current is the boundary of a chain. We shall also make use of de Rham’s cor- respondence between cohomology of forms and homology of chains in which exterior products of cohomology classes correspond to algebraic intersections of homology classes. Let PCn 1 be a hyperplane of PCn PCn 1 is a cycle. PCn 1 deter- − − − mines an element (, 0) of the 2nd cohomology class (of currents). So Ω is homologous (in the sense of the currents) to k PCn 1 where k is a − real number: Ω k PCn 1. Actually k > 0. For, ≈ − n n 1 n 1 Ω = Ω, Ω − = k Ω − n n 1 PC PC − n n 1 Since PC and PC − are Kahler¨ manifolds (with the associated exterior 2-form Ω),

n n 1 Ω > 0 and Ω − > 0 so that k > 0. n n 1 PC PC − Considering n hyperplanes in general position (whose intersection is a 143 point PC0) we find that

Ωn knPC0 ≈ kn x point (as a chain or current). ≈ Therefore Ωn = kn PC0, 1 = kn. PCn So n k = n Ω . PCn If we choose the Kahlerian¨ metric whose associated 2-form is

n Ω′ =Ω/ n Ω PCn 128 Lecture 22

we have Ω p PCn p. In other words if the associated 2-form Ω of a ′ ≈ − Kahlerian¨ metric satisfies the relation

Ωn = 1 PCn

then p n p Ω PC − . ≈ The volume of PCn with respect to the volume element given by such a metric is n!. The Betti numbers of PCn verify, of course, all the properties of the Betti numbers of a compact Kahlerian¨ manifold. We have

b0,0 = b1,1 = ... = bn,n = 1.

144 and all the other double Betti-numbers are zero. In particular bp,0 = 0 for p , 0. So on the complex projective space there are no holomorphic differentials of any degree except the degree zero (in which case the holomorphic forms are constant functions).

Example of a compact complex manifold which is not Kahlerian¨

Let us consider in Cn the shell between the two spheres of radii 1 and 2:(1 z 2). Identify the two points on the spheres which are on the ≤ | |≤ same radius. Let us denote by V the space obtained after this identification. Let G denote the properly discontinuous group of analytic automorphisms of (Cn 0) consisting of the homothetic transformations. − k k (z1,..., zn) (2 z1,..., 2 zn) → k running over all integers, positive, negative or zero. V is a fundamen- tal domain for this group. Since we can always introduce a complex Lecture 22 129 analytic structure in the quotient space obtained from a properly discon- tinuous group, V can be endowed with a natural analytic structure. V is compact. V is not Kahlerian¨ for n > 1. It can be proved that, for n > 1,

b0 = b1 = 1, bi = 0 for 2 i 2(n 1). ≤ ≤ − 2n 1 2n b − = 1, b = 1.

For n 3, the Betti-numbers bp are not greater than 1 for even p. For 145 ≥ n 2 odd dimensional Betti numbers are not even. ≥ If P(z1,..., zn) and Q(z1,..., zn) are homogeneous polynomials of the same degree P(z1,..., zn)/Q(z1,..., zn) defines a meromorphic func- tion on V. In this connection we may remark that there exist compact complex analytic manifolds which admit of no non-constant meromor- phic function. Examples are provided by certain torii.

Lecture 23

Cousin’s Problem 146 Cousin’s problem for meromorphic functions in the complex plane is Mittag-Leffler’s problem. Mittag-Leffler’s problem in the plane is as fol- lows: Given a discrete set of points in the plane and polar developments at each of these points, construct a function meromorphic in the whole plane having the given points as poles and the given developments as the polar developments. We know that this problem always admits of a solution. Let ai be the given points and = Ci,1 + Ci,2 + + Ci, pi Pi(1/z ai) p − (z ai) (z ai) (z ai) i − − − be the polar development at ai. Then we can find Qi(z) such that the 1 (Pi Qi(z)) z ai − − coverges absolutely and uniformly on every compact set not containing anyone of the ai. The limit function gives a solution of the problem. The solution is unique upto an additive entire function. The indeterminacy in the problem is thus an entire function. Let us consider the corresponding problem on the Riemann sphere. 1 We have a finite number of points ai and polar developments Pi z ai − at ai , and a polar development P(z) (a polynomial without constant ∞ term) at . Then the function ∞ Pi(1/z ai) + P(z) − 131 132 Lecture 23

147 is meromorphic on the Riemann sphere and has the prescribed devel- opment at the given points. The solution is unique upto an additive constant. The method of construction of the solution corresponds to the classical decomposition of a rational function into partial fractions. We want to consider similar problems on complex analytic mani- folds. We shall first define a . f (z ,..., z ) Suppose we consider in Cn a ‘function’ ϕ = 1 n which is g(z1,..., zn) the quotient of two holomorphic functions f and g. The variety of poles is the set of zeros of g(z1,..., zn). The variety of poles is not, in general, a true manifold; if all the partial derivatives of g vanish at a point this point is a singular point for this variety. (For example the variety defined 2 + 2 + 2 = by z1 z2 z3 0 has a singularity at the origin). We shall call the set of zeros of an analytic function an analytic subset. At points where the analytic subsets defined by the zeros of the functions f and g intersect the quotient f /g is indeterminate. (In the case n = 1, if f and g are coprime at a point a it is impossible to have f (a) = 0 and g(a) = 0 simultaneously. However for n 2 this is possible; for example the ≥ functions z1 and z2 are coprime at the origin and vanish simultaneously 148 at the origin). So the function f /g is defined only in the complement of certain analytic subsets. It can be shown that an analytic subset is a set of measure zero. It is not good to define a meromorphic function on V(n) to be the quotient of two holomorphic functions on V(n); as then the only meromorphic functions on a compact complex analytic manifold would be constants! These considerations lead to the following definition of a meromorphic function on a complex analytic manifold. A meromorphic function, ϕ, on a complex analytic manifold is a complex valued function defined almost everywhere on the manifold such that for every point a of the manifold there exists a neighbourhood Ua of a such that, in Ua, ϕ is almost everywhere equal to the quotient of two holomorphic functions defined everywhere on Ua. A meromorphic form of degree p is a field of covectors defined al- most everywhere on the manifold such that every point has a neighbour- hood in which the form is almost everywhere equal to the quotient of a holomorphic form of degree p by a holomorphic function. Suppose we are given an open covering (Ui) of the manifold and in Lecture 23 133

p every Ui a meromorphic differential form ωi of degree p such that the form ωi j = ωi ω j is almost everywhere in Ui U j equal to a holo- − ∩ morphic form defined everywhere in Ui j = Ui U j. These constitute ∩ a Cousin’s datum. Cousin’ s problem is to find a meromorphic form p ω on the whole manifold such that ω ωi is a holomorphic form in − Ui (i.e., ω ωi) is almost everywhere equal to a holomorphic form de- − fined everywhere in Ui). The indeterminacy of the problem is evidently 149 an additive holomorphic form (defined over the whole manifold). ωi are “singular parts” of ω in Ui. Cousin’s datum for the Mittag-Leffler’s problem is the following: With each ai we associate an open set Ui such that Ui U j is empty for i , j. In Ui we take for ωi the meromor- ∩ phic function given by the polar development. In the complementary set U0 of the points ai we take the function ω0 0. All these together ≡ constitute the Cousin datum for the Mittag-Leffler problem. We now wish to formulate the Cousin datum and problem in terms of currents. If n = 1, a meromorphic function (or a form) can be con- 1 sidered as a current. In C1 the function defines a current in the (z a) 1 − usual manner. But the function , k > 1, is not summable in any (z a)k neighbourhood of a. However if− we take the Cauchy principal value, 1 defines a current: for every ϕ D (z a)k ∈ − 1 ϕdx dy ,ϕ = lim (z a)k ǫ 0 " (z a)k − → z a ǫ − | − |≥ This enables us to consider meromorphic functions or forms as currents when n = 1; considering meromorphic functions and forms as currents in this way leads to good solutions of problems on a compact Riemann surface. However, it becomes difficult to associate canonically a current with an arbitrary meromorphic function in higher dimensions. If the an- alytic subset defined by the singularities is a true manifold it is possible to associate canonically a current with the meromorphic function. In the 150 case of a general meromorphic function it has not yet been possible to associate canonically a current with the meromorphic function. 134 Lecture 23

We shall introduce currents in the problem in some other way. We p,0 can find currents Ti (of bidegree (p, 0)) in Ui such that Ti T j = ωi j (as − a current) in Ui j. The indeterminacy in the choice of the Ti is a current defined on the whole manifold. If S is a current on the whole manifold T = Ti + S also possess the same property: T T = ωi j in Ui j, i′ i′ − ′j conversely if Ti and Ti′ are two systems of currents such that Ti and Ti′ are defined in Ui and

Ti T j = ωi j in Ui j, − T ′ T ′ = ωi j in Ui j, i − j = + there exists a current S on the whole manifold such that Ti′ Ti S . We define the current S by ‘piecing’ together the currents S i defined in Ui by S i = T Ti. The currents S i define a single current on the whole i′ − manifold as we have, in Ui j,

S i S j = (T ′ Ti) (T ′ T j) − i − − j − = (T ′ T ′) (Ti T j) i − j − − = ωi j ωi j − = 0

151 To find the currents Ti we proceed as follows. We choose a partition of unity αi subordinate to the covering system Ui . We put { } { }

Ti = αkωik (the summation being over all k for which Ui Uk is non-empty) where ∩ αkωik is the C∞ form defined in Ui as:

αkωik, in Ui Uk αkωik = ∩ 0 in Ui (complement of support of αk)  ∩  The definition of αkωik is consistent, as

αkωik = 0 on Ui Uk (complement of support of αk). ∩ ∩ Lecture 23 135

Ti is a C∞ form in Ui. Now

Ti T j = αk(ωik ω jk) in Ui j. − − αk(ωik ω jk) is zero outside Ui jk = Ui U j Uk. − ∩ ∩ In Ui jk we have the relation

ωi j + ω jk + ωki = 0 and, in Uik, the relation ωik + ωki = 0 so that 152

Ti T j = αk(ωik ω jk) − − k

= αkωi j k

=  αk ωi j k    = ω i j, in Ui j as αk = 1. k This result is a particular case of a more general one. Suppose we have an open covering Ui and a system of currents ωi j defined in Ui j = { } Ui U j such that ∩ ωi j + ω jk + ωki = 0 in Ui U j Uk ∩ ∩ and ω + ω = 0 in U . i j ji i j

Then it is possible to find a system of currents Ti in Ui such that

Ti T j = ωi j in Ui j − and Ti is given explicitly by

Ti = αkωik 136 Lecture 23

where αk is a partition of unity subordinate to the covering Ui. If ωi j are C∞ forms, Ti also can be chosen to be C∞ forms. [One can also consider the following problem which is more gen- 153 eral: given an open covering Ui and a system of holomorphic forms { } ωi j in Ui j such that

ωi j + ω jk + ωki = 0 in Ui jk

and ωi j + ω ji = 0 in Ui j,

to find holomorphic forms hi in Ui such that

hi h j = ωi j in Ui j] −

By means of the currents Ti we define a current of bidegree (p, 1) defined on the whole manifold. We put Ri = dzTi in Ui. In Ui j,

Ri R j = d (Ti T j) = d ωi j = 0 − z − z

(as ωi j is holomorphic). So the currents Ri define a single current defined p,1 on the whole manifold, which we denote by R . Since the currents Ri p,1 p,1 are z closed, the current R is also closed. ( R is locally a coboundary but need not be a coboundary in the large). If we replace Ti by a { } system T having the same properties as Ti, T = Ti + S , S a current { i′} i′ defined on the whole manifold, the ‘R’ corresponding to Ti′ would be p,1 R + dzS . So we can associate canonically with the Cousin datum a whole z cohomology class of bidegree (p, 1). This z cohomology class is the ‘obstruction’ to the solution of Cousin’s problem. 154 We shall prove that in order that Cousin’s problem be solvable it is necessary and sufficient that the z cohomology class associated with the Cousin datum is the zero class. Suppose Cousin’s problem is solvable and let ω be a solution. Let further ω ωi = hi, hi holomorphic in Ui. − Take Ti = hi. In Ui j, −

Ti T j = hi + h j − − Lecture 23 137

= (ω ωi) + (ω ω j) − − − = ωi ω j − = ωi j.

dz Ti = d hi = 0 − − z

So the current of bidegree (p, 1) associated with the system Ti is zero, or the z cohomology class associated with the Cousin datum is the zero class. Conversely if the associated z cohomology class is the zero class, Cousin’s problem is solvable. In this case we can find Ti such that the p,1 associated R is zero (we may have to adjust the S suitably). That is, we find Ti such that

Ti T j = ωi j in Ui j and d Ti = 0. − z 0 Then by the ellipticity of dz (on D′), Ti is a holomorphic form hi. Then a solution is given by the form ω = ωi hi. In Ui j, ωi hi = ω j h j so − − − that ω is a meromorphic form well-defined on the manifold. p,1 p,1 Let R be the current defined by: R = dzTi in Ui. If R = dzS , a 155 solution of the Cousin’s problem is given by the form ω:

ω = ωi + S Ti in Ui. − In a compact Kahlerian¨ manifold d and z cohomologies coincide and with a Cousin datum we have an ordinary cohomology class. As an example let us consider the Riemann sphere. Here b′ = 0; so b0,1 = 0. So Cousin’s problem for meromorphic functions is solvable for any Cousin datum.

Lecture 24

Cousin’s Problem on Cn 156 n (p,q) = It has been proved recently that for C (the complex n-space) Hz 0 for q > 0. So Cousin’s problem for meromorphic forms is solvable in Cn for any Cousin datum.

Cousin’s problem on a compact Kahlerian¨ mani- fold: pseudo-solution

We can give a more explicit solution in the case of a compact Kahlerian¨ p,1 manifold. A necessary and sufficient condition for R to be a z cobound- p,1 ary is that π1R = 0. (This gives a system of b linear conditions for R). Now

R = π1R + 2dz∂zGR + 2∂zdzGR

= π1R + 2dz∂zGR.

We can choose S = 2∂zGR. Then ω = ωi + 2∂ GR Ti in Ui z − is a solution of the Cousin’s problem. Even in the case when Cousin’s problem has no solution, ω defined as above makes sense. Even if Ti are currents, (2∂ GR Ti) is a C form; for z − ∞ d (2∂ GR Ti) = 2d ∂ GR d Ti z z − z z − z 139 140 Lecture 24

= (2d ∂ GR R) z z − = π1R. −

157 π1R is a C form and (2∂ GR Ti) is a current of bidegree (p, 0); by ∞ z − the ellipticity of d , (2∂ GR Ti) is a C form. ω is thus a sum of a z z − ∞ meromorphic form and a C∞ form. When Cousin’s problem is solvable (2∂ GR Ti) is a holomorphic form and ω is a solution of Cousin’s z − problem. In any case, we call ω a pseudo-solution. When Cousin’s problem is solvable, the pseudo-solution is actually a solution.

Cousin’s problem on PCn

For p , 1 Cousin’s problem has a solution for any Cousin datum, since bp,1 = 0 for p , 1. The solution is unique if p 2 and is determined ≥ upto an additive constant when p = 0, as holomorphic differentials of degree p 1 are zero and those of degree zero are constants. For p = 1, ≥ for Cousin’s problem to be solvable it is necessary and sufficient that R be orthogonal to all closed C forms of bidegree (n 1, n 1). But Ωn 1 ∞ − − − is a generator of the cohomology classes of bidegree (n 1, n 1). So − − the condition is n 1 R, Ω − = 0. When the solution exists, the solution is unique.

Cousin’s problem on a compact Riemann Surface

We shall now consider Cousin’s problem on a compact Riemann sur- face, which we assume to be connected. We shall first consider the Cousin’s problem for meromorphic differential forms of degree 1. For 1,1 Cousin’s problem to be solvable, it is necessary and sufficient that R 1,1 158 ( R is a current of the associated cohomology class) be orthogonal to all closed zero forms i.e., constants: or

1,1 R , 1 = 0. Lecture 24 141

We shall now interpret this condition in terms of residues. Let ω be a meromorphic differential form of degree 1. We define the residue of ω at a pole a of ω, denoted by Resa ω, as follows: We choose a local coordinate system (z) at a ω = f (z)dz, where f (z) is a meromorphic function of z. Then Resa ω is defined to be the residue of f (z)dz at z(a). This definition is intrinsic. If we choose a regular curve C, contained in the domain of a map, winding around a once in the positive sense 1 Resa( f (z)dz) = f (z)dz 2πi C 1 = ω. 2πi C Since the Riemann surface has no boundary, the sum of the residues of a meromorphic differential of degree 1 is zero. This gives a trivial necessary condition on the Cousin datum, for Cousin’s problem to be solvable. We shall see that this condition is also sufficient. Let a1,..., am be a finite number points given on the Riemann sur- face. At each a we have a map (Wa,ϕa) such that Wa Wb is empty for ∩ a , b. Let Ua be a neighbourhood of a such that Ua Wa and ϕa(Ua) ⊂ is a closed disc. In each Wa we are given a meromorphic form of degree 159 1, ωa, having a pole only at a. Let U0 be the complement of the set of points a1,..., am. In U0 we put ω0 0. With this Cousin datum we ≡ associate currents Ta in Ua and T0 in U0 defined as follows:

Ta = 0 in Ua

T0 = ωa in U0 − where ωa in Ua ωa = 0 outside Ua.   The form ωa has discontinuities along the boundary of Ua; but ωa is locally summable in U0 and defines a current in U0. In Ua U0, ∩ Ta T0 = ωa = ωa − 142 Lecture 24

The current d T in U R = z a a  dzT0 in U0   is the “obstruction”. R = 0 in Ua. Since T0 is of bidegree (1, 0), dzT0 = dT0 =. If ϕ is a C∞ function with compact support in U0 (i.e., if the support of ϕ does not contain the points ai),

1,0 d T0,ϕ = dT0,ϕ z = T0, dϕ = ωa, dϕ − = ωa dϕ − " ∧ a Ua = d(ω ϕ) " a a Ua

= ωaϕ. a bUa

160 Consequently R = (bUa) ωa ∧ in U0 and this relation is also true in Ua. [ωa is a C∞ form in a neigh- bourhood of the support of bUa. So multiplication of bUa by ωa is pos- sible]. Now the necessary and sufficient condition for Cousin’s problem to have a solution is R, 1 = 0 or

ωa = 0 a bUa i.e., Resa ωa = 0. a Lecture 24 143

In particular if we take ωa without residues a solution of the problem always exists. The indeterminacy in the solution is a holomorphic differential form 161 of degree 1. Thus in this problem we have one condition of compatibility and g degrees of indeterminacy. If we choose a Kahlerian¨ metric on the Riemann surface, an explicit solution is given by

ω + 2∂ GR in U ω = a z a  2∂zGR + ωa in U0.   We may write 

ω = ωa + 2∂ G[ bUa ωa ] z { ∧ } a a as ωa = ωa in Ua. This expression has always a meaning and gives a solution of the Cousin’s problem when the problem is solvable. The general solution of Cousin’s problem is given by

ω = ωa + 2∂ G[ (bUa ωa)] + h z ∧ a a where h is a holomorphic differential form of degree 1 on the Riemann surface.

Cousin’s problem for meromorphic functions (Compact Riemann Surface)

Here in each Ua we have a meromorphic function fa instead of a mero- morphic differential form. In U0 we take the function f0 = 0. Let f˜a be the function defined on the surface by

fa in Ua fa = 0 outside U  a   We consider the currents Ta = 0 in Ua and Ta = fa in U0. Let 162 − 144 Lecture 24

0,1 1,0 R be the associated current of bidegree (0, 1). If ϕ is a C∞ form with compact support in U0

1,0 d T0, ϕ = T, d ϕ z − z = fa, d ϕ z a = d ( f ϕ) " z a a Ua = d( f ϕ) " a Ua

= faϕ. bUa Consequently, 0,1 R = (bUa) fa. a A necessary and sufficient condition for this Cousin’s problem to have a solution is that R be orthogonal (with respect to , ) to all harmonic forms of bidegree (1, 0) i.e., R be orthogonal to all holomorphic forms of degree 1 or

fah = 0 bUa for every holomorphic form h of degree 1. For Cousin’s problem for meromorphic functions we have g inde- pendent compatibility conditions and 1 degree of indeterminacy while 163 for Cousin’s problem for meromorphic forms we had one compatibility condition and g degress of indeterminacy. This suggests a sort of duality between Cousin’s problems for meromorphic functions and forms. This duality will be made precise in the theorem of Riemann-Roch. The above results prove the existence of lots of meromorphic func- tions and forms on a compact Riemann surface. Lecture 25

Some applications 164 Before proving the Riemann-Roch theorem we shall make some appli- cations of the existence theorems proved above. We shall first prove that every compact connected Riemann Surface, V, of genus zero is analytically homeomorphic to the Riemann sphere, S 2. Since the genus of V is zero, there are no compatibility conditions on any Cousin’s problem for meromorphic functions. So we can construct a meromorphic function, f , having a simple pole at a point a and regu- lar elsewhere. Every meromorphic function on a compact (connected) Riemann surface assumes every value in S 2 the same number of times. So f assumes every value in S 2 exactly once and maps V conformally onto S 2. In the case of a torus (g = 1) there is no distinction between mero- morphic functions and meromorphic forms of degree 1 (because of the existence of dz). Evidently, the problem of finding an elliptic func- tion with prescribed periods and singularities is a problem of finding a meromorphic function or a meromorphic differential of degree 1 with prescribed singularities on a torus. Therefore our theorem yields the existence of elliptic functions for which singularities are prescribed in the fundamental parallelogram with the restriction that the sum of the residues at the singularities is zero. The function is determined uniquely upto an additive constant by the singularities. In particular if we pre- scribe the principal part 1/z2 at the origin and choose the constant term 165 in the Laurent development at the origin to be zero, we obtain the Weier-

145 146 Lecture 25

strass P-function.

The Riemann-Roch Theorem

Let V(1) be a compact connected Riemann Surface of genus g. A divi- sor D on V is a formal linear combination of points of V with integer coefficients such that all but a finite number of coefficients are zero:

D = αp p βqq; − p, q are points of V, αp, βq integers, αp > 0, βq > 0. The degree of the divisor D is defined to be the integer ( αp βq). We write A = αp, = − B βq. Meromorphic function f on V is said to be a multiple of the divisor D if at every point p f has a zero of order αp and at every ≥1 point q, f has a pole of order βq. For example in C a function f is a ≤ multiple of the divisor D if and only if

αp β f = h (z p) (z q)− q − − where h is an entire function. To find a meromorphic function which is a multiple of a given divisor is to find a meromorphic function for which the maximum number of poles with the maximum order at each pole and the minimum number of zeros with the minimum order at each zero are 166 prescribed. Since the order of a zero or a pole of a meromorphic differ- ential form of degree 1 has an invariant meaning we can similarly define what it means to say that a differential form is a multiple of D. (The 0 order is defined by means of a map). Let m(D) denote the dimension of the vector space of the meromorphic function which are multiples of D 1 and m(D) that of the vector space of the meromorphic differential forms which are multiples of D. The Riemann-Roch theorem asserts that

0 1 m( D) m(D) = d g + 1 − − − (d is the degree of the divisor D; D is the inverse of D). − Lecture 25 147

To prove the Riemann-Roch theorem let us first consider the mero- morphic functions which are multiples of D. The singular part of such − a function at a point p can be written in a map as

j fp = C j,p1/zp jαp ≤

(zp is the function defining the map). The singular parts fp give a 1 Cousin’s datum. If we denote by [ j ] a pseudo-solution (which de- zp pends on the choice of Ui and the metric) associated with the Cousin 1 datum j , a pseudo-solution associated with the Cousin datum fp is zp 1 ω = + C j,p[ j ] C p,j αp zp ≤ and this gives the general (true) solution when the Cousin problem has 167 a solution. As conditions for the Cousin datum to be compatible we obtain

j C j,p Resp(hν/z ) = 0, (ν = 1, 2,..., g) − p p,j αp ≤ where hν is a basis for the space of holomorphic forms of degree 1. { } (The minus sign on the left side is introduced for later convenience). To write the condition that the solution has a zero of order βq at q it is ffi k ≥ su cient to write that the product of ω by dzq/zq has zero residue at q for k βq. This residue exists if ω is a meromorphic form, otherwise it ≤ is defined by 1 dzq ω k 2iπ zq bUq So we obtain the equations

1 dzq 1 C dzq C j,p [ ] + = 0 2iπ zk j 2iπ zk p,j αp q zp q ≤ bUq bUq 148 Lecture 25

for every q and for every k βq. We have here A + 1 unknowns, C j,p ≤ and C, and g + B equations:

g(ν = 1, 2,..., g) B(q, k βq). ≤ A = αpB = βq 168 Next let us consider the meromorphic forms of degree 1 which are multiples of D. The singular part of such a differential in a map at q can be written as dzq dk,q zk q,k βq q ≤ A general pseudo-solution of the associated Cousin problem is

dz dk,q + eνhν  zk  q,k βq q ≤     If we write down the condition for the solution of the Cousin problem to exist, we obtain

dzq dk,q Resq = 0.  zk  q,k βq q ≤     That the solution should have a zero of order αp at p gives the condi- ≥ tions

1 1 dzq j  dk,q + eν Resp(hν/zp) = 0. −  2iπ j  zk   q,k zp  q    bUp          In this case we have B + g unknowns, dk,q and eν, and A + 1 equations (compatibility condition and A(p, j αp) equations). We shall now ≤ show that the systems of equations

1 j C j,p Resp(hν/z ) = 0(ν = 1,..., g), − 2πi p (I)  C 1 dzq 1 C dzq = 0  j,p 2πi zk j 2πi zk p, j p zp q  bU q bU q   Lecture 25 149

169 and

d Res dzq = 0 k,q q zk  q (II)   1 1 dzq + hν =  dk,q j k eν Resp j 0   2iπ z zq z  − q,k βq bU p p   ≤ p     are transposes  of each other. We have only to verify (the rest of the verification being trivial) that dz dz 1 q = q 1 − j  zk  zk  j  zp  q  q zp  bUp   bUq       Now,

1 1 bUp = + 2∂zG z j  z j  z j   p  p  p      dzq  dz  bUq  dzq = + 2∂zG ∧  zk  zk  zk   q  q  q      (Refer to the last lecture).    1 dz Since = 0 on bUq and = 0 on bUp is remains only to verify 170 j zk zp p that

1 (bUq dzq) 2∂ G ∧ j z k − zp zq bUp

dzq 1 = 2∂ G bU zqk z  p j   zp  bUq     or

1 dzq (bUp) , 2∂zG bUq − j  ∧ zk  zp  q    1  dzq = 2∂ G bU , bU . z  p j  q k zp ∧ zq     150 Lecture 25

If we put 1 dzq S = bU and T = bU . p j q k zp ∧ zq We have to verify that

1 1 2 0 S ,∂ GT = T,∂ GS − z z Write 2∂zGS = U, 2∂zGT = V We have then S = π1S + dzU, T = π1T + dzV. (as S and T are z closed). So we have to verify that

π1S + d U, V = U, π1T + d V . − z z 171 Now π1S, V = 0 as π1S is harmonic and V is a ∂ coboundary. Simi- z larly U, π1T = 0. It remains to show that d U, V = d V, U . − z z We define the set of singularities of a current T to be the smallest closed subset of the manifold in the (open) complement of which T is an indefinitely differentiable form (such a set exists, for, if T, is a form in a family of open subsets then it is a form in their union). Now let S and T be two currents whose sets of singularities have no common point; then it is possible to give a meaning to S, T . It is possible to find decompositions.

S = S 1 + S 2, T = T1 + T2

where S 2 and T2 are C∞ forms and S 1 and T1 are currents whose sup- ports have no common point. (For example, let 1 = α + β be a partition of unity subordinate to ∁F, ∁G where F and G are the supports of S and T respectively. We may take S 1 = βS , S 2 = αS , T1 = T2 = T). By definition we put:

S, T = S 1, T2 + S 2, T1 + S 2, T2 ; Lecture 25 151

each bracket on the right side has a meaning since S 2 and T2 are C∞ forms ( S 1, T1 is not written; we take it by definition to be zero which is natural because the supports of S and T have no common point). This definition is correct provided we prove i) it is independent of the choice 172 of the choice of the decompositions of S and T and ii) it gives the usual result if S or T is a form. Suppose we have another decomposition S i′, = ffi Ti′(i 1, 2). It is su cient to prove that the decompositions S i, Ti and S i, Ti′ give the same result; for then the decompositions S i, Ti′ and S i′, Ti′ will also give the same result for an analogous reason. We have to show that

S 1, T ′ T2 + S 2, T ′ T1 + S 2, T ′ T2 2 − 1 − 2 − is zero. The last two brackets can be added by linearity because S 2 is a C form and their sum is S 2, 0 = 0. It remains to prove that S 1, T ∞ 2′ − T2 = 0. We shall add to S 1, T T2 the expression S 1, T T1 2′ − 1′ − which has a meaning because T T1 is a form ((T T1) + (T T2) 1′ − 1′ − 2′ − is the 0 form and T T2 is a form) and which is equal to zero since the 2′ − supports of S 1 and T T1 have no common point. But now the sum 1′ − S 1, T T1 + S 1, T T2 is equal, by linearity, to S, 0 = 0. So 1′ − 2′ − S 1, T T2 = 0. This proves (i). (ii) is trivial because if S is a form we 2′ − may take the decomposition S = 0+S ; T = T +0. Now we have proved the existence of the expression when the sets of singularities of S and T have no common point. (We did not pay any attention to the question of supports, assuming the manifold to be compact; if the manifold is not compact, we have to assume further that the intersection of the supports of S and T is compact). That the sets of singularities of S and T have 173 no common point can be expressed in the following way: every point of the manifold has an open neighbourhood in which one at least of the currents S and T (not necessarily the same one) is a form. In this case the properties of differentiation such as the formula

d S, T = ( 1)p+1 S, d T z − z (S is of total degree p) remain obviously valid as is seen immediately by means of a decomposition S i, Ti of S , T. In our case since the sets 152 Lecture 25

of singularities of U and V have no common point

d U, V = d V, U − z z or S,∂ GT = T,∂ GS − z z Thus the systems (I) and (II) are transposes of each other. Two systems which are transposes of each other have the same rank ρ, ρ = number of unknowns - degree of indeterminacy. So we have

0 1 (A + 1) m( D) = (B + g) m(D) − − − or

0 1 m( D) m(D) = (A B) g + 1 − − − − = d g + 1. − This completes the proof of the Riemann-Roch theorem. Let D be a positive divisor, D = αp p, αp 0. All the differentials ≥ of degree 1 which are multiples of D are holomorphic. Now an holo- morphic differential form of degree 1 has exactly 2g 2 zeros. (This can − 174 be proved by using Poincare’s´ theorem on vector fields or deduced from 1 the Riemann Roch theorem). Consequently if αp > 2g 2, m(D) = 0 − and 0 m( D) = d g + 1(d > 2g 2) − − − i.e., if we give a sufficient number of poles the g conditions of compati- bility are all independent. Appendix

Kahlerian¨ structure on the complex projective space 175 The unit sphere S 2n+1 in Cn can be considered in a natural way as a fibre bundle over PCn with circles as fibres (if z is a point of S 2n+1 the points eiθz, θ real, will constitute the fibre through z). We shall first determine how J operates on the tangent spaces of PCn. Let π : S 2n+1 PCn denote the projection map. We shall denote → the differential of π also by π. The differential map π maps the tangent bundle of S 2n+1 onto that of PCn. Let X = πU, U tangent to S 2n+1, be a vector tangent to PCn, say at a. J is uniquely determined by:

JX, d f = i X, d f JX, d f = i X, d f − for every f holomorphic in a neighbourhood of a. If zn+1 , 0, zk/zn+1 is a local coordinate system; we may suppose that a belongs to the domain of this coordinate system. J is uniquely determined by:

JX, d(zk/zn+1) = i X, d(zk/zn+1) JX, d(zk/zn+1) = i X, d(zk/zn+1) . −

(u1,..., un) being the coordinates of U, let (iU) denote the vector whose coordinates are (iu1,..., iun). The vector π(iU) has the properties 176

π(iU), d(zk/zn+1) = iU, d(zk/zn+1) = i X, d(zk/zn+1) ; 153 154 Appendix

and π(iU), d(zk/zn+1) = i X, d(zk/zn+1) . − Therefore π(iU) = JX So, to find JX we multiply U by i and take its image by π. Consider now the form

ω = zkdzk of bidegree (1, 0) in Cn+1. Put

H = dzkdzk ωω. − This Hermitian form on Cn+1 induces on S 2n+1 a semi-definite Hermi- tian form. To prove that this form is semi-definite on S 2n+1, let U = 2n+1 2n+1 (u1,..., un) be a vector tangent to S at z S , z = (z1,..., zn). ∈ By Schwarz’s inequality

zkuk z U = U | | ≤ | | | | | | (since z is on the unit sphere). Similarly

zkuk U . | | ≤ | | So H(U, U) = ukuk zkuk zkuk 0. − ≥ Moreover H(U, U) = 0 if and only if uk = zk, complex. Actually = iλ, where λ is real. For, since z is on S 2n+1 and U is tangent to the unit sphere at z (zkuk + zkuk) = 0 177 or zkzk ( + ) = 0 or Rl = 0. Thus H(U, U) = 0 if and only if U = iλz, λ real i.e., if and only if U is tangent to the fibre, that is if π(U) = 0. If U is tangent to the fibre at z and V = (v1,..., vn) is any vector tangent to S 2n+1 at z, H(U, V) = 0. For by Schwarz’s inequality

H(U, V) H(U, U)H(V, V) = 0 | |≤ Appendix 155

Moreover H is invariant under the operations z eiθz, θ real. These → two facts imply that H defines quadratic forms H on tangent spaces of PCn (if X and Y are tangent to PCn at a we choose vectors U and V 2n+1 1 tangent to S at some point in π− (a) such that πU = X and πV = Y and define H(X, Y) = H(U, V). The two properties proved above imply that this definition defines H invariantly). H are J Hermitian forms. i) Evidently H is R-bilinear

ii) H(JX, Y) = H(X, JY) = iH(X, Y) − For, H(JX, Y) = H(iU, V) = iH(X, Y) as we have seen that JX = π(iU). Similarly

H(X, JY) = iH(X, Y) − iii) H˜ (X, X) > 0 for X , 0. This follows from the fact that H(U, U) = 178 0 if and only if U is tangent to the fibre. It remains to prove that the exterior form Ω associated with H is closed. If Ω is the form associated with H,

2iΩ= dzk dzk ω ω − ∧ − ∧ = dω ω ω − − ∧ Ω= π 1Ω. But on S 2n+1, ω = ω; for, the relation − − zkzk = 1 yields (dzkzk + zkdz) = 0 Consequently 2iΩ= dω; so Ω is closed. − Now dΩ is the reciprocal image of dΩ; therefore Ω is also closed. This proves that PCn is Kahlerian.¨ 156 Appendix

It may be remarked that even though Ω is a coboundary, Ω is not a coboundary, (as PCn is Kahlerian!);¨ ω is not a reciprocal image. Another method to prove that PCn is Kahlerian¨ would be to consider PCn as symmetric space with respect to the unitary group, Un+1. Since Un+1 is compact we may construct an invariant Hermitian metric by the averaging process. The associated 2-form Ω will be an invariant form. But in a symmetric space any invariant form is closed.

BIBLIOGRAPHY

179 We do not give here an exhaustive bibliography on the subject. We mention some articles and books which may help the reader of these notes to obtain some additional information on the subject or which con- tain other methods of exposition. Differentiable manifolds may be studied in (7), (18); differential forms, exterior differentiation and Stokes’ formula, in (7), (15), (18); Poincare’s´ theorem on differential forms in RN, in (3. p.72), (16), (18); Currents and distributions, in (18), (19); de Rham’s theorems and Poincare’s´ duality theorem, in (5-a), (16), (18), (27); harmonic forms, in (10), (11), (17), (18); elliptic character of ∆, in (18); Hahn-Banach theo- rem and Banach’s closed graph theorem, in (2); Kahlerian¨ manifolds in (5-b), (11), (12), (17), (24), (26), (28); the osculating Euclidean metric in (4. p. 90), (5-b); Cousin’s problem, in (5-b), (17), (21), (22), (25); z cohomology, in (9), (12), (13), (21), (22), (24); Riemann surfaces and the Riemann-Roch theorem, in (8), (14), (21), (29). Bibliography

[1] S. Bochner and W. T. Martin: Several Complex Variables, Prince- 180 ton University Press, 1948.

[2] N. Bourbaki: Espaces vectoriels topologiques, Paris, Hermann, 1953.

[3] E. Cartan: Lec¸ons sur les invariants integraux,´ Paris, Gauthiers- Villars, 1922.

[4] E. Cartan: Lec¸ons sur la geom´ etrie´ des espaces de Riemann, Paris, Gauthiers-Villars, 1946.

[5] H. Cartan: Seminaire´ de l’Ecole Normale Superieure:

(a) 1950-51 (b) 1951-52 (c) 1953-54

[6] H. Cartan and J. P. Serre: Un theor´ eme` de finitude concernant les variet´ es´ analytiques compactes, Comptes Rendus de l’Academie´ des Sciences (Paris), 237(1953) pp. 128-130.

[7] C. Chevalley: Theory of Lie Groups, Princeton University Press, 1946.

[8] C. Chevalley: Introduction to the theory of algebraic functions of one variable, Americal Mathematical Society, 1951.

157 158 BIBLIOGRAPHY

[9] P. Dolbeault: Sur la cohomologie des variet´ es´ analytiques com- plexes, Comptes Rendus de l’Academie´ des Sciences, Paris, 236 (1935), pp. 175-177; pp. 2203-2205.

[10] G. F. D. Duff and D. C. Spencer: Harmonic tensors on Riemannian manifolds with boundary, Annals of Mathematics, 56(1952) pp. 128-156.

181 [11] W. V. D. Hodge: The theory and applications of harmonic integrals, Cambridge University Press, 1952.

[12] W. V. D. Hodge: Differential forms on Kahler¨ manifolds, Proceed- ings of the Cambride Philosophical Society, 47 (1951), pp. 504- 517.

[13] K. Kodaira: On a differential geometric method in the theory of analytic stacks, Proceedings of the National Academy of Sciences (Washington), 39 (1953), pp. 1268-73.

[14] K. Kodaira: Some results in the transcendental theory of algebraic varieties, Annals of Mathematics, 59 (1954), pp. 86-134.

[15] A. Lichnerovicz: Algebre` et Analyse Lineaires,´ Paris, Masson, 1947.

[16] G. de Rham: Sur l’analysis situs des variet´ es´ a` n dimensions, Jour- nal de Mathematiques´ pures et appliquees,´ 10 (1931), 115-120.

[17] G. de Rham and K. Kodaira: Harmonic Integrals, Institute for Ad- vanced Study, Princeton, 1950.

[18] G. de Rham: Variet´ es´ differentiables´ (Formes, courants, formes harmoniques), Paris, Hermann, 1955.

[19] L. Schwartz: Theorie des Distributions, Vol. I, II, Paris, Hermann, 1950-51.

[20] L. Schwartz: Homomorphismes et applications completement` con- tinues, Comptes Rendus de l’Academie´ des Sciences (Paris), 236 (1953), pp. 2472-74. BIBLIOGRAPHY 159

[21] L. Schwartz: Courant attache´ a` une forme differentielle´ meromor- 182 phe sur une variet´ e´ analytique complexe, Colloque de Geom´ etrie´ Differentielle,´ Strasbourg, 1953, pp. 185-195.

[22] J. P. Serre: Quelques problemes` globaux relatifs aux variet´ es´ de Stein, Colloque sur les functions plusieurs variables complexes, Bruxelles, 1953, pp. 57-68.

[23] J. P. Serre: Un theor´ eme` de dualite,´ Commentarii Mathematici Helvetici, 29 (1955), pp. 9-26.

[24] D. C. Spencer: Real and complex operators on Manifolds, Con- tributions to the theory of Riemann surfaces, Princeton University Press, 1951, pp. 203-227.

[25] K. Stein: Analytische Funktionen mehrerer kumplexer Veran- derlichen zu vorgegebenen Periodizitatsmoduln¨ und das zweite Cousinsche Problem, Mathematische Annalen, 123 (1951), pp. 201-222.

[26] A. Weil: Sur la theorie´ des formes differentialles´ attachees´ a` une variet´ e´ analytique complexe. Commentarii Mathematici Helvetici, 20 (1947), pp. 110-116.

[27] A. Weil: Sur les theor´ emes` de de Rham, Commentarii Mathematici Helvetici, 26 (1952), pp. 119-145.

[28] A. Weil: Theorie der Kahlerschen¨ Mannigfaltigkeiten, Universitat¨ Gottingen,¨ 1953.

[29] H. Weyl: Die Idee der Riemannschen Flache,¨ Leipzig, 1913 and 1923; Stuttgart, 1955.